GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core...

27
Research Paper 237 Pape et al. | Structural reconstruction, Spruce Mountain, Nevada GEOSPHERE | Volume 12 | Number 1 Structural reconstruction and age of an extensionally faulted porphyry molybdenum system at Spruce Mountain, Elko County, Nevada James R. Pape 1, *, Eric Seedorff 1 , Tyler C. Baril 2 , and Tommy B. Thompson 2 1 Lowell Institute for Mineral Resources, Department of Geosciences, University of Arizona, 1040 East Fourth Street, Tucson, Arizona 85721-0077, USA 2 Ralph J. Roberts Center for Research in Economic Geology, Mail Stop 169, University of Nevada, Reno, Reno, Nevada 89557-0232, USA ABSTRACT This study integrates a cross-sectional restoration of normal faults and por- phyry-related hydrothermal alteration zones with six new U-Pb zircon dates to constrain the ages of mineralization and large-magnitude Cenozoic extension at Spruce Mountain, northeastern Nevada (USA). Paleozoic sedimentary rocks and sparse rhyolitic intrusive rocks host a porphyry molybdenum deposit dated as ca. 38 Ma that contains associated skarn, carbonate replacements, fissure veins, and jasperoid. At least six crosscutting sets of normal faults with variable dip di- rections (east, west, and north) and angles (<15° to >60°) are identified at Spruce Mountain, reflecting overprinting phases of extension. Based on the restored cross section, normal faulting at Spruce Mountain resulted in ~6.9 km (120%) of total extension and ~35° of net eastward tilting. The first four fault sets, late Eocene (ca. 38 Ma) or older, collectively accommodate most of the total exten- sion (~5.4 km). Later faults probably were active in the Miocene and Quaternary. All restored faults had initial dips of 45° or greater, and the restored preexten- sional structure of Spruce Mountain consists of west-vergent folds and gentle westward dips of Paleozoic rocks. Spruce Mountain is classified as a rhyolitic porphyry Mo or Climax-type deposit, with extensive skarn. Eocene rhyolite porphyry dikes associated with porphyry Mo mineralization locally intrude the early generations of normal faults at Spruce Mountain. However, many normal faults also postdate molybdenum mineralization and have partially dismem- bered the system. Cross-sectional restoration of the postmineralization faulting suggests that the original deposit had an ~6 km long by ~3 km deep elliptical pattern with carbonate replacement deposits surrounding skarn. INTRODUCTION Cenozoic extension has played an important role in the formation, preser- vation, and exposure of ore deposits in the Basin and Range province (western United States and Mexico; Dreier, 1984; Spencer and Welty, 1986; Seedorff, 1991a). However, postmineralization extension can also complicate the ex- ploration and development of mineralized systems by dismembering and tilt- ing ore deposits, thus distorting their original structural configurations (e.g., Proffett, 1977; Wilkins and Heidrick, 1995; Rahl et al., 2002; Begbie et al., 2007; Stavast et al., 2008). When applied to structurally dismembered mineralized systems, cross-sectional structural reconstructions can be used to predict where fault-offset fragments of the system are located and to delineate poten- tial exploration targets (e.g., Lowell, 1968; Seedorff, 1991b; Dilles and Proffett, 1995). In addition, structural restorations can lead to a better understanding of the preextensional geology of a region and provide insights into mechanisms of crustal extension, particularly when they are integrated with absolute age constraints (e.g., Proffett, 1977; Gans and Miller, 1983; Dilles and Gans, 1995; Surpless, 2012). This study focuses on the Spruce Mountain district of northeastern Nevada, a topographically rugged area with >1 km of local relief and complex structure produced by faults with opposing senses of displacement (Hope, 1972). Spruce Mountain is ~30 km east of the Ruby Mountains–East Humboldt Range meta- morphic core complex and a few kilometers from the southwestern tip of the adjacent Pequop Mountains (Fig. 1). The Spruce Mountain mining district cov- ers the summit and northern flank of Spruce Mountain and extends northward onto Spruce Mountain Ridge. The district contains porphyry molybdenum and associated skarn, carbonate replacements, fissure veins, and jasperoid and has been explored for Carlin-type gold mineralization (LaPointe et al., 1991; Wolverson, 2010; E.M. Struhsacker, 2014, personal commun.). The Spruce Mountain district is ~40 km south-southwest of the Pequop mining district, where Carlin-type gold deposits were discovered recently (Bedell et al., 2010; Felder et al., 2011; Smith et al., 2013). This region has undergone significant amounts of both Mesozoic crustal shortening and subsequent Cenozoic extension (e.g., Camilleri and Chamber- lain, 1997), but there is ongoing debate over the total amount of extension in eastern Nevada and the relative importance of late Eocene versus middle Mio- cene extension (e.g., Seedorff, 1991a; Miller et al., 1999; Sullivan and Snoke, 2007; Henry, 2008; Colgan et al., 2010). We present a detailed, map-based structural analysis and cross-sectional reconstruction of Spruce Mountain coupled with new U-Pb zircon dates of GEOSPHERE GEOSPHERE; v. 12, no. 1 doi:10.1130/GES01249.1 8 figures; 2 tables CORRESPONDENCE: [email protected] CITATION: Pape, J.R., Seedorff, E., Baril, T.C., and Thompson, T.B., 2016, Structural reconstruction and age of an extensionally faulted porphyry molyb- denum system at Spruce Mountain, Elko County, Nevada: Geosphere, v. 12, no. 1, p. 237–263, doi:10 .1130/GES01249.1. Received 10 August 2015 Revision received 24 September 2015 Accepted 17 November 2015 Published online 23 December 2015 For permission to copy, contact Copyright Permissions, GSA, or [email protected]. © 2015 Geological Society of America *Present address: ExxonMobil Exploration Company, 22777 Springwoods Village Parkway, Spring, Texas 77389, USA

Transcript of GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core...

Page 1: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

237Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Structural reconstruction and age of an extensionally faulted porphyry molybdenum system at Spruce Mountain, Elko County, NevadaJames R. Pape1,*, Eric Seedorff1, Tyler C. Baril2, and Tommy B. Thompson2

1Lowell Institute for Mineral Resources, Department of Geosciences, University of Arizona, 1040 East Fourth Street, Tucson, Arizona 85721-0077, USA2Ralph J. Roberts Center for Research in Economic Geology, Mail Stop 169, University of Nevada, Reno, Reno, Nevada 89557-0232, USA

ABSTRACT

This study integrates a cross-sectional restoration of normal faults and por-phyry-related hydrothermal alteration zones with six new U-Pb zircon dates to constrain the ages of mineralization and large-magnitude Cenozoic extension at Spruce Mountain, northeastern Nevada (USA). Paleozoic sedimentary rocks and sparse rhyolitic intrusive rocks host a porphyry molybdenum deposit dated as ca. 38 Ma that contains associated skarn, carbonate replacements, fissure veins, and jasperoid. At least six crosscutting sets of normal faults with variable dip di-rections (east, west, and north) and angles (<15° to >60°) are identified at Spruce Mountain, reflecting overprinting phases of extension. Based on the restored cross section, normal faulting at Spruce Mountain resulted in ~6.9 km (120%) of total extension and ~35° of net eastward tilting. The first four fault sets, late Eocene (ca. 38 Ma) or older, collectively accommodate most of the total exten-sion (~5.4 km). Later faults probably were active in the Miocene and Quaternary. All restored faults had initial dips of 45° or greater, and the restored preexten-sional structure of Spruce Mountain consists of west-vergent folds and gentle westward dips of Paleozoic rocks. Spruce Mountain is classified as a rhyo litic porphyry Mo or Climax-type deposit, with extensive skarn. Eocene rhyolite porphyry dikes associated with porphyry Mo mineralization locally intrude the early generations of normal faults at Spruce Mountain. However, many normal faults also postdate molybdenum mineralization and have partially dismem-bered the system. Cross-sectional restoration of the postmineralization faulting suggests that the original deposit had an ~6 km long by ~3 km deep elliptical pattern with carbonate replacement deposits surrounding skarn.

INTRODUCTION

Cenozoic extension has played an important role in the formation, preser-vation, and exposure of ore deposits in the Basin and Range province (western United States and Mexico; Dreier, 1984; Spencer and Welty, 1986; Seedorff,

1991a). However, postmineralization extension can also complicate the ex-ploration and development of mineralized systems by dismembering and tilt-ing ore deposits, thus distorting their original structural configurations (e.g., Proffett, 1977; Wilkins and Heidrick, 1995; Rahl et al., 2002; Begbie et al., 2007; Stavast et al., 2008). When applied to structurally dismembered mineralized systems, cross-sectional structural reconstructions can be used to predict where fault-offset fragments of the system are located and to delineate poten-tial exploration targets (e.g., Lowell, 1968; Seedorff, 1991b; Dilles and Proffett, 1995). In addition, structural restorations can lead to a better understanding of the preextensional geology of a region and provide insights into mechanisms of crustal extension, particularly when they are integrated with absolute age constraints (e.g., Proffett, 1977; Gans and Miller, 1983; Dilles and Gans, 1995; Surpless, 2012).

This study focuses on the Spruce Mountain district of northeastern Nevada, a topographically rugged area with >1 km of local relief and complex structure produced by faults with opposing senses of displacement (Hope, 1972). Spruce Mountain is ~30 km east of the Ruby Mountains–East Humboldt Range meta-morphic core complex and a few kilometers from the southwestern tip of the adjacent Pequop Mountains (Fig. 1). The Spruce Mountain mining district cov-ers the summit and northern flank of Spruce Mountain and extends northward onto Spruce Mountain Ridge. The district contains porphyry molybdenum and associated skarn, carbonate replacements, fissure veins, and jasperoid and has been explored for Carlin-type gold mineralization (LaPointe et al., 1991; Wolverson, 2010; E.M. Struhsacker, 2014, personal commun.). The Spruce Mountain district is ~40 km south-southwest of the Pequop mining district, where Carlin-type gold deposits were discovered recently (Bedell et al., 2010; Felder et al., 2011; Smith et al., 2013).

This region has undergone significant amounts of both Mesozoic crustal shortening and subsequent Cenozoic extension (e.g., Camilleri and Chamber-lain, 1997), but there is ongoing debate over the total amount of extension in eastern Nevada and the relative importance of late Eocene versus middle Mio-cene extension (e.g., Seedorff, 1991a; Miller et al., 1999; Sullivan and Snoke, 2007; Henry, 2008; Colgan et al., 2010).

We present a detailed, map-based structural analysis and cross-sectional reconstruction of Spruce Mountain coupled with new U-Pb zircon dates of

GEOSPHERE

GEOSPHERE; v. 12, no. 1

doi:10.1130/GES01249.1

8 figures; 2 tables

CORRESPONDENCE: [email protected]

CITATION: Pape, J.R., Seedorff, E., Baril, T.C., and Thompson, T.B., 2016, Structural reconstruction and age of an extensionally faulted porphyry molyb­denum system at Spruce Mountain, Elko County, Nevada: Geosphere, v. 12, no. 1, p. 237–263, doi: 10 .1130 /GES01249.1.

Received 10 August 2015Revision received 24 September 2015Accepted 17 November 2015Published online 23 December 2015

For permission to copy, contact Copyright Permissions, GSA, or [email protected].

© 2015 Geological Society of America

*Present address: ExxonMobil Exploration Company, 22777 Springwoods Village Parkway, Spring, Texas 77389, USA

Page 2: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

238Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

igneous dikes that provide constraints on (1) the age of extension and por-phyry molybdenum mineralization, and (2) the original geometry of mineral-ized zones prior to subsequent extensional deformation. We conclude that at least 6 individual sets of crosscutting normal faults accommodated ~6.9 km (120%) of extension at Spruce Mountain, and that the first 4 fault sets, which collectively accommodate 5.4 km of extension, were active at or before ca. 38 Ma. Although the oldest normal faults currently dip at <15°, based on the structural reconstruction, all faults had initial dips of ~45° or greater. We also discuss implications of this cross-sectional reconstruction for the timing and style of extension in the northeastern Great Basin and its importance to miner-alization processes and mineral exploration at Spruce Mountain.

GEOLOGIC SETTING

Spruce Mountain is part of a belt of Precambrian to Mesozoic continental margin strata and Mesozoic to Cenozoic igneous rocks situated to the east of contractional allochthons emplaced during the mid-Paleozoic to mid-Meso-zoic (Figs. 1 and 2). It is significant that this belt of rocks is also within a zone of meta morphic core complexes that characterizes the eastern fringe of the hinterland of the Mesozoic Sevier fold and thrust belt (Fig. 1; Dickinson, 2011; Cashman et al., 2011). Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic rocks, are the iconic landforms of extended crust (Crittenden

116° 00’ W 115° 30’ 115° 00’

40° 3

0’ N

41° 0

0’

114° 30’

Area of Fig. 3

Ruby Vall

eyRuby

Mou

ntai

ns

Adob

e Ra

nge

Maverick SpringsRange

Huntingto

nVall

ey

Spru

ce

Mou

ntai

n

Elko

Carlin

Cherry Creek

Range

Medicine

RangeDelcerButtes

Wells

Wendover

Gos

hute

-Toa

no R

ange

Pequ

op M

ount

ains

Emigrant Pass

East Hum

boldt

Range

Wood Hills

Carlin Trend

40 km

I-80

US 93

US 93A

Pequop Syncline

Ant

ler B

elt

Nevada

Idaho

UtahSevier BeltRM

SR

AR

Figure 1. Regional location map. Location of the Peqoup syncline, outline of Figure 3, and Carlin trend are shown. Outcrops of pre-Cenozoic rocks are outlined in gray. Inset: Index map of Nevada, Utah, and southern Idaho showing location of the regional map (gray box) relative to Ruby Mountains–East Humboldt (RM), Albion–Raft River–Grouse Creek (AR), and Snake Range (SR) metamor-phic core complexes (black). The eastern limits of the Sevier fold and thrust belt and allochthonous western facies rocks of the Antler orogenic belt are also shown (modified from Howard, 2003).

Page 3: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

239Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

et al., 1980; Sullivan and Snoke, 2007). Spruce Mountain is located in the mid-dle of three core complexes, Ruby Mountains–East Humboldt Range, Albion–Raft River–Grouse Creek, and Snake Range (Fig. 1).

In proximity of Spruce Mountain, crustal thickening events, perhaps Late Jurassic to Late Cretaceous in age, are recorded by thrust faults and regional metamorphic rocks exposed in the Ruby Mountains–East Humboldt Range to the northwest, in the Wood Hills to the north, and in the Pequop Mountains to the northeast (Thorman, 1970; Howard, 1980; Snoke and Miller, 1988; Hudec, 1992; Camilleri and Chamberlain, 1997). Thermobarometry and thermochro-nology data from the East Humboldt Range and the Wood Hills suggest that a decompression event may have occurred there as early as the Late Creta-ceous, which may reflect the onset of crustal extension in northeastern Nevada (Hodges et al., 1992; McGrew and Snee, 1994; McGrew et al., 2000; Druschke et al., 2009; Hallett and Spear, 2014). However, work has also underscored the importance of middle Miocene extension in the region (e.g., Miller et al., 1999; Colgan and Henry, 2009).

Spruce Mountain exposes Ordovician to Permian miogeoclinal sedimen-tary rocks (Fig. 2) in numerous fault-bounded blocks (Fig. 3A). Lower Paleozoic rocks exposed at the deepest structural levels may have reached greenschist facies metamorphic grade (Camilleri and Chamberlain, 1997). Throughout most of the district, the Paleozoic rocks dip to the east at moderate to steep angles, but in some fault blocks in the northern part of the district, Paleozoic rocks dip consistently westward (Fig. 3A), similar to the nearest exposures of Paleozoic rocks 10 km west of the district (Hope, 1972).

No Mesozoic sedimentary rocks crop out at Spruce Mountain, but ~1.0–1.2 km of lower Triassic marine shale and limestone are exposed ~10 km east of Spruce Mountain within the core of the Pequop syncline (Fraser et al., 1986; Swenson, 1991). Cenozoic volcanic and sedimentary rocks unconformably overlie Paleozoic rocks at Spruce Mountain. A southeast-dipping Cenozoic dacite tuff crops out near Cole Creek on the eastern side of the district and overlies Permian and upper Pennsylvanian rocks (Fig. 3A). Sedimentary rocks that crop out in the southwestern corner of the district (Fig. 3A) dip gently to moderately to the east-northeast (Harlow, 1956; Hope, 1972) and have been assigned to the Miocene Humboldt Formation (Harlow, 1956; Coats, 1987).

Although intrusive rocks constitute only a small fraction of the surface ex-posures at Spruce Mountain, a variety of intrusive rocks have been observed in the district (Fig. 3A; Table 1). Prior to this study, there were no radiometric ages on intrusive rocks from Spruce Mountain. However, in nearby ranges, most igneous rocks range from Jurassic through late Eocene in age (e.g., Wright and Snoke, 1993; Henry, 2008; Bedell et al., 2010).

INTRUSIVE ROCKS AND U-Pb GEOCHRONOLOGY

The mapped distribution of intrusive rocks at Spruce Mountain is shown in Figure 3A. Six samples that represent many of the igneous rocks known at Spruce Mountain were selected for U-Pb zircon geochronology (Baril, 2013)

Sil.

Dev

.O

rd.

Mis

s.P

enn.

Tria

ssic

Per

mia

n

C

ambr

ian

Pro

tero

zoic

Stra

tigra

phic

Thi

ckne

ss (k

m)

Thaynes Formation

Park City Group

Loray Formation

Pequop Formation

Rib Hill FormationRiepe Spring FormationEly Limestone

Diamond Peak Formation Chainman Formation Joana LimestoneGuilmette FormationSimonson DolomiteLaketown Dolomite and Lone Mountain Dolomite

Eureka Quartzite Pogonip GroupWindfall FormationDunderberg ShaleHamburg Formation

Secret Canyon Formation

El Dorado FormationPioche Shale

Prospect Mountain Quartzite

McCoy Creek Group

11

Proterozoic-Mesozoic Stratigraphy

Cenozoic Stratigraphy

Triassic to Ordovician Aged Sedimentary rocks

Oligocene to Miocene volcanic and sedimentary rocks including Humboldt Fm. Eocene Volcanic Rocks

12

13

14

10

6

7

8

9

5

1

2

3

4

Quaternary alluvial deposits

Figure 2. Generalized stratigraphic column for the Spruce Mountain area. Based on data from Thorman (1970), Hope (1972), Stewart (1980), Coats (1987), and Swenson (1991).

