ArticlePDF Available

Crystal preferred orientations of olivine, orthopyroxene, serpentine, chlorite, and amphibole, and implications for seismic anisotropy in subduction zones: a review

Authors:

Abstract and Figures

This study provides a comprehensive review of the crystal preferred orientation (CPO) of olivine and orthopyroxene in the upper mantle, and of several hydrous minerals in the mantle wedge and at the slab-mantle interface. It discusses the seismic anisotropy of those minerals. Water-induced CPOs of olivine produced by previous experimental studies under high pressure and temperature conditions were found in many natural rocks. It is emphasized that the strong CPOs of hydrous minerals such as serpentine, chlorite, and amphibole, play an important role in interpreting the anomalously strong seismic anisotropy observed in subduction zones.
Content may be subject to copyright.
Review
Vol. 21, No. 6, p. 9851011, December 2017
http://dx.doi.org/10.1007/s12303-017-0045-1
pISSN 1226-4806 eISSN 1598-7477
Geosciences Journal
GJ
Crystal preferred orientations of olivine, orthopyroxene,
serpentine, chlorite, and amphibole, and implications
for seismic anisotropy in subduction zones: a review
Haemyeong Jung*
Tectonophysics Laboratory, School of Earth and Environmental Sciences, Seoul National University, Seoul 08826, Republic of Korea
ABSTRACT: This study provides a comprehensive review of the crystal preferred orientation (CPO) of olivine and orthopyroxene
in the upper mantle, and of several hydrous minerals in the mantle wedge and at the slab-mantle interface. It discusses the seismic
anisotropy of those minerals. Water-induced CPOs of olivine produced by previous experimental studies under high pressure and
temperature conditions were found in many natural rocks. It is emphasized that the strong CPOs of hydrous minerals such as ser-
pentine, chlorite, and amphibole, play an important role in interpreting the anomalously strong seismic anisotropy observed in sub-
duction zones.
Key words: crystal preferred orientation, olivine, orthopyroxene, serpentine, chlorite, amphibole, seismic anisotropy
Manuscript received June 17, 2017; Manuscript accepted September 27, 2017
1. INTRODUCTION
Rocks inside the Earths crust and mantle are deformed under
differential stress at high pressure and temperature conditions,
and those in the deep interior go through plastic deformation.
The minerals constituting rocks are deformed and elongated,
forming a crystal preferred orientation (CPO) or a lattice preferred
orientation (LPO). If a mineral is elastically very anisotropic, the
CPO of its aggregates can cause a significant seismic anisotropy
in the crust and the mantle. Seismic anisotropy has been observed
worldwide in the Earths interior (Fig. 1) and can provide important
information for understanding the evolution of the upper
mantle, mantle flow pattern, mantle dynamics, and tectonics
(Hess, 1964; Nicolas and Christensen, 1987; Silver, 1996; Ben
Ismail and Mainprice, 1998; Savage, 1999; Jung and Karato,
2001; Smith et al., 2001; Jung et al., 2006; Mainprice, 2007;
Karato et al., 2008; Long and Silver, 2008; Long and Silver, 2009;
Di Leo et al., 2012; Jung, 2012; Long, 2013; Tommasi and Vauchez,
2015; Skemer and Hansen, 2016; Zhao et al., 2016). Therefore,
studying the CPO of elastically anisotropic minerals that are
dominant in the crust and the upper mantle is essential.
In the upper mantle, the dominant minerals inside rocks are
olivine and orthopyroxene (Fig. 2a). Olivine is the most abundant
mineral and is elastically anisotropic (Birch, 1960; Verma, 1960;
Abramson et al., 1997) (Table 1). The CPO of olivine is key to
understanding the seismic anisotropy and flow pattern of the
upper mantle (Hess, 1964; Nicolas and Christensen, 1987; Ben
Ismail and Mainprice, 1998; Jung et al., 2006; Karato et al., 2008;
Cao et al., 2015; Michibayashi et al., 2016; Cao et al., 2017). In
the mantle wedge and at the slab-mantle interface, hydrous
minerals such as serpentine, chlorite, and amphibole (Figs. 2b–
d) can form from the fluids generated by the dehydration of
hydrous minerals in the subducting slab (Ulmer and Trommsdorff,
1995; Peacock and Hyndman, 1999; Pawley, 2003; Fumagalli
and Poli, 2005; van Keken et al., 2011). Many of hydrous minerals
such as chlorite, amphibole, and serpentine can be formed after
olivine meets water during exhumation in the mantle wedge
and at the interface between slab and mantle wedge. Those
hydrous minerals are known to be stable at a wide range of
pressure-temperature conditions in subdution zone (Schimidt
and Poli, 1998; Ulmer and Trommsdorf, 1995; Fumagali and
Poli, 2005) (Fig. 3). Those hydrous minerals are also known as
elastically anisotropic minerals (Aleksandrov and Ryzhova, 1961a;
Mainprice and Ildefonse, 2009; Bezacier et al., 2010; Mookherjee
*Corresponding author:
Haemyeong Jung
Tectonophysics Laboratory, School of Earth and Environmental Sci-
ences, Seoul National University, Seoul 08826, Republic of Korea
Tel: +82-2-880-6733, Fax: +82-2-871-3269, E-mail: hjung@snu.ac.kr
©
The Association of Korean Geoscience Societies and Springer 2017
986 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
Fig. 1. Examples of seismic anisotropy observed worldwide. This figure shows a summary map of mantle wedge splitting of S-waves (Long
and Wirth, 2013). Arrows indicate the first‐order patterns in average fast direction; where multiple arrows are present, this indicates a spatial
transition in observed orientation (φ). Arrows are color coded by fast direction observations; magenta arrows indicate dominantly trench‐par-
allel φ, blue arrows indicate dominantly trench‐perpendicular φ, yellow arrows indicate complex and variable φ, red arrows indicate a tran-
sition from trench‐parallel φ close to the trench to trench‐perpendicular φ farther away, and green arrows indicate the opposite transition
(from trench‐perpendicular φ close to the trench to trench‐parallel φ farther away). Beneath the name of each subduction zone, the range
of observed delay times is indicated.
Tab le 1 . Seismic (elastic) anisotropy of olivine and orthopyroxene (opx)
Mineral Single crystal Poly crystal Reference
AVp (%) max. AVs (%) AVp (%) max. AVs (%)
Olivine (forsterite)
23.8 16.0 Mainprice (2007)
(a)
4.9–11.1 4.9–7.4 Jung and Karato (2001)
(a)
8.3–13.3 5.3–9.3 Michibayashi et al. (2006)
(a)
2.7–6.0 1.8–4.8 Katayama and Karato (2006)
(a)
2.6 2.9 Skemer et al. (2006)
(a)
12.0 7.0–10.4 Hidas et al. (2007)
(a)
4.5–9.8 3.2–7.0 Tasaka et al. (2008)
(a)
7.8–13.6 5.4–8.2 Jung et al. (2009a)
(a)
3.2–7.2 2.2–5.5 Jung (2009)
(a)
1.8–7.3 2.2–5.5 Ohuchi et al. (2011)
(a)
6.1–8.6 3.9–5.6 Michibayashi et al. (2012)
(a)
2.7–5.9 2.5–5.0 Jung et al. (2013)
(a)
8.8–8.9 5.3–5.7 Park et al. (2014)
(a)
1.2–3.6 2.1–3.7 Watanabe et al. (2014)
(a)
2.2–11.6 1.9–7.5 Park and Jung (2015)
(a)
1.8–3.8 1.7–2.7 Kim and Jung (2015)
(a)
5.2–9.1 8.4–8.6 Boneh et al. (2015)
(a)
11.0–13.9 10.5–14.2 Lee and Jung (2015)
(a)
1.8–7.5 1.1–5.2 Kang and Jung (2017)
(a)
Opx (enstatite)
13.7 17.6 Mainprice (2007)
(b)
2.9 2.7 Skemer et al. (2006)
(b)
1.8–5.9 2.4–4.0 Jung et al. (2010)
(b)
3.4 3.9 Jung et al. (2013)
(b)
1.2–2.3 1.5–2.5 Park and Jung (2015)
(b)
AVp: anisotropy of P-wave velocity. AVp = 100 (%) × [(Vpmax – Vpmin)/(0.5 × (Vpmax + Vpmin))], where Vpmax is the maximum P-wave velocity
and Vpmin is the minimum P-wave velocity (Birch, 1960).
AVs: anisotropy of S-wave velocity. AVs = 100 (%) × [(Vs1 – Vs2)/(0.5 × (Vs1 + Vs2))], where Vs1 is the fast S-wave velocity and Vs2 is the slow
S-wave velocity.
(a)
Elastic constants of olivine(forsterite) from Abramson et al. (1997) were used.
(b)
Elastic constants of orthopyroxene(enstatite) from Chai et al. (1997) were used.
Crystal preferred orientations of olivine and hydrous minerals 987
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
and Mainprice, 2014) (Table 2); their CPOs are important for
interpreting seismic anisotropy in many subduction zones
(Katayama et al., 2009; Jung, 2011; Brownlee et al., 2013; Morales
et al., 2013; Kim and Jung, 2015; Ko and Jung, 2015; Nagaya et
al., 2016; Kang and Jung, submitted). In addition, one of the
important minerals in the middle and lower crust is amphibole,
which is also elastically anisotropic (Aleksandrov and Ryzhova,
1961a; Siegesmund et al., 1989; Weiss et al., 1999; Kaczmarek
and Tommasi, 2011; Lloyd et al., 2011; Ji et al., 2013; Ko and
Jung, 2015; Brown and Abramson, 2016; Almqvist and Mainprice,
2017). A recent experimental study under simple shear at high
pressure and temperature showed that the CPO of amphibole
(hornblende) might cause large seismic anisotropy (Ko and
Jung, 2015). Amphibole and chlorite in hydrated peridotites in
the lower part of the mantle wedge also play a key role in the
interpretation of seismic anisotropy in subduction zones (Kim
and Jung, 2015; Kang and Jung, submitted).
This work provides a comprehensive review of the experimental
studies on the CPO of olivine first. The effect of water, stress,
temperature, pressure, shear strain, and deformation history on
the CPO of olivine is presented along with the recent findings
on the CPO of olivine in natural rocks. Next, the CPOs of
orthopyroxene and serpentine are described, followed by a review
of the recent advances in the study of the CPO of hydrous minerals
including chlorite and hornblende. Finally, geophysical implications
of the results of the recent studies on the CPOs of hydrous minerals
Fig. 2. Optical photomicrographs showing deformation microstructures of natural rocks (a–e) and an experimental sample (f). (a) Mantle
xenolith (SVF-04) from Svalbard, Arctic showing olivine and orthopyroxene (Jung et al., 2009a). White arrows indicate Kink bands in olivine.
Scale bar represents 2 mm. (b) Serpentinite (VM3) from Val Malenco, Italy (Jung, 2011). Scale bar represents 0.2 mm. (c) Chlorite peridotite
(436) from Amklovdalen, SW Norway (Kim and Jung, 2015). Scale bar represents 1 mm. (d) Amphibole peridotite from Bjorkedalen, SW Nor-
way (Kang and Jung, submitted). Scale bar represents 0.5 mm. (e) Yugu peridotite (YG-10) showing ultramylonite texture in Yugu, South
Korea (Park and Jung, 2017). Scale bar represents 3 mm. (f) A back-scattered electron image of a deformed amphibolite at a high pressure
of 1 GPa and temperature of 600 °C (Ko and Jung, 2015). Arrows show the sense of shear. Ol: olivine, Opx: orthopyroxene, Ant: antigorite,
Mgt: magnetite, Chl: chlorite, Amp: amphibole, Sp: spinel, Cpx: clinopyroxene, Srp: serpentine, Hb: hornblende, Pl: plagioclase, and Il: ilmenite.
988 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
and olivine, and the resultant seismic anisotropy are discussed.
2. STUDIES ON THE SLILP SYSTEM AND CPOs
OF OLIVINE
Early experiments on the deformation of single crystal olivine
at high pressure were performed by Rayleigh (Raleigh, 1965, 1968)
under uniaxial compression and identified active slip systems of
olivine by observing detailed microstructures such as kink bands,
slip bands, and deformation lamellae. Slip system is defined by a
slip plane and a slip direction (i.e., (010)[100] slip system refers
to a slip occurring at the (010) plane and along the [100] direction).
At low temperatures (less than 1,000 °C), the slip systems of
olivine were {110}[001], (100)[001], and (100)[010], while at the
Tab l e 2. Seismic (elastic) anisotropy of hydrous minerals.
Mineral Single crystal Poly crystal Reference
AVp (%) max. AVs (%) AVp (%) max. AVs (%)
Serpentine (antigorite)
71.2 67.6 Mainprice and Ildefonse (2009)
(a)
46.0 66.5 Bezacier et al. (2010)
(b)
38.4–46.3 24.2–32.4 Katayama et al. (2009)
(c)
10.4–31.3 8.9–36.0 Hirauchi et al. (2010)
(b)
36.8 50.5 Bezacier et al. (2010)
(b)
32.9 31.0 Soda and Takagi (2010)
(b)
23.6–31.4 22.8–36.5 Jung (2011)
(b)
7.6–8.2 14.0–36.2 Nishii et al. (2011)
(b)
5.7–27.6 4.9–33.2 Brownlee et al. (2013)
(b)
13.9–25.7 15.0–24.0 Watanabe et al. (2014)
(b)
34.2 - Nagaya et al. (2016)
(b)
Chlorite (clinochlore)
35.5 76.2 Mainprice and Ildefonse (2009)
(d)
35.5 76.2 Kim and Jung (2015)
(d)
14.9–21.1 14.5–31.7 Kim and Jung (2015)
(d)
10.3–15.2 10.6–22.8 Kim and Jung (2015)
(e)
22.3–25.2 31.6–46.2 Kang and Jung (2017)
(d)
Amphibole (hornblende)
27.2 - Kitamura (2006)
(f)
27.1 30.7 Mainprice and Ildefonse (2009)
(f)
27.1 30.7 Lloyd et al. (2011)
(f)
27.1 30.7 Ko and Jung (2015)
(f)
9.5–11.1 7.0–8.0 Siegesmund and Vollbrecht (1991)
(g)
11.4 - Barruol and Kern (1996)
(f),(g)
3.2–7.7 - Kitamura (2006)
(f),(g)
3.6–6.0 3.8–6.9 Tatham et al. (2008)
(f),(g)
7.5–14.0 4.2–8.5 Ji et al. (2013)
(g),(h)
10.2–13.5 6.9–11.2 Jung et al. (2014a)
(f)
9.0–14.6 7.6–12.1 Ko and Jung (2015)
(f)
5.7–14.0 4.1–8.5 Ji et al. (2015)
(g),(i)
11.4 6.9 Lamarque et al. (2016)
(f)
10.0–15.2 7.5–11.9 Kang and Jung (2017)
(f)
AVp: anisotropy of P-wave velocity. AVp = 100 (%) × [(Vpmax – Vpmin)/(0.5 × (Vpmax + Vpmin))], where Vpmax is the maximum P-wave velocity
and Vpmin is the minimum P-wave velocity (Birch, 1960).
AVs: anisotropy of S-wave velocity. AVs = 100 (%) × [(Vs1 – Vs2)/(0.5 × (Vs1 + Vs2))], where Vs1 is the fast S-wave velocity and Vs2 is the slow
S-wave velocity.
(a)
Elastic constants of serpentine (antigorite) from Pellenq et al. (submitted) were used.
(b)
Elastic constants of serpentine (antigorite) from Bezacier et al. (2010) were used.
(c)
Elastic constants of serpentine (lizardite) from Auzende et al. (2006).
(d)
Elastic constants of chlorite (clinochlore) from Alexandrov and Ryzhova (1961b).
(e)
Elastic constants of chlorite (clinochlore) from Mookherjee and Mainprice (2014).
(f)
Elastic constants of amphibole (hornblende) from Alexandrov and Ryzhova (1961a).
(g)
Seismic anisotropy of whole rock, amphibolite.
(h)
Elastic constants of amphibole(hornblende) from Hearmon (1984).
(i)
Elastic constants of amphibole(hornblende) from Alexandrov et al. (1974).
Crystal preferred orientations of olivine and hydrous minerals 989
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
high temperature of 1,000°C, the slip system was {0kl}[100],
which is the pencil glide in [100] direction (Raleigh, 1968). Later
experiments at higher pressure (up to P = 3 GPa) and temperature
(up to 1,400 °C) conditions (Carter and Avé Lallemant, 1970)
identified the slip system of olivine as (010)[100] at temperatures in
excess of 1,100 °C.
The first experimental study to determine the CPO of olivine
aggregates was conducted in the uniaxial compression mode
using a Griggs apparatus (Fig. 4a) under dry conditions (A
Lallemant and Carter, 1970). Olivine [010] axes were found to
be aligned subparallel to the maximum principal stress (strain
axis) orientation. Both [100] and [001] axes were aligned in a
girdle subnormal to the maximum principal stress (strain axis)
orientation. In 1995, the first simple shear deformation experiment
of olivine aggregates at the pressure of 300 MPa and temperatures of
1,200–1,300 °C was conducted using a gas-medium Paterson
apparatus under dry conditions (Zhang and Karato, 1995) and
mor e exp erimental data were pub lish ed la ter ( Zhang et a l., 200 0).
These simple shear experiments showed that olivine [010] axes
were aligned subnormal to the shear plane and [100] axes were
aligned subparallel to the shear direction under dry conditions.
2.1. Effect of Water and Stress on the CPOs of Olivine
2.1.1. Experimental study
Water has been known to affect the CPO of olivine since the
pioneering experimental study (Jung and Karato, 2001). To
understand the effect of water and stress on the CPO of olivine,
deformation experiments were conducted on olivine aggregates
Fig. 3. Stability field of hydrous minerals
in subduction zone. Stippled lines are
isotherms by thermal model (Furukawa,
1993) and bold arrows indicate flow
lines in the mantle wedge. Blue line
indicates the stability boundary of chlo-
rite in peridotite. Example of strong
CPO of chl (Kim and Jung, 2015), serp
(Jung, 2011), and amp (Ko and Jung,
2015) is shown, which is important for
the interpretation of anomalously
strong seismic anisotropy in subduction
zones. chl: chlorite, amp: amphibole
(hornblende), serp: serpentine (antigor-
ite), opx: orthopyroxene, cld: chloritoid,
lws: lawsonite, and phg: phengite. Mod-
ified after Schmidt and Poli (1998).
Fig. 4. Equipment used for high-temperature, high-pressure defor-
mation experiments on minerals and rocks. (a) Griggs apparatus (2
GPa), and (b) modified Griggs apparatus (5 GPa) at the Tectonophys-
ics Laboratory, Seoul National University, South Korea.
