Figure 11 - uploaded by Katsuyoshi Michibayashi
Content may be subject to copyright.
FTIR spectra of garnet (a) and clinopyroxene (b). (a) The 3570 cm-1 peak is from structural OH in the garnets. (b) The 3460, 3540 and 3640 cm − 1 peaks are from structural OH in the clinopyroxene 

FTIR spectra of garnet (a) and clinopyroxene (b). (a) The 3570 cm-1 peak is from structural OH in the garnets. (b) The 3460, 3540 and 3640 cm − 1 peaks are from structural OH in the clinopyroxene 

Source publication
Article
Full-text available
Garnet clinopyroxenites occur within foliated dunite in the Higashi-akaishi peridotite mass, located within the subduction-type high-pressure/low-temperature Sanbagawa metamorphic belt. The garnet clinopyroxenites contain 3-80% garnet, and garnet and clinopyroxene are homogeneously distributed. Garnet crystals contain extensive, regular dislocation...

Contexts in source publication

Context 1
... point, with the color reflecting the lattice orientation of garnet. In the case that the <100>, <110>, and <111> axes of garnet were oriented parallel to the lineation (X), the grain was colored red, green, or blue, respectively. White areas in the maps are points that were not indexed, or minerals other than garnet. Black lines, representing grain boundaries, are drawn between any adjacent points with a misorientation > 10 . The orientation maps (Fig. 7c, d) show the occurrence of grain boundaries and fractures. The grain size of garnet is similar to that of clinopyroxene (~0.3 mm). Garnet is flattened and elongated in the garnet clinopyroxenites (Fig. 7c, d), and some grain boundaries consist of interfingering sutures (Fig. 7c, d; Stipp et al ., 2002). Some garnet grains contain low-angle boundaries (misorientation angles of 2 – 9°), and aspect ratios are generally ~2, regardless of modal composition. Garnet clinopyroxenites are pervasively cracked by extensional fractures that are generally oriented perpendicular to the stretching lineation (Figs. 6 and 7). Such lineation- normal fractures are common in granulite facies mylonites and UHP metamorphic rocks, and are thought to form during the final stages of exhumation (Ji et al ., 1997, 2003). Single crystals of garnet were hand-picked from crushed samples of garnet clinopyroxenite for TEM observations. TEM samples were thinned using the ion- bombardment technique, and analyzed using a JEOL JEM-2010 (Hiroshima University, Japan) transmission electron microscope at an accelerating voltage of 200 kV. TEM observations of garnet focused on geometrical dislocation microstructures, dislocation densities, and the Burgers vectors of dislocations. The analyzed garnet grains contain extensive, regular dislocation arrays and dislocation networks. The dislocation arrays (Fig. 8a and b) can be interpreted as tilt subgrain boundaries, while the dislocation networks (Fig. 8c and d) indicate the activation of at least two slip systems. These well-organized dislocation microstructures indicate the occurrence of an efficient diffusion-assisted recovery mechanism such as dislocation climb or cross slip. The density of free dislocations within garnet ranges from 3 to 6 10 – 2 cm (Fig. 9). Dislocation junctions, which are the complex intersections (tangles) of several dislocations, were also observed in each sample (Fig. 8f), indicating interaction between dislocations and therefore the operation of multiple slip systems (Voegelé et al ., 1998; Wang and Ji, 1999; Ji et al ., 2003). The above observations are in agreement with previous TEM investigations on naturally deformed garnets (Ando et al ., 1993; Doukhan et al ., 1994; Ji and Martignole, 1994; Voegele ́ et al ., 1998; Prior et al ., 2000; Ji et al ., 2003), all of which reported that garnet can deform plastically given appropriate conditions of temperature, pressure, differential stress, and strain rate. The Burgers vectors of dislocations were identified using the g ⋅ b = 0 and g ⋅ b u = 0 criteria. The majority of dislocations have a Burgers vector b = 1⁄2<111>. We analyzed the chemical compositions of garnet grains within four samples of garnet clinopyroxenites (Fig. 10) using a JEOL electron microprobe (JXA733) housed at the Centre for Instrumental Analysis, Shizuoka University, Japan. Analytical conditions were 15 kV accelerating voltage, 12 nA probe current, and beam diameter of 20 