Full Text View

Volume 22 Issue 3 (March 2012)

GSA Today

Bookmark and Share

Article, pp. 4-11 | Abstract | PDF (3.8MB)

Detrital zircon facies of Cordilleran terranes in western North America

Table of Contents
Search GoogleScholar for

Search GSA Today


 

Todd A. LaMaskin*

Department of Geography and Geology, University of North Carolina Wilmington, 601 South College Road, Wilmington, North Carolina 28403-5944, USA

Abstract

Paleozoic–Mesozoic basins in Cordilleran terranes of western North America contain detrital zircon U-Pb age distributions that vary over 10–100 Ma in a systematic and predictable manner. A minimum of four detrital zircon age distributions, here termed “detrital zircon facies,” are present: (1) Paleoproterozoic and Archean facies, chiefly found in Paleozoic and early Mesozoic accretionary complexes, is defined by late Archean–early Proterozoic (ca. 2.7–2.3 Ga) and late Proterozoic ages (ca. 2.0–1.6 Ga) with variable quantities of Paleozoic and early Mesozoic ages. (2) Mixed Proterozoic and Phanerozoic facies is found in Early–Late Jurassic basins and is defined by grains spanning ca. 2.0 Ga–160 Ma, derived from eastern-southwestern Laurentian transcontinental sources and enriched by western U.S. and eastern Mexican early Mesozoic plate-margin magmatism. (3) Triassic and Jurassic facies, found in Late Jurassic–Early Cretaceous basins, is defined by Late Jurassic ages (peak ca. 155 Ma) with a subordinate proportion of Triassic ages (peak ca. 230 Ma). (4) Jurassic and Early Cretaceous facies is found in late Early–early Late Cretaceous marginal basins and is defined by Jurassic and Early Cretaceous ages (ca. 200–130 and ca. 130–100 Ma). Detrital zircon U-Pb ages from terranes of western North America record stages of basin formation during phases of the supercontinent cycle and reflect second-order variability in the tectonic setting of an active continental plate margin. At this temporal and spatial scale, the integrated evolution of orogenic, erosion, and sediment-transport systems controls sediment provenance.

*Email:

Manuscript received 2 Nov. 2011; accepted 6 Jan. 2012.

DOI: 10.1130/GSATG142A.1

Introduction

On a global time-integrated scale, detrital zircon U-Pb age distributions reflect episodic magmatic accretion of continental crust (i.e., first-order scale, 100–1000 Ma, e.g., Condie and Aster, 2009). At an order of magnitude finer scale (i.e., second-order, 10–100 Ma, 100,000 km2 of rock), detrital zircon U-Pb age distributions in continental-margin basins should reflect the long-term distribution of orogenic belts, long-lived Andean-type magmatic arcs, and continental sediment-dispersal systems (e.g., Leeder, 1988; Patchett et al., 1999). Supporting evidence for this scale of tectonic control on sedimentary provenance in the ancient rock record exists, but is generally sparse and typically does not document transitional stages of paleotectonics (e.g., Rainbird et al., 1992; Riggs et al., 1996; Dickinson and Gehrels, 2003; Tyrrell et al., 2007; Druschke et al., 2011).

In this paper, my objectives are to use detrital zircon U-Pb data in terranes of western North America to (1) assess long-term provenance links to the Laurentian craton, (2) identify multi-stage sediment sources through time, and (3) emphasize second-order scale observation in detrital zircon U-Pb age studies. I suggest that this observational scale is critical for understanding changes in tectonic setting along active plate margins, as has been shown for neodymium isotopic values in continental sedimentary rocks (Patchett et al., 1999). In western North America, analysis of the detrital zircon record at this scale elucidates the transition from a marginal-basin regime (i.e., Karig, 1974; Tarney et al., 1981) during most of Paleozoic and early Mesozoic time, to an Andean-type integrated margin by Late Mesozoic time (i.e., McClelland et al., 1992; Saleeby and Busby-Spera, 1992; Dickinson et al., 1996; DeCelles, 2004) and argues against models invoking exotic ribbon continents (e.g., Johnston, 2008; Hildebrand, 2009).

Detrital Zircon Age Distributions in Western North America

Over the past ~20 years, the application of U-Pb geochronology to detrital zircon grains has yielded significant insight concerning sediment sources to western North American basins. In particular, terranes of western North America have been extensively studied using detrital zircon U-Pb ages with numerous contrasting interpretations (e.g., Ross and Bowring, 1990; Miller and Saleeby, 1995; Gehrels and Kapp, 1998; Gehrels et al., 2000; Brown and Gehrels, 2007; Wright and Wyld, 2007; Grove et al., 2008; Scherer and Ernst, 2008; Snow and Ernst, 2008; Piercey and Colpron, 2009; LaMaskin et al., 2011).

Here, I compile published and new detrital zircon U-Pb ages from pre-Devonian–early Late Cretaceous, arc-related basins in terranes of western North America. The compilation includes terranes with sufficient available data from west of the 87/86Sri = 0.706 line (Armstrong et al., 1977), from the southern California Coast Range to the Yukon-Tanana terrane in the north. My compilation, observations, and interpretations are specific to the time frame in this geographic range only; they are not intended to be a comprehensive review of Cordilleran provenance and tectonics, but rather, to serve as a starting point for continued investigation at this scale.

The overwhelming fundamental observation from the data is that regardless of interpreted terrane association, at a stratigraphic scale of 10–100 Ma, distinct age distributions are present in the same stratigraphic order along strike of the western North American margin. I recognize a minimum of four distinct detrital zircon age distributions in pre-Devonian–early Late Cretaceous clastic sedimentary successions in western North America. These age distributions vary systematically based on the depositional age and tectonic setting of the basin (cf. Gehrels, 2003).

