Quick viewing(Text Mode)

Paleoproterozoic Orogenesis and Quartz-Arenite Deposition in the Little Chino Valley Area, Yavapai Tectonic Province, Central Ar

Paleoproterozoic Orogenesis and Quartz-Arenite Deposition in the Little Chino Valley Area, Yavapai Tectonic Province, Central Ar

Research Paper

GEOSPHERE Paleoproterozoic orogenesis and - deposition in the Little Chino Valley area, Yavapai tectonic province, central GEOSPHERE; v. 12, no. 6 Arizona, USA doi:10.1130/GES01339.1 Jon E. Spencer1, Mark E. Pecha1, George E. Gehrels1, William R. Dickinson1,*, Kenneth J. Domanik2, and Jay Quade1 14 figures; 3 supplemental files 1Department of Geosciences, University of Arizona, Tucson, Arizona 85721, USA 2Lunar and Planetary Laboratory, University of Arizona, Tucson, Arizona 85721, USA

CORRESPONDENCE: jklmspenc@​dakotacom​.net

CITATION: Spencer, J.E., Pecha, M.E., Gehrels, ABSTRACT strata in the area, and it is probably significantly younger. We infer that the G.E., Dickinson, W.R., Domanik, K.J., and Quade, J., 2016, Paleoproterozoic orogenesis and quartz-­arenite physically immature but chemically super-mature Del Rio was de- deposition in the Little Chino Valley area, Yavapai New field mapping and laboratory studies of Paleoproterozoic rock units posited during a time of extreme weathering during a hot, humid climate with tectonic province, central Arizona, USA: Geosphere, around Little Chino Valley in central Arizona clarify the timing of magmatism, exceptionally high atmospheric CO concentrations and associated corrosive v. 12, no. 6, p. 1–21, doi:10.1130/GES01339.1. 2 deformation, and sedimentation in part of the Yavapai tectonic province and rainwater rich in carbonic acid. yield new insights into sources of and weathering environments. Mafic Received 29 March 2016 Revision received 22 July 2016 lavas, calc-silicate rocks, and pelitic and psammitic strata in the Jerome Can- Accepted 24 August 2016 yon area west of Little Chino Valley were deposited, deformed, and intruded INTRODUCTION by the 1736 ± 21 (2σ) Ma Williamson Valley Granodiorite. U-Pb geochronologic analysis of detrital zircons from a sample of psammitic strata yielded a maxi­ Laurentia is the Paleoproterozoic ancestor of the North American continent. mum depositional age of ca. 1738 Ma. Approximately 25% of the detrital-­ Much of it formed by 1.8 Ga as a result of tectonic assembly of Paleoprotero- zircon grains were derived from a ca. 2480 Ma source, as previously identified zoic and Archean cratonic elements into the Hudsonian craton (Hoffman, 1988). in Grand Canyon schist units. Kolmogorov-Smirnov statistical comparison of Over the following 200 million years, magmatism, sedimentation, and tec- the Jerome Canyon detrital-zircon analyses with Grand Canyon schist analy- tonic accretion added the Transcontinental Proterozoic provinces to its (now) ses indicates that three of the 12 samples analyzed by Shufeldt et al. (2010) southern margin (Condie, 1982; Karlstrom and Bowring, 1993; Van Schmus and are statistically indistinguishable from the Jerome Canyon sample at the 95% Bickford, 1993; Karlstrom et al., 2001; Whitmeyer and Karlstrom, 2007). These confidence level and supports the concept that the Jerome Canyon sequence tectonic provinces represent genesis of much of the continental crust beneath and Paleoproterozoic schists in the eastern and western Grand Canyon are the United States, including that beneath all of Arizona and New Mexico (Fig. 1). part of the same tectonostratigraphic terrane. Genesis of continental crust, and changes to crustal genesis processes over The Del Rio Quartzite on the northeast side of Little Chino Valley, previously geologic time, are major topics of geologic interest. Earth’s greater radiogenic considered an outlier of Mazatzal Quartzite, consists of poorly sorted quartz heat production in the past has been identified as the likely cause of different arenite, pebbly , and conglomerate deposited in a braided-­ tectonic and magmatic processes associated with crustal genesis (e.g., Ham- stream environment. Microscope examination of 32 thin sections stained for ilton, 2007), including rates of accretion in forearcs (Condie, 2007). The Chino potassium and calcium failed to identify any , mica, or mafic silicate Valley area in central Arizona is in the Yavapai tectonic province, which was grains. Similarly, conglomerate clasts consist entirely of vein quartz and less created by magmatic activity, sedimentation, and tectonic accretion at ca. 1.7– abundant argillite and jasper. A rock unit interpreted as a paleosol beneath the 1.8 Ga (Fig. 2; Karlstrom and Bowring, 1988, 1993) and perhaps modified by rift- Del Rio Quartzite contains no surviving minerals except quartz, some of which ing and related magmatism (Duebendorfer et al., 2006; Bickford and Hill, 2007). is embayed and rounded as in corrosive saprolitic soils. U-Pb geochrono­ logic­ Two 7 ½′ quadrangles in the Chino Valley area, mapped in detail as part of analyses of detrital zircons from the 1400-m-thick Quartzite indicate maxi- the STATEMAP program (a component of the National Geologic Mapping Act mum depositional ages of ca. 1745 Ma for the base and ca. 1737 Ma for the of 1992), encompass areas of Paleoproterozoic magmatism, sedimentation, top. The unit is folded but is unaffected by the penetrative deformation and and deformation. This new mapping and associated U-Pb geochronologic, metamorphism that affected other Paleoproterozoic volcanic and sedimentary geochemical, and petrographic analyses yielded insights into Yavapai prov- For permission to copy, contact Copyright ince genesis. A folded section of quartzite and quartz-cobble conglomerate in Permissions, GSA, or [email protected]. *Published posthumously the same area has been mapped as an outlier of the Mazatzal Quartzite. The

© 2016 Geological Society of America

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 1 Research Paper

quartzite is not affected by the metamorphism and penetrative deformation that affected other Paleoproterozoic strata in the area and is likely younger. The quartz-rich composition of the quartzite and a deeply weathered paleosol at its base support previous interpretations of extreme chemical weathering during latest Paleoproterozoic quartz-arenite deposition across the Transcontinental Paleoproterozoic provinces.

Background

Since genesis, the Transcontinental Proterozoic tectonic provinces have been greatly modified and largely buried, but they are fairly well preserved and exposed in a 500-km-long, northwest-southeast belt across Arizona that is perpendicular to the dominant structural and lithologic fabric of the prov- inces (Fig. 1). This belt is within a part of Arizona known as the Transition Zone. Unlike the Colorado Plateau to the northeast, its Paleozoic cover has been mostly removed by erosion, and unlike the Basin and Range province to the southwest, it is not severely affected by Phanerozoic tectonism and magmatism. Paleoproterozoic rocks of the Transition Zone consist largely of granitoids and weakly to strongly metamorphosed volcanic and clastic sedimentary rocks (Fig. 2; Wrucke and Conway, 1987; Karlstrom and Bowring, 1988; DeWitt et al., 2008). Geochronologic studies suggest that assembly of the Proterozoic tectonic provinces occurred from northwest to southeast (Condie, 1982; Karl- strom et al., 1987, 2001; Van Schmus et al., 1993; Eisele and Isachsen, 2001; Meijer, 2014), although in detail, igneous-rock ages are only weakly correlated with position in the orogen. Isotopic studies indicate that the newly formed Paleoproterozoic crust was largely juvenile, with little incorporation of older material (Bennett and DePaolo, 1987; Wooden and DeWitt, 1991; Iriondo et al., 2004). Central Arizona Paleoproterozoic rocks are notable for their prominent north- to northeast-striking shear zones that separate multiple tectonic blocks, as identified by Karlstrom and Bowring (1993). The well-developed Shylock shear zone, a north-south–striking zone of strong, subvertical foliation that is exposed over ~60 km, bounds the Chino Valley area on the east (Fig. 2). The

steep, planar fabric in the shear zone (S2 of Darrach et al., 1991), with steeply plunging lineations and fold axes, is interpreted to represent shortening with dip-slip displacement. The fabric is both intruded by, and affects, a 1699 Ma pluton (Karlstrom, 1989; Darrach et al., 1991). The Mesa Butte shear zone, a feature that includes a fault on the west side of Chino Valley, was proposed to be part of the tectonic boundary between the Hualapai block to the northwest and the Green Gulch block to the southeast (Bergh and Karlstrom, 1992; Karl- strom and Bowring, 1993). The geologic histories of each block have been used to infer the approximate timing of juxtaposition of the blocks, with a plausi- ble interpretation of juxtaposition indicated by synchronous deformation and Figure 1. Geologic map of the interior of southwestern North America showing location of Precambrian bedrock including Mazatzal and Uncompahgre and similar units. Also shown are named shear zones in central Arizona and extrapolated boundaries from Arizona magmatism in adjacent blocks (Karlstrom and Bowring, 1988, 1993). to the Rocky Mountains. Modified from Karlstrom and Bowring (1993) and Jones et al. (2009). Note that the Mojave, Yavapai, and Mazatzal Upper Paleoproterozoic to lower Mesoproterozoic quartz-rich clastic rocks provinces are part of the Transcontinental Proterozoic provinces, whereas the Wyoming province is part of the Hudsonian craton. are present in widely scattered exposures across the Transcontinental Protero­

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 2 Research Paper

Figure 2. Geologic map of pre-Neogene rocks in central Arizona (modified from Richard et al., 2000), showing locations of Figures 3 and 7. Inset shows map location within Yavapai-Mazatzal tectonic province (adapted from Hoffman, 1988).

zoic provinces of the American Southwest (Fig. 1) and the upper Midwest that are Mesoproterozoic in age and hence younger than, and unrelated­ to, (Cox et al., 2002b; Dott, 2003; Medaris et al., 2003; Jones et al., 2009; Doe the Mazatzal orogeny (Doe et al., 2012; Mako et al., 2015). This has brought et al., 2012; Jones and Thrane, 2012; Mako et al., 2015). At least two aspects into question the existence of the Mazatzal orogeny because it had been of these rocks are significant. Deformation of these strata in Arizona and New partly defined on the basis of its effect on quartz-rich clastic sequences that Mexico had been attributed to the 1.6–1.7 Ga Mazatzal orogeny (e.g., Wilson, were thought to be Paleoproterozoic. Secondly, the unusually quartz-rich 1939; Karlstrom and Bowring, 1988, 1993; Doe and Karlstrom, 1991). Recent character of the clastic deposits is difficult to reconcile with their inferred syn- detrital-zircon geochronology has identified some quartz-rich clastic units to late-orogenic setting, and led to the interpretation of unusually effective

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 3 Research Paper

weathering that eliminated most non-quartzose minerals from the weather- determinations are based on 206Pb/207Pb for ages older than ca. 1.0 Ga. Each ing environment and produced deeply altered paleosols (Medaris et al., 2003; age-probability plot (also referred to as a probability-density plot or age-distri- Jones et al., 2009). bution curve) is the sum of all ages represented on the corresponding histo- gram with each age determination represented by a normal probability distri- bution with equal area (vertical scale is arbitrary). The age-distribution curves Purpose more accurately represent raw data because, unlike the histograms, they in- clude representation of analytical uncertainties. The histograms, in contrast, Geologic mapping and related laboratory analyses of two areas flanking provide a ready reference to determine how many grains are represented by a Chino Valley in central Arizona (Fig. 2), done as part of a long-term geologic probability peak or set of peaks. mapping program by the Arizona Geological Survey, identified features rele- vant to the origin and evolution of the Transcontinental Paleoproterozoic prov- inces in the Southwest. One purpose of this paper is to present evidence on JEROME CANYON AREA the nature and timing of crustal genesis in this part of the Yavapai tectonic province and to present evidence that it occurred partly through genesis of an Paleoproterozoic rocks in the Jerome Canyon area northwest of Prescott extensive tectonostratigraphic terrane of supracrustal rocks. Another purpose consist of a sequence of basaltic lava and breccia, calc-silicate, phyllite, and of this paper is to outline the geology of an ~1400-m-thick sequence of Paleo­ psammite. The sequence is intruded by the Williamson Valley Granodiorite proterozoic­ quartz arenite and conglomerate on the northeast side of Chino and the Mint Wash Granodiorite (Fig. 3; Krieger, 1967; DeWitt et al., 2008; Valley. The complete absence of feldspar and feldspar-bearing clasts in this Spencer and Young, 2011; Ferguson and Pearthree, 2013). Basaltic rocks are rock unit and in an underlying paleosol is attributed to an incompletely under­ interbedded with phyllite and psammite, plus the following rock types, none stood environment of extreme chemical weathering. Widespread, quartz-­ of which contain grains or mica, and all of which are suspected to con- arenite deposition across the Southwest and mid-continent regions has been sist at least partially of chemical precipitates from Paleoproterozoic seawater: referred to as a “signature event” of the late Paleoproterozoic (Jones et al., (1) Dark-gray, resistant, ferruginous, massive to laminated silica, with thin 2009), a concept that we support with an extreme example. (<10 mm) white and dark layers. (2) Hard, epidote-silica rock with pervasive epidote-green color, locally with elongate pits and thin (<10 mm) layers that likely contained, or contain, carbonate. (3) Thinly interlayered, mottled silica U-Pb ISOTOPE ANALYSIS METHODS and gray carbonate. (4) Laminated, talc and/or pyrophyllite paper-schist with

