Cortical Surface-Based Analysis - FreeSurfer · Cortical Surface-BasedAnalysis II: Inflation,...

13
Cortical Surface-Based Analysis II: Inflation, Flattening, and a Surface-Based Coordinate System Bruce Fischl,* Martin I. Sereno,² and Anders M. Dale* ,1 *Nuclear Magnetic Resonance Center, Massachusetts General Hosp/Harvard Medical School, Building 149, 13th Street, Charlestown, Massachusetts 02129; and ²Department of Cognitive Science, University of California at San Diego, Mailcode 0515, 9500 Gilman Drive, La Jolla, California 92093-0515 Received May 27, 1998 The surface of the human cerebral cortex is a highly folded sheet with the majority of its surface area buried within folds. As such, it is a difficult domain for computational as well as visualization purposes. We have therefore designed a set of procedures for modify- ing the representation of the cortical surface to (i) inflate it so that activity buried inside sulci may be visualized, (ii) cut and flatten an entire hemisphere, and (iii) transform a hemisphere into a simple param- eterizable surface such as a sphere for the purpose of establishing a surface-based coordinate system. r 1999 Academic Press Key Words: Cortical surface reconstruction, flatten- ing, coordinate systems, atlas. 1. INTRODUCTION Currently, the most widely used method of analyzing functional brain imaging data is to project the func- tional data from a sequence of slices onto a standard- ized anatomical 3-D space. The most common of these procedures is based on the Talairach atlas (Talairach et al., 1967; Talairach and Tournoux, 1988; see, e.g., Collins et al., 1994, for an automated procedure). While this type of approach has certain advantages (ease of use, widespread acceptance, applicability to subcortical structures), it also has significant drawbacks. These drawbacks derive from the fact that the intrin- sic topology of the cerebral cortex is that of 2-D sheet with a highly folded and curved geometry. Estimates of the amount of ‘‘buried’’ cortex range from 60 to 70% (Zilles et al., 1988; Van Essen and Drury, 1997), imply- ing that distances measured in 3-D space between two points on the cortical surface will substantially under- estimate the true distance along the cortical sheet, particularly in cases where the points lie on different banks of a sulcus. From a functional standpoint, nonhuman primate neocortex is composed of a mosaic of visual, auditory, somatosensory, and motor areas, with visual areas alone occupying more than half of the total cortical surface area (Felleman and Van Essen, 1991; Kaas and Krubitzer, 1991; Sereno and Allman, 1991). The bulk of the remaining half is composed of auditory, somatosen- sory, motor, and limbic areas, each occupying about 1 /8 of the total neocortex (Morel and Kaas, 1992; Stepniewska et al., 1993). The majority of these areas are defined by their topographic maps of the sensory periphery (e.g., retino- topic, tonotopic, somatotopic). Typically, the metric encoding the relationship between these maps and the sensory periphery which they represent is not known (see, Schwartz, 1977, 1980, for a notable exception). However, the two-dimensional nature of the maps as well as their topographic arrangement strongly suggest that a two-dimensional surface-based metric is more appropriate for analyzing their functional properties than the more typically used volume-based metrics. The highly folded nature of the cortical surface also makes it difficult to view functional activity in a meaningful way. The typical means of visualization of this type of data is the projection of functional activa- tion onto a set of orthogonal slices. This procedure is problematic as regions of activity which are close together in the volume may be relatively far apart in terms of the distance measured along the cortical surface. In addition, the naturally two-dimensional organization of cortical maps is largely obscured by the imposition of an external coordinate system in the form of orthogonal slices. These problems have led an increas- ing number of studies to make use of surface-based techniques for visualization (Tootell et al., 1995; DeYoe et al., 1996; Engel et al., 1997; Reppas et al., 1997; Talavage et al., 1997; Van Essen and Drury, 1997; Hadjikhani et al., 1998; Moore et al., 1998). 1 To whom correspondence and reprint requests should be ad- dressed. Fax: (617) 726-7422. E-mail: [email protected]. NeuroImage 9, 195–207 (1999) Article ID nimg.1998.0396, available online at http://www.idealibrary.com on 195 1053-8119/99 $30.00 Copyright r 1999 by Academic Press All rights of reproduction in any form reserved.

Transcript of Cortical Surface-Based Analysis - FreeSurfer · Cortical Surface-BasedAnalysis II: Inflation,...

  • fbchiivaeeA

    i

    ftipaCtus

    swt(ipe

    d

    NeuroImage 9, 195–207 (1999)Article ID nimg.1998.0396, available online at http://www.idealibrary.com on

    Cortical Surface-Based AnalysisII: Inflation, Flattening, and a Surface-Based Coordinate System

    Bruce Fischl,* Martin I. Sereno,† and Anders M. Dale*,1

    *Nuclear Magnetic Resonance Center, Massachusetts General Hosp/Harvard Medical School, Building 149, 13th Street,Charlestown, Massachusetts 02129; and †Department of Cognitive Science, University of California at San Diego,

    Mailcode 0515, 9500 Gilman Drive, La Jolla, California 92093-0515

    Received May 27, 1998

    pb

    nsasKtste

    ttes(Hwtat

    mmttpttsoioite

    The surface of the human cerebral cortex is a highlyolded sheet with the majority of its surface areauried within folds. As such, it is a difficult domain foromputational as well as visualization purposes. Weave therefore designed a set of procedures for modify-

    ng the representation of the cortical surface to (i)nflate it so that activity buried inside sulci may beisualized, (ii) cut and flatten an entire hemisphere,nd (iii) transform a hemisphere into a simple param-terizable surface such as a sphere for the purpose ofstablishing a surface-based coordinate system. r 1999cademic Press

    Key Words: Cortical surface reconstruction, flatten-ng, coordinate systems, atlas.

    1. INTRODUCTION

    Currently, the most widely used method of analyzingunctional brain imaging data is to project the func-ional data from a sequence of slices onto a standard-zed anatomical 3-D space. The most common of theserocedures is based on the Talairach atlas (Talairach etl., 1967; Talairach and Tournoux, 1988; see, e.g.,ollins et al., 1994, for an automated procedure). While

    his type of approach has certain advantages (ease ofse, widespread acceptance, applicability to subcorticaltructures), it also has significant drawbacks.These drawbacks derive from the fact that the intrin-

    ic topology of the cerebral cortex is that of 2-D sheetith a highly folded and curved geometry. Estimates of

    he amount of ‘‘buried’’ cortex range from 60 to 70%Zilles et al., 1988; Van Essen and Drury, 1997), imply-ng that distances measured in 3-D space between twooints on the cortical surface will substantially under-stimate the true distance along the cortical sheet,

    1 To whom correspondence and reprint requests should be ad-

    ressed. Fax: (617) 726-7422. E-mail: [email protected].

    195

    articularly in cases where the points lie on differentanks of a sulcus.From a functional standpoint, nonhuman primate

    eocortex is composed of a mosaic of visual, auditory,omatosensory, and motor areas, with visual areaslone occupying more than half of the total corticalurface area (Felleman and Van Essen, 1991; Kaas andrubitzer, 1991; Sereno and Allman, 1991). The bulk of

    he remaining half is composed of auditory, somatosen-ory, motor, and limbic areas, each occupying about 1⁄8 ofhe total neocortex (Morel and Kaas, 1992; Stepniewskat al., 1993).The majority of these areas are defined by their

    opographic maps of the sensory periphery (e.g., retino-opic, tonotopic, somatotopic). Typically, the metricncoding the relationship between these maps and theensory periphery which they represent is not knownsee, Schwartz, 1977, 1980, for a notable exception).owever, the two-dimensional nature of the maps asell as their topographic arrangement strongly suggest

    hat a two-dimensional surface-based metric is moreppropriate for analyzing their functional propertieshan the more typically used volume-based metrics.

