Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls...

24
Inferring crustal structure in the Aleutian island arc from a sparse wide-angle seismic data set Harm J. A. Van Avendonk University of Texas at Austin, Institute for Geophysics, 4412 Spicewood Springs Road, Austin, Texas 78759, USA ([email protected]) Donna J. Shillington, W. Steven Holbrook, and Matthew J. Hornbach Department of Geology and Geophysics, University of Wyoming, P.O. Box 3006, Laramie, Wyoming 82071, USA [1] Compressional seismic travel times from a relatively sparse wide-angle data set hold key information on the structure of a 800 km long section of the central Aleutian arc. Since the source and receiver locations form a swath along the arc crest that is 50 km wide, we trace rays in 3-D for a collection of 8336 seismic refraction and reflection arrivals. We investigate variations in seismic velocity structure parallel to the Aleutian arc, assuming that our result represents average crustal structure across the arc. We explore seismic velocity models that consist of three crustal layers that exhibit smooth variations in structure in the 2-D vertical plane. We consider the influence of additional constraints and model parameterization in our search for a plausible model for Aleutian arc crust. A tomographic inversion with static corrections for island stations reduces the data variance of a 1-D starting model by 91%. Our best model has seismic velocities of 6.0–6.5 km/s in the upper crust, 6.5–7.3 km/s in the middle crust, and 7.3 – 7.7 km/s in the lower crust and a total crustal thickness of 35 – 37 ± 1 km. A resolution analysis shows that features having a horizontal scale less than 20 km cannot be imaged, but at horizontal length scales of 50 km most model features are well resolved. The study indicates that the Aleutian island arc crust is thick compared to other island arcs and strongly stratified and that only the upper 60% of the arc crust has seismic velocities that are comparable to average seismic velocities in continental crust. Components: 12,601 words, 22 figures, 2 tables. Index Terms: 3025 Marine Geology and Geophysics: Marine seismics (0935); 8105 Tectonophysics: Continental margins and sedimentary basins (1212); 8150 Tectonophysics: Plate boundary—general (3040). Received 11 November 2003; Revised 23 April 2004; Accepted 21 June 2004; Published 25 August 2004. Van Avendonk, H. J. A., D. J. Shillington, W. S. Holbrook, and M. J. Hornbach (2004), Inferring crustal structure in the Aleutian island arc from a sparse wide-angle seismic data set, Geochem. Geophys. Geosyst., 5, Q08008, doi:10.1029/ 2003GC000664. 1. Introduction [2] Seismic refraction profiling has contributed much to our understanding of the deep structure of the Earth’s crust [Holbrook et al. , 1992; Christensen and Mooney , 1995]. Wide-angle seis- mic imaging can deliver compelling results be- cause of its ability to investigate large-scale average properties of the crust but both the strengths and limitations of this technique must be recognized. Pitfalls and successful strategies for modeling seismic refraction data have been discussed by Zelt [1999] in a general sense, but the variety in geological settings and experimental design is large. New insight can therefore be gained by exploring modeling options for a single data set. [3] In 1994 a seismic refraction experiment in the Aleutian islands shed a new light on island arc systems, and the role that magmatic arcs may play G 3 G 3 Geochemistry Geophysics Geosystems Published by AGU and the Geochemical Society AN ELECTRONIC JOURNAL OF THE EARTH SCIENCES Geochemistry Geophysics Geosystems Article Volume 5, Number 8 25 August 2004 Q08008, doi:10.1029/2003GC000664 ISSN: 1525-2027 Copyright 2004 by the American Geophysical Union 1 of 24

Transcript of Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls...

Page 1: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Inferring crustal structure in the Aleutian island arc from asparse wide-angle seismic data set

Harm J. A. Van AvendonkUniversity of Texas at Austin, Institute for Geophysics, 4412 Spicewood Springs Road, Austin, Texas 78759, USA([email protected])

Donna J. Shillington, W. Steven Holbrook, and Matthew J. HornbachDepartment of Geology and Geophysics, University of Wyoming, P.O. Box 3006, Laramie, Wyoming 82071, USA

[1] Compressional seismic travel times from a relatively sparse wide-angle data set hold key informationon the structure of a 800 km long section of the central Aleutian arc. Since the source and receiverlocations form a swath along the arc crest that is �50 km wide, we trace rays in 3-D for a collection of8336 seismic refraction and reflection arrivals. We investigate variations in seismic velocity structureparallel to the Aleutian arc, assuming that our result represents average crustal structure across the arc. Weexplore seismic velocity models that consist of three crustal layers that exhibit smooth variations instructure in the 2-D vertical plane. We consider the influence of additional constraints and modelparameterization in our search for a plausible model for Aleutian arc crust. A tomographic inversion withstatic corrections for island stations reduces the data variance of a 1-D starting model by 91%. Our bestmodel has seismic velocities of 6.0–6.5 km/s in the upper crust, 6.5–7.3 km/s in the middle crust, and7.3–7.7 km/s in the lower crust and a total crustal thickness of 35–37 ± 1 km. A resolution analysis showsthat features having a horizontal scale less than 20 km cannot be imaged, but at horizontal length scales of�50 km most model features are well resolved. The study indicates that the Aleutian island arc crust isthick compared to other island arcs and strongly stratified and that only the upper 60% of the arc crust hasseismic velocities that are comparable to average seismic velocities in continental crust.

Components: 12,601 words, 22 figures, 2 tables.

Index Terms: 3025 Marine Geology and Geophysics: Marine seismics (0935); 8105 Tectonophysics: Continental margins

and sedimentary basins (1212); 8150 Tectonophysics: Plate boundary—general (3040).

Received 11 November 2003; Revised 23 April 2004; Accepted 21 June 2004; Published 25 August 2004.

Van Avendonk, H. J. A., D. J. Shillington, W. S. Holbrook, and M. J. Hornbach (2004), Inferring crustal structure in the

Aleutian island arc from a sparse wide-angle seismic data set, Geochem. Geophys. Geosyst., 5, Q08008, doi:10.1029/

2003GC000664.

1. Introduction

[2] Seismic refraction profiling has contributedmuch to our understanding of the deep structureof the Earth’s crust [Holbrook et al., 1992;Christensen and Mooney, 1995]. Wide-angle seis-mic imaging can deliver compelling results be-cause of its ability to investigate large-scaleaverage properties of the crust but both thestrengths and limitations of this technique must

be recognized. Pitfalls and successful strategiesfor modeling seismic refraction data have beendiscussed by Zelt [1999] in a general sense, butthe variety in geological settings and experimentaldesign is large. New insight can therefore begained by exploring modeling options for a singledata set.

[3] In 1994 a seismic refraction experiment in theAleutian islands shed a new light on island arcsystems, and the role that magmatic arcs may play

G3G3GeochemistryGeophysics

Geosystems

Published by AGU and the Geochemical Society

AN ELECTRONIC JOURNAL OF THE EARTH SCIENCES

GeochemistryGeophysics

Geosystems

Article

Volume 5, Number 8

25 August 2004

Q08008, doi:10.1029/2003GC000664

ISSN: 1525-2027

Copyright 2004 by the American Geophysical Union 1 of 24

Page 2: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

in the evolution of continental crust (Figure 1).Current working hypotheses for the magmafluxes and the construction of volcanic arcs havemostly been tested against petrological data [Kay,1980; Myers et al., 1986; Grove and Kinzler,1986; Kelemen, 1995], but seismic constraints onthe deeper sections of the Aleutian arc offer analternative method for quantifying the volume ofmagmatic additions to the crust and inferring itsbulk composition. The 1994 wide-angle surveyconsisted of two transects across the volcanic arc,and a profile along the strike of the arc connect-ing the other two profiles (Figure 1b). The twoseismic velocity profiles perpendicular to theisland arc, labeled A1 and A3 in Figure 1b,resulted in detailed cross sections of the arc andsubducting Pacific plate [Holbrook et al., 1999;

Lizarralde et al., 2002]. The arc-parallel transectA2, published by Fliedner and Klemperer[1999], was designed to capture variability ofthe crustal structure along the arc between thetwo cross-arc transects. With thirteen receiversdeployed along an 800 km long section of thearc (Figure 1), a seismic velocity model of lineA-2 cannot be interpreted without knowledge ofits nonuniqueness. We used the travel times froma new interpretation of the seismic refraction dataof line A-2 presented in a future paper byShillington et al. [2004]. In our paper we inves-tigated uncertainties in seismic velocity imagingthat result from travel time errors, sparse datacoverage, model parameterization, and the choiceof additional constraints for tomographic inver-sions of the line A-2 data set. Shillington et al.

Figure 1. (a) Study area covers the central Aleutian islands, southwest Alaska, from the islands of Unimak to Atka.The ship track of line A-2 is shown in red. (b) Wide-angle data of line A2 cover an area of 830 � 160 km along theisland arc. The model kilometers used in this plot are consistent with figures of seismic velocity models further in thepaper. Two of the instruments (triangles), A2A3 and A2C3, were ocean bottom seismometers; all others were Reftekland seismometers. The sites Atka and Fort Glenn were each occupied with two instruments. The ship tracks areshown as a dashed line.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

2 of 24

Page 3: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[2004] compare our imaging results with the tec-tonic segmentation and major element geochemis-try of the Aleutian arc. In order to use empiricalrelationships between rock composition and seis-mic velocity [Behn and Kelemen, 2003] we mustunderstand the errors in our velocity model well.

[4] A number of tomographic inversion schemessearch for a seismic velocity model that ade-quately satisfies travel time constraints alongwith other considerations [e.g., Hole, 1992;Koch, 1993; Zelt et al., 1999; Van Avendonket al., 2001b]. Ideally, we should formulatethese additional criteria in a way that incorpo-rates other geological information from theAleutian volcanic arc. Such an implementationis not practical because the seismic velocitystructure of the Aleutian arc, as other types ofcrust [Goff and Holliger, 2003], will vary onlength scales that are much smaller than whatwe can resolve with a seismic refraction study.Our best option is to solve for a minimum-structure seismic velocity model and to acknowl-edge that our image averages the structure of thetrue Earth over horizontal and vertical distances[Backus and Gilbert, 1970; Chou and Booker,1979]. Estimates of these averaging widths areessential for the interpretation of seismic velocitymodels.

[5] Receivers on line A2 recorded up to fourpairs of reflected and refracted wave arrivals thatoriginate from different levels in the crust anduppermost mantle [Shillington et al., 2004].Segmentation of seismic travel time branchesoffers a diagnostic clue to the large-scale struc-ture of the Aleutian arc. Shillington et al. [2004]found that these travel time branches correlatewell between receivers, such that a model con-sisting of three discrete layers over a half-space,extending over the full length of line A2, bestexplains the seismic arrivals in the wide-angledata of line A2. We therefore exclusively ex-plored layered seismic velocity models to fit thetravel times of the line A2 data set. The assump-tion that seismic reflections are associated withmajor structural boundaries is not new, but stud-ies of crustal reflectivity have indicated that smallheterogeneities, unrelated to macroscopic velocitystructure, can give rise to apparently coherentseismic reflections [Levander and Gibson, 1991;Smithson et al., 2000]. Fliedner and Klemperer[1999] interpreted the wide-angle reflections fromwithin the arc crust as local events, and insteadof a layered structure they presented a model

without first-order discontinuities in seismicvelocity.

[6] The source-receiver layout of line A2 is some-times referred to as a crooked line geometry [Zeltand Zelt, 1998; Zelt, 1999]. The ship track of theR/V Maurice Ewing curved around the southernshores of the islands, and a number of land stationsrecorded the air gun shots from 20 to 40 km northof the shot line. The experimental geometryrequires 3-D ray tracing to properly calculate raypaths and travel times for a seismic velocity model.An inversion for 2-D seismic velocity variationsalong the strike of the volcanic arc may be valid ifthe source-receiver paths do not sample significantvariations in structure perpendicular to the arc. Thiscondition depends on the ray geometry and lateralheterogeneity in the model. We followed the 3-Dray tracing/2-D inversion approach, named 2.5-Dinversion by Zelt [1999], which considerablyreduces the nonuniqueness of the modeling solu-tion. However, after 2.5-D inversion we mustevaluate whether lateral velocity variations areweak enough to justify this approach.

[7] Even if we limit our search to 2-D models, thelarge distances between instruments on line A2leaves much of the model space poorly sampled.An iterative least squares inversion method [VanAvendonk et al., 2001b] gives the opportunity tofind formal expressions of the model covarianceand resolution [Jackson, 1979; Tarantola, 1987].Within the limits of the model resolution we foundthat the crust of the Aleutian arc is characterized byseismic velocities that are much higher than inaverage continental crust. Our result therefore givesmore evidence for the argument that the crust ofisland arcs must undergo significant changes dur-ing collision with continents if they have contrib-uted to the continental volume since the Archean[McLennan and Taylor, 1982; Rudnick, 1995].

2. Refraction and ReflectionTomography

[8] Seismic travel time tomography is generallyregarded as an objective tool for interpreting traveltime data [Iyer and Hirahara, 1993; Thurber,2003], although differences exist between imple-mentations [Phillips and Fehler, 1991; Zelt, 1999].The technique starts with the choice of an appro-priate reference model in which rays are tracedbetween the sources and receivers. The travel timeresiduals dti, the difference between observed andpredicted travel time for source-receiver pair i,

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

3 of 24

Page 4: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

form the input in a linearized inversion. UsingFermat’s principle, dti can be expressed in a linearcombination of perturbations du of slowness (thereciprocal of seismic velocity) and changes in thedepth of layer boundaries drk at each layer k,provided that these model perturbations are small.

dti �Zi

du x; zð ÞdsþXk

Gi;k xð Þdrk xð Þ ð1Þ

The integration of du(x, z) is carried out over the ithray path in the reference model. Expressions for theFrechet derivatives Gi,k(x) for reflector depth rk(x)have been derived by a number of authors [Bishopet al., 1985; Nowack and Lyslo, 1989; Sambridge,1990].

[9] Travel time branches observed in the line A2wide-angle data [Shillington et al., 2004] can beexplained by a model with three crustal layers.Seismic refractions and reflections emanating fromthe layer boundaries were picked on most receivergathers (Table 1). The emergence of middle andlower crustal refractions P2 and P3 as first arrivalsat source-receiver offsets of �100 km in somereceiver gathers (Shillington et al., submitted man-uscript, 2004) indicates that the average seismicvelocity increases across every layer boundary. Wetherefore adopt a strategy to search for a seismicvelocity model with first-order discontinuities atthe layer boundaries rk(x) (Figure 2). This assump-tion is fairly common for minimum-parametertravel time inversions [Zelt and Smith, 1992; Wangand Braile, 1996], but in many fine-gridded traveltime inversion methods the reflecting boundariesrk(x) are ‘‘floating’’ in a continuous seismic veloc-ity model [Bishop et al., 1985; Hole et al., 1992;Delprat-Jannaud and Lailly, 1995; Zhang and

Toksoz, 1998; Zelt et al., 1999]. McCaughey andSingh [1997], Ditmar et al. [1999], and Hobro etal. [2003] allowed velocity discontinuities by let-ting the seismic velocity layers overlap across rk(x).As an alternative we propose a model parameter-ization where the step in seismic velocity acrossthese boundaries is explicitly defined. The materialslowness u(x, z) is split in two components:

u x; zð Þ ¼ u0 x; zð Þ þ uk xð Þ ð2Þ

Gradual slowness variations, or vertical gradientsin u(x, z) within a layer, are represented by u0(x, z) inequation (2). The uk(x) in equation (2) is a step inu(x, z) across the kth layer boundary in the model.Substituting equation (2) in equation (1) we get

dti �Zi

du0 x; zð ÞdsþXk¼1

Z

k;i

duk xð ÞdsþXk

Gi;k xð Þdrk xð Þ

ð3Þ

The integration of du0(x, z) is carried out over theentire ray path i. The integration of the perturba-

Figure 2. Schematic overview of our interpretation ofseven ray path groups shows that they are consistentwith a three-layer crust overlying the mantle.

Table 1. Number of Travel Time Picks for Each Instrument and Phase

Instrument P1 P2P P2 P3P P3 PMP Pn All

Atka 0 0 0 64 291 188 0 543A2A3 120 41 0 51 0 40 0 252A2C3 113 0 0 0 0 0 0 113Nikolski 382 202 288 115 13 177 54 1231Fort Glenn 556 97 34 70 0 198 0 955Chernofski 351 131 276 62 0 285 30 1135Kashega 315 156 168 147 245 100 26 1157Captain’s Bay 364 170 305 0 48 143 38 1068Beaver Inlet 201 77 124 50 0 0 0 452Akutan 352 87 136 62 26 0 0 663Akun 219 142 250 48 0 108 0 767Total 2973 1103 1581 669 623 1239 148 8336Percent of picks 35.7 13.2 19.0 8.0 7.5 14.9 1.8 100.0

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

4 of 24

Page 5: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

tions duk(x), however, is carried out over thesegment of the ith ray path that lies within layer k(Figure 3). We choose bilinear basis functions forthe model parameters u(x, z) and rk(x), such thatequation (3) can be expressed in matrix form[Thurber, 1983].

d ¼ Gdm ð4Þ

The data vector d contains the travel time residuals,matrix G is the Frechet matrix, and the modelperturbation dm is a vector that includes du0, duk,and drk. Figure 3 shows the model parameters thatare influenced by one ray path. Since ray paths arethin and the parameterization is local, matrix G issparse [Nolet, 1985]. The mixed parameterization ofthe model can pose a problem if we solve forslowness and interface depth simultaneously. Wesolve all for model parameters together, but wescale dm with fixed constants for slowness andinterface depth.

[10] We employ a fine grid for slowness andinterface depth, such that the size of dm (N) willbe considerably larger than the size of d (M). Inthis approach, infinitely many models can fitequation (4) and a minimum structure model ispreferred [e.g., Constable et al., 1987]. In regular-ized inversions we minimize a cost function F(dm)that includes the data misfit equation (4) inconjunction with a norm for dm that limits themodel roughness [Tikhonov and Arsenin, 1977;

Spakman and Nolet, 1988; Zhang and Toksoz,1998]:

F dmð Þ ¼

C�1

2

d Gdm� dð Þ

lD mþ dmð Þ

gIdm

����������

����������

����������

����������

2

ð5Þ

The a priori data covariance Cd is a diagonal matrixwith data variances si

2 if the data errors can beassumed independent. The roughness operator Dcan be a matrix with first or second derivatives,depending on whether a flat or a smooth finalmodel is preferred [e.g., Zhang and Toksoz, 1998;Zelt, 1999]. By letting D operate on the futuremodel m + dm we do not impose constraints on theroughness of dm, which will help us find a newmodel that is not biased toward the reference modelm [Shaw and Orcutt, 1985; Scales et al., 1990].The Tikhonov trade-off parameters l and g mustbe tuned such that minimizing equation (5) leads toa c2 data misfit close to 1 [Scales et al., 1990; VanAvendonk et al., 1998]. The parameter l limits themodel roughness in the inversion. It may also benecessary to suppress modeling artifacts by limit-ing the size of dm with damping matrix gI. We findan approximate least squares estimate dm byminimizing equation (5) using a sparse matrixsolver such as LSQR [Paige and Saunders, 1982;Nolet, 1985; Spakman and Nolet, 1988]. Afterupdating the model m with dm we can evaluate thedata fit by comparing picked and calculated traveltimes:

c2 ¼ 1

M

XMi¼1

C�1d tpick � tcalc� �2 ð6Þ

However, the ray-tracing process is computation-ally too expensive to explore the tuning parametersl and g for an acceptable solution dm. Wetherefore approximate c2 by substituting dm fordm in equation (4) and take the Eucledian norm:

c2 � 1

MC

�12

d Gdm� dð Þ���

������

���2 ð7Þ

The inversion may not give a satisfactory solutionif the starting model m is not close to the truemodel, in which case the approximations inequation (1) are not accurate. Iteration of the raytracing and norm minimization should lead tobetter a data fit, and the model perturbations dmrequired to fit the data should become small after afew iterations [e.g., White, 1989; Hole, 1992].

