Topography, the geoid, and compensation mechanisms for the southern Rocky Mountains

Similar documents
FORWARD MODELING OF THE GEOID ANOMALY USING SPHERICAL HARMONICS: APPLICATIONS IN THE SIERRA NEVADA. Alissa C Scire

Isostacy: Compensation of Topography and Isostatic Geoid Anomalies

GRAVITY AND ISOSTASY

Gravitational constraints

Evolution of Continents Chapter 20


Lecture 2: Gravity, Isostasy and Flexure

Introduction to: The Rocky Mountain Region An Evolving Lithosphere: Tectonics, Geochemistry, and Geophysics

12 Gravity and Topography

Small scale convection at the edge of the Colorado Plateau?

The influence of short wavelength variations in viscosity on subduction dynamics

Constraints on Mantle Structure from Surface Observables

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth.

Why is the North America Cordillera high? Hot backarcs, thermal isostasy and

Plate Tectonics. Why Continents and Ocean Basins Exist

Defining the former elevation and shape of the lithosphere, in particular the elevation of the Earth s surface,

exerted by the convective flow on the base of the lithosphere, may constraint on the way these tectonic forces interact is provided by

by west-dipping subduction of the North American Plate

1. A few words about EarthScope and USArray. 3. Tomography using noise and Aki s method

5. Gravity. 5.1 Geoid Variations. The Australian Continent: A Geophysical Synthesis Gravity

B6 Isostacy. B6.1 Airy and Pratt hypotheses. Geophysics 210 September 2008

Mountain Building. Mountain Building

9. Density Structure. The Australian Continent: A Geophysical Synthesis Density Structure

Isostasy and Tectonics Lab Understanding the Nature of Mobile Floating Lithospheric Plates

Plate Tectonics. entirely rock both and rock

Plume-induced topography and geoid anomalies and their implications for the Tharsis rise on Mars

Geodynamics Lecture 10 The forces driving plate tectonics

GEO-DEEP9300 Lithosphere and Asthenosphere: Composition and Evolution

scale Three-dimensional Modelling

Selected Presentation from the INSTAAR Monday Noon Seminar Series.

NC Earth Science Essential Standards

4-D Geodynamic Modeling With Data Assimilation: Subduction and Continental Evolution

Edge Driven Convection and Iceland

Total gravitational field is sum of contributions from all masses.

L.O: THE CRUST USE REFERENCE TABLE PAGE 10

High resolution receiver function imaging reveals Colorado Plateau lithospheric architecture and mantle supported topography

topography half is commonly represented in maps, and how it interacts with

Seismotectonics of intraplate oceanic regions. Thermal model Strength envelopes Plate forces Seismicity distributions

Earth Systems Science Chapter 7. Earth Systems Science Chapter 7 11/11/2010. Seismology: study of earthquakes and related phenomena

Marine Science and Oceanography

The Earth s Structure from Travel Times

GEOLOGY 101 LABORATORY LAB #2 Isostacy and Plate Tectonics How Density and Gravity Control Plate Dynamics

Chapter 2. The Planet Oceanus

Plaattektoniek en Mickey Mouse: de bewegingen van de Aarde en de geologie van Marokko. G. Bertotti - TUDelft

Surface changes caused by erosion and sedimentation were treated by solving: (2)

Summary and Conclusions

Dynamic Crust Practice

NAME HOMEWORK ASSIGNMENT #4 MATERIAL COVERS CHAPTERS 19, 20, 21, & 2

Plate Tectonics. A. Continental Drift Theory 1. Early development 2. Alfred Wegener s mechanism

Copyright McGraw-Hill Education. All rights reserved. No reproduction or distribution without the prior written consent of McGraw-Hill Education

6. In the diagram below, letters A and B represent locations near the edge of a continent.

High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation

44031 Crustal Compensation Rate In Central Anatolian Region And Comparison Of Seismic Activity

Forces That Shape Earth. How do continents move? What forces can change rocks? How does plate motion affect the rock cycle?

Petrology. Petrology: the study of rocks, especially aspects such as physical, chemical, spatial and chronoligic. Classification:

Earth s Continents and Seafloors. GEOL100 Physical Geology Ray Rector - Instructor

Petrology. Petrology: the study of rocks, especially aspects such as physical, chemical, spatial and chronoligic. Associated fields include:

Strike-Slip Faults. ! Fault motion is parallel to the strike of the fault.

Physics of the Earth

EARTH S ENERGY SOURCES

(b) What is the amplitude at the altitude of a satellite of 400 km?

Nathan A. Simmons Lawrence Livermore National Laboratory, California Alessandro M. Forte Université du Québec à Montréal Lapo Boschi

USU 1360 TECTONICS / PROCESSES

Geophysical Journal International

4. Basins due to orogenic loading

OCN 201 Physiography of the Seafloor

Lab 8: Gravity and Isostasy (35 points)

Dynamic Subsidence and Uplift of the Colorado Plateau. Supplementary Material

We have compiled bathymetry and earthquake data of the Sunda Shelf to study the

Continent-sized anomalous zones with low seismic velocity at the base of Earth s mantle

Crustal Boundaries. As they move across the asthenosphere and form plate boundaries they interact in various ways. Convergent Transform Divergent

Marine Geophysics. Plate tectonics. Dept. of Marine Sciences, Ocean College, Zhejiang University. Nov. 8, 2016

Lecture notes Bill Engstrom: Instructor Earth s Interior GLG 101 Physical Geology

Earth s s Topographic Regions

Topic 12: Dynamic Earth Pracatice

Geotherms. Reading: Fowler Ch 7. Equilibrium geotherms One layer model

Important information from Chapter 1

Previously Documented Basin-localized Extension on Mercury

Correction to Localized gravity/topography admittance and correlation spectra on Mars: Implications for regional and global evolution

North America subducted under Rubia. Are there modern analogs for Hildebrand s model of North America subducting under Rubia?

D) outer core B) 1300 C A) rigid mantle A) 2000 C B) density, temperature, and pressure increase D) stiffer mantle C) outer core

Structural models of a Venusian rift system from Magellan gravity and topography

Mapping the admittance and correlation functions on the Martian surface using a spherical wavelet filter

Lab 2: Visualizing the Large-Scale Earth (35 points)

Earthscope in the Northern Rockies Workshop

Gravity/topography admittance inversion on Venus using niching genetic algorithms

Directed Reading. Section: How Mountains Form MOUNTAIN RANGES AND SYSTEMS. Skills Worksheet

DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a

Earth s Interior: Big Ideas. Structure of the Earth

Definition: Plate tectonics: the theory that the earth s surface consists of a mosaic of internally rigid plates that move relative to each other

Week Five: Earth s Interior/Structure

A Rootless Rockies - Support and Lithospheric Structure of the Colorado Rocky Mountains Inferred from CREST and TA Seismic Data

Standard 2, Objective 1: Evaluate the source of Earth s internal heat and the evidence of Earth s internal structure.