    The highly folded nature of the cortical surface alsoakes it difficult to view functional activity in aeaningful way. The typical means of visualization of

    his type of data is the projection of functional activa-ion onto a set of orthogonal slices. This procedure isroblematic as regions of activity which are closeogether in the volume may be relatively far apart inerms of the distance measured along the corticalurface. In addition, the naturally two-dimensionalrganization of cortical maps is largely obscured by themposition of an external coordinate system in the formf orthogonal slices. These problems have led an increas-ng number of studies to make use of surface-basedechniques for visualization (Tootell et al., 1995; DeYoet al., 1996; Engel et al., 1997; Reppas et al., 1997;

    Talavage et al., 1997; Van Essen and Drury, 1997;

    Hadjikhani et al., 1998; Moore et al., 1998).

    1053-8119/99 $30.00Copyright r 1999 by Academic Press

    All rights of reproduction in any form reserved.

  • nfdpDt

    a

    aswe

    wnn

    bt1tas1vCpa1hpaacm(

    csslCmm

    aaoctssteopbirplra

    mah

    b1sai

    196 FISCHL, SERENO, AND DALE

    In order to facilitate the use of surface-based tech-iques for both display and analysis of structural andunctional properties of the cerebral cortex, we haveeveloped a unified procedure which begins with areviously reconstructed cortex (Dale and Sereno, 1993;ale et al., 1998) and modifies it in order to achieve

    hree separate but related goals:(1) The ‘‘inflation’’ of the cortical surface so that

    ctivity occurring inside sulci may be easily visualized.(2) The flattening of an entire hemisphere so that the

    ctivity across the hemisphere may be seen from aingle view, and so that computational procedureshich are not tractable on arbitrary manifolds may bemployed in the analysis of the cerebral cortex.(3) The ‘‘morphing’’ of a hemisphere into a surface,hich maintains the topological structure2 of the origi-al surface, but has a natural (i.e., closed-form) coordi-ate system.Using the methods described in this paper we have

    een able to map out the detailed topographic organiza-ion of human retinotopic visual areas (Sereno et al.,995; Tootell et al., 1996a,b, 1997, 1998a,b), the tono-opic structure of primary auditory cortex (Talavage etl., 1996, 1997a,b, 1998a,b), the topography of primaryomatosensory cortex (Moore et al., 1997, 1998; Moore,998), as well as areas involved in the processing ofisual motion (Tootell et al., 1995; Reppas et al., 1997;ulham et al., 1998), color (Hadjikhani et al., 1998),erception of faces and objects (Halgren et al., 1998),nd processing the meaning of words (Halgren et al.,998). In addition, the cortical surface reconstructionas been used to constrain the EEG/MEG inverseroblem (Dale and Sereno, 1993; Liu et al., 1996; Liu etl., 1998a,b). The cortical surface reconstruction, visu-lization, and analysis tools described here and in aompanion article (Dale et al., 1998) are based on theethods previously described by Dale and Sereno

    1993).

    2. MAPPING OF THE CORTICAL SURFACETO PARAMETERIZABLE SHAPES

    Because of the varying intrinsic curvature of theortical surface it is not possible to map it onto otherignificantly smoother surfaces (such as planes orpheres) without introducing some metric and/or topo-ogical distortion into the surface representations (doarmo, 1976). A mapping between two surfaces with noetric distortion is called an isometry. Finding such aapping from the sphere to the plane has been called

    2 The term topological structure is frequently used to refer to theorder of a domain as opposed to its global topology (Mortenson,997). For example, once an incision has been made in the corticalurface it is topologically equivalent to a plane. Further incisionslter its topological structure, but not its topology (unless they result

    mn multiple disconnected components).

    the mapmaker’s problem and was shown to be impos-sible by Gauss (1828), as the surfaces in question havediffering intrinsic (or Gaussian) curvature. Neverthe-less, for representations to be useful for either visualiza-tion or computational purposes, metric distortion mustbe minimized. Toward that end, we have developed ageneral procedure for minimizing metric distortion in avariety of contexts, such as surface inflation, flattening,as well as mapping to other parameterizable surfacessuch as a sphere.

    Constructing this type of mapping is a difficult taskdue to the complex and highly folded nature of theoriginal surface, which requires a fine-scale tessella-tion in order to capture its metric and topologicalproperties. One attractive means of flattening thesurface is the method employed by Schwartz andcolleagues (Schwartz and Merker, 1986; Schwartz etal., 1989; Wolfson and Schwartz, 1989), in which thematrix of distances of each vertex to all other vertices isconstructed in order to represent the metric propertiesof the original surface. The surface is then randomlyprojected onto a plane and unfolded in such a way as tominimize the mean-squared error between the originaldistance matrix and that of the flattened surface. Whilethis method is more than adequate for flattening smallpatches of the cortical surface, such as primary visualcortex to which it was originally applied, the computa-tional requirements of the procedure in terms of bothmemory and time become prohibitive as the patch sizegrows.

    A different type of method was employed by Dale andSereno (1993), and later by Carman et al. (1995) as wellas Drury and Van Essen (Drury et al., 1996; Van Essennd Drury, 1997; Van Essen et al., 1998). In thispproach, a variety of local forces are constructed inrder to encourage the preservation of local area andonformality (i.e., angle), while also forcing the surfaceo unfold onto a plane. These techniques have beenuccessfully applied to entire cortical hemispheres, butuffer from a number of drawbacks. First, they requirehe use of terms such as a spring force in order toliminate folds, which results in surfaces that are notptimal with respect to the preservation of any metricroperty. In addition, they treat the vertices on theorders of the flattened surface differently than thosen the interior, thus constraining the shape of theesulting surface. Finally, they preserve only localroperties of the surface and therefore do not rule outarge-scale distortions caused by locally correlated er-ors, although the use of multiresolution techniquesddresses this concern to some degree.Part of the problem with applying the Schwartzethod is that relatively long-range distances must be

    ccounted for in order to unfold patches of cortex whichave been folded by the projection process. They esti-

    ate that a procedure incorporating distance con-

  • sVrsmdbc

    esivpwTtptsvutmdttl

    sfiisTi

    tieuaoufitapias

    pm

    dsccattpwftfted

    2

    sd3tnm

    wnvtno

    wveifcr

    2

    a

    J

    tf

    197CORTICAL SURFACE-BASED ANALYSIS II

    traints on the order of 1 cm suffices to unfold monkey1 (Schwartz et al., 1989). Unfortunately, the distanceequired to smooth out a fold grows with the size of theurface (and the fold), quickly requiring untenableemory usage. Using a random subsampling of the

    istance matrix alleviates this problem to some extent,ut not to the degree required to flatten an entireortical hemisphere.This problem occurs because distances are unori-

    nted, and therefore mirror image configurations repre-ent local minima in the energy functional. To see this,magine a piece of paper folded exactly along a string ofertices. If only nearest neighbor distances are beingreserved, this represents an optimal configurationith the same energy as the completely unfolded state.he inclusion of neighborhoods which are small rela-ive to the size of the entire sheet will not aid theroblem, as the majority of the nodes on the surface arehen beyond the neighborhood of the fold. This type ofituation thus represents a local minimum, as movingertices along the fold will increase the metric errorntil the rest of the surface expands. In order to causehe surface to unfold, a sufficient number of verticesust be included in the distance matrix so that the

    ecrease in error caused by removing the fold morehan offsets the increase in error of the region outsidehe fold, a solution that is not viable for as complex andarge a surface as an entire cortical hemisphere.

