<<

Two-stage formation of Death

Ian Norton* Jackson School of Geosciences, Institute for Geophysics, University of Texas at Austin, J.J. Pickle Research Campus, Bldg. 196, 10100 Burnet Road (R2200), Austin, Texas 78758-4445, USA

ABSTRACT lar to the core complex at Tucki , (1966), on the basis of the morphology of the at the northern end of the range. The Basin valley and occurrence of strike-slip faults, sug- Extension in is usually inter- and Range extensional detachment tracks gested that Death Valley was formed as a strike- preted as a combination of low-angle Basin over the top of the range, with extensional slip pull-apart basin, with the basin forming and Range–style extension and strike slip allochthons perched on the eastern fl anks of between the Northern Death Valley–Furnace associated with the developing Pacifi c-North the range. This modifi ed model for evolution Creek to the north and the Southern America plate boundary in western North of Death Valley suggests a strong link between to the south (Fig. 2). America, with these two tectonic regimes timing and style of deformation in the basin This idea has been incorporated into several operating synchronously in Death Valley. with the developing Pacifi c- models for the structural evolution of Death Examination of structural, stratigraphic, plate boundary, particularly eastward propa- Valley. Miller and Pavlis (2005) divided these and timing relationships in the region sug- gation of this boundary. models into two categories, depending on how gests that this interpretation needs revision. strike-slip and low-angle detachment faulting Evolution of Death Valley is best described INTRODUCTION are combined. The fi rst category described by as a two-stage process. In the fi rst stage, last- Miller and Pavlis (2005) is the “Rolling Hinge” ing from ca. 18 to 5 Ma, low-angle Basin and Interpretations of the geology of Death Val- model as proposed by Stewart (1983), Hamil- Range extension transported allochthons ley have played an important role in the devel- ton (1988), Wernicke et al. (1988a), and Snow consisting of Late Proterozoic through Early opment of models of continental extension, and Wernicke (2000). In these models, the low- Paleozoic miogeoclinal section along detach- particularly for models that incorporate large- angle detachment faults are dominant, with the ment surfaces that, as extension continued, magnitude extension accommodated by low- moving 80 km west from an were exhumed from mid-lower crustal levels angle detachment faults (Wright and Troxel, original position east of the Black to the surface. Near the end of this extensional 1973; Hamilton, 1988; Wernicke et al., 1988a; and strike-slip faults as upper crustal edges of phase and lasting until ca. 3 Ma, deposition Snow and Wernicke, 2000; Hayman et al., the detachment system. In the second category, of a thick sequence of volcanics, clastics, and 2003). Detachment fault surfaces are exposed in based on the pull-apart model developed by some lacustrine carbonates signaled a period several places in the Death Valley region, with Burchfi el and Stewart (1966), strike-slip faults of relative tectonic quiescence, with sediments the best studied being the antiformal structures penetrate deeply into the crust and drive exten- in some areas covering the exhumed detach- known as turtlebacks in the Black Mountains sion between their terminations (Wright and ment surfaces. At ca. 3 Ma, initiation of the on the east side of the valley (Figs. 1 and 2; Troxel, 1984; Topping, 1993; Serpa and Pavlis, East Shear Zone started develop- Miller and Pavlis, 2005). The detachment fault 1996; Miller and Prave, 2002). The low-angle ment of the present-day topographic depres- in the Black Mountains is known as the Amar- detachment surfaces in the latter category are sion of Death Valley, formed as a pull-apart gosa detachment (Wright et al., 1974). Another normal faults linking the strike-slip faults. The basin associated with this strike slip. Faulting important detachment surface is exposed at pull-apart concept has also been applied to the broke the older, inactive, Basin and Range Tucki Mountain, located at the north end of , the basin on the west side detachment surfaces, with high-angle trans- the Panamint Range on the west side of Death of the Panamint Range, which links via the tensional faulting along the Black Mountains Valley (TM on Fig. 2; Hodges et al., 1987; Hunter Mountain strike-slip fault northwards front. The Black Mountains were elevated as Wernicke et al., 1988b). Stewart (1983) proposed into (Burchfi el et al., 1987; Lee a result of footwall uplift, with the well-known that the Panamint Range forms the hanging wall et al., 2009). In all current models, strike-slip turtleback structures being megamullions of an extensional system that transported the and low-angle extensional faults are regarded along these bounding faults. These mega- Panamint Range westward from on top or east as synchronous. mullions are similar to those seen at oceanic of the Black Mountains, with the latter form- To understand more about structural evolu- spreading centers. The Panamint Range has ing the footwall of this extensional system. In tion of Death Valley, it is useful to consider the previously been interpreted as an extensional this interpretation, motion was accommodated area’s position in relation to structural prov- allochthon, with the entire range transported along the Amargosa detachment fault. A predic- inces that have been defi ned in western North from on top of or east of the Black Moun- tion of this model is that there is a detachment America. Death Valley is located in three partly tains. A new interpretation presented here is fault underneath the Panamint Range; this will overlapping structural provinces (Fig. 1). It that the range is a large core complex simi- be addressed later in this paper. is located in the Basin and Range extensional Prior to recognition of low-angle extensional province (Sonder and Jones, 1999; Burchfi el *E-mail: [email protected] faulting in Death Valley, Burchfi el and Stewart et al., 1992), in the belt (Stewart,

Geosphere; February 2011; v. 7; no. 1; p. 171–182; doi: 10.1130/GES00588.1; 11 fi gures; 1 table.

For permission to copy, contact [email protected] 171 © 2011 Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

120°W 118°W 116°W Although the ECSZ today accommodates 25% of Pacifi c-North America relative motion, timing of initiation of this shear zone is poorly known. Dokka and Travis (1990b) favored an 38°N Sierra Basin and Range 38°N initiation age of 10–6 Ma for the Mojave por-

WM tion of the ECSZ. In the White Mountains northwest of Death Valley, Stockli et al. (2003) found two main phases of tectonism. In the Walker fi rst phase, up to 8 km of uplift of the White IR Lane Mountains (WM, Fig. 1) occurred in the Middle Miocene as a result of footwall uplift associated with east-west extension. In the second phase, Death transtensional faulting began east of the White Valley

36°N San Andreas Mountains at 6 Ma, followed at 3 Ma by initia-

NV 36°N CA tion of strike slip in the Owens Valley (Fig. 1) on the west side of the White Mountains. The Inyo FaultEastern Mountains, south of the White Mountains, are California bounded to the east by the East Inyo fault zone, Garlock Shear which links via the Hunter Mountain fault to the Fault Zone Panamint Valley. Lee et al. (2009), in a detailed analysis of the Inyo Mountains, report a phase of normal faulting starting at 15.6 Ma and initia- tion of strike-slip faulting on the Hunter Moun-

34°N tain fault at 2.8 Ma. The Hunter Mountain fault SJFZ 34°N Elsinore FZ connects to strike-slip faults on the west side of the Panamint Range; Burchfi el et al. (1987) docu- Salton ment 8–10 km of offset on this fault system. Trough These recent studies indicate that strike-slip faulting and extension occurred in separate 120°W 118°W 116°W phases. In this paper, I suggest that Death Valley also formed in two phases, with low-angle Basin Figure 1. Regional topography of western North America, showing and Range–style extension in the Miocene fol- location of Death Valley and tectonic domains. WM—White Moun- lowed by pull-apart basin development during tains; IR—Inyo Range. Topography is gray-shaded with illumina- strike-slip deformation in the last three million tion from the north for areas above sea level. Areas below sea level years. This inference is based on a compilation are in blue. Data from Smith and Sandwell (1997). of age data for the Death Valley region, com- bined with a new interpretation of the geometry of detachment surfaces. In this new interpreta- 1988), and also in the Shear of these domains, satellite geodesy (GPS) has tion, a detachment surface tracks over the crest Zone (ECSZ; Dokka and Travis, 1990a). The shown that the Walker Lane belt and ECSZ are of the Panamint Range, rather than underneath Basin and Range is characterized by block together accommodating ~25% of present-day the range, and the Black Mountains turtlebacks faulting and, in areas of large-magnitude exten- plate motion between the Pacifi c and North are megamullion structures formed as a result of sion like Death Valley, by exhumation of middle American plates (Miller et al., 2001; Hammond strike slip. The rocks now exposed in the turtle- and lower crustal level rocks in core complexes and Thatcher, 2007). The Walker Lane belt and backs were exhumed from deep crustal levels as (Davis, 1980; Armstrong, 1982; Stewart, 1998; ECSZ therefore together form a single strike- a result of Basin and Range extension; a thick Dickinson, 2002). Both the Walker Lane and slip zone, which will be referred to as the ECSZ series of Pliocene and younger sediments were ECSZ are characterized by north- to northwest- in this paper. This shear zone developed as a deposited on these surfaces before renewed trending topography with active dextral strike- result of the evolving tectonics of western North uplift and exhumation to the present-day surface slip and normal faults (Oldow et al., 1994, 2008; America, particularly eastward propagation of as a result of formation of the Death Valley pull- Henry et al., 2007; Lee et al., 2009; Dokka and the boundary between the Pacifi c and North apart basin. Travis, 1990a, 1990b). The Walker Lane belt America plates. A recent phase of this propaga- was originally defi ned by Stewart (1988) as tion was opening of the Gulf of California and GEOLOGIC FRAMEWORK OF THE the triangular zone of high topographic relief initiation of the as the plate DEATH VALLEY REGION separating the from the Great boundary at 5–6 Ma (Lonsdale, 1989; Atwater Basin, with its southern limit the . and Stock, 1998). This eastward propagation When Burchfi el and Stewart (1966) fi rst sug- The ECSZ was originally defi ned by Dokka continues today with strike slip in the ECSZ, gested that Death Valley was formed as a pull- and Travis (1990a) based on mapping of strike- which will become the full plate boundary if apart basin, little was known about timing of slip faults in the Mojave , and included the San Andreas fault ceases activity in western tectonic events in the region. Their paper was the Death Valley area (i.e., north of the Gar- California and the plate boundary continues its published in the same year as Hunt and Mabey’s lock fault) as its northern limit. Since defi nition eastward migration (Faulds et al., 2005). (1966) seminal paper that mapped the geology

