A diamictite dichotomy: Glacial conveyor belts and ... - GeoScienceWorld

4 downloads 147 Views 1MB Size Report
3 Nov 2016 - Death Valley, California (USA), rifting of Rodinia occurred concomitantly with a major gla- cial event that
A diamictite dichotomy: Glacial conveyor belts and olistostromes in the Neoproterozoic of Death Valley, California, USA D.P. Le Heron*, S. Tofaif, T. Vandyk, and D.O. Ali Department of Earth Sciences, Royal Holloway and Bedford New College, University of London, Egham, Surrey TW20 0EX, UK ABSTRACT Multiple intercalations of glacially derived and slope-derived diamictites testify to the drawbacks of correlating Neoproterozoic diamictites more widely, but shed new light on the close interrelationship of these processes in the Cryogenian world. In the Neoproterozoic of Death Valley, California (USA), rifting of Rodinia occurred concomitantly with a major glacial event that deposited the Kingston Peak Formation. A new sedimentologic investigation of this formation in the Silurian Hills demonstrates, for the first time, that some diamictites are ultimately of glacial origin. Abundant dropstone textures occur in interstratified heterolithic deposits, with clasts of identical composition (gneiss, schist, granite, metabasite, quartzite) to those of boulder-bearing diamictites suggesting a common source (the glacial conveyor belt). In stark contrast, megaclast-bearing diamictites, yielding clasts of carbonate and siliciclastic preglacial strata as much as 100 m across, are interpreted as olistostromes. The occurrence of syn-sedimentary faults within the succession allows glacial versus slope-derived material to be distinguished for the first time. INTRODUCTION The distinction between diamictites ultimately of glacial origin and those of mass flows derived from slope collapse is of fundamental paleoclimatic importance and has remained a prominent part of the research agenda in Neoproterozoic studies for over 50 yr (e.g., Schermerhorn, 1974; Eyles and Januszczak, 2004; Domack and Hoffman, 2011; Nascimento et al., 2016). Misinterpreting slope deposits as glacial, or vice versa, means that the number of glacial cycles in Earth’s history could be either under- or overemphasized (Arnaud and Etienne, 2011). Such an understanding is of profound importance in the Cryogenian, with putative panglacial (sensu Hoffman, 2009) episodes in the Neoproterozoic ~720–635 m.y. ago (Spence et al., 2016). In Death Valley (California, USA), the Kingston Peak Formation (KPF) was deposited in a regional sediment trap that received sediment from multiple directions (Mahon et al., 2014). A glacial influence has long been suspected (Hazzard, 1939). Mid-oceanic ridge basalt (MORB)–type pillow lavas on the western Death Valley flank (Labotka et al., 1980; Miller, 1985), en echelon growth faults in the Kingston Range (Walker et al., 1986), major lateral thickness variations (Prave, 1999), and olistostrome complexes (Le Heron et al., 2014) all underscore the view that the sediments record glaciation superimposed on a rift event, or vice versa (Mrofka and Kennedy, 2011; Petterson et al., 2011). The Silurian Hills (SH) are one of the Death Valley outcrop belts that have received *E-mail: daniel​.le​-heron@​rhul​.ac​.uk

very little study in the past 40 yr. Their exposures yield critical insight into the competing influence of rift and glacial process in the Cryogenian. In the SH, Basse (1978) entertained two working hypotheses for the diamictite-bearing strata. The first was a mass-flow origin, supported by co-occurrence with repetitive turbidite deposits in the succession. The second was the contention that the deposits were glaciomarine, but the absence of strong indicators for glacial deposition made this “inconclusive” (Basse, 1978, p. 51). In this paper, we resolve this controversy, with widespread implications for how rift-related and glacially related diamictites can be distinguished in the Cryogenian. The sections occur near the westward limit of the Basin and Range province and are cut through by a series of fault systems in multiple orientations that include Miocene extensional overprint of older fault zones (Ferrill et al., 2012). In spite of the rocks having locally reached amphibolite grade (Kupfer, 1960), delicate sedimentary textures are well preserved throughout the succession. METHODS A new geological facies map of the KPF in the SH (Fig. 1) was completed during two field seasons over 2015 and 2016. Most lithological boundaries on Kupfer’s (1960) map are extremely accurate, and thus our mapping focused on documenting the context of megaclasts (sensu Terry and Goff, 2014), their relationship to faults, and diamictite units. Broadly, Kupfer’s (1960) mapping units p8–p11 correspond to preglacial strata of the Beck Spring– Horse Thief Spring undifferentiated unit (see

