Medical Image Analysis (1996) volume 1, number 2, pp 91–108
c Oxford University Press
Deformable models in medical image analysis: a survey
Tim McInerney∗ and Demetri Terzopoulos
Department of Computer Science, University of Toronto, Toronto, ON, Canada M5S 3H5
Abstract
This article surveys deformable models, a promising and vigorously researched computerassisted medical image analysis technique. Among model-based techniques, deformable models
offer a unique and powerful approach to image analysis that combines geometry, physics
and approximation theory. They have proven to be effective in segmenting, matching and
tracking anatomic structures by exploiting (bottom-up) constraints derived from the image
data together with (top-down) a priori knowledge about the location, size and shape of these
structures. Deformable models are capable of accommodating the significant variability of
biological structures over time and across different individuals. Furthermore, they support highly
intuitive interaction mechanisms that, when necessary, allow medical scientists and practitioners
to bring their expertise to bear on the model-based image interpretation task. This article reviews
the rapidly expanding body of work on the development and application of deformable models to
problems of fundamental importance in medical image analysis, including segmentation, shape
representation, matching and motion tracking.
Keywords: deformable models, matching, motion tracking, segmentation, shape modeling
Received December 4, 1995; revised March 15, 1996; accepted March 18, 1996
1. INTRODUCTION
The rapid development and proliferation of medical imaging
technologies is revolutionizing medicine. Medical imaging
allows scientists and physicians to glean potentially lifesaving information by peering non-invasively into the human
body. The role of medical imaging has expanded beyond the
simple visualization and inspection of anatomic structures.
It has become a tool for surgical planning and simulation,
intra-operative navigation, radiotherapy planning, and for
tracking the progress of disease. For example, ascertaining the
detailed shape and organization of anatomic structures enables
a surgeon preoperatively to plan an optimal approach to some
target structure. In radiotherapy, medical imaging allows the
delivery of a necrotic dose of radiation to a tumor with minimal
collateral damage to healthy tissue.
With medical imaging playing an increasingly prominent
role in the diagnosis and treatment of disease, the medical
image analysis community has become preoccupied with
the challenging problem of extracting—with the assistance
∗ Corresponding
author
(e-mail: tim@vis.toronto.edu)
of computers—clinically useful information about anatomic
structures imaged through CT, MR, PET, and other modalities
(Stytz et al., 1991; Robb, 1994; Ayache, 1995a, b; Bizais
et al., 1995). Although modern imaging devices provide
exceptional views of internal anatomy, the use of computers to
quantify and analyze the embedded structures with accuracy
and efficiency is limited. Accurate, repeatable, quantitative
data must be efficiently extracted in order to support the
spectrum of biomedical investigations and clinical activities
from diagnosis, to radiotherapy, to surgery.
Segmenting structures from medical images and reconstructing a compact geometric representation of these structures is difficult due to the sheer size of the datasets and the
complexity and variability of the anatomic shapes of interest.
Furthermore, the shortcomings typical of sampled data, such
as sampling artifacts, spatial aliasing and noise, may cause
the boundaries of structures to be indistinct and disconnected.
The challenge is to extract boundary elements belonging
to the same structure and integrate these elements into a
coherent and consistent model of the structure. Traditional
low-level image-processing techniques which consider only
local information can make incorrect assumptions during this
integration process and generate infeasible object boundaries.
92
T. McInerney and D. Terzopoulos
As a result, these model-free techniques usually require
considerable amounts of expert intervention. Furthermore,
the subsequent analysis and interpretation of the segmented
objects is hindered by the pixel- or voxel-level structure
representations generated by most image-processing operations.
This article surveys deformable models, a promising and
vigorously researched model-based approach to computerassisted medical image analysis. The widely recognized
potency of deformable models stems from their ability to
segment, match and track images of anatomic structures by
exploiting (bottom-up) constraints derived from the image
data together with (top-down) a priori knowledge about the
location, size and shape of these structures. Deformable
models are capable of accommodating the often significant
variability of biological structures over time and across different individuals. Furthermore, deformable models support
highly intuitive interaction mechanisms that allow medical
scientists and practitioners to bring their expertise to bear on
the model-based image interpretation task when necessary.
We will review the basic formulation of deformable models
and survey their application to fundamental medical image
analysis problems, including segmentation, shape representation, matching and motion tracking.
2. MATHEMATICAL FOUNDATIONS OF
DEFORMABLE MODELS
The mathematical foundations of deformable models represent the confluence of geometry, physics and approximation
theory. Geometry serves to represent object shape, physics
imposes constraints on how the shape may vary over space
and time, and optimal approximation theory provides the
formal underpinnings of mechanisms for fitting the models
to measured data.
Deformable model geometry usually permits broad shape
coverage by employing geometric representations that involve
many degrees of freedom, such as splines. The model remains
manageable, however, because the degrees of freedom are
generally not permitted to evolve independently, but are governed by physical principles that bestow intuitively meaningful
behavior upon the geometric substrate. The name ‘deformable
models’ stems primarily from the use of elasticity theory at
the physical level, generally within a Lagrangian dynamics
setting. The physical interpretation views deformable models
as elastic bodies which respond naturally to applied forces
and constraints. Typically, deformation energy functions
defined in terms of the geometric degrees of freedom are
associated with the deformable model. The energy grows
monotonically as the model deforms away from a specified
natural or ‘rest shape’ and often includes terms that constrain
Figure 1. Snake (white) attracted to cell membrane in an EM
photomicrograph (Carlbom et al., 1994).
the smoothness or symmetry of the model. In the Lagrangian
setting, the deformation energy gives rise to elastic forces
internal to the model. Taking a physics-based view of classical
optimal approximation, external potential energy functions
are defined in terms of the data of interest to which the
model is to be fitted. These potential energies give rise to
external forces which deform the model such that it fits the
data.
Deformable curve, surface and solid models gained popularity after they were proposed for use in computer vision (Terzopoulos et al., 1988) and computer graphics (Terzopoulos and
Fleischer, 1988) in the mid-1980s. Terzopoulos introduced the
theory of continuous (multidimensional) deformable models
in a Lagrangian dynamics setting (Terzopoulos, 1986a), based
on deformation energies in the form of (controlled-continuity)
generalized splines (Terzopoulos, 1986b). Ancestors of the
deformable models now in common use include Fischler
and Elshlager’s spring-loaded templates (1973) and Widrow’s
rubber mask technique (1973).
The deformable model that has attracted the most attention
to date is popularly known as ‘snakes’ (Kass et al., 1988).
Snakes or ‘deformable contour models’ represent a special
case of the general multidimensional deformable model theory
(Terzopoulos, 1986a). We will review their simple formulation in the remainder of this section in order to illustrate with
a concrete example the basic mathematical machinery that is
present in many deformable models.
Snakes are planar deformable contours that are useful
in several image analysis tasks. They are often used to
approximate the locations and shapes of object boundaries
in images based on the reasonable assumption that boundaries
are piecewise continuous or smooth (Figure 1). In its basic
form, the mathematical formulation of snakes draws from the
theory of optimal approximation involving functionals.
Deformable models in medical image analysis: a survey
93
2.1. Energy-minimizing deformable models
Geometrically, a snake is a parametric contour embedded in
the image plane (x, y) ∈ ℜ2 . The contour is represented
as v(s) = (x(s), y(s))⊤ , where x and y are the coordinate
functions and s ∈ [0, 1] is the parametric domain. The shape
of the contour subject to an image I (x, y) is dictated by the
functional
E(v) = S(v) + P(v).
(1)
The functional can be viewed as a representation of the energy
of the contour and the final shape of the contour corresponds
to the minimum of this energy. The first term of the functional,
S(v) =
1
0
2 2
2
∂ v
∂v
w1 (s) + w2 (s) 2 ds,
∂s
∂s
(2)
is the internal deformation energy. It characterizes the deformation of a stretchy, flexible contour. Two physical parameter
functions dictate the simulated physical characteristics of the
contour: w1 (s) controls the ‘tension’ of the contour while
w2 (s) controls its ‘rigidity’a . The second term in (1) couples
the snake to the image. Traditionally,
P(v) =
1
P(v(s)) ds,
(3)
0
where P(x, y) denotes a scalar potential function defined
on the image plane. To apply snakes to images, external
potentials are designed whose local minima coincide with
intensity extrema, edges and other image features of interest.
For example, the contour will be attracted to intensity edges
in an image I (x, y) by choosing a potential P(x, y) =
−c|∇[G σ ∗ I (x, y)]|, where c controls the magnitude of the
potential, ∇ is the gradient operator, and G σ ∗ I denotes the
image convolved with a (Gaussian) smoothing filter whose
characteristic width σ controls the spatial extent of the local
minima of P.
In accordance with the calculus of variations, the contour
v(s) which minimizes the energy E(v) must satisfy the Euler–
Lagrange equation
∂
∂ 2v
∂v
∂2
−
w1
+ 2 w2 2 + ∇ P(v(s, t)) = 0. (4)
∂s
∂s
∂s
∂s
This vector-valued partial differential equation expresses the
balance of internal and external forces when the contour rests
a The
values of the non-negative functions w1 (s) and w2 (s) determine the
extent to which the snake can stretch or bend at any point s on the snake. For
example, increasing the magnitude of w1 (s) increases the ‘tension’ and tends
to eliminate extraneous loops and ripples by reducing the length of the snake.
Increasing w2 (s) increases the bending ‘rigidity’ of the snake and tends to
make the snake smoother and less flexible. Setting the value of one or both of
these functions to zero at a point s permits discontinuities in the contour at s.
Figure 2. Snake deforming towards high gradients in a processed
cardiac image, influenced by ‘pin’ points and an interactive ‘spring’
which pulls the contour towards an edge (McInerney and Terzopoulos, 1995a).
at equilibrium. The first two terms represent the internal
stretching and bending forces respectively, while the third
term represents the external forces that couple the snake to
the image data. The usual approach to solving (4) is through
the application of numerical algorithms (section 2.3).