Chapter 10: Volcanoes and Other Igneous Activity Section 1: The Nature of Volcanic Eruptions I. Factors Affecting Eruptions Group # Main Idea:

OCN 201 Physiography of the Seafloor

A) B) C) D) 4. Which diagram below best represents the pattern of magnetic orientation in the seafloor on the west (left) side of the ocean ridge?

Constraints on lithosphere and mantle rheology from in-situ observations

Continental Drift and Plate Tectonics

Shape Earth. Plate Boundaries. Building. Building

Peer Reviewed Publications

Transcription:

Article Volume 12, Number 4 1 April 2011 Q04002, doi: ISSN: 1525 2027 Topography, the geoid, and compensation mechanisms for the southern Rocky Mountains D. Coblentz Geodynamics Group, Los Alamos National Laboratory, MS D443, Los Alamos, New Mexico 87545, USA (coblentz@lanl.gov) C. G. Chase Department of Geosciences, University of Arizona, Tucson, Arizona 85721, USA K. E. Karlstrom Department of Earth and Planetary Sciences, University of New Mexico, Albuquerque, New Mexico 87131, USA J. van Wijk Department of Earth and Atmospheric Sciences, University of Houston, Houston, Texas 77204 5007, USA [1] The southern Rockies of Colorado are anomalously high (elevations greater than 2800 m), topographically rough (implying active uplift), and underlain by significant low velocity anomalies in the upper mantle that suggest an intimate relationship between mantle geodynamic processes and the surface topography. The region is in isostatic equilibrium (i.e., near zero free air gravity anomaly); however, the poor correlation between the high topography and crustal thickness makes the application of simple compensation models (e.g., pure Heiskanen or Pratt Hayford) problematic. Knowledge of how the current topography of the Rockies is isostatically compensated could provide constraints on the relative role of sublithospheric buoyancy versus lithospheric support. Here we evaluate the geoid and its relationship to the topography (using the geoid to elevation ratio (GTR) in the spatial domain and the admittance in the frequency domain) to constrain the mechanism of compensation. We separate the upper mantle geoid anomalies from those with deeper sources through the use of spherical harmonic filtering of the EGM2008 geoid. We exploit the fact that at wavelengths greater than the flexural wavelength where features are isostatically compensated, the geoid/ topography ratio can be used to estimate the depth of compensation and the elastic thickness of the lithosphere. The results presented below indicate that the main tectonic provinces of the western United States have moderate geoid/topography ratios between 3.5 and 5.5 m/km ( 3.9 for the southern Rockies, 4.25 for the Colorado Plateau, and 5.2 for the Northern Basin and Range) suggesting shallow levels of isostatic compensation. In terms of the elastic thickness of the lithosphere, our results indicate an elastic thickness of less than 20 km. These value support the notion that a major portion of the buoyancy that has driven uplift resides at depths less than 100 km and that upper mantle processes such as small scale convection may play a significant role in the buoyant uplift of the southern Rockies (as well as other actively uplifting areas of the western United States). Further support for this hypothesis is provided by high coherence for the geoidtopography relationship for nearly all wavelengths between 50 and 1000 km. Components: 9800 words, 8 figures. Keywords: topography; geoid; Rocky Mountains. Copyright 2011 by the American Geophysical Union 1 of 18

Index Terms: 8120 Tectonophysics: Dynamics of lithosphere and mantle: general (1213); 8175 Tectonophysics: Tectonics and landscape evolution; 8122 Tectonophysics: Dynamics: gravity and tectonics. Received 2 December 2010; Revised 17 February 2011; Accepted 22 February 2011; Published 1 April 2011. Coblentz, D., C. G. Chase, K. E. Karlstrom, and J. van Wijk (2011), Topography, the geoid, and compensation mechanisms for the southern Rocky Mountains, Geochem. Geophys. Geosyst., 12, Q04002, doi:. 1. Introduction [2] Our understanding of how the high topography of mountain ranges is compensated at depth is incomplete except for the most simplistic cases (e.g., balanced one dimensional columns and cartoon illustrations of the lithosphere). Reality is a much more complicated mix of the principal compensation mechanisms such as thickening a constant density crust (Airy), lateral changes in density of crust (Pratt) and the support of topographic loads by elastic stresses within the lithospheric plate overlying a weak, fluid asthenosphere (flexural support). The dominance of the Airy type compensation in earlier studies stems in large part from early crustal structure studies that indicated that most high mountain ranges have a thick root/crust. In such a situation the free air gravity anomaly is small and approaches zero for the longest wavelengths. In contrast, the Bouguer gravity anomaly is nonzero, reflecting the crustal root, and strongly correlated with topography at long wavelengths. To a first order these observation hold true for a majority of the Earth s mountain ranges. It is only with the arrival of recent, more detailed, seismic studies (e.g., the EarthScope project in the western United States) and evaluations of the intraplate stress field [Zoback and Mooney, 2003] that sufficient information has become available to raise serious question about the applicability of the Airy model of compensation. Thus, while it is generally accepted that mountain belts are underlain by roots that are less dense than the surrounding mantle, it is becoming more and more apparent that the roots are not large enough to appeal to a purely Airy mechanism of compensation. Indeed, one of the principal conclusions of Zoback and Mooney [2003] was that for most continental areas there exists no clear relationship between elevation and crustal thickness and pure Airy type correlation between elevation and crustal thickness is limited to continental regions with either very thick (>55 km) or very thin (<20 km) crust. The implication is that mantle buoyancy may be a significant contributor to surface elevation, a suggestion made more than 40 years ago [Thompson and Talwani, 1964]. [3] This is particularly the case for the actively deforming western U.S. Cordillera. The western United States is characterized by regionally high elevation typically exceeding 1.5 km and recent deformation and uplift which invites evaluation of the relationship between topography and the underlying geodynamics driving continental uplift. The southern Rocky Mountains in Colorado (SRM) and the Colorado Plateau (CP) regions are particularly intriguing, given the debate surrounding the uplift dynamics of the CP and the anomalously high and actively deforming SRM. Both of these provinces are characterized by a present day lithosphere that is a composite of structures that formed during lithospheric assembly and during later modification, including active tectonism. Recent modeling results suggest that while the CP and the SRM are quite distinct in geology and physiography, they share recent uplift which is hypothesized to be driven by small scale and relatively shallow mantle flow and buoyancy [van Wijk et al., 2010]. This influence from the upper mantle is consistent with the observation that the Tertiary epeirogenies in the western United States are coincident with mafic igneous activity and suggests that continental swells such as the southern Rockies are associated with sublithospheric processes (possibly akin to a variation on mantle hot spots) in the same manner as oceanic swells [Crough, 1979; Eaton, 1986, 2008; Suppe et al., 1975]. [4] Much of the high elevation of the SRM coincides with thin or attenuated continental crust, which suggests that anomalous buoyancy of the mantle plays an important role in the topographic support. An emerging hypothesis from the recently completed CREST (Colorado Rocky Mountain Experiment and Seismic Transect) experiment [Aster et al., 2009] is that 100 km scale mantle velocity domains correlate with regions undergoing differential uplift in the SRM and CP and, more generally, that surface topography is being shaped by mantle flow and buoyancy at various scales. Significant upper mantle low velocity features beneath the Aspen region and San Juan Mountains appear to coincide with Proterozoic lithospheric structures within the Colorado 2of18