990 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
in simple shear (Fig. 5), using a Griggs apparatus (Fig. 4a) at high
pressure (0.3–2.1 GPa) and temperature (1,190–1,300 °C), discovering
olivine Type-B and -C CPOs (Fig. 6a) and defining olivine Type-A,
-B, -C, and -D CPOs (Jung and Karato, 2001). The starting materials
were San Carlos olivine aggregates and single crystal olivine that
contained 10% Fe, as a representative sample of the upper mantle.
The sample was hot-pressed at 300 MPa and 1,200 °C. Experiments
were conducted under both dry and wet conditions. In wet
conditions, water was added to the sample by the dehydration of
the mixture of talc and brucite at high temperature. Sample
assembly for shear deformation experiment is shown in Figure
5. The sample was deformed at constant strain rates (9.5 × 10
–4
5.6 × 10
–6
s
–1
) with a shear strain of 0.6–6.3 (Jung and Karato,
2001; Katayama et al., 2004; Jung et al., 2006). Dry conditions
were defined as a water content of less than 200 ppm H/Si in single
crystal olivine without any cracks, inclusions, or grain boundaries
in the sample.
The representative CPOs of olivine produced in a simple
shear at high pressures under both dry and wet conditions are
summarized in Figure 6a and a fabric diagram of olivine is
shown in Figure 6b. Several types of olivine CPOs were found:
Type-A CPO of olivine was formed under low-stress and dry
conditions; water content in the single crystal olivine was less
than 200 ppm H/Si (Jung and Karato, 2001). The Type-A CPO
of olivine is characterized by olivine [010] axes aligned subnormal
to the shear plane and [100] axes aligned subparallel to the shear
direction, having a dominant slip system of (010)[100]. Type-B
CPO of olivine was found under high-stress and varied water
content (200 C
OH
1200 ppm H/Si) conditions and characterized
as [010] axes aligned subnormal to the shear plane and [001] axes
aligned subparallel to the shear direction, having a dominant
slip system of (010)[001] (Jung and Karato, 2001). Type-C CPO
of olivine was observed under low-stress and water-rich conditions
with a water content of C
OH
700 ppm H/Si and characterized
Fig. 5. Central portion of sample assembly for a shear deformation
experiment at high pressure and temperature using a Griggs apparatus.
Fig. 6. (a) Typical CPOs of olivine showing Type-A, -B, -C, -D, -E, and
-AG (modified from Jung et al., 2006). Type-A (MIT23), Type-B (JK21),
Type-C (JK11), Type-D (SVF-49: Jung et al., 2009a), Type-E (GA25),
and Type-AG (Jung et al., 2014b). Pole figures are presented in the
lower hemisphere using an equal area projection. The sense of shear
is presented by arrows. The north (south) poles correspond to the
shear plane normal. Crystallographic orientations of ~200–650
grains of each sample were measured manually by the EBSD tech-
nique. A half-width of 30° was used to draw the pole figure. The
color coding refers to the density of data points (contours in the pole
figures correspond to the multiples of uniform distribution). (b) A
fabric diagram of olivine at P = 0.3–2.1 GPa and T ~ 1200 °C showing
the dominant fabrics as a function of water content (C
OH
) and stress
(after Jung et al., 2006). σ: differential stress, μ: shear modulus.
Crystal preferred orientations of olivine and hydrous minerals 991
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
as [100] axes aligned subnormal to the shear plane and [001]
axes aligned subparallel to the shear direction, having a dominant
slip system of (100)[001] (Jung and Karato, 2001). Type-D CPO
of olivine was formed under high-stress and dry conditions
and characterized as [100] axes aligned subparallel to the
shear direction and both [010] and [001] axes aligned as a
girdle subnormal to shear direction, having a dominant slip
system of {0kl}[100] (Bystricky et al., 2000; Zhang et al., 2000).
Finally, Type-E CPO of olivine was found under low-stress
and moderate water content (200 C
OH
< 700 ppm H/Si)
conditions and characterized as [001] axes aligned subnormal
to the shear plane and [100] axes aligned subparallel to the
shear direction, having a dominant slip system of (001)[100]
(Katayama et al., 2004; Jung et al., 2006).
Based on the study of natural rocks, one more CPO of
olivine, Type-AG (Mainprice, 2007; Michibayashi et al., 2016),
sometimes called as axial [010] pattern (Tommasi and
Vauchez, 2015), was reported. It is characterized by [010] axes
strongly aligned subnormal to the foliation and both [100] and
[001] axes aligned subparallel to the foliation (Fig. 6a). The
Type-AG CPO of olivine was also called type A + B CPO ( type
A + type B CPO) because it can be formed under a mixed
condition (deformed in dry condition and later in wet condition,
vice versa) (Jung et al., 2014b).
2.1.2. Observation of water-induced CPOs of olivine
in natural rocks
Early studies of olivine CPOs in natural rocks were conducted
by Nicolas and Christensen (1987) and Ben Ismail and Mainprice
(1998). Since water-induced CPOs of olivine were discovered in
experimental studies (Jung and Karato, 2001; Katayama et al.,
2004; Jung et al., 2006), there has been active research on the
petrofabrics of olivine in natural rocks (Table 3). Mizukami et al.
(2004) reported on the water-induced Type-B CPO of olivine in
the peridotites in Higashiakaishiyama, southwest Japan. Frese et
al. (2003) found Type-C CPOs of olivine in prograde garnet
peridotites in Cima di Gagnone, in the Central Alps, and reported
that the fabrics formed at high water activity. Skemer et al. (2006)
also reported the Type-B CPO of olivine in the peridotites in
Cima di Gagnone in the presence of water. Tasaka et al. (2008)
reported water-induced Type-B CPO in highly depleted dunites
from the Imono peridotite body within the subduction-type
Sanbagawa metamorphic belt in southwest Japan. Jung (2009)
found both Type-B and Type-E CPOs of olivine in serpentinized
peridotites in Val Malenco, Italy. Skemer et al. (2013) reported
the water-induced Type-E CPO of olivine from Josephine
peridotites in southwest Oregon, in the United States. Michibayashi
and Oohara (2013) reported C- and E-type CPOs of olivine in a
hydrated ductile shear zone from the Fizh massif, Oman ophiolite.
Tab l e 3. Water-induced CPOs of olivine found in natural rocks
Fabric type Rock type Location Reference
B-type Dunite Higashiakaishiyama, SW Japan Mizukami et al. (2004)
Garnet peridotite Cima di Gagnone, Central Alps Skemer et al. (2006)
Dunite Southern Mariana Trench Michibayashi et al. (2007)
Depeleted dunite Snabagawa metamorphic belt, Japan Tasaka et al. (2008)
Serpentinized peridotite Val Malenco, Italy Jung (2009)
Mylonitic peridotite Bergen arc, Western Norway Jung et al. (2014b)
Spinel peridotite Shanwang, Eastern China Park & Jung (2015)
Chlorite peridotite Amklovdalen, WGR, Western Norway Kim & Jung (2015)
Mylonitic peridotite Navajo volcanic fiend, Colorado Plateau, USA Behr & Smith (2016)
Spinel lherzolitie El Aprisco, Calatrava volcanic field, Spain Puelles et al. (2016)
C-type Garnet peridotite Cima di Gagnone, Central Alps Frese et al. (2003)
Garnet peridotite Otroy Island, WGR, Western Norway Katayama et al. (2005)
Dunite Fizh massif, Oman ophiolite Michibayashi & Oohara (2013)
Garnet peridotite North Quidam UHP belt, NW China Jung et al. (2013)
Spinel peridotite Adam's Diggings, Rio Grande rift, USA Park et al. (2014)
Spinel lherzolitie El Aprisco, Calatrava volcanic field, Spain Puelles et al. (2016)
E-type Serpentinized peridotite Val Malenco, Italy Jung (2009)
Harzburgite Josephine peridotite, SW Oregon, USA Skemer et al. (2013)
Dunite Fizh massif, Oman ophiolite Michibayashi & Oohara (2013)
Mylonitic peridotite Bergen arc, Western Norway Jung et al. (2014b)
Spinel peridotite Shanwang, Eastern China Park & Jung (2015)
Chlorite peridotite Amklovdalen, WGR, Western Norway Kim & Jung (2015)
Spinel peridotite Yugu, South Korea Park & Jung (2017)
Mylonitic peridotite Ronda massif, Southern Spain Précigout et al. (2017)
992 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
Jung et al. (2013) found water-induced strong Type-C CPOs in
ultra-high-pressure (UHP) garnet peridotites from the North
Qaidam UHP collision belt in northwest China. In that study
the olivine contained water up to C
OH
= 1,130 ± 50 ppm H/Si, the
highest water content in olivine with the Type-C CPO in natural rock.
Park et al. (2014) reported Type-C CPOs of olivine in the
mantle xenoliths from the Adam’s Diggings at a rift shoulder in
the Rio Grande rift in the US, where spinel peridotites showed
evidence of rock deformation under wet conditions. Jung et al.
(2014b) found Type-B and -E CPOs of olivine in the peridotite
mylonites under a wet environment in the Bergen arc in western
Norway. Park and Jung (2015) reported Type-B and -E CPOs of
olivine in the spinel peridotites from mantle xenoliths in Shanwang,
eastern China. Kim and Jung (2015) studied chlorite peridotites
from Amklovdalen in the western gneiss region in western Norway
and reported both Type-B and Type-E CPOs of olivine. Behr
and Smit h (2016) re ported Type-B C POs of olivine in t he mantle
xenoliths from a mantle wedge setting in the Navajo volcanic
field in the Four Corners region of the Colorado Plateau, the
US, confirming that this fabric does form in natural subduction
zones. Czertowicz et al. (2016) reported dislocation glide on E-
type slip system in the coarse pre-mylonitised grains of olivine
under hydrous condition in Anita shear zone in New Zealand,
probably within hydrated sub-arc mantle lithosphere. Puelles et
al. (2016) reported both Type-B and -C CPOs of olivine in the
mantle xenoliths in El Aprisco in the Calatrava volcanic field in
central Spain. Recent study on microstructures of Pinatubo
peridotite xenoliths Yamamoto et al. (2017) found some dislocations
with the (001)[100] slip system in olivine which could have
been formed by the deformation under moderate water content
and low-temperature conditions. Park and Jung (2017) recently
reported Type-E CPOs of olivine in the ultramylonites (Figs. 2e
and 7) from the Yugu peridotite body in Yugu, South Korea.
Precigout et al. (2017) also found Type-E CPOs of olivine in the
mylonitic peridotites in the Ronda massif, Southern Spain.
2.2. Effect of Temperature and Pressure on the
CPOs of Olivine
2.2.1. Effect of temperature on the CPOs of olivine
Katayama and Karato (2006) investigated the effect of temperature
on the CPO of olivine at P = 2 GPa and T = 1,000–1,100 °C.
Samples were deformed at the constant strain rates of 3.8 × 10
–4
3.2 × 10
–5
s
–1
under water-saturated conditions using the Griggs
apparatus. The Type-B CPO of olivine occurred at higher stresses
than the Type-C CPO. In addition, the stress magnitude at which
the Type-B to -C CPO transition occurs decreased significantly
with decreasing temperature.
2.2.2. Effect of pressure on the CPOs of olivine
Couvy et al. (2004) studied the effect of pressure on the CPO
of olivine, performing relaxation experiments on forsterite (Mg
2
SiO
4
)
aggregates using a multi-anvil apparatus at P = 11 GPa and T =
1,400 °C. In that study, shear strain was limited to γ = 0.2 because
of the conditions of the experimental facility, but a weak Type-C
CPO of olivine was observed, suggesting that high pressure
makes Type-C fabric. However, this result is associated with the
following uncertainties: (1) the analysis of water content of olivine
Fig. 7. Microstructural evolution and CPO evolution of olivine in the Yugu peridotites in Yugu, South Korea (Park and Jung, 2017). From proto-
mylonite (PM) via mylonite (M) to ultra-mylonite (UM), both shear strain and water activity are increased whereas recrystallized grain-size of
olivine is decreased. Crystal preferred orientations (CPOs) of olivine are changed with the textural types of peridotites, from A-type (PM) via
D-type (M) to E-type (UM).
Crystal preferred orientations of olivine and hydrous minerals 993
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
after the experiments using Fourier Transformation Infrared
(FTIR) spectroscopy revealed that the samples contained a large
amount of water (1,500–2,500 ppm H/Si). Deformation of olivine
in wet conditions (C
OH
700 ppm H/Si) can produce Type-C
CPO of olivine (Jung and Karato, 2001); (2) stress was varied
during the experiment from 100 to 1,500 MPa because of the
stress-relaxation experiment. High stress favors [001] slip in olivine
(Carter and Avé Lallemant, 1970) and can induce changes in
olivine fabric (Jung et al., 2006; Katayama and Karato, 2006); (3)
the sample of forsterite used contained no Fe in olivine, whereas
olivine in nature is Fe-bearing in the upper mantle (~10%); and
(4) the shear strain was so small (γ = 0.2) that it is not certain
that the CPO represents the real fabric of the Earth where shear
strain is considered large.
More experimental studies of olivine aggregates at high pressure
were performed by exercising great care on the water content,
stress, and shear strain and pressure was found to induce changes
in the olivine CPO from Type-A to Type-B above 3 GPa (Jung et
al., 2009b). Deformation experiments of natural olivine aggregates
((Mg
0.9
Fe
0.1
)
2
SiO
4
) were conducted using a modified Griggs
apparatus at P =2.5–3.6 GPa and T = 1,300°C under dry conditions
(C
OH
90 ppm H/Si) with a large shear strain (γ=36). Samples
were d eforme d at cons tant s trai n rate s (2 × 10
–4
–6 × 10
–5
s
–1
) and
the CPO of olivine was measured using the SEM/EBSD (scanning
electron microscope/electron backscattered diffraction) technique
at the Seoul National University in South Korea. Olivine Type-B
CPO was developed at the pressures of 3.1 GPa and 3.6 GPa and
the temperature of 1,300 °C whereas Type-A CPO was formed
at the low pressure of 2.5 GPa with other conditions similar to
the higher-pressure experiments (Fig. 8) (Jung et al., 2009b). The
pressure-induced Type-B CPO of olivine occurred at the pressure
greater than 3 GPa. A subsequent study on the deformation of
olivine aggregates using the D-Dia (deformation-dia), a multi-
anvil high-pressure apparatus, confirmed that similar Type-B
CPO of olivine was formed at 5 GPa (Ohuchi et al., 2011).
Recently, more experiments were performed on the deformation
of olivine aggregates at high pressure under wet and dry conditions
(Ohuchi and Irifune, 2013, 2014; Ohuchi et al., 2015). To
Fig. 8. Pole figures of olivine showing the effect of pressure on the CPO of olivine (after Jung et al., 2009b). Arrows show the shear sense,
and north-south poles correspond to the shear plane normal. Contours correspond to multiples of a uniform distribution of data points. In
all three specimens represented here, the water content was undetectable. (a) T = 1300 °C, γ = 3, differential stress = 120 MPa. (b) T = 1300 °C,
γ = 3, differential stress = 150 MPa. (c) T = 1300 °C, γ = 6, differential stress = 390 MPa. A half-width of 20° was used to draw the pole figures.
MPD = maximum pole figure density.
994 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
compare the olivine CPOs in those studies, one needs to pay
attention to the measurement of olivine water content. The water
content of olivine (Jung and Karato, 2001; Jung et al., 2006) was
originally measured inside single crystal olivine without any
cracks, inclusions, or grain boundaries. Recently, Soustelle and
Manthilake (2017) deformed olivine-orthopyroxene aggregates
under simple shear at high pressures using the multi-anvil press
with shear strains of γ = 0.5–1.3. They found that the Type-B
CPO of oliv ine d evel oped at hig h pre ssure s of 3, 5 , and 8 GPa f or
olivine plus orthopyroxene content of 12.5% and 25% at the
temperatures of 1,300, 1,400, and 1,500 °C, respectively.
2.2.3. Natural example of pressure-induced CPOs of
olivine
Deformation fabrics of olivine from diamond-bearing garnet
peridotites in Finsch, South Africa were studied (Lee and Jung,
2015). They found strong Type-B CPOs of olivine in the samples
originated from a depth of 120 km (P = 4 GPa, T = 1,000 °C)
based on the analysis of the chemical composition of minerals
(thermobarometry data from the electron microprobe analysis)
and the existence of diamonds in garnet peridotites. Measurements
of the water content of both olivine and orthopyroxene using
FTIR spectroscopy revealed that both olivine and orthopyroxenes
are dry, indicating that strong Type-B CPOs of olivine were
formed under dry and high-pressure conditions. These observations
indicate that the strong Type-B CPOs in garnet peridotites of
Finsch, South Africa is the first natural example of pressure-induced
fabrics of olivine.
2.3. Effect of Large Shear Strain and Deformation
History on the CPOs of Olivine
2.3.1. Effect of large shear strain on the CPOs of olivine
The first shear deformation experiment of olivine aggregates
under dry conditions using the torsion apparatus for a large shear
strain (up to γ = 5) was conducted by Bystricky et al. (2000) at
the pressure of 300 MPa and 1,200 °C. They reported a Type-D
CPO of olivine at shear strains of γ = 2, 4, and 5. Later, Hansen et
al. (2012) deformed olivine aggregates (Fo
50
) for a large shear
strain (up to g ~10) using a torsion apparatus under dry conditions
at P = 250 MPa and 1200 °C. The Type-D CPO of olivine was
observed at a shear strain of γ ~3.5, but the Type-A CPO was
observed at a shear strain of γ ~10. To examine the systematics of
the evolution of olivine crystallographic fabrics at high strain,
Hansen et al. (2014) conducted more torsion experiments on olivine
aggregates (Fo
50
) under dry conditions at the pressures of 250
and 300 MPa and at 1200 °C with shear strains of up to γ ~19. A
steady-state fabric was not reached until the shear strain was
greater than 10. The authors showed that Type-D CPOs of olivine
developed with increasing strains up to γ = 9 and γ = 14, and
showe d Type-A CPOs at shear strains of γ = 7, 10, 11, and 19. These
results indicate that more study is probably needed to understand
the development of CPOs of olivine with large shear strain.
2.3.2. Effect of deformation history on the CPOs of
olivine
Recently, the effect of deformation history on the CPO of
olivine was studied by Boneh and Skemer (2014) using natural
samples of Åheim dunite that had a pre-existing CPO of olivine.
Deformation experiments were conducted using the Griggs
apparatus at 1 GPa and 1,200 °C with uniaxial strain (0.7). Samples
were deformed in three orientations, in which the shortening was
imposed perpendicular, oblique, and parallel to the pre-existing
foliation. The evolution of the CPO of olivine was distinct for
three initial orientations and none of the CPOs was observed to
reach steady state.