172 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Two-stage formation of Death Valley

118°°W 117°°W 116°°W the detachment surfaces now exposed in the 37°°N 37°°N NDVFZ Basin and Range Black Mountains (Holm et al., 1992; Holm and Owens Valley IFZ GM Dokka, 1993) and, outside Death Valley, exten- sion in the Cottonwood Mountains (Snow and CM SWNVF SV Lux, 1999), the Inyo Range (Lee et al., 2009), IR and the Kingston Peak–Tecopa Basin area HM HMF (Calzia and Rämö, 2000). Neogene tectonic TM activity reported by Lee et al. (2009) consists Death Valley of exhumation and normal faulting initiated FCFZ Sierra Nevada SM along the eastern Inyo fault zone (IFZ in Fig. 2) DP at 15.6 Ma, followed by a period of slow uplift PR PV BM AR then renewed normal slip at 2.8 Ma. The event NR that commenced at 15.6 Ma included 16° west- Tecopa ward tilt of the Inyo Range and 5300 m of uplift 36°°N NV 36°°N Basin of the eastern fl ank, at a rapid rate of 1.9 mm/a. CA SDVFZ Lee et al. (2009) suggest that the later event at KP 2.8 Ma dates onset of strike-slip faulting along the Hunter Mountain fault zone and initiation of strike slip in this portion of the ECSZ. This two- AM stage structural evolution of extensional faulting Garlock Fault followed by strike slip is also seen in the Pana- mint Valley (Andrew and Walker, 2009) and, 118°°W 117°°W 116°°W north of the Inyo Range, in the White Mountains Figure 2. Digital elevation model (DEM) topography of the Death Valley region, from USGS (Stockli et al., 2003) and central Walker Lane DEM (2009). Faults are from the USGS faults database (2009). Topography is belt (Oldow et al., 2008). colored with a nonlinear scale ranging from pale blues below sea level through greens and Interpretations of tectonic settings in Figure 3 browns to white of the Sierra Nevada peaks. The Nevada–California state line is labeled NV/ come from Snow and Lux (1999) for formations CA. Labels are colored for visibility only. AM—Avawatz Mountains; AR—; in the Cottonwood Mountains (Tub, Tpn, Tnv), BM—Black Mountains; CM—Cottonwood Mountains; DP—Darwin Plateau; DV—Death from Knott et al. (2005) for the Ubehebe– Valley; FM—; FCFZ—; GM—Grapevine Rogers Formation (U-LR), also in the Cotton- Mountains; HM—Hunter Mountain; HMF—Hunter Mountain fault; IFZ—Inyo fault woods area, and from Snow and Lux (1999) for zone; IR—Inyo Range; KP—Kingston Peak; NDVFZ—Northern Death Valley fault zone; the Bat Mountain Formation. The Bat Moun- NR—; OV—Owens Valley; PR—Panamint Range; PV—Panamint Valley; tain Formation is an interlayered conglomerate SM— Mountain; SDVFZ—Southern Death Valley fault zone; SV—Saline Valley; and sandstone section found at the east end of SWNVF—Southwest Nevada Volcanic Field; TM—Tucki Mountain. the Funeral Mountains (Fig. 2; Cemen et al., 1999). It overlies Mississippian carbonates and has not been directly dated. It overlies a 19.8 ± 0.2 Ma (K/Ar age) and is in turn overlain by of Death Valley. This paper established a robust 2009), and tectonic and sediment fi ll compila- a 13.7 ± 0.4 Ma (K/Ar age) tuff (Cemen et al., stratigraphy, including detailed relative stratig- tions (Blakely and Ponce, 2002). 1999). This age is an important constraint on ini- raphy of Neogene sediments. These sediments Figure 3 is a tectonostratigraphic chart show- tiation of Basin and Range extension, as the Bat are important recorders of tectonic develop- ing a compilation of Neogene geologic events Mountain Formation is interpreted to have been ment of Death Valley, but their chronostratigra- in the Death Valley region. The area covered deposited in a basin formed during the early phy was poorly known in 1966 because of the ranges from Kingston Peak to the southeast of stages of extension (Cemen et al., 1999). A litho- mostly endemic fauna in these deposits and lim- Death Valley, northeast to the Southwest Nevada logic correlation with a similar conglomerate- ited radiometric age control. Recent work, espe- Vol canic Field (SWNVF), and as far west as the sandstone succession at Tucki Mountain has cially tephrochronology (Sarna-Wojcicki et al., Darwin Plateau on the west side of Panamint also been used to constrain both the age and 2001) and argon-argon radiometric dating, has Valley. For the SWNVF, Sawyer et al. (1994) amount of extension along the Furnace Creek greatly increased our knowledge of the ages of provide an interpretation of eruption volume ver- fault zone (Wernicke et al., 1988b). Cemen these deposits. Summaries have been published sus time, which the sawtooth pattern in Figure 3 et al. (1999) question the correlation because by Snow and Lux (1999), Calzia and Rämö schematically follows. Ash layers, important for the Funeral Mountains section is underlain by a (2000), and Knott et al. (2005). Also in the last age control on sediments, are included in the Miocene conglomerate unit that does not occur few years, several important data sets relevant to igneous activity column, although the volcanism at Tucki Mountain and there are differences in a regional understanding of the geology of the that produced these ash beds was usually well structures in the underlying Paleozoic section Death Valley area have become available. These outside the Death Valley area (Knott et al., 2005). (a thrust fault in the Mississippian strata in the data sets include digital geologic maps pub- The last two columns in Figure 3 show tec- Funeral Mountains is not seen at Tucki Moun- lished by the U.S. Geological Survey (USGS) tonic events and interpreted tectonic regimes. tain). These observations imply that details of (Workman et al., 2002) and the California Geo- Events associated with Basin and Range exten- the structural connection between Tucki Moun- logic Survey (Saucedo et al., 2000), digital ele- sion occurred in the Miocene to very early tain and the east end of the Funeral Mountains vation model (DEM) topography (USGS DEM, Pliocene. These events include exhumation of must be reexamined, although a probable Early

Geosphere, February 2011 173

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

Figure 3. Tectonostratigraphic chart for the Death Valley area. To accommodate as many events as possible, the time scale (from Gradstein et al., 2004) is shown with a variable vertical scale, logarithmic from 1000 yr to 1 m.y., then a linear scale from 1 to 5 m.y. and a different linear scale from 5 to 25 m.y. From the left, the columns show the igneous events, stratigraphy, tectonic events, and regional tectonic regime. See Table 1 for data and explanations of abbreviations used. The stratigraphic column illustrates age ranges of named sedimentary forma- tions, with different patterns indicating whether or not formations have been interpreted as syntectonic (Snow and Lux, 1999). The Artist Drive Formation is shown by a different symbol as these rocks in their northern Black Mountains type area are primarily lava fl ows or volcaniclastics. The naming convention for sedimentary formations follows that established by Hunt and Mabey (1966) and subsequently modifi ed by Snow and Lux (1999) and Knott et al. (2005). Also shown are the three Quaternary pluvial events when Death Valley was fl ooded (Anderson and Wells, 2003). In the stratigraphic column, formations from the north and west of the area are on the left so that left to right is equivalent to north and west to south and east. DP—Darwin Plateau (Snyder and Hodges, 2000); WSP—Willow Springs Pluton (Calzia nd Rämö, 2000); SMG—Smith Mountain Granite (Miller and Pavlis, 2005); MSG—granite of Miller Spring (Calzia and Rämö, 2000); SWNVF—Southwest Nevada Volcanic Field (Sawyer et al., 1994); TC+RT—Trail Canyon volcanics and Rhodes Tuff/ Sheepshead andesite (McKenna and Hodges, 1990).

Miocene age of the Tucki Mountain section can Mountains and Furnace Creek Basin, the Artist [2005] include the Artist Drive and Greenwater still be used to date extension at Tucki Mountain Drive, Greenwater, Furnace Creek, and Funeral in the Furnace Creek), but these differences are (Cemen et al., 1999). Formations reach a cumulative 4000 m (Hunt not substantial to this paper. Deposition of these Neogene sediments preserved within Death and Mabey, 1966; Greene, 1997; Wright et al., formations initiated near the end of events asso- Valley record the latest phases of structural 1999). Naming conventions for these formations ciated with Basin and Range extension, starting evolution of the basin. In the northern Black vary among different authors (e.g., Knott et al. with the predominantly volcanic Artist Drive

174 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Two-stage formation of Death Valley

