Frictional properties of low-angle normal fault gouges and
Transcription
Frictional properties of low-angle normal fault gouges and
Earth and Planetary Science Letters 408 (2014) 57–65 Contents lists available at ScienceDirect Earth and Planetary Science Letters www.elsevier.com/locate/epsl Frictional properties of low-angle normal fault gouges and implications for low-angle normal fault slip Samuel Haines a,b,∗ , Chris Marone a , Demian Saffer a a b Department of Geosciences, Pennsylvania State University, 522 Deike Building, University Park, PA 16802, USA Chevron ETC, 1500 Louisiana St., Houston, TX 77002, USA a r t i c l e i n f o Article history: Received 27 June 2014 Received in revised form 19 September 2014 Accepted 20 September 2014 Available online xxxx Editor: P. Shearer Keywords: rock friction clay gouge authigenic gouge low-angle normal faults Cordillera a b s t r a c t The mechanics of slip on low-angle normal faults (LANFs) remain an enduring problem in structural geology and fault mechanics. In most cases, new faults should form rather than having slip occur on LANFs, assuming values of fault friction consistent with Byerlee’s Law. We present results of laboratory measurements on the frictional properties of natural clay-rich gouges from low-angle normal faults (LANF) in the American Cordillera, from the Whipple Mts. Detachment, the Panamint range-front detachment, and the Waterman Hills detachment. These clay-rich gouges are dominated by neoformed clay minerals and are an integral part of fault zones in many LANFs, yet their frictional properties under in situ conditions remain relatively unknown. We conducted measurements under saturated and controlled pore pressure conditions at effective normal stresses ranging from 20 to 60 MPa (corresponding to depths of 0.9–2.9 km), on both powdered and intact wafers of fault rock. For the Whipple Mountains detachment, friction coefficient (μ) varies depending on clast content, with values ranging from 0.40 to 0.58 for clast-rich material, and 0.29–0.30 for clay-rich gouge. Samples from the Panamint range-front detachment were clay-rich, and exhibit friction values of 0.28 to 0.38, significantly lower than reported from previous studies on fault gouges tested under room humidity (nominally dry) conditions, including samples from the same exposure. Samples from the Waterman Hills detachment are slightly stronger, with μ ranging from 0.38 to 0.43. The neoformed gouge materials from all three localities exhibits velocity-strengthening frictional behavior under almost all of the experimental conditions we explored, with values of the friction rate parameter (a − b) ranging from −0.001 to +0.025. Clast-rich samples exhibited frictional healing (strength increases with hold time), whereas clay-rich samples do not. Our results indicate that where clay-rich neoformed gouges are present along LANFs, they provide a mechanically viable explanation for slip on faults with dips <20◦ , requiring only moderate (P f < σ3 ) overpressures and/or correcting for ∼5◦ of footwall tilting. Furthermore, the low rates of frictional strength recovery and velocity-strengthening frictional behavior we observe provide an explanation for the lack of observed seismicity on these structures. We suggest that LANFs in the upper crust (depth <8 km) slip via a combination of a) reaction-weakening of initially high-angle fault zones by the formation of neoformed clay-rich gouges, and b) regional tectonic accommodation of rotating fault blocks. © 2014 Elsevier B.V. All rights reserved. 1. Introduction Normal faults with low dips (<30◦ ) have been recognized in the field for over a century (e.g. Ransom et al., 1910; Longwell, 1945), but the shallow observed dips stand in apparent contradiction with two fundamental predictions from rock mechanics. For typical values of rock internal friction (∼0.6–0.8), nor- * Corresponding author at: Department of Geosciences, Pennsylvania State University, 522 Deike Building, University Park, PA 16802, USA. E-mail addresses: shhaines@umich.edu, faultgougeguy@gmail.com (S. Haines). http://dx.doi.org/10.1016/j.epsl.2014.09.034 0012-821X/© 2014 Elsevier B.V. All rights reserved. mal faults should not form at dips <∼50◦ ; for typical values of sliding friction (μ = 0.6) those faults should not then slip at dips much below 30◦ . The crust should fail instead by formation of new higher-angle faults (e.g., Anderson, 1951; Byerlee, 1978; Collettini and Sibson, 2001) (Fig. 1). Several explanations for LANF slip have been offered. One set of explanations argues that the low dips observed in the field are caused by post-deformational rotation of the fault plane, either by the passage of a ‘rolling hinge’ (Wernicke and Axen, 1988; Buck, 1988), or by late-stage ‘domino’-style’ rotation of normal fault blocks (e.g. Wong and Gans, 2008). A separate set 58 S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 Fig. 1. A) Map of the southwestern US showing metamorphic core complexes with low-angle normal faults (LANFs). Core complexes with LANFs highlighted in red are those sampled in this study: PM: Panamint range-front LANF, WM: Whipple Mountains Detachment, WH: Waterman Hills LANF. B) Panamint Detachment outcrop; this is the same outcrop sampled by Numelin et al. (2007). Footwall is Neoproterozoic schist, hanging wall is weakly consolidated conglomerate. Clay gouge zone is 0.5–2.0 m thick. Inset is clay-rich wafer used in our experiments. C) WM: Whipple Mountains detachment fault exposed at Bowman’s Wash. Hanging wall is quartz-cemented quartzite breccia (interpreted megabreccia deposits – Forshee and Yin, 1995). Footwall is Cretaceous granite with characteristic chlorite/epidote ‘microbreccia’ texture and alteration. Clay gouge zone is 0.5–2.0 m thick. D) Waterman Hills detachment fault. Hanging wall is Miocene rhyolite, footwall is Miocene granodiorite (Fletcher and Bartley, 1994). Clay gouge zone is 1.0–2.0 m thick. Grey material is smectite-dominated, brown is illite-dominated. Illite-dominated material was sampled. E) Field photo of Whipple Moutain Detachment at Bowman’s Wash, showing clay-rich gouge horizons we sampled (white arrows). F) Whipple Mountain Detachment at Bowman’s Wash, showing clast-rich gouge (inset is a wafer used in our experiments). Scale in insets in 1B and 1F are the same, upper scale units are in 10’s of mm. Photo annotations: HW: Hanging wall. FW: Footwall. G: Gouge. of explanations argues for the presence of unusual conditions within the fault core itself, through either: (a) rotation of principal stresses in the immediate vicinity of the fault core (Chery, 2002; Lecomte et al., 2012), (b) locally elevated pore-fluid pressures in the fault zone (Rice, 1992; Axen, 1992), or (c) the presence of low-friction materials in the fault core (Hayman et al., 2003; Numelin et al., 2007; Collettini et al., 2009a, 2009b). Many LANF in the American Cordillera contain well-developed clay-rich gouges, ranging in thickness from cm’s to >10 m (Haines and van der Pluijm, 2012, Fig. 1). These