on the thermal structure of oceanic transform faults9 structure beneath oceanic transform faults. we...

17
* Corresponding Author, Dept. of Geology and Geophysics, Woods Hole Oceanographic Institution, 360 Woods Hole Road MS #22, Woods Hole, MA 02543, email: [email protected], phone: 508-289-3637, fax: 508-457-2187. On the thermal structure of oceanic transform faults 1 Mark D. Behn 1,* , Margaret S. Boettcher 1,2 , and Greg Hirth 1 2 1 Department of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, MA, USA 3 2 U.S. Geological Survey, Menlo Park, CA, USA 4 5 6 Abstract: 7 We use 3-D finite element simulations to investigate the upper mantle temperature 8 structure beneath oceanic transform faults. We show that using a rheology that 9 incorporates brittle weakening of the lithosphere generates a region of enhanced mantle 10 upwelling and elevated temperatures along the transform, with the warmest temperatures 11 and thinnest lithosphere predicted near the center of the transform. In contrast, previous 12 studies that examined 3-D advective and conductive heat transport found that oceanic 13 transform faults are characterized by anomalously cold upper mantle relative to adjacent 14 intra-plate regions, with the thickest lithosphere at the center of the transform. These 15 earlier studies used simplified rheologic laws to simulate the behavior of the lithosphere 16 and underlying asthenosphere. Here, we show that the warmer thermal structure 17 predicted by our calculations is directly attributed to the inclusion of a more realistic 18 brittle rheology. This warmer upper mantle temperature structure is consistent with a 19 wide range of geophysical and geochemical observations from ridge-transform 20 environments, including the depth of transform fault seismicity, geochemical anomalies 21 along adjacent ridge segments, and the tendency for long transforms to break into a series 22 of small intra-transform spreading centers during changes in plate motion. 23 Key Words: Oceanic transform faults, mid-ocean ridges, fault rheology, intra-transform 24 spreading centers 25 26

Upload: others

Post on 23-Oct-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

  • *Corresponding Author, Dept. of Geology and Geophysics, Woods Hole Oceanographic Institution, 360 Woods Hole Road MS #22, Woods Hole, MA 02543, email: [email protected], phone: 508-289-3637, fax: 508-457-2187.

    On the thermal structure of oceanic transform faults 1 Mark D. Behn1,*, Margaret S. Boettcher1,2, and Greg Hirth1 2 1Department of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, MA, USA 3 2U.S. Geological Survey, Menlo Park, CA, USA 4 5 6 Abstract: 7 We use 3-D finite element simulations to investigate the upper mantle temperature 8

    structure beneath oceanic transform faults. We show that using a rheology that 9

    incorporates brittle weakening of the lithosphere generates a region of enhanced mantle 10

    upwelling and elevated temperatures along the transform, with the warmest temperatures 11

    and thinnest lithosphere predicted near the center of the transform. In contrast, previous 12

    studies that examined 3-D advective and conductive heat transport found that oceanic 13

    transform faults are characterized by anomalously cold upper mantle relative to adjacent 14

    intra-plate regions, with the thickest lithosphere at the center of the transform. These 15

    earlier studies used simplified rheologic laws to simulate the behavior of the lithosphere 16

    and underlying asthenosphere. Here, we show that the warmer thermal structure 17

    predicted by our calculations is directly attributed to the inclusion of a more realistic 18

    brittle rheology. This warmer upper mantle temperature structure is consistent with a 19

    wide range of geophysical and geochemical observations from ridge-transform 20

    environments, including the depth of transform fault seismicity, geochemical anomalies 21

    along adjacent ridge segments, and the tendency for long transforms to break into a series 22

    of small intra-transform spreading centers during changes in plate motion. 23

    Key Words: Oceanic transform faults, mid-ocean ridges, fault rheology, intra-transform 24 spreading centers 25

    26

  • -2-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    1. Introduction 26

    Oceanic transform faults are an ideal environment for studying the mechanical 27

    behavior of strike-slip faults because of the relatively simple thermal, kinematic, and 28

    compositional structure of the oceanic lithosphere. In continental regions fault zone 29

    rheology is influenced by a combination of mantle thermal structure, variations in crustal 30

    thickness, and heterogeneous crustal and mantle composition. By contrast, the rheology 31

    of the oceanic upper mantle is primarily controlled by temperature. In the ocean basins 32

    effective elastic plate thickness (e.g., Watts, 1978) and the maximum depth of intra-plate 33

    earthquakes (Chen and Molnar, 1983; McKenzie et al., 2005; Wiens and Stein, 1983) 34

    correlate closely with the location of the 600ºC isotherm as calculated from a half-space 35

    cooling model. Similarly, recent studies show that the maximum depth of transform fault 36

    earthquakes corresponds to the location of the 600ºC isotherm derived by averaging the 37

    half-space thermal structures on either side of the fault (Abercrombie and Ekström, 2001; 38