Figure 1. P and S velocity model depth slices at 60 km (first row) and cross sections from 0 to 660 km (second through fourth rows) from linear inversion of relative traveltime delays across a network of 167 CREST, USArray, and other broadband seismic stations located in central Colorado (J. R. MacCarthy, personal communication, 2011). These high resolution images of the upper mantle reveal large seismic velocity variations on a 50 100 km scale and are suggestive of interpretations relating the velocity anomalies to surface topographic features that are of the same spatial scale. mineral belt. This region is characterized by high and rough topography, recent (5 10 Ma), rapid exhumation, and the lowest Bouguer gravity anomalies in the western United States. Recent integration of EarthScope Transportable Array (TA) and PASSCAL/Flexible Array seismic data beneath the SRM and High Plains reveals an extraordinary image of the velocity complexities in the upper mantle including a fascinating array of blobs and dipping pipe like features extending to the 410 km discontinuity beneath Gunnison, CO and Chama, NM [Hansen et al., 2011]. Details of the upper mantle velocity anomalies along several profiles across the central Colorado Rockies from the CREST data are shown in Figure 1. These highresolution 3 D upper mantle images reveal large seismic velocity variations on a 50 100 km scale and are suggestive of interpretations relating the velocity anomalies to surface topographic features that are of the same spatial scale. [5] Several studies have noted that the topography of the SRM correlates better with the seismic velocity anomalies in the upper mantle than with crustal thicknesses [Humphreys and Dueker, 1994; Karlstrom and Humphreys, 1998] which supports the notion that isostatic compensation in this region takes place largely in the mantle. A more recent analysis of the receiver function images collected during the CREST experiment [Hansen et al., 2011] has shown that the SRM is not associated with thickened crust, suggesting upper mantle support of the high topography. Furthermore, it appears that the Colorado Plateau and SRM have an average lithospheric thickness of about 80 km which is thinner than generally accepted continental litho- 3of18

spheric average of 100 km which is observed further to the east under the Great Plains [Abt et al., 2010; Yuan and Romanowicz, 2010]. The relationship between the surface topography, crustal thickness and upper mantle processes in central Colorado remains poorly understood and constraints on the compensation mechanism provided by an analysis of the geoid is germane for resolving the contentious issues surrounding the lithospheric structure. Other studies, in particular the analysis of Rayleigh wave tomography [Li et al., 2002], have concluded the relationship between the elevated topography with mantle velocities is not as strong as the correlation with crustal anomalies. [6] In this study we build on the previous investigations into the coupling between elevation and thermal field in the context of isostasy and its role in continental deformation in convergent orogens [Sandiford and Powell, 1990; Zhou and Sandiford, 1992] and the lithospheric density moment as a source of intraplate stress and associated deformation at both global and tectonic plate scales [Coblentz et al., 1994]. We present the results of an evaluation of the upper mantle geoid and its relationship to the topography of the SRM in an effort to help resolve the controversy of how the topography is compensated and constrain the various geodynamic models being postulated to explain the observed upper mantle velocity anomalies beneath the SRM. We evaluate the geoid to topography ratio (GTR) for various compensation models using one dimensional isostatically balanced lithospheric density models to place constraints on the plausible lithospheric structure. This is followed by an interpretation of the coherence and admittance of the relationship between the topography and the gravity/ geoid fields in an attempt to distinguish between topography that is compensated by lithospheric variations and those that are supported by flexure of the lithospheric plate [Dorman and Lewis, 1970; Forsyth, 1985]. Our estimates for the mechanism and depth of compensation beneath SRM places strong constraints on the depth at which the buoyancy resides and help resolve the attendant issues surrounding uplift mechanisms in the western United States. 2. Topography, Gravity, and the Geoid [7] The high topography of the SRM is a dominant feature of the western United States (Figure 2a) and is part of a larger north south striking, longwavelength topographic swell (topography with l > 250 km is shown in Figure 2b) that extends for more than 1000 km from central New Mexico into Wyoming. This long wavelength swell cuts across major structural features in the crust including the northeast trending paleosubduction sutures inherited from the assembly of the North American plate, the northwest striking trends of the Ancestral Rocky Mountains, and the northeast trending Colorado Mineral Belt (see discussions by Eaton [1986, 2008]). From a geophysical point of view the lithosphere in the SRM region has been more intensely studied than any other orogenic system (with the possible exception of the Alps and parts of the Andes) and provides the data necessary to evaluate and understand how the topographic load of the southern Rockies is compensated. [8] As is well established by the principle of isostasy, topographic loads at the Earth s surface are compensated by mass deficits and excesses within the Earth s interior; Watts [2001] provides a comprehensive review of the development of the concept of isostasy. The small free air gravity anomaly throughout the western United States indicates that most of the tectonic provinces, including the Colorado Plateau and the Rockies are in a state of isostatic equilibrium. Thus, given the very long wavelength topographic features of the CP and the SRM, simple isostatic models can be employed to constrain the first order compensation mechanisms. The principle of parsimony suggests beginning with consideration of the Airy Heiskanen compensation (with purely crustal thickness variations) as the dominant compensation mechanism. For the simplest crustal model, a surface elevation of 2800 m will correspond with a crustal thickness of more than 48 km (assuming a crustal density of 2800 kg/m 3 and mantle lithospheric density of 3300 kg/m 3 ). However, the analysis of recent seismic imaging [Hansen et al., 2011] indicates that the crustal thickness under the high topography in the Aspen region of central Colorado is about 42 km or more than 5 km thinner than predicted by Airy Heiskanen compensation. This crustal thickness, combined with the lack of correlation between the highest topography and the largest Bouguer gravity anomalies, is evidence that the compensation mechanism beneath central Colorado is more complex. The situation calls for either a combination of both Airy Heiskanen and Pratt Hayford isostasy (see discussions by Martinec [1994a, 1994b]) or involvement of the whole lithosphere in the isostatic compensation (that is, a depth of compensation in the range of 100 150 km as suggested by Martinec [1994b]). As a result, in order to achieve more 4of18