    We therefore construct a means of encouraging theurface to unfold which satisfies three criteria: (1) Thenal surface should be optimal with respect to minimiz-

    ng metric distortion. (2) The borders of the cut surfacehould be treated no differently than the interior. (3)he resulting surface should have only minimal fold-

    ng.The first two criteria exclude the use of spring terms

    o ‘‘regularize’’ the mesh, which are typically introducedn order to prevent folding. Instead, we construct annergy functional that employs only a distance term fornfolded or positive regions of the surface, but appliesn additional term to folded or negative regions inrder to cause the surface to unfold. This term makesse of the embedding space to give the normal vectoreld of the surface a consistent orientation (positive z inhe plane, radially outward on the sphere). Any tri-ngles in the tessellation for which the ordered cross-roduct of its legs is antiparallel to the normal directions then assigned a negative area.3 Constraining therea to be positive everywhere eliminates folds in theurface.The spherical and flattening transformations thus

    roceed as follows. First, geodesic distances are esti-ated on the original (folded) surface, after the intro-

    3 This oriented area term can be seen as the determinant of the

    acobian matrix of the transformation. a

    uction of cuts in the flattened case. Note that after theurface has been cut, only the remaining vertices areonsidered in the distance calculations. Next, the corti-al manifold is projected onto the target surface andssigned a normal vector field with a consistent orienta-ion (the inflation procedure does not require a projec-ion step as the energy functional used in the inflationrocess contains a term which drives the surface to-ard the target configuration). The potentially large

    olds and metric distortions introduced by the projec-ion process are then removed by minimizing an energyunctional, which contains terms representing thesewo factors separately. The resulting surfaces havessentially no remaining folds and only minimal metricistortion.

    .1. Minimizing Metric Distortions

    The term that minimizes metric distortions is con-tructed as follows. Consider a mesh of V verticesistributed irregularly over a surface S embedded in a-D Cartesian space. Denoting the distance betweenhe ith and jth vertices at iteration number t of theumerical optimization procedure by dij

    t , we construct aean-squared energy functional Jd:

    Jd 51

    4V oi51V

    on[N(i)

    (dint 2 din

    0 )2, dint 5 \xi

    t 2 xnt \ (1)

    here xit is the (x, y, z) position of vertex i at iteration

    umber t, din0 is the distance between the ith and nth

    ertices on the original cortical surface before projec-ion, and N(i) is the set of vertices defined to be in theeighborhood of vertex i.4 Taking the partial derivativef Jd with respect to the kth vertex results in

    Jdxk

    51

    V on[N(k) (dknt 2 dkn

    0 )ekn, (2)

    here ekn is a unit vector pointing from vertex k toertex n. This term is similar to the ‘‘longitudinal’’ onemployed by Carman et al. (1995), except here it isdentified as the gradient of a well-defined energyunction, and the distances we employ are along signifi-ant sections of the manifold, as opposed to simplyepresenting the spacing of the mesh.

    .2. Long Range-Distance Calculation

    The problem of computing geodesic distances on anrbitrary manifold is a difficult one. One solution is to

    4 The 14 scaling factor removes the factor of 2 introduced by takinghe derivative of the quadratic distance term as well as an additionalactor of 2 that accounts for all the symmetric terms in which xk

    ppears as a neighbor of another vertex.

  • upWrncdmrbacs

    ttt(gsact

    cncta22ttpphitgsctdn

    2

    uscf

    fiessimsrtsat

    ftnivottnotoan

    sepa

    J

    wtrt

    nft

    198 FISCHL, SERENO, AND DALE

    se the faces of the polyhedral approximation to com-ute the exact geodesic distances, as suggested byolfson and Schwartz (1989). Unfortunately, the time

    equirements of this algorithm are exponential in theumber of triangles in the tessellation, which makes itomputationally untenable if millimeter resolution isesired for the surface. While decimation techniquesay be employed to reduce the size of the surface

    epresentation, it is not clear that the reduction woulde sufficient to make the use of this algorithm feasible,s on the order of 300,000 triangles are required toover a typical human cortex and single millimeter-ized structures are common.A simpler technique that is computationally trac-

    able is to take a dynamic programming approach tohe calculation of distances and employ an algorithmypically used to calculate minimal distances in a graphDijkstra, 1959). Here we modify it slightly, as theraph in question is the tessellation of the corticalurface. This requires a minor correction in the form ofscaling factor, so that the distance estimates on the

    ortical manifold are essentially unbiased with respecto the geodesic distances.

    The Dijkstra algorithm for computing distances pro-eeds as follows. For each vertex we label each of itsearest neighbors as 1-neighbor and compute the Eu-lidean distance from them to the central vertex. Wehen label each neighbor of a 1-neighbor that is notlready labeled (and is not the central vertex) as a-neighbor. Next, we examine the neighbors of each-neighbor and find the one with the shortest distanceo the central vertex. We then compute the distance tohe 2-neighbor as the distance of the optimal neighborlus the length of the edge connecting them. Thisrocedure is then applied iteratively with the neighbor-ood size expanding at each iteration. One point to note

    s that the resulting distances are similar to a Manhat-an metric in that they typically overestimate theeodesic distance. In order to alleviate this problem wecale the resulting distances by ((1 1 sqrt(2))/2). Thisorrection factor results in an estimate which is essen-ially zero mean,5 with the variation from the trueistance acting as white noise, a factor we will show isegligible in section 6.

    .3. Unfolding Using Oriented Area

    As noted previously, causing the surface to unfoldsing only a distance term is not feasible for largeurfaces. This is due to the fact that mirror-imageonfigurations are not directly penalized, resulting inolded states that are local minima of the energy

    5 The scaling factor we use is representative of the rectangularature of the tessellation, as it is constructed directly from voxel

    aces. A different scaling factor is required for other tessellation

    echniques. M

    unctional. These local minima are caused by thenherently unoriented nature of distances that do notxplicitly distinguish between folded and unfoldedtates. In order to resolve this problem, we thereforeeek an oriented metric property that discourages foldsn the surface. The two obvious candidates are confor-

    ality and areal terms. While both can be employeduccessfully in this context, the use of an angle termesults in a gradient that is dependent on the square ofhe inverse of the vertex spacing and is thereforeomewhat numerically unstable. In contrast, the use ofn oriented area results in a quadratic energy func-ional.

    In order to define the areal term of the energyunctional we consider the ith triangle in the surfaceessellation depicted in Fig. 1, with unit normal vectori and edges ai and bi connecting the vertex xi to two of

    ts neighbors (note that bold-faced symbols denoteector quantities). The unit normal ni is given on theriginal manifold by the normalized cross product ofhe edges ai and bi, while the area of the triangle is halfhe cross product of ai and bi dotted with the unitormal (i.e., the triple scalar product). However, forther representations such as spherical and flattened,he normal vector field can be given a consistentrientation on the surface6 using the embedding space,nd Ai becomes an oriented area, which may take onegative values, indicating folds in the surface.Given this description of the metric properties of the

    urface through the triangular tessellation, we form annergy functional Ja that penalizes negative area inroportion to the difference between the current areand the original area occupied by each triangle:

    a 51

    2T oi51T

    P(Ait)(Ai

    t 2 Ai0)2, P(Ai

    t) 5 51, Ait # 0

    0, otherwise,(3)

    here, as before, superscripts denote time, with 0 beinghe areal values on the original cortical surface, Tefers to the number of triangles in the tessellation, andhe functional dependence of the Ais on the position of

    6 This is always possible except in pathological cases such as the

    FIG. 1. Metric properties of the triangular tessellation.

