11
© BRAND X PICTURES T he availability of multiple views of a scene makes possible new and exciting applica- tions ranging from 3-D and free-viewpoint television to robust scene interpretation and object tracking. The hardware for multicamera systems is developing fast and is already being deployed for multimedia, security, and industrial applications. However, there are still some challenging issues in terms of processing, primarily due to the sheer amount of data involved when the number of cameras becomes very large. It is therefore a primordial point to understand how the information is structured and how to take advantage of the inherent redundancy that results when the cameras are looking at the same scene. This article provides insights on the nature of the data in multiview imaging systems, partic- ularly in terms of structure and coherence. Using this structure, we derive a multidimensional variational framework for the extraction of coherent regions and occlusion boundaries, which is an important issue in numerous multiview image processing applications such as view interpo- lation, compression, and scene understanding. SEEING IN SEVEN DIMENSIONS Our visual perception sense (i.e., our eyes) enables us to view the world in three dimensions. One might also say that time is a fourth dimension we are able to perceive. One way to under- stand why this is the case is to say that an eye captures two spatial dimensions describing where IEEE SIGNAL PROCESSING MAGAZINE [34] NOVEMBER 2007 1053-5888/07/$25.00©2007IEEE [ Jesse Berent and Pier Luigi Dragotti ] Plenoptic Manifolds [ Exploiting structure and coherence in multiview images ] Digital Object Identifier 10.1109/MSP.2007.905883

Plenoptic Manifolds - Imperial College London

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Plenoptic Manifolds - Imperial College London

©B

RA

ND

X P

ICT

UR

ES

The availability of multiple views of a scene makes possible new and exciting applica-tions ranging from 3-D and free-viewpoint television to robust scene interpretationand object tracking. The hardware for multicamera systems is developing fast and isalready being deployed for multimedia, security, and industrial applications. However,there are still some challenging issues in terms of processing, primarily due to the

sheer amount of data involved when the number of cameras becomes very large. It is therefore aprimordial point to understand how the information is structured and how to take advantage ofthe inherent redundancy that results when the cameras are looking at the same scene.

This article provides insights on the nature of the data in multiview imaging systems, partic-ularly in terms of structure and coherence. Using this structure, we derive a multidimensionalvariational framework for the extraction of coherent regions and occlusion boundaries, which isan important issue in numerous multiview image processing applications such as view interpo-lation, compression, and scene understanding.

SEEING IN SEVEN DIMENSIONSOur visual perception sense (i.e., our eyes) enables us to view the world in three dimensions.One might also say that time is a fourth dimension we are able to perceive. One way to under-stand why this is the case is to say that an eye captures two spatial dimensions describing where

IEEE SIGNAL PROCESSING MAGAZINE [34] NOVEMBER 2007 1053-5888/07/$25.00©2007IEEE

[Jesse Berent and Pier Luigi Dragotti]

Plenoptic Manifolds[Exploiting structure and coherence in multiview images]

Digital Object Identifier 10.1109/MSP.2007.905883

Page 2: Plenoptic Manifolds - Imperial College London

it is looking and another dimension for time. Our second eyeprovides the fourth dimension for the location of the viewingpoint. In the case of a camera array however, the number of“eyes” and their position is unlimited.

Studying the data in multiview camera systems from animage processing point of view means adding more degrees offreedom to the problem, and this leads to new difficulties, notthe least that the data have more dimensions than most of usare able to visualize. In fact, the number of dimensions goes upto seven when all the degrees of freedom are taken into account.Indeed, the visual information captured depends on the viewingposition (Vx, Vy, Vz), the viewing direction (θ, φ), the wave-length λ and the time t if dynamic scenes are considered. In [1],Adelson and Bergen gather all these parameters into a singlefunction P = P7(θ, φ, λ, t, Vx, Vy, Vz) called the plenoptic func-tion. Usually, it is represented with the cartesian coordinatesused in numerous computer vision and image processing algo-rithms. It therefore becomes

P = P7(x, y, λ, t, Vx, Vy, Vz) , (1)

where x and y are analogous to the coordinates on the image plane.It is far from trivial to deal with the seven dimensions of the

plenoptic function. However, there are some solutions to over-come this obstacle. First, several assumptions can be made toreduce the dimensionality. As we will see in the next section,these assumptions include dropping the wavelength, consideringstatic scenes, or constraining the camera locations. Second, theplenoptic function in all its parameterizations has a high degreeof regularity under the assumption of photoconsistency. Clearly,the analysis of multiview images calls for multidimensional sig-nal processing algorithms that take advantage of this inherentregularity for compression, interpolation, and interpretation.

In order to understand the regularity involved in plenopticdata, consider the more easily visualized caseof the video. Looking at the full space-timevolume such as the one illustrated in Figure3(a) reveals that an object or a layer movingaround in the scene carves out a 3-D volumeor an object tunnel [2]. Since this volume isconstructed with images of the same object,the information inside it is highly regular.Similarly, consider a dense set of multibase-line stereo images that are collated such thatthey form a 3-D data set (also known as theepipolar plane image (EPI) volume [3]). Again,looking at the whole volume of data, such asthat represented in Figure 3(b), shows thatEPI-tubes [4] are carved out by objects at dif-ferent depths in very much the same way as inthe video. The purpose of this article is toemphasize that, just like the tunnels carvedout by objects in space-time or the tubes inEPI volumes, hypervolumes are carved out inthe plenoptic function.

In an effort to generalize the notion and inspired by Adelsonand Bergen [1], we introduce, with a slight abuse of terminolo-gy, the concept of plenoptic manifolds. (See the section“Capturing the Plenoptic Function.”) With the term plenopticmanifold, we mean the hypervolume carved out by an object inthe plenoptic domain. Since these manifolds capture the coher-ence of the plenoptic function in all dimensions, their extractionis a very useful step in numerous multiview imaging applica-tions such as layer-based representations [5], [6], MPEG-4–likeobject-based coding [7], disparity-compensated and shape-adap-tive wavelet coding [8], and image-based rendering (IBR) [9],[10], especially in the case of occlusions and large depth varia-tions. Other applications include scene interpretation andunderstanding [11]. All these applications make it very attractiveto develop methods that are able to extract such manifolds.

In this article, we go through some common parameteriza-tions of the plenoptic function and recall the shape constraintsimposed on the plenoptic manifolds in some simple camerasetups. Then, we focus mainly on the light field parameteriza-tion [12] and derive a global multidimensional variationalframework based on [13], [14] for the extraction of these plenop-tic manifolds. Finally, we demonstrate some experimentalresults and applications in IBR.