Figure 2. Spatial variations in the (a) topography, (b) filtered topography (low pass l > 250 km), (c) upper mantlelithospheric geoid (see text for details), and (d) Bouguer gravity anomaly in the southern Rockies region of the western United States. White lines delineate the principal tectonic boundaries of the Basin and Range, Colorado Plateau, and southern Rockies. accurate estimates of the vertical distribution of compensating mass anomalies, isostatic calculations must consider the mantle lithosphere as well as the crust [Sandiford and Powell, 1990]. [9] The question of how the topographic load of the Rockies is compensated cannot be resolved by consideration of the Bouguer gravity anomalies alone. While gravity analyses have long been used 5of18

to constrain crustal structure in the western United States [Li et al., 2002; Thompson and Zoback, 1979], use of geoid anomalies (which are variations in the gravitational potential) are much better suited to the purpose of detecting where the compensating masses are located in the subsurface. Geoid anomalies caused by long wavelength isostatically compensated continental topography are proportional to the elevation multiplied by the mean depth of compensation (that is, they are proportional to the local dipole moment of the density depth distribution [Haxby and Turcotte, 1978]) and can be used to evaluate and constrain the subsurface distribution of mass associated with various topographic loads and features. Because the geoid reflects the gravitational dipole moment it is more sensitive to deeper sources than the gravity field, reflecting the fact that the geoid anomaly observed at the surface caused by a point mass buried at depth d decreases in amplitude as 1/d, whereas the gravity anomaly of the same point mass decays as 1/d 2. This makes the study of geoid anomalies more useful for discerning depths of isostatic compensation. In continental settings, the geoid has primarily been used to deduce regional tectonic stresses from the variations of gravitational potential energy [Jones et al., 1996, 1998; Sonder and Jones, 1999]. Recently, geoid anomalies have been used to constrain deep continental structure as well [Doin et al., 1996; Ebbing et al., 2001]. [10] There is a trade off, however, to the advantages of the depth sensitivity of geoid anomalies. Mantle heterogeneities produce long wavelength, highamplitude anomalies that dominate the geoid and consequently the observed geoid anomaly is the combination of mass anomalies in both the lithosphere and deeper mantle. Thus it is imperative to separate the upper mantle geoid anomalies from those with deeper sources, especially those that arise in the lower mantle before using the geoid for understanding the distribution of masses in the lithosphere. We accomplish this by removing all spherical harmonic terms of degree less than 7 and applying a one sided cosine taper to harmonic terms from degree 7 to 11 (see Chase et al. [2002] for details). This filter allows removal of all anomalies with spatial wavelengths greater than 3000 km and limits the sources of the anomaly to depths less than about 600 km. This geoid field (which we refer to as the upper mantle lithospheric geoid ) is shown in Figure 2c and appears to be appropriate for the southwestern North America, in that the average geoid value with crust near sea level is zero both on the west coast and the Gulf of Mexico (see Chase et al. [2002, Figure 4] for a more regional view of this geoid). It is important to note that this geoid is also consistent with the natural choice for a reference lithospheric column that has a near zero geoid associated with zero elevation; a feature that is exploited below in the discussion of partitioning deformational strain between the crust and mantle lithosphere. [11] Haxby and Turcotte [1978] showed that in the limit of long wavelength and complete isostatic compensation, the geoid anomaly caused by density variations above the depth of compensation is proportional to the vertical dipole moment of the mass distribution above the depth of compensation: DN ¼ 2G g Z D S zdðþdz z where DN is the geoid anomaly; G, Newton s gravitational constant; g, the value of gravitational acceleration on the reference ellipsoid; D, the depth of compensation (the depth below geoid at which pressure becomes equal beneath columns in isostatic balance, or here, where lateral density differences cease); S, the surface elevation; z, the depth; and Dr(z) the lateral difference in density from a reference lithosphere, including crust. There are two important implications of this formulation that are germane to this study: (1) the geoid anomaly over continents should increase both with increasing elevation and increasing average depth of compensation [Turcotte and Mcadoo, 1979; Turcotte and Schubert, 2002] and (2) for a particular elevation, the greater the depth extent of the isostatic root (or the deeper the source of buoyancy) the larger the resulting geoid anomaly. The linear relationship between gravity and topography can be used in a variety of ways for geodynamic applications. Two application of interest are of interest here. First, for a region where both the topography and gravity are measured over an area that is several times greater than the flexural wavelength (i.e., > 100 km), the gravity topography transfer function relationship (also referred to as the admittance or isostatic response function ) can be evaluated to extract information about the isostatic state of the lithosphere. This approach has also been used to evaluate the flexural rigidity of the lithosphere [McNutt and Menard, 1982] and to estimate the elastic thickness of the lithosphere [Banks et al., 1977; Dorman and Lewis, 1972; McKenzie and Bowin, 1976; McNutt and Menard, 1979; Watts, 1978]. A second application is for wavelengths greater than the flexural wavelength and where features are iso- 6of18