Numerical modeling studies on the olivine CPO development
were performed previously, producing a strong CPO of olivine
(Wenk and Tomé, 1999; Tommasi et al., 2000; Kaminski and
Ribe, 2001, 2002; Kaminski et al., 2004; Castelnau et al., 2008;
Castelnau et al., 2009; Castelnau et al., 2010). Recent numerical
studies on the development of olivine CPO using the dynamic
recrystallization-induced LPO (D-Rex) and the viscoplastic
self-consistent (VPSC) models (Boneh et al., 2015) identified
similar trends suggesting that a large strain is required to reach a
steady state CPO. In many cases, models initiated with a pre-
existing CPO require greater strain (about 3–5 times) to reach
steady state CPO than models initiated with a uniform CPO.
Signorelli and Tommasi (2015) modified a viscoplastic self-
consistent code to simulate the effects of subgrain rotation
recrystallization on the CPO of olivine. They found that the easy
slip system rotates fast towards parallelism with imposed shear
and that steady-state CPO (orientation and intensity) is reached
at shear strains of γ > 5.
2.4. CPOs of Olivine in Natural Shear Zones
CPOs of olivine in natural shear zones were studied in Josephine
Peridotite in southeastern Oregon, US (Warren et al., 2008;
Skemer et al., 2010; Hansen and Warren, 2015), in the Red Hills,
New Zealand (Webber et al., 2008), in the Hilti massif, Semail
ophiolite, Oman (Linckens et al., 2011), in an extensional shear
zone in the mantle, Lanzo massif, Italy (Kaczmarek and Tommasi,
2011), in an oblique-slip low-angle shear zone in the Beni Bousera
peridotite (Rif Belt, Morocco) (Frets et al., 2014), in the mylonitic
peridotites in the Ronda massif, Southern Spain (Precigout et
al., 2017), and in Yugu peridotites in South Korea (Fig. 7) (Park
and Jung, 2017). These studies suggested that the re-orientation
Crystal preferred orientations of olivine and hydrous minerals 995
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
of olivine in rocks with pre-existing textures requires large
strains. For example, a shear strain of γ > 4 was needed to re-orient
the CPOs of olivine with respect to the shear zone kinematics
(Hansen and Warren, 2015).
2.5. Other Possibilities for the Formation of the CPOs
of Olivine
Experimental study at the pressure of 300 MPa and temperature
of 1200 °C (Holtzman et al., 2003) reported, in the presence of
melt, a Type-B like CPO of olivine which is characterized as a
strong concentration of [010] axes aligned subnormal to the
shear plane and a weak girdle of both [100] and [001] axes aligned
subparallel to the shear plane. However, this type of melt-induced
CPO of olivine has been rarely reported in natural rocks. There
is also a report of the development of CPO of iron-free olivine
(forsterite) which was experimentally produced during diffusion
creep at the temperature of 1,200–1,350 °C (Miyazaki et al., 2013).
Fabric analysis of olivine in spinel peridotite xenoliths of Jeju Island
in South Korea (Yang et al., 2010) showed a weak activation of {0kl}
[100] slip system in the porphyrocalstic and mylonitic peridotites.
The results of the trace element analysis in that study revealed
that the smaller the grain size and weaker the fabric, the more
enriched in LREE and HFSE are the peridotites, indicating a strong
relationship between metasomatic agents and mantle shear zones.
A weak Type-C CPO of olivine was reported in Xugo UHP
garnet peridotites from the southern Sulu UHP terrane in China
(Wang et al., 2013b). The Type-C CPO of olivine was generated
by a dominant dislocation-accommodated grain boundary sliding
(DisGBS) with a minor contribution from the diffusion creep at
elevated pressure. Nagaya et al. (2014) reported that the Type-B
CPO of olivine can form as a result of the static topotactic growth
of olivine after the high-temperature breakdown of foliated
serpentinite. Wang et al. (2013a) reported Type-B CPO of olivine
in dunite from Raudkleivane, Amklovdalen and Type–C CPO
of olivine in garnet harzburgite from Ugelvik, Otroy in Norway.
Based on low water content in olivine, they interpreted that those
CPOs were produced by high stress or high pressure. However,
there is a high possibility that water in olivine was lost during
exhumation process because of high diffusivity of hydrogen in
olivine (Mackwell and Kohlstedt, 1990).
The Type-B CPO of olivine was also reported in peridotites
from the Ronda massif in Spain (Precigout and Hirth, 2014). The
Type-B CPO was suggested to be resulted from the enhancement
of the grain boundary sliding (GBS) with decreasing grain size.
The effect of finite strain geometry on CPO of olivine was
investigated using naturally deformed mantle rocks (Chatzaras
et al., 2016). They showed that finite strain geometry controls
the development of axial-type olivine CPO; axial-[010] and axial-
[100] CPOs form in relation to oblate and prolate fabric ellipsoids,
respectively. Cao et al. (2017) reported CPOs of olivine similar
to Types C and B in Songshugou peridotites and suggested that
those CPOs were formed by a diffusion-accommodated grain
boundary sliding (DifGBS) at high temperatures.
2.6. Scaling to the Nature
Because some of the conditions in experimental studies (e.g.,
the strain-rates) are different from those in the Earth, we need to
understand the scaling law for fabric transitions if experimental
results were to be applied to the Earth. This issue was already
discussed in previous study (Jung et al., 2006) and is summarized
below. Any fabric transition occurs when the rates of two
processes (e.g., strain-rates of two slip systems) become similar.
Therefore a generic equation to define a fabric boundary is
A
1
(T, P, C
OH
, σ) = A
2
(T, P, C
OH
, σ), (1)
where A
1
and A
2
are the rates of processes responsible for a
fabric development (e.g., strain-rate), T is the temperature, P
is the pressure, C
OH
is the water content, and σ is the stress.
Therefore, the transition conditions between two types of fabric
are given by a hyper-surface defined as,
f (T, P, C
OH
, σ) = 0. (2)
In most cases, the rates of these processes can be given by a set
of non-dimensional variables as
,
(3)
and consequently
. (4)
Therefore, the boundaries between different types of CPO
can be given by a hyper-surface in a multi-dimensional space
that does not include strain-rate explicitly. Therefore, the
fabric boundaries do not explicitly depend on strain-rates so
that the fabric boundaries determined by laboratory experiments
can be applied to the Earth’s interior where deformation occurs
at much slower strain-rates. The only difference between
laboratory and Earth is that because laboratory experiments
are conducted at much faster strain-rates than those in Earth,
laboratory studies can explore only limited range of parameter
space (e.g., relatively high stress regions).
3. CPOs OF ORTHOPYROXENE
Orthopyroxene is the second dominant mineral in the upper
A
1
T
T
m
P()
-------------- C
OH
σ
μPT,()
-----------------,,
⎝⎠
⎛⎞
A
2
T
T
m
P()
-------------- C
OH
σ
μPT,()
-----------------,,
⎝⎠
⎛⎞
=
fT
T
m
P()
-------------- C
OH
σ
μPT,()
-----------------,,
⎝⎠
⎛⎞
0=
996 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
mantle and constitutes most of the upper mantle along with
olivine (Ringwood, 1970). To better understand the seismic
anisotropy in the upper mantle, one needs to understand the
CPO of orthopyroxene as well (Jung et al., 2010). However,
experimental studies on the CPO of orthopyroxene are very
limited. According to a review on the CPOs of orthopyroxene in
natural mantle rocks (Christensen and Lundquist, 1982), the
CPO of orthopyroxene is characteriz ed as [100] axes al igned normal
to foliation and [001] axes aligned parallel to the lineation,
indicating a major slip system of (100)[001]. This type of
orthopyroxene CPO was classified and defined as Type-AC CPO
(Fig. 9) (Jung et al., 2010), which has been the most commonly
observed CPO in natural rocks (Ishii and Sawaguchi, 2002;
Skemer et al., 2006; Xu et al., 2006; Hidas et al., 2007; Tommasi
et al., 2008; Soustelle et al., 2009; Jung et al., 2010; Puelles et al.,
2012). There are three other types of orthopyroxene CPOs
found in the study of mantle xenoliths in the Spitsbergen, Svalbard
in the Arctic (Jung et al., 2010). Figure 9 shows the Type-AC
CPO of orthopyroxene as well as three others (Type-AB, -BC,
and -ABC) (Jung et al., 2010). The Type-AB CPO of orthopyroxe ne
was defined as [100] axes aligned subnormal to the foliation and
[010] axes aligned subparallel to the lineation. The Type-BC CPO
of orthopyroxene was defined as [010] axes aligned subnormal
to the foliation and [001] axes aligned as subparallel to the lineation.
The Type-ABC CPO of orthopyroxene was defined as both [100]
and [010] axes aligned as a girdle subnormal to the lineation
and [001] axes aligned subparallel to the lineation. The reason
for the occurrence of different CPO types of orthopyroxene was
considered to be a modal proportion of orthopyroxene in the
sample (Jung et al., 2010), but it is still not clearly understood.
Effects of Al and water on the CPO of MgSiO
3
enstatite,
anhydrous and hydrous aluminous enstatite had been investigated
at the pressure of 1.5 GPa and temperature of 1,100 °C using D-
Dia apparatus (Manthilake et al., 2013). In MgSiO
3
enstatite and
hydrous aluminous enstatite, dislocations showing (100)[001]
slip systems (i.e., Type-AC CPO) was observed. However, EBSD
analysis of anhydrous aluminous enstatite indicated operation
of (010)[001] slip system which produces Type-BC CPO. There
are two other experimental studies on the deformation of two-
phase mixtures of orthopyroxene and olivine. Sundberg and
Cooper (2008) conducted deformation experiments of samples
at 1.6 GPa and 1,200 °C, using the Griggs apparatus, and reported
that orthopyroxene shows Type-AC CPO (Fig. 9) for samples
with an olivine:orthopyroxene ratio of (35:65%). Recently, at
Fig. 9. Representative pole figures of
orthopyroxene (opx: enstatite). Four
types of CPOs of orthopyroxene are
shown after Jung et al. (2010). Pole fig-
ures of the crystallographic orientation
of enstatite are presented in the upper
hemisphere using an equal area projec-
tion. The color coding refers to the den-
sity of data points (the numbers in the
legend correspond to the multiples of
uniform distribution). Type-AB: sample
SVF-04, N = 172. Type-AC: sample SVF-
49, N = 176. Type-BC: sample SVF-30, N
= 138. Type-ABC: sample SVF-71, N =
168. A half scatter width of 30° was
used. S: foliation, L: lineation.
Crystal preferred orientations of olivine and hydrous minerals 997
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
high pressures of 3, 5, and 8 GPa and high temperatures of
1,300, 1,400, and 1,500 °C, respectively, Soutelle and Manthilake
(2017) deformed samples (a mixture of olivine and orthopyroxene)
using D-Dia at shear strains of 0.5–1.3. They found that CPOs
of orthopyroxenes are ABC-type (Jung et al., 2010) for samples
with olivine:orthopyroxene ratios of (50:50%), characterized as
[001] axes aligned subparallel to the shear direction and both
[100] and [010] axes aligned subnormal to the shear direction.
4. CPOs OF SERPENTINE (ANTIGORITE)
To understand the CPO development of serpentine, deformation
experiments were conducted on serpentinite under simple shear
at the high pressure of 1 GPa and temperatures of 300 and 400 °C
using a Griggs apparatus (Katayama et al., 2009). The sample
was serpentinite consisting mostly of antigorite. The authors found
that the antigorite CPO was characterized as [001] axes aligned
subnormal to the shear plane and [100] axes aligned subparallel
to the shear direction. Similar CPOs of serpentine were observed in
few natural samples (Bezacier et al., 2010; Brownlee et al., 2013).
On the other hand, CPOs of serpentine (antigorite) in natural
rock were investigated using serpentinites from Val Malenco and
Punta Bettolina in northwest Italy (Jung, 2011). The amount of
serpentine in the samples varied between 40% and 93%. The
CPO of antigorite is characterized as [001] axes aligned subnorma l
to the foliation but [010] axes aligned subparallel to the lineation
(Figs. 10a–c), consistent with other observations of the CPOs of
antigorite in many natural rocks (Hirauchi et al., 2010; Soda and
Takagi, 2010; Nishii et al., 2011; Brownlee et al., 2013; Nagaya et
al., 2014; Watanabe et al., 2014). However, for reasons that are
not yet understood, the alignment of [010] axes subparallel to the
lineation is different from the CPO produced by an experimental
study (Katayama et al., 2009).
5. CPOs OF CHLORITE
Deformation fabrics of chlorite were studied by Kim and Jung
(2015) using chlorite peridotites in Amklovdalen, Western Gneiss
Fig. 10. Typical CPOs of serpentine
(antigorite) in serpentinites from Val
Malenco (a and b) and Punta Bettolina
(c and d), northern Italy (Jung, 2011). (a)
Sample VM1 consists of Ant: 40%, Ol:
52%, Di: 5%, and Mgt: 3% where Ant:
antigorite, Ol: olivine, Di: diopside, and
Mgt: magnetite. (b) Sample VM3 con-
sists of Ant: 87%, Ol: 8%, and Mgt: 5%.
(c) Sample 12B consists of Ant: 93%, Ol:
5%, and Mgt: 2%. (d) Sample 12M con-
sists of Ant: 78%, Ol: 4%, and Mgt: 18%.
Pole figures of serpentine are presented
in equal-area and lower-hemisphere
projections. The E-W direction corresponds
to lineation (x). The N-S direction (z) rep-
resents the direction normal to foliation.
The color coding represents the density
of the data points. The contours corre-
spond to multiples of a uniform distri-
bution. A half-width of 30° was used to
draw the pole figures. n: number of
grains analyzed.
998 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
Region (WGR) in southwest Norway. Two types of chlorite
CPOs (Type-1 and -2) were found. Figure 11a shows that
chlorite [001] axes are strongly aligned subnormal to foliation and
both [100] axes and (010) poles subparallel to foliation. This
type of CPO is defined as the Type-1 CPO of chlorite (Kim and
Jung, 2015). Figure 11b shows that chlorite [001] axes are aligned
as a girdle subnormal to the lineation. This type of CPO is defined
as the Type-2 CPO of chlorite. Type-1 CPO of chlorite was also
observed in natural rocks (Puelles et al., 2012; Morales et al.,
2013; Padrón-Navarta et al., 2015; Kang and Jung, submitted).
Type-2 CPO of chlorite was less commonly observed and reported
in natural samples (Padrón-Navarta et al., 2015; Wallis et al.,
2015). The reason for the occurrence of these two chlorite CPO
types is currently unknown, and may be a topic of further research.
6. CPOs OF AMPHIBOLE (HORNBLENDE)
6.1. CPOs of Amphibole Experimentally Produced
at High Pressure and Temperature
To understand the CPO development of the amphibole,
deformati on experiments on amphibole (hornbl ende) in amphibo lit e
were conducted under simple shear at the high pressure of
1 GPa and temperatures of 480, 500, 600, 700 °C using the
modified Griggs apparatus (Fig. 4b) (Ko and Jung, 2015). The
sample was natural amphibolite consisting mainly of hornblende,
plagioclase, and minor ilmenite (Fig. 2f). Ko and Jung found
three types of amphibole (hornblende) CPOs (Fig. 12a) depending
on temperature and stress (Fig. 12b). Type-I CPO of amphibole
was found at low temperatures and varied stress conditions. It
was characterized as [001] axes aligned subparallel to the shear
direction and (100) poles aligned subnormal to the shear plane.
Type-II CPO of amphibole was found at an intermediate
temperature range (550–700 °C) and high-stress conditions,
and characterized as (010) poles aligned subparallel to the shear
direction and (100) poles aligned subnormal to the shear plane.
Type-III CPO of amphibole was found at high temperature,
low-stress conditions and characterized as both (010) poles and
[001] axes aligned subparallel to the shear plane and (100) poles
aligned subnormal to the shear plane (Fig. 12).
In a study of amphibole fabric formation during diffusion
creep (Getsinger and Hirth, 2014), basalt powder was used as the
starting material to which water was added to produce fine-grained
amphiboles. Shear deformation experiments were conducted at
the pressure of 1 GPa and high temperature of 800 °C. The CPO
of amphibole was characterized as [001] axes aligned subparallel
to the shear direction and [100] axes aligned subnormal to the
shear plane, which is a Type-I CPO (Fig. 12) (Ko and Jung, 2015).
6.2. CPOs of Amphibole found in Natural Rocks
Early studies of the CPO of hornblende were conducted
by Schwerdtner (1964) and Mainprice and Nicolas (1989).
Amphibole with Type-I CPO was found in many places in the
world: lower crustal rocks from the Ivrea Zone, Italy (Barruol and
Kern, 1996), Bergell tonalite in the Central Alps (Berger and
Stünitz, 1996), amphibolite mylonites from the Diancang Shan
in southwest Yunnan, China (Cao et al., 2010), metabasites from
the Aracena metamorphic belt in southwest Spain (Díaz Aspiroz
et al., 2007), the Lewisian in northwest Scotland, a classic tonalitic–
trondjhemite–granodioritic (TTG) gneiss complex (Tatham et al.,
2008), amphibolites from the East Athabasca mylonite triangle
in Saskatchewan, Canada (Ji et al., 2013), and in hornblende schist
(Ji et al., 2015). Recent study of amphibole fabric in hydrated
peridotites from Bjorkedalen, SW Norway also showed the
Type-I CPO of amphibole (Kang and Jung, submitted).
Type-II CPO of amphibole was found in amphibolite mylonites
from the Diancang Shan in southwest Yunnan, China (Cao et
al., 2010) and in amphibolite in Yeoncheon, South Korea (Jung
et al., 2014a). Type-III CPO of amphibole was reported in the
Acebuches metabasites in southwestern Spain (Díaz Aspiroz et
al., 2007), fine-grained amphibolite from the East Athabasca
Fig. 11. Typical CPOs of chlorite pre-
sented in the lower hemisphere using
an equal-area projection (modified from
Kim and Jung, 2015). (a) Type-1 CPO of
chlorite. Sample 436, N = 302. (b) Type-
2 CPO of chlorite. Sample 438, N = 395.
N represents the number of grains. The
white line (S) indicates the foliation and
the red dot (L) indicates the lineation. A
half-scatter width of 20° was used for
the contours. The red color represents
the high density of data points, and the
numbers in the legend correspond to
the multiples of uniform distribution.
Crystal preferred orientations of olivine and hydrous minerals 999
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
mylonite triangle in Saskatchewan, Canada (Ji et al., 2013), and
quartz-hornblende-biotite schist (Ji et al., 2015). Type-IV CPO
is defined as amphibole [001] axes aligned subparallel to the
lineation and both (100) and (110) poles aligned as a girdle
subnormal to the lineation. This type of CPO of amphibole was
reported in amphibolites from the Ryoke metamorphic belt in
southwest Japan (Imon et al., 2004), metabasites from the Aracena
metamorphic belt in southwest Spain (Díaz Aspiroz et al., 2007),
Fig. 12. Pole figures and fabric diagram of amphibole (hornblende). (a) Pole figures of three types of CPOs of the deformed hornblende in
simple shear at high pressure and high temperature conditions and (b) their fabric diagram (Ko and Jung, 2015). The x and z direction cor-
respond to the shear direction and the shear plane normal, respectively. The arrows indicate the sense of shear. The pole figures are pre-
sented in equal-area and lower-hemisphere projection with a half-width of 20°. The contours indicate the m.u.d. for the density of poles.