2.2. Dynamic deformable models
While it is natural to view energy minimization as a static
problem, a potent approach to computing the local minima of
a functional such as (1) is to construct a dynamical system
that is governed by the functional and allow the system to
evolve to equilibrium. The system may be constructed by
applying the principles of Lagrangian mechanics. This leads
to dynamic deformable models that unify the description of
shape and motion, making it possible to quantify not just
static shape, but also shape evolution through time. Dynamic
models are valuable for medical image analysis, since most
anatomical structures are deformable and continually undergo
non-rigid motion in vivo. Moreover, dynamic models exhibit
intuitively meaningful physical behaviors, making their evolution amenable to interactive guidance from a user (Figure 2).
A simple example is a dynamic snake which can be
represented by introducing a time-varying contour v(s, t) =
(x(s, t), y(s, t))⊤ with a mass density µ(s) and a damping
density γ (s). The Lagrange equations of motion for a snake
with the internal energy given by (2) and external energy given
by (3) is
∂ 2v
∂2
∂ 2v
∂v ∂
∂v
µ 2 +γ −
w1
+ 2 w2 2 = −∇ P(v(s, t)).
∂t
∂t ∂s
∂s
∂s
∂s
(5)
94
T. McInerney and D. Terzopoulos
The first two terms on the left-hand side of this partial
differential equation represent inertial and damping forces.
Referring to (4), the remaining terms represent the internal
stretching and bending forces, while the right-hand side
represents the external forces. Equilibrium is achieved when
the internal and external forces balance and the contour
comes to rest (i.e. ∂v/∂t = ∂ 2 v/∂t 2 = 0), which yields the
equilibrium condition (4).
2.3. Discretization and numerical simulation
In order to compute a minimum energy solution numerically,
it is necessary to discretize the energy E(v). The usual
approach is to represent the continuous geometric model v
in terms of linear combinations of local-support or globalsupport basis functions. Finite elements (Zienkiewicz and
Taylor, 1989), finite differences (Press et al., 1992) and
geometric splines (Farin, 1993) are local representation methods, whereas Fourier bases (Ballard and Brown, 1982) are
global representation methods. The continuous model v(s) is
represented in discrete form by a vector u of shape parameters
associated with the basis functions. The discrete form of
energies such as E(v) for the snake may be written as
E(u) = 21 u⊤ Ku + P(u)
(6)
where K is called the stiffness matrix, and P(u) is the discrete
version of the external potential. The minimum energy
solution results from setting the gradient of (6) to 0, which
is equivalent to solving the set of algebraic equations
Ku = −∇P = f
(7)
where f is the generalized external force vector.
The discretized version of the Lagrangian dynamics equation (5) may be written as a set of second-order ordinary
differential equations for u(t):
Mü + Cu̇ + Ku = f ,
(8)
where M is the mass matrix and C is a damping matrix. The
time derivatives in (5) are approximated by finite differences
and explicit or implicit numerical time-integration methods are
applied to simulate the resulting system of ordinary differential
equations in the shape parameters u.
2.4. Probabilistic deformable models
An alternative view of deformable models emerges from
casting the model fitting process in a probabilistic framework.
This permits the incorporation of prior model and sensor
model characteristics in terms of probability distributions.
The probabilistic framework also provides a measure of the
uncertainty of the estimated shape parameters after the model
is fitted to the image data (Szeliski, 1990).
Let u represent the deformable model shape parameters
with a prior probability p(u) on the parameters. Let p(I |u)
be the imaging (sensor) model—the probability of producing
an image I given a model u. Bayes’ theorem
p(u|I ) =
p(I |u) p(u)
p(I )
(9)
expresses the posterior probability p(u|I ) of a model given
the image, in terms of the imaging model and the prior
probabilities of model and image.
It is easy to convert the internal energy measure (2) of
the deformable model into a prior distribution over expected
shapes, with lower energy shapes being the more likely. This
is achieved using a Boltzmann (or Gibbs) distribution of the
form
1
exp(−S(u)),
(10)
p(u) =
Zs
where S(u) is the discretized version of S(v) in (2) and Z s is a
normalizing constant (called the partition function). This prior
model is then combined with a simple sensor model based on
linear measurements with Gaussian noise
p(I |u) =
1
exp(−P(u)),
ZI
(11)
where P(u) is a discrete version of the potential P(v) in (3),
which is a function of the image I (x, y).
Models may be fitted by finding u which locally maximize
p(u|I ) in (9). This is known as the maximum a posteriori
solution. With the above construction, it yields the same result
as minimizing (1), the energy configuration of the deformable
model given the image.
The probabilistic framework can be extended by assuming
a time-varying prior model, or system model, in conjunction
with the sensor model, resulting in a Kalman filter. The
system model describes the expected evolution of the shape
parameters u over time. If the equations of motion of the
physical snakes model (8) are employed as the system model,
the result is a sequential estimation algorithm known as
‘Kalman snakes’ (Terzopoulos and Szeliski, 1992).
3. MEDICAL IMAGE ANALYSIS WITH
DEFORMABLE MODELS
Although originally developed for application to problems
in computer vision and computer graphics, the potential of
deformable models for use in medical image analysis has been
quickly realized. They have been applied to images generated
by imaging modalities as varied as X-ray, CT, angiography,
MR and ultrasound. Two-dimensional and three-dimensional
deformable models have been used to segment, visualize,
Deformable models in medical image analysis: a survey
(a)
(b)
(c)
(d)
95
(e)
(f)
Figure 3. (a) Intensity CT image slice of LV. (b) Edge detected image. (c) Initial snake. (d)–(f) Snake deforming towards LV boundary, driven
by ‘inflation’ force. (McInerney and Terzopoulos, 1995a).
track and quantify a variety of anatomic structures ranging
in scale from the macroscopic to the microscopic. These
include the brain, heart, face, cerebral, coronary and retinal
arteries, kidney, lungs, stomach, liver, skull, vertebra, objects
such as brain tumors, a fetus, and even cellular structures such
as neurons and chromosomes. Deformable models have been
used to track the non-rigid motion of the heart, the growing tip
of a neurite and the motion of erythrocytes. They have been
used to locate structures in the brain, and to register images of
the retina, vertebra and neuronal tissue.
In the following sections, we review and discuss the application of deformable models to medical image interpretation
tasks including segmentation, matching and motion analysis.
3.1. Image segmentation with deformable curves
The segmentation of anatomic structures—the partitioning of
the original set of image points into subsets corresponding
to the structures—is an essential first stage of most medical
image analysis tasks, such as registration, labeling and
motion tracking. These tasks require anatomic structures
in the original image to be reduced to a compact, analytic
representation of their shapes. Performing this segmentation
manually is extremely labor-intensive and time-consuming. A
primary example is the segmentation of the heart, especially
the left ventricle (LV), from cardiac imagery. Segmentation
of the left ventricle is a prerequisite for computing diagnostic
information such as ejection-fraction ratio, ventricular volume
ratio, heart output, and for wall motion analysis which provides information on wall thickening etc. (Singh et al., 1993).
Most clinical segmentation is currently performed using
manual slice editing. In this scenario, a skilled operator, using
a computer mouse or trackball, manually traces the region of
interest on each slice of an image volume. Manual slice editing
suffers from several drawbacks. These include the difficulty in
achieving reproducible results, operator bias, forcing the operator to view each 2-D slice separately to deduce and measure
the shape and volume of 3-D structures, and operator fatigue.
Segmentation using traditional low-level image-processing
techniques, such as region growing, edge detection and
mathematical morphology operations, also requires consid-
erable amounts of expert interactive guidance. Furthermore,
automating these model-free approaches is difficult because
of the shape complexity and variability within and across
individuals. In general, the underconstrained nature of the
segmentation problem limits the efficacy of approaches that
consider local information only. Noise and other image artifacts can cause incorrect regions or boundary discontinuities
in objects recovered by these methods.
A deformable-model-based segmentation scheme, used in
concert with image preprocessing, can overcome many of
the limitations of manual slice editing and traditional imageprocessing techniques. These connected and continuous geometric models consider an object boundary as a whole and can
make use of a priori knowledge of object shape to constrain the
segmentation problem. The inherent continuity and smoothness of the models can compensate for noise, gaps and other
irregularities in object boundaries. Furthermore, the parametric representations of the models provide a compact, analytical
description of object shape. These properties lead to a robust
and elegant technique for linking sparse or noisy local image
features into a coherent and consistent model of the object.
Among the first and primary uses of deformable models
in medical image analysis was the application of deformable
contour models, such as snakes (Kass et al., 1988), to segment
structures in 2-D images (Berger, 1990; Cohen, 1991; Ueda
and Mase, 1992; Rougon and Prêteux, 1993; Cohen and
Cohen, 1993; Leitner and Cinquin, 1993; Carlbom et al., 1994;
Gupta et al., 1994; Lobregt and Viergever, 1995; Davatzikos
and Prince, 1995). Typically users initialized a deformable
model near the object of interest (Figure 3) and allowed it
to deform into place. Users could then use the interactive
capabilities of these models and manually fine-tune them.
Furthermore, once the user is satisfied with the result on
an initial image slice, the fitted contour model may then be
used as the initial boundary approximation for neighboring
slices. These models are then deformed into place and again
propagated until all slices have been processed. The resulting
sequence of 2-D contours can then be connected to form a
continuous 3-D surface model (Lin and Chen, 1989; Chang
et al., 1991; Cohen, 1991; Cohen and Cohen, 1993).
96
T. McInerney and D. Terzopoulos
Figure 4. Image sequence of a clipped angiogram of a retina showing an automatically subdividing snake flowing and branching along a vessel
(McInerney and Terzopoulos, 1995b).
(a)
(b)
(c)
(d)
Figure 5. Segmentation of a cross sectional image of a human vertebra phantom with a topologically adaptable snake (McInerney and
Terzopoulos, 1995b). The snake begins as a single closed curve and becomes three closed curves.
