• Keine Ergebnisse gefunden

Chapter 7 / 8 (Summary and Conclusion / Zusammenfassung und Schlussfolgerung) These chapters contain the major findings of this work and their implications for

5. Conclusions

99

Acknowledgments

Data to support this article are available in the supporting information (Tables S1–S5)*.

We thank M. Wong, D. A. Foster, A. Manning (USGS), and the Associate Editor for their insightful reviews that helped to significantly improve this manuscript. K.M., A.M., and A.G. acknowledge support through the LOEWE funding program of Hesse’s Ministry of Higher Education, Research, and the Arts. We gratefully acknowledge support from National Science Foundation grants EAR 0838541 (CT), EAR 1019648 (CPC, AM), and EAR 0610098 (MW), as well as Swiss grants 200020-126973/1 (CT) and SNF-200021-103674 (MC). We thank T. Vennemann (Lausanne) and C. Wenske (Hannover) for laboratory support. Any use of trade, product, or firm names is for descriptive purposes only and does not imply endorsement by the U.S. Government.

*Appendix 4 (Tab. A4.1-A4.5) in this thesis.

101 References

References

Armstrong, R. L., 1968, Sevier orogenic belt in Nevada and Utah. Geol Soc Am Bull 79(4), 429-458, doi:10.1130/0016-7606(1968)79[429:sobina]2.0.co;2.

Armstrong, R. L., 1982, Cordilleran metamorphic core complexes—From Arizona to Southern Canada.

Annu Rev Earth Planet Sci 10(1), 129-154, doi:10.1146/annurev.ea.10.050182.001021.

Campani, M., A. Mulch, O. Kempf, F. Schlunegger, and N. Mancktelow, 2012, Miocene paleotopography of the Central Alps, Earth Planet Sci Lett 337-338, 174-185, doi:10.1016/j.epsl.2012.05.017.

Chacko, T., X. Hu, T. K. Mayeda, R. N. Clayton, and J. R. Goldsmith, 1996, Oxygen isotope fractionations in muscovite, phlogopite, and rutile. Geochim Cosmochim Ac 60(14), 2595-2608, doi:10.1016/0016-7037(96)00112-3.

Compton, R. R., 1980, Fabrics and strains in quartzites of a metamorphic core complex Raft River Mountains, Utah. Geol Soc Am Mem 153, 385-398, doi:10.1130/MEM153-p385.

Compton, R. R., V. R. Todd, R. E. Zartman, and C. W. Naeser, 1977, Oligocene and Miocene metamorphism, folding, and low-angle faulting in northwestern Utah. Geol Soc Am Bull 88(9), 1237-1250, doi:10.1130/0016-7606(1977)88<1237:oammfa>2.0.co;2.

Coney, P. J., 1980, Cordilleran metamorphic core complexes: An overview. Geol Soc Am Mem 153, 7-31, doi:10.1130/MEM153-p7.

Coney, P. J., and T. A. Harms, 1984, Cordilleran metamorphic core complexes: Cenozoic extensional relics of Mesozoic compression. Geology 12(9), 550-554, doi:10.1130/0091-7613(1984) 12<550:cmccce>2.0.co;2.

Constenius, K. N., 1996, Late Paleogene extensional collapse of the Cordilleran foreland fold and thrust belt. Geol Soc Am Bull 108(1), 20-39, doi:10.1130/0016-7606(1996)108<0020:lpecot>2.3.co;2.

Cosca, M., H. Stunitz, A.-L. Bourgeix, and J. P. Lee, 2011, 40Ar* loss in experimentally deformed muscovite and biotite with implications for 40Ar/39Ar geochronology of naturally deformed rocks, Geochim. Cosmochim. Ac 75(24), 7759-7778, doi:10.1016/j.gca.2011.10.012.

DeCelles, P. G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland basin system, western USA. Am J Sci 304(2), 105-168, doi:10.2475/ajs.304.2.105.

Dickinson,W. R., 2002, The basin and range province as a composite extensional domain. Int Geol Rev 44(1), 1-38, doi:10.2747/0020-6814.44.1.1.

Drury, M. R., and J. L. Urai, 1990, Deformation-related recrystallization processes. Tectonophysics 172(3-4), 235-253, doi:10.1016/0040-1951(90)90033-5.

Druschke, P., A. D. Hanson, and M. L. Wells, 2009, Structural, stratigraphic, and geochronologic evidence for extension predating Palaeogene volcanism in the Sevier hinterland, east-central Nevada. Int Geol Rev 51(7-8), 743-775, doi:10.1080/00206810902917941.

Egger, A. E., T. A. Dumitru, E. L. Miller, C. F. I. Savage, and J. L. Wooden, 2003, Timing and nature of tertiary plutonism and extension in the Grouse Creek Mountains, Utah. Int Geol Rev 45(6), 497-532, doi:10.2747/0020-6814.45.6.497.