Supplemental Table 1. U-Pb analytical data. AnalysisU206Pb U/Th 206Pb*±207Pb*±206Pb*± error206Pb*±207Pb* ±206Pb*±Best age± Conc a soapstone-like character (Spencer and Young, 2011). These basaltic and fine- (ppm)204Pb 207Pb*(%) 235U*(%) 238U (%)corr.238U* (Ma) 235U (Ma) 207Pb* (Ma) (Ma) (Ma) (%) 1 1 1 1 Sample 2-14-11-1, Mint Wash Granodiorite (strikethrough data not used in final age determination due to high analytical uncertainty or low percent concordance) (n=25 for age) 2-14-11-1-8R 10550990 3.19.8589 1.74.15971.8 0.2974 0.70.401678.610.81666.115.11650.531.31650.531.3101.7 2-14-11-1-8C 1121102492.9 9.7781 1.04.14641.2 0.2941 0.50.471661.88.0 1663.59.5 1665.718.91665.718.999.8 2-14-11-1-18C 13372211 1.59.7714 0.84.21771.4 0.2989 1.10.831685.916.91677.511.31667.014.31667.014.3101.1 Samples of two granitoids and three were crushed and zircon grained rocks form an ~15-km-long, steeply dipping, north-northeast–striking 2-14-11-1-10C 11995977 4.39.7609 1.04.23181.3 0.2996 0.90.671689.213.01680.210.71669.018.01669.018.0101.2 2-14-11-1-19R 15317775 2.79.7389 0.74.25371.2 0.3005 1.00.831693.614.91684.49.9 1673.112.41673.112.4101.2 2-14-11-1-16R 13398287 1.99.7324 1.14.33241.3 0.3058 0.70.541720.110.41699.610.61674.420.11674.420.1102.7 2-14-11-1-24R 1541743221.4 9.7249 0.54.28291.9 0.3021 1.80.971701.627.21690.115.41675.88.5 1675.88.5 101.5 2-14-11-1-23R 16381845.0 9.7231 0.93.96561.2 0.2797 0.80.691589.612.01627.29.9 1676.116.21676.116.294.8 2-14-11-1-13R 12672530 2.59.7222 0.44.28730.7 0.3023 0.60.841702.78.7 1690.95.7 1676.37.0 1676.37.0 101.6 grains separated by density and magnetic susceptibility. Zircon grains were belt (Fig. 3). 2-14-11-1-9R 1961414162.6 9.7107 0.54.32961.3 0.3049 1.20.921715.717.81699.010.61678.59.3 1678.59.3 102.2 2-14-11-1-9C 12466320 3.29.7101 0.84.26151.1 0.3001 0.80.701691.911.41686.08.9 1678.614.21678.614.2100.8 2-14-11-1-11R 11667919 2.99.7061 0.94.24851.6 0.2991 1.30.821686.719.31683.413.11679.417.01679.417.0100.4 2-14-11-1-10R' 10355303 1.49.7054 0.64.06015.1 0.2858 5.00.991620.572.01646.341.21679.510.41679.510.496.5 2-14-11-1-22R 11165551 2.89.6959 0.94.32721.2 0.3043 0.80.671712.612.51698.610.21681.316.91681.316.9101.9 then embedded in epoxy along with a geochronologic standard zircon from Sri Metasedimentary rocks are moderately to strongly cleaved to weakly 2-14-11-1-26R 13746182.3 9.6880 1.14.37912.2 0.3077 1.90.861729.428.71708.418.11682.820.31682.820.3102.8 2-14-11-1-7R 1851153014.1 9.6862 0.54.19970.7 0.2950 0.50.671666.67.0 1674.05.9 1683.19.9 1683.19.9 99.0 2-14-11-1-20R 11375192 2.59.6786 0.74.24171.1 0.2978 0.90.771680.212.91682.19.4 1684.613.51684.613.599.7 2-14-11-1-21R 15624641 1.49.6689 0.84.25752.3 0.2986 2.20.951684.232.71685.219.21686.413.81686.413.899.9 2-14-11-1-2R 1831650872.2 9.6650 0.64.28431.1 0.3003 0.90.841692.913.41690.48.9 1687.210.91687.210.9100.3 2-14-11-1-1C 10321411.1 9.6634 1.03.83671.9 0.2689 1.60.841535.221.51600.515.11687.518.61687.518.691.0 Lanka (Sri-Lanka 2; see Gehrels et al., 2008, and supplemental file 1 inGehrels ­ ­schistose. Bedding is apparent but probably transposed by penetrative defor- 2-14-11-1-27R 13786560 1.59.6627 0.94.37282.0 0.3064 1.80.881723.226.51707.216.41687.617.11687.617.1102.1 2-14-11-1-3R 22557883.0 9.6614 0.73.85457.5 0.2701 7.51.001541.2102.2 1604.260.41687.912.11687.912.191.3 2-14-11-1-25R 15557302.1 9.6573 0.84.26521.2 0.2987 0.90.771685.113.61686.79.8 1688.714.11688.714.199.8 2-14-11-1-6R 646 554 2.0 9.6247 3.4 3.8390 19.9 0.2680 19.6 0.99 1530.5 267.5 1601.0 161.8 1694.9 62.7 1694.9 62.7 90.3 2-14-11-1-1R 12064587 2.49.6135 1.14.40163.4 0.3069 3.20.941725.448.31712.628.01697.020.81697.020.8101.7 and Pecha, 2014, for additional details on laser ablation–multicollector–induc- mation. Where bedding is apparent, strata are almost everywhere plane bed- 2-14-11-1-25R' 13513837 2.49.6038 1.24.27391.8 0.2977 1.30.741679.919.41688.414.51698.921.81698.921.898.9 2-14-11-1-29R 186 34891 2.1 9.5725 3.2 4.3875 3.9 0.3046 2.2 0.56 1714.1 32.5 1710.0 32.0 1704.9 59.2 1704.9 59.2 100.5 2-14-11-1-5R 306 828 2.5 9.5055 2.5 3.9031 2.6 0.2691 0.5 0.19 1536.1 6.8 1614.3 20.9 1717.8 46.6 1717.8 46.6 89.4 2-14-11-1-12R 192 1363 2.2 9.4840 4.2 4.3495 4.4 0.2992 1.3 0.29 1687.3 19.1 1702.8 36.5 1722.0 77.8 1722.0 77.8 98.0 2-14-11-1-28R 377 905 2.8 9.4784 1.3 3.0855 8.4 0.2121 8.3 0.99 1240.0 93.3 1429.1 64.3 1723.1 24.8 1723.1 24.8 72.0 2-14-11-1-3C 152 816 3.0 9.4579 3.3 3.7238 5.0 0.2554 3.8 0.76 1466.4 49.9 1576.5 40.2 1727.1 60.0 1727.1 60.0 84.9 tively coupled plasma–mass spectrometry [LA-MC-ICP-MS] analytical proce- ded and typically thin bedded to laminated. rarely contains graded 2-14-11-1-17R 150 3229 2.8 9.2437 9.4 3.8766 11.3 0.2599 6.1 0.55 1489.3 81.7 1608.8 91.2 1769.0 172.9 1769.0 172.9 84.2 dures and data reduction and presentation). The epoxy mounts were then beds or ripple cross laminations. Stratigraphic top direction was recognized 1 Supplemental Table 1. Isotopic data and calculated abraded to a depth of ~20 µm, polished, imaged, and cleaned before place- at only three locations based on truncation of low-angle cross beds or ripple ages for all zircon grains dated during this study. Please visit http://​dx.doi​ .org​ /10​ ​.1130/GES01339​ .S1​ ment in the laser chamber. An ~35 µm spot on each zircon grain was ­ablated cross laminations, and these were conflicting (two with tops to the east). Fold- or the full-text article on www​.gsapubs.org​ to view with an excimer laser, mobilized in a helium carrier gas, and accelerated down ing probably had a significant effect on this rock sequence and may account Supplemental Table 1. the mass-spectrometer flight tube in static magnetic mode with multiple Fara­ for the conflicting facing directions. day collectors, one for each analyzed isotope (238U, 232Th, 208Pb, 207Pb, 206Pb, Cleavage, schistosity, and bedding are all steep and northeast strik-

Supplemental Table 2. U-Pb and paleosol-geochemistry sample locations 204 202 1 Sample no. Latitude Longitude UTMN UTMEDatum zone Notes GPS Station Pb, and Hg) (analytical data are included in Supplemental Table 1 ; sample ing. The average dips of cleavage and schistosity are slightly steeper than 2-15-11-3 34.707008-112.572745 3841678355968 NAD83 12 Medium to fine grain sandstone near Jerome Canyon JES-11-193 Mint Wash granodiorite of DeWitt - Sample of fresh 2-14-11-1 34.682364-112.560197 3838927357075 NAD83 12 granodiorite from blasted road cut JES-11-133 2 4-15-11-1 34.7426983 -112.56363845623 356867 NAD83 12 Williamson Valley Granodiorite of DeWitt JES-11-481 locations are included in Supplemental Table 2 ). aver­age bedding (Fig. 4). Metamorphic mica is parallel to cleavage within 11-4-08-9 34.823058-112.422938 3854343369872 NAD83 12 Del Rio Quartzite, top of section JES-09-034 11-5-08-1 34.813694-112.403918 3853280371597 NAD83 12 Del Rio Quartzite, base of section JES-09-043 6-20-15-1 34.8143132 -112.4031643853349 371666 NAD83 12 Paleosol JES-15-164 207 206 6-20-15-2 Paleosol sample from between 6-20-15-1 and 6-20-15-4 All of the ages reported here are derived from measurement of Pb/ Pb. ~1–2 km of the irregular intrusive contact with the undeformed Williamson 6-20-15-3 Paleosol sample from between 6-20-15-1 and 6-20-15-5 6-20-15-4 34.8144608 -112.4028959 3853365371691 NAD83 12 Paleosol JES-15-165 Because of low concentrations, 235U is not measured but is calculated from Valley Granodiorite. Within a few tens of meters of this contact, however, 2Supplemental Table 2. Location information for U-Pb measured 238U (235U = 238U/137.82). 207Pb/235U age is calculated from mea- metamorphic mica is coarser, preferred orientation is less pronounced, and geochronology samples and paleosol geochemis- sured 206Pb/238U and measured 207Pb/206Pb [207Pb/235U = (206Pb/(238U/137.82))/ well-defined schistosity is lacking, reflecting conditions of static recrystalli- try samples. Please visit http://dx​ ​.doi​.org/10​ ​.1130​ 206 207 206 207 /GES01339.S2​ or the full-text article on www​.gsapubs​ ( Pb/ Pb)]. Two-sigma uncertainty in measured Pb/ Pb age is ~1%–2% zation. Directly adjacent to the NNE-striking, linear shear-zone contact with .org to view Supplemental Table 2. for zircon crystals older than ca. 1 Ga (Gehrels et al., 2008). Histograms of age Mint Wash Granodiorite (Fig. 3), units have neither sec-

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 4 Research Paper

Figure 3. Geologic map of the Jerome Canyon area north of Prescott (see Fig. 2 for location). Figure 4. Stereonets of poles to cleavage, schistosity, and bedding in the deformed metasedimentary rocks of the Jerome Canyon area Adapted from Spencer and Young (2011). (data plotted and eigenvalues calculated using stereonet software from Allmendinger et al. [2013] and Cardozo and Allmendinger [2013]).

ondary mica nor phyllite sheen, so it is inferred that displacement along Two samples of granitic rocks and one of metasandstone were analyzed for this contact juxtaposed the two units (farther south, the contact is intrusive U-Pb geochronology at the Arizona Laserchron Center at the University of Ari­ but was not examined in detail). Strong mylonitic fabrics are present lo- zona. Fourteen U-Pb LA-ICP-MS spot analyses of zircons from the Williamson cally along the contact, primarily in the granite but locally in metasedimen- Valley Granodiorite yielded an age of 1736 ± 21 Ma (2s uncertainty including tary rocks. Lineation is insufficiently developed to determine displacement all internal and external uncertainties). Twenty-five U-Pb analyses of zircons direction. from the Mint Wash Granodiorite yielded an age of 1680 ± 16 Ma (2s; Fig. 5).