The application of snakes and other similar deformable
contour models to extract regions of interest is, however, not
without limitations. For example, snakes were designed as
interactive models. In non-interactive applications, they must
be initialized close to the structure of interest to guarantee
good performance. The internal energy constraints of snakes
can limit their geometric flexibility and prevent a snake from
representing long tube-like shapes or shapes with significant
protrusions or bifurcations. Furthermore, the topology of the
structure of interest must be known in advance since classical
deformable contour models are parametric and are incapable
of topological transformations without additional machinery.
Various methods have been proposed to improve and further automate the deformable contour segmentation process.
Cohen and Cohen (1993) used an internal ‘inflation’ force
to expand a snakes model past spurious edges towards the
real edges of the structure, making the snake less sensitive
to initial conditions [inflation forces were also employed in
Terzopoulos et al. (1988)]. Amini et al. (1990) used dynamic
programming to carry out a more extensive search for global
minima. Poon et al. (1994) and Grzeszczuk and Levin
(1994) minimized the energy of active contour models using
simulated annealing which is known to give global solutions
and allows the incorporation of non-differentiable constraints.
Poon et al. (1994) also used a discriminant function
to incorporate region-based image features into the image
forces of their active contour model. The discriminant
function allows the inclusion of additional image features
in the segmentation and serves as a constraint for global
segmentation consistency (i.e. every image pixel contributes
to the discriminant function). The result is a more robust
energy functional and a much better tolerance to deviation of
the initial guess from the true boundaries. Others researchers
(Rougon and Prêteux, 1991; Herlin et al., 1992; Chakraborty
et al., 1994; Gauch et al., 1994; Chakraborty and Duncan,
1995; Mangin et al., 1995) have also integrated region-based
information into deformable contour models in an attempt to
decrease sensitivity to insignificant edges and initial model
placement.
Recently, several researchers (Leitner and Cinquin, 1991;
Caselles, et al., 1993, 1995; Whitaker, 1994; Malladi et al.,
1995; Mc Inerney and Terzopoulos, 1995b; Sapiro et al., 1995)
have been developing topology-independent shape modeling
schemes that allow a deformable contour or surface model
not only to represent long tube-like shapes or shapes with
bifurcations (Figure 4), but also dynamically to sense and
change its topology (Figure 5).
3.2. Volume image segmentation with deformable surfaces
Segmenting 3-D image volumes slice by slice, either manually
or by applying 2-D contour models, is a laborious process
and requires a post-processing step to connect the sequence
of 2-D contours into a continuous surface. Furthermore, the
resulting surface reconstruction can contain inconsistencies or
show rings or bands. The use of a true 3-D deformable surface
model, on the other hand, can result in a faster, more robust
segmentation technique which ensures a globally smooth and
coherent surface between image slices. Deformable surface
models in 3-D were first used in computer vision (Terzopoulos
et al., 1988). Many researchers have since explored the use
of deformable surface models for segmenting structures in
Deformable models in medical image analysis: a survey
(a)
97
(b)
Figure 6. (a) Deformable ‘balloon’ model embedded in volume image deforming towards LV edges. (b) Reconstruction of LV (McInerney and
Terzopoulos, 1995a).
medical image volumes. Miller (1991) constructed a polygonal approximation to a sphere and geometrically deformed
this ‘balloon’ model until the balloon surface conforms to the
object surface in 3-D CT data. The segmentation process
is formulated as the minimization of a cost function where
the desired behavior of the balloon model is determined by a
local cost function associated with each model vertex. The
cost function is a weighted sum of three terms: a deformation
potential that ‘expands’ the model vertices towards the object
boundary, an image term that identifies features such as edges
and opposes the balloon expansion, and a term that maintains
the topology of the model by constraining each vertex to
remain close to the centroid of its neighbors.
Cohen and Cohen (1992b; 1993) and McInerney and
Terzopoulos (1995a) used finite-element and physics-based
techniques to implement an elastically deformable cylinder
and sphere, respectively. The models are used to segment
the inner wall of the LV of the heart from MR or CT image
volumes (Figure 6). These deformable surfaces are based on a
thin-plate under tension surface spline, the higher dimensional
generalization of Equation (2), which controls and constrains
the stretching and bending of the surface. The models are
fitted to data dynamically by integrating Lagrangian equations
of motion through time in order to adjust the deformational
degrees of freedom. Furthermore, the finite-element method is
used to represent the models as a continuous surface in the form
of weighted sums of local polynomial basis functions. Unlike
Miller’s (1991) polygonal model, the finite element method
provides an analytic surface representation and the use of highorder polynomials means that fewer elements are required to
represent an object accurately. Pentland and Sclaroff (1991)
and Nastar and Ayache (1993a) also developed physics-based
Figure 7. The result of matching a labeled deformable atlas to a
morphologically preprocessed MR image of the brain (Sandor and
Leahy, 1995).
models but used a reduced modal basis for the finite elements
(see section 3.5).
Staib and Duncan (1992b) described a 3-D surface model
used for geometric surface matching to 3-D medical image
data. The model uses a Fourier parameterization which
decomposes the surface into a weighted sum of sinusoidal
basis functions. Several different surface types are developed
such as tori, open surfaces, closed surfaces and tubes. Surface
finding is formulated as an optimization problem using gradient ascent which attracts the surface to strong image gradients
98
T. McInerney and D. Terzopoulos
in the vicinity of the model. An advantage of the Fourier
parameterization is that it allows a wide variety of smooth
surfaces to be described with a small number of parameters.
That is, a Fourier representation expresses a function in terms
of an orthonormal basis and higher indexed basis functions
in the sum represent higher spatial variation. Therefore, the
series can be truncated and still represent relatively smooth
objects accurately.
In a different approach, Szeliski et al. (1993) used a
dynamic, self-organizing oriented particle system to model
surfaces of objects. The oriented particles, which can be
visualized as small, flat disks, evolve according to Newtonian
mechanics and interact through external and interparticle
forces. The external forces attract the particles to the data
while interparticle forces attempt to group the particles into
a coherent surface. The particles can reconstruct objects
with complex shapes and topologies by ‘flowing’ over the
data, extracting and conforming to meaningful surfaces. A
triangulation is then performed which connects the particles
into a continuous global model that is consistent with the
inferred object surface.
Other notable work involving 3-D deformable surface
models and medical image applications can be found in
Delingette et al. (1992), Whitaker (1994), Tek and Kimia
(1995), Davatzikos and Bryan (1995) as well as several models
described in the following sections.
3.3. Incorporating a priori knowledge
In medical images, the general shape, location and orientation
of objects is known and this knowledge may be incorporated
into the deformable model in the form of initial conditions,
data constraints, constraints on the model shape parameters, or
into the model fitting procedure. The use of implicit or explicit
anatomical knowledge to guide shape recovery is especially
important for robust automatic interpretation of medical
images. For automatic interpretation, it is essential to have
a model that not only describes the size, shape, location and
orientation of the target object but that also permits expected
variations in these characteristics. Automatic interpretation
of medical images can relieve clinicians from the laborintensive aspects of their work while increasing the accuracy,
consistency and reproducibility of the interpretations.
A number of researchers have incorporated knowledge of
object shape into deformable models by using deformable
shape templates. These models usually use ‘hand-crafted’
global shape parameters to embody a priori knowledge of
expected shape and shape variation of the structures and have
been used successfully for many applications of automatic
image interpretation. The idea of deformable templates can
be traced back to the early work on spring-loaded templates
by Fischler and Elshlager (1973). An excellent example
in computer vision is the work of Yuille et al. (1992) who
constructed deformable templates for detecting and describing
features of faces, such as the eye. In medical image analysis,
Lipson et al. (1990) noted that axial cross sectional images
of the spine yield approximately elliptical vertebral contours
and consequently extracted the contours using a deformable
ellipsoidal template.
Deformable models based on superquadrics are another
example of deformable shape templates that are gaining in
popularity in medical image research. Superquadrics contain
a small number of intuitive global shape parameters that can
be tailored to the average shape of a target anatomic structure.
Furthermore, the global parameters can often be coupled
with local shape parameters such as splines resulting in a
powerful shape representation scheme. For example, Metaxas
and Terzopoulos (1993) employed a dynamic deformable
superquadric model (Terzopoulos and Metaxas, 1991) to
reconstruct and track human limbs from 3-D biokinetic
data. Their models can deform both locally and globally by
incorporating the global shape parameters of a superellipsoid
with the local degrees of freedom of a membrane spline in
a Lagrangian dynamics formulation. The global parameters
efficiently capture the gross shape features of the data, while
the local deformation parameters reconstruct the fine details
of complex shapes. Using Kalman filtering theory, they
developed and demonstrated a biokinetic motion tracker based
on their deformable superquadric model.
Vemuri and Radisavljevic (1993, 1994) constructed a
deformable superquadric model in an orthonormal wavelet
basis. This multiresolution basis provides the model with the
ability to transform continuously from local to global shape
deformations thereby allowing a continuum of shape models
to be created and to be represented with relatively few parameters. They applied the model to segment and reconstruct
anatomical structures in the human brain from MRI data.
As a final example, Bardinet et al. (1995, 1996a, b)
fitted a deformable superquadric to segmented 3-D cardiac
images and then refined the superquadric fit using a volumetric
deformation technique known as free form deformations
(FFDs). FFDs are defined by tensor product trivariate splines
and can be visualized as a rubber-like box in which the object
to be deformed (in this case the superquadric) is embedded.
Deformations of the box are automatically transmitted to
embedded objects. This volumetric aspect of FFDs allows two
superquadric surface models to be simultaneously deformed
in order to reconstruct the inner and outer surfaces of the LV of
the heart and the volume in between these surfaces. Further examples of deformable superquadrics can be found in Pentland
and Horowitz (1991) and Chen et al. (1994) (see section 3.5).
Several researchers cast the deformable model fitting
process in a probabilistic framework (see section 2.4) and
Deformable models in medical image analysis: a survey
include prior knowledge of object shape by incorporating prior
probability distributions on the shape variables to be estimated
(Staib and Duncan, 1992a; Worring et al., 1993; Vemuri and
Radisavljevic, 1994). For example, Staib and Duncan (1992a)
used a deformable contour model on 2-D echocardiograms
and MR images to extract the LV of the heart and the corpus
callosum of the brain, respectively. This closed contour model
is parameterized using an elliptic Fourier decomposition and
a priori shape information is included as a spatial probability
expressed through the likelihood of each model parameter.