m, using a count time of 20 s and 10 s background. Table 3 lists the results of analyses of 10 representative garnet grains. The grains have similar compositions (Fig. 10a). Microprobe analyses of 75 garnet grains from the four samples yielded the following compositional ranges: 47.3 – 54.8% pyrope, 27.7 – 30.9% almandine, 16.7 – 20.2% grossular, and 0.8 – 1.6% spessartine, with an average composition of Alm 29.9±1.4 Prp 50.5±2.6 Grs 18.6±1.8 Spe 1.0±0.3 . Most of the grains show no significant compositional zoning, although the almandine component shows a slight increase around grain boundaries and cracks, whereas the pyrope component shows a slight decrease. Clinopyroxene grains are near-homogeneous, contain a diopside component (Fig. 10b), and are compositionally similar among samples. We measured the water content in garnet grains using Fourier-transform infrared (FTIR) spectroscopy. For each sample, a polished thin section ( ~ 0.25 mm thick) was prepared and dried in an oven at 120 °C for 4 h. The infrared spectrum was obtained using a Perkin Elmer, Spectrum 2000 (Geochemical Laboratory, University of Tokyo) at room temperature and in the wavenumber range of 740 – 7800 cm – 1 . A series of 100 scans was – 1 averaged for each spectrum with a resolution of 1 cm . For analysis, we carefully selected 2 crack-free and unaltered regions of 60 60 m in size. We used FTIR to analyze 38 grains (29 of garnet, 9 of cpx) from four samples. In the typical OH vibration region (3400 – 3800 cm – 1 ), all grains show several absorption bands. Figure 11 shows typical FTIR spectra of garnet and clinopyroxene from the Higashi-akaishi garnet clinopyroxenite. The infrared (IR) spectra of all garnet grains show – 1 – 1 a sharp peak at ~3570 cm (Fig. 11a); one sample (GM01) shows a peak at ~3430 cm , which is typical of the stretching vibrations (ν3+ν1) of molecular water, which may occur in submicroscopic fluid inclusions within garnet. Following previous studies of natural garnets (Rossman and Smyth, 1990; Bell and Rossman, 1992), we ascribed this group of – 1 bands to submicroscopic fluid inclusions. The 3570 cm peak is produced by structural OH in the garnets. The IR spectra of clinopyroxene show sharp peaks at 3460, 3540, and – 1 3640 cm (Fig. 11b), produced by structural OH (Bell et al ., 1995). We calculated the water content (H 2 O ppm wt.) of both minerals using the Beer – Lambert law (absorbance = absorbance coefficient thickness water concentration). Absorbance is expressed as the integrated absorbance areas of OH. We used the following 2 integrated molar absorbance coefficients from Bell et al . (1995): 1.39 ppm H 2 O/cm for 2 garnet, 7.09 ppm H 2 O/cm for clinopyroxene. Table 4 provides detailed information on peak positions, absorbance, and the calculated water content. H 2 O contents in garnet range from 17 to 1000 ppm (mainly ~60 ppm) (Fig. 12a). The H 2 O content of Higashi-akaishi garnet varies both among and within samples. In contrast, H 2 O contents in clinopyroxene are relatively homogeneous within and among samples (~70 ppm)(Fig. 12b). We measured the crystal-preferred orientations (CPOs) of olivine, garnet, and clinopyroxene grains from highly polished thin sections, using a scanning electron microscope equipped with an electron-backscatter diffraction system (JEOL JSM6300 with HKL Channel5), housed at the Centre for Instrumental Analysis, Shizuoka University, Japan. We measured about 200 crystal orientations per sample, visually confirming the computerized indexation of the diffraction pattern for each orientation. The measured CPOs are presented on equal-area, lower-hemisphere projections in the structural (XZ) reference frame (Figs. 13 – 17). To characterize the CPOs, we determined the fabric strength and distribution density of the principal crystallographic axes (e.g., Michibayashi and Mainprice, 2004). The rotation matrix between crystal and sample co- ordinates is used to describe the orientation g of a grain or crystal in sample co-ordinates. In practice, it is convenient to describe the rotation by a triplet of Euler angles; e.g., g = ( φ 1 , , φ 2 ), as used by Bunge (1982). The orientation distribution function (ODF), f ( g ), is defined as the volume fraction of orientations in the interval between g and g + d g in a space containing all possible orientations, as given ...
