+ All Categories
Home > Documents > Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging...

Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging...

Date post: 14-Mar-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
22
Geosphere doi: 10.1130/GES00790.1 published online 16 November 2012; Geosphere H. Gilbert, Y. Yang, D.W. Forsyth, C.H. Jones, T.J. Owens, G. Zandt and J.C. Stachnik Imaging lithospheric foundering in the structure of the Sierra Nevada Email alerting services articles cite this article to receive free e-mail alerts when new www.gsapubs.org/cgi/alerts click Subscribe to subscribe to Geosphere www.gsapubs.org/subscriptions/ click Permission request to contact GSA http://www.geosociety.org/pubs/copyrt.htm#gsa click official positions of the Society. citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect presentation of diverse opinions and positions by scientists worldwide, regardless of their race, includes a reference to the article's full citation. GSA provides this and other forums for the the abstracts only of their articles on their own or their organization's Web site providing the posting to further education and science. This file may not be posted to any Web site, but authors may post works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent their employment. Individual scientists are hereby granted permission, without fees or further Copyright not claimed on content prepared wholly by U.S. government employees within scope of Notes articles must include the digital object identifier (DOIs) and date of initial publication. priority; they are indexed by GeoRef from initial publication. Citations to Advance online prior to final publication). Advance online articles are citable and establish publication yet appeared in the paper journal (edited, typeset versions may be posted when available Advance online articles have been peer reviewed and accepted for publication but have not © Geological Society of America as doi:10.1130/GES00790.1 Geosphere, published online on 16 November 2012 as doi:10.1130/GES00790.1 Geosphere, published online on 16 November 2012
Transcript
Page 1: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Geosphere

doi: 10.1130/GES00790.1 published online 16 November 2012;Geosphere

 H. Gilbert, Y. Yang, D.W. Forsyth, C.H. Jones, T.J. Owens, G. Zandt and J.C. Stachnik Imaging lithospheric foundering in the structure of the Sierra Nevada  

Email alerting servicesarticles cite this article

to receive free e-mail alerts when newwww.gsapubs.org/cgi/alertsclick

Subscribe to subscribe to Geospherewww.gsapubs.org/subscriptions/click

Permission request to contact GSAhttp://www.geosociety.org/pubs/copyrt.htm#gsaclick

official positions of the Society.citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflectpresentation of diverse opinions and positions by scientists worldwide, regardless of their race, includes a reference to the article's full citation. GSA provides this and other forums for thethe abstracts only of their articles on their own or their organization's Web site providing the posting to further education and science. This file may not be posted to any Web site, but authors may postworks and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequenttheir employment. Individual scientists are hereby granted permission, without fees or further Copyright not claimed on content prepared wholly by U.S. government employees within scope of

Notes

articles must include the digital object identifier (DOIs) and date of initial publication. priority; they are indexed by GeoRef from initial publication. Citations to Advance online prior to final publication). Advance online articles are citable and establish publicationyet appeared in the paper journal (edited, typeset versions may be posted when available Advance online articles have been peer reviewed and accepted for publication but have not

© Geological Society of America

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012 as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 2: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Imaging lithospheric foundering in the structure of the Sierra Nevada

H. Gilbert1, Y. Yang2, D.W. Forsyth3, C.H. Jones4, T.J. Owens5, G. Zandt6, and J.C. Stachnik7

1Department of Earth, Atmospheric, and Planetary Sciences, Purdue University, 550 Stadium Mall Drive, West Lafayette, Indiana 47907, USA2ARC Centre of Excellence for Core to Crust Fluid Systems/GEMOC, Department of Earth and Planetary Sciences, Macquarie University, North Ryde, NSW 2109, Australia3Department of Geological Sciences, Brown University, 324 Brook Street, Box 1846, Providence, Rhode Island 02906, USA4Department of Geological Sciences, University of Colorado, Boulder, Colorado 80309, USA5Department of Earth and Ocean Sciences, University of South Carolina, 701 Sumter Street, EWS 617, Columbia, South Carolina 29208, USA6Department of Geosciences, University of Arizona, 1040 East 4th Street, Tucson, Arizona 85721, USA7Geophysical Institute, University of Alaska Fairbanks, 903 Koyukuk Drive, Fairbanks, Alaska 99775, USA

ABSTRACT

Tomographic studies of the mantle of southern California (USA) commonly found evidence for seismically high speed mate-rial, known as the Isabella anomaly, extend-ing from near the base of the crust of the southwestern Sierra Nevada foothills into the asthenosphere. This anomaly has been inter-preted to mark downwelling lithospheric material that had been removed from the southern Sierra Nevada. Using data from the Sierra Nevada EarthScope Project (SNEP) array, we investigate the lithosphere of the Sierra Nevada and surrounding region to better understand the process by which batholiths form dense lithospheric roots that become unstable and founder. Inverting phase velocities of fundamental mode Ray-leigh waves for shear wave speeds provides observations of the distribution of high and low wave-speed anomalies, which correspond to portions of the batholith that formed an intact lithospheric root, and where seismi-cally slower shallow asthenosphere marks areas where lithosphere has been removed. Our results corroborate previous observa-tions that the southern Sierra Nevada has thin crust underlain by shallow astheno-sphere. High shear wave velocity (Vs) mate-rial in the mantle beneath the southwestern foothills marks the location of the Isabella anomaly, to the east of which is a region of low Vs mantle where asthenosphere has risen to replace the delaminating root. Farther north, near the latitude of Long Valley, low velocities at shallow depths beneath the high elevations of the eastern Sierra indicate the presence of asthenosphere close to the base of

the crust. Thicker high-speed material, how-ever, underlies the western foothills of the Sierra Nevada at this latitude and dips to the east where it extends to depths of ~100 km or more, giving it the appearance of a portion of lithosphere that has detached from the east but remains attached to the west as it is cur-rently peeling off. The structure of the Sierra Nevada changes near the latitude of Lake Tahoe, where thinner lithosphere extends between depths of 40 and 80 km, but does not reach greater depths. It appears that the lithospheric material of the Sierra Nevada from latitudes close to Lake Tahoe, and con-tinuing to the north, is not being removed, indicating a change between the structure and evolution of the southern and northern Sierra Nevada.

INTRODUCTION

The process of lithospheric foundering or removal, involving the detachment of negatively buoyant portions of the lithosphere that then sink into the underlying mantle, remains poorly docu mented. Lithospheric removal has been pro-posed to have played a role in the development of a wide range of regions, including expan-sive areas such as the Tibetan Plateau (England and Houseman, 1989), the Altiplano (Kay and Mahlberg-Kay, 1991; Beck and Zandt, 2002), and the Alboran Sea (Seber et al., 1996; Cal-vert et al., 2000; Dündar et al., 2011), and more focused locations such as the Basin and Range (Platt and England, 1994; Humphreys, 1995; West et al., 2009), the Colorado Plateau (Bird, 1979; Karlstrom et al., 2008; Levander et al., 2011), the Wallowa Mountains in northeastern Oregon (Hales et al., 2005), and the southern

portion of the Sierra Nevada batholith (Ducea and Saleeby, 1996; Saleeby, 2003; Zandt et al., 2004; Le Pourhiet et al., 2006; Saleeby et al., 2012). The process may also play a signifi cant role in continental crust becoming more felsic as denser mafi c material recycles back into the mantle (Kay and Mahlberg-Kay, 1991; Saleeby et al., 2003). The Sierra Nevada is of particu-lar interest because there is evidence that some portion of its uplift is young. If its uplift were related to the loss of a dense residue root dur-ing the past 10 Ma, the lithospheric foundering process would have occurred recently enough (and may be ongoing) that evidence related to the removal process likely remains. Studying the structure of the Sierra Nevada could pro-vide insight into the signature of material being removed from the continental lithosphere and the conditions under which this process occurs.

Here we present a model of shear wave veloci-ties of the crust and upper mantle of the Sierra Nevada based on phase velocities of teleseismic Rayleigh waves. Using a dense array of seis-mometers in eastern California to record these Rayleigh waves allows us to produce a model that includes variations in velocities over short distances. As Rayleigh waves are sensitive to structures in the lower crust and upper mantle they are particularly well suited for studying lithospheric foundering. Because colder litho-sphere has higher shear wave velocities than warm asthenosphere, differences in shear wave velocity anomalies provide insights into which locations of the Sierra Nevada have higher speed lithospheric material and where lithosphere has been replaced by lower speed, warm buoyant asthenosphere. This distinction helps illuminate how pervasive lithospheric foundering has been across the Sierra Nevada. Identifying patterns

For permission to copy, contact [email protected]© 2012 Geological Society of America

1

Geosphere; December 2012; v. 8; no. 6; p. 1–21; doi:10.1130/GES00790.1; 11 fi gures; 1 table; 4 supplemental fi les.Received 17 January 2012 ♦ Revision received 21 July 2012 ♦ Accepted 6 August 2012 ♦ Published online 16 November 2012

Geodynamics and Consequences of Lithospheric Removal in the Sierra Nevada, California themed issue as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 3: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

2 Geosphere, December 2012

in the velocity structure and comparing those to geologic observations (e.g., Ducea and Saleeby, 1996; Manley et al., 2000; Cassel et al., 2009; Cecil et al., 2012; Saleeby et al., 2013) and dynamic models (e.g., Le Pourhiet et al., 2006; Saleeby et al., 2012) may help to discriminate whether the removal process occurred over a wide area during a single event or if it gradually progressed across the Sierra Nevada. Answers to these questions will constrain the conditions under which lithospheric foundering occurs.

SIERRA NEVADA BACKGROUND

As the Farallon plate subducted beneath North America during the Mesozoic, leading to an active arc along the margin of North Amer-ica, accretion of island arcs welded volcanic and sedimentary rocks onto the continental margin. Plutonic bodies later intruded these accreted terranes, forming the Sierra Nevada batholith (Dickinson, 1981; Schweickert, 1981; Dickin-son, 2008). A dense concentration of granitic plutons composes the southern and central sections of the Sierra Nevada batholith; more sparsely distributed plutons are to the north (Bateman and Eaton, 1967; Bateman, 1988). The composition of the batholith varies from quartz diorite or granodiorite in the east to more mafi c lithologies in the west (Bateman and Eaton, 1967). A large part of the magmatism responsible for the formation of the southern portion of the batholith was primarily confi ned to two distinct episodes, 160–150 and 100–85 Ma (Ducea, 2001). Between those episodes, the westernmost zone of the Sierra Nevada was emplaced during the Early Cretaceous from 140 to 130 Ma along the transition between the western foothills of the Sierra Nevada and the Great Valley (Saleeby, 2007). Subsequently, the Fine Gold Intrusive Suite, one of the larg-est intrusive complexes in the Sierra Nevada batholith, was emplaced between 124 and 105 Ma (Lackey et al., 2012). During that time there were pulses of magmatism in the northern Sierra between ca. 130 and 100 Ma (Cecil et al., 2012). These events serve to illustrate that the Sierra Nevada batholith formed continuously over an extended time period through the Cre-taceous (Cecil et al., 2012).

The overall composition of the Sierra Nevada changes northward as the propor-tion of intruded plutons within the country rock decreases. In addition, the tectonic set-ting into which the batholith was emplaced changes from continental rocks in the south to oceanic rocks in the north. The northern part of the Sierra Nevada has a lower mean eleva-tion than the south, and the deepest portions of the batholith are exposed in the south (Ague

and Brimhall, 1988; Cassel et al., 2009). The pattern of heat fl ow aligns with topography; areas of high heat fl ow are within the higher elevations, while lower elevations correspond to areas of lower heat fl ow (e.g., Saltus and Lachenbruch, 1991; Blackwell and Richards, 2004). Paleozoic and Mesozoic metavolcanic and metasedimentary roof pendants, which are present in the central Sierra, and Cenozoic volcanic rocks, which are well preserved in the north, are all absent in the south.

During the formation of granitic batholiths like the Sierra Nevada, in which granitoid plu-tons compose the majority of the upper crust, there must be a complementary production of a thick, deeper section of garnet pyroxenites that form as residues of the felsic granitoids (Ducea, 2001). Xenolith data indicate that the residue material becomes signifi cantly denser at ~1.5 GPa (depths of ~45 km) as it transitions from a granulite to an eclogite facies assem-blage (Ducea, 2002). Therefore only suffi ciently thick arcs should form an unstable lithospheric root, while thinner arcs would not. The term “arclogite,” developed from the work of Ander-son (2005), refers to the combined eclogitic cumulates and interlayered spinel-garnet peri-dotites that form within arcs. The existence of an arclogite root has been demonstrated for the southern Sierra from garnet-rich xenoliths entrained in 12–8 Ma basalt. No lavas younger than 4 Ma in the southern Sierra have been found to host arclogite xenoliths, suggesting delamination of the dense lower crust between 10 and 4 Ma (Ducea and Saleeby, 1996). The timing, consequences, and dynamics of litho-spheric removal have been explored in a series of thermomechanical models by Le Pourhiet et al. (2006) and Saleeby et al. (2012) that were constructed for the specifi c case of the Sierra Nevada. The seismic observations presented here can be compared to predictions of these models to gain insight into the processes that led to the current distribution of seismic velocities.

The southern edge of the subducting Juan de Fuca plate migrated northward past the south-ern Sierra Nevada at 20–15 Ma (Atwater and Stock, 1998), the time when the ancestral Cas-cade arc became active in the area to the north of Yosemite National Park and south of Lake Tahoe (Fig. 1), as indicated by the westward sweep of basaltic andesite and dacite magmas (Busby et al., 2008; Cousens et al., 2008; Busby and Putirka, 2009). The southern edge of the Juan de Fuca plate would have migrated past the central Sierra by 5 Ma, and currently is to the north of Lake Tahoe (Atwater and Stock, 1998). Once the subducting Farallon plate migrated north, the lower crust and lithosphere of the southern Sierra Nevada would no longer

be supported from below and could mechani-cally detach into the mantle with buoyant asthenosphere then rising to replace it. With the northward migration of the southern edge of the Juan de Fuca plate, the west coast of California transitioned from a convergent to a transform margin. Eruptions of basaltic magmas begin-ning in the Miocene in the southern and central Sierra and adjacent Basin and Range support the presence of shallow, upwelling astheno-spheric material in the area (e.g., Manley et al., 2000). Episodes of potassium-rich volcanism during the Pliocene in the central and eastern Sierra may indicate melting and remobilization of the Precambrian mantle lithosphere. Alterna-tively, the melting of metasomatized peridotite beneath the lithosphere has also been proposed to produce the potassic volcanism (Elkins-Tanton and Grove, 2003), which was followed by basaltic and bimodal eruptions across the eastern Sierra and Walker Lane (Manley et al., 2000; Farmer et al., 2002).