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 5 Research Paper

A C

B D

Figure 5. (A, C) 206Pb/207Pb dates of zircon phenocrysts from the Williamson Valley and Mint Wash granodiorites. (B, D) Concordia diagrams for analyses shown in A and C. Plots made using Isoplot 3.6 (Ludwig, 2008). For the Williamson Valley Granodiorite, one analysis was excluded from the age calculation because 1σ uncer- tainty was greater than 100 Ma. For the Mint Wash Granodiorite, seven analyses were excluded from the age calculation because 1σ uncertainty was greater than 50 Ma or concordance was less than 90%.

Seventy-five detrital-zircon grains from a sample of fine-grained sandstone older than known local rock units. Also, 19 of 75 dated grains are within the were analyzed for U-Pb age. The youngest peak in the probability density narrow age range of 2467–2485 Ma (mean 2480 Ma; Fig. 6). This narrow range plot, at 1738 Ma (Fig. 6), indicates the maximum possible depositional age. of dates, representing ~25% of the analyzed zircon grains, suggests a single, Detrital-zircon analyses yielded abundant dates typical of the Yavapai Province unidentified igneous source that is ca. 740 Ma older than typical dated igneous (e.g., Karlstrom et al., 1987) as well as seven dates at 1800–1900 Ma that are rocks in the Yavapai Province in Arizona.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 6 Research Paper

could be significantly greater. The rock unit was originally considered an out- lier of Mazatzal Quartzite (Wilson, 1922, 1939; Krieger, 1965; Bradshaw, 1975; A ­Trevena, 1979), which is well exposed in the Mazatzal Mountains to the south- east (Fig. 2); however, doubts that the unit can be confidently correlated over the intervening ~100 km led to informal designation as the Del Rio Quartzite (Bayne, 1987; Chamberlain et al., 1991; Gootee et al., 2010). Conglomerate and conglomeratic sandstone of the Del Rio Quartzite are poorly to moderately sorted, with clasts of white vein quartz and less abun- dant red jasper and argillite within a matrix of poorly to moderately sorted, reddish-brown sand and pebbly sand. Clasts are typically subrounded to sub- angular (Fig. 8A). Quartzite and pebbly quartzite are typically cross stratified in medium- to thick-bedded trough, wedge, and tabular sets, commonly with sweeping asymptotic basal cross strata with black-sand laminations (Fig. 8B). The upper conglomerate unit in northwestern exposures is generally plane bedded, with massive and graded beds of cobble and pebble conglomerate (Fig. 8C). Even thin pebble beds are commonly planar and laterally continuous for many meters (Figs. 8D and 8E). Channels are rare and, where present, are broad and shallow (Fig. 8E). Some coarse sandstone and pebbly conglomerate B are characterized by planar cross beds (Fig. 8F). In thin section, quartzite that makes up most of the Del Rio Quartzite con- sists of quartz grains with variably abundant, commonly rounded opaque grains and a sparse (generally <10%) microcrystalline to cryptocrystalline pseudomatrix consisting largely of very fine secondary mica. Examination of 32 thin sections stained for plagioclase and K-feldspar did not identify any de- trital grains of feldspar, mica, or mafic silicates. The hardest and most pure quartzite consists almost entirely of quartz grains with sutured, complex grain boundaries (Fig. 9A). Quartz overgrowths are common, with subrounded cores and crystallographically continuous overgrowths (Fig. 9B). In other samples, quartz grains are poorly sorted and consist of mixed grains that vary from sub- rounded to angular (Fig. 9C). Some angular quartz grains are surrounded by fine secondary matrix (Fig. 9D). Abundant heavy-mineral laminations in sand- stone consist of rounded opaque grains generally 75–500 mm in diameter (Figs. 9E and 9F). Electron microprobe analysis indicates that the opaque grains are Figure 6. (A) Histogram and age-probability plot of 206Pb/207Pb dates of detrital-zircon hematite or maghemite, and their very weak magnetic susceptibility strongly grains from a sample of Proterozoic metasandstone from the Jerome Canyon area. Histo­ gram age bins are 20 m.y. (B) Concordia diagram for analyses shown in (A). Plots made suggests they are hematite. Some grains contain up to 20% solid solution to- using Isoplot 3.6 (Ludwig, 2008). ward the ilmenite end member (Fig. 10). Some grains also exhibit exsolved lamellae and rims of rutile and minor lamellae of ilmenite. At one location, hematite grains form much or most of a black-sand unit several meters thick DEL RIO QUARTZITE (Gootee et al., 2010). Detrital-zircon grains from quartzite samples from the base and top of the The Paleoproterozoic Del Rio Quartzite along lower Granite Creek near Little­ Del Rio Quartzite yielded age spectra dominated by ca. 1.65–1.95 Ga zircons, Chino Valley consists of a heterogeneous sequence of interbedded quartzite, with a scattering of grains as old as 3.4 Ga (Fig. 11). The youngest peak in pebbly quartzite, pebble to cobble conglomerate, and sparse, variably silty the age-probability plot for the stratigraphically lowest sample, at 1745 Ma, argillite­ (Fig. 7; Krieger, 1965; Trevena, 1979; Bayne, 1987; Gootee et al., 2010). is the maximum possible depositional age for strata near the base of the se- A slightly disrupted, 1400-m-thick section is exposed in the northwestern limb quence (Fig. 11A). The youngest peak in the probability-density plot for the of an anticline in the northwestern area of quartzite exposures (Fig. 7; Brad- stratigraphically highest sample, at 1737 Ma, is the maximum possible deposi- shaw, 1975). The top of the Quartzite is covered and total thickness of the unit tional age for strata near the top of the sequence (Fig. 11C).

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 7 Research Paper

Figure 7. Geologic map and cross section of the Del Rio Quartzite and associated rocks (see Fig. 1 for location). Note that a deformation zone near the center of the map results in uncertainty regarding strati- graphic relations across this zone. Offset across this zone is not shown in the cross section but could be significant. The base of the Quartzite is shown as if the outcrop at the core of the anticline, here interpreted as a paleosol, marks the stratigraphic base of the Quartzite. The basal contact is, how- ever, concealed. The strike of bedding in Paleozoic­ strata roughly follows the con- tact with the under­lying Proterozoic rocks and dips gently away from exposures of Del Rio Quartzite, as if differential compac- tion caused the variable dips.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 8 Research Paper

A B

C D Figure 8. Del Rio Quartzite: (A) Conglomer- ate clasts, dominated by vein quartz, range from subrounded to angular (red arrows point to angular clasts). (B) Channel mar- gin in pebbly quartzite with black-sand laminations in lower conglomerate. (C) The upper conglomerate unit in northwestern exposures is generally plane bedded, with massive and graded beds of ­cobble and pebble conglomerate. (D, E) Red arrows point to thin pebble beds, which are com- monly planar and laterally continuous for many meters. (E) Channels are rare in the upper conglomerate and, where present, are broad and shallow (channel is ~3 m wide). (F) Some coarse sandstone and pebbly conglomerate are characterized by planar cross beds in tabular sets.

E F

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 9 Research Paper

A B

C D Figure 9. (A) Del Rio Quartzite quartz grains commonly have sutured, complex grain boundaries (1-mm scale bar, crossed nicols). (B) Quartz overgrowths are not uncommon, with subrounded cores and epitaxial overgrowths (1-mm scale bar, crossed nicols). (C) Poorly sorted quartzite with grains that vary from subrounded to angular (1-mm scale bar, crossed nicols). (D) Some angular quartz grains are sur- rounded by fine secondary matrix and ap- pear to be primary angular grains and not products of postdepositional fracturing and crushing (1-mm scale bar, plain light). (E) Abundant opaque-mineral laminations in sandstone consist of rounded hematite grains (1-mm scale bar, reflected light). (F) Amalgamated heavy-mineral lamina- tions at the base of a cross-bed set.

E F

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 10 Research Paper

is covered and is possibly depositional. Discordance between the ledge and bedding to the southeast suggests that the paleosol is faulted on the southeast, although this contact is also concealed. The paleosol consists of ~20%, <7 mm quartz and clumps of multiple quartz grains in a mixture of very fine, opaque and translucent minerals or mineraloids (Fig. 12). Em- bayed boundaries and rounded margins characterize some quartz grains (Figs. 12C and 12D). Abundant microfracturing in some quartz grains (Fig. 12E) is similar to such fracturing in the quartzite. Relict grains of uncertain protolith have no extinction under crossed nicols and are possibly clay minerals derived from feldspar (Fig. 12F). We sampled the paleosol for geochemical analysis at four locations along the length of the outcrop belt

Supplemental Table 3. Geochemical analyses of Del Rio paleosol Bureau Veritas Commodies Canada Ltd. Final Report Client: Arizona Geological Survey and at variable but poorly constrained relative paleodepths. The average File Created: 16-Jul-15 Job Number:VAN15001606 Number of Samples: 4 Project: STATEMAP 2015 Received:2-Jul-15 of four very similar chemical analyses indicates a chemical composition of Method WGHT LF200 LF200 LF200 LF200 LF200 LF200LF200 LF200LF200 LF200LF200 LF200LF200 LF200

AnalyteWeightSiO 2 Al 2O 3 Fe 2O 3 MgOCaO Na 2OK2OK2O/Na 2OTiO 2 P 2O 5 MnOCr 2O 3 Ni Sc LOI Unit KG %%%%%%%%%%%PPM PPM % MDL0.010.010.010.040.010.010.01 0.01 0.01 0.01 0.010.002 20 1-5.1 SampleType 6-20-15-1Rock0.3879.87 10.484.880.070.030.18 1.43 7.9 0.18 0.02 0.03<0.002<20 9 2.7 silica (79.2% SiO2), aluminum (10.7% Al2O3), iron (5.1% Fe2O3), potassium 6-20-15-2Rock0.4779.17 9.34 7.03 0.08 0.03 0.15 1.28 8.5 0.17 0.02 0.04<0.002<20 9 2.6 6-20-15-3Rock0.4679.36 10.015.870.120.060.16 1.72 10.8 0.18 0.01 0.02<0.002<20 11 2.4 6-20-15-4Rock0.4679.23 10.665.060.120.050.17 1.66 9.8 0.18 0.01 0.02<0.002<20 9 2.7 average79.41 10.125.710.100.040.17 1.52 9.2 0.180.015 0.0289.5 2.6 Pulp Duplicates (1.66% K O), and volatiles lost on melting (2.7% loss on ignition) that to- 6-20-15-4Rock0.4679.23 10.665.060.120.050.17 1.66 9.8 0.18 0.01 0.02<0.002<20 9 2.7 2 6-20-15-4REP Reference Materials STD GS311-1STD STD GS910-4STD STD SO-18STD 57.9114.15 7.72 3.46.383.67 2.16 0.69 0.8 0.40.557 50 24 1.9 gether make up 99.4% of the unit. Analytical data are included in Supple- STD SO-19STD 60.4613.95 7.55 2.87 5.89 4.08 1.3 0.69 0.31 0.130.494 47227 1.9 STD DS10 STD STD OREAS45EASTD BLKBLK 3 BLKBLK <0.01<0.01 <0.04 <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 <0.002<20 <1 0 BLKBLK mental Table 3 . Prep Wash ROCK-VANPrep Blank 70.7814.14 30.832.4 4.52 2.1 0.36 0.08 0.08<0.002<20 7 1.5

3Supplemental Table 3. Major and trace element geo- chemistry results for four paleosol samples below INTERPRETATION the Del Rio Quartzite. Please visit http://dx​ .doi​ .org​ ​ /10​.1130/GES01339​ .S3​ or the full-text article on www​ Figure 10. Electron microprobe analyses of opaque grains indicate entirely hematitic .gsapubs.org​ to view Supplemental Table 3. compositions, including some with high titanium and some with exsolved rutile. Geochronology of the Jerome Canyon Area