Figure 3. Detail of a velocity model with a param-eterization that represents two layers of the crust be-neath water and sediments. A ray path passes throughthe first layer and turns in the second layer. The threetypes of model parameters are continuous slownessvariations u0 (circles), interface boundaries rk (squares),and slowness discontinuities uk at these boundaries(triangles). Model parameters that are sampled by theray path are shown in white; others are shown in gray.The ray path activates the r and u0 parameters in itsvicinity and all uk parameters at the interface directlyabove the ray.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

5 of 24

Page 6: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[11] Interpretation of the new model m + dmrequires insight in the model fidelity. We canexpress modeling uncertainties in the a posterioricovariance matrix Cm and in the amount of spatialaveraging in the resolution matrix R [e.g., Jackson,1979; Menke, 1984; Tarantola, 1987]. We assumethat a series of linearized inversions yields a stablesolution m0, and that subsequent model perturba-tions dm have zero mean and obey Gaussianstatistics. If no damping is used in the lastlinearized inversion we set g = 0 in equation (5)and express the optimal solution dm in the normalequations:

dm ¼ GTC�1d Gþ l2DTD

� ��1GTC

�12

d d ð8Þ

Jackson [1979] also showed that the covariance ofresolving errors can be written as

Cm ¼ GTC�1d Gþ l2DTD

� ��1¼ ATA� ��1

; ð9Þ

where

A C

�12

d G

lD

0@

1A ð10Þ

The calculation of the inverse of the large, sym-metric matrix in equation (9) requires an iterativestrategy, which is outlined in Appendix A. Bycombining equation (4) in equation (8) we canexpress the imaged dm in the ‘‘true’’ dm:

dm ¼ GTC�1d Gþ l2DTD

� ��1GTC�1

d Gdm ð11Þ

The resolution matrix

R ¼ GTC�1d Gþ l2DTD

� ��1GTC�1

d G

¼ CmGTC�1

d G ð12Þ

is an averaging window between dm and dm that isoften used to show the degree of nonuniqueness inseismic velocity models.

[12] The full N � N model covariance and resolu-tion matrices are large, and the calculation ofequations (9) and (12) requires O(N3) operations.When dealing with large models it may thereforebe necessary to limit the computation of Cm and Rto those matrix elements that represent theinteraction between model parameters that arephysically close.

[13] When the relevant parts of Cm and R arecalculated we must represent them in a form thatshows the reliability of the model features of

interest. We therefore aim to find the uncertaintiesin spatial averages of the model vector m. Weconsider an averaging window

a ¼

0

..

.

w

..

.

0

0BBBBBBBBBBBBB@

1CCCCCCCCCCCCCA

; ð13Þ

with kak = 1, that finds the average materialslowness ua over a region A:

ua ¼ a �m ð14Þ

If b is a similar averaging window for region B, wecan express the covariance between the slownessaverages ua and ub by applying a and b to themodel covariance matrix (9):

Cab ¼ a � Cmb ð15Þ

Similarly, we find covariances Caa and Cbb anddefine

Cab ¼Caa Cab

Cab Cbb

0@

1A ð16Þ

If Gaussian statistics apply for ua and ub, we canassign a probability density function to variationsin average slowness:

P dua; dubð Þ / exp � 1

2

dua

dub

0@

1A � C�1

ab

dua

dub

0@

1A

8<:

9=; ð17Þ

[14] We also calculate the joint probability densityfunction for slowness and boundary depth. Theprobability density function for variations in over-lying slowness du and interface depth drk takes theform

P du; drkð Þ / exp � 1

2

du

drk

0@

1A � C�1

ur

du

drk

0@

1A

8<:

9=;; ð18Þ

where

Cur ¼Cuu Cur

Cur Crr

0@

1A ð19Þ

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

6 of 24

Page 7: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[15] The summation of Cm is an example of howthe N � N covariance and resolution matrices canbe presented in a graphical form. The matrices Cm

and R provide important information on thegoodness of the model solution, but it isimportant that the assumptions in their derivationare understood.

3. Seismic Velocity Structure

[16] The seismic refraction data of line A2 werediscussed by Shillington et al. [2004]. They incor-porated multichannel seismic (MCS) data to con-strain the basement depth, and picked 8336 traveltimes in the wide-angle data (Table 1). Errorsassigned to the wide-angle data ranged from 100to 200 ms. In the tomographic inversion presentedin this paper, we did not model the MCS andwide-angle data together, because the acousticbasement relief varies on length scales muchshorter than what we can resolve in the crustwith the wide-angle data. We took the basementdepth and seismic velocities in the overlyingsediments from Shillington et al. (submitted manu-script, 2004) and constructed a starting model thatis consistent with the travel time phases (Figure 2).In this model, the seismic velocity varies from 6.2km/s below basement to 7.5 km/s in the lower-most crust and 8.1 km/s in the upper mantle,similar to the structure around Unimak [Abers,1994] (Figure 4). Since the data coverage isspread fairly evenly over the model layers(Table 1) we did not apply layer stripping in thecrust [Zelt, 1999], but inverted all wide-angledata simultaneously to update the crustal seismicvelocity structure.

[17] We traced ray paths in the 3-D volume usingthe shortest path method [Moser, 1991] and raybending [Um and Thurber, 1987; Van Avendonk etal., 2001a]. The combination of widely spacedinstruments and a virtually continuous shot lineresulted in a row of intertwined ray fans (Figure 5).Comparing the picked travel times with traveltimes calculated in the starting model, we founda c2 of 16.34 before inversion. As discussed in theIntroduction, we reduced the number of degrees offreedom by carrying out a 2.5-D tomographicinversion [Zelt and Zelt, 1998], but we must becautious with the 3-D structure in the Aleutian arc.The paths between the shot line and the thirteenstations mostly lie on the arc platform (Figure 1),and we therefore assumed that the deeper phasesP3, PMP and Pn all sample the Aleutian arc where itis thick [Holbrook et al., 1999; Lizarralde et al.,

2002]. The azimuthal distribution in ray paths(Figure 6) shows that 66% of all paths lie within30� of the strike of the volcanic arc. The volcanicislands, on which all 11 land stations weredeployed, may exhibit a shallow structure that isdifferent from other parts of the arc (Figure 1).After examining the travel residuals in the startingmodel, it appeared that arrivals recorded by landstations were consistently �0.4 s early, but thisconstant time shift was not observed in the OBSsA2A3 and A2C3. The 0.4 s was subtracted fromthe observed travel times for all land instrumentsbefore the tomographic inversion, and reduced theinitial c2 data misfit to 4.89.

[18] We parameterized u0(x, z) with a grid with7.5 km spacing along the strike of the arc, and avertical grid spacing that varied from 1.5 km nearthe surface increasing linearly with depth to 3.5 kmin the upper mantle. The depth of the reflectingboundaries rk(x) and slowness discontinuities uk(x)were parameterized with the same horizontal spac-ing (Figure 3). Of the 3277 model parameters,2792 were sampled by at least one ray path and

Figure 4. The starting model for the tomographicinversion is a nearly 1-D seismic velocity structurehanging from the acoustic basement in the rectangulararea shown in Figure 1b. The sediment and basementstructures were determined from reflection seismic data(Shillington et al., submitted manuscript, 2004) andremain fixed in all inversions presented in this paper.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

7 of 24

Page 8: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

therefore included in our model vector m. Later inthis section we show the role of the modeldiscretization in the tomographic inversion. Aslong as the grid spacing is smaller than the lengthscales that can be resolved in the inversion, theparameterization should not significantly restrictthe model space.

[19] We carried out nine iterations of the linearized2-D inversion followed by 3-D ray tracing in theupdated model. We employed a roughness matrixD that contains both first and second derivatives toregularize the inversion. Horizontal derivativeswere given a 4.0 times larger weight than verticalderivatives to account for the fact that we expect theseismic velocity structure to be rougher in thevertical direction [Van Avendonk et al., 1998].Damping was applied to the first three iterations,but in later iterations dm was sufficiently small toset g = 0 in equation (5). By minimizing equation(5) for a few values of l we found a range of dm,and we used equation (7) to obtain the correspond-ing c2 without ray tracing in the new model. Fromthis suite of solutions we created the L curve[Lawson and Hanson, 1974], which shows thetrade-off between the model norm and data misfit(Figure 7). The well-behaved curve made it easy tofind the appropriate l for any desired c2 by usingthe secant method [Van Avendonk et al., 1998].Fortuitously, the model with c2 = 1.00 lies nearthe corner of the L curve, where the curvature islarge [Hansen, 1992; Reginska, 1996]. This modelreproduces travel time residuals that are similar insize to the assigned data errors.

[20] The c2 error statistic plays an important rolein our model search, but the errors that affect theinversion are not easy to estimate. Even if errors inthe travel time picks can be determined, errors inthe travel time phase interpretation, ray tracing, ormodeling assumptions also contribute to the over-all data misfit. If the sum of errors is systematicallyunderestimated by a factor of

ffiffiffi2

p, c2 doubles in

size. We therefore examined models for someadditional c2 values. The model corresponding toc2 = 1.00 (Figure 8a) exhibits significant varia-tions in seismic velocity over length scales of�30 km. Given the 3-D geometry of the ray paths(Figure 5), the inversion is susceptible to imagingseismic structure across the arc into the arc-parallelvertical plane. We therefore relaxed the data fit totarget c2 of 1.40 and 1.80. The correspondingvelocity models are shown in Figures 8b and 8c,respectively. The three models of Figure 8 show aclear trend toward less complicated seismic struc-ture for higher c2 misfit, consistent with the Lcurve (Figure 7). We chose the more conservativemodel with c2 = 1.40 (Figure 8b) over c2 = 1.00.The misfit of the preferred model was verified byray tracing and equation (6). The actual c2 cameout a little higher at 1.46, presumably because ofray bending effects.

[21] We list the average normalized and RMS datamisfit for each travel time phase for the model ofFigure 8b in Table 2. All travel time phases fit at orclose to the estimated accuracy of travel time picks,except forPn. These Pn picks may not all correspondto minimum arrival times in the mantle, and the

Figure 5. Ray paths traced between shots and receivers in the starting model are shown in perspective view. Adifferent color is used for each instrument (inverted triangles) to illustrate the fan geometry of the source-receiverpaths. The map on the bottom of the model volume shows the location of the instruments on a map that outlines theislands of Umnak and Unalaska.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

8 of 24

Page 9: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

mantle portion of our model is therefore not reliable.The distribution of scaled data residuals beforeinversion is asymmetric, but after inversion itapproaches a normal distribution (Figure 9). Thischange in shape gives us confidence that muchof the remaining misfit is noise, and that theleast squares inversion is not dominated by out-

liers in the data. Shillington et al. [2004] dis-cussed the data fit of individual receiver gathers,but Figure 10 shows all travel time fits and raypaths projected on a 2-D profile along the strikeof the arc. Note that the calculated paths for eachray group reflects and refracts from the appro-priate model boundary as stated in Figure 2. Thisis an important condition to the validity of thelinearization (equation (1)), but it is not strictlyimposed in the iterative tomographic inversionscheme.