Famin, V., P. Philippot, L. Jolivet, and P. Agard, 2004, Evolution of hydrothermal regime along a crustal shear zone, Tinos Island, Greece. Tectonics 23(5), TC5004, doi:10.1029/2003TC001509.

Foster, D. A., P. T. Doughty, T. J. Kalakay, C. M. Fanning, S. Coyner, W. C. Grice, and J. Vogl, 2007, Kinematics and timing of exhumation of metamorphic core complexes along the Lewis and Clark fault zone, northern Rocky Mountains USA.GeolSocAmSpecPap434,207-232, doi:10.1130/2007.2434(10).

Foster, D. A., W. C. Grice, and T. J. Kalakay, 2010, Extension of the Anaconda metamorphic core complex: 40Ar/39Ar thermochronology and implications for Eocene tectonics of the northern Rocky Mountains and the Boulder batholith. Lithosphere 2(4), 232–246, doi:10.1130/l94.1.

Fricke, H. C., S.M.Wickham, and J.R.O’Neil, 1992, Oxygen and hydrogen isotope evidence for meteoric water infiltration during mylonitization and uplift in the Ruby Mountains-East Humboldt Range core complex Nevada. Contrib Mineral Petrol 111, 203–221, doi:10.1007/BF00348952.

Gébelin, A., A. Mulch, C. Teyssier, M. Heizler, T. Vennemann, and N. C. A. Seaton, 2011, Oligo-Miocene extensional tectonics and fluid flow across the Northern Snake Range detachment system, Nevada.

Tectonics 30, TC5010, doi:10.1029/2010TC002797.

Gébelin, A., A. Mulch, C. Teyssier, C. Page Chamberlain, and M. Heizler, 2012, Coupled basin-detachment systems as paleoaltimetry archives of the western North American Cordillera. Earth Planet Sci Lett 335-336, 36-47, doi:10.1016/j.epsl.2012.04.029.

Gébelin, A., A. Mulch, C. Teyssier, M. J. Jessup, R. D. Law, and M. Brunel, 2013, The Miocene elevation of Mount Everest, Geology, 41, 799-802, doi:10.1130/g34331.1.

Gébelin, A., C. Teyssier, M. Heizler, and A. Mulch, 2014, Meteoric water circulation in a rolling-hinge detachment system (northern Snake Range core complex, Nevada). Geol Soc Am Bull 127(1-2), 149-161, doi:10.1130/b31063.1.

Gottardi, R., and C. Teyssier, 2013, Thermomechanics of an extensional shear zone, Raft River metamorphic core complex, NW Utah. J Struct Geol 53, 54-69, doi:10.1016/j.jsg.2013.05.012.

Gottardi, R., C. Teyssier, A. Mulch, T. W. Vennemann, and M. L. Wells, 2011, Preservation of an extreme transient geotherm in the Raft River detachment shear zone. Geology 39(8), 759-762, doi:10.1130/g31834.1.

Gottardi, R., C. Teyssier, A. Mulch, J. W. Valley, M. J. Spicuzza, T. W. Vennemann, A. Quilichinni, and M. Heizler, 2015, Strain and permeability gradients traced by stable isotope exchange in the Raft River detachment shear zone, Utah. J Struct Geol 71, 41-57, doi:10.1016/j.jsg.2014.10.005.

Harrison, T. M., J. Célérier, A. B. Aikman, J. Hermann, and M. T. Heizler, 2009, Diffusion of 40Ar in muscovite. Geochim Cosmochim Ac 73(4), 1039–1051, doi:10.1016/j.gca.2008.09.038.

Hetzel, R., H. Zwingmann, A. Mulch, K. Gessner, C. Akal, A. Hampel, T. Güngör, R. Petschick, T. Mikes, and F. Wedin, 2013, Spatiotemporal evolution of brittle normal faulting and fluid infiltration in detachment fault systems: A case study from the Menderes Massif, western Turkey. Tectonics 32, 364-376, doi:10.1002/tect.20031.

Hoisch, T. D., M. L. Wells, and L. M. Hanson, 2002, Pressure-temperature paths from garnet-zoning:

Evidence for multiple episodes of thrust burial in the hinterland of the Sevier orogenic belt. Am Mineral 87(1), 115-131.

Jessell, M. W., 1987, Grain-boundary migration microstructures in a naturally deformed quartzite. J Struct Geol 9(8), 1007-1014, doi:10.1016/0191-8141(87)90008-3.

Kerrich, R., 1988, Detachment zones of Cordilleran metamorphic core complexes: Thermal, fluid and metasomatic regimes. Geol Rundsch 77(1), 157-182, doi:10.1007/bf01848682.