  1. Paleoproterozoic and Archean distribution (PPA; Figs. 1 and 2A) is defined by an age distribution of late Archean–early Proterozoic (ca. 2.7–2.3 Ga) and late Proterozoic ages (ca. 2.0–1.6 Ga), with variable quantities of Paleozoic and Early Mesozoic ages dependent on depositional age of the rocks. The majority of samples do not include Mesoproterozoic ages (ca. 1.5–1.0 Ga). PPA is dominant in Paleozoic rocks from California to Alaska, in terranes typically defined as subduction-accretionary complexes or subduction mélange.
  2. Mixed Proterozoic and Phanerozoic distribution (MPP; Figs. 1 and 2B) is defined by a multimodal age distribution spanning 2.0–0.16 Ga, including distinct age ranges of (1) late Paleoproterozoic (ca. 2.0–1.6 Ga); (2) Mesoproterozoic (ca. 1.5–1.0 Ga); (3) Neoproterozoic (ca. 0.8–0.6 Ga); (4) early Paleozoic (ca. 0.5–0.35); (5) late Paleozoic (ca. 350–250 Ma); and (6) early Mesozoic (ca. 250–160 Ma). MPP is present in Early–Late Jurassic samples, typically in forearc/intra-arc basins in terranes defined as island-arc complexes.
  3. Triassic and Jurassic distribution (TrJ; Figs. 1 and 3A) is defined by a dominant Middle–Late Jurassic (ca. 175–145 Ma) age distribution, with variable quantities of Triassic ages (ca. 250–220 Ma), and a general lack of Precambrian ages. (Note that Fig. 3A is truncated at 300 Ma.) TrJ is present in samples from an extensive belt of Late Jurassic–Early Cretaceous basins that has long been recognized as correlative but of unclear tectonic setting (i.e., Cowan and Brandon, 1981; Garver, 1988; McClelland et al., 1992; Miller and Saleeby, 1995).
  4. Jurassic and Early Cretaceous distribution (JeK; Figs. 1 and 3B) is defined by an approximately bimodal age distribution of Jurassic–Early Cretaceous (ca. 200–130 Ma) and Early Cretaceous ages (ca. 130–100 Ma). JeK is found in samples from late Early–early Late Cretaceous basins, generally recognized as the forearc of the western North American Andean-style margin (i.e., Great Valley forearc) (Ingersoll, 1979; Degraaff-Surpless et al., 2002; Brown and Gehrels, 2007; Jacobson et al., 2011).

Fig. 1Figure 1

Map illustrating present-day location of terranes discussed in this paper, age distribution of Laurentian Precambrian crust, and detrital zircon sample locations: SEYTT—Southeastern Yukon-Tanana terrane; YTT-CM—Yukon-Tanana terrane in Coast Mountains; GRAV—Gravina Belt; MT—Methow-Tyaughton; EFLC—Easton-Fidalgo-Lummi-Constitution; YA—Yellow Aster; IG—Ingalls Graywacke; T-N—Tonga-Nooksack; LM—Lookout Mountain; CH—Coon Hollow; BT—Baker terrane; MI—Mitchell Inlier; SL—Snowshoe and Lonesome fms.; KRC—Klamath River Conglomerate; GAL—Galice; AMQ—Antelope Mountain Quartzite; EHT—Eastern Hayfork terrane; NFT—North Fork terrane; U-GVG—Upper Great Valley Group; T-GVG—“Tithonian” Great Valley Group; LDC-P—Lang-Duncan-Culberton allochthons and Picayune Valley Fm.; M-GVG—Middle Great Valley Group; MAR—Mariposa; JEK—Jurassic–Early Cretaceous; MPP—Mixed Proterozoic and Phanerozoic; and PPA—Paleoproterozoic and Archean. See Table DR1 for specific data sources 1 . Adapted from Gehrels (2001); Wyld and Wright (2001); DeGraaff-Surpless et al. (2002); Wyld et al. (2006); Brown and Gehrels (2007); Nelson and Gehrels (2007); base modified from Whitmeyer and Karlstrom (2007).

1 GSA supplemental data item 2012079, Table DR1: Data sources, is available in the online data repository. You can also request a copy from GSA Today, P.O. Box 9140, Boulder, CO 80301-9140, USA; .

Fig. 2Figure 2

Detrital zircon U-Pb age data from (A) Paleozoic–Mesozoic subduction-accretionary complexes and (B) Early to Late Jurassic basins in island-arc complexes. Specific data sources are shown in Table DR1 (see footnote 1).

Fig. 3Figure 3

Detrital zircon U-Pb age data from (A) Late Jurassic–Early Cretaceous basins and (B) late Early to early Late Cretaceous basins. Specific data sources are shown in Table DR1 (see footnote 1).

Discussion

The compilation presented here shows sequential regularity in distinct detrital zircon U-Pb ages at a scale of 10–100 Ma. The detrital zircon age distributions represent a definable aspect of large bodies of rock, are observation based, and allow for distinction between adjacent units. As such, they are here designated “detrital zircon facies.” The age ranges present in a given detrital zircon facies (i.e., their recognition criteria) relate directly to known regional source areas of both primary and recycled grains. In the following sections, I review each detrital zircon facies and implications for development of the western Laurentian margin (Fig. 4).

Fig. 4Figure 4

Chronostratigraphic distribution of detrital zircon facies in western North American terranes, as well as referenced samples from the western U.S. interior. GVG—Great Valley Group; PPA—Paleoproterozoic and Archean. Specific data sources are shown in Table DR1 (see footnote 1).

Paleozoic–Early Mesozoic Time

The provenance of PPA facies is either (1) crystalline sources in northwestern Laurentia (Gehrels et al., 1995, 2000), (2) rifted and translated crustal fragments of the Precambrian–Paleozoic northwestern Laurentian miogeocline (Nelson and Gehrels, 2007; cf. Bradley et al., 2007; Beranek et al., 2010a), or
(3) originally peri-Gondwanan/Avalonian crust that was tectonically emplaced along the southern Laurentian margin in early–mid-Paleozoic time and subsequently translated along the plate margin in mid–late Paleozoic time (Wright and Wyld, 2006; Grove et al., 2008). Thus, rocks bearing PPA facies may reflect sediment derivation from northwestern Laurentian sources enhanced by plate-margin magmatism or multicycle sediment reworking and tectonic translation of crustal fragments allochthonous to western North America.