[22] The purpose of the difference operator D inequation (5) is to select a model of minimumstructure with acceptable c2 data misfit. The sizeand shape of D depends on the discretization of themodel m. However, the outcome of the regularizedinversion should be independent of the gridspacing, provided that the grid is fine enough tocapture the seismic velocity structure that isrequired to fit the data. We performed tomographicinversions with different grid spacings Dx and Dz,but with the same mix of first and second spatialderivatives in D, and for each inversion we chosec2 = 1.40. As expected, decreasing Dx and Dz by afactor of two does not change the seismic velocitystructure significantly (Figure 11a). Increasing Dxand Dz by a factor of two from Figure 8b also givesa similar result (Figure 11b), but care must be takenthat small details in the velocity structure are notlost by coarsening the grid [e.g., Toomey and

Figure 6. All 8336 source-receiver paths are sorted byazimuth. Small angles indicate paths subparallel to thestrike of the volcanic arc.

Figure 7. Trade-off between model roughness and data misfit explored with Tikhonov parameter l. After thedesired c2 misfit is set, the L curve is searched for the appropriate l using the secant method. The long- and short-dashed line corresponds to misfit levels c2 = 1.4, while the other two dashed lines are at c2 = 1.0 and c2 = 1.8.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

9 of 24

Page 10: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Foulger, 1989]. The test shows that the griddiscretization used in Figure 8 appears to beadequate for the inversion of the line A-2 data set.

[23] The derivatives of operatorD apply to both theseismic slownesses and the interface boundaries inthe model. Physical scaling of these parametergroups helps to balance the inversion [e.g., Scaleset al., 1990]. We accounted for the difference inphysical dimension by scaling these parameterswith a characteristic slowness (0.30 s/km) andinterface depth (20 km). More rigorous schemesto balance different types of parameters havebeen developed to address this issue in mixedinverse problems [Pavlis and Booker, 1980;Kennett et al., 1988; Zhang and Toksoz, 1998].Even when the inversion matrix is scaled, therelative strength of regularization of slownessesand boundary depths is arbitrary to some degree.We carried out tomographic inversions with thesame grid spacings as in Figure 8b and we selectedagain a c2 of 1.40. In the first test we placed aheavy penalty on spatial variations in the slownessand weak smoothness constraints on the layerboundaries. By carrying out this inversion weobtained a model with subdued velocity gradientsbut pronounced topography on the layer bound-aries (Figure 12a). Conversely, we were also ableto create a model with virtually flat layerboundaries and strong seismic velocity anomalies

(Figure 12b). The two models of Figure 12 span asuite of seismic structures that all minimize anobjective function F with different operators D.The model of Figure 8b exhibits lateral variationsin velocity and boundary depth because thatinversion balanced smoothing constraints on theslownesses and boundary depths [e.g., Scales et al.,1990]. Whether any of these models is preferableover others may depend on additional geologicalinformation on the crustal structure. We select themodel of Figure 8b since it does not contain stronglateral variations in either seismic velocity orboundary topography that may be misinterpreted.

[24] Another subjective choice in the design ofdifference operator D is the use of first- orsecond-order derivative constraints. We applied

Table 2. Misfit After Inversion

PhaseRMS

Misfit, msNormalized

MisfitAssignedError, ms

P1 101 1.04 100P2P 131 1.68 100P2 145 2.11 100P3P 119 1.27 110P3 157 1.42 150PmP 130 0.97 130Pn 351 6.57 120All 133 1.46 110

Figure 8. Inversion results for three different levels of misfit. Instrument locations are shown as red triangles.(a) The misfit c2 = 1.0 corresponds to travel time residuals approximately equal to the assigned data errors. (b and c)More conservative models have higher c2 values (1.4 and 1.8) and exhibit less seismic structure. We prefer the modelin Figure 8b with c2 = 1.4 over the model in Figure 8a because the influence of modeling errors on the inversion isuncertain.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

10 of 24

Page 11: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

both types of derivative constraints with equalweight in the model of Figure 8b. We alsoproduced models with c2 = 1.40 in an inversionwhere only first or second derivatives were used inthe regularization. Minimizing only the firstderivatives (Figure 13a) will lead to the flattestmodel, with relatively small peak to peak varia-tions in both the seismic velocities and boundarytopography. A drawback of minimizing firstderivatives is that the method does not impose

constraints on the curvature of reflecting interfacesor velocity contours. Sharp corners in the modelmake the linearization (equation (1)) inaccurate.The other end-member is a model with minimalsecond derivatives (Figure 13b), which behavesbetter in the inversion [Delprat-Jannaud andLailly, 1993]. However, the amplitudes of theseismic velocity anomalies in this minimumcurvature model are larger than in the minimumgradient model of Figure 13a. The model of

Figure 9. Travel time residuals relative to the current model show the goodness of fit. The distribution of travel timeresiduals changes during the tomographic inversion. All residuals are scaled by their assigned errors.

Figure 10. The data fit and ray paths for all data are shown in different colors that represent the travel time phase.(a) The travel times are reduced by 7.0 km/s. Travel time picks are shown as solid lines, while model predictions areshown by dashed lines. (b) Corresponding ray paths traced in the final model and projected on a 2-D plane show thediversity of observed phases.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

11 of 24

Page 12: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Figure 8b finds a compromise by adding both firstand second derivative constraints to the objectivefunction (5).

[25] The interpretation of travel time branches[Shillington et al., 2004] is also subjective. Theray paths shown in Figure 10 were forced to refractand reflect from horizons that were tied to thephase interpretation (Figure 2). To investigate thelinkage between the model parameterization andtravel time interpretation, we perform an inversionwithout wide-angle reflections (Figure 14). Only60% of our travel time data are first-arrival refrac-tions (Table 1), and most of these have shortsource-receiver offsets. Consequently, the first-arrival inversion cannot constrain the deep crustalstructure as well as the inversions that includedreflections, but first-arrival time inversions aremore robust [Zelt et al., 2003]. The shallow struc-ture in Figure 14 is very similar to the model that

includes wide-angle reflections (Figure 8b), andsome of the high seismic velocities in the lowercrust between 400 km and 650 km were alsoreproduced. We therefore conclude that the phaseinterpretation of Shillington et al. (submitted manu-script, 2004) is consistent with the first-arrival timeconstraints.

4. Model Covariance and Resolution

[26] We demonstrated that different regularizationmatrices D affect the outcome of the inversion.These models (Figures 8b, 12, and 13) all fit theseismic refraction data with c2 = 1.4, so thedifferences in seismic structure indicate an un-certainty in the solution that stems from thearbitrariness of the regularization. This ‘‘geomet-rical ambiguity’’ [Bohm et al., 2000] is impossibleto quantify, because the number of solutions to our

Figure 11. Inversion results for different inversion parameterizations. We nearly reproduce the result of our earliermode by grid spacings that are (a) half the size and (b) twice the size of the grid used for the model of Figure 8b.

Figure 12. Two end-member models illustrate the ambiguity in minimum structure models of mixedparameterization. Besides the model of Figure 8b, models with c2 = 1.4 can be constructed with (a) relativelysmooth velocities and rough boundaries or (b) rough velocities and smooth boundaries.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

12 of 24

Page 13: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

inverse problem is infinite unless we define our apriori expectations of Earth structure by means ofoperator D in equation (5). By choosing a rough-ness penalty for D we ignored the fact that geologyof the Aleutian arc can be very diverse even on thescale of a single magma center [e.g., Miller et al.,1992; Myers et al., 2002]. We assume, however,that our velocity model of Figure 8b represents agood large-scale averaged seismic velocity modelfor the Aleutian arc. We will estimate the modelcovariance Cm and model resolution R usingequations (9) and (12), but these quantitieswill underestimate the uncertainty in small-scalefeatures.

[27] To get a first impression of the modelquality we show the ray paths of all wide-angledata projected on the 2-D profile along the arc(Figure 15a). The wide instrument spacing in thewestern half of the model leaves large gaps inthe coverage of the crustal layers, but to the eastthe crust is more evenly sampled. The density ofray paths can be obtained by summing thecolumns in the Frechet matrix G, scaled byassigned data errors (Figure 15b). The so-called

derivative weight sum (DWS) gives an indicationof model fidelity [Scales, 1987; Toomey andFoulger, 1989]. The model parameters between400 km and 750 km in the model are visited bymany ray paths, but outside this area the datacoverage is more limited.

[28] The effect of sparse data coverage and thepropagation of travel time errors in the model canbe evaluated by a covariance and resolution anal-ysis. We use equation (17) to translate Cm toconfidence ellipses for the average seismic velocityin a 10 km wide window of the middle and lowercrust. Between model distances 420 km and 680 kmthe seismic velocity varies laterally by �0.6 km/sin the middle crust, and by �0.4 km/s in the lowercrust. The 1s error ellipses show that the trends inseismic velocity between the islands Umnak,Unalaska and Akutan are significantly largerthan the modeling errors presented through Cm

(Figure 16). The uncertainties in lower crustalvelocities are slightly higher than in the middlecrust, which may result from fewer data and largerdata errors for the travel time phases that samplethe lower crust.

Figure 13. Our preferred model of Figure 8b is obtained with both first- and second-derivative regularization.Seismic velocity models with only (a) first- or (b) second-derivative D only differ in detail. The large-scale featuresappear to be very similar.

Figure 14. Inversion of first-arrival times for a smooth seismic velocity model without discontinuities reproducesthe high seismic velocities around 600 km in the lower crust.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

13 of 24

Page 14: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[29] The ambiguity in resolving seismic reflectordepth and seismic velocity of the overlying layeris a classic problem in reflection seismology [e.g.,Ross, 1994]. In wide-angle data, the presence ofboth seismic reflections and refractions that areobserved over a wide range of source-receiveroffsets helps to constrain both seismic velocitiesand interface depths, but the lower crustal layerand Moho depth are both largely dependent on

PmP arrivals (Table 1). The trade-off between duand drk is represented by matrix element Cur inequation (19), and it causes skewness of the 1serror ellipses for lower crustal seismic velocityand Moho depth averaged over a 100 km widewindow (Figure 17). Beside the positive velocity-depth trade-off and scatter in lower crustal veloc-ities we observe a slight eastward thickening ofthe magmatic arc.

Figure 15. (a) Ray paths, traced in a 3-D volume around the Aleutian arc, projected on a vertical plane along thearc. (b) Ray density is highest in the central and eastern portion of the model, as shown by the derivative weight sum(DWS).