TABLE 1. DATA AND EXPLANATIONS OF ABBREVIATIONS USED IN FIGURE 3 turtleback (Drewes, 1963; Holm et al., 1994). Codes for Death Valley area key ash layers Further south, between the Copper Canyon and Code Ash layer name Age (Ma) Reference 1 Mono Craters 0.014–0.025 Klinger (2001) Mormon Point turtlebacks, the Mormon Point 2 Lava Creek B, Bishop 0.62 Knott et al. (2005) Formation is also in fault contact with basement 3a Bishop 0.77 Knott et al. (2005) rocks, in this case the Mormon Point turtleback 3b Upper Glass Mountain 0.8–1.2 Knott et al. (2005) 4 Middle Glass Mountain 1.5 Knott et al. (2005) (Hayman, 2006; MP, Fig. 4). These sediments 5 Lower Glass Mountain 1.7–1.9 Knott et al. (2005) are younger than the Copper Canyon Forma- 6 Huckleberry Ridge 2.07 Knott et al. (2005) tion, with the 0.77 Ma Bishop tuff (number 3 in 7 Upper Mesquite Spring 3.1 Knott et al. (2005) 8 Lower Mesquite Spring 3.35 Knott et al. (2005) Fig. 3) found near the top of the section (Knott 9 QTf, southern Black Mountains 5.2 Topping (1993) et al., 2005). South of Mormon Point, another Formation abbreviations important data point for structural evolution is Abbreviation Formation Reference in the Confi dence Hills (CH, Fig. 4). These low Qmp Mormon Point Formation Knott et al. (2005) Qtf Funeral Formation Knott et al. (2005) hills are composed of clastic sediments of the Qtch Confidence Hills Formation Knott et al. (2005 ) Confi dence Hills Formation (Wright and Troxel, Tfc Furnace Creek Formation Knott et al. (2005) 1984), with an age range of 1.7–2.2 Ma (Knott Tn Nova Formation Snyder and Hodges (2000) Tcc Copper Canyon Formation Nyborg and Buchheim (2009) et al., 2005). Like the Black Mountains sections, Tv Artist Drive Formation Greene (1997) these sediments show little evidence of syndepo- Tg Greenwater Formation Greene (1997) sitional tec tonism, yet are now highly deformed, U-LR Ubehebe–Lake Rogers Knott et al. (2005) BM Bat Mountain Snow and Lux (1999) with almost vertical dips in places (Beratan et al., Tnv Navadu Formation Snow and Lux (1999) 1999). This deformation is associated with the Tpn Panuga Formation Snow and Lux (1999) Tub Ubehebe Formation Snow and Lux (1999) Southern Death Valley strike-slip fault (Dooley Tec Tecopa Basin Hillhouse (1987) and McClay, 1996) and gives a qualitative indi- DV Death Valley fi ll (lacustrine, playa, and alluvial fans) Hunt and Mabey (1966) cation of the large amount of post-1.7 Ma motion Codes for Tectonic Events column there has been on this fault. Code Event Reference East of Death Valley, the Tecopa Basin (TB, 1a Inyo Range faulting (older) Lee et al. (2009) 1b Inyo Range faulting (younger) Lee et al. (2009) Fig. 4) contains some useful timing information. 2 Extension in Cottonwoods Snow and Lux (1999) The oldest outcropping basin deposit consists of 3 Unroofing in central BM Holm et al. (1992) 4 Black Mountains exhumation Holm and Dokka (1993) fl at-lying basalt as old as 5.12 Ma (Calzia and 5 Extension east of DV—starts in Kingston Peak area, Calzia and Rämö (2000) Rämö, 2000). These are overlain by sediments ends in area of ancestral Lake Tecopa, which include the 6 Strike slip along Furnace Creek Fault Blair and Raynolds (1999) Lava Creek B tuff (0.62 Ma; number 2 in Fig. 3) near the top of the section (Hillhouse, 1987). Seismic refl ection data (Louie et al., 1992) Formation (Greene, 1997). Structural inferences Black Mountains formations, the Artist Drive, is show ~140 m of fl at-lying section below the from younger formations include: (1) two cor- not seen. The Artist Drive Formation is in fault 5.12 Ma basalts, this section lying unconform- related outcrops of the Green water Formation in contact with Proterozoic gneiss and Miocene ably on tilted and faulted sediments. This older the Natural Bridge area are separated by a nor- intrusives of the Badwater turtleback structure. faulting is interpreted as the fi nal phase of Basin mal fault with 2000 m of throw (Greene, 1997); Rocks along the highest points of the northern and Range extension in the area (Louie et al., (2) conglomerates in the lower Furnace Creek Black Mountains are volcanics of the Artist 1992; Calzia and Rämö, 2000). Extrapolation Formation include granite boulders derived from Drive formation, including 6.5-m.y.-old rhyo- of measured sedimentary rates in the upper the northwest, possibly Hunter Mountain (Hunt lite at the popular tourist overlook, Dantes View fl at-lying section yields an age of between 5 and Mabey, 1966; Wright et al., 1999), suggest- (Greene, 1997). and 7 Ma for the unconformity above the fault ing that the Death Valley basin did not yet exist Other sedimentary formations that contain blocks, suggesting that Basin and Range exten- at 5 Ma; (3) fan-deltas in the upper part of the structurally relevant data are found along the sion ceased in the Late Miocene in the Tecopa Funeral Formation are offset ~8 km by dextral Black Mountains. North of Copper Canyon Basin (Calzia and Rämö, 2000). strike slip along the Furnace Creek fault zone, turtleback (CC, Fig. 4), on the fl anks of the The above summary of younger deposits in implying this much strike slip in the last 2 m.y. Black Mountains, the Copper Canyon forma- the Death Valley region highlights some of the (Blair and Raynolds, 1999); (4) although there tion preserves 1800 m of mostly fl uvial to lacus- structurally important inferences that can be are angular uncon formi ties within these forma- trine sedi ments, including some distinctive tufa drawn from these sections. These will be used tions, there are no data such as progressive tilt mounds (Nyborg and Buchheim, 2009). Ages to build a structural model for evolution of the (Snow and Lux, 1999) to suggest that they were of basalt within the succession constrain the basin, but before doing that, it is useful to exam- deposited in active . This relative timing age of the formation to between 5 and 3.3 Ma ine the geometry of Basin and Range detach- can be seen in Figure 3, where events associ- (Nyborg and Buchheim, 2009). The younger age ment faults in the basin. ated with Basin and Range extension predate is further confi rmed by occurrence of one of the deposition of most of the Artist Drive through Mesquite Spring tuffs (3.1–3.35 Ma; numbers 7 BASIN AND RANGE DETACHMENT Funeral Formations. Today, however, these for- and 8 in Fig. 3) near the top of the beds (Knott FAULTS mations all show signifi cant structural dip, such et al., 2005). Presumably deposited in a basin like as the 30° NE dip of the Furnace Creek Forma- present-day Death Valley, these sediments are Figure 4 shows surface geology (Workman tion seen at (Greene, 1997). The today located up to 800 m above the valley fl oor. et al., 2002) and Figure 5 the thickness of Ceno- depositional base of the oldest of these northern They are in fault contact with the Copper Canyon zoic sediments (Blakely and Ponce, 2002) in the

Geosphere, February 2011 175

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

117°30′W NDVFZ117°W116°30′W 116°W consists of Proterozoic basement and metasedi- Quaternary ments. At Mosaic Canyon, for instance, where CM Igneous Quaternary GM Sedimentary Tertiary the detachment surface is exposed (TD, Fig. 4), TM Volcanic Tertiary the canyon cuts through mylonized dolo- Intrusive Tertiary HM Sedimentary Mesozoic mite. This dolomite has been correlated to the A TD 36°30′N FM NV Igneous Mesozoic 36°30′N Noonday Dolomite (Hunt and Mabey, 1966; CA Sedimentary Paleozoic DP Proterozoic Johnnie fm Williams et al., 1976), which is also found in DV Proterozoic Pahrump gp PV AP Proterozoic undivided the footwall of the Black Mountains detachment EF Proterozoic metamorphics fault at Copper Canyon and Mormon Point BW EM (Williams et al., 1976), although Miller and B HFHC AR BM GR Pavlis (2005) question details of this correlation. TP The Noonday Dolomite (Fig. 7) is a platformal C CC NR MP carbonate deposited unconformably across a Neoproterozoic rift basin that marks the onset 36°N 36°N of Proterozoic through Paleozoic miogeoclinal CH SDFZ TB deposits of the western North American passive KR margin (Wright et al., 1978). The Proterozoic– Cambrian boundary is within the (Corsetti and Hagadorn, 2000). The Wood Canyon also marks the base of the Paleo- 117°30′W117°W 116°30′W 116°W zoic succession that, in the Death Valley region, commonly forms Basin and Range extensional Figure 4. Geology of the Death Valley area, from Workman et al. (2002) and, for the allochthons like the Funeral Mountains and area west of the Panamint Range, from Saucedo et al. (2000). Faults are from the USGS Nopah Range (McAllister, 1976; Burchfi el Quaternary faults database. Cross sections in Figures 6, 8, and 9 are labeled A, B, and et al., 1983; Hamilton, 1988). As shown in Fig- C. AP—Aguereberry Point; AR—Argus Range; BW—Badwater; BM—Black Mountains; ure 6, this section is also seen in the extensional CC—Copper Canyon; CH—Confi dence Hills; CM—Cottonwood Mountains; DP—Dar- allochthons on the east fl ank of Tucki Mountain. win Plateau; DV—Death Valley; EM—Eagle Mountain; FM—Funeral Mountains; FP— Detachment surfaces in the Death Valley region Funeral Peak; GM—; GR—; HC—Hanaupah are commonly located in the upper Proterozoic Canyon; HM—Hunter Mountain; IR—Inyo Range; KR—; MP—Mormon section (Fig. 7), probably because this section is Point; NDFZ—Northern Death Valley fault zone; NR—Nopah Range; PR—Panamint the oldest that covers the entire area. The Pah- Range; PV—Panamint Valley; SDFZ—Southern Death Valley fault zone; TB—Tecopa rump Group is found in the western Panamint Basin; TC—Trail Canyon; TM—Tucki Mountain; TP—. Detachment faults Range and southern Black Mountains region, in (italicized): TD—Tucki Detachment; EF—Emigrant Detachment; HF—Harrisburg fault; a confi guration that Wright et al. (1976) relate to HC—Hanaupah Detachment. a Proterozoic aulacogen. As mentioned in the Introduction, Stewart (1983) suggested that the Panamint Range Death Valley area. Rocks exposed in the Pana- structural relationship at Tucki Mountain (Hunt forms the hanging wall of a fault system that mint Range include Proterozoic metamorphic and Mabey, 1966, p. A99). Their cross section transported the Panamints westward from a basement, a Neoproterozoic through Paleozoic across Tucki Mountain nevertheless remains position on top or east of the Black Mountains. sedimentary section, and younger Tertiary sedi- a classic illustration of the geometry of a core This implies that there should be a detachment mentary and volcanic rocks. The northern nose complex. Figure 6 is a cross section constructed fault underneath the Panamint Range. Along the of the Panamints is formed by an extensional from the data in Figures 2 and 4, approximately eastern base of the Panamint Range, between core complex, Tucki Mountain (Hodges et al., along the line illustrated by Hunt and Mabey Trail and Hanaupah Canyons (TC and HC, 1987; Wernicke et al., 1988b). Along the east (1966), profi le A in Figures 4 and 5. The cross Fig. 4), several outcrops of Proterozoic gneissic side of Tucki Mountain, extension is marked section was made by extracting topography basement were mapped by Hunt and Mabey by Paleozoic sedimentary rocks in fault contact from the DEM data (Fig. 2) and geology from (1966) as the footwall of the “Amargosa Thrust.” with Proterozoic rocks that have been tectoni- the digital map of Workman et al. (2002; Fig. 4). This Proterozoic gneiss is overlain by east-tilted cally exhumed from mid-crustal (greenschist; Dips of the Paleozoic strata at the eastern end of Basin and Range fault blocks consisting of Hodges et al., 1987) levels. This extension the section are from Hunt and Mabey (1966), Wood Canyon Formation through Ordovician occurred in the Miocene (Cemen et al., 1999; as is the location of the detachment below these section, like other extensional allochthons in the Wernicke et al., 1988b). The main extensional strata; Figure 7 is a summary of stratigraphic region and also structurally similar to the east detachment surface is gently folded into the nomenclature for these older sediments. Along fl ank of Tucki Mountain (Fig. 6). The gneiss arched topographic surface of Tucki Moun- the western side of Tucki Mountain, rocks in the displays mylonitic fabric consistent with west- tain (Wernicke et al., 1988b). Hunt and Mabey hanging wall of the detachment are Miocene– ward displacement of the extensional alloch- (1966) recognized this structure and correlated Pliocene sediments and volcanics of the Nova thons (McKenna and Hodges, 1990). Hamilton it with the “Amargosa Thrust” of Noble (1941). Basin (Tn in Fig. 3; Snyder and Hodges, 2000). (1988) examined the structures at Hanaupah This unfortunate thrust fault term was used by As summarized in Figure 6, interpretations Canyon and correlated this Hanaupah detach- Hunt and Mabey, even though they recognized of the extensional structures at Tucki Mountain ment (Hamilton’s more appropriate term; HD, the extensional origin of the younger-over-older show that the footwall of the detachment surface Fig. 4) with that at Tucki Mountain. McKenna

176 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Two-stage formation of Death Valley

and Hodges (1990) mapped extensional struc- in the southern Black Mountains (TC+RT, ment are Late Proterozoic Pahrump Group and tures at Trail Canyon. McKenna and Hodges Fig. 3). This correlation implies 25–55 km of Johnnie Formation. These rocks are found along (1990) use the name “Eastern Panamint fault post-eruption offset between the two locations, the full length of the Panamint Range (Fig. 4) system” for a series of faults that include a basal the distance depending on details of how the and form the crest of the range, just like at Tucki detachment and higher faults that collectively intervening faults are restored (McKenna and Mountain. There is no apparent reason for these form an extensional duplex. The Paleozoic Hodges, 1990). rocks to be in the footwall of the detachment at extensional allochthons are located above this If the detachment system along the east side Tucki Mountain and in the hanging wall further duplex. Also in this area, McKenna and Hodges of the Panamints is the one along which the south along the range, as they would need to be (1990) mapped the 8.6–9.4 Ma Trail Canyon Panamints moved, as in the model of Stewart if the entire Panamint Range is allochthonous. volcanic sequence (Fig. 3), which yields some (1983), then there is a correlation problem. At A better interpretation is that the Hanaupah important constraints on extensional geometry. Tucki Mountain, the detachment tracks over the detachment does indeed correlate with the Tucki These volcanics correlate across Death Valley top of the anticline that forms Tucki Mountain detachment and that this detachment tracks up with the Rhodes Tuff and Sheepshead Andesite (Fig. 6). Rocks in the footwall of the detach- the fl ank and over the top of the Panamint Range, just like the Tucki detachment tracks over the top of Tucki Mountain. This is illustrated in Figure 8, a cross section across the Pana mint ′ 117°W 116°30 W 116°W Range near Trail Canyon. This cross section, like Figure 6, uses topographic data from the DEM and outcrop geology from Workman et al. 5 km (2002), in this case supplemented from the map of Hunt and Mabey (1966). Bedding dips are 0 from Hunt and Mabey (1966). In this fi gure, Prot bsmnt ′ A 36°30 N 36°30′N the detachment surface is shown schematically tracking across the top of the Panamint Range, 4420 rather than underneath it. It lies underneath the extensional allochthons on the east side of the –86 range, surfacing along the west side of these allochthons in the east-dipping Harrisburg B fault zone mapped by Hodges et al. (1990).

C

36°N NV 36°N CA

117°W 116°30′W 116°W

Figure 5. Thickness of Cenozoic sediments from the digital compilation of Blakely and Ponce (2002) in color. Topography (USGS DEM, 2009) in gray shades fi lls in spaces where there are no Cenozoic sediments. Dark red lines in Death Valley are faults from Blakely et al. (1999). Purple polygons (Prot bsmnt) are exposures of Proterozoic basement, generally denoting detachment footwalls, from Workman et al. (2002).

Figure 6. Cross section across Tucki Mountain along profi le A in previous location maps. Figure 7. Stratigraphic nomenclature for This profi le is close to the one shown by Hunt and Mabey (1966). It is derived from the the Proterozoic and Early Paleozoic section topography (Fig. 2) and outcrop geology from Workman et al. (2002) and Hunt and Mabey of the Death Valley area. Compiled from (1966). Dashed line shows inferred location of the detachment fault across Tucki Mountain. Wright et al. (1976) and McAllister (1976).

Geosphere, February 2011 177

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

fault underlying Death Valley that would link the Black Mountains and Panamint Range, with the Black Mountains the footwall of the detach- ment and the Panamints the hanging wall. Also not consistent with this model is the fact that the Panamint Range is much higher in elevation (highest point is Telescope Peak, 3368 m) than the Black Mountains (highest point is Funeral Peak, 1946 m); this would mean that the hang- ing wall of the detachment would be higher than the footwall, a structurally unlikely scenario. Figure 8. Cross section across the Panamint Range near Aguereberry Point and Trail Canyon, The Black Mountains display an asymmetric profi le B in location maps Figures 4 and 5. Proterozoic basement consists of metamorphic east-tilted geometry that is typical of a rift-fl ank basement and sediments of the Pahrump Group and Johnnie Formation. Late Protero- uplift (Watts, 2001). The Panamint Range is zoic sediment in the hanging wall of the detachment consists of Wood Canyon Formation. more symmetric, with perhaps a slightly steeper Dashed line shows inferred location of the detachment fault across the Panamint Range. western fl ank. Along the profi le of Figure 9, Tv—Tertiary volcanics. the west fl ank of the range averages a 10° dip and the east fl ank 9.5°. There does not appear to be a high-angle fault along the west side of This interpretation is similar to one shown by dip on the eastern side of the basin is 50°. This the Panamints. Basement mapped by Blakely Hunt and Mabey (1996) in a cross section near matches well with fault dips of 45°–54° mapped and Ponce (2002) shows only a shallow basin in Hanaupah Canyon (their fi g. 108). As noted by Miller (1991) along the base of the Badwater Panamint Valley, west of the Panamint Range. above, Late Proterozoic rocks make up the turtleback, so it is logical to assume that the As mentioned above in the discussion on Tucki crest of the Panamint Range. These consist of eastern fl ank of the basin is a high-angle fault. Mountain, the west side of Tucki Mountain is sediments of the Pahrump Group, Noonday The west side of the basin shows a steep seg- today a low-angle fault. Further south, a strike- Dolomite, and Johnnie Formation, along with ment bordering the main basin and a gently dip- slip fault has been mapped by Zhang et al. some outcrops of Proterozoic basement (Fig. 4; ping surface that joins to the Panamint Range. (1990) along the west side of the range, but the the Noonday Dolomite is included in the John- The dip of the steeper portion of the west side dip of this fault is not known. Cichanski (2000) nie Formation in this digital compilation), with of the basin is ~30°, and the surface projection mapped the southern portion of the range front the highest peak of the range at Telescope Peak of this portion is close to faults at the base of and found generally low-angle faults (dips 15°– consisting of argillites of the Johnnie Forma- the alluvial fans on the west side of Death Val- 34°), with indications that these fault surfaces tion (Albee et al., 1981). As shown in Figure 7, ley (Hunt and Mabey, 1966; USGS Quaternary have been cut by steeper faults, but the vertical these rocks are found in the footwall of detach- faults database), again suggesting that the basin offset on these faults is minor. Cichanski (2000) ments elsewhere in the Death Valley region and is fault-bounded, as shown in Figure 9. Along suggests that the higher-angle faults are part of their greenschist-grade metamorphism in the this profi le, then, Death Valley is an asymmet- the Panamint Valley strike-slip fault. Panamint Range (Labotka and Albee, 1990) is ric with a steep (50°) eastern fault and Although the cross section in Figure 9 shows also consistent with deep burial and subsequent less steep (30°) western fault. This fault pattern a simple graben structure, the three-dimensional exhumation like at these other detachments. As is not consistent with a low-angle detachment basement structure of Death Valley is complex. Figure 4 shows, outcrops of Paleozoic sedi- ments along the eastern side of the range are continuous from Tucki Mountain southward along the eastern slope of the Panamints, and these Paleozoic sediments are interpreted to be Basin and Range–style extensional allochthons. Further justifi cation for the detachment geom- etry shown in Figure 7 comes from the south- western fl anks of the Panamint Range. Here, detachment faults that dip to the west, like the Emigrant fault at Tucki Mountain, have been mapped by Cichanski (2000). So both fl anks of the Panamint Range display fault geometries that indicate detachment surfaces tracking over the top of the range, rather than underneath it. The structural connection between the Pana- mint Range and Black Mountains is illustrated in Figure 9. This cross section includes the thickness of Cenozoic sediment fi ll as deter- mined by Blakely and Ponce (2002; Fig. 4). Basement geometry below Death Valley shows Figure 9. Cross section across the Panamint Range and Black Mountains, profi le C in previ- a steep-sided basin nearly 5000 m deep. The ous location maps. LCS—Little Chief Stock (10.6 Ma granite; Hodges et al., 1990).