Mahon et al. [2014] for the latter unit) on our map. By comparison, what we recognize as the KPF encompasses units p12–p19 of Kupfer (1960) and corresponds to division II of Basse (1978), while the Noonday Dolomite (representing the well-established postglacial cap carbonate unit elsewhere in Death Valley; Creveling et al., 2016) corresponds to unit p20 of Kupfer (1960). Detailed measured sections (drawn to a resolution of 10 cm) underpin the simplified sedimentary log herein, which demonstrates an almost 1.4-km-thick succession (Fig. 2) (see the GSA Data Repository1). Not all diamictites are mappable; hence, the greater number of these are shown on the simplified sedimentary log than on the map. DATA DESCRIPTION Lonestone-Bearing Heterolithics Facies Association The lonestone-bearing heterolithics facies association comprises normally graded sandstones and mudrocks, sandwiched between diamictite intervals (Figs. 1 and 2). Lonestones in these strata have not historically been interpreted as dropstones: “the penetrative deformation criterion is not satisfied” (Basse, 1978, p. 51). However, careful reinvestigation by the authors found several examples of lonestones that do deflect and penetrate delicate underlying laminations and cross-laminae, and these are overlain by undeformed laminae. These are spread throughout the succession, first appearing 6 m above the base of the KPF (Fig. 3A), last appearing in the top 30 m of the formation (Fig. 3B), and appearing at multiple intervals in between (Fig. 2). They include gneiss, schist, granite, and quartzite pebbles and cobbles. Boulder-Bearing Diamictite Facies Association The boulder-bearing diamictite facies association comprises structureless clast-poor to clastrich muddy diamictites. Clasts are subrounded to rounded and gneissose, schistose (Fig. 3C), 1  GSA Data Repository item 2017008, complete set of sedimentary logs used to construct Figure 2, with some lateral control, is available online at http://​ www​.geosociety​.org​/pubs​/ft2016​.htm or on request from editing@​geosociety​.org.

GEOLOGY, January 2017; v. 45; no. 1; p. 31–34  |  Data Repository item 2017008  |  doi:​10​.1130​/G38460​.1  |  Published online 3 November 2016 © 2016 The Authors. Access: This paper is published under the terms of the CC-BY license. GEOLOGY  45  |Open   Number 1  | www.gsapubs.org |  VolumeGold

31

Noonday Dolomite

116° 05′ N a ad ev rnia ifo al

C

3B 1300

Study area (main map)

76

N App

rox. fie Fig. ld of vie 3H, H w, ′

*

1000

88

0

200

76

35° 32′

Thickness (m)

1750

Figure 1. Geological facies map of Silurian Hills, Death Valley area, California (USA). Map was compiled over course of two field seasons from 2015 to 2016, with aim of differentiating boulder-bearing and megaclast-bearing diamictites. Map does not show member subdivisions, but rather details repeated intercalation of these two types of diamictites, thought to have been deposited through different processes.

82

Facies associations

85

Quaternary cover

Kingston Peak

Stirling Quartzite equivalent Johnnie Formation equivalent Noonday Dolomite

Carbonate

m 100 200 300 400 500 Faults

Measured sections used Sandstone unit to construct Fig 2 Megaclast diamictite Lonestone-bearing heterolithics

Boulder-bearing diamictite Preglacial

Igneous rocks Metabasite Aplite

3H taken from here looking north * FigTopographic contours are shown in feet.