    Boettcher, 2005). These observations are consistent with extrapolations from laboratory 39

    studies on olivine that indicate the transition from stable to unstable frictional sliding 40

    occurs around 600ºC at geologic strain-rates (Boettcher et al., 2006). Furthermore, 41

    microstructures observed in peridotite mylonites from oceanic transforms show that 42

    localized viscous deformation occurs at temperatures around 600–800ºC (e.g., Jaroslow 43

    et al., 1996; Warren and Hirth, 2006). 44

    While a half-space cooling model does a good job of predicting the maximum depth 45

    of transform earthquakes, it neglects many important physical processes that occur in the 46

    Earth’s crust and upper mantle (e.g., advective heat transport resulting from temperature-47

    dependent viscous flow, hydrothermal circulation, and viscous dissipation). Numerical 48

    models that incorporate 3-D advective and conductive heat transport indicate that the 49

    upper mantle beneath oceanic transform faults is anomalously cold relative to a half-50

    space model (Forsyth and Wilson, 1984; Furlong et al., 2001; Phipps Morgan and 51

    Forsyth, 1988; Shen and Forsyth, 1992). This reduction in mantle temperature results 52

    from a combination of two effects: 1) conductive cooling from the adjacent old, cold 53

    lithosphere across the transform fault, and 2) decreased mantle upwelling beneath the 54

    transform. Together these effects can result in up to a ~75% increase in lithospheric 55

    thickness beneath the center of a transform fault relative to a half-space cooling model, as 56

  • -3-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    well as significant cooling of the upper mantle beneath the ends of the adjacent spreading 57

    centers. This characteristic ridge-transform thermal structure has been invoked to explain 58

    the focusing of crustal production toward the centers of ridge segments (Magde and 59

    Sparks, 1997; Phipps Morgan and Forsyth, 1988; Sparks et al., 1993), geochemical 60

    evidence for colder upper mantle temperatures near segment ends (Ghose et al., 1996; 61

    Niu and Batiza, 1994; Reynolds and Langmuir, 1997), increased fault throw and wider 62

    fault spacing near segment ends (Shaw, 1992; Shaw and Lin, 1993), and the blockage of 63

    along-axis flow of plume material (Georgen and Lin, 2003). 64

    However, correlating the maximum depth of earthquakes on transform faults with this 65

    colder thermal structure implies that the transition from stable to unstable frictional 66

    sliding occurs at temperatures closer to ~350ºC, which is inconsistent with both 67

    laboratory studies and the depth of intra-plate earthquakes. Moreover, if oceanic 68

    lithosphere is anomalously cold and thick beneath oceanic transform faults, it is difficult 69

    to explain the tendency for long transform faults to break into a series of en echelon 70

    transform zones separated by small intra-transform spreading centers during changes in 71

    plate motion (Fox and Gallo, 1984; Lonsdale, 1989; Menard and Atwater, 1969; Searle, 72

    1983). 73

    To address these discrepancies between the geophysical observations and the 74

    predictions of previous numerical modeling studies, we investigate the importance of 75

    fault rheology on the thermo-mechanical behavior of oceanic transform faults. Earlier 76

    studies that incorporated 3-D conductive and advective heat transport used simplified 77

    rheologic laws to simulate the behavior of the lithosphere and underlying asthenosphere. 78

    Using a series of 3-D finite element models, we show that brittle weakening of the 79

    lithosphere strongly reduces the effective viscosity beneath the transform, resulting in 80

    enhanced upwelling and thinning of the lithosphere. Our calculations suggest that the 81

    thermal structure of oceanic transform faults is more similar to that predicted from half-82

    space cooling, but with the warmest temperatures located near the center of the 83

    transform. These results have important implications for the mechanical behavior of 84

    oceanic transforms, melt generation and migration at mid-ocean ridges, and the long-term 85

    response of transform faults to changes in plate motion. 86

  • -4-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    2. Model Setup 87

    To solve for coupled 3-D incompressible mantle flow and thermal structure 88

    surrounding an oceanic transform fault we use the COMSOL 3.2 finite element software 89