The changing composition of xenoliths that erupted over time provides strong evidence that some portions of the Sierra Nevada lithosphere were removed. Xenoliths that erupted within the central part of the range during the Late Mio-cene indicate that, at that time, the lithosphere had a thick section of arclogite. This arclogite would have formed at depths of 40–90 km as a residue, or cumulate, of the granitoids that com-pose the upper portions of the batholith. The composition of xenoliths entrained in younger volcanics that erupted along the eastern side of the range in the Pliocene and Pleistocene dif-fers from the older xenoliths in that they do not contain garnet pyroxenites, instead having high-temperature spinel peridotites that originated from similar depths (Ducea and Saleeby, 1996, 1998). Such a compositional change could result from the younger xenoliths sampling warm asthenospheric material instead of lithosphere. Questions persist regarding how the lithosphere beneath the Sierra Nevada has been modifi ed to lead to the differences in xenolith composition.

Asymmetry of the Sierra Nevada structure is evident at the surface as well as at depth. High elevations along the southeastern portion of the mountain range abruptly drop into the much lower Owens Valley. Conversely, on its western side elevations gradually diminish, with a near constant slope from the high Sierra down to the Great Valley. Well cores from the Great Valley reveal that the batholithic basement extends at least half the width of the valley (Saleeby et al., 2009). Controlled source seismic data collected across the southern Sierra Nevada show low crustal velocities to depths of ~30 km and have been interpreted to indicate that the granitic plutons found at the surface are deep seated

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 4: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 3

(Fliedner and Ruppert, 1996; Ruppert et al., 1998; Fliedner et al., 2000). The signature of the Moho changes from a sharp boundary at 30 km beneath relatively thin crust on the eastern side, to a thick gradual transition at 50 km beneath the western foothills (Zandt et al., 2004; Fras-setto et al., 2011). Low mantle velocities have been identifi ed beneath the eastern Sierra along its margin with Owens Valley (e.g., Jones, 1987; Schmandt and Humphreys, 2010). Tomo-graphic studies of the mantle (Aki, 1982; Biasi and Humphreys , 1992; Benz and Zandt, 1993; Boyd et al., 2004; Yang and Forsyth, 2006a; Burdick et al., 2009; Biasi, 2009; Schmandt and Humphreys , 2010) found evidence for a high-velocity body, known as the Isabella anomaly (Fig. 1), extending from near the base of the crust of the southwestern Sierra foothills to >200 km deep. Magnetotelluric studies simi-larly identifi ed a resistive region that extends from the base of the crust to depths of ~250 km near the boundary between the Sierra Nevada and the Great Valley (Park, 2004). This high-velocity, resistive body has been interpreted to mark downwelling mantle and lithospheric material that had been removed from the south-ern Sierra (Boyd et al., 2004; Zandt et al., 2004). The Isabella anomaly has also been interpreted as a dehydrated remnant piece of the subducted Farallon plate (Forsyth et al., 2011). A similarly high velocity anomaly, originally called the

Redding anomaly, appears in the upper mantle of northern California beneath the Sierra and northernmost Great Valley. Body wave tomog-raphy shows that it is at the southern end of the subducting Juan de Fuca slab (e.g., Benz and Zandt, 1993; Burdick et al., 2009). However, in Jones et al. (2004) it was noted that the increased magnitude of this anomaly compared to other portions of the slab farther to the north could result from the Redding anomaly also contain-ing descending Sierra Nevada lithosphere.

The lack of thick crust beneath the south-eastern portion of the Sierra Nevada suggests the need for buoyant mantle to explain much of the ~150 mGal negative Bouguer anomaly associated with the high Sierra (Oliver, 1977).

Early geophysical studies suggested a shallow lithosphere-asthenosphere boundary beneath the Moho of the eastern Sierra Nevada based on seismic and heat-fl ow observations (e.g., Crough and Thompson, 1977; Jones, 1987). Explanation of these and other geophysical signatures in the southern and central Sierra Nevada involved the removal of an arclogite root from beneath the southern Sierra Nevada since ca. 10 Ma, leading to the observed tectonic and magmatic response. However, the details of the lithospheric removal process remain unclear: what conditions contribute to the for-mation of an unstable root? Does the lithosphere delaminate and peel away, or does it drip off as a Rayleigh-Taylor type instability? Has removal

−124° −123° −122° −121° −120° −119° −118° −117°

34°

35°

36°

37°

38°

39°

40°

41°

42°

Long Long

ValleyValley

Ow

ens Valley

Ow

ens Valley

Great Valley

Garlock Fault

Garlock Fault

Basin and

Range

Mt.Lassen

LakeTahoe

Yosemite

Yosemite

N. P.N. P.

eastern C

alifornia

shear zon

e

Walker Lane

Go

rda

slab

western Foothills

San And

reas Fault

Kern CanyonKern CanyonFaultFault

NevadaCalifornia

Mojave

Transverse Ranges

Coast Ranges

San Francisco

Bay

Redding

anomaly

Isabella

anomaly

Figure 1. Topographic map of the Sierra Nevada study spanning from north of the Mendocino Triple Junction to the Salton Trough in the south. Red triangles—loca-tions of SNEP (Sierra Nevada EarthScope Project) stations contributing data; inverted yellow triangles—SPE (Sierra Paradox Experiment); blue diamonds—other net-works including the USArray Transport-able Array, Caltech Regional Network, Berkeley Digital Seismic Network, Leo Brady Network, Nevada Regional Network, and the United States National Seismic Net-work. Red dashed circles show the locations of the previously identifi ed Redding and Isabella anomalies. The locations of major tectonic provinces mentioned in the text are labeled as well as locations of prominent seismic anomalies. Black triangles outlined in red indicate the locations of recent vol-canism in Long Valley, Big Pine, Golden Trout, and Coso from north to south. Thick gray lines indicate the locations of the San Andreas, Garlock, and Kern Canyon faults. Red shading shows the location of Yosemite National Park (N.P).

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 5: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

4 Geosphere, December 2012

occurred along the entire Sierra Nevada or is it confi ned to the southern part? If removal does extend beyond the southern Sierra, did it occur all at once, or is it a time-transgressive feature?

The Sierra Nevada EarthScope Project (SNEP) array was designed to investigate lithospheric structure of the Sierra Nevada and further our understanding of the lithospheric removal process. Using data from the SNEP array, we identify which parts of the batholith formed a lithospheric root that remains intact and where lithosphere has been replaced by shal-low asthenosphere. The dense station spacing of the SNEP array permits identifying variations in seismic structures in more detail than would be possible with an array of more sparsely spaced stations. We are particularly interested in distin-guishing between areas of cold, dense, seismi-cally fast lithosphere and arclogite from seis-mically slower, warm, buoyant asthenosphere.

SURFACE WAVE DATA

We investigate lithospheric structure using fundamental mode Rayleigh waves from more than 150 earthquakes with magnitudes >5.0 recorded by various arrays in eastern California with epicentral distances between 30° and 120°. The Rayleigh waves were recorded by stations from permanent networks and past temporary arrays that spanned the Sierra Nevada, includ-ing the Sierra Paradox Experiment (Jones et al., 1994; Boyd et al., 2004) and SNEP (Gilbert

et al., 2007) temporary arrays and the neigh-boring Transportable Array component of the EarthScope USArray. The SNEP array com-prised more than 50 broadband stations, the majority of which occupied 2 locations dur-ing a 30 month recording interval between the spring of 2005 and the fall of 2007. The average spacing between stations across the SNEP array area is ~25 km and spans the area covering the Sierra Nevada batholith from Mount Lassen in the north to just north of the Garlock fault in the south and extends between the Great Valley and the Basin and Range (Fig. 1).

In order to measure phase velocities at a range of frequencies, we apply a series of nar-row band-pass fi lters centered at frequencies between 10 and 50 mHz (periods between 20 and 100 s) to vertical seismograms that have been recorded by the seismometers used in this study. Because a heterogeneous collection of seismometers recorded the data analyzed here, all of the waveforms were transferred to the instrument responses of an STS-2 broadband seismometer. Following visual inspection of each of the waveforms, we only retain traces for which the signal:noise ratio of the Rayleigh wave was at least 10:1. More details of the data selection and processing procedures were given in Yang and Forsyth (2006a). Maps of ray cov-erage illustrate the dense sampling achieved in this study from all backazimuths, leading to numerous crossing rays within and along the perimeter of the array of stations (Fig. 2). Fol-

lowing removal of noisy seismograms, the sam-pling of the clean data set varies from period to period, the longest and shortest periods having fewer events with suffi cient signal to make reli-able phase velocity measurements (Table 1). Differences in the ray coverage between differ-ent frequencies illustrate the variations in sam-pling from the moderately well sampled shorter period data (30 s), to the well-sampled inter-mediate periods (50 s), and the more sparsely sampled longer periods (100 s) (Fig. 2).

Waveforms from surface waves that have tra-versed tectonic boundaries or through tectonically active areas exhibit changes in amplitude across the array of stations. These amplitude differences

−124° −123° −122° −121° −120° −119° −118° −117° −124° −123° −122° −121° −120° −119° −118° −117°

50 s period50 s period50 s period50 s period 100 s period100 s period100 s period100 s period−124° −123° −122° −121° −120° −119° −118° −117°

34°

35°

36°

37°

38°

39°

40°

41°

42°

30 s period30 s period30 s period30 s period

BBB CCCCAAA

Figure 2. Map of the great circle ray paths for periods used in this study. (A) 30 s. (B) 50 s. (C) 100 s. Note the dense crossing paths within the array of stations and the reduction in number of ray paths due to fewer events producing Rayleigh waves with clear waveforms at periods above and below 50 s.

TABLE 1. PHASE VELOCITIES FOR PERIODS BETWEEN 20 AND 100 s MEASURED BY

COMBINING DATA FROM THE ENTIRE STUDY AREA

Period(s)

Nobs Mean phase velocity (km/s)

std(km/s)

20 2626 3.429 2.5 × 10–3

22 3023 3.489 2.3 × 10–3

25 2632 3.566 2.6 × 10–3

27 4600 3.605 2.0 × 10–3

30 3811 3.656 2.1 × 10–3

34 5668 3.701 1.7 × 10–3

40 6002 3.744 1.8 × 10–3

45 6921 3.773 1.7 × 10–3

50 6938 3.791 1.9 × 10–3

59 6479 3.821 1.9 × 10–3

67 6162 3.855 2.0 × 10–3

77 6102 3.892 2.3 × 10–3

87 5646 3.928 2.8 × 10–3

100 5072 3.974 3.5 × 10–3

Note: Nobs—number of raypaths observed; STD—standard deviation.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 6: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 5

can result from multipathing, scatter ing, or focusing and defocusing of the wavefi eld as a result of topography or velocity hetero geneities along the ray path. Figure 3 illustrates seismo-grams recorded by Transportable Array stations for an earthquake that occurred ~9800 km south-west of the study area. The effects of focusing and defocusing of the wavefi eld can be observed in the changing amplitudes of the waveforms that vary between smaller arrivals to the north and larger arrivals in the south.

To account for complexities of the incoming Rayleigh waves, the processing employed here treats the incoming wavefi eld as two interfer-ing plane waves; this accounts for the wavefi eld arriving off its projected great circle path and the interference effects that produce amplitude variations. We construct our two-dimensional phase velocity maps following the methods in Yang and Forsyth (2006b; methods built on those in Forsyth and Li, 2005), wherein the amplitude and phase of Rayleigh waves are inverted for lateral variations in phase veloc-ity and azimuthal anisotropy on a grid of nodes distributed across the study area. We use two-dimensional sensitivity kernels that account for the fi nite frequency effects on both the phase and amplitude variations of the surface waves; this leads to increased resolution (Yang and Forsyth, 2006b). The kernels account for the concentrated sensitivity of the Rayleigh waves within the fi rst two Fresnel zones that decreases as the kernels broadens along the ray path with increasing distance from the station. Higher order Fresnel zones are relatively unimportant

because the windowing of the seismograms leads to a fi nite effective bandwidth at each frequency producing destructive interference in the outer zones. The phase velocity map is constructed through an iterative process start-ing with a grid of nodes initially spaced by 1° and refi ned during successive iterations to 0.8°, 0.6°, and 0.4°, or ~44 km. The grid of nodes extends beyond the array of stations such that any traveltime anomalies from outside the array that are not accounted for in the two-plane wave approach can be mapped onto these outer nodes. The grid nodes used to solve for azimuthal anisotropy terms remain fi xed with a spacing of 1°. Because the dominant anisotropic terms for Rayleigh waves are the 2θ variations (Smith

and Dahlen, 1973), in which θ is the direction of propagation, we restrict the inversion for azi-muthal anisotropy to these terms.

DISTRIBUTION OF PHASE VELOCITIES

Rayleigh wave phase velocities provide information of the distribution of hetero-geneous seismic velocities across the Sierra Nevada. To accurately account for rapid large lateral changes in seismic wave speeds asso-ciated with variations in crustal thickness or other heterogeneities, we construct an initial three-dimensional starting model within the study area based on previous seismological

−124° −123° −122° −121° −120° −119° −118° −117°

34°

35°

36°

37°

38°

39°

40°

41°

42°Figure 3. Example Rayleigh waveforms from an earthquake at a distance of ~9800 km, with a magnitude of 6.1, and a back azimuth to the southwest used in this study and fi l-tered between 0.015–0.025 Hz (correspond-ing to a period of 50 s). The waveforms are plotted at the locations where they were recorded. In order to avoid plotting overlap-ping waveforms, only arrivals recorded by Transportable Array stations that were used for this event are shown. Each trace is ori-ented along its great circle path between the station where it was recorded and the event. A common 400 s time window is presented for each trace and each trace is plotted at the same scale. Note the earlier arrivals (toward the right side of the waveforms) at the closer stations. The effects of focusing and defocusing can be observed from the variations in the amplitudes of the Rayleigh waves between the northern and southern parts of the array.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 7: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

6 Geosphere, December 2012

studies of crustal velocities (Moschetti et al., 2010) and thicknesses (Frassetto et al., 2011; Gilbert, 2012). Using this initial model, we calculate predicted phase velocities for each period at each node location throughout the region, assuming the mantle everywhere to uniformly have the structure of the TNA (Tectonic North America) model (Grand and Helmberger, 1984). The phase velocity maps are then calculated using these predicted phase velocities as the starting model at each node location for each period (Rau and Forsyth, 2011). This procedure minimizes inaccura-cies resulting from mapping crustal features into the mantle and helps identify structural changes over shorter distances than could be identifi ed using a uniform starting model dur-ing the phase velocity inversion (Wang et al., 2009; Rau and Forsyth, 2011).