granitic, or quartzitic in composition. In strati- within a black-weathering, locally folded (Fig. graphic section, uninterrupted 100-m-thick 3F) diamictite matrix. Rarely, above meta­basite intervals of diamictite are recognized (Fig. 2). “sills” (Kupfer, 1960; Basse, 1978), clasts of In some cases, diamictites show no trends in metabasite of identical character occur within matrix character over such intervals, whereas the megaclast diamictite facies association (Fig. in others, upward-coarsening and rare upward- 3G). Megaclasts are angular and measure up to fining patterns can be observed. At the meter 100 m in length (Fig. 1). Gneiss, schist, granite, scale, intercalation of diamictite beds with and quartzite pebbles and cobbles are present, hetero­lithic, graded and rippled sandstones but only within intraclasts of the boulder-bearoccurs (Figs. 2 and 3D). Most observed beds ing diamictite facies association. At Rattlesnake are massive, though rare occurrences of poorly stratified diamictites with wispy lamination (discontinuous, undulose bed partings) are noted. Megaclast Diamictite Facies Association Appraisal of both the map (Fig. 1) and the log (Fig. 2) shows that the megaclast diamictite facies association occurs at multiple stratigraphic levels. These intervals are typically >100 m each in thickness; some are faultbounded (Fig. 1; see the Data Repository). They are characterized by megaclasts of orange dolo­ stone (Fig. 3E), sandstone, and limestone, set

Figure 2. Simplified sedimentary log throughout Kingston Peak Formation (California, USA). Color scheme and facies associations correspond to those shown on geologic map in Figure 1, but note that there is greater detail shown on log (i.e., some intervals on log are below mapping resolution or contain laterally discontinuous diamictites). Note abundance of ice-rafted debris (shown by dropstone symbol to right of log). Numbers and letters to right-hand side correspond to photos in Figure 3.

Lonestonebearing heterolithics Megaclast Diamictite Boulder Diamictite Lithologies

3G

Conglomerate

3D

Diamictite (dominated by gneiss, schist, granite, quartzite clasts) Diamictite (abundant carbonate 500 megaclasts)

3F 3E

Sandstone Calcareous sandstone Mudrock Deformed mudrock Dolostone Limestone Sandy limestone Metabasite Sedimentary structures Parallel lamination

Kingston Peak Formation

80

1500

Intrabed folds Dropstones

3C 3A 0

Preglacial

32 www.gsapubs.org  |  Volume 45  |  Number 1  |  GEOLOGY

Ridge, bed thicknesses change across a series of normal faults; these faults are sealed by a thick package of the megaclast diamictite (Figs. 3H and 3H′; see also the Data Repository). This includes the largest megaclast in the SH (Figs. 1, 3H, and 3H′); previously interpreted as a Cambrian klippe (Basse, 1978), it is compositionally identical to orange dolostone and limestone clasts elsewhere in the succession. DATA INTERPRETATION AND MODEL The observation that gneiss, schist, granite, and quartzite clasts penetrate, puncture, and deform underlying laminae suggests that these clasts are dropstones. Moreover, the intercalation of these with delicately laminated and rippled deposits indicates that the clasts are hydro­ dynamically incongruent (i.e., were deposited by a different process than the encasing laminae; Le Heron, 2015). Multiple ice-rafted debris (IRD) intervals are thus recognized (Fig. 4). These textures, ­coupled with the graded beds described in detail by Basse (1978) and interpreted as turbidites, testify to a subaqueous origin (Fig. 4). Allied to this, the boulder-bearing diamictite facies association is not considered to record primary ice-contact sedimentation (i.e., tillites), but rather debris-flow deposits. This interpretation carries an important caveat: the compositional similarity of the clasts in these debrites to the dropstones, in tandem with their intercalation with IRD, strongly supports the suggestion that they are glaciogenic debris-flow deposits (GDFs) ultimately sourced from an ice sheet or glacier. While no striations on clasts are reported, comparison to a near-identical facies association some 30 km to the north at Sperry Wash (Busfield and Le Heron, 2016) shows that where the metamorphic grade is lower, delicate striations occur. We offer a very different interpretation for the megaclast diamictite facies association. The highly angular nature of the megaclasts, predominantly composed of preglacial sedimentary material, suggests a local source and minimal transport. The assemblage bears close similarity to the olistostrome complex widely reported from the Kingston Range (Mac­donald et al., 2013; Le Heron et al., 2014), and a similar mechanism is proposed here. The field evidence points to fault arrays that were active during deposition because thickness changes occur across the faults, andsome fault arrays are sharply overlain by megaclast diamictites, some of which are themselves fault bounded (Figs. 1, 3H, and 3H′; see also the Data Repository). Therefore, unroofing and toppling of material from footwall blocks in the hinterland may well account for the size, shape, and composition of olistoliths (Fig. 4). This interpretation is strengthened by the concentration of meta­ basite clasts immediately above metabasite bodies, implying reworking of extruded igneous materials (rather than sill bodies as previously