    package (Figure 1). In all simulations, flow is driven by imposing horizontal velocities 90

    parallel to a 150-km transform fault along the top boundaries of the model space 91

    assuming a full spreading rate of 6 cm/yr. The base of the model is open to convective 92

    flux without resistance from the underlying mantle. Symmetric boundary conditions are 93

    imposed on the sides of the model space parallel to the spreading direction, and the 94

    boundaries perpendicular to spreading are open to convective flux. The temperature 95

    across the top and bottom of the model space is set to Ts = 0ºC and Tm = 1300ºC, 96

    respectively. Flow associated with temperature and compositional buoyancy is ignored. 97

    To investigate the importance of rheology on the pattern of flow and thermal structure 98

    at oceanic transform faults we examined four scenarios with increasingly realistic 99

    descriptions of mantle rheology: 1) constant viscosity, 2) temperature-dependent 100

    viscosity, 3) temperature-dependent viscosity with an pre-defined weak zone around the 101

    transform, and 4) temperature-dependent viscosity with a visco-plastic approximation for 102

    brittle weakening. In all models we assume a Newtonian mantle rheology. In Models 2–103

    4 the effect of temperature on viscosity is calculated by: 104

    !

    " ="oexp Qo /RT( )exp Qo /RTm( )

    #

    $ %

    &

    ' ( (1) 105

    where ηo is the reference viscosity of 1019 Pa⋅s, Qo is the activation energy, and R is the 106

    gas constant. In all simulations we assume an activation energy of 250 kJ/mol. This 107

    value represents a reduction of a factor of two relative to the laboratory value as a linear 108

    approximation for non-linear rheology (Christensen, 1983). The maximum viscosity is 109

    not allowed to exceed 1023 Pa⋅s. 110

    3. Influence of 3-D Mantle Flow on Transform Thermal Structure 111

    Figure 2 illustrates the thermal structure calculated at the center of the transform fault 112

    from Models 1–4, as well as the thermal structure determined by averaging half-spacing 113

    cooling models on either side of the transform. Because the thermal structure calculated 114

    from half-space cooling is equal on the adjacent plates, the averaging approach predicts 115

  • -5-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    the same temperature at the center of the transform fault as for the adjacent intra-plate 116

    regions. This model, therefore, provides a good reference for evaluating whether our 3-D 117

    numerical calculations predict excess cooling or excess heating below the transform. The 118

    temperature solution for a constant viscosity mantle (Model 1) was determined 119

    previously by Phipps Morgan and Forsyth (1988); our results agree with theirs to within 120

    5% throughout the model space. The coupled temperature, mantle flow solution predicts 121

    a significantly colder thermal structure at the center of the transform than does half-space 122

    cooling, with the depth of the 600ºC isotherm increasing from ~7 km for the half-space 123

    model to ~12 km for the constant viscosity flow solution (Figure 2A). As noted by 124

    Phipps Morgan and Forsyth (1988) this reduction in temperature is primarily the result of 125

    decreased mantle upwelling beneath the transform fault relative to enhanced upwelling 126

    under the ridge axis. Shen and Forsyth (1992) showed that incorporating temperature-127

    dependent viscosity (e.g., Model 2) produces enhanced upwelling and warmer 128

    temperatures beneath the ridge axis relative to a constant viscosity mantle (Figure 3). 129

    However, away from the ridge axis the two solutions are quite similar and result in 130

    almost identical temperature-depth profiles at the center of the transform (Figures 2A & 131

    3). 132

    4. Influence of Fault Zone Rheology on Transform Thermal Structure 133

    Several lines of evidence indicate that oceanic transform faults are significantly 134

    weaker than the surrounding lithosphere. Comparisons of abyssal hill fabric observed 135

    near transforms to predictions of fault patterns from numerical modeling suggest that the 136

    mechanical coupling across the fault is very weak on geologic time scales (Behn et al., 137

    2002; Phipps Morgan and Parmentier, 1984). Furthermore, dredging in transform valleys 138

    and valley walls frequently returns serpentinized peridotites (Cannat et al., 1991; Dick et 139

    al., 1991), which may promote considerable frictional weakening along the transform 140

    (Escartín et al., 2001; Moore et al., 1996; Rutter and Brodie, 1987). Finally, in 141

    comparison to continental strike-slip faults, seismic moment studies show that oceanic 142

    transforms have high seismic deficits (Boettcher and Jordan, 2004; Okal and 143

    Langenhorst, 2000), suggesting that oceanic transforms may be characterized by large 144

    amounts of aseismic slip. 145

  • -6-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    In Model 3 we simulate the effect of a weak fault zone, by setting the viscosity to 1019 146