    öbius strip which are said to be nonorientable (do Carmo 1976).

  • tstdwts

    r

    Ey

    Tcpt

    2

    d

    wimvoautetpbbag1

    m

    fTmmstatstTwtep

    wvfmrs

    Jaaadsc

    irmgmsFt

    wtcd

    a

    199CORTICAL SURFACE-BASED ANALYSIS II

    he vertex and its neighbors has been suppressed forimplicity of notation. This term ensures that theransformation is one-to-one and therefore has a well-efined inverse, by preventing folds in the surfacehich would represent the mapping of multiple points

    o the same location in the embedding coordinateystem.In order to minimize Ja, we take the gradient with

    espect to the vertex positions xk:

    Jaxk

    51

    T oi51T

    (Ait 2 Ai

    0)Ai

    t

    xk. (4)

    xpanding the partial derivative using the chain ruleields:

    Ait

    xk5

    Ait

    ai

    aixk

    1Ai

    t

    bi

    bixk

    ,Ai

    t

    ai5 bi 3 ni

    Ait

    bi5 ni 3 ai. (5)

    he partials of the change in the legs with respect to ahange in the vertex position are dependent on whatosition the vertex in question occupies in a givenriangle:

    ai

    xk5 5

    [21,21,21]T, k 5 i

    [1,1,1], k 5 l

    0, otherwise

    ,bixk

    5 5[21,21,21]T, k 5 i

    [1,1,1], k 5 j

    0, otherwise,

    (6)

    .4. The Complete Energy Functional

    The complete energy functional incorporating bothistance and areal terms is given by

    J 5 ldJd 1 laJa, (7)

    here the la and ld coefficients define the relativemportance of unfolding versus the minimization of

    etric distortions respectively. Initially, la takes onalues much larger than ld, and gradually decreasesver time as the surface successfully unfolds. Onedditional point to note is that we smooth the gradientssing iterative averaging during the numerical integra-ion. This allows entire regions that are compressed orxpanded to move coherently in the appropriate direc-ion, and is similar to decimation followed by upsam-ling with interpolation. We allow each scale (definedy the number of iterations in the averaging) to equili-rate before reducing the scale and continuing. Thectual minimization of J(x) is accomplished usingradient descent with line minimization (Press et al.,994), as detailed in the appendix.

    3. SURFACE INFLATION

    The high degree of folding of the cortical surface

    akes it desirable to inflate the reconstructed surface b

    or visualization purposes (Dale and Sereno, 1993).his renders the interior of sulci visible, as well asaking the surface-based distance between regionsore apparent to visual inspection. The purpose of the

    urface inflation is thus to provide a representation ofhe cortical hemisphere that retains much of the shapend metric properties of the original surface, but allowshe visualization of functional activity occurring withinulci. For this purpose, we define an energy functional,he minimization of which results in the desired shape.his functional consists of two terms, a spring forcehich smooths the surface, and the metric-preserva-

    ion term described in Section 2.1, which constrains thevolving surface to retain as much of the original metricroperties as possible:

    Js 51

    2V 1oi51V

    on[N1(i)

    \xi 2 xn \22 1 ldJd, (8)here Nl denotes the set of nearest neighbors of eachertex, Jd is as defined in Section 2.1, and a value of 0.1or the coefficient ld yields smooth surfaces with mini-al metric distortion. Larger values of ld result in

    educed metric distortion at the cost of generating lessmooth surfaces.We use Euler’s method with momentum to integrate

    s until the surface has achieved a desired smoothnesss measured by the goodness-of-fit of the polyhedralpproximation.7 Note that, in contrast to the flatteningnd spherical transformation, the surface inflationoes not require a projection step, as the spring termerves to drive the surface toward the desired smootheronfiguration.One further point to note is that the inflation process

    s driven by the average convexity or concavity of aegion. That is, points which lie in concave regionsove outwards over time, while points in convex re-

    ions move inwards. Thus, integrating the normalovement of a point during inflation provides a mea-

    ure of average convexity or concavity at that point.ormally, the average convexity C(xk

    0) at position xk onhe original surface is given by:

    C (xk0) 5 e 1

    Jsxk

    t· n(k)2 dt, (9)

    here n(k) is the unit normal vector to the surface athe kth vertex in the tessellation at time t. Vertices thatonsistently move outwards, or parallel to the normalirection, will have a large positive value of C, while

    7 We integrate the inflation functional until the normalized aver-ge distance of the neighbors of each vertex from its tangent plane is

    elow a prespecified threshold.

  • pa

    fgtsatFcctfsasevtaaai(

    matoeopomtt(citdppiwrcsbss

    c1

    rtttat[trvpfl

    sfe1Fcscefintcmatirotooaocidde

    tsd

    200 FISCHL, SERENO, AND DALE

    oints that move antiparallel to the normal will bessigned large negative values.The average convexity is useful for quantifying the

    olding pattern of a surface, as C captures large-scaleeometric features, while being relatively insensitive tohe small folds that typically occur on the banks of aulcus. This is in contrast to mean curvature whichttains equally high values for small secondary andertiary folds in a surface as for the primary folds.igure 3 illustrates this difference between the meanurvature of the folded cortex (left) and the averageonvexity as quantified by C (right), painted onto bothhe gray/white matter (top), and inflated (bottom) sur-aces. Note how accurately the average convexity repre-ents only the primary folding patterns. The majornatomical features of this surface, such as the centralulcus (CS), superior temporal sulcus (STS), intrapari-tal sulcus (IPS), and sylvian fissure (SF), are clearlyisible, while the secondary and tertiary folding pat-erns apparent in the mean curvature are largelybsent. In our standard analysis procedure we use C asmeans for quantifying the folding pattern of a surfaces it is insensitive to noise in the form of small wrinklesn a surface and relatively stable across individualssee Fischl et al., 1998).

    4. FLATTENING

    In order to flatten a cortical hemisphere with mini-al distortion we make a number of cuts on the medial

    spect of the original surface—one in a region aroundhe corpus callosum to remove all midbrain structures,ne down the fundus of the calcarine sulcus, a set ofqually spaced radial cuts, as well as a sagittallyriented cut around the temporal pole (see Fig. 4). Thisattern of cuts removes most of the intrinsic curvaturef the surface, allowing it to be flattened with onlyinor distortion, while preserving the topological struc-

    ure of the lateral aspect of the surface. Several alterna-ive cutting schemes have been suggested by othersDeYoe et al., 1996; Drury et al., 1996). The optimalhoice of cuts depends on which parts of the surface ones most interested in preserving for a particular applica-ion. For example, in our standard analysis of visualata, we make a planar cut that detaches the posteriorart of the brain including all of the occipital lobe, andarts of the parietal and temporal lobes. We thenntroduce a cut down the fundus of the calcarine sulcus,hich separates the upper and lower visual fields of the

    etinotopic areas and removes most of the intrinsicurvature of the remaining surface. Currently, theurface cuts are made manually by a trained operator,ased on anatomical landmarks. In the future, ithould be possible to automate the cutting process, by

    pecifying the coordinates of the cuts in a surface-based 2

    oordinate system (Fischl et al., 1998; Thompson et al.,998; Sereno et al., 1996; Van Essen et al., 1997).Once the desired cuts have been made, we project the

    esulting surface onto a plane whose normal is given byhe average surface normal of the cut surface (that is,he portion of the original surface which remains afterhe cutting process). After the projection has beenccomplished, we give the flattened surface a consis-ent orientation by setting the normal vector field to0,0,1]T and allow the surface to unfold by minimizinghe energy given in Section 2.4, using angularly spacedandomly sampled distances in a 0.8 cm radius of eachertex as the neighborhood N(i).8 The result of thisrocedure is shown in Fig. 5, which depicts threeattened left hemispheres.