gouges are dominated by authigenic clay minerals (illite, illite–smectite, smectite and chlorite–smectite phases) formed at temperatures of 60–180 ◦ C. Depth constraints on the formation of the neoformed clay gouges we sampled from three well-studied detachments for this work (the Whipple Mountains, Panamint range-front, and Waterman Hills detachments) are harder to establish, given uncertainties in heat flow and temperature gradients within exhuming LANF footwalls, but range from <2 to >8 km (Haines and van der Pluijm, 2012). The neoformation of clay-rich fault gouge, a lowtemperature metasomatic process confined to fault zones, thus has a potential impact on the frictional strength of LANFs in the upper crust, where low-angle slip is mechanically most problematic. Although reaction-softening has been discussed as a possible explanation for LANF slip (e.g. Manatschal, 1999), the actual frictional properties of these neoformed clay-rich gouges and their implications for LANF slip remain poorly characterized. Frictional properties of natural fault zones are commonly extrapolated or estimated from laboratory tests on synthetic gouges of similar composition (e.g. crushed illite shale as a proxy for illitic gouges; Brown et al., 2003; Ikari et al., 2009). Work comparing frictional properties of chlorite separated from natural fault gouge with crushed chlorite schist has highlighted dramatic differences in both the frictional strength and stability between phyllosilicate materials that have identical compositions, but very different parti- S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 59 Fig. 2. Hand specimen and SEM images of gouge materials sampled. A) Hand sample of gouge from the Panamint Detachment. B) Hand sample of clay-rich gouge from the Whipple Detachment. Light irregular lines on surface are cracks in epoxy coating surface and not vein mineralization. C) Hand sample of grey illitic gouge from the Waterman Hills Detachment. D) Wafers of clast-rich gouge from the Whipple Detachment. Compare with B. Wafers are slightly undersized for testing and were not used. Upper scale is in millimeters. E) SEM image of gouge from the Panamint Detachment. Note scaly fabric in the ‘P’ orientation of Logan et al. (1979) and narrow through-going polished Riedel surface. Sense of shear is top to the left. F) SEM image of gouge from the Waterman Hills Detachment. Note polished and striated Riedel surface. Sense of shear is shown surface to bottom of image (missing other side of Riedel surface to top). cle morphologies (Haines et al., 2013). While tests of natural fault rocks are now common (e.g. Tembe et al., 2009; Collettini et al., 2009a, 2009b; Carpenter et al., 2011), LANF gouges remain understudied relative to similarly mechanically problematic strike-slip systems such as the San Andreas and Alpine Fault (e.g. Tembe et al., 2009; Boulton et al., 2012) and seismogenic thrust-fault systems such as the Nankai Trough and Longman Shan (e.g. Verbane et al., 2010; Ikari and Saffer, 2011). Fault rocks from the Zuccale Fault, an exhumed LANF in central Italy, have been extensively studied (Collettini et al., 2009a, 2009b; Smith and Faulkner, 2010; Tesei et al., 2012), but the Zuccale LANF fault rocks are dominated by talc and chlorite-dominated phyllonites mineralogically and structurally distinct from the illite- and smectite-dominated clayey gouges common to the LANFs of the American Cordillera (Haines and van der Pluijm, 2012). Numelin et al. (2007) measured frictional strengths and stability of gouges from the Panamint Detachment, but worked exclusively with dry materials that were powdered and sieved. Their experimental conditions leave unresolved questions about the in situ frictional properties of these fault gouges, given the importance of fluid saturation, pore pressure and effective stress, and fault zone fabric in governing the frictional behavior of clay-rich gouge (e.g., Ikari et al., 2009, 2011; Collettini et al., 2009a, 2009b). A need thus exists to measure frictional properties of both intact and powdered LANF natural gouges under saturated and controlled pore pressure conditions. Here, we report on shearing experiments designed to measure the frictional properties of natural LANF neoformed clay-rich fault gouges. Specifically, we (1) measure the frictional strength of natural gouges, using intact gouge wafers where possible; (2) measure rate/state friction constitutive properties including the rate of frictional healing and the velocity dependence of sliding friction to address the question of whether LANF seismicity is expected; and (3) discuss these results in the context of LANF evolution, geometry, and seismicity. 2. Experimental samples and methods We collected natural gouges from exposures of three LANF detachment faults in SE California: the Whipple Detachment, the Panamint Detachment, and the Waterman Hills Detachment of the Central Mojave Metamorphic Core Complex (CMMCC – Fig. 1). The Panamint Detachment lacks a mylonitic footwall and was apparently active entirely in the brittle regime (Cichanski, 2000) during the Miocene to Pleistocene (Andrew and Walker, 2009). We sampled the detachment in South Park Canyon (N36◦ 00 01.8 ), W117◦ 12 02.0 , NAD 1927) where it locally dips 18◦ to the west and juxtaposes likely Pliocene or latest Miocene weakly indurated conglomerates against a footwall of Proterozoic schists. This exposure is the same exposure as ‘Location A’ of Numelin et al. (2007) and was also characterized mineralogically by Haines and van der Pluijm (2012) (their sample “South Park, central gouge”). At this outcrop, clay gouge material is reddish brown or yellowish, is relatively free of clasts larger than pebble size, and is only very crudely foliated in hand specimen (Fig. 2A and inset in Fig. 1B). Visual inspection of clasts in the gouge indicate that they consist of partially metasomatized schist and granitic material, indicating the neoformed clays in the gouge are apparently derived from footwall lithologies (Haines and van der Pluijm, 2012). XRD analysis indicates that both reddish and yellowish materials are dominated by neoformed illite (specifically the disordered 1Md polytype) derived from the breakdown of potassium feldspar, along with catclastically-derived quartz and potassium feldspar, and minor gypsum (Haines and van der Pluijm, 2012). SEM images show a scaly fabric with thin polished Riedel shear surfaces. The likely temperature of gouge formation, based on the neoformed clay mineral assemblage in the gouge, is 60–160◦ C (Haines and van der Pluijm, 2012). The Whipple Detachment was active in the mid-Miocene (Davis, 1988). We sampled the detachment at Bowman’s Wash 60 S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 (N34◦ 17 15.3 , W114◦ 17 30.3 ) where it juxtaposes cemented monolithic quartzose breccia in the hanging wall and pervasively brecciated and epidote–chlorite metasomatized Cretaceous granite in the footwall (Forshee and Yin, 1995; Selverstone et al., 2012) The Bowman’s Wash locality