Figure 16. Errors in local averages of seismic velocity of both the middle and lower crust shown simultaneously bya series of error ellipses. Arrows show the trend in these average velocities along the strike of the arc. Seismicvelocities are averaged over a 10 km wide window in the middle and lower crust, swinging from relatively low at420 km in our model profile to higher values beneath the islands of Umnak and Unalaska and back to lower seismicvelocities beneath Akutan.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

14 of 24

Page 15: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[30] Our model of the Aleutian arc (Figure 8b)exhibits lateral variations in seismic velocity of0.2–0.5 km/s over length scales of �50 km, and anear absence of vertical velocity gradients withineach of the three crustal layers. Since the gridspacing varies between 1.5 and 7.0 km (Figure 3),these seismic velocity variations extend over atleast a few model parameters. We define a hypo-thetical model feature by a Gaussian function,whose width is greater than its vertical extent(Figure 18). We examine how much of the bell-shaped averaging window is mapped onto itself byresolution matrix R.

[31] The window can be centered on any location(x, z) to measure the local resolution value forthe length scale of interest. In Figure 19 we plotresolution values on our velocity model (Figure 8b)for a suite of length scales to show the reliability oflarge model features. Resolution values of 0.5,meaning that half of the bell shape is recovered,are found throughout the model if the anomaly sizeis 36 km � 16 km. The resolution is thereforesufficient for almost all noticeable structure inFigure 8b.

5. Discussion

5.1. Model and Nonuniqueness

[32] We obtained a simple three–layer seismicvelocity model for line A2 with only moderatelateral variation in layer thickness and seismicvelocity (Figure 8b). Our preferred seismic velocitymodel shows a crust that thickens slightly from�35 km in the west near Amlia island to �37 kmnear Unimak in the east (Figure 1). The modelconsists of three crustal layers (Figure 8b) withseismic velocities that increase progressively from6.0–6.5 km/s in the upper crust to 6.5–7.3 km/s inthe middle crust, and 7.3–7.7 km/s in the lowercrust. Both the seismic velocities and layer bound-aries vary gradually over distances of 50–100 km,length scales that are well resolved according toour resolution tests (Figure 19). Unfortunately, themixed parameterization of the model makes itimpossible to define an unambiguous minimumstructure solution to the inverse problem. This isillustrated by the two test inversions of Figure 12.However, we feel that the model of Figure 8b fitsour expectations of the crustal structure of islandarc crust.

[33] The test inversions in the paper give us confi-dence that the model of Figure 8b does not dependon arbitrary modeling decisions such as the inver-

sion grid spacing or details in the design of roughnessoperatorD. Nevertheless, our result for line A-2appears very different from the model of Fliednerand Klemperer [1999], who used the method ofHole [1992] and Hole and Zelt [1995]. Comparedto our result their model has a thinner crust withlower average seismic velocity, and although itdoes not have velocity discontinuities inside thecrust, their model exhibits more heterogeneity onwavelengths less than 30 km (Figure 20). Some ofthe differences in model roughness can beattributed to the level of data misfit. The modelof Fliedner and Klemperer [1999] fits the first-arrival picks with 73 ms RMS error, and PmParrivals are matched with 104 ms RMS error. Theseerrors are slightly smaller than those for our modelof Figure 8b (Table 2), but comparable to ourrougher model of Figure 8a.

[34] We recognize five deviations in the modelingapproaches taken by us and Fliedner and Klemperer[1999] that can explain the large differences be-tween these models: (1) Shillington et al. (submit-ted manuscript, 2004) used multichannel seismicobservations to constrain the shallow sedimentvelocities and basement depth on the arc platform.This information was used in our starting model(Figure 4). The seismic refraction data of line A-2alone do not provide sufficient ray coverage toimage the uppermost basement and overlying sedi-ments. (2) The interpretation of deep wide-anglereflected phases varied between the two studies,which may explain the difference in crustal thick-ness. (3) Fliedner and Klemperer [1999] con-structed a continuous velocity model from theseafloor to the Moho. We assumed velocity dis-continuities at the reflecting layer boundaries. Thepattern of wide-angle phases (Shillington et al.,submitted manuscript, 2004) is consistent with ourlayered model (Figure 2), but such prominentvelocity discontinuities may not be required to fittravel time data. (4) We accommodated variationsin seismic velocity across the island arc by largestatic time corrections for the land seismic stations.Fliedner and Klemperer [1999] performed a full3-D inversion of the data instead. (5) We performeda regularized inversion with first- and second-derivative constraints to treat the underdeterminedparts of the model. Fliedner and Klemperer [1999]used the method of Hole [1992] which usesback projection to invert the data for a model solu-tion. Backprojection is computationally fast, butcan lead to models that are rough compared toresults from regularized inversion [Zelt and Barton,1998].

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

15 of 24

Page 16: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[35] Compilations of seismic refraction data[Spudich and Orcutt, 1980; Christensen andMooney, 1995] show that seismic velocities in-crease with depth in the Earth’s crust. From themodels presented in this paper and by Fliedner andKlemperer [1999] it is clear that this also holds truefor the Aleutian arc. However, in our model thevelocity increases instantaneously at the layerboundaries, while the model of Fliedner andKlemperer [1999] has strong gradients. Could weaugment our objective function (5) to control therelative contribution of vertical velocity gradientsand velocity discontinuities? The vertical seismicvelocity gradients within the three crustal layers inthe model of Figure 8b are small, and most of theincrease in seismic velocity from the surface to theMoho occurs at the layer boundaries. Changingthe aspect ratio of the vertical and lateral smooth-ing lengths would therefore not have a large effecton the outcome of the inversion. Alternatively, wecan impose damping on the magnitude of the stepin slowness across the layer boundaries. Weachieve that by minimizing an objective functionof the following form:

F dmð Þ ¼

C�1

2

d Gdm� dð Þ

lD mþ dmð Þ

gIdu mþ dmð Þ

����������

����������

����������

����������

2

; ð20Þ

where Idu = diag{0,. . ., 0, 1,. . ., 1} is nonzero onlywhere the diagonal elements multiply into themodel parameters that represent slowness disconti-nuities duk. In contrast to equation (5), damping inthe objective function in equation (20) applies tothe final model m + dm instead of the modelperturbation dm. Increasing g in equation (20)decreases the magnitude of duk. Keeping g = 0gives us the model of Figure 8, but by increasing g

we would obtain a crustal velocity structure that ismore similar to the result of Fliedner andKlemperer [1999]. Unfortunately, the interpretationof P2 and P3 as two distinct deep crustal refractions[Shillington et al., 2004] prevents us from applyingequation (20) to the travel time data. These twophases would not be observed if the two velocitydiscontinuities in the crust were absent or veryweak. If we seek a model with a gradual increase inseismic velocity with depth we cannot interpret theP2 and P3 picks as shown in Figure 2. Atomographic inversion of only first-arrival picks(Figure 14) does not impose a phase interpretation,but it also lacks the depth resolution that thereflection tomography inversion achieves.

[36] The relatively small number of instruments inthis study limits our resolution to at best a fewtens of km (Figure 19). Our model represents thelocally averaged seismic velocity structure of theisland arc, which is relevant for comparison withlarge-scale trends in geochemistry [Shillington etal., 2004]. Since travel times are essentially pathintegrals over the velocity model, the formalerrors in spatial averages of the seismic velocitystructure can be small (Figures 16 and 17. How-ever, the assumptions we made in the calculationof Cm and R do not apply to individual elements ofmodel vector m. The problem lies in the role ofthe regularization D in the calculation of Cm

(equation (9)). The regularization merely servesto stabilize our inversion and does not representour a priori knowledge of the Earth [Nolet et al.,1999]. Consequently, the a posteriori covarianceCm gives us low formal errors for individualmodel parameters, which does not account forthe arbitrariness in our choice of model parameter-ization or the effects of inadequate sampling ofour model [Trampert and Snieder, 1996]. Thenonlinearity of equation (1) can also cause smallmodeling errors [Zhang et al., 1998; Korenaga etal., 2000]. These errors complicate the interpreta-tion of detailed model features, but we assume thatthey do not affect the large-scale properties, such asaverage crustal velocity and thickness. Perhaps

Figure 17. Trade-off between seismic velocity in thelower crust and depth of the Moho is shown by eight 1serror ellipses calculated in 100 km wide windowsspaced equally over the length of our model. Theseismic velocity is averaged over the entire lower crustallayer in this calculation. The errors are relatively smallin the eastern half of the model, where the data coverageis better.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

16 of 24

Page 17: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

even the largest uncertainty in these average crustalproperties may be due to variations in crustalstructure perpendicular to the arc that are notrepresented in our velocity model and erroranalysis.

5.2. Crooked Line Geometry

[37] We chose to perform 3-D ray tracing and 2-Dinversion to treat the crooked line geometry. Ouruse of a 2-D inversion, with time corrections of�0.4 s applied to travel times observed on islandstations, raises two questions: (1) Can lateral var-iations in the shallow structure of the Aleutian arcgive rise to such large time delays? (2) Is the deepcrustal structure of the arc sufficiently 2-D tovalidate this approach?

[38] The static correction for island stations waseffective, since it reduced c2 from 16.34 to 4.89(70% reduction), before the tomographic inversionreduced it further to 1.46 (91% overall variancereduction). The large delays cannot be explainedby the absence of low-velocity sediments on theislands, because the sediment cover is not thickenough on the submerged arc platform to accountfor 0.4 s. The delays indicate a high-velocityanomaly in the upper crust beneath the volcanicislands, where ray paths emerge at subverticalangles (Figure 5). Such high velocities near thesurface of Aleutian islands were previouslyreported by Grow [1973] and Abers [1994]. Thesehigh-velocity anomalies are consistent with thepresence of large plutons and relatively fresh,low-porosity lava flows around these volcanic cen-ters [Kay and Kay, 1985]. The upper crust else-

where on the platform may be more fractured and oflower seismic velocity. The shallow depth of �5km inferred by InSAR for a major magma reservoirbeneath Okmok suggests that these volcanic centersextend into the upper crust [Mann et al., 2002],thereby disrupting the layering of the island arc.Layering of the volcanic arc is evidenced by abun-dant midcrustal wide-angle reflections (Figure 21).These reflection bounce points occur in the subsidedplatform of the Aleutian arc, which we thereforeconsider representative for our 2-D seismic velocitymodel (Figure 8b).

[39] To validate a 2.5-D inversion we assumed thatour ray paths do not sample a significant amount ofcrustal seismic velocity heterogeneity perpendicu-lar to the arc [Zelt and Zelt, 1998]. The 3-D raypaths cover a �50 km wide swath in the uppercrust, and a narrow arching band deeper in themodel (Figure 21). Velocity models of the crosssection of the Aleutian arc [Holbrook et al., 1999;Lizarralde et al., 2002] show that the structure ofthe magmatic arc is fairly constant over a widthof �100 km, which supports our assumption. The2-D seismic velocity model must be interpreted asan averaged along-arc profile that can be used tocompare with the bulk composition of the Aleutianarc crust (Shillington et al., submitted manuscript,2004).