Page 4: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

240Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Wes

t fau

lt

West fault

East

faul

t

Spring fault Sout

h Pe

ak fa

ult

Nor

th P

eak

faul

t

Prospect fault

North Peak fault

Bann

er H

ill fa

ult

Coyote fault

Min

e Sh

aft f

ault

654

3

21

Spruce

Sadd

le fa

ult

Map Units

Pp

Qa

Th

Tt

Tqfp

Prh

Mdpc

MDgu

DSdu

Op

bx

PpcMap Symbols Fault Sets

(Fault traces dashed and queried where relative age relationships are uncertain)36

Youngest

Oldest

0 km 2 km

Strike and dip of bedding

Contact

High-angle normal fault (ball and bar on downthrown side)Low-angle normal fault(hachures on hanging wall )

Quaternary deposits

Tertiary Humboldt Formation

Tertiary quartz-feldsparporphyry

Tertiary dacite tu�

Iron-stained quartz breccia

Permian Park City Group

Permian Pequop Formation

Permian Rib Hill Formation

Pennsylvanian-PermianRiepe Spring Limestone

Pennsylvanian Ely Limestone

Mississippian Chainman/Diamond Peak, undivided

Devonian-Mississppian Guilmette/Pi-lot-Shale/Joanna Limestone, undivided

Silurian-Devonian Dolomite,undivided

Ordovician Pogonip Group

N

High-angle transfer fault (Mesozoic?)

Jf Jurassic felsic dikes

U-Pb sample location

Trp Tertiary rhyolite porphyry P rs

e

2525

69

36

25

25

60

18

17

34

31

32

25

3346

35

41 35

45

36

29

36

7846

25

2520

25

25

45

30

25

25

20

38

2539

27

27

12

22

27

47

59

3534

48

34

2535

50

25

40

2327

65

40

25

25

53

32

Qa

Qa

Qa

Qa

Op

Pp

Pp

Pp

Pp

PpPp

Pp

Pp

Pp

TtTt

Th

Qa

MDgu

DSdu

Mdpc

MDgu

Mdpc

Mdpc

Mdpc

Pp

Prh

Prh

PrhPp

Prh

Prh

Prh

Prh

DSdu

Ppc

Qa

Qa

Qa

bx

North Peakof Spruce Mountain

South Peakof Spruce Mountain

114° 53′ 00″ W 114° 50′ 30″ 114° 48′ 00″ 114° 45′ 30″

40° 3

3′ 0

0″40

° 30′

30″

N

53

49

50

45

18

7

33

50

38

20

33

3420

49

58

28

46

33

47

33

42

25

30

3079

5655

35

57

27

76

32

45

50

35

4050

35

64

12

25

24

25

42

18

20

31

20

10

47

50

39

15

80

42

BannerHill

Cole Creek

SMIg44

SMIg45

SMIg46

SMIg47

SMIg49SMIg51

Trp

Trp

Tqfp

Jf

Jf

P rs

P rsP rs

P rs

P rs

P rs

P rs

P rs

e

e

e

e

e

e

e

e

e

e

e

e

ee

A

Figure 3 (on this and following page). Geo-logic maps of the Spruce Mountain area. (A) Map shows lithologic units, exten-sional faults color coded by relative ages, locations of samples for U-Pb dates with ages, and line of reconstructed cross sec-tion (see text). Map compiled and modi-fied from Hope (1972), Fraser et al. (1986), and Swenson (1991), incorporating infor-mation from Harlow (1956), Coats (1987), and this study. Revisions in local areas based on footnotes in proof on map of Hope (1972), mapping by Saunders and Bresnahan (2010), and our field mapping and observations.

Page 5: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

241Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Wes

t fau

lt

West fault

East

faul

tSpring

fault

Sout

h Pe

ak fa

ult

Nor

th P

eak

faul

t

Prospect fault

North Peak fault

Bann

er H

ill fa

ult

Coyote fault

Min

e Sh

aft f

ault

654

3

21

Spruce

Sadd

le fa

ult

Iron-stained quartz breccia

Quartz-molydenite veins at depth

Skarn +/- hornfels with copper

Jasperoid

Carbonate replacement deposits with Pb-Zn-Ag-Cu

Map SymbolsHigh-angle normal fault (ball and bar on downthrown side)Low-angle normal fault(hachures on hanging wall )

High-angle transfer fault(Mesozoic?)Historical mine site

12345

Ada H MineBlack Forest MineKille MineMonarch MineStandard Mine

0 km 2 km

Fault Sets

(Fault traces dashed and queried where relative age relationships are uncertain)

Youngest

Oldest

6 Banner Hill Mine7 Keystone Mine8 Paramount Mine9 Spence Mine

Alteration Zones

TqfpTertiary quartz-feldsparporphyry

Jurassic felsic dikes

Trp Tertiary rhyolite porphyry

Jf

Map Units

North Peakof Spruce Mountain

South Peakof Spruce Mountain

114° 53′ 00″ W 114° 50′ 30″ 114° 48′ 00″ 114° 45′ 30″

40° 3

3′ 0

0″40

° 30′

30″

N

1

23

54

6

7

8

9

B

Figure 3 (continued ). (B) Map showing simplified distribution of selected alter-ation-mineralization features and num-bered locations of selected historical mines. Extensional faults color coded by relative ages, intrusive rocks, and the line of the reconstructed cross section are shown for reference. Sericitic alter-ation (not shown) occurs only locally in porphyry dikes. Map compiled from Hope (1972), LaPointe et al. (1991), American CuMo Mining Corporation (2014), and our observations.

Page 6: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

242Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

and are described in Appendix 1. Igneous rocks with a fine-grained ground-mass are assigned volcanic rock names (whether extrusive or shallow intru-sive), whereas coarser grained porphyritic and equigranular rocks are given plutonic rock names. Hypabyssal rocks may have either a volcanic or plutonic name, depending on grain size. Sample locations are listed in Table 1 and plot-ted in Figure 3A. Table 1 also lists the stratigraphic hosts of each sample; this bears on the possible depth of emplacement of the intrusions, as well as the alteration-mineralization characteristics of each sample, which are relevant to evaluating the age of alteration mineralization. U-Pb dates were obtained from microanalysis of zircons by the laser ablation–inductively coupled plasma–mass spectrometry (LA-ICP-MS) method at Boise State University (Idaho). Ana lytical methods and reporting procedures are described in Appendix 2, and results are summarized in Table 1.

Late Jurassic Dacite

A sample from a small dike of Late Jurassic porphyritic dacite that crops out on the North Peak of Spruce Mountain (SMIg47) is dated as 156 ± 2 Ma (Table 1). In addition, a sample from a small exposure of Late Jurassic flow-banded dacite (SMIg46), dated as 155 ± 4 Ma (Table 1), crops out ~1 km north of the Monarch mine (Fig. 3B).

Late Eocene Rhyolite

An east-west–trending dike or band of dikes of late Eocene rhyolite por-phyry and quartz-rich breccias cuts across the northern side of Spruce Moun-tain, crossing the range at a pass near the Kille mine (Figs. 3A and 3B; Hill, 1916; Hope, 1972). Although poorly exposed, surface and dump samples indicate

the presence of several varieties of rhyolite porphyry (72%–74% SiO2) dikes, including altered porphyry dikes with ~25–30 vol% phenocrysts of quartz, K-feldspar, plagioclase, and minor biotite set in an aplitic groundmass and less altered varieties with ~10–15 vol% phenocrysts and an aplitic groundmass (E. Seedorff, personal data; Baril, 2013). A sample of coarse-grained rhyo lite porphyry with ~30 vol% phenocrysts was collected near the Black Forest mine in the north-central part of the district (SMIg49). This sample is potassically altered and cut by quartz veins with minor molybdenite (Figs. 3A, 3B), and has a U-Pb zircon age of 39.1 ± 0.6 Ma (Table 1). Dates (40Ar/39A) on K-feldspar and biotite from the same sample, SMIg49, are similar to the U-Pb age, indicating rapid cooling in the late Eocene (C.D. Henry, 2013, written commun.). A sample of rhyolite porphyry with ~15 vol% phenocrysts (SMIg51) that was collected farther east of the Black Forest mine has a U-Pb zircon age of 37.5 ± 0.4 Ma (Table 1). Samples from the vicinity of the Standard mine on the western side of the district (Figs. 3A, 3B) yield U-Pb zircon ages of 37.7 ± 0.5 Ma on a flow-banded porphyritic rhyolite dike (SMIg44) and 37.3 ± 0.4 Ma on a rhyolite por-phyry with ~30 vol% phenocrysts (Table 1).

Undated Intrusive Rocks

In addition to the intrusive rocks dated as part of this study, several other occurrences of undated igneous rocks have been reported at Spruce Moun-tain. A lamprophyre dike of unknown age was observed in the underground workings of the Black Forest mine (Fig. 3B) by Schrader (1931), and an undated stock of hornblende diorite intrudes Permian strata near the northern end of Spruce Mountain Ridge 19 km beyond the northern boundary shown in Figure 3. Small, altered dikes also crop out on Banner Hill (Fig. 3), some of which contain hydrothermal chlorite and epidote (Hill, 1916).

TABLE 1. SUMMARY OF U-Pb AGES OF IGNEOUS ROCKS IN THE SPRUCE MOUNTAIN DISTRICT

SampleUTM

easting*UTM

northing*Rock type

(phenocryst content, vol%)Local stratigraphic hosts of intrusion Alteration, mineralization

Age (Ma)

2σ(Ma)

SMIg44 679937 4492126 Porphyritic rhyolite (5) Pennsylvanian Ely Limestone

Sericite and kaolinite after feldspars

37.7 0.5

SMIg45 680682 4490741 Rhyolite porphyry (30) Pennsylvanian Ely Limestone

Sericite and kaolinite after feldspars

37.3 0.4

SMIg46 683420 4493380 Dacite (none) Pennsylvanian–Permian Riepe Spring Limestone

Sericite after plagioclase 155 4

SMIg47 684690 4491810 Porphyritic dacite (15) Mississippian Chainman Shale

Propylitic alteration (chlorite, calcite, sericite, kaolinite)

156 2

SMIg49 685957 4492716 Coarse-grained rhyolite porphyry (30)

Devonian Guilmette Formation and Pennsylvanian Ely Limestone

Quartz ± molybdenite veinlets, potassic alteration, minor chalcopyrite and pyrite

39.1 0.6

SMIg51 686910 4492715 Rhyolite porphyry (15) Pennsylvanian Ely Limestone

Potassic alteration and weak argillization; elsewhere can contain quartz veins

37.5 0.4

*Universal Transverse Mercator North American Datum (NAD27) datum, UTM zone 11N.

Page 7: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

243Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

MINERAL DEPOSITS AND THE AGE OF PORPHYRY Mo MINERALIZATION

The Spruce Mountain district contains a porphyry molybdenum deposit asso ciated with rhyolite poryphyry intrusions and may also host Carlin-type gold mineralization. Our geochronologic data and mapping indicate that the porphyry moldybdenum deposit formed in the late Eocene, ca. 38 Ma, whereas the age of Carlin-type mineralization in the district remains undetermined.

Skarn and carbonate replacement deposits were first discovered in the Spruce Mountain district in 1869, and there was at least small production from these ores for most years between 1869 and 1961 (Hill, 1916; Schrader, 1931; Smith, 1976; LaPointe et al., 1991). The district historically has been mined for lead and silver, with subordinate zinc, copper, and gold, from small under-ground mines, including the Ada H, Banner Hill, Black Forest, Keystone, Kille, Monarch, Paramount, and Standard (Fig. 3B). The principal hypogene ore min-erals are argentiferous galena, tetrahedrite, sphalerite, chalcopyrite, pyrite, arseno pyrite, and minor bornite (Schrader, 1931). Scheelite occurs at the Atlan-tic prospect, which is a few hundred meters west of the Black Forest mine (Stager and Tingley, 1988). Supergene phases include cerussite, anglesite, malachite, and chrysocolla along with lesser amounts of wulfenite, calamine, and smithsonite. Oxidation extends to depths of 70 m or more, and most of the production was from supergene-enriched carbonate ores that contain higher silver contents than sulfide ores (LaPointe et al., 1991).

A map of alteration-mineralization products is shown in Figure 3B, which complements the companion geologic map (Fig. 3A). As shown on these maps, a belt of skarn, quartz-rich breccias, and local hornfels is discontin-uously exposed at the surface along the east-west–trending band of silicic porphyry dikes, but hornfels is absent and skarn is mostly supplanted by jasperoid and carbonate replacement and fissure vein deposits in calcare-ous rocks at the eastern end of the band (Hope, 1972; LaPointe et al., 1991). At the surface, the porphyry dikes are variably altered but have rare quartz-molyb denite veinlets with associated potassic alteration and locally exhibit quartz-sericite-pyrite alteration. In contrast, the skarns consist principally of garnet, pyroxene, actinolite, and fluorite (Granger et al., 1957; E. Seedorff, per-sonal data; LaPointe et al., 1991; American CuMo Mining Corporation, 2014). Skarn is localized around the porphyry dikes. Copper and local tungsten are mostly restricted to skarn; zinc grades generally decrease toward porphyry dikes (Schrader, 1931). Lead and silver grades are highest in fissure fillings, commonly in marble or limestone (LaPointe et al., 1991). Both skarn and al-tered igneous rocks are enriched in tin, tungsten, and fluorine (E. Seedorff, personal data; Tingley, 1981a, 1981b).

From the 1950s through the early 1980s, the district was explored for por-phyry mineralization, and at the end of this period a resource of ~80 Mt of low-grade molybdenum-copper mineralized rock (average grade not reported) was delineated in the southwestern part of the district by a joint venture of Freeport Exploration and AMAX (LaPointe et al., 1991). Complete results are not publicly available, although published cross sections showing selected results suggest

that a well-mineralized area occurs near the Kille mine (American CuMo Min-ing Corporation, 2014) in the footwall of a steeply dipping normal fault (proba-bly the West fault; Fig. 3B) and in the footwall of the gently dipping North Peak fault. In recent years, this part of the district has been controlled by Mosquito Gold Mines Limited, now known as American CuMo Mining Corporation; this company reports molybdenum intercepts (copper grades not reported) from 5 of the historic drill holes that are 105–366 m in length (American CuMo Min-ing Corporation, 2014). The average grades in each of the 5 intercepts range from 0.05% to 0.17% MoS2 (~0.03%–0.1% Mo). A length-weighted average of these intercepts, which might be a rough estimate of the average grade of the resource, is 0.089% MoS2 (~0.053% Mo). Molybdenum is preferentially lo-calized in porphyry dikes, whereas copper is concentrated in adjacent skarns (American CuMo Mining Corporation, 2014) and appears to be best developed in Pennsylvanian Ely Limestone. Lead, zinc, and silver preferentially occur in carbonate replacement deposits and fissure veins in limestones and other cal-careous and dolomitic rocks, especially of Pennsylvanian–early Permian age. Descriptions of underground exposures (e.g., Schrader, 1931) indicate that porphyry dikes at least locally intruded along normal faults, and a porphyry dike is mapped as intruding along the North Peak fault (Fig. 3A). Likewise, porphyry-related skarn mineralization at least locally occurs adjacent to dikes and along faults (Schrader, 1931), and brecciated quartz veins occur within the Prospect fault (Figs. 3A and 3B).

There are several textural varieties of porphyritic rhyolite and rhyolite por-phyry at Spruce Mountain. A phenocryst-rich (~30 vol%) rhyolite porphyry that is dated as 39.1 ± 0.6 Ma contains quartz veins and molybdenite (though the veins and molybdenum in the dated intrusion could be related to a younger intrusion). An intrusion with 15 vol% phenocrysts dated as 37.5 ± 0.4 Ma locally contains quartz veins, and intrusions as young as 37.3 ± 0.4 Ma with 30 vol% phenocrysts underwent strong hypogene sericite alteration. Analogies with porphyry molybdenum deposits elsewhere in the world (e.g., Wallace et al., 1968; Carten et al., 1988) suggest that Spruce Mountain may have several, pos-sibly spatially separated, mineralizing intrusions.

Although some crosscutting relationships between faults and rhyolite in-trusions are known (Table 2), other relationships remain to be determined. The spatial and genetic relationship between various rhyolite porphyry dikes and porphyry-related mineralization also is only weakly constrained (see Table 2). Therefore, it is not possible to determine whether porphyry-related mineraliza-tion was genetically related to intrusions as old as 39.1 ± 0.6 Ma, to intrusions as young as 37.3 ± 0.4 Ma, or to multiple intrusions of different ages. Conse-quently, we regard porphyry molybdenum mineralization at Spruce Mountain as having formed in the late Eocene, ca. 38 Ma.

In the northwestern part of the Spruce Mountain district, six holes drilled by Santa Fe Mining in the mid-1980s intersected significant gold-bearing in-tervals, and Renaissance Gold has explored the western side of the district for Carlin-type gold mineralization in the Pilot Shale and for gold skarn mineral-ization (Wolverson, 2010). The U-Pb dates place no constraints on the age of Carlin-type gold occurrences that are reported in the district.

Page 8: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

244Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

PLAN-VIEW ANALYSIS OF FAULTING AT SPRUCE MOUNTAIN

As part of a larger study (Pape, 2010), a map-based, plan-view analysis of extensional faulting at Spruce Mountain was done to help constrain the rela-tive timing relationships among crosscutting normal faults and to determine the present-day geometries of low-angle faults in the district (Figs. 3 and 4). The surface geologic map of the Spruce Mountain quadrangle of Hope (1972) is the primary resource for this analysis, and additional data are incorporated from other sources, including our new field mapping and checks of key geo-logic relationships. The names of faults follow those of previous workers where possible; certain previously unnamed faults have been given names here for ease of reference.

Crosscutting Fault Relationships

Numerous crosscutting high- and low-angle normal faults have accom-modated extensional strain at Spruce Mountain. These crosscutting faults are grouped into six sets of similar relative age and orientation, and are numbered 1 through 6, from oldest to youngest (Table 2; Fig. 3). In some cases, particu-larly in the southeastern part of the study area where numerous faults inter-sect, the crosscutting relationships among faults are uncertain. In these cases, faults have been assigned to a fault set, but the fault traces in Figure 3 are dashed and queried to reflect this uncertainty.