statically compensated, in which case the geoid/ topography ratio (GTR) can be used to estimate the depth of compensation in the spatial domain [Haxby and Turcotte, 1978]. A low GTR value (in the range of 2 4 m/km) indicates the topography is shallowly supported by thickened crust (dominantly Airy compensation), and intermediate GTR (in the range of 5 7 m/km) suggests deeper compensation that is related to lithospheric thinning [Crough, 1978] and a high GTR value (greater than 7 m/km) is reflective of dynamic support from sublithospheric sources such as mantle convection. As will be discussed below, however, these general values for GTR and corresponding depths of compensation (without reference to the surface elevation) are applicable only to Pratt Hayford isostasy; for other mechanisms of compensation including Airy Heiskanen and pure lithospheric thinning there is a linear correlation between the GTR and surface elevation. 3. Compensation Mechanisms in the Western United States [12] The near zero free air gravity anomaly throughout the western United States is indicative of long wavelength isostatic equilibrium and allows the use of isostatically balanced lithospheric columns for understanding the relative geometries of the crust and mantle lithosphere that result in plausible models for compensating the topography. It is clear that compensation is a process that incorporates the whole lithosphere; see, for example, the discussion about the case of the Colorado Plateau in [Coblentz et al., 2007]. As a result, in order to achieve more accurate estimates of the vertical distribution of compensating mass anomalies, isostatic calculations must consider the mantle lithosphere as well as the crust. [13] The f c f l diagram [Coblentz and Stüwe, 2002; Sandiford and Powell, 1990] provides a convenient way to evaluate how vertical strains in the lithosphere may be decoupled between crustal and mantle components. Following Sandiford and Powell [1990] we defined the f c as the ratio of deformed to referenced crustal thickness and f l as the ratio of the deformed to reference total lithospheric thickness. This approach using 1 D vertical columns to represent the lithosphere so deformed in this case is the vertical strain of the lithospheric column. For example, f c = 2.0 implies a doubling of the crustal thickness from the reference thickness to the deformed thickness. The origin of the f c f l plane (where f c = f l = 1.0) corresponds to an undeformed equilibrium lithosphere where there is no vertical strain. This is also the state that the crust and lithosphere will tend toward in the absence of deformational forces [Sandiford and Powell, 1990]. Completely undeformed lithosphere lies at the center of the of the f c f l plane with larger and smaller values of f c and f l corresponding to thicker and thinner crustal or lithospheric thicknesses, respectively. Here we assume a reference lithospheric column as 30 km of 2800 kg/m 3 crust overlying mantle lithosphere with a density of 3300 kg/m 3 extending to a depth of 100 km. [14] Deformational paths in the f c f l plane tend to be closed paths, starting at the origin and returning there eventually. Three typical deformational paths are shown in Figure 3. Deformation that involves only a change in crustal thickness without a change in the mantle lithospheric thickness, follows path A with f l = 1 until the crust is nearly doubled ( f c approaches 2.0). The surface elevation in response to this crustal thickening increases to a height of 4 km with a corresponding geoid anomaly of about 20 m. Alternatively, the case of thermal thinning (Path B) leads to a reduction in the mantle lithosphere thickness with no substantial change in crustal thickness. In this case, f c = 1 and f l < 1 and the deformation path in the f c f l plane is vertical, parallel to the f c axis. Path B in Figure 3 illustrates the response in the surface elevation ( 1.6 km) and geoid anomaly (5 m) associated with this type of lithospheric deformation. It clear from many recent studies (see review by Coblentz and Stüwe [2002]) that lithospheric deformation often involves a hybrid of these two cases. In the case of isostatic continental plateaus, lithospheric strain is typically partioned between crustal thickening ( f c > 1) and overall lithospheric thinning ( f l < 1). Path C in Figure 3 illustrates this type of composite deformation ( f c = 1.6 and f l = 0.61, defining crustal thickening to 48 km and mantle lithosphere thinning to 61 km). The resulting surface elevation is 3.7 km with a geoid anomaly 20 m. Below, we employ the f c f l approach to quantifying lithospheric strain of balanced 1 D isostatic columns (an appropriate approach given the very long wavelength topographic features under consideration) to evaluate the compensation mechanism for the southern Rockies and its neighboring tectonic provinces. [15] Figure 4 illustrates the GTR elevation values for the four fundamental compensation mechanisms: (1) Pratt, (2) pure Airy ( f c >1,f l = 1, corresponding to only crustal thickening, path A in Figure 3), (3) thermal thinning of the mantle litho- 7of18

Figure 3. (a) The f c f l space [Sandiford and Powell, 1990; Coblentz and Stüwe, 2002] contoured for surface elevation, geoid, and geoid elevation ratio based on deformation of the reference column shown in Figure 3b. Three deformation paths are illustrated: path A is pure crustal thickening ( f c >1and f l = 1), path B is pure mantle lithospheric thinning ( f c = 1 and f l < 1), and path C is a distributed strain case with both crustal thickening ( f c > 1) and mantle lithospheric thinning (f l < 1). (b) The lithospheric crust and mantle geometries resulting from these three cases. See text for discussion. spheric ( f c =1,f l < 1, path B in Figure 3), and (4) a combined strain path ( f c >1,f l < 1, corresponding to combined crustal thickening and mantle lithospheric thinning, path C in Figure 3). Also plotted in Figure 4 are the relative GTR elevation values for the Colorado Plateau (CP), southern Rocky Mountains (SRM), Northern Basin and Range (NBR), and Yellowstone (YS) to facilitate the evaluation of plausible compensation mechanisms. [16] Pratt type compensation (Figure 4a) yields a constant GTR value independent of surface elevation and for the mantle density contrast assumed here (r = 3300 km) a low GTR value ( 3 m/km) indicates shallow compensation supported by thickened crust (depth of compensation = 50 km), an intermediate GTR of 4 5 m/km corresponds to deeper compensation ( 75 km), and a GTR value 6 m/km corresponds to deep support (depth of compensation = 100 km). In the context of Pratt compensation, the topography of Yellowstone requires a very deep compensation depth of greater than about 120 km while the Colorado Plateau and southern Rockies can be supported with near crustal compensation depths of 55 km and 68 km, respec- 8of18

Figure 4. The predicted GTR elevation values for the four fundamental compensation mechanisms: (a) Pratt, (b) pure Airy (f c >1,f l = 1, corresponding to only crustal thickening (path A in Figure 3)), (c) thinning of the mantle lithosphere (thermal thinning) with f c =1,f l < 1 (path B in Figure 3), and (d) a combined strain path (f c >1,f l < 1, corresponding to combined crustal thickening and mantle lithospheric thinning (path C in Figure 3)). Also plotted are the relative GTRelevation values for the Colorado Plateau (CP), southern Rocky Mountains (SRM), Northern Basin and Range (NBR), and Yellowstone (YS). tively. We note that the Northern Basin and Range has a significantly deeper Pratt depth of compensation of about 87 km. [17] In contrast to Pratt compensation, the other three compensation mechanisms considered have a linear relationship between the GTR values and the surface elevation (z). The GTR for Airy compensation is proportional to h 2 and therefore the Airy GTR is a function of h with a relatively small slope (Figure 4b). The Airy GTR is also dependent on the L o (the reference lithospheric thickness) with a GTR variation of about 1.25 m/km as L o ranges from 60 km to 120 km. We note that given the crustal and mantle lithospheric density variations used here (r c = 2750 kg/m 3 and r c = 3300 kg/m 3 ), only the NBR province has GTR and surface elevation values that fall with the plausible ranges of Airy GTR. That is, the geoid for SRM is too low relative to the surface elevation for Airy compensation to be a viable compensation mechanism for the Rockies. [18] The GTR for thermal thinning (Figure 4c; mantle lithospheric thinning, f l < 1; and constant crustal thickness, f c = 1) differs significantly from the Airy GTR and is characterized by a steep, negative slope with respect to z. This results from the fact that the thermal density contrast resulting from 9of18