Static recrystallization of phyllitic Jerome Canyon strata near the un­ Contacts between the Del Rio Quartzite and other Proterozoic units are deformed Williamson Valley Granodiorite resulted in mica growth parallel to covered or faulted, and the Quartzite is depositionally overlain by flat-lying to cleavage at low levels of recrystallization but with increasingly random ori- gently tilted Phanerozoic strata (Fig. 7). Nearby Proterozoic units include psam- entations of metamorphic mica with greater levels of recrystallization closer mitic, biotite-sericite schist that records penetrative deformation that did not to the intrusion. It remains uncertain, however, if the Jerome Canyon strata ­affect the Del Rio Quartzite, and an isolated exposure of pillow ­basalt unaffected­ were tilted to near vertical before cleavage development, or if gently inclined by penetrative deformation (Krieger, 1965; Gootee et al., 2010). Although the cleavage developed before tilting to steep dips. The preponderance of steep, Del Rio Quartzite is folded, it is not affected by cleavage development or meta- generally east-northeast–striking cleavage in Paleoproterozoic rocks across morphic mineral growth except that recrystallization of pseudomatrix in sand- central and southeastern Arizona suggests that steep cleavage is a general stone and silty argillite is apparent in thin section but not significant enough feature reflecting consolidation of the newly formed Paleoproterozoic crust, to impart cleavage or micaceous sheen in hand samples. Silty argillite from with steep cleavage perpendicular to shortening direction and approximately the Quartzite contains “incompletely ordered illite, kaolinite, and pyrophyllite, parallel to the convergent continental margin. We conclude that deposition compatible with temperatures of 280–400 °C” (Gillentine et al., 1991). Further- of the Jerome Canyon sandstone, tilting to steep dips, and later cleavage de- more, Gillentine et al. (1991) state that if the pyrophyllite is detrital rather than velopment, all occurred after the ca. 1738 Ma maximum depositional age of metamorphic, peak temperatures were lower, at 120–200 °C. Folding of the Del the sandstone as indicated by the youngest age-probability peak, and before Rio Quartzite, as is apparent on the cross section in Figure 7, likely occurred intrusion of the 1736 ± 21 Ma (2s) Williamson Valley Granodiorite. It is possible during latest Paleoproterozoic orogenesis (Karlstrom and Bowring, 1993) or that strata that are stratigraphically below the analyzed sample, including the during the Mesoproterozoic (Doe et al., 2012; Mako et al., 2015). basalts, are significantly older. The ca. 1738 Ma maximum depositional age and the 1736 ± 21 Ma (2s) Paleosol below the Del Rio Quartzite age of the Williamson Valley Granodiorite are essentially identical, yet the true ages of these units are separated by the time necessary for complete deposi- A massive, reddish-brown, uncleaved rock unit interpreted as a paleo­ tion of the clastic sequence, tilting to steep dips, and cleavage development. sol is exposed below the Quartzite (Fig. 7) in an ~5-m-wide ledge that paral- All this could have occurred within a few million years, within an active con- lels bedding in the Quartzite to the northwest. The contact between the two vergent plate margin, and within the uncertainties of the analyses (which are

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 11 Research Paper

A C

B D

Figure 11. (A, C) Histogram and age-probability plots of 206Pb/207Pb dates of detrital-zircon grains from the Del Rio Quartzite. Histogram age bins are 50 m.y. (B, D) Concordia diagrams for analyses shown in (A) and (C). Plots made using Isoplot 3.6 (Ludwig, 2008).

unspecified for the age-probability peak but probably similar to that for the Significance of circa 2480 Ma Zircon Grains Granodiorite). It is also possible that metamorphic zircons or metamorphic zir- in the Jerome Canyon Sandstone con overgrowths have displaced the maximum depositional age to a younger age than the actual age of deposition (e.g., Jacobson et al., 2015). We consider Nineteen of 75 U-Pb zircon dates from the Jerome Canyon sandstone fall this unlikely, however, because the youngest U-Pb ages in the sandstone were in the range 2467–2485 Ma. All but one have individual analytical uncertainty not derived from the rims of zircon grains, and U/Th for zircon grains younger less than 8 Ma (1s), with an average of 4.7 Ma. The age and uncertainty of than 1750 Ma are all less than 3 and are not anomalous compared to older the age-probability peak (Fig. 6), 2480.2 ± 27.4 Ma, includes consideration of grains. The short period of sedimentation, tilting, deformation, and granitoid systematic error, but considering only the uncertainty associated with each intrusion reveals a rapid pace of what appears to be genesis of a small part of individual analysis (without consideration of systematic error), the weighted Yavapai province crust. mean and uncertainty is 2480.2 ± 1.9 Ma. In other words, reproducibility and

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 12 Research Paper

AABB

CCDD

Figure 12. Photomicrographs of the Del Rio paleosol, with crossed nicols in all images. (A) Large, chemically(?) rounded quartz grain with two embayments at lower right. (B) Euhedral quartz crystal (lower). (C) Cluster of quartz crystals with rounded corners, embayments, and straight bound- aries that could have been the margins of adjacent feldspar crystals before alteration. (D) Embayed quartz crystal. (E) Highly fractured quartz crystal as with some Del Rio Quartzite. (F) Relict minerals in bright, finely mottled area with no extinction in crossed nicols.

EFE F

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 13 Research Paper

precision are very high compared to accuracy. The mean square of weighted for which 1s = 24–45 Ma). The three low-uncertainty analyses by Shufeldt deviations (MSWD) of the nineteen analyses is 0.65, which means that the et al. (2010) include a subset of zircons that cluster similarly tightly around peak in the age-probability plot is narrower than would be expected for the 2475–2480 Ma, and all of the others contain a sub-population of about this analytical uncertainties associated with the individual analyses (the individual age. Kolmogorov-Smirnov (K-S) statistical comparison of all U-Pb dates from analyses are “underdispersed”). This in turn suggests that analytical uncer- the 13 samples (Fig. 14) determined a P value of 0.51 for comparison of the tainty associated with individual analyses is overestimated for earliest Protero- Jerome Canyon sample and the westernmost schist sample (Fig. 13) analyzed zoic zircon grains. The similarity and precision of these dates suggest that a by Shufeldt et al. (2010). This means that the two samples can’t be statistically single igneous rock body or cluster of related units was the source for a large distinguished from each other with even 50% confidence and could have been fraction of the sand grains in the Jerome Canyon sandstone. derived from the same source. A distinctive population of similar-age zircons was also identified in ten Kolmogorov-Smirnov statistical analysis compares populations to deter- samples of Vishnu Schist (Ilg et al., 1996; Karlstrom et al., 2012) from ~150 km mine if they can be distinguished from each other with some level of con- to the north-northeast in eastern Grand Canyon, and in two samples of fidence. The null hypothesis is that two samples can’t be statistically dis- schist from western Grand Canyon (Billingsley et al., 2006), ~140 km to the tinguished from each other. This would be the situation for small data sets northwest (Fig. 13; Shufeldt et al., 2010). The 1s uncertainty for zircon U-Pb where there are too little data to determine if two samples were derived from analyses from the Jerome Canyon sample and for three of the 12 samples different sources (the null hypothesis can’t be disproved). With larger data analyzed by Shufeldt et al. (2010) are 4.5–5.5 Ma (unlike the other samples, sets, statistically different populations can be better distinguished. In gen- eral, a 95% confidence in statistical distinctness is considered definitive with detrital-zircon U-Pb dates (Guynn and Gehrels, 2010). A P value of 0.05 corre- sponds to 95% confidence that two samples can’t have come from the same population. Below this confidence limit (P > 0.05), samples appear more sim- ilar and may have been derived from the same source. It is noteworthy that P values of 0.093 and 0.068, representing comparison of the Jerome Canyon sample with two samples of Vishnu Schist in eastern Grand Canyon, indicate that these samples can’t be distinguished with 95% confidence, although they can be distinguished with 90% confidence. Calculation of K-S values reported in Figure 14 includes uncertainty in individual analyses, and these values are a better test of similarity or difference if analytical uncertainty is small. One of the two Vishnu Schist samples that is statistically indistinguish- able from the Jerome Canyon sample at the 95% confidence level (but not at the 90% level) also has very low analytical uncertainty associated with individual analyses. Paleoproterozoic Vishnu Schist from the eastern Grand Canyon is a fine- grained psammitic and pelitic unit that is associated with the Brahma and Rama schists (Ilg et al., 1996). These units consist of mafic and felsic schists and gneisses that are interlayered with hypabyssal, volcanic, or volcaniclastic layers dated at 1750 ± 2 Ma and 1741 ± 1 Ma. The three units are described as follows by Hawkins et al. (1996): “comparable units of the Rama, Vishnu, and Brahma Schists are interlayered at several localities in the Upper Gorge transect, suggesting that these units compose a complex volcanic and sedi- mentary package characterized by spatial and temporal lithologic variation.” These lithologies match well with the Jerome Canyon strata with the minor difference that Jerome Canyon strata contains significant iron-silica rocks and calc-silicate chemical sediments. Furthermore, the maximum depositional age of the Jerome Canyon psammite is essentially identical to the eastern Grand Figure 13. Map of Proterozoic bedrock exposures in central and northwest Arizona and Canyon supracrustal rocks. adjacent areas showing named shear zones and detrital-zircon sample locations as re- ported here for the Jerome Canyon psammite (number 1) and reported by Shufeldt et al. Because of detrital-zircon statistical similarity of the 12 Grand Canyon (2010) for 12 detrital-zircon samples of Grand Canyon schists (numbers 2–13). samples, Shufeldt et al. (2010) concluded that the name “Vishnu Schist”

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 14 Research Paper

Figure 14. P values from Kolmogorov-Smirnov (K-S) statistical analysis, using error in the cumulative distribution function, of a detrital-zircon sample of Jerome Canyon psammite and 12 detrital-­ zircon samples of Grand Canyon schists collected and analyzed by Shufeldt et al. (2010). P values less than 0.05 (white boxes) are sample-pair comparisons with greater than 95% statistical confidence that these could not have been derived from the same source (the null hypothesis of statistical indistinguishability is disproved). P values greater than 0.05 correspond to sample pairs that can’t be statistically distinguished from each other with more than 95% confidence (the null hypothesis is not disproved). Samples in yellow boxes with red numbers cannot be distinguished from each other with more than 50% confidence, and are most similar to each other. Note that Jerome Canyon sample 1 and lower Grand Canyon sample 2 WGC-1 (sample-number prefixes) are statistically distinguishable with no more than 49% confidence (P = 0.51). P values and statistical confidence in distinctness are inversely related.

could be applied to all of the sampled Grand Canyon schists. As shown in Depositional Environment of the Del Rio Quartzite Figure 13, this schist terrane extends for ~150 km across regional tectonic strike and broadly corresponds to the Hualapai block of Karlstrom and Bow- Asymptotic and planar cross beds and hematite laminations are common in ring (1993). The zircon population from the Jerome Canyon sample contains Del Rio sandstones. Sorting is typically poor, with abundant white pebbles and a higher proportion of ca. 2480 Ma grains than all but one of the 12 Grand granules in poorly sorted sands (Fig. 8). Channels are sparse in sandstones. Canyon samples and a lower proportion of Archean (>2500 Ma) grains than Some channels and cross beds contain coarse basal debris. All of these fea- all but one (a different one) of the 12 Grand Canyon samples. The ratio of ca. tures are consistent with near-shore marine and braided-stream depo­sitional 2480 Ma to >2500 Ma dates for the Jerome Canyon sample is larger (2.2) than environments. However, poor to very poor sorting, especially in conglomerate for any of the Grand Canyon samples (0.2–1.2). These characteristics of the units, supports a fluvial environment. Planar cross beds of pebble to granule Jerome Canyon detrital zircons suggest that these strata were closer to the conglomeratic sandstone with basal cobble lag (Fig. 8F) are readily interpreted source of the ca. 2480 Ma zircon grains than most or all of the Grand Canyon as the result of aggrading bars in a braided stream environment. The sparse- schists, and Jerome Canyon strata were not simply derived from erosion and ness of channels in the upper conglomerate unit, with planar, laterally con- dispersal of strata now represented by the Vishnu Schist in Grand Canyon. tinuous pebble and cobble conglomerate beds (Figs. 8C–8E) indicates rapid Considering the lithologic, geochronologic, and detrital-zircon–population changes in flow velocity on a planar stream bed. similarities between the Jerome Canyon strata and the Vishnu Schist, we Quartzite and conglomerate in the Del Rio Quartzite are unusual because suggest that the Jerome Canyon strata could be correlated with the Vishnu they are highly mature chemically, containing little but quartz, jasper, and Schist, and that the entire area encompassed by the 13 samples is a single hematite,­ but are physically immature, as indicated by angular quartz sand tectonostratigraphic terrane. One difference between Jerome Canyon strata grains and angular conglomerate clasts. We attribute this to unusually effec- and Vishnu Schist is that the dominant, steep fabric in Jerome Canyon strata tive chemical weathering in a surficial environment reminiscent of modern, predates intrusion of the 1736 ± 21 Ma (2s) Williamson Valley Granodiorite, tropical fluvial environments. For example, where sediments have traveled

whereas similar fabric in the eastern Grand Canyon (S2 of Ilg et al. [1996] and many hundreds of kilometers in low-energy river systems such as the Amazon­ Dumond et al. [2007]) formed at 1713–1685 Ma (Hawkins et al., 1996). This and Orinoco, chemical destruction is especially effective where sands are inter­ difference in timing of deformations, however, is unrelated to correlation of mittently stored in point-bar deposits with variable degrees of soil develop- the schists. ment (Johnsson and Meade, 1990; Johnsson et al., 1991). The Del Rio Quartzite