178 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Two-stage formation of Death Valley

Blakely et al. (1999), using an earlier version tory of the range is uncertain with present data. of the gravity-derived sediment thickness Rocks now in the crest of the range were meta- map of Blakely and Ponce (2002), showed a morphosed to greenschist facies, representing system of northwest-striking strike-slip faults ~10 km burial, in the Jurassic, with retrograde and northeast-striking normal faults that is illus- metamorphism associated with granitic intru- trated in Figure 5. Blakely et al. (1999) relate sions in the Cretaceous (Labotka and Albee, this fault system to the pull-apart fault system 1990). Extension along the Harrisburg fault that Burchfi el and Stewart (1966) postulated in (HF, Fig. 4) commenced prior to intrusion of the their model for formation of Death Valley. Pull- 10.6 Ma Little Chief stock, which was intruded apart basins are characterized by steeply dipping to ~3 km below the contemporary surface. This bounding faults (e.g., the Dead Sea graben, Gar- implies ~7 km of unroofi ng between the Creta- funkel and Ben-Avraham, 1996). The steeply ceous and 10.6 Ma, with some of this unroofi ng dipping faults seen in Figure 9 make interpreta- Figure 10. Schematic evolution of the cross due to the extension (Labotka and Albee, 1990). tion of Death Valley as a pull-apart basin more section of Figure 9. See text for details. Lack of preserved basinal section equivalent to realistic than the low-angle fault interpretation. the Black Mountains’ Artist Drive Formation The age of fi ll in Death Valley is not directly and younger formations across the top of the observed. Hunt and Mabey (1966) summarize Panamint Range suggests that the range was a data from three exploration wells, the perched on the eastern fl ank of the range. In the high by 8 Ma, although the northwestern fl anks deepest reaching 300 m, drilled in Death Val- Black Mountains, the detachment surface tracks were low enough at 10 Ma to begin accumula- ley. From these data, Hunt and Mabey (1966) over the top of the Badwater turtleback and is tion of Nova Basin sediments (Fig. 3). “assume that about a third of the fi ll [of Death covered by Late Miocene to Pliocene sediments Valley] is Quaternary and that the rest is Ter- of the Artist Drive Formation. The 3 Ma resto- TURTLEBACK STRUCTURES IN THE tiary” (Hunt and Mabey, 1966, p. A72). Using ration is shown in two phases for convenience. BLACK MOUNTAINS this very rough estimate of 1.81 Ma (base of The fi rst restoration at 3 Ma removes the graben Quaternary) for a third of the section and assum- formed as a result of pull-apart basin formation; The three turtleback structures in the Black ing a constant sedimentation rate, the base of strike slip, which would be perpendicular to Mountains, as pointed out by Miller and Pavlis the section would be at 5.4 Ma. A more recent the page (Topping, 1993), is not incorporated. (2005), are keys to understanding tectonic evo- drill hole near Badwater reached a total depth of The concept illustrated here is that faulting due lution of the area. Miller and Pavlis (2005) give 187 m (Lowenstein et al., 1999). Sediment age to strike slip broke and separated the original an excellent overview of current data and under- at the base of this well was estimated by Low- Basin and Range detachment surface. In the standing of the turtlebacks. In summary, the enstein et al. (1999) as 200 ± 10 kA. Extrapolat- second 3 Ma restoration, the topography that turtlebacks are antiformal structures that strike ing this age to the 4.7 km maximum depth in formed as a result of young transtensional strike and dip toward the northwest, exposed along the (Blakely and Ponce, 2002) slip is removed. The suggestion illustrated here the range front bordering Death Valley. They are yields an age of 4.8–5.3 Ma for the base of the is that the Black Mountains, with their present- made up of Proterozoic basement gneiss, early section. These age estimates are very uncertain, day asymmetric structure typical of footwall Tertiary pegmatitic intrusives, and Late Mio- but do indicate that the bulk of the fi ll of Death uplift geometry, formed as a result of footwall cene plutonic rocks. The Late Miocene intru- Valley could be Pliocene and younger, i.e., post– uplift associated with the Death Valley pull- sives consist of the mafi c Willow Springs pluton Basin and Range extension (Fig. 3). apart graben. As discussed earlier, the present- and the felsic Smith Mountain granite. A fea- day contact between the Artist Drive Formation ture of these intrusives is that they are intruded FORMATION OF DEATH VALLEY and basement of the Badwater surface is faulted, along the structural top of each turtleback, with but the geometry shown in this restoration the Willow Springs diorite structurally beneath This model, using the preceding data and shows that the Artist Drive Formation could the younger (Fig. 3) Smith Mountain granite. interpretations, includes two tectonic phases have been originally deposited on basement, As shown in the tectonostratigraphic summary separated by a short interval of relative tectonic with faulting seen today a result of postdeposi- of Figure 3, the Miocene Black Mountain intru- quiescence, as summarized in the timing chart tion emergence of the Badwater turtleback. As sives were formed during Basin and Range in Figure 3. The fi rst phase, lasting from ca. also discussed earlier, the Panamint Range is extension, as indicated by their ages relative 18 to 5 Ma, involved Basin and Range–style a large, almost symmetric anticlinal structure, to age ranges of exhumation associated with low-angle extension with fault blocks made up with no high-angle faults along the western side extensional tectonic denudation. Toward the end predominantly of Paleozoic sediments moving of the range. Dips of the fl anks of the Pana- of this episode of tectonic denudation, deposi- on detachment faults that exhumed rocks from mint Range are shallow, averaging ~10°. Tucki tion of the predominantly volcanic Artist Drive mid crustal levels. The second phase was the Mountain at the north end of the range (Fig. 6) Formation began. In Pliocene time, the Furnace development of Death Valley as a pull-apart is an antiformal structure formed by arching of Creek and Funeral Formations were deposited basin, with higher-angle faults and transten- the exposed detachment surface; I suggest that in the northern Black Mountains, these sections sional motion. this arching also forms the rest of the Panamint being predominantly clastic with little evidence Figure 10 is an illustration of the evolution of Range, making the entire range a core complex. for syndepositional tectonism. At the same time Death Valley. Figure 10A, “Present day,” shows The scale of this core complex, 30–40 km wide and further south along the Black Mountains, the geometry of the Basin and Range detach- by 60 km long with an amplitude of ~4 km, is the Copper Canyon and Mormon Point Forma- ment surface. This surface tracks over the top larger than other well-known core complexes tions were deposited. All these formations were of the Panamint Range, with extensional alloch- like the Snake Range and Whipple core com- deposited in basins formed by Basin and Range thons consisting of an Early Paleozoic section plexes (Block and Royden, 1990). Uplift his- low-angle extension; in the model presented