    5. SPHERICAL TRANSFORMATIONS AND ASURFACE-BASED COORDINATE SYSTEM

    Identifying corresponding points on different corticalurfaces requires the establishment of a uniform sur-ace-based coordinate system (Drury et al., 1996; Serenot al., 1996; Thompson et al., 1996; Thompson and Toga,996; Davatzikos, 1997; Van Essen and Drury, 1997;ischl et al., 1998). This is in contrast to volume-basedoordinate systems in which a point on the corticalurface in one volume will typically not lie on theortical surface of a different volume. In order tostablish a surface-based coordinate system, we trans-orm the reconstructed cortical surface into a parameter-zable surface, as the parameterization then provides aatural coordinate system. The surface we choose forhis purpose is a sphere for a number of reasons. Thishoice is primarily motivated by the fact that theapping of the cortical hemisphere onto a sphere

    llows the preservation of the topological structure ofhe original surface (i.e., the local connectivity). This isn contrast to the use of a flattened surface, whichequires cuts to be introduced prior to flattening inrder to minimize distortion. These cuts change theopological structure of the surface, resulting in pointsn opposite sides of a cut, which are close to each othern the original cortical surface, becoming quite farpart in the final flattened representation. The choicef the sphere also allows us to retain much of theomputational attractiveness of a flat space, facilitat-ng the calculation of metric properties such as geodesicistances, areas, and angles, properties that are moreifficult to compute on less symmetric surfaces such asllipsoids.

    8 At each distance we enforce a minimal angular spacing betweenhe sampled points. This spacing is based on the number of neighborselected at each extent. For the flattening we sample 8 points at eachistance out to 8 cm, forcing the points to be separated by as close to

    p/8 radians as possible.

  • siltgo

    wsttorstpubtrveic

    lcig(dsctiosboTitbio

    aam

    mdttrai

    sssoagdtpifausdpfltsatiFiacfsaft

    ssuts(tflc

    201CORTICAL SURFACE-BASED ANALYSIS II

    The process of unfolding the cortical surface on aphere is identical to the flattening procedure outlinedn Section 4, except that distances on the sphere are noonger Euclidean, but rather must be computed usinghe geodesics of the sphere. In the spherical case, weive the surface a consistent orientation by using anutwards pointing normal vector field:

    ni 5xi

    \xi \(10)

    here xi refers to the vector from the center of thephere to the ith vertex. In addition, the lack of freedomo modify the shape of the unfolding surface necessi-ates the use of longer range distances than in the casef the flattening. In order to generate the sphericalepresentation, we first project the inflated corticalurface onto the unit sphere by moving each vertex inhe tessellation of the inflated surface to the closestoint on the sphere. Next, we allow the surface tonfold by minimizing the metric distortions introducedy the inflation and projection procedures. These distor-ions are estimated based on an angularly spacedandom sampling of distances in a 1-cm radius of eachertex. The surface is projected back onto the sphere atach step in the minimization procedure. Figure 6llustrates the result of applying this procedure to threeortical hemispheres.Once the spherical representation has been estab-

    ished, we can use any of the standard sphericaloordinates systems (e.g., longitude and colatitude) tondex a point on any of the surface representations for aiven subject (Fig. 7). Furthermore, in other workSereno et al., 1996; Fischl et al., 1998), we haveeveloped a procedure for aligning cortical hemi-pheres with an average surface, based on the averageonvexity measure C defined by Eq. (9). By maximizinghe correlation of the convexity measure between thendividual and the average, the procedure computes anptimal mapping to a ‘‘canonical’’ surface and hence aurface-based coordinate system. This type of surface-ased approach, which a number of groups are workingn (Drury et al., 1996, 1997; Thompson et al., 1996;hompson and Toga, 1996; Davatzikos, 1997), can

    ncrease the accuracy of localizing anatomically consis-ent functional areas by a factor of three over volume-ased techniques (Fischl et al., 1998), as well as provid-ng a means for high-resolution intersubject averagingf functional data occurring on the cortical surface.

    6. ERROR ANALYSIS

    There are a number of concerns which must beddressed in regard to using the corrected Dijkstralgorithm as an estimator of geodesic distances. The

    ost important of these concerns the accuracy of the

    etric properties of surfaces flattened using theseistances. Another consideration is to what degreehese surfaces diverge from surfaces flattened usingrue geodesic distances as targets. A third question is inegard to the accuracy of the flattening procedurepplied to surfaces which contain varying amounts ofntrinsic curvature.

    In order to assess these issues, we flattened a set ofurfaces for which analytic geodesic distance expres-ions exist, namely a plane and various portions of aphere. The entire set of surfaces were flattened twice,nce using the corrected Dijkstra distances as targets,nd once with the targets computed using the trueeodesic distances. The results of this analysis areepicted in Fig. 8. The x-axis in this plot is a measure ofhe curvature of the original surfaces, beginning with alane at the far left (x 5 0), and proceeding to increas-ng portions of a sphere, with a full hemisphere at thear right (x 5 50%). The y-axis represents the percent-ge distance error of the flattened surfaces, measuredsing analytic expressions for geodesics on the originalurfaces (great circles on the sphere and Cartesianistances in the plane). The solid line in this figure is alot of distance error versus curvature for the surfacesattened using the corrected Dijkstra distances, whilehe dotted line depicts the distance error for the sameurfaces flattened using the actual geodesic distancess targets. As can be seen, the difference between thewo is small (the maximum difference is 1.37%), indicat-ng the overall accuracy of the flattening procedure.urthermore, the convergence of the plots indicates the

    nsensitivity of the corrected Dijkstra distances, as wells the flattening procedure in general, to increasingurvature. The difference in percentage distance erroror the two flattened hemispheres is less than 0.35%,uggesting that the error is dominated by the unavoid-ble distortions introduced by flattening a curved sur-ace, rather than inaccuracies resulting from the use ofhe corrected Dijkstra distances as targets.

    Finally, we present the errors of the flattening andpherical transformation on a set of 10 human hemi-pheres. All percentages in this section are computedsing an L1 norm as it is less sensitive to outliers thanhe L2 norm. The transformation of the cortex into aphere results in the largest metric distortion19.4 6 0.67%), presumably due to the lack of freedomo manipulate the borders on the closed shape. Theattening of the entire hemisphere gives rise to signifi-antly smaller distortion (11.7 6 0.42%), while the flat-

    tening of the posterior third of cortex, useful for analyz-ing visual areas, results in the smallest averagedistortion (9.6 6 0.65%). The spatial distribution ofmetric distortions introduced by the flattening andspherical transformations are shown in Figs. 9 and 10,respectively, together with a histogram of the distor-

    tions in Fig. 11. The regions of high distortion for the

  • s

    FIG. 2. Inflated representations of the three cortical surfaces (sulFIG. 3. Mean curvature (left) and average convexity (right) painte

    ubject’s cortical surface.FIG. 4. Medial view of an inflated surface after the introduction o

    ci are red and green are light).d onto folded (top) and inflated (bottom) representations of an individual

    f cuts. Yellow regions indicate the borders of the cut surface.