The model parameter probability distributions are derived
from a set of example object boundaries and serve to bias
the contour model towards expected or more likely shapes.
Szekely et al. (1996) have also developed Fourier parameterized models. Furthermore, they have added elasticity to
their models to create ‘Fourier snakes’ in 2-D and elastically
deformable Fourier surface models in 3-D. By using the
Fourier parameterization followed by a statistical analysis
of a training set, they define mean organ models and their
eigen-deformations. An elastic fit of the mean model in
the subspace of eigenmodes restricts possible deformations
and finds an optimal match between the model surface and
boundary candidates.
Cootes et al. (1994) and Hill et al. (1993) presented a
statistically based technique for building deformable shape
templates and use these models to segment various organs from
2-D and 3-D medical images. The statistical parameterization
provides global shape constraints and allows the model to
deform only in ways implied by the training set. The shape
models represent objects by sets of landmark points which are
placed in the same way on an object boundary in each input
image. For example, to extract the LV from echocardiograms,
they chose points around the ventricle boundary, the nearby
edge of the right ventricle, and the top of the left atrium.
The points can be connected to form a deformable contour.
By examining the statistics of training sets of hand-labeled
medical images, and using principal component analysis, a
shape model is derived that describes the average positions
and the major modes of variation of the object points. New
shapes are generated using the mean shape and a weighted
sum of the major modes of variation. Object boundaries
are then segmented using this ‘point distribution model’ by
examining a region around each model point to calculate the
displacement required to move it towards the boundary. These
displacements are then used to update the shape parameter
weights.
3.4. Matching
Matching of regions in images can be performed between the
representation of a region and a model (labeling) or between
the representation of two distinct regions (registration). Reg-
99
istration of 2-D and 3-D medical images is necessary in order
to study the evolution of a pathology in an individual, or to
take full advantage of the complementary information coming
from multimodality imagery. Recent examples of the use of
deformable models to perform medical image registration are
found in Bookstein (1989), Moshfeghi (1991), Moshfeghi et
al. (1994), Gueziec and Ayache (1994), Feldmar and Ayache
(1994), Thirion (1994), Hamadeh et al. (1995), Lavallée
and Szeliski (1995). These techniques primarily consist in
constructing highly structured descriptions for matching. This
operation is usually carried out by extracting regions of interest
with an edge-detection algorithm, followed by the extraction
of landmark points or characteristic contours (or curves on
extracted boundary surfaces in the case of 3-D data). In
3-D, these curves usually describe differential structures such
as ridges, or topological singularities. An elastic matching
algorithm can then be applied between corresponding pairs
of curves or contours where the ‘start’ contour is iteratively
deformed to the ‘goal’ contour using forces derived from local
pattern matches with the goal contour (Moshfeghi, 1991).
An example of matching where the use of explicit a priori
knowledge has been embedded into deformable models is the
extraction and labeling of anatomic structures in the brain,
primarily from MR images. The anatomical knowledge
is made explicit in the form of a 3-D brain atlas. The
atlas is modeled as a physical object and is given elastic
properties. After an initial global alignment, the atlas deforms
and matches itself onto corresponding regions in the brain
image volume in response to forces derived from image
features. The assumption underlying this approach is that
at some representational level, normal brains have the same
topological structure and differ only in shape details. The
idea of modeling the atlas as an elastic object was originated
by Broit (1981), who formulated the matching process as
a minimization of a cost function. Subsequently, Bajcsy
and Kovacic (1989) implemented a multiresolution version
of Broit’s system where the deformation of the atlas proceeds
step-by-step in a coarse-to-fine strategy, increasing the local
similarity and global coherence. The elastically deformable
atlas technique has since become a very active area of research
and is being explored by several researchers (Evans et al.,
1991; Bookstein, 1991; Bozma and Duncan, 1992; Gee et al.,
1993; Delibasis and Undrill, 1994; McDonald et al., 1994;
Christensen et al., 1995; Declerck et al., 1995; Sandor and
Leahy, 1995; Snell et al., 1995; Subsol et al., 1995; Davatzikos
et al., 1996).
There are several problems with the deformable atlas
approach. The technique is sensitive to initial positioning
of the atlas—if the initial rigid alignment is off by too much,
then the elastic match may perform poorly. The presence of
neighboring features may also cause matching problems—the
100
T. McInerney and D. Terzopoulos
atlas may warp to an incorrect boundary. Finally, without
user interaction, the atlas can have difficulty converging
to complicated object boundaries. One solution to these
problems is to used image preprocessing in conjunction with
the deformable atlas. Sandor and Leahy (1995) used this
approach automatically to label regions of the cortical surface
that appear in 3-D MR images of human brains (Figure 7).
They automatically matched a labeled deformable atlas model
to preprocessed brain images, where preprocessing consists
of 3-D edge detection and morphological operations. These
filtering operations automatically extract the brain and sulci
(deep grooves in the cortical surface) from an MR image
and provide a smoothed representation of the brain surface
to which their 3-D B-spline deformable surface model can
rapidly converge.
3.5. Motion tracking and analysis
The idea of tracking objects in time-varying images using
deformable models was originally proposed in the context
of computer vision (Kass et al., 1988; Terzopoulos et al.,
1988). Deformable models have been used to track non-rigid
microscopic and macroscopic structures in motion, such as
blood cells (Leymarie and Levine, 1993) and neurite growth
cones (Gwydir et al., 1994) in cine-microscopy, as well as
coronary arteries in cine-angiography (Lengyel et al., 1995).
However, the primary use of deformable models for tracking in
medical image analysis is to measure the dynamic behavior of
the human heart, especially the LV. Regional characterization
of the heart wall motion is necessary to isolate the severity
and extent of diseases such as ischemia. Magnetic resonance
and other imaging technologies can now provide time-varying
3-D images of the heart with excellent spatial resolution and
reasonable temporal resolutions. Deformable models are well
suited for this image analysis task.
In the simplest approach, a 2-D deformable contour model
is used to segment the LV boundary in each slice of an
initial image volume. These contours are then used as the
initial approximation of the LV boundaries in corresponding
slices of the image volume at the next time instant and
are then deformed to extract the new set of LV boundaries
(Ueda and Mase, 1992; Ayache et al., 1992; Herlin and
Ayache, 1992; Singh et al., 1993; Geiger et al., 1995). This
temporal propagation of the deformable contours dramatically
decreases the time taken to segment the LV from a sequence
of image volumes over a cardiac cycle. Singh et al. (1993)
reported a time of 15 min to perform the segmentation,
considerably less than the 1.5–2 h that a human expert takes for
manual segmentation. McInerney and Terzopoulos (1995a)
have applied the temporal propagation approach in 3-D using
a 3-D dynamic deformable ‘balloon’ model to track the LV
(Figures 8 and 9).
In a more involved approach, Amini and Duncan (1992)
used bending energy and surface curvature to track and
analyze LV motion. For each time instant, two sparse
subsets of surface points are created by choosing geometrically
significant landmark points, one for the endocardial surface
and the other for the epicardial surface of the LV. Surface
patches surrounding these points are then modeled as thin,
flexible plates. Making the assumption that each surface patch
deforms only slightly and locally within a small time interval,
for each sampled point on the first surface they construct a
search area on the LV surface in the image volume at the next
time instant. The best matched (i.e. minimum bending energy)
point within the search window on the second surface is taken
to correspond to the point on the first surface. This matching
process yields a set of initial motion vectors for pairs of LV
surfaces derived from a 3-D image sequence. A smoothing
procedure is then performed using the initial motion vectors
to generate a dense motion vector field over the LV surfaces.
Cohen et al. (1992a) also employed a bending energy
technique in 2-D and attempt to improve on this method by
adding a term to the bending energy function that tends to
preserve the matching of high curvature points. Goldgof et
al. (Goldgof et al., 1988; Mishra et al., 1991; Huang and
Goldgof, 1993; Kambhamettu and Goldgof, 1994) have also
been pursuing surface shape matching ideas primarily based
on changes in Gaussian curvature and assume a conformal
motion model (i.e. motion which preserves angles between
curves on a surface but not distances).
An alternative approach is that of Chen et al. (1994),
who used a hierarchical motion model of the LV constructed
by combining a globally deformable superquadric with a
locally deformable surface using spherical harmonic shape
modeling primitives. Using this model, they estimated the LV
motion from angiographic data and produced a hierarchical
decomposition that characterizes the LV motion in a coarseto-fine fashion.
Pentland and Horowitz (1991) and Nastar and Ayache
(1993a, b) were also able to produce a coarse-to-fine characterization of the LV motion. They used dynamic deformable
models to track and recover the LV motion and make use of
modal analysis, a well-known mechanical engineering technique, to parameterize their models. This parameterization is
obtained from the eigenvectors of a finite element formulation
of the models. These eigenvectors are often referred to as
the ‘free vibration’ modes and variable detail of LV motion
representation results from varying the number of modes used.
The heart is a relatively smooth organ and consequently
there are few reliable landmark points. The heart also
undergoes complex non-rigid motion that includes a twisting
(tangential) component as well as the normal component
of motion. The motion recovery methods described above
Deformable models in medical image analysis: a survey
101
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
Figure 8. Sagittal slice of successive CT volumes over one cardiac cycle (1–16) showing motion of canine LV (McInerney and Terzopoulos,
1995a).
are, in general, not able to capture this tangential motion
without additional information. Recently, magnetic resonance
techniques, based on magnetic tagging (Axel and Dougherty,
1989) have been developed to track material points on the
myocardium in a non-invasive way. The temporal correspondence of material points that these techniques provide
allow for quantitative measurement of tissue motion and
deformation including the twisting component of the LV
motion. Several researchers have applied deformable models
to image sequences of MR tagged data (Young et al., 1993,
1995; Ducan et al., 1994; Kumar and Goldgof, 1994; Amini
et al., 1995; Kraitchman et al., 1995; Park et al., 1996).