Faults from the two relatively youngest sets, fault sets 5 and 6 (Table 2), do not intersect each other within the study area, and thus their relative ages cannot be firmly established. Faults of set 6, which crop out on the western side of the study area, include the present-day range-bounding faults and the relatively large offset West fault. Faults in this set have northwest to northeast strikes and dip at high to moderate angles to the northwest and southwest. Northeast-striking, southeast-dipping, high-angle normal faults of small offset in the southeasternmost part of the study area are assigned to set 5. Faults of both of these sets cut faults grouped into the next oldest set, set 4 (Table 2). Fault set 4 consists of a group of generally north-striking, east-dipping, mod-erate- to high-angle normal faults that crop out throughout the study area. Fault set 3 consists of nearly east-west–striking, north-dipping high-angle faults that traverse the central portion of the study area. Fault set 2 consists of generally south- to southwest-striking, west- to northwest-dipping, high- to moderate-angle normal faults that crop out throughout the study area. Certain west-dipping faults grouped within fault set 2 may, in detail, cut each other. This is especially true in the southeastern part of the study area, where numer-ous faults intersect and crosscutting relationships are difficult to determine. Despite this complexity, these faults have similar orientations and display simi lar crosscutting relationships relative to faults from other fault sets; there-fore, for simplicity, they have been grouped into fault set 2.

The oldest fault set includes all low-angle normal faults that are present within the study area. Where they crop out, these low-angle faults are always cut by high- to moderate-angle faults. However, like faults assigned to fault set

TABLE 2. FAULTS AND FAULT SETS AT SPRUCE MOUNTAIN

Fault setExamples of named faults Strike direction

Present-day dip of faults

Expected direction of concurrent tilting of fault blocks

Exposed crosscutting relationships with faults of older sets

Relationship to Cenozoic stratigraphic and intrusive units Notes

6 (youngest) West fault Northwest to northeast

Northwest and southwest, steep to moderate

Eastward Faults cut and offset faults of sets 4 and 1; no spatial overlap with faults of set 5

Some faults locally mantled by Quaternary alluvium, but faults on western range front active in Quaternary

Restricted to western part of area

5 Unnamed faults only Northeast Southeast, steep Westward Faults cut and offset faults of sets 4 and 2; no spatial overlap with faults of set 6

Faults mantled by Quaternary alluvium

Restricted to southeastern-most part of area

4 Banner Hill, East, Mine Shaft faults

North East, moderate to steep

Westward Faults cut and offset faults of sets 3, 2, and 1

Intruded by late Eocene (ca. 38 Ma) rhyolite porphyry dikes; Miocene sedimentary rocks mantle faults of set 4 and seem to have fanning upward dips from ~30° to 10°E

3 Unnamed faults only East-west North, steep Southward Faults seemingly cut and offset faults of set 2

Uncertain

2 Saddle, Coyote, and Prospect faults

South to southwest

West to northwest, moderate to steep

Eastward Faults seemingly cut and offset faults of set 1

Tertiary tuff mantles faults of set 2; dips ~35°SE

1 (oldest) North Peak, South Peak, Spruce Spring faults

Variable; North Peak fault strikes northeast

Variable and shallow; North Peak fault dips 13°–17° northwest

Variable; eastward for North Peak fault

N.A. (oldest set) Intruded by late Eocene (ca. 38 Ma) rhyolite porphyry

Spruce Spring fault cuts and offsets South Peak fault

Note: N.A.–not applicable.

Page 9: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

245Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

2, individual low-angle faults within fault set 1 in some cases cut and offset other faults grouped in the same set. Collectively, these low-angle faults are the oldest extensional structures that crop out at Spruce Mountain (Table 2) and have, therefore, been included within the same fault set.

Although not strictly included as part of the structural analysis presented here, a final set of small-offset, high-angle faults crops out in the northeastern-most portion of Figure 3. These faults are part of a family of transfer faults that occur as subsidiary structures to the Pequop syncline and appear to incorpo-rate a large component of strike-slip offset (Fraser et al., 1986; Swenson, 1991). These faults likely accommodated localized differential strains perpendicular

to the axis of the syncline during Mesozoic folding. Given the probable com-pressional origin of these faults and the fact that they are not intersected by the restored cross section, they are not analyzed further here.

Structure Contour Maps of Low-Angle Faults

Three major low-angle faults (the North Peak, South Peak, and Spruce Spring faults) that are assigned to fault set 1 (Table 1) crop out on the peak of Spruce Mountain (Fig. 3). Structure contour maps of these faults were gen-erated from the intersection of their mapped traces with topography to bet-ter constrain their dips and three-dimensional geometries (Fig. 4). All three low-angle faults appear to have relatively planar geometries where they crop out at the surface (Fig. 4). Dips of faults calculated from these structure contour maps show that the Spruce Spring fault dips 15°–18° to the southwest, the South Peak fault dips south ~20°, and the North Peak fault dips northwest at 13°–17°. A subsidiary low-angle fault was originally mapped by Hope (1972) structurally above the North Peak fault on the North Peak of Spruce Mountain that juxtaposed the Ely Limestone with the Diamond Peak Formation and/or Chainman Shale (undivided). Field observations documented in Pape (2010), however, provide no evidence for this subsidiary fault. Instead, the contact be-tween the Diamond Peak Formation and overlying Ely Limestone is interpreted to be a depositional contact.

The direction of slip on these low-angle faults has not been directly deter-mined. Strata in the upper and lower plates of these faults dip fairly uniformly eastward, and stratigraphic section is always missing between the hanging walls and footwalls of the faults, which place younger strata on older strata. A westward, normal-sense, hanging-wall transport direction can explain these observations without requiring exceptionally large amounts of slip on these faults (Hope, 1972).

Mapped relationships among these faults (Fig. 3) indicate that the Spruce Spring fault is structurally above and truncates the South Peak fault (Hope, 1972). The relationship between the North Peak fault and the Spruce Spring and South Peak faults is less certain. The North Peak fault may be truncated by the northward projections of the Spruce Spring and South Peak faults (Hope, 1972); alternatively, the South Peak fault and North Peak fault may be the along-strike continuation of the same fault.

GEOLOGIC CROSS SECTION

Stratigraphic and Structural Constraints

As part of our structural analysis of the Spruce Mountain district, we con-structed a northwest-southeast–trending cross section through the Spruce Mountain area showing our interpretation of the subsurface distribution of faults and stratigraphic units (Fig. 5A) as well as an overlay schematically show-ing zones of alteration (Fig. 5B). Thicknesses of units are based primarily on

West fault

East

faul

t

Prospect fault

North Peak fault

Min

e Sh

aft f

ault bx

1km

e ee

e

e

Mdpc

Qa

Qa

Mdpc

DSdu

DSdu Mdpc

Qa Pp

PpPpPrh

Op

9600

9200

8800

9600

10000

North Peak fault

South Peak fault

9200

8800

Spruce Spring

fault9600

920088008400

8000

7600

7200

Figure 4. Structure contours on gently dipping Spruce Spring, South Peak, and North Peak faults constructed from intersections of mapped fault traces with topography. Contours are elevations in feet at 200 ft (~60.9 m) intervals; contours are dashed where projected beyond the fault plane intersection with the land surface. Map unit outlines and abbreviations correspond to those of Figure 3.

Page 10: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

246Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

measurements reported by Hope (1972). Faults belonging to sets 6, 4, 2, and 1 project into the cross section (Fig. 5). Publicly available exploration drilling data in the Spruce Mountain district are sparsely distributed and limited to relatively shallow depths (typically <0.5 km). Therefore, it not possible to meaningfully show the distribution of igneous rocks on the cross section of Figure 5A.

Within the Spruce Mountain district, attitudes of bedding typically do not change substantially between the footwall and hanging-wall blocks across faults, suggesting that these faults are planar or curviplanar in the subsurface. There are few direct measurements of fault dips in the Spruce Mountain area because of poor exposure of fault surfaces in the district. Hope (1972) reported that the East fault and West fault dip ~45° to the east and west, respectively, near where these faults intersect the cross section (Fig. 5). In addition, the

North Peak fault projects into the cross section with an apparent northwest dip of ~5° (Fig. 5). Otherwise, dips on faults are based on relationships between topography and the surface traces of faults.

It is uncertain how the Spruce Spring and South Peak faults may project into the cross section (Figs. 3 and 4). Hope (1972) interpreted both of these faults as steepening northward and truncating the North Peak fault south of the cross section. Alternatively, the South Peak fault could represent the along-strike continuation of the North Peak fault if the projection of this fault is bowed over the peak of Spruce Mountain. It is also possible that the Spruce Spring fault may project into the cross section at high structural and stratigraphic levels. However, given the uncertainties in its three-dimensional geometry, the Spruce Spring fault was not projected into the section of Figure 5.

OCl

OCl

OCl

OCl OCl

DSd

OCl

DSd

DSd

DSd

Ds

Ds

Ds

Mine Shaft fault

Pros

pect f

ault

West fault

East fault

North Peak fault

Banner Hill fault

Coyo

te fa

ult

Dg

Dg

DgDg

Dg

Mdcp

MdcpMdcp

Mdcp North Peak fault

Dg

Mdcp

Prh

Prh PrhPp Pp

WNW ESE

0

OCl

DSd

Ds

MDgu

MdcpPp

Prh

Fault Sets (Faults of sets 5 and 3 do not project into

section)

Youngest

Oldest

6421

North Peak of Spruce Mountain

3 kmEl

evat

ion

(km

)

1

2

3

Sadd

le fa

ult

DSdu

Permian Pequop Formation

Permian Rib Hill Formation

Pennsylvanian-PermianRiepe Spring Limestone

Pennsylvanian Ely Limestone

Ely LimestoneUpper Member

Ely LimestoneLower Member

Silurian-Devonian Dolomite, undivided

Devonian-Mississppian Guilmette/Pi-lot-Shale/Joanna Limestone, undivided

Mississippian Chainman/Diamond Peak, undivided

Ordovician Pogonip Group

Devonian Simonson Dolomite

Silurian Devonian Simonson Dolomite

el

el

eu

eu

euel

el P rsP rs

P rs

P rs

eu

ele

eu eu

P rs

A

Figure 5 (on this and following page). Geologic cross section across the Spruce Mountain area along line of section shown in Figure 3, with no vertical exaggeration. (A) Cross section showing litho-logic units and extensional faults color coded by relative ages, with projected locations of measured bedding attitudes shown near the surface. To aid in constraining the fault restoration, certain lithologic units have been subdivided relative to the associated map (Fig 3B); where this is the case, unit correlations are shown in the legend of the cross section. Note that the igneous rocks are omitted because of poor three-dimensional control (see text).

Page 11: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

247Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Present-Day Distribution of Structures

The cross section shown in Figure 5 represents the interpretation of the present-day structure of the Spruce Mountain area that best fits available data. A critical aspect of the subsurface structure is the geometry of the low-angle North Peak fault, which is cut and downdropped on either side of the peak of Spruce Mountain by the East and West faults (Fig. 5). In Figure 5, the North Peak fault is inferred to maintain a shallow west-dipping orientation in the sub-surface and to root westward in the apparent hanging-wall transport direction. The eastern continuation of the North Peak fault is interpreted to project above the present-day land surface in the eastern part of the cross section, where it is cut by the small-offset Coyote fault, and to project above the southern Pequop Range east of Spruce Mountain (Figs. 3 and 5).

This interpretation differs substantially from the previous structural in-terpretation along the same line of section by Hope (1972), who depicted the North Peak fault as abruptly changing orientations from a subhorizontal dip

on the peak of Spruce Mountain to dipping ~40° E in the eastern portion of the cross section after being downdropped by the East fault. Hope’s (1972) struc-tural interpretation results in the North Peak fault projecting beneath the south-ern Pequop Range, where no surface exposures of major extensional faults are present (Fraser et al., 1986; Swenson, 1991), and, therefore, requires the entire southern Pequop Range to be underlain by the North Peak fault.

STRUCTURAL RECONSTRUCTION

Methodology

A stepwise cross-sectional restoration of extensional faulting illustrates the Cenozoic structural evolution of the Spruce Mountain area (Fig. 6). The primary geologic markers employed for restoring fault offsets are Ordovician through Permian sedimentary rocks that are assumed to maintain relatively

B

Mine Shaft fault

Pros

pect f

ault

West fault

East fault Banner Hill fault

Coyo

te fa

ult

North Peak fault

WNW ESE

0 Fault Sets

(Faults of sets 5 and 3 do not project into section)

Youngest

Oldest

6421

North Peak of Spruce Mountain

3 kmEl

evat

ion

(km

)

1

2

3

Sadd

le fa

ult

Iron-stained quartz breccia

Quartz-molybdenite veins at depth

Skarn +/- hornfels with copper

Jasperoid

Carbonate replacement deposits with Pb-Zn-Ag-Cu

Alteration ZonesFigure 5 (continued ). (B) Overlay showing distribution of selected alteration-mineralization features, with extensional faults shown for reference from A, color coded by relative age. By comparison with the associated map (Fig. 3B), note that the line of section departs somewhat from the overall trend of the known distribution of alteration mineralization; in those portions of the section, alteration-mineralization features are structurally projected into the cross section. Given the lim-ited information about crosscutting relationships between individual faults, alteration-mineralization features, and intru-sions and their subsurface distribution, this overlay should be regarded as a preliminary, largely speculative interpretation.

Page 12: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Rese

arch

Pape

r

248

Pap

e et

al.

| Str

uctu

ral r

econ

stru

ctio

n, S

pru

ce M

ount

ain,

Nev

ada

GE

OS

PH

ER

E |

Volu

me

12 |

Num

ber

1

Permian Limestone and Siltstone

Pennsylvanian Limestone

Missippian Sandstone, Siltstone, Conglomerate and Shale

Devonian and Silurian Limestone and Dolomite

Ordovician - Cambrian, undivided

OCl

DSd

Ds

MDgu

Mdcp

Pp

Prh

Fault Sets (Faults of sets 5 and 3 do not project into

section) Youngest

Oldest

6421

0 km

~5.7 km

~12.6 kmA

B

C

D

E

4 km

4 km

0 km

eu

el

P rs

WNW ESE

Minimum relative paleosurface elevation allowed at each reconstruc-tion step (ignores paleotopography)

Present-day structure

Fault set 6 restored

Fault set 4 restored

Fault set 2 restored

Fault set 1 restored

(pre-extension)

Figure 6. Stepwise structural reconstruction of the central Spruce Mountain area, based on cross section of Figure 5. (A) Present-day structure and pin lines along cross section, with faults and key stratigraphic horizons projected above the topographic profile. Horizons projected above present-day topography are shown with 50% transparency. (B–E) Sequential restoration of crosscutting normal faults of sets 6, 4, 2, and 1 (faults of sets 3 and 5 not present in cross section). The dashed gray horizontal line shown in each restoration panel represents the minimum relative paleosurface elevation allowable at each restoration step (ignoring local topography). The magnitude of fault block rotation applied to each panel relative to immediately preceding panel is as follows: (B) 22°WNW (C) 33°ESE (D) 20°WNW (E) 27°WNW.

Page 13: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

249Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

constant gross thicknesses within the study area. No attempt was made in the reconstructed cross section to differentiate stratigraphic units older than the Pogonip Group, which is the oldest marker employed in the reconstruction, and igneous units are omitted for reasons cited herein.

Roughly simultaneous movement on subparallel faults is expected from geometrical considerations (e.g., Davison, 1989), and certain active fault sys-tems in the Basin and Range province behave in that manner. For example, major steeply east dipping faults at Rainbow Mountain, Dixie Valley, and Fair-view Peak, western Nevada, exhibited coordinated movement during a series of earthquakes in 1954 that accommodated westward tilting of the interven-ing basins and ranges. In addition, concurrent movement was observed on the west-dipping, antithetic Westgate and Gold King faults immediately east of that fault system (e.g., Slemmons and Bell, 1987; Hodgkinson et al., 1996; Caskey et al., 1996). Detailed studies of areas with multiple sets of Cenozoic domino-style faults are also consistent with coordinated movement on major subparallel normal faults and synthetic splays (e.g., Proffett, 1977; Gans and Miller, 1983).

For Spruce Mountain, we utilize a stepwise approach to restoration, restor-ing movement on all faults of a given set concurrently, beginning with the youngest set (set 6), then working backward in time, set by set (Fig. 6). The movement on any given fault is assumed to become locked when the stress required to produce continued slip along them is greater than required to ini-tiate a new set of high-angle faults. Once a fault is cut and offset by a fault of a younger set, the older fault also is buttressed against any renewed movement. Given the nature of the geologic complexity coupled with limited three-dimen-sional exposure at Spruce Mountain, however, it is not possible to preclude the possibility that some faults with opposing dips that are members of adjacent fault sets might have been active concurrently. The offsets on faults in the cross section (Fig. 5A), which belong to sets 6, 4, 2, and 1, are restored in a stepwise fashion from youngest to oldest (Fig. 6). The cross section is restored as a series of rigid fault blocks, ignoring internal deformation (e.g., folding) within individual fault blocks. These assumptions can lead to space problems where fault blocks are differentially tilted. However, these space problems are small relative to the uncertainties of subsurface fault dips and are assumed to be accommodated by slight curvature of faults and/or internal deformation within major fault blocks.

Evidence for Tilting

Several lines of evidence suggest that extensional faulting was accompa-nied by fault block tilting, ultimately resulting in a net eastward tilting of fault blocks in the Spruce Mountain area. (1) Outcrops of (Miocene?) sedimentary rocks within the southern portion of the study area display generally eastward dips between 15° and 30°. (2) The trace of the basal contact of the undated (likely either Eocene or Miocene) Cenozoic dacite tuff relative to topography (Fig. 3) suggests that this tuff dips gently to moderately eastward in the central

Spruce Mountain district. Moreover, although no strike and dip measurements directly constrain the bedding orientation of this tuff in the immediate study area, a single dip measurement of 35°SE was obtained on a small outcrop of the tuff ~3 km beyond the northern edge of the area in Figure 3 (Hope, 1972), and is consistent with the observation that the tuff unit as a whole appears to have been tilted eastward. (3) Paleozoic strata within the Spruce Mountain area dominantly dip eastward, which is consistent with eastward extensional tilting if these strata were horizontal or gently dipping prior to Cenozoic extension. This is a reasonable assumption if the preextensional structure in the Spruce Mountain area was characterized by broad open folding of a style similar to the adjacent Pequop syncline immediately to the east. (4) East-dipping Paleozoic strata in the footwalls and hanging walls of the normal faults that currently dip at low angles generally display moderate to high (>40°) cutoff angles with these faults where they are exposed at the surface, suggesting that these faults initiated at high to moderate angles. (5) The inferred westward hanging-wall transport directions of the currently low-angle faults is consistent with east-ward fault block tilting if these faults initiated as higher angle, west-dipping normal faults.

Given this evidence, rotation consistent with the sense of slip of each re-stored fault set is applied to the stepwise cross-sectional restoration (Fig. 6). Although the magnitude of tilting associated with each fault set is uncertain, there are some loose constraints on the overall history of tilting (Table 2). The undated southeastward-dipping Cenozoic tuff mantles faults of sets 1 and 2, implying that most or all of the southeast-directed tilting of this unit resulted from movement on faults of younger sets. In addition, the sedi-mentary rocks in the southeastern part of the map area, probably belonging to the middle Miocene Humboldt Formation, lap across faults of set 4 and appear to have fanning-upward dips from 30° to 15°E (Table 2). This sug-gests that fault blocks in the southern part of the district have undergone a net eastward tilting of as much as 30° since the mid-Miocene. Given the relative age constraints, and considering that both the Cenozoic dacite tuff and sedimentary rocks display similar eastward dips in the central Spruce Mountain district, it is likely that most of the eastward tilting of these rocks occurred in association with slip on west-dipping faults grouped with fault set 6 (Figs. 3 and 6). Because cutoff-angle relationships suggest that the currently low-angle faults in the Spruce Mountain area most likely initiated at high to moderate angles, the final reconstructed panel (Fig. 6E) is shown tilted 36° west relative to the present-day cross section. This restores the North Peak fault, which currently dips northwest at 13°–17° (Fig. 4), to an ini-tial northwestward dip of ~45°–50° (Fig. 6E). Although the Spruce Mountain area contains faults of varying polarity over time (Table 2), the net eastward tilting of ~35° suggests that movement on faults of the three sets of down-to-the-west faults, of the first, second, and sixth generations of normal faults, dominated the overall tilting history.

Further constraint on the approximate history of fault block tilting associ-ated with each fault set comes from the probable amount of differential burial and/or exhumation of the Paleozoic marker strata at each time step in the

Page 14: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

250Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

restoration. Apatite fission track and (U-Th)/He analyses from probable Ju-rassic granite plutons that intrude Permian strata at Delcer Buttes (~25 km southwest of Spruce Mountain) are consistent with these strata having re-mained at low temperatures (<60 °C) throughout the Cenozoic (Colgan et al., 2010). Given the proximity of the Spruce Mountain district to Delcer Buttes (Fig. 1) as well as the presence of probable Paleogene rocks that unconform-ably overlie Permian and Pennsylvanian rocks in the Spruce Mountain district, it is reasonable to assume that Permian strata were not deeply buried (i.e., no greater than 3 km, and probably less) at Spruce Mountain at the onset of Cenozoic extension. These Permian rocks most likely remained at relatively shallow depths throughout active Cenozoic extension as well. Because rocks at or below the present-day land surface must have also been at or below the paleosurface throughout the deformational history of Spruce Mountain, the cross-sectional restoration provides constraint on the deepest possible paleo-surface elevation relative to the restored Paleozoic strata at each time step. This paleosurface constraint is depicted in Figures 6B–6E as a straight, dashed line (ignoring any paleotopographic relief). At each restoration step, tilting magnitudes were applied to the section that maintain consistency with prob-able shallow burial depths of Permian and younger strata at Spruce Mountain during the Cenozoic. The restoration shown in Figure 6 does not require more than 2 km of burial for Permian marker rocks at any given step in the resto-ration (although greater burial depths are still permissible within the available constraints) and is also consistent with other tilting constraints based on dips of rocks in the district.

Restored Preextensional Structure of the Spruce Mountain Area

The magnitude of extension restored in the cross section, which is the dif-ference in final and initial distances between reference lines of 12.6 and 5.7 km, respectively, is ~6.9 km, which is equivalent to 120% extension (Fig. 6). Al-though the total amount of extension across the section is large, it was distrib-uted among numerous faults. The largest amount of slip on any one fault in the restored section, the Banner Hill fault, is 1.9 km; however, the North Peak fault, West fault, Mine Shaft fault, and Saddle fault all have offsets >1 km. Some low-angle faults mapped in the area, such as the Spruce Spring fault, were not projected into the reconstructed cross section; if, alternatively, they are present in the section, then this would increase the estimate of extension across the Spruce Mountain area, such that the estimates of the amount of extension based on Figure 6 may be closer to a minimum value.

Note that reconstruction of Figure 6 implies that some beds rotated through horizontal as a result of tilting associated with multiple generations of normal faults; for example, Paleozoic rocks on the eastern side of the map area that currently dip 20°–30°SE (Figs. 3 and 6A) would have dipped ~10°–15° northwest prior to faulting (Fig. 6E). As a result, the attitudes of these Paleozoic rocks in the restored section imply west-vergent folding and gentle westward dips (~15°). The overall westward dip of beds in the restored section is inter-

preted to reflect that the Spruce Mountain district was involved in large-scale, open folding of a character similar to the Pequop syncline immediately east of Spruce Mountain. Spruce Mountain may have been on the western limb of an anticline that connected to the Pequop syncline prior to being dismembered by extensional faulting and may have connected to another upright syncline further west. The kilometer-scale, west-vergent anticline shown in the restored cross section (Fig. 6E) is consistent with the scale and character of west-ver-gent folding and thrust faulting that occurred in the southern Pequop Moun-tains during development of the Pequop syncline (Swenson, 1991) and with the notion that the Spruce Mountain area was similarly deformed during the same contractional deformational event.

Regional studies of Mesozoic deformation and metamorphism in the Pequop Mountains–Wood Hills–East Humboldt Range region to the north of the Spruce Mountain area and in the Ruby Mountains to the west have demon-strated that significant crustal thickening must have occurred in northeastern Nevada during the Mesozoic (e.g., Thorman, 1970; Snoke and Miller, 1988; Camilleri and Chamberlain, 1997). Metamorphic mineral assemblages and sheared fabrics locally present in impure limestones of the Pogonip Group on the southern side of Spruce Mountain have been interpreted to indicate that these rocks reached lower greenschist facies metamorphic conditions during the Mesozoic and that significant structural thickening occurred in the Spruce Mountain area during the Mesozoic (Camilleri and Chamberlain, 1997). However, no thrust faults have been mapped locally in the Spruce Mountain district, and the reconstructed cross section (Fig. 6) suggests that significant structural duplication of stratigraphic section between Ordovician and lower Permian strata has not occurred at Spruce Mountain. Thus, any significant structural thickening and overthrusting at Spruce Mountain during the Meso-zoic was likely limited either to deeper or, possibly, shallower structural and stratigraphic levels than are currently exposed at the surface.

Constraints on Age of Extensional Faulting at Spruce Mountain

Several observations constrain the timing of faulting at Spruce Mountain. First, all normal faults in the Spruce Mountain district deform Paleozoic rocks and dismember older Mesozoic contractional structures and, given regional considerations regarding the onset of extension in the Great Basin (e.g., Wer-nicke et al., 1987; Seedorff, 1991a; Dickinson, 2006), are most likely all Cenozoic structures, although ages as old as Late Cretaceous are possible. Eocene rhyo-lite porphyry dikes (ca. 38 Ma; Table 1) crop out throughout the central portion of the Spruce Mountain district (Fig. 3). These dikes cut or intrude the North Peak fault (a member of fault set 1) and locally intrude the Mine Shaft fault (a member of fault set 4) (Fig. 3; D. Pace, 2013, personal commun.). Thus, both of these faults, and by analogy other faults of these sets, were active during or prior to the late Eocene. Sedimentary rocks that crop out in the southern part of the study area, which probably correlate with the middle Miocene Hum-boldt Formation (Harlow, 1956; Coats, 1987), dip eastward and appear to dis-

Page 15: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

251Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

play fanning-eastward dips, indicating that at least some extensional faulting within the Spruce Mountain district occurred during the middle to late Mio-cene. Based on relative age constraints, some of the faults that are grouped in sets 5 and 6 were probably active in the middle to late Miocene. The ages of faults that were active during the Quaternary have been determined mostly on the basis of photogeologic evidence (e.g., Dohrenwend et al., 1991, 1996). As shown by dePolo (2008), several normal faults mapped on the eastern sides of Spruce Mountain Ridge and Spruce Mountain ruptured in the Quaternary (i.e., in the past 1.8 m.y.), and the latest slip on normal faults mapped along the western sides of Spruce Mountain Ridge and Spruce Mountain probably is even more recent (latest Quaternary, i.e., <130 k.y.). Notably, several faults grouped in fault set 6 appear to have been active during the Quaternary, in-cluding the relatively large offset West fault. Given the current absolute and relative age constraints at Spruce Mountain, it is not possible to determine rigorously which faults grouped in sets 5 and 6 initiated during the probable middle to late Miocene extensional event and which faults initiated later. It is possible (and even likely) that some faults included in these relatively young-est fault sets initiated during the middle to late Miocene and subsequently were active more recently, whereas other faults also grouped in these sets may have initiated much later.

Restoration of Magmatic-Hydrothermal Features at Spruce Mountain

The reconstruction in Figure 6 and crosscutting relationships can be uti-lized to infer the original geometry of mineralized zones at Spruce Mountain. Because porphyry dikes intrude faults of sets 1 and 4, and quartz veins appear to have formed along the Prospect fault of set 2, it appears that at least some, if not all, of the hydrothermal alteration features formed after movement on faults of sets 1, 2, and 3, and perhaps during movement on faults of set 4. Thus, the hydrothermal system might have been subjected to extension and associated tilting related to movement on faults of sets 5 and 6 and possibly some portion of the movement on faults of set 4. Consequently, the alteration zones were probably formed after the time represented in the restoration by Figure 6C, but before that in Figure 6B. Figure 7 shows the restored distribution of interpreted hydrothermal features (Figs. 3B and 5B) prior to movement on faults of sets 4 through 6.

Given the large geologic uncertainties, the distribution of features is plausi-bly consistent with a relatively upright hydrothermal system that developed as faults of set 4 began to move. Based on this reconstruction, the original two-di-mensional geometry of mineralized zones was a ~6 km long by ~3 km deep el-liptical pattern with carbonate replacement deposits surrounding skarn. In ad-dition, given our assumptions regarding the timing of mineralization relative to extensional faulting, the stepwise reconstruction of Figure 6 indicates that the hydrothermal system underwent ~10° of westward net tilting and ~2.7 km of extension associated with postmineralization deformation by fault sets 4 through 6.

DISCUSSION

Comparison and Classification of Mineral Deposits at Spruce Mountain

The presence of rhyolite and granite porphyry dikes, quartz-molybdenite veins, and the geochemical enrichment in Sn-W-F suggest that the mineral deposits at Spruce Mountain, except for the Carlin-type prospect, are various manifestations of a porphyry molybdenum deposit. The limited subsurface information available suggests that the proximal part of the system contains quartz-molybdenite veins and is relatively copper free, in contrast to porphyry molybdenum deposits of the quartz monzonitic-granitic porphyry Mo-Cu sub-class such as the Hall (Nevada Moly) and Buckingham deposits in Nevada (Seedorff et al., 2005). The best copper grades at Spruce Mountain seem to be somewhat more distal, located in skarns that are developed fairly close to contacts with rhyolite and granite porphyry.

The best analogues for Spruce Mountain are the Mount Hope deposit in Eureka County 145 km to the southwest (Westra and Riedell, 1996) and several molybdenum prospects in White Pine County to the south (Seedorff, 1991a). Spruce Mountain and Mount Hope are assigned to the rhyolitic porphyry Mo category (of Seedorff et al., 2005), which includes end-member deposits such as Climax and Henderson (Colorado), but also deposits that have less evolved igneous compositions (e.g., lower SiO2, Rb, and Nb and higher Sr and Zr con-tents) and metals with higher overall Cu:Mo ratios, such as Mount Hope. The latter deposits were referred to as transitional deposits by Westra and Keith (1981, Table 2 therein) and Carten et al. (1993, Table II therein). In the classifica-tion of Ludington and Plumlee (2009), the Spruce Mountain deposit would be regarded as a Climax-type deposit.

Spruce Mountain may be the best example of a rhyolitic porphyry Mo or Climax-type porphyry molybdenum deposit with extensively developed skarn. The overall copper content of Spruce Mountain and certain prospects in White Pine County such as the Ellison district (Johnson, 1983) may be somewhat higher than other rhyolitic porphyry molybdenum deposits where carbonate rocks are largely to entirely absent, such as Climax, Henderson, Mount Hope, and Pine Grove (Utah). Carbonate wall rocks probably promoted efficient pre-cipitation of copper in skarn at Spruce Mountain and the occurrences in White Pine County compared to other rhyolitic porphyry Mo deposits.

Significance of Ages of Igneous Rocks and Ore Deposits at Spruce Mountain

With some notable exceptions, including a large body of work in the Ruby Mountains–East Humboldt metamorphic core complex (e.g., Hodges et al., 1992; Wright and Snoke, 1993; Brooks et al., 1995; Henry, 2008; Bedell et al., 2010), there are few radiometric ages in this part of northeastern Nevada, where exposures of igneous rocks generally are sparse. U-Pb ages reported here for igneous rocks at Spruce Mountain are Late Jurassic and late Eocene.

Page 16: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

252Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

The Jurassic rocks at Spruce Mountain occur as small dikes that are dated as 155–156 Ma. Large Jurassic intrusions occur in the Dolly Varden Moun-tains, Delcer Buttes, Ruby Mountains, and Snake Range (Miller et al., 1988; Miller and Hoisch, 1995; Barton et al., 2011), and small dikes of Jurassic age are widespread but sparsely present in other parts of northeastern Nevada, such as in the Pequop Mountains (Bedell et al., 2010; Camilleri, 2010; Wyld and Wright, 2015). At some localities such as Spruce Mountain, the Jurassic rocks

are barren, whereas they are at least variably mineralized at the present levels of exposure at other sites, such as the Delcer district (LaPointe et al., 1991) and the Victoria skarn breccia pipe in the Dolly Varden district (Atkinson et al., 1982; Zamudio and Atkinson, 1995).

Dated rhyolite porphyries at Spruce Mountain range in age from ca. 39 to ca. 37 Ma (Table 1) and have ages similar to those of granite and rhyolite porphyry bodies in nearby parts of eastern Nevada (Stewart and Carlson, 1976;

Iron-stained quartz breccia

Quartz-molybdenite veins at depth

Skarn +/– hornfels with copper

Jasperoid

Carbonate replacement deposits with Pb-Zn-Ag-Cu

Alteration Zones

0 3 km

Fault Sets

Youngest

Oldest

21

Figure 7. Interpretive preliminary reconstructed cross section of selected alteration-mineralization features at Spruce Mountain, showing how present-day distribution of features from Figure 5B might be distributed prior to movement of faults of set 4 in geologic reconstruction of Figure 6.

Page 17: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

253Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Seedorff, 1991a). The Spruce Mountain porphyry molybdenum deposit, dated here as ca. 38 Ma, has an age similar to that of the Mount Hope porphyry molybdenum deposit in Eureka County (ca. 37 Ma; Westra and Riedell, 1996).

Although ore deposits have formed in the Great Basin over many inter-vals of geologic time, especially numerous epithermal deposits in the Miocene (e.g., John, 2001), the late Eocene was an important period of ore formation in the Great Basin, including Carlin-type gold deposits, porphyry copper and related skarn, replacement, vein and distal disseminated deposits, porphyry molybdenum, and a few adularia-sericite–type or low sulfidation–type epither-mal silver-gold deposits (e.g., Seedorff, 1991a).

Structural Restoration Uncertainties and Constraints

Section balancing and restoration techniques were originally developed in contractional settings (Dahlstrom, 1969) and have since been widely applied in the structural analysis of fold and thrust belts (e.g., Hossack, 1979; Mitra, 1992; Wilkerson and Dicken, 2001). Cross-sectional reconstructions are less commonly attempted in extensional environments but have been increasingly applied in settings of continental extension, particularly in the Basin and Range province (e.g., Proffett, 1977; Wernicke and Axen, 1988; Smith et al., 1991; See-dorff, 1991a; McGrew, 1993; Brady et al., 2000; Colgan et al., 2008, 2010; Nick-erson et al., 2010). As these extensional restorations become more commonly applied, it is important to acknowledge key uncertainties that are inherent to cross-section reconstructions, particularly in areas of polyphase deformation such as northeastern Nevada.

A common uncertainty in cross-section restorations, and one of the most important uncertainties in the Spruce Mountain restoration presented here, is the slip direction of the individual restored faults and the possibility of fault movement out of the plane of the restored cross section. No reliable piercing points or slickenline data are available from within the Spruce Mountain dis-trict to fully constrain the kinematic histories of each fault, and as a simplifying assumption for the cross-sectional restoration, all restored faults in Figure 6 are assumed to have moved in a pure dip-slip sense, parallel to the line of sec-tion. Although this is probably a reasonable approximation of the major fault slip directions given that the line of section was chosen to extend orthogonal to the strike orientations of the major mapped normal faults, this simplified kinematic history is unlikely to be true in detail. This simplifying assumption of pure dip-slip motion along the line of the restored cross section effectively adds uncertainty to the estimates of offset on individual faults, as well as the extension estimates at each step of the cross-section restoration.

In addition, uncertainties regarding present-day fault dips and the magni-tude of tilting applied to each step of the restored cross section have a direct impact on extension estimates at each step of the restoration. Of these two variables, the net tilting magnitude uncertainty is the most easily quantified. For example, we restore ~35° of net eastward tilting between the present day and reconstructed cross sections (Figs. 6A, 6E). Although this tilting magnitude satisfies all available geologic constraints, it is likely close to a minimum esti-

mate of the overall net tilting associated with extensional faulting at Spruce Mountain. As much as 10° of additional east-directed tilting could reasonably be restored in the final reconstructed cross section (Fig. 6E) without requiring excessive burial depths of Permian sedimentary rocks. Restoring an additional 10° of eastward tilting in Figure 6E would increase the overall extension esti-mate by 0.8 km to 7.7 km. Moreover, the tilting uncertainties of the intermediate restoration steps are more poorly constrained than the overall net tilting his-tory, and based on burial and/or exhumation and fault geometry constraints, we estimate a tilting uncertainty of ±10° for each intermediate step. Although it is a straightforward exercise to quantify the impact of fault dip uncertainty on extension estimates for an individual curviplanar normal fault or even a single set of kinematically linked normal faults (e.g., Wernicke and Burchfiel, 1982), for an intensely faulted area with multiple overprinting, crosscutting fault sets, such as Spruce Mountain, this exercise becomes more challenging. In part this is because, in a complexly faulted area like Spruce Mountain, the crosscutting fault relationships themselves provide considerable constraint on the range of structurally viable fault dips and geometries in the subsurface, and individual fault geometries frequently cannot be varied independently and still result in viable retrodeformable cross sections. Thus, in structurally complex areas like Spruce Mountain, accurate quantification of the impact of fault dip uncertainty on extension estimates would necessitate the construction of multiple alterna-tive cross-section reconstructions.