    202

  • r

    FIG. 5. Three flattened left hemispheres (sulci are red and greenFIG. 6. Lateral view of three left hemispheres after spherical tranFIG. 7. Spherical coordinate system painted onto a variety of surf

    are light).sformation.ace representations.

    FIG. 9. Spatial distribution of metric distortion introduced by the flattening, painted onto flattened (left), and inflated (right)epresentations (top—lateral view, bottom—medial view).

    203

  • flsstcvsctt

    pcpcgua

    t

    (r

    204 FISCHL, SERENO, AND DALE

    attening are largely confined to the borders of the cuturface, with the interior having a relatively uniformmall degree of distortion. The spherical transforma-ion results in higher distortion, but this too is mainlyonfined to noncortical regions such as the lateralentricle and the basal ganglia, where the surface isomewhat arbitrary. Thus, the actual mean error forortical regions is probably a few percent lower thanhe numbers cited above, as evidenced by the mode ofhe two histograms in Fig. 11.

    FIG. 8. Percentage distance error as a function of increasing curvarue geodesic distances (dashed).

    espectively.

    7. CONCLUSION

    In this paper we have presented a unified set ofrocedures for transforming a previously reconstructedortical surface. These transformations achieve tworimary goals. First, they facilitate visualization ofortical activation patterns, including the detailed topo-raphic organization of cortical areas, as well as distrib-ted activity occurring across an entire hemisphere. Inddition, such transformations enable two-dimensional

    e for surfaces flattened using corrected Dijkstra distances (solid) and

    tur

    FIG. 10. Spatial distribution of metric distortion introduced by the spherical transformation, painted onto flattened (left), inflatedcenter), and spherical (right) representations of the same surface. Lateral and medial views are shown in the top and bottom rows,

  • aspoftciroc

    8

    mutp

    Sipsmsskcbc

    wTamtktwqtpi

    8

    lstfsdtip

    ienfifptudts

    205CORTICAL SURFACE-BASED ANALYSIS II

    nalysis techniques to be applied to the functional andtructural properties of the cortical surface. The map-ing procedures we have presented have the advantagef being optimal with respect to a well-defined energyunctional that measures the amount of metric distor-ion and degree of folding of the transformed surface. Inonjunction with segmentation methods, as describedn the companion paper, these procedures allow theoutine use of surface-based visualization and analysisf functional and structural properties of the humanortex.

    8. APPENDIX

    .1. Numerical Integration

    The numerical integration scheme we use is a form ofultiscale line minimization that allows the surface tonfold robustly. Given an error functional Jt(x) at time, and its gradient =J with respect to the vertexositions, we will find the time step k which minimizes:

    Jt11(x) 5 Jt (x 1 k=J ). (11)

    ince the surfaces typically have large folds due to thenitial projection, simply setting k to a small value anderforming gradient descent can be very costly. In-tead, we search for the optimal k in a multiscaleanner which permits macroscopic changes in the

    urface at each time step. Specifically, we first con-train k to be in the range [kmin, kmax], where kmin andmax are given by the smallest (0.1 mm) and largest (20m) allowable vertex movements, respectively, dividedy the mean gradient magnitude. We then find theonstant k1 5 kmin 10l (l 5 0,1,2,. . .) which minimizes:

    t11 t l

    FIG. 11. Histogram of distance errors for the flattening (

    J (x) 5 J (x 1 kmin10 =J ), (12) g

    ith the search terminating when kmin 10l exceeds kmax.hat is, we sample Jt(x 1 k=J ) in the range [kmin, kmax]t order of magnitude intervals. Once we have deter-ined k1, and hence the proper scale to be searching for

    he optimal k, we then sample Jt(x 1 k=J ) around k1 at2 5 1.5*k1 and k3 5 0.5*k1. Next, we fit a quadratic tohese three values and compute the time step k4, whichould be optimal if the error surface were locallyuadratic, as is frequently the case. Finally, we choosehe time step k from the set [0 k1 k2 k3 k4], whichroduces the smallest value of Jt(x 1 k=J ), thus ensur-ng that J(x) is monotonically decreasing with time.

    .2. Integration Schedule

    As noted above, the surfaces frequently start witharge folds due to the projection process. Unfolding theurface in this state usually requires that the arealerm be large relative to the metric term in the energyunctional J(x) (that is, la : ld). Conversely, after theurface has unfolded we will minimize the metricistortions and have the areal term be small. Thus, inhe numerical integration we set the ratio la/ld to benitially large, and let it decrease as the integrationroceeds.Another factor to consider is that the distortion

    ntroduced by the projection usually results in a gradi-nt which is not locally coherent. This prevents theumerical integration scheme outlined above fromnding large times steps which would unfold the sur-

    ace in few iterations. One means of alleviating thisroblem would be to first unfold a decimated surfacehen interpolate to the full resolution and continuenfolding (Drury et al., 1996). In this approach theecimation must be done in a manner which respectshe local geometry of the surface. Instead, we take aimpler approach, and use a smoothed version of the

    ) and spherical transformation (right) of a typical surface.

    left

    radient in the numerical integration, with the degree

  • oTtcrt

    eahgffslgfwtse

    iistpflto

    arLgt

    A

    B

    C

    C

    C

    D

    D

    D

    D

    D

    D

    d

    D

    D

    D

    E

    F

    F

    G

    H

    H

    H

    K

    L

    L

    206 FISCHL, SERENO, AND DALE

    f smoothing decreasing as the integration proceeds.he smoothing of the gradient has an effect similar tohat of decimation: large regions of the surface moveoherently. Since the surface has no well-defined met-ic, we are forced to use iterative averaging to performhe smoothing, a time consuming process.

    Thus, the integration proceeds in epochs, where eachpoch has a fixed la/ld ratio. Typically we let la/ld startt 1000 and decrease by factors of 10 until the surfaceas converged, with 5 epochs usually being sufficient toenerate a near-optimal surface. We then set la/ld largeor one last epoch in order to smooth out any remainingolds in the surface (usually less than 0.05% of theurface is folded at this point). Within each epoch, weet the integration proceed using a fixed amount ofradient smoothing, until the decrease in the errorunctional asymptotes. After the error has asymptoted,e reduce the amount of smoothing and continue with

    he integration until the integration with the un-moothed gradient has asymptoted, which signals thend of an epoch.All the surfaces shown in this paper were generated

    n this manner, with each epoch beginning with 1024terations of gradient smoothing, and the amount ofmoothing decreasing by factors of 4 until the integra-ion using the unsmoothed gradient asymptoted. Thisrocedure typically requires on the order of 15 h toatten a full cortical hemisphere on a 266 MHz Pen-ium II, although slightly suboptimal surfaces can bebtained in about half that time.

    ACKNOWLEDGMENTS

    We are indebted to Eric Schwartz for many useful discussionsbout flattening. Thanks also go to Kevin Hall for testing andetesting the surface transformation routines, as well as to Arthuriu for his numerous suggestions. And finally, we express ourratitude to one of the anonymous reviewers for many suggestionshat helped make this a better paper.