5.3. Interpretation

[40] Shillington et al. [2004] interpret the layeringin our model as a result of distinct magmaticprocesses at different depth levels. The upper crustmay largely consist of submarine lava flows, vol-

Figure 18. Resolution of model features of different sizes is estimated by bell-shaped averaging windows ofvariable size. We use the resolution matrix to map this synthetic structure and compare with the original to verify howmuch is recovered. The model parameters are shown circles and triangles. The window slides over the model toexamine spatial variations in resolution.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

17 of 24

Page 18: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

caniclastic sediments, and small plutons [Scholl etal., 1983; Kay and Kay, 1985; Kelemen, 1995].Variations in seismic velocity in this layer maydepend more on rock porosity [Carlson and Gangi,

1985; Berge et al., 1992] than composition. Muchof the middle crust may be formed by the relictoceanic crust of the upper plate. The Aleutian arccrust was built on a fragment of the Kula plate

Figure 19. The resolution of both seismic velocity and layer boundary depth increase if the averaging window isincreased. (a) The resolution of 9 km � 4 km objects is negligible (<0.1) except for patches of layer boundaries wherewide-angle reflections bounce. (b) At 18 km � 8 km we find resolution values of 0.5 in the upper crust, but theseobjects are not well recovered in the middle and lower crust. (c) At 36 km � 16 km, the length scale of most of theheterogeneity in the middle and lower crust, we find that resolution values are larger than 0.5 and thereforesufficiently resolved. (d) At 72 km � 32 km most of the 2-D model is well resolved.

Figure 20. Seismic velocity models of (a) our study and (b) Fliedner and Klemperer [1999] shown on the samescale. Since the analysis of Fliedner and Klemperer [1999] was a 3-D inversion, the profile in Figure 20b is a verticalalong-arc slice through their model.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

18 of 24

Page 19: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

[Marlow and Cooper, 1983; Debiche et al., 1987].Magmatic intrusives added to the middle crust maybe andesitic where midcrustal velocities are 6.5–7.0 km/s, and more gabbroic where velocities arehigher [Shillington et al., 2004]. The lower crustalvelocities are consistent with both mafic to ultra-mafic cumulates and garnet granulites [Kay andKay, 1985; DeBari and Sleep, 1991; Miller andChristensen, 1994; Muntener et al., 2004]. Lateralvariability in arc composition can give us insight inthe scale of magmatic processes and the relativeinfluence of changes in age and character of thesubducting crust (Shillington et al., submittedmanuscript, 2004). The lateral velocity variationsimaged in our model with horizontal length scalesof �50 km appear significant within our resolution(Figure 16), but they appear subdued compared tolateral variability found by Fliedner and Klemperer[1999] for line A2 (Figure 20). The variations inseismic velocity found by Fliedner and Klemperer[1999] are not implausible given the geochemicalheterogeneity of the Aleutian arc lavas [e.g., Milleret al., 1992; Singer et al., 1992].

[41] The Moho is well defined in our model, owingto the abundance of PmP observations (Table 1).The crustal thickness of 35–37 ± 1 km is large com-pared to arcs in the southwest Pacific [Dimalanta etal., 2002]. This can be explained by the lack of backarc spreading in the Aleutians, because the Aleutianarc is relatively narrow. According to our model,thickening of the lower crust in the east can accountfor the crustal thickness variation. The subtle thick-ness variations are difficult to verify with gravitybecause the subducting slab and its dynamic effects

dominate the free air gravity signal [Abers, 1994].Moreover, Jull and Kelemen [2001] have suggestedthat the lower crust of an island arc may be as denseas the underlying mantle. In other geologic settingsgravity anomalies have been associated with mag-matic, underplated crust with seismic velocities inbetween 7 and 8 km/s [e.g., ten Brink and Brocher,1987].

[42] The large difference in average seismic veloc-ity between this island arc and the average seismicvelocity of continental crust [Christensen andMooney, 1995; Holbrook et al., 1999] is confirmedby our study. The underplated lower crust in ourmodel, which composes 40% of the crust, mayrepresent ultramafic cumulates or garnet granulitethat do not survive when the arc is accreted to acontinent [Shillington et al., 2004]. Seismic veloc-ities in the middle and lower crust are highestbeneath the islands of Umnak and Unalaska(Figure 21), which may indicate a larger amountof melt fractionation [Kay and Kay, 1985]. Becauseof our improving understanding of the complexrelationship between the seismic velocity structureand the composition of igneous crust [Kelemen andHolbrook, 1995; Kern et al., 2001; Korenaga et al.,2002; Behn and Kelemen, 2003], seismic refractiondata will continue to provide key results in studiesof island arcs.

6. Conclusions

[43] We did an extensive investigation of the seis-mic velocity structure of the central Aleutian arcusing a relatively sparse wide-angle seismic refrac-

Figure 21. Land instruments, indicated by inverted triangles, for line A2 are located around the major volcaniccenters of the Aleutians, while the shot line (thick solid line) runs well south of the islands of Umnak, Unalaska, andAkutan. The observed bounce points of 3-D ray paths between the shot and instruments therefore lie on the southernedge of the arc platform. Reflections from the top of the middle crust are shown as small circles, reflections from thebottom of the lower crust as diamonds.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

19 of 24

Page 20: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

tion data set. The data offer new insight in thedeep structure of this island arc and the mag-matic processes that shaped it. A description ofthe data and the implications of this study forisland arc composition and magmatic processesare described in more detail by Shillington et al.[2004].

[44] The first goal of this paper was to describea tomographic inversion that handles seismicrefraction data recorded in a crooked line geom-etry. Although 3-D ray tracing between sourcesand receivers was necessary to account for theout-of-plane effects, the sparseness of the data setsuited a 2-D inversion with static time correc-tions for the island stations. The resulting seismicmodel consists of a smoothly varying velocityfield divided by a few velocity discontinuities,which are consistent with the reflected andrefracted arrivals. Second, we carried out a fewtests to examine the dependence of the imagedseismic structure on the model parameterization.These tests show that the model discretizationand regularization do not lead to arbitrary seis-mic structure that may be wrongly interpreted.However, a certain trade-off remains in theamount of complexity in seismic velocity andlayer boundary structure. Once we accept theperformance of the model regularization, esti-mates of the model covariance and resolutioncan show the model goodness.

[45] Our seismic velocity model consists of a 35–37 km thick arc crust with three layers that increasesteadily in seismic velocity with depth from.6.5 km/s near the surface to 7.7 km/s at the baseof the crust. Anomalous shallow structure in thevicinity of large volcanic centers may have givenrise to �0.4 s faster seismic arrivals than at the arcplatform. Although horizontal length scales of�20 km are resolved in some areas, in most ofthe model horizontal features less than �50 km arenot well resolved. The seismic velocities of the arcare significantly higher than the velocity of averagecontinental crust. The increase in seismic velocitywith depth in our model occurs mostly at theboundaries of these three layers, although thatcharacteristic may be exaggerated by the modelparameterization. The three seismic layers mayrepresent an upper layer with volcaniclastic sedi-ments, extrusives and small plutons. The middlecrust may be the original oceanic upper plate of theAleutian subduction zone, but is thickened andmodified by andesitic and gabbroic intrusions.The high seismic velocities in the lower crust are

consistent with amphibole- and garnet-bearingrocks.

Appendix A

[46] The calculation of the a posteriori covariancematrix Cm requires us to invert the symmetricmodel connectivity matrix ATA, where A is theinversion matrix. Although A is sparse, ATA can bea dense N � N matrix. Fortunately, we cancalculate the inverse without explicitly formingATA. We exploit the sparseness of A by usingLanczos’ [1950] algorithm. This iterative techniqueproduces a singular value decomposition (SVD) intwo steps. First, ATA is transformed into atridiagonal system

ATA QTQT ; ðA1Þ

where Q is an N � N orthogonal matrix and

T ¼

a1 b2 0

b2 a2. ..

. .. . .

.bn

0 bn an

0BBBBBBBBB@

1CCCCCCCCCA: ðA2Þ

The coefficients ai and bi and vectors qi that formthe orthogonal columns of Q satisfy a simplerecursive relationship that only requires two vectormultiplications of A for each triplet, and memoryrequirements are modest. A detailed description ofthis algorithm and the numerical caveats are givenby Simon [1984], Cullum and Willoughby [1985],and Vasco et al. [1999]. The matrix T can bediagonalized by computing the eigenvalues andeigenvectors through inverse iteration [Golub andVan Loan, 1996; Parlett, 1998]. Starting with arandom vector h1 of length N and a guess foreigenvalue s1, we obtain our first approximationfor eigenvector w1 by solving the tridiagonalsystem

T� s1Ið Þw1 ¼ h1 ðA3Þ

by back substitution [e.g., Press et al., 1986]. Wereplace h1 with w1 and replace s1 with kTw1k andrepeat equation (A3) until w1 and h1 converge. Ass1 approaches an eigenvalue of T, equation (A3)becomes singular, at which time the iteration isstopped, usually after 5 to 7 cycles. Othereigenpairs (si, wi) can be found by the sameprocedure, although the starting vector hi must be

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

20 of 24

Page 21: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

chosen perpendicular to the eigenvectors w thathave already been found by inverse iteration. Inthis manner we develop an orthogonal basis ofeigenvectors, and erroneous duplication of eigen-pairs can be avoided even when the eigenvaluesbecome degenerate in the tail of the spectrum(Figure A1). The eigenvalue decomposition

T ¼ W2WT; ðA4Þ

where2 = diag{s1, s2,. . ., sn} and wi compose thecolumns of W, can be used to invert T and ATA.The a posteriori model covariance matrix is givenby

Cm ¼ QTT�1Q ¼ QTWT2�1WQ: ðA5Þ

The tridiagonal matrix T is usually not expanded toa N � N matrix, but truncated at a smaller size P �P if enough eigenvalues of the submatrix Tp, calledRitz values, have converged to eigenvalues of T[Berry, 1992; Minkoff, 1996; Vasco et al., 1999]. Ifthe number of Lanczos iterations P is too small, the

vectors wi will not span enough of the modelspace, which can lead to significant errors in thecalculation of model assessment [Deal and Nolet,1996]. In our 2-D application the size of the modelspace is small enough to carry out the requisitenumber of Lanczos iterations. However, if the fullspectrum of 2 is used, small si can causeinstabilities in equation (A5). Although small sican be removed explicitly from equation (A4), wefix the problem by introducing a small amount ofdamping to assure that all si are finite (Figure A1).