The polyphase deformational history of northeastern Nevada further complicates the structural restoration of the Spruce Mountain district (Figs. 6 and 7). Specifically, the primary structural markers used for restoring fault off-sets are Paleozoic passive margin strata, but these strata had an earlier history of contractional deformation during the Mesozoic and possibly in the Paleozoic (Snyder et al., 1991; Carpenter et al., 1993; Camilleri and Chamberlain, 1997; Howard, 2003; Trexler et al., 2003, 2004; Dickinson, 2006; Long, 2012, 2015; Rhys et al., 2015). At Spruce Mountain, Cenozoic extension has dissected and obscured the earlier contractional structures. Hence, there are uncertainties in both the nature of the extensional deformation, as well as the configuration of the preextensional structure. Consequently, an iterative approach is used in this study to produce a viable cross-sectional restoration, wherein the pre-exten sional model is progressively modified to produce a restored extensional structural geometry that satisfies all of the available geologic constraints. This process results in a degree of internal circularity within the retrodeform-able cross section because the extensional structure and the preextensional structural model are not mutually independent, i.e., uncertainties regarding the structure of both the extended and restored states of the cross sections effec tively increase the number of potentially valid retrodeformable cross-sec-tional models.

Additional uncertainties are also introduced when attempting to recon-struct mineralized systems. Reconstruction of the geometry of magmatic-hydro thermal systems requires constraints on the ages of various generations of faults relative to the ages of intrusion and associated alteration mineraliza-tion, but the geometric constraints applicable to preextensional hydrothermal

Page 18: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

254Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

systems, such as Yerington (Proffett, 1977; Richardson and Seedorff, 2015) or Teacup (Nickerson et al., 2010), are more straightforward than for synexten-sional hydrothermal systems such as Spruce Mountain. The observation that a given intrusive rock intrudes a fault requires that the intrusion postdate at least the initial movement on the fault; if the intrusion (e.g., a dike) is sheared or brecciated along the fault, then the same fault may have continued moving after the dike intruded the fault zone. Thus, the amount of postemplacement deformation of the dike associated with movement on the fault may range from 0% to nearly 100% of the total deformation associated with movement on a given fault, and this may introduce considerable uncertainty regarding the amount of postemplacement extension. Analogous reasoning applies to the relative ages and amounts of postmineral extension of hydrothermal systems that may be related to intrusions.

Despite these uncertainties, every step in a viable reconstruction must sat-isfy certain firm constraints: they cannot imply erosion of units that are still present in the bedrock today, nor can bedrock units exposed at the surface today be buried to depths during any stage in the structural evolution that are inconsistent with their present-day textures and thermochronologic charac-teristics. Moreover, the restoration must result in reasonable fault geometries and imply plausible fault kinematics. The reconstruction presented here meets these tests, although there are uncertainties in the assignments of the amount of extension to any generation of faults. The most robust reconstructions are possible from locales where there are abundant, widely distributed, pre-ore, syn-ore, and post-ore strata and intrusions that create compelling crosscutting relationships and that contain radiometrically datable units.

Distribution, Polarity, Amount, and Style of Large-Magnitude Extension

Cenozoic upper crustal extensional strain was generally directed east-west to northwest-southeast within northeastern Nevada and is distributed hetero-geneously, with areas of high extensional strain separated by relatively low strain areas (Seedorff, 1991a; Colgan and Henry, 2009). This heterogeneous distribution of upper crustal extensional strain is a characteristic feature of extension in the Great Basin (e.g., Gans and Miller, 1983). Combined with a relatively uniform crustal thickness of 30–35 km across the eastern Great Basin (e.g., Allmendinger et al., 1987), this heterogeneity requires large-scale middle to lower crustal flow of rocks from domains that have undergone little upper crustal extension into regions of large upper crustal extension (e.g., Gans, 1987; Smith et al., 1991; MacCready et al., 1997).

Consistent with these regional observations, normal faulting and large extensional strain was directed east-west to northwest-southeast at Spruce Mountain and is more intense at Spruce Mountain than in the nearby south-ern Pequop Range, which is an essentially unextended structural block that preserves Mesozoic compressional structures (Swenson, 1991; Camilleri and Chamberlain, 1997). Our reconstruction of Spruce Mountain indicates that the 5.7 km distance between reference lines in the restored cross section (Fig. 6E)

was extended by 6.9 km, resulting in a present-day length of ~12.6 km (Fig. 6A). This is equivalent to 120% extension between the fully restored and exten-sionally deformed states. Spruce Mountain, however, differs somewhat from other nearby highly extended areas in northeastern Nevada in that the large extensional strain in the central Spruce Mountain district is distributed among numerous, closely spaced, small to moderate offset (<2 km) faults in multiple, superimposed sets of faults. By contrast, extension in nearby southern Ruby Mountains appears to have been accommodated primarily by large offset (13–22 km) on a single west-dipping fault that produced eastward tilting (Colgan et al., 2010). Although Colgan et al. (2010) did not specify the amount of exten-sion in the southern Ruby Mountains and vicinity, our choice of reference lines in their reconstruction (Colgan et al., 2010, fig. 11 therein) implies only ~45% extension, albeit over a present-day width of ~19.3 km. The relative amount of extension at Spruce Mountain is roughly three times that of the southern Ruby Mountains, although the absolute amount of extension at Spruce Mountain (~6.9 km) is somewhat smaller than that in southern Ruby Mountains (~8.6 km).

Notably, there is at least one fault within the southern part of the Spruce Mountain district, the Spruce Spring fault, which appears to have much larger offset than any other mapped faults in the district. This currently low-angle fault (which is grouped as part of fault set 1) places Permian on Ordovician rocks in the southern part of the Spruce Mountain district (Fig. 3). Assuming a previously unbroken stratigraphic thickness of ~4.5 km for the Ordovician to Permian sequence (Fig. 2), a ~50° initial westward dip for the Spruce Spring fault, and a gentle westward preextensional dip of 15° of the Paleozoic rocks at Spruce Mountain (Fig. 6), a minimum ~8 km of offset would be required on the Spruce Spring fault to place Permian on Ordovician strata. However, the Spruce Spring fault cuts off the South Peak fault, so the stratigraphic section was not unbroken; therefore, the 8 km estimate may considerably overesti-mate the amount of slip on the Spruce Spring fault.

The more distributed overall style of brittle extension at Spruce Mountain nonetheless contrasts with the style of extension observed in the southern Ruby Mountains. Precisely why extensional strains are in some cases distrib-uted among numerous, closely spaced faults and elsewhere concentrated on a few large-offset faults is not well understood. Pervasive and distributed ex-tensional faulting of bidirectional polarity can occur in transfer zones between major normal fault systems (e.g., Faulds and Varga, 1998). Geophysical mod-els also suggest that the style of extension may be strongly affected by the thermal structure of the lithosphere and that normal faults that undergo large amounts of hydrothermal cooling during extension may have a tendency to accumulate greater amounts of offset (Lavier and Buck, 2002).

Initial Dips of Normal Faults

The angle of initiation of normal faults in continental extensional tecton-ics is a persistent controversy (e.g., Proffett, 1977; Wernicke, 1981; Gans et al., 1985; Roberts and Yielding, 1994; Wong and Gans, 2008), and structural re-construction is an important method for constraining initial dips of faults. The

Page 19: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

255Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

reconstruction of Spruce Mountain presented here as well as a published re-construction of the southern Ruby Mountains (Colgan et al., 2010) indicate that the near-surface dips of major normal faults in this region generally initiated at moderate to high angles (i.e., >45° and commonly ~60°). These faults rotated to lower angles as they moved and, in many cases, subsequently were passively rotated to still lower angles (or higher angles) by movement and associated tilting by crosscutting younger faults. Although uncertainties remain in these reconstructions, the evidence nonetheless suggests that the normal faults in this region, including all of the faults in the Spruce Mountain area, initiated with moderate to steep dips at least through the seismogenic upper crust.

Structural Complexity and Changing Polarity of Faulting

Many extended regions contain more than one generation of normal faults, each of which constituted a normal fault system that was active for a finite period of time. For example, the Yerington district, western Nevada ( Proffett, 1977), the Royston district, west-central Nevada (Seedorff, 1991b), and the Hunter district, east-central Nevada (Gans, 1982) all are characterized by sev-eral generations of faults with northward strikes and down-to-the-east off-sets. As maps and cross sections of the Yerington district reveal, the resulting geol ogy is complex (Proffett and Dilles, 1984; Richardson and Seedorff, 2015). The structure of such extended regions is nonetheless structurally relatively simple, in the sense that most of the faults with significant displacement have similar strikes and senses of displacement.

The geology of the Spruce Mountain area is also complex (Figs. 3 and 5), but in contrast to many areas within the Basin and Range province such as Yerington, the polarity of faulting changed over time: three sets of down-to-the-west faults that produced large amounts of tilting, two sets of down-to-the-east faults that produced small amounts of westward tilting, and one set of down-to-the-north faults with small amounts of southward tilting (Table 2). Periods of incremental eastward tilting were interspersed with periods of in-cremental westward tilting, but time-integration of crustal extension resulted in ~35 ° of net eastward tilting of the district (Fig. 6).

Areas with superimposed sets of faults with variable strike directions or differing senses of displacement produce less net tilting of strata for compara-ble amounts of extension (Axen, 1986). Few reconstructions (e.g., Shaver and McWilliams, 1987) have been previously attempted in regions that have faults of opposing polarities, such as Spruce Mountain, where the structural geome-tries are especially challenging to reconstruct.

Timing of Upper Crustal Extension

Extension across the Basin and Range province is time transgressive (e.g., Wernicke et al., 1987; Seedorff, 1991a; Dickinson, 2002). Within the Great Basin, extension initiated at least as early as the Eocene in the north-central Great Basin, and the most intense extensional deformation in the past 5–10 m.y. may have been in the southwestern Great Basin in the Death Valley–Beatty

region. Sound field, geochronologic, and thermochronologic evidence has shown that there was an important middle Miocene extensional event at many sites around the Basin and Range (e.g., Miller et al., 1999; Stockli et al., 2002; Surpless et al., 2002; Colgan and Henry, 2009). However, not only has there been a spatial migration in the initiation of extension in a given area, but some areas are extended again following a hiatus in normal faulting (e.g., Seedorff, 1991a; Seedorff and Richardson, 2014). Even in places such as the greater Yering ton district where most of the extension is mid-Miocene or younger, extension was partitioned temporally between three intervals: a period of ex-treme extension ca. 15–13 Ma and two less significant periods ca. 11–8 Ma and ca. 4 Ma to present (Dilles and Gans, 1995; Stockli et al., 2002; Surpless, 2012). At other locales, such as the Robinson district in east-central Nevada, many fault generations may have been active within only a few million years or less, without a significant hiatus (Gans et al., 2001).

Given the available age constraints, at least three periods of extension within the Spruce Mountain district can be recognized: an important phase of extension during or before the late Eocene, another period that likely occurred in the middle to late(?) Miocene, and ongoing Quaternary extension. The earli-est period of extension occurred after Mesozoic contraction and before and/or broadly concurrent with intrusion of rhyolite porphyry dikes in the district ca. 38 Ma. This first period of extension coincides with movement on the first four normal fault sets in the Spruce Mountain district (Table 2) and may have been a single continuous event or may have been composed of multiple, discrete extensional episodes separated by periods of relative inactivity. Based on the structural reconstruction (Fig. 6), this early extensional period produced the largest amount of extension at Spruce Mountain (5.4 km). A later period, likely of middle to late Miocene age, produced the east-directed tilting of sedimen-tary rocks in the southeastern part of the district when faults grouped in sets 5 and 6 were active. The most recent period of extension occurred with the latest movement on faults grouped in set 6, which includes range-front fault scarps mapped along the western sides of Spruce Mountain Ridge and Spruce Mountain that offset Quaternary deposits (dePolo, 2008).

In spite of the regional importance of middle Miocene extension in north-eastern Nevada, other periods of extension cannot be overlooked. Extension at ca. 38 Ma or earlier appears to have resulted in considerably more extensional strain at Spruce Mountain than later episodes of extension. Late Eocene ex-tension may have contributed more extensional strain than the mid-Miocene event in other parts of the Great Basin, including areas of east-central Nevada that are south of Spruce Mountain, such as the northern Egan Range and the Robinson district (Gans and Miller, 1983; Gans et al., 1989, 2001). Even in areas where the magnitude of middle Miocene extension overwhelms the impor-tance of older extension, the earlier events nonetheless require geo dynamic explanation. Moreover, late Eocene extension, regardless of magnitude, certainly played a role in localizing ore formation in many locales (e.g., late Eocene ores at Bingham, Utah, Redmond and Einaudi, 2010; at Cove-McCoy, Nevada, Johnston et al., 2008; at Battle Mountain, Nevada, Keeler and See-dorff, 2007; and at many sites in White Pine County, Nevada, Seedorff, 1991a).

Page 20: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

256Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Implications of Crustal Extension and Structural Restorations for Mineral Exploration

Preexisting mineralized systems can be dismembered and tilted by nor-mal faults (e.g., Proffett, 1977; Wilkins and Heidrick, 1995; Rahl et al., 2002; Begbie et al., 2007; Stavast et al., 2008), thereby aiding scientific understand-ing of the systems by creating greater exposures of their vertical and lateral extents (e.g., Dilles and Einaudi, 1992; Seedorff et al., 2005, 2008). Mineralized systems may also provide additional geologic markers for reconstructions, such as hydrothermal alteration or metal zoning patterns, which may aid structural reconstructions, especially where there are large areas with uni-form rock types (Maher, 2008; Nickerson et al., 2010; Houston and Dilles, 2013; Seedorff et al., 2015). Mineralized systems or fault-bound fragments of them, especially those in the footwall blocks of large-displacement normal faults, may be uplifted and eroded during extension, yet in other cases, ore may be downdropped, buried, and preserved under postmineral rocks and may constitute an exploration target (Maher, 2008; Richardson and Seedorff, 2015). Numerical modeling indicates that crustal extension at the scale of an oro-genic belt enhances the preservation of mineral deposits formed at relatively shallow levels of the crust, such as porphyry and epithermal systems; i.e., it postpones their erosion (Barton, 1996). Crustal extension also creates per-meability and drives fluid circulation in the brittle upper crust (Ilchik and Bar-ton, 1997), favoring the formation of certain types of deposits during periods of crustal extension, including epithermal silver-gold deposits, Carlin-type gold deposits, iron oxide–copper–gold deposits including detachment-style Fe-Au-(Cu) deposits), and certain classes of porphyry systems (Dreier, 1984; Spencer and Welty, 1986; Seedorff, 1991a; Tosdal and Richards, 2001; John, 2001; Barton et al., 2011; Coolbaugh et al., 2011). The geometric constraints on determining the amount of deformation of synextensional hydrothermal systems such as Spruce Mountain, however, are less straightforward than for preextensional deposits (see preceding).

Although the precise amount of deformation of the porphyry molybdenum system at Spruce Mountain system is not tightly constrained, it is clear that the Spruce Mountain system has been structurally dismembered, with deeper lev-els generally exposed on the western side and shallower levels on the eastern side of the district. Abrupt increases in grade with depth occur beneath some post-ore faults, and some of the highest molybdenum grades intersected to date are in the footwall of a steeply dipping post-ore normal fault. Thus, sev-eral of the generations of faults postdate ore formation, and the net tilting of the highly segmented hydrothermal system is ~10° west. It is highly unlikely, therefore, that the mineralizing plutons are entirely upright or fully intact. Early generations of normal faults are locally intruded by porphyry dikes and host tabular bodies of quartz-rich breccia. Mineralized skarn is spatially associated with the margins of porphyry dikes and occurs in certain premineral to syn-mineral normal faults.

The structural and magmatic setting of Spruce Mountain may be analo-gous to the Hunter district in White Pine County, where a rhyolitic center

was emplaced at the onset of extension in that area during the late Eocene. Hydrothermal activity accompanied emplacement of porphyry dikes along normal faults, and faults continued to slip after crystallization of the dikes. In the Hunter district, most, if not all, normal fault–related tilting postdated in-trusion of the dikes and extrusion of coeval volcanic rocks (Gans, 1982; Gans and Miller, 1983). Both the Hunter and Spruce Mountain districts are synexten-sional porphyry systems. Occurrences of Carlin-style mineralization at Spruce Mountain may belong to a genetically unrelated, earlier or later, system that partially overlaps the porphyry molybdenum system spatially. Spatial super-position of genetically unrelated systems is common in the Great Basin (e.g., Keeler and Seedorff, 2007) and is especially common for Carlin-type gold sys-tems (e.g., Seedorff, 1991a; Seedorff and Barton, 2004; Hastings, 2008; Lubben et al., 2012; Hoge et al., 2015).

The early sets of faults at Spruce Mountain are prospective because they could have influenced the locations of porphyry intrusive centers, and they were conduits for hydrothermal fluid flow that produced skarn and car-bonate replacement mineralization. In contrast, later generations of faults have dismembered the system and contributed to its post-ore tilting; hence, de-tailed structural restorations may be used to predict where mineralization is in a given block, where it is displaced by a post-ore fault, and thus where it might be discovered in nearby fault blocks. Results of future exploration activities will permit refinement or revision of this two-dimensional structural interpretation. With greater three-dimensional exposure (e.g., access to drilling results), an attempt at a three-dimensional structural restoration might be justified.

CONCLUSIONS

Complexly faulted Ordovician through Permian miogeoclinal sedimentary rocks, Cenozoic volcanic and sedimentary rocks, and sparse intrusive rocks are exposed at Spruce Mountain. Skarn, carbonate replacements, fissure veins, and jasperoid that were exploited by small underground mines in the district, principally for lead and silver, originally were the distal expressions of a por-phyry molybdenum system at depth, although the system is now tilted and dismembered. The porphyry system, here dated by U-Pb zircon as late Eocene (ca. 38 Ma), is related to sparsely exposed rhyolite porphyry and granite por-phyry intrusions. Spruce Mountain is a rhyolitic porphyry Mo or Climax-type porphyry molybdenum deposit with extensive associated skarn. The district also contains a Carlin-type gold prospect of unknown age.