    REFERENCES

    tkins, M. S., and Mackiewich, B. T. 1996. Automated Segmentationof the Brain in MRI. The 4th International Conference no Visualiza-tion in Biomedical Computing, Hamburg, Germany.ullmore, E., Brammer, M., Rouleau, G., Everitt, B., Simmons, A.,Sharma, T., Frangou, S., Murray, R., and Dunn, G. 1995. Computer-ized Brain tissue classification of magnetic resonance images: Anew approach to the problem of partial volume artifact. NeuroIm-age 2:133–147.arman, G. J., Drury, H. A., and Van Essen, D. C. 1995. Computa-tional Methods for reconstructing and unfolding the cerebralcortex. Cerebral Cortex.

    ollins, D. L., Neelin, P., Peters, T. M., and Evans, A. C. 1994. Data inStandardized Talairach Space. J. Comput. Assist. Tomogr. 18(2):292–205.ulham, J. C., Brandt, S. A., Cavanagh, P., Kanwisher, N. G., Dale,A. M., and Tootell, R. B. H. 1998. Cortical fMRI activation produced

    by attentive tracking of moving targets. J. Neurophysiol., in press.

    ale, A. M., Fischl, B., and Sereno, M. I. 1998. Cortical Surface-BasedAnalysis I: Segmentation and Surface Reconstruction. NeuroIm-age, in Press.ale, A. M., and Sereno, M. I. 1993. Improved localization of corticalactivity by combining EEG and MEG with MRI cortical surfacereconstruction: A linear approach. J. Cogn. Neurosci. 5(2):162–176.avatzikos, C. 1997. Spatial Transformation and Registration ofBrain Images Using Elastically Deformable Models. Comput.Vision Image Understand. 66(2):207–222.avatzikos, C., and Bryan, R. N. 1996. Using a Deformable SurfaceModel to Obtain a Shape Representation of the Cortex. IEEETrans. Med. Imag. 15:785–795.eYoe, E. A., Carman, G. J., Bandettini, P., Glickman, S., Wieser, J.,Cox, R., Miller, D., and Neitz, J. 1996. Mapping striate andextrastriate visual areas in human cerebral cortex. Proc. Natl.Acad. Sci. USA 93(6):2382–2386.ijkstra, E. W. 1959. A note on two problems in connexion withgraphs. Numerische Mathematik 1:269–271.

    o Carmo, M. 1976. Differential Geometry of Curves and Surfaces.Prentice-Hall, Englewood Cliffs, NJ.rury, H. A., Van Essen, D. C., Anderson, C. H., Lee, C. W., Coogan,T. A., and Lewis, J. W. 1996. Computerized Mappings of theCerebral Cortex: A Multiresolution Flattening Method and aSurface-Based Coordinate System. J. Cogn. Neurosci. 8(1):1–28.rury, H. A., Van Essen, D. C., Joshi, S. C., and Miller, M. I. 1996.Analysis and comparison of areal partitioning schemes usingtwo-dimensional fluid deformations. NeuroImage 3(S130).rury, H. A., Van Essen, D. C., Snyder, A. Z., Shulman, G. L.,Akbudak, E., Ollinger, J. M., Conturo, T. E., Raichle, M., andCorbetta, M. 1997. Warping fMRI activation patterns onto thevisible man atlas using fluid deformations of cortical flat maps.NeuroImage 5(S421).ngel, S. A., Glover, G. H., and Wandell, B. A. 1997. RetinotopicOrganization In Human Visual Cortex and the Spatial Precision OfFunctional MRI. Cerebral Cortex 7(2):181–192.

    elleman, D., and Van Essen, D. C. 1991. Distributed hierarchicalprocessing in primate cerebral cortex. Cerebral Cortex 1:1–47.

    ischl, B., Dale, A. M., Sereno, M. I., Tootell, R. B. H., and Rosen, B. R.1998. A coordinate system for the cortical surface. NeuroImage7(4):S740.auss, K. F. 1828. Disquisitiones generales circa superficies curvas.Comm. Soc. Gottingen Bd 6:1823–1827.adjikhani, N. K., Liu, A. K., Dale, A. M., Cavanagh, P., and Tootell,R. B. H. 1998. Retinotopy and color sensitivity in human visualcortical area V8. Nature Neurosci. 1:235–241.algren, E., Dale, A. M., Buckner, R. L., Marinkovic, K., Destrieux,C., Fischl, B., and Rosen, B. R. 1998. Cortical location of implicitrepetition effects in a size-judgment task to visually-presentedwords. NeuroImage, submitted.algren, E., Dale, A. M., Sereno, M. I., Tootell, R. B. H., Marinkovic,K., and Rosen, B. R. 1998. Location of human face-selective cortexwith respect to retinotopic areas. Hum. Brain Map., submitted.aas, J. H., and Krubitzer, L. A. 1991. The organization of extrastri-ate visual cortex. In Neuroanatomy of Visual Pathways and theirRetinotopic Organization (B. Dreher and S. R. Robinson, Eds.), Vol.3, pp. 302–359. Macmillan, London.

    iu, A. K., Belliveau, J. W., and Dale, A. M. 1998a. Spatiotemporalimaging of human brain activity using fMRI constrained MEGdata: Monte Carlo simulations. Proc. Natl. Acad. Sci. USA 95:8945–8950.

    iu, A. K., Dale, A. M., Ahlfors, S., Aronen, H., Huotilainen, M.,Ilmoniemi, R., Korvenoja, A., Simpson, G., Tootell, R. B. H.,Virtanen, J., and Belliveau, J. W. 1998b. Spatiotemporal Imaging

    of Motion Selective Areas in Human Cortex Using Combined fMRI

  • L

    M

    M

    M

    M

    M

    MP

    R

    S

    S

    S

    S

    S

    S

    S

    S

    S

    T

    T

    TT

    T

    T

    T

    T

    T

    T

    T

    T

    T

    T

    T

    T

    T

    V

    V

    W

    W

    W

    Z

    207CORTICAL SURFACE-BASED ANALYSIS II

    and MEG. Seventh Annual Meeting of the International Society forMagnetic Resonance in Medicine, Sydney, Australia. ISMRM,Berkeley, CA.

    iu, A. K., Dale, A. M., Sereno, M. I., Rosen, B. R., and Belliveau, J. W.1996. fMRI-Constrained Linear Estimation of Cortical Activityfrom MEG and EEG Measurements: A Model Study. Tenth Interna-tional Conference on Biomagnetism, Santa Fe, NM. SpringerVerlag.acDonald, D. 1998. A Method for Identifying Geometrically SimpleSurfaces from Three Dimensional Images. Montreal NeurologicalInstitute, McGill University, Montreal.oore, C. I. 1998. Some principles of somatosensory cortical organiza-tion in rats and humans. Brain Cogn. Sci. MIT, Boston.oore, C. I., Gehi, A., Corkin, S., Rosen, B. R., Stern, C., and Dale, A.1997. Basic and Fine Somatotopy in Human SI. The CognitiveNeuroscience Society, San Francisco, CA.oore, C. I., Stern, C. E., Corkin, S., Gray, A., Thelusma, F., Rosen,B. R., and Dale, A. M. 1998. Segregation of Multiple SomatosensoryMaps within the Human Postcentral Gyrus Using fMRI. NeuroIm-age 7(4).orel, A., and Kaas, J. J. 1992. Subdivisions and connections ofauditory cortex in owl monkeys. J. Comp. Neurol. 318:27–63.ortenson, M. E. 1997. Geometric Modeling. Wiley, New York.ress, W. H., Teukolsky, S. A., Vetterling, W. T., and Flannery, B. P.1994. Numerical Recipes in C. Cambridge Univ. Press, Cambridge.eppas, J. B., Niyogi, S., Dale, A. M., Sereno, M. I., and Tootell,R. B. H. 1997. Representation of motion boundaries in retinotopichuman visual cortical areas. Nature 388:175–179.

    chwartz, E. L. 1977. Spatial mapping in the primate sensoryprojection: Analytic structure and relevance to perception. Biol.Cybernet. 25:181–194.