Geosphere, February 2011 179

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

here, the fl oor of these basins was in some areas tural restorations comes from a thrust complex, the exhumed Basin and Range detachment sur- preserved in one of the Basin and Range alloch- face. When the Death Valley pull-apart basin thons in the northern Panamint Range, which started to form ca. 3 Ma, these basins were correlates with a thrust complex of similar uplifted in the footwall of the Black Mountains geometry in the Nopah Range (Snow and Wer- fault block, which is why they now lie high nicke, 2000). In Snow and Wernicke’s (2000) above the valley fl oor. reconstruction, this correlation is used to restore With the northwest strike of the turtleback the Panamint Range to a position adjacent to the structures, parallel to the inferred direction Nopah Range. The suggestion in this paper that of strike slip, an origin of these features as the Panamints are not a far-traveled extensional a result of the strike slip is suggested. Wright allochthon means that this reconstruction must et al. (1974) showed that the turtlebacks include Figure 11. The cross section of Figure 9 with, be reexamined. The correlation between thrust many linear structures parallel to the direction of superimposed at Badwater, a generalized complexes is still valid, but instead of the entire strike slip, ranging from “minute slickenslides cross section across an oceanic megamullion Panamint Range being moved, only the exten- to fault mullions tens or hundreds of meters structure (megamullion data from Tucholke sional allochthons need to be moved. Further in amplitude” (Wright et al., 1974, p. 54). The and Lin, 1998). implications of this update will be discussed slickenslides are on the southwest fl anks of the elsewhere. turtleback surfaces and trend northwest with a The change from Basin and Range extension plunge of 10° to 15° (Troxel and Wright, 1987); to strike slip between 8 and 3 Ma seen in the these structures imply both normal and strike- the fl anks of the Death Valley pull-apart basin. Death Valley region is very close to age ranges slip components of motion along the fl anks They also lie on the fl ank of a Proterozoic rift of changes in similar tectonic settings seen in of the turtlebacks. Wright et al. (1974) do not basin fi lled with Pahrump Group sediments, several other localities in the ECSZ (Stockli suggest that the turtlebacks themselves are mul- as mapped by Wright et al. (1976). This mul- et al., 2003; Faulds et al., 2005; Oldow et al., lion structures, but a comparison with oceanic tiphase history of the structures plus their loca- 2008; Lee et al., 2009; Andrew and Walker, megamullions suggests that this is what they tion along a preexisting rift may explain why the 2009). It also coincides with fi nal opening of are. Megamullions in oceanic crust have been turtlebacks are unique features of the western the northern Gulf of California and initiation mapped since the development of multibeam North America extensional province. of the southern portion of the San Andreas fault bathymetry in the 1980s. The term is applied to (Dorsey et al., 2007). The strong link between large domed structures formed in inside-corner CONCLUSIONS Pacifi c-North America plate motion and struc- settings at the intersections of spreading centers tural evolution of the ECSZ is emphasized by and transform faults (Tucholke and Lin, 1998). The interpretations presented above offer this synchronous timing. It would appear that, The structures are usually corrugated with mul- some new ideas for the evolution of Death Val- as the southern portion of the San Andreas fault lion structures parallel to spreading direction ley. The Panamint Range, rather than being a joined up to the northern Gulf of California, the and Tucholke and Lin (1998) relate them to con- large fault block detached from above the Black large-scale restraining bend in the San Andreas tinental metamorphic core complexes. Tucholke Mountains, is a core complex like the north- fault formed by the Mojave Block (Fig. 1) devel- and Lin (1998) compiled dimensional data on ern end of the range at Tucki Mountain. Basin oped and has forced deformation to the east, into 17 oceanic megamullions. A simple average of and Range extensional allochthons occur along the ECSZ. This eastward propagation is likely to their compilation yields dimensions of 15.2 km the east fl ank of the range; these formed dur- continue with the Pacifi c-North America plate wide by 1550 m relief. In Figure 11 a sinusoidal ing Miocene–Early Pliocene Basin and Range boundary ultimately moving to the ECSZ, with function with these dimensions is included in extension. Death Valley is a pull-apart basin, as motion on the San Andreas fault ceasing (Faulds the cross section across Badwater. It is intrigu- originally proposed by Burchfi el and Stewart et al., 2005). ing that the ocean megamullion matches fairly (1966), formed in the last 3 m.y. Prior to this ACKNOWLEDGMENTS well with the shape of the basement surface time, a thick sedimentary sequence was depos- of the Badwater turtleback. In map view, Fig- ited in what is now the northern Black Moun- This paper was sponsored by the PLATES project ure 5, the turtlebacks are seen to be adjacent to tains, with early sedimentation being mostly at UTIG. Dave Reynolds, Larry Lawver, and espe- inferred strike-slip faults and in an “inside cor- volcanics. These volcanics were deposited cially an anonymous reviewer provided excellent sug- ner” position similar to that of oceanic mega- partly on the exhumed detachment surface from gestions for improving the manuscript. I would like to thank Dave Reynolds and Stefan Boettcher for intro- mullions (Tucholke and Lin, 1998). Based on the earlier phase of Basin and Range extension. ducing me to Death Valley geology. these comparisons, I suggest that the turtlebacks The turtlebacks are megamullions very similar are megamullions formed as a result of strike- to megamullions adjacent to seafl oor spread- REFERENCES CITED slip faulting in Death Valley. This means that the ing centers. The turtleback structures formed in Albee, A.L., Labotka, T.C., Lanphere, M.A., and McDowell, structures are young, created in the last 3 m.y., the last 3 m.y., although the rocks making up S.D., 1981, Geologic map of the Telescope Peak although the rocks that make up the turtleback the turtlebacks were exhumed from mid-crustal quadrangle, California: U.S. Geological Survey Map GQ-1532. megamullions formed the footwall to the Basin levels during Basin and Range extension. Fault Anderson, D.E., and Wells, S.G., 2003, Latest and Range extensional system and were tec- dips along the southwest fl anks of the turtle- lake highstands in Death Valley, California, in Enzel, tonically exhumed in the Miocene. The turtle- backs are steep, e.g., 50° at Badwater as shown Y., Wells, S.G., and Lancaster, N., eds., Paleoenviron- ments and paleohydrology of the Mojave and southern backs are therefore formed of rocks that were in Figure 9. : Boulder, Colorado, Geological exhumed due to extension and at the same time The Death Valley region has played a key role Society of America Special Paper 368, p. 115–128. Andrew, J.E., and Walker, J.D., 2009, Reconstructing late were intruded by a variety of igneous rocks. in development of models of Basin and Range Cenozoic deformation in central Panamint Valley, They were then cut by strike-slip faults, forming extension. One of the data points used in struc- California: Evolution of slip partitioning in the Walker

180 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Two-stage formation of Death Valley