    Pa⋅s in a 5-km wide region surrounding the transform that extends downward to a depth 147

    of 20 km. This results in a narrow fault zone that is 3–4 orders of magnitude lower 148

    viscosity than the surrounding regions. Our approach is similar to that used by Furlong et 149

    al. (2001) and van Wijk and Blackman (2004), though these earlier studies modeled 150

    deformation in a visco-elastic system in which the transform fault was simulated as a 151

    shear-stress-free plane using the slippery node technique of Melosh and Williams (1989). 152

    Our models show that the incorporation of a weak fault zone produces slightly warmer 153

    conditions along the transform than for either a constant viscosity mantle (Model 1) or 154

    temperature-dependent viscosity without an imposed fault zone (Model 2). However, the 155

    predicted temperatures from the fault zone model remain colder than those calculated by 156

    the half-space cooling model (Figures 2A & 3). Varying the maximum depth of the 157

    weak zone does not significantly influence the predicted thermal structure. 158

    Representing the transform as a pre-defined zone of uniform weakness clearly over-159

    simplifies the brittle processes occurring within the lithosphere. In Model 4, we 160

    incorporate a more realistic formulation for fault zone behavior by using a visco-plastic 161

    rheology to simulate brittle weakening (Chen and Morgan, 1990). In this formulation, 162

    brittle strength is approximated by defining a frictional resistance law (e.g., Byerlee, 163

    1978): 164

    !

    "max

    = Co

    + µ#gz (2) 165

    in which Co is cohesion (10 MPa), µ is the friction coefficient (0.6), ρ is density (3300 166

    kg/m3), g is the gravitational acceleration, and z is depth. Following Chen and Morgan 167

    (1990), the maximum effective viscosity is then limited by: 168

    !

    " =#

    max

    2˙ $ II

    (3) 169

    where

    !

    ˙ " II is the second-invariant of the strain-rate tensor. The effect of adding this brittle 170

    failure law is to limit viscosity near the surface where the temperature dependence of 171

    Equation 1 produces unrealistically high mantle viscosities and stresses (Figure 2B). 172

    The inclusion of the visco-plastic rheology results in significantly warmer thermal 173

    conditions beneath the transform than predicted by Models 1–3 or half-space cooling 174

    (Figures 2A & 3). The higher temperatures result from the brittle weakening of the 175

  • -7-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    lithosphere, which reduces the effective viscosity by up to 2 orders of magnitude in a 10-176

    km wide region surrounding the fault zone. Unlike Model 3 the width of this region is 177

    not predefined and develops as a function of the rheology and applied boundary 178

    conditions. The zone of decreased viscosity enhances passive upwelling beneath the 179

    transform, which in turn increases upward heat transport, warming the fault zone and 180

    further reducing viscosity (Figure 4). The result is a characteristic thermal structure in 181

    which the transform fault is warmest at its center and cools towards the adjacent ridge 182

    segments (Figures 2C). Moreover, rather than the transform being a region of 183

    anomalously cold lithosphere relative to a half-space cooling model and the surrounding 184

    intra-plate mantle, the center of the transform is warmer than adjacent lithosphere of the 185

    same age. 186

    Although we have not explicitly modeled the effects of non-linear rheology, several 187

    previous studies have examined the importance of a non-linear viscosity law on mantle 188

    flow and thermal structure in a segmented ridge-transform system (Furlong et al., 2001; 189

    Shen and Forsyth, 1992; van Wijk and Blackman, 2004). Without the effects of brittle 190

    weakening, these earlier studies predicted temperatures below the transform that were 191

    significantly colder than a half-space cooling model. Thus, we conclude that the 192

    inclusion of a visco-plastic rheology is the key factor for producing the warmer transform 193

    fault thermal structure illustrated in Model 4. 194

    5. Implications for the Behavior of Oceanic Transform Faults 195

    Our numerical simulations indicate that brittle weakening plays an important role in 196

    controlling the thermal structure beneath oceanic transform faults. Specifically, 197

    incorporating a more realistic treatment of brittle rheology (as shown in Model 4), results 198

    in an upper mantle temperature structure that is consistent with a wide range of 199

    geophysical and geochemical observations from ridge-transform environments. The 200

    temperatures below the transform fault predicted in Model 4 are similar to the half-space 201

    cooling model, indicating that the maximum depth of transform fault seismicity is indeed 202

    limited by the ~600ºC isotherm as shown by Abercrombie and Ekström (2001). This 203

    temperature is consistent with the transition from velocity-weakening to velocity-204

    strengthening frictional behavior extrapolated from laboratory experiments (Boettcher et 205