Konstantinou, A., A. Strickland, E. L. Miller, and J. P. Wooden, 2012, Multistage Cenozoic extension of the Albion–Raft River–Grouse Creek metamorphic core complex: Geochronologic and stratigraphic constraints. Geosphere 8(6), 1429–1466, doi:10.1130/ges00778.1.

Konstantinou, A., A. Strickland, E. Miller, J. Vervoort, C. M. Fisher, J. Wooden, and J. Valley, 2013, Synextensional magmatism leading to crustal flow in the Albion–Raft River–Grouse Creek meta-morphic core complex, northeastern Basin and Range.Tectonics 32, 1384-1403, doi:10.1002/tect.20085.

Lister, G. S., and A. W. Snoke, 1984, S-C mylonites. J Struct Geol 6(6), 617-638, doi:10.1016/0191-8141(84)90001-4

Malavieille, J., 1987, Kinematics of compressional and extensional ductile shearing deformation in a metamorphic core complex of the northeastern Basin and Range. J Struct Geol 9(5-6), 541-554, doi:10.1016/0191-8141(87)90139-8.

Malavieille, J., and A. Taboada, 1991, Kinematic model for postorogenic Basin and Range extension.

Geology 19(6), 555-558, doi:10.1130/0091-7613(1991)019<0555:kmfpba>2.3.co;2.

Manning, A. H., and J. M. Bartley, 1994, Postmylonitic deformation in the Raft River metamorphic core complex, northwestern Utah: Evidence of a rolling hinge. Tectonics 13(3), 596–612, doi:10.1029/93TC03321.

McFadden, R. R., A. Mulch, C. Teyssier, N. C. Seaton, and L. Tokle, 2010, The relationship between microstructure and hydrogen isotopes in the Wildhorse detachment, Pioneer Mountains, Idaho, Abstract

#T41B-2152, AGU Fall Meeting, San Francisco, Ca.

103 References

Menzies, C. D., D. A. H. Teagle, D. Craw, S. C. Cox, A. J. Boyce, C. D. Barrie, and S. Roberts, 2014, Incursion of meteoric waters into the ductile regime in an active orogen. Earth Planet Sc Lett 399, 1-13, doi:10.1016/j.epsl.2014.04.046.

Miller, E. L., A. Konstantinou, and A. Strickland, 2012, Comment on “Geodynamics of synconvergent extension and tectonic mode switching: Constraints from the Sevier-Laramide orogen” by Michael L.

Wells et al. Tectonics 31, TC4015, doi:10.1029/2012TC003103.

Morrison, J., and J. L. Anderson, 1998, Footwall refrigeration along a detachment fault: Implications for the thermal evolution of core complexes. Science, 279(5347), 63-66, doi:10.1126/science.279.5347.63.

Mulch, A., and C. P. Chamberlain, 2007, Stable isotope paleoaltimetry in orogenic belts-The silicate record in surface and crustal geological archives. Rev Mineral Geochem 66(1), 89-118, doi:10.2138 /rmg.2007.66.4.

Mulch, A., and M. A. Cosca, 2004, Recrystallization or cooling ages: In situ UV-laser 40Ar/39Ar geochronology of muscovite in mylonitic rocks.J Geol Soc London 161(4), 573–582, doi:10.1144/0016-764903-110.

Mulch, A., M. A. Cosca, and M. Handy, 2002, In-situ UV-laser 40Ar/39Ar geochronology of a micaceousmylonite: An example of defect-enhanced argon loss. Contrib Mineral Petrol 142(6), 738-752, doi:10.1007/s00410-001-0325-6.

Mulch, A., C. Teyssier, M. A. Cosca, O. Vanderhaeghe, and T. W. Vennemann, 2004, Reconstructing paleoelevation in eroded orogens. Geology 32(6), 525-528, doi:10.1130/g20394.1.

Mulch, A., M. A. Cosca, A. Andresen, and J. Fiebig, 2005, Time scales of deformation and exhumation in extensional detachment systems determined by high-spatial resolution in situ UV-laser 40Ar/39Ar dating.

Earth Planet Sci Lett 233(3-4), 375-390, doi:10.1016/ j.epsl.2005.01.042.

Mulch, A., C. Teyssier, M. A. Cosca, and T. W. Vennemann, 2006, Thermomechanical analysis of strain localization in a ductile detachment zone. J Geophys Res 111, B12405, doi:10.1029/2005JB004032.

Mulch, A., C. Teyssier, M. A. Cosca, and C. P. Chamberlain, 2007, Stable isotope paleoaltimetry of Eocene core complexes in the North American Cordillera. Tectonics 26, TC4001, doi:10.1029/2006TC001995.