Acknowledgments

[47] We thank Moritz Fliedner and Simon Klemperer for

making their seismic refraction data available through the

IRIS Data Management Center. Some of the work on this

project was carried out while HJV was a visitor at Utrecht

University Institute for Earth Sciences. We thank K. Dueker,

P. Kelemen, O. Muntener, and W. Spakman for useful dis-

cussions on this work. The constructive reviews of Associate

Editor J. Gaherty and reviewers T. Brocher and R. Carbonell

greatly improved the manuscript. This work was supported by

the University of Texas Institute for Geophysics. This is UTIG

contribution 1698.

References

Abers, G. A. (1994), Three-dimensional inversion of regionalP and S arrival times in the East Aleutians and sources ofsubduction zone gravity highs, J. Geophys. Res., 99, 4395–4412.

Backus, G. E., and F. Gilbert (1970), Uniqueness in the inver-sion of inaccurate gross Earth data, Philos. Trans. R. Soc.London, 266, 123–192.

Behn, M. D., and P. B. Kelemen (2003), Relationship betweenseismic P-wave velocity and the composition of anhydrousigneous and meta-igneous rocks, Geochem. Geophys. Geo-syst., 4(5), 1041, doi:10.1029/2002GC000393.

Berge, P. A., G. J. Fryer, and R. H. Wilkens (1992), Velocity-porosity relationships in the upper oceanic crust: Theoreticalconsiderations, J. Geophys. Res., 97, 15,239–15,254.

Berry, M. (1992), Large scale singular value computations, Int.J. Supercomput. Appl., 6, 13–49.

Bishop, T. N., K. P. Bube, R. T. Cutler, R. T. Langan, P. L.Love, J. R. Resnick, R. T. Shuey, D. A. Spindler, andH. W. Wyld (1985), Tomographic determination of velocityand depth in laterally varying media, Geophysics, 50, 903–923.

Bohm, G., P. Galuppo, and A. Vesnaver (2000), 3D adaptivetomography using Delaunay triangles and Voronoi polygons,Geophys. Prospect., 48, 723–744.

Carlson, R. L., and A. F. Gangi (1985), Effects of cracks on thepressure dependence of P wave velocities in crystallinerocks, J. Geophys. Res., 90, 8675–8684.

Chou, C. W., and J. R. Booker (1979), A Backus-Gilbert ap-proach to inversion of travel-time data for three-dimensionalvelocity structure, Geophys. J. R. Astron. Soc., 59, 325–344.

Christensen, N. I., and W. D. Mooney (1995), Seismic velocitystructure and composition of the continental crust: A globalview, J. Geophys. Res., 100, 9761–9788.

Figure A1. (a) For the calculation of the inverse ofATA we perform an SVD and invert the singular valuesstored in S. This process leads to numerical instabilitiesif some of the singular values s are nearly zero. Weprevent this problem by introducing a small amount ofdamping. (b) The damping does not affect the largesingular values.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

21 of 24

Page 22: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Constable, S. C., R. L. Parker, and C. G. Constable (1987),Occam’s inversion: A practical algorithm for generatingsmooth models from electromagnetic sounding data, Geo-physics, 52, 289–300.

Cullum, J. K., and R. A. Willoughby (1985), Lanczos Algo-rithms for Large Symmetric Eigenvalue Computations, vol. 1,Theory, 268 pp., Birkhauser Boston, Cambridge, Mass.

Deal, M. M., and G. Nolet (1996), Comment on ‘‘Estimationof resolution and covariance for large matrix inversions’’ byJ. Zhang and G. A. McMechan, Geophys. J. Int., 127, 245–250.

DeBari, S. M., and N. H. Sleep (1991), High-Mg, low-Al bulkcomposition of the Talkeetna island arc, Alaska: Implicationsfor primary magmas and the nature of arc crust, Geol. Soc.Am. Bull., 103, 37–47.

Debiche, M. G., A. Cox, and D. C. Engebretson (1987), TheMotion of Allochthonous Terranes Across the North PacificBasin, Geol. Soc. Am. Spec. Pap., 207, 49 pp.

Delprat-Jannaud, F., and P. Lailly (1993), Ill-posed and well-posed formulations of the reflection traveltime tomographyproblem, J. Geophys. Res., 98, 6589–6605.

Delprat-Jannaud, F., and P. Lailly (1995), Reflection tomogra-phy: How to handle multiple arrivals?, J. Geophys. Res.,100, 703–715.

Dimalanta, C., A. Taira, G. P. Yumul, H. Tokuyama, andK. Mochizuki (2002), New rates of western Pacific islandarc magmatism from seismic and gravity data, Earth Planet.Sci. Lett., 202, 105–115.

Ditmar, P., J. Penopp, R. Kasig, and J. Makris (1999), Inter-pretation of shallow refraction seismic data by reflection/refraction tomography, Geophys. Prospect., 47, 871–901.

Fliedner, M. M., and S. L. Klemperer (1999), Structure of anisland-arc: Wide-angle seismic studies in the eastern Aleu-tian islands, Alaska, J. Geophys. Res., 104, 10,667–10,694.

Goff, J. A., and K. Holliger (2003), Heterogeneity in the Crustand Upper Mantle, 349 pp., Kluwer Acad., New York.

Golub, G. H., and C. F. Van Loan (1996), Matrix Computa-tions, 694 pp., Johns Hopkins Univ. Press, Baltimore, Md.

Grove, T. L., and R. J. Kinzler (1986), Petrogenesis of ande-sites, Annu. Rev. Earth Planet. Sci., 14, 417–454.

Grow, J. A. (1973), Crustal and upper mantle structure of thecentral Aleutian arc, Geol. Soc. Am. Bull., 84, 2169–2192.

Hansen, P. C. (1992), Analysis of discrete ill-posed problemsby means of the L-curve, SIAM Rev., 54, 561–580.

Hobro, J. W. D., S. C. Singh, and T. A. Minshull (2003),Three-dimensional tomographic inversion of combined re-flection and refraction seismic traveltime data, Geophys. J.Int., 152, 79–93.

Holbrook, W. S., W. D. Mooney, and N. I. Christensen(1992), The seismic velocity structure of the deep continentalcrust, in The Lower Continental Crust, edited by D. M.Fountain, R. Arculus, and R. W. Kay, pp. 1–43, ElsevierSci., New York.

Holbrook, W. S., D. Lizarralde, S. McGeary, N. Bangs, andJ. Diebold (1999), Structure and composition of the Aleutianisland arc and implications for continental crustal growth,Geology, 27, 31–34.

Hole, J. A. (1992), Nonlinear high-resolution three-dimensionalseismic travel time tomography, J. Geophys. Res., 97, 6553–6562.

Hole, J. A., and B. C. Zelt (1995), 3-D finite-difference reflec-tion traveltimes, Geophys. J. Int., 121, 427–434.

Hole, J. A., R. M. Clowes, and R. M. Ellis (1992), Interfaceinversion using broadside seismic refraction data and three-dimensional travel time calculations, J. Geophys. Res., 97,3417–3429.

Iyer, H. M., and K. Hirahara (1993), Seismic Tomography:Theory and Practice, 842 pp., Chapman and Hall, New York.

Jackson, D. D. (1979), The use of a priori data to resolve non-uniqueness in linear inversion, Geophys. J. R. Astron. Soc.,57, 137–157.

Jull, M., and P. B. Kelemen (2001), On the conditions forlower crustal convective instability, J. Geophys. Res., 106,6423–6446.

Kay, R. W. (1980), Volcanic arc magmas: Implications of amelting-mixing model for element recycling in the crust-upper mantle system, J. Geol., 88, 497–522.

Kay, S. M., and R. W. Kay (1985), Role of crystal cumulatesand the oceanic crust in the formation of the lower crust ofthe Aleutian arc, Geology, 13, 461–464.

Kelemen, P. B. (1995), Genesis of high Mg# andesites and thecontinental crust, Contrib. Mineral. Petrol, 120, 1–19.

Kelemen, P. B., and W. S. Holbrook (1995), Origin of thick,high-velocity crust along the U.S. East Coast Margin,J. Geophys. Res., 100, 10,077–10,094.

Kennett, B. L. N., M. S. Sambridge, and P. R. Williamson(1988), Subspace methods for large inverse problems withmultiple parameter classes, Geophys. J., 94, 237–247.

Kern, H., T. Popp, F. Gorbatschevich, A. Zharikov, K. V.Lobanov, and Y. P. Smirnov (2001), Pressure and tempera-ture dependence of Vp and Vs in rocks from the superdeepwell and from surface analogues at Kola and the nature ofvelocity anisotropy, Tectonophysics, 338, 113–134.

Koch, M. (1993), Simultaneous inversion for 3-D crustal struc-ture and hypocentres including direct, refracted and reflectedphases—Development, I., validation and optimal regulariza-tion of the method, Geophys. J. Int., 112, 385–412.

Korenaga, J., W. S. Holbrook, G. M. Kent, P. B. Kelemen, R. S.Detrick, H.-C. Larsen, J. R. Hopper, and T. Dahl-Jensen(2000), Crustal structure of the southeast Greenland marginfrom joint refraction and reflection seismic tomography,J. Geophys. Res., 105, 21,591–21,614.

Korenaga, J., P. B. Kelemen, and W. S. Holbrook (2002),Methods for resolving the origin of large igneous provincesfrom crustal seismology, J. Geophys. Res., 107(B9), 2178,doi:10.1029/2001JB001030.

Lanczos, C. (1950), An iterative method for the solution of theeigenvalue problem of linear differential and integral opera-tors, J. Res. Natl. Bur. Stand. U. S., 45, 255–282.

Lawson, C. L., and R. J. Hanson (1974), Solving Least SquaresProblems, 340 pp., Prentice-Hall, Old Tappan, N. J.

Levander, A. R., and B. S. Gibson (1991), Wide-angle seismicreflections from two-dimensional random target zones,J. Geophys. Res., 96, 10,251–10,260.

Lizarralde, D., W. S. Holbrook, S. McGeary, N. Bangs,and J. Diebold (2002), Crustal construction of a volcanicarc, wide-angle seismic results from the western AlaskaPeninsula, J. Geophys. Res., 107(B8), 2164, doi:10.1029/2001JB000230.

Mann, D., J. Freymueller, and Z. Lu (2002), Deforma-tion associated with the 1997 eruption of Okmok volcano,Alaska, J. Geophys. Res., 107(B4), 2072, doi:10.1029/2001JB000163.