Because the lateral resolution for Rayleigh waves at longer periods is not as fi ne as at shorter periods, the longer periods are effectively more heavily damped when inverted on a fi ne grid. Subsequent inversions for vertical shear veloc-ity structure can then yield artifacts in anoma-lous regions because the relative amplitude of the phase velocity anomalies at different peri-ods will be altered. Phase velocities are there-fore fi rst determined on a grid of nodes spaced 1° apart with linear interpolation of velocities between nodes. At this spacing all periods have roughly the same resolution as measured by the rank of the resolution matrix. The phase veloci-ties are inverted point by point to construct a new three-dimensional shear velocity model with the initial model serving as the starting model. The phase velocities are then iteratively refi ned on grids for which the spacing between nodes decreases to 0.8°, 0.6°, and 0.4° (Fig. 4). The starting phase velocity model at each subse-quent step is the set of phase velocities predicted for the new shear velocity model calculated at each step, rather than the actual observed phase velocities of the previous step. Using predicted rather than observed velocities as starting mod-els reduces the period to period oscillations that can accumulate in an iterative inversion, effec-tively emphasizing the better resolved portions of the model. A minimum length constraint and a strong minimum curvature constraint that penalize variations from the starting model are imposed during both the phase and shear veloc-ity inversions. At each step, the phase velocities are retained only for those periods with similar lateral resolution (defi ned as having rank ≥80% of the maximum rank for phase velocities at any period). Consequently, the longer periods are progressively eliminated as the grid is refi ned, although information from the longer periods is retained through its infl uence on the refi ned

starting models. At the fi nal grid spacing of 0.4°, only periods of 50 s and less are retained. This iterative approach has the advantages of largely eliminating underestimates of velocity anoma-lies caused by damping and of increasing lateral resolution at shallow depths constrained by the shorter periods. However, it must be recognized that both vertical and lateral resolution decay with increasing depth.

The effects of closer grid spacing and the use of iteratively refi ned starting models on the phase velocity inversions can be seen comparing the 30 s phase velocities inverted on a grid with nodes spaced from 1.0°, 0.8°, 0.6°, and 0.4° (Figs. 4A–4D). Smaller features that do not appear in inversions using more largely spaced grids (Fig. 4A) can be detected in the grids using fi ner spacing and the velocity gradients sharpen, such as in the vicinity of the Garlock fault (Fig. 4D). The large-scale trend of higher phase veloc-ities to the west of, and within, the Sierra Nevada compared to the east remains.

Variations in phase velocities that are greater than approximately two times their standard deviation from the starting model can be viewed as statistically signifi cant. We limit our inter-pretation of the phase velocity maps to areas that are suffi ciently well sampled that the stan-dard deviation of velocities for each grid node is <0.05 km/s for the well-sampled 30 s phase velocity map produced using the fi nal 0.4° grid (Figs. 5A–5C). For periods sampled by a large number of Rayleigh wave paths, the phase velocity inversions have regions that satisfy our standard deviation cutoff that encompass an area extending beyond California into Nevada (Fig. 5B). However, for the shortest and longest peri-ods used here, where larger amounts of noise and scattering lead to less useable data, the area with acceptable standard deviations is within a smaller region centered on the densely sampled Sierra Nevada (Fig. 5C). The standard devia-tions increase with increasing fi neness of the grid, which means that in the iterative inversion for shear velocity structure, the a priori starting models are effectively given progressively more relative weight. Because the standard deviations do not account for correlated noise between stations, which could particularly be the case in areas sampled by the closely spaced SNEP stations, the standard deviation values presented here may slightly underestimate the actual stan-dard deviations (e.g., Wagner et al., 2010).

For the entire study area, the average phase velocities range between 3.43 km/s at 20 s period and 3.97 km/s at 100 s (Table 1; Fig. 6A). At shorter periods, the average values measured here are smaller than the average values mea-sured for southern California (Yang and Forsyth , 2006a), but our measurements approach similar

velocities to those of southern California at peri-ods >~50 s. At a period of ~50 s, the dispersion curve changes slope switching from concave down to concave up (Fig. 6A). Changes in the slope of phase velocity plots can be indicative of a shift in sensitivity from lithospheric to astheno spheric structures.

Differences between four dispersion curves created for locations interpolated between nodes on the 0.4° grid for periods up to and including 50 s, on the 0.6° grid for periods of 59–87 s, and on the 1.0° grid for 100 s period illustrate the effects of structural heterogeneity on phase velocities within the Sierra Nevada (Fig. 6A). Several signatures indicative of that heteroge neity can be identifi ed based on these dispersion curves. Phase velocities that are greater than those observed in other locations characterize the western foothills for periods between 33 and 67 s. Both the Great Valley and the western foothills have phase velocities that are higher than those found in the central Sierra Nevada or Owens Valley for periods <50 s. Between 30–40 s, which are most sensi-tive to the 40–60 km depth range, the low phase velocities in Owens Valley are reduced by more than ~3%–4% compared to the higher velocities in the western foothills. The dispersion curves converge at longer periods, indicative of the decreasing level of lateral variations at depth.

The distribution of high and low phase veloci-ties produced on the 0.4° grid for periods of 20, 27, 40, and 50 s and on the 0.6° grid for periods of 67 and 77 s provides insight into crustal and upper mantle structure within and surrounding the Sierra Nevada. At short periods between 20 and 27 s, the pattern of phase velocities displays high values along the western side of the Sierra Nevada and lower phase velocities along the eastern Sierra Nevada. The higher phase veloci-ties along the western portion of the batholith extend into the Great Valley (Figs. 7A, 7B). The northern extent of the batholith can be identifi ed at shorter periods where higher phase velocities marking the northern Sierra give way to lower phase velocities in the active arc near Mount Lassen (Figs. 4D and 7B). The high-velocity anomaly located to the north of the southern end of the Sierra Nevada at 36°N, 119.5°W straddles the boundary between the Sierra Nevada and the Great Valley at periods to 77 s (Fig. 7F). This anomaly extends to the west into the Great Val-ley and continues to the coast at shorter periods (Figs. 7A–7E).

The high phase velocities present in the western foothills of the central Sierra at shorter periods shift toward the northern and southern ends of the Sierra Nevada for periods >50 s (Figs. 7E, 7F). The locations of high phase velocities at these longer periods straddle the

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 8: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 7

A B

C D

Figure 4. (A) Map illustrating the distribution of phase velocities at a period of 30 s using a grid of nodes spaced 1°. Using a more fi nely spaced grid of nodes allows for smaller scale anomalies and sharper gradients to be detected that are not apparent on the more coarsely spaced grids. (B) 0.8°. (C) 0.6°. (D) 0.4°. To view a higher-resolution version of Figure 4, please visit http://dx.doi.org/10.1130/GES00790.S1.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 9: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

8 Geosphere, December 2012

boundary between the western Sierra Nevada and the Great Valley, toward the Gorda slab and Redding anomaly in the north and the Isabella anomaly in the south (Fig. 7E). The location of the northern anomaly shifts to the east for peri-ods >67 s, reaching 121°W at 77 s (Fig. 7F). The eastward shift of elevated phase velocities with increasing periods appears to indicate that the high-velocity material responsible for produc-ing these increased phase velocities dips to the east. The location of the highest phase veloci-ties associated with the Isabella anomaly also shifts to the east for periods >40 s, the small eastward drift indicating only a slight dip of the shear velocity anomaly in that direction (Figs. 7C–7F).

In contrast to the batholith, low phase veloci-ties mark the eastern boundary of the southern Sierra, where reduced phase velocities appear to wrap from the eastern side of the southern Sierra around to the west. Several areas that have undergone recent volcanism to the east of the Sierra Nevada are located in areas of reduced phase velocities in Owens Valley close to the area of the Long Valley Caldera (Figs. 7A, 7B). Lower phase velocities are to the north of the Sierra Nevada at shorter periods (e.g., Figs. 7A, 7B) and shift eastward with increasing period (Figs. 7C, 7D). The manner in which these low velocities remain east of the zone of high phase velocities that likely mark the Gorda slab sug-gests that they may correspond to low-velocity material above the slab.

In the Great Valley greater than average phase velocities continue from the western foothills into the valley and reach the coast at short peri-ods of 20 and 25 s (Figs. 7A, 7B). The elevated phase velocities in this region at shorter periods actually appear to increase to the west across the Great Valley into the central portion of the Coast Ranges. High phase velocities are also present in the area south of the Garlock fault and con-tinue southward into the Mojave province for periods between 20 and ~50 s (Figs. 7A–7D) and into the Transverse Ranges at greater peri-ods (Figs. 7E, 7F).

OVERVIEW OF FEATURES IN SHEAR WAVE VELOCITY MODEL

Because of the sensitivity of phase veloci-ties to the integrated Earth structures sampled by Rayleigh waves, phase velocities need to be inverted for shear wave speeds to obtain direct information of physical properties of the Earth at specifi c depths. Our phase velocity maps for 14 periods between 20 and 100 s serve as input to invert for shear wave velocity, Vs, at each grid node across the study area. Because the Rayleigh waves used in this study are primarily

−12

4°−

123°

−12

2°−

121°

−12

0°−

119°

−11

8°−

117°

0.03

0.03

0.03

0.030.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.06

0.06

0.06

0.06

0.060.06

0.06

0.07 0.0

7

0.08

0.090.1

0.11

12

−12

4°−

123°

−12

2°−

121°

−12

0°−

119°

−11

8°−

117°

0.04

0.04

0.04

0.040.04

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.06

0.06

0.06

0.06

0.06

0.060.06

0.06

0.06

0.06

0.06

0.07

0.070.08

0.090.1

0.11

50 s

per

iod

50 s

per

iod

50 s

per

iod

50 s

per

iod

77 s

per

iod

77 s

per

iod

77 s

per

iod

77 s

per

iod

BBBCCC

−12

4°−

123°

−12

2°−

121°

−12

0°−

119°

−11

8°−

117°

34°

35°

36°

37°

38°

39°

40°

41°

42°

0.02

0.02

0.02

0.03

0.03

0.03

0.03

0.03

0.03

0.03

0.03

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.04

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.05

0.06

0.07

0.08

0.09

30 s

per

iod

30 s

per

iod

30 s

per

iod

30 s

per

iod

AAA

Fig

ure

5. P

lots

of s

tand

ard

erro

r fo

r ph

ase

velo

city

inve

rsio

ns. (

A) A

t 30

s us

ing

the

0.4°

gri

d. (B

) At 5

0 s

usin

g th

e 0.

4° g

rid.

(C) A

t 77

s us

ing

the

0.6 °

gri

d. S

tand

ard

erro

rs r

ange

fro

m <

0.04

km

/s in

wel

l-sa

mpl

ed a

reas

in t

he c

ente

r of

the

arr

ay t

o >0

.1 k

m/s

in t

he le

ss-w

ell-

sam

pled

ar

eas

arou

nd t

he p

erim

eter

of

the

arra

y. N

ote

the

incr

ease

d ar

ea o

f lo

wer

sta

ndar

d er

rors

in t

he m

ap f

or t

he 5

0 s

data

as

com

pare

d to

the

m

ap f

or 7

7 s

data

. Her

e w

e co

nsid

er o

nly

area

s w

ith

stan

dard

err

ors

<0.0

5 km

/s o

n th

e 30

s m

ap u

sing

the

0.4

° gr

id.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 10: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 9

sensitive to vertical shear wave speeds (Vsv), we are only able to solve for Vsv and are insen-sitive to horizontal shear wave speeds (Vsh). To account for the slight sensitivity of Rayleigh waves to crustal compressional wave velocity, Vp, and density variations, we assume that these parameters vary proportionally to Vs scaled by factors of 1.2 and 0.3 (kg m–3)/(km s–1), respectively, instead of leaving them as poorly resolved free parameters (Rau and Forsyth, 2011). Shear wave speeds are determined by interpolating the phase velocities observed at each period onto a more fi nely spaced grid of 0.25°. The dispersion curve for each grid node is then inverted for the Vs profi le at that loca-tion using the model that was used to generate the predicted phase velocities at that period as the starting model. The inversion determines the necessary perturbations to the Vs starting model that will provide the best fi t between observed and predicted phase velocities (Saito, 1988). The rank of the resolution matrices and the reso-lution kernels for the shear wave inversion using the 1.0° spaced grid provide a rough indica-tion of the vertical resolution of the shear wave model presented here. In the well-sampled area in the vicinity of the SNEP stations, the rank of the resolution matrices is between 5 and 6, with the Rayleigh wave data contributing 2 pieces of information about crustal structure and 3–4 about mantle structure. The regions of peak sen-sitivity of the resolution kernels are ~20–30 km wide at a depth of 50 km and the kernels widen to ~50 km for 110 km depth.

The variations observed in dispersion curves for the locations along a profi le highlighted here

lead to changes in the velocity profi les from the Vs inversion (Fig. 6B). The velocity profi le for the TNA model (Grand and Helmberger, 1984) is plotted for comparison; it has a shallow high-velocity lid where Vs increases to >4.4 km/s at depths between 40 and 60 km, below which is a low-velocity zone where velocities reduce to 4.2 km/s between depths of 100 and 150 km. A sim-ilar high-velocity lid can be found in the upper mantle of the profi les for the Great Valley and the western foothills. However, these are clearly different from the profi les for Owens Valley and the eastern Sierra Nevada, which do not have the same high-velocity lid. The Vs in the upper mantle of Owens Valley is <4.2 km/s to depths of 70 km and gradually increases to a maximum of 4.25 km/s at 100 km depth, where it becomes indistinguishable from the other example pro-fi les. All 4 Vs profi les presented here have low-velocity zones where Vs decreases to a minimum of ~4.1 km/s at depths between 140 and 180 km. This comparison illustrates the relatively low wave speeds of the Sierra Nevada and California compared to the TNA model at the top of the mantle and between depths of 150–200 km.