CAPTURING THE PLENOPTIC FUNCTIONThe plenoptic function was introduced by Adelson and Bergen[1] in order to describe the visual information available fromany point space. It is characterized by seven dimensions, name-ly, the viewing position (Vx, Vy, Vz), the viewing direction(x, y), the time t and the wavelength λ. Usually the wavelengthis omitted by considering separate channels for color images orone channel for grayscale images. There are many differentways to capture the plenoptic function and most of the popularsensing devices, some of which are illustrated in Figure 1, do

[FIG1] Capturing the plenoptic function. From the still image camera to the videocamera or multiview imaging systems, all the sensing devices illustrated sample theplenoptic function with a varying number of degrees of freedom.

2-D (x,y)

4-D (x,y,Vx,Vy)

3-D (x,y,t)

5-D (x,y,Vx,Vy,Vz)

3-D (x,y,Vx)

(b)

(a)

(d)

(c) (e)

IEEE SIGNAL PROCESSING MAGAZINE [35] NOVEMBER 2007

Page 3: Plenoptic Manifolds - Imperial College London

not necessarily sample all the dimensions. The still image cam-era, for instance, fixes the viewing point and the time. Only the(x, y) dimensions remain. The video camera is able to captureimages at different times and therefore captures the (x, y, t)dimensions. Another case of a three-dimensional plenoptic func-tion can be obtained by giving one degree of freedom to the cam-era location such that (x, y, Vx) is sampled. Higher-dimensionalcases add more degrees of freedom to the viewing position suchthat (x, y, Vx, Vy) or even (x, y, Vx, Vy, Vz) can be captured.

PLENOPTIC TRAJECTORIES AND ASSOCIATED MANIFOLDSGiven a known camera setup, a point in space is projected ontothe images in a particular fashion dictated by the geometry ofthe array of cameras and the movement of the objects. The cap-tured plenoptic function therefore has a structure that dependson the camera setup sampling it. In this section, we illustratemore precisely the properties of the plenoptic function for somesimple camera setups starting from the most basic and going onto the higher-dimensional cases (see also the section “Capturingthe Plenoptic Function”).

Before going through some common representations of theplenoptic function and their associated plenoptic manifolds, welay out the assumptions made in this article. First, we assumethat the cameras follow the basic pinhole camera model as illus-trated in Figure 2, and we will use the cartesian coordinate sys-tem for the plenoptic function as in (1). Second, the wavelengthparameter λ is dropped by considering grayscale images orseparate red, green, and blue channels for color data. Finally, weassume that the scenes are made of opaque Lambertian surfaces.

SINGLE-VIEW CAMERASA single still image camera samples the plenoptic functionwhere the viewing position and the time are fixed (e.g., inVx = Vy = Vz = t = 0). Only the x and y dimensions remain,which are the image coordinates. The pinhole camera modelsays that points in the world coordinates �X = (X, Y, Z) aremapped onto the image plane (x, y) in the point where the line

connecting �X and the camera center intersects with the imageplane [15]. The focal length f measures the distance separatingthe camera center and the image plane. By using similar trian-gles, it can be shown that the mapping is given by

⎛⎝ X

YZ

⎞⎠ �→

(xy

)=

(X/ZY/Z

),

where we assume that the focal length f is unity and that theprincipal point is located at the origin. The extraction of coher-ent regions in this case can based on color or texture (i.e., strict-ly spatial coherence in the x and y dimensions), and, althoughthis problem is extremely interesting, it is not the point we wishto put forward in this article. Rather, we wish to portray thecoherence involved when several images of the same objects atdifferent locations or different times are available.

A single viewpoint imaging system can sample a 3-Dplenoptic function if it is able to capture the scene at differ-ent times. This is the case of the video or moving image. Thepoint in space �X is free to move in time and its mapping ontothe video data becomes

⎛⎝ X(t)

Y(t)Z(t)

⎞⎠ �→

⎛⎝ x

yt

⎞⎠ =

⎛⎝ X(t)/Z(t)

Y(t)/Z(t)t

⎞⎠

which is the parameterization of a trajectory in the 3D plenop-tic domain. Note that the intensity along the trajectory remainsfairly constant if the radiance of the point does not change intime. Furthermore, assuming the scene is made of movingobjects, neighboring points in space will generate similarneighboring trajectories in the video data. Hence, apart fromthe object boundaries, the information captured varies mainlyin a smooth fashion. This observation has motivated the seg-mentation of videos into coherent regions such as the onesundergoing similar motion. Recent methods for segmentation

and motion estimation haveshown that added robustness isachieved by considering thewhole space-time volume asopposed to one or a few consecu-tive frames (see [11] for a recentpresentation of space-time videoanalysis). The analysis of video asa 3-D function enables to imposecoherence throughout the stackof images and gain insights onlong term effects such as occlu-sions. In [2], Ristivojevic andKonrad show that tunnels arecarved out by objects in the dataas illustrated in Figure 3(a).Note that in the general case,the volumes do not have much

[FIG2] The pinhole camera model. Light rays from points in real-world coordinates �X = (X, Y, Z)

generate intensities on the image plane in the point (x, y) where the line connecting �X and thecamera center (Vx, Vy, Vz) intersects the image plane. The focal length f measures the distanceseparating the camera center and the image plane.

X

Y

Z (X,Y,Z)

Image Plane

Principal Axis (Vx,Vy,Vz)CameraCenter x

y

Z

X

f

f (X-Vx)/(Z-Vz)

Image Plane

(Vx,Vy,Vz)

(X,Y,Z)

x

Principal Axis

f

(a) (b)

IEEE SIGNAL PROCESSING MAGAZINE [36] NOVEMBER 2007

Page 4: Plenoptic Manifolds - Imperial College London

structure. Indeed, there is no real prior constraining the shapeof the tunnel carved out unless some assumptions are made onthe movement and the rigidity of the objects. Nevertheless, innatural videos, assuming a certain degree of smoothness andtemporal coherence is usually a valid assumption.