Figure 5. The location in f c f l space of plausible lithospheric geometries (indicated by the diamonds for the NBR, circles for the CP, triangles for YS, and squares for the SRM) for the four principal tectonic provinces of the western United States at consistent observed GTR elevation values. The initial total lithospheric thickness (L o ) for the initial reference column is indicted. While plausible geometries for most of the tectonic provinces are tightly clustered, the NBR is poorly constrained in terms of plausible mantle lithospheric thicknesses. Furthermore, we note that in contrast to the neighboring provinces, the southern Rockies province requires a thin initial lithospheric thickness of about 80 km. f l < 1 is much smaller than the crustal density contrast. This compensation mechanism has an upper elevation limit (designated by the termination of the curves in Figure 4c) that corresponds to the f l value that produces a mantle lithospheric thickness equal to the reference crustal thickness (30 km), i.e., where the mantle lithosphere disappears entirely. In the case of a reference lithospheric with a thickness of 100 km f l is limited to 0.30 and the maximum surface elevation is about 1.6 km. As illustrated in Figure 4c, the GTR elevation values of the NBR, CP and YS fall within the plausible ranges of the thermal thinning model. For example, the 2.4 km average surface elevation and 7 m/km GTR of Yellowstone can be produced by pure mantle lithospheric thinning of a 175 km reference lithosphere with a thinning factor of f l = 0.44 (producing a final lithospheric thickness, L d, of 77 km with an undeformed crust of 30 km). While the same deformational mechanism operating on a thinner reference lithosphere (L o = 120 km) results in GTR elevation values for the NBR and the CP that are consistent with observed geoid and elevation values, the required crustal (30 km) and mantle lithospheric geometries (NBR L d = 54 km, CP L d = 36 km) are inconsistent with known crustal and mantle lithospheric thicknesses for these provinces (i.e., z crust = 35 km in the NBR and 43 km in the CP and z lithosphere = 65 km and 98 km in the NBR and CP, respectively [Coblentz et al., 2007]). In the case of the southern Rockies, the observed surface elevation cannot be achieved through thinning of the mantle lithosphere alone, eliminating the plausibility of this compensation mechanism for supporting the topography of the Rockies. [19] As discussed in the Introduction, and highlighted by the results of the GTR elevation values for the pure mantle thinning case, the most likely lithospheric deformation scenarios in the western United States involve both crustal and mantle lithospheric strain. Relative to the reference lithospheric column, the tectonic provinces in the western United States are characterized by a combination of crustal thickening (f c > 1) and lithospheric thinning (f l < 1), and follow a deformation along path C in Figure 3. The GTR values are dependent on both the initial reference lithospheric thickness and the f c f l deformation path. The GTR elevation values for the path with f c varying from 1.0 to 1.5 as f l varies from 1.0 to 0.4 (roughly path C in Figure 3) are shown in Figure 4d. The lithospheric thinning results in a negative slope of the GTR as a function of elevation but the slope is less pronounced than for the pure thermal thinning case. For this particular deformation path, the GTR elevation values of the four provinces can be reproduced by varying the initial lithospheric thickness (L o ). [20] Plausible crustal and mantle lithosphere strains (as indicated by locations in f c f l space) that correspond with the observed GTR elevation values for the Colorado Plateau, Northern Basin and Range, Yellowstone, and the southern Rocky Mountains is shown in Figure 5. Here we have explored the solution space by evaluating a suite of initial lithospheric values (L o ) in the range of 60 to 140 km and calculating the f c f l ranges that generate the observed GTR elevation values for the four tectonic provinces. Lithospheric thickness values (L o ) vary from more than 140 km for Yellowstone to more intermediate values for the NBR (L o = 110 120 km) and the CP (L o =90 100 km) and relatively thin values for the SRM (L o = 80 km). While the range of plausible f c f l values for the provinces are generally tightly clustered, the exception is the Northern Basin and Range where a large number of f c f l combinations satisfies the observed GTR values. We note that crustal thickness in the range of 32 to 37 km and 10 of 18

overall lithospheric thickness values of about 52 to 90 km are indicated. Because the generally accepted lithospheric thickness in the Northern Basin and Range is about 65 km, the range of f c f l values delineated by the green filled circles (corresponding to a lithospheric than thinned from 120 km to a range of 60 to 72 km with) are the most plausible. The crustal thicknesses of the other three provinces considered fall in the range of 40 to 44 km which is consistent with the minimum end of the generally accepted values. This analysis indicates that with the exception of the Colorado Plateau, the major tectonic provinces of the western United States have all undergone more than 40 km of lithospheric thinning during their tectonic evolution. 4. Coherence and Admittance Analysis [21] Because compensation mechanisms eventually produce responses at different wavelengths it is not an easy task to separate the relative role of short (isostatic) versus long wavelength (flexure) compensation of topography using the 1 D analysis in section 3. To help estimate the relative roles of these two mechanisms, we turn next to an evaluation of the admittance and coherence functions (or the spectral measure of the gravity/geoid to topography relationship in the frequency domain). [22] The spectral relationship between topography and gravity has been used in many studies to evaluate the form of isostatic compensation in continental regions [Banks et al., 1977; Bechteletal., 1990, 1999; Dorman and Lewis, 1970; Forsyth, 1985; Londe, 1990; McKenzie and Bowin, 1976; Pilkington, 1990] in particular to help constrain estimates of the elastic thickness of the lithosphere. A similar analysis using the geoid has been commonly used to investigate the depth and mechanism of isostatic compensation in ocean plateaus and swells [Cserepes et al., 2000; Grevemeyer, 1999; Sandwell and Renkin, 1988; Sandwell and MacKenzie, 1989; Wessel, 1993]. [23] Other works have been focused on the viscous mantle flow associated with thermal convection and induced dynamic topography [Panasyuk and Hager, 2000]. Here we evaluate the coherence and admittance between the gravity (free air and Bouguer) and geoid fields and the topography along four profiles through the southern Rockies (38 N, 38.5 N, 39 N, and 39.5 N) extending from 117 W to 100 W. A composite image of the profiles (Figure 6a) illustrates the strong (qualitative) correlation between the Bouguer gravity, the geoid, the upper mantle velocity anomaly at 110 km [Schmandt and Humphreys, 2010], and topography across the southern Rockies (that is, the highest topography corresponds to the lowest Bouguer gravity anomaly, a pronounced negative Vp velocity anomaly, and the highest geoid anomaly). As noted previously, the free air gravity anomaly is near zero across most of this region (indicative of isostatic equilibrium). In order to include as much short wavelength geoid information as possible, we have used the harmonic expansion from EGM2008 [Pavils et al., 2008] which contains spectral information out to degree/order 2190 (corresponding to a spatial wavelength of about 18 km). Because this geoid is sensitive to sources in the upper mantle (harmonic depth cutoff of about 700 km), we have renamed this geoid the Upper Mantle Lithospheric Geoid. The spectral power of the Bouguer gravity, free air gravity, geoid and topography along the 38.5 N is shown in Figure 6b. All four share a characteristic red spectra (i.e., greater power at longer wavelengths) with the gravity fields (on the order of tens to hundreds of mgal) having significantly more spectral power than the topography (in the range of 0 to 4 km) and geoid (in the range of 0 to 10 m). We note that the spectral power of both the gravity and the topography geoid fields converge at wavelengths on the order of the lithospheric thickness; at 110 km for the Bouguer and free air gravity fields and 90 km for the geoid topography fields. [24] Any load on the lithosphere can, in general, be expressed as a Fourier series of different wavelength loads [Dorman and Lewis, 1970]. In the Fourier domain, the admittance, Z(k), is the ratio between the gravity anomaly, G(k), caused by a topographic load on an elastic plate and the magnitude of the topography, H(k), where G(k) and H(k) are the Fourier transforms of the Bouguer anomaly and the topography, respectively, and k is the wave number (k = 2p/l): GðkÞ ¼ Zk ð ÞHk ð Þ Thus the admittance represents a ratio of the amplitudes between gravity anomaly and topography. The coherence, on the other hand, is a measure of the correlation between the gravity anomaly and the topography; or more formally, the square of the correlation coefficient in the frequency domain between the two signals. The coherence and admittance and between the gravity/geoid fields and the topography for the four profiles are shown in Figure 7. 11 of 18