  • -8-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    al., 2006) and the depth of oceanic intra-plate earthquakes (McKenzie et al., 2005). In 206

    addition, a combination of microstructural and petrological observations indicate that the 207

    transition from brittle to ductile processes occurs at a temperature of ~600ºC in oceanic 208

    transform faults. Microstructural analyses of peridotite mylonites recovered from 209

    oceanic fracture zones indicate that strain localization results from the combined effects 210

    of grain size reduction, grain boundary sliding and second phase pinning (Warren and 211

    Hirth, 2006). Jaroslow et al. (1996) estimated a minimum temperature for mylonite 212

    deformation of ~600ºC, based on olivine-spinel geothermometry. Furthermore, the 213

    randomization of pre-existing lattice preferred orientation (LPO) in the finest grained 214

    areas of these mylonites indicates the grain size reduction promotes a transition from 215

    dislocation creep processes to diffusion creep, consistent with extrapolation of 216

    experimental olivine flow laws to temperatures of 600–800ºC. 217

    While the inclusion of a visco-plastic rheology results in significant warming beneath 218

    the transform fault, the temperature structure at the ends of the adjacent ridge segments 219

    changes only slightly relative to the solution for a constant viscosity mantle. In 220

    particular, both models predict a region of cooling along the adjacent ridge segments that 221

    extends 15–20 km from the transform fault (Figure 2C). This transform “edge effect” 222

    has been invoked to explain segment scale variations in basalt chemistry (Ghose et al., 223

    1996; Niu and Batiza, 1994; Reynolds and Langmuir, 1997) and increased fault throw 224

    and fault spacing toward the ends of slow-spreading ridge segments (Shaw, 1992; Shaw 225

    and Lin, 1993). In addition, this along-axis temperature gradient provides an efficient 226

    mechanism for focusing crustal production toward the centers of ridge segments (Magde 227

    and Sparks, 1997; Phipps Morgan and Forsyth, 1988; Sparks et al., 1993). 228

    The elevated temperatures near the center of the transform in Model 4 may also 229

    account for the tendency of long transform faults to break into a series of intra-transform 230

    spreading centers during changes in plate motion (Fox and Gallo, 1984; Lonsdale, 1989; 231

    Menard and Atwater, 1969). This “leaky transform” phenomenon has been attributed to 232

    the weakness of oceanic transform faults relative to the surrounding lithosphere (Fox and 233

    Gallo, 1984; Lowrie et al., 1986; Searle, 1983). However, the leaky transform hypothesis 234

    is in direct conflict with the thermal structure predicted from Models 1–3, which show 235

    the transform to be a region of anomalously cold, thick lithosphere. In contrast, the 236

  • -9-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    thermal structure predicted by Model 4 indicates that transforms are hottest and weakest 237

    near their centers (Figure 2C). Thus, perturbations in plate motion, which generate 238

    extension across the transform, should result in rifting and enhanced melting in these 239

    regions. 240

    The incorporation of the brittle rheology also promotes strain localization on the plate 241

    scale. In particular, if transforms were regions of thick, cold lithosphere (as predicted by 242

    Models 1–3) then over time deformation would tend to migrate outward from the 243

    transform zone into the adjacent regions of thinner lithosphere. However, the warmer 244

    thermal structure that results from the incorporation of a visco-plastic rheology will tend 245

    to keep deformation localized within the transform zone on time-scales corresponding to 246

    the age of ocean basins. 247

    In summary, brittle weakening of the lithosphere along oceanic transform faults 248

    generates a region of enhanced mantle upwelling and elevated temperatures relative to 249

    adjacent intra-plate regions. The thermal structure is similar to that predicted by a half-250

    space cooling model, but with the warmest temperatures located at the center of the 251

    transform. This characteristic upper mantle temperature structure is consistent with a 252

    wide range of geophysical and geochemical observations, and provides important 253

    constraints on the future interpretation of microseismicity data, heat flow, and basalt and 254

    peridotite geochemistry in ridge-transform environments. 255

    Acknowledgements 256

    We thank Jeff McGuire, Laurent Montési, Trish Gregg, Jian Lin, Henry Dick, and Don 257