Lane: Geosphere, v. 5, no. 3, p. 172–198, doi: 10.1130/ ogy, v. 28, p. 299–302, doi: 10.1130/0091-7613(2000)28 Hodges, K.V., Walker J.D., and Wernicke, B.P., 1987, GES00178.1. <299:PTDVUS>2.0.CO;2. Footwall structural evolution of the Tucki Mountain Armstrong, R.L., 1982, Cordilleran metamorphic core com- Davis, G.H., 1980, Structural characteristics of metamorphic detachment system, Death Valley region, southeastern plexes from to southern Canada: Annual Reviews core complexes, southern Arizona, in Crittenden, Jr., California, Geological Society [London] Special Pub- of Earth Sciences, v. 10, p. 129–154, doi: 10.1146/ M.D., Coney, P.J., and Davis, G.H., eds., Cordilleran lications 1987, v. 28, p. 393–408. annurev.ea.10.050182.001021. Metamorphic Core Complexes: Boulder, CO, Geologi- Hodges, K.V., McKenna, L.W., and Harding, M.B., 1990, Atwater, T., and Stock, J., 1998, Pacifi c-North America plate cal Society of America Memoir 153, p. 35–78. Structural unroofi ng of the central Panamint Moun- tectonics of the Neogene southwestern : Dickinson, W.R., 2002, The as a tains, Death Valley region, southeastern California, in An update: International Geology Review, v. 40, composite extensional domain: International Geology Wernicke, B.P., ed., Basin and Range extensional tec- p. 375–402, doi: 10.1080/00206819809465216. Review, v. 44, p. 1–38, doi: 10.2747/0020-6814.44.1.1. tonics near the latitude of , Nevada: Geologi- Beratan, K.K., Hsieh, J., and Murray, B., 1999, Pliocene- Dokka, R.K., and Travis, C.J., 1990a, Late Cenozoic strike- cal Society of America Memoir 176, p. 377–390. Pleistocene stratigraphy and depositional environments, slip faulting in the : Tectonics, v. 9, Holm, D.K., and Dokka, R.K., 1993, Interpretation and tec- southern Confi dence Hills, Death Valley, California, p. 311–340, doi: 10.1029/TC009i002p00311. tonic implications of cooling histories: An example in Wright, L.A., and Troxel, B., eds., Cenozoic Basins Dokka, R.K., and Travis, C.J., 1990b, Role of the Eastern from the Black Mountains, Death Valley extended of the Death Valley Region, Boulder, CO, Geological California Shear Zone in accommodating Pacifi c-North terrane, California: Earth and Planetary Science Society of America Special Paper 333, p. 289–300. American plate motion: Geophysical Research Letters, Letters, v. 116, p. 63–80, doi: 10.1016/0012-821X Blair, T.C., and Raynolds, R.G., 1999, Sedimentology and v. 17, p. 1323–1326, doi: 10.1029/GL017i009p01323. (93)90045-B. tectonic implications of the Neogene synrift Hole Dooley, T.P., and McClay, K.R., 1996, Strike-slip deforma- Holm, D.K., Snow, J.K., and Lux, D.R., 1992, Thermal and in the Wall and Wall Front members, Furnace Creek tion in the Confi dence Hills, southern Death Valley barometric constraints on the intrusive and unroofi ng Basin, Death Valley, California, in Wright, L.A., and fault zone, eastern California, USA: Journal of the history of the Black Mountains: Implications for tim- Troxel, B.W., eds., Cenozoic Basins of the Death Val- Geological Society, v. 153, p. 375–387, doi: 10.1144/ ing, initial dip, and kinematics of detachment faulting ley Region: Boulder, Colorado, Geological Society of gsjgs.153.3.0375. in the Death Valley region, California: Tectonics, v. 11, America Special Paper 333, p. 127–168. Dorsey, R.J., Fluette, A., McDougall, K., Housen, B.A., p. 507–522, doi: 10.1029/92TC00211. Blakely, R.J., Jachens, R.C., Calzia, J.P., and Langenheim, Janecke, S.U., Axen, G.J., and Shirvell, C.R., 2007, Holm, D.K., Fleck, R.J., and Lux, D.R., 1994, The Death V.E., 1999, Cenozoic basins of the Death Valley Chronology of Miocene–Pliocene deposits at Split Valley Turtlebacks reinterpreted as Miocene-Pliocene extended terrane as refl ected in regional scale gravity Mountain Gorge, : A record of folds of a major detachment surface: The Journal of anomalies, in Wright, L.A., and Troxel, B.W., eds., regional tectonics and evolution: Geol- Geology, v. 102, p. 718–727, doi: 10.1086/629715. Cenozoic Basins of the Death Valley Region: Boul- ogy, v. 35, p. 57–60, doi: 10.1130/G23139A.1. Hunt, C.B., and Mabey, D.R., 1966, Stratigraphy and struc- der, Colorado, Geological Society of America Special Drewes, H., 1963, Geology of the Funeral Peak quadrangle, ture, Death Valley, California: U.S. Geological Survey Paper 333, p. 1–16. California, on the east fl ank of Death Valley: U.S. Geo- Professional Paper 494-A, 162 p. Blakely, R.J., and Ponce, D.A., 2002, Map Showing Depth logical Survey Professional Paper 413, 78 p. Klinger, R.E., 2001, Late Quaternary volcanism of Ubehebe to Pre-Cenozoic Basement in the Death Valley Ground- Faulds, J.E., Henry, C.D., and Hinz, N.H., 2005, Kinemat- Crater, in Machette, M.N., Johnson, M.L., and Slate, water Model Area, Nevada and California, in USGS ics of the northern Walker Lane: An incipient trans- J.L., eds., Quaternary and Late Pliocene Geology MF2381, http://pubs.usgs.gov/mf/2002/mf-2381/. form fault along the Pacifi c–North American plate of the Death Valley Region: Recent Observations on Block, L., and Royden, L.H., 1990, Core complex geome- boundary: Geology, v. 33, p. 505–508, doi: 10.1130/ Tectonics, Stratigraphy, and Lake Cycles (Guidebook tries and regional scale fl ow in the lower crust: Tecton- G21274.1. for the 2001 Pacifi c Cell—Friends of the Pleistocene ics, v. 9, p. 557–567, doi: 10.1029/TC009i004p00557. Garfunkel, Z., and Ben-Avraham, Z., 1996, The structure of Fieldtrip): U.S. Geological Survey Open-File Report Burchfi el, B.C., and Stewart, J.H., 1966, ‘Pull-apart’ origin the Dead Sea basin: Tectonophysics, v. 266, p. 155– 01–51, p. A21–A24. of the central segment of Death Valley, California: 176, doi: 10.1016/S0040-1951(96)00188-6. Knott, J.R., Sarna-Wojcicki, A.M., Machette, M.N., and Geological Society of America Bulletin, v. 77, p. 439– Gradstein, F.M., Ogg, J.G., Smith, A.G., Bleeker, W., and Klinger, R.E., 2005, Upper Neogene stratigraphy and 442, doi: 10.1130/0016-7606(1966)77[439:POOTCS] Lourens, L.J., 2004, A new Geologic Time Scale, with tectonics of Death Valley—A review: Earth-Science 2.0.CO;2. special reference to Precambrian and Neogene: Epi- Reviews, v. 73, p. 245–270, doi: 10.1016/j.earscirev Burchfi el, B.C., Cowan, D.S., and Davis, G.A., 1992, sodes, v. 27, p. 83–100. .2005.07.004. Tectonic overview of the Cordilleran Orogen in the Greene, R.C., 1997, Geology of the northern Black Moun- Labotka, T.C., and Albee, A.L., 1990, Uplift and exposure , in Burchfi el, B.C., Lipman, tains, Death Valley, California: U.S. Geological Sur- of the Panamint metamorphic complex, California, in P.W., and Zoback, M.L., eds., The Cordilleran Oro- vey Open-File Report OF 97–79, scale 1:24,000, Wernicke, B.P., ed., Basin and Range extensional tec- gen, conterminous U.S., Geology of North America, 110 p. tonics near the latitude of Las Vegas, Nevada: Geologi- Volume G3: Boulder, Colorado, Geological Society of Hamilton, W.B., 1988, Detachment faulting in the Death cal Society of America Memoir 176, p. 345–362. America, p. 407–479. Valley region, California and Nevada, in Carr, M.D., Lee, J., Stockli, D.F., Owen, L.A., Finkel, R.C., and Kislitsyn , Burchfi el, B.C., Hodges, K.V., and Royden, L.H., 1987, and Yount, J.C., eds., Geologic and hydrologic inves- R., 2009, Exhumation of the Inyo Mountains, Califor- Geology of Panamint Valley-Saline Valley pull-apart tigations of a potential nuclear waste disposal site at nia: Implications for the timing of extension along the system, California: Palinspastic evidence for low angle , southern Nevada: U. S. Geological western boundary of the Basin and Range Province and geometry of a Neogene range-bounding fault: Journal Survey Bulletin, Report B1790, p. 51–85. distribution of dextral fault slip rates across the eastern of Geophysical Research, v. 92, p. 10,422–10,426, doi: Hammond, W.C., and Thatcher, W., 2007, Crustal deforma- California shear zone: Tectonics, v. 28, p. TC1001, doi: 10.1029/JB092iB10p10422. tion across the Sierra Nevada, northern Walker Lane, 10.1029/2008TC002295. Burchfi el, B.C., Hamill, G.S., and Wilhelms, D.E., 1983, Basin and Range transition, western United States Lonsdale, P., 1989, Geology and tectonic history of the Gulf Structural geology of the Montgomery Mountains measured with GPS, 2000–2004: Journal of Geo- of California, in Winterer, E.L., et al., eds., The eastern and the northern half of the Nopah and Resting Spring physical Research, v. 112, p. B05411, doi: 10.1029/ Pacifi c Ocean and Hawaii: Boulder, Colorado, Geo- Ranges, Nevada and California: Geological Society of 2006JB004625. logi cal Society of America, Geology of North Amer- America Bulletin, v. 94, p. 1359–1376, doi: 10.1130/ Hayman, N.W., 2006, Shallow crustal fault rocks from the ica, v. N, p. 499–521. 0016-7606(1983)94<1359:SGOTMM>2.0.CO;2. Black Mountain detachments, Death Valley, CA: Jour- Louie, J., Cetintas, A., Chekuri, V., Corchuelo, W., Lei, Y., Calzia, J.P., and Rämö, O.T., 2000, Late Cenozoic crustal nal of Structural Geology, v. 28, p. 1767–1784, doi: Li, L., Mekala, G., Ozalaybey, S., and Raskulinecz, J., extension and magmatism, southern Death Valley 10.1016/j.jsg.2006.06.017. 1992, 7 Ma age of Tecopa Lake and tectonic quiescence region, California, in Lageson, D.R., Peters, S.G., and Hayman, N.W., Knott, J.R., Cowan, D.S., Nemser, E., and from seismic stratigraphy east of Death Valley, CA: Eos Lahren, M.M., eds., Great Basin and Sierra Nevada: Sarna-Wojcicki, A.M., 2003, Quaternary low-angle (Transactions, American Geophysical Union), v. 73, Boulder, Colorado, Geological Society of America slip on detachment faults in Death Valley, California: Issue 43, p. 548. Also see http://www.seismo.unr.edu/ Field Guide 2, p. 135–164. Geology, v. 31, p. 343–346, doi: 10.1130/0091-7613 ftp/pub/louie/papers/tecopa/tecopa.html. Cemen, I., Wright, L.A., and Prave, A.R., 1999, Stratigraphy (2003)031<0343:QLASOD>2.0.CO;2. Lowenstein, T.K., Li, J., Brown, C., Roberts, S.M., Ku, and tectonic implications of the latest Oligocene and Henry, C.D., Faulds, J.E., and dePolo, C.M., 2007, Geom- T.-L., and Luo, S., 1999, 200 k.y. paleoclimate record early Miocene sedimentary succession, southernmost etry and timing of strike-slip and normal faults in the from Death Valley salt core: Geology, v. 27, p. 3–6, Funeral Mountains, Death Valley region, California, northern Walker Lane, northwestern Nevada and north- doi: 10.1130/0091-7613(1999)027<0003:KYPRFD> in Wright, L.A., and Troxel, B.W., eds., Cenozoic eastern California: Strain partitioning or sequential 2.3.CO;2. Basins of the Death Valley Region: Boulder, Colorado, extensional and strike-slip deformation?, in Till, A.B., McAllister, J.F., 1976, Geologic maps and sections of a strip Geological Society of America Special Paper 333, Roeske, S.M., Sample, J.C., and Foster, D.A., eds., from Pyramid Peak to the southeast end of the Funeral p. 65–86. Exhumation Associated with Continental Strike-Slip Mountains, Ryan Quadrangle, California, in Troxel, Cichanski, M., 2000, Low-angle, range-fl ank faults in the Fault Systems: Boulder, CO, Geological Society of B.W. and Wright, L.A., eds., Geologic features, Death Panamint, Inyo, and Slate ranges, California: Implica- America Special Paper 434, p. 59–79, doi: 10.1130/ Valley, California: California Division of Mines and tions for recent tectonics of the Death Valley region, 2007.2434(04). Geology Special Report 106, p. 63–66. Geological Society of America Bulletin, v. 112, Hillhouse, J.W., 1987, Late Tertiary and Quaternary geol- McKenna, L.W., and Hodges, K.V., 1990, Constraints on p. 871–883, doi:10.1130/0016-7606 ogy of the Tecopa basin, southeastern California: U.S. the kinematics and timing of late Miocene–Recent Corsetti, F.A., and Hagadorn, J.W., 2000, The Precambrian- Geological Survey Miscellaneous Investigations Map extension between the Panamint and Black Mountains, Cambrian transition: Death Valley, United States: Geol- I-1728, scale 1:48,000. southeastern California, in Wernicke, B.P., ed., Basin