Person, M., A. Mulch, C. Teyssier, and Y. Gao, 2007, Isotope transport and exchange within metamorphic core complexes. Am J Sci 307(3), 555-589, doi:10.2475/03.2007.01.

Rey, P. F., C. Teyssier, and D. L. Whitney, 2009, Extension rates, crustal melting, and core complex dynamics. Geology 37(5), 391-394, doi:10.1130/g25460a.1.

Saltzer, S. D., and K. V. Hodges, 1988, The Middle Mountain shear zone, southern Idaho: Kinematic analysis of an early Tertiary high-temperature detachment. Geol Soc Am Bull 100(1), 96-103, doi:10.1130/0016-7606(1988)100<0096:tmmszs>2.3.co;2.

Sheely, J. C., M. L. Wells, and T. L. Spell, 2001, Timing and kinematics of an Eocene-Miocene shear zone:

Grouse Creek Mountains, Utah. Geol Soc Am Abstr Programs 33(6), 148.

Sheppard, S. M. F., 1986, Characterization and isotopic variations in natural waters. Rev Mineral Geochem 16(1), 165-183.

Simpson, C., and S. M. Schmid, 1983, An evaluation of criteria to deduce the sense of movement in sheared rocks. Geol Soc Am Bull 94(11), 1281-1288, doi:10.1130/0016-7606(1983)94<1281:aeoctd>2.0.co;2.

Sonder, L. J., and C. H. Jones, 1999, Western United States extension: How the West was widened. Annu Rev Earth Planet Sci 27(1), 417-462, doi:10.1146/annurev.earth.27.1.417.

Stipp,M., H. Stünitz, R. Heilbronner, and S. M. Schmid, 2002, The eastern Tonale fault zone: A‘natural laboratory’ for crystal plastic deformation of quartz over a temperature range from 250 to 700°C. J Struct Geol 24(12), 1861-1884, doi:10.1016/S0191-8141(02)00035-4.

Strickland, A., E. L. Miller, and J. L. Wooden, 2011a, The timing of Tertiary metamorphism and deformation in the Albion–Raft River–Grouse Creek metamorphic core complex, Utah and Idaho. J Geol 119(2), 185-206, doi:10.1086/658294.

Strickland, A., E. L. Miller, J. L. Wooden, R. Kozdon, and J. W. Valley, 2011b, Syn-extensional plutonism and peak metamorphism in the Albion–Raft River–Grouse Creek metamorphic core complex. Am J Sci 311(4), 261-314, doi:10.2475/04.2011.01.

Sullivan, W. A., 2008, Significance of transport-parallel strain variations in part of the Raft River shear zone, Raft River Mountains, Utah, USA. J Struc Geol 30(2), 138-158, doi:10.1016/j.jsg.2007.11.007.

Sullivan, W. A., and A. W. Snoke, 2007, Comparative anatomy of core-complex development in the northeastern Great Basin, U.S.A. Rocky Mt. Geol 42(1), 1-29, doi:10.2113/gsrocky.42.1.1.

Suzuoki, T., and S. Epstein, 1976, Hydrogen isotope fractionation between OH-bearing minerals and water.

Geochim Cosmochim Ac 40(10), 1229-1240, doi:10.1016/0016-7037(76)90158-7.

Taylor, H. P., Jr., 1978, Oxygen and hydrogen isotope studies of plutonic granitic rocks. Earth Planet Sci Lett 38(1), 177-210, doi:10.1016/0012-821X(78)90131-0.

Teyssier, C., E. C. Ferré, D. L. Whitney, B. Norlander, O. Vanderhaeghe, and D. Parkinson, 2005, Flow of partially molten crust and origin of detachments during collapse of the Cordilleran orogen. Geol Soc London Spec Publ 245(1), 39-64, doi:10.1144/gsl.sp.2005.245.01.03.

Vanderhaeghe, O., C. Teyssier, I. McDougall, and W. J. Dunlap, 2003, Cooling and exhumation of the Shuswap Metamorphic Core Complex constrained by 40Ar/39Ar thermochronology. Geol Soc Am Bull 115(2), 200-216, doi:10.1130/0016-7606(2003)115<0200:caeots>2.0.co;2.

Vogl, J. J., D. A. Foster, C. M. Fanning, K. A. Kent, D. W. Rodgers, and T. Diedesch, 2012, Timing of extension in the Pioneer metamorphic core complex with implications for the spatial-temporal pattern of Cenozoic extension and exhumation in the northern U.S. Cordillera. Tectonics 31, TC1008, doi:10.1029/2011TC002981.

Wells, M. L., 1997, Alternating contraction and extension in the hinterlands of orogenic belts: An example from the Raft River Mountains, Utah. Geol Soc Am Bull 109(1), 107-126, doi:10.1130/0016-7606(1997)109<0107:acaeit>2.3.co;2.

Wells, M. L., 2001, Rheological control on the initial geometry of the Raft River detachment fault and shear zone, western United States. Tectonics 20(4), 435-457, doi:10.1029/2000TC001202.

Wells, M. L., and T. D. Hoisch, 2012, Reply to comment by E. L. Miller et al. on “Geodynamics of synconvergent extension and tectonic mode switching: Constraints from the Sevier-Laramide orogen”.

Tectonics 31, TC4016, doi:10.1029/2012TC003136.

Wells, M. L., R. D. Dallmeyer, and R. W. Allmendinger, 1990, Late Cretaceous extension in the hinterland of the Sevier thrust belt, northwestern Utah and southern Idaho. Geology 18(10), 929-933, doi:10.1130/0091-7613(1990)018<0929:lceith>2.3.co;2.

Wells, M. L., T. D. Hoisch, M. T. Peters, D.M. Miller, E. D.Wolff, and L.M. Hanson, 1998, The Mahogany Peaks fault, a Late Cretaceous-Paleocene(?) Normal fault in the Hinterland of the Sevier Orogen. J Geol 106(5), 623-634, doi:10.1086/516046.

Wells, M. L., L. W. Snee, and A. E. Blythe, 2000, Dating of major normal fault systems using thermochronology: An example from the Raft River detachment, Basin and Range, western United States. J Geophys Res 105(B7), 16,303-16,327, doi:10.1029/2000JB900094.

Wells, M. L., J. C. Sheeley, T. L. Spell, E. D. Kelly, and T. D. Hoisch, 2004, Eocene extension in northwestern Utah-southern Idaho: Early motion on the polyphase Middle Mountain shear zone. Rocky Mountain (56th Annual) and Cordilleran (100th Annual) Joint Meeting (May 3–5, 2004).

Wells, M. L., T. D. Hoisch, A. M. Cruz-Uribe, and J. D. Vervoort, 2012, Geodynamics of synconvergent extension and tectonic mode switching: Constraints from the Sevier-Laramide orogen. Tectonics 31, TC1002, doi:10.1029/2011TC002913.

Wernicke, B., 1981, Low-angle normal faults in the Basin and Range Province: Nappe tectonics in an extending orogen. Nature 291(5817), 645-648, doi:10.1038/291645a0.

Whitney, D. L., C. Teyssier, P. Rey, and W. R. Buck, 2013, Continental and oceanic core complexes. Geol Soc Am Bull 125(3/4), 273-298, doi:10.1130/b30754.1.

Wickham, S. M., M. T. Peters, H. C. Fricke, and J. R. O’Neil, 1993, Identification of magmatic and meteoric fluid sources and upward- and downward-moving infiltration fronts in a metamorphic core complex. Geology 21(1), 81-84, doi:10.1130/0091-7613(1993)021<0081:iomamf>2.3.co;2.

Wong, M. S., J. Singleton, J. Baughman, and K. C. Bunting, 2013, Evidence for Miocene reactivation of a late Cretaceous to early Tertiary shear zone in the Harcuvar and Buckskin-rawhide metamorphic core complexes, Arizona. GSA Annual Meeting, pp. 523, Geol Soc of Am, Denver.

105 Figures

Figures

Fig. 1. (A) Simplified map of the western United States showing Cenozoic extension (dark grey regions) and metamorphic core complexes (red). Modified from Foster et al.

(2007) and Sonder and Jones (1999), (B) simplified geological map of the Raft River-Albion-Grouse Creek Metamorphic Core Complex with the top-to-the-east Raft River detachment shear zone (RRDSZ) in the Raft River Mountains and the top-to-the-WNW Middle Mountain shear zone in the Albion and Grouse Creek Mountains (modified from Wells (1997),Wells et al. (2012), and Konstantinou et al. (2013)).White dots mark Pine Creek Canyon, Clear Creek Canyon, and Ten Mile Canyon sampling localities.

Fig. 2. (A) Section of variably deformed Elba Quartzite at Pine Creek Canyon showing the location of hydrogen isotope (circles), 40Ar/39Ar geochronology (stars), and oxygen isotope exchange thermometry (pentagons) samples. (B) Range of typical δ2Hms values indicative of interaction with meteoric and metamorphic fluids (notice the different scale compared to Fig.2C). (C) Muscovite hydrogen isotope values (δ2Hms) of Elba Quartzite at Pine Creek Canyon. Multiple data points show replicates and different grain size fractions.