LINEAR MULTIVIEW CAMERA ARRAYSA linear camera array gives one degree of freedom to the posi-tion of the camera (e.g., in Vx). That is, parallax information isavailable. In this case, the plenopticfunction reduces to the Epipolarplane image (EPI) volume firstintroduced by Bolles et al. [3]. It canbe acquired either by translating acamera along a rail or by a linearcamera array (such as the one illus-trated in Figure 1). According to thepinhole camera model and assuming Vy = Vz = t = 0, pointsin real-world coordinates are mapped onto the EPI volume as afunction of Vx according to

⎛⎝

XYZ

⎞⎠ �→

⎛⎝

xyVx

⎞⎠ =

⎛⎝

X/Z − Vx/ZY/ZVx

⎞⎠ ,

where we notice that a point in space generates a line.Furthermore, the slope of the line is inversely proportionalto the depth of the point Z. Therefore, the data in this para-meterization, as opposed to the video, have a very particularstructure, which is noticeable in Figure 3(b). The occurrenceof occlusions, for example, is predictable since a line with alarger slope will always occlude a line with a smaller slope.This property follows naturally from the fact that points clos-

er to the image plane will occlude points that are furtheraway. The example illustrated in Figure 4 portrays this prop-erty with natural images.

In the original EPI analysis paper [3], this particular struc-ture is used to infer depth information in a scene by finding theslopes of the lines in the EPI volume. It is emphasized that bylooking at the problem in this manner, all the images are takeninto account simultaneously. However, the problem of densesegmentation was not dealt with. This problem was studied

much later by Criminisi et al. [4]where horizontal slices of the EPIvolume are analyzed in order togather lines with similar slopes. Thissegmentation generates coherentvolumes that are called EPI-tubesfor their obvious tube-like appear-ance (see Figure 3(b)). As opposed to

the method in [4], which analyses the data slice by slice, themethod presented in [13], [14], which we describe in moredetail below in “Extracting Plenoptic Manifolds in MultiviewData,” exploits coherence in the three dimensions; that is, thewhole stack of images is analyzed in a global manner.

The concept of EPI analysis is not necessarily restricted tothe case of cameras placed along a line, and has been extendedby Feldmann et al. [16] with image cube trajectories (ICTs). Theauthors show that other one-dimensional camera setups, suchas the circular case illustrated in Figure 3(c), generate particulartrajectories in the plenoptic domain and occlusion-compatibleorders can be defined. While the image cubes, EPI volumes, andvideos are all three-dimensional, more dimensions can be added.For example, the case where the sensors are video camerasalong a line leads to a four-dimensional parameterization thatincludes the time dimension.

[FIG3] Illustration of some 3-D plenoptic manifolds. (a) The volume carved out by a flat object in the space-time volume. (b) Thevolumes carved out by two flat objects in a linear camera array, where Vx denotes the position of the cameras along a line. (c) Thevolumes generated by two objects in the case of a circular camera array, where θ denotes the angle of the camera position around thecircle. Note that the shape of the volume in (a) depends on the movement of the objects which means it is not necessarily structured. Inboth the other cases (linear and circular still image camera arrays), the shape of the manifold is constrained by the camera setup andocclusion events can be predicted.

xy

t

(a)

xy

Vx

(b)

y

x

θ

(c)

IEEE SIGNAL PROCESSING MAGAZINE [37] NOVEMBER 2007

THIS ARTICLE PROVIDES INSIGHTSON THE NATURE OF THE DATA INMULTIVIEW IMAGING SYSTEMS,

PARTICULARLY IN TERMS OFSTRUCTURE AND COHERENCE.

Page 5: Plenoptic Manifolds - Imperial College London

PLANAR AND UNCONSTRAINED CAMERA ARRAYSPlanar camera arrays such as the one illustrated in Figure 1 givetwo degrees of freedom to camera locations (e.g., in Vx and Vy).The structure that governs the light rays in this case has beenvery well explored in the popular light field [12] or lumigraph[17] parameterizations introduced by Levoy and Hanrahan andby Gortler et al., respectively. In the planar case and assumingVz = t = 0, a point in space is mapped onto a four-dimensionaltrajectory according to

⎛⎝

XYZ

⎞⎠ �→

⎛⎜⎜⎝

xyVx

Vy

⎞⎟⎟⎠ =

⎛⎜⎜⎝

X/Z − Vx/ZY/Z − Vy/Z

Vx

Vy

⎞⎟⎟⎠ , (2)

where Vx and Vy are variable. By extension of the EPI volume, 4-Dhypervolumes are carved out by objects at different depths inthe scene and, just as in the EPI, their shape is constrained. In asimilar spirit to the plenoptic manifolds, the pop-up light fieldmethod [18] makes use of the segmentation of the data intocoherent regions (or layers) for view interpolation purposes. Thecontours of layers are semimanually extracted on one image andpropagated to the other views (i.e., the two other dimensions Vx

and Vy) by applying a user-defined depth map. By performingthe segmentation in this manner, the coherence of the layers isenforced in all the views.

Other planar camera setups include the dynamic light fieldsthat are 5-D since the time dimension is captured as well. Theconcentric mosaic [19] introduced by Shum and He also givestwo degrees of freedom to the camera locations. In that paper, a1-D camera (i.e., capturing slit images) is free to move along acircle with variable radius. The data are therefore parameterized

with three dimensions, namely the rotation angle, theradius, and the y axis of the image.

The case when the camera locations are uncon-strained (i.e., free to move in Vx, Vy, and Vz) gives riseto the 5-D plenoptic modeling function [20] firstintroduced by McMillan and Bishop. An even moregeneral case including the wavelength and the timedimension has been called the surface plenoptic func-tion [10] by Zhang and Chen. It contains six dimen-sions since it is assumed that radiance along a lightray does not change unless it is occluded. While it ismore difficult to visualize, a point in space generates aparticular trajectory in these parameterizations andobjects generate multidimensional hypervolumes.

PLENOPTIC MANIFOLDSAs we saw in the previous sections, points in spaceare mapped onto trajectories in the plenoptic func-tion. Objects that are made of neighboring points inspace are therefore mapped onto volumes (or moregenerally hypervolumes) that are made of neighbor-ing trajectories. This collection of trajectories gener-ates a multidimensional manifold M which we will

call plenoptic manifold. Note that the concept of plenopticmanifold shares many ideas with the coherent layers in [18]and the IBR objects in [7].

There are two important elements to retain from the para-meterizations described above. First, the multidimensionaltrajectories are constrained by the camera setup. This is illus-trated by the way points in space are mapped onto the plenop-tic domain. In the following, we will refer to this prior as thegeometry constraint. Second, there is a well-defined occlu-sion ordering. Points at different depths generate differenttrajectories and will intersect in the event of an occlusion.The study of these trajectories determines which point willocclude the other. We will refer to this prior as the occlusionconstraint. There are several benefits in considering theextraction of the whole manifold carved out by objects. Theprocedure enables a global vision of the problem and operateson the entire available data. That is, all the images are takeninto account simultaneously and the segmentation is consis-tent throughout all the views, which increases robustness. Aspointed out in [18], this consistency is also beneficial forapplications such as view interpolation.