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 15 Research Paper

was not, however, deposited by a meandering tropical river, as indicated by derived from granitic and/or metamorphic host rocks, but the absence of clasts clast angularity and large size, poor sorting, and absence of point-bar deposits. of such host rocks suggests that weathering destroyed these rocks before they Complete destruction of feldspar occurs in some poorly drained tropi­cal soils, reached the surface and were dispersed by fluvial processes. but sands derived from such soils typically carry a small to moderate compo- nent of feldspar and lithic grains (Suttner et al., 1981; van Hattum et al., 2006; Hall and Smyth, 2008; Garzanti et al., 2013). While Del Rio sediments have Paleoenvironmental Implications of the Del Rio Paleosol some similarities to modern tropical sediments, the striking contrast between extreme chemical maturity and physical immaturity in these braided stream The Del Rio paleosol has a distinctive bulk chemical composition consistent deposits suggests chemical weathering of greater effectiveness than in any with extreme weathering of the protolith. What is most diagnostic and unusual

modern fluvial environment. is elevated average SiO2 (79%–80%) and Fe2O3 (5%–7%) and very low CaO

Like the Del Rio Quartzite, Mazatzal Group Quartzite and similar late Paleo­ (<0.10%) and Na2O (<0.20%). Several weathering indices provide a measure of proterozoic quartzite in the southern Rocky Mountains and the mid-continent weathering effects by comparing the ratio of mobile (or weatherable elements region include units that are devoid or nearly devoid of feldspar and lithic such as Ca and Na) to immobile elements (such as Ti or Zr). For example, the fragments but have interstitial alteration products interpreted as relict feld- widely used chemical index of alteration minus potassium, or CIA – K = 100 ×

spar and lithic grains (Dott, 1983; Soegaard and Eriksson, 1989; Cox et al., Al2O3 /( Al2O3 + CaO + Na2O), returns very high values of 97.8–98.1, consistent 2002b; Medaris et al., 2003; Jones et al., 2009). The quartz-rich character of with almost total loss of mobile elements (Supplemental Table 3 [see footnote Mazatzal Group quartzites has been attributed to unusually effective dia­ 3]). Potassium is excluded here because of suspected potassium metasomatism genetic alteration of feldspar and lithic grains (Cox et al., 2002a), but thin-sec- indicated by K/Na of ~10 (e.g., Driese and Medaris, 2008). tion examination of some Del Rio Quartzite samples reveals almost no inter­ The presence of embayed and rounded quartz in the Del Rio paleosol stitial material. If feldspar had been present before diagenetic alteration, (Figs. 12A, 12C, and 12D) is also consistent with effective chemical weather- interstitial alteration products likely would be present in greater abundance ing. Saprolitic soils in the southeastern United States are known for embayed and would form relict grains. The quartz-rich composition of mid-continent and chemically rounded quartz grains in soils developed on metamorphic and (Baraboo) quartzites has been attributed to derivation from deeply weath- igneous basement (Cleary and Conolly, 1971, 1972), as well as in Australia ered bedrock in which feldspar and mafic silicates had been altered to clays (Crook, 1968) and in buried paleosols (Lander et al., 1991). Embayed quartz and oxides by weathering and saprolitic soil development, possibly aided by has also been recognized in glacial till where corrosion is thought to be the near-surface physical stabilization by microbial mats for extended periods of result of repeated periods of evaporation that increase the pH of residual aque- time (Dott, 2003; Medaris et al., 2003; Driese and Medaris, 2008). We propose ous fluids (Krauskopf, 1956; May, 1980). Embayed and rounded quartz are also that similar if not identical weathering and soil development, followed by present in igneous rocks due to mixing of magmas with appropriate compo- erosion, sediment dispersal, and concentration of quartz grains by winnow- sitions (Vernon, 1986; Donaldson and Henderson, 1988; Watt et al., 1997) and ing of fines, is responsible for the composition of the Del Rio Quartzite. Unlike have been recognized in some Paleoproterozoic metamorphic rocks in nearby the Baraboo interval quartzites, however, we doubt that the Del Rio Quartzite areas (Anderson and Creasey, 1958). We consider it possible but unlikely that was deposited on a low-relief, stable platform. Rather, significant relief at the Del Rio paleosol was derived from igneous rocks with embayed quartz, in the time of deposition seems likely, both to provide a source of coarse sand part because igneous quartz phenocrysts with embayments as large and well and conglomerate and to accommodate deposition of >1400 meters of clastic developed as shown in Figure 12D are rare and we know of none in Arizona sediments. Paleoproterozoic rocks. As noted above, conglomerate clasts in the Del Rio Quartzite consist largely We infer that the protolith of the Del Rio paleosol was a granitoid, as indi- of vein quartz, with less common jasper and argillite, and with no evidence of cated by abundant large quartz grains and clusters of quartz grains in a matrix diagenetic replacement of other clast types with alteration products. Common that is plausibly dominated by highly altered feldspar. This interpretation is and highly visible hematitic laminations in pebbly sandstone and conglom- also supported by the straight margins of some quartz-grain clusters (Fig. 12C) erate were not deformed by the mechanical collapse of voids following or that are interpreted to reflect boundaries with earlier-crystallized feldspar in accompanying diagenetic elimination of granitoid and other feldspar-bearing a granitoid. The chemical composition of the paleosol, compared to 1.7 Ga conglomerate clasts. Similarly, no secondary-mineral void fillings were iden- granitoids in central and northern Arizona, suggests that pedogenic chemical

tified where originally feldspar-bearing clasts had been. We infer that nearly processes reduced Na2O by 94%–95% and CaO by 96%–98%. Such weather- pure quartz compositions of conglomerate clasts are primary and that alter- ing losses are consistent with modern, low-latitude, warm, and generally wet ation and destruction of feldspar and lithic fragments were highly effective tropical environments (e.g., Gardner and Walsh, 1996; Guan et al., 2001; Li and in near-surface environments before clast transport and deposition. It is par- Yang, 2010; Betard, 2012). Comparable losses are also observed in the 2.2 Ga ticularly significant that abundant vein-quartz clasts likely would have been Hekpoort Paleosol (Rye and Holland, 2000) and have been interpreted to indi-

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 16 Research Paper

cate intense weathering not due to wet climate but to elevated atmospheric In summary, 43 of 44 U-Pb dates from these areas define the time of magma-

pCO2 in the Early Proterozoic (Sheldon, 2006). We can’t distinguish the cause tism in the Yavapai tectonic province in central Arizona at 1755–1662 Ma. To

(high H2O, high pCO2, or both) with our analysis, but our results demonstrate the southeast, in areas that are considered to be part of the Mazatzal tectonic intense weathering conditions just before Del Rio Quartzite deposition in cen- province or at least close to it, U-Pb dates are largely similar or somewhat tral Arizona. Indeed, weathering and dispersal of soils such as that represented younger. Eleven of 13 dates listed by Karlstrom et al. (1987) are 1695–1710 Ma, by the Del Rio paleosol would yield abundant quartz grains without feldspar with two younger granitoids dated at 1640 and 1630 Ma. Five granitoids within or mafic silicates. 70 km of the city of Phoenix yielded U-Pb dates of 1632–1644 Ma (Isachsen et al., 1999; Spencer et al., 2003). In conclusion, igneous rocks in a wide region around the Del Rio Quartzite could have supplied 1750–1630 Ma zircons to the Age of Del Rio Quartzite Deposition and Folding Quartzite and allow for the possibility that the Quartzite is significantly younger than its maximum depositional age as determined by age-probability peaks. As noted above, the maximum depositional age of the Del Rio Quartzite is The timing of folding of the Del Rio Quartzite is poorly constrained. Over- indicated by the 1745 Ma age-probability peak for the sample near the base of lying and surrounding Paleozoic strata are generally flat lying and not mod- the Quartzite and the 1737 Ma age-probability peak for the sample near the top erately to strongly folded like the Quartzite (Fig. 7). Directly adjacent to the (Fig. 11). The large range of individual zircon dates, the significant number of Del Rio Quartzite, Paleozoic strata appear to be slightly folded, possibly due multiple peaks on the age-probability plots, and numerous older dates extend- to differential compaction (Fig. 7; Krieger, 1965). Paleozoic strata in the Transi- ing back to the Archean, all indicate that the zircon grains were derived from tion Zone of central Arizona were affected locally by Phanerozoic deformation. diverse sources (or possibly from a sedimentary unit that had previously accu- Paleozoic strata to the northwest of the Del Rio Quartzite are deformed into mulated zircons from diverse sources). The fact that the youngest age-proba- open, northwest-striking folds of probable Laramide (50–70 Ma) age (Ferguson bility peak is older for the sample at the stratigraphic base than for the sample et al., 2012; Pearthree and Ferguson, 2012) that are nearly at right angles to the near the top suggests that the age peaks approximate depositional ages, but northeast-striking folds in the much more strongly folded Del Rio Quartzite. In the two peaks are so close in age that the difference is probably not statistically eastern and southeastern Arizona, the widespread Mesoproterozoic Apache significant. The fact that 17 of 93 zircon dates from the stratigraphically higher Group, dated near its base at 1328 ± 5 Ma (Stewart et al., 2001), is flat lying ex- sample are 1626–1695 Ma suggests that at least the upper part of the Quartzite cept locally where affected by Mesoproterozoic diabase intrusion or by much is younger than the 1737 Ma age peak. Furthermore, 10 of these 17 dates are later and more severe Phanerozoic deformation (Shride, 1967; Wrucke, 1989). more than 70 Ma younger than the age peak, which is notable because the Extrapolation of post–1328 Ma tectonic stability to central Arizona suggests average 2s analytical uncertainty for these dates is ±70 Ma. Deformation and that the Del Rio Quartzite was folded before ca. 1328 Ma. In conclusion, folding metamorphism of the nearby Jerome Canyon area at ca. 1740 Ma probably likely occurred after the main phase of Yavapai orogenesis at ca. 1700–1740 Ma occurred before deposition of the nearby uncleaved and unmetamorphosed (Karlstrom and Bowring, 1993) and exhumation of deformed metamorphic Del Rio Quartzite and genesis of the underlying paleosol, although these two rocks, and before ca. 1328 Ma. Folding plausibly occurred during the early areas could have been separated by greater distance at that time. Directly to Mesoprotero­ zoic­ as it did in areas to the southeast in the Mazatzal province the northeast of the main exposure of Quartzite is a psammitic quartz-sericite (e.g., Doe et al., 2012; Mako et al., 2015). schist similar to the Jerome Canyon strata (Krieger, 1965; Gootee et al., 2010). In addition to the young zircon grains in the Quartzite, the contrast in levels of deformation and metamorphism, both locally and regionally, also suggests Implications for the Late Paleoproterozoic Atmosphere that the Del Rio Quartzite is significantly younger than ca. 1740 Ma. The broad range of zircon grains in the Del Rio Quartzite that are younger The extreme effectiveness of chemical weathering during deposition of the than ca. 1750 Ma plausibly represents derivation from igneous rocks in the sur- Del Rio Quartzite, with similarities to modern tropical weathering conditions, rounding region, as indicated by the following: (1) South and east of the Chino could be interpreted as the result of higher temperatures than in modern en- Valley area, eight of nine U-Pb dates listed by Karlstrom et al. (1987) indicate vironments. This is problematic, however, because at 1.7 Ga solar radiation igneous activity from 1755 Ma to 1699 Ma. The ninth date (ca. 1800 Ma), on was 87% of modern levels (Gough, 1981; Bahcall et al., 2001; Ribas, 2009). the Cleopatra Rhyolite near the United Verde mine (Fig. 2; Lindberg, 2008), Indeed, a primary paleoclimatological problem of Precambrian Earth has been was more recently dated at 1738.5 ± 0.5 Ma (Slack et al., 2007). (2) In the area to determine how Earth avoided snowball or ice-age conditions during most of around the Bagdad mine ~80 km to the west-southwest, Paleoproterozoic rock its history (e.g., Feulner, 2012). Conditions that could have elevated late Paleo-

units yielded six U-Pb isochron dates of 1720–1677 Ma (Bryant et al., 2001). proterozoic temperatures include greater concentrations of CO2 (e.g., Kasting, (3) In the Grand Canyon area ~150 km to the north, 28 of 29 dates of Paleo­ 1987; Kiehl and Dickinson, 1987; Sheldon, 2006; Bekker and Kaufman, 2007),

proterozoic­ rocks listed by Karlstrom et al. (2012) range from 1750 to 1662 Ma. CH4 (Pavlov et al., 2003; Kasting, 2005), climatic consequences of Earth’s past