(Type-I) Sample JH54, P = 1 GPa, T = 480 °C, n = 202. (Type-II) Sample JH46, P = 1 GPa, T = 600 °C, n = 206. (Type-III) Sample JH74, P = 1
GPa, T = 640 °C, n = 216. n: number of measured grains. Differential stress: peak value of differential stress after yielding.
1000 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
the Lewisian of northwest Scotland (Tatham et al., 2008),
amphibolites in Cabo Ortegal, Spain (Llana-Fúnez and Brown,
2012), retrogressed Limo harzburgites, Limo massif, Cabo Ortegal
in northwest Spain (Puelles et al., 2012), and fine-grained amphibolite
and clinopyroxene amphibolite from the East Athabasca mylonite
triangle in Saskatchewan, Canada (Ji et al., 2013).
7. GEOPHYSICAL IMPLICATIONS
Only Type-A CPO of olivine had been used to interpret
seismic anisotropy observed in the upper mantle before the
paper (Jung and Karato, 2001) was published (Mercier, 1985;
Nicolas and Christensen, 1987; Silver, 1996; Savage, 1999;
Fig. 13. Seismic anisotropy of olivine for
different types of CPOs of olivine in Fig-
ure 6a (modified after Jung et al., 2006).
Seismic anisotropy is shown on a ste-
reographic projection in which the cen-
ter of a plot is the direction normal to
the shear plane, and the E-W direction
corresponds to the shear direction.
P-wave anisotropy, the amplitude of
shear wave splitting (dVs) and the direc-
tion of polarization of the faster shear
wave (Vs1) are shown.
Crystal preferred orientations of olivine and hydrous minerals 1001
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
Mainprice et al., 2000; Smith et al., 2001) because no other types
of olivine CPOs were known at that time. However, there were
difficulties in explaining seismic anisotropies observed in
collision zones and hotspot regions such as Hawaii using the
Type-A CPO of olivine (Jung and Karato, 2001; Park and Levin,
2002). The seismic velocities and anisotropies of different types
of olivine CPOs are shown in Figure 13. Type-B, -C, -D, -E CPOs
of olivine produce different seismic signatures from Type-A
CPO (Fig. 13) (see also Jung et al., 2006; Karato et al., 2008). The
water content of olivine in the mid-ocean ridge basalt (MORB),
originating from the oceanic asthenosphere, was reported as
500–1,000 ppm H/Si (Hirth and Kohlstedt, 1996). In addition, a
higher water content exists in basalt originating from hotspots
(Wallace, 1998; Jamtveit et al., 2001). Therefore, the anomalous
seismic anisotropy observed below hotspots (Montagner and
Guillot, 2000; Gaherty, 2001), can be explained by water-induced
Type-C CPO of olivine. Seismic anisotropies produced by the
four orthopyroxene CPOs are shown in Figure 14; their seismic
signatures are also different (Jung et al., 2010). The seismic
properties of orthopyroxene should be carefully considered to
better understand the overall seismic anisotropy in subduction
zones.
Seismic anisotropy of the P- and S-wave was observed in the
mantle wedge in many subduction zones worldwide (i.e., Fig. 1)
(Savage, 1999; Park and Levin, 2002; Long and Silver, 2008;
Long, 2013; Long and Wirth, 2013; Wang and Zhao, 2013; Zhao
et al., 2016). Trench-parallel seismic anisotropy of the fast S-
wave was observed in many fore-arc areas and sub-slabs in
subduction zones (Ando et al., 1983; Fouch and Fischer, 1996;
Margheriti et al., 1996; Smith et al., 2001; Nakajima and Hasegawa,
2004; Long and van der Hilst, 2005; Abt et al., 2010; Long and
Becker, 2010; Di Leo et al., 2012; Long and Wirth, 2013; Wagner
Fig. 14. Seismic anisotropy of orthopy-
roxene for different types of CPOs of
orthopyroxene in Figure 9 (after Jung et
al., 2010). The east-west direction corre-
sponds to the lineation (L), and the
north-south corresponds to the folia-
tion normal. Azimuthal anisotropy of P-
waves (Vp) and polarization anisotropy
of S-waves are shown (AVs is a contour
plot of the magnitude of shear wave
polarization anisotropy and Vs1 is a plot
of the polarization direction of fast S-
waves along different orientations of
propagation).
1002 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
et al., 2013; Lynner and Long, 2014). The source of this trench-
parallel seismic anisotropy was proposed as: (1) the trench-
parallel mantle flow due to trench-roll back (Russo and Silver,
1994; Long and Silver, 2008; Long and Silver, 2009); (2) Type-B
CPO of olivine (Jung and Karato, 2001; Kneller et al., 2005; Jung
et al., 2006; Katayama and Karato, 2006; Kneller and van Keken,
2007; Karato et al., 2008); (3) the 3-D mantle flow around the
slab (Faccenda and Capitanio, 2012, 2013; Li et al., 2014; Lynner
Fig. 15. Seismic anisotropy of serpentine which was calculated from the CPOs of serpentine in Figure 10. The effects of the degree of ser-
pentinization and composition on seismic anisotropy are shown in equal-area and lower hemisphere projections (from Jung, 2011). The seis-
mic anisotropy of serpentine in all the samples is shown in the horizontal flow, where the Y and Z axes are rotated 90° relative to Figure 10.
The X-direction corresponds to the direction of lineation (flow direction). The Z-direction represents the direction normal to foliation (flow
plane). 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 orientations of propagation. The center of the stereonet corresponds to vertical propagation. For the vertical prop-
agation of S-waves, the polarization direction of fast S-waves is nearly perpendicular to the flow direction (lineation). Ant: antigorite, Mgt:
magnetite.
Crystal preferred orientations of olivine and hydrous minerals 1003
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
et al., 2017); (4) fault or crack-induced anisotropy in the slab
(Faccenda et al., 2008; Healy et al., 2009); (5) strong radial
anisotropy (Song and Kawakatsu, 2012, 2013); (6) CPO of
serpentine in the slab-mantle interface and in the mantle wedge
(Katayama et al., 2009; Jung, 2011; Nagaya et al., 2016); and (7)
CPOs of chlorite (Kim and Jung, 2015; Kang and Jung, submitted)
and amphibole in the lower part of mantle wedge (Ko and Jung,
2015; Kang and Jung, submitted). Since olivine in the mantle
wedge is hydrated by the fluids from the dehydration of minerals
in the subducting slab (Peacock and Hyndman, 1999; van Keken
et al., 2011), and the stress is high in a collision zone in the
subduction zones, it is highly likely that Type-B CPO of olivine
is formed in the mantle wedge. This hypothesis is supported by
numerical modeling studies of the effect of Type-B olivine CPO
on seismic anisotropy in the subduction zone (Kneller et al.,
2005, 2007). It is also supported by many observations of Type-
B olivine CPOs in mantle xenoliths (Park and Jung, 2015) and
peridotites in subduction zone settings (Table 3) (Mizukami et al.,
2004; Skemer et al., 2006; Tasaka et al., 2008; Jung, 2009; Jung et
al., 2014b; Kim and Jung, 2015; Behr and Smith, 2016).
It is interesting to see the changes in seismic anisotropy patterns;
from trench-parallel S-wave anisotropy in the fore-arc area to
the trench-normal S-wave anisotropy in the back-arc area (Smith
et al., 2001; Nakajima and Hasegawa, 2004; Long and van der
Hilst, 2006; Long and Wirth, 2013). This pattern of seismic
anisotropy may be explained by the change in olivine CPO from
Type-B in the fore-arc to Type-A, -C (or -E) in the back-arc
(Kneller et al., 2005; Katayama and Karato, 2006; Kneller et al.,
2007; Karato et al., 2008; Jung, 2012). The trench-parallel seismic
anisotropy was also observed in the subducting slab and below
the slab at a depth greater than 100 km in the subduction zone
(Russo and Silver, 1994; Müller et al., 2008; Long and Silver,
2009; Abt et al., 2010; Di Leo et al., 2014; Lynner and Long, 2014;
Lynner et al., 2017). Because of the relatively dry conditions
below the slab (Karato et al., 2008), one possible explanation of
this seismic anisotropy is the Type-B olivine CPO that can be
produced by high pressure in dry conditions (Jung et al., 2009b;
Ohuchi et al., 2011; Soustelle and Manthilake, 2017). This is
supported by a recent study on Type-B olivine CPOs found in
diamond-bearing peridotites in Finsch, South Africa (Lee and
Jung, 2015), which is the first report of a natural example of
pressure-induced Type-B CPO of olivine.
In addition, anomalously long delay times (1–4 s) of S-waves
and strong trench-parallel anisotropy have been observed in
some subduction zones such as in the Ryukyu, Izu-Bonin, and
Tonga-Kermadec arcs (Smith et al., 2001; Anglin and Fouch,
2005; Long and van der Hilst, 2005; Greve et al., 2008). This
strong seismic anisotropy may be caused by the strong CPOs of
hydrous minerals in the lower part of the mantle wedge and at
the interface between slab and mantle wedge. Hydrous minerals
Fig. 16. Schematic diagram showing
the effect of the dipping angle of the
slab on the seismic anisotropy caused
by the CPO of chlorite (Fig. 11a), assum-
ing 2-D corner flow (from Kim and Jung,
2015). (a) Trench-normal seismic anisot-
ropy in the low-angle subduction zone
(θ ≤ 45°). (b) Trench-parallel seismic
anisotropy in the high-angle subduc-
tion zone (θ > 50°). (c) Change in the
seismic anisotropy corresponding to
the change in the dipping angle of the
slab. Blue and red bars represent
trench-normal and trench-parallel seis-
mic anisotropy of the fast S-wave,
respectively. (a) Low-angle (warm) sub-
duction zone. (b) High-angle (cold) sub-
duction zone. (c) Angle-changing
subduction zone.
1004 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
such as serpentine, chlorite, and amphibole can be stable at high
pressure and temperature conditions down to ~200 km (Ulmer
and Trommsdorff, 1995; Pawley, 2003; Fumagalli and Poli, 2005).
Shear deformation experiments of those hydrous minerals
produce strong CPOs at high pressures (Katayama et al., 2009;
Ko and Jung, 2015). Recent studies on the deformation fabrics
of hydrous minerals (serpentine, chlorite, and amphibole) also
showed that strong CPOs of the hydrous minerals in natural rocks
are formed (Figs. 10 and 11) and the CPOs of those serpentine,
chlorite, and ampibole can produce a strong trench-parallel
seismic anisotropy (Figs. 15, 16, and 17b) (Jung, 2011; Kim and
Jung, 2015; Ko and Jung, 2015; Kang and Jung, submitted) and
long delay times of S-waves (Fig. 18). The seismic velocity and
anisotropy pattern of serpentine, chlorite, and amphibole were
investigated in detail (Figs. 15–17), revealing that the strong
trench-parallel seismic anisotropy in the slab and at the slab-
mantle interface where the hydrous minerals are stable can be
explained by the CPOs of the hydrous minerals when the
subduction angle of the slab is high (θ≥45°) (Jung, 2011; Kim
and Jung, 2015; Ko and Jung, 2015). Figure 18 shows the thickness
(L) of the anisotropic layer (i.e., antigorite, chlorite, hornblende,
and olivine) that explains a wide range of delay times (dt) of S-
waves. The figure shows that long delay times (large seismic
anisotropy) of S-waves in some subduction zones can be explained
by a thin anisotropic layer of hydrous minerals such as chlorite,
antigorite, and hornblende.
Fig. 17. Seismic signatures from three different CPO types of amphibole are shown in (a) continental crust and (b) subduction zone (from
Ko and Jung, 2015). (a) For horizontal flow in the continental crust, the anisotropy contours of the S-waves (AVs) show that the orientation
of high anisotropy depends on the CPO types of amphibole. (b) For the flow dipping at 45° in the subduction zone, seismic anisotropy is
strong, and the Vs1 polarization direction (blue bar) is parallel to the trench for all three CPO types of amphibole for a vertically propagating
S-wave (SKS). The blue and red bars indicate the polarization directions of the fast shear wave (Vs1) and slow shear wave (Vs2), respectively.
Hb, hornblende.
Fig. 18. A diagram showing the relationship between delay time (dt)
and anisotropic layer thickness (L) of individual mineral. The follow-
ing equation (Silver and Chan, 1991; Mainprice and Silver, 1993) was
used: dt/L = AVs/<Vs>, where dt is the delay time of the shear waves,
AVs is the anisotropy of the S-wave for a specific propagation direc-
tion, and <Vs> is the average velocity of the fast and slow S-wave
velocities. To calculate delay time, seismic anisotropy data of olivine
(sample JK8; AVs = 5%) (Jung and Karato, 2001), antigorite (sample
VM3; AVs = 34%) (Jung, 2011), antigorite (sample HKB-B; AVs = 24%)
(Watanabe et al., 2014), chlorite (sample 1194-2; AVs = 46%) (Kang
and Jung, submitted), and hornblende (sample JH65; AVs = 12%) (Ko
and Jung, 2015) were used.
Crystal preferred orientations of olivine and hydrous minerals 1005
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
A part of seismic anisotropy in subduction zones can be
attributed to the other hydrous minerals such as glaucophane,
lawsonite, and epidote in subducting slabs. There are some
previous reports on the study of CPO and resultant seismic
anisotropy of glaucophane (Cao et al., 2013; Kim et al., 2013; Cao
and Jung, 2016), lawsonite (Kim et al., 2013, 2016; Cao et al.,
2014; Cao and Jung, 2016), and epidote (Cao et al., 2011, 2013).
However, these data are very limited so far and more detail studies
on the development of CPOs of those minerals and resultant
seismic signatures are needed to better understand the seismic
anisotropy in the subducting slabs.
8. CONCLUSIONS
Crystal preferred orientations (CPOs) of olivine, orthopyroxene,
serpentine, chlorite, and amphibole in previous experimental
research and in natural rocks were reviewed, and the seismic
anisotropies of those minerals were discussed. Water-induced
CPOs of olivine, such as Type-B, -C, and -E found in experimental
studies, were also observed in many natural rocks and could be
very important for interpreting seismic anisotropies in the upper
mantle. Hydrous minerals such as amphibole, chlorite, and
serpentine in the lower part of the mantle wedge and at the slab-
mantle interface can also produce strong CPOs, which can be
used to interpret anomalously strong seismic anisotropies in
some subduction zones. In the future, hydrous minerals in the
subducting slab (i.e., glaucophane, epidote, and lawsonite)
should be investigated in detail to better understand the seismic
anisotropy in the slab in subduction zones.
ACKNOWLEDGMENTS
H.J. would like to thank colleagues at the Tectonophysics
Laboratory at the Seoul National University and other institutions
for their help in various projects. H.J. wishes to thank Y. Park
and S. Jung for preparing Tables and S. Choi for his help in
preparing Figure 3. H.J. would also like to express sincere gratitude
to Dr. Simon Wallis and an anonymous reviewer for their valuable
comments and suggestions. This study was supported by the NRF
grant of Korea (NRF-2017R1A2B2004688) and by KMIPA2017-
9020.
REFERENCES
Abramson, E.H., Brown, J.M., Slutsky, L.J., and Zaug, J., 1997, The elas-
tic constants of San Carlos olivine to 17 GPa. Journal of Geophysi-
cal Research: Solid Earth, 102, 12253–12263.
Abt, D.L., Fischer, K.M., Abers, G.A., Protti, M., González, V., and
Strauch, W., 2010, Constraints on upper mantle anisotropy sur-
rounding the Cocos slab from SK(K)S splitting. Journal of Geo-
physical Research: Solid Earth, 115, B06316.
Aleksandrov, K., Alchikov, U., Belikov, B., Zaslavskii, B., and Krupnyi,
A., 1974, Velocities of elastic waves in minerals at atmospheric
pressure and increasing precision of elastic constants by means of
EVM. Izvestiya Akademii Nauk SSSR, Seriya Geologicheskaya, 10,
15–24. (in Russian)
Aleksandrov, K.S. and Ryzhova, T.V., 1961a, The elastic properties of
rock-forming minerals, I: pyroxenes and amphiboles. Bulletin of
the Academy of Sciences of the USSR: Geophysics series, 871–875,
1339–1344.
Aleksandrov, K. and Ryzhova, T., 1961b, Elastic properties of rock-
forming minerals II: layered silicates. Bulletin of the Academy of
Sciences of the USSR: Geophysics series, English translation, 12,
1165–1168.
Almqvist, B.S.G. and Mainprice, D., 2017, Seismic properties and
anisotropy of the continental crust: predictions based on mineral
texture and rock microstructure. Reviews of Geophysics, 55, 367–
433.
Ando, M., Ishikawa, Y., and Yamazaki, F., 1983, Shear wave polariza-
tion anisotropy in the upper mantle beneath Honshu, Japan. Jour-
nal of Geophysical Research, 88, 5850–5864.
Anglin, D.K. and Fouch, M.J., 2005, Seismic anisotropy in the Izu-
Bonin subduction system. Geophysical Research Letters, 32, L09307.
doi:10.1029/2005GL022714
Auzende, A.-L., Pellenq, R.-M., Devouard, B., Baronnet, A., and
Grauby, O., 2006, Atomistic calculations of structural and elastic
properties of serpentine minerals: the case of lizardite. Physics and
Chemistry of Minerals, 33, 266–275.
Avé Lallemant, H.G. and Carter, N.L., 1970, Syntectonic recrystalliza-
tion of olivine and modes of flow in the upper mantle. Bulletin of
the Geological Society of America, 81, 2203–2220.
Barruol, G. and Kern, H., 1996, Seismic anisotropy and shear-wave
splitting in lower-crustal and upper-mantle rocks from the Ivrea
Zone – experimental and calculated data. Physics of the Earth and
Planetary Interiors, 95, 175–194.
Behr, W.M. and Smith, D., 2016, Deformation in the mantle wedge
associated with Laramide flat-slab subduction. Geochemistry, Geo-
physics, Geosystems, 17, 2643–2660.
Ben Ismail, W. and Mainprice, D., 1998, An olivine fabric database: an
overview of upper mantle fabrics and seismic anisotropy. Tectono-
physics, 296, 145–157.
Berger, A. and Stünitz, H., 1996, Deformation mechanisms and reac-
tion of hornblende: examples from the Bergell tonalite (Central
Alps). Tectonophysics, 257, 149–174.