Comparing the Vs values observed here to those observed by seismic investigations else-where and measurements of various lithologies allows us to discriminate between areas under-lain by cool lithospheric material and those underlain by warm asthenosphere. The Vs in the lithosphere is >4.5 km/s at ~100 km depth in the IASP91 (International Association of Seismology and Physics of the Earth’s Interior) seismic velocity model (Kennett and Engdahl,

1991). However, in the seismically slower tec-tonically active portion of North America, Vs in the lithosphere reaches just >4.4 km/s (TNA; Grand and Helmberger, 1984). Lithospheric shear velocities are further reduced in south-ern California where Yang and Forsyth (2006a) identifi ed average velocities of ~4.3 km/s above a lower velocity asthenosphere where velocities decrease to <4.1 km/s at 125 km depth. Simi-larly low shear wave velocities in the upper mantle of California have also been found by previous investigations of arrivals from local events (e.g., Savage et al., 1994) and surface wave studies (e.g., Moschetti et al., 2010; Lin et al., 2011). Shear velocities in the mantle <~4.3 km/s are diffi cult to generate as a result of only temperature or compositional varia-tions (e.g., Wang et al., 2009), unless the sheer quality factor (Qµ) for the area is 30 or less. On average, this low of a value of Qµ is not found in the asthenosphere beneath southern Califor-nia (Yang and Forsyth, 2008) or beneath the East Pacifi c Rise (Yang et al., 2007), which has similarly low Vs velocities. This pattern of low shear velocities with moderate attenuation (Qµ values of 50–60 in the asthenosphere) suggests that the velocities are affected by attenuation outside the band of seismic frequencies used here, most likely by the melt squirt mechanism, which would require <~1% melt (Hammond and Humphreys, 2000). Either melt (Stixrude and Lithgow-Bertelloni, 2005) or a solid-state mechanism that would lead to relaxation of the elastic moduli outside the seismic frequency band could explain the observed low mantle velocities.

Figure 6. (A) Examples of dis-persion curves, with their asso-ciated standard deviations, for four locations across the Sierra Nevada indicating the variabil-ity in phase velocities across the southern Sierra from the Great Valley (35.5°N, 119.5°W; blue), to the western foothills (36.0°N, 119.0°W; purple), to the mid-Sierra (36.0°N, 118.5°W; green), and Owens Valley (36.5°N, 118.0°W; red). Locations of sam-ple points are shown as gray cir-cles outlined in red in Figure 8A. The average dispersion curve calculated for the entire study area along with its standard deviation is in black. (B) Cor-responding shear wave velocity (Vs) profi les determined by inverting each of the dispersion curves. The color of each Vs profi le matches the dispersion curve for each location. The TNA (Tectonic North America; Grand and Helmberger, 1984) reference model is plotted in black for comparison.

36.0, –119.0western foothills

35.5, –119.5Great Valley

36.0, –118.5mid-Sierra

36.5, –118.0Owens Valley

A B

20 30 40 50 60 70 80 90 100

3.4

3.5

3.6

3.7

3.8

3.9

4

period (s)

phas

e ve

loci

ty (

km/s

)

2.7 2.9 3.1 3.3 3.5 3.7 3.9 4.1 4.3 4.5

0

20

40

60

80

100

120

140

160

180

200

Vs km/s

dept

h (k

m)

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 11: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

10 Geosphere, December 2012

DE

F

AB

C

Fig

ure

7. M

aps

of R

ayle

igh

wav

e ph

ase

velo

citi

es f

or s

ix p

erio

ds. (

A)

20 s

. (B

) 27

s. (

C)

40 s

. (D

) 50

s. (

E)

67 s

. (F

) 77

s. A

nom

alie

s di

scus

sed

in t

he t

ext

are

pre-

sent

ed in

the

pha

se v

eloc

ity

map

s. A

reas

of

posi

tive

pha

se v

eloc

ity

anom

alie

s ar

e sh

own

as b

lue

and

purp

le, a

nd a

reas

of

nega

tive

ano

mal

ies

are

show

n as

red

. To

vie

w a

hig

her-

reso

luti

on v

ersi

on o

f F

igur

e 7,

ple

ase

visi

t ht

tp:/

/dx.

doi.o

rg/1

0.11

30/G

ES0

0790

.S2.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 12: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 11

Crustal Structure of the Sierra Nevada and Surrounding Areas

In the upper crust, there are two primary features revealed by the Rayleigh wave tomog-raphy: high velocities in the western Sierra Nevada and low velocities in the sedimentary basins of the Great Valley. We show velocities in our model at 10 km (Fig. 8A), but note that given the depth resolution, these lateral veloc-ity anomalies represent an integrated effect from the surface to ~15 km. Some of the model struc-ture at this depth comes from the initial starting model that incorporates information from sur-face waves at shorter periods than we measure obtained by ambient noise analysis (Moschetti et al., 2010). The anomalies around the periph-

ery of the region are somewhat questionable, because at this depth, the primary control comes from the shortest period data, which are less numerous (Fig. 2), more scattered, and less well distributed in azimuth. The disruption of the Great Valley low-velocity anomaly at ~37°N is due to thinning of the sediment cover at that latitude. The high-velocity Sierra Nevada block terminates to the north in the vicinity of the Mount Lassen volcanic fi eld. In the south, it is truncated by the Garlock fault. Lower velocities lie east of the Kern Canyon fault where numerous extensional features have deformed the basement rocks of the Sierra Nevada and the seismicity indicates ongoing extension (Jones and Dollar , 1986; Mahéo et al., 2009). The highest velocity gradients generally are on the eastern side in the

vicinity of the normal faults just to the east of the crest of the Sierra Nevada. The high velocities of the shallow crust in the western Sierra Nevada can be attributed to lack of sediments and an intact, largely unfaulted crustal block with meta-morphic and plutonic rocks exposed at the sur-face. The Vs variations within, and surrounding, the Sierra Nevada can be viewed in cross sections perpendicular (Figs. 9A–9L) and parallel (Figs. 10W–10Z) to the axis of the mountain range.

At a depth of 25 km, representing the middle to lower crust, an area of high Vs (>3.9 km/s) extends along the western foothills of the cen-tral portion of the Sierra Nevada and continues across the Great Valley and Coast Ranges (Fig. 8B). Higher Vs values in the western foothills extend from the shallow crust to the mid-crust,

A B

Figure 8. Maps of crustal shear wave speeds. (A) At depth of 10 km. Gray circles with red outlines indicate locations of dispersion curves and velocity profi les in Figure 6. White lines show locations of cross-sections A–L and W–Z in Figures 9 and 10. (B) At depth of 25 km. Black circles show locations of magmatism younger than 1 Ma. The dashed white line marks the boundary identifi ed as the Mesozoic tectonic Moho by Frassetto et al. (2011) and the dashed black line marks the outline of the area they identifi ed to have the Mesozoic petrologic Moho. Note that different color scales are used in A and B. To view a higher-resolution version of Figure 8, please visit http://dx.doi.org/10.1130/GES00790.S3.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 13: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

12 Geosphere, December 2012

2.6

2.6

33

33

33.

43.

43.

43.

43.

43.

83.

83.

83.

83.

84

4

44

4

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

long

itude

(de

gree

s)

depth (km)

−12

1−

120

−11

9−

118

0 25 50 75 100

125

2.6

2.6

33

33

33.

43.

43.

43.

43.

43.

83.

83.

83.

83.

84

44

44

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

depth (km)

−12

0−

119

−11

8−

117

0 25 50 75 100

125

2.6

2.6

33

33

3.4

3.4

3.4

3.8

3.8

3.8

44

44

4.2

4.2

4.2

4.2

4.2

4.2

4.4

depth (km)

−12

0−

119

−11

8−

117

0 25 50 75 100

125

2.6

2.6

33

33

33.

43.

43.

43.

43.

43.

83.

83.

8

3.8

3.8

44

4

44

4.2

4.2

42

2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

4.4

−12

1−

120

−11

9−

118

2.6

2.6

2.6

33

33

33.

43.

43.

43.

43.

43.

83.

83.

8

3.8

3.8

44

4

44

4.2

42

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

−12

1−

120

−11

9−

118

2.6

2.6

33

33

3.4

3.4

3.4

3.4

3.4

3.8

3.8

3.8

3.8

3.8

44

4

44

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

long

itude

(de

gree

s)−

122

−12

1−

120

−11

9

BB′

CC′

EE′

FF′

DD′

AA′

KC

F

KC

F

IA

IA

GV

GV

SN

IA

AS

TH

AS

TH

OV

OV

SN

2.6

2.8

33.

23.

43.

63.

84

4.2

4.4

shea

r w

ave

velo

city

(km

/s)

Fig

ure

9 (o

n th

is a

nd f

ollo

win

g pa

ge).

(A

-A′–

L-L

′) C

ross

sec

tion

s ex

tend

ing

from

the

sou

ther

n to

nor

ther

n po

rtio

ns o

f th

e Si

erra

Nev

ada

orie

nted

from

the

sout

hwes

t to

the

nort

heas

t. H

oriz

onta

l lin

es a

re p

lott

ed a

t dep

ths

of 3

0 an

d 50

km

for

refe

renc

e. R

ed li

ne in

the

cros

s se

c-ti

ons

is t

he d

epth

of

the

Moh

o ba

sed

on c

rust

al t

hick

ness

obs

erva

tion

s of

Fra

sset

to e

t al

. (20

11)

and

Gilb

ert

(201

2) in

put

into

the

sta

rtin

g m

odel

for

the

she

ar w

ave

velo

city

(V

s) in

vers

ion.

Loc

atio

ns o

f cr

oss

sect

ions

are

in F

igur

e 8A

. AST

H—

asth

enos

pher

e, G

V—

Gre

at V

alle

y,

IA—

Isab

ella

ano

mal

y, K

CF

—K

ern

Can

yon

faul

t, L

ITH

—lit

hosp

here

, SN

—Si

erra

Nev

ada.

Not

e th

at th

e co

lor

scal

e us

ed o

n cr

oss

sect

ions

is

not

the

sam

e as

tha

t us

ed o

n th

e m

aps.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 14: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 13

Fig

ure

9 (c

ontin

ued)

.

2.6

2.6

33

33

3.4

3.4

3.4

3.4

3.4

3.8

3.8

3.8

3.8

3.8

44

44

4

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

depth (km)

−12

2−

121

−12

0−

119

0 25 50 75 100

125

2.6

2.6

33

33

33.

43.

43.

4

3.4

3.4

3.8

3.8

3.8

3.8

3.8

44

44

4

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

4.4

4.4

depth (km)

−12

2−

121

−12

0−

119

0 25 50 75 100

125

2.6

2.6

33

33

3.4

3.4

3.4

3.4

3.4

3.8

3.8

3.8

3.8

3.8

44

44

4

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

4.4

long

itude

(de

gree

s)

depth (km)

−12

3−

122

−12

1−

120

0 25 50 75 100

125

2.6

2.6

33

33

33.

43.

43.

43.

43.

4

3.8

3.8

3.8

3.8

3.8

44

44

4

4

4

4

4.2

4.2

4.24.

24.

24.

2

4.2

4.2

4.2

4.2

4.4

4.4

−12

3−

122

−12

1−

120

2.6

2.6

33

33

33.

43.

43.

43.

43.

4

3.8

3.8

3.8

3.8

3.8

44

44

4

4.2

4.2

4.2

4.24.2

4.2

4.2

4.2

4.4

4.4

−12

3−

122

−12

1−

120

2.6

2.6

2.6

33

33

33.

43.

43.

43.

43.

4

3.8

3.8

3.8

3.8

3.8

44

44

4

4.2

4.2

4.24.2

4.2

4.2

4.2

4.4

4.4

4.4

4.4

4.4

long

itude

(de

gree

s)−

124

−12

3−

122

−12

1

HH′

II′

KK′

LL′

JJ′

GG′

GS

GS

LIT

H LIT

H

LIT

H

2.6

2.8

33.

23.

43.

63.

84

4.2

4.4

shea

r w

ave

velo

city

(km

/s)

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 15: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

14 Geosphere, December 2012

2.6 3

33

33

3.4

3.4

3.4

3.4

3.4

3.8

3.8

3.8

3.8

3.8

44

44

4

42

4

4

4.2

4.24.

2

4.2

4.2

2

4.2

4.2

4.2

4.2

4.2

depth (km)

W

4140

3938

3736

35

0 25 50 75 100

125

W′

33

33

33.

43.

43.

43.

43.

43.

4

3.8

3.8

3.8

3.8

3.8

44

44

4

4

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

depth (km)

X

4140

3938

3736

35

0 25 50 75 100

125

X′

2.6

33

33

33.

43.

43.

43.

43.

43.

83.

83.

83.

83.

84

44

44

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

4.4

4.4

4.4

4.4

4.4

depth (km)

Y

4039

3837

3635

0 25 50 75 100

125

Y′

2.6

2.6

2.6

33

33

3.4

3.4

3.4

3.8

3.8

3.8

44

44

44.

2

4.2

4.2

4.2

4.2

4.2

4.2

4.2

4.4

4.4

4.4

4.4

latit

ude

(deg

rees

)

depth (km)

Z

4039

3837

3635

34

0 25 50 75 100

125

Z′

IA

AS

TH

AS

TH

MO

J

MO

J

MO

J

MO

J

GS GS

GS

IA

LIT

H

LIT

H

Fig

ure

10. (

W-W

′–Z

-Z′)

Cro

ss s

ecti

ons

exte

ndin

g fr

om t

he w

este

rn t

o ea

ster

n po

rtio

ns o

f th

e Si

erra

Nev

ada

orie

nted

fro

m t

he n

orth

wes

t to

the

sou

thea

st. H

ori-

zont

al li

nes

are

plot

ted

at d

epth

s of

30

and

50 k

m f

or r

efer

ence

. Red

line

is t

he d

epth

of

the

Moh

o in

put

into

the

sta

rtin

g m

odel

for

the

she

ar w

ave

velo

city

(V

s)

inve

rsio

n. L

ocat

ions

of

cros

s se

ctio

ns a

re i

n F

igur

e 8A

. AST

H—

asth

enos

pher

e, G

S—G

orda

sla

b, I

A—

Isab

ella

ano

mal

y, L

ITH

—lit

hosp

here

, M

OJ—

Moj

ave

anom

aly.