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 17 Research Paper

lower land area and higher rotation rate (Jenkins et al., 1993), a thicker atmo- maturity of quartzite and conglomerate is moderate, however, with a range of sphere (Goldblatt et al., 2009; Som et al., 2012), and reduced oxygen levels rounding and the presence of angular and subangular sand grains and con-

(Poulsen et al., 2015). Only higher CO2 levels will obviously contribute to the glomerate clasts. Chemical maturity with physical immaturity in these braided-­ corrosiveness of the atmosphere. We suggest that hot, humid greenhouse stream deposits supports the concept of exceptionally effective chemical

conditions associated with very high atmospheric CO2 concentration produced weathering in late Paleoproterozoic time, as does the severe alteration of the exceptionally corrosive, carbonic-acid–bearing rainwater that caused rapid paleosol. This aspect of Paleoproterozoic quartzites has been recognized pre- and effective chemical weathering during deposition of the Del Rio Quartzite. viously (e.g., Cox et al., 2002b; Jones et al., 2009), but unlike Cox et al. (2002b), The Del Rio Quartzite is exceptional in representing these conditions because we attribute chemical maturation to surficial rather than diagenetic processes of its proximal fluvial depositional setting and is an example of the “signature because of the absence of alteration products in some orthoquartzites and, es- lithology” of syntectonic quartzites that appear to represent unique climatic pecially, because conglomerate clasts consist entirely of vein quartz and jasper conditions during the late Paleoproterozoic (Jones et al., 2009). without feldspar, mica, or mafic silicates.

ACKNOWLEDGMENTS CONCLUSION J. Spencer thanks Mike Doe for discussions of Proterozoic geology, Charles Ferguson for com- ments on an earlier draft, Jay Holberg for discussions regarding stellar evolution, and Beth Nichols Several conclusions derived from this study are as follows: Boyd and Diane Love for assistance in the field. We thank Associate Editor Mike Williams, reviewer Jamey Jones, and an anonymous reviewer for thorough reviews that resulted in substantial im- (1) In the Jerome Canyon area, deposition of a sequence of siltstone, sand- provement. U-Pb geochronologic analyses of single zircon grains were done by LA-ICP-MS at stone, calc-silicate, and basalt, with a 1738 Ma maximum depositional age of a the Arizona Laserchron Center at the University of Arizona. LaserChron Center facilities support psammite sample, was followed by tilting to near vertical and deformation re- was provided by National Science Foundation grant EAR-1338583. Field mapping and laboratory studies were supported by the U.S. Geological Survey National Cooperative Geologic Mapping sulting in cleavage and weak schistosity. This in turn was followed by intrusion Program under STATEMAP assistance awards 08HQAG0093 and G10AC00428. The views and of the undeformed 1736 ± 21 Ma (2s) Williamson Valley Granodiorite. conclusions contained in this document are those of the authors and should not be interpreted as (2) Jerome Canyon sandstone was derived in part from an unidentified ca. necessarily representing the official policies, either expressed or implied, of the U.S. Government. This manuscript is submitted for publication with the understanding that the United States Gov- 2480 Ma rock unit, most likely an intrusion or amalgamation of similar-age ernment is authorized to reproduce and distribute reprints for governmental use. intrusions. This rock unit was also a source of zircons in the Vishnu Schist in eastern Grand Canyon and similar schist in western Grand Canyon (Shufeldt REFERENCES CITED et al., 2010). This suggests that the Yavapai-Mojave orogenic collage locally Allmendinger, R.W., Cardozo, N.C., and Fisher, D., 2013, Structural Geology Algorithms: Vectors included fault blocks of much older rock. & Tensors: Cambridge, UK, Cambridge University Press, 289 p. (3) The statistical similarity of U-Pb dates from detrital zircons in the ­Jerome Anderson, C.A., and Creasey, S.C., 1958, Geology and ore deposits of the Jerome area, Yavapai Canyon sample with three samples of schist from the Grand Canyon, and simi- County, Arizona: U.S. Geological Survey Professional Paper 308, 185 p., with plates. Bahcall, J.N., Pinsonneault, M.H., and Basu, S., 2001, Solar models: Current epoch and time de- larities among the Grand Canyon schist samples (Shufeldt et al., 2010), support pendencies, neutrinos, and helioseismological properties: The Astrophysical Journal, v. 555, the existence of a Paleoproterozoic tectonostratigraphic terrane that encom- p. 990–1012, doi:​10​.1086​/321493​. passes all the samples and is approximately coincident with the Hualapai block Bayne, B.J., 1987, Depositional analysis of conglomerates in the Mazatzal Group and related strata, central Arizona [M.S. thesis]: Flagstaff, Northern Arizona University, 182 p. of Karlstrom and Bowring (1993). Bekker, A., and Kaufman, A.J., 2007, Oxidative forcing of global climate change: A biogeochemi­ (4) Quartzite and conglomerate that make up the Del Rio Quartzite were de- cal record across the oldest Paleoproterozoic ice age in North America: Earth and Planetary posited sometime after ca. 1745 Ma (base) to ca. 1737 Ma (top). The significant Science Letters, v. 258, p. 486–499, doi:​10​.1016​/j​.epsl​.2007​.04​.009​. Bennett, V.C., and DePaolo, D.J., 1987, Proterozoic crustal history of the western United States as fraction of younger ages represented by the broad left slope of the age-prob- determined by neodymium isotopic mapping: Geological Society of America Bulletin, v. 99, ability plots for the two samples suggests that the actual age of the quartzite p. 674–685, doi:​10​.1130​/0016​-7606​(1987)99​<674:​PCHOTW>2​.0​.CO;2​. is significantly younger than the two maximum depositional ages, as does the Bergh, S.G., and Karlstrom, K.E., 1992, The Chaparral shear zone: Deformation partitioning and fact that the Quartzite is not affected by the metamorphism and penetrative heterogeneous bulk crustal shortening during Proterozoic orogeny in central Arizona: Geo- logical Society of America Bulletin, v. 104, p. 329–345, doi:​10.1130​ /0016​ ​-7606(1992)104​ <0329:​ ​ deformation that affected rocks in surrounding areas. TCSZDP>2​.3​.CO;2​. (5) Quartzite in the Del Rio Quartzite consists almost exclusively of quartz Betard, F., 2012, Spatial variations of soil weathering processes in a tropical mountain environ- grains with minor hematite. Conglomerate contains clasts of vein quartz and ment: The Baturite massif and its piedmont (Ceara, NE Brazil): Catena, v. 93, p. 18–28, doi:​ 10​.1016​/j​.catena​.2012​.01​.013​. less common argillite and jasper, with a complete absence of clasts contain- Bickford, M.E., and Hill, B.M., 2007, Does the arc accretion model adequately explain the Paleo- ing feldspar, mica, or mafic silicate minerals. Similarly, an associated paleosol proterozoic evolution of southern Laurentia?: An expanded interpretation: Geology, v. 35, contains only quartz grains in a microcrystalline to cryptocrystalline matrix, is p. 167–170, doi:​10​.1130​/G23174A​.1​. Billingsley, G.H., Block, D.L., and Dyer, H.C., 2006, Geologic Map of the Peach Springs 30′ × 60′ severely leached of sodium and calcium, and contains embayed and rounded Quadrangle, Mohave, and Coconino Counties, Northwestern Arizona: U.S. Geological Sur- quartz grains that are characteristic of some deeply weathered soils. Physical vey Scientific Investigations Map SIM-2900, scale 1:100,000.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 18 Research Paper