Regardless of the ultimate source of age-characteristic zircon grains, PPA facies sand was present in arc-basin complexes on the western North American plate margin by mid-Paleozoic time (Harding et al., 2000; Spurlin et al., 2000; cf. Lindsley-Griffin et al., 2006), and was subsequently recycled along the margin (Fig. 4) (e.g., Scherer et al., 2010; LaMaskin et al., 2011). Noteworthy, non-PPA age distributions in Paleozoic accretionary-subduction complexes of the Klamath Mountains and Sierra Nevada may represent exotic crust, structurally intercalated with PPA-bearing rocks, or may suggest other explanations (e.g., Harding et al., 2000; Wright and Wyld, 2007; Grove et al., 2008).

Early–Late Jurassic Time

The age distribution in MPP facies represents transcontinental sand shed from the greater Ouachita-Appalachian orogeny and enriched by southwestern Laurentian sources, as well as early Mesozoic, plate-margin magmatism in the western U.S. and eastern Mexico (Fig. 4; cf. Dickinson and Gehrels, 2003, 2009; Rahl et al., 2003). The presence of a transcontinental signature in each of these Early–Late Jurassic basins (Fig. 2B; cf. Izsak et al., 2007; Dickinson and Gehrels, 2008a, 2009; LaMaskin et al., 2011) suggests proximity to North America and the modern southwestern U.S. in early Mesozoic time, and that active orogenic structures and the plate-margin arc itself were not barriers to sediment transfer from the craton to the arc.
Existing data suggest that these transcontinental sediments were not incorporated into western North American peripheral-arc systems until Early Jurassic time (Fig. 4; ca. 190–185 Ma, Klamath Mountains, North Fork terrane) (Scherer and Ernst, 2008). It is not clear why transcontinental sediment was not delivered to arc-basin systems of the western U.S. during late Paleozoic time coincident with onset of the Alleghanian orogeny in eastern Laurentia (Hatcher, 2010). Additional data is needed from rocks of Late Paleozoic–Early Jurassic age to assess the timing of delivery of transcontinental sands to plate-margin basins and the transition from PPA to MPP facies.

Despite interpretations of a forearc setting for most Early–Late Jurassic basins (e.g., Dickinson, 1979; MacDonald, 2006; Scherer and Ernst, 2008; Snow and Ernst, 2008), MPP facies may also represent deposition (1) in a flexural basin adjacent to the uplifted western Nevadan back-arc basin (i.e., Jurassic Luning-fencemaker fold-thrust belt) (Wyld, 2002; Dorsey and LaMaskin, 2007; LaMaskin et al., 2011); (2) in extensional basins along the northward-deepening plate-margin arc (e.g., Busby-Spera, 1988; Barth et al., 2004; Dickinson and Gehrels, 2009; LaMaskin et al., 2011); or (3) in suprasubduction zone basins during arc extension and subsequent closure (e.g., Snoke, 1977; Harper, 1980; Hacker et al., 1995). Along-strike variability in modern southeast Pacific active margins suggests that these alternatives are not mutually exclusive. Proposed Cretaceous dextral-transpressive shear along the Mojave-Snow-Nevada-Idaho shear zone would have resulted in northern displacement of these originally southwestern U.S. basins (Wyld and Wright, 2001).

Late Jurassic–Early Cretaceous Time

In Late Jurassic–Early Cretaceous basins bearing TrJ facies, the paucity, or complete absence, of Precambrian zircon grains, in conjunction with a voluminous record of Middle–Late Jurassic magmatism, is interpreted to represent initial isolation of basins from the Laurentian craton due to nascent construction of the Andean-type margin (Figs. 3A and 4). Where Precambrian ages are present, age distributions mimic underlying MPP facies. A decrease in the relative abundance of Precambrian grains as compared to older samples discussed here may also simply represent an overwhelming increase in Mesozoic grains; however, for large sample sizes (e.g., Methow basin), Precambrian grains are not represented, suggesting that the decrease is not due to simple dilution.

The stratigraphic transition from MPP to TrJ facies is interpreted to record stepwise growth of orogenic highlands, as sediment pathways that were connected to the craton were cut off and new ones established along the rising plate margin. I propose that the transition from MPP to TrJ facies illustrates that, despite retro-arc thrusting and associated elevation gain during Middle Jurassic time (Wyld, 2002; Fuentes et al., 2009), there was not a contiguous mountain belt (i.e., topographically integrated) along the Pacific margin until Late Jurassic–Early Cretaceous time. According to this model, continental-scale drainage patterns in the western U.S. were reversed between Late Jurassic–Early Cretaceous time and transcontinental sands no longer entered western U.S. marginal basins. Instead, sediment sources to marginal basins became restricted to active arc and older igneous basement rocks in the now high-standing continental-margin, Andean-type arc.

The presence of TrJ facies in rocks as old as ca. 153–150 Ma (Fig. 4; Galice Formation, Klamath Mountains Province; Miller et al., 2003; Saleeby et al., 1982) suggests that plate reorganization from a complex, marginal-basin regime to an Andean-style margin may have initiated in middle Late Jurassic time (cf. Miller and Saleeby, 1995; Dumitru et al., 2010). A continuous basinal record of this Late Jurassic tectonic reorganization may be found in the western Klamath Mountains, where the Galice Formation includes both a transcontinental MPP facies (i.e., Izsak, et al., 2007) and a TrJ facies (i.e., Miller et al., 2003). Syndepositional, suprasubduction compression of the Galice basin in the western Klamath Mountains (Snoke, 1977; Wyld and Wright, 1988; Harper et al., 1994) may represent the earliest manifestation of plate reorganization. This compilation of detrital zircon U-Pb age data supports the idea of a “common origin for coeval strata on differing basement terranes” (Miller and Saleeby, 1995, p. 18,057), suggesting along-strike integration of disparate substrate crust by Late Jurassic time.

Late Early–Early Late Cretaceous Time

In late Early–early Late Cretaceous basins containing JeK facies, the general lack of Precambrian zircon grains represents continued isolation of marginal basins from the craton by the high-standing Andean-style margin (Figs. 3B and 4). Considering the analytical precision of SHRIMP (sensitive high-resolution ion microprobe) and LA-ICPMS (laser-ablation–inductively coupled plasma mass spectroscopy) methods, the nearly bimodal age distribution is a clear record of the two main magmatic phases of the Mesozoic Andean-type arc (i.e., Late Jurassic and Late Cretaceous; Ducea, 2001; Irwin and Wooden, 2001; Irwin, 2003). Any true differences in age modes and variance represent expected variability in the timing of arc magmatism along the Andean-type arc.

Conclusions and Implications

In western North America, detrital zircon U-Pb age distributions vary in a systematic and predictable manner and are interpreted to reflect second-order variability in the tectonic setting of an active plate margin. I suggest that at a second-order, 10–100 Ma scale, detrital zircon ages are governed by plate-tectonic setting, in a similar manner to controls on neodymium isotopic values from continental sedimentary rocks (Patchett et al., 1999). At the second-order scale, the integrated evolution of orogenic, erosion, and sediment-transport systems controls detrital zircon U-Pb age distributions and, accordingly, sediment provenance.

Critical evaluation of the model presented here, and integration with rapidly emerging data sets from the miogeocline and interior of the western U.S. and Canada (e.g., Dickinson and Gehrels, 2008a, 2008b, 2009; Scherer et al., 2008; Dickinson et al., 2010; Beranek et al., 2010b; Druschke et al., 2011; Fuentes et al., 2010; Leier and Gehrels, 2011) from tectonically enigmatic thin-skinned sheets such as the Roberts Mountain and Golconda allochthons (e.g., Riley et al., 2000; Gehrels et al., 2000; Wright and Wyld, 2006), and roof pendants within the Sierra Nevada batholith (e.g., Memeti et al., 2010), may help resolve the affinity of numerous western Laurentian basins through time. Continued integration of data sets will lead to a better understanding of the pace, areal extent, and along-strike variability of North American plate-margin tectonics with inference for global plate-tectonic processes.

The model proposed here unifies a large data set collected over thousands of kilometers and representing hundreds of millions of years of sedimentation, sets forth predictions for new data collection, and is inherently testable both regionally and on other continents. Identification of age distributions that do not fit the model presented here is critical and may point to a truly allochthonous origin for rocks within western North America (e.g., Harding et al., 2000; Wright and Wyld, 2007; Grove et al., 2008). A global evaluation of detrital zircon U-Pb age distributions at this scale may provide important information for understanding the pace and spatial scale of crustal growth via arc and terrane accretion.

Acknowledgments

My work in the Blue Mountains was supported by NSF grants to R. Dorsey (0537691) and J. Vervoort (EAR-0537913). I thank B. Housen, J. Miller, and J. MacDonald, and J. Schwartz for the use of data previously published in abstract form. Early versions of this manuscript were reviewed by M. Benedetti, R. Dorsey, R. Hildebrand, T. Lawton, N. Niemi, and J. Schwartz. I thank GSA Today editor R.D. Nance, an anonymous reviewer, and C. Busby for reviews of this manuscript. My interpretations would not be possible without the work of W. Dickinson, G. Ernst, G. Gehrels, W. McLelland, J. Saleeby, K. Surpless, S. Wyld, J. Wright, and many others.

REFERENCES CITED

  1. Alexander, R.S., and Schwartz, J.J., 2009, Detrital zircon geochronology of Permian-Triassic metasedimentary rocks in the Baker terrane, Blue Mountains Province, NE Oregon: Geological Society of America Abstracts with Programs, v. 41, no. 7, p. 294.
  2. Armstrong, R.L., Taubeneck, W.H., and Hales, P.O., 1977, Rb-Sr and K-Ar geochronometry of Mesozoic granitic rocks and their Sr isotopic composition, Oregon, Washington, and Idaho: GSA Bulletin, v. 88, p. 397–411, doi: 10.1130/0016-7606(1977)88<397:RAKGOM>2.0.CO;2.
  3. Barth, A.P., Wooden, J.L., Jacobson, C.E., and Probst, K., 2004, U-Pb geochronology and geochemistry of the McCoy Mountains Formation, southeastern California: A Cretaceous retroarc foreland basin: GSA Bulletin, v. 116, p. 142–153, doi: 10.1130/B25288.1.
  4. Beranek, L.P., Mortensen, J.K., Lane, L.S., Allen, T.L., Fraser, T.A., Hadlari, T., and Zantvoort, W.G., 2010a, Detrital zircon geochronology of the western Ellesmerian clastic wedge, northwestern Canada: Insights on Arctic tectonics and the evolution of the northern Cordilleran miogeocline: GSA Bulletin, v. 122, p. 1899–1911, doi: 10.1130/B30120.1.
  5. Beranek, L.P., Mortensen, J.K., Orchard, M.J., and Ullrich, T., 2010b, Provenance of North American Triassic strata from west-central and southeastern Yukon: Correlations with coeval strata in the Western Canada Sedimentary Basin and Canadian Arctic Islands: Canadian Journal of Earth Sciences, v. 47, p. 53–73, doi: 10.1139/E09-065.
  6. Bradley, D.C., McClelland, W.C., Wooden, J.L., Till, A.B., Roeske, S.M., Miller, M.L., Karl, S.M., and Abbott, J.G., 2007, Detrital zircon geochronology of some Neoproterozoic to Triassic rocks in interior Alaska, in Ridgway, K.D., et al., eds., Tectonic Growth of a Collisional Continental Margin: Crustal Evolution of Southern Alaska: Geological Society of America Special Paper 431, p. 155–189, doi: 10.1130/2007.2431(07).
  7. Brown, E.H., and Gehrels, G.E., 2007, Detrital zircon constraints on terrane ages and affinities and timing of orogenic events in the San Juan Islands and North Cascades, Washington: Canadian Journal of Earth Sciences, v. 44, p. 1375–1396, doi: 10.1139/E07-040.
  8. Busby-Spera, C., 1988, Speculative tectonic model for the early Mesozoic arc of the southwest Cordilleran United States: Geology, v. 16, p. 1121–1125, doi: 10.1130/0091-7613(1988)016<1121:STMFTE>2.3.CO;2.
  9. Condie, K.C., and Aster, R.C., 2009, Zircon age episodicity and growth of continental crust: EOS (Transactions, American Geophysical Union), v. 90, doi: 10.1029/2009EO410003.
  10. Cowan, D.S., and Brandon, M.T., 1981, Contrasting facies in upper Mesozoic strata of Pacific Northwest (abs.): AAPG Bulletin, v. 65, p. 913–914, doi: 10.1306/2F919D00-16CE-11D7-8645000102C1865D.
  11. DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland basin system, western USA: American Journal of Science, v. 304, p. 105–168, doi: 10.2475/ajs.304.2.105.
  12. DeGraaff-Surpless, K., Graham, S.A., Wooden, J.L., and McWilliams, M.O., 2002, Detrital zircon provenance analysis of the Great Valley Group, California: Evolution of an arc-forearc system: GSA Bulletin, v. 114, p. 1564–1580, doi: 10.1130/0016-7606(2002)114<1564:DZPAOT>2.0.CO;2.
  13. Dickinson, W.R., 1979, Mesozoic forearc basin in central Oregon: Geology, v. 7, no. 4, p. 166–170, doi: 10.1130/0091-7613(1979)7<166:MFBICO>2.0.CO;2.
  14. Dickinson, W.R., and Gehrels, G.E., 2003, U-Pb ages of detrital zircons from Permian and Jurassic eolian sandstones of the Colorado Plateau, USA: Paleogeographic implications: Sedimentary Geology, v. 163, p. 29–66, doi: 10.1016/S0037-0738(03)00158-1.
  15. Dickinson, W.R., and Gehrels, G.E., 2008a, U-Pb ages of detrital zircons in relation to paleogeography: Triassic paleodrainage networks and sediment dispersal across southwest Laurentia: Journal of Sedimentary Research, v. 78, p. 745–764, doi: 10.2110/jsr.2008.088.
  16. Dickinson, W.R., and Gehrels, G.E., 2008b, Sediment delivery to the Cordilleran foreland basin: Insights from U-Pb ages of detrital zircons in Upper Jurassic and Cretaceous strata of the Colorado Plateau: American Journal of Science, v. 308, p. 1041–1082, doi: 10.2475/10.2008.01.
  17. Dickinson, W.R., and Gehrels, G.E., 2009, U-Pb ages of detrital zircons in Jurassic eolian and associated sandstones of the Colorado Plateau: Evidence for transcontinental dispersal and intraregional recycling of sediment: GSA Bulletin, v. 121, p. 408–433, doi: 10.1130/B26406.1.
  18. Dickinson, W.R., Hopson, C.A., and Saleeby, J.B., 1996, Alternate origins of the Coast Range ophiolite (California): Introduction and implications: GSA Today, v. 6, no. 2, p. 1–2.
  19. Dickinson, W.R., Lawton, T.F., and Gehrels, G.E., 2009, Recycling detrital zircons: A case study from the Cretaceous Bisbee Group of southern Arizona: Geology, v. 37, p. 503–506, doi: 10.1130/G25646A.1.
  20. Dickinson, W.R., Gehrels, G.E., and Stern, R.J., 2010, Late Triassic Texas uplift preceding Jurassic opening of the Gulf of Mexico: Evidence from U-Pb ages of detrital zircons: Geosphere, v. 6, p. 641–662, doi: 10.1130/GES00532.1.
  21. Dorsey, R.J., and LaMaskin, T.A., 2007, Stratigraphic record of Triassic-Jurassic collisional tectonics in the Blue Mountains Province, northeastern Oregon: American Journal of Science, v. 307, p. 1167–1193, doi: 10.2475/10.2007.03.
  22. Druschke, P., Hanson, A.D., Wells, M.L., Gehrels, G.E., and Stockli, D., 2011, Paleogeographic isolation of the Cretaceous to Eocene Sevier hinterland, east-central Nevada: Insights from U-Pb and (U-Th)/He detrital zircon ages of hinterland strata: GSA Bulletin, v. 123, p. 1141–1160, doi: 10.1130/B30029.1.
  23. Ducea, M., 2001, The California arc: Thick granitic batholiths, eclogitic residues, lithospheric-scale thrusting, and magmatic flare-ups: GSA Today, v. 11, p. 4–10, doi: 10.1130/1052-5173(2001)011<0004:TCATGB>2.0.CO;2.
  24. Dumitru, T.A., Wakabayashi, J., Wright, J.E., and Wooden, J.L., 2010, Early Cretaceous transition from nonaccretionary behavior to strongly accretionary behavior within the Franciscan subduction complex: Tectonics, v. 29, doi: 10.1029/2009TC002542.
  25. Fuentes, F., DeCelles, P.G., and Gehrels, G.E., 2009, Jurassic onset of foreland basin deposition in northwestern Montana, USA: Implications for along-strike synchroneity of Cordilleran orogenic activity: Geology, v. 37, p. 379–382, doi: 10.1130/G25557A.1.
  26. Fuentes, F., DeCelles, P.G., Constenius, K.N., and Gehrels, G.E., 2010, Evolution of the Cordilleran foreland basin system in northwestern Montana, U.S.A.: GSA Bulletin, v. 123, p. 507–533, doi: 10.1130/B30204.1.
  27. Garver, J.I., 1988, Fragment of the Coast Range ophiolite and the Great Valley sequence in the San Juan Islands, Washington: Geology, v. 16, p. 948–951, doi: 10.1130/0091-7613(1988)016<0948:FOTCRO>2.3.CO;2.
  28. Gehrels, G.E., 2001, Geology of the Chatham Sound region, southeast Alaska and coastal British Columbia: Canadian Journal of Earth Sciences, v. 38, p. 1579–1599, doi: 10.1139/e01-040.
  29. Gehrels, G.E., 2003, Detrital zircon constraints on sediment dispersal patterns in western North America: Geological Society of America Abstracts with Programs, v. 35, no. 6, p. 389.
  30. Gehrels, G.E., and Kapp, P.A., 1998, Detrital zircon geochronology and regional correlation of metasedimentary rocks in the Coast Mountains, southeastern Alaska: Canadian Journal of Earth Sciences, v. 35, p. 269–279, doi: 10.1139/e97-114.
  31. Gehrels, G.E., Dickinson, W.R., Ross, G.M., Stewart, J.H., and Howell, D.G., 1995, Detrital zircon reference for Cambrian to Triassic miogeoclinal strata of western North America: Geology, v. 23, p. 831–834, doi: 10.1130/0091-7613(1995)023<0831:DZRFCT>2.3.CO;2.
  32. Gehrels, G.E., Dickinson, W.R., Darby, B.J., Harding, J.P., Manuszak, J.D., Riley, B.C.D., Spurlin, M.S., Finney, S.C., Girty, G.H., Harwood, D.S., Miller, M.M., Satterfield, J.I., Smith, M.T., Snyder, W.S., Wallin, E.T., and Wyld, S.J., 2000, Tectonic implications of detrital zircon data from Paleozoic and Triassic strata in western Nevada and northern California, in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern California: Geological Society of America Special Paper 347, p. 133–150, doi: 10.1130/0-8137-2347-7.133.
  33. Grove, M., Gehrels, G.E., Cotkin, S.J., Wright, J.E., and Zou, H., 2008, Non-Laurentian cratonal provenance of Late Ordovician eastern Klamath blueschists and a link to the Alexander terrane, in Wright, J.E., and Shervais, J.W., eds., Ophiolites, Arcs, and Batholiths: A Tribute to Cliff Hopson: Geological Society of America Special Paper 438, p. 223–250, doi: 10.1130/2008.2438(08).
  34. Hacker, B., Donato, M., Barnes, C., McWilliams, M.O., and Ernst, W.G., 1995, Timescales of orogeny: Jurassic construction of the Klamath Mountains: Tectonics, v. 14, p. 677–703, doi: 10.1029/94TC02454.
  35. Harding, J.P., Gehrels, G.E., Harwood, D.S., and Girty, G.H., 2000, Detrital zircon geochronology of the Shoo Fly Complex, northern Sierra terrane, northeastern California, in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern California: Geological Society of America Special Paper 347, p. 43–55, doi: 10.1130/0-8137-2347-7.43.
  36. Harper, G., 1980, The Josephine ophiolite—Remains of a Late Jurassic marginal basin in northwestern California: Geology, v. 8, p. 333–337, doi: 10.1130/0091-7613(1980)8<333:TJOOAL>2.0.CO;2.
  37. Harper, G.D., Saleeby, J.B., and Heizler, M., 1994, Formation and emplacement of the Josephine ophiolite and the Nevadan orogeny in the Klamath Mountains, California-Oregon: U/Pb zircon and 40Ar/39Ar geochronology: Journal of Geophysical Research, v. 99, no. B3, p. 4293–4321, doi: 10.1029/93JB02061.
  38. Harper, G.D., Miller, R.B., MacDonald, J.H., Jr., Miller, J.S., and Mlinarevic, A.N., 2003, Evolution of a polygenetic ophiolite: The Jurassic Ingalls Ophiolite, Washington Cascades, in Swanson, T.W., ed., Western Cordillera and Adjacent Areas: Geological Society of America Field Guide 4, p. 251–265, doi: 10.1130/0-8137-0004-3.251.
  39. Hatcher, R.D., Jr., 2010, The Appalachian orogen: A brief summary, in Tollo, R.P., Bartholomew, M.J., Hibbard, J.P., and Karabinos, P.M, eds., From Rodinia to Pangea: The Lithotectonic Record of the Appalachian Region: Geological Society of America Memoir 206, p. 1–19, doi: 10.1130/2010.1206(01).
  40. Hildebrand, R.S., 2009, Did Westward Subduction Cause Cretaceous–Tertiary Orogeny in the North American Cordillera?: Geological Society of America Special Paper 457, 71 p., doi: 10.1130/2009.2457.
  41. Ingersoll, R.V., 1979, Evolution of the Late Cretaceous forearc basin, northern and central California: GSA Bulletin, v. 90, p. 813–826, doi: 10.1130/0016-7606(1979)90<813:EOTLCF>2.0.CO;2.
  42. Irwin, W.P., 2003, Correlation of the Klamath Mountains and Sierra Nevada: U.S. Geological Survey Open-File Report 02-0490, 2 sheets.
  43. Irwin, W.P., and Wooden, J.L., 2001, Map showing plutons and accreted terranes of the Sierra Nevada, California, with a tabulation of U/Pb isotopic ages: U.S. Geological Survey Open-File Report 01-299, 1 sheet.
  44. Izsak, G., Wright, J.E., Wyld, S., Kimbrough, D., and Grove, M., 2007, Paleogeographic implications of similar detrital zircon signatures in mostly Jurassic and Cretaceous strata from Idaho to southern Baja California: Geological Society of America Abstracts with Programs, v. 39, no. 4, p. 8.
  45. Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsaglia, K.M., Gehrels, G.E., and Nourse, J.A., 2011, Late Cretaceous–early Cenozoic tectonic evolution of the southern California margin inferred from provenance of trench and forearc sediments: GSA Bulletin, v. 123, p. 485–506, doi: 10.1130/B30238.1.
  46. Johnston, S.J., 2008, The Cordilleran ribbon continent of North America: Annual Review of Earth and Planetary Sciences, v. 36, p. 495–530, doi: 10.1146/annurev.earth.36.031207.124331.
  47. Karig, D.E., 1974, Evolution of arc systems in the western Pacific: Annual Review of Earth and Planetary Sciences, v. 2, p. 51–75, doi: 10.1146/annurev.ea.02.050174.000411.
  48. LaMaskin, T.A., Vervoort, J.D., and Dorsey, R.J., 2011, Early Mesozoic paleogeography and tectonic evolution of the western United States: Insights from detrital zircon U-Pb geochronology of the Blue Mountains Province, northeastern Oregon, U.S.A.: GSA Bulletin, v. 123, p. 1939–1965, doi: 10.1130/B30260.1.
  49. Leeder, M.R., 1988, Devono-Carboniferous river systems and sediment dispersal from the orogenic belts and cratons of NW Europe, in Harris, A.L., and Fettes, D.J., eds., The Caledonian-Appalachian Orogen: Geological Society of London Special Publication 38, p. 549–558.
  50. Leier, A.L., and Gehrels, G.E., 2011, Continental-scale detrital zircon provenance signatures in Lower Cretaceous strata, western North America: Geology, v. 39, p. 399–402, doi: 10.1130/G31762.1.
  51. Lindsley-Griffin, N., Griffin, J.R., Farmer, J.D., Sivers, E.A., Bruckno, B., and Tozer, M.K., 2006, Ediacaran cyclomedusoids and the paleogeographic setting of the Neoproterozoic–early Paleozoic Yreka and Trinity terranes, eastern Klamath Mountains, California, in Snoke, A.W., and Barnes, C.G., eds., Geological Studies in the Klamath Mountains Province, California and Oregon: A Volume in Honor of William P. Irwin: Geological Society of America Special Paper 410, p. 411–431, doi:10.1130/2006.2410(20).
  52. MacDonald, J.H., Jr., 2006, Petrology, petrogenesis, and tectonic setting of Jurassic rocks of the Central Cascades, Washington, and Western Klamath Mountains, California-Oregon [Ph.D. diss.]: Albany, State University of New York at Albany, 415 p.
  53. Martin, A.J., Wyld, S.J., Wright, J.E., and Bradford, J.H., 2009, The Lower Cretaceous King Lear Formation, northwest Nevada: Implications for Mesozoic orogenesis in the western U.S. Cordillera: GSA Bulletin, v. 122, p. 537–562, doi: 10.1130/B26555.1.
  54. McClelland, W.C., Gehrels, G.E., and Saleeby, J.B., 1992, Upper Jurassic-Lower Cretaceous basinal strata along the Cordilleran Margin: Implications for the accretionary history of the Alexander-Wrangellia-Peninsular Terrane: Tectonics, v. 11, p. 823–835, doi: 10.1029/92TC00241.
  55. Memeti, V., Gehrels, G.E., Paterson, S.R., Thompson, J.M., Mueller, R.M., and Pignotta, G.S., 2010, Evaluating the Mojave-Snow Lake fault hypothesis and origins of central Sierran metasedimentary pendant strata using detrital zircon provenance analyses: Lithosphere, v. 2, p. 341–360, doi: 10.1130/L58.1.
  56. Miller, J.S., Miller, R.B., Wooden, J.L., and Harper, G.D., 2003, Geochronologic links between the Ingalls Ophiolite, North Cascades, Washington and the Josephine Ophiolite, Klamath Mts., Oregon and California: Geological Society of America Abstracts with Programs, v. 35, no. 6, p. 113.
  57. Miller, M.M., and Saleeby, J.B., 1995, U-Pb geochronology of detrital zircon from Upper Jurassic synorogenic turbidites, Galice Formation, and related rocks, western Klamath Mountains: Correlation and Klamath Mountains provenance: Journal of Geophysical Research, v. 100, B9, p. 18,045–18,058, doi: 10.1029/95JB00761.
  58. Nelson, J.A., and Gehrels, G.E., 2007, Detrital zircon geochronology and provenance of the southeastern Yukon-Tanana terrane: Canadian Journal of Earth Sciences, v. 44, p. 297–316, doi: 10.1139/e06-105.
  59. Patchett, P.J., Ross, G.M., and Gleason, J.D., 1999, Continental drainage in North America during the Phanerozoic from Nd isotopes: Science, v. 283, p. 671–673, doi: 10.1126/science.283.5402.671.
  60. Piercey, S.J., and Colpron, M., 2009, Composition and provenance of the Snowcap assemblage, basement to the Yukon-Tanana terrane, northern Cordillera: Implications for Cordilleran crustal growth: Geosphere, v. 5, p. 439–464, doi: 10.1130/GES00505.1.
  61. Rainbird, R.H., Hearnan, L.M., and Young, G., 1992, Sampling Laurentia: Detrital zircon geochronology offers evidence for an extensive Neoproterozoic river system originating from the Grenville orogeny: Geology, v. 20, p. 351–354, doi: 10.1130/0091-7613(1992)020<0351:SLDZGO>2.3.CO;2.
  62. Rahl, J.M., Reiners, P.W., Campbell, I.H., Nicolescu, S., and Allen, C.M., 2003, Combined single-grain (U-Th)/He and U/Pb dating of detrital zircons from the Navajo Sandstone, Utah: Geology, v. 31, p. 761–764, doi: 10.1130/G19653.1.
  63. Riggs, N.R., Lehman, T.M., Gehrels, G.E., and Dickinson, W.R., 1996, Detrital zircon link between headwaters and terminus of the Upper Triassic Chinle-Dockum paleoriver system: Science, v. 273, p. 97–100, doi: 10.1126/science.273.5271.97.
  64. Riley, B.C.D., Snyder, W.S., and Gehrels, G.E., 2000, U-Pb detrital zircon geochronology of the Golconda Allochthon, Nevada, in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern California: Geological Society of America Special Paper 347, p. 65–76.
  65. Ross, G.M., and Bowring, S.A., 1990, Detrital zircon geochronology of the Windermere Supergroup and the tectonic assembly of the southern Canadian Cordillera: The Journal of Geology, v. 98, p. 879–893, doi: 10.1086/629459.
  66. Saleeby, J.B., and Busby-Spera, C., 1992, Early Mesozoic tectonic evolution of the Western U.S. Cordillera, in Burchfiel, B.C., Lipman, P.W., and Zoback, M.L., eds., The Cordilleran Orogen: Conterminous U.S.: Boulder, Colorado, Geological Society of America, Geology of North America v. G-3, p. 107–168.
  67. Saleeby, J.B., Harper, G.D., Snoke, A.W., and Sharp, W.D., 1982, Time relations and structural-stratigraphic patterns in ophiolite accretion, west central Klamath Mountains, California: Journal of Geophysical Research, v. 87, B5, p. 3831–3848, doi: 10.1029/JB087iB05p03831.
  68. Scherer, H.H., and Ernst, W.G., 2008, North Fork terrane, Klamath Mountains, California: Geologic, geochemical, and geochronologic evidence for an early Mesozoic forearc, in Wright, J.E., and Shervais, J.W., eds., Ophiolites, Arcs, and Batholiths: A Tribute to Cliff Hopson: Geological Society of America Special Paper 438, p. 289–309.
  69. Scherer, H.H., Ernst, W.G., and Hanson, R.B., 2008, Geologic implications of new zircon U-Pb ages from the White Mountain Peak Metavolcanic Complex, eastern California: Tectonics, v. 27, TC2002, doi: 10.1029/2007TC002141.
  70. Scherer, H.H., Ernst, W.G., and Wooden, J.L., 2010, Regional detrital zircon provenance of exotic metasandstone blocks, Eastern Hayfork Terrane, Western Paleozoic and Triassic Belt, Klamath Mountains, California: The Journal of Geology, v. 118, p. 641–653, doi: 10.1086/656352.
  71. Snoke, A.W., 1977, A thrust plate of ophiolitic rocks in the Preston Peak area, Klamath Mountains, California: Geological Society of America Bulletin, v. 88, p. 1641–1659, doi: 10.1130/0016-7606(1977)88<1641:ATPOOR>2.0.CO;2.
  72. Snow, C.A., and Ernst, W.G., 2008, Detrital zircon constraints on sediment distribution and provenance of the Mariposa Formation, central Sierra Nevada foothills, California, in Wright, J.E., and Shervais, J.W., eds., Ophiolites, Arcs, and Batholiths: A Tribute to Cliff Hopson: Geological Society of America Special Paper 438, p. 311–330, doi: 10.1130/2008.2438(11).
  73. Spurlin, M.S., Gehrels, G.E., and Harwood, D.S., 2000, Detrital zircon geochronology of upper Paleozoic and lower Mesozoic strata of the northern Sierra terrane, northeastern California, in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern California: Geological Society of America Special Paper 347, p. 89–98.
  74. Surpless, K.D., and Augsburger, G.A., 2009, Provenance of the Pythian Cave conglomerate, northern California: Implications for mid-Cretaceous paleogeography of the U.S. Cordillera: Cretaceous Research, v. 30, p. 1181–1192, doi: 10.1016/j.cretres.2009.05.005.
  75. Surpless, K.D., Graham, S.A., Covault, J.A., and Wooden, J.L., 2006, Does the Great Valley Group contain Jurassic strata? Reevaluation of the age and early evolution of a classic forearc basin: Geology, v. 34, p. 21–24, doi: 10.1130/G21940.1.
  76. Tarney, J., Windley, B.F., and Krö;ner, A., 1981, Marginal basins through geological time (and discussion): Royal Society of London Philosophical Transactions, ser. A, v. 301, p. 217–232, doi: 10.1098/rsta.1981.0107.
  77. Tyrrell, S., Haughton, P.D.W., and Daly, J.S., 2007, Drainage reorganization during breakup of Pangea revealed by in-situ Pb isotopic analysis of detrital K-feldspar: Geology, v. 35, p. 971–974, doi: 10.1130/G4123A.1.
  78. Whitmeyer, S.J., and Karlstrom, K.E., 2007, Tectonic model for the Proterozoic growth of North America: Geosphere, v. 3, p. 220–259, doi: 10.1130/GES00055.1.
  79. Wright, J.E., and Wyld, S.J., 2006, Gondwanan, Iapetan, Cordilleran interactions: A geodynamic model for the Paleozoic tectonic evolution of the North American Cordillera, in Haggart, J.E., Enkin, R.J., and Monger, J.W.H., eds., Paleogeography of the North American Cordillera: Evidence for and against Large-Scale Displacements: Geological Association of Canada Special Paper 46, p. 377–408.
  80. Wright, J.E., and Wyld, S.J., 2007, Alternative tectonic model for Late Jurassic through Early Cretaceous evolution of the Great Valley Group, California, in Cloos, M., Carlson, W.D., Gilbert, M.C., Liou, J.G., and Sorensen, S.S., eds., Convergent Margin Terranes and Associated Regions: A Tribute to W.G. Ernst: Geological Society of America Special Paper 419, p. 81–95.
  81. Wyld, S.J., 2002, Structural evolution of a Mesozoic backarc fold-and-thrust belt in the U.S. Cordillera: New evidence from northern Nevada: Geological Society of America Bulletin, v. 114, p. 1452–1468, doi: 10.1130/0016-7606(2002)114<1452:SEOAMB>2.0.CO;2.
  82. Wyld, S.J., and Wright, J.E., 1988, The Devils Elbow ophiolite remnant and overlying Galice Formation: New constraints on the Middle to Late Jurassic evolution of the Klamath Mountains, California: Geological Society of America Bulletin, v. 100, p. 29–44, doi: 10.1130/0016-7606(1988)100<0029:TDEORA>2.3.CO;2.
  83. Wyld, S.J., and Wright, J.E., 2001, New evidence for Cretaceous strike-slip faulting in the United States Cordillera and implications for terrane-displacement, deformation patterns, and plutonism: American Journal of Science, v. 301, p. 150–181, doi: 10.2475/ajs.301.2.150.
  84. Wyld, S.J., Umhoefer, P.J., and Wright, J.E., 2006, Reconstructing northern Cordilleran terranes along known Cretaceous and Cenozoic strike-slip faults: Implications for the Baja British Columbia hypothesis and other models, in Haggart, J.W., Enkin, R.J., and Monger, J.W.H., eds., Paleogeography of the North American Cordillera: Evidence for and against Large-Scale Displacements: Geological Association of Canada Special Paper 46, p. 277–298.

top