EXTRACTING PLENOPTIC MANIFOLDS IN MULTIVIEW DATAThe extraction of coherent regions in plenoptic data is essen-tially a segmentation problem and closely related to that ofextracting layers. Semiautomatic schemes such as those pro-posed in [18], [7] can be very accurate but necessitate theuser's input. Other methods are unsupervised, in which casethe end result is usually obtained by initializing a set ofregions and using an iterative method that converges towardthe desired segmentation.

IEEE SIGNAL PROCESSING MAGAZINE [38] NOVEMBER 2007

[FIG4] The Epipolar Plane Image (EPI) volume. Cameras are constrained to aline resulting in a 3-D plenoptic function where x and y are the imagecoordinates and Vx denotes the position of the camera. Points in space areprojected onto lines where the slope of the line is inversely proportional tothe depth of the point. The volume illustrated is sliced in order to show theparticular structure of the data.

x

yVx

Vx

(a) (b)

Page 6: Plenoptic Manifolds - Imperial College London

Some layer extraction methods include k-means clustering [5]and linear subspace approaches [21]. Other common methodolo-gies use the Expectation-Maximization (EM) algorithm or energyminimization with graph-based methods such as Graph Cuts [22].These methods can be very efficient but are sometimes difficult toformulate. Alternative approaches such as active contours [23] arebased on the computation of the gradient of an energy functionaland use a steepest-descent methodology to converge toward aminimum. As we will see in the next section, we set the problemof extracting plenoptic manifolds in the variational framework,since it scales naturally to multidimensional signals and is flexiblein terms of functionals to minimize.

A VARIATIONAL APPROACHSince their introduction in the late 1980s, active contours (a.k.a.snakes) [23], along with active surfaces and variational frame-works in general, have enjoyed a huge popularity in numerouscomputer vision and image processing algorithms. Part of theirsuccess is due to the introduction of the level set method [24](see also box “The Level Set Method Basics”), which solved someof the issues such as numerical stability and topology depend-ence. Another key advantage lies in the formulation of the ener-gy minimization, which can be extended to any number ofdimensions [25], [26]. Applications of the 3D case include volu-metric data segmentation [22] and space-time video segmenta-

tion [11]. In [25], dynamic 3D modeling is performed by using a4D framework in order to impose coherence in the time dimen-sion as well. The trend is definitely going toward higher dimen-sional analysis, and these methods are capable of dealing with itin an elegant fashion. Since the plenoptic function has sevendimensions, there is a role to play for variational methods in theanalysis of multiview data.

The problem is formulated as follows: Start with a surface��(�σ) = (x (�σ), y(�σ), . . . ) ⊂ RN for points �σ ∈ RN−1 andmake it evolve until it converges to the boundaries of thesought-after region. In order to do this, the surface must bemade dependent on an evolution parameter τ such that ��(�σ, τ)

and assigned a speed function F(��(�σ, τ)) in order to evolveaccording to the following partial differential equation [29], [28]

∂ ��(�σ, τ)

∂τ= F (��(�σ, τ))�n�(�σ, τ), (3)

where ∂ ��/∂τ = �v� is the velocity, �n� is the outward normalvector and the initial condition ��(�σ, 0) is the starting pointdefined by the initialization of the algorithm. The velocity func-tion F will be chosen such that the surface converges toward thedesired segmentation when τ → ∞. In practice, F can be arbi-trarily designed or it can be derived from an energy functionalto minimize. In the latter case, the optimal speed in a steepestdescent sense can be determined.

Consider the problem of evolving a boundary�γ (s, τ ) =(x(s, τ ), y(s, τ )) ⊂ R

2 with a speed F in its outwardnormal direction �n. This evolution can be described with thefollowing partial differential equation

∂ �γ (s, τ )

∂τ= F ( �γ (s, τ ))�n(s, τ ),

where the initial condition is the curve in �γ (s, 0). A naturalway to implement the evolution is to discretize the curvewith a set of connected points and compute the displace-ment for each point according to the speed F. While thisapproach seems natural, it has some drawbacks. First, thepoints may move in such a way that they are closer and clos-er together or farther and farther away, which can lead tonumerical instabilities in the computation of derivatives.Second, a curve may be separated into two regions by thespeed function or, inversely, two curves could merge (i.e.,topological changes). These cases are difficult to deal with,since they require reparameterizing the curve. This proce-dure becomes even more problematic as the number ofdimensions increases. The level set method [24] addressesthese issues by embedding the curve �γ as the zero level of ahigher-dimensional surface z = φ(x, y, τ ) ⊂ R

3 and evolvingthe surface as opposed to the curve itself. In order to derivean evolution equation for the surface that will solve theoriginal problem, φ( �γ (s, τ ), τ ) = 0 needs to hold for all sand at all iterations τ . In other terms, the partial derivativesof φ( �γ (s, τ ), τ ) with respect to s and τ must be zero since

the function is constant. The application of the chain rulefor both cases leads to

{ ∂φ

∂τ+ �∇φ( �γ (s, τ ), τ ) · ∂ �γ

∂τ= 0

�∇φ

| �∇φ| = −�n

where �∇ is the gradient operator. Putting the two togetheralong with the definition of the speed of the original contour∂ �γ /∂τ = F �n enables us to write the level set equation

∂φ(x, y, τ )

∂τ= F(x, y)| �∇φ(x, y, τ )|.

The level set surface φ is free to expand, shrink, rise, andfall in order to generate the deformations of the originalcurve, and topological changes are naturally handled.Moreover, since this evolution equation is defined over thewhole domain, there is no need to parameterize the curvewith individual points. The numerical computations areperformed by using finite differences on a fixed cartesiangrid (thus solving the stability issue). Furthermore, themethodology extends naturally to evolving surfaces andgenerally hypersurfaces in any number of dimensions.These advantages, however, clearly come at the cost ofcomputational complexity, since the gradients and thespeed functions need to be computed for all the levels ofφ. Some solutions to reduce the number of computationshave been developed, such as the fast marching and nar-rowband methods [24].