Bradshaw, E.C., 1975, Structure in the Mazatzal Quartzite, Del Rio, Arizona [M.S. thesis]: Flag- crustal architecture, processes at 25-km-deep levels, and unroofing of orogens: Geological staff, Northern Arizona University, 134 p. Society of America Bulletin, v. 119, p. 202–220, doi:​10.1130​ /B25903​ ​.1​. Bryant, B., Wooden, J.L., and Nealey, L.D., 2001, Geology, geochronology, geochemistry, and Eisele, J., and Isachsen, C.E., 2001, Crustal growth in southern Arizona: U-Pb geochronologic and Pb-isotopic compositions of Proterozoic rocks, Poachie region, west-central Arizona—A Sm-Nd isotopic evidence for addition of the Paleoproterozoic Cochise block to the Mazatzal study of the east boundary of the Proterozoic Mojave crustal province: U.S. Geological Sur- province: American Journal of Science, v. 301, p. 773–797, doi:​10.2475​ /ajs​ ​.301​.9​.773​. vey Professional Paper 1639, 54 p. Ferguson, C.A., and Pearthree, P.A., 2013, Geologic map of the Sullivan Buttes 7 1/2′ Quadrangle, Cardozo, N., and Allmendinger, R.W., 2013, Spherical projections with OSXStereonet: Com­ Yavapai County, Arizona: Arizona Geological Survey Digital Geologic Map DGM-102, scale puters & Geosciences, v. 51, no. 0, p. 193–205, doi:10​ .1016​ /j​ ​.cageo​.2012.07​ ​.021​. 1:24,000. Chamberlain, K.R., Karlstrom, K.E., and Bowring, S.A., 1991, U-Pb age constraints on deposition Ferguson, C.A., Gootee, B.F., Pearthree, P.A., and Cook, J.P., 2012, Geologic map of the Paulden of the quartzite at Del Rio, and 1.72 Ga deformation in the Chino Valley area in northern 7. 5 ′ quadrangle, Yavapai County, Arizona: Arizona Geological Survey Digital Geologic Map Yavapai County, Arizona: Geological Society of America Abstracts with Programs, v. 23, p. 11. DGM-91, scale 1:24,000. Cleary, W.J., and Conolly, J.R., 1971, Distribution and genesis of quartz in a piedmont-coastal Feulner, G., 2012, The faint young sun problem: Reviews of Geophysics, v. 50, RG2006, doi:10​ ​ plain environment: Geological Society of America Bulletin, v. 82, no. 10, p. 2755–2766, doi:​10​ .1029​/2011RG000375​. .1130​/0016​-7606​(1971)82​[2755:​DAGOQI]2​.0​.CO;2​. Gardner, R., and Walsh, R., 1996, Chemical weathering of metamorphic rocks from low eleva- Cleary, W.J., and Conolly, J.R., 1972, Embayed quartz grains in soils and their significance: Jour- tions in the southern Himalaya: Chemical Geology, v. 127, p. 161–176, doi:​10.1016​ /0009​ ​-2541​ nal of Sedimentary , v. 42, p. 899–904. (95)00089​-5​. Condie, K.C., 1982, Plate tectonics model for Proterozoic continental accretion in the southwest- Garzanti, E., Padoan, M., Andò, S., Resentini, A., Vezzoli, G., and Lustrino, M., 2013, Weathering ern United States: Geology, v. 10, p. 37–42, doi:​10​.1130​/0091​-7613​(1982)10​<37:​PMFPCA>2​.0​ and relative durability of detrital minerals in equatorial climate: Sand petrology and geo- .CO;2​. chemistry in the East African Rift, v. 121, p. 547–580. Condie, K.C., 2007, Accretionary orogens in space and time, in Hatcher, R.D., Jr., Carlson, M.P., Gehrels, G.E., and Pecha, M., 2014, Detrital zircon U-Pb geochronology and Hf isotope geochem- McBride, J.H., and Martínez Catalán, J.R., eds., 4-D Framework of Continental Crust: Geo­ istry of Paleozoic and Triassic passive margin strata of western North America: Geosphere, logi­cal Society of America Memoir 200, p. 145–158, doi:10​ .1130​ /2007​ ​.1200​(09)​. v. 10, no. 1, p. 49–65, doi:10​ .1130​ /GES00889​ ​.1​. Cox, R., Gutmann, E.D., and Hines, P.G., 2002a, Diagenetic origin for quartz-pebble conglomer- Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy, efficiency, and ates: Geology, v. 30, p. 323–326, doi:​10​.1130​/0091​-7613​(2002)030​<0323:​DOFQPC>2​.0​.CO;2​. spatial resolution of U-Pb ages by laser ablation–multicollector–inductively coupled plasma– Cox, R., Martin, M.W., Comstock, J.C., Dickerson, L.S., Ekstrom, I.L., and Sammons, J.H., 2002b, mass spectrometry: Geochemistry Geophysics Geosystems, v. 9, Q03017, doi:​10.1029​ ​ Sedimentology, stratigraphy, and geochronology of the Proterozoic Mazatzal Group, central /2007GC001805.​ Arizona: Geological Society of America Bulletin, v. 114, no. 12, p. 1535–1549, doi:​10.1130​ ​ Gillentine, J.M., Karlstrom, K.E., Parnell, R.A., Jr., and Puls, D., 1991, Constraints on temperatures /0016​-7606​(2002)114​<1535:​SSAGOT>2​.0​.CO;2​. of Proterozoic metamorphism in low-grade rocks of central Arizona, in Karlstrom, K.E., ed., Crook, K.A.W., 1968, Weathering and roundness of quartz sand grains: Sedimentology, v. 11, Proterozoic Geology and Ore Deposits of Arizona: Arizona Geological Society Digest, v. 19, p. 171–182, doi:​10​.1111​/j​.1365​-3091​.1968​.tb00851​.x​. p. 165–180. Darrach, M.E., Karlstrom, K.E., Argenbright, D.N., and Williams, M.L., 1991, Progressive defor- Goldblatt, C., Claire, M.W., Lenton, T.M., Matthews, A.J., Watson, A.J., and Zahnle, K.J., 2009, mation in the early Proterozoic Shylock shear zone, central Arizona, in Karlstrom, K.E., ed., Nitrogen-­enhanced greenhouse warming on early Earth: Nature Geoscience, v. 2, p. 891– Proterozoic Geology and Ore Deposits of Arizona: Arizona Geological Society Digest 19, 896, doi:​10​.1038​/ngeo692​. p. 97–116. Gootee, B.F., Ferguson, C.A., Spencer, J.E., and Cook, J.P., 2010, Geologic map of the Chino Valley DeWitt, E., Langenheim, V., Force, E., Vance, R.K., Lindberg, P.A., and Driscoll, R.L., 2008, Geo- North 7 1/2′ Quadrangle, Yavapai County, Arizona: Arizona Geological Survey Digital Geo- logic Map of the Prescott National Forest and the Headwaters of the Verde River, Yavapai logic Map DGM-80, scale 1:24,000. and Coconino Counties, Arizona: U.S. Geological Survey Scientific Investigations Map 2996, Gough, D.O., 1981, Solar interior structure and luminosity variations: Solar Physics, v. 74, p. 21– scale 1:100,000, with 100 p. text. 34, doi:​10​.1007​/BF00151270​. Doe, M.F., and Karlstrom, K.E., 1991, Structural geology of an Early Proterozoic foreland thrust Guan, P., Ng, C.W.W., Sun, M., and Tang, W., 2001, Weathering indices for rhyolitic tuff and granite belt, Mazatzal Mountains, Arizona, in Karlstrom, K.E., ed., Proterozoic Geology and Ore De- in Hong Kong: Engineering Geology, v. 59, p. 147–159, doi:10​ .1016​ /S0013​ ​-7952(00)00071​ ​-5.​ posits of Arizona: Arizona Geological Society Digest 19, p. 181–192. Guynn, J., and Gehrels, G., 2010, Comparison of detrital zircon age distributions using the K-S Doe, M.F., Jones, J.V., III, Karlstrom, K.E., Thrane, K., Frei, D., Gehrels, G., and Pecha, M., 2012, test: Arizona Laserchron Center, 16 p. Basin formation near the end of the 1.60–1.45 Ga tectonic gap in southern Laurentia: Meso- Hall, R., and Smyth, H.R., 2008, Cenozoic arc processes in Indonesia: Identification of the key proterozoic Hess Canyon Group of Arizona and implications for ca. 1.5 Ga supercontinent influences on the stratigraphic record in active volcanic arcs, in Draut, A.E., Clift, P.D., and configurations: Lithosphere, v. 4, p. 77–88, doi:10​ .1130​ /L160​ .1​ ​. Scholl, D.W., eds., Formation of the Sedimentary Record in Arc Collision Zones: Geological Donaldson, C.H., and Henderson, C.M.B., 1988, A new interpretation of round embayments in Society of America Special Paper 436, p. 27–54, doi:10​ .1130​ /2008​ ​.2436​(03)​. quartz crystals: Mineralogical Magazine, v. 52, p. 27–33, doi:10​ .1180​ /minmag​ .1988​ .052​ .364​ .02​ .​ Hamilton, W.B., 2007, Earth’s first two billion years—The era of internally mobile crust, in Dott, R.H., Jr., 1983, The Proterozoic red quartzite enigma in the north-central United States: Re- Hatcher, R.D., Jr., Carlson, M.P., McBride, J.H., and Martínez Catalán, J.R., eds., 4-D Frame- solved by plate collision?, in Medaris, L.G., Jr., Early Proterozoic Geology of the Great Lakes work of Continental Crust: Geological Society of America Memoir 200, p. 233–296, doi:​10​ Region: Geological Society of America, Memoir 160, p. 129–143, doi:​10.1130​ /MEM160​ -p129​ .​ .1130​/2007​.1200​(13)​. Dott, R.H., Jr., 2003, The importance of eolian abrasion in supermature quartz sandstones and Hawkins, D.P., Bowring, S.A., Ilg, B.R., Karlstrom, K.E., and Williams, M.L., 1996, U-Pb geo­chrono­ the paradox of weathering on vegetation-free landscapes: The Journal of Geology, v. 111, logic constraints on the Paleoproterozoic crustal evolution of the Upper Granite Gorge, p. 387–405, doi:​10​.1086​/375286​. Grand Canyon, Arizona: Geological Society of America Bulletin, v. 108, no. 9, p. 1167–1181, Driese, S.G., and Medaris, L.G., Jr., 2008, Evidence for biological and hydrological controls on doi:​10​.1130​/0016​-7606​(1996)108​<1167:​UPGCOT>2​.3​.CO;2​. the development of a Paleoproterozoic paleoweathering profile in the Baraboo Range, Wis- Hoffman, P.F., 1988, United plates of America, the birth of a craton: Early Proterozoic assembly consin, U.S.A.: Journal of Sedimentary Research, v. 78, p. 443–457, doi:​10.2110​ /jsr​ ​.2008​.051​. and growth of Laurentia: Annual Review of Earth and Planetary Sciences, v. 16, p. 543–603, Duebendorfer, E.M., Chamberlain, K.R., and Fry, B., 2006, Mojave-Yavapai boundary zone, south- doi:​10​.1146​/annurev​.ea​.16​.050188​.002551​. western United States: A rifting model for the formation of an isotopically mixed crustal Ilg, B.R., Karlstrom, K.E., Hawkins, D.P., and Williams, M.L., 1996, Tectonic evolution of Paleo­ boundary zone: Geology, v. 34, p. 681–684, doi:10​ .1130​ /G22581​ ​.1​. proterozoic­ rocks in the Grand Canyon: Insights into middle-crustal processes: Geologi- Dumond, G., Mahan, K., Williams, M.W., and Karlstrom, K.E., 2007, Metamorphism in middle cal Society of America Bulletin, v. 108, p. 1149–1166, doi:10​ .1130​ /0016​ ​-7606(1996)108​ <1149:​ ​ continental crust, Upper Granite Gorge, Grand Canyon, Arizona: Implications for segmented TEOPRI>2​.3​.CO;2​.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 19 Research Paper