    Forsyth for fruitful discussions that helped motivate this work. Funding was provided by 258

    NSF grants EAR-0405709 and OCE-0443246. 259

    260

  • -10-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    Figure Captions 260

    Figure 1: Model setup for numerical simulations of mantle flow and thermal structure at 261

    oceanic transform faults. All calculations are performed for a 150-km long transform and 262

    a full spreading rate of 6 cm/yr. Locations of cross-sections used in Figures 2–4 are 263

    shown in grey. 264

    Figure 2: (A) Thermal structure and (B) stress calculated versus depth calculated at the 265

    center of a 150-km long transform fault assuming a full spreading rate of 6 cm/yr. (C) 266

    Location of the 600ºC and 1200ºC isotherms along the plate boundary for the half-space 267

    model (grey), Model 1 (black), and Model 4 (red). 268

    Figure 3: Cross-sections of mantle temperature at a depth of 20 km for (A) Model 1: 269

    constant viscosity of 1019 Pa⋅s, (B) Model 2: temperature-dependent viscosity, (C) Model 270

    3: temperature-dependent viscosity with a weak fault zone, and (D) Model 4: 271

    temperature-dependent viscosity with a frictional failure law. Black arrows indicate 272

    horizontal flow velocities. Grey lines show position of plate boundary. Location of 273

    horizontal cross-section is indicated in Figure 1. Note that Model 4 incorporating 274

    frictional resistance predicts significantly warmer temperatures along the transform than 275

    Models 1–3. 276

    Figure 4: Vertical cross-sections through the center of the transform fault showing (left) 277

    strain-rate, and (right) temperature and mantle flow for Models 1–4. The location of the 278

    cross-sections is indicated in Figure 1. Note the enhanced upwelling below the transform 279

    results in warmer thermal structure for Model 4 compared to Models 1–3. 280

    281

  • -11-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    References 281

    Abercrombie, R.E., and Ekström, G., 2001, Earthquake slip on oceanic transform faults: 282 Nature, v. 410, p. 74-77. 283

    Behn, M.D., Lin, J., and Zuber, M.T., 2002, Evidence for weak oceanic transform faults: 284 Geophys. Res. Lett., v. 29, doi:10.1029/2002GL015612. 285

    Boettcher, M.S., 2005, Slip on Ridge Transform Faults: Insights from Earthquakes and 286 Laboratory Experiments [Ph.D. thesis], MIT/WHOI Joint Program. 287

    Boettcher, M.S., Hirth, G., and Evans, B., 2006, Olivine friction at the base of oceanic 288 seismogenic zones: J. Geophys. Res., submitted. 289

    Boettcher, M.S., and Jordan, T.H., 2004, Earthquake scaling relations for mid-ocean 290 ridge transform faults: J. Geophys. Res., v. 109, doi:10.1029/2004JB003110. 291

    Byerlee, J., 1978, Friction of rocks: Pure Appl. Geophys., v. 116, p. 615-626. 292 Cannat, M., Mamaloukas-Frangoulis, V., Auzende, J.-M., Bideau, D., Bonatti, E., 293

    Honnorez, J., Lagabrielle, Y., Malavielle, J., and Mevel, C., 1991, A geological 294 cross-section of the Vema fracture zone transverse ridge, Atlantic Ocean: J. 295 Geodyn., v. 13, p. 97-118. 296

    Chen, W.-P., and Molnar, P., 1983, Focal depths of intracontinental and intraplate 297 earthquakes and their implications for the thermal and mechanical properties of 298 the lithosphere: J. Geophys. Res., v. 88, p. 4183-4214. 299

    Chen, Y., and Morgan, W.J., 1990, A nonlinear rheology model for mid-ocean ridge axis 300 topography: J. Geophys. Res., v. 95, p. 17583-17604. 301

    Christensen, U., 1983, Convection in a variable-viscosity fluid: Newtonian versus power-302 law rheology: Earth Planet. Sci. Lett., v. 64, p. 153-162. 303

    Dick, H.J.B., Schouten, H., Meyer, P.S., Gallo, D.G., Bergh, H., Tyce, R., Patriat, P., 304 Johnson, K.T.M., Snow, J., and Fisher, A., 1991, Tectonic evolution of the 305 Atlantis II Fracture Zone, in Von Herzen, R.P., and Robinson, P.T., eds., Proc. 306 ODP, Sci. Results, Volume 118: College Station, Ocean Drilling Program, p. 359-307 398. 308