Bezacier, L., Reynard, B., Bass, J.D., Sanchez-Valle, C., and Van de
Moortele, B.V., 2010, Elasticity of antigorite, seismic detection of
serpentinites, and anisotropy in subduction zones. Earth and Plan-
etary Science Letters, 289, 198–208.
Birch, F., 1960, The velocity of compressional waves in rocks to 10 kilo-
bars: 1. Journal of Geophysical Research, 65, 1083–1102.
Boneh, Y., Morales, L.F.G., Kaminski, E., and Skemer, P., 2015, Model-
ing olivine CPO evolution with complex deformation histories:
implications for the interpretation of seismic anisotropy in the
mantle. Geochemistry, Geophysics, Geosystems, 16, 3436–3455.
1006 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
Boneh, Y. and Skemer, P., 2014, The effect of deformation history on
the evolution of olivine CPO. Earth and Planetary Science Letters,
406, 213–222.
Brown, J.M. and Abramson, E.H., 2016, Elasticity of calcium and cal-
cium-sodium amphiboles. Physics of the Earth and Planetary Inte-
riors, 261, 161–171.
Brownlee, S.J., Hacker, B.R., Harlow, G.E., and Seward, G., 2013, Seis-
mic signatures of a hydrated mantle wedge from antigorite crystal-
preferred orientation (CPO). Earth and Planetary Science Letters,
375, 395–407.
Bystricky, M., Kunze, K., Burlini, L., and Burg, J.P., 2000, High shear
strain of olivine aggregates: rheological and seismic consequences.
Science, 290, 1564–1567.
Cao, S., Liu, J., and Leiss, B., 2010, Orientation-related deformation
mechanisms of naturally deformed amphibole in amphibolite mylon-
ites from the Diancang Shan, SW Yunnan, China. Journal of Struc-
tural Geology, 32, 606–622.
Cao, Y. and Jung, H., 2016, Seismic properties of subducting oceanic
crust: constraints from natural lawsonite-bearing blueschist and
eclogite in Sivrihisar Massif, Turkey. Physics of the Earth and Plan-
etary Interiors, 250, 12–30.
Cao, Y., Jung, H., and Song, S., 2013, Petro‐fabrics and seismic proper-
ties of blueschist and eclogite in the North Qilian suture zone, NW
China: implications for the low‐velocity upper layer in subducting
slab, trench‐parallel seismic anisotropy, and eclogite detectability in
the subduction zone. Journal of Geophysical Research: Solid Earth,
118, 3037–3058.
Cao, Y., Jung, H., and Song, S., 2014, Microstructures and petro-fabrics
of lawsonite blueschist in the North Qilian suture zone, NW China:
implications for seismic anisotropy of subducting oceanic crust.
Tectonophysics, 628, 140–157.
Cao, Y., Jung, H., and Song, S.G., 2017, Olivine fabrics and tectonic
evolution of fore-arc mantles: a natural perspective from the Song-
shugou dunite and harzburgite in the Qinling orogenic belt, cen-
tral China. Geochemistry, Geophysics, Geosystems, 18, 907–934.
Cao, Y., Jung, H., Song, S.G., Park, M., Jung, S., and Lee, J., 2015, Plastic
deformation and seismic properties in fore-arc mantles: a petrofab-
ric analysis of the Yushigou Harzburgites, North Qilian Suture
Zone, NW China. Journal of Petrology, 56, 1897–1943.
Cao, Y., Song, S.G., Niu, Y.L., Jung, H., and Jin, Z.M., 2011, Variation of
mineral composition, fabric and oxygen fugacity from massive to
foliated eclogites during exhumation of subducted ocean crust in
the North Qilian suture zone, NW China. Journal of Metamorphic
Geology, 29, 699–720.
Carter, N.L. and Avé Lallemant, H.G., 1970, High temperature flow of
dunite and peridotite. Bulletin of the Geological Society of Amer-
ica, 81, 2181–2202.
Castelnau, O., Blackman, D.K., and Becker, T.W., 2009, Numerical sim-
ulations of texture development and associated rheological anisot-
ropy in regions of complex mantle flow. Geophysical Research
Letters, 36, L12304. doi:10.1029/2009GL038027
Castelnau, O., Blackman, D.K., Lebensohn, R.A., and Ponte Castañeda,
P., 2008, Micromechanical modeling of the viscoplastic behavior of
olivine. Journal of Geophysical Research: Solid Earth, 113, B09202.
doi:10.1029/2007JB005444
Castelnau, O., Cordier, P., Lebensohn, R.A., Merkel, S., and Raterron, P.,
2010, Microstructures and rheology of the Earth's upper mantle
inferred from a multiscale approach. Comptes Rendus Physique,
11, 304–315.
Chai, M., Brown, J.M., and Slutsky, L.J., 1997, The elastic constants of
an aluminous orthopyroxene to 12.5 GPa. Journal of Geophysical
Research: Solid Earth, 102, 14779–14785.
Chatzaras, V., Kruckenberg, S.C., Cohen, S.M., Medaris, L.G., Jr., With-
ers, A.C., and Bagley, B., 2016, Axial-type olivine crystallographic
preferred orientations: the effect of strain geometry on mantle tex-
ture. Journal of Geophysical Research: Solid Earth, 121, 4895–4922.
Christensen, N.I. and Lundquist, S.M., 1982, Pyroxene orientation within
the upper mantle. Geological Society of America Bulletin, 93, 279–
288.
Couvy, H., Frost, D.J., Heidelbach, F., Nyilas, K., Ungar, T., Mackwell, S.,
and Cordier, P., 2004, Shear deformation experiments of forsterite
at 11 GPa-1400 degrees C in the multianvil apparatus. European
Journal of Mineralogy, 16, 877–889.
Czertowicz, T.A., Toy, V.G., and Scott, J.M., 2016, Recrystallisation, phase
mixing and strain localisation in peridotite during rapid extrusion
of sub-arc mantle lithosphere. Journal of Structural Geology, 88, 1–19.
Di Leo, J.F., Walker, A.M., Li, Z.H., Wookey, J., Ribe, N.M., Kendall,
J.M., and Tommasi, A., 2014, Development of texture and seismic
anisotropy during the onset of subduction. Geochemistry, Geo-
physics, Geosystems, 15, 192–212.
Di Leo, J.F., Wookey, J., Hammond, J.O.S., Kendall, J.M., Kaneshima, S.,
Inoue, H., Yamashina, T., and Harjadi, P., 2012, Deformation and
mantle flow beneath the Sangihe subduction zone from seismic
anisotropy. Physics of the Earth and Planetary Interiors, 194, 38–54.
Díaz Aspiroz, M., Lloyd, G.E., and Fernández, C., 2007, Development
of lattice preferred orientation in clinoamphiboles deformed under
low-pressure metamorphic conditions. A SEM/EBSD study of metaba-
sites from the Aracena metamorphic belt (SW Spain). Journal of
Structural Geology, 29, 629–645.
Faccenda, M., Burlini, L., Gerya, T.V., and Mainprice, D., 2008, Fault-
induced seismic anisotropy by hydration in subducting oceanic plates.
Nature, 455, 1097–1100.
Faccenda, M. and Capitanio, F.A., 2012, Development of mantle seis-
mic anisotropy during subduction-induced 3-D flow. Geophysical
Research Letters, 39, L11305. doi:10.1029/2012GL051988
Faccenda, M. and Capitanio, F.A., 2013, Seismic anisotropy around
subduction zones: insights from three-dimensional modeling of
upper mantle deformation and SKS splitting calculations. Geo-
chemistry, Geophysics, Geosystems, 14, 243–262.
Fouch, M.J., Fischer, K.M., 1996, Mantle anisotropy beneath northwest
Pacific subduction zones. Journal of Geophysical Research: Solid
Earth, 101, 15987–16002.
Frese, K., Trommsdorff, V., and Kunze, K., 2003, Olivine [100] normal
to foliation: lattice preferred orientation in prograde garnet peridot-
ite formed at high H
2O activity, Cima di Gagnone (Central Alps).
Contributions to Mineralogy and Petrology, 145, 75–86.
Frets, E.C., Tommasi, A., Garrido, C.J., Vauchez, A., Mainprice, D., Tar-
guisti, K., and Amri, I., 2014, The Beni Bousera Peridotite (Rif Belt,
Morocco): an oblique-slip low-angle shear zone thinning the sub-
continental mantle lithosphere. Journal of Petrology, 55, 283–313.
Crystal preferred orientations of olivine and hydrous minerals 1007
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
Fumagalli, P. and Poli, S., 2005, Experimentally determined phase rela-
tions in hydrous peridotites to 6.5 GPa and their consequences on
the dynamics of subduction zones. Journal of Petrology, 46, 555–
578.
Furukawa, Y., 1993, Magmatic processes under arcs and formation of
the volcanic front. Journal of Geophysical Research, 98, 8309–8319.
Gaherty, J.B., 2001, Seismic evidence for hotspot-induced bouyant flow
beneath the Peykjanes ridge. Science, 293, 1645–1647.
Getsinger, A.J. and Hirth, G., 2014, Amphibole fabric formation during
diffusion creep and the rheology of shear zones. Geology, 42, 535–
538.
Greve, S.M., Savage, M.K., and Hofmann, S.D., 2008, Strong variations
in seismic anisotropy across the Hikurangi subduction zone, North
Island, New Zealand. Tectonophysics, 462, 7–21.
Hansen, L., Zimmerman, M., and Kohlstedt, D., 2012, Laboratory mea-
surements of the viscous anisotropy of olivine aggregates. Nature,
492, 415–418.
Hansen, L.N. and Warren, J.M., 2015, Quantifying the effect of pyrox-
ene on deformation of peridotite in a natural shear zone. Journal of
Geophysical Research: Solid Earth, 120, 2717–2738.
Hansen, L.N., Zhao, Y.-H., Zimmerman, M.E., and Kohlstedt, D.L.,
2014, Protracted fabric evolution in olivine: implications for the
relationship among strain, crystallographic fabric, and seismic
anisotropy. Earth and Planetary Science Letters, 387, 157–168.
Healy, D., Reddy, S.M., Timms, N.E., Gray, E.M., and Brovarone, A.V.,
2009, Trench-parallel fast axes of seismic anisotropy due to fluid-
filled cracks in subducting slabs. Earth and Planetary Science Let-
ters, 283, 75–86.
Hearmon, R., 1984, The elastic constants of crystals and other anisotro-
pic materials. In: Hellwege K.H. (ed.), Landolt-Bornstein Tables,
III/18. Springer-Verlag, Berlin, p. 1–154.
Hess, H.H., 1964, Seismic anisotropy of the uppermost mantle under
oceans. Nature, 203, 629–631.
Hidas, K., Falus, G., Szabo, C., Szabo, P.J., Kovacs, I., and Foldes, T.,
2007, Geodynamic implications of flattened tabular equigranular
textured peridotites from the Bakony-Balaton Highland Volcanic
Field (Western Hungary). Journal of Geodynamics, 43, 484–503.
Hirauchi, K., Michibayashi, K., Ueda, H., and Katayama, I., 2010, Spa-
tial variations in antigorite fabric across a serpentinite subduction
channel: insights from the Ohmachi Seamount, Izu-Bonin frontal
arc. Earth and Planetary Science Letters, 299, 196–206.
Hirth, G. and Kohlstedt, D.L., 1996, Water in the oceanic upper man-
tle: implications for rheology, melt extraction and the evolution of
the lithosphere. Earth and Planetary Science Letters, 144, 93–108.
Holtzman, B.K., Kohlstedt, D.L., Zimmerman, M.E., Heidelbach, F.,
Hiraga, T., and Hustoft, J., 2003, Melt segregation and strain parti-
tioning: implications for seismic anisotropy and mantle flow. Sci-
ence, 301, 1227–1230.
Imon, R., Okudaira, T., and Kanagawa, K., 2004, Development of shape-
and lattice-preferred orientations of amphibole grains during ini-
tial cataclastic deformation and subsequent deformation by disso-
lution-precipitation creep in amphibolites from the Ryoke metamorphic
belt, SW Japan. Journal of Structural Geology, 26, 793–805.
Ishii, K. and Sawaguchi, T., 2002, Lattice- and shape-preferred orienta-
tion of orthopyroxene porphyroclasts in peridotites: an application
of two-dimensional numerical modeling. Journal of Structural
Geology, 24, 517–530.
Jamtveit, B., Brooker, R., Brooks, K., Larsen, L.M., and Pedersen, T.,
2001, The water content of olivines from the North Atlantic Volca-
nic Province. Earth and Planetary Science Letters, 186, 401–415.
Ji, S., Shao, T., Michibayashi, K., Long, C., Wang, Q., Kondo, Y., Zhao,
W., Wang, H., and Salisbury, M.H., 2013, A new calibration of seis-
mic velocities, anisotropy, fabrics, and elastic moduli of amphibole-rich
rocks. Journal of Geophysical Research: Solid Earth, 118, 4699–
4728.
Ji, S., Shao, T., Michibayashi, K., Oya, S., Satsukawa, T., Wang, Q., Zhao,
W., and Salisbury, M.H., 2015, Magnitude and symmetry of seis-
mic anisotropy in mica- and amphibole-bearing metamorphic rocks
and implications for tectonic interpretation of seismic data from
the southeast Tibetan Plateau. Journal of Geophysical Research:
Solid Earth, 120, 6404–6430.
Jung, H., 2009, Deformation fabrics of olivine in Val Malenco peridot-
ite found in Italy and implications for the seismic anisotropy in the
upper mantle. Lithos, 109, 341–349.
Jung, H., 2011, Seismic anisotropy produced by serpentine in mantle
wedge. Earth and Planetary Science Letters, 307, 535–543.
Jung, H., 2012, Rock deformation and formation of LPO of minerals in
the upper mantle: implications for seismic anisotropy. Journal of
Petrological Society of Korea, 21, 229–241. (in Korean with English
abstract)
Jung, H., Jung, S., Ko, B., and Lee, J., 2014a, Crystal preferred orienta-
tion of amphibolites found at Yeoncheon and Chuncheon area in
South Korea. Proceedings of the Joint Conference of the Geologi-
cal Science and Technology of Korea, Busan, Apr. 20–23, p. 146–
147.
Jung, H. and Karato, S., 2001, Water-induced fabric transitions in oliv-
ine. Science, 293, 1460–1463.
Jung, H., Katayama, I., Jiang, Z., Hiraga, T., and Karato, S., 2006, Effect
of water and stress on the lattice-preferred orientation of olivine.
Tectonophysics, 421, 1–22.
Jung, H., Lee, J., Ko, B., Jung, S., Park, M., Cao, Y., and Song, S., 2013,
Natural Type-C olivine fabrics in garnet peridotites in North Qaidam
UHP collision belt, NW China. Tectonophysics, 594, 91–102.
Jung, H., Mo, W., and Choi, S.H., 2009a, Deformation microstructures
of olivine in peridotite from Spitsbergen, Svalbard and implications
for seismic anisotropy. Journal of Metamorphic Geology, 27, 707–
720.
Jung, H., Mo, W., and Green, H.W., 2009b, Upper mantle seismic
anisotropy resulting from pressure-induced slip transition in oliv-
ine. Nature Geoscience, 2, 73–77.
Jung, H., Park, M., Jung, S., and Lee, J., 2010, Lattice preferred orienta-
tion, water content, and seismic anisotropy of orthopyroxene. Jour-
nal of Earth Science, 21, 555–568.
Jung, S., Jung, H., and Austrheim, H., 2014b, Characterization of oliv-
ine fabrics and mylonite in the presence of fluid and implications
for seismic anisotropy and shear localization. Earth, Planets and
Space, 66, 41–21.
Kaczmarek, M.-A. and Tommasi, A., 2011, Anatomy of an extensional
shear zone in the mantle, Lanzo massif, Italy. Geochemistry, Geo-
physics, Geosystems, 12, Q0AG06. doi:10.1029/2011GC003627
1008 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
Kaminski, E. and Ribe, N.M., 2001, A kinematic model for recrystalli-
zation and texture development in olivine polycrystals. Earth and
Planetary Science Letters, 189, 253–267.
Kaminski, E. and Ribe, N.M., 2002, Timescales for the evolution of seis-
mic anisotropy in mantle flow. Geochemistry, Geophysics, Geosys-
tems, 3. doi:10.1029/2001GC000222
Kaminski, E., Ribe, N.M., and Browaeys, J.T., 2004, D-Rex, a program
for calculation of seismic anisotropy due to crystal lattice preferred
orientation in the convective upper mantle. Geophysical Journal
International, 158, 744–752.
Karato, S., Jung, H., Katayama, I., and Skemer, P., 2008, Geodynamic
significance of seismic anisotropy of the upper mantle: new
insights from laboratory studies. Annual Review of Earth and Plan-
etary Sciences, 36, 59–95.
Katayama, I., Hirauchi, H., Michibayashi, K., and Ando, J., 2009,
Trench-parallel anisotropy produced by serpentine deformation in
the hydrated mantle wedge. Nature, 461, 1114–1118.
Katayama, I., Jung, H., and Karato, S.I., 2004, New type of olivine fabric
from deformation experiments at modest water content and low
stress. Geology, 32, 1045–1048.
Katayama, I. and Karato, S., 2006, Effect of temperature on the B- to C-
type olivine fabric transition and implication for flow pattern in
subduction zones. Physics of the Earth and Planetary Interiors,
157, 33–45.
Kim, D. and Jung, H., 2015, Deformation microstructures of olivine
and chlorite in chlorite peridotites from Almklovdalen in the West-
ern Gneiss Region, southwest Norway, and implications for seis-
mic anisotropy. International Geology Review, 57, 650–668.
Kim, D., Katayama, I., Michibayashi, K., and Tsujimori, T., 2013, Defor-
mation fabrics of natural blueschists and implications for seismic
anisotropy in subducting oceanic crust. Physics of the Earth and
Planetary Interiors, 222, 8–21.
Kim, D., Wallis, S., Endo, S., and Ree, J.-H., 2016, Seismic properties of
lawsonite eclogites from the southern Motagua fault zone, Guate-
mala. Tectonophysics, 677, 88–98.
Kitamura, K., 2006, Constraint of lattice-preferred orientation (LPO)
on Vp anisotropy of amphibole-rich rocks. Geophysical Journal
International, 165, 1058–1065.
Kneller, E.A. and van Keken, P.E., 2007, Trench-parallel flow and seis-
mic anisotropy in the Mariana and Andean subduction systems.
Nature, 450, 1222–1225.
Kneller, E.A., van Keken, P.E., Karato, S., and Park, J., 2005, B-type oliv-
ine fabric in the mantle wedge: insights from high-resolution non-
Newtonian subduction zone models. Earth and Planetary Science
Letters, 237, 781–797.