Col

or s

cale

is a

s in

Fig

ure

9. N

ote

that

the

col

or s

cale

use

d on

cro

ss s

ecti

ons

is n

ot t

he s

ame

as t

hat

used

on

the

map

s.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 16: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 15

where Vs is >3.9 km/s in locations of the central and northern foothills between lat ~36.5°N and 39°N (Fig. 8B). Within the Sierra Nevada there are increased Vs values in areas where crustal thicknesses are >50 km beneath the south and central portions of the western foothills (Fras-setto et al., 2011; Figs. 9B–9G). Receiver func-tions sampling the area of reduced velocity contrast resulting from the juxtaposition of high Vs lower crust overlying mantle with similar Vs would exhibit a diminished Moho signal (e.g., Figs. 9C–9F), as observed for the region along the western side of the Sierra (e.g., Zandt et al., 2004; Frassetto et al., 2011). In the mid-crust the region of elevated Vs continues to the west of the Sierra Nevada, through the Great Valley, and into the Coast Ranges, where Vs is >4.0 km/s (Fig. 8B). Elevated Vs values continuing from the western foothills into the Great Valley support previous observations that the Sierra Nevada batholith extends beneath the Great Val-ley (Saleeby et al., 2009).

While we are confi dent that the zone of high crustal Vs continues some distance into the Great Valley, the crust thins to the west to 30 km or even less. Because Rayleigh waves cannot resolve sharp discontinuities and have a vertical resolving length of ~20 km at this 25 km target depth, the high velocities in the lower crust in the western Great Valley and Coastal Ranges could be due to blurring of mantle structure into the lower crust. The importance of this blurring depends on the correctness of the depth and velocity contrast at the Moho in the initial start-ing model. The uncertainty in structures identi-fi ed to the west also increases due to the sparse station coverage outside of the Sierra Nevada.

The north-south extent of the 3.95 km/s contour along the western edge of the Sierra Nevada closely matches the north-south extent of the area Frassetto et al. (2011) identifi ed as the petro logic Moho formed in the Mesozoic (Fig. 8B), characterized by a weak velocity con-trast between the high-velocity, high-density mafi c lower crust and the mantle. The Mesozoic petro logic Moho in this area would result from a suffi cient amount of material being emplaced during batholith formation to produce a thick arclogite layer. East of the area where Vs in the mid-crust is >3.9 km/s, much of the eastern Sierra has low Vs, corresponding to the areas where the mountain ranges reach higher eleva-tions and the denser, high-velocity mafi c mate-rial has probably been removed. The lack of an arclogite root in this area supports the modeling predictions of Saleeby et al. (2012) that the root was removed from this area. Farther to the north along the western foothills, the northern limit of the 3.9 km/s contour appears to coincide with the termination of the region Frassetto et al. (2011)

labeled as having Mesozoic tectonic Moho, which they interpreted as corresponding to the original Moho prior to the intrusion of the Sierra Nevada batholith. The correlation between the location where the original Moho can still be identifi ed and the location of elevated Vs sug-gests that the crust in that area may be represen-tative of the crustal structure that the batholith intruded into and was not subsequently modifi ed by extension or lithospheric removal.

The other major features at 25 km are the pro-nounced low-velocity anomalies in the vicinity of recent magmatism (Fig. 8B), including the southern Cascades and the Long Valley, Big Pine, Golden Trout, and Coso volcanic fi elds east of the Sierra Nevada. The crust in these areas thins to between 30 and 40 km. Subdued low velocities continue to the south of Coso beneath the southernmost Sierra Nevada. The southern extent of the zone of low wave speeds continues along the eastern side of the Sierra and spreads to the southwest, where a zone of low Vs underlies areas of high elevation in the southern Sierra Nevada. Lake Tahoe marks the northern extent of the area of reduced Vs along the eastern portion of the Sierra Nevada. The east-west asymmetry becomes less pro-nounced to the north of Lake Tahoe where the northern Sierra lacks the signature of high Vs to the west and low Vs to the east (Figs. 9H, 9I). The northern Sierra does not have thick crust characterized by high wave speeds similar to that found in the western foothills farther south.

Upper Mantle Structure of the Sierra Nevada

The mantle beneath the Sierra Nevada also has asymmetric structures between the eastern and western sides of the mountain range. Simi-lar to the crust, differences in mantle structures vary from being pronounced in the southern Sierra to subtler in the northern Sierra Nevada. High upper mantle velocities where Vs is >4.4 km/s underlie the zone of thickened crust along the boundary between the southern Sierra and the Great Valley (Figs. 9B–9D). This is the loca-tion of the Isabella anomaly, which appears prominently in this Vs model and has zones of increased Vs that are >4.5 km/s at depths near 80 km beneath the western foothills. The high velocities associated with this anomaly extend from the base of the crust to depths of ~125 km beneath the southern Sierra between ~36.5°N and 37.5°N (Fig. 11E). The magnitude of the anomaly diminishes with depth and the highest velocities shift slightly to the east with increas-ing depth. Along the western part of the southern portion of the Sierra Nevada, these high veloci-ties appear to mark thick, high-wave-speed

material that remains in contact with the overly-ing batholith. In the 55 km depth range, a zone of high Vs extends northward from the southern Sierra along the western foothills to close to the latitude of Yosemite National Park (~38°N), where it shifts northward and extends toward the central portion of the batholith to latitudes just north of Lake Tahoe (Figs. 10Y, 10X, and 11A). The northward extent of the Isabella anomaly at depths between 55 and 70 km reaches ~37.5°N, as can be seen in the northwest-southeast–trend-ing cross section of the depth to the 4.3 km/s contour (Fig. 10Y) and in map view (Figs. 11A, 11B). Waveform modeling of P (compressional) waves from regional events indicates low veloci ties in the mantle beneath the southern Sierra Nevada and Walker Lane underlain by an abrupt transition to higher velocities at a depth of 75–100 km (Savage et al., 2003), thus a sepa-ration in that region between the batholith and high-velocity lithosphere.

Higher wave-speed material (Vs > 4.2 km/s) is in the uppermost mantle at depths of 60 km and greater beneath the western foothills under the area of thickened crust to the area north of the southern Sierra (Figs. 9B–9D and 11A–11E). The eastward dip of the Isabella anomaly can be discerned from the area of increased Vs reaching farther to the east at greater depths and shifting from the Great Valley to beneath the high elevations toward the central portion of the Sierra Nevada at depths >90 km for latitudes south of 37°N (Figs. 9C–9E and 11A–11C). Lower wave-speed mantle with Vs just >4 km/s is to the east of the Isabella anomaly and under-lies the region of high elevations in the eastern Sierra from the base of the crust to depths shal-lower than 70 km (Figs. 9B–9F, 11A, and 11B). Higher wave-speed mantle underlies the area of shallow lower wave speeds in the mantle along the eastern portion of the southern and central Sierra Nevada to as far north as Lake Tahoe (Figs. 10W and 11B–11E). These zones of lower wave speeds suggest the presence of partial melt and appear to correspond to the location of shal-low asthenosphere that replaced portions of the lithospheric root following its removal. This dis-tribution of velocities appears similar to predic-tions from thermomechanical modeling of the distribution of lithosphere and asthenosphere following the removal of the dense root of the Sierra Nevada. Following removal, the root would be beneath lower velocity asthenosphere that ascended to replace it (Le Pourhiet et al., 2006; Saleeby et al., 2012). Within the north-ern portion of Owens Valley at depths of 70 km and more, Vs remains slightly >4.2 km/s, while the area of lower Vs that is reduced to <4.2 km/s does not extend beneath the entire length of Owens Valley as it does at 55 km, and instead is

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 17: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

16 Geosphere, December 2012

DE

AB

C Fig

ure

11.

Map

s of

man

tle

shea

r w

ave

spee

ds.

(A)

At

dept

h of

55

km.

(B)

At

dept

h of

70

km.

(C)

At

dept

h of

90

km. (

D)

At

dept

h of

110

km

. (E

) A

t de

pth

of 1

25 k

m. G

S—G

orda

sla

b, I

A—

Isab

ella

ano

mal

y, M

oj—

Moj

ave

anom

aly,

TR

—T

rans

vers

e R

ange

ano

mal

y. T

o vi

ew a

hig

her-

reso

luti

on v

ersi

on o

f Fig

ure

11, p

leas

e vi

sit h

ttp:

//dx

.doi

.org

/10.

1130

/GE

S007

90.S

4.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 18: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 17

restricted to its southern end (Figs. 11B–11E). The presence of high Vs values at depths near 70 km beneath a body with lower Vs may cor-respond to lithospheric material at this depth range that is not present at shallower levels. The presence of descending portions of lithospheric material that have been removed and are now sinking into the asthenosphere (e.g., Le Pourhiet et al., 2006) could be responsible for this distri-bution of wave speeds.

A change in lithospheric structure coincides with the location where the Kern Canyon fault would project into the mantle, which appears to mark the western limit of a region of exten-sion in the eastern Sierra (Mahéo et al., 2009) (Figs. 1, 8B, 9B, 9C, and 11A). This location correlates with the point where the velocity contrast between the crust based on receiver function amplitudes changes from the large contrast observed in the eastern Sierra Nevada to the diminished contrast to the west (Frassetto et al., 2011). To the east of the Kern Canyon fault Vs diminishes in both the crust and upper-most mantle (Figs. 9B, 9C), suggesting a link between the shallow extension and alteration of the mantle lithosphere.

To the north of Yosemite, close to the lati-tude of Lake Tahoe, the difference in shear wave speeds between the eastern and western sides of the Sierra Nevada diminishes. The Vs increases to >4.2 km/s at depths to 70 km on both the western and eastern sides of the north-ern portion of the Sierra (Figs. 9I, 9J, 11A, and 11B). Similar to the pattern of Vs observed in the crust, the lack of low velocities in the mantle on the eastern side of the northern Sierra offers evidence that processes that shaped the structure of the southern Sierra Nevada did not operate in the same manner in the northern Sierra Nevada. The lack of low Vs in to the east of the northern Sierra is consistent with the decreased level of recent magmatic activity in the area to the north of Lake Tahoe. In addition, the high Vs litho-spheric material in the mantle to the north does not extend as deep as in the south, extending to only ~90 km at its deepest (Figs. 10W, 10X).

Toward the northern portion of the Sierra Nevada, areas of higher Vs in the mantle coin-cide with the location of the Gorda slab (Figs. 9J–9L and 11B–11E). The anomaly associated with the Gorda slab remains consistent with increasing depth (Figs. 9K, 9L, and 11A), in contrast to the Isabella anomaly, which appears as a stronger Vs anomaly at shallow depths and weakens in amplitude with greater depth. To the east of the high Vs anomaly associated with the Gorda slab is a region of decreased Vs that con-sistently shifts to the east as the position of the slab progresses to the east (Figs. 11C–11E). The fl ux of volatiles from the slab into the overly-

ing mantle wedge could be responsible for pro-ducing this low Vs region. The areas of high Vs imaged in the mantle here appear to be related to the Gorda slab and not to other high-speed material associated with another body that has been proposed to produce the Redding anomaly.

Structures Surrounding the Sierra Nevada

Toward the top of the upper mantle, to the south of the Sierra Nevada, the area of high Vs associated with the Isabella anomaly in the southern part of the Great Valley continues southward through the Mojave province and into the Transverse Ranges at a depth of 55 km (Figs. 10Z and 11A). The high Vs values also continue to the northwest from the Isabella anomaly into the Coast Ranges. High-wave-speed material has previously been identifi ed in the upper mantle beneath the Transverse Ranges (e.g., Hadley and Kanamori, 1977; Humphreys et al., 1984; Kohler et al., 2003; Yang and For-syth, 2006a). Similar to the results in Yang and Forsyth (2006a), the anomaly beneath the Transverse Ranges presented here diverges into two separate bodies that extend toward the east-ern and western ends of the range with increas-ing depths. The eastern anomaly migrates to beneath the Mojave province at greater depths. The form of the Transverse Range anomaly pre-sented here has a gap as the regions where the anomaly reaches its greatest amplitude diverge into two separate bodies at depths >70 km (Fig. 11B). This form of the Transverse Range anomaly supports the idea that it represents localized drips that could entrain, and remove, lithospheric material from the surrounding area (Yang and Forsyth, 2006a).

High-wave-speed material extends northward from the Transverse Range and across the Mojave province at depths between ~50 and 70 km (Figs. 10W, 10X, and 11A). Seismic observations of high subcrustal wave speeds beneath the Mojave province have been suggested to result from an increasing amount of oceanic crust within the North American lithosphere with depth (e.g., Fuis et al., 2003; Yan et al., 2005). Mantle xeno-liths from the central Mojave province also sug-gest that its lower lithosphere represents tectoni-cally subcreted and imbricated material from an oceanic protolith (Luffi et al., 2009). This sub-creted material could be a portion of the oceanic Farallon slab that accreted during low-angle subduction and currently composes part of the lithosphere beneath the Mojave province (Luffi et al., 2009). The region to the north of San Fran-cisco Bay is also an area of high Vs at the top of the upper mantle; however, the distribution of stations used here has better resolution to the east of the Great Valley. Still, it is interesting to

note the location of this anomaly, as results from earlier body wave tomography also identifi ed high seismic velocities in this area and attributed them to the presence of a small portion of the Pacifi c plate that subducted beneath the Coast Ranges (Benz et al., 1992).

The Vs images presented here indicate areas of higher Vs extending westward from the Isa-bella anomaly at the southern end of the Great Valley (e.g., Figs. 11A, 11B), particularly com-pared to those observed farther north in the valley (e.g., Figs. 11B–11D). Though the high velocities surrounding the Isabella anomaly do not appear to continue to the coast in the same manner as the high Vs region to the north of San Francisco Bay and the area associated with the Gorda slab, they appear to spread across the San Andres fault at depths of 70 km or less (e.g., Figs. 11A, 11B). The lack of sampling and diminished resolution along the coast (Figs. 5A–5C) prohibit us from determining whether the zone of high Vs extending to the west of the Isabella anomaly is a stalled segment of sub-ducting lithosphere (Forsyth et al., 2011). It is notable, however, that the stalled slab explana-tion for the high Vs material does not explain the stratigraphic history observed in the southern Great Valley (Saleeby and Foster, 2004; Saleeby et al., 2013) that fi ts well with the foundering root hypothesis. Connecting the crustal structure of the Great Valley to the Franciscan complex in the adjacent Coast Ranges suggests that a signifi cant slab of Franciscan eclogite and blue-schist accreted against the western margin of the arclogitic root and was subsequently drawn westward back out of the base of the Late Cre-taceous subduction zone (Saleeby et al., 2012). The presence of Franciscan eclogite and blue-schist material beneath the crust of the Great Valley offers an alternative explanation for the shallow zone of high Vs extending to the west of the Isabella anomaly.

DISCUSSION

As outlined above in the description of fea-tures in the Vs model, the Isabella anomaly imaged here extends to ~125 km depth and appears to dip to the east (Figs. 9C–9E, 11B, and 11C). Along the vertical profi le in the Vs model where the Isabella anomaly is close to the crust at shallow depths, the velocities actually decrease at greater depths (Figs. 6B and 9B). While there are concerns of oscillating patterns of high and low velocities when inverting phase velocities for Vs, that does not appear to be the case here; the dispersion curve for the western foothills in this area shows that phase veloci-ties for periods >40 s do not increase as rapidly with increasing period as they do for shorter

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 19: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

18 Geosphere, December 2012

periods (Fig. 6A). It therefore appears that Vs does not continue to increase with increasing depth below the shallow portion of the Isabella anomaly, suggesting a dip to the anomaly.

Observations of the dip of a the shallow por-tion of the Isabella anomaly do not agree with P- and S-wave tomography that identifi ed the portion of the Isabella anomaly where the mag-nitude of the anomaly is greater to be more of a vertical structure to depths of ~150 km and dip to the northeast at greater depths (e.g., Schmandt and Humphreys, 2010). A hetero-geneous distribution of anisotropy characterizes southern Cali fornia along the boundary between the Sierra Nevada and the Great Valley. Changes from high to low SKS splitting times (Bastow et al., 2006; Zandt and Humphreys, 2008) and discrepancies between the velocities of short-period Rayleigh and Love waves (Moschetti et al., 2010) indicate that this region has pro-nounced variations in azimuthal and radial anisotropy. Additional evidence for anisotropy in the mantle comes from phase velocities of longer period Love and Rayleigh waves in southern California that also exhibit a clear discrepancy that cannot be explained by an iso-tropic model (Polet and Kanamori, 1997). The presence of laterally varying anisotropy could contribute to different results between isotropic Vp and Vs inversions.

The regional Vs values observed, i.e., 4.3–4.4 km/s for the lithosphere, can be compared to the lower Vs values expected for melt-rich asthenosphere. These differences in Vs facili-tate discriminating between high-speed areas of the upper mantle that likely correspond to lithosphere or arclogite that can be contrasted with areas of lower Vs, that correspond to loca-tions of infl owing asthenosphere. Calculating seismic properties for rocks with compositions of the garnet-bearing xenoliths from the south-ern Sierra Nevada, Frassetto et al. (2011) deter-mined that Vs of arclogite would be between ~4.3 and 4.4 km/s at pressures corresponding to the 30–50 km depth range. Exposed cumulates in the lower crust of the Talkeetna arc of south-ern Alaska (Behn and Kelemen, 2006) have a composition analogous to that expected for the arclogite layer underlying the Sierra Nevada granitic batholith, and can provide estimates of the range of Vs expected for the arclogite. These rocks exhibit a large change in velocity at pressures corresponding to the depth range of 25–75 km where the gabbro-eclogite phase change occurs and Vs increases from 4.0 to 4.4 km/s. Guided by these values and observations of the distribution of shear wave speeds, it is possible to identify areas where the lithospheric root of the Sierra Nevada remains and where it was replaced by melt-rich asthenosphere.

Where Lithosphere Was Removed

Low shear velocities (<4.2 km/s) extend along the eastern portion of the southern Sierra Nevada at depths from the base of the crust to depths >55 km. The low velocities that charac-terize much of the eastern Sierra Nevada indi-cate the lack of signifi cant amounts of litho-sphere or arclogite in the upper mantle. Instead, Vs values near 4 km/s, closer to expectations for asthenosphere, are directly beneath the crust. The area of low wave speeds below 4.2 km/s between the Moho and the ~80 km depth range beneath the eastern Sierra extends from Long Valley and continues south through Owens Val-ley. No area of similarly low velocities is at the top of the upper mantle beneath the northern Sierra Nevada. This suggests that if the removal process is responsible for producing zones of low-velocity material in the crust and upper mantle, then removal has been more extensive in the southern Sierra Nevada than in the north.

Within the area of low shear wave speeds along the eastern Sierra Nevada, focused zones where Vs is further reduced in the crust and upper mantle coincide with locations of recent volcanism (e.g., Long Valley, Golden Trout, Big Pine, and Coso; Manley et al., 2000) (Figs. 8B and 11A). The low velocities could mark the presence of partial melt, magma chambers, or fl uids associated with volcanic activity. Par-tial melt would likely accompany infl owing, upwelling asthenosphere. Comparing obser-vations from the Long Valley caldera to those from the Yellowstone caldera, the low Vs values observed here, in which Vs is ~4 km/s at the Moho, do not extend as deep as those found beneath Yellowstone, where Vs decreases to 3.9 km/s at 70 km depth (Schutt et al., 2008). This difference could result from lower temperatures in the upwelling mantle beneath Long Valley leading to shallower onset of pressure release melting than beneath the Yellowstone caldera.

The modeling of both Le Pourhiet et al. (2006) and Saleeby et al. (2012) found the removal of the Sierra Nevada mantle lithosphere as a Rayleigh-Taylor instability to be a necessary condition for the arclogite root to subsequently delaminate. This makes the removal of arclogite from the base of a batholith a two-stage process. This modeling further predicts that fragments of arclogite and lithosphere could then potentially reascend while entrained in the asthenosphere that is rising in the Rayleigh-Taylor return fl ow. The presence of reascended lithospheric mate-rial offers an explanation for the isostatic model-ing results of Schulte-Pelkum et al. (2011), who identifi ed intact lithosphere present beneath the California-Nevada border, while the seis-mic images presented here show asthenosphere

above the lithosphere in that area. As illustrated by the modeling of Saleeby et al. (2012), the majority of descending and reascending mate-rial becomes laterally displaced relative to where it was located prior to being removed. This can explain why the high-velocity anoma-lies in the central Sierra Nevada appear to dip to the east and extend beneath Owens Valley, as imaged here and by previous investigations of seismic velocities (e.g., Savage et al., 2003; Boyd et al., 2004; Yang and Forsyth, 2006b).

Summary of Where the Lithosphere Remains Intact

Xenoliths that erupted within the central part of the range during the Late Miocene indicate that the lithosphere then had a thick section of arclogite. These arclogites would have formed at depths of 40–90 km as a residue, or cumu-late, of the granitoids that compose the upper portions of the batholith. Based on observations from xenoliths that erupted during the Late Miocene indicating the presence of a thick sec-tion of arclogite, areas in the mantle with Vs of 4.2 km/s or more appear to correspond to either portions of residuum or where preserved litho-sphere remains intact.

Locations where high Vs material remains connected to the overlying crust include the western foothills and localized regions near the center of the southern and central Sierra Nevada and the central and western portions of the northern Sierra Nevada (Figs. 11A, 11B). The thickness of the remaining lithosphere or arclogite varies along the length of the mountain range, reaching depths of ~100 km or more in the southern part of the western foothills (Figs. 10X, 10Y, and 11D). The continuation of the high Vs anomaly beneath the southern Great Valley at depths between 55 and 90 km make it appear that the batholith underlies portions of the Great Valley, in agreement with inferences based on basement cores from the southern part of the valley (Saleeby et al., 2009).

Northward, the zone of higher wave-speed material within the upper mantle thins beneath the central and northern Sierra Nevada, where velocities only exceed 4.2 km/s to depths of ~70–80 km to the north of Lake Tahoe near 39.5°N (Figs. 10W, 10X, 11B, and 11C). Geo-chemical evidence indicates that at least part of the northern Sierra Nevada batholith equili-brated with arclogite cumulates, although the northern batholith is more spatially dispersed, and was emplaced into thinner preexisting crust; this appears to have inhibited the accumula-tion of a substantial arclogite layer (Cecil et al., 2012). Uncertainties remain regarding the con-tinuation of the arclogite root into the northern

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 20: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 19

Sierra due to differences between the northern and southern Sierra Nevada. Because the width of the Sierra Nevada arc increases in the north, it is unlikely that it would have formed a thick arclogitic root similar to that of the narrower southern Sierra Nevada. Instead the increased width to the north could have led to the produc-tion of no arclogitic root or a thinner root distrib-uted over a broader area. In addition, the greater proportion of metamorphic rocks in the north-ern Sierra Nevada may have caused the more sparsely emplaced granites to produce a smaller amount of arclogite during the formation of a root (Bateman and Eaton, 1967). We observe high velocities from the base of the crust to 80 km depth in the northern Sierra Nevada (Figs. 9G–9I, 11C, and 11D); this could indicate that a root formed beneath a limited portion of the northern Sierra to the west of Lake Tahoe and that it has not yet been removed. Perhaps it lacks the negative buoyancy necessary for the removal process to initiate due to its reduced thickness. Alternatively, the more sparsely emplaced batholith during the Cretaceous may have intruded the crust in the northern Sierra in such a diffuse manner that it did not disrupt sig-nifi cantly the mantle wedge in this area. In this scenario the observed high velocities beneath the crust in the northern Sierra Nevada would mark a remnant of the peridotitic mantle wedge and a lack of signifi cant arclogite cumulates.

Places Where Lithosphere Is Currently Being Removed—Peeling Off

The eastern portion of the central Sierra Nevada lacks high-velocity material at the base of the crust, suggesting that lithospheric material that had previously been present was removed. Instead a layer of lower Vs mate-rial extends to depths between 55 and 70 km beneath Owens Valley along the eastern Sierra and extends southward to the southeast portion of the batholith. The distribution of low veloci-ties along the eastern Sierra largely follows the extent of the high velocities along the western Sierra Nevada. However, zones of higher Vs exist at depths >70 km that connect to bodies of high Vs at the base of the crust beneath the western foothills (Figs. 9B–9D and 11A–11C). This appears to be the location of the high-speed arclogite that has been removed from the east-ern part of the central Sierra Nevada. The low Vs material above the high Vs layer would cor-respond to buoyant asthenosphere that rose and replaced the lithosphere that was removed; this agrees with modeling predictions (Le Pourhiet et al., 2006; Saleeby et al., 2012). Though par-tially removed, the remaining arclogite root along the western foothills near the latitude of

Yosemite National Park appears to be actively delaminating. Stresses originating from delami-nation in this area offer a potential explanation for the enigmatic lower crustal earthquakes occurring beneath the western foothills in this area at depths as great as 38 km (Gilbert et al., 2007; Frassetto et al., 2011). The interlayer-ing of asthenosphere and mantle lithosphere between lat 37°N and 38°N in Figures 10W and 10X appears to be similar to the complex inter-layering structures predicted by the modeling. However, note that these layers are nearly 25 km thick, near the limit of the resolution of the Vs model in this depth range.

A consistent result in all the modeling runs of both Le Pourhiet et al. (2006) and Saleeby et al. (2012) is that the higher topography of the Sierra Nevada and Nevada Plano relative to the Great Valley led to their greater gravitational potential driving lower crustal fl ow westward. Following this pattern of fl ow, the arclogite along the eastern Sierra Nevada could then delaminate in the central portion of the Sierra Nevada and, once removed from the base of the batholith, would lead to the area of low Vs in the shallow mantle along the eastern Sierra between 36°N and 38°N (Figs. 11A, 11B). Saleeby et al. (2012) speculated that a portion of this delaminating arclogite eventually necked off and dropped into the mantle. This would leave partially delaminated arclogite beneath the western Sierra Nevada and explain the zone of high Vs beneath the western foothills that dips to the east (Fig. 9C–9F). Material peeling away in this manner aligns more with expectations of lithospheric delamination (Le Pourhiet et al., 2006; Saleeby et al., 2012) than with those of a Rayleigh-Taylor type convective instability, which would predict material descending as vertical drips (e.g., Elkins-Tanton, 2005).

Due to the lack of high topography to the south of the Sierra Nevada, no gravitational potential drove northward-directed lower crustal fl ow, which thereby inhibited an initial south-north progression in delamination. This would also explain why delamination did not follow the opening of the slab window (Saleeby et al., 2012). It is likely that the high Vs body imaged here as the Isabella anomaly also followed an east-west path while initially delaminating. However, following the foundering of a portion of arclogite from the central Sierra Nevada, the root in the south would be able to peel away from south to north. The distribution of low Vs mate-rial to the south and east of the Isabella anomaly at depths >70 km is nearly collocated with the location of the thermal transient and distribu-tion of ca. 4 Ma or younger volcanic rocks pre-sented in Saleeby et al. (2012). The locations of the high and low Vs regions presented here,

combined with their additional geomorphic and stratigraphic evidence, further support the sug-gestion of Saleeby et al. (2012) that the southern part of the Sierra Nevada is a location of active east to west and south to north delamination.

CONCLUSIONS

The shear velocity model presented here includes features within the Sierra Nevada cor-responding to various stages in the process of the removal of a lithospheric root. Areas with very little or no high-speed lithospheric material pres-ent represent the fi nal stage of removal; these are places where the lithosphere has been removed and replaced by asthenosphere, as in the east-ern Sierra Nevada spanning from its southern limit to ~38.5°N. The shallow low velocities found directly below the base of the crust in the southern and eastern Sierra Nevada indicate the lack of high-velocity lithosphere. The infl owing asthenosphere in these areas would provide the necessary buoyancy to support the high eleva-tions of the southern and eastern Sierra Nevada.

The juxtaposition of high-velocity mantle material along the western portion of the Sierra Nevada with lower velocities in the shallow mantle to the east that are underlain by higher velocities marks the location of ongoing litho-spheric removal. The high-velocity material at greater depth to the east corresponds to litho-sphere that was removed and now underlies infl owing astheno sphere that replaced descend-ing dense material. The removal process at this stage appears to be progressing westward and to the north for the southern portion of the Sierra Nevada. This would predict removal of the high Vs material currently observed beneath the west-ern foothills of the Sierra Nevada in the future.

Thinner high Vs lithospheric material under-lies the northern Sierra Nevada than that found beneath the western foothills along the central portion of the mountain range. Without addi-tional information it is diffi cult to determine the origin of this thinner lithosphere. One possibil-ity is that because granitic plutons have a more limited distribution in the northern portion of the batholith than in the south, the northern portion of the batholith would correspondingly form a thinner arclogite root. This portion of lithosphere would then lack the necessary negative buoy-ancy to be removed. Another possibility is that the high-velocity material marks lithosphere that was present prior to the emplacement of the batholith. Regardless of the origin of the litho-spheric material in the northern Sierra Nevada, there is no clear evidence that lithospheric removal occurred to the north of Lake Tahoe. There are no regions of pronounced low veloci-ties in the shallow crust or at the top of the upper

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 21: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Gilbert et al.

20 Geosphere, December 2012

mantle, as in the southern Sierra Nevada. Com-paring the distribution of high Vs crustal material to variations in Moho signal, it appears that the extent of the petrologic Moho marking where a thick arclogite root formed, and remains intact, mirrors the north-south limits of low crustal Vs beneath the eastern Sierra Nevada. This would suggest that the removal of the thickened root is required for the upwelling low Vs material to rise up in the mantle and produce basaltic melt that intrudes the crust.

ACKNOWLEDGMENTS

We are grateful for thoughtful reviews by David Hill, Brandon Schmandt, and Jason Saleeby that improved this manuscript. We thank those involved in the Sierra Nevada EarthScope Project (SNEP) fi eld program and the private landowners and public offi cials who pro-vided access to the land that hosted SNEP stations. Grants EAR-0454554, EAR-0454524, and EAR-0454535 from the National Science Foundation (NSF) EarthScope program to the Universities of Arizona, Colorado, and South Carolina, and Purdue University helped support this research. This is contribution 203 of the ARC (Australian Research Council) Centre of Excellence for Core to Crust Fluid Systems (http://www.ccfs.mq.edu.au) and contribution 845 of the GEMOC (Geochemical Evolution and Metallogeny of Continents) Key Centre (http://www.gemoc.mq.edu.au). The instruments used in the SNEP fi eld deploy-ment were provided by the EarthScope FlexArray , which is managed by the Incorporated Research Insti-tutions for Seismology (IRIS) through the PASSCAL (Program for Array Seismic Studies of the Continental Lithosphere) Instrument Center. The collected data are available through the IRIS Data Management Center. The NSF under cooperative agreement EAR-0004370 supports the facilities of the IRIS Consortium. Data from the Transportable Array network were made freely available as part of the EarthScope USArray facility supported by the NSF Major Research Facility program under cooperative agreement EAR-0350030.

REFERENCES CITED

Ague, J.J., and Brimhall, G.H., 1988, Magmatic arc asym-metry and distribution of anomalous plutonic belts in the batholiths of California: Effects of assimilation, crustal thickness and depth of crystallization: Geological Soci-ety of America Bulletin, v. 100, p. 912–927, doi:10.1130/0016-7606(1988)100<0912:MAAADO>2.3.CO;2.

Aki, K., 1982, Three-dimensional seismic inhomogeneities in the lithosphere and asthenosphere: Evidence for decoupling in the lithosphere and fl ow in the astheno-sphere: Reviews of Geophysics and Space Physics, v. 20, p. 161–170, doi:10.1029/RG020i002p00161.

Anderson, D.L., 2005, Large igneous provinces, delamina-tion, and fertile mantle: Elements, v. 1, p. 271–275, doi:10.2113/gselements.1.5.271.

Atwater, T., and Stock, J., 1998, Pacifi c–North America plate tectonics of the Neogene southwestern United States: An update: International Geology Review, v. 40, p. 375–402, doi:10.1080/00206819809465216.

Bastow, I.D., Owens, T.J., Zandt, G., Jones, C., Gilbert, H. and Julia, J., 2006, SKS splitting analyses from the Sierra Nevada Earthscope Project: Insights into lithospheric foundering: Eos (Transactions, American Geophysical Union), v. 87, fall meeting supplement, abs. S43A–1361.

Bateman, P.C., 1988, Constitution and genesis of the central Sierra Nevada batholith, California: U.S. Geological Survey Open-File Report 88–382, 284 p.

Bateman, P.C., and Eaton, J.P., 1967, Sierra Nevada Batho-lith: Science, v. 158, p. 1407–1417, doi:10.1126/science.158.3807.1407.

Beck, S.L., and Zandt, G., 2002, The nature of orogenic crust in the central Andes: Journal of Geophysical Research, v. 107, 2230, doi:10.1029/2000JB000124.

Behn, M.D., and Kelemen, P.B., 2006, Stability of arc lower crust: Insights from the Talkeetna arc section, south central Alaska, and the seismic structure of modern arcs: Journal of Geophysical Research, v. 111, B11207, doi:10.1029/2006JB004327.

Benz, H.M., and Zandt, G., 1993, Teleseismic tomography, lithospheric structure of the San Andreas fault system in northern and central California, in Iyer, H.M., and Hira-hara, K., eds., Seismic tomography, theory and practice: New York, Chapman and Hall, p. 440–465.

Benz, H.M., Zandt, G., and Oppenheimer, D.H., 1992, Lithospheric structure of northern California from tele-seismic images of the upper mantle: Journal of Geo-physical Research, v. 97, p. 4791–4807, doi:10.1029/92JB00067.

Biasi, G., 2009, Lithospheric evolution of the Pacifi c–North American plate boundary considered in three dimen-sions: Tectonophysics, v. 464, p. 43–59, doi:10.1016/j.tecto.2008.10.018.

Biasi, G.P., and Humphreys, E.D., 1992, P-wave image of the upper mantle structure of central California and southern Nevada: Geophysical Research Letters, v. 19, p. 1161–1164, doi:10.1029/92GL00439.

Bird, P., 1979, Continental delamination and the Colorado Plateau: Journal of Geophysical Research, v. 84, no. B13, p. 7561–7571, http://peterbird.name/publications/1979_delamination/1979_delamination.htm.

Blackwell, D.D., and Richards, M.C., 2004, Geothermal map of North America: American Association of Petro-leum Geologists, scale 1:6,500,000.

Boyd, O.S., Jones, C.H., and Sheehan, A.F., 2004, Foundering lithosphere imaged beneath the southern Sierra Nevada, California, USA: Science, v. 305, p. 660–662, doi:10.1126/science.1099181.

Burdick, S., van der Hilst, R.D., Vernon, F.L., Martynov, V., Cox, T., Eakins, J., Astiz, L., and Pavlis, G.L., 2009, Model update December 2008: Upper mantle hetero-geneity beneath North America from P-wave travel time tomography with global and USArray transport-able array data: Seismological Research Letters, v. 80, p. 638–645, doi:10.1785/gssrl.80.4.638.

Busby, C.J., and Putirka, K., 2009, Miocene evolution of the western edge of the Nevadaplano in the central and northern Sierra Nevada: Paleocanyons, magmatism, and structure: International Geology Review, v. 51, p. 670–701, doi:10.1080/00206810902978265.

Busby, C.J., DeOreo, S.B., Skilling, I., Gans, P.B., and Hagan, J.C., 2008, Carson Pass–Kirkwood paleocanyon system: Paleogeography of the ancestral Cascades arc and implications for landscape evolution of the Sierra Nevada (California): Geological Society of America Bulletin, v. 120, p. 274–299, doi:10.1130/B25849.1.

Calvert, A., Sandvol, E., Seber, D., Barazangi, M., Roecker, S., Mourabit, T., Vidal, F., Alguacil, G., and Jabour, N., 2000, Geodynamic evolution of the lithosphere and upper mantle beneath the Alboran region of the western Mediterranean: Constraints from travel time tomography: Journal of Geophysical Research, v. 105, p. 10871–10898, doi:10.1029/2000JB900024.

Cassel, E.L., Graham, S.A., and Chamberlain, C.P., 2009, Cenozoic tectonic and topographic evolution of the northern Sierra Nevada, California, through stable iso-tope paleoaltimetry in volcanic glass: Geology, v. 37, p. 547–550, doi:10.1130/G25572A.1.

Cecil, M.R., Rothberg, G.L., Ducea, M.N., Saleeby, J.B., and Gehrels, G.E., 2012, Magmatic growth and batholithic root development in the northern Sierra Nevada, California: Geosphere, v. 8, p. 592–606, doi:10.1130/GES00729.1.

Cousens, B., Prytulak, J., Henry, C., Alcazar, A., and Brown-rigg, T., 2008, Geology, geochronology, and geochem-istry of the Miocene–Pliocene Ancestral Cascades arc, northern Sierra Nevada, California and Nevada: The roles of the upper mantle, subducting slab, and the Sierra Nevada lithosphere: Geosphere, v. 4, p. 829–853, doi:10.1130/GES00166.1.

Crough, S.T., and Thompson, G.A., 1977, Upper mantle origin of the Sierra Nevada uplift: Geology, v. 5, p. 396–399, doi:10.1130/0091-7613(1977)5<396:UMOOSN>2.0.CO;2.

Dickinson, W.R., 1981, Plate tectonics and the continental margin of California, in Ernst, W.G., ed., The geo-tectonic development of California: Rubey volume 1: Englewood Cliffs, New Jersey, Prentice-Hall, p. 1–28.

Dickinson, W.R., 2008, Accretionary Mesozoic–Cenozoic expansion of the Cordilleran continental margin in Cali-fornia and adjacent Oregon: Geosphere, v. 4, p. 329–353, doi:10.1130/GES00105.1.

Ducea, M.N., 2001, The California arc: Thick granitic batholiths, eclogitic residues, lithospheric-scale thrusting, and mag-matic fl are-ups: GSA Today, v. 11, p. 4–10, doi:10.1130/1052-5173(2001)011<0004:TCATGB>2.0.CO;2.

Ducea, M.N., 2002, Constraints on the bulk composition and root foundering rates of continental arcs: A Califor-nia arc perspective: Journal of Geophysical Research, v. 107, 2304, doi:10.1029/2001JB000643.

Ducea, M.N., and Saleeby, J.B., 1996, Buoyancy sources for a large, unrooted mountain range, the Sierra Nevada, California: Evidence from xenoliths thermobarometry: Journal of Geophysical Research, v. 101, p. 8229–8244, doi:10.1029/95JB03452.

Ducea, M.N., and Saleeby, J.B., 1998, A case for delami-nation of the deep batholithic crust beneath the Sierra Nevada, California: International Geology Review, v. 40, p. 78–93, doi:10.1080/00206819809465199.

Dündar, S., and 14 others, 2011, Receiver function images of the base of the lithosphere in the Alboran Sea region: Geophysical Journal International, v. 187, p. 1019–1026, doi:10.1111/j.1365-246X.2011.05216.x.

Elkins-Tanton, L.T., 2005, Continental magmatism caused by lithospheric delamination, in Foulger, G.R., et al., eds., Plates, plumes, and paradigms: Geological Soci-ety of America Special Paper 388, p. 449–461.

England, P., and Houseman G., 1989, Extension during conti-nental convergence, with application to the Tibetan Pla-teau: Journal of Geophysical Research, v. 94, no. B12, p. 17,561–17,579, doi:10.1029/JB094iB12p17561.

Elkins-Tanton, L.T., and Grove, T.L., 2003, Evidence for deep melting of hydrous, metasomatized mantle: Plio-cene high potassium magmas from the Sierra Nevadas: Journal of Geophysical Research, v. 108, 2350, doi:10.1029/2002JB002168.

Farmer, G.L., Glazner, A.F., and Manley, C.R., 2002, Did lithospheric delamination trigger late Cenozoic potas-sic volcanism in the southern Sierra Nevada, Califor-nia?: Geological Society of America Bulletin, v. 114, p. 754–768, doi:10.1130/0016-7606(2002)114<0754:DLDTLC>2.0.CO;2.

Fliedner, M.M., Ruppert, S., and the SSCD Working Group, 1996, Three-dimensional crustal structure of the south-ern Sierra Nevada from seismic fan profi les and grav-ity modeling: Geology, v. 24, p. 367–370, doi:10.1130/0091-7613(1996)024<0367:TDCSOT>2.3.CO;2.

Fliedner, M.M., Klemperer, S.L., and Christensen, N.I., 2000, Three-dimensional seismic model of the Sierra Nevada arc, California, and its implications for crustal and upper mantle composition: Journal of Geophysical Research, v. 105, p. 10899–10921, doi:10.1029/2000JB900029.

Forsyth, D.W., and Li, A., 2005, Array analysis of two-dimen-sional variations in surface wave phase velocity and azimuthal anisotropy in the presence of multipathing interference, in Levander, A., and Nolet, G., eds., Seis-mic Earth: Array analysis of broadband seismograms: American Geophysical Union Geophysical Monograph 157, p. 81–97, doi:10.1029/157GM06.

Forsyth, D.W., Wang, Y., Rau, C.J., Savage, B., Eichenbaum-Pikser, J., and Carriero, N., 2011, Fossil slabs attached to unsubducted fragments of the Farallon plate: Geo-logical Society of America Abstracts with Programs, v. 43, no. 5, p. 319.

Frassetto, A., Zandt, G., Gilbert, H., Owens, T.J., and Jones, C., 2011, Lithospheric structure of the Sierra Nevada from receiver functions and implications for litho-spheric foundering: Geosphere, v. 7, p. 898–921, doi:10.1130/GES00570.1.

Fuis, G.S., and 14 others, 2003, Fault systems of the 1971 San Fernando and 1994 Northridge earthquakes, southern California: Relocated aftershocks and seismic images from LARSE II: Geology, v. 31, p. 171–174, doi:10.1130/0091-7613(2003)031<0171:FSOTSF>2.0.CO;2.

Gilbert, H., 2012, Crustal structure and signatures of recent tectonism as influenced by ancient terranes in the

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012

Page 22: Geosphere - Macquarie Universityeps.mq.edu.au/~yingjie/publication/2012-Gilbert-Geosphere.pdfImaging lithospheric foundering in the structure of the Sierra Nevada H. Gilbert1, Y. Yang

Lithospheric foundering in the Sierra Nevada

Geosphere, December 2012 21

western United States: Geosphere, v. 8, p. 141–157, doi:10.1130/GES00720.1.

Gilbert, H.J., Jones, C., Owens, T.J., Zandt, G., 2007, Imaging Sierra Nevada lithospheric sinking: Eos (Trans actions, American Geophysical Union), v. 88, p. 225, doi:10.1029/2007EO210001.