Fig. 3. (A) Simplified geological map and cross section of the Raft River Mountains (with sampling localities: B = Pine Creek Canyon, C = Clear Creek Canyon, and D = Ten Mile Creek) adapted from Wells et al. (2000). (B–D) Sections through Elba Quartzite with δ2Hms values (versus VSMOW) from mylonitic quartzite (grey circles), mica-rich schist (orange circles), regolith (grey square), oxygen exchange-based temperatures (orange pentagons), and 40Ar/39Ar data (yellow stars) from each sampling locality: (Fig. 3B) Pine Creek Canyon, (Fig. 3C) Clear Creek Canyon, and (Fig. 3D) Ten Mile Creek Canyon. The composite section as well as δ2Hms values from mylonitic quartzite (grey circles) of the Clear Creek Canyon are adapted from Gottardi et al. (2011). The succession of profiles from west to east reveals the thinning of the Elba Quartzite in the direction of shearing.

Fig. 4. Macroscopic and microscopic structures of the deformed Elba Quartzite at Pine Creek Canyon: (a) Stereogram of foliation poles (circles) and stretching lineation (triangles). (b) Mylonitic foliation with C′-shear bands indicating top-to-the east sense of shear. (c) Flat and elongate muscovite parallels C-plane with very high aspect ratios;

grain-size reduction of quartz occurs preferentially in mica-rich parts (RR-07-25, 125 m above basement). (d) Grain boundary bulging and migration at the boundaries of large quartz grains (arrows) and castellated grain shape of quartz (box) (RR-07-34, 207 m). (e) Deformation lamellae indicative of high differential stress in large quartz grains (RR-07-34, 207 m). Castellate grain boundaries (arrow 1) and grain boundary pinning as well as aligned fluid inclusions (arrow 2) point to mobile grain boundaries. (f) Elba Quartzite sample with high-temperature grain boundary migration recrystallization microstructures, lacking evidence for low-temperature deformation microstructures such as grain boundary bulging, deformation bands, deformation lamellae, or undulose extinction (RR91-20, 75 m).

109 Figures

Fig. 5. 40Ar/39Ar age spectra of white mica from (a) Pine Creek Canyon and (b) Clear Creek Canyon (CC) and Ten Mile Canyon (TM) localities. Black spectra are bulk furnace heating results (Wells et al., 2000), white and grey spectra are laser heating experiments of 2–4 grains (this study). Results from samples of this study do not form 40Ar/39Ar plateau ages by any standard criteria, rather preferred weighted mean ages for each sample are reported from contiguous incremental heating steps with > 50 % of 39Ar released that have similar 40Ar*/39Ar ratios.

Fig. 6. Comparison of δ2Hms values of white mica from Pine Creek Canyon (blue), Clear Creek Canyon (green), and Ten Mile Canyon (yellow) across the Raft River detachment shear zone. Part of the δ2Hms data of the Clear Creek Canyon is adapted from Gottardi et al. (2011) (see Fig. 3 and Appendix).

111 Figures

Fig. 7. Tectonic models illustrating the exhumation and fluid flow history of the RAG-MCC (adapted and modified from Wells [2001]). Based on microstructural, 40Ar/39Ar geochronological, and stable isotopic evidence, we suggest two different exhumation and fluid infiltration events with distinct fluid compositions: (a) Mid-Eocene (45–40 Ma) exhumation and top-to-the-east shearing was accompanied by infiltration meteoric fluids with very low δ2H values (δ2Hffluid ≥ -126 ‰) resulting in δ2Hms values as low as -154 ‰, (b) Mid-Miocene extensional top-to-the-east shearing along the Raft River detachment fault led to localized hydrogen isotope exchange with meteoric water attaining δ2Hfluid

values of ≥ ~ -100 ‰ resulting in δ2Hms ≥ -125 ‰. PC = Pine Creek; CC = Clear Creek;

TM = Ten Mile Canyon. Ar Tc = Closure temperature to 40Ar diffusion in muscovite;

BDT = brittle-ductile transition of quartz.

113

Chapter 5

Eo-Oligocene proto-Cascades topography revealed by clumped ( ∆

47

) and oxygen isotope ( δ

18

O) geochemistry (Chumstick Basin, WA, USA)

Methner K., J. Fiebig, U. Wacker, P. Umhoefer, C. Page Chamberlain, A. Mulch Submitted to: Tectonics, August 2015, pending revisions

Abstract

The topography of the present-day Washington Cascades impacts atmospheric circulation and precipitation patterns in the Pacific Northwest, introducing a pronounced orographic rain-shadow in the lee of the mountain range. The temporal development of Cascade topography, however, remains largely unconstrained for the Early Cenozoic. Based on coupled clumped isotope (∆47) and oxygen isotope (δ18O) measurements we present δ18O values of Eocene near surface groundwater (δ18Owater) from the Chumstick basin (central Washington) located in the present-day Cascade rain shadow. ∆47 (paleo)thermometry indicates a systematic change in basin burial temperatures from 110°C to 70°C depending on burial depth in the basin. These data are in good agreement with low-T thermochronological and vitrinite reflectance data, and further constrain the basin burial and exhumation history. In concert with field observations, microstructural analysis, and δ18O values of the analyzed carbonates, we suggest that the ∆47 temperatures reflect carbonate formation early during burial diagenesis from near-surface groundwater.