contains both laterally continuous clay-rich gouge (Fig. 1E) and lenses of clast-rich material with clasts to pebble and locally boulder size (Figs. 1C, 1F, 1F inset, and Fig. 2). Gouge clasts are fresh sub-rounded to sub-angular granitic material, indicating that the gouge was derived from footwall lithologies, and not from the hanging wall quartzose monomict breccia. Notably, gouge clasts are fresher than the epidote-altered granitic material in the immediate footwall (Fig. 1C), indicating formation from metasomatism of materials located some distance away from the immediate outcrop sampled, presumably up-dip on the footwall. XRD analysis of the gouge indicates that both the clay-rich and clast-rich horizons (shown in Figs. 1E, 1F, and Figs. 2B, 2D) are dominated by neoformed illite (specifically the disordered 1Md polytype) derived from the breakdown of potassium feldspar, along with catclastically-derived quartz and potassium feldspar. The likely temperature of gouge formation, based on the neoformed clay mineral assemblage in the gouge, is 60–160 ◦ C (Haines and van der Pluijm, 2012). The epidote-rich metasomatism in the footwall (visible in Fig. 1C as a greenish layer in the immediate footwall) formed at significantly higher temperatures (380–420 ◦ C, Selverstone et al., 2012), at depths of ∼9 km and likely earlier in the fault’s history. Both clay-rich and clast-rich materials were sampled for this study. The clay-rich material was too friable to prepare wafers, so samples were disaggregated gently with a mortar and pestle, and then sieved to <106 μm. Clastrich material was cut into wafers ∼8–10 mm × 60 mm × 60 mm (Fig. 1F) for our tests. The Waterman Hills detachment fault was sampled in the northern Waterman Hills (Walker et al., 1990; Anderson, 2007) where it juxtaposes Tertiary rhyolites in the hanging wall and mylonitic granodiorite in the footwall (Fig. 1D). The detachment is a late brittle fault that is part of the ductile-to-brittle Central Mojave metamorphic core complex (Dokka, 1989; Walker et al., 1990). The temporal and kinematic relationship of the brittle fault to the mylonitic detachment is complex, in that kinematic indicators associated with the brittle fault record north–south extension as opposed to the northeast–southwest mid-Miocene extension recorded by the mylonites (Walker et al., 1990; Anderson, 2007). At outcrop, the gouge consists of grey illitic layers and brown smectite-dominated layers (Haines and van der Pluijm, 2012; Figs. 1D and 2C). We sampled grayish illitic material for this study, as it is laterally continuous throughout the outcrop. The illite in the Waterman Hills Detachment is also neoformed, of the disordered 1Md polytype, and derived from the breakdown of potassium feldspar, similar to the Panamint Detachment and Whipple Detachment localities, and also likely formed at temperatures of 60–160 ◦ C (Haines and van der Pluijm, 2012). Sample material was disaggregated gently with a mortar and pestle and sieved to <106 μm. Representative material from all 3 exposures was examined by scanning electron microscopy (SEM) to image fault rock microfabric, and material from the Panamint and Waterman Hills exposures was characterized by high-resolution X-ray texture goniometry (XTG) to quantify the degree of preferred orientation of phyllosilicates. Details of the XTG method are found in van der Pluijm et al. (1994). SEM images show a scaly fabric with welldeveloped polished and striated narrow (<10 μm) Riedel shear surfaces, and with phyllosilicates weakly aligned in the aligned in the P orientation (Figs. 2E and 2F). These fabrics are similar to those described in experimental clay-rich fault rocks (Haines et al., 2009, 2013), and the very thin nature of the Riedel surfaces is similar to those found in both natural and experimental fault gouges Fig. 3. Data for a representative shearing experiment run at an effective normal stress of 20 MPa. Friction (shear stress normalized by normal stress) is plotted vs. shear strain determined by normalizing shear displacement by layer thickness. A shear displacement of ∼16 mm corresponds to a shear strain of 10. Larger inset: Frictional healing (μ) plotted versus hold time for example clast-rich and clayrich samples from the Whipple Detachment. Lines indicate best-fit healing rates (μ/decade). Smaller inset: Example measurement of frictional healing parameter μ for a single slide–hold-slide test. and faulted mudrocks (Haines et al., 2013; Laurich et al., 2014). XTG analysis indicate that the degree of preferred orientation of phyllosilicates in the Panamint material is 2.3 and 3.1 multiples of a random distribution (MRD), and that material in the Waterman Hills exposure has a preferred orientation of 2.5 and 3.4 MRD units. These low values are very similar to those reported from clay rich fault gouges elsewhere in the Cordillera and globally (Haines et al., 2009). We conducted our experiments on intact wafers of gouge from the Panamint Detachment (clay-rich gouge) and the Whipple Detachment (clast-rich lenses), and on powdered gouges from all three localities. Sample conditions did not allow tests of intact wafers of clay-rich gouge from the Waterman Hills or Whipple Detachments. Gouge material was either disaggregated gently using a mortar and pestle to preserve natural particle morphology and sieved to <106 μm (for tests on powdered material), or cut into wafers ∼8–10 mm × 60 mm × 60 mm using an oil-cooled tile saw (for intact wafers). Wafers were cut parallel to the in situ fault-zone shear direction and fault zone fabric (Fig. 1B). The plasticity and saturated nature of the sheared materials from all sites prevented removal of intact samples post-test for fabric studies to compare with pre-experiment fabric (shown in Fig. 2). We sheared samples in a double-direct shear configuration inside a pressure vessel, under saturated and controlled pore pressure conditions at room temperature. Details of the experimental configuration are given in Samuelson et al. (2009) and Ikari et al. (2009). Samples were jacketed and saturated with distilled water at a pore fluid pressure of 5 MPa and a confining pressure of 6 MPa, prior to application of effective normal stresses ranging from 18 to 60 MPa. Samples were then sheared at 10 μm/s until reaching a steady-state friction value, followed by velocity-step tests to measure frictional rate dependence (a − b), and slide–holdslide tests to measure the rate of frictional healing (β ) (Fig. 3). We report friction values as the ratio of shear stress to normal stress (assuming no cohesion), taken at angular shear strains of 2–3, past the point at which the stress–strain curve ‘rolls over’ and at which (in dry materials sheared in the same apparatus) Riedel shears and a weak preferred orientation of phyllosilicates in the P orientation (of Logan et al., 1979; Marone, 1998) is developed (Haines et al., 2013). The change in friction coefficient with sliding velocity is described by the friction rate parameter (a − b) = μss / ln