B

A

3 cm

C

15 cm

2 cm

E

D

G

F

H

H′

Looking NE

Rattlesnake Ridge

e ictit

MB

m

Kingston Range

e strom

Dia

MB

o

Olist

me ostro Olist Diamictite

Olistolith

Cambrian

Regional dip: 70-80° N Sheet heterolithics (foreground)

Figure 3. Images of Kingston Peak Formation in Silurian Hills (California, USA). A,B: Lonestones, showing disruption and puncturing of underlying laminae beneath clasts. C: Cobble-sized clasts in boulder-bearing diamictite facies association. Visible clasts include gneiss and schist in this photograph. D: Boulder-bearing diamictite (hammer [33 cm long], for scale, circled) to left of photo, overlain by lonestone-bearing heterolithics to right of photo. E: Bright orange dolostone clast (hammer for scale, circled) in megaclast diamictite facies association. F,G: Folded matrix to megaclast diamictite facies association (F) with metabasite clast (G) directly above the upper metabasite (hammer for scale). H,H′: Photo and line interpretation of megaclast diamictite facies association (i.e., olistostrome deposits) and relationships to en echelon faults, a metabasite (MB), and intercalated boulder-bearing diamictite (glaciogenic debris flow deposits). View is looking north (see Fig. 1 for location).

Glaciogenic debris flows Turbidites

IRD (distal)

Olistostrome

Crystalline basement (gneiss, schist, granitoid) Preglacial stratigraphy IRD (proximal)

Fault scarp (exposing Beck Spring Dolomite)

Syn-depositional faulting

Figure 4. Simple model showing envisaged mechanisms to produce two dramatically different diamictite deposits in Death Valley (California, USA) area, focused specifically on Silurian Hills. First type (boulder-bearing diamictite) comprises glaciogenic debris flow deposits shed from ice margin and dominated by crystalline clast types. Second type (megaclast diamictite facies association) was produced by syn-sedimentary extensional faulting, yielding locally derived cover strata of preglacial stratigraphy, and locally metabasite. Intercalation of all these facies with turbidites and ice-rafted debris (IRD)–bearing heterolithics testifies to subaqueous sedimentation within basin.

GEOLOGY  |  Volume 45  |  Number 1  | www.gsapubs.org

33

described; Kupfer, 1960) during olistostrome emplacement. These interpretations suggest that both glacial and slope processes vied for prominence in the sedimentary record of the KPF in the SH. Evidence for ice rafting is unequivocal, and the compositional similarity between clasts as IRD and clasts in the boulder-bearing diamictite facies association suggests that the glacial conveyor belt was dominated by crystalline basement clasts. IMPLICATIONS AND CONCLUSIONS Based on the data herein, it is proposed that at least four principal pulses of boulder-bearing diamictite—which we interpret as GDFs—represent multiple phases of sediment release from the paleo–ice margin. Dovetailed with observations in the southern Kingston Range some 30 km to the east (Le Heron et al., 2014), this suggests that the glacial record is much more complicated, with potentially many more glacial phases recognized, than the popular two-phase model for Death Valley which seeks to incorporate putative Sturtian and Marinoan panglacials (Macdonald et al., 2013; Smith et al., 2016). In addition to establishing the case for Neoproterozoic glaciation in the SH for the first time, the intercalation between olistostromes and GDFs illustrates the competition between these processes in building the Cryogenian record of Death Valley. They also underscore local sensitivity to tectonic processes: the stratigraphic contrast with one olistostrome complex sandwiched between two diamictites in the Kingston Range depocenter (Macdonald et al., 2013) is striking. Given that the global age constraints for Cryogenian glaciation are permissive of a continuous Neoproterozoic record (Spence et al., 2016) and that the KPF is still undated (see Smith et al., 2016), we speculate that the diamictites are diach­ronous from outcrop belt to outcrop belt within the Death Valley area (cf. Prave, 1999). A pulsed record of intercalated GDFs and olistostrome deposits—the diamictite dichotomy—records a highly complex interplay of glacially related and rift-related sedimentation. Being from one of Death Valley’s thickest occurrences of the KPF (Prave, 1999), which hence potentially records one of the greatest number of sedimentary events in the basin, the data also revive the possibility that the sedimentary record actually records many more than the two glacial episodes that are popularly suggested (e.g., Kennedy et al., 1998; Spence et al., 2016). ACKNOWLEDGMENTS Le Heron thanks Tony Prave for suggesting the SH as a study area and for a continually collegial attitude to the Death Valley geology, and thanks the Geological Society of London for multiple small grants that made this work possible. We are also very grateful to Débora B. Nascimento and Gene Domack for reviewing the manuscript, and to Judith Totman Parrish for her editorial work.