Reconstructed mean δ18Owater values of -7 ‰ (mid-Eocene) to ~ -9 ‰ (late Eocene/early Oligocene) are consistent with a low-elevation source of the corresponding meteoric waters that permeated the sandstone/conglomerate members of the Eocene/Oligocene sedimentary units. In light of the paleogeographic setting of the Chumstick Basin, the reconstructed δ18Owater values agree well with Pacific-derived moisture that did not experience strong rainout. The absence of an Eocene/Oligocene rain shadow effect therefore permits only moderate elevation at least for the southern part the Washington proto-Cascades.

1. Introduction

The topography of high mountain ranges exerts a strong impact on atmospheric circulation and precipitation patterns and is a key element in understanding the interplay of Earth surface processes and lithospheric-scale geodynamics. The topographic history of the world’s major orogens, therefore, remains one of the key challenges when discussing the interactions among tectonics, climate, and Earth surface processes (e.g., Ruddiman and Kutzbach, 1989; Molnar and England, 1990; Chamberlain et al., 2012).

The Cascade Range in western North America is a prominent example where present-day topography induces a strong orographic rain-shadow in the lee of the mountain range (e.g., Siler et al., 2012). Today the Cascades interfere with Pacific-derived air masses and focus precipitation along their western flanks while enforcing dry climate conditions in the eastern continental interior. Thus, the spatial and temporal development of topography and relief in the Cascades and their interaction with Pacific-derived moisture over time has gained continuous attention (e.g., Smith, 1903; Smiley, 1963; Leopold and Denton, 1987; Kohn et al., 2002; Reiners et al., 2002, 2003; Takeuchi and Larson, 2005; Kohn and Law, 2006; Mitchell and Montgomery, 2006; Mitchell et al., 2009; Mustoe and Leopold, 2014; see also compilation in Mitchell and Montgomery, 2006), yet the early Cenozoic topographic history of the Cascades is still largely unconstrained.

Paleotopography and paleoclimate (modeling) studies of the early Cenozoic North American Cordillera mostly focused on the Sevier fold and thrust belt and the Sevier hinterland. Collectively stable isotope paleoaltimetry and paleobotanical data reveal that extensive regions of western North America occupied a high elevation, plateau configuration by the end of the Eocene (e.g., Chase et al., 1998; Wolfe et al., 1998;

Horton et al., 2004; Mulch et al., 2004, 2006, 2007; Fan and Dettman, 2009; Hren et al., 2010; Huntington et al., 2010; Mix et al., 2011; Chamberlain et al., 2012; Feng et al., 2013). The location of the western plateau margin, however, is still poorly constrained. A proto-Cascade range, therefore, might have been the “western edge” of a Cordilleran plateau or alternatively, represented arc-/wedge-related topography above the Pacific subduction system. Systematic changes in the stable oxygen (δ18O) and hydrogen (δ2H) isotope compositions of precipitation provide a robust tool to reconstruct paleoelevation of mountain ranges if the isotopic composition of orographic precipitation is preserved in the geological record and the competing effects of topography, climate, and atmospheric circulation on stable isotopes in precipitation can be adequately assessed (e.g., Chamberlain and Poage, 2000; Garzione et al., 2000; Mulch et al., 2004; Quade et al., 2007; Rowley and Garzione, 2007; Mulch, in rev.). Stable isotope paleoaltimetry is based on the systematic decrease of δ18O and δ2H values in precipitation as air-masses ascend and traverse an orographic barrier. As a result, the windward flanks of many orogens

115 1. Introduction

typically show a change in δ18O with elevation (e.g., Ingraham and Taylor, 1991;

Garzione et al., 2000; Schemmel et al., 2013) that averages -0.28 ‰/100 m (Poage and Chamberlain, 2001). Along the North American Pacific margin, similar systematic relationships are documented in the Sierra Nevada and Washington Cascades, both regions where the average elevation scales with δ18O of precipitation. This isotopes in precipitation effect translates into the leeward rain shadow of these ranges (e.g., Kendall and Coplen, 2001; Dutton et al., 2005).

Paleogene oxygen isotope (δ18O) precipitation maps still lack records from the Pacific Northwest and thus are unable to constrain the E-W extent of the Eocene plateau (c.f., Mix et al., 2011; Chamberlain et al., 2012). Therefore, stable isotope paleoaltimetry records from the Pacific Northwest would substantially contribute to our understanding of early Cenozoic patterns of surface uplift in the North American Cordillera. Decreasing δ18O values in Mio-/Pliocene authigenic clays from paleosols and altered tuffs, however, suggest that surface uplift of the present-day Washington Cascades started no later than

~16 Ma ago and continued until ~ 4 Ma (Takeuchi and Larson, 2005), potentially some 10 m.y. after onset of rain shadow conditions in the Oregon Cascades (Kohn et al., 2002).