    Escartín, J., Hirth, G., and Evans, B., 2001, Strength of slightly serpentinized peridotites: 309 Implications for the tectonics of oceanic lithosphere: Geology, v. 29, p. 1023-310 1026. 311

    Forsyth, D.W., and Wilson, B., 1984, Three-dimensional temperature structure of a ridge-312 transform-ridge system: Earth Planet. Sci. Lett., v. 70, p. 355-362. 313

    Fox, P.J., and Gallo, D.G., 1984, A tectonic model for ridge-transform-ridge plate 314 boundaries: Implications for the structure of oceanic lithosphere: Tectonophys., v. 315 104, p. 205-242. 316

    Furlong, K.P., Sheaffer, S.D., and Malservisi, R., 2001, Thermal-rheological controls on 317 deformation within oceanic transforms, in Holdsworth, R.E., Strachan, R.A., 318 Magloughlin, J.F., and Knipe, R.J., eds., The Nature and Tectonic Significance of 319 Fault Zone Weakening, Volume 186: London, Geology Society, p. 65-84. 320

    Georgen, J.E., and Lin, J., 2003, Plume-transform interactions at ultra-slow spreading 321 ridges: Implications for the Southwest Indian Ridge: Geochem., Geophys., 322 Geosys., v. 4, 9106, doi:10.1029/2003GC000542. 323

  • -12-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    Ghose, I., Cannat, M., and Seyler, M., 1996, Transform fault effect on mantle melting in 324 the MARK area (Mid-Atlantic Ridge south of the Kane transform): Geology, v. 325 24, p. 1139-1142. 326

    Jaroslow, G.E., Hirth, G., and Dick, H.J.B., 1996, Abyssal peridotite mylonites: 327 Implications for grain-size sensitive flow and strain localization in the oceanic 328 lithosphere: Tectonophys., v. 256, p. 17-37. 329

    Lonsdale, P., 1989, Segmentation of the Pacific-Nazca spreading center, 1ºN-20ºS: J. 330 Geophys. Res., v. 94, p. 12,197-12,225. 331

    Lowrie, A., Smoot, C., and Batiza, R., 1986, Are oceanic fracture zones locked and 332 strong or weak? New evidence for volcanic activity and weakness: Geology, v. 333 14, p. 242-245. 334

    Magde, L.S., and Sparks, D.W., 1997, Three-dimensional mantle upwelling, melt 335 generation, and melt migration beneath segment slow spreading ridges: J. 336 Geophys. Res., v. 102, p. 20571-20583. 337

    McKenzie, D., Jackson, J., and Priestley, K., 2005, Thermal structure of oceanic and 338 continental lithosphere: Earth Planet. Sci. Lett., v. 233, p. 337-349. 339

    Melosh, H.J., and Williams, C.A., Jr., 1989, Mechanics of graben formation in crustal 340 rocks: A finite element analysis: J. Geophys. Res., v. 94, p. 13,961-13,973. 341

    Menard, H.W., and Atwater, T., 1969, Origin of fracture zone topography: Nature, v. 342 222, p. 1037-1040. 343

    Moore, D.E., Lockner, L.D.A., Summers, R., Shengli, M., and Byerlee, J.D., 1996, 344 Strength of chrysotile-serpentinite gouge under hydrothermal conditions: Can it 345 explain a weak San Andreas fault? Geology, v. 24, p. 1041-1044. 346

    Niu, Y., and Batiza, R., 1994, Magmatic processes at a slow spreading ridge segment: 347 26ºS Mid-Atlantic Ridge: J. Geophys. Res., v. 99, p. 19,719-19,740. 348

    Okal, E.A., and Langenhorst, A.R., 2000, Seismic properties of the Eltanin Transform 349 System, South Pacific: Phys. Earth Planet. Inter., v. 119, p. 185-208. 350

    Phipps Morgan, J., and Forsyth, D.W., 1988, Three-dimensional flow and temperature 351 perturbations due to a transform offset: Effects on oceanic crust and upper mantle 352 structure: J. Geophys. Res., v. 93, p. 2955-2966. 353

    Phipps Morgan, J., and Parmentier, E.M., 1984, Lithospheric stress near a ridge-354 transform intersection: Geophys. Res. Lett., v. 11, p. 113-116. 355

    Reynolds, J.R., and Langmuir, C.H., 1997, Petrological systematics of the Mid-Atlantic 356 Ridge south of Kane: Implications for ocean crust formation: J. Geophys. Res., v. 357 102, p. 14,915-14,946. 358

    Rutter, E.H., and Brodie, K.H., 1987, On the mechanical properties of oceanic transform 359 faults: Ann. Tecton., v. 1, p. 87-96. 360

    Searle, R.C., 1983, Multiple, closely spaced transform faults in fast-slipping fracture 361 zones: Geology, v. 11, p. 607-611. 362

    Shaw, P.R., 1992, Ridge segmentation, faulting and crustal thickness in the Atlantic 363 Ocean: Nature, v. 358, p. 490-493. 364

    Shaw, P.R., and Lin, J., 1993, Causes and consequences of variations in faulting style at 365 the Mid-Atlantic Ridge: J. Geophys. Res., v. 98, p. 21,839-21,851. 366

    Shen, Y., and Forsyth, D.W., 1992, The effects of temperature- and pressure-dependent 367 viscosity on three-dimensional passive flow of the mantle beneath a ridge-368 transform system: J. Geophys. Res., v. 97, p. 19,717-19,728. 369

  • -13-

    Behn et al: Thermal Structure of Oceanic Transform Faults, Submitted to Geology March 2006

    Sparks, D.W., Parmentier, E.M., and Phipps Morgan, J., 1993, Three-dimensional mantle 370 convection beneath a segmented spreading center: Implications for along-axis 371 variations in crustal thickness and gravity: J. Geophys. Res., v. 98, p. 21,977-372 21,995. 373

    van Wijk, J.W., and Blackman, D.K., 2004, Deformation of oceanic lithosphere near 374 slow-spreading ridge discontinuities: Tectonophys., v. 407, p. 211-225. 375

    Warren, J.M., and Hirth, G., 2006, Grain size sensitive deformation mechanisms in 376 naturally deformed peridotites: Earth Planet. Sci. Lett., v. in revision. 377

    Watts, A.B., 1978, An analysis of isostasy in the World's Oceans 1. Hawaiian-Emperor 378 Seamount Chain: J. Geophys. Res., v. 83, p. 5989-6004. 379

    Wiens, D.A., and Stein, S., 1983, Age dependence of oceanic intraplate seismicity and 380 implications for lithospheric evolution: J. Geophys. Res., v. 88, p. 6455-6468. 381

    382 383

  • 250 km

    150 km

    50 km

    100 km

    100 km

    3 cm/yr

    3 cm/yr Ts = 0ºC

    Tm = 1300ºC

    Fig. 4

    Fig. 3

    Figure 1

  • 0 100 200 300Stress (MPa)

    B)Center ofTransform

    Temperature (oC)

    Dep

    th (k

    m)

    Half−Space ModelModel 1: Constant ηModel 2: η(T)Model 3: η(T) + FaultModel 4: η(T, friction)

    A)Center ofTransform

    0 500 1000−35

    −30

    −25

    −20

    −15

    −10

    −5

    0

    Distance Along Plate Boundary (km)

    Dep

    th (k

    m)

    Transform FaultRidge RidgeAxis Axis

    600oC Isotherm ~ B/D Transition

    1200oC Isotherm ~ Solidus

    C)

    Figure 2

    −100 −75 −50 −25 0 25 50 75 100

    −30

    −20

    −10

    0

  • Alon

    g−R

    idge

    Dis

    tanc

    e (k

    m)

    A. Model 1: Constant η

    −40

    −20

    0

    20

    40 B. Model 2: η(T)

    Along−Transform Distance (km)

    C. Model 3: η(T) + Fault

    −100 −50 0 50 100−40

    −20

    0

    20

    40

    Along−Transform Distance (km)

    D. Model 4: η(T, friction)

    Figure 3

    −100 −50 0 50 100

    Temperature (oC)1000 1100 1200 1300

  • Dep

    th (k

    m)

    A. Model 1: Constant η

    −80

    −60

    −40

    −20

    0

    Dep

    th (k

    m)

    B. Model 2: η(T)

    −80

    −60

    −40

    −20

    0

    Dep

    th (k

    m)

    C. Model 3: η(T) + Fault

    −80

    −60

    −40

    −20

    0

    Dep

    th (k

    m)

    D. Model 4: η(T, friction)

    log10 srII (1/s)

    −40 −20 0 20 40−80

    −60

    −40

    −20

    0

    −15 −14 −13 −12

    Across−Transform Distance (km)

    Temperature (oC)

    Figure 4

    −40 −20 0 20 40

    0 400 800 1200