14
Max-flow segmentation of the left ventricle by recovering subject-specific distributions via a bound of the Bhattacharyya measure Ismail Ben Ayed a,b,, Hua-mei Chen b , Kumaradevan Punithakumar a , Ian Ross b , Shuo Li a,b a GE Healthcare, London, ON, Canada b University of Western Ontario, London, ON, Canada article info Article history: Received 17 May 2010 Received in revised form 11 May 2011 Accepted 12 May 2011 Available online 26 May 2011 Keywords: Left ventricle segmentation Bhattacharyya measure Max-flow optimization Cardiac magnetic resonance images (cardiac MRI) abstract This study investigates fast detection of the left ventricle (LV) endo- and epicardium boundaries in a car- diac magnetic resonance (MR) sequence following the optimization of two original discrete cost func- tions, each containing global intensity and geometry constraints based on the Bhattacharyya similarity. The cost functions and the corresponding max-flow optimization built upon an original bound of the Bhattacharyya measure yield competitive results in nearly real-time. Within each frame, the algorithm seeks the LV cavity and myocardium regions consistent with subject-specific model distributions learned from the first frame in the sequence. Based on global rather than pixel-wise information, the proposed formulation relaxes the need of a large training set and optimization with respect to geometric transfor- mations. Different from related active contour methods, it does not require a large number of iterative updates of the segmentation and the corresponding computationally onerous kernel density estimates (KDEs). The algorithm requires very few iterations and KDEs to converge. Furthermore, the proposed bound can be used for several other applications and, therefore, can lead to segmentation algorithms which share the flexibility of active contours and computational advantages of max-flow optimization. Quantitative evaluations over 2280 images acquired from 20 subjects demonstrated that the results cor- relate well with independent manual segmentations by an expert. Moreover, comparisons with a related recent active contour method showed that the proposed framework brings significant improvements in regard to accuracy and computational efficiency. Ó 2011 Elsevier B.V. All rights reserved. 1. Introduction The standardized left ventricle (LV) segmentation in Cerqueira et al. (2002) prescribes the use of representative 2D cardiac slices to generate 17 standardized LV segments relevant to the clinical assessment of regional (localized) heart motion abnormalities. Such standard 2D segments are commonly used for regional anal- ysis and quantification of the LV function. An essential component in the diagnosis of cardiovascular diseases related to localized regions with movement abnormalities, obtaining these segments requires accurate detection of the LV endo- and epicardium bound- aries in cardiac Magnetic Resonance (MR) sequences (Zhu et al., 2010; Ben Ayed et al., 2009a; Spottiswoode et al., 2009; Hautvast et al., 2006). The problem amounts to segmenting each frame into three target regions: the LV cavity, myocardium, and background (refer to the examples in Fig. 4). Manual delineation in all the images of a subject is prohibitively time-consuming and, as such, automatic or semi-automatic algorithms are highly desired. Albeit an impressive research effort has been devoted to this problem, current algorithms are still not sufficiently fast and flexible for rou- tine clinical use, mainly because of the challenges inherent to MR cardiac images (Jolly, 2008). For instance, cardiac regions, such as the papillary muscles and the myocardium, are connected and have almost the same intensity profile (cf. the typical example in Fig. 3). Furthermore, appropriate intensity and geometry models of the LV are hard to learn from a finite training set (Ben Ayed et al., 2009a; Hautvast et al., 2006; Jolly, 2008) given the substan- tial variations in the LV shape and intensity between subjects, par- ticularly those with pathological patterns. Current algorithms are based, among others, on active con- tours (Ben Ayed et al., 2009a,c; Hautvast et al., 2006; Fradkin et al., 2008; Lynch et al., 2008; Pluempitiwiriyawej et al., 2005; Sun et al., 2005; El-Berbari et al., 2007; Kaus et al., 2004; Fritscher et al., 2005), active appearance and shape models (Andreopoulos and Tsotsos, 2008; Mitchell et al., 2002; Mitchell et al., 2001; Zambal et al., 2006; Van Assen et al., 2008), classification using probabilistic atlases (Lorenzo-Valdés et al., 2004), graph cuts (Zhu-Jacquot and Zabih, 2008; Ben Ayed et al., 2009), Bayesian fil- 1361-8415/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.media.2011.05.009 Corresponding author at: GE Healthcare, London, ON, Canada. E-mail address: [email protected] (I. Ben Ayed). Medical Image Analysis 16 (2012) 87–100 Contents lists available at ScienceDirect Medical Image Analysis journal homepage: www.elsevier.com/locate/media

Medical Image Analysisprofs.etsmtl.ca/ibenayed/Ismail-Media-12.pdf · (2009a,c), Hautvast et al. (2006) build subject-specific models from a user-provided segmentation of one frame

  • Upload
    phamtu

  • View
    215

  • Download
    0

Embed Size (px)

Citation preview

Medical Image Analysis 16 (2012) 87–100

Contents lists available at ScienceDirect

Medical Image Analysis

journal homepage: www.elsevier .com/locate /media

Max-flow segmentation of the left ventricle by recovering subject-specificdistributions via a bound of the Bhattacharyya measure

Ismail Ben Ayed a,b,⇑, Hua-mei Chen b, Kumaradevan Punithakumar a, Ian Ross b, Shuo Li a,b

a GE Healthcare, London, ON, Canadab University of Western Ontario, London, ON, Canada

a r t i c l e i n f o a b s t r a c t

Article history:Received 17 May 2010Received in revised form 11 May 2011Accepted 12 May 2011Available online 26 May 2011

Keywords:Left ventricle segmentationBhattacharyya measureMax-flow optimizationCardiac magnetic resonance images(cardiac MRI)

1361-8415/$ - see front matter � 2011 Elsevier B.V. Adoi:10.1016/j.media.2011.05.009

⇑ Corresponding author at: GE Healthcare, London,E-mail address: [email protected] (I. Ben Ay

This study investigates fast detection of the left ventricle (LV) endo- and epicardium boundaries in a car-diac magnetic resonance (MR) sequence following the optimization of two original discrete cost func-tions, each containing global intensity and geometry constraints based on the Bhattacharyya similarity.The cost functions and the corresponding max-flow optimization built upon an original bound of theBhattacharyya measure yield competitive results in nearly real-time. Within each frame, the algorithmseeks the LV cavity and myocardium regions consistent with subject-specific model distributions learnedfrom the first frame in the sequence. Based on global rather than pixel-wise information, the proposedformulation relaxes the need of a large training set and optimization with respect to geometric transfor-mations. Different from related active contour methods, it does not require a large number of iterativeupdates of the segmentation and the corresponding computationally onerous kernel density estimates(KDEs). The algorithm requires very few iterations and KDEs to converge. Furthermore, the proposedbound can be used for several other applications and, therefore, can lead to segmentation algorithmswhich share the flexibility of active contours and computational advantages of max-flow optimization.Quantitative evaluations over 2280 images acquired from 20 subjects demonstrated that the results cor-relate well with independent manual segmentