Tectonophysics 429 (2007) 1 – 20 www.elsevier.com/locate/tecto Continuation of the San Andreas fault system into the upper mantle: Evidence from spinel peridotite xenoliths in the Coyote Lake basalt, central California Sarah J. Titus ⁎, L. Gordon Medaris Jr., Herbert F. Wang, Basil Tikoff Department of Geology and Geophysics, University of Wisconsin-Madison, 1215 W. Dayton, Madison, WI 53706, USA Received 28 July 2005; received in revised form 12 June 2006; accepted 12 July 2006 Available online 16 November 2006 Abstract The Coyote Lake basalt, located near the intersection of the Hayward and Calaveras faults in central California, contains spinel peridotite xenoliths from the mantle beneath the San Andreas fault system. Six upper mantle xenoliths were studied in detail by a combination of petrologic techniques. Temperature estimates, obtained from three two-pyroxene geothermometers and the Al-inorthopyroxene geothermometer, indicate that the xenoliths equilibrated at 970–1100 °C. A thermal model was used to estimate the corresponding depth of equilibration for these xenoliths, resulting in depths between 38 and 43 km. The lattice preferred orientation of olivine measured in five of the xenolith samples show strong point distributions of olivine crystallographic axes suggesting that fabrics formed under high-temperature conditions. Calculated seismic anisotropy values indicate an average shear wave anisotropy of 6%, higher than the anisotropy calculated from xenoliths from other tectonic environments. Using this value, the anisotropic layer responsible for fault-parallel shear wave splitting in central California is less than 100 km thick. The strong fabric preserved in the xenoliths suggests that a mantle shear zone exists below the Calaveras fault to a depth of at least 40 km, and combining xenolith petrofabrics with shear wave splitting studies helps distinguish between different models for deformation at depth beneath the San Andrea fault system. © 2006 Elsevier B.V. All rights reserved. Keywords: Peridotite xenoliths; San Andreas fault; Calaveras fault; Geothermometry; Lattice preferred orientation; Seismic anisotropy 1. Introduction Collision of the East Pacific Rise with the western margin of North America at ∼ 29 Ma generated the Mendocino triple junction, a transform–transform–trench intersection, whose subsequent northwesterly migration along coastal California produced the San Andreas ⁎ Corresponding author. Department of Geology, Carleton College, Northfield, MN 55057, USA. E-mail address: [email protected] (S.J. Titus). 0040-1951/$ - see front matter © 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.tecto.2006.07.004 transform system (Atwater, 1970; Dickinson and Snyder, 1979). Migration of the Mendocino triple junction generated an expanding slab window (Thorkelson and Taylor, 1989; Severinghaus and Atwater, 1990), which was accompanied by several pulses of volcanism, thought to have been triggered by decompression melting as upwelling mantle flowed into the slab window (Johnson and O'Neil, 1984; Dickinson, 1997). These volcanic rocks are exposed along a 280-km long belt in the eastern Coast Ranges and are generally older in the south and younger in the north (e.g. Fox et al., 1985; Dickinson, 2 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 1997). The migratory volcanic centers are predominantly intermediate to silicic in composition and have relatively high δ18O values and 87Sr/86Sr ratios, indicating a significant degree of crustal anatexis in their petrogenesis (Johnson and O'Neil, 1984). Among the eastern Coast Range volcanic fields, the Coyote Lake basalt, which is located near the intersection of the Hayward and Calaveras faults about 100 km southeast of San Francisco (Fig. 1), is noteworthy for several reasons. First, the Coyote Lake basalt represents a clear exception to the northward younging trend of the volcanic fields. In contrast to other eastern Coast Range volcanic fields whose eruptions followed closely after passage of the Mendocino triple junction, the Coyote Lake basalt was erupted at 2.5–3.6 Ma (Nakata et al., 1993), well after passage of the triple junction at this latitude (Dickinson, 1997). Instead, eruption of the Coyote Lake basalt volcanic rocks has been ascribed to near-fault normal extension along the Calaveras fault in Pliocene time (Jové and Coleman, 1998). Second, in contrast to other siliceous volcanic centers of the eastern Coast Ranges (Johnson and O'Neil, 1984), this small volcanic field covering an area of ∼ 11 km2 consists of a sequence of basaltic tuffs, breccias and flows of alkaline affinity (Nakata, 1977). Lastly, the Coyote Lake basalt is unique among Cenozoic volcanic rocks in the Coast Ranges because of the inclusion of mantle xenoliths, providing the only known samples from the mantle beneath the San Andreas fault system. This study examines six mantle xenoliths in detail from the Coyote Lake basalt to document the mineralogy, petrology and microstructures preserved in these samples. Geothermometry coupled with thermal modeling allows us to estimate the temperatures, pressures and depths at which these xenoliths equilibrated in the mantle. Petrofabric analysis, using the lattice preferred orientation of olivine and the modal composition of each xenolith, allows calculation of the seismic properties. The xenolith petrofabrics and the associated seismic anisotropy calculations place constraints on the thickness of the anisotropic layer observed in modern shear wave splitting studies in central California. 2. Description of mantle xenoliths 2.1. Occurrence and lithology Small (1–3 cm diameter), subangular to subrounded xenoliths of lower crustal and upper mantle material are abundant in massive basalt flows and feeders of the Fig. 1. Map of central California showing the locations of Tertiary volcanic fields. Dark arrows indicate the position of the Mendocino triple junction through time in Ma. Upper inset shows the location within the state of California. Lower inset shows a more detailed map of the Coyote Lake basalt with a star denoting the xenolith sample site. Modified from McLaughlin et al. (1996) and Jové and Coleman (1998). S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Coyote Lake basalt southwest of Coyote Lake. Lower crustal xenoliths include metagabbronorite, which consists of plagioclase, olivine, orthopyroxene, clinopyroxene and Fe–Ti oxides, and andesine megacrysts. Upper mantle xenoliths include Cr-diopside lherzolite, harzburgite, wehrlite and dunite, and Al-augite clinopyroxenite and clinopyroxene megacrysts (following the classification for mantle xenoliths proposed by Wilshire and Shervais, 1975). Spinel is the characteristic aluminous phase in the ultramafic rocks, although plagioclase is present in some clinopyroxenite and has been reported in some lherzolite (Nakata, 1977). Garnet is absent in these xenoliths. Six spinel-bearing ultramafic xenoliths were selected for detailed investigation, including three samples of Crdiopside lherzolite (3a, 7a, 12a), two of harzburgite (2a, 7c) and one Al-augite clinopyroxenite-wehrlite composite (6a); mineral modes for the analyzed samples are summarized in Table 1. Because of the uncertainty in determining the identity of each silicate mineral grain in these ultramafic rocks during conventional point-counting for modal analysis, modes were obtained from false-color images prepared from Al, Ca and Fe X-ray maps, which provide unequivocal discrimination among olivine, orthopyroxene and clinopyroxene, and yield a relative precision from b 1% for major phases to 4% for minor phases. The five Cr-diopside peridotite samples contain a four-phase assemblage of olivine (64–83%), orthopyroxene (14– 23%), clinopyroxene (3–15%) and spinel (0.2–1.9%). The variation in modes (and mineral compositions described below) among the five samples most likely reflects different degrees of partial fusion ranging from 3.0% to 8.5%, as estimated from the compositions of constituent spinel (Table 1; Hellebrand et al., 2001). Composite sample 6a consists of Al-augite olivine-spinel clinopyrox- 3 enite, which includes feldspar (plagioclase, 8%) and apatite (2%), and adjoining spinel wehrlite, which also includes feldspar (alkali feldspar, 4%) and apatite (1%). 2.2. Microstructures The grain shapes and grain boundary textures observed in all xenoliths are quite similar and are described using the terminology of Passchier and Trouw (1998; Fig. 2). Grain shapes vary from polygonal to interlobate, although polygonal shapes with relatively straight grain boundaries meeting in 120° triple junctions are more common. All xenoliths, except 6a, are inequigranular, where the matrix (primarily olivine) has an average grain size of ∼ 1 mm and larger grains (both olivine and pyroxene) are ∼ 4 mm in diameter. Sample 6a, in contrast, has a bimodal distribution of grain sizes (0.5 mm and 1 mm) and no larger porphyroclasts. Sample 6a is also noteworthy because of the abundance of fine-grained, subhedral spinel grains. Undulose extinction is common in many olivine grains and, more rarely, deformation bands and deformation lamellae are present. Pyroxenes vary in size; generally, they are the same size as the olivine matrix, although some larger grains (3–4 mm) are present in sample 3a. Extremely fine-grained material, which is localized along the xenolith margins and grain boundaries, especially those of clinopyroxene, is the result of the early stages of partial fusion during entrainment and ascent of the xenoliths in basalt. No shape preferred orientation of olivine or pyroxene grains is observed in any of the xenoliths studied in detail. This, however, may be related to the orientation of the xenolith thin sections, which were created to maximize sizes and not relative to structural fabrics within the xenolith. Table 1 Modal analysis of CLB mantle xenoliths Sample 2a 3a 6a 6a⁎ 7a 7c 12a Vol.% Harzburgite Lherzolite Wehrlite Pyroxenite Lherzolite Harzburgite Lherzolite Olivine Orthopyroxene Clinopyroxene Spinel Apatite Alkali feldspar Plagioclase Sum F1 81 15 4 0.5 – – – 100 6.4 64 23 12 0.5 – – – 100 3.5 68 11 16 0.9 1 4 – 100 – 68 24 7 2 – – – 100 3.0 83 14 3 0.2 – – – 100 8.5 66 19 15 0.4 – – – 100 3.1 6 – 70 14 2 – 8 100 – Modal analyses of xenoliths based on image analysis of X-ray maps, described in more detail in the text. F denotes the percent of partial melting of each sample where F = 10 × ln(Cr#)spinel + 24 (Hellebrand et al., 2001). Sample 6a, the composite xenolith, has two entries in this table for the wehrlite and pyroxenite portions of the xenolith. 4 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 2.3. Mineral compositions Mineral compositions were determined by means of a Cameca SX 50 electron microprobe, using an accelerating voltage of 15 kV, a beam current of 20 nA, a suite of analyzed natural minerals as standards and the ϕ(ρz) data reduction program (Armstrong, 1988). Mineral grains were analyzed in several domains in each sample, and no significant compositional differences were found in the peridotites on the interor intragranular scale, except locally at the rims of some clinopyroxene and spinel grains near the xenolith borders, reflecting changes due to xenolith excavation and entrainment. In contrast to the peridotites, the composite sample 6a shows a marked compositional variation in minerals between clinopyroxenite and adjoining wehrlite, as described below. Representative analyses of constituent minerals are summarized in Table 2. Mineral compositions in the Coyote Lake basalt Crdiopside xenoliths are typical of those in Cr-diopside xenoliths elsewhere. Olivine is magnesian, with Mg# being restricted to values between 89.3 and 90.4, where Mg# is defined as 100 × molecular Mg/(Mg + Fe). Spinel is rich in MgAl2O4 component (Fig. 3a), with Mg# ranging from 78.5 to 84.8 and Cr# from 12.4 to 21.2, where Cr# is defined as 100 × molecular Cr/(Cr + Al). The most Cr-rich spinel occurs in the two samples of harzburgite, for which the highest degrees of partial melting, 6.4% and 8.5%, are calculated (Fig. 3a). Orthopyroxene and clinopyroxene are also magnesian and contain relatively large R2O3 contents (0.42– 0.62 wt.% Cr2O3 and 4.23–5.05 wt.% Al2O3 in orthopyroxene; 0.76–1.19 wt.% Cr2O3 and 5.45– 6.94 wt.% Al2O3 in clinopyroxene). The R2O3 contents of coexisting ortho- and clinopyroxene and Cr#'s in associated spinel are shown in Fig. 3b, demonstrating the positive correlation between Cr# in spinel and Cr/Al ratio in ortho- and clinopyroxene (i.e. an increase in spinel Cr# is accompanied by a continuous shift in position and increase in slope of the R2O3 tie lines in pyroxene). The two harzburgite samples contain pyroxene with the highest Cr/Al ratio. Minerals in composite xenolith 6a are less magnesian and chromiferous than those in the Cr-diopside peridotites. In addition, mineral compositions differ between the two lithologies of the composite xenolith, with an increase in Mg/Fe and Cr/Al ratios and a decrease in TiO2 content from clinopyroxenite to wehrlite. For example, the Mg# in clinopyroxene increases from 83.4 in clinopyroxenite to 88.5 in wehrlite, TiO2 decreases from 1.68 to 0.94 wt.% and Cr2O3 increases from 0 to 0.68 wt.%, with a corresponding decrease in Al2O3 from 8.62 to 6.58 wt.%. A concomitant variation occurs in spinel, which is Cr-free in clinopyroxenite and Cr-bearing (Cr# = 6.3) in wehrlite. Such variations between mineral compositions in the clinopyroxenite and wehrlite of sample 6a are illustrated graphically in Fig. 3, where these differences may be compared with the mineral compositions in the five peridotite xenoliths that cluster more closely on each graph. 3. Geothermometry Equilibration temperatures for the Coyote Lake mantle xenoliths were determined by application of three different calibrations of the two-pyroxene geothermometer (Bertrand and Mercier, 1985; Brey and Köhler, 1990; Taylor, 1998) and by the Al-inorthopyroxene geothermometer (Witt-Eickschen and Seck, 1991). Temperatures were calculated at an assumed pressure of 10 kbar, because of the uncertainty in estimating pressures for spinel peridotites (see below). Results from the three two-pyroxene geothermometers are in excellent agreement (Table 3, Fig. 4), although the Brey and Köhler (1990) calibration yields temperatures that are consistently about 50 °C higher than those from the Bertrand and Mercier (1985) and Taylor (1998) calibrations. Least squares analysis of temperatures from the three geothermometers, taking the Taylor calibration as the independent variable, yields: TBertrandMercier ð-CÞ ¼ 17:5 þ 0:985 TTaylor ð-CÞ R2 ¼ 0:977 TBreyKPhler ð-CÞ ¼ 151 þ 0:896 TTaylor ð-CÞ R2 ¼ 0:969 Using the Bertrand–Mercier and Taylor calibrations, the mean temperatures for the six analyzed samples range from 970 to 1060 °C (Table 3). Temperatures were also calculated for three additional peridotite xenoliths, which were analyzed in grain mounts prepared from crushed samples (shown in Fig. 4, but not included in Table 3). Results from two of these xenoliths, 1030 and 1045 °CTaylor, lie within the range of the previous six samples, and the third, 1090 °CTaylor, extends the range to a slightly higher temperature. Recalculation of temperature estimates from three other Coyote Lake basalt peridotite xenoliths, based on mineral compositions reported in Jové and Coleman S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 (1998), yield 925, 965 and 1085 °CTaylor. Based on all available results, the Coyote Lake basalt mantle suite appear to have been derived from a relatively restricted temperature range of 175 °C. Results from the Al-in-orthopyroxene geothermometer are comparable to those from the two-pyroxene geothermometers (Table 3), being within 40 °C of the temperatures obtained from the Bertrand–Mercier and Taylor calibrations. Such correspondence between the 5 two types of geothermometers and the regular distribution of sub-parallel pyroxene tie lines in Fig. 3b indicates the attainment and preservation of equilibrium compositions in pyroxene with respect to Ca, Mg, Fe, Al and Cr. 4. Estimated depths In general, spinel peridotites are assigned to an intermediate pressure range from ∼ 8 to ∼ 18 kbar Fig. 2. Photomicrographs of xenoliths taken under crossed-polarizers. Inset in each figure taken under plane-polarized light illustrates the texture of spinel. Width of each photograph is 1 cm. Numbers indicate temperature (°C) and depth (km) estimates, based on the thermal model in Fig. 5. 6 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Table 2 Representative mineral analyses for Coyote Lake basalt mantle xenoliths Wt.% SiO2 TiO2 Al2O3 Cr2O3 V2O3 FeO MnO MgO ZnO NiO CaO Na2O Sum Si Ti Al Cr V Fe Mn Mg Zn Ni Ca Na Sum Mg# Cr# Olivine Spinel 2a 3a 6a 7a 7c 12a 40.31 40.52 39.87 40.07 40.37 40.87 3a 0.07 54.42 12.54 0.08 11.21 0.11 20.96 0.08 0.33 0.22 60.42 5.79 0.08 12.22 0.09 21.08 0.08 0.30 99.80 100.28 10.37 0.15 48.33 9.91 0.14 48.48 12.03 0.16 47.07 9.37 0.16 49.31 9.61 0.14 48.97 9.71 0.13 48.78 0.32 0.36 0.27 0.37 0.36 0.34 0.14 50.66 15.54 0.10 12.90 0.13 20.12 0.08 0.27 99.48 99.41 99.40 99.28 99.45 99.82 99.94 Cations per 4 oxygen atoms 0.997 1.000 0.995 6a⁎ 2a 6a 7a 7c 12a 0.26 65.25 0.00 0.10 14.93 0.11 20.13 0.07 0.17 0.13 56.22 11.93 0.08 9.07 0.11 21.91 0.07 0.33 0.16 46.38 18.56 0.10 14.72 0.13 19.36 0.07 0.30 0.09 54.08 11.44 0.07 12.71 0.10 20.76 0.09 0.36 101.02 99.85 99.78 99.70 Cations per 4 oxygen atoms 0.990 0.996 1.003 0.214 0.003 1.782 0.205 0.003 1.784 0.251 0.003 1.751 0.194 0.003 1.816 0.198 0.003 1.800 0.199 0.003 1.785 0.006 0.007 0.005 0.007 0.007 0.007 3.003 89.3 3.000 89.7 3.005 87.5 3.010 90.3 3.004 90.1 2.997 90 0.003 1.598 0.329 0.002 0.289 0.003 0.803 0.002 0.006 0.001 1.684 0.260 0.002 0.246 0.002 0.821 0.002 0.007 0.004 1.824 0.117 0.002 0.262 0.002 0.805 0.002 0.006 0.005 1.939 0.000 0.002 0.315 0.002 0.757 0.001 0.003 0.003 1.716 0.244 0.002 0.196 0.002 0.846 0.001 0.007 0.003 1.498 0.402 0.002 0.337 0.003 0.791 0.001 0.007 0.002 1.684 0.239 0.001 0.281 0.002 0.818 0.002 0.008 3.033 73.5 17.1 3.025 76.9 13.4 3.024 75.4 6.0 3.025 70.6 0.0 3.017 81.2 12.4 3.045 70.1 21.2 3.036 74.4 12.4 Upper tabulation shows weight percent oxides of representative analyses of olivine, spinel, orthopyroxene and clinopyroxene from Orthopyroxene Clinopyroxene * each xenolith sample. 6a has two different lithologies: 6a denotes the wehrlite portion portion. 2a 3a 6a 7a 7c 12a 2a 3a 6a and 6a 6a⁎ denotes 7a the pyroxenite 7c 12a Lower tabulation shows the number of cations per 4 oxygen atoms for olivine and spinel, and per 6 oxygen atoms for orthopyroxene and clinopyroxene. The bottom two rows show the Mg# and Cr# for each mineral. The Mg# is defined for silicates as the molecular percent Mg/(Mg + Fetotal) and for spinel as Mg/(Mg + Fe2+), where Fe2+ is calculated by stoichiometry. Cr# is similarly defined as the molecular percent Cr/(Cr + Al). (depending on bulk-rock Mg# and Cr#), between the stability fields of plagioclase peridotite and garnet peridotite, and maximum pressures are estimated for individual samples from the Cr content of spinel (O'Neill, 1981). However, no viable procedure exists at present for specific pressure determinations of spinel peridotites. Although a geobarometer based on the Ca content of olivine was proposed by Köhler and Brey (1990), and accurate and precise determinations of trace amounts of Ca in olivine can be made by electron microprobe under appropriate operating conditions, application of the geobarometer produces spurious results, especially for lower temperature xenoliths (O'Reilly et al., 1997; Medaris et al., 1999). Despite the present lack of a geobarometer suitable for spinel peridotites, depth estimates for spinel peridotite xenoliths can be obtained by combining equilibration temperatures with a geotherm associated with the eruptive site at the time of xenolith transferal to the surface. The method and results of a numerical model for thermal evolution of the Coyote Lake basalt site are presented in the following section. 5. Thermal modeling and evolution The thermal evolution of lithosphere beneath the eastern Coast Ranges was estimated from a thermal model, which simulates upwelling of 1300 °C asthenosphere that fills in the slabless window following the northward movement of the Mendocino triple junction (Lachenbruch and Sass, 1980). The one-dimensional, finite difference model extended from the surface to a S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Orthopyroxene 2a 3a 7 Clinopyroxene 6a 7a 12a 2a 3a 6a 6a⁎ 7a 7c 12a 53.90 0.01 4.49 0.48 53.59 0.07 5.05 0.50 53.21 0.30 5.17 0.26 54.02 0.13 4.94 0.43 54.31 0.06 4.23 0.62 54.19 0.12 5.05 0.42 51.73 0.19 5.55 1.19 51.35 0.33 5.93 0.76 50.04 0.94 6.58 0.68 48.02 1.68 8.62 0.00 51.56 0.47 6.04 0.78 51.75 0.27 5.45 1.08 51.05 0.36 6.94 0.85 6.87 0.15 33.15 6.39 0.14 32.91 7.97 0.19 32.21 6.20 0.14 33.49 6.35 0.14 33.37 6.39 0.14 32.84 3.34 0.12 16.69 3.12 0.10 16.25 3.66 0.12 15.84 5.20 0.16 14.67 2.66 0.07 16.11 3.07 0.10 16.56 3.25 0.10 16.15 0.94 0.08 100.07 1.51 0.13 100.29 0.84 0.07 100.22 1.06 0.07 100.48 1.07 0.08 100.23 1.07 0.08 100.30 20.06 1.04 99.91 20.76 0.94 99.54 20.73 0.91 99.50 19.90 0.98 99.23 20.97 1.04 99.70 20.39 1.12 99.79 19.78 1.16 99.62 Cations per 6 oxygen atoms 1.873 1.858 1.855 0.000 0.002 0.008 0.184 0.206 0.212 0.013 0.014 0.007 7c 1.864 0.003 0.201 0.012 1.880 0.002 0.173 0.017 1.873 0.003 0.206 0.011 Cations per 6 oxygen atoms 1.878 1.871 1.833 0.005 0.009 0.026 0.238 0.255 0.284 0.034 0.022 0.020 1.774 0.047 0.375 0.000 1.873 0.013 0.259 0.022 1.881 0.007 0.233 0.031 1.855 0.010 0.297 0.024 0.200 0.004 1.717 0.185 0.004 1.701 0.232 0.006 1.674 0.179 0.004 1.722 0.184 0.004 1.722 0.185 0.004 1.692 0.101 0.004 0.903 0.095 0.003 0.883 0.112 0.004 0.865 0.161 0.005 0.808 0.081 0.002 0.872 0.093 0.003 0.897 0.099 0.003 0.875 0.035 0.005 4.031 89.6 6.6 0.056 0.009 4.035 90.2 6.4 0.031 0.005 4.030 87.8 3.2 0.039 0.005 4.029 90.6 5.6 0.040 0.005 4.026 90.3 8.9 0.040 0.005 4.018 90.1 5.1 0.780 0.073 4.017 89.9 12.5 0.811 0.066 4.015 90.3 7.9 0.814 0.065 4.022 88.5 6.6 0.788 0.070 4.027 83.4 0.0 0.816 0.073 4.011 91.5 7.8 0.794 0.079 4.019 90.6 11.7 0.770 0.081 4.015 89.8 7.5 depth of 81 km. The initial condition consisted of a steady-state crustal geotherm with exponential decay of heat production from the surface to a depth of 30 km. The steady-state geotherm was based on a reduced heat flow of 20 mW/m2 and a heat production contribution of 25 mW/m2, yielding an initial surface heat flow of 45 mW/m2. The reduced heat flow reflects the insulating effect of the subducted Farallon plate underlying the North American plate (Furlong, 1984) and the initial surface heat flow is comparable to that used by Zandt and Furlong (1982) and Furlong (1984). The asthenospheric upwelling was represented by a step temperature discontinuity to 1300 °C followed by an adiabatic gradient of 1 °C/km to 81-km depth. The boundary conditions were specified as temperatures of 0 °C at the surface and 1351 °C at 81-km depth. Other parameters used were a thermal conductivity of 2.5 W/m K and thermal diffusivity of 10− 6 m2/s. The calculated geotherms at 3.0 m.y. after upwelling (Fig. 5) match closely those shown by Furlong (1984) in his Fig. 3. The predicted surface heat flow changes through time, with 73 mW/m2 at a model time of 6 m.y., a peak value of 78 mW/m2 at 8 m.y. and 77 mW/m2 at 9 m.y. (0 Ma), compared to a present-day observed surface heat flow of 80 mW/m2. These results for the one-dimensional thermal model lie between the convective model I and conductive model II of Zandt and Furlong (1982, their Fig. 3c). Depths for the six Coyote Lake basalt xenoliths were estimated by intersecting their temperatures (using results from the Taylor two-pyroxene geothermometer and taking into account its pressure dependence) with the 8 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 the Coyote Lake basalt xenoliths occurs at depths between 37.8 and 42.5 km. Thus, the Coyote Lake basalt spinel peridotite xenoliths were likely derived from a relatively narrow depth interval, ∼ 5 km, in the upper part of the lithospheric mantle. This estimate is consistent with their phase petrology and well within the maximum pressures allowed by the compositions of constituent spinel (Table 2). If the higher temperatures from the Brey and Köhler two-pyroxene geothermometer are used, the depth interval remains constant at ∼ 5 km, but the depth estimates increase by about 3 km representing depths of approximately 40–45 km. Thus, the spinel peridotite xenoliths from the Coyote Lake basalt allow us to evaluate the physical properties of the mantle associated with the San Andreas fault system, from a depth of ∼ 40 km at 3 Ma, approximately 6 m.y. after passage of the Mendocino triple junction. 6. Olivine petrofabrics Fig. 3. Graphical representations of geochemical data. (a) Compositions of spinel in mantle xenoliths in the Coyote Lake basalt. Such Mgrich and Cr-poor spinel compositions are typical for relatively undepleted mantle peridotite. The numbers adjacent to the spinel data points indicate the Mg#'s of coexisting olivine. Note the relatively Fe-rich and Cr-poor compositions of spinel in the composite sample, 6a: pd, peridotite (wehrlite); pxite, pyroxenite. (b) R2O3 contents of orthopyroxene (opx) and clinopyroxene (cpx) in mantle xenoliths in the Coyote Lake basalt. The sub-parallel tie lines between opx and cpx indicate chemical equilibrium among pairs of pyroxenes. Cr#'s of coexisting spinels are indicated by the italicized numbers adjacent to clinopyroxene compositional points. Higher Cr/Al ratios in pyroxenes (and corresponding spinels) reflect higher degrees of partial melting. Note the lower Cr/Al ratios of pyroxene in the composite sample; pd, peridotite; pxite, pyroxenite. geotherm calculated at a model age of 6 m.y. (Table 3, Fig. 5). This geotherm corresponds to the eruption of the Coyote Lake basalt at 3 Ma. The range of temperatures of The lattice preferred orientation (LPO) of olivine was measured in the five peridotite xenolith samples (but not the composite pyroxenite/wehrlite sample 6a) using a five-axis universal stage microscope following the method of Emmons (1943). For each thin section, the orientation of each individual olivine grains was measured. Due to variations in the physical size of each xenolith, between 70 and 165 grain orientations were determined in each thin section. The LPO was not measured for the orthopyroxene and clinopyroxene grains in thin section since the small volume fraction of pyroxene represented in each sample (Table 1) makes it impossible to determine a statistically significant LPO pattern for these minerals. Olivine LPO data are presented in Fig. 6a. On each lower hemisphere, equal-area projection, olivine a-axes, [100], are oriented horizontal and east–west, while the caxes, [001], are approximately vertical. This orientation was chosen for several reasons. First, LPO data are typically plotted with a vertical east–west foliation and a horizontal lineation, where the foliation and lineation are determined independently based on the shape preferred orientation of the constituent minerals. Because no shape preferred orientation was apparent in these samples, the selected orientation provides a common framework to compare the LPO patterns of each sample in the suite, and to compare the patterns with those from other studies of mantle xenoliths lacking a strong shape preferred orientation (e.g. Christensen et al., 2001; Vauchez et al., 2005). Second, this orientation is reasonable based on the bulk kinematics of the San Andreas fault system, where a vertical foliation with a horizontal lineation is S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 9 Table 3 Estimated temperatures, maximum pressures, depths and calculated seismic anisotropy values of Coyote Lake basalt mantle xenoliths Sample 6a 7a 3a 7c 2a 12a Average T at 10 kbar 2-Px, T 2-Px, BM 2-Px, BK Al-in-Opx 970 975 1000 1005 1040 1060 1016 975 980 1005 1010 1040 1060 1019 1025 1020 1045 1055 1080 1100 1060 930 990 1005 1045 1010 1030 1016 Pmax (kbar) Temp. (°C) Depth (km) Vp max (km/ s) Vp min (km/ s) AVp (%) Vs1 max (km/ s) Vs1 min (km/ s) Vs 2 max (km/ s) Vs2 min (km/ s) AVs (%) δt (s) 16.1 18.5 21.2 19 20.5 19.6 20 975 980 1015 1020 1055 1070 1028 37.8 37.9 39.6 39.8 41.7 42.5 40.3 – 8.71 8.61 8.84 8.80 8.53 8.70 – 7.96 8.03 8.01 7.97 8.17 8.03 – 9.0 7.0 9.9 9.9 4.3 8.0 – 5.01 4.95 5.01 5.02 4.94 4.99 – 4.83 4.84 4.87 4.82 4.83 4.84 – 4.87 4.88 4.92 4.90 4.87 4.89 – 4.68 4.70 4.64 4.68 4.74 4.69 – 6.8 5.2 7.6 7.0 4.1 6.1 – 1.4 1.1 1.6 1.4 0.9 1.3 The left section of the table shows estimated temperatures, maximum pressures and depths based on geothermometry and thermal modeling. Four thermometers were used to estimate temperatures at an assumed pressure of 10 kbar: three two pyroxene thermometers: T—Taylor (1998), BM— Bertrand and Mercier (1985), BK—Brey and Köhler (1990) and one Al-in-opx geothermometer from Witt-Eickschen and Seck (1991). Maximum pressures are calculated from the composition of spinel (O'Neill, 1981). Temperatures and depths are based on a combination of two-pyroxene geothermometry and the thermal model presented in Fig. 5. The right section of the table shows the calculated seismic anisotropy values, based on the LPO fabrics presented in Fig. 6a and computed using software by Mainprice (1990). The anisotropy of P-waves, AVp, is defined as (Vp max − Vp min) / (average Vp) from Mainprice and Silver (1993); the shear wave anisotropy, AVs, is similarly defined. Shear wave delay times (δt) are reported for a 100-km path defined as AVs normalized by the average S-wave velocity. anticipated due to dominantly wrench deformation at the plate boundary. Although no shape preferred orientation was visible in these samples, olivine a-axes in many naturally and experimentally deformed peridotites are sub-parallel to the lineation direction and a- and c-axes define a plane parallel to the foliation (e.g. Zhang and Karato, 1995; Ben-Ismail and Mainprice, 1998; Tommasi et al., 2004). Third, this orientation produces maximum values of shear wave splitting for vertically Fig. 4. Comparison of temperature estimates (at 10 kbar) for the Coyote Lake basalt mantle xenoliths from the T—Taylor (1998), BM— Bertrand and Mercier (1985) and BK—Brey and Köhler (1990) twopyroxene geothermometers. incident shear waves, described in more detail in the following section. The LPO patterns in all the Coyote Lake basalt xenoliths display strong point distributions of olivine a- Fig. 5. Calculated thermal evolution of the Coyote Lake area, showing the estimated temperatures and depths for the Coyote Lake basalt mantle xenoliths. Samples symbols as in Fig. 4. 10 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 axes except 12a, in which a-axes form a girdle. Olivine b-axes are characterized by slightly weaker point distributions (fewer contours) and in 7c, a weak girdle is developed. The c-axes show the most diffuse LPO patterns, which is typical in both naturally deformed peridotites from ophiolite massifs (e.g. Boudier and Coleman, 1981; Salisbury and Christensen, 1985), xenoliths (e.g. Soedjatmiko and Crhistensen, 2000), experimentally deformed peridotites (e.g. Carter and Avé Lallement, 1970; Nicolas et al., 1973) and in modeled olivine LPO patterns from many different kinematic environments (e.g. Tommasi et al., 1999). There is no systematic change in olivine LPO patterns as a function of depth, such as that observed in other xenolith suites (Christensen et al., 2001; Kennedy et al., 2002; Kobussen, 2005). The olivine LPO patterns can be compared to predicted LPO patterns under a variety of imposed kinematic constraints and based on numerical modeling. The pattern observed in the Coyote Lake basalt xenoliths is more consistent with deformation in simple shear or in pure shear, both of which predict a-axis point maxima Fig. 6. (a) Lower hemisphere, equal area projections of olivine a-, b- and c-axes measured on a universal stage microscope; contours are multiples of uniform distribution with the lowest contour indicated by the dashed line. (b) Seismic anisotropy values calculated from the olivine LPO patterns. Compressional wave velocity (Vp) in km/s, shear wave splitting (AVs) in % and polarization directions for shear waves with different incidence angles. The black square indicates the maximum velocity or anisotropy and the white circle indicates the minimum; CI stands for contour interval. The numerical valves for Vp and Vs for each xenolith sample are included in Table 3. S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 (Tommasi et al., 1999). The LPO patterns developed in transpression, which might be expected given the oblique convergence along the San Andreas fault system, are characterized by a-axis girdles (Tommasi et al., 1999), which are only observed in sample 12a. Because an independent measure of foliation and lineation is not available in these xenoliths, we cannot definitively determine the predominant slip system. However, the tendency towards orthorhombic point distributions with strong a-axes concentrations suggests that deformation may best be described by dislocation creep at relatively high temperatures, N1000 °C, on the slip system (010)[100] (e.g. Nicolas and Poirier, 1976; Carter and Avé Lallement, 1970). This is one of the most common slip systems for olivine from xenolith suites (e.g. Ben-Ismail and Mainprice, 1998; Tommasi et al., 2000) and is consistent with the temperatures determined by geothermometry. 7. Seismic anisotropy calculations The small sample sizes of these xenoliths require calculation, rather than direct measurement, of seismic properties. Seismic velocities are calculated using the LPO, density and the elastic stiffness coefficient for each mineral in the rock (e.g. Crossen and Lin, 1971; Mainprice and Silver, 1993). These calculations assume samples are free from cracks and alteration, and use the modal composition of each xenolith (Table 1) for the volume fraction of each phases. We used published elastic stiffness coefficients for olivine (Abramson et al., 1997), enstatite (Duffy and Vaughan, 1998) and diopside (Collins and Brown, 1998), and our calculations were made using updated software by Mainprice (based on Mainprice, 1990; Mainprice and Humbert, 1994) utilizing the Voigt averaging technique. Choice of this average is based on the agreement between laboratory measurements of seismic velocities and calculated values from LPO patterns for the Twin Sisters dunite (Crossen and Lin, 1971), and is useful for comparison with other xenolith petrofabric studies, the majority of which also use the Voigt average (Table 4). The averaging technique affects the absolute velocities but not the anisotropy values (Mainprice and Silver, 1993). Because pyroxene LPOs were not measured in our samples, but their LPOs tend to decrease the overall anisotropy (Christensen and Lundquist, 1982), we approximated this contribution by assuming that pyroxene LPO is uniformly distributed (i.e. isotropic to seismic waves). The small contribution from minor phases like spinel (b 5% of the total volume) was ignored in the calculations. Delay times were computed for a 100-km 11 thick slab with an anisotropy equal to the aggregate anisotropy of each sample. These assumptions, which are common in many petrofabric studies (e.g. Mainprice and Silver, 1993), generally produce anisotropies that are ∼ 1% higher than for seismic anisotropy measured in rocks in situ (e.g. Christensen, 2002). The results of the seismic anisotropy calculations are presented in Fig. 6b. The orientation of each sample is the same as in Fig. 6a. Regardless of the LPO fabric strength (or total number of grains), the seismic anisotropy values are similar for all five xenoliths (Table 3). The average values for the compressional wave velocity (Vp) and anisotropy (AVp) are 8.4 km/s and 8.0%; the average shear wave velocity (Vs) and anisotropy (AVs) are 4.9 km/s and 6.1%; and the average delay time (δt) is 1.3 s for a 100-km path. It is important to note that the maximum shear wave splitting is near vertical in Fig. 6b, suggesting that with this particular orientation, delay times for vertically incident seismic waves will be close to the maximum. These results can be compared with data from similar studies of mantle xenoliths from different tectonic environments, which have been compiled in Table 4. For this compilation, we recompute seismic anisotropy parameters from each paper, where necessary, to create a common framework for comparison between studies. For example, we weight each xenolith with normal mantle composition equally when calculating the average seismic anisotropy parameters. We do not use reported seismic anisotropy values calculated from aggregate LPO data (e.g. Ji et al., 1994; Saruwatari et al., 2001); thus, the values in this table may not agree with stated values in the original source. For completeness, Table 4 also summarizes the modal variations of major minerals, as well as temperatures and pressures provided for each xenolith suite. The modal composition of each individual xenolith or the regional modal average was used in the seismic anisotropy calculations except in the studies by Pera et al. (2003), Christensen et al. (2001) and Kobussen (2005), where seismic anisotropy calculations assumed xenoliths were composed of 100% olivine and therefore represent maximum values. Additionally, the reported temperature ranges in some instances are based on older geothermometers (e.g. Boyd, 1973; Wells, 1977) that have been superseded by more recent geothermometers (e.g. Bertrand and Mercier, 1985; Brey and Köhler, 1990; Taylor, 1998). Depth estimates are commonly based on the inferred geothermal gradients in each tectonic setting, since estimation of pressures from spinel peridotites is problematic (e.g. Medaris et al., 1999). 12 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Table 4 Comparison of mantle xenolith compositions and calculated seismic anisotropy values from different tectonic settings Location Lithology Cratonic environments Kaapvaal craton Peridotite ± kimberlites, garnet South Africa Kaapvaal craton Peridotite ± kimberlites, garnet South Africa Kaapvaal craton Peridotite ± kimberlites, spinel/ South Africa garnet Labait volcano, Spinel/ Tanzania garnet peridotite Convergent setting N. British Spinel Columbia, peridotite Canada and Alaska, USA S. British Spinel Columbia, peridotite Canada Peridotite Torre Alfina, Northern Apennines, Italy Method ol opx cpx sp gt Ave. Vp Vs AVp AVs (%) (%) (%) (%) (%) (km/s) (km/s) (%) (%) 5 U 56– 22– – 80 44 – 3– V 8 8.4 NA 5.4 3.7 ∼ 1 900– 1050 120– 170 6 U 33– 10– b4 60 32 – 7– H 12 8.3 4.8 5.4 4.4 0.9 980 150 48 U/ EBSD NA NA NA NA NA V 8.2 4.7 3.0 2.6 0.6 1200– 170– 1400 200 Mainprice and Silver (1993) Long and Christensen (2000) Ben-Ismail et al. (2001) 13 EBSD 60– 4– 94 28 1 – b15 VRH 8.3 3 NA 6.5 5.1 N1 1000– 70– 1400 140 Vauchez et al. (2005) 13 U 64– 21– 6– 68 26 11 1– – 2 12 U 54– 25– 12– b3 61 34 21 15 EBSD Extension setting (rifts) Vitim, Baikal Peridotite + 4 U region, Russia spinel/garnet Cima volcanic Spinel 15 U field, California, peridotite USA Kozákov volcano, Czech Republic As Shamah volcanic field, Syria Hotspot track Polynesian hotspots, South Pacific 1– 7 NA 6.3 4.5 1.0 900– 1100 45– 59 Ji et al. (1994) – V 8.0 4.6 7.9 6.6 1.5 900– 1100 45– 67 Saruwatari et al. (2001) N90 NA NA NA – V 8.2 4.6 13.5 11.5 1.8 1000– 50– 1080 60⁎ 8.2 4.4⁎ 5.8 4.4⁎ 0.4 8.3 NA 7.4 5.2 40– 90 30– 45 680– 1065 900– 1100 57– 86 54– 93 10– FIX b3 b6 NA 26 6– 1– ≤1 – H 38 7 49– 82 65– 90 0– 34 0– 25 Kern et al. (1996) Soedjatmiko and Crhistensen, 2000 Christensen et al. (2001) Kobussen (2005) 1– – 6 2– – 4 NA 8.6 4.9 NA 8.5 4.9 4–8 2 . 4 – 1.0 5.0 8.5 6.3 1.3 4– 37 NA – VRH 8.3 4.8 6.0 4.2 0.9 b1100 NA Tommasi et al. (2004) b2 V 4.9 8.0 6.1 1.3 970– 1100 This study 20 EBSD 45– 0– 100 30 5 64– 11– 4– 83 23 16 U 925– 1200 NA 955– 1060 Pera et al. (2003) 5 – 13 5– 8 Spinel peridotite U Depth Citation (km) 8.0 15 U 9 T (°C) V Spinel lherzolite Spinel peridotite Transform setting Calaveras fault, Spinel California, USA peridotite δt (s) N – 8.4 32– 70 NA 38– 43 Compilation of xenolith composition data and calculated seismic anisotropy values based on LPO measurements from different tectonic environments. N represents the total number of xenoliths studied. The method of LPO measurement is either by using a U-stage (U) or electron backscatter diffraction patterns (EBSD). Range of modal percentages is reported for olivine (ol), orthopyroxene (opx), clinopyroxene (cpx), spinel (sp) and garnet (gt). The averaging technique for seismic anisotropy calculations (Ave.) is abbreviated as Voigt (V), Hill (H) or Voigt–Reuss–Hill (VRH). Average seismic anisotropy values are reported for P-wave velocity (Vp), S-wave velocity (Vs), P-wave anisotropy (AVp), S-wave anisotropy (AVs) and shear wave delay time (δt) for a 100-km ray path. The range of temperatures (T) and estimated depths are reported for each xenolith suite. The S-wave velocity and anisotropy that are starred (Kern et al., 1996) are from experimentally determined values (not calculated from the xenoliths' LPOs). NA indicates that specific data were not available. S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 The seismic velocities of compressional (Vp) and shear waves (Vs) in the Coyote Lake basalt xenoliths are similar to those observed from other tectonic environments, while the anisotropy of compressional (AVp) and shear waves (AVs) is generally higher. In particular, the 6% shear wave anisotropy from the Coyote Lake basalt xenoliths is higher than the 4–5% anisotropies calculated for xenoliths from cratons and rift environments but similar to values from orogenic belts that range from 4.5% to 11%. This higher anisotropy significantly affects the interpretations of modern shear wave splitting studies, which generally assume 4% shear wave anisotropy for all tectonic environments, based primarily on the LPOs of kimberlite xenoliths (e.g. Silver and Chan, 1991; Silver and Savage, 1994). The implications of the strong LPO patterns and associated seismic anisotropy values of the Coyote Lake basalt xenoliths on the tectonics of the San Andreas fault system are discussed below. 8. Discussion The xenoliths from the Coyote Lake basalt represent the only known samples of the mantle beneath the San Andreas fault system. The variation in mineral modes and compositions of the xenoliths (Tables 1 and 2) reflects compositional variability within a vertically restricted (∼ 5 km) region in the mantle. Note, however, that the xenoliths investigated here do not represent the full suite of mantle xenoliths from the Coyote Lake basalt; they were chosen because geothermometry was possible based on their coexisting mineral phases. The seismic anisotropy values calculated from petrofabrics in the Coyote Lake basalt peridotite xenoliths are significantly higher than those calculated from mantle xenoliths in most other tectonic settings, especially from xenoliths that sample the mantle beneath cratonic environments (Table 4). This suggests that the LPO strength is related to deformation associated with the San Andreas fault system; thus the Coyote Lake basalt xenoliths represent the first non-geophysical documentation of a shear zone in the mantle beneath the San Andreas fault. Because geothermometry and thermal modeling of the mantle xenoliths entrained in the Coyote Lake basalt indicate that xenoliths originated from depths of approximately 40 km, the San Andreas fault system presumably continues into the mantle as a shear zone to at least 40-km depth. This depth is approximately 10–15 km below the crust–mantle boundary in west–central California, based on seismic refraction profiles across the San Andreas fault system (Fuis and Mooney, 1990 and references therein). Because of their location and history within the deforming plate boundary system, the Coyote Lake 13 basalt xenoliths provide important information about the lithospheric structure along the San Andreas fault system. The LPO patterns aid the interpretation of shear wave splitting values from central California, and together these data sets shed light on the patterns of deformation in the mantle beneath the San Andreas fault system, as discussed in more detail below. 8.1. Comparison with shear wave splitting values While analysis of petrofabrics from xenoliths represents one of the few ways to directly study continental mantle LPO, albeit with small sample sizes, shear wave splitting measurements provide a large-scale picture that is interpreted to reflect LPO patterns in the mantle (e.g. Vinnik et al., 1989; Silver and Chan, 1991). If the mantle has a consistent anisotropy over a wide region due to the alignment of anisotropic crystals (primarily olivine), shear waves traveling through the medium will be split into two orthogonal polarized waves traveling at different speeds. Shear wave splitting studies generally use this birefringence in SKS and SKKS waveforms, which have near-vertical incidence, and measure the difference between arrival times of the fast and slow directions of these waves (e.g. Silver, 1996). The delay time between S-wave arrivals is indicative of the degree of anisotropy sampled by the waveform as well as the thickness of the anisotropic layer. For a given thickness of anisotropic mantle composed primarily of olivine, and assuming that olivine crystallographic axes track finite strain, the largest delay times are observed when wave propagation is parallel to the intermediate finite strain axis, with the fast wave traveling parallel to the maximum finite strain axis and the slow wave parallel to the minimum finite strain axis (e.g. McKenzie, 1979; Ribe, 1992). For typical olivine LPO patterns, this translates into propagation parallel to the c-direction with the fast wave parallel to the a-axis and the slow wave parallel to the b-axis. Anisotropy in the crust can also contribute to the observed shear wave splitting at the surface, but is often neglected since its maximum contribution is estimated to be 0.1–0.2 s per 10 km depending on the composition and orientation of bulk crustal fabrics (e.g. Barroul and Mainprice, 1993). Shear wave splitting delay times from central California are generally interpreted to result from two different anisotropic layers in the mantle: a deep asthenospheric layer responsible for EW oriented splitting and a lithospheric mantle layer with splitting directions parallel to the San Andreas fault (Silver and Savage, 1994; Ozalaybey and Savage, 1994, 1995; Hartog and Schwartz, 2001; Polet and Kanamori, 2002). The average delay 14 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 times for the lower asthenospheric layer are generally 0.85–1.7 s and 0.5–1.25 s for the upper lithospheric layer (Hartog and Schwartz, 2001). The observed delay times using a two-layer model are illustrated in Fig. 7a. The deeper asthenospheric splitting is attributed to asthenospheric flow within the slab window (Ozalaybey and Savage, 1995; Hartog and Schwartz, 2000) or to absolute motion of the Sierra Nevada–Great Valley block (Hartog and Schwartz, 2001). The shallow lithospheric fabric is generally attributed to strain in the upper mantle along the plate boundary, the lateral extent of which seems to extend approximately 55 km to S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 the west of the San Andreas fault and up to 80 km to the east (Ozalaybey and Savage, 1995). The magnitude of delay times from the upper anisotropic layer tends to decrease towards the east (Fig. 7a; Hartog and Schwartz, 2001), although the orientation remains parallel to the San Andreas fault. This is in marked contrast to the rotation of fast shear wave splitting directions observed across the Alpine fault in New Zealand (e.g. Molnar et al., 1999). Fault-parallel fabrics are also observed in central California, determined by Pn tomography, which sample only the upper mantle (Hearn, 1996). The twolayer anisotropy models based on shear wave splitting values (Ozalaybey and Savage, 1995; Polet and Kanamori, 2002) generally assume a 4% shear wave anisotropy in both layers of the mantle (e.g. Silver and Chan, 1991; Silver and Savage, 1994). These models therefore require 100–150 km of lithospheric mantle with a fault-parallel fast polarization direction in central California to account for the observed shear wave splitting of the upper layer (Ozalaybey and Savage, 1995; Hartog and Schwartz, 2001). The calculated depths of the Coyote Lake basalt xenoliths place the xenoliths within the upper lithospheric mantle layer in the two-layer model. The average 6% shear wave anisotropy calculated from the xenolith LPO patterns can be used to estimate the thickness of the upper anisotropic layer depending on the orientation of LPO in the mantle. The fabric strength can also be compared to the fault-parallel fast direction and the kinematics of the fault system in order to understand fabric development in these xenoliths. 8.2. Lithospheric thickness If foliation (defined here by olivine a- and c-axes) is vertical and lineation (a-axes) is horizontal (Fig. 6a), the maximum shear wave splitting is nearly vertical and maximum delay times will be observed at the surface for vertically propagating waves with the fast wave parallel to a-axes (Fig. 6b). The observed shear wave splitting delay times require 40–95 km of anisotropic mantle with a 15 consistent orientation. If instead, foliation and lineation are both horizontal (horizontal a- and c-axes), the calculated shear wave anisotropies for vertical incidence are slower due to the orthorhombic LPO symmetry in four out of the five xenoliths. Only sample 12a predicts a maximum vertical shear wave anisotropy for a horizontal foliation—the four other xenoliths would have shear wave anisotropies approximately half to two-thirds as great, suggesting a required lithospheric mantle thickness of 60–150 km. Seismological data (Zandt, 1981; Hill, 1989) and thermal modeling (Zandt and Furlong, 1982) indicate that, although there are variations in depth to the asthenosphere in central California, the total lithospheric thickness is probably less than 80 km and may be as little as 40–60 km. For such a thin lithosphere, the delay times from shear wave splitting measurements at the surface are more consistent with an anisotropic mantle in which foliation is vertical and lineation is horizontal. 8.3. LPO development Olivine crystallographic axes are interpreted to track finite strain orientations (e.g. McKenzie, 1979; Ribe, 1992) but at high strains may rotate towards the shear direction faster than the finite strain axis (e.g. Nicolas and Christensen, 1987; Zhang et al., 2000). In most situations, the olivine a-axes are aligned with the lineation direction (see however Mizukami and Wallis, 2005; Katayama et al., 2005); thus, the observed fast shear wave polarization direction should parallel the maximum finite strain axis and reflect the orientation of olivine a-axes in the lithospheric mantle. Presently, the orientation of fast shear waves in the lithospheric mantle layer is sub-parallel (b 10°) to the San Andreas fault (Fig. 7a), and we can calculate the shear strain required to produce this subparallel alignment of olivine a-axes based on the kinematics of the San Andreas fault system. For a simplified calculation, we assume that the plate boundary is deforming in simple shear (instead of intranspression) and that the width of the deforming Fig. 7. (a) Map of central California with shear wave splitting measurements. Arrow color reflects the two-layer anisotropy model discussed in the text, where white arrows show modeled shear wave splitting for the deeper asthenospheric mantle layer and dark arrow reflect splitting in the upper lithospheric mantle layer. Each arrow is scaled by delay time length and oriented parallel to the modeled fast direction. Modified from Hartog and Schwartz (2001). (b) Two cross-sections through San Francisco Bay showing possible orientations of major fault strands within the crust based on seismic reflection studies. The first model illustrates a horizontal decollément linking the different faults of the San Andreas fault system; the second model shows vertical continuations of all major faults. The width of the shear zones in the mantle may either be discrete or reflect distributed shearing at depth. See text for details. Adapted from Brocher et al. (1994) and Bürgmann (1997). (c) Three-dimensional model for the lithosphere in central California, including the preferred orientation for olivine crystallographic axes to produce fault-parallel fast shear wave directions shown in (a). A vertical foliation with horizontal lineation produces the greatest shear wave splitting values for vertically propagating waves based on the LPO patterns from Coyote Lake basalt mantle xenoliths, thus requiring less than 100 km of anisotropic mantle to produce the observed delay times at the surface. Modified from Teyssier and Tikoff (1998). 16 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 zone in the mantle is approximately 130 km wide, based on the extent of fault-parallel polarization directions observed at the surface (Fig. 1). To align the long-axis of finite strain (and therefore olivine a-axes) within 10° of the San Andreas fault requires a simple shear component of γ ≈ 5 averaged over the whole zone (e.g. Tikoff and Fossen, 1993). This translates into 650 km of faultparallel displacement that must be accommodated after passage of the Mendocino triple junction at ∼ 10 Ma in central California (Fig. 1). Plate reconstructions suggest 605 ± 67 km of net transform displacement on the San Andreas fault system since 10.9 Ma (Dickinson, 1996) in good agreement with this simple forward calculation. This calculation suggests that there has been sufficient fault-parallel displacement since 10 Ma to align or reorient olivine a-axes in the mantle parallel to the strike of the San Andreas fault simply through the rotation of olivine crystals. Thus, the mantle beneath the San Andreas fault may have had a pre-existing LPO that was reoriented after strike-slip deformation began, or the fabric may have been erased by the high temperatures induced in the slab window trailing the triple junction. In either case, LPO development related to San Andreas fault kinematics may have been aided by dynamic recrystallization, which can speed up the process of LPO development (e.g. Tommasi et al., 1999; Zhang et al., 2000) so that olivine a-axes may align parallel to the San Andreas fault at lower finite strains. 8.4. Strike-slip shear zones in the mantle In order to understand the lithospheric structure beneath central California, it is necessary to combine all available geophysical and geological information. Based on the geophysical data sets, primarily seismic reflection/refraction surveys and gravity and magnetic measurements, there are several models for the manner in which plate boundary deformation of the San Andreas fault may be accommodated across major strike-slip faults, as summarized below. Shear wave splitting studies (e.g. Ozalaybey and Savage, 1995; Hartog and Schwartz, 2001; Polet and Kanamori, 2002) and xenolith LPO patterns provide additional information that may help distinguish between the different models. In the first type of model (Fig. 7b), plate boundary deformation in the crust is decoupled from the mantle lithosphere by horizontal mid-crustal detachments (e.g. Namson and Davis, 1988). Applying this model to central California suggests that there is a decollément horizon in the mid-crust linking a shallowly dipping San Andreas fault to vertical Hayward and Calaveras faults (e.g. Furlong et al., 1989; Brocher et al., 1994; Holbrook et al., 1996; Bürgmann, 1997). In this model, the Hayward and Calaveras faults represent the major plate boundary suture, and we would expect either a vertical or horizontal foliation in the upper mantle beneath these faults depending on which sections of the mantle were sampled by xenoliths. An example of this type of model is illustrated for San Francisco Bay (Fig. 7b), slightly to the north of the Coyote Lake basalt locale. In the second type of model (Fig. 7b), mantle deformation may be accommodated by discrete narrow shear zones beneath the major crustal faults (e.g. Savage and Burford, 1970; Thatcher, 1989), in effect blurring plate tectonic movements in the mantle (e.g. Tapponnier et al., 1982). Variations in crustal thickness on either side of the fault have been used as evidence for offset of the Moho in both northern (Henstock et al., 1997; Hole et al., 2000) and southern California (Zhu, 2000), implying that the plate boundary structure continues as a discrete feature in the upper mantle. Narrow zones, perhaps less than 5 km wide (e.g. Wilson et al., 2004), with vertical foliation and horizontal lineation would be predicted in this style of model beneath the major faults in the San Andreas fault system. A third possibility is that deformation is distributed across a broad region in the lower crust and upper mantle (Fig. 7b; e.g. Prescott and Nur, 1981; McKenzie and Jackson, 1983; Lamb, 1994; Bourne et al., 1998; Molnar et al., 1999; Wilson et al., 2004). A distributed deformation model predicts subhorizontal anisotropy in the lower crust (e.g. Wilson et al., 2004) and vertical foliations with horizontal lineations in the upper lithospheric mantle reflecting the predominance of strike-slip plate motion in the region (e.g. Molnar et al., 1999). Applying this model to central California suggests that the major sub-parallel faults of the San Andreas fault system merge at depth into a wider mantle shear zone (e.g. Parsons and Hart, 1999) and is somewhat similar to the previous model, except that much wider interconnected shear zones are anticipated in the mantle. It is likely that the upper mantle beneath the San Andreas fault has a vertical foliation and horizontal lineation based on the strike-slip kinematics and the necessity to create sufficiently large shear wave splitting delay times with a relatively thin lithosphere. Vertical foliations are predicted in all of the above models, although the width of strongly foliated areas varies in each model. Because consistent fault-parallel shear wave splitting is observed across a 130-km wide zone in central California (Fig. 1), the petrofabric and shear wave splitting data are most consistent with the third model of broad distributed deformation, as shown by the schematic three-dimensional cartoon in Fig. 7c. S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Fig. 7c illustrates the merging of the Calaveras and San Andreas faults at depth into a wider shear zone in the mantle. In detail, the upper crust south of San Francisco shows the major vertical strike-slip faults flanked by foldand-thrust belts. The mid-crustal foliations are highly interpretive (e.g. Jones et al., 1994), but are consistent with the seismic reflection data of Parsons and Hart (1999) and are based on an interpretation of coupling between the mid- and upper-crust predicted by numerical models of distributed deformation at plate boundaries (Wilson et al., 2004). Mantle fabrics in this schematic cartoon are based on shear wave splitting data (Fig. 7a) and illustrate the two layers of anisotropic mantle (Ozalaybey and Savage, 1995; Hartog and Schwartz, 2001). EW striking lineations with horizontal foliations are inferred for the asthenospheric mantle, similar to the model proposed by Silver and Holt (2002). These fabrics are not only observed along the San Andreas fault system but also east of the Great Valley (Ozalaybey and Savage, 1995; Polet and Kanamori, 2002). Fault-parallel vertical foliations with horizontal lineations are shown for the upper lithospheric mantle. The spacing of foliations in the model indicates the fabric strength, where more closely spaced lines reflect stronger alignment of olivine LPO, consistent with the longer delay times observed closer to the San Andreas fault (Fig. 7a). This model is specifically for central California, and may not hold for other areas of California where fault-parallel shear wave splitting fabrics have not been observed (Ozalaybey and Savage, 1995; Hartog and Schwartz, 2001; Polet and Kanamori, 2002). 9. Conclusions The Coyote Lake basalt, which is located at the intersection of the Hayward and Calaveras faults in central California, erupted more than 6 m.y. after the passage of the Mendocino triple junction. This delay in time suggests that volcanism at Coyote Lake was not related to the slab window trailing the migrating triple junction. Entrained in this alkali-basalt are the only known peridotite xenoliths from the mantle beneath the San Andreas fault system. Six upper mantle xenoliths were studied in detail by a combination of petrologic techniques. The compositions of the xenoliths vary and include Cr-diopside lherzolite, harzburgite, and one composite sample of wehrlite and clinopyroxenite. Mineral compositions within the xenoliths are generally similar, except for the composite sample. Xenolith equilibration temperatures are estimated to have been 970–1100 °C, based on two-pyroxene and Al- 17 in-orthopyroxene geothermometry. A thermal model was used in conjunction with the results from geothermometry to estimate depths of 38 to 43 km for the xenoliths. Olivine LPO patterns measured in five xenolith samples show orthorhombic symmetry with strong point distributions, which is consistent with deformation at relatively high temperatures (N1000 °C). Calculated seismic anisotropy values, based on the LPO patterns and modal compositions, indicate shear wave anisotropies of approximately 6%. This value is higher than those commonly calculated for mantle xenoliths from a variety of tectonic settings and higher than the 4% anisotropy used to calibrate many shear wave splitting studies. In order to impart the strong LPO patterns onto the mantle, the Calaveras fault probably extends to at least 40 km into the lithospheric mantle as a shear zone. To explain the observed shear wave splitting delay times in a two-layer model requires 40–95 km of anisotropic upper mantle, assuming 6% shear wave anisotropy and a vertical foliation (c-axes vertical) with a horizontal lineation (a-axes parallel to the fault). This orientation is consistent with the kinematics in a primarily strike-slip plate boundary and agrees with lithospheric thickness estimates based on seismology and thermal modeling. The 130-km wide area in central California that displays fault-parallel shear wave splitting values is most consistent with distributed deformation in the mantle, in contrast to deformation restricted to narrow horizontal or vertical mantle shear zones. Acknowledgments We thank Clark Johnson for samples from the Coyote Lake basalt and Scott Giorgis for comments on an earlier version of this manuscript. The manuscript was improved by the comments of two anonymous reviewers. This material is based upon work supported under a National Science Foundation Graduate Research Fellowship (Titus), a Packard Fellowship (Tikoff) and NSF-EAR0337498 (Tikoff). References Abramson, E.H., Brown, J.M., Slutsky, L.J., Zaug, J., 1997. The elastic constants of San Carlos olivine to 17 GPa. Journal of Geophysical Research 102, 12,253–12,263. Armstrong, J.T., 1988. Quantitative analysis of silicate and oxide materials: comparison of Monte Carlo, ZAF, and ϕ(ρz) procedures. In: Newbury, D.E. (Ed.), Microbeam Analyses, Proceedings of the 23rd Annual Conference of the Microbeam Analysis Society. San Francisco Press, San Francisco, CA, pp. 239–246. Atwater, T., 1970. Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Geological Society of America Bulletin 81, 3513–3536. 18 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Barroul, R., Mainprice, D., 1993. A quantitative evaluation of the contribution of crustal rocks to the shear wave splitting of teleseismic SKS waves. Physics of the Earth and Planetary Interiors 78, 281–300. Ben-Ismail, W., Mainprice, D., 1998. An olivine fabric database: an overview of upper mantle fabrics and seismic anisotropy. Tectonophysics 296, 145–157. Ben-Ismail, W., Barruol, G., Mainprice, D., 2001. The Kaapvaal craton seismic anisotropy: petrophysical analyses of upper mantle kimberlite nodules. Geophysical Research Letters 28, 2497–2500. Bertrand, P., Mercier, J.-C.C., 1985. The mutual solubility of coexisting ortho- and clinopyroxene: toward an absolute geothermometer for the natural system? Earth Planetary Science Letters 76, 109–122. Boudier, F., Coleman, R.G., 1981. Cross section through the peridotite in the Samail ophiolite, southeastern Oman Mountains. Journal of Geophysical Research 86, 2573–2592. Bourne, S.J., England, P.C., Parsons, B., 1998. The motion of crustal blocks driven by flow of the lower lithosphere and implications for slip rates of continental strike-slip faults. Nature 391, 655–659. Boyd, F.R., 1973. A pyroxene geotherm. Geochemica et Cosmochimica Acta 37, 2533–2546. Brey, G.P., Köhler, T., 1990. Geothermobarometry in four-phase lherzolites: II. New thermobarometers and practical assessment of existing thermobarometry. Journal of Petrology 31, 1352–1378. Brocher, T.M., McCarthy, J., Hart, P.E., Holbrook, W.S., Furlong, K.P., McEvilly, T.V., Hole, J.A., Klemperer, S.L., 1994. Seismic evidence for a lower-crustal detachment beneath San Francisco Bay, California. Science 265, 1436–1439. Bürgmann, R., 1997. Active detachment faulting in the San Francisco Bay area? Geology 25, 1135–1138. Carter, N.L., Avé Lallement, H.G., 1970. High temperature flow of dunite and peridotite. Geological Society of America Bulletin 81, 2181–2202. Christensen, N.I., 2002. Continental mantle seismic anisotropy: a new look at the Twin Sisters massif. Tectonophysics 355, 163–170. Christensen, N.I., Lundquist, S.M., 1982. Pyroxene orientation within the upper mantle. Geological Society of America Bulletin 93, 279–288. Christensen, N.I., Medaris Jr., L.G., Wang, H.F., Jelínek, E., 2001. Depth variation of seismic anisotropy and petrology in central European lithosphere: a tectonothermal synthesis from spinel lherzolite. Journal of Geophysical Research 106, 645–664. Collins, M.D., Brown, J.M., 1998. Elasticity of an upper mantle clinopyroxene. Physics and Chemistry of Minerals 26, 7–13. Crossen, R.S., Lin, J.W., 1971. Voight and Reuss prediction of anisotropy elasticity of dunite. Journal of Geophysical Research 79, 570–578. Dickinson, W.R., 1996. Kinematics of transrotational tectonism in the California transverse ranges and its contribution to cumulative slip along the San Andreas transform fault system. Geological Society of America Special Paper, vol. 305. 46 pp. Dickinson, W.R., 1997. Tectonic implications of Cenozoic volcanism in coastal California. Geological Society of America Bulletin 109, 936–954. Dickinson, W.R., Snyder, W.S., 1979. Geometry of subducted slabs related to San Andreas transform. Journal of Geology 87, 609–627. Duffy, T.S., Vaughan, M.T., 1998. Elasticity of enstatite and its relationship to crystal structure. Journal of Geophysical Research 93, 383–391. Emmons, R.C., 1943. The Universal Stage. Memoirs of the Geological Society of America, vol. 8. 205 pp. Fox Jr., K.F., Fleck, R.J., Curtis, G.H., Meyer, C.E., 1985. Implications of the northwestwardly younger age of the volcanic rocks of west- central California. Geological Society of America Bulletin 96, 647–654. Fuis, G.S., Mooney, W.D., 1990. Lithospheric structure and tectonics from seismic-refraction and other data. In: Wallace, R.E. (Ed.), The San Andreas Fault System, California. U.S. Geological Survey Professional Paper, vol. 1515, pp. 207–236. Furlong, K.P., 1984. In: Neugebauer, H.J., Olson, P. (Eds.), Lithospheric behavior with triple junction migration; an example based on the Mendocino triple junction: In Quantitative Geodynamics; Lithospheric Processes, Convection and Global Dynamics. Physics of the Earth and Planetary Interiors, vol. 36, pp. 213–223. Furlong, K.P., Hugo, W.D., Zandt, G., 1989. The geometry and evolution of the San Andreas fault zone in northern California. Journal of Geophysical Research 94, 3100–3110. Hartog, R., Schwartz, S.Y., 2000. Subduction-induced strain in the upper mantle east of the Mendocino triple junction, California. Journal of Geophysical Research 105, 7909–7930. Hartog, R., Schwartz, S.Y., 2001. Depth-dependent mantle anisotropy below the San Andreas fault system; apparent splitting parameters and waveforms. Journal of Geophysical Research 106, 4155–4167. Hearn, T.M., 1996. Anisotropic Pn tomography in the western United States. Journal of Geophysical Research 101, 8403–8414. Hellebrand, E., Snow, J.E., Dick, H.J.B., Hofmann, A.W., 2001. Coupled major and trace elements as indicators of the extent of melting in mid-ocean-ridge peridotites. Nature 410, 677–681. Henstock, T.J., Levander, A.R., Hole, J.A., 1997. Deformation in the lower crust of the San Andreas fault system in northern California. Science 278, 650–653. Hill, D.P., 1989. Seismic evidence for the structure and Cenozoic tectonics of the Pacific Coast states. In: Smith, R.B., Eaton, G.P. (Eds.), Cenozoic Tectonics and Regional Geophysics of the Western Cordillera. Geological Society of America Memoir, vol. 152, pp. 145–174. Holbrook, W.S., Brocher, T.M., ten Brink, U.S., Hole, J.A., 1996. Crustal structure of a transform plate boundary: San Francisco Bay and the central California continental margin. Journal of Geophysical Research 101, 22,311–22,334. Hole, J.A., Beaudoin, B.C., Klemperer, S.L., 2000. Vertical extent of the newborn San Andreas fault at the Mendocino triple junction. Geology 28, 1111–1114. Ji, S., Zhao, X., Francis, C., 1994. Calibration of shear-wave splitting in the subcontinental upper mantle beneath active orogenic belts using ultramafic xenoliths from the Canadian Cordillera and Alaska. Tectonophysics 239, 1–27. Johnson, C.M., O'Neil, J.R., 1984. Triple junction magmatism: a geochemical study of Neogene volcanic rocks in western California. Earth and Planetary Science Letters 71, 241–262. Jones, D.L., Graymer, R., Wang, C., McEvilly, T.V., Lomax, A., 1994. Neogene transpressive evolution of the California Coast Ranges. Tectonics 13, 561–574. Jové, C.F., Coleman, R.G., 1998. Extension and mantle upwelling within the San Andreas fault zone, San Francisco Bay area, California. Tectonics 17, 883–890. Katayama, I., Karato, S.