THE LEVEL SET METHOD BASICS

IEEE SIGNAL PROCESSING MAGAZINE [39] NOVEMBER 2007

Page 7: Plenoptic Manifolds - Imperial College London

The energy functional to minimize is usually written as afunction of the surface

Etot(��) = Edata(��) + Esmooth(��), (4)

where the first term Edata measures the consistency of the seg-mentation with the data and the second term Esmooth ensuressmooth surfaces in order to computederivatives and reject outliers. Theseterms are also referred to as the exter-nal and internal energies [23], sincethe former is measured by the dataand the latter is derived from the prop-erties of the curve itself. Assumingthat the surface � = ∂M is theboundary of a region M, the data con-tribution can be written as a regioncompetition term [29]

Edata =∫M(τ)

din(�x)d �x +∫M(τ)

dout(�x)d �x,

where �x ∈ RN , M is the outside of M and din(�x) and dout(�x)are descriptors measuring the consistency with their respec-tive regions. The smoothness constraint is written as aboundary-based term

Esmooth =∫

∂M(τ)

μd �σ,

where μ is a constant weighting factor determining the influ-ence of Esmooth. Minimizing the total energy thus involves com-puting the derivative of the total functional with respect to τand evolving the boundary in a steepest descent fashion suchthat the energy converges to a minimum. It is possible to showthat the gradient of the energy Etot is given by [29], [26]

dE tot(τ)

dτ=

∫∂M(τ)

[din(�x) − dout(�x ) + μκ(�x )](�v� · �n�)d �σ,

where κ is the mean curvature of �� and · denotes the scalarproduct. From this equation, we deduce that the steepest descentof the energy yields the following partial differential equation:

�v� = [dout(�x ) − din(�x ) − μκ(�x )]�n�, (5)

where, by comparing (5) with (3), we now have an explicitform for the speed F of the evolving interface. Note that the so-called competition formulation is now clear. A point thebelongs to the inside of the sought-after region has a small d in

and a large dout, thus resulting in a positive speed. The pointwill therefore be incorporated. Inversely, a point belonging tothe outside has a small dout and a large d in, resulting in a neg-ative speed thus causing the point to be rejected. The curva-ture term helps to smooth the contour by straightening the

curve in places where the curvature is large. On the basis ofphotoconsistency, some common descriptors used to extractcoherent regions are related to the intensity differencesbetween two frames or views and in a more global way, thevariance along a plenoptic trajectory [29], [2].

While the equation in (5) driving the evolution of the hyper-surfaces is valid in the general case, it does not take into

account the geometry and occlusionconstraints that are inherent to theplenoptic function. In the followingsections, we study the cases of the EPIand the light field parameterizations(see “Linear Multiview Camera Arrays”and “Planar and UnconstrainedMultiview Camera Arrays,” above).These representations are popular,easier to visualize, and enable a clearimposition of the constraints. Someextensions to other camera setups are

possible. Recall that, under these parameterizations, points inspace are mapped onto lines in the plenoptic domain and theslopes of the lines are inversely proportional to the depth of thepoints. The next sections show how the evolution of the sur-faces or hypersurfaces can be modified in order to take intoaccount these constraints.

IMPOSING THE GEOMETRY CONSTRAINTConsider a scene with a single object or layer that carves out aplenoptic manifold M in a light field. The problem of extractingthe plenoptic manifold consists in finding the hypersurface�� ⊂ R4 that delimits the contour of the object on all the views.Let �γ (s, τ ) = (x0(s, τ ), y0(s, τ )) be the 2D contour defined bythe intersection of the hypersurface and the image plane inVx = Vy = 0. That is, it represents the contour of the object on asingle image. For simplicity, we assume that the depth map ofthe object is strictly frontoparallel, hence the depth Z = Z0 isconstant. According to (2), the boundary plenoptic manifoldunder these assumptions can be parameterized as

��(s, Vx, Vy, τ ) =

⎛⎜⎜⎝

x0(s, τ ) − Vx/Z0

y0(s, τ ) − Vy/Z0

Vx

Vy

⎞⎟⎟⎠

and is completely determined by the curve �γ (s, τ ) if we assumethat Z0 is known. It is therefore possible to propagate the posi-tion and the shape of �γ on all the other images. That is, theshape variations in the hypersurface �� are completely deter-mined by the shape variations of the curve �γ in the two-dimen-sional subspace. An explicit derivation of the normal andvelocity vectors shows that the projection of �v� · �n� onto thesubspace is equal to �vγ · �nγ and therefore does not depend onthe location of the camera (Vx, Vy). The intuition behind thisproperty is easily grasped. Given the fact that the depth map isconstant, the layer's contour will simply be a translated version

IEEE SIGNAL PROCESSING MAGAZINE [40] NOVEMBER 2007

THERE ARE STILL SOMECHALLENGING ISSUES INTERMS OF PROCESSING,PRIMARILY DUE TO THE

SHEER AMOUNT OF DATAINVOLVED WHEN THENUMBER OF CAMERASBECOMES VERY LARGE.

Page 8: Plenoptic Manifolds - Imperial College London

of itself on all the other images. Hence the gradient of the dataterm can be rewritten in the form

dEdata(τ)

dτ=

∫γ

[D in(s) − Dout(s)](�vγ · �nγ )ds,

where Din(s) and Dout(s) are the original descriptors integratedover the plenoptic trajectories. Therefore the optimal velocity vectordriving the evolution of the contour �γ in the subspace becomes

�vγ = [Dout(s) − Din(s) − μκ(s)]�nγ , (6)

where a smoothness term is added to insure that the 2Dcontour stays regular. In the case of regions that are notfrontoparallel, the same intuition holds; however, a weight-ing factor must be introduced in order to compensate forthe shape changes between theviews. In a more general sense, therelation between the normal veloci-ty of original boundary �� and thenormal velocity of the contour �γbecomes �v� · �n� = α(�σ)(�vγ · �nγ ) ,where α(�σ) is the weighting func-tion depending on the depth mapand the camera setup.

There are several advantages tousing evolution equation (6) as opposed to (5). First, it con-strains the shape of the manifold according to the camerasetup. Second, it is implemented as an active contour insteadof an active hypersurface, which reduces the computationalcomplexity. However, it comes at the cost of having tocompute the geometry of the object (i.e., the slope ofthe lines or in general the parameters of the plenoptictrajectories).

Inspired by stereo computer vision methods [30], wemodel the depth map as a linear combination of bicubicsplines and use classical nonlinear optimization methodsto find such parameters. The advantage of this model liesin the great variety of smooth depth maps that can beestimated. In addition, the bicubic splines can be forcedto model simplified geometry if an accurate depth recon-struction is not necessary. For instance, frontoparallelregions can be extracted by imposing that all the weightsof the linear combination are the same. The weights areestimated by minimizing the same functional (4) wherethe shape of the curves are kept constant. The overalloptimization is done by an iterative approach in whichthe contours are estimated while filxing the depth mapsand the depths are estimated while fixing the contoursuntil there is no significant decrease in energy.