For example, Amini et al. (1995) and Kumar and Goldgof
(1994) used a 2-D deformable grid to localize and track
SPAMM (spatial modulation of magnetization) tag points on
the LV tissue. Park et al. (1995, 1996) fitted a volumetric
physics-based deformable model to MRI-SPAMM data of
the LV. The parameters of the model are functions which
can capture regional shape variations of the LV such as
bending, twisting and contraction. Based on this model, the
authors quantitatively compared normal hearts and hearts with
hypertrophic cardiomyopathy.
Another problem with most of the methods described above
is that they model the endocardial and epicardial surfaces of the
LV separately. In reality the heart is a thick-walled structure.
Duncan et al. (1994) and Park et al. (1995, 1996) developed
models which consider the volumetric nature of the heart wall.
These models use the shape properties of the endocardial
and epicardial surfaces and incorporate mid-wall displacement
information of tagged MR images. By constructing 3-D finite
element models of the LV with nodes in the mid-wall region as
well as nodes on the endocardial and epicardial surfaces, more
accurate measurements of the LV motion can be obtained.
Young and Axel (1992, 1995) and Creswell (1992) have also
102
T. McInerney and D. Terzopoulos
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
Figure 9. Tracking of the LV motion during one cardiac cycle (1–16) using deformable balloon model (McInerney and Terzopoulos, 1995a).
constructed 3-D finite element models from the boundary
representations of the endocardial and epicardial surfaces.
4. DISCUSSION
In the previous sections we have surveyed the considerable
and rapidly expanding body of work on deformable models
in medical image analysis. The survey has revealed several
issues that are relevant to the continued development of the
deformable model approach. This section summarizes the key
issues and indicates some promising research directions.
• Automation versus control
Interactive (semiautomatic) algorithms and fully automatic algorithms represent two alternative approaches
to computerized medical image analysis. Certainly
automatic interpretation of medical images is a desirable, albeit very difficult, long-term goal, since it can
potentially increase the speed, accuracy, consistency
and reproducibility of the analysis. However, the
interactive or semiautomatic methodology is likely to
remain dominant in practice for some time to come, especially in applications where erroneous interpretations are
unacceptable. Consequently, the most immediately successful deformable-model-based techniques are likely be
those that drastically decrease the labor-intensiveness
of medical image processing tasks through partial automation, while still allowing for interactive guidance or
editing by the medical expert. Although fully automatic
Deformable models in medical image analysis: a survey
techniques based on deformable models are unlikely to
reach their full potential for some time to come, they can
be of immediate value in specific application domains
where interactive techniques are either impractical (e.g.
3-D reconstruction of complex vascular structures) or
for non-critical tasks (e.g. segmentation of healthy tissue
surrounding a pathology for enhanced visualization).
• Generality versus specificity
Ideally a deformable model should be capable of representing a broad range of shapes and be useful in a
wide array of medical applications. Generality is the
basis of deformable model formulations with local shape
parameters such as snakes. Alternatively, highly specific,
‘hand-crafted’ or constrained deformable models appear
to be useful in applications such as tracking the non-rigid
motion of the heart (section 3.5), automatically matching
and labeling structures in the brain from 3-D MR images
(section 3.4), or segmenting very noisy images such
as echocardiograms. Certainly attempts to automate
completely the processing of medical images would
require a high degree of application and model specificity.
A promising direction for future study appears to be
techniques for learning ‘tailored’ models from simple
general purpose models. The work of Cootes et al. (1994)
may be viewed as an example of such a strategy.
• Compactness versus geometric coverage versus topological flexibility
A geometric model of shape may be evaluated based
on the parsimony of its formulation, its representational
power and its topological flexibility. Generally, parameterized models offer the greatest parsimony, freeform (spline) models feature the broadest coverage, and
implicit models have the greatest topological flexibility.
Deformable models have been developed based on each
of these geometric classes. Increasingly, researchers
are turning to the development of hybrid deformable
models that combine complementary features. For
objects with a simple, fixed topology and without
significant protrusions, parameterized models coupled
with local (spline) and/or global deformations schemes
appear to provide a good compactness–descriptiveness
tradeoff (Terzopoulos and Metaxas, 1991; Pentland and
Horowitz, 1991; Vemuri and Radisavljevic, 1994; Chen
et al., 1994). On the other hand, the segmentation
and modeling of complex, multipart objects such as
arterial or bronchial ‘tree’ structures, or topologically
complex structures such as vertebrae, is a difficult task
with these types of models. Polygon-based or particlebased deformable modeling schemes seem promising in
segmenting and reconstructing such structures. Polygonbased models may be made compacted by removing
103
and ‘retiling’ (Turk, 1992; Gourdon, 1995) polygons in
regions of low shape variation, or by replacing a region
of polygons with a single, high-order finite element or
spline patch. A possible research direction is to develop
alternative models that blend or combine descriptive
primitive elements, such as flexible cylinders, into a
global structure.
• Curve versus surface versus solid models
The earliest deformable models were curves and
surfaces. Anatomic structures in the human body,
however, are either solid or thick-walled. To support
the expanding role of medical images into tasks such
as surgical planning and simulation, and the functional
modeling of structures such as bones, muscles, skin
or arterial blood flow, may require volumetric or solid
deformable models rather than surface models. For
example, the planning of facial reconstructive surgery
requires the extraction and reconstruction of the skin,
muscles and bones from 3-D images using accurate
solid models. It also requires the ability to simulate
the movement and interactions of these structures in
response to forces, the ability to move, cut and fuse
pieces of the model in a realistic fashion, and the ability
to stimulate the simulated muscles of the model to predict
the effect of the surgery. Several researchers have begun
to explore the use of volumetric or solid deformable
models of the human face and head for computer
graphics applications (Essa et al., 1993; Lee et al., 1995)
and for medical applications, particularly reconstructive
surgery (Waters, 1992; Pieper et al., 1992; Geiger, 1992;
Delingette et al., 1994) and there is much room for further
research. Researchers have also begun to use volumetric
deformable models to track and analyze LV motion
more accurately (Young and Axel, 1992; Creswell et al.,
1992; Duncan et al., 1994; Park et al., 1996).
• Accuracy and quantitative power
Ideally it should be possible to measure and control the
accuracy of a deformable model. The most common
accuracy control mechanisms are the global or local subdivision of model basis functions (Miller et al., 1991), or
the repositioning of model points to increase their density
in regions of the data exhibiting rapid shape variations
(Vasilescu and Terzopoulos, 1992). Other mechanisms
that warrant further research are the local control and
adaptation of model continuity, parameter evolution
(including the rate and scheduling of the evolution), and
the automation of all accuracy control mechanisms. The
parametric formulation of a deformable model should
not only yield an accurate description of the object, but
it should also provide quantitative information about
the object in an intuitive, convenient form. That is,
104
T. McInerney and D. Terzopoulos
the model parameters should be useful for operations
such as measuring, matching, modification, rendering,
and higher-level analysis or geometric reasoning. This
‘parameter descriptiveness’ criterion may be achieved
in a postprocessing step by adapting or optimizing
the parameterization to match the data more efficiently
or more descriptively. However, it is preferable to
incorporate the descriptive parameterization directly into
the model formulation. An example of this strategy is the
deformable model of Park et al. (1995, 1996).
• Robustness
Ideally, a deformable model should be insensitive to
initial conditions and noisy data. Deformable models are
able to exploit multiple image attributes and high-level or
global information to increase the robustness of shape recovery. For example, many snakes models now incorporate region-based image features as well as the traditional
edge-based features (section 3.1). Strategies worthy of
further research include the incorporation of shape constraints into the deformable model that are derived from
low-level image-processing operations such as thinning,
medial axis transforms or mathematical morphology. A
classical approach to improve the robustness of model
fitting is the use of multiscale image preprocessing techniques (Kass et al., 1988; Terzopoulos et al., 1988), perhaps coupled with a multiresolution deformable model
(Bajcsy and Kovacic, 1989). A multiresolution technique
that merits further research in the context of deformable
models is the use of wavelet bases (Strang and Nguyen,
1996) for deformations (Vemuri et al., 1993; Vemuri and
Radisavljevic, 1994). A deformable model should be
able to easily incorporate added constraints and any other
a priori anatomic knowledge of object shape and motion.
Section 3.3 reviewed several of the most promising techniques to incorporate a priori knowledge. For example,
for LV motion tracking, a promising research direction is
the incorporation of biomechanical properties of the heart
and the inclusion of the temporal periodic characteristics
of the heart motion. Future directions include modeling schemes that incorporate reasoning and recognition
mechanisms using techniques from artificial intelligence
such as rule-based systems or neural networks.
5. CONCLUSION
The increasingly important role of medical imaging in the
diagnosis and treatment of disease has opened an array of
challenging problems centered on the computation of accurate geometric models of anatomic structures from medical
images. Deformable models offer an attractive approach to
tackling such problems, because these models are able to
represent the complex shapes and broad shape variability of
anatomical structures. Deformable models overcome many of
the limitations of traditional low-level image-processing techniques, by providing compact and analytical representations
of object shape, by incorporating anatomic knowledge, and by
providing interactive capabilities. The continued development
and refinement of these models should remain an important
area of research into the foreseeable future.
ACKNOWLEDGEMENTS
We would like to thank Stephanie Sandor and Richard
Leahy of the USC Signal and Image Processing Institute for
the deformable brain atlas figure, as well as the following
individuals for providing citation information that improved
the completeness of the bibliography: Amir Amini, Nicholas
Ayache, Ingrid Carlbom, Chang Wen Chen, James Duncan,
Dmitry Goldgof, Thomas Huang, Stephane Lavallee, Francois
Leitner, Gerard Medioni, Dimitri Metaxas, Alex Pentland,
Stan Sclaroff, Ajit Singh, Richard Szeliski, Baba Vemuri,
Alistair Young and Alan Yuille. T.M. is grateful for the
financial support of an NSERC postgraduate scholarship. D.T.
is a fellow of the Canadian Institute for Advanced Research.