V , where μss is the steady-state friction coefficient and V is the sliding velocity (Fig. 3). Positive values of (a − b) reflect velocity-strengthening behavior (i.e. friction coefficient increases with increasing sliding velocity), whereas negative values of (a − b) reflect velocity-weakening behavior. The rate of frictional healing (β ) is given by μ/ log t h , based on the S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 61 Fig. 4. A) Coefficient of sliding friction for clay gouges from (A) Panamint Detachment; (B) Waterman Hills Detachment; and (C) Whipple Mts. Detachment. Center key applies to all panels. Panel (A) also shows data collected at room humidity from Numelin et al. (2007) for comparison, including results for powders comprised of <50% phyllosilicates (orange shaded area), sample A3 (57% phyllosilicates; gray triangles), and sample B6 (63% phyllosilicates; gray circles). Numelin ‘A3’ was taken from the same Panamint outcrop as we sampled. Note our wafers and powders are substantially weaker than reported by Numelin et al. (2007) for >50% clay gouges from same outcrop sheared at room humidity. ‘B6’ is different clay mineralogy than we studied (predominantly smectite) and from a different Panamint outcrop. Whipple Detachment samples are separated into clay-rich and clast-rich materials. Note Whipple clast-rich wafers and powders are 0.1 μ weaker than Numelin ‘<50% clay’ powders, although samples are compositionally similar. Fig. 5. A) Friction rate parameter (a − b) as function of sliding velocity for our complete dataset. No significant variation of (a − b) between outcrops was observed, so data from all localities are plotted together. B) Rate of frictional re-strengthening (β) (μ after hold −μ before hold) measured during slide–hold-slide tests plotted as a function of friction coefficient (μ). Key is the same as for Fig. 4A. change in measured peak friction (μ) before and after a hold of length t h (Fig. 3, inset). Positive healing rates indicate frictional restrengthening with hold time. Both velocity-weakening friction and positive healing rates are pre-requisites for initiation of failure in repeated stick-slip events (i.e. earthquakes) (e.g., Marone, 1998). 3. Frictional strength and properties of low-angle normal fault gouges The clay-rich gouges from all three detachment faults are frictionally weak, with steady-state friction values of μ = 0.28–0.43 (Fig. 3). In contrast, the clast-rich gouges are stronger, with friction coefficients of 0.43–0.56. In general, wafers and powders of the same material exhibit similar frictional strengths. The friction coefficient for wafers of clay-rich gouge from the Panamint Detachment was μ = 0.28–0.38, whereas μ for powders of the same material ranged from 0.29 to 0.36. Wafers of clast-rich gouge from the Whipple Detachment were slightly weaker (μ = 0.43–0.52) than powders of the same material (μ = 0.47–0.56). Powdered clay-rich gouge from the Waterman Hills Detachment was slightly stronger (μ = 0.38–0.44) than clay-rich gouges from the other two localities (Fig. 3). We observe no systematic trends in friction coefficient with effective normal stress for either wafers or powders (Figs. 4A–4C). Gouge materials from all three detachments exhibit velocitystrengthening behavior, with the friction rate parameter (a − b) ranging from +0.0002 to +0.0250 (Fig. 5A). Examination of the data reveals two distinct trends. First, (a − b) generally increases with upstep sliding velocity for all of the gouges (Fig. 5A). Second, clay-rich materials have slightly higher values of (a − b) at upstep velocities <100 μm/s than for clast-rich (clay-poor) materials (Fig. 5B). The parameter a − b does not vary significantly with normal stress, or between wafers and powders. Healing rates vary strongly with both clay content and frictional strength for all gouges. Values of β range from −0.039μ/decade (negative healing) for clay-rich powdered gouge from the Panamint Detachment, to +0.069μ/decade for clast-rich gouge from the Whipple Detachment. In general, the clay-rich gouges exhibit systematically lower healing rates than the clast-rich gouges; a clear trend of decreasing healing rate with lower frictional strength of the clay gouge is also evident in Fig. 4B. Overall, we do not observe a significant difference in healing between wafers and powders. Notably, all three detachments had at least 1 clay-rich gouge sample that was weak (μ < 0.4), velocity-strengthening, and characterized by negative healing rates (Fig. 4B), including wafers of gouge from the Panamint Detachment. Overall, we do not observe a significant difference in healing between wafers and powders. 4. Discussion The LANF gouges we tested are weak relative to Byerlee’s Law, and are comparable in frictional strength to previously studied natural and synthetic clay-rich gouges from other major fault systems (e.g. Carpenter et al., 2011; Ikari et al., 2011). Our observation of similar frictional strength between intact wafers and powdered material is consistent with previous studies of intact vs. powdered clay gouges (Ikari et al., 2009, 2011; Carpenter et al., 2011, 2012) and contrasts with large differences in frictional strength between powders and intact wafers in visibly foliated and mineralogically layered fault rocks (Collettini et al., 2009a, 2009b). Clay-rich fault gouges lack strong fabrics at outcrop and fabric intensity studies using X-ray texture goniometry and synchrotron X-rays both demonstrate that they are characterized by weak preferred orientation of phyllosilicates relative to visibly-foliated fault rocks (Haines et al., 2009; Buatier et al., 2011). Notably, our results suggest that in situ LANF gouges are significantly weaker than reported by Numelin et al. (2007) for similar, or in some cases, the same materials. This is consistent with previous work demonstrating lower frictional strengths for saturated versus nominally dry clay-rich gouges of the same material (e.g., Ikari et al., 2007; Behnsen and Faulkner, 2012). For example, at the effective normal stress conditions we investigated (17–60 MPa), Numelin et al., reported μ values ranging from μ = 0.48–0.64 for samples with <50% phyllosilicates, and for sample A3 (57% phyllosilicates, and from the same outcrop we sample) that drop systematically with increasing normal stress from μ = 0.65 at 5 MPa to μ = 0.42 at 70 MPa (Fig. 4A). Our results for saturated samples 62 S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 from the same Panamint outcrop indicate μ = 0.28–0.38, with no systematic decrease in frictional strength with increasing normal stress. The frictional properties of clay rich gouges, both natural and synthetic, change markedly as a function of temperature. Previous works show that clay gouges strengthen markedly and in some cases becoming frictionally unstable at temperatures >200 ◦ C (Tembe et al., 2009; Den Hartog et al., 2012a, 2012b). These results were obtained to understand the behavior of either: (1) subduction-zone faults, where clay-rich