Iriondo, A., Premo, W.R., Martínez-Torres, L.M., Budahn, J.R., Atkinson, W.W., Jr., Siems, D.F., Krauskopf, K.B., 1956, Dissolution and precipitation of silica at low temperatures: Geochimica et and Guarás-González, B., 2004, Isotopic, geochemical, and temporal characterization of Cosmochimica Acta, v. 10, p. 1–26, doi:​10​.1016​/0016​-7037​(56)90009​-6​. Proterozoic­ basement rocks in the Quitovac region, northwestern Sonora, Mexico: Implica- Krieger, M.H., 1965, Geology of the Prescott and Paulden Quadrangles, Arizona: U.S. Geological tions for the reconstruction of the southwestern margin of Laurentia: Geological Society of Survey Professional Paper 467, 127 p., with maps, scale 1:48,000. America Bulletin, v. 116, no. 1/2, p. 154–170, doi:​10.1130​ /B25138​ .1​ ​. Krieger, M.H., 1967, Reconnaissance geologic map of the Iron Springs quadrangle, Yavapai Isachsen, C.E., Gehrels, G.E., Riggs, N.R., Spencer, J.E., Ferguson, C.A., Skotnicki, S.J., and Rich- County, Arizona [Iron Springs, Skull Valley, Mount Josh, and Jerome Canyon 7.5 min]: U.S. ard, S.M., 1999, U-Pb geochronologic data from zircons from eleven granitic rocks in central Geological Survey Miscellaneous Geologic Investigations Map I-504, 1 sheet, scale 1:62,500. and western Arizona: Arizona Geological Survey Open-File Report 99-5, 27 p. Lander, R.H., Bloch, S., Mehta, S., and Atkinson, C.D., 1991, Burial diagenesis of paleosols in Jacobson, C.E., Hourigan, J., Haxel, G.B., Grove, M.J., and Hoyt, J.J., 2015, Orocopia Schist at the Giant Yacheng gas field, People’s Republic of China: Bearing on illite reaction pathways: Cemetery Ridge, Arizona: A type exposure of subducted rocks accreted during Laramide flat Journal of Sedimentary Petrology, v. 61, p. 256–268. subduction, then incorporated into a Miocene Cordilleran core complex: Geological Society Li, C., and Yang, S., 2010, Is chemical index of alteration (CIA) a reliable proxy for chemical of America Abstracts with Programs, v. 47, no. 7, p. 432. weathering in global drainage basins: American Journal of Science, v. 310, p. 111–127, doi:​ Jenkins, G.S., Marshall, H.G., and Kuhn, W.R., 1993, Precambrian climate: The effects of land 10​.2475​/02​.2010​.03​. area and Earth’s rotation rate: Journal of Geophysical Research, v. 98, p. 8785–8791, doi:10​ ​ Lindberg, P.A., 2008, Early Proterozoic volcanogenic massive sulfide ore deposits, Jerome, Arizona, .1029​/93JD00033​. USA, in Spencer, J.E., and Titley, S.R., eds., Ores and Orogenesis: Circum-Pacific Tectonics,­ Johnsson, M.J., and Meade, R.H., 1990, Chemical weathering of fluvial sediments during alluvial Geologic Evolution, and Ore Deposits: Arizona Geological Society Digest 22, p. 601–610. storage: The Macuapanim Island point bar, Solimões River, Brazil: Journal of Sedimentary Ludwig, K.R., 2008, Isoplot 3.6: Berkeley Geochronology Center, Special Publication 4, 77 p. Petrology, v. 60, p. 827–842. Mako, C.A., Williams, M.L., Karlstrom, K.E., Doe, M.F., Powicki, D., Holland, M.E., Gehrels, G., Johnsson, M.J., Stallard, R.F., and Lundberg, N., 1991, Controls on the composition of fluvial and Pecha, M., 2015, Polyphase Proterozoic deformation in the Four Peaks area, central Ari­ sands from a tropical weathering environment: Sands of the Orinoco River drainage basin, zona, and relevance for the Mazatzal orogeny: Geosphere, v. 11, p. 1975–1995, doi:10​ ​.1130​ Venezuela and Columbia: Geological Society of America Bulletin, v. 103, p. 1622–1647, doi:10​ ​ /GES01196​.1​. .1130​/0016​-7606​(1991)103​<1622:​COTCOF>2​.3​.CO;2​. May, R.W., 1980, The formation and significance of irregularly shaped quartz grains in till: Sedi- Jones, J.V., III, and Thrane, K., 2012, Correlating Proterozoic synorogenic metasedimentary suc- mentology, v. 27, p. 325–331, doi:​10​.1111​/j​.1365​-3091​.1980​.tb01182​.x​. cessions in southwestern Laurentia: New insights from detrital zircon U-Pb geochronology Medaris, L.G., Jr., Singer, B.S., Dott, R.H., Jr., Naymark, A., Johnson, C.M., and Schott, R.C., 2003, of Paleoproterozoic quartzite and metaconglomerate in central and northern Colorado, Late Paleoproterozoic Climate, Tectonics, and Metamorphism in the Southern Lake Superior U.S.A.: Rocky Mountain Geology, v. 47, no. 1, p. 1–35, doi:10​ .2113​ /gsrocky​ ​.47​.1​.1​. Region and Proto–North America: Evidence from Baraboo Interval Quartzites: The Journal Jones, J.V., III, Connelly, J.N., Karlstrom, K.E., Williams, M.L., and Doe, M.F., 2009, Age, prove- of Geology, v. 111, p. 243–257, doi:​10.1086​ /373967​ ​. nance, and tectonic setting of Paleoproterozoic quartzite successions in the southwestern Meijer, A., 2014, The Pinal Schist of southern Arizona: A Paleoproterozoic forearc complex with United States: Geological Society of America Bulletin, v. 121, no. 1/2, p. 247–264, doi:​10​ evidence of spreading ridge–trench interaction at ca. 1.65 Ga and a Proterozoic arc obduc- .1130​/B26351​.1​. tion event: Geological Society of America Bulletin, v. 126, p. 1145–1163, doi:10​ .1130​ /B31002​ ​.1​. Karlstrom, K.E., 1989, Early recumbent folding during Proterozoic orogeny in central Arizona, Pavlov, A.A., Hurtgen, M.T., Kasting, J.F., and Arthur, M.A., 2003, Methane-rich Proterozoic atmo­ in Grambling, J.A., and Tewksbury, B.J., Proterozoic Geology of the Southern Rocky Moun- sphere?: Geology, v. 31, p. 87–90, doi:​10​.1130​/0091​-7613​(2003)031​<0087:​MRPA>2​.0​.CO;2​. tains: Geological Society of America Special Paper 235, p. 155–172, doi:10​ .1130​ /SPE235​ ​ Pearthree, P.A., and Ferguson, C.A., 2012, Geologic map of the Wineglass Ranch 7.5′ quadrangle, -p155​. Yavapai County, Arizona: Arizona Geological Survey Digital Geologic Map DGM-90, scale Karlstrom, K.E., and Bowring, S.A., 1988, Early Proterozoic assembly of tectonostratigraphic 1:24,000. terranes in southwestern North America: The Journal of Geology, v. 96, p. 561–576, doi:​10​ Poulsen, C.J., Tabor, C., and White, J.D., 2015, Long-term climate forcing by atmospheric oxygen .1086​/629252​. concentrations: Science, v. 348, p. 1238–1241, doi:​10​.1126/science​ ​.1260670.​ Karlstrom, K.E., and Bowring, S.A., 1993, Proterozoic orogenic history of Arizona, in Van Schmus, Ribas, I., 2009, The Sun and stars as the primary energy input in planetary atmospheres, in W.R., and Bickford, M.E., eds., Transcontinental Proterozoic Provinces, in Reed, J.C., Jr., Kosovichev, A.G., Andrei, A.H., and Rozelot, J.-P., eds., Solar and Stellar Variability: Impact et al., eds., Precambrian: Conterminous U.S.: Geological Society of America, Decade of on Earth and Planets: Proceedings of the International Astronomical Union symposium North America Geology, v. C-2, p. 188–211. no. 264, 16 p. Karlstrom, K.E., Bowring, S.A., and Conway, C.M., 1987, Tectonic significance of an Early Protero- Richard, S.M., Reynolds, S.J., Spencer, J.E., and Pearthree, P.A., 2000, Geologic map of Arizona: zoic two-province boundary in central Arizona: Geological Society of America Bulletin, v. 99, Arizona Geological Survey Map 35, scale 1:1,000,000. p. 529–538, doi:​10​.1130​/0016​-7606​(1987)99​<529:​TSOAEP>2​.0​.CO;2​. Rye, R., and Holland, H.D., 2000, Geology and geochemistry of paleosols developed on the Hek- Karlstrom, K.E., Åhäll, K.-I., Harlan, S.S., Williams, M.L., McLelland, J., and Geissman, J.W., 2001, poort basalt, Pretoria Group, South Africa: American Journal of Science, v. 300, p. 85–141, Long-lived (1.8–1.0 Ga) convergent orogen in southern Laurentia, its extensions to Australia doi:​10​.2475​/ajs​.300​.2​.85​.

and Baltica, and implications for refining Rodinia: Precambrian Research, v. 111, p. 5–30, doi:​ Sheldon, N.D., 2006, Precambrian paleosols and atmospheric CO2 levels: Precambrian Research, 10​.1016​/S0301​-9268​(01)00154​-1​. v. 147, p. 148–155, doi:​10​.1016​/j​.precamres​.2006​.02​.004​. Karlstrom, K.E., Ilg, B.R., Hawkins, D., Williams, M.L., Dumond, G., Mahan, K., and Bowring, S.A., Shride, A.F., 1967, Younger Precambrian geology in southern Arizona: U.S. Geological Survey 2012, Vishnu basement rocks of the Upper Granite Gorge: Continent formation 1.84 to 1.66 Professional Paper 566, 89 p. billion years ago, in Timmons, J.M., and Karlstrom, K.E., eds., Grand Canyon Geology: Two Shufeldt, O.P., Karlstrom, K.E., Gehrels, G.E., and Howard, K.E., 2010, Archean detrital zircons in Billion Years of Earth’s History: Geological Society of America Special Paper 489, p. 7–24, the Proterozoic Vishnu Schist of the Grand Canyon, Arizona: Implications for crustal archi- doi:​10​.1130​/2012​.2489​(01)​. tecture and Nuna supercontinent reconstructions: Geology, v. 38, no. 12, p. 1099–1102, doi:​ Kasting, J.F., 1987, Theoretical constraints on oxygen and carbon dioxide concentrations in the 10​.1130​/G31335​.1​. Precambrian atmosphere: Precambrian Research, v. 34, p. 205–229, doi:10​ .1016​ /0301​ ​-9268​ Slack, J.F., Grenne, T., Bekker, A., Rouxel, O.J., and Lindberg, P.A., 2007, Suboxic deep seawater (87)90001​-5​. in the late Paleoproterozoic: Evidence from hematitic chert and iron formation related to Kasting, J.F., 2005, Methane and climate during the Precambrian era: Precambrian Research, seafloor-hydrothermal sulfide deposits, central Arizona, USA: Earth and Planetary Science v. 137, p. 119–129, doi:​10​.1016​/j​.precamres​.2005​.03​.002​. Letters, v. 255, p. 243–256, doi:​10​.1016​/j​.epsl​.2006​.12​.018​. Kiehl, J.T., and Dickinson, R.E., 1987, A study of the radiative effects of enhanced atmospheric Soegaard, K., and Eriksson, K.A., 1989, Origin of thick, first-cycle quartz arenite successions:

CO2 on early earth surface temperatures: Journal of Geophysical Research, v. 92, p. 2991– Evidence from the 1.7 Ga Ortega Group, northern New Mexico: Precambrian Research, v. 43, 2998, doi:​10​.1029​/JD092iD03p02991​. p. 129–141, doi:​10​.1016​/0301​-9268​(89)90008​-9​.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 20 Research Paper

Som, S.M., Catling, D.C., Harnmeijer, J.P., Polivka, P.M., and Buik, R., 2012, Air density 2.7 bil- terminous U.S.: Geological Society of America, Decade of North America Geology (DNAG), lion years ago limited to less than twice modern levels by fossil raindrop imprints: Nature, v. C-2. p. 270–281. v. 484, p. 359–362, doi:​10​.1038​/nature10890​. Vernon, R.H., 1986, Evaluation of the “quartz-eye” hypothesis: Economic Geology and the Bul- Spencer, J.E., and Young, J.J., 2011, Geologic map of the Jerome Canyon 7 1/2′ quadrangle, letin of the Society of Economic Geologists, v. 81, p. 1520–1527, doi:​10​.2113/gsecongeo​ ​.81​ Yavapai County, Arizona: Arizona Geological Survey Digital Geologic Map DGM-94, version .6​.1520​. 1.0, scale 1:24,000. Watt, G.R., Wright, P., Galloway, S., and McLean, C., 1997, Cathodoluminescence and trace ele­ Spencer, J.E., Isachsen, C.E., Ferguson, C.A., Richard, S.M., Skotnicki, S.J., Wooden, J., and ment zoning in quartz phenocrysts and xenocrysts: Geochimica et Cosmochimica Acta, Riggs, N.R., 2003, U-Pb isotope geochronologic data from 23 igneous rock units in central v. 61, p. 4337–4348, doi:​10​.1016​/S0016​-7037​(97)00248​-2​. and southeastern Arizona: Arizona Geological Survey Open-File Report 03-08, 40 p. Whitmeyer, S.J., and Karlstrom, K.E., 2007, Tectonic model for the Proterozoic growth of North Stewart, J.H., Gehrels, G.E., Barth, A.P., Link, P.K., Christie-Blick, N., and Wrucke, C.T., 2001, America: Geosphere, v. 3, p. 220–259, doi:10​ .1130​ /GES00055​ ​.1​. ­Detrital zircon of Mesoproterozoic to Cambrian in the western United Wilson, E.D., 1922, Proterozoic Mazatzal Quartzite of central Arizona: Pan-American Geology, States and northwestern Mexico: Geological Society of America Bulletin, v. 113, no. 10, v. 38, p. 299–312. p. 1343–1356, doi:​10​.1130​/0016​-7606​(2001)113​<1343:​DZPOMT>2​.0​.CO;2​. Wilson, E.D., 1939, Pre-Cambrian Mazatzal revolution in central Arizona: Geological Society of Suttner, L.J., Basu, A., and Mack, G.H., 1981, Climate and the origin of quartz arenites: Journal of America Bulletin, v. 50, p. 1113–1164, doi:​10​.1130​/GSAB​-50​-1113​. Sedimentary Petrology, v. 51, p. 1235–1246. Wooden, J.L., and DeWitt, E., 1991, Pb isotopic evidence for the boundary between the Early Trevena, A.S., 1979, Studies in sandstone petrology: Origin of the Precambrian Mazatzal Quartzite Proterozoic­ Mojave and central Arizona crustal provinces in western Arizona, in Karlstrom, and provenance of detrital feldspar [Ph.D. thesis]: Salt Lake City, University of Utah, 390 p. van Hattum, M.W.A., Hall, R., Pickard, A.L., and Nichols, G.J., 2006, SE Asian sediments not K.E., ed., Proterozoic Geology and Ore Deposits of Arizona: Arizona Geological Society from Asia: Provenance and geochronology of North Borneo sandstones: Geology, v. 34, ­Digest 19, p. 27–50. p. 589–592, doi:​10​.1130​/G21939​.1​. Wrucke, C.T., 1989, The middle Proterozoic Apache Group, Troy Quartzite, and associated dia­ Van Schmus, W.R., and Bickford, M.E., eds., 1993, Transcontinental Proterozoic provinces, in base of Arizona, in Jenney, J.P., and Reynolds, S.J., eds., Geologic Evolution of Arizona: Reed, J.C., Jr., et al., eds., Precambrian: Conterminous U.S.: Geological Society of America, Arizona Geological Society Digest 17, p. 239–258. Decade of North America Geology (DNAG), v. C-2. p. 171–334. Wrucke, C.T., and Conway, C.M., 1987, Geologic map of the Mazatzal Wilderness and Contigu- Van Schmus, W.R., Bickford, M.E., and Condie, K.C., 1993, Transcontinental Proterozoic prov- ous Roadless Area, Gila, Maricopa, and Yavapai Counties, Arizona: U.S. Geological Survey inces: Early Proterozoic crustal evolution, in Reed, J.C., Jr., et al., eds., Precambrian: Con- Open-File Report 87-664, 22 p., scale 1:48,000.

GEOSPHERE | Volume 12 | Number 6 Spencer et al. | Orogenesis and quartzite deposition, Yavapai tectonic province 21