Surface uplift of the Washington Cascades leading to present-day Cascade topography has been suggested to have started in the middle to late Miocene (Reiners et al., 2002;

Takeuchi and Larson, 2005) or even as late as the Pliocene (Mustoe and Leopold, 2014).

It also has been suggested that the Washington Cascades experienced a complex polygenetic topographic history with spatial differences between the southern (south of Snoqualmie Pass) and the northern Washington Cascades, which might have already been at higher elevations during the Miocene (Mitchell and Montgomery, 2006).

Here, we provide a δ18O record of near-surface groundwater (δ18Owater) as preserved in terrestrial sedimentary sequences of the Eocene Chumstick basin, which today is located in the rain shadow of the northern Washington Cascades. Combining information on the basin burial history (Evans, 1988, 1994; Enkelmann et al., 2015) we use clumped isotope (∆47) thermometry to reconstruct δ18O values of Eocene groundwater in the Chumstick basin. Our ∆47 thermometry (T(∆47)) results indicate that the Chumstick basin sediments experienced carbonate growth at temperatures between 70 °C and 110 °C depending on burial depth, a result in accord with thermochronological and vitrinite reflectance data. In concert with field observations and microstructural analysis of the analyzed carbonates, we suggest that these carbonates formed early during burial diagenesis from near-surface groundwater. Combining T(∆47) with δ18O values of the carbonates (δ18Ocarbonate), the resulting fluid δ18O values indicate a meteoric origin of these near-surface groundwaters with δ18Owater = -5.2 ‰ to -11.3 ‰. When compared to the Miocene precipitation records,

such δ18Owater values agree well with Pacific-derived moisture that did not experience strong orographic rainout.

2. Geological setting

Following mid-Cretaceous crustal shortening, thrusting and pluton emplacement along the southern Coast Belt (Washington Cascades and British Columbia), plate boundary dynamics changed from arc-normal contraction to transpression and transtension due to northward oblique convergence of the Kula plate, which resulted in dextral strike-slip and thrust faulting starting at about ~85 Ma (e.g., Johnson, 1985; Umhoefer and Miller, 1996;

Paterson et al., 2004). The resulting dextral fault systems were active until the end of the Paleocene, accounting for a total offset of about 215 km to 255 km (Umhoefer and Miller, 1996). After this transpressive phase, Eocene deformation was dominated by dextral faulting and extension (e.g., Miller and Bowring, 1990; Umhoefer and Miller, 1996), leading to basin development in the North Cascades region during the Paleogene (Johnson, 1985; Heller et al., 1987).

The Chumstick basin, situated to the east of the modern Cascade Range (Fig. 1a), is an Eocene fault-bounded terrestrial basin, confined by the Entiat Fault Zone (EFZ) to the east and the Leavenworth Fault Zone (LFZ) to the west. It is separated into two subbasins by the Eagle Creek Fault Zone (ECFZ) (Fig. 1b). To the south, the basin termination is covered by the middle Miocene Columbia River Basalt Group (CRBG) and it has been speculated that the Chumstick sediments continue beneath the CRBG (Evans, 1988).

Models of the basin evolution include (a) an (oblique) strike-slip origin (pull-apart basin) (e.g., Johnson, 1985; 1996; Taylor et al., 1988), (b) a two-stage model comprising extensional half-graben formation bounded to the east by the ECFZ (> 51 Ma) followed by dextral strike-slip deformation (approximately 44-42 Ma) (Evans, 1994), and (c) a regional syncline without Eocene extension-related basin formation (Cheney and Hayman, 2009).

Chumstick Formation (Fm.) sediments comprise syntectonically deposited fluvial, alluvial and lacustrine units mainly derived from proximal sources (Gresens et al., 1981;

Johnson, 1985; Evans, 1988, 1994). Deposition of the Chumstick Fm. was rapid and largely tectonically controlled resulting in apparently large sedimentary thicknesses.

However, the total stratigraphic thickness of the Chumstick Fm. remains debated.

Thermal maturity (Evans, 1988) and geophysical data (Silling, 1979) suggest a basin thickness of ≤ 2 km, whereas the composite stratigraphic thickness is reported to be 5.2 km (Cheney and Hayman, 2009), ~ 8 km (Gresens et al., 1981), or up to 12 km due to the lateral offset of sediment source regions during deposition (Fig. 1c) (Evans, 1988, 1991a).

117