Geosphere, February 2011 181

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021 Norton

and Range extensional tectonics near the latitude of Smith, W.H.F., and Sandwell, D.T., 1997, Global seafl oor Watts, A.B., 2001, Isostasy and Flexure of the Lithosphere: Las Vegas, Nevada: Geological Society of America topography from satellite altimetry and ship depth Cambridge, UK, Cambridge University Press, 458 p. Memoir 176, p. 363–376. soundings: Science, v. 277, p. 957–1962. Version 12.1 Wernicke, B.P., Axen, G.J., and Snow, J.K., 1988a, Basin Miller, M.G., 1991, High angle origin of the currently low- used in this paper, see ftp://topex.ucsd.edu/pub/ and Range extensional tectonics at the latitude of Las angle Badwater Turtleback fault, Death Valley, Califor- global_topo_1min. Vegas, Nevada: Geological Society of America Bul- nia: Geology, v. 19, p. 372–375, doi: 10.1130/0091-7613 Snow, J.K., and Lux, D.R., 1999, Tectono-sequence stratig- letin, v. 100, p. 1738–1757, doi: 10.1130/0016-7606 (1991)019<0372:HAOOTC>2.3.CO;2. raphy of Tertiary rocks in the Cottonwood Mountains (1988)100<1738:BARETA>2.3.CO;2. Miller, M.M., Johnson, D.J., Dixon, T.H., and Dokka, R.K., and northern Death Valley area, California and Nevada, Wernicke, B.P., Walker, J.D., and Hodges, K.V., 1988b, 2001, Refi ned kinematics of the eastern California in Wright, L.A., and Troxel, B., eds., Cenozoic Basins Field guide to the northern part of the Tucki Mountain shear zone from GPS observations, 1993–1994: Jour- of the Death Valley Region: Boulder, Colorado, Geo- fault system, Death Valley region, California, in This nal of Geophysical Research, v. 106, p. 2245–2263, logical Society of America Memoir Special Paper 333, extended land: Geologic journeys in the southern Basin doi: 10.1029/2000JB900328. p.17–64. and Range, fi eld trip guidebook: Geological Society Miller, M., and Prave, A.R., 2002, Rolling hinge or fi xed basin?: Snow, J.K., and Wernicke, B.P., 2000, Cenozoic tectonism of America Cordilleran Section Meeting, Las Vegas, A test of continental extensional models in Death Valley, in the central Basin and Range: Magnitude, rate, and Nevada, p. 58–63. California, United States: Geology, v. 30, p. 847–850, distribution of upper crustal strain: American Jour- Williams, E.G., Wright, L.A., and Troxel, B.W., 1976, The doi: 10.1130/0091-7613(2002)030<0847:RHOFBA> nal of Science, v. 300, p. 659–719, doi: 10.2475/ Noonday Dolomite and equivalent stratigraphic units, 2.0.CO;2. ajs.300.9.659. southern Death Valley region, California, in Troxel, Miller, M.B., and Pavlis, T.L., 2005, The Black Mountains Snyder, N.P., and Hodges, K.V., 2000, Depositional and tec- B.W., and Wright, L.A., eds., Geologic features, Death turtlebacks: Rosetta stones of Death Valley tectonics: tonic evolution of a supradetachment basin: 40Ar/39Ar Valley, California: California Division of Mines and Earth-Science Reviews, v. 73, p. 115–138, doi: 10.1016/ geochronology of the Nova Formation, Panamint Geology Special Report 106, p. 45–50. j.earscirev.2005.04.007. Range, California: Basin Research, v. 12, p. 19–30, Workman, J.B., Menges, C.M., Page, W.R., Taylor, E.M., Noble, L.F., 1941, Structural features of the Virgin Spring doi: 10.1046/j.1365-2117.2000.00108.x. Ekren, E.B., Rowley, P.D., Dixon, G.L., Thompson, area, Death Valley, California: Geological Society of Sonder, L.J., and Jones, C.H., 1999, Western United States R.A., and Wright, L.A., 2002, Geologic Map of the America Bulletin, v. 52, p. 942–1000. extension: How the west was widened: Annual Review Death Valley Ground-water Model Area, Nevada and Nyborg, T., and Buchheim, P., 2009, Lacustrine Tufa of Earth and Planetary Sciences, v. 27, p. 417–462, doi: California, in USGS MF2381, http://pubs.usgs.gov/ Mounds of the Miocene-Pliocene Copper Canyon 10.1146/annurev.earth.27.1.417. mf/2002/mf-2381/ (accessed January 2010). “Formation,” Death Valley, California: Relationship to Stewart, J.H., 1983, Extensional tectonics in the Death Wright, L.A., and Troxel, B.W., 1973, Shallow-fault inter- Organism Abundance and Distribution: AAPG Search Valley area, California: Transport of the Panamint pretation of Basin and Range structure, in DeJong, and Discovery Article 90090, ©2009 AAPG Annual Range structural block 80 km northwestward: Geology, K.A., and Scholten, R., eds., Gravity and Tectonics: Convention and Exhibition, Denver, Colorado. v. 11, p. 153–157, doi: 10.1130/0091-7613(1983)11 New York, John Wiley and Sons, p. 397–407. Oldow, J.S., Kohler, G.K., and Donelick, R.A., 1994, Late <153:ETITDV>2.0.CO;2. Wright, L.A., and Troxel, B.W., 1984, Geology of the North Cenozoic extensional transfer in the Walker Lane Stewart, J.H., 1988, Tectonics of the Walker Lane belt, west- 1/2 Confidence Hills 15′ quadrangle, Inyo County, strike-slip belt, Nevada: Geology, v. 22, p. 637–640, ern Great Basin: Mesozoic and Cenozoic deformation Cali fornia: Sacramento, California, California Divi- doi: 10.1130/0091-7613(1994)022<0637:LCETIT> in a zone of shear, in Ernst, W.G., ed., Metamorphism sion of Mines and Geology. 2.3.CO;2. and Crustal Evolution of the Western United States, Wright, L.A., Otton, J., and Troxel, B.W., 1974, Turtle- Oldow, J., Geissman, J., and Stockli, D., 2008, Evolution Rubey, vol. VII: Englewood Cliffs, New Jersey, Pren- back surfaces of Death Valley viewed as phenomena and strain reorganization within Late Neogene struc- tice Hall, p. 683–713. of extension: Geology, v. 2, p. 53–54, doi: 10.1130/ tural stepovers linking the Central Walker Lane and Stewart, J.H., 1998, Regional characteristics, tilt domains, 0091-7613(1974)2<53:TSODVV>2.0.CO;2. Northern Eastern California Shear Zone, Western and extensional history of the late Cenozoic Basin and Wright, L.A., Troxel, B.W., Williams, E.G., Roberts, M.T., Great Basin: International Geology Review, v. 50, Range province, western North America, in Faulds, and Diehl, P.E., 1976, Precambrian sedimentary envi- p. 270–290, doi: 10.2747/0020-6814.50.3.270. J.E., and Stewart, J.H., eds., Accommodation zones ronments of the Death Valley region, eastern Califor- Sarna-Wojcicki, A.M., Machette, M.N., Knott, J.R., Klinger, and transfer zones: The regional segmentation of the nia, in Troxel, B.W., and Wright, L.A., eds., Geologic R.E., Fleck, R.J., Tinsley, J.C., Troxel, B., Budahn, Basin and Range province: Geological Society of features, Death Valley, California: California Division J.R., and Walker, J.P., 2001, Weaving a temporal and America Special Paper 323, p. 47–74. of Mines and Geology Special Report 106, p. 7–18. spatial framework for the Late Neogene of Death Val- Stockli, D.F., Dumitru, T.A., McWilliams, M.O., and Farley, Wright, L.A., Williams, E.G., and Cloud, P., 1978, Algal ley—Correlation and dating of Pliocene and Quater- K.A., 2003, Cenozoic tectonic evolution of the White and cryptalgal structures and platform environments nary units using tephrochronology, 40Ar/39Ar dating, Mountains, California and Nevada: Geological Society of the late pre-Phanerozoic Noonday Dolomite, eastern and other dating methods, in Machette, M.N., Johnson, of America Bulletin, v. 115, p. 788–816, doi: 10.1130/ California: Geological Society of America Bulletin, M.L., and Sand late, J.L., eds., Quaternary and Late 0016-7606(2003)115<0788:CTEOTW>2.0.CO;2. v. 89, p. 321–333, doi: 10.1130/0016-7606(1978)89 Pliocene Geology of the Death Valley Region: Recent Topping, D.J., 1993, Paleogeographic reconstruction of the <321:AACSAP>2.0.CO;2. Observations on Tectonics, Stratigraphy, and Lake Death Valley extended region: Evidence from Miocene Wright, L.A., Greene, R.C., Cemen, I., Johnson, F.C., Cycles (Guidebook for the 2001 Pacifi c Cell—Friends large rock avalanche deposits in the Amargosa Chaos Prave, A.R., and Drake, R.E., 1999, Tectonostrati- of the Pleistocene Fieldtrip): Denver, Colorado, U.S. Basin, California: Geological Society of America Bul- graphic development of the Miocene-Pliocene Fur- Geological Survey, p. 121–135. letin, v. 105, p. 1190–1213, doi: 10.1130/0016-7606 nace Creek Basin and related features, Death Valley Saucedo, G.J., Bedford, D.R., Raines, G.L., Miller, R.J., (1993)105<1190:PROTDV>2.3.CO;2. region, California, in Wright, L.A., and Troxel, B.W., and Wentworth, C.M., 2000, Digital Data for the Geo- Troxel, B.W., and Wright, L.A., 1987, Tertiary extensional eds., Cenozoic Basins of the Death Valley Region: logic Map of California: California Geological Survey, features, Death Valley region, eastern California: Geo- Geological Society of America Special Paper 333, CD-ROM. logical Society of America Centennial Field Guide- p. 115–126. Sawyer, D.A., Hudson, M.R., Fleck, R.J., Lanphere, M.A., Cordilleran Section, p. 121–132. Zhang, P., Ellis, M., Slemmons, D.B., and Mao, F., 1990, Warren, R.G., and Broxton, D.E., 1994, Episodic cal- Tucholke, B.E., and Lin, J., 1998, Megamullions and mul- Right-lateral displacements and the Holocene slip dera volcanism in the Miocene southwestern Nevada lion structure defining oceanic metamorphic core rate associated with prehistoric along the volcanic fi eld: Revised stratigraphic framework, complexes on the Mid-Atlantic Ridge: Journal of Geo- Southern Panamint Valley Fault Zone: Implications for 40Ar/39Ar geochronology, and implications for mag- physical Research, v. 103, p. 9857–9866, doi: 10.1029/ Southern Basin and Range tectonics and coastal Cali- matism and extension: Geological Society of America 98JB00167. fornia deformation: Journal of Geophysical Research, Bulletin, v. 106, p. 1304–1319, doi: 10.1130/0016-7606 U.S. Geological Survey DEM, 2009, http://seamless.usgs.gov/ v. 95, p. 4857–4872, doi: 10.1029/JB095iB04p04857. (1994)106<1304:ECVITM>2.3.CO;2. index.php 1 arc-second (30 meter) version (accessed Serpa, L., and Pavlis, T.L., 1996, Three-dimensional model January 2010). of the late Cenozoic history of the Death Valley region, U.S. Geological Survey Quaternary faults database, 2009, MANUSCRIPT RECEIVED 11 FEBRUARY 2010 southeastern California: Tectonics, v. 15, p. 1113– http://earthquake.usgs.gov/regional/qfaults/imsintro.php REVISED MANUSCRIPT RECEIVED 3 AUGUST 2010 1128, doi: 10.1029/96TC01633. (accessed January 2010). MANUSCRIPT ACCEPTED 4 AUGUST 2010

182 Geosphere, February 2011

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/7/1/171/3339711/171.pdf by guest on 28 September 2021