Figure 6. (a) Variations in the free air and Bouguer gravity anomalies, topography, upper mantle velocity anomaly at 110 km [Schmandt and Humphreys, 2010], and the upper mantle lithospheric geoid (see text for explanation) along four profiles centered on 38.5 N, 38.5 N, 39 N, and 39.5 N. The composite profiles illustrate the strong correlation between the high topography, upper mantle velocity anomaly, geoid, and Bouguer gravity across the southern Rockies. (b) The spectral power for the Bouguer gravity, free air gravity, geoid, and topography along the 38.5 N profile illustrating the characteristic red spectra of these fields. [25] The coherence between the Bouguer gravity and topography is particularly useful for evaluating the elastic thickness of the lithosphere (Te). Theoretical curves for various values of the elastic thickness (assuming a uniform load on the surface of the lithosphere and a crustal thickness of 30 km) are shown as the light blue lines in Figure 7a for a range of Te values (20 to 50 km). Large topographic loads will flex even very strong plates (i.e., Te = 50 km); isostatic compensation produces a large Bouguer anomaly and a coherence that approaches unity (for very long wavelengths). Shorter wavelength features, in contrast, will be adequately supported by the plate s mechanical strength, generating little or no associated Bouguer anomaly. Thus the coherence approaches zero for short wavelengths. The wavelength corresponding to the transition between coherent and incoherent gravity and topography provides a direct indication of the flexural rigidity (see discussion by Forsyth [1985], particularly on the subject of surface versus internal loading of the lithosphere). The wavelength corresponding to a coherence of about 0.5 provides an estimate of the length scale at which isostatic compensation begins 12 of 18

Figure 7. The (left) coherence and (right) admittance between the potential fields and topography: (a) Bouguer gravity, (b) free air gravity, and (c) geoid along the four profiles shown in Figure 6. Light blue lines in the Bouguer gravitytopography coherence plot are theoretical curves for various values of the elastic thickness (assuming a uniform load on the surface of the lithosphere and Moho thickness of 30 km). Theoretical admittance curves are shown in the Bouguer gravity topography admittance plot for values of r o = 2700 kg/m 3, dr = 670 kg/m 3, and z c = 35 km. The dark blue curve is the admittance curve for the pure Airy compensation case which corresponds to a flexural rigidity of zero. (d) Coherence between the Bouguer and free air gravity anomaly (blue) and Bouguer gravity anomaly and the geoid (red) illustrating the sharp dropoff in the coherence at wavelengths greater than 100 km. to prevail over mechanical support (as load size increases). As indicated by the theoretical curves in Figure 7, the transition wavelength for strong/thick plate can be as large as 600 km (for Te = 50 km) and will be considerable less for weak/thin plates (e.g., about 275 km for Te = 20 km). In the case of the profiles considered here, loads with l 225 km delineate the transition from mechanical to isostatic compensation. Referencing the theoretical curves in Figure 7, this is suggestive of an elastic thickness of less than 20 km. [26] The admittance also depends on the elastic thickness and can be viewed as a filter which predicts the gravity anomaly given the topography [McKenzie and Bowin, 1976]. As discussed by Forsyth [1985] the isostatic response function for 13 of 18

a model of a thin elastic plate loaded on top by topography of density r o is: where Qk ð Þ ¼ 2 o G expð kz c Þ= ¼ 1 þ Dk 4 = g where G is defined as the gravitational constant; D, the flexural rigidity of the plate; g, the gravitational acceleration; dr, the density contrast at the compensation interface (e.g., the crust mantle density contrast), and z c the depth to the density contrast (e.g., the crustal thickness). Theoretical admittance curves for values of r o = 2700 kg/m 3, dr = 670 kg/m 3, and z c = 35 km, are shown as light blue curves in the admittance plot in Figure 7. The dark blue curve is the admittance curve for the pure Airy compensation case which corresponds to a flexural rigidity of zero. The gravity anomaly produced by topography with l < 200 km is less than about 10 mgal/km and is consistent with a Te < 20 km. [27] For wholly uncompensated topography the free air anomaly rather than the Bouguer gravity will correlate with topography [Lambeck et al., 1988; McKenzie and Fairhead, 1997]. We note, however, that the coherence between the free air gravity and the topography (Figure 7b) is similar to Bouguer gravity coherence (particularly at wavelengths less than 200 km). For wavelengths greater than 300 km, however, the coherence of the free air and topography is significantly less than the Bouguer gravity topography coherence. Consistent with the general observation that the free air gravity anomaly is near zero for most of the region [Thompson and Zoback, 1979] the free air admittance is near zero for all wavelengths. [28] In contrast to the gravity fields, the coherence between the geoid and topography is high (greater than 0.75 for most wavelengths along the four profiles) substantiating the notion that there is a strong correlation between the geoid and topography. Figure 7 illustrates that this is the case for nearly all wavelengths of topography (we do note, however, there is considerable scatter in the coherence for the most northern of the profiles considered (39.5 N). The admittance between the geoid and topography is low (<2 m/km) for wavelengths less than about 200 km, and increases significantly to the range of 3 4 m/km at longer wavelengths. This is suggestive of significant, deep sources for the topographygeoid relationship most likely located in the upper mantle. 5. Discussion [29] Using the fact that at long wavelengths the GTR can be used to constrain the depth and mode of compensation [Sandwell and Renkin, 1988] we have evaluated various mechanisms by which the high topography of the southern Rocky Mountains can be isostatically compensated. Within the framework of the neighboring tectonic provinces there are several commonalities that have important tectonic implications. [30] We first considered the geoid elevation ratios based on first order 1 D isostatically balanced lithospheric columns. Using this approach, several generalities can be applied. In general, features with a GTR greater than 6 m/km are sublithospherically compensated and are dynamically supported by convective stresses [McKenzie et al., 1980]; that is, supported by dynamic topography generated by convection. The relatively high GTR value computed for Yellowstone ( 7 m/km) is consistent with this scenario. Low GTR values (<2 m/km) are associated with shallowly (<50 km) Airy or Pratttype compensated plateaus [Angevine and Turcotte, 1983]. We note that none of the western U.S. tectonic provinces considered here fit this category. Finally, intermediate GTR values in the range of 2 6 m/km reflect average compensation depths of 50 80 km and suggests reheating of the lower lithosphere [Crough, 1978]. Given that we find such GTR values for the western United States ( 3.9 for the southern Rockies, 4.25 for the Colorado Plateau, and 5.2 for the Northern Basin and Range), some kind of mantle lithospheric thinning in conjunction with crustal thickening is the most plausible compensation mechanism in the western United States. The notion that lithospheric strain is partioned between the crust and mantle lithosphere is supported by previous investigations [Sandiford and Powell, 1990]. [31] While isostatic elevation changes can be the result of either crustal or mantle lithosphere mechanisms, our analysis of the geoid topography relationship indicates that variations in the thickness of both the crust and mantle lithosphere are required to produce the modern elevation of the western United States in general and the southern Rockies in particular. In the case of the central Rockies, a combination of a thick crust ( 45 km) in conjunction with significant thinning of the mantle lithosphere 14 of 18