    chwartz, E. L. 1980. Computational anatomy and functional archi-tecture of striate cortex: A spatial mapping approach to perceptualcoding. Vision Res. 20:645–669.

    chwartz, E. L., and Merker, B. 1986. Computer-Aided Neuro-anatomy: Differential Geometry of Cortical Surfaces and an Opti-mal Flattening Algorithm. IEEE Comp. Graph. Appl. 6:36–44.

    chwartz, E. L., Shaw, A., and Wolfson, E. 1989. A numerical solutionto the generalized mapmaker’s problem: Flattening nonconvexpolyhedral surfaces. IEEE Transactions on Pattern Analysis andMachine Intelligence 11:1005–1008.

    ereno, M. I., and Allman, J. M. 1991. Cortical visual areas inmammals. The Neural Basis of Visual Function (A. G. Leventhal,Ed.), pp. 160–172. Macmillan, London.

    ereno, M. I., Dale, A. M., Liu, A., and Tootell, R. B. H. 1996. A Sur-face-based Coordinate System for a Canonical Cortex. NeuroImage.

    ereno, M. I., Dale, A. M., Reppas, J. B., Kwong, K. K., Belliveau,J. W., Brady, T. J., Rosen, B. R., and Tootell, R. B. H. 1995. Bordersof multiple visual areas in humans revealed by functional magneticresonance imaging. Science 268(May 12):889–893.

    tepniewska, I., Preuss, T. M., and Kaas, J. H. 1993. Architectonics,somatotopic organization, and ipsilateral cortical connections ofthe primary motor area (M1) in owl monkeys. J. Comp. Neurol.330:238–271.

    zékely, G., Kelemen, A., Brechbühler, C., and Gerig, G. 1995.Segmentation of 3D Objects from MRI Volume Data Using Con-strained Elastic Deformations of Flexible Fourier Surface Models.First International Conference on Computer Vision, Virtual Real-ity and Robotics in Medicine, CVRMed’95, Nice, France.

    alairach, J., Szikla, G., Tournoux, P., Prosalentis, A., Bordas-Ferrier,M., Covello, L., Iacob, M., and Mempel, E. 1967. Atlas d’AnatomieStereotaxique du Telencephale. Masson, Paris.

    alairach, J., and Tournoux, P. 1988. Co-Planar Stereotaxic Atlas ofthe Human Brain. Thieme Medical Publishers, NY.

    alavage. 1998. Primary Auditory Cortex. HST. MIT, Boston.alavage, T., Ledden, P., MI, S., Rosen, B., and Dale, A. 1997a.Multiple phase-encoded tonotopic maps in human auditory cortex.NeuroImage 5(S8).

    alavage, T., Ledden, P., Sereno, M., Benson, R., Melcher, J., Rosen,B., and Dale, A. 1997b. Phase-encoded tonotopic maps in humanauditory cortex. ISMRM 5:6.

    alavage, T., Ledden, P., Sereno, M., Benson, R., and Rosen, B. 1996.Preliminary fMRI evidence for tonotopicity in human auditorycortex. NeuroImage 3(S355).

    alavage, T. M. 1998. Functional magnetic resonance imaging of thefrequency organization of human auditory cortex. Health Sci.Technol. Boston, MIT.

    eo, P. C., Sapiro, G., and Wandell, B. A. 1997. Creating ConnectedRepresentations of Cortical Gray Matter for Functional MRIVisualization. IEEE Trans. Med. Imag. 16(6):852–863.

    hompson, P., Schwartz, C., Lin, R. T., Khan, A. A., and Toga, A. W.1996. Three-dimensional statistical analysis of sulcal variability inthe human brain. J. Neurosci. 16(13):4261–4274.

    hompson, P. M., and Toga, A. W. 1996. A surface-based technique forwarping 3-dimensional images of the brain. IEEE Trans. Med.Imag. 15:1–16.

    ootell, R. B. H., Dale, A. M., Mendola, J. D., Reppas, J. B., andSereno, M. I. 1996a. fMRI analysis of human visual cortical areaV3A. NeuroImage 3:S358.

    ootell, R. B. H., Hadjikhani, N. K., Vanduffel, W., Liu, A. K.,Mendola, J. D., Sereno, M. I., and Dale, A. M. 1998a. Functionalanalysis of primary visual cortex (V1) in humans. Proc. Natl. Acad.Sci. USA 95:811–817.

    ootell, R. B. H., Mendola, J. D., Hadjikhani, N. K., Ledden, P. J., Liu,A. K., Reppas, J. B., Sereno, M. I., and Dale, A. M. 1997. Functionalanalysis of V3A and related areas in human visual cortex. J.Neurosci. 17:7060–7078.

    ootell, R. B. H., Mendola, J. D., Hadjikhani, N. K., Liu, A. K., andDale, A. M. 1998b. The representation of the ipsilateral visual fieldin human cerebral cortex. Proc. Natl. Acad. Sci. USA 95:818–824.

    ootell, R. B. H., Reppas, J. B., Dale, A. D., Look, R. B., Malach, R.,Jiang, H.-J., Brady, T. J., Rosen, B. R., and Belliveau, J. W. 1995a.Visual motion aftereffect in human cortical area MT/V5 revealed byfunctional magnetic resonance imaging. Nature 375:139–141.

    ootell, R. B. H., Reppas, J. B., Kwong, K. K., Malach, R., Born, R. T.,Brady, T. J., Rosen, B. R., and Belliveau, J. W. 1995b. Functionalanalysis of human MT and related visual cortical areas usingmagnetic resonance imaging. J. Neurosci. 15(14):3215–3230.

    ootell, R. H. B., Dale, A. M., Sereno, M. I., and Malach, R. M. 1996b.New images from human visual cortex. Trends Neurosci. 19:481–489.

    an Essen, D. C., and Drury, H. A. 1997. Structural and FunctionalAnalyses of Human Cerebral Cortex Using a Surface-Based Atlas.J. Neurosci. 17(18):7079–7102.

    an Essen, D. C., Drury, H. A., Joshi, S., and Miller, M. I. 1998.Functional and Structural Mapping of Human Cerebral Cortex:Solutions are in the Surfaces. Proc. Natl. Acad. Sci. USA.ells, W., Grimson, W., Kikinis, R., and Jolesz, F. 1996. AdaptiveSegmentation of MRI Data. IEEE Trans. Med. Imag. 15(4):429–442.ells, W., Kikinis, R., and Jolesz, F. A. 1994. Statistical IntensityCorrection and Segmentation of Magnetic Resonance Image Data.Proceedings of the Third Conference on Visualization in Biomedi-cal Computing VBC’94.olfson, E., and Schwartz, E. L. 1989. Computing minimal distanceson polyhedral surfaces. IEEE Trans. Pattern Anal. Machine Intel.11:1001–1005.

    illes, K., Armstrong, E., Schleicher, A., and Kretschmann, H.-J.1988. The human pattern of gyrification in the cerebral cortex.

    Anat. Embryol. 179:173–179.

    1. INTRODUCTION2. MAPPING OF THE CORTICAL SURFACE TO PARAMETERIZABLE SHAPESFIG. 1.

    3. SURFACE INFLATIONFIG. 2.FIG. 3.

    4. FLATTENINGFIG. 4.FIG. 5.

    5. SPHERICAL TRANSFORMATIONS AND A SURFACE-BASED COORDINATE SYSTEMFIG. 6.FIG. 7.

    6. ERROR ANALYSISFIG. 8.FIG. 9.FIG. 10.FIG. 11.

    7. CONCLUSION8. APPENDIXACKNOWLEDGMENTSREFERENCES