material is entrained in the subduction zone and transported to regions of higher temperatures and pressures; or (2) continental-scale strike-slip faults. In contrast to those types of fault systems, exposure of LANF gouges to higher temperatures and pressures after their formation is exceedingly unlikely, given the clear extensional nature of LANFs. The temperature at the time of neoformed LANF gouge formation is likely the highest temperature experienced by the gouge, and is constrained by the clay mineral assemblage to be 60–160 ◦ C. Although our results were obtained at room temperature, they are likely valid for LANF fault-zone conditions, as the gouges likely never experienced temperatures >200 ◦ C at which clay-rich gouges with similar illite-dominated mineralogies are reported to strengthen and become frictionally unstable as reported elsewhere (Den Hartog et al., 2012a, 2012b). To explore the implications of our lab data for the paradox of slip on shallowly-dipping (and thus severely misoriented) LANF, we define a relationship between the dip of faults on which slip would be permitted, and the effective horizontal stress (S, normalized to the overburden) based on simple re-arrangement of the Coulomb– Mohr failure criterion (after Abers, 2001): S= σ3 1 − μ cot θ = (1 − λ) ρ gz 1 + μ cot θ (1) where σ3 is the least principal effective stress, ρ is the mean bulk density of the overburden, g is gravitational acceleration, z is depth, λ is the ratio of pore fluid pressure to lithostatic stress, and θ is the dip of the detachment (Fig. 6C). The lower limit for effective stress labeled ‘hydrofracture’ in Fig. 5C is slightly below 0, reflecting an assumed rock tensile strength of 10 MPa and shown for depths of 2, 5 and 8 km, assuming a mean bulk density of 2300 kg m−3 for the overburden. Slip is permissible at dips as low as 16◦ (μ = 0.28) to 22◦ (μ = 0.43) (Fig. 6C) and at dips of >20◦ with overpressures similar to those found in sedimentary basins elsewhere (Figs. 6 and 7, Gaarenstroom et al., 1993; Converse et al., 2000). Alternatively, overpressure could be locally developed in the clay-rich and lowpermeability gouge layer itself and not require a regionally overpressured basin. Although slip is not permitted at hydrostatic pore fluid pressure for μ = 0.28, slip is possible at the range of dips we observe in the field for the Panamint Detachment with the introduction of only modest overpressure ( P f equivalent to a gradient of 11.3 MPa/km, vs. a hydrostatic gradient of 10.1 MPa). For the friction coefficients we measured, slip would be permitted on LANF with dips as low as ∼16◦ (μ = 0.28) to 22◦ (μ = 0.43) (Fig. 4) with higher overpressures (nearer lithostatic). For μ = 0.43, moderate to high overpressures (but still P f < σ3 ) are required for slip to be permitted (Figs. 6–7). Estimates of fluid pressures from field evidence in the vicinity of low-angle normal faults vary with crustal depth. For LANFs active in the upper few km of the crust, reliable estimates of pore pressure are sparse. By contrast, LANF fault rocks exhumed from the midcrustal brittle–plastic transition at temperatures of 250 to 420 ◦ C, several LANFs (the Woodlark Basin detachment, Zuccalle Fault, and the Whipple Detachment) exhibit compelling field evidence for overpressure, such as fault-parallel tensile veins (Roller et al., 2001; Collettini et al., 2006; Selverstone et al., 2012) Fig. 6. Mechanical analysis of measured friction results. A) Measurements of the modern dip angle of the Panamint range-front LANF (after Andrew, 2002). B) Mohr circle diagram for a stress state at failure consistent with our laboratory findings and a mean fault dip of 21◦ . Circles are for our measured minimum and maximum friction values. C) Minimum horizontal effective stress (σ3 − P f ) at failure normalized to the overburden (σ1 ; given by ρ gz), as a function of fault dip for frictional failure, assuming hydrostatic conditions (λ = 0.4) and σ3 = 0.7 × σ1 (after Abers, 2001). Coefficients of friction from our experimental results are shown in bold. Dots are Mohr circles shown in panel B. Shaded area below each curve of μ defines the range of mechanically viable fault dips and stress states that would allow slip at that value of μ. Pink area is mean fault dip shown in panel (A), showing ±1σ . Brown areas define fields of hydrofracture for surrounding rock with tensile strength of 10 MPa at depths of 2, 5 and 8 km. For comparison, the Panamint Detachment has dips ranging from 6◦ to 44◦ (mean = 21◦ , 1σ variation = 8.8◦ ) (Andrew, 2002, Fig. 4A), and sub-horizontal horizontal lacustrine strata in the hanging wall indicate that it slipped at or near its current dips. Previous studies of LANF dips indicate that the Panamint Detachment mean dip of 21◦ is representative of many LANFs (e.g., Collettini, 2011). The amount of footwall tilting in the vicinity of the Panammint Detachment is likely minor. We note that even for well-exposed and undisturbed detachments such as the Panamint Detachment, some footwall tilting (∼5◦ ) may have occurred that is not resolvable. With 5◦ of footwall tilting, slip is possible at any fluid pressure for μ = 0.28 and at moderate to high overpressures for μ = 0.43 (Fig. 7). S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 Fig. 7. Minimum horizontal effective stress ( S ) at failure normalized to the overburden stress as a function of fault dip as in Fig. 5C, showing conditions required for LANF slip at current Panamint Detachment dip. Stresses are calculated using Eq. (1) and friction data we report. Mean fault dip for Panamint Detachment ± 1 standard deviation shown in vertical transparent grey box. Effective stress ratios at failure are calculated holding μ and σ3 constant (for values of μ we measure, μ = 0.28 and μ = 0.43, and σ3 = 0.7 × σ1 ) and varying λ. Note, by contrast, in Fig. 5C we hold λ constant (λ = 0.44) and calculate states of stress for different values of μ. A) μ = 0.28 and varying λ, dipping at present day dips. Hydrostatic fracture window is same as 5 km window in Fig. 5C. B) μ = 0.43, varying λ, and dipping at present day dips. Note that with moderate or high (but sub-σ3 ) overpressures, slip is possible at the observed dips. The dip of the Waterman Hills detachment fault (present-day 10◦ ) is more difficult to establish at the time of slip, as field relations are significantly complicated by later high-angle normal faulting and strike-slip faulting associated with the onset of the San Andreas strike-slip system. An estimate based on apatite and zircon fission-track dating and assumptions of a geothermal gradient elsewhere in the greater Central Mojave metamorphic core complex put the dip at the time of slip to be 20–27◦ (Dokka, 1995), but is strongly sensitive to the choice of early Miocene geothermal gradient. For a postulated 20–27◦ dip at the time of slip, slip is similarly possible at μ = 0.38 to μ = 0.44 with moderate to high overpressures. The dip of the Whipple detachment is also difficult to establish at the time of slip. At our sampling locality, the detachment surface dips 15◦ to the SW, and is clearly back-rotated at outcrop, as the detachment has a clear overall top to the NE sense of displacement (Davis, 1988). The detachment has undergone significant late doming due to isostatic rebound owing to a probably ‘rolling-hinge’ evolution (Dorsey and Becker, 1995; Axen, 2004) but was likely never above 45◦ . In the least-disrupted northeastern portion of the detachment (10 km north of our sampling locality), the detachment dips at 12–28◦ (Yin and Dunn, 1992). For the clay-rich material at the Whipple detachment, slip 63 is possible with our measured friction values at the higher end of the observed dip range for most values of overpressure, but slip is impossible at the lowest dip with all but near-lithostatic values of overpressure (Figs. 6–7). Two explanations have been suggested to explain slip on the Whipple Detachment: high fluid pressures or low-friction gouge material (Axen, 2004). Our data lend support to the low-friction gouge material interpretation or a combination of low friction and moderately elevated pore pressures. As paleo-dip constraints for the Waterman Hill and Whipple Detachments are significantly poorer than for the Panamint Detachment, we do not treat them quantitatively in Fig. 7. Clay-rich gouges are found at multiple exposures of many Cordilleran LANFs (Haines and van der Pluijm, 2012). Our results indicate that the widespread occurrence of these mechanically weak fault zone materials, together with moderately-elevated pore fluid pressures and/or correcting for only minor footwall tilting, are a viable explanation for LANF slip at low dips. In particular, field relations at the Whipple Detachment require slip at low dips during the final period of activity on the detachment (Dorsey and Becker, 1995; Axen, 2004). The low frictional strength of LANF clay-rich gouge correlates with both velocity-strengthening behavior and near-zero healing rates, similar to both synthetic clay-rich gouges (Bos and Spiers, 2000) and those recovered from the SAFOD borehole (Carpenter et al., 2011, 2012). The negative healing rates we observe indicate that grain contacts in the clay-rich LANF gouges we studied do not strengthen with hold time, and together with the rate strengthening behavior we observe (a − b > 0), provide direct experimental evidence suggesting that LANFs should slip aseismically. One of the most compelling arguments against low-angle normal fault slip has been the absence of an unequivocal focal mechanism for a low-angle slip event (Wernicke, 1995), although isolated examples have been reported (Abers, 2001; Chairaluce et al., 2007). The most compelling modern example is the Alto Tibernia Fault in the Italian Appenines, which has both documented microseismicity (Chairaluce et al., 2007) and apparent aseismic creep apparent from GPS measurements (Hreinsdóttir and Bennett, 2009). Our data suggest that an absence of LANF focal mechanisms for large earthquakes should not be considered evidence of absence of LANF slip. We note that the velocity-strengthening behavior of the clay gouges we sampled indicates that rupture nucleation in clay-rich gouges is likely impossible, but does not preclude the possibility of permitting propagation of a rupture initiated elsewhere on the detachment surface. Our data thus indicate that if LANF faults are capable of seismicity, slip events must nucleate at depths greater than those at which clay-rich gouges form, and suggest that the depth of rooting of a low-angle normal fault may control its propensity for seismicity and shallow coseismic slip. Another key issue in the debate surrounding low-angle normal faults has been the mechanical problem of initiating a normal fault at low dips (e.g. Axen, 2004). While our data do not shed light directly on the problem of LANF initiation, neoformation of frictionally-weak clay-rich gouges offers a potential solution to the problem if the fault initiated at a higher dip. Neoformation of clay-rich fault gouge can be thought of as a low-temperature metasomatic process confined to fault zones. Our data indicate that clay gouge formation, a process that can be dated directly to occur during periods of fault activity (Haines and van der Pluijm, 2008, 2010, Duvall et al., 2011), is a process that produces materials in LANF fault cores with low frictional strength, velocity-strengthening behavior and little to no healing. These materials, once formed, can act to weaken a pre-existing fault surface, and thus permit it to slip at dips lower than it otherwise could. Our data are consistent with field observations indicating that some low-angle normal faults initially formed at high angles (∼60◦ ) and later rotated to their current low dips (e.g. 64 S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 Wong and Gans, 2008). Faults that are metasomatically weakened by the growth of neoformed clay could thus continue to slip to the low dips observed, as long as larger-scale crustal extensional processes continue to create accommodation space for the ends of the ‘domino blocks’ to rotate. For faults that initiated at anomalously low dips (at or <45◦ ), the formation of weak fault zone materials by clay metasomatism will still permit slip at dips significantly lower than would be permitted by typically assumed frictional values of μ = ∼0.6. Cordilleran LANF faults are observed to ‘cluster’ in space and time, suggesting larger-scale crustal processes, such as gravitational collapse of thick lithosphere triggered by thermal weakening of subducted oceanic lithosphere and passage of the Mendocino fracture zone, play a role in their formation (Axen et al., 1993). Examples include the Miocene–Pliocene ‘Turtleback’ LANFs in the Death Valley area and the Panamint Range in CA (Badwater, Mormon Point, and Copper Canyon Turtlebacks, Amargosa Detachment, and Panamint Detachment), and the Middle Miocene Colorado River Extensional Corridor LANFs, (Sacramento Detachment, Chemehuevi Detachment, Whipple Detachment, Buckskin–Rawhide Detachment, Harcuvar and Harquahala Detachments). Older examples include the Oligo-Miocene LANFs in eastern Nevada (RubyHumboldt Mts LANF, North and South Snake Range Detachments, Schell Creek Detachment, Albion–Grouse Creek and Raft River detachments) and the Eocene-age belt of metamorphic core complexes of western Canada and the NW United States (Valhalla MCC, Kettle Dome, Shushwap MCC, and Bitterroot Detachment). We suggest that LANFs may be the result of a combination of widespread reaction-softening of the fault surface by formation of clay-rich gouges during fault evolution together with larger-scale crustal processes that create new accommodation space for fault block rotation. LANF formation and slip may thus be the result of a confluence of distinct processes, one metasomatic, the other crustal in scale. Acknowledgements This research was funded by NSF Award #0911589 to DS, CM, and E. Kirby; and awards EAR054570, EAR0746192, and OCE-0648331 to CM, and a grant from the GDL Foundation to S. Haines. We are grateful to Gary Axen and Cristiano Collettini for thorough and constructive reviews. We thank Steve Swavely for technical assistance in the laboratory and B. Kaproth, K. Bhadra, B. Carpenter, M. Ikari, and A. Rathbun for field and laboratory support. Appendix A. Supplementary data Supplementary data associated with this article can be found in the online version at http://dx.doi.org/10.1016/j.epsl.2014.09.034. These data include the Google map of the most important areas described in this article. References Abers, G., 2001. Evidence for seismogenic normal faults at shallow dips in continental rifts. In: Wilson, R., Whitmarsh, R., Taylor, B., Froitzheim, N. (Eds.), Non-Volcanic Rifting of Continental Margins: A Comparison of Evidence from Land and Sea. In: Geol. Soc. (Lond.) Spec. Publ., vol. 187, pp. 305–318. Anderson, E.M., 1951. The Dynamics of Faulting. Olivier and Boyd, Edinburgh. 206 pp. Anderson, R., 2007. A blister hypothesis for the Central Mojave metamorphic core complex near Barstow, California. GSA Specialty Meetings Abstracts With Programs 39, 227. Andrew, J., 2002. The mesozoic and tertiary tectonics of the Panamint Range and Quail Mountains, California. PhD thesis. University of Kansas. 400 pp. Andrew, J., Walker, J.D., 2009. Reconstructing late Cenozoic deformation in central Panamint Valley, California: evolution of slip partitioning in the Walker Lane. Geosphere 5, 172–198. Axen, G., 1992. Pore pressure, stress increase and fault weakening in low-angle normal faulting. J. Geophys. Res. B, Solid Earth Planets 97, 8979–8991. Axen, G., 2004. Mechanics of low-angle normal faults. In: Karner, G., Taylor, B., Driscoll, N., Kohlstedt, D. (Eds.), Rheology and Deformation of the Lithosphere at Continental Margins. Columbia Univ. Press, New York, pp. 46–91. Axen, G., Taylor, W., Bartley, J., 1993. Space–time patterns and tectonic controls of Tertiary extension and magmatism in the Great Basin of the western United State. Geol. Soc. Am. Bull. 105, 56–76. Behnsen, J., Faulkner, D., 2012. The effect of mineralogy and effective normal stress on frictional strength of phyllosilicates. J. Struct. Geol. 42, 49–61. Bos, B., Spiers, C., 2000. Effect of phyllosilicates on fluid-assisted healing of gougebearing faults. Earth Planet. Sci. Lett. 184, 199–210. Boulton, C., Carpenter, B.M., Toy, V., Marone, C., 2012. Physical properties of surface outcrop cataclastic fault rocks, Alpine Fault, New Zealand. Geochem. Geophys. Geosyst. 13, Q01018. http://dx.doi.org/10.1029/2011GC003872. Brown, K., Kopf, A., Underwood, M., Weinberger, J., 2003. Compositional and fluid pressure controls on the state of stress on the Nankai subduction thrust: a weak plate boundary. Earth Planet. Sci. Lett. 214, 589–603. Buatier, M.D., Chauvet, A., Kanitpanyacharoen, W., Wenk, H.R., Ritz, J.F., Jolivet, M., 2011. Origin and behavior of clay minerals in the Bogd fault gouge, Mongolia. J. Struct. Geol. 34, 77–90. Buck, W., 1988. Flexural rotation of normal faults. Tectonics 7, 959–974. Byerlee, J., 1978. Friction of rocks. Pure Appl. Geophys. 116, 615–626. Carpenter, B., Marone, C., Saffer, D., 2011. Frictional strength of the San Andreas fault from laboratory measurements of SAFOD drill samples. Nat. Geosci. 10. http://dx.doi.org/10.1038/ngeo1089. Carpenter, B.M., Saffer, D.M., Marone, C., 2012. Frictional properties and sliding stability of the San Andreas Fault from deep drill core. Geology 40, 759–762. http:// dx.doi.org/10.1130/G33007.1. Chairaluce, L., Chiarabba, C., Collettini, C., Piccinini, D., Cocco, M., 2007. Architecture and mechanics of an active low-angle normal fault: Alto Tibernia Fault, northern Apennines, Italy. J. Geophys. Res. B, Solid Earth Planets 112, B10310. http:// dx.doi.org/10.1029/2007JB005015. Chery, J., 2002. Core complex mechanics: from the Gulf of Corinth to the Snake Range. Geology 29, 439–442. Cichanski, M., 2000. Low-angle range-flank faults in the Panamint, Inyo and Slate ranges, California: implications for recent tectonics of the Death Valley region. Geol. Soc. Am. Bull. 112, 871–883. Collettini, C., 2011. The mechanical paradox of low-angle normal faults: current understanding and open questions. Tectonophysics 510, 253–268. Collettini, C., Sibson, R., 2001. Normal faults, normal friction? Geology 29, 927–930. Collettini, C., De Paola, N., Goulty, N., 2006. Swithces in the minimum compressive stress direction induced by overpressure beneath a low-permeability fault zone. Terra Nova 18, 224–231. Collettini, C., Niemeijer, A., Viti, C., Marone, C., 2009a. Fault zone fabric and fault weakness. Nature 462, 907–910. http://dx.doi.org/10.1038/nature08585. Collettini, C., Viti, C., Smith, S., Holdsworth, R., 2009b. The development of interconnected talc networks and weakening of continental low-angle normal faults. Geology 37, 567–570. Converse, D., Nicholson, P., Pottorf, R., Miller, T., 2000. Controls on overpressure in rapidly subsiding basins and implications for failure of top seal. In: Mello, M., Katz, B. (Eds.), Petroleum Systems of South Atlantic Margins, AAPG Memoir, vol. 73, pp. 133–150. Davis, G., 1988. Rapid upward transport of mid-crustal mylonitic gneisses in the footwall of a Miocene detachment fault, Whipple Mountains, southeastern California. Geol. Rundsch. 77, 191–209. Den Hartog, S., Neimeijer, A., Spiers, C., 2012a. New constraints on megathrust stability under subduction zone P –T conditions. Earth Planet. Sci. Lett. 353–354, 240–252. Den Hartog, S., Peach, C., Matthjis de Winter, D., Spiers, C., Shimamoto, T., 2012b. Frictional properties of megathrust fault gouges at low sliding velocities: new data on effects of normal stress and temperature. J. Struct. Geol. 38, 156–171. Dokka, R., 1989. The Mojave extensional belt of southern California. Tectonics 8, 363–390. Dokka, R., 1995. Original dip and subsequent modification of a Cordilleral detachment fault, Mojave extensional belt, California. Geology 21, 711–714. Dorsey, R., Becker, U., 1995. Evolution of a large Miocene growth structure in the upper plate of the Whipple detachment fault, northeastern Whipple Mountains, California. Basin Res. 7, 151–163. Duvall, A., Clark, M., van der Pluijm, B., Li, Chuanyou, 2011. Direct dating of Eocene reverse faulting in northeastern Tibet using Ar-dating of fault clays and lowtemperature thermochronometry. Earth Planet. Sci. Lett. 304, 520–526. Fletcher, J., Bartley, J., 1994. Constrictional strain in a non-coaxial shear zone: implications for fold and rock fabric development, central Mojave mertamorphic core complex, California. J. Struct. Geol. 16, 555–570. Forshee, E., Yin, A., 1995. Evolution of monolithic breccia deposits in supradetachment basins, Whipple Mountains, California. Basin Res. 7, 181–197. Gaarenstroom, L., Tromp, A., DeJong, M., Brandenburg, A., 1993. Overpressure in the Central North Sea: implications for trap integrity and drilling safety. In: Parker, J.P. (Ed.), Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference. Geological Society, London, pp. 1305–1313. S. Haines et al. / Earth and Planetary Science Letters 408 (2014) 57–65 Haines, S., van der Pluijm, B., 2008. Clay quantification and Ar–Ar dating of synthetic and natural gouge: application to the Miocene Sierra Mazatán detachment fault, Sonora, Mexico. J. Struct. Geol. 30, 525–538. Haines, S., van der Pluijm, B., 2010. Dating the detachment fault system of the Ruby Mountains, Nevada: significance for the kinematics of low-angle normal faults. Tectonics 29 (TC4028). http://dx.doi.org/10.1029/2009TC002552. Haines, S., van der Pluijm, B., 2012. Patterns of mineral transformations in clay gouge, with examples from low-angle normal fault rocks in the western USA. J. Struct. Geol. 43, 2–32. Haines, S., Ikari, M., van der Pluijm, B., Marone, C., Saffer, D., 2009. Clay fabric intensity in natural and artificial fault gouges: implications for brittle fault zone processes and sedimentary basin clay fabric evolution. J. Geophys. Res. B, Solid Earth Planets 114. http://dx.doi.org/10.1029/2008JB005866. Haines, S., Kaproth, B., Marone, C., Saffer, D., van der Pluijm, B., 2013. Shear zones in clay-rich fault gouge: a laboratory study of fabric and evolution. J. Struct. Geol. 51, 206–225. Hayman, N., Knott, J., Cowan, D.S., Nemser, E., Sarna-Wojcicki, A., 2003. Quaternary lowangle slip on detachment faults in Death Valley, California. Geology 31, 343–346. Hreinsdóttir, S., Bennett, R., 2009. Active aseismic creep on the Alto Tibernia lowangle normal fault, Italy. Geology 37, 683–686. Ikari, M., Saffer, D., 2011. Comparison of frictional strength and velocity dependence between fault zones in the Nankai accretionary complex. Geochem. Geophys. Geosyst. 12. http://dx.doi.org/10.1029/2010GC003442. Ikari, M., Saffer, D., Marone, C., 2007. Effect of hydration state on the frictional properties of montmorillonite-based fault gouge. J. Geophys. Res. B, Solid Earth Planets 112. http://dx.doi.org/10.1029/2006JB004748. Ikari, M., Saffer, D., Marone, C., 2009. Frictional and hydrologic properties of a major splay fault system, Nankai subduction zone. Geophys. Res. Lett. 36, L20313. http://dx.doi.org/10.1029/2009GL040009. Ikari, M., Niemeijer, A., Marone, C., 2011. The role of fault zone fabric and lithification state on frictional strength, constitutive behavior, and deformation microstructure. J. Geophys. Res. B, Solid Earth Planets 116. http://dx.doi.org/ 10.1029/2011JB008264. Laurich, B., Urai, J., Desbois, G., Vollmer, C., Nussbaum, C., 2014. Microstructural evolution of an incipient fault zone in Opalinus Clay: insights from an optical and electron microscopic study of ion-beam polished samples from the Main Fault in the Mt-Terri Underground Research Laboratory. J. Struct. Geol. 67, 107–128. Lecomte, E., Le Pourhiet, L., Lacomb, O., 2012. Mechanical basis for slip along lowangle normal faults. Geophys. Res. Lett. 39, L03307. http://dx.doi.org/10.1029/ 2011GL050756. Logan, J., Freidman, M., Higgs, M., Dengo, C., Shimamoto, T., 1979. Experimental studies of simulated fault gouge and their application to studies of natural fault zones. In: Proc. Conf. VIII, Analysis of Actual Fault Zones in Bedrock. U.S. Geological Survey, Menlo Park, CA, pp. 305–343. Longwell, C., 1945. Low-angle normal faults in the Basin and Range province. Eos 26, 1087–1118. Manatschal, G., 1999. Fluid- and reaction-assisted low-angle normal faulting: evidence from rift-related brittle fault rocks in the Alps (Err Nappe, eastern Switzerland). J. Struct. Geol. 21, 777–793. 65 Marone, C., 1998. Laboratory-derived friction laws and their application to seismic faulting. Annu. Rev. Earth Planet. Sci. 26, 643–696. Numelin, T., Marone, C., Kirby, E., 2007. Frictional properties of natural gouge from a low-angle normal fault, Panamint Vallet, California. Tectonics 26, TC2004. http://dx.doi.org/10.1029/2005TC001916. Ransom, F., Emmons, W., Garrey, G., 1910. Geology and ore deposits of the Bullfrog District, Nevada. U.S. Geol. Surv. Bull. 407, 1–130. Rice, J., 1992. Fault stress states, pore pressure distributions, and the weakness of the San Andreas Fault. In: Evans, B., Wong, T.-F. (Eds.), Fault Mechanics and Transport Properties of Rocks; a Festschrift in Honor of W.F. Brace. Academic Press, San Diego, CA, pp. 475–503. Roller, S., Behrmann, J., Kpof, A., 2001. Deformational fabrics of faulted rocks, and some syntectonic stress estimates from the active Woodlark Basin detachment zone. In: Geol. Soc. (Lond.) Spec. Publ., vol. 187, pp. 319–334. Samuelson, J., Elsworth, D., Marone, C., 2009. Shear-induced dilatancy of fluidsaturated faults: experiment and theory. J. Geophys. Res. B, Solid Earth Planets 114, B12404. http://dx.doi.org/10.1029/2008JB006273. Selverstone, J., Axen, G., Luther, A., 2012. Fault localization controlled by fluid infiltration into mylonites: formation and strength of low-angle normal faults in the midcrustal brittle-plastic transition. J. Geophys. Res. 117, B06210. http:// dx.doi.org/10.1029/2012JB009171. Smith, S., Faulkner, D., 2010. Laboratory measurements of the frictional properties of the Zuccale low-angle normal fault, Elba Island, Italy. J. Geophys. Res. B, Solid Earth Planets 115, B02407. http://dx.doi.org/10.1029/2008JB006274. Tembe, S., Lockner, D., Wong, Teng-Fong, 2009. Constraints on the stress state of the San Andreas Fault with analysis based on core and cuttings from San Andreas Fault Observatory at Depth (SAFOD) drilling phases 1 and 2. J. Geophys. Res. 114, B11401. http://dx.doi.org/10.1029/2008JB005883. Tesei, T., Collettini, C., Carpenter, B., Viti, C., Marone, C., 2012. Frictional strength and healing behavior of phyllosilicate-rich faults. J. Geophys. Res. B, Solid Earth Planets 117, B09402. http://dx.doi.org/10.1029/2012JB09402. van der Pluijm, B., Ho, N., Peacor, D., 1994. High-resolution X-ray texture goniometry. J. Struct. Geol. 16, 1029–1032. Verbane, B.A., He, C., Spiers, C.J., 2010. Frictional properties of sedimentary rocks and natural fault gouge from the Longmen Shan Fault Zone, Sichuan, China. Bull. Seismol. Soc. Am. 100, 2767–2790. Walker, J., Bartley, B., Glazner, A., 1990. Large-magnitude Miocene extension in the central Mojave desert: implications for Paleozoic to Tertiary paleogeography and tectonics. J. Geophys. Res. B, Solid Earth Planets 95, 557–569. Wernicke, B., 1995. Low-angle normal faults and seismicity: a review. J. Geophys. Res. B, Solid Earth Planets 100, 20159–20174. Wernicke, B., Axen, G., 1988. On the role of isostasy in the evolution of low-angle normal fault systems. Geology 16, 848–851. Wong, M., Gans, P., 2008. Geologic, structural, and thermochronologic constraints on the tectonic evolution of the Sierra Mazatán core complex, Sonora, Mexico: new insights into metamorphic core complex formation. Tectonics 27, TC4013. http://dx.doi.org/10.1029/2007TC002173. Yin, A., Dunn, J., 1992. Structural and stratigraphic development of the Whipple– Chemehuevi detachment system, southeastern California: implications for the geometrical evolution of domal and basinal low-angle normal faults. Geol. Soc. Am. Bull. 104, 659–674.