IMPOSING THE OCCLUSION CONSTRAINTIn the previous section, we have seen how to constrainthe shape of the plenoptic manifolds according to thegeometry of the camera setup. The second main factor

to consider is occlusions and occlusion ordering. That is,sometimes the full manifold is not available, since it is occlud-ed in some of the views. To account for this case, we denote asMn the full manifold (as if it were not occluded) and M⊥

n asthe available manifold (i.e., excluding the occluded regions).Assuming the camera centers lie on a line or a plane (as in theEPI or light field parameterizations), this occlusion orderingstays constant throughout the views. Therefore, if the Mn’sare ordered from front (n = 1) to back (n = N), the occlusionconstraint [13], [14] can be written as

M⊥n = Mn ∩

n−1∑i=1

M⊥i , (7)

where the superscript perp denotes that the plenoptic manifoldhas been geometrically orthogonalized such that the occluding

manifolds carve through the back-ground ones (see Figure 5). A common-ly used approach to deal withocclusions in EPI analysis is to start byextracting the frontmost regions (orlines) and removing them from furtherconsideration [16], [4]. That is, wheneach Mn is extracted, the

∑n−1i=1 M⊥

iis known. While this approach isstraightforward, it has some drawbacks.

First, the extraction of occluded objects will depend on how wellthe occluding objects were extracted. Second, it does not enablea proper competition formulation since the background regionsare not known at the time of the extraction of the front ones.

[FIG5] The occlusion constraint with two 3-D plenoptic manifoldsunder the EPI parameterization. When put together, plenoptic manifoldsM1 and M2 become M⊥

1 and M⊥2 . The occlusion constraint

says that M⊥1 = M1 and M⊥

2 = M2 ∩ M⊥1

.

xy

Vx

y x

Vx Vx

y x

M2M1

M ⊥1

M ⊥2

(a) (b)

(c)

IEEE SIGNAL PROCESSING MAGAZINE [41] NOVEMBER 2007

THE PLENOPTIC FUNCTIONWAS INTRODUCED BY ADELSON

AND BERGEN TO DESCRIBETHE VISUAL INFORMATION

AVAILABLE FROM ANYPOINT SPACE.

Page 9: Plenoptic Manifolds - Imperial College London

An alternative approach consists insetting up a competition formulationbetween the front and backgroundregions and using an iterative approach.The energy in this case becomes

Edata =N∑

n=1

∫M⊥

n

dn(�x)d �x,

which can be minimized by iteratively evolving one M⊥n while

fixing the others. This leads to the evolution equation in (6)where all the other manifolds are gathered in M⊥

n . The idea isthat each iteration will contribute to minimize the total energy.Note that, as a result of the occlusion constraint in (7), evolvingM⊥

n will modify not only it but also all the manifolds it occludes(i.e., M⊥

i where i goes from n + 1 to N). Therefore these mani-folds will contribute to the dEdata/dτ and influence the compe-tition. That is, a manifold competes with all the regions it willocclude throughout all the views. However, a background mani-

fold evolving does not affect the shape ofthe foreground ones (i.e., M⊥

i where igoes from 1 to n − 1) and therefore theywill not contribute to dEdata/dτ andhence do not compete. An overview ofthe algorithm that shows how the twoconstraints are combined is presented inTable 1. In the next section, we showsome experimental results with partial

and total occlusions in order to demonstrate the benefits ofapplying this interplay between occluding and occluded regions.

APPLICATIONS IN IMAGE-BASED RENDERINGImage-based rendering is essentially the study of the sampling,interpolation, and extrapolation of the plenoptic function. It hasattracted a lot of attention recently thanks to its ability torecreate visually pleasing and realistic virtual viewpoints from aset of multiview images. This ability is without a doubt one ofthe most important issues when it comes to free-viewpoint visu-al media systems. In this section, we analyze some natural mul-tiview data sets with the variational framework presented aboveand discuss some of the applications related to IBR that benefitfrom the extraction of plenoptic manifolds.

As with all partial differential equation based—methods, ini-tialization is an important issue. For the results shown, theinitialization of the regions is performed in an unsupervisedfashion by computing local directions in the EPI images andmerging regions with similar slopes (i.e., similar depths). Inmore complicated camera setups, however, a more sophisticat-ed method such as a stereo disparity estimation algorithmcould be used. As in [2], [4], the descriptors used minimize thevariance along plenoptic trajectories. The depth model adoptedis piecewise constant. Using the competition formulation andimposing the geometry and occlusion constraints, the algo-rithm applied to the sequence in Figure 4 automaticallyextracts the plenoptic manifolds depicted in Figure 6. The dataconsist of 32 images containing a background and threeobjects, two of which are partially or totally occluded in numer-ous views. Note that occlusions and disocclusions are correctlycaptured. This is most obvious in the “cat” and “owl” layers.

The global nature of the segmentation scheme also suppress-es some of the discontinuities visible when individual slices ofthe EPI volume are analyzed, as in [4]. We refer to [13], [14] formore experimental results. The running time is approximately1000 seconds when the classical implementation of the level setmethod is used, although improvements of several orders ofmagnitude can be expected with faster implementations [24].

Several applications are possible once the multiview datahas been segmented into individual plenoptic manifolds. Forinstance, one of the typical issues is view interpolation. Thisproblem was studied by Chai et al. [9], using a classical signalprocessing framework, and was further generalized by Zhangand Chen in [10]. Both showed that, thanks to the particularstructure of the data, the band of the plenoptic function isapproximately bound by the minimum and maximum depths

[FIG6] Automatically extracted 3-D plenoptic manifolds fromthe EPI volume in Figure 4. (a) The first row illustrates theindividual volumes carved out by the different layers in thescene, and (b) the second row shows how the coherentregions fit together in the original data. Note how theforeground objects carve through the background ones.

(a)

(b)

[TABLE 1] OVERVIEW OF THE PLENOPTIC MANIFOLD EXTRACTION ALGORITHM.

STEP 1: INITIALIZE A SET OF PLENOPTIC MANIFOLDS STEP 2: ESTIMATE DEPTH PARAMETERSSTEP 3: UPDATE OCCLUSION ORDERING AND UPDATE MANIFOLDSSTEP 4: FOR EACH MANIFOLD

FIX THE OTHER MANIFOLDS COMPUTE SPEED FUNCTION WITH COMPETITION TERMSEVOLVE BOUNDARY

STEP 5: GO TO STEP 2 OR STOP WHEN THERE IS NO SIGNIFICANT DECREASE IN ENERGY

IEEE SIGNAL PROCESSING MAGAZINE [42] NOVEMBER 2007

IMAGE-BASED RENDERINGIS ESSENTIALLY THE STUDY

OF THE SAMPLING,INTERPOLATION, AND

EXTRAPOLATION OF THEPLENOPTIC FUNCTION.

Page 10: Plenoptic Manifolds - Imperial College London

in the scene. This fact makes it pos-sible to give an answer to the mini-mum sampling rate needed (i.e., thenumber of cameras) in order to havean aliasing-free rendering. However,when the scene has somewhat largedepth variations, this rate becomesvery large and the number of cam-eras required may be cumbersome.

In [18], the scene is segmented into coherent regions thatcan be individually rendered, free of aliasing. The method usesa coherence matting approach to blend the layers and alleviatesome of the errors caused by oversegmentation or underseg-mentation. In a similar spirit, we use the plenoptic manifoldsegmentation scheme to interpolate new viewpoints. Figure7(a) illustrates a linearly interpolated viewpoint, using a con-stant depth plane at the optimal depth [9]. The same renderedview in which three extracted plenoptic manifolds are individ-ually interpolated by using their estimateddepths is depicted in Figure 7(b) for compari-son. The blurring is greatly reduced while thenatural aspect of the images is maintained. Aspointed out in [9], the aliasing is reducedbecause the individual depth variations in eachplenoptic manifold are much smaller than inthe whole scene, hence fewer cameras are need-ed for an aliasing free rendering.

The extraction of the plenoptic manifoldsmay also provide a first step in scene under-standing. The fact that occlusions and objectboundaries are known, for instance, may be uti-lized in object recognition algorithms. Thisunderstanding also allows to manipulate themultiview data in a coherent fashion. Newscenes can be created by combining the plenop-tic manifolds in different ways. Forexample, occluded regions may beextrapolated by using the availableplenoptic trajectories and their intensi-ties. New images are generated wherebackground objects are disoccluded.Other plenoptic functions may be con-structed by inserting the plenopticmanifolds of external objects (capturedby a camera array or synthetically cre-ated) into the scene. Figure 8 illus-trates some of these manipulations.Despite the simplified depth modelused, the objects still show their origi-nal shapes in the rendered images (seethe duck’s beak, for instance). This isbecause the whole plenoptic manifoldsare recombined instead of using a layerrepresentation (i.e., alpha map, texture,and plane or motion parameters).

CONCLUSIONSIn this article, we looked into thecoherence of multiview images fromthe plenoptic function point of view,emphasizing that looking at the prob-lem from this angle provides a niceframework for studying the data in aglobal manner and imposing a coher-ent segmentation. Using this repre-

sentation, we looked into the nature of the function, such asthe structure, and suggested that in extension to the objecttunnels in videos and EPI-tubes in multibaseline stereo data,objects carve multidimensional hypervolumes in the plenopticfunction that we called plenoptic manifolds. Just as in thethree-dimensional cases, the manifolds contain highly regularinformation, since they are constructed with images of thesame objects. There is therefore clearly potential for robustanalysis and efficient representation.

[FIG8] “To duck or not to duck, or maybe to teapot?”: (a) illustrates some of theoriginal image taken from a multiview image sequence, (b) shows the same imageswith the manifold carved out by the duck removed and the background manifoldsextrapolated. Note that there are some incomplete regions, since they are nevervisible in the entire stack of images, and (c) illustrates the insertion of a syntheticmanifold generated by a teapot.

(a) (b) (c)

[FIG7] Image-based rendering by plenoptic manifold interpolation. (a) The linearlyinterpolated viewpoint using a single plane at the optimal constant depth. Notethat the image is blurred since there are not enough viewpoints for an aliasingfree rendering. (b) The rendered image obtained by interpolating the data in theextracted plenoptic manifolds on the basis of their individual estimated depths.

(a) (b)

IEEE SIGNAL PROCESSING MAGAZINE [43] NOVEMBER 2007

THE ANALYSIS OF MULTIVIEWIMAGES CALLS FOR

MULTIDIMENSIONAL SIGNALPROCESSING ALGORITHMS THAT

TAKE ADVANTAGE OF THEINHERENT REGULARITY.

Page 11: Plenoptic Manifolds - Imperial College London

We then looked into extracting these manifolds in simplecamera setups, using a variational framework that is flexiblein terms of the number of dimensions (depending on thecamera setup), the depth estimation, and the descriptors used.This flexibility is important for several reasons. First, thesame framework can be used for different camera setups.Second, some applications in image-based rendering do notalways necessitate an accurate depth reconstruction. Third,possible extensions to take into account large texturelessregions and specular effects, for instance, may be incorporat-ed into the descriptors. Future work may explore more com-plicated and unstructured camera setups as well as dynamicscenes and nonrigid objects. This would eventually lead to theextraction of seven dimensional manifolds.

ACKNOWLEDGMENTSThe authors wish to thank the Audiovisual CommunicationsLaboratory (LCAV) at the Swiss Federal Institute of Technology(EPFL) for providing the equipment to capture the data setsshown in this article. They also thank Yizhou Wang for helpingto prepare the images in Figure 3. Jesse Berent acknowledgesthe Engineering and Physical Sciences Research Council(EPSRC) for the doctoral training grant. The work presented isalso funded in part by the Royal Society.

AUTHORSJesse Berent ([email protected]) received his mas-ter’s degree in microengineering from the Swiss FederalInstitute of Technology (EPFL) in Lausanne, Switzerland in2004. He is currently with the Communications and SignalProcessing Group at Imperial College London where he isworking on his Ph.D. thesis. In 2006, he was a visiting researchstudent at the Audiovisual Communications Laboratory atEPFL, Switzerland. His research interests include multiviewimaging, biomedical image processing and sampling theory. Heis a Student Member of the IEEE.

Pier Luigi Dragotti ([email protected]) is a seniorlecturer in the Electrical and Electronic EngineeringDepartment at Imperial College, London. He received theLaurea in electrical engineering from the University Federico II,Naples, in 1997, the master’s degree in communications systemsfrom EPFL, Lausanne, in 1998, and the Ph.D. degree from EPFLin 2002. He was a visiting student at Stanford University and asummer researcher in the Mathematics of CommunicationsDepartment at Bell Labs, Lucent Technologies. His researchinterests include wavelet theory, image and video processing andcompression, joint source-channel coding, and signal process-ing for sensor networks. He is a Member of the IEEE.

REFERENCES[1] E.H. Adelson and J.R. Bergen, “The plenoptic function and the elements ofearly vision,” in Computational Models of Visual Processing. Cambridge, MA: MITPress, 1991, pp. 3–20.

[2] M. Ristivojevic and J. Konrad, “Space-time image sequence analysis: Objecttunnels and occlusion volumes,” IEEE Trans. Image Processing, vol. 15, pp. 364–376, Feb. 2006.

[3] R. Bolles, H.H. Baker, and D. Marimont, “Epipolar-plane image analysis: Anapproach to determining structure from motion,” Int. J. Comput. Vis., vol. 1, no. 1,pp. 7–55, 1987.

[4] A. Criminisi, S.B. Kang, R. Swaminathan, R. Szeliski, and P. Anandan, “Extractinglayers and analyzing their specular properties using epipolar-plane-image analysis,”Comp. Vis. Image Understanding, vol. 97, no. 1, pp. 51–85, Jan. 2005.

[5] J.Y.A. Wang and E.H. Adelson, “Representing moving images with layers,” IEEETrans. Image Processing, Special Issue on Image Sequence Compression, vol. 3,pp. 625–638, Sept. 1994.

[6] J. Shade, S. Gortler, L.W. He, and R. Szeliski, “Layered depth images,” in Proc.Comput. Graphics (SIGGRAPH ’98), 1998, pp. 231–242.

[7] Z.F. Gan, S.C. Chan, K.T. Ng, and H.Y. Shum, “An object-based approach toplenoptic videos,” in IEEE Int. Symp. Circuits and Syst., 2005, vol. 4, pp.3435–3438.

[8] C.L. Chang, X. Zhu, P. Ramanathan, and B. Girod, “Light field compressionusing disparity-compensated lifting and shape adaptation,” IEEE Trans. ImageProcessing, vol. 15, no. 4, pp. 793–806, Apr. 2006.

[9] J.-X. Chai, S.-C. Chan, H.-Y. Shum, and X. Tong, “Plenoptic sampling,” in Proc.Comput. Graphics (SIGGRAPH ’00), 2000, pp. 307–318.

[10] C. Zhang and T. Chen, “Spectral analysis for sampling image-based renderingdata,” IEEE Trans. Circuits Syst. Video Technol., vol. 13, pp. 1083–1050, Nov. 2003.

[11] J. Konrad, “Videopsy: Dissecting visual data in space-time,” IEEE Commun.Mag., vol. 45, no. 1, pp. 34–42, 2007.

12] M. Levoy and P. Hanrahan, “Light field rendering,” in Proc. Comput. Graphics(SIGGRAPH ’96), 1996, pp. 31–42.

[13] J. Berent and P.L. Dragotti, “Segmentation of epipolar plane image volumeswith occlusion and dissocclusion competition,” in Proc. IEEE Int. WorkshopMultimedia Signal Processing, Oct. 2006, pp. 182–185.

[14] “Unsupervised extraction of coherent regions for image based rendering,” inProc. British Machine Vision Conf., 2007.

[15] R. Hartley and A. Zisserman, Multiple View Geometry in Computer Vision,2nd ed. Cambridge, UK: Cambridge Univ. Press, 2004.

[16] I. Feldmann, P. Eisert, and P. Kauff, “Extension of epipolar image analysis tocircular camera movements,” in Proc. IEEE Int. Conf. Image Processing, 2003, pp.697–700.

[17] S.J. Gortler, R. Grzeszczuk, R. Szeliski, and M.F. Cohen, “The lumigraph,” inProc. Comput. Graphics (SIGGRAPH ’96), 1996, pp. 43–54.

[18] H.Y. Shum, J. Sun, S. Yamazaki, Y. Li, and C.K. Tang, “Pop-Up light field: Aninteractive image-based modeling and rendering system,” ACM Trans. Graph., vol.23, no. 2, pp. 143–162, Apr. 2004.

[19] H.-Y. Schum and L.-W. He, “Rendering with concentric mosaics,” in Proc.Comput. Graphics (SIGGRAPH ’99), 1999, pp. 299–306.

[20] L. McMillan and G. Bishop, “Plenoptic modeling: an image-based renderingsystem,” in Proc. Comput. Graphics (SIGGRAPH ’95), 1995, pp. 39–46.

[21] Q. Ke and T. Kanade, “A subspace approach to layer extraction,” in Proc. IEEEConf. Comput. Vision and Pattern Recognition, vol. 1, 2001, pp. 255–262.

[22] Y. Boykov, O. Veksler, and R. Zabih, “Fast approximate energy minimizationvia graph cuts,” IEEE Trans. on Pattern Analysis and Machine Intelligence, vol.23, no. 11, pp. 1222–1239, Nov. 2001 [26] pp. 419–430.

[23] M. Kass, A. Witkin, and D. Terzopoulos, “Snakes: Active contour models,” Int.J. Comput. Vis., vol. 1, no. 4, pp. 321–331, Jan. 1988.

[24] J. Sethian, Level Set Methods. Cambridge, UK: Cambridge Univ. Press, 1996.

[25] B. Goldluecke, I. Ihrke, C. Linz, and M. Magnor, “Weighted minimal hypersur-face reconstruction,” IEEE Trans. Pattern Anal. Machine Intell., vol. 29, no. 7, pp.1194–1208, July 2007.

[26] J. Solem and N. Overgaard, “A geometric formulation of gradient descent forvariational problems with moving surfaces,” in Proc. Int. Conf. Scale Space andPDE Methods Comput. Vision, 2005, pp. 419–430.

[27] V. Caselles, R. Kimmel, G. Sapiro, and C. Sbert, “Minimal surfaces basedobject segmentation,” IEEE Trans. Pattern Anal. Machine Intell., vol. 19, pp. 394–398, Apr. 1997.

[28] V. Caselles, R. Kimmel, and G. Sapiro, “Geodesic active contours,” Int. J.Comput. Vis., vol. 1, no. 22, pp. 61–79, 1997.

[29] S. Jehan-Besson, M. Barlaud, and G. Aubert, “DREAMS: Deformable regionsdriven by an Eulerian accurate minimization method for image and video segmen-tation,” Int. J. Comput. Vis., vol. 53, no. 1, pp. 45–70, 2003.

[30] M. Lin and C. Tomasi, “Surfaces with occlusions from layered stereo,” in Proc.IEEE Conf. Comput. Vision and Pattern Recognition, 2003, pp. 710–717.

IEEE SIGNAL PROCESSING MAGAZINE [44] NOVEMBER 2007

[SP]