Ko, B. and Jung, H., 2015, Crystal preferred orientation of an amphi-
bole experimentally deformed by simple shear. Nature Communi-
cations, 6, 6586. doi:10.1038/ncomms7586
Lamarque, G., Bascou, J., Maurice, C., Cottin, J.-Y., Riel, N., Riel, and
Ménot, R.-P., 2016, Microstructures, deformation mechanisms and
seismic properties of a Palaeoproterozoic shear zone: the Mertz
shear zone, East-Antarctica. Tectonophysics, 680, 174–191.
Lee, J. and Jung, H., 2015, Lattice-preferred orientation of olivine found
in diamond-bearing garnet peridotites in Finsch, South Africa and
implications for seismic anisotropy. Journal of Structural Geology,
70, 12–22.
Li, Z.-H., Di Leo, J.F., and Ribe, N.M., 2014, Subduction-induced man-
tle flow, finite strain, and seismic anisotropy: numerical modeling.
Journal of Geophysical Research: Solid Earth, 119, 5052–5076.
Linckens, J., Herwegh, M., Muntener, O., and Mercolli, I., 2011, Evolu-
tion of a polymineralic mantle shear zone and the role of second
phases in the localization of deformation. Journal of Geophysical
Research: Solid Earth, 116, B06210. doi:10.1029/2010JB008119
Llana-Fúnez, S. and Brown, D., 2012, Contribution of crystallographic
preferred orientation to seismic anisotropy across a surface analog
of the continental Moho at Cabo Ortegal, Spain. Geological Soci-
ety of America Bulletin, 124, 1495–1513.
Lloyd, G.E., Butler, R.W.H., Casey, M., Tatham, D.J., and Mainprice, D.,
2011, Constraints on the seismic properties of the middle and
lower continental crust. Geological Society of London, Special
Publication, 360, 7–32.
Long, M.D., 2013, Constraints on subduction geodynamics from seis-
mic anisotropy. Reviews of Geophysics, 51, 76–112.
Long, M.D. and Becker, T.W., 2010, Mantle dynamics and seismic
anisotropy. Earth and Planetary Science Letters, 297, 341–354.
Long, M.D. and Silver, P.G., 2008, The subduction zone flow field from
seismic anisotropy: a global view. Science, 319, 315–318.
Long, M.D. and Silver, P.G., 2009, Shear wave splitting and mantle
anisotropy: measurements, interpretations, and new directions.
Surveys in Geophysics, 30, 407–461.
Long, M.D. and van der Hilst, R.D., 2005, Upper mantle anisotropy
beneath Japan from shear wave splitting. Physics of the Earth and
Planetary Interiors, 151, 206–222.
Long, M.D. and Wirth, E.A., 2013, Mantle flow in subduction systems:
the mantle wedge flow field and implications for wedge processes.
Journal of Geophysical Research: Solid Earth, 118, 583–606.
Lynner, C. and Long, M.D., 2014, Sub-slab anisotropy beneath the
Sumatra and circum-Pacific subduction zones from source-side
shear wave splitting observations. Geochemistry, Geophysics, Geo-
systems, 15, 2262–2281.
Lynner, C., Long, M.D., Tissen, C.J., Paczkowski, K., and Montesi,
L.G.J., 2017, Evaluating geodynamic models for sub-slab anisot-
ropy: effects of olivine fabric type. Geological Society of America
Bulletin, 13, 247–259.
Mackwell, S.J. and Kohlstedt, D.L., 1990, Diffusion of hydrogen in oliv-
ine – implications for water in the mantle. Journal of Geophysical
Research: Solid Earth and Planets, 95, 5079–5088.
Mainprice, D., 2007, Seismic anisotropy of the deep Earth from a min-
eral and rock physics perspective, treatise on geophysics. In:
Schubert, G. (ed.), Treatise in Geophysics Volume 2: Mineral
Physics. Elsevier, Amsterdam, p. 437–491.
Mainprice, D., Barruol, G., and Ismail, W.B., 2000, The seismic anisot-
ropy of the earth's mantle from single crystal to polycrystal. In:
Karato, S., Forte, A.M., Liebermann, R.C., Masters, G., and Stix-
rude, L. (eds.), Earth's Deep Interior. American Geophysical Union,
Geophysical Monograph, 117, p. 237–264.
Mainprice, D. and Ildefonse, B., 2009, Seismic anisotropy of subduc-
tion zone minerals – contribution of hydrous phases. In: Lallemand,
S. and Funiciello, F. (eds.), Subduction Zone Geodynamics, Fron-
tiers in Earth Sciences. Springer-Verlag, Berlin, p. 63–84.
Crystal preferred orientations of olivine and hydrous minerals 1009
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
Mainprice, D. and Nicolas, A., 1989, Development of shape and lattice
preferred orientations: application to the seismic anisotropy of the
lower crust. Journal of Structural Geology, 11, 175–189.
Mainprice, D. and Silver, P.G., 1993, Interpretation of SKS-waves using
samples from the subcontinental lithosphere. Physics of the Earth
and Planetary Interiors, 78, 257–280.
Manthilake, M.A.G.M., Miyajima, N., Heidelbach, F., Soustelle, V., and
Frost, D.J., 2013, The effect of aluminum and water on the develop-
ment of deformation fabrics of orthopyroxene. Contributions to
Mineralogy and Petrology, 165, 495–505.
Margheriti, L., Nostro, C., Cocco, M., and Amato, A., 1996, Seismic
anisotropy beneath the Northern Apennines (Italy) and its tec-
tonic implications. Geophysical Research Letters, 23, 2721–2724.
Mercier, J.C., 1985, Olivine and pyroxene. In: Wenk, H.R. (ed.), Pre-
ferred Orientation in Deformed Metals and Rocks: An Introduc-
tion to Modern Texture Analysis. Academic Press, New York, p.
407–430.
Michibayashi, K., Ina, T., and Kanagawa, K., 2006, The effect of dynamic
recrystallization on olivine fabric and seismic anisotropy: insight
from a ductile shear zone, Oman ophiolite. Earth and Planetary Sci-
ence Letters, 244, 695–708.
Michibayashi, K., Kusafuka, Y., Satsukawa, T., and Nasir, S.J., 2012, Seis-
mic properties of peridotite xenoliths as a clue to imaging the litho-
spheric mantle beneath NE Tasmania, Australia. Tectonophysics,
522–523, 218–223.
Michibayashi, K., Mainprice, D., Fujii, A., Uehara, S., Shinkai, Y., Kondo,
Y., Ohara, Y., Ishii, T., Fryer, P., Bloomer, S.H., Ishiwatari, A., Haw-
kins, J.W., and Ji, S., 2016, Natural olivine crystal-fabrics in the western
Pacific convergence region: a new method to identify fabric type.
Earth and Planetary Science Letters, 443, 70–80.
Michibayashi, K. and Oohara, T., 2013, Olivine fabric evolution in a
hydrated ductile shear zone at the Moho Transition Zone, Oman
Ophiolite. Earth and Planetary Science Letters, 377–378, 299–310.
Michibayashi, K., Tasaka, M., Ohara, Y., Ishii, T., Okamoto, A., and
Fryer, P., 2007, Variable microstructure of peridotite samples from
the southern Mariana Trench: evidence of a complex tectonic evo-
lution. Tectonophysics, 444, 111–118.
Miyazaki, T., Sueyoshi, K., and Hiraga, T., 2013, Olivine crystals align
during diffusion creep of Earth’s upper mantle. Nature, 502, 321–
326.
Mizukami, T., Wallis, S.R., and Yamamoto, J., 2004, Natural examples of
olivine lattice preferred orientation patterns with a flow-normal a-
axis maximum. Nature, 427, 432–436.
Montagner, J.P. and Guillot, L., 2000, Seismic anisotropy in the Earths
mantle. In: Boschi, E., Dziewonski, A.M., Morelli, A., and Ekström,
G. (eds.), Problems in Geophysics for the New Millennium. Com-
positori, Bologna, p. 213–253.
Mookherjee, M. and Mainprice, D., 2014, Unusually large shear wave
anisotropy for chlorite in subduction zone settings. Geophysical
Research Letters, 41, 1506–1513.
Morales, L.F., Mainprice, D., and Boudier, F., 2013, The influence of
hydrous phases on the microstructure and seismic properties of a
hydrated mantle rock. Tectonophysics, 594, 103–117.
Müller, C., Bayer, B., Eckstaller, A., and Miller, H., 2008, Mantle flow in
the South Sandwich subduction environment from source-side
shear wave splitting. Geophysical Research Letters, 35, L03301.
doi:10.1029/2007GL032411
Nagaya, T., Walker, A.M., Wookey, J., Wallis, S.R., Ishii, K., and Kend-
all, J.M., 2016, Seismic evidence for flow in the hydrated mantle
wedge of the Ryukyu subduction zone. Scientific Reports, 6, 29981.
doi:10.1038/srep29981
Nagaya, T., Wallis, S.R., Kobayashi, H., Michibayashi, K., Mizukami, T.,
Seto, Y., Miyake, A., and Matsumoto, M., 2014, Dehydration break-
down of antigorite and the formation of B-type olivine CPO. Earth
and Planetary Science Letters, 387, 67–76.
Nakajima, J. and Hasegawa, A., 2004, Shear-wave polarization anisot-
ropy and subduction-induced flow in the mantle wedge of north-
eastern Japan. Earth and Planetary Science Letters, 225, 365–377.
Nicolas, A. and Christensen, N.I., 1987, Formation of anisotropy in upper
mantle peridotites: a review. American Geophysical Union, 16, 111–
123.
Nishii, A., Wallis, S.R., Mizukami, T., and Michibayashi, K., 2011, Sub-
duction related antigorite CPO patterns from forearc mantle in the
Sanbagawa belt, southwest Japan. Journal of Structural Geology, 33,
1436–1445.
Ohuchi, T. and Irifune, T., 2013, Development of A-type olivine fabric
in water-rich deep upper mantle. Earth and Planetary Science Let-
ters, 362, 20–30.
Ohuchi, T. and Irifune, T., 2014, Crystallographic preferred orientation
of olivine in the Earth's deep upper mantle. Physics of the Earth
and Planetary Interiors, 228, 220–231.
Ohuchi, T., Kawazoe, T., Nishihara, Y., Nishiyama, N., and Irifune, T.,
2011, High pressure and temperature fabric transitions in olivine
and variations in upper mantle seismic anisotropy. Earth and Plan-
etary Science Letters, 304, 55–63.
Ohuchi, T., Nishihara, Y., Seto, Y., Kawazoe, T., Nishi, M., Maruyama,
G., Hashimoto, M., Higo, Y., Funakoshi, K.-I., Suzuki, A., Kikegawa,
T., and Irifune, T., 2015, In situ observation of crystallographic pre-
ferred orientation of deforming olivine at high pressure and high
temperature. Physics of the Earth and Planetary Interiors, 243, 1–21.
Padrón-Navarta, J.A., Tommasi, A., Garrido, C.J., and Mainprice, D.,
2015, On topotaxy and compaction during antigorite and chlorite
dehydration: an experimental and natural study. Contributions to
Mineralogy and Petrology, 169, 35.
Park, J. and Levin, V., 2002, Seismic anisotropy: tracing plate dynamics
in the mantle. Science, 296, 485–489.
Park, M. and Jung, H., 2017, Microstructural evolution of the Yugu
peridotites in the Gyeonggi Massif, Korea: implications for olivine
fabric transition in mantle shear zones. Tectonophysics, 709, 55–68.
Park, M., Jung, H., and Kil, Y., 2014, Petrofabrics of olivine in a rift axis
and rift shoulder and their implications for seismic anisotropy beneath
the Rio Grande rift. Island Arc, 23, 299–311.
Park, Y. and Jung, H., 2015, Deformation microstructures of olivine and
pyroxene in mantle xenoliths in Shanwang, eastern China, near the
convergent plate margin, and implications for seismic anisotropy.
International Geology Review, 57, 629–649.
Pawley, A., 2003, Chlorite stability in mantle peridotite: the reaction
clinochlore + enstatite = forsterite + pyrope + H2O. Contributions
to Mineralogy and Petrology, 144, 449–456.
Peacock, S.M. and Hyndman, R.D., 1999, Hydrous minerals in the
1010 Haemyeong Jung
http://dx.doi.org/10.1007/s12303-017-0045-1
http://www.springer.com/journal/12303
mantle wedge and the maximum depth of subduction thrust earth-
quakes. Geophysical Research Letters, 26, 2517–2520.
Precigout, J. and Hirth, G., 2014, B-type olivine fabric induced by grain
boundary sliding. Earth and Planetary Science Letters, 395, 231–
240.
Précigout, J., Prigent, C., Palasse, L., and Pochon, A., 2017, Water pumping
in mantle shear zones. Nature Communications, 8, 15736. doi:10.1038/
ncomms15736
Puelles, P., Ábalos, B., Gil Ibarguchi, J.I., Sarrionandia, F., Carracedo,
M., and Fernández-Armas, S., 2016, Petrofabric and seismic prop-
erties of lithospheric mantle xenoliths from the Calatrava volcanic
field (Central Spain). Tectonophysics, 683, 200–215.
Puelles, P., Gil Ibarguchi, J.I., Beranoaguirre, A., and Ábalos, B., 2012,
Mantle wedge deformation recorded by high-temperature peridot-
ite fabric superposition and hydrous retrogression (Limo massif,
Cabo Ortegal, NW Spain). International Journal of Earth Sciences,
101, 1835–1853.
Raleigh, C.B., 1965, Glide mechanisms in experimentally deformed
minerals. Science, 150, 739–741.
Raleigh, C.B., 1968, Mechanisms of plastic deformation of olivine. Jour-
nal of Geophysical Research, 73, 5391–5406.
Ringwood, A.E., 1970, Phase transformations and the constutution of
the mantle. Physics of the Earth and Planetary Interiors, 3, 109–155.
Russo, R.M. and Silver, P.G., 1994, Trench-parallel flow beneath the
Nazca plate from seismic anisotropy. Science, 263, 1105–1111.
Savage, M.K., 1999, Seismic anisotropy and mantle deformation: what
have we learned from shear wave splitting? Reviews of Geophysics,
37, 65–106.
Schmidt, M.W. and Poli, S., 1998, Experimentally based water budgets
for dehydrating slabs and consequences for arc magma generation.
Earth and Planetary Science Letters, 163, 361–379.
Schwerdtner, W.M., 1964, Preferred orientation of hornblende in a
banded hornblende gneiss. American Journal of Science, 262, 1212–
1229.
Siegesmund, S., Takeshita, T., and Kern, H., 1989, Anisotropy of Vp and
Vs in an amphibolite of the deeper crust and its relationship to the
mineralogical, microstructural and textural characteristics of the
rock. Tectonophysics, 157, 25–38.
Siegesmund, S. and Vollbrecht, A., 1991, Complete seismic properties
obtained from microcrack fabrics and textures in an amphibolite
from the Ivrea zone, Western Alps, Italy. Tectonophysics, 199, 13–24.
Signorelli, J. and Tommasi, A., 2015, Modeling the effect of subgrain
rotation recrystallization on the evolution of olivine crystal preferred
orientations in simple shear. Earth and Planetary Science Letters,
430, 356–366.
Silver, P.G., 1996, Seismic anisotropy beneath the continents: probing
the depths of geology. Annual Review of Earth and Planetary Sci-
ences, 24, 385.
Silver, P.G. and Chan, W.W., 1991, Shear-wave splitting and subconti-
nental mantle deformation. Journal of Geophysical Research: Solid
Earth, 96, 16429–16454.
Skemer, P. and Hansen, L.N., 2016, Inferring upper-mantle flow from
seismic anisotropy: an experimental perspective. Tectonophysics,
668, 1–14.
Skemer, P., Katayama, I., and 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, 43–51.
Skemer, P., Warren, J.M., Hansen, L.N., Hirth, G., and Kelemen, P.B.,
2013, The influence of water and LPO on the initiation and evolu-
tion of mantle shear zones. Earth and Planetary Science Letters,
375, 222–233.
Skemer, P., Warren, J.M., Kelemen, P.B., and Hirth, G., 2010, Micro-
structural and rheological evolution of a mantle shear zone. Jour-
nal of Petrology, 51, 43–53.
Smith, G.P., Wiens, D.A., Fischer, K.M., Dorman, L.M., Webb, S.C., and
Hildebrand, J.A., 2001, A complex pattern of mantle flow in the
Lau backarc. Science, 292, 713–716.
Soda, Y. and Takagi, H., 2010, Sequential deformation from serpentinite
mylonite to metasomatic rocks along the Sashu fault, SW Japan.
Journal of Structural Geology, 32, 792–802.
Song, T.-R.A. and Kawakatsu, H., 2012, Subduction of oceanic astheno-
sphere: evidence from sub-slab seismic anisotropy. Geophysical
Research Letters, 39, L17301. doi:10.1029/2012GL052639
Song, T.-R.A. and Kawakatsu, H., 2013, Subduction of oceanic astheno-
sphere: a critical appraisal in central Alaska. Earth and Planetary
Science Letters, 367, 82–94.
Soustelle, V. and Manthilake, G., 2017, Deformation of olivine-ortho-
pyroxene aggregates at high pressure and temperature: implica-
tions for the seismic properties of the asthenosphere. Tectonophysics,
694, 385–399.
Soustelle, V., Tommasi, A., Demouchy, S., and Ionov, D.A., 2009, Defor-
mation and fluid-rock interaction in the supra-subduction mantle:
microstructures and water contents in peridotite xenoliths from the
Avacha volcano, Kamchatka. Journal of Petrology, 51, 363–394.
Sundberg, M. and Cooper, R.F., 2008, Crystallographic preferred orien-
tation produced by diffusional creep of harzburgite: effects of
chemical interactions among phases during plastic flow. Journal of
Geophysical Research: Solid Earth, 113, B12208. doi:10.1029/
2008JB005618
Tasaka, M., Michibayashi, K., and 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, 747–757.
Tatham, D.J., Lloyd, G.E., Butler, R.W.H., and Casey, M., 2008, Amphi-
bole and lower crustal seismic properties. Earth and Planetary Sci-
ence Letters, 267, 118–128.
Tommasi, A., Mainprice, D., Canova, G., and Chastel, Y., 2000, Visco-
plastic self-consistent and equilibrium-based modeling of olivine lat-
tice preferred orientations: implications for the upper mantle seismic
anisotropy. Journal of Geophysical Research: Solid Earth, 105,
7893–7908.
Tommasi, A. and Vauchez, A., 2015, Heterogeneity and anisotropy in
the lithospheric mantle. Tectonophysics, 661, 11–37.
Tommasi, A., Vauchez, A., and Ionov, D.A., 2008, Deformation, static
recrystallization, and reactive melt transport in shallow subconti-
nental mantle xenoliths (Tok Cenozoic volcanic field, SE Siberia).
Earth and Planetary Science Letters, 272, 65–77.
Ulmer, P. and Trommsdorff, V., 1995, Serpentine stability to mantle
depths and subduction-related magmatism. Science, 268, 858–861.
van Keken, P.E., Hacker, B.R., Syracuse, E.M., and Abers, G.A., 2011,
Crystal preferred orientations of olivine and hydrous minerals 1011
http://www.springer.com/journal/12303
http://dx.doi.org/10.1007/s12303-017-0045-1
Subduction factory: 4. Depth-dependent flux of H2O from sub-
ducting slabs worldwide. Journal of Geophysical Research: Solid
Earth, 116, B01401. doi:10.1029/2010JB007922
Verma, R.K., 1960, Elasticity of some high-density crystals. Journal of
Geophysical Research, 65, 757–766.
Wagner, L.S., Fouch, M.J., James, D.E., and Long, M.D., 2013, The role
of hydrous phases in the formation of trench parallel anisotropy:
evidence from Rayleigh waves in Cascadia. Geophysical Research
Letters, 40, 2642–2646.
Wallace, P.J., 1998, Water and partial melting in mantle plumes: infer-
ences from the dissolved H2O concentrations of Hawaiian basaltic
magmas. Geophysical Research Letters, 25, 3639–3642.
Wallis, D., Lloyd, G.E., Phillips, R.J., Parsons, A.J., and Walshaw, R.D.,
2015, Low effective fault strength due to frictional-viscous flow in
phyllonites, Karakoram Fault Zone, NW India. Journal of Struc-
tural Geology, 77, 45–61.
Wang, J. and Zhao, D., 2013, P-wave tomography for 3-D radial and
azimuthal anisotropy of Tohoku and Kyushu subduction zones.
Geophysical Journal International, 193, 1166–1181.
Wang, Q., Xia, Q.K., O'Reilly, S.Y., Griffin, W.L., Beyer, E.E., and Bruec-
kner, H.K., 2013a, Pressure- and stress-induced fabric transition in
olivine from peridotites in the Western Gneiss Region (Norway):
implications for mantle seismic anisotropy. Journal of Metamor-
phic Geology, 31, 93–111.
Wang, Y., Zhang, J., and Shi, F., 2013b, The origin and geophysical
implications of a weak C-type olivine fabric in the Xugou ultrahigh
pressure garnet peridotite. Earth and Planetary Science Letters,
376, 63–73.
Warren, J.M., Hirth, G., and Kelemen, P.B., 2008, Evolution of olivine
lattice preferred orientation during simple shear in the mantle. Earth
and Planetary Science Letters, 272, 501–512.
Watanabe, T., Shirasugi, Y., and Michibayashi, K., 2014, A new method
for calculating seismic velocities in rocks containing strongly
dimensionally anisotropic mineral grains and its application to
antigorite-bearing serpentinite mylonites. Earth and Planetary Sci-
ence Letters, 391, 24–35.
Webber, C.E., Little, T., Newman, J., and Tikoff, B., 2008, Fabric super-
position in upper mantle peridotite, Red Hills, New Zealand. Jour-
nal of Structural Geology, 30, 1412–1428.
Weiss, T., Siegesmund, S., Rabbel, W., Bohlen, T., and Pohl, M., 1999,
Seismic velocities and anisotropy of the lower continental crust: a
review. Pure and Applied Geophysics, 156, 97–122.
Wenk, H.R. and Tomé, C.N., 1999, Modeling dynamic recrystallization
of olivine aggregates deformed in simple shear. Journal of Geo-
physical Research: Solid Earth, 104, 25513–25527.
Xu, Z.Q., Wang, Q., Ji, S.C., Chen, J., Zeng, L.S., Yang, J.S., Chen, F.Y.,
Liang, F.H., and Wenk, H.R., 2006, Petrofabrics and seismic prop-
erties of garnet peridotite from the UHP Sulu terrane (China):
implications for olivine deformation mechanism in a cold and dry
subducting continental slab. Tectonophysics, 421, 111–127.
Yamamoto, T., Ando, J.-i., Tomioka, N., and Kobayashi, T., 2017, Defor-
mation history of Pinatubo peridotite xenoliths: constraints from
microstructural observation and determination of olivine slip sys-
tems. Physics and Chemistry of Minerals, 44, 247–262.
Yang, K., Hidas, K., Falus, G., Szabo, C., Nam, B., Kovacs, I., and Hwang,
B., 2010, Relation between mantle shear zone deformation and
metasomatism in spinel peridotite xenoliths of Jeju Island (South
Korea): evidence from olivine CPO and trace elements. Journal of
Geodynamics, 50, 424–440.
Zhang, S.Q. and Karato, S., 1995, Lattice preferred orientation of oliv-
ine aggregates deformed in simple shear, Nature, 375, 774–777.
Zhang, S.Q., Karato, S., Fitz Gerald, J.D., Faul, U.H., and Zhou, Y., 2000,
Simple shear deformation of olivine aggregates. Tectonophysics, 316,
133–152.
Zhao, D., Yu, S., and Liu, X., 2016, Seismic anisotropy tomography: new
insight into subduction dynamics. Gondwana Research, 33, 24–43.
... Объяснить это можно следующими главными причинами: 1) в породе значительное содержание ортопироксена, который является более компетентной («сильной») фазой по сравнению с оливином [Carter, 1976;Yamamoto et al., 2008;, 2) из сочетания двух условий, указанных в первом пункте, следует, что макроскопические структурные элементы породы (S, L), определяемые ориентировкой минералов по форме, определяются пироксеном, а ориентировка зерен оливина является результатом их «приспособления» к таковой пироксена. В ортопироксене по текстурным диаграммам четко определяется активная система трансляционного скольжения (010) [001], характерная для типа BC [Jung, 2017]. ...
... Активные системы скольжения позволяет выявить локальный анализ разориентировок порфирокластов, они соответствуют двум основным: (010) [100] и (010) [001]. Первая система скольжения характеризует тип A, который является наиболее распространенным типом текстур в деплетированных реститовых ультрамафитах [Nicolas et al., 1971;Carter, 1976;, а вторая -тип B, довольно редкий, но интенсивно исследованный в последние десятилетия Jung, 2017]. Считается, что тип A текстуры Рис. 5. Набор микроструктурных карт для двух исследованных участков (66 и 67) в образце Ке-112-11 ...
Article
Full-text available
At the example of ultramafic rocks and chromitites of the Urals, the paper describes the possibilities of microstructural study of geological samples using one of the modern microstructural methods - electron backscattered diffraction (EBSD). This method has a number of advantages compared to previously used optical methods using a universal stage. The main advantages of the method: high accuracy, automation, objectivity, wide range of materials studied. However, there are some difficulties, which consist mainly in the susceptibility of geological materials to secondary changes and superimposed metamorphic processes. The article discusses various possibilities for processing primary materials and obtaining representative data on the microstructure of samples of ultramafic rocks and chromitites from various massifs of the Urals.
... Neoblast formation in both the tectonitic mixed matrix and around orthopyroxene porphyroclasts at the tectonite-mylonite transition show weak dependence on the foliation. Additionally, tectonitic mixed-matrix orthopyroxene neoblasts have a CPO with [001] subperpendicular to the foliation, which is atypical for deformation-induced CPOs (e.g., Jung, 2017). Instead, the CPO of opx neoblasts is in most cases strongly connected to the parent clast orientation. ...
... An increasing number of neoblasts and the stabilization of their small grain size by mixing in turn enhances the share of the grain-size-sensitive deformation mechanism corroborated by the decrease in the CPO strength of olivine. Concomitant with the increase in the neoblast tails, the dominant olivine CPO changes within ∼ 150 m distance to the NW-B from an A-type CPO, which indicates low water and intermediate stress conditions, to an AG-type CPO or occasionally a B-type CPO, which indicates increased water content and high stress (e.g., Jung, 2017). The increased presence of olivine B-type CPOs towards the NW-B was formerly interpreted to result from grain boundary sliding (GBS) rather from a change in the dominant slip system (Précigout and Hirth, 2014). ...
Article
Full-text available
Strain localization in upper-mantle shear zones by grain size reduction and the activation of grain-size-sensitive deformation mechanisms is closely linked to phase mixing. With its mylonitic grain size (50–100 µm) and well-mixed phase assemblage, the kilometer-scale shear zone at the northwestern boundary of the Ronda peridotite is, in this respect, no exception. In transects across the high-strain mylonitic into the low-strain tectonitic part of this shear zone, the following four dominant microstructural domains were identified: (1) olivine-rich matrix, (2) mixed matrix, (3) neoblast tails of clinopyroxene porphyroclasts, and (4) neoblast tails of orthopyroxene porphyroclasts. In these domains, phase mixing and its impact on strain localization were investigated by a combination of microstructural (optical microscopy), textural (EBSD), and geochemical (EPMA) analysis. The dominant microstructural domain of all samples is the mixed matrix composed of olivine, orthopyroxene, and clinopyroxene. Its homogenous distribution of interstitial pyroxenes contradicts mechanical mixing. Instead, extensive phase mixing under near-steady-state conditions is documented by the constant grain size and by phase boundary percentages > 60 % for the entire mylonitic unit and all the microstructural domains. Lobate phase boundaries, homogenous phase mixing, and secondary-phase distribution, as well as continuous geochemical trends that are independent of the microstructural domain, point to a reaction-driven, metasomatic formation of the mixed matrix and pyroxene porphyroclast tails in the entire shear zone. An OH-bearing metasomatism by small fractions of evolved melts is indicated by amphibole abundance in pyroxene neoblast tails, olivine B-type-crystallographic-preferred orientations (CPOs), and the microstructural consistency of the garnet–spinel (grt–spl) mylonites from both major peridotite massifs of the Gibraltar arc, Ronda, and Beni Bousera (Morocco). The established syn-deformational temperature of 800–900 ∘C at 1.95–2.00 GPa suggests that the metasomatism did not reset the equilibrium temperatures. Consistent geochemistry and phase assemblage in mylonites and tectonites but a change from equiaxial (tectonites) to wedge-shaped pyroxenes aligned parallel to the foliation (mylonites) point to a pre- to syn-deformational metasomatism, with the potential annealing of the tectonites. For the mylonitic mixed matrix, wedge-shaped pyroxenes, and neoblast tail formation in pyroxene porphyroclast stress shadows point to the activity of incongruent dissolution–precipitation creep. Apart from the dissolution–precipitation creep, strong CPOs of all major phases (ol, opx, and cpx) suggest dislocation creep as being the major deformation mechanism in the entire shear zone.
... Antigorite has a highly anisotropic phyllosilicate structure 33 , which is very compressible in the cross-plane direction (the [001] direction), and relatively stiff in the in-plane direction 34,36 . Therefore, antigorite is characterized by a weak rheology and, during shear deformation, it likely forms a CPO, in which the [001] direction tends to align perpendicularly to the shear plane/foliation 38,39,59 . In other words, the serpentinized rocks of the slab accommodate most of the shear deformation developed during subduction 60 by orienting antigorite's [001] direction perpendicularly to slab's dip 38,39 . ...
Article
Full-text available
Double seismic zones (DSZs) are a feature of some subducting slabs, where intermediate-depth earthquakes (~70–300 km) align along two separate planes. The upper seismic plane is generally attributed to dehydration embrittlement, whereas mechanisms forming the lower seismic plane are still debated. Thermal conductivity of slab minerals is expected to control the temperature evolution of subducting slabs, and therefore their seismicity. However, effects of the potential anisotropic thermal conductivity of layered serpentine minerals with crystal preferred orientation on slab’s thermal evolution remain poorly understood. Here we measure the lattice thermal conductivity of antigorite, a hydrous serpentine mineral, along its crystallographic b- and c-axis at relevant high pressure-temperature conditions of subduction. We find that antigorite’s thermal conductivity along the c-axis is ~3–4 folds smaller than the b-axis. Our numerical models further reveal that when the low-thermal-conductivity c-axis is aligned normal to the slab dip, antigorite’s strongly anisotropic thermal conductivity enables heating at the top portion of the slab, facilitating dehydration embrittlement that causes the seismicity in the upper plane of DSZs. Potentially, the antigorite’s thermal insulating effect also hinders the dissipation of frictional heat inside shear zones, promoting thermal runaway along serpentinized faults that could trigger intermediate-depth earthquakes.
... These minerals are elastically anisotropic (Bezacier et al., 2010;Mao et al., 2007;Sinogeikin et al., 2000). As the crystal-preferred orientation (CPO) of elastically anisotropic minerals can cause seismic anisotropy (Almqvist and Mainprice, 2017;Jung, 2017), deformed blueschists have been studied to understand the seismic anisotropies of subducting slabs (Cao and Jung, 2016;Cao et al., 2014;Fujimoto et al., 2010;Ha et al., 2019;Kim et al., 2013b). Furthermore, few studies investigated the role of blueschist on the formation of low-velocity layers observed in subducting slabs (Cao and Jung, 2016;Cao et al., 2013;Fujimoto et al., 2010;Kim et al., 2013b;Park and Jung, 2022). ...
... Seismic anisotropy is produced in large part by crystallographic preferred orientation (CPO) of anisotropic minerals at depths below the closure of open cracks (Almqvist & Mainprice, 2017;Brownlee et al., 2017;Jung, 2017;Mainprice & Humbert, 1994;Mainprice & Nicolas, 1989). CPO formation is typical of the viscous deformation that occurs along the plate interface below the seismogenic zone (Mainprice & Nicolas, 1989), and thus, parts of the subducting slab may develop significant seismic anisotropy if CPOs are developed in anisotropic phases such as amphibole or mica. ...
Article
Full-text available
Seismic anisotropy constitutes a useful tool for imaging the structure along the plate interface in subduction zones, but the seismic properties of mafic blueschists, a common rock type in subduction zones, remain poorly constrained. We applied the technique of electron backscatter diffraction (EBSD) based petrofabric analysis to calculate the seismic anisotropies of 14 naturally deformed mafic blueschists at dry, ambient conditions. The ductilely deformed blueschists were collected from terranes with inferred peak P‐T conditions applicable to subducting slabs at or near the plate interface in active subduction zones. Epidote blueschists display the greatest P wave anisotropy range (AVp ∼7%–20%), while lawsonite blueschist AVp ranges from ∼2% to 10%. S wave anisotropies generate shear wave splitting delay times up to ∼0.1 s over a thickness of 5 km. AVp magnitude increases with glaucophane abundance (from areal EBSD measurements), decreases with increasing epidote or lawsonite abundance, and is enhanced by glaucophane crystallographic preferred orientation (CPO) strength. Two‐phase rock recipe models provide further evidence of the primary role of glaucophane, epidote, and lawsonite in generating blueschist seismic anisotropy. The symmetry of P wave velocity patterns reflects the deformation‐induced CPO type in glaucophane—an effect previously observed for hornblende on amphibolite P wave anisotropy. The distinctive seismic properties that distinguish blueschist from other subduction zone rock types and the strong correlation between anisotropy magnitude/symmetry and glaucophane CPO suggest that seismic anisotropy may be a useful tool in mapping the extent and deformation of blueschists along the interface, and the blueschist‐eclogite transition in active subduction zones.
Article
Strain localization processes in the continental crust generate faults and ductile shear zones over a broad range of scales affecting the long-term lithosphere deformation and the mechanical response of faults during the seismic cycle. Seismic anisotropy originated within the continental crust can be applied to deduce the kinematics and structures within orogens and is widely attributed to regionally aligned minerals, e. g., hornblende. However, naturally deformed rocks commonly show various structural layers (e.g., strain localization layers). It is necessary to reveal how both varying amphibole contents and fabrics in the structural layers of strain localization impact seismic property and its interpretations in terms of deformation. We present microstructures, petrofabrics, and calculate seismic properties of deformed amphibolite with the microstructures ranging from mylonite to ultramylonite. The transition from mylonite to ultramylonite is accompanied by a slight decrease of amphibole grain size, a disintegration of amphibole and plagioclase aggregates, and amphibole aspect ratio increase (from 1.68 to 2.23), concomitant with the precipitation of feldspar and/or quartz between amphibole grains. The intensities of amphibole crystallographic preferred orientations (CPOs) show a progressively increasing trend from mylonitic layers to homogeneous ultramylonitic layers, as indicated by the JAm index increasing from 1.9–4.0 for the mylonitic layers and 4.0–4.8 for the transition layer, to 5.1–6.9 for the ultramylonitic layers. The CPO patterns are nearly random for plagioclase and quartz. Polycrystalline amphibole aggregates in the amphibolitic mylonite deform by diffusion, mechanical rotation, and weak dislocation creep, and develop CPOs collectively. The polymineralic matrix (such as quartz and plagioclase) of the mylonite and the ultramylonite deform dominantly by dissolution-precipitation, combined with weak dislocation creep. The mean P and S wave velocities are estimated to be 6.3 and 3.5 km/s, respectively, for three layers of the mylonitic amphibolite. The respective maximum P and S anisotropies are 1.5%–6.4% and 1.8%–4.5% for the mylonite layers of the mylonitic amphibolite, and 6.0%–6.9% and 4.5%–5.0% for the transition layers; but for the ultramylonite layers, these values increase significantly to 8.0%–9.1% and 5.1%–6.0%, respectively. Furthermore, increasing strain (strain localization) generates significant variations in the geometry of the seismic anisotropy. This effect, coupled with the geographical orientations of structures in the Hengshan-Wutai-Fuping complex terrains, can generate substantial variations in the orientation and magnitude of seismic anisotropy for the continental crust as measured by the existing North China Geoscience Transect. Thickened amphibolitic layers by extensively folding or thrusting in the middle crust can explain the strong shear wave splitting and the tectonic boundary parallel fast shear wave polarization beneath the Hengshan-Wutai-Fuping complex terrains. Therefore, signals of seismic anisotropy varying with depth in the deforming continent crust need not deduce depth-varying kinematics or/and tectonic decoupling.
Article
This study investigates the relationship between olivine fabric transitions and seismic anisotropy in mantle shear zones, focusing on the Yugu peridotite body in the Korean Peninsula. Olivine, a key mineral in the upper mantle, influences seismic anisotropy through deformation fabrics. The Yugu peridotite body provides insights into these processes within mantle shear zones. Based on microstructures and olivine fabric transitions, this study categorizes peridotites into three groups (PM: proto-mylonite, M: mylonite, and UM: ultra-mylonite) and explores their seismic properties. The findings highlight a direct correlation between olivine fabric strength and seismic anisotropy. Group PM peridotites exhibited higher seismic anisotropy compared to those in Group M and UM peridotites. This study emphasizes that variations in seismic anisotropy within mantle shear zones are primarily driven by the strength of olivine fabric, with additional influences from fabric type and rock composition. The calculated anisotropic layer thickness supports the observation that seismic anisotropy is significantly larger away from the UM peridotites. These insights contribute to understanding of the complex interplay among olivine fabric, seismic anisotropy, and geological processes within mantle shear zones. The implications of this study extend to the improved interpretation of seismic data related to shear localization during peridotite exhumation.