Context 2
... Kleinschrodt and McGrew, 2000; Ji et al ., 2003; Michibayashi et al ., 2004; Terry and Heidelbach, 2004; Okamoto and Michibayashi, 2005; Storey and Prior, 2005; Bestmann et al ., 2008). Some of these studies have suggested that dislocation creep is the dominant deformation mechanism for garnet, based on TEM observations of dislocation networks and subgrain walls (e.g., Ando et al ., 1993; Ji and Martignole, 1994; Ji et al ., 2003); however, the problem exists that although numerical simulations (performed using the VPSC model) of CPO development in garnet produce characteristic CPOs for both axial compression and simple shear deformation (Mainprice et al ., 2004), very weak or random CPOs patterns are obtained for naturally occurring elongate garnet (Ji et al ., 2003; Storey and Prior, 2005). Consequently, alternative deformation mechanisms have been proposed. Storey and Prior (2005) suggested that plastic deformation of garnet is dominated by grain- boundary sliding accompanied by subgrain formation and rotation, rather than dislocation creep. Similarly, in an analysis of experimentally deformed eclogites, Zhang and Green (2007) reported the repeated ‘sliding off’ into the foliation of superficial layers of recrystallized garnet. However, we consider it unlikely that garnet within the Higashiakaishi mass was deformed in this way, as it shows contrasting textures to those reported in the above studies. For example, Storey and Prior (2005) reported fine recrystallized garnet (~50 m) around coarse garnet (~1 mm), whereas garnet clinopyroxenites of the Higashi-akaishi mass contain elongate garnet crystals of homogeneous grain size (0.3 mm) (Fig. 7). In addition, some of the grain boundaries observed in the present study consist of interfingering sutures, and the grain size is much coarser than that generally associated with grain-boundary sliding. Therefore, it is difficult to explain the observed garnet deformation in terms of grain-boundary sliding. Some of the analyzed garnet grains contain low-angle internal boundaries. TEM observations indicate that the garnet grains contain extensive, regular dislocation arrays and dislocation networks, suggesting that the low-angle boundaries are sub-grain boundaries. The density of free dislocations within the analysed garnet grains ranges from 7 – 2 3 to 6 10 cm (Fig. 9), which is relatively high (Ando et al ., 1993); however, garnet has low strength and shows complex CPO patterns. Garnet also has a high degree of symmetry and 12 potential slip systems (Ji et al ., 2003); consequently, any given slip plane needs only undergo a small amount of rotation before the resolved shear stress reaches high levels on a different slip system, such as 1/2 <111> {110}. Although slip occurs predominantly on {110} planes, it is important to realize that three {110}-type planes intersect in a [111] direction and that screw dislocations with a 1/2 <111> Burgers vector may migrate randomly on {111} planes with high resolved shear stress. Thus, the weakness of garnet CPOs does not provide unequivocal evidence for diffusion creep or against dislocation creep as a deformation mechanism within garnet (Ji et al ., 2003). Therefore, we propose that dislocation creep was the dominant deformation mechanism in garnet within the Higashi-akaishi garnet clinopyroxenites. The studied garnet clinopyroxenites are typically bimineralic, being composed of garnet and diopside. Figure 18b shows grain size and aspect ratio data for garnet and diopside within these rocks, with respect to modal composition. Grain sizes and aspect ratios in garnet are comparable with those in clinopyroxene, regardless of the modal composition. Since the garnet clinopyroxenites were plastically sheared along with the foliated dunites, these observations reveal that the two minerals deformed under similar degree of plasticity. The fabric strength ( M -index and J -index) of a deformed rock is related to finite strain, as the M -index ( J -index) increases with finite plastic strain (Tommasi et al ., 2000; Skemer et al ., 2005). Figure 18a shows that the M -index of both garnet and clinopyroxene increases with increasing modal composition of garnet. Because the distribution of garnet and clinopyroxene is homogeneous in the analyzed samples (Fig. 6) , the dominant phase controls the deformation of the rock. If all of the garnet clinopyroxenites had deformed under the same stress conditions, the relationship shown in Fig. 18a suggests that garnet- dominated part has been more strained than clinopyroxene-diminated part. However, the results of an experimental study performed at high temperature and pressure (1500 K, 3 GPa) revealed that garnet is three to four times as strong as omphacite (Jin et al ., 2001). The olivine fabrics measured in the present study suggest that the garnet clinopyroxenites were plastically sheared under wet conditions, and FTIR analyses reveal H 2 O contents in garnet of 17 – 1000 ppm (mostly ~60 ppm; Fig. 11; Table 4). It is possible that the presence of water influences the deformation of garnet. Indeed, Katayama and Karato (2008) reported that the creep rate of Mg-rich garnet (Alm 19 Prp 68 Grs 12 ) is sensitive to water. Figure 22 shows the relation between stress and strain rate for garnet and diopside under dry and wet conditions. Here, the water contents of wet and dry garnet