REFERENCES CITED Arnaud, E., and Etienne, J.L., 2011, Recognition of glacial influence in Neoproterozoic sedimentary successions, in Arnaud, E., et al., eds., The Geological Record of Neoproterozoic Glaciations: Geological Society of London Memoir 36, p. 39– 50, doi:​10​.1144​/M36​.3​. Basse, R.A., 1978, Stratigraphy, sedimentology, and depositional setting of the late Precambrian Pahrump Group, Silurian Hills, California [M.S. thesis]: Stanford, California, Stanford University, 86 p. Busfield, M.E., and Le Heron, D.P., 2016, A Neoproterozoic ice advance sequence, Sperry Wash, California: Sedimentology, v. 63, p. 307–330, doi:​ 10​.1111​/sed​.12210​. Creveling, J.R., Bergmann, K.D., and Grotzinger, J.P., 2016, Cap carbonate platform facies model, Noonday Formation, SE California: Geological Society of America Bulletin, v. 128, p. 1249– 1269, doi:​10​.1130​/B31442​.1​. Domack, E.W., and Hoffman, P.F., 2011, An ice grounding-line wedge from the Ghaub glaciation (635 Ma) on the distal foreslope of the Otavi carbonate platform, Namibia, and its bearing on the snowball Earth hypothesis: Geological Society of America Bulletin, v. 123, p. 1448–1477, doi:​10​.1130​/B30217​.1​. Eyles, N., and Januszczak, N., 2004, ‘Zipper-rift’: A tectonic model for Neoproterozoic glaciations during the breakup of Rodinia after 750 Ma: Earth-Science Reviews, v. 65, p. 1–73, doi:​10​ .1016​/S0012​-8252​(03)00080​-1​. Ferrill, D.A., Morris, A.P., Stamatakos, J.A., Waiting, D.J., Donelick, R.A., and Blythe, A.E., 2012, Constraints on exhumation and extensional faulting in southwestern Nevada and eastern California, U.S.A., from zircon and apatite thermochronology: Lithosphere, v. 4, p. 63–76, doi:​10​.1130​ /L171​.1​. Hazzard, J.C., 1939, Possibility of a pre-Cambric glaciation in southeastern California: Pan-American Geologist, v. 71, p. 47–48. Hoffman, P.F., 2009, Pan-glacial—A third state in the climate system: Geology Today, v. 25, p. 100– 107, doi:​10​.1111​/j​.1365​-2451​.2009​.00716​.x​. Kennedy, M.J., Runnegar, B., Prave, A.R., Hoffmann, K.-H., and Arthur, M.A., 1998, Two or four Neoproterozoic glaciations?: Geology, v. 26, p. 1059– 1063, doi:​10​.1130​/0091​-7613​(1998)026​2​.3​.CO;2​. Kupfer, D.H., 1960, Thrust faulting and chaos structure, Silurian Hills, San Bernardino County, California: Geological Society of America Bulletin, v. 71, p. 181–214, doi:​10​.1130​/0016​-7606​ (1960)71​[181:​TFACSS]2​.0​.CO;2​. Labotka, T.C., Albee, A.L., Lanphere, M.A., and McDowell, S.D., 1980, Stratigraphy, structure, and metamorphism in the central Panamint Mountains (Telescope Peak quadrangle), Death Valley area, California: Geological Society of America Bulletin, v. 91, p. 843–933, doi:​10​.1130​/GSAB​ -P2​-91​-843​. Le Heron, D.P., 2015, The significance of ice-rafted debris in Sturtian glacial successions: Sedimentary Geology, v. 322, p. 19–33, doi:​10​.1016​/j​ .sedgeo​.2015​.04​.001​. Le Heron, D.P., Busfield, M.E., and Prave, A.R., 2014, Neoproterozoic ice sheets and olistoliths: Multi­ ple glacial cycles in the Kingston Peak Formation, California: Journal of the Geological Society, v.  171, p.  525–538, doi:​10​.1144​/jgs2013​-130​. Macdonald, F.A., Prave, A.R., Petterson, R., Smith, E.F., Pruss, S.B., Oates, K., Trotzuk, D., and ­Fallick, A.E., 2013, The Laurentian record of Neoproterozoic glaciation, tectonism, and

eukaryotic evolution in Death Valley, California: Geological Society of America Bulletin, v. 125, p.  1203–1223, doi:​10​.1130​/B30789​.1​. Mahon, R.C., Dehler, C.M., Link, P.K., Karlstrom, K.E., and Gehrels, G.E., 2014, Detrital zircon provenance and paleogeography of the Pahrump Group and overlying strata, Death Valley, California: Precambrian Research, v. 251, p. 102–117, doi:​10​.1016​/j​.precamres​.2014​.06​.005​. Miller, J.M.G., 1985, Glacial and syntectonic sedimentation: The upper Proterozoic Kingston Peak Formation, southern Panamint Range, eastern California: Geological Society of America Bulletin, v. 96, p. 1537–1553, doi:​10​.1130​/0016​-7606​ (1985)96​2​.0​.CO;2​. Mrofka, D., and Kennedy, M., 2011, The Kingston Peak Formation in the eastern Death Valley region, in Arnaud, E., et al., eds., The Geological Record of Neoproterozoic Glaciations: Geological Society of London Memoir 36, p. 449–458, doi:​10​.1144​/M36​.40​. Nascimento, D.B., Ribeiro, A., Trouw, R.A.J., Schmitt, R.S., and Passchier, C.W., 2016, Stratigraphy of the Neoproterozoic Damara Sequence in northwest Namibia: Slope to basin sub-marine masstransport deposits and olistolith fields: Precambrian Research, v. 278, p. 108–125, doi:​10​.1016​ /j​.precamres​.2016​.03​.005​. Petterson, R., Prave, A.R., and Wernicke, B.P., 2011, Glaciogenic and related strata of the Neo­protero­ zoic Kingston Peak Formation in the Panamint Range, Death Valley region, California, in ­Arnaud, E., et al., eds., The Geological Record of Neoproterozoic Glaciations: Geological Society of London Memoir 36, p. 459–465, doi:​10​ .1144​/M36​.41​. Prave, A.R., 1999, Two diamictites, two cap carbonates, two d13C excursions, two rifts: The Neo­ protero­zoic Kingston Peak Formation, Death Valley, California: Geology, v. 27, p. 339–342, doi:​ 10​.1130​/0091​-7613​(1999)027​2​ .3​.CO;2​. Schermerhorn, L.J.G., 1974, Late Precambrian mixites: Glacial and/or nonglacial?: American Journal of Science, v. 274, p. 673–824, doi:​10​.2475​ /ajs​.274​.7​.673​. Smith, E.F., Macdonald, F.A., Crowley, J.L., ­Hodgin, E.B., and Schrag, D.P., 2016, Tectonostratigraphic evolution of the c. 780–730 Ma Beck Spring Dolomite: Basin formation in the core of Rodinia, in Li, Z. X., et al., eds., Super­conti­ nent Cycles Through Earth History: Geological Society of London Special Publication 424, p.  213–239, doi:​10​.1144​/SP424​.6​. Spence, G.H., Le Heron, D.P., and Fairchild, I.J., 2016, Sedimentological perspectives on climatic, atmospheric and environmental change in the Neo­ proterozoic Era: Sedimentology, v. 63, p. 253– 306, doi:​10​.1111​/sed​.12261​. Terry, J.P., and Goff, J., 2014, Megaclasts: Proposed revised nomenclature at the coarse end of the Udden-Wentworth grain-size scale for sedimentary particles: Journal of Sedimentary Research, v.  84, p.  192–197, doi:​10​.2110​/jsr​.2014​.19​. Walker, J.D., Klepacki, D.W., and Burchfiel, B.C., 1986, Late Precambrian tectonism in the ­Kingston Range, southern California: Geology, v. 14, p.  15–18, doi:​10​.1130​/0091​-7613​(1986)14​2​.0​.CO;2​. Manuscript received 9 August 2016 Revised manuscript received 9 October 2016 Manuscript accepted 10 October 2016 Printed in USA

34 www.gsapubs.org  |  Volume 45  |  Number 1  |  GEOLOGY