-I., Brandon, M., 2005. Evidence of high water content in the deep upper mantle inferred from deformation microstructures. Geology 33, 1045–1048. Kennedy, L.A., Russel, J.K., Kopylova, M.G., 2002. Mantle shear zones revisited: the connection between the cratons and mantle dynamics. Geology 30, 419–422. Kern, H., Burlini, L., Ashcehpkov, I.V., 1996. Fabric-related seismic anisotropy in upper-mantle xenoliths: evidence from measurements S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 and calculations. Physics of the Earth and Planetary Interiors 95, 195–209. Kobussen, A.F., 2005. Xenolith and petrophysical perspectives of continental mantle seismic anisotropy: case studies in Syria and Siberia. [M.S. thesis] University of Wisconsin-Madison, 87 p. Köhler, T.P., Brey, G.P., 1990. Calcium exchange between olivine and clinopyroxene calibrated as a geothermobarometer for natural peridotites from 2 to 60 kb with applications. Geochemica et Cosmochimica Acta 31, 2375–2388. Lachenbruch, A.H., Sass, J.H., 1980. Heat flow and energetics of the San Andreas fault zone. U.S. Geological Survey Open file Report, vol. 80–625, pp. 47–146. Lamb, S., 1994. Behavior of the brittle crust in wide plate boundary zones. Journal of Geophysical Research 99, 4457–4483. Long, C., Christensen, N.I., 2000. Seismic anisotropy of South African upper mantle xenoliths. Earth and Planetary Science Letters 179, 551–565. Mainprice, D., 1990. An efficient Fortran program to calculate seismic anisotropy from the lattice preferred orientation of minerals. Computers & Geosciences 16, 385–393. Mainprice, D., Humbert, M., 1994. Methods of calculating petrophysical properties from lattice preferred orientation data. Surveys in Geophysics 15, 575–592. Mainprice, D., Silver, P.G., 1993. Interpretation of SKS-waves using samples from the subcontinental lithosphere. Physics of the Earth and Planetary Interiors 78, 257–280. McKenzie, D., 1979. Finite deformation during fluid flow. Geophysical Journal of the Royal Astronomical Society 58, 689–715. McKenzie, D., Jackson, J., 1983. The relationship between strain rates, crustal thickening, paleomagnetism, finite strain and fault movements within a deforming zone. Earth and Planetary Science Letters 65, 182–202. McLaughlin, R.J., Sliter, W.V., Sorg, D.H., Russell, P.C., Sarna-Wojcicki, A.M., 1996. Large-scale right-slip displacement on the East San Francisco Bay region fault system, California: implications for location of late Miocene to Pliocene Pacific plate boundary. Tectonics 15, 1–18. Medaris Jr., L.G., Wang, H.F., Fournelle, J.F., Zimmer, J.H., Jelínek, E., 1999. A cautionary tale of spinel peridotite thermobarometry: an example from xenoliths of Kozákov Volcano, Czech Republic. Geolines 9, 92–96. Mizukami, T., Wallis, S.R., 2005. Structural and petrological constraints on the tecotnic evolution of the garnet-lherzolite facies Higashi-Akaishi peridotite body, Sanbagawa Belt, SW Japan. Tectonics 24, 17 pp. Molnar, P., Anderson, H.J., Audoine, E., Eberhart-Phillips, D., Gledhill, K.R., Klosko, E.R., McEvilly, T.V., Okaya, D., Savage, M.K., Stern, T., Wu, F.T., 1999. Continuous deformation versus faulting through the continental lithosphere of New Zealand. Science 286, 516–519. Nakata, J.K., 1977. Distribution and petrology of the AndersonCoyote Reservoir volcanic rocks, [M.S. thesis] San Jose State University, 105 p. Nakata, J.K., Sorg, D.H., Russell, P.C., Meyer, C.E., Wooden, J., Lanphere, M.A., McLaughlin, R.J., Sarna-Wojcicki, A.M., Saburomaru, J.Y., Pringle, M.S., Drinkwater, J., 1993. New radiometric ages and tephra correlations from the San Jose and the northeastern part of the Monterey 1:100,000 map quadrangles, California. Isochron/West 60, 19–32. Namson, J.S., Davis, T.L., 1988. Structural transect of the western Transverse Ranges, California: implications for lithospheric kinematics and seismic risk evaluation. Geology 16, 675–679. 19 Nicolas, A., Christensen, N., 1987. Formation of anisotropy in upper mantle peridotites—a review. In: Fuchs, K., Froideveaux, C. (Eds.), Composition, Structure and Dynamics of the Ltihosphere– Asthenosphere System. American Geophysical Union Geodynamic Series, vol. 16, pp. 111–123. Nicolas, A., Poirier, J.P., 1976. Crystalline Plasticity and Solid Sate Flow in Metamorphic Rocks. John Wiley, London. 444 pp. Nicolas, A., Boudier, F., Boullier, A.M., 1973. Mechanisms of flow in naturally and experimentally deformed peridotites. American Journal of Science 273, 853–876. O'Neill, H.St.C., 1981. The transition between spinel lherzolite and garnet lherzolite, and its use as a geobarometer. Contributions to Mineralogy and Petrology 77, 185–194. O'Reilly, S.Y., Chen, D., Griffin, W.L., Ryan, C.G., 1997. Minor elements in olivine from spinel lherzolite xenoliths: implications for thermobarometry. Mineralogical Magazine 61, 257–269. Ozalaybey, S., Savage, M.K., 1994. Double-layer anisotropy resolved from S phases. Geophysical Journal International 177, 653–664. Ozalaybey, S., Savage, M.K., 1995. Shear wave splitting beneath western United States in relation to plate tectonics. Journal of Geophysical Research 100, 18135–18149. Parsons, T., Hart, P.E., 1999. Dipping San Andreas and Hayward faults revealed beneath San Francisco Bay, California. Geology 27, 839–842. Passchier, C.W., Trouw, R.A.J., 1998. Microtectonics. Springer, Berlin. 289 pp. Pera, E., Mainprice, D., Burlini, L., 2003. Anisotropic seismic properties of the upper mantle beneath the Torre Alfina area (Northern Apennines, central Italy). Tectonophysics 370, 11–30. Polet, J., Kanamori, H., 2002. Anisotropy beneath California: shear wave splitting measurements using a dense broadband array. Geophysical Journal International 149, 313–327. Prescott, W.H., Nur, A., 1981. The accommodation of relative motion at depth on the San Andreas fault system in California. Journal of Geophysical Research 86, 999–1004. Ribe, N.M., 1992. On the relation between seismic anisotropy and finite strain. Journal of Geophysical Research 97, 8737–8747. Salisbury, M.H., Christensen, N.I., 1985. Olivine fabrics in the Bay of Islands Ophiolite: implications for oceanic mantle structure and anisotropy. Canadian Journal of Earth Sciences 22, 1757–1766. Saruwatari, K., Ji, S., Long, C., Salisbury, M.H., 2001. Seismic anisotropy of mantle xenoliths and constraints on upper mantle structure beneath the southern Canadian Cordillera. Tectonophysics 339, 403–426. Savage, J., Burford, R., 1970. Accumulation of tectonic strain in California. Bulletin of the Seismological Society of America 60, 1877–1896. Severinghaus, J., Atwater, T., 1990. Cenozoic geometry and thermal state of the subducting slabs beneath North America. In: Wernicke, B.P. (Ed.), Basin and Range Extensional Tectonics Near the Latitude of Las Vegas, Nevada. Geological Society of America Memoir, vol. 176, pp. 1–22. Silver, P.G., 1996. Seismic anisotropy beneath the continents; probing the depths of geology. Annual Review of Earth and Planetary Sciences 24, 385–432. Silver, P.G., Chan, W.W., 1991. Shear wave splitting and subcontinental mantle deformation. Journal of Geophysical Research 96, 16429–16454. Silver, P.G., Holt, W.E., 2002. The mantle flow field beneath western North America. Science 295, 1054–1058. 20 S.J. Titus et al. / Tectonophysics 429 (2007) 1–20 Silver, P.G., Savage, M.K., 1994. The interpretation of shear-wave splitting parameters in the presence of two anisotropic layers. Geophysical Journal International 119, 949–963. Soedjatmiko, B., Crhistensen, N.I., 2000. Seismic anisotropy under extended crust: evidence from upper mantle xenoliths, Cima Volcanic Field, California. Tectonophysics 321, 279–296. Tapponnier, P., Peltzer, G., Ledain, A.Y., Armijo, R., Cobbold, P., 1982. Propagating extrusion tectonics in Asia—new insights from simple experiments with plasticine. Geology 10, 611–616. Taylor, W.R., 1998. An experimental test of some geothermometer and geobarometer formulations for upper mantle peridotites with application to the thermobarometry of fertile lherzolite and garnet websterite. Neues Jahrbuch fur Mineralogie Abhandlungen 172, 381–408. Teyssier, C., Tikoff, B., 1998. Strike-slip partitioned transpression of the San Andreas fault system: a lithospheric-scale approach. Geological Society Special Publications 135, 143–158. Thatcher, W., 1989. Systematic inversion of geodetic data in central California. Journal of Geophysical Research 84, 2283–2295. Thorkelson, D.J., Taylor, R.P., 1989. Cordilleran slab windows. Geology 17, 833–836. Tikoff, B., Fossen, H., 1993. Simultaneous pure and simple shear; the unifying deformation matrix. Tectonophysics 217, 267–283. Tommasi, A., Tikoff, B., Vauchez, A., 1999. Upper mantle tectonics: three-dimensional deformation, olivine crystallographic fabrics and seismic properties. Earth and Planetary Science Letters 168, 173–186. Tommasi, A., Mainprice, D., Gilles, C., Chastal, Y., 2000. Viscoplastic self-consistent and equilibrium-based modeling of olivine lattice preferred orientations; implications for the upper mantle seismic anisotropy. Journal of Geophysical Research 105, 7893–7908. Tommasi, A., Godard, M., Coromina, G., Dautria, J.-M., Barsczus, H., 2004. Seismic anisotropy and compositionally induced velocity anomalies in the lithosphere above mantle plumes: a petrological and microstructural study of mantle xenoliths from French Polynesia. Earth and Planetary Science Letters 227, 539–556. Vauchez, A., Dineur, F., Rudnick, R., 2005. Microstructure, texture and seismic anisotropy of the lithospheric mangle above a mantle plume: Insights from the Labait volcano xenoliths (Tanzania). Earth and Planetary Science Letters 232, 295–314. Vinnik, L.P., Farra, V., Romanowicz, B., 1989. Aziumthal anisotropy in the Earth from observations of SKS at GEOSCOPE and NARS broadband stations. Bulletin of the Seismological Society of America 79, 1542–1558. Wells, R.A., 1977. Pyroxene thermometry in simple and complex systems. Contributions to Mineralogy and Petrology 62, 129–139. Wilshire, H.G., Shervais, J.W., 1975. Al-augite and Cr-diopside ultramafic xenoliths in basaltic rocks from western United States. Physics and Chemistry of the Earth 9, 257–272. Wilson, C.K., Jones, C.H., Molnar, P., Sheehan, A.F., Boyd, O.S., 2004. Distributed deformation in the lower crust and upper mantle beneath a continental strike-slip fault zone: Marlborough fault system, South Island, New Zealand. Geology 32, 837–840. Witt-Eickschen, G., Seck, H.A., 1991. Solubility of Ca and Al in orthopyroxene from spinel peridotite: an improved version of an empirical geothermometer. Contributions to Mineralogy and Petrology 106, 431–439. Zandt, G., 1981. Seismic images of the deep structures of the San Andreas fault system, central Coast Ranges, California. Journal of Geophysical Research 86, 5039–5052. Zandt, G., Furlong, K.P., 1982. Evolution and thickness of the lithosphere beneath coastal California. Geology 10, 376–381. Zhang, S., Karato, W.I., 1995. Lattice preferred orientation of olivine aggregates deformed in simple shear. Nature 375, 774–777. Zhang, S., Karato, S.-I., Fitz Gerald, F., Faul, U.H., Zhou, Y., 2000. Simple shear deformation of olivine aggregates. Tectonophysics 316, 133–152. Zhu, L.P., 2000. Crustal structure across the San Andreas fault, southern California, from teleseismic converted waves. Earth and Planetary Science Letters 179, 183–190.
© Copyright 2026 Paperzz