Based on a plan-view analysis of geologic maps and a reconstructed cross section, at least six crosscutting sets of normal faults have been delineated with contrasting polarities of faulting. From oldest to youngest, these sets have senses of displacement of (1) down to the west, (2) down to the west, (3) down to the north, (4) down to the east, (5) down to the east, and (6) down to the west. The present-day fault network reflects the overprinting of multiple phases of extension, ranging in age from late Eocene or older to Quaternary, in which periods of incremental eastward tilting were interspersed with periods

Page 21: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

257Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

of westward tilting. Based on age constraints from six new U-Pb age dates of igneous dikes within the district, some of which cut and intrude the oldest fault sets, the earliest phase of extension occurred at or before ca. 38 Ma. Early extension was accommodated by the oldest four fault sets, all of which are late Eocene or older in age, and resulted in the largest extensional strain in the district (5.4 km). Probable middle to late Miocene sedimentary rocks ex-posed in the southeastern part of the district were tilted during a later phase of normal faulting, which (along with minor additional extension during the Quaternary) resulted in ~1.5 km of additional extension in the central part of the district. Thus, we estimate that Cenozoic normal faulting accommodated a total of 6.9 km of extension (120%) across the central portion of the Spruce Mountain district. Although the oldest normal faults that crop out at Spruce Mountain currently dip <15°, based on the structural reconstruction, all faults in the reconstructed section had initial dips of ~45° or greater. The recognition of large-magnitude late Eocene or older extension at Spruce Mountain is im-portant in that Paleogene extension in this region of northeastern Nevada has previously been regarded by many as relatively insignificant compared with later phases of extension during the middle Miocene. Our restorations and age dates at Spruce Mountain show that Eocene extension was at least locally significant in northeastern Nevada.

Within the central Spruce Mountain district, large extensional strains have been distributed among numerous normal faults. Stepwise, fault by fault res-toration of the normal faults at Spruce Mountain yields a restored section that implies west-vergent folding and gentle westward dips of miogeoclinal strata. This geometry is interpreted to reflect that the Spruce Mountain area was on the western limb of an anticline that connected to the Pequop syncline to the east prior to being dismembered by extensional faulting. The structural recon-struction indicates that the entire extensional history of the district produced net eastward tilting of ~35°, which suggests that movement on faults of the three sets of down-to-the-west faults of the first, second, and sixth generations dominated the tilting history of the area.

The earliest four sets of normal faults at Spruce Mountain are pre-ore or syn-ore faults relative to hydrothermal alteration-mineralization related to the porphyry molybdenum system. Faults of these sets are prospective as they probably were conduits for hydrothermal fluid flow that produced skarn and carbonate replacement mineralization. However, the Spruce Mountain system has also been structurally dismembered by younger sets of normal faults. Net tilting of the hydrothermal system by postmineral faults is probably ~10° west-ward, and deeper levels of the system are generally exposed on the western side of the district. Cross-sectional restoration of the postmineralization exten-sional faulting and associated tilting suggests that the original deposit had a ~6 km long by ~3 km deep elliptical pattern in 2 dimensions, with carbonate replacement deposits surrounding skarn. Spruce Mountain joins Bingham, Utah, Cove-McCoy, Nevada, several deposits in the Battle Mountain district, Nevada, and many sites in White Pine County, Nevada, as mineral deposits where late Eocene extension, regardless of magnitude, played a role in local-izing ore formation.

ACKNOWLEDGMENTS

This paper is a collaboration that incorporates results from two independently funded M.S. thesis projects. Pape’s study and associated field work at the Lowell Institute of Mineral Resources at the University of Arizona was financially supported by an Arthur A. Meyerhoff Memorial Grant from the American Association of Petroleum Geologists Foundation, a Hugh E. McKinstry Student Research Award from the Society of Economic Geologists, and Science Foundation Arizona. Baril’s study was supported by Renaissance Gold Inc. through the Ralph J. Roberts Center for Research in Economic Geology at the University of Nevada, Reno, with radiometric dating conducted at the Isotope Geology Laboratory at Boise State University. We gratefully acknowledge all sources of financial support, and analytical support from the staff at the Isotope Geology Laboratory at Boise State University (Idaho).

We thank Roy Johnson, Clay Postlethwaite, Phil Nickerson, and Chris Henry for helpful dis-cussions and critical comments, and Mark Pecha for assistance with interpretation of the U-Pb results. We also thank William Pape, who assisted in field mapping; Mark Barton, Joe Colgan, Pete DeCelles, Dave John, Allen McGrew, Julie Rowland, and an anonymous reviewer for comments on earlier versions of the manuscript; Eric Struhsacker of Renaissance Gold Inc. for leads regard-ing recent exploration at Spruce Mountain; and Phil Gans for discussions of the geology of Spruce Mountain with Seedorff decades ago. We also acknowledge a helpful review of an earlier version of the manuscript by Dan Pace of Renaissance Gold Inc. that led to the collaboration.

APPENDIX 1. DESCRIPTIONS OF SAMPLES DATED BY U-Pb ZIRCON METHOD

SMIg44SMIg44 is an Eocene porphyritic rhyolite dike ~1.1 km northwest of the Standard mine; it is

50 cm wide and strikes north-south. The texture is porphyritic with 5% phenocrysts and distinct flow banding along strike. The phenocryst assemblage includes 50% plagioclase 0.1–1 mm, 25% resorbed quartz 1–4 mm, 20% potassium feldspar 1–2 mm, and 5% biotite 0.1–0.5 mm. Sericite and kaolinite have partially to completely replaced the feldspars.

SMIg45

SMIg45 is an Eocene rhyolite porphyry intrusion ~0.6 km southeast of the Standard mine and 0.4 km west of the Ada H mine, and occurs as a circular body of float ~100 m wide. The rock is white to light gray with a porphyritic texture and shows pervasive alteration effects, including bleached and clay-altered feldspars and abundant iron oxides that may represent oxidized pyrite. Phenocrysts represent 30% of the rock and consist of 40% partially resorbed quartz 1–3 mm, 40% potassium feldspar 2–10 mm, and 20% plagioclase 1–2 mm. The matrix is a very fine grained mixture of quartz and feldspars. Sericite and kaolinite have completely replaced the feldspars.

SMIg46

SMIg46 is a Jurassic dacite ~1.5 km northwest of the Kille mine and is a light gray compe-tent rock that forms large outcrops and exhibits distinct flow banding. The texture is fine-grained equigranular. The rock is composed of 70% plagioclase 1–2 mm and 30% quartz 0.5–2 mm; mafic minerals have been destroyed. Sericite alteration of plagioclase is pervasive. Calcite locally over-prints sericite alteration.

SMIg47

SMIg47 is a Jurassic porphyritic dacite ~1.2 km south of the Kille mine that occurs in an elon-gate body of pale green rock 200 m long by 100 m wide. The texture is porphyritic with 15% pheno-crysts in a fine-grained matrix. The phenocryst assemblage includes 60% plagioclase 0.5–2 mm, 20% quartz 0.5–2 mm, and 20% biotite 0.5–1 mm. The groundmass is composed of plagio clase, bio tite, and quartz. This rock has been strongly propylitized, with complete alteration of plagio-clase and biotite to chlorite and calcite. Plagioclase phenocrysts also display minor sericite and kaolinite alteration.

Page 22: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Rese

arch

Pape

r

258

Pap

e et

al.

| Str

uctu

ral r

econ

stru

ctio

n, S

pru

ce M

ount

ain,

Nev

ada

GE

OS

PH

ER

E |

Volu

me

12 |

Num

ber

1

3 0

3 2

3 4

3 6

3 8

4 0

4 2

4 4

0 .0 0 4 6

0 .0 0 5 0

0 .0 0 5 4

0 .0 0 5 8

0 .0 0 6 2

0 .0 0 6 6

0 .0 0 7 0

0 .0 0 0 .0 2 0 .0 4 0 .0 6 0 .0 8 0 .1 0

20

6P

b/2

38U

2 0 7Pb/2 3 5U

da ta -po in t e rro r e ll ipses a re 2 σ

Sample SMIg45

3 4

3 8

4 2

4 6

0 .0 0 4 8

0 .0 0 5 2

0 .0 0 5 6

0 .0 0 6 0

0 .0 0 6 4

0 .0 0 6 8

0 .0 0 7 2

0 .0 2 5 0 .0 3 5 0 .0 4 5 0 .0 5 5 0 .0 6 5

206 P

b/23

8 U

2 0 7Pb/2 3 5U

da ta -po in t e rro r e ll ipses a re 2 σ

Sample SMIg44

1 0 0 0

1 4 0 0

1 8 0 0

2 2 0 0

2 6 0 0

0 .0

0 .1

0 .2

0 .3

0 .4

0 .5

0 2 4 6 8 1 0 1 2 1 4

206 P

b/23

8 U

2 0 7Pb/2 3 5U

da ta -po in t e rro r e ll ipses a re 2 σ

Sample SMIg46

130

140

150

160

170

180

190

0 .0 2 0

0 .0 2 2

0 .0 2 4

0 .0 2 6

0 .0 2 8

0 .0 3 0

0 .1 2 0 .1 4 0 .1 6 0 .1 8 0 .2 0 0 .2 2 0 .2 4 0 .2 6

20

6P

b/2

38U

2 0 7Pb/2 3 5U

d ata-p o int erro r ellip s es are 2σ

Sample SMIg47

3 0

3 4

3 8

4 2

4 6

5 0

0 .0 0 4 5

0 .0 0 5 5

0 .0 0 6 5

0 .0 0 7 5

0 .0 2 0 .0 3 0 .0 4 0 .0 5 0 .0 6 0 .0 7 0 .0 8

20

6P

b/2

38U

2 07P b /235U

da ta -po in t e rro r e ll ipses a re 2 σ

Sample SMIg49

3 6

4 0

4 4

4 8

0 .0 0 5 0

0 .0 0 5 4

0 .0 0 5 8

0 .0 0 6 2

0 .0 0 6 6

0 .0 0 7 0

0 .0 0 7 4

0 .0 2 5 0 .0 3 5 0 .0 4 5 0 .0 5 5 0 .0 6 5 0 .0 7 5

206 P

b/23

8 U

2 0 7Pb/2 3 5U

da ta -po in t e rro r e ll ipses a re 2 σ

Sample SMIg51

Figure A1. Concordia diagrams for U-Pb dates. See Appendix 2 for interpretations.

Page 23: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

259Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

SMIg49

SMIg49 is an Eocene coarse-grained rhyolite porphyry ~0.5 km southeast of the Black Forest mine; it is light gray and contains 30% phenocrysts set in a fine-grained matrix of quartz and feldspar. Phenocrysts are 20% resorbed quartz 1–6 mm, 50% potassium feldspar 3–9 mm, 20% plagioclase 1–3 mm, and 10% biotite 0.5–1.5 mm. Very minor disseminated pyrite and chalcopyrite are present. Quartz veins 1–20 mm wide are common, some of which contain fine-grained molyb-denite. The rock has been potassically altered with pervasive K-feldspar.

SMIg51

The second rhyolite porphyry unit occurs along a roughly linear trend across the entire study area that may be a large dike; the dated rock was collected on the south side of Banner Hill ~1.3 km east of the Black Forest mine. Exposures are commonly either masses of float or highly weath-ered rock exposed by prospect pits. The rock is bright white with a porphyritic texture. Parallel quartz veins 1–3 mm wide ~10/m are present in outcrops on the western edge of the study area. Phenocrysts make up 15% of the rock and consist of 60% quartz 1–3 mm, 20% potassium feldspar 1–2 mm, and 20% plagioclase 1–2 mm. Alteration is typically weak argillization, although one out-crop on the eastern edge of the study area contains potassium feldspar flooding, 1–2 mm fluorite veins, and minor copper oxides.

APPENDIX 2. LASER ABLATION–INDUCTIVELY COUPLED PLASMA–MASS SPECTROMETRY U-Pb ZIRCON GEOCHRONOLOGIC METHODS

Sample Preparation

Zircon grains were separated from rocks using standard techniques and mounted in epoxy and polished until the centers of the grains were exposed. Cathodoluminescence (CL) images were obtained with a JEOL JSM-1300 scanning electron microscope and Gatan MiniCL. CL images guided the placement sites for microanalysis. U-Pb dates were obtained from spot analyses by laser ablation–inductively coupled plasma–mass spectrometry (LA-ICP-MS) at the Isotope Geology Laboratory at Boise State University (Idaho).

Analytical Methods and Reporting Procedures

U-Pb isotope systematics and trace element compositions were analyzed by LA-ICP-MS using a ThermoElectron X-Series II quadrupole ICP-MS and New Wave Research UP-213 Nd:YAG UV (213 nm) laser ablation system. In-house analytical protocols, standard materials, and data reduction software were used for simultaneous acquisition and real-time calibration of U-Pb dates and a suite of high field strength elements (HFSE) and rare earth elements (REE) using the high sensitiv-ity and unique properties of the interface (Xs cones), extraction lens, and quadrupole analyzer of the X-Series II. Zircons were ablated with a laser diameter of 25 mm using fluence and pulse rates of 5 J/cm2 and 10 Hz, respectively, during a 30 s analysis (15 s gas blank, 45 s ablation) that exca-vated a pit ~25 µm deep. Ablated material was carried by a 1.2 L/min He gas stream to the nebulizer flow of the plasma. Dwell times were 5 ms for Si and Zr, 100 ms for 49Ti and 207Pb, 40 ms for 238U, 232Th, 202Hg, 204Pb, 206Pb and 208Pb isotopes, and 10 ms for all other HFSEs and REEs. Background count rates for each analyte were obtained prior to each spot analysis and subtracted from the raw count rate for each analyte.

For U-Pb dates, instrumental fractionation of the background-subtracted 206Pb/238U and 207Pb/206Pb ratios was corrected, and dates were calibrated with respect to interspersed measure-ments of the Plešovice zircon standard (Sláma et al., 2008). Signals at mass 204 were indistin-guishable from zero following subtraction of mercury backgrounds measured during the gas blank (<1000 cps 202Hg), and thus dates are reported without common Pb correction. Radiogenic isotope ratio and age error propagation for each spot includes uncertainty contributions from counting statistics and background subtraction. For concentration calculations, background-subtracted count rates for each analyte were internally normalized to 29Si and calibrated with respect to NIST SRM-610 and SRM-612 glasses as the primary standards.

Errors on weighted mean 206Pb/238U dates given in the following are presented at 2s as follows: weighted mean date ± x/y, where x is the internal error and y is the error including the uncer-tainty on the standard calibration, propagated in quadrature. Weighted mean calculations were performed using Isoplot 3.0 (Ludwig, 2003). A standard calibration uncertainty of 1.0% (2s) is used because that is the average of these experiments. In Table 1 errors on single dates do not include uncertainties on the standard calibration.

We experimented with other schemes to reject certain individual spot analyses. Although these routines in some cases resulted in slightly different dates, all dates were well within the uncertainties reported here.

Zircon standard AUSZ2 was measured as an unknown during the experiment as a quality control standard; 23 analyses of AUSZ2 yielded a weighted mean 206Pb/238U date of 38 ± 0.5 Ma, in agreement with the chemical abrasion–thermal ionization mass spectrometry weighted mean date of 38.86 ± 0.01 Ma (2s internal error, J. Crowley, 2012, personal commun.).

Results

Analytical results are summarized in Table 1; Figure A1 shows concordia diagrams for the six analyzed samples.

Sample SMIg44 has an interpreted age of crystallization of 37.7 ± 0.5 Ma and exhibits minor normal discordance, reflecting several spot analyses that exhibit lead loss.

Sample SMIg45 has an interpreted age of crystallization of 37.3 ± 0.4 Ma and exhibits normal and reverse discordance, probably reflecting lead loss for some of the analyzed spots.

Sample SMIg46 has an interpreted Jurassic age of crystallization of 155 ± 4 Ma and has sev-eral spot analyses that exhibit lead loss, but it also shows the greatest evidence of inheritance, likely with several different Precambrian age cores with ages of ca. 1.1 Ga and ca. 2.7 Ga.

Sample SMIg47 has an interpreted Jurassic age of crystallization of 156 ± 2 Ma and exhibits minor discordance, reflecting several spot analyses that exhibit lead loss.

Sample SMIg49 has an interpreted age of crystallization of 39.1 ± 0.6 Ma and exhibits normal and reverse discordance, reflecting lead loss for some of the analyzed spots.

Sample SMIg51 has an interpreted age of crystallization of 37.5 ± 0.4 Ma and exhibits discor-dance that likely results from lead loss for some of the analyzed spots.

REFERENCES CITED

Allmendinger, R.W., Hauge, T.A., Hauser, E.C., Potter, C.J., Klemperer, S.L., Nelson, K.D., Knuep-fer, P., and Oliver, J.E., 1987, Overview of the COCORP 40°N transect, western United States: The fabric of an orogenic belt: Geological Society of America Bulletin, v. 98, p. 308–319, doi: 10 .1130 /0016 -7606 (1987)98 <308: OOTCNT>2 .0 .CO;2 .

American CuMo Mining Corporation, 2014, Projects—Spruce Mountain: Vancouver, British Co-lumbia, American CuMo Mining Corporation, http:// cumoco .com /projects /spruce -mountain/ (27 January 2014).

Atkinson, W.W., Jr., Kaczmarowski, J.H., and Erickson, A.J., Jr., 1982, Geology of a skarn-breccia orebody at the Victoria mine, Elko County, Nevada: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 77, p. 899–918, doi: 10 .2113 /gsecongeo .77 .4 .899 .

Axen, G.J., 1986, Superposed normal faults in the Ely Springs Range, Nevada: Estimates of extension: Journal of Structural Geology, v. 8, p. 711–713, doi: 10 .1016 /0191 -8141 (86)90076 -3 .

Baril, T.C., 2013, Igneous geochemistry and geochronology of the Spruce Mountain district, Elko County, Nevada [M.S. thesis]: Reno, University of Nevada, Reno, 89 p.

Barton, M.D., 1996, Granitic magmatism and metallogeny of southwestern North America, in Brown, M., et al., eds., The Third Hutton Symposium on the origin of granites and related rocks: Geological Society of America Special Paper 315, p. 261–280, doi: 10 .1130 /0 -8137 -2315 -9 .261 .

Barton, M.D., Girardi, J.D., Kreiner, D.C., Seedorff, E., Zurcher, L., Dilles, J.H., Haxel, G.B., and Johnson, D.A., 2011, Jurassic igneous-related metallogeny of southwestern North America, in Steininger, R.C., and Pennell, W.M., eds., Great Basin evolution and metallogeny: Pro-ceedings, Geological Society of Nevada Symposium, v. 1, p. 373–396.