This work was made possible by the financial support of the
Information Technologies Research Center of Ontario.
REFERENCES
Amini, A. A. and Duncan, J. S. (1992). Bending and stretching
models for LV wall motion analysis from curves and surfaces.
Image and Vision Computing, 10(6), 418–430.
Amini, A. A., Weymouth, T. E. and Jain, R. C. (1990). Using dynamic
programming for solving variational problems in vision. IEEE
Trans. Pattern Anal. Machine Intelligence, 12(9), 855–867.
Amini, A. A., Curwen, R. W., Klein, A. K., Egglin, T. K., Pollak, J.,
Lee, F. and Gore, J. C. (1995). Physics based snakes, Kalman
snakes, and snake grids for feature localization and tracking in
medical images. In (Bizais et al., 1995), pp. 363–364.
Axel, L. and Dougherty, L. (1989). Heart wall motion: Improved
method of spatial modulation of magnetization for MR imaging.
Radiology, 172, 349–350.
Ayache, N. (1995a). Medical computer vision, virtual reality and
robotics. Image and Vision Computing, 13(4), 295–313.
Ayache, N. (ed.) (1995b). Proc. First International Conf. on
Computer Vision, Virtual Reality and Robotics in Medicine
(CVRMed’95), Nice, France, April, 1995. Lectures Notes in
Computer Science, Vol. 905. Berlin, Germany: Springer–Verlag.
Ayache, N., Cohen, I. and Herlin, I. L. (1992). Medical image
tracking. Ch. 17 Blake, A. and Yuille, A. (eds), Active Vision.
Cambridge, MA: MIT Press.
Bajcsy, R. and Kovacic, S. (1989). Multiresolution elastic matching.
Comput. Vision, Graphics Image Process., 46, 1–21.
Ballard, D. and Brown, C. (1982). Computer Vision. Englewood
Cliffs, NJ: Prentice–Hall.
Deformable models in medical image analysis: a survey
Bardinet, E., Cohen, L. D. and Ayache, N. (1995). Superquadrics
and free-form deformations: A global model to fit and track 3D
medical data. In (Ayache, 1995b), pp. 319–326.
Bardinet, E., Cohen, L. D. and Ayache, N. (1996a). Analyzing the
deformation of the left ventricle of the heart with a parametric
deformable model. Research report 2797. INRIA, SophiaAntipolis, France.
Bardinet, E., Cohen, L. D. and Ayache, N. (1996b). A parametric
deformable model to fit unstructured 3D data. Computer Vision
and Image Understanding. In press. Also research report 2617,
INRIA, Sophia-Antipolis, France.
Berger, M. O. (1990). Snake growing. In Faugeras, O. (ed.),
Computer Vision – Proc. First European Conf. on Computer
Vision (ECCV’90), Antibes, France, April, 1990, pp. 570–572,
Lectures Notes in Computer Science. Springer–Verlag.
Bizais, Y., Barillot, C. and Paola, R. Di (eds) (1995). Information
Processing in Medical Imaging: Proc. 14th Int. Conf. (IPMI’95),
Ile de Berder, France, June, 1995. Computational Imaging and
Vision, Vol. 3. Dordrecht, The Netherlands: Kluwer Academic.
Bookstein, F. L. (1989). Principal warps: Thin-plate splines and
the decomposition of deformations. IEEE Trans. Pattern Anal.
Machine Intelligence, 11(6), 567–585.
Bookstein, F. L. (1991). Thin-plate splines and the atlas problem
for biomedical images. In Barret, H.H. and Gmitro, A.F. (eds),
Information Processing in Medical Imaging: Proc. 12th Int.
Conf. (IPMI’91), Wye, UK, July, 1991, pp. 326–342, Lectures
Notes in Computer Science. Springer–Verlag.
Bozma, I. and Duncan, J. S. (1992). A modular system for
image analysis using a game theoretic framework. Image Vision
Comput., 10(6), 431–443.
Broit, C. (1981). Optimal Registration of Deformed Images. Ph.D.
thesis, Computer and Information Science Dept., University of
Pennsylvania, Philadelphia, PA.
Carlbom, I., Terzopoulos, D. and Harris, K. (1994). Computerassisted registration, segmentation, and 3D reconstruction from
images of neuronal tissue sections. IEEE Trans. Med. Imag.,
13(2), 351–362.
Caselles, V., Catte, F., Coll, T. and Dibos, F. (1993). A geometric
model for active contours. Numerische Mathematik, 66.
Caselles, V., Kimmel, R. and Sapiro, G. (1995). Geodesic active
contours. In Proc. Fifth International Conf. on Computer Vision
(ICCV’95), Cambridge, MA, June, 1995, pp. 694–699, Los
Alamitos, CA: IEEE Computer Society Press.
Chakraborty, A. and Duncan, J. S. (1995). Integration of boundary
finding and region-based segmentation using game theory. In
(Bizais et al., 1995), pp. 189–200.
Chakraborty, A., Staib, L. H. and Duncan, J. S. (1994). Deformable
boundary finding influenced by region homogeneity. In Proc.
Conf. on Computer Vision and Pattern Recognition (CVPR’94),
Seattle, WA, June, 1994, Los Alamitos, CA, pp. 624–627. IEEE
Computer Society Press.
Chang, L. W., Chen, H. W. and Ho, J. R. (1991). Reconstruction
of 3D medical images: A nonlinear interpolation technique for
reconstruction of 3D medical images. Comput. Vision, Graphics
Image Process., 53(4), 382–391.
105
Chen, C. W., Huang, T. S. and Arrott, M. (1994). Modeling,
analysis and visualization of left ventricle shape and motion by
hierarchical decomposition. IEEE Trans. Pattern Anal. Machine
Intelligence, 16(April), 342–356.
Christensen, G., Rabbitt, R. D., Miller, M. I., Joshi, S. C., Grenander,
U., Coogan, T. A. and van Essen, D. C. (1995). Topological
properties of smooth anatomic maps. In (Bizais et al., 1995),
pp. 101–112.
Cohen, I., Ayache, N. and Sulger, P. (1992a). Tracking points on
deformable objects using curvature information. In Sandini,
G. (ed.), Computer Vision – Proc. Second European Conf. on
Computer Vision (ECCV’92), Santa Margherita Ligure, Italy,
May, 1992, pp. 458–466, Lectures Notes in Computer Science.
Springer–Verlag.
Cohen, I., Cohen, L. D. and Ayache, N. (1992b). Using deformable
surfaces to segment 3D images and infer differential structures.
CVGIP: Image Understanding, 56(2), 242–263.
Cohen, L. D. (1991). On active contour models and balloons. CVGIP:
Image Understanding, 53(2), 211–218.
Cohen, L. D. and Cohen, I. (1993). Finite element methods for active
contour models and balloons for 2D and 3D images. IEEE Trans.
Pattern Anal. Machine Intelligence, 15(11), 1131–1147.
Cootes, T., Hill, A., Taylor, C. and Haslam, J. (1994). The use of
active shape models for locating structures in medical images.
Image Vision Comput., 12(6), 355–366.
Creswell, L. L., Wyers, S. G., Pirolo, J. S., Perman, W. H., Vannier,
M. W. and Pasque, M. K. (1992). Mathematical modelling of
the heart using magnetic resonance imaging. IEEE Trans. Med.
Imag., 11(4), 581–589.
Davatzikos, C. and Bryan, R. N. (1995). Using a deformable
surface model to obtain a mathematical representation of the
cortex. In International Symp. on Computer Vision, Coral
Gables, FL, November, 1995, pp. 212–217, Los Alamitos, CA:
IEEE Computer Society Press.
Davatzikos, C. A. and Prince, J. L. (1995). An active contour model
for mapping the cortex. IEEE Trans. Med. Imag., 14(1), 65–80.
Davatzikos, C. A., Prince, J. L. and Bryan, R. N. (1996). Image
registration based on boundary mapping. IEEE Trans. Med.
Imag., 15(1), 112–115.
Declerck, J., Subsol, G., Thirion, J. P. and Ayache, N. (1995).
Automatic retrieval of anatomic structures in 3D medical images.
In (Ayache, 1995b), pp. 153–162.
Delibasis, K. and Undrill, P. E. (1994). Anatomical object recognition
using deformable geometric models. Image Vision Comput.,
12(7), 423–433.
Delingette, H., Hebert, M. and Ikeuchi, K. (1992). Shape representation and image segmentation using deformable surfaces. Image
Vision Comput., 10(3), 132–144.
Delingette, H., Subsol, G., Cotin, S. and Pignon, J. (1994). Virtual
reality and craniofacial surgery simulation. In (Robb, 1994),
pp. 607–618.
Duncan, J. et al. (1994). Towards reliable, noninvasive measurement
of myocardial function from 4D images. In Medical Imaging
1994: Physiology and Function from Multidimensional Medical
106
T. McInerney and D. Terzopoulos
Images, SPIE Proc., Vol. 2168, pp. 149–161, Bellingham, WA:
SPIE.
Essa, I., Sclaroff, S. and Pentland, A. P. (1993). Physically-based
modeling for graphics and vision. In Martin, R. (ed.), Directions
in Geometric Computing. Information Geometers, U.K.
Evans, A. C., Dai, W., Collins, L., Neelin, P. and Marrett, S.
(1991). Warping of a computerized 3D atlas to match brain
image volumes for quantitative neuroanatomical and functional
analysis. In Medical Imaging V: Image Processing, SPIE Proc.,
Vol. 1445, pp. 236–246, Bellingham, WA: SPIE.
Farin, G. (1993). Curves and Surfaces for CAGD. New York, NY:
Academic Press.
Feldmar, J. and Ayache, N. (1994). Locally affine registration of freeform surfaces. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’94), Seattle, WA, June, 1994, pp. 496–501,
Los Alamitos, CA: IEEE Computer Society Press.