are 80 – 200 ppm and < 30 ppm, respectively (Katayama and Karato, 2008). The results reveal contrasting influences of water on the deformation of garnet and diopside: under wet conditions compared with dry, the strain rate increases by two orders of magnitude for garnet but by an order of magnitude for diopside. Given the influence of water on the creep strength of garnet, garnet within the Higashi-akaishi mass may have become significantly as weak as clinopyroxene during deformation. It should be, however, noted that the creep strength of minerals may be also influenced by some other effects such as stress or pressure as well as the effect of water (e.g., Chen et al., 2006; Li et al., 2006; Katayama and Karato, 2008; Amiguet et al., 2009). The rheological behavior of garnet is an important factor in the deformation of subducted oceanic crust, as oceanic crust that has been deeply subducted (> 400 km depth) is composed largely of garnet (Ringwood, 1982). The nominally anhydrous mineral phases (NAMs; olivine, pyroxene, and garnet) in both subducting oceanic crust and the overlying mantle wedge can carry a significant amount of H 2 O to the deep mantle (Forneris and Holloway, 2003; Iwamori, 2007). If water has a stronger influence on the creep strength of garnet than on that of clinopyroxene, as shown above, oceanic crust would be expected to weaken with ongoing subduction. This hypothesis is the opposite to that suggested in previous studies (e.g., Karato et al ., 1995; Jin et al ., 2001) and may have implications for our understanding of mantle convection. Dunites in the Higashi-akaishi peridotite mass, located in the subduction-type Sanbagawa metamorphic belt, record two stages of deformation (D1 and D2; Mizukami and Wallis, 2005) and contain various microstructures, ranging from coarse-grained to porphyroclastic. At Gongen Pass, dunites contain fine-grained textures with weak fabric strength, suggesting that the outcrop represents the center of a D2 shear zone. CPO patterns for olivine are B-type regardless of texture. Accordingly, dunite within the Higashi-akaishi mass is interpreted to have been deformed under high-stress and wet conditions. Garnet clinopyroxenites that occur within foliated dunite at Gongen Pass contain 3 – 80% garnet, and garnet and clinopyroxene within these rocks have a homogeneous distribution. Garnet contains extensive, regular dislocation arrays and dislocation networks, suggesting that dislocation creep was the dominant deformation mechanism. Analyses of orientation maps reveal that garnet and clinopyroxene have similar grain sizes and aspect ratios, regardless of modal composition. These findings indicate that the two minerals were deformed under similar conditions of plasticity. In addition, M -index values for both garnet and clinopyroxene increase with increasing modal composition of garnet. During deformation, garnet was possibly weaker than clinopyroxene. The obtained olivine CPOs indicate deformation under wet conditions, and the water content of garnet is ~60 ppm. It is possible that the presence of water helped to induce garnet deformation, and flow laws indicate that under water-rich conditions (relative to dry conditions), the strain rate for garnet increases by two orders of magnitude, whereas for diopside it increases by an order of magnitude. This finding may have significant implications for our understanding of mantle convection, as garnet may in fact be weaker than clinopyroxene. The authors thank Hiroyuki Kagi of University of Tokyo for assistance with FTIR analyses and M. Satish-Kumar of Shizuoka University for assistance with EPMA analyses. We would like to thank Toshiaki Masuda of Shizuoka University for help in using the Ion thinning technique, and Yasuro Kugimiya for providing samples for analysis. MM wishes to thank Takumi Nanbu, Masahiro Hara, Kyoko Kanayama, and members of the Michibayashi Laboratory (Yumiko Harigane, Takako Satsukawa, Hisashi Imoto, Ayano Fujii, Shigeki Uehara, Tatsuya Ohara, Makoto Suzuki, Naohiko Ueta, Naoaki Komori, and Yuri Shinkai) for their helpful advice and discussions. Amiguet, E., Raterron, P., Cordier, P., Couvy, H. and Chen, J., 2009. Deformation of diopside single crystal at mantle pressure, 1, Mechanical data. Physics of the Earth and Planetary Interiors , 177 , 122-129. Ando, J., Fujino, K. and Takeshita, T., 1993. Dislocation microstructures in naturally deformed silicate garnets. Physics of the Earth and Planetary Interiors, 80 , 105-116. Aoya, M., 2001. P-T-D path of eclogite from the Sambagawa belt deduced from combination of ...

Similar publications

Article
Full-text available
The deformation behaviour of minerals in upper mantle rocks plays an important role in controlling the dynamics of the Earth. The garnet peridotite bodies on the island of Otrøy in the Western Gneiss Region of Norway originate from extreme depth (up to 390 km). The deformation structures in those garnet peridotite bodies will therefore provide a un...
Article
Full-text available
Ultramylonitic peridotites collected from a high-pressure metamorphic terrain in the Yushugou Massif at the northern margin of the south Tianshan orogenic belt (Xinjiang, China), in which strongly elongate and folded orthopyroxene (En90) porphyroclasts embedded in a fine-grained (d=0.01 mm ), recrystallized olivine (Fo90) matrix, were optically inv...