Figure 8. Spatial correlation between (a) the topography, (b) crustal thickness (a composite based on CRUST 5.1 and seismic results from analysis of the EarthScope transportable Array (TA), the PASSCAL/Flexible Array, and the CREST experiment), (c) the upper mantle Vp velocity anomalies at 110 km [Schmandt and Humphreys, 2010], and (d) the calculated geoid elevation ratios from this study. We note that the crustal thickness under the southern Rockies is nonthickened in the sense that observed values of 42 45 km are at odds with the high standing topography. since the Oligocene (30 Ma) (presumably related to delamination of a denser upper mantle or thermal erosion) can explain the recent and pronounced uplift. In terms of the neighboring provinces, we find that the Colorado Plateau topography can be supported by a combination of relatively undisturbed mantle lithosphere with significant crustal thickening (Figure 5); in the Northern Basin and Range, 15 of 18

in contrast, a wide range of mantle lithospheric strains in combination with moderate crustal thickening satisfies the observed GTR information. As expected, mantle thinning plays a large role for Yellowstone (with more than 50 km of thinning of initial very thick lithosphere. In general we find that with the exception of the Colorado Plateau, the major tectonic provinces of the western United States have all undergone more than 40 km of lithospheric thinning during their tectonic evolution. [32] Recent studies that evaluated the role of the crustal portion of lithospheric buoyancy and concluded that pure Airy type correlation between elevation and crustal thickness (i.e., thicker crust associated with higher elevation), exists only where the crust is very thick (>55 km) or very thin, less than 20 km [Zoback and Mooney, 2003]. The important tectonic implication is that non Airy lithospheric compensation is more common than traditionally thought. This is indeed the case for the southern Rockies where the emerging picture of a tectonic setting where the observed crustal thicknesses (Figure 8b) are not big enough to compensated the high topography (Figure 8a) by purely Airy compensation. The lack of sufficient crustal roots beneath the central Rockies in conjunction with the existence of profound upper mantle velocity variations (Figure 8c) and high coherence between the geoid and the topography (Figure 7) supports the notion that the anomalous upper mantle in the western United States (and the associated mantle buoyancy) is a significant contributor to surface elevation (a notion dating back many years to Thompson and Talwani [1964]). [33] Our conclusion that upper mantle dynamic processes play an important role in the support of the topography of the central Rocky Mountains raises the question of supporting, but plausible, geodynamic scenarios for the region. An assumption that observed upper mantle velocity variations (Figure 8c) are purely thermal, anelastic velocitytemperature mapping of velocity anomalies constrains the anomalies to be 200 300 C warmer than a 1300 C mantle adiabat for a pyrolitic mantle composition [Cammarano et al., 2003; Moucha et al., 2009; van Wijk et al., 2010]. Such thermal heterogeneity at short scales requires advective transport of warmer than normal mantle rising from a deep thermal boundary layer. At larger scales, inversion of a global seismic traveltime and geodynamic surface observables for a mantle thermal/ composition model [Simmons et al., 2009] find the southwestern United States to be a region of upwelling warm mantle from the base of the mantle [Moucha et al., 2009]. A plausible hypothesis then, is that complex small scale upper mantle convection and resulting buoyancy variations are driving of dynamic uplift (on the order of 400 800 m [van Wijk et al., 2010]) in many parts of the western United States and is actively reshaping the physiographic characteristics of western U.S. tectonic provinces. This observation is consistent with recent Apatite fission track thermochronology that shows rapid exhumation in the Rockies starting 10 6 Ma at modern elevations ranging from 1.5 to 3 km, requiring Neogene tectonic uplift [Karlstrom et al., 2008, 2010]. 6. Conclusions [34] Within the framework of the assumptions made for this study, we draw two principal conclusions: (1) a major portion of the buoyancy that has driven uplift in the southern Rockies resides at depths less than 100 km and that upper mantle processes such as small scale convection may play a significant role in the buoyant uplift of the southern Rockies (as well as other actively uplifting areas of the western United States) and (2) the high coherence for the geoid topography relationship for nearly all wavelengths between 50 and 1000 km suggests an intimate relationship between upper mantle dynamic processes and the surface topographic character. Acknowledgments [35] This work has benefited from productive conversations with a number of investigators. In particular, we acknowledge the insightful comments and suggestions provided by the CREST working group (including but not limited to Ken Dueker, Rick Aster, Jonathan MacCarthy, Andres Aslan, and Sheri Kelley). Randy Keller and an anonymous reviewer are thanked for insightful comments. Jonathan MacCarthy is thanked for providing Figure 1. This manuscript is released under Los Alamos National Laboratory LAUR 10 08105. References Abt,D.L.,K.M.Fischer,S.W.French,H.A.Ford,H.Y. Yuan, and B. Romanowicz (2010), North American lithospheric discontinuity structure imaged by Ps and Sp receiver functions, J. Geophys. Res., 115, B09301, doi:10.1029/ 2009JB006914. Angevine, C. L., and D. L. Turcotte (1983), Correlation of geoid and depth anomalies over the Agulhas Plateau, Tectonophysics, 100(1 3), 43 52, doi:10.1016/0040-1951(83) 90177-4. Aster,R.,J.MacCarthy,M.T.Heizler,S.A.Kelley, K. Karlstrom, L. Crossey, and K. Dueker (2009), Experiment 16 of 18