Fischler, M. and Elschlager, R. (1973). The representation and
matching of pictorial structures. IEEE Trans. on Computers,
22(1), 67–92.
Gauch, J. M., Pien, H. H. and Shah, J. (1994). Hybrid boundarybased and region-based deformable models for biomedical image
segmentation. In Mathematical Methods in Medical Imaging III,
SPIE Proc., Vol. 2299, pp. 72–83, San Diego, CA: SPIE.
Gee, J., Reivich, M. and Bajcsy, R. (1993). Elastically deforming
3D atlas to match anatomical brain images. Journal of Computer
Assisted Tomography, 17(2), 225–236.
Geiger, B. (1992). Three dimensional simulation of delivery for
cephalopelvic disproportion. In First International Workshop on
Mechatronics in Medicine and Surgery, Costa del Sol, Spain,
October, 1992, pp. 146–152.
Geiger, D., Gupta, A., Costa, L. A. and Vlontzos, J. (1995). Dynamic
programming for detecting, tracking and matching deformable
contours. IEEE Trans. Pattern Anal. Machine Intelligence, 17(3),
294–302.
Goldgof, D. B., Lee, H. and Huang, T. S. (1988). Motion analysis of
nonrigid surfaces. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’88), Ann Arbor, MI, June, 1988, pp. 375–
380, Los Alamitos, CA: IEEE Computer Society Press.
Gourdon, A. (1995). Simplification of irregular surface meshes in
3D medical images. In (Ayache, 1995b), pp. 413–419.
Grzeszczuk, R. P. and Levin, D. N. (1994). Brownian strings:
Segmenting images with stochastically deformable contours. In
(Robb, 1994), pp. 72–89.
Gueziec, A. and Ayache, N. (1994). Smoothing and matching of 3D
space curves. Int. J. Comput. Vision, 12(1), 79–104.
Gupta, A., O’Donnell, T. and Singh, A. (1994). Segmentation
and tracking of cine cardiac MR and CT images using a 3D deformable model. In Proc. IEEE Conf. on Computers in
Cardiology, September, 1994.
Gwydir, S. H., Buettner, H. M. and Dunn, S. M. (1994). Non-rigid
motion analysis and feature labelling of the growth cone. In IEEE
Workshop on Biomedical Image Analysis, Seattle, WA, June,
1994, pp. 80–87, Los Alamitos, CA: IEEE Computer Society
Press.
Hamadeh, A., Lavallee, S., Szeliski, R., Cinquin, P. and Peria,
O. (1995). Anatomy-based registration for computer-integrated
surgery. In (Ayache, 1995b), pp. 212–218.
Herlin, I. L. and Ayache, N. (1992). Features extraction and analysis
methods for sequences of ultrasound images. Image Vision
Comput., 10(10), 673–682.
Herlin, I. L., Nguyen, C. and Graffigne, C. (1992). A deformable
region model using stochastic processes applied to echocardiographic images. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’92), Urbana, IL, June, 1992, pp. 534–539,
Los Alamitos, CA: IEEE Computer Society Press.
Hill, A., Thornham, A. and Taylor, C. J. (1993). Model-based
interpretation of 3D medical images. In Proc. 4th British Machine
Vision Conf. (BMVC’93), Surrey, UK, September, 1993, pp. 339–
348, BMVA Press.
Huang, W. C. and Goldgof, D. B. (1993). Adaptive-size meshes for
rigid and nonrigid shape analysis and synthesis. IEEE Trans.
Pattern Anal. Machine Intelligence, 15(3).
Kambhamettu, C. and Goldgof, D. B. (1994). Point correspondence
recovery in nonrigid motion. CVGIP: Image Understanding,
60(1), 26–43.
Kass, M., Witkin, A. and Terzopoulos, D. (1988). Snakes: Active
contour models. Int. J. Comput. Vision, 1(4), 321–331.
Kraitchman, D. L., Young, A. A., Chang, C. N. and Axel, L. (1995).
Semi-automatic tracking of myocardial motion in MR tagged
images. IEEE Trans. Med. Imag., 14(3), 422–432.
Kumar, S. and Goldgof, D. (1994). Automatic tracking of SPAMM
grid and the estimation of deformation parameters from cardiac
MR images. IEEE Trans. Med. Imag., 13(1), 122–132.
Lavallée, S. and Szeliski, R. (1995). Recovering the position and
orientation of free-form objects from image contours using 3D
distance maps. IEEE Trans. Pattern Anal. Machine Intelligence,
17(4), 378–390.
Lee, Y., Terzopoulos, D. and Waters, K. (1995). Realistic modeling
for facial animation. In Proc. SIGGRAPH’95, Los Angeles, CA,
August, 1995, in Computer Graphics Proc., Annual Conf. Series
1995, pp. 55–62, New York, NY: ACM SIGGRAPH.
Leitner, F. and Cinquin, P. (1991). Complex topology 3D objects
segmentation. In Model-Based Vision Development and Tools,
SPIE Proc., Vol. 1609, pp. 16–26, Bellingham, WA: SPIE.
Leitner, F. and Cinquin, P. (1993). From splines and snakes to
Snakes Splines. In Laugier, C. (ed.), Geometric Reasoning: From
Perception to Action, Lectures Notes in Computer Science, Vol.
708, pp. 264–281, Springer–Verlag.
Lengyel, J., Greenberg, D. P. and Popp, R. (1995). Timedependent three-dimensional intravascular ultrasound. In Proc.
SIGGRAPH’95, Los Angeles, CA, August, 1995, in Computer
Graphics Proc., Annual Conf. Series 1995, pp. 457–464, New
York, NY: ACM SIGGRAPH.
Leymarie, F. and Levine, M. (1993). Tracking deformable objects
in the plane using an active contour model. IEEE Trans. Pattern
Anal. Machine Intelligence, 15(6), 635–646.
Lin, W. C. and Chen, S. Y. (1989). A new surface interpolation
technique for reconstructing 3D objects from serial cross-
Deformable models in medical image analysis: a survey
sections. Comput. Vision, Graphics Image Process., 48(Oct.),
124–143.
Lipson, P., Yuille, A. L., O’Keefe, D., Cavanaugh, J., Taaffe, J. and
Rosenthal, D. (1990). Deformable templates for feature extraction from medical images. In Faugeras, O. (ed.), Computer Vision
– Proc. First European Conf. on Computer Vision (ECCV’90),
Antibes, France, April, 1990, pp. 477–484, Lectures Notes in
Computer Science. Springer–Verlag.
Lobregt, S. and Viergever, M. A. (1995). A discrete dynamic contour
model. IEEE Trans. Med. Imag., 14(1), 12–24.
Malladi, R., Sethian, J. and Vemuri, B. C. (1995). Shape modeling
with front propagation: A level set approach. IEEE Trans. Pattern
Anal. Machine Intelligence, 17(2), 158–175.
Mangin, J. F., Tupin, F., Frouin, V., Bloch, I., Rougetet, R., Regis,
J. and Lopez-Krahe, J. (1995). Deformable topological models
for segmentation of 3D medical images. In (Bizais et al., 1995),
pp. 153–164.
McDonald, D., Avis, D. and Evans, A. (1994). Multiple surface
identification and matching in magnetic resonance images. In
(Robb, 1994), pp. 160–169.
McInerney, T. and Terzopoulos, D. (1995a). A dynamic finite element
surface model for segmentation and tracking in multidimensional
medical images with application to cardiac 4D image analysis.
Computerized Medical Imaging and Graphics, 19(1), 69–83.
McInerney, T. and Terzopoulos, D. (1995b). Topologically adaptable
snakes. In Proc. Fifth International Conf. on Computer Vision
(ICCV’95), Cambridge, MA, June, 1995, pp. 840–845, Los
Alamitos, CA: IEEE Computer Society Press.
Metaxas, D. and Terzopoulos, D. (1993). Shape and nonrigid motion
estimation through physics-based synthesis. IEEE Trans. Pattern
Anal. Machine Intelligence, 15(6), 580–591.
Miller, J. V., Breen, D. E., Lorensen, W. E., O’Bara, R. M. and
Wozny, M. J. (1991). Geometrically deformed models: A method
for extracting closed geometric models from volume data. In
Computer Graphics (Proc. SIGGRAPH’91 Conf., Las Vegas, NV,
July, 1991), Vol. 25(4), pp. 217–226.
Mishra, S. K., Goldgof, D. B. and Huang, T. S. (1991). Nonrigid motion analysis and epicardial deformation estimation from
angiography data. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’91), Maui, HI, June, 1991, pp. 331–336, Los
Alamitos, CA: IEEE Computer Society Press.
Moshfeghi, M. (1991). Elastic matching of multimodality medical
images. CVGIP: Graphical Models and Image Processing, 53,
271–282.
Moshfeghi, M., Ranganath, S. and Nawyn, K. (1994). Threedimensional elastic matching of volumes. IEEE Trans. on Image
Processing, 3, 128–138.
Nastar, C. and Ayache, N. (1993a). Fast segmentation, tracking, and
analysis of deformable objects. In Proc. Fourth International
Conf. on Computer Vision (ICCV’93), Berlin, Germany, May,
1993, pp. 275–279, Los Alamitos, CA: IEEE Computer Society
Press.
Nastar, C. and Ayache, N. (1993b). Non-rigid motion analysis in
medical images: A physically based approach. In Colchester,
A.C.F. and Hawkes, D.J. (eds), Information Processing in
107
Medical Imaging: Proc. 13th Int. Conf. (IPMI’93), Flagstaff,
AZ, June, 1993, pp. 17–32, Lectures Notes in Computer Science.
Springer–Verlag.
Park, J., Metaxas, D. and Axel, L. (1995). Volumetric deformable
models with parameter functions: A new approach to the 3D
motion analysis of the LV from MRI–SPAMM. In Proc. Fifth
International Conf. on Computer Vision (ICCV’95), Cambridge,
MA, June, 1995, pp. 700–705, Los Alamitos, CA: IEEE
Computer Society Press.
Park, J., Metaxas, D. and Axel, L. (1996). Analysis of left ventricular
wall motion based on volumetric deformable models and MRISPAMM. Medical Image Analysis, 1(1), 53–72.
Pentland, A. and Horowitz, B. (1991). Recovery of nonrigid motion
and structure. IEEE Trans. Pattern Anal. Machine Intelligence,
13(7), 730–742.
Pentland, A. and Sclaroff, S. (1991). Closed-form solutions for
physically based shape modelling and recognition. IEEE Trans.
Pattern Anal. Machine Intelligence, 13(7), 715–729.
Pieper, S., Rosen, J. and Zeltzer, D. (1992). Interactive graphics
for plastic surgery: A task-level analysis and implementation.
In Proc. ACM 1992 Symposium on Interactive 3D Graphics,
pp. 127–134.
Poon, C. S., Braun, M., Fahrig, R., Ginige, A. and Dorrell, A. (1994).
Segmentation of medical images using an active contour model
incorporating region-based images features. In (Robb, 1994),
pp. 90–97.
Press, W. H., Teukolsky, S. A., Vetterling, W. T. and Flannery, B. P.
(1992). Numerical Recipes in C. Cambridge University Press.
Robb, R. A. (ed.) (1994). Proc. Third Conf. on Visualization
in Biomedical Computing (VBC’94), Rochester, MN, October,
1994. SPIE Proc., Vol. 2359, Bellingham, WA: SPIE.
Rougon, N. and Prêteux, F. (1991). Deformable markers: Mathematical morphology for active contour models control. In Image
Algebra and Morphological Image Processing II SPIE Proc.,
Vol. 1568, pp. 78–89, Bellingham, WA: SPIE.
Rougon, N. and Prêteux, F. (1993). Directional adaptive deformable
models for segmentation with application to 2D and 3D medical
images. In Medical Imaging 93: Image Processing, SPIE Proc.,
Vol. 1898, pp. 193–207, Bellingham, WA: SPIE.
Sandor, S. and Leahy, R. (1995). Towards automatic labelling of the
cerebral cortex using a deformable atlas model. In (Bizais et al.,
1995), pp. 127–138.
Sapiro, G., Kimmel, R. and Caselles, V. (1995). Object detection
and measurements in medical images via geodesic deformable
contours. In Vision Geometry IV, SPIE Proc., Vol. 2573, pp. 366–
378, Bellingham, WA: SPIE.
Singh, A., von Kurowski, L. and Chiu, M. Y. (1993). Cardiac MR
image segmentation using deformable models. In Biomedical
Image Processing and Biomedical Visualization, SPIE Proc.,
Vol. 1905, pp. 8–28, Bellingham, WA: SPIE.
Snell, J. W., Merickel, M. B., Ortega, J. M., Goble, J. C., Brookeman,
J. R. and Kassell, N. F. (1995). Model-based boundary estimation
of complex objects using hierarchical active surface templates.
Pattern Recog., 28(10), 1599–1609.
108
T. McInerney and D. Terzopoulos
Staib, L. H. and Duncan, J. S. (1992a). Boundary finding with
parametrically deformable models. IEEE Trans. Pattern Anal.
Machine Intelligence, 14(11), 1061–1075.
Staib, L. H. and Duncan, J. S. (1992b). Deformable Fourier
models for surface finding in 3D images. In Robb, R. A. (ed.),
Proc. Second Conf. on Visualization in Biomedical Computing
(VBC’92), Chapel Hill, NC, October, 1992, SPIE Proc., Vol.
1808, pp. 90–104, Bellingham, WA: SPIE.
Strang, G. and Nguyen, T. (1996). Wavelets and Filter Banks.
Wellesley, MA: Wellesley-Cambridge Press.
Stytz, M., Frieder, G. and Frieder, O. (1991) Three-dimensional
medical imaging: Algorithms and computer systems. ACM
Computing Surveys, 23, 421–499.
Subsol, G., Thirion, J.Ph. and Ayache, N. (1995). A general scheme
for automatically building 3D morphometric anatomical atlases:
Application to a skull atlas. In Proc. Second International
Symp. on Medical Robotics and Computer Assisted Surgery
(MRCAS’95), Baltimore, MD, November, 1995, pp. 226–233.
Székely, G., Kelemen, A., Brechbuhler, Ch. and Gerig, G. (1996).
Segmentation of 2-D and 3-D objects from MRI volume data
using constrained elastic deformations of flexible Fourier surface
models. Medical Image Analysis, 1(1), 19–34.
Szeliski, R. (1990). Bayesian modeling of uncertainty in low-level
vision. Int. J. Comput. Vision, 5, 271–301.
Szeliski, R., Tonnesen, D. and Terzopoulos, D. (1993). Modeling
surfaces of arbitrary topology with dynamic particles. In Proc.
Conf. on Computer Vision and Pattern Recognition (CVPR’93),
New York, NY, June, 1993, pp. 82–87, Los Alamitos, CA: IEEE
Computer Society Press.
Tek, H. and Kimia, B. (1995). Shock-based reaction-diffusion
bubbles for image segmentation.
In (Ayache, 1995b),
pp. 434–438.
Terzopoulos, D. (1986a). On Matching Deformable Models to
Images. Tech. rept. 60. Schlumberger Palo Alto Research.
Reprinted in Topical Meeting on Machine Vision, Technical
Digest Series, Vol. 12, (Optical Society of America, Washington,
DC) 1987, 160–167.
Terzopoulos, D. (1986b). Regularization of inverse visual problems
involving discontinuities. IEEE Trans. Pattern Anal. Machine
Intelligence, 8(4), 413–424.
Terzopoulos, D. and Fleischer, K. (1988). Deformable models. The
Visual Computer, 4(6), 306–331.
Terzopoulos, D. and Metaxas, D. (1991). Dynamic 3D models with
local and global deformations: Deformable superquadrics. IEEE
Trans. Pattern Anal. Machine Intelligence, 13(7), 703–714.
Terzopoulos, D. and Szeliski, R. (1992). Tracking with Kalman
snakes. In Blake, A. and Yuille, A. (eds), Active Vision, pp. 3–20,
Cambridge, MA: MIT Press.
Terzopoulos, D., Witkin, A. and Kass, M. (1988). Constraints on
deformable models: Recovering 3D shape and nonrigid motion.
Artif. Intelligence, 36(1), 91–123.
Thirion, J. P. (1994). Extremal points: Definition and application
to 3D image registration. In Proc. Conf. on Computer Vision
and Pattern Recognition (CVPR’94), Seattle, WA, June, 1994,
pp. 587–592, Los Alamitos, CA: IEEE Computer Society Press.
Turk, G. (1992). Re-tiling polygonal surfaces. In Computer Graphics
(Proc. SIGGRAPH’92 Conf., Chicago, IL, July, 1992), Vol. 26(2),
pp. 55–64, ACM SIGGRAPH.
Ueda, N. and Mase, K. (1992). Tracking moving contours using
energy-minimizing elastic contour models. In Sandini, G. (ed.),
Computer Vision – Proc. Second European Conf. on Computer
Vision (ECCV’92), Santa Margherita Ligure, Italy, May, 1992,
pp. 453–457, Lectures Notes in Computer Science. Springer–
Verlag.
Vasilescu, M. and Terzopoulos, D. (1992). Adaptive meshes and
shells: Irregular triangulation, discontinuities and hierarchical
subdivision. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’92), Urbana, IL, June, 1992, pp. 829–832,
Los Alamitos, CA: IEEE Computer Society Press.
Vemuri, B. C. and Radisavljevic, A. (1994). Multiresolution
stochastic hybrid shape models with fractal priors. ACM Trans.
on Graphics, 13(2), 177–207.
Vemuri, B. C., Radisavljevic, A. and Leonard, C. (1993). Multiresolution 3D stochastic shape models for image segmentation.
In Colchester, A. C. F. and Hawkes, D. J. (eds), Information
Processing in Medical Imaging: Proc. 13th Int. Conf. (IPMI’93),
Flagstaff, AZ, June, 1993, pp. 62–76, Lectures Notes in Computer
Science. Springer–Verlag.
Waters, K. (1992). A physical model of facial tissue and muscle
articulation derived from computer tomography data. In Robb,
R. A. (ed.), Proc. Second Conf. on Visualization in Biomedical
Computing (VBC’92), Chapel Hill, NC, October, 1992, SPIE
Proc., Vol. 1808, pp. 574–583, Bellingham, WA: SPIE.
Whitaker, R. (1994). Volumetric deformable models. In (Robb,
1994).
Widrow, B. (1973). The rubber mask technique, part I. Pattern
Recognition, 5(3), 175–211.
Worring, M., Smeulders, A. W. M., Staib, L. H. and Duncan, J. S.
(1993). Parameterized feasible boundaries in gradient vector
fields. In Colchester, A. C. F. and Hawkes, D. J. (eds), Information
Processing in Medical Imaging: Proc. 13th Int. Conf. (IPMI’93),
Flagstaff, AZ, June, 1993, pp. 48–61, Lectures Notes in Computer
Science. Springer–Verlag.
Young, A. and Axel, L. (1992). Non-rigid wall motion using
MR tagging. In Proc. Conf. on Computer Vision and Pattern
Recognition (CVPR’92), Urbana, IL, June, 1992, pp. 399–404,
Los Alamitos, CA: IEEE Computer Society Press.
Young, A. A., Axel, L., Dougherty, L., Bogen, D. K. and Parenteau,
C. S. (1993). Validation of tagging with MR imaging to estimate
material deformation. Radiology, 188, 101–108.
Young, A. A., Kraitchman, D. L., Dougherty, L. and Axel, L. (1995).
Tracking and finite element analysis of stripe deformation in
magnetic resonance tagging. IEEE Trans. Med. Imag., 14(3),
413–421.
Yuille, A. L., Hallinan, P. W. and Cohen, D. S. (1992). Feature
extraction from faces using deformable templates. Int. J. Comput.
Vision, 8, 99–111.
Zienkiewicz, O. C. and Taylor, R. L. (1989). The Finite Element
Method. New York, NY: McGraw–Hill.