Bedell, R., Struhsacker, E.M., Craig, L., Miller, M.S., Coolbaugh, M.F., Smith, J.L., and Parratt, R.L., 2010, The Pequop mining district, Elko County, Nevada: An evolving new gold dis-trict, in Goldfarb, R.J., et al., eds., The challenge of finding new mineral resources: Global metal logeny, innovative exploration, and new discoveries: Society of Economic Geologists Special Publication 15, p. 29–56.

Page 24: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

260Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Begbie, M.J., Spörli, K.B., and Mauk, J.L., 2007, Structural evolution of the Golden Cross epi-thermal Au-Ag deposit, New Zealand: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 102, p. 873–892, doi: 10 .2113 /gsecongeo .102 .5 .873 .

Brady, R.J., Wernicke, B., and Fryxell, J., 2000, Kinematic evolution of a large-offset continental normal fault system, South Virgin Mountains, Nevada: Geological Society of America Bulle-tin, v. 112, p. 1375–1397, doi: 10 .1130 /0016 -7606 (2000)112 <1375: KEOALO>2 .0 .CO;2 .

Brooks, W.E., Thorman, C.H., and Snee, L.W., 1995, The 40Ar/39Ar ages and tectonic setting of the middle Eocene northeast Nevada volcanic field: Journal of Geophysical Research, v. 100, p. 10,403–10,416, doi: 10 .1029 /94JB03389 .

Camilleri, P.A., 2010, Geologic map of the northern Pequop Mountains, Elko County, Nevada: Nevada Bureau of Mines and Geology Map 171, scale 1:48,000, 15 p.

Camilleri, P.A., and Chamberlain, K.R., 1997, Mesozoic tectonics and metamorphism in the Pequop Mountains and Wood Hills region, northeast Nevada: Implications for the architecture and evolution of the Sevier orogen: Geological Society of America Bulletin, v. 109, p. 74–94, doi: 10 .1130 /0016 -7606 (1997)109 <0074: MTAMIT>2 .3 .CO;2 .

Carpenter, J.A., Carpenter, D.G., and Dobbs, S.W., 1993, Structural analysis of the Pine Valley area, Nevada, in Gillespie, C.W., ed., Structural and stratigraphic relationships of Devonian reservoir rocks, east central Nevada: Nevada Petroleum Society 1993 Field Conference Guidebook, p. 9–49.

Carten, R.B., Geraghty, E.P., Walker, B.M., and Shannon, J.R., 1988, Cyclic development of igneous features and their relationship to high-temperature hydrothermal features in the Hender son porphyry molybdenum deposit, Colorado: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 83, p. 266–296, doi: 10 .2113 /gsecongeo .83 .2 .266 .

Carten, R.B., White, W.H., and Stein, H.J., 1993, High-grade granite-related molybdenum sys-tems: Classification and origin, in Kirkham, R.V., et al., eds., Mineral deposit modeling: Geo-logical Association of Canada Special Paper 40, p. 521–554.

Cashman, P.H., Trexler, J.H., Jr., Davydov, V.I., and Snyder, W.S., 2011, Tectonostratigraphy of the Great Basin from the Antler orogeny through Pennsylvanian time, in Steininger, R.C., and Pennell, W.M., eds., Great Basin evolution and metallogeny: Proceedings, Geological Society of Nevada Symposium, v. 1, p. 299–311.

Caskey, S.J., Wesnousky, S.G., Zhang, P., and Slemmons, D.B., 1996, Surface faulting of the 1954 Fairview Peak (Ms 7.2) and Dixie Valley (Ms 6.8) earthquakes, central Nevada: Seismological Society of America Bulletin, v. 86, p. 761–787.

Coats, R.R., 1987, Geology of Elko County, Nevada: Nevada Bureau of Mines and Geology Bulle-tin 101, map scale 1:250,000, 112 p.

Colgan, J.P., and Henry, C.D., 2009, Rapid Miocene collapse of the Mesozoic orogenic plateau in north-central Nevada: International Geology Review, v. 51, p. 920–961, doi: 10 .1080 /00206810903056731 .

Colgan, J.P., John, D.A., Henry, C.D., and Fleck, R.J., 2008, Large-magnitude Miocene extension of the Eocene Caetano caldera, Shoshone and Toiyabe Ranges, Nevada: Geosphere, v. 4, p. 107–130, doi: 10 .1130 /GES00115 .1 .

Colgan, J.P., Howard, K.A., Fleck, R.J., and Wooden, J.L., 2010, Rapid middle Miocene extension and unroofing of the southern Ruby Mountains, Nevada: Tectonics, v. 29, TC6022, doi: 10 .1029 /2009TC002655 .

Coolbaugh, M.F., Vikre, P.G., and Faulds, J.E., 2011, Young (≤7 Ma) gold deposits and active geothermal systems of the Great Basin: Enigmas, questions, and exploration potential, in Steininger, R.C., and Pennell, W.M., eds., Great Basin evolution and metallogeny: Proceed-ings, Geological Society of Nevada Symposium, v. 2, p. 845–859.

Crittenden, M.D., Jr., Coney, P.J., and Davis, G.H., eds., 1980, Cordilleran metamorphic core com-plexes: Geological Society of America Memoir 153, 490 p., doi: 10 .1130 /MEM153 .

Dahlstrom, C.D.A., 1969, Balanced cross sections: Canadian Journal of Earth Sciences, v. 6, p. 743–757, doi: 10 .1139 /e69 -069 .

Davison, I., 1989, Extensional domino fault tectonics: Kinematics and geometrical constraints: Annales Tectonicae, v. 3, p. 12–24.

dePolo, C.M., 2008, Quaternary faults in Nevada: Nevada Bureau of Mines and Geology Map 167, scale 1:1,000,000.

Dickinson, W.R., 2002, The Basin and Range province as a composite extensional domain: Inter-national Geology Review, v. 44, p. 1–38, doi: 10 .2747 /0020 -6814 .44 .1 .1 .

Dickinson, W.R., 2006, Geotectonic evolution of the Great Basin: Geosphere, v. 2, p. 353–368, doi: 10 .1130 /GES00054 .1 .

Dickinson, W.R., 2011, The place of the Great Basin in the Cordilleran orogen, in Steininger, R.C., and Pennell, W.M., eds., Great Basin evolution and metallogeny: Proceedings, Geological Society of Nevada Symposium, v. 1, p. 419–436.

Dilles, J.H., and Einaudi, M.T., 1992, Wall-rock alteration and hydrothermal flow paths about the Ann-Mason porphyry copper deposit, Nevada—A 6-km vertical reconstruction: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 87, p. 1963–2001, doi: 10 .2113 /gsecongeo .87 .8 .1963 .

Dilles, J.H., and Gans, P.B., 1995, The chronology of Cenozoic volcanism and deformation in the Yerington area, western Basin and Range and Walker Lane: Geological Society of America Bulletin, v. 107, p. 474–486, doi: 10 .1130 /0016 -7606 (1995)107 <0474: TCOCVA>2 .3 .CO;2 .

Dilles, J.H., and Proffett, J.M., 1995, Metallogenesis of the Yerington batholith, Nevada, in Pierce, F.W., and Bolm, J.G., eds., Porphyry copper deposits of the American Cordillera: Arizona Geological Society Digest Volume 20, p. 306–315.

Dohrenwend, J.C., Schell, B.A., and Moring, B.C., 1991, Reconnaissance photogeologic map of young faults in the Elko 1° by 2° quadrangle, Nevada and Utah: U.S. Geological Survey Miscellaneous Field Studies Map MF-2179, scale 1:250,000.

Dohrenwend, J.C., Schell, B.A., Menges, C.M., Moring, B.C., and McKittrick, M.A., 1996, Recon-naissance photogeologic map of young (Quaternary and late Tertiary) faults, in Singer, D.A., ed., An analysis of Nevada’s metal-bearing mineral resources: Nevada Bureau of Mines and Geology Open-File Report 96-2, p. 9-1–9-12, scale 1:1,000,000.

Dreier, J.E., 1984, Regional tectonic control of epithermal veins in the western United States and Mexico, in Wilkins, J., Jr., ed., Gold and silver deposits of the Basin and Range province, western U.S.A.: Arizona Geological Society Digest Volume 15, p. 28–50.

Druschke, P.A., Hanson, A.D., and Wells, M.L., 2009, Structural, stratigraphic, and geochronologic evidence for extension predating Palaeogene volcanism in the Sevier hinterland, east-central Nevada: International Geology Review, v. 51, p. 743–775, doi: 10 .1080 /00206810902917941 .

Faulds, J.E., and Varga, R., 1998, The role of accommodation zones and transfer zones in the regional segmentation of extended terranes, in Faulds, J.E., and Stewart, J.H., eds., Accom-modation zones and transfer zones; the regional segmentation of the Basin and Range Prov-ince: Geological Society of America Special Paper 323, p. 1–45, doi: 10 .1130 /0 -8137 -2323 -X .1 .

Felder, R.P., Struhsacker, E.M., and Miller, M.S., 2011, The history of exploration and discovery of the Long Canyon gold deposit, Elko County, Nevada, USA, in Steininger, R.C., and Pennell, W.M., eds., Great Basin evolution and metallogeny: Proceedings, Geological Society of Ne-vada Symposium, v. 1, p. 141–152.

Fraser, G.D., Ketner, K.B., and Smith, M.C., 1986, Geologic map of the Spruce Mountain 4 quad-rangle, Elko County, Nevada: U.S. Geological Survey Miscellaneous Field Studies Map MF-1846, scale 1:24,000.

Gans, P.B., 1982, Multiple intrusion and faulting in the Hunter district, White Pine County, Nevada [M.S. thesis]: Stanford, California, Stanford University, 179 p.

Gans, P.B., 1987, An open-system, two-layer crustal stretching model for the eastern Great Basin: Tectonics, v. 6, p. 1–12, doi: 10 .1029 /TC006i001p00001 .

Gans, P.B., and Miller, E.L., 1983, Style of mid-Tertiary extension in east-central Nevada, in Nash, W.P., and Gurgel, K.D., eds., Geological excursions in the overthrust belt and metamorphic core complexes of the Intermountain region: Utah Geological and Mineral Survey Special Studies 59, p. 107–139.

Gans, P.B., Miller, E.L., McCarthy, J., and Ouldcott, M.L., 1985, Tertiary extensional faulting and evolving ductile-brittle transition zones in the northern Snake Range and vicinity: New in-sights from seismic data: Geology, v. 13, p. 189–193, doi: 10 .1130 /0091 -7613 (1985)13 <189: TEFAED>2 .0 .CO;2 .

Gans, P.B., Mahood, G.A., and Schermer, E., 1989, Synextensional magmatism in the Basin and Range province; a case study from the eastern Great Basin: Geological Society America Special Paper 233, 53 p., doi: 10 .1130 /SPE233 .

Gans, P.B., Seedorff, E., Fahey, P.L., Hasler, R.W., Maher, D.J., Jeanne, R.A., and Shaver, S.A., 2001, Rapid Eocene extension in the Robinson district, White Pine County, Nevada: Geology, v. 29, p. 475–478, doi: 10 .1130 /0091 -7613 (2001)029 <0475: REEITR>2 .0 .CO;2 .

Granger, A.E., Bell, M.M., Simmons, G.C., and Lee, F., 1957, Geology and mineral resources of Elko County, Nevada: Nevada Bureau of Mines Bulletin 54, 190 p.

Hallett, B.W., and Spear, F.S., 2014, The P-T history of anatectic pelites of the northern East Hum-boldt Range, Nevada: Evidence for tectonic loading, decompression, and anatexis: Journal of Petrology, v. 55, p. 3–36, doi: 10 .1093 /petrology /egt057 .

Harlow, G.R., 1956, The stratigraphy and structure of the Spruce Mountain area, Elko County, Nevada [M.S. thesis]: Seattle, University of Washington, 72 p.

Hastings, M.H., 2008, Relationship of base-metal skarn mineralization to Carlin-type gold min-eralization at the Archimedes gold deposit, Eureka, Nevada [M.S. thesis]: Reno, University of Nevada, Reno, 109 p.

Page 25: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

261Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Henry, C.D., 2008, Ash-flow tuffs and paleovalleys in northeastern Nevada: Implications for Eo-cene paleogeography and extension in the Sevier hinterland, northern Great Basin: Geo-sphere, v. 4, p. 1–35, doi: 10 .1130 /GES00122 .1 .

Hill, J.M., 1916, Notes on some mining districts in eastern Nevada: U.S. Geological Survey Bul-letin 648, 214 p.

Hodges, K.V., Snoke, A.W., and Hurlow, H.A., 1992, Thermal evolution of a portion of the Sevier hinterland: The northern Ruby Mountains–East Humboldt Range and Wood Hills, northeast-ern Nevada: Tectonics, v. 11, p. 154–164, doi: 10 .1029 /91TC01879 .

Hodgkinson, K.M., Stein, R.S., and King, G.C.P., 1996, The 1954 Rainbow Mountain–Fairview Peak–Dixie Valley earthquakes: A triggered normal faulting sequence: Journal of Geophysi-cal Research, v. 101, p. 25,459–25,471, doi: 10 .1029 /96JB01302 .

Hoge, A.K., Seedorff, E., Barton, M.D., Richardson, C.A., and Favorito, D.A., 2015, The Jack-son-Lawton-Bowman normal fault system and its relationship to Carlin-type gold miner-alization, Eureka district, Nevada, in Pennell, W.M., and Garside, L.J., eds., New concepts and discoveries: Proceedings, Geological Society of Nevada Symposium, v. 2, p. 967–1000.

Hope, R.A., 1972, Geologic map of the Spruce Mountain quadrangle, Elko County, Nevada: U.S. Geological Survey Quadrangle Map GQ-942, scale 1:62,500, 3 p.

Hossack, J.R., 1979, The use of balanced cross-sections in the calculation of orogenic contrac-tion: A review: Journal of the Geological Society [London], v. 136, p. 705–711, doi: 10 .1144 /gsjgs .136 .6 .0705 .

Houston, R.A., and Dilles, J.H., 2013, Structural geologic evolution of the Butte district, Montana: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 108, p. 1397–1424, doi: 10 .2113 /econgeo .108 .6 .1397 .

Howard, K.A., 1980, Metamorphic infrastructure in the northern Ruby Mountains, Nevada, in Crittenden, M.D., Jr., et al., eds., Cordilleran metamorphic core complexes: Geological Soci-ety of America Memoir 153, p. 335–347, doi: 10 .1130 /MEM153 -p335 .

Howard, K.A., 2003, Crustal structure in the Elko-Carlin region, Nevada, during Eocene gold min-eralization: Ruby–East Humboldt metamorphic core complex as a guide to the deep crust: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 98, p. 249–268, doi: 10 .2113 /gsecongeo .98 .2 .249 .

Hudec, M.R., 1992, Mesozoic structural and metamorphic history of the central Ruby Moun-tains metamorphic core complex, Nevada: Geological Society of America Bulletin, v. 104, p. 1086–1100, doi: 10 .1130 /0016 -7606 (1992)104 <1086: MSAMHO>2 .3 .CO;2 .

Ilchik, R.P., and Barton, M.D., 1997, An amagmatic origin of Carlin-type gold deposits: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 92, p. 269–288, doi: 10 .2113 /gsecongeo .92 .3 .269 .

John, D.A., 2001, Miocene and early Pliocene epithermal gold-silver deposits in the northern Great Basin, western United States: Characteristics, distribution, and relationship to mag-matism: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 96, p. 1827–1853.

Johnson, L.C., 1983, The Ellison district: Alteration-mineralization associated with a mid-Tertiary intrusive complex at Sawmill Canyon, White Pine County, Nevada [M.S. thesis]: Tucson, University of Arizona, 123 p.

Johnston, M.K., Thompson, T.B., Emmons, D.L., and Jones, K., 2008, Geology of the Cove mine, Lander County, Nevada, and a genetic model for the McCoy-Cove hydrothermal system: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 103, p. 759–782, doi: 10 .2113 /gsecongeo .103 .4 .759 .

Keeler, D.A., and Seedorff, E., 2007, Structural dismemberment of Copper Basin, Battle Mountain mining district, north-central Nevada [abs.], in Ores and orogenesis: A symposium honor-ing the career of William R. Dickinson, Program with Abstracts: Tucson, Arizona Geological Society, p. 168.

LaPointe, D.D., Tingley, J.V., and Jones, R.B., 1991, Mineral resources of Elko County, Nevada: Nevada Bureau of Mines and Geology Bulletin 106, 236 p.

Lavier, L.L., and Buck, W.R., 2002, Half graben versus large-offset low-angle normal fault: Im-portance of keeping cool during normal faulting: Journal of Geophysical Research, v. 107, no. B6, doi: 10 .1029 /2001JB000513 .

Long, S.P., 2012, Magnitudes and spatial patterns of erosional exhumation in the Sevier hinter-land, eastern Nevada and western Utah, USA: Insights from a Paleogene paleogeologic map: Geosphere, v. 8, p. 881–901, doi: 10 .1130 /GES00783 .1 .

Long, S.P., 2015, An upper-crustal fold province in the hinterland of the Sevier orogenic belt, eastern Nevada, U.S.A.: A Cordilleran Valley and Ridge in the Basin and Range: Geosphere, v. 11, p. 404–424, doi: 10 .1130 /GES01102 .1 .

Lowell, J.D., 1968, Geology of the Kalamazoo orebody, San Manuel district, Arizona: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 63, p. 645–654, doi: 10 .2113 /gsecongeo .63 .6 .645 .

Lubben, J.D., Cline, J.S., and Barker, S.L.L., 2012, Ore fluid properties and sources from quartz-asso ciated gold at the Betze-Post Carlin-type gold deposit, Nevada, United States: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 107, p. 1351–1385, doi: 10 .2113 /econgeo .107 .7 .1351 .

Ludington, S., and Plumlee, G.S., 2009, Climax-type porphyry molybdenum deposits: U.S. Geo-logical Survey Open-File Report 2009-1215, 16 p.

Ludwig, K.R., 2003, User’s manual for Isoplot/Ex, rev. 3.00: A geochronological toolkit for Micro-soft Excel: Berkeley, California, Berkeley Geochronology Center Special Publication 4, 70 p.

MacCready, T., Snoke, A.W., Wright, J.E., and Howard, K.A., 1997, Mid-crustal flow during Tertiary extension in the Ruby Mountains core complex, Nevada: Geological Society of America Bul-letin, v. 109, p. 1576–1594, doi: 10 .1130 /0016 -7606 (1997)109 <1576: MCFDTE>2 .3 .CO;2 .