Grand, S.P., and Helmberger, D.V., 1984, Upper mantle shear structure of North America: Royal Society of London Geophysical Journal, v. 76, p. 399–438, doi:10.1111/j.1365-246X.1984.tb05053.x.

Hadley, D.M., and Kanamori, H., 1977, Seismic structure of the Transverse Ranges, California: Geological Society of America Bulletin, v. 88, p. 1469–1478, doi:10.1130/0016-7606(1977)88<1469:SSOTTR>2.0.CO;2.

Hales, T.C., Abt, D., Humphreys, E., and Roering, J., 2005, A lithospheric instability origin for Columbia River fl ood basalts and Wallowa Mountains uplift in north-east Oregon: Nature, v. 438, p. 842–845, doi:10.1038/nature04313.

Hammond, W.C., and Humphreys, E.D., 2000, Upper man-tle seismic wave attenuation: Effects of realistic partial melt distribution: Journal of Geophysical Research, Solid Earth, v. 105, p. 10,987–10,999.

Humphreys, E.D., 1995, Post-Laramide removal of the Farallon slab, western United States: Geology, v. 23, p. 987–990, doi:10.1130/0091-7613(1995)023<0987:PLROTF>2.3.CO;2.

Humphreys, E.D., Clayton, R.W., and Hager, B.H., 1984, A tomographic image of mantle structure beneath south-ern California: Geophysical Research Letters, v. 11, p. 625–627, doi:10.1029/GL011i007p00625.

Jones, C.H., 1987, Is extension in Death Valley accommo-dated by thinning of the mantle lithosphere beneath the Sierra Nevada, California?: Tectonics, v. 6, p. 449–473, doi:10.1029/TC006i004p00449.

Jones, C.H., Kanamori, H., and Roecker, S.W., 1994, Miss-ing roots and mantle “drips”: Regional Pn and teleseis-mic arrival times in the southern Sierra Nevada and vicinity, California: Journal of Geophysical Research, v. 99, p. 4567–4601, doi:10.1029/93JB01232.

Jones, C.H., Farmer, G.L., and Unruh, J., 2004, Tectonics of Pliocene removal of lithosphere of the Sierra Nevada, California: Geological Society of America Bulletin, v. 116, p. 1408–1422, doi:10.1130/B25397.1.

Jones, L.M., and Dollar, R.S., 1986, Evidence of basin-and-range extensional tectonics in the Sierra Nevada: The Durrwood Meadows swarm, Tulare County, California (1983–1984): Seismological Society of America Bul-letin, v. 76, p. 439–461.

Karlstrom, K., Crow, R., Crossey, L.J., Coblentz, D., and Van Wijk, J.W., 2008, Model for tectonically driven incision of the younger than 6 Ma Grand Canyon: Geology, v. 36, p. 835–838, doi:10.1130/G25032A.1.

Kay, R.W., and Mahlburg-Kay, S., 1991, Creation and destruction of lower continental crust: Geolo-gische Rundschau, v. 80, p. 259–278, doi:10.1007/BF01829365.

Kennett, B.L.N., and Engdahl, E.R., 1991, Traveltimes for global earthquake location and phase identifi cation: Geophysical Journal International, v. 105, p. 429–465, doi:10.1111/j.1365-246X.1991.tb06724.x.

Kohler, M.D., Magistrale, H., and Clayton, R.W., 2003, Mantle heterogeneities and the SCEC reference three-dimensional seismic velocity model version 3: Seismo-logical Society of America Bulletin, v. 93, p. 757–774, doi:10.1785/0120020017.

Lackey, J.S., Cecil, M.R., Windham, C.J., Frazer, R.E., Binde-man, I.N., and Gehrels, G.E., 2012, The Fine Gold Intru-sive Suite: The roles of basement terranes and magma source development in the Early Cretaceous Sierra Nevada batholith: Geosphere, v. 8, p. 292–313, doi:10.1130/GES00745.1.

Le Pourhiet, L., Gurnis, M., and Saleeby, J., 2006, Mantle instability beneath the Sierra Nevada Mountains in California and Death Valley extension: Earth and Plan-etary Science Letters, v. 251, p. 104–119, doi:10.1016/j.epsl.2006.08.028.

Levander, A., Schmandt, B., Miller, M.S., Liu, K., Karlstrom, K.E., Crow, R.S., Lee, C.T.A., and Humphreys, E.D., 2011, Continuing Colorado plateau uplift by delamina-tion style convective lithospheric downwelling: Nature, v. 472, p. 461–465, doi:10.1038/nature10001.

Lin, F., Ritzwoller, M.H., Yang, Y., Moschetti, M.P., and Fouch, M.J., 2011, Complex and variable crustal and upper-most mantle seismic anisotropy in the western United States: Nature Geoscience, v. 4, p. 55–61, doi:10.1038/ngeo1036.

Luffi , P., Saleeby, J.B., Lee, C.-T.A., and Ducea, M.N., 2009, Lithospheric mantle duplex beneath the central Mojave Desert revealed by xenoliths from Dish Hill, Califor-nia: Journal of Geophysical Research, v. 114, B03202, doi:10.1029/2008JB005906.

Mahéo, G., Saleeby, J., Saleeby, Z., and Farley, K.A., 2009, Tectonic control on southern Sierra Nevada topogra-phy, California: Tectonics, v. 28, TC6006, doi:10.1029/2008TC002340.

Manley, C.R., Glazner, A.F., and Farmer, G.L., 2000, Timing of volcanism in the Sierra Nevada of California: Evidence for the Pliocene delamination of the batholithic root?: Geology, v. 28, p. 811–814, doi:10.1130/0091-7613(2000)28<811:TOVITS>2.0.CO;2.

Moschetti, M.P., Ritzwoller, M.H., Lin, F., and Yang, Y., 2010, Seismic evidence for widespread western US deep-crustal deformation caused by extension: Nature, v. 464, p. 885–889, doi:10.1038/nature08951.

Oliver, H.W., 1977, Gravity and magnetic investigations of the Sierra Nevada batholith, California: Geological Soci-ety of America Bulletin, v. 88, p. 445–461, doi:10.1130/0016-7606(1977)88<445:GAMIOT>2.0.CO;2.

Park, S., 2004, Mantle heterogeneity beneath eastern Cali-fornia from magnetotelluric measurements: Journal of Geophysical Research, v. 109, B09406, doi:10.1029/2003JB002948.

Platt, J.P., and England, P.C., 1994, Convective removal of lithosphere beneath mountain belts: Thermal and mechanical consequences: American Journal of Sci-ence, v. 294, p. 307–336, doi:10.2475/ajs.294.3.307.

Polet, J., and Kanamori, H., 1997, Upper-mantle shear velocities beneath southern California determined from long-period surface waves: Seismological Society of America Bulletin, v. 87, p. 200–209.

Rau, C.J., and Forsyth, D.W., 2011, Melt in the mantle beneath the amagmatic zone, southern Nevada: Geol-ogy, v. 39, p. 975–978, doi:10.1130/G32179.1.

Ruppert, S., Fliedner, M.M., and Zandt, G., 1998, Thin crust and active upper mantle beneath the southern Sierra Nevada in the western United States: Tectonophysics, v. 286, p. 237–252, doi:10.1016/S0040-1951(97)00268-0.

Saito, M., 1988, DISPER80: A subroutine package for the calculation of seismic normal mode solutions, in Doornbos, D.J., ed., Seismological algorithms: Com-putational methods and computer programs: New York, Elsevier, p. 293–319.

Saleeby, J., 2003, Segmentation of the Laramide slab—Evi-dence from the southern Sierra Nevada region: Geologi-cal Society of America Bulletin, v. 115, p. 655–668, doi:10.1130/0016-7606(2003)115<0655:SOTLSF>2.0.CO;2.

Saleeby, J., 2007, Western extent of the Sierra Nevada batho-lith in the Great Valley basement and its signifi cance in underlying mantle dynamics: Eos (Transactions, American Geophysical Union), v. 88, abs. T31E-02.

Saleeby, J., and Foster, Z., 2004, Topographic response to mantle lithosphere removal in the southern Sierra Nevada region, California: Geology, v. 32, p. 245–248, doi:10.1130/G19958.1.

Saleeby, J., Ducea, M., and Clemens-Knott, D., 2003, Pro-duction and loss of high-density batholithic root, south-ern Sierra Nevada, California: Tectonics, v. 22, 1064, doi:10.1029/2002TC001374.

Saleeby, J., Saleeby, Z., Nadin, E., and Maheo, G., 2009, Step-over in the structure controlling the regional west tilt of the Sierra Nevada microplate: Eastern escarpment system to Kern Canyon system: International Geology Review, v. 51, p. 634–669, doi:10.1080/00206810902867773.

Saleeby, J., Le Pourhiet, L., Saleeby, Z., and Gurnis, M., 2012, Epeirogenic transients related to mantle lithosphere removal in the southern Sierra Nevada region, California, part I: Implications of thermomechanical modeling: Geo-sphere, doi:10.1130/GES00746.1 (in press).

Saleeby, J., Saleeby, Z., and Le Pourhiet, L., 2013, Epeiro-genic transients related to mantle lithosphere removal in the southern Sierra Nevada region, California, part II: Implications of rock uplift and basin subsidence rela-tions: Geosphere, doi:10.1130/GES00816.1 (in press).

Saltus, R.W., and Lachenbruch, A.H., 1991, Thermal evo-lution of the Sierra Nevada: Tectonic implications of new heat fl ow data: Tectonics, v. 10, p. 325–344, doi:10.1029/90TC02681.

Savage, B., Ji, C., and Helmberger, D.V., 2003, Velocity vari-ations in the uppermost mantle beneath the southern Sierra Nevada and Walker Lane: Journal of Geophysical Research, v. 108, 2325, doi:10.1029/2001JB001393.

Savage, M.K., Li, L., Eaton, J.P., Jones, C.H., and Brune, J.N., 1994, Earthquake refraction profi les of the root of the Sierra Nevada: Tectonics, v. 13, p. 803–817, doi:10.1029/93TC03488.

Schmandt, B., and Humphreys, E., 2010, Seismic heterogene-ity and small-scale convection in the southern California upper mantle: Geochemistry Geophysics Geosystems, v. 11, Q05004, doi:10.1029/2010GC003042.

Schulte-Pelkum, V., Biasi, G., Sheehan, A., and Jones, C., 2011, Differential motion between upper crust and litho-spheric mantle in the central Basin and Range: Nature Geoscience, v. 4, p. 619–623, doi:10.1038/ngeo1229.

Schutt, D.L., Dueker, K., and Yuan, H., 2008, Crust and upper mantle velocity structure of the Yellowstone hot spot and surroundings: Journal of Geophysical Research, v. 113, B03310, doi:10.1029/2007JB005109.

Schweickert, R.A., 1981, Tectonic evolution of the Sierra Nevada Range, in Ernst, W.G., ed., The geotectonic development of California: Rubey volume 1: Engle-wood Cliffs, New Jersey, Prentice-Hall, pp. 87–131.

Seber, D., Barazangi, M., Ibenbrahim, A., and Demnati, A., 1996, Geophysical evidence for lithospheric delamina-tion beneath the Alboran Sea and Rif-Betic mountains: Nature, v. 379, p. 785–790, doi:10.1038/379785a0.

Smith, M.L., and Dahlen, F.A., 1973, The azimuthal depen-dence of Love and Rayleigh wave propagation in a slightly anisotropic medium: Journal of Geophysical Research, v. 78, p. 3321–3333, doi:10.1029/JB078i017p03321.

Stixrude, L., and Lithgow-Bertelloni, C., 2005, Mineralogy and elasticity of the oceanic upper mantle: Origin of the low-velocity zone: Journal of Geophysical Research, v. 110, B03204, doi:10.1029/2004JB002965.

Wagner, L.S., Forsyth, D.W., Fouch, M.J., and James, D.E., 2010, Detailed three-dimensional shear wave velocity structure of the northwestern United States from Rayleigh wave tomography: Earth and Planetary Science Letters, v. 299, p. 273–284, doi:10.1016/j.epsl.2010.09.005.

Wang, Y., Forsyth, D.W., and Savage, B., 2009, Convective upwelling in the mantle beneath the Gulf of California: Nature, v. 462, p. 499–501, doi:10.1038/nature08552.

West, J.D., Fouch, M.J., Roth, J.B., and Enkins-Tanton, L.T., 2009, Vertical mantle fl ow associated with a litho-spheric drip beneath the Great Basin: Nature Geosci-ence, v. 2, p. 439–444, doi:10.1038/ngeo526.

Yan, Z.M., Clayton, R.W., and Saleeby, J., 2005, Seismic refrac-tion evidence for steep faults cutting highly attenuated continental basement in the central Transverse ranges, California: Geophysical Journal International, v. 160, p. 651–666, doi:10.1111/j.1365-246X.2005.02506.x.

Yang, Y., and Forsyth, D.W., 2006a, Regional tomographic inversion of the amplitude and phase of Rayleigh waves with 2-D sensitivity kernels: Geophysical Jour-nal International, v. 166, p. 1148–1160, doi:10.1111/j.1365-246X.2006.02972.x.

Yang, Y., and Forsyth, D.W., 2006b, Rayleigh wave phase velocities, small-scale convection, and azimuthal anisot-ropy beneath southern California: Journal of Geophysical Research, v. 111, B07306, doi:10.1029/2005JB004180.

Yang, Y., and Forsyth, D.W., 2008, Attenuation in the upper mantle beneath southern California: Physical state of the lithosphere and asthenosphere: Journal of Geophysical Research, v. 113, B03308, doi:10.1029/2007JB005118.

Yang, Y., Forsyth, D.W., and Weeraratne, D.S., 2007, Seismic attenuation near the East Pacifi c Rise and the origin of the low-velocity zone: Earth and Planetary Science Letters, v. 258, p. 260–268, doi:10.1016/j.epsl.2007.03.040.

Zandt, G., and Humphreys, E., 2008, Toroidal mantle fl ow through the western US slab window: Geology, v. 36, p. 295–298, doi:10.1130/G24611A.1.

Zandt, G., Gilbert, H., Owens, T.J., Ducea, M., Saleeby, J., and Jones, C.H., 2004, Active foundering of a conti-nental arc root beneath the southern Sierra Nevada in California: Nature, v. 431, p. 41–46, doi:10.1038/nature02847.

as doi:10.1130/GES00790.1Geosphere, published online on 16 November 2012


Recommended