Chapter
Full-text available
Numerical models of subduction are presented to investigate the relative role of a Newtonian versus composite viscosity, maximum yield stress, and the initial slab dip on the upper‐mantle viscosity structure and velocity. The results show that using the experimentally derived flow law to define the Newtonian viscosity (diffusion creep deformation m...
Preprint
Full-text available
Geodynamic numerical models often employ solely grain-size-independent dislocation creep to describe upper mantle dynamics. However, observations from nature and rock deformation experiments suggest that shear zones can transition to a grain-size-dependent creep mechanism due to dynamic grain size evolution, with important implications for the over...

Citations

... Olivine and clinopyroxene are two of the main minerals in ultramafic rocks and can form dunite, wehrlite, olivine clinopyroxenite, and clinopyroxenite in the order of increasing clinopyroxene percentage. Although not as common as orthopyroxene, clinopyroxene is an important mineral in mantle rocks [1][2][3]. Wehrlite is the most common peridotite with clinopyroxene, which is generated by either a melt-rock reaction or interaction with SiO 2 -undersaturated magmas [4][5][6]. Although the solid second phase may inhibit the development of the olivine crystallographic preferred orientation (CPO) [7,8], wehrlites in different areas and geological environments can generate various olivine CPO types (B, E, and AG-type; [9][10][11]). ...
... (Olivine) clinopyroxenites are usually banded or veined in natural outcrops [18][19][20]. The main generation mechanisms of clinopyroxenites include (1) oceanic basaltic crust recycling by subduction [21][22][23], (2) basaltic melt crystallization [24,25], and (3) melting and the melt-rock reaction [26,27]. The (010) [001] CPO is common in clinopyroxenites of different generations [14,19,20,24]. ...
Article
Full-text available
The microstructural relationship between olivine and clinopyroxene is significant in recovering the mantle evolution under clinopyroxene-saturated melting conditions. This study focuses on olivine/clinopyroxene-related ultramafic rocks (dunite, wehrlite, olivine clinopyroxenite, and clinopyroxenite) in the Ells Stream Complex of the Red Hills Massif. (Olivine) clinopyroxenites have an A/D-type olivine crystallographic preferred orientation (CPO) whereas peridotites have various olivine CPO types. B-type olivine CPO was newly discovered, which may have been generated under hydrous conditions. The discovery of B-type CPO means that all six olivine CPO types could exist in a single research area. Clinopyroxene CPOs also vary and have weaker deformation characteristics (e.g., lower M index and weaker intracrystalline deformation) than olivine; thus, they probably melted and the clinopyroxene-rich ultramafic bands existed as melt veins. Irregular clinopyroxene shapes in the peridotites and incoherent olivine and clinopyroxene CPOs ([100] OL and [001] CPX are not parallel) also indicate a melted state. The dominant orthorhombic and LS-type CPOs in olivine and clinopyroxene imply that simple shear was the main deformation mechanism. Such complicated microstructural characteristics result from the overprinted simple shear under high temperatures (>1000 • C) and hydrous melting environments until the melt-frozen period. This case study is helpful to better understand the olivine and clinopyroxene relationship.
... These macroscopic indicators are found only in clinopyroxene and are not common in the other minerals (garnet, amphibole) that comprise the arclogites. In eclogite pyroxenites, dry diopside is rheologically the weakest mineral unless garnet is particularly hydrous (Muramoto et al., 2011). We find that in many xenoliths without well-defined fabrics, mineral modes are dominated by garnet (SB3ME-02, CCME-16), suggesting that garnet-rich samples may not have recorded as much deformation. ...
Article
Full-text available
The microstructural properties of deep arc cumulates (arclogites) are poorly understood, but are essential in gaining a comprehensive picture of the rheology of continental lithosphere. Here, we analyze 16 arclogite xenoliths, comprising a low MgO and a high MgO suite, from Arizona, USA using electron backscatter diffraction to map microstructures, clinopyroxene shape preferred orientations (SPO), and clinopyroxene crystallographic preferred orientations (CPO). The lower pressure (∼1 GPa) low MgO arclogites show a variety of different clinopyroxene fabrics (S, L, and LS‐type), whereas the high pressure (>2 GPa) high MgO arclogites show predominantly LS‐type fabrics. Furthermore, clinopyroxenes in low MgO arclogites all show a pronounced correspondence between the long axis of their grain shape ellipsoids with the [001] crystal direction, indicating an SPO control on the CPO. In contrast, high MgO arclogite clinopyroxenes lack such a correspondence. We propose that both arclogite types originated as igneous cumulates, consistent with previous studies, but that the high MgO suite experienced substantial recrystallization which diminished the original igneous SPO‐induced CPO. Using strain rates appropriate for arc settings, we calculate a strength profile for the lithosphere and argue that the deepest arclogite textures are consistent with lithospheric foundering through ductile deformation under high shear strain (10⁻¹⁴–10⁻¹² s⁻¹). Our study shows that there is a high degree of shear strain localization in deep arc roots while shallower portions are relatively undeformed.
... The AG-type olivine fabric is different from the other olivine fabrics, and forms via least two slip systems: (010) [100] and (010)[001] (Figure 1). AG-type olivine fabric has been reported from many localities in various tectonic settings (Bascou et al., 2008;Ben-Ismail et al., 2001;Michibayashi & Mainprice, 2004;Muramoto et al., 2011;Figure 12. Lower-hemisphere equal-area projections of the olivine fabrics in the garnet lherzolites. XZ is the sample section, X is the lineation, and Z is normal to the foliation. ...
Article
Full-text available
The different olivine fabrics in ultramafic rocks have been widely used to discuss past tectonic settings, given that the olivine fabrics vary with pressure, temperature and water content. However, there are no research related to whether and how the olivine fabrics transform at different metamorphic stages in a natural rock during the process of deep subduction and exhumation. Yinggelisayi garnet lherzolites from South Altyn have experienced deep continental subduction and exhumation. The garnet lherzolites contain well-preserved residual protolith minerals, and near-peak (M1), granulite-facies retrograde (M2), and amphibolite-facies retrograde (M3) metamorphic mineral assemblages. Olivine grains in M1 formed at P-T conditions of 2.52–3.08 GPa, 1,095–1,136°C and low water contents (183–213 ppm H/Si), and showed [010] axes sub-normal to the foliation and [001] axes subparallel to the lineation, which is characteristic of B-type fabric ((010)[001]). Olivine grains in M2 formed at P-T conditions of 1.31–1.80 GPa, 851–893°C and also low water contents (93–139 ppm H/Si), and exhibited [010] axes sub-normal to the foliation and [100] axes subparallel to the lineation, which is characteristic of A-type fabric ((010)[100]). These observations suggest that olivine fabrics in high pressure-ultrahigh pressure metamorphosed ultramafic rocks are different in the near-peak and retrograde metamorphic stages, and also that the olivine fabrics can be transformed during deep continental subduction and exhumation. Therefore, the dispersed or no clear olivine fabric is probably caused by multi-stage deformation and metamorphism, and the distinct olivine fabrics can also be used as a clue to identify geological processes and better understand metamorphism and deformation during subduction and exhumation.
... The AG-type olivine fabric is different from the other olivine fabrics, and forms via least two slip systems: (010) [100] and (010)[001] (Figure 1). AG-type olivine fabric has been reported from many localities in various tectonic settings (Bascou et al., 2008;Ben-Ismail et al., 2001;Michibayashi & Mainprice, 2004;Muramoto et al., 2011;Figure 12. Lower-hemisphere equal-area projections of the olivine fabrics in the garnet lherzolites. XZ is the sample section, X is the lineation, and Z is normal to the foliation. ...
... The AG-type olivine fabric is different from the other olivine fabrics, and forms via least two slip systems: (010) [100] and (010)[001] (Figure 1). AG-type olivine fabric has been reported from many localities in various tectonic settings (Bascou et al., 2008;Ben-Ismail et al., 2001;Michibayashi & Mainprice, 2004;Muramoto et al., 2011;Figure 12. Lower-hemisphere equal-area projections of the olivine fabrics in the garnet lherzolites. XZ is the sample section, X is the lineation, and Z is normal to the foliation. ...
... In the case of garnet, the <100>(010) and ½<111>(110) easy slip systems reported for naturally deformed garnet (cf. Karato, Zhang, & Wenk, 1995;Voegel e, Ando, Cordier, & Liebermann, 1998) likely operated combined with grain boundary sliding and diffusion-assisted recovery (Ando, Fujino, & Takeshita, 1993;Ji & Martignole, 1994;Muramoto, Michibayashi, Ando, & Kagi, 2011;Prior, Wheeler, Brenker, Harte, & Matthews, 2000;Storey & Prior, 2005;Voegel e et al., 1998). ...