Maher, D.J., 2008, Reconstruction of middle Tertiary extension and Laramide porphyry copper systems, east-central Arizona [Ph.D. thesis]: Tucson, University of Arizona, 328 p.

McGrew, A.J., 1993, The origin and evolution of the southern Snake Range décollement, east central Nevada: Tectonics, v. 12, p. 21–34, doi: 10 .1029 /92TC01713 .

McGrew, A.J., and Snee, L.W., 1994, 40Ar/39Ar thermochronologic constraints on the tectono thermal evolution of the northern East Humboldt Range metamorphic core complex, Nevada: Tectono-physics, v. 238, p. 425–450, doi: 10 .1016 /0040 -1951 (94)90067 -1 .

McGrew, A.J., Peters, M.T., and Wright, J.E., 2000, Thermobarometric constraints on the tec-tonothermal evolution of the East Humboldt Range metamorphic core complex, Nevada: Geological Society of America Bulletin, v. 112, p. 45–60, doi: 10 .1130 /0016 -7606 (2000)112 <45: TCOTTE>2 .0 .CO;2 .

Miller, D.M., and Hoisch, T.D., 1995, Jurassic tectonics of northeastern Nevada and northwest-ern Utah from the perspective of barometric studies, in Miller, D.M., and Busby, C., eds., Jurassic magmatism and tectonics of the North American Cordillera: Geological Society of America Special Paper 299, p. 267–294, doi: 10 .1130 /SPE299 -p .267 .

Miller, E.L., Gans, P.B., Wright, J.E., and Sutter, J.F., 1988, Metamorphic history of the east-central Basin and Range province: Tectonic setting and relationship to magmatism, in Ernst, W. G., ed., Metamorphism and crustal evolution of the western United States (Rubey Volume VII): Englewood Cliffs, New Jersey, Prentice Hall, p. 649–682.

Miller, E.L., Dumitru, T.A., Brown, R.W., and Gans, P.B., 1999, Rapid Miocene slip on the Snake Range–Deep Creek Range fault system, east-central Nevada: Geological Society of America Bulletin, v. 111, p. 886–905, doi: 10 .1130 /0016 -7606 (1999)111 <0886: RMSOTS>2 .3 .CO;2 .

Mitra, G., 1992, Balanced structural interpretations in fold and thrust belts, in Mitra, G., and Fisher, G.W., eds., Structural geology of fold and thrust belts: Baltimore, Maryland, Johns Hopkins University Press, p. 53–77.

Nickerson, P.A., Barton, M.D., and Seedorff, E., 2010, Characterization and reconstruction of multi-ple copper-bearing hydrothermal systems in the Tea Cup porphyry system, Pinal County, Ari zona, in Goldfarb, R.J., et al., eds., The challenge of finding new mineral resources: Global metallogeny, innovative exploration, and new discoveries: Society of Economic Geol o gists Special Publication 15, p. 299–316.

Pape, J.R., 2010, Reconstructions of Cenozoic extensional faulting within the southern Ruby Mountains metamorphic core complex and adjacent areas, northeastern Nevada [M.S. thesis]: Tucson, University of Arizona, 99 p.

Proffett, J.M., Jr., 1977, Cenozoic geology of the Yerington district, Nevada, and implications for the nature and origin of Basin and Range faulting: Geological Society of America Bulletin, v. 88, p. 247–266, doi: 10 .1130 /0016 -7606 (1977)88 <247: CGOTYD>2 .0 .CO;2 .

Proffett, J.M., Jr., and Dilles, J.H., 1984, Geologic map of the Yerington district, Nevada: Nevada Bureau of Mines and Geology Map 77, scale 1:24,000.

Rahl, J.M., McGrew, A.J., and Foland, K.A., 2002, Transition from contraction to extension in the northeastern Basin and Range: New evidence from the Copper Mountains, Nevada: Journal of Geology, v. 110, p. 179–194, doi: 10 .1086 /338413 .

Redmond, P.B., and Einaudi, M.T., 2010, The Bingham Canyon porphyry Cu-Mo-Au deposit. I. Se-quence of intrusions, vein formation, and sulfide deposition: Economic Geology and the Bul-letin of the Society of Economic Geologists, v. 105, p. 43–68, doi: 10 .2113 /gsecongeo .105 .1 .43 .

Rhys, D.A., Valli, F., Burgess, R., Heitt, D.G., Greisel, G., and Hart, K., 2015, Controls of fault and fold geometry on the distribution of gold mineralization on the Carlin trend, in Pennell, W.M., and Garside, L.J., eds., New concepts and discoveries: Proceedings, Geological Soci-ety of Nevada Symposium, v. 1, p. 333–389.

Page 26: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

262Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Richardson, C.A., and Seedorff, E., 2015, Reconstruction of normal fault blocks in the Ann-Mason and Blue Hill areas, Yerington district, Lyon County, western Nevada, in Pennell, W.M., and Garside, L.J., eds., New concepts and discoveries: Proceedings, Geological Society of Ne-vada Symposium, v. 2, p. 1153–1178.

Roberts, A., and Yielding, G., 1994, Continental extensional tectonics, in Hancock, P. L., ed., Con-tinental deformation: Oxford, Pergamon Press, p. 223–250.

Saunders, F.T., and Bresnahan, J., 2010, Geologic map of Spruce Mountain property, Renaissance Exploration, Inc., internal company report, in Wolverson, N. J., ed., Technical report on the Spruce Mountain property, Elko County, Nevada, USA: Technical Report NI 43–101, prepared for Renaissance Gold, Inc., p. 36.

Schrader, F.C., 1931, Spruce Mountain district, Elko County, and Cherry Creek (Egan Canyon) district, White Pine County, Nevada: University of Nevada Bulletin, v. 25, no. 7, 39 p.

Seedorff, E., 1991a, Magmatism, extension, and ore deposits of Eocene to Holocene age in the Great Basin—Mutual effects and preliminary proposed genetic relationships, in Raines, G.L., et al., eds., Geology and ore deposits of the Great Basin: Proceedings, Geological Society of Nevada Symposium, v. 1, p. 133–178.

Seedorff, E., 1991b, Royston district, western Nevada: A Mesozoic porphyry copper system that was tilted and dismembered by Tertiary normal faults, in Raines, G.L., et al., eds., Geology and ore deposits of the Great Basin: Proceedings, Geological Society of Nevada Sympo-sium, v. 1, p. 359–391.

Seedorff, E., and Barton, M.D., 2004, Enigmatic origin of Carlin-type deposits: An amagmatic solution?: SEG Newsletter, no. 59, p. 14–18.

Seedorff, E., and Richardson, C.A., 2014, Diverse geometries and temporal relationships of nor-mal faults and fault systems in continental extension: A perspective from the Basin and Range province and mineral deposits [abs.], in Geometry and growth of normal faults: Geo-logical Society of London Conference, Programme and Abstract Volume, p. 20–22.

Seedorff, E., Dilles, J.H., Proffett, J.M., Jr., Einaudi, M.T., Zurcher, L., Stavast, W.J.A., Johnson, D.A., and Barton, M.D., 2005, Porphyry deposits: Characteristics and origin of hypogene features, in Hedenquist, J.W., et al., eds., Economic Geology 100th Anniversary Volume, p. 251–298.

Seedorff, E., Barton, M.D., Stavast, W.J.A., and Maher, D.J., 2008, Root zones of porphyry sys-tems: Extending the porphyry model to depth: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 103, p. 939–956, doi: 10 .2113 /gsecongeo .103 .5 .939 .

Seedorff, E., Richardson, C.A., and Maher, D.J., 2015, Fault surface maps: Three-dimensional structural reconstructions and their utility in exploration and mining, in Pennell, W.M., and Garside, L.J., eds., New concepts and discoveries: Proceedings, Geological Society of Ne-vada Symposium, v. 2, p. 1179–1206.

Shaver, S.A., and McWilliams, M., 1987, Cenozoic extension and tilting recorded in Upper Creta-ceous and Tertiary rocks at the Hall molybdenum deposit, northern San Antonio Mountains, Nevada: Geological Society of America Bulletin, v. 99, p. 341–353, doi: 10 .1130 /0016 -7606 (1987)99 <341: CEATRI>2 .0 .CO;2 .

Sláma, J., et al., 2008, Plešovice zircon—A new natural reference material for U-Pb and Hf iso-topic microanalysis: Chemical Geology, v. 249, p. 1–35, doi: 10 .1016 /j .chemgeo .2007 .11 .005 .

Slemmons, D.B., and Bell, J.W., 1987, 1954 Fairview Peak earthquake area, Nevada, in Hill, M.L., ed., Cordilleran Section of the Geological Society of America: Boulder, Colorado, Geological Society of America Centennial Field Guide, v. 1, p. 73–76.

Smith, D.L., Gans, P.B., and Miller, E.L., 1991, Palinspastic restoration of Cenozoic extension in the central and eastern Basin and Range province at latitude 39–40° N, in Raines, G.L., et al., eds., Geology and ore deposits of the Great Basin: Proceedings, Geological Society of Ne-vada Symposium, v. 1, p. 75–86.

Smith, M.T., Rhys, D.A., Ross, K., Lee, C., and Gray, J.N., 2013, The Long Canyon deposit: Anat-omy of a new off-trend sedimentary rock-hosted gold discovery in northeastern Nevada: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 108, p. 1119–1145, doi: 10 .2113 /econgeo .108 .5 .1119 .

Smith, R.M., 1976, Mineral resources of Elko County, Nevada: U.S. Geological Survey Open-File Report 76–56, 205 p.

Snoke, A.W., and Miller, D.M., 1988, Metamorphic and tectonic history of the northeastern Great Basin, in Ernst, W.G., ed., Metamorphism and crustal evolution of the western United States (Rubey Volume VII): Englewood Cliffs, New Jersey, Prentice Hall, p. 606–648.

Snyder, W.S., Spinosa, C., and Gallegos, D.M., 1991, Pennsylvanian–Permian tectonism on the western U.S. continental margin, in Raines, G.L., et al., eds., Geology and ore deposits of the Great Basin: Proceedings, Geological Society of Nevada Symposium, v. 1, p. 5–19.

Spencer, J.E., and Welty, J.W., 1986, Possible controls of base- and precious-metal mineralization associated with Tertiary detachment faults in the lower Colorado River trough, Arizona and California: Geology, v. 14, p. 195–198, doi: 10 .1130 /0091 -7613 (1986)14 <195: PCOBAP>2 .0 .CO;2 .

Stager, H.K., and Tingley, J.V., 1988, Tungsten deposits in Nevada: Nevada Bureau of Mines and Geology Bulletin 105, 256 p.

Stavast, W.J.A., Butler, R.F., Seedorff, E., Barton, M.D., and Ferguson, C.A., 2008, Tertiary tilting and dismemberment of the Laramide arc and related hydrothermal systems, Sierrita Moun-tains, Arizona: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 103, p. 629–636, doi: 10 .2113 /gsecongeo .103 .3 .629 .

Stewart, J.H., 1980, Geology of Nevada: Nevada Bureau of Mines and Geology Special Publica-tion 4, 136 p.

Stewart, J.H., and Carlson, J.E., 1976, Cenozoic rocks of Nevada: Nevada Bureau of Mines and Geology Map 52, scale 1:1,000,000, 5 p.

Stockli, D.F., Surpless, B.E., Dumitru, T.A., and Farley, K.A., 2002, Thermochronological con-straints on the timing and magnitude of Miocene and Pliocene extension in the central Was-suk Range, western Nevada: Tectonics, v. 21, doi: 10 .1029 /2001TC001295 .

Sullivan, W.A., and Snoke, A.W., 2007, Comparative anatomy of core-complex development in the northeastern Great Basin, U.S.A.: Rocky Mountain Geology, v. 42, p. 1–29, doi: 10 .2113 /gsrocky .42 .1 .1 .

Surpless, B.E., 2012, Cenozoic tectonic evolution of the central Wassuk Range, western Nevada, USA: International Geology Review, v. 54, p. 547–571, doi: 10 .1080 /00206814 .2010 .548117 .

Surpless, B.E., Stockli, D.F., Dumitru, T.A., and Miller, E.L., 2002, Two-phase westward encroach-ment of Basin and Range extension into the northern Sierra Nevada: Tectonics, v. 21, doi: 10 .1029 /2000TC001257 .

Swenson, R.F., 1991, Analysis of a fault-fold system in the southern Pequop Mountains, Elko County, Nevada [M.S. thesis]: Laramie, University of Wyoming, 131 p.

Thorman, C.H., 1970, Metamorphosed and nonmetamorphosed Paleozoic rocks in the Wood Hills and Pequop Mountains, northeast Nevada: Geological Society of America Bulletin, v. 81, p. 2417–2448, doi: 10 .1130 /0016 -7606 (1970)81 [2417: MANPRI]2 .0 .CO;2 .

Tingley, J.V., 1981a, Results of geochemical sampling within the Wells Resource Area, Elko County, Nevada (portions of the Wells and Elko 2° sheets): Nevada Bureau of Mines and Geology Open-File Report 81-3, 79 p.

Tingley, J.V., 1981b, Summary report, mineral inventory of the Wells Resource Area, Elko district, Elko County, Nevada: Nevada Bureau of Mines and Geology Open-File Report 81-4, 50 p.

Tosdal, R.M., and Richards, J.P., 2001, Magmatic and structural controls on the development of porphyry Cu ± Mo ± Au deposits: Reviews in Economic Geology, v. 14, p. 157–181.

Trexler, J.H., Jr., Cashman, P.H., Cole, J.C., Snyder, W.S., Tosdal, R.M., and Davydov, V.I., 2003, Widespread effects of middle Mississippian deformation in the Great Basin of western North America: Geological Society of America Bulletin, v. 115, p. 1278–1288, doi: 10 .1130 /B25176 .1 .

Trexler, J.H., Cashman, P.H., Snyder, W.S., and Davydov, V.I., 2004, Late Paleozoic tectonism in Nevada: Timing, kinematics, and tectonic significance: Geological Society of America Bulle-tin, v. 116, p. 525–538, doi: 10 .1130 /B25295 .1 .

Wallace, S.R., Muncaster, N.K., Jonson, D.C., MacKenzie, W.B., Bookstrom, A.A., and Surface, V.E., 1968, Multiple intrusion and mineralization at Climax, Colorado, in Ridge, J.D., ed., Ore deposits of the United States, 1933–1967 (Graton-Sales Volume): New York, American Institute Mining Metallurgy and Petroleum Engineers, p. 605–640.

Wernicke, B.P., 1981, Low-angle normal faults in the Basin and Range province: Nappe tectonics in an extending orogen: Nature, v. 291, p. 645–648, doi: 10 .1038 /291645a0 .

Wernicke, B.P., and Axen, G.J., 1988, On the role of isostasy in the evolution of normal fault systems: Geology, v. 16, p. 848–851, doi: 10 .1130 /0091 -7613 (1988)016 <0848: OTROII>2 .3 .CO;2 .

Wernicke, B.P., and Burchfiel, B.C., 1982, Modes of extensional tectonics: Journal of Structural Geology, v. 4, p. 105–115, doi: 10 .1016 /0191 -8141 (82)90021 -9 .

Wernicke, B.P., Christiansen, R.L., England, P.C., and Sonder, L.J., 1987, Tectonomagmatic evo-lution of Cenozoic extension in the North American Cordillera, in Coward, M.P., et al., eds., Continental extensional tectonics: Geological Society, London, Special Publication 28, p. 203–221, doi: 10 .1144 /GSL .SP .1987 .028 .01 .15 .

Westra, G., and Keith, S.B., 1981, Classification and genesis of stockwork molybdenum deposits: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 76, p. 844–873, doi: 10 .2113 /gsecongeo .76 .4 .844 .

Westra, G., and Riedell, K.B., 1996, Geology of the Mount Hope stockwork molybdenum deposit, Eureka County, Nevada, in Coyner, A.R., and Fahey, P.L., eds., Geology and ore deposits of the American Cordillera: Proceedings, Geological Society of Nevada Symposium, v. 3, p. 1639–1666.

Page 27: GEOSPHERE Structural reconstruction and age of an ...€¦ · Cordilleran metamorphic core complexes, which are exposures of formerly deeply buried mid-crustal and lower crustal metamor-phic-plutonic

Research Paper

263Pape et al. | Structural reconstruction, Spruce Mountain, NevadaGEOSPHERE | Volume 12 | Number 1

Wilkerson, M.S., and Dicken, C.L., 2001, Quick-look techniques for evaluating two-dimensional cross sections in detached contractional settings: American Association of Petroleum Geol-ogists Bulletin, v. 85, p. 1759–1770.

Wilkins, J., Jr., and Heidrick, T.L., 1995, Post-Laramide extension and rotation of porphyry cop-per deposits, southwestern United States, in Pierce, F.W., and Bolm, J.G., eds., Porphyry copper deposits of the American Cordillera: Arizona Geological Society Digest Volume 20, p. 109–127.

Wolverson, N.J., 2010, Technical report on the Spruce Mountain property, Elko County, Nevada, USA: Technical Report NI 43–101, prepared for Renaissance Gold, Inc., 79 p.

Wong, M.S., and Gans, P.B., 2008, Geologic, structural, and thermochronologic constraints on the tectonic evolution of the Sierra Mazatán core complex, Sonora, Mexico: New insights into metamorphic core complex formation: Tectonics, v. 27, TC4013, doi: 10 .1029 /2007TC002173 .

Wright, J.E., and Snoke, A.W., 1993, Tertiary magmatism and mylonitization in the Ruby–East Humboldt metamorphic core complex, northeastern Nevada: U-Pb geochronology and Sr, Nd, and Pb isotope geochemistry: Geological Society of America Bulletin, v. 105, p. 935–952, doi: 10 .1130 /0016 -7606 (1993)105 <0935: TMAMIT>2 .3 .CO;2 .

Wyld, S.J., and Wright, J.E., 2015, Effects of Jurassic slab break-off in the northern Great Basin: Geochemistry and geochronology of back-arc magmatism; structural and stratigraphic rela-tions; and possible implication for younger gold mineralization [abs.], in New concepts and discoveries: Geological Society of Nevada Symposium Program with Abstracts, p. 111–112.

Zamudio, J.A., and Atkinson, W.W., Jr., 1995, Mesozoic structures of the Dolly Varden Mountains and Currie Hills, Elko County, Nevada, in Miller, D.M., and Busby, C., eds., Jurassic magma-tism and tectonics of the North American Cordillera: Geological Society of America Special Paper 299, p. 295–312, doi: 10 .1130 /SPE299 -p295 .