Article
Increasing evidence from seismic methods shows that anisotropy within subduction zones should consist of multiple layers. To test this, we calculate and model shear wave splitting across the Alaska-Aleutians Subduction Zone (AASZ), where previous studies have argued for separate layers of anisotropy in the subslab, slab, and mantle wedge. We present an updated teleseismic splitting catalog along the span of the AASZ, which has many broadband seismometers recently upgraded to three components. Splitting observations are sparse in the Western Aleutians, and fast directions are oriented generally trench parallel. There are significantly more splitting measurements further east along the AASZ. We identify six regions in the Central and Eastern Aleutians, Alaskan Peninsula, and Cook Inlet with a high density of splits suitable for multilayered anisotropy analyses. These regions were tested for multilayer anisotropy, and for five of the six regions we favor multiple layers over a single layer of anisotropy. We find that the optimal setup for our models is one with a dipping middle layer oriented parallel to paleospreading. A prominent feature of our modeling is that fast directions above and below the dipping layer are generally oriented parallel to the strike of the slab. Additionally, we lay out a framework for robust and statistically reliable multilayer shear wave splitting modeling.
Article
Sheared peridotite xenoliths are snapshots of deformation processes that occur in the cratonic mantle shortly before their entrainment by kimberlites. The process of deformation that caused the shearing has, however, been highly debated since the 1970s and remains uncertain. To investigate the processes involved in the deformation, we have studied twelve sheared peridotites from Late Cretaceous (90 Ma) kimberlites in northern Lesotho, on the southeast margin of the Kaapvaal craton. Various deformation textures are represented, ranging from porphyroclastic to fluidal mosaic. Our sample suite consists of eleven garnet peridotites, with various amounts of clinopyroxene, and one garnet-free spinel peridotite with a small amount of clinopyroxene. All of the peridotites are depleted in Fe, and the Mg# of olivine and orthopyroxene range from 91 – 94. Three groups of sheared peridotites are present and have been identified primarily on the basis of Ca contents of olivine and orthopyroxene. The porphyroclasts preserve pre-deformation P-T conditions of 3.5 – 4.5 GPa and 900 – 1100°C (Group I), 5 – 5.5 GPa and 1200 – 1250°C (Group II) and 6±0.5 GPa and 1400±50°C (Group III). Group III samples lie above the 40mW/m² conductive geothermal gradient, indicating thermal perturbation prior to deformation. The sheared peridotites from Lesotho were affected by various metasomatic events. Pre-deformation metasomatism, involving melts and fluids, is recorded in the porphyroclasts. In Group II and III samples the clinopyroxene porphyroclasts have similar compositions to Cr-rich and Cr-poor clinopyroxene megacrysts, respectively, that have previously described from southern African kimberlites. This suggests a relationship between them. Younger pre-deformation metasomatism is preserved in a zoned garnet from Group II (enrichment in Ti, Zr, Y+HREE) and orthopyroxene in a Group I sample. The latter exhibits a complex zonation, with a highly-enriched (Fe, Ti) inner rim and a less-enriched outer rim. These enrichments must have occurred shortly before deformation. Metasomatism during deformation is revealed by the complex chemical changes recorded in olivine neoblasts with, depending on the sample, increasing or decreasing contents of Ti, Ca, Al, Cr, Mn and Na. Crystallographic preferred orientations of olivine neoblasts are consistent with bimodal, B, C, E, AG-type fabrics and indicate the presence of a hydrous metasomatic agent. We suggest that, akin to the shallower sheared peridotites (Group I), Group II and III were influenced by early (proto-)kimberlite melt pulses and propose the following model: (Proto-)kimberlitic melts invaded the lower lithosphere. These melts followed narrow shear zone networks, produced by deformation at the lithosphere-asthenosphere-boundary, heated and metasomatized the surrounding peridotites and were responsible for megacryst crystallization. Sheared peridotites from close to the melt conduits (Group III) have compositions comparable to Cr-poor megacrysts, while those located at a greater distance (Group II) resemble Cr-rich megacrysts. Reactive infiltration of volatile-rich proto-kimberlite melts caused rheologically weakening of olivine in the lithospheric mantle. The consequence of this positive feedback mechanism of metasomatism, weakening and deformation -- due to the high magmatic and metasomatic activity in the Late Cretaceous -- is the progressive perforation of the lower Kaapvaal lithosphere by rheologically weak zones and the destruction of the protecting dry and depleted layer at its base. This could have caused the observed thinning and destabilization of the lower lithosphere below the southern margin of the Kaapvaal craton.
Article
Full-text available
Water plays an important role in geological processes. Providing constraints on what may influence the distribution of aqueous fluids is thus crucial to understanding how water impacts Earth’s geodynamics. Here we demonstrate that ductile flow exerts a dynamic control on water-rich fluid circulation in mantle shear zones. Based on amphibole distribution and using dislocation slip-systems as a proxy for syn-tectonic water content in olivine, we highlight fluid accumulation around fine-grained layers dominated by grain-size-sensitive creep. This fluid aggregation correlates with dislocation creep-accommodated strain that localises in water-rich layers. We also give evidence of cracking induced by fluid pressure where the highest amount of water is expected. These results emphasize long-term fluid pumping attributed to creep cavitation and associated phase nucleation during grain size reduction. Considering the ubiquitous process of grain size reduction during strain localisation, our findings shed light on multiple fluid reservoirs in the crust and mantle.
Article
Full-text available
To advance our understanding of deformation characteristics, rheological behaviors and tectonic evolution of the forearc lithospheric mantle, we analyzed mineral fabrics for a large spinel-bearing ultramafic massif in the Songshugou area in the Qinling orogenic belt, central China. In the spinel-poor coarse-grained dunite, stronger A-/D-type and weaker C-type-like fabrics were found, whereas the spinel-rich coarse-grained dunite displayed a comparatively stronger B-type-like fabric. These olivine fabrics are high-T fabrics influenced by the presence of melt, in which B- and C-type-like fabrics are inferred to be produced by melt-assisted grain boundary sliding during synkinematic high-T melt‒rock reactions. In contrast, the spinel-poor porphyroclastic and fine-grained dunites present weak AG- and B-type-like fabrics, respectively. Their olivine fabrics (low-T fabrics) are inferred to transform from A-/D-type fabric in their coarse-grained counterparts possibly through mylonitization process assisted by low-T fluid‒rock reactions, during which strain was accommodated by the fluid-enhanced dislocation slip and/or fluid-assisted grain boundary sliding processes. Combined with the tectonic results of our previous work [Cao et al., 2016], the high-T olivine fabrics are probably related to a young and warm forearc mantle where intense partial melting and high-T boninitic melt‒rock reactions prevalently occurred, whereas the low-T olivine fabrics likely reflect the evolving tectonic settings through the cooling forearc mantle to a continental lower crust in a collisional orogeny where low-T fluid‒rock reactions were pervasively activated. These low-T olivine fabrics imply that, though cold, the forearc lithospheric mantle may be locally weak (∼20‒30 MPa), allowing ductile deformation to occur at a geologically significant strain rate. This article is protected by copyright. All rights reserved.
Article
Full-text available
The deformation history of the Pinatubo peridotite xenoliths was estimated on the basis of the microstructural observations and the determination of olivine slip systems. The latter was performed by using three methods: lattice-preferred orientation (LPO), crystallographic analysis of subgrain boundaries, and direct characterization of dislocations. The Pinatubo peridotites are composed of coarse olivine grains containing numerous fluid inclusions and some fine aggregates of orthopyroxene and amphibole grains, which implies intense fluid–rock interaction. The development of euhedral fine recrystallized olivine grains along the healed cracks within the coarse olivine grains suggests that the strain-free grains were nucleated and grew during static recovery. The LPO patterns and the analyses of subgrain boundaries indicate the activation of a [100]{0kl} slip system that developed under high temperature, low pressure, and dry deformation conditions. Although dislocations showing the [100]{0kl} slip system are dominantly observed, the other slip systems which could be formed by the deformation under moderate–high water content and lower-temperature conditions are also developed. The discrepancy between the results of dislocation characterization and the other two methods might have been caused by fulfilling the von Mises criterion or overprinting dislocation microstructures. Either way, the possible deformation history of the Pinatubo peridotites can be explained by the following scenario. The peridotites plastically moved from the back-arc to the fore-arc adjacent region, where CO2-rich saline fluid was trapped, by the corner flow of a mantle wedge. They were then annealed and metasomatized during entrapment of the upwelling magma.
Article
Full-text available
It is widely accepted that water-rich serpentinite domains are commonly present in the mantle above shallow subducting slabs and play key roles in controlling the geochemical cycling and physical properties of subduction zones. Thermal and petrological models show the dominant serpentine mineral is antigorite. However, there is no good consensus on the amount, distribution and alignment of this mineral. Seismic velocities are commonly used to identify antigorite-rich domains, but antigorite is highly-anisotropic and depending on the seismic ray path, its properties can be very difficult to distinguish from non-hydrated olivine-rich mantle. Here, we utilize this anisotropy and show how an analysis of seismic anisotropy that incorporates measured ray path geometries in the Ryukyu arc can constrain the distribution, orientation and amount of antigorite. We find more than 54% of the wedge must consist of antigorite and the alignment must change from vertically aligned to parallel to the slab. This orientation change suggests convective flow in the hydrated forearc mantle. Shear wave splitting analysis in other subduction zones indicates large-scale serpentinization and forearc mantle convection are likely to be more widespread than generally recognized. The view that the forearc mantle of cold subduction zones is dry needs to be reassessed.
Article
Large-scale emplaced peridotite bodies may provide insights into plastic deformation process and tectonic evolution in the mantle shear zone. Due to the complexity of deformation microstructures and processes in natural mantle rocks, the evolution of pre-existing olivine fabrics is still not well understood. In this study, we examine well-preserved transitional characteristics of microstructures and olivine fabrics developed in a mantle shear zone from the Yugu peridotite body, the Gyeonggi Massif, Korean Peninsula. The Yugu peridotite body predominantly comprises spinel harzburgite together with minor lherzolite, dunite, and clinopyroxenite. We classified highly deformed peridotites into four textural types based on their microstructural characteristics: proto-mylonite; proto-mylonite to mylonite transition; mylonite; and ultra-mylonite. Olivine fabrics changed from A-type (proto-mylonite) via D-type (mylonite) to E-type (ultra-mylonite). Olivine fabric transition is interpreted as occurring under hydrous conditions at low temperature and high strain, because of characteristics such as Ti-clinohumite defects (and serpentine) and fluid inclusion trails in olivine, and a hydrous mineral (pargasite) in the matrix, especially in the ultra-mylonitic peridotites. Even though the ultra-mylonitic peridotites contained extremely small (24–30 μm) olivine neoblasts, the olivine fabrics showed a distinct (E-type) pattern rather than a random one. Analysis of the lattice preferred orientation strength, dislocation microstructures, recrystallized grain-size, and deformation mechanism maps of olivine suggest that the proto-mylonitic, mylonitic, and ultra-mylonitic peridotites were deformed by dislocation creep (A-type), dislocation-accommodated grain-boundary sliding (D-type), and combination of dislocation and diffusion creep (E-type), respectively.
Article
Progress in seismic methodology and ambitious large-scale seismic projects are enabling high-resolution imaging of the continental crust. The ability to constrain interpretations of crustal seismic data is based on laboratory measurements on rock samples and calculations of seismic properties. Seismic velocity calculations and their directional dependence are based on the rock micro fabric, which consists of mineral aggregate properties including crystallographic preferred orientation (CPO), grain shape and distribution, grain boundary distribution, and misorientation within grains. Single mineral elastic constants and density are crucial for predicting seismic velocities, preferably at conditions that span the crust. However, high temperature and pressure properties are not as common as elastic constants at standard temperature and pressure (STP) at atmospheric conditions. Continental crust has a very diverse mineralogy, however a select number appear to dominate seismic properties because of their high volume fraction contribution. Calculations of micro fabric-based seismic properties and anisotropy are performed with averaging methods that in their simplest form takes into account the CPO and modal mineral composition. More complex methods can take into account other microstructural characteristics, including the grain shape and distribution of mineral grains, and cracks and pores. A challenge for the geophysics and rock physics communities is the separation of intrinsic factors affecting seismic anisotropy, due to properties of crystals within a rock and apparent sources due to extrinsic factors like cracks, fractures and alteration. This is of particular importance when trying to deduce the state of crustal composition and deformation from seismic parameters.
Article
We examine spatially varying patterns of sub-slab anisotropy derived from geodynamic models of subduction beneath Central America and Tonga. Invoking a variety of anisotropic fabrics, we compare the predicted sub-slab anisotropy of these models against source-side, shear-wave splitting observations using realistic ray paths. We find that in both regions fabric type has a strong impact on predicted shear-wave splitting. In Tonga, where three-dimensional (3D) return flow dominates, E-type olivine lattice-preferred orientation (LPO) fabric predicts a sub-slab mantle anisotropy that best matches observations. In Central America, where entrained flow dominates, anisotropy from C-type LPO fabric yields the best fit. This highlights the importance of fabric type when interrogating geodynamic models because different regions may be characterized by different LPO fabrics. A primary controller of fabric type is water content. Taken at face value, these results then suggest the sub-slab mantle beneath Tonga is less well hydrated than that beneath Central America.
Article
The effect of pressure, temperature and composition on the development of crystal preferred orientations (CPO) and seismic properties of olivine-orthopyroxene aggregates were investigated using samples containing olivine and 12.5, 25 and 50 vol.% of orthopyroxene. The samples were deformed in simple-shear at a constant strain-rate of 10− 4 s− 1 with total shear strains between 0.5 and 1.3, at pressures of 3, 5 and 8 GPa and temperatures of 1300, 1400 and 1500 °C, respectively. Olivine's CPO vary as a function of the orthopyroxene content. All samples have their olivine [010] axes normal to the foliation. Samples with 12.5 and 25% orthopyroxene have their [001] axes parallel to the lineation (B-type), whereas the samples with 50% orthopyroxene have their [100] axes oriented parallel to the lineation (A-type). At 3 GPa, we propose that olivine CPO may result from a variation between two types of diffusion accommodated grain boundary sliding (difGBS) mechanisms. At higher pressure, the relative contribution of difGBS and dislocation related mechanisms depends on the volume of secondary phases. For low orthopyroxene contents, dislocation related mechanisms prevail and induce the development of B-type CPO, whereas for higher amount of orthopyroxene difGBS controls the deformation and leads to A-type CPO. Orthopyroxene's CPO strength increases with increasing pressure and temperature and is characterized by the concentration of [100] and [010] axes normal to the foliation and [001] close to the lineation. The seismic properties show that deformation in pyroxene-poor and rich peridotites are consistent with the seismic anisotropy observed in intraplate regions where the mantle flow is horizontal. Conversely, only pyroxene-rich peridotites deformed through difGBS could explain the Vsh/Vsv < 1 observed below mid-oceanic ridges.
Article
Measurements of single-crystal elastic moduli under ambient conditions are reported for nine calcium to calcium-sodium amphiboles that lie in the composition range of common crustal constituents. Velocities of body and surface acoustic waves measured by Impulsive Stimulated Light Scattering (ISLS) were inverted to determine the 13 moduli characterizing these monoclinic samples. Moduli show a consistent pattern: C33>C22>C11 and C23>C12>C13 and C44>C55∼C66 and for the uniquely monoclinic moduli, |C35|>>C46∼|C25|>|C15|∼0. Most of the compositionally-induced variance of moduli is associated with aluminum and iron content. Seven moduli (C11 C12 C13 C22 C44 C55 C66) increase with increasing aluminum while all diagonal moduli decrease with increasing iron. Three moduli (C11, C13 and C44) increase with increasing sodium and potassium occupancy in A-sites. The uniquely monoclinic moduli (C15 C25 and C35) have no significant compositional dependence. Moduli associated with the a∗ direction (C11 C12 C13 C55 and C66) are substantially smaller than values associated with structurally and chemically related clinopyroxenes. Other moduli are more similar for both inosilicates. The isotropically averaged adiabatic bulk modulus does not vary with iron content but increases with aluminum content from 85 GPa for tremolite to 99 GPa for pargasite. Increasing iron reduces while increasing aluminum increases the isotropic shear modulus which ranges from 47 GPa for ferro-actinolite to 64 GPa for pargasite. These results exhibit far greater anisotropy and higher velocities than apparent in earlier work. Quasi-longitudinal velocities are as fast as ∼9 km/s and (intermediate between the a∗- and c-axes) are as slow as ∼6 km/s. Voigt-Reuss-Hill averaging based on prior single crystal moduli resulted in calculated rock velocities lower than laboratory measurements, leading to adoption of the (higher velocity) Voigt bound. Thus, former uses of the upper Voigt bound can be understood as an ad hoc decision that compensated for inaccurate data. Furthermore, because properties of the end-member amphiboles deviate substantially from prior estimates, all predictions of rock velocities as a function of modal mineralogy and elemental partitioning require reconsideration.
Article
The microstructural and petrofabric study of peridotite xenoliths from the El Aprisco (Neogene Calatrava Volcanic Field) has provided new information on deformation mechanisms, ambient conditions and seismic properties of the central Iberian subcontinental mantle. Olivine, orthopyroxene, clinopyroxene, amphibole and spinel constitute the mineral assemblage in equilibrium. Their microstructure indicates that they accommodated crystal-plastic deformation under high water fugacity conditions. Crystallographic preferred orientation patterns of key minerals were determined with the EBSD technique. The xenoliths exhibit B, C and A olivine fabrics. B-type fabrics, involving the (010)[001] slip system, may develop in domains where deformation occurs under comparatively lower temperature, higher water-content and faster strain rates. They are interpreted here as the result of deformation in a suprasubduction mantle setting triggered by changing conditions imposed by a cooler subducting slab that incorporated fluids into the system. Xenoliths with olivine C-type fabrics involve activation of the dominant (100)[001] slip system, denote intracrystalline slip at higher temperatures and water-contents. They are here interpreted to sample lithospheric mantle domains where the impact of those new conditions was not so strong. Finally, the A-type fabrics, characteristic of the (010)[100] slip system, are frequent in the mantle under moderate to high temperature. These fabrics are considered here as characteristic of the mantle prior to subduction. The olivine fabrics constrain heterogeneous seismic properties. Propagation orientation of P waves (8.27–8.51 km/s) coincides with olivine [100] axis concentrations, whereas the fastest S1 waves (5.13–5.22 km/s) propagate parallel to [010] axis minima. The maximum shear wave birefringence (VS1-VS2 = 0.17–0.37 km/s) is close to the direction of the macroscopic lineation. Heterogeneity of calculated seismic properties would concur with preservation of fossil flow perturbations, reworking and metasomatism in the subcontinental lithospheric mantle of central Spain, likely related to changes in the ambient conditions close to paleosubductions.