Marlow, M. S., and A. K. Cooper (1983), Wandering terranesin southern Alaska: The Aleutia Microplate and implicationsfor the Bering Sea, J. Geophys. Res., 88, 3439–3446.

McCaughey, M., and S. C. Singh (1997), Simultaneous velocityand interface tomography of normal-incidence and wide-aper-ture seismic traveltime data, Geophys. J. Int., 131, 87–99.

McLennan, S. M., and S. R. Taylor (1982), Geochemical con-straints on the growth of the continental crust, J. Geol., 90,347–361.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

22 of 24

Page 23: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Menke, W. (1984), Geophysical Data Analysis: Discrete In-verse Theory, 289 pp., Academic, San Diego, Calif.

Miller, D. J., and N. I. Christensen (1994), Seismic signatureand geochemistry of an island arc: A multidisciplinary studyof the Kohistan accreted terrane, northern Pakistan, J. Geo-phys. Res., 99, 11,623–11,642.

Miller, D. M., C. H. Langmuir, S. L. Goldstein, and A. L.Franks (1992), The importance of parental magma composi-tion to calc-alkaline and tholeiitic evolution: Evidence fromUmnak Island in the Aleutians, J. Geophys. Res., 97, 321–343.

Minkoff, S. E. (1996), A computationally feasible approximateresolution matrix for seismic inverse problems, Geophys. J.Int., 126, 345–359.

Moser, T. J. (1991), Shortest path calculation of seismic rays,Geophysics, 56, 59–67.

Muntener, O., R. Perez-Alonso, and P. Ulmer (2004), Phaserelations of garnet, amphibole and plagioclase in H2O under-saturated andesite liquids at high pressure and implicationsfor the genesis of lower arc crust, Geophys. Res. Abstr., 6,5183.

Myers, J. D., C. D. Frost, and C. L. Angevine (1986), A test ofa quartz eclogite source for parental Aleutian magmas: Amass balance approach, J. Geol., 94, 811–828.

Myers, J. D., B. D. Marsh, C. D. Frost, and J. A.Linton (2002), Petrological constraints on the spatialdistribution of crustal magma chambers, Atka volcaniccenter, central Aleutian arc, Contrib. Mineral. Petrol.,143, 567–586.

Nolet, G. (1985), Solving or resolving inadequate andnoisy tomographic systems, J. Comput. Phys., 61,463–482.

Nolet, G., R. Montelli, and J. Virieux (1999), Explicit, approx-imate expressions for the resolution and a posteriori covar-iance of massive tomographic systems, Geophys. J. Int., 138,36–44.

Nowack, R. L., and J. A. Lyslo (1989), Frechet derivatives forcurved interfaces in the ray approximation, Geophys. J., 97,497–509.

Paige, C. C., and M. A. Saunders (1982), LSQR: Sparse linearequations and least squares problems, Assoc. Comput. Mach.Trans. Math. Software, 8, 195–209.

Parlett, B. N. (1998), The Symmetric Eigenvalue Problem,397 pp., Soc. for Ind. and Appl. Math., Philadelphia, Pa.

Pavlis, G. L., and J. R. Booker (1980), The mixed discrete-continuous inverse problem: Application to the simultaneousdetermination of earthquake hypocenters and velocity struc-ture, J. Geophys. Res., 85, 4801–4810.

Phillips, W. S., and M. C. Fehler (1991), Traveltime tomogra-phy: A comparison of popular methods, Geophysics, 56,1639–1649.

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P.Flannery (1986), Numerical Recipes in Fortran: The Art ofScientific Computing, 818 pp., Cambridge Univ. Press, NewYork.

Reginska, T. (1996), A regularization parameter in discrete ill-posed problems, SIAM J. Sci. Comput., 17, 740–749.

Ross, W. S. (1994), The velocity-depth ambiguity in seismictraveltime data, Geophysics, 59, 830–843.

Rudnick, R. L. (1995), Making continental crust, Nature, 378,571–578.

Sambridge, M. S. (1990), Non-linear arrival time inversion:Constraining velocity anomalies by seeking smooth modelsin 3-D, Geophys. J. Int., 102, 653–677.

Scales, J. A. (1987), Tomographic inversion via the conjugategradient method, Geophysics, 52, 179–185.

Scales, J. A., P. Docherty, and A. Gersztenkorn (1990),Regularisation of nonlinear inverse problems: Imaging thenear-surface weathering layer, Inverse Problems, 6, 115–131.

Scholl, D. W., T. L. Vallier, and A. J. Stevenson (1983), Arc,forearc, and trench sedimentation and tectonics; Amlia cor-ridor of the Aleutian Ridge, AAPG Mem., 34, 413–439.

Shaw, P. R., and J. A. Orcutt (1985), Waveform inversion ofseismic refraction data and applications to young Pacificcrust, Geophys. J. R. Astron. Soc., 82, 375–414.

Shillington, D. J., H. J. A. Van Avendonk, W. S. Holbrook,P. B. Kelemen, and M. J. Hornbach (2004), Composition andstructure of the central Aleutian island arc from arc-parallelwide-angle seismic data, Geochem. Geophys. Geosyst., 5,doi:10.1029/2004GC000715, in press.

Simon, H. D. (1984), The Lanczos algorithm with partial re-orthogonalization, Math. Comput., 42, 115–142.

Singer, B. S., J. D. Myers, and C. D. Frost (1992), Mid-Pleistocene basalts from the Seguam volcanic center, centralAleutian arc, Alaska: Local lithospheric structures andsource variability in the Aleutian arc, J. Geophys. Res., 97,4579–4586.

Smithson, S. B., F. Wenzel, Y. V. Ganchin, and I. B. Morozov(2000), Seismic results at Kola and KTB deep scientificboreholes: Velocities, reflections, fluids, and crustal compo-sition, Tectonophysics, 329, 301–317.

Spakman, W., and G. Nolet (1988), Imaging algorithms, accu-racy and resolution in delay time tomography, in Mathema-tical Geophysics, edited by N. J. Vlaar et al., pp. 155–187,D. Reidel, Norwell, Mass.

Spudich, P., and J. A. Orcutt (1980), A new look at the seismicvelocity structure of the oceanic crust, Rev. Geophys., 18,627–645.

Tarantola, A. (1987), Inverse Problem Theory, 613 pp., Else-vier Sci., New York.

ten Brink, U. S., and T. M. Brocher (1987), Multichannelseismic evidence for a subcrustal intrusive complex underOahu and a model for Hawaiian volcanism, J. Geophys. Res.,92, 13,687–13,707.

Thurber, C. H. (1983), Earthquake locations and three-dimen-sional crustal structure in the Coyote Lake area, centralCalifornia, J. Geophys. Res., 88, 8226–8236.

Thurber, C. H. (2003), Seismic tomography of the litho-sphere with body waves, Pure Appl. Geophys., 160,717–737.

Tikhonov, A. N., and V. Y. Arsenin (1977), Solutions of Ill-Posed Problems, 258 pp., John Wiley, Hoboken, N. J.

Toomey, D. R., and G. R. Foulger (1989), Tomographic inver-sion of local earthquake data from the Hengill-Grensdalurcentral volcano complex, Iceland, J. Geophys. Res., 94,17,497–17,510.

Trampert, J., and R. Snieder (1996), Model estimations biasedby truncated expansions: Possible artifacts in seismic tomog-raphy, Science, 271, 1257–1260.

Um, J., and C. Thurber (1987), A fast algorithm for two-pointray tracing, Bull. Seismol. Soc. Am., 77, 972–986.

Van Avendonk, H. J. A., A. J. Harding, J. A. Orcutt, and J. S.McClain (1998), A two-dimensional tomographic study ofthe Clipperton transform fault, J. Geophys. Res., 103,17,885–17,899.

Van Avendonk, H. J. A., A. J. Harding, J. A. Orcutt, and W. S.Holbrook (2001a), Hybrid shortest path and ray bendingmethod for traveltime and raypath calculations, Geophysics,66, 648–653.

Van Avendonk, H. J. A., A. J. Harding, J. A. Orcutt, and J. S.McClain (2001b), Contrast in crustal structure across the

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

23 of 24

Page 24: Geochemistry 3 Volume 5 Geophysics Geosystemsdjs/donna_ldeo/publications_files/van... · Pitfalls and successful strategies ... a new interpretation of the seismic refraction data

Clipperton transform fault from travel time tomography,J. Geophys. Res., 106, 10,961–10,981.

Vasco, D. W., L. R. Johnson, and O. Marques (1999), GlobalEarth structure: Inference and assessment, Geophys. J. Int.,137, 381–407.

Wang, B., and L. W. Braile (1996), Simultaneous inversion ofreflection and refraction data and application to field datafrom the northern Rio Grande rift, Geophys. J. Int., 125,443–458.

White, D. J. (1989), Two-dimensional seismic refraction tomo-graphy, Geophys. J., 97, 223–245.

Zelt, C. A. (1999), Modelling strategies and model assessmentfor wide-angle seismic traveltime data, Geophys. J. Int., 139,183–204.

Zelt, C. A., and P. J. Barton (1998), Three-dimensional seismicrefraction tomography: A comparison of two methods ap-plied to data from the Faeroe Basin, J. Geophys. Res., 103,7187–7210.

Zelt, C. A., andR. B. Smith (1992), Seismic traveltime inversionfor 2-D crustal velocity structure,Geophys. J. Int.,108, 16–34.

Zelt, C. A., and B. C. Zelt (1998), Study of out-of-plane effectsin the inversion of refraction/wide-angle reflection travel-times, Tectonophysics, 286, 209–221.

Zelt, C. A., A. M. Hojka, E. R. Flueh, and K. D. McIntosh(1999), 3D simultaneous seismic refraction and reflectiontomography of wide-angle data from the central Chileanmargin, Geophys. Res. Lett., 26, 2577–2580.

Zelt, C. A., K. Sain, J. V. Naumenko, and D. S. Sawyer (2003),Assessment of crustal velocitymodels using seismic refractionand reflection tomography, Geophys. J. Int., 153, 609–626.

Zhang, J., and M. N. Toksoz (1998), Nonlinear refrac-tion traveltime tomography, Geophysics, 63, 1726–1737.

Zhang, J., U. S. ten Brink, and M. N. Toksoz (1998), Nonlinearrefraction and reflection travel time tomography, J. Geophys.Res., 103, 29,743–29,757.

GeochemistryGeophysicsGeosystems G3G3

van avendonk et al.: inferring structure aleutian arc 10.1029/2003GC000664

24 of 24