Article
Deformation partitioning is identified as the fingerprint of late Palaeozoic continental subduction that affected various lithologies whose field relationship, thermobarometric and petrofabric features are closely related. Different modes of deformation partitioning can be identified within medium temperature, high-pressure eclogite lenses, between them and the host gneisses, and within the latter. Development of foliations and lineations with a coherent attitude in all these rocks and their related structural petrology demonstrate that eclogite enclosures and their country rocks underwent a common, pervasive deformational event. The published P-T stability fields of the eclogite phases that define the microscopic fabric are used to define the metamorphic conditions prevailing during the deformation event and relate it to the subduction process. The mineral equilibria of the gneisses (ortho- and paragneisses) fail to record the full range of those P-T conditions, but the field relationships show that eclogites were originally basic dykes emplaced in acid igneous rocks and demonstrate that the eclogites and gneisses shared a common tectonometamorphic evolution. Deformation partitioning within the latter occurred at variable scales and involved (1) meso macroscale preservation of virtually undeformed metagranite bodies, surrounded by (2) pervasively foliated and lineated gneisses, and (3) the simultaneous microscale operation in the latter of ductile and brittle-ductile mechanisms at conditions above 500°C and below 1.5 GPa. A subduction channel tectonic setting is proposed to explain the subduction of upper to mid-crustal igneous rocks and exhumation subsequent to high-pressure metamorphism. Its currently accessible dimensions, and its organization into several lithotectonic units mapped as nappes support tectonic amalgamation of units several km³ in volume. Maximum burial in the subduction channel likely reached depths shallower than the lithostatic pressure implied by geobarometric calculations, possibly conditioned by a sudden pressure drop during the initial retrogression stages accompanying exhumation.
... Amiguet et al., 2010;Bystricky and Mackwell, 2001;Mauler et al., 2000) and studies of natural pyroxene-rich samples (e.g. AvéLallemant, 1978;Frets et al., 2012;Helmstaedt et al., 1972;Muramoto et al., 2011) have addressed the deformation of pyroxenes but more are required to fully understand the processes at work. ...
Article
In the Herbeira massif, Cabo Ortegal Complex, Spain, a well exposed assemblage of deformed dunites and pyroxenites offers a unique opportunity to investigate key upper mantle tectonic processes. Four types of pyroxenites are recognized: clinopyroxenites with enclosed dunitic lenses (type-1), massive websterites (type-2), foliated and commonly highly amphibolitized clinopyroxenites (type-3) and orthopyroxenites (type-4). Field and petrological observations together with EBSD analysis provide new insights on the physical behavior of the pyroxenes and their conditions of deformation and reveal the unexpected journey of the Cabo Ortegal pyroxenites.
... In the Sanbagawa metamorphic belt several studies provide fabric data, e.g., garnet and clinopyroxene CPO from the Higashi-Akaishi peridotite body (Muramoto et al., 2011), inclusion-trail geometry of albite in basic schists (Okamoto, 1998), and misorientation of garnet aggregates in vein in the mafic schists (Okamoto & Michibayashi, 2006 research provides the first report on the fabric of eclogitefacies and retrograde-stage minerals as well as useful insights into the rheological mineral behaviour during eclogite formation and exhumation. ...
... The fabric indicating a dominant slip direction along [001] was probably activated within the subduction-related high-stress regime. These results are consistent with the fabric developed in the Higashi-Akaishi peridotite massif (Muramoto et al., 2011). This evidence clearly indicates that the Sanbagawa eclogites experienced the same deformational event as that observed in the Higashi-Akaishi peridotite massif when these lithologies were in the mantle-wedge. ...
Article
Electron back-scattered diffraction (EBSD) maps and crystal-preferred orientation (CPO) of eclogite-facies (omphacite and garnet) and amphibolite-facies (hornblende and actinolite) phases are reported for understanding the rheological behaviour of crust during subduction. Two types of eclogites from the subduction-related high-pressure/low-temperature type Sanbagawa metamorphic belt, Japan, have been investigated. Type-I eclogite (sample Sb-1) is composed of garnet, omphacite, secondary actinolite and hornblende. Type-II eclogite (samples Sb-2 and Sb-2a) are mainly composed of omphacite, garnet, and retrograde hornblende with no actinolite. Omphacite, the peak eclogite-facies phase, exhibits L-type CPO (maximum density of [001] axes parallel to and high density of {110} poles normal to the lineation) in Type-I eclogite, suggesting intra-crystalline plasticity with [001] {110} and 〈110i{110} active slip systems, indicating a constrictive strain regime at mantle depths. Omphacite in Type-II eclogite exhibits a similar fabric but with much weaker CPO. Using the LS-index symmetry analysis (one for the end-member L-type, zero for the end-member S-type, and intermediate values for LS-types), a progressive change in LS index of 0.80 for Type-I and 0.61 to 0.44 for Type-II eclogites is observed. These values suggest a transition from axial extension parallel to the lineation for Sb-1 and weaker CPO associated with pure or simple shear for Sb-2 and Sb-2a. Garnet, the second dominant phase in the eclogite-facies stage, exhibits weak and complex fabric patterns in all eclogite types, behaved like rigid bodies and does not show plastic deformation. Amphibolite-facies phases (e.g., hornblende and actinolite) exhibit more than two types of CPO. Hornblende and actinolite in Type-I eclogite have a strong CPO along [001] axes aligned parallel to the lineation, indicating homotactic crystal growth probably by the replacement of omphacite during the early stages of retrogression. Type-II eclogites have weak CPO in hornblende but with characteristic alignment of [001] parallel to the lineation and other poles to planes (100), (010), and {110} normal to the lineation. This fabric might have resulted from a cataclastic deformation and could be related to the late-“D1” deformation stage in the Sanbagawa metamorphic belt.
... 7-11) are presented on equal-area, lower hemisphere projections in which the plane of projection contains the stretching lineation (X) and the pole to the foliation (Z). The CPO strength for each pole figure is described by the pfJ-index, which is a volume-averaged integral of the squared orientation densities (Bunge, 1982) and has a value of unity for a random distribution and a value of infinity for a single crystal (Michibayashi and Mainprice, 2004) although the maximum value is usually much less due to truncation of the spherical harmonic series at an expansion of 22 (Muramoto et al., 2011). ...
... The peridotite xenolith data are from Ichinomegata (Michibayashi et al., 2006a(Michibayashi et al., , 2006b, Avacha (Michibayashi et al., 2009a), Knippa (Satsukawa et al., 2010), Kilbourne Hole (Satsukawa et al., 2011), a petitspot (Harigane et al., 2011b), and Tasmania (Michibayashi et al., 2012) (green points in Fig. 3). The data for the peridotites in HP metamorphic rocks are from the Imono (Tasaka et al., 2008) and Higashi-Akaishi (Muramoto et al., 2011) massifs (cyan points in Fig. 3). Fig. 4 shows the olivine fabrics from our database plotted on a Vp-Flinn diagram. ...
... One peridotite xenolith (H11) had been entrained in magma erupted from petit-spot volcano, so it could have been derived from the lithosphere-asthenosphere boundary (Hirano et al., 2001(Hirano et al., , 2004(Hirano et al., , 2006, representing AG type (Harigane et al., 2011b). The peridotites entrained in the HP metamorphic rocks (T08 and M11) were possibly derived from the deep mantle wedge above the subducting slab, where B type may be dominant (Mizukami et al., 2004;Tasaka et al., 2008;Muramoto et al., 2011). ...
Article
Full-text available
Crystallographic preferred orientations (CPOs) of olivine within natural peridotites are commonly depicted by pole figures for the [100], [010], and [001] axes, and they can be categorized into five well-known fabric types: A, B, C, D, and E. These fabric types can be related to olivine slip systems: A with (010)[100], B with (010)[001], C with (001)[001], D with {0kl}[100], and E with (001)[100]. In addition, an AG type is commonly found in nature, but its origin is controversial, and could involve several contributing factors such as complex slip systems, non-coaxial strain types, or the effects of melt during plastic flow. In this paper we present all of our olivine fabric database published previously as well as new data mostly from ocean floor, mainly for the convergent margin of the western Pacific region, and we introduce a new index named Fabric-Index Angle (FIA), which is related to the P-wave property of a single olivine crystal. The FIA can be used as an alternative to classifying the CPOs into the six fabric types, and it allows a set of CPOs to be expressed as a single angle in a range between −90° and 180°. The six olivine fabric types have unique values of FIA: 63° for A type, −28° for B type, 158° for C type, 90° for D type, 106° for E type, and 0° for AG type. We divided our olivine database into five tectonic groups: ophiolites, ridge peridotites, trench peridotites, peridotite xenoliths, and peridotites enclosed in high-pressure metamorphic rocks. Our results show that although our database is not yet large enough (except for trench peridotites) to define the characteristics of the five tectonic groups, the natural olivine fabrics vary in their range of FIA: 0° to 150° for the ophiolites, 40° to 80° for the ridge peridotites, −40° to 100° for the trench peridotites, 0° to 100° for the peridotite xenoliths, and −40° to 10° for the peridotites enclosed in high-pressure metamorphic rocks. The trench peridotites show a statistically unimodal distribution of FIA consisting of the high peak equivalent of the A type, but with some FIAs close to the AG and D types. The variations in the olivine fabrics in the trench peridotites could result from variations in deformation within the supra-subduction uppermost mantle, possibly related to evolution of the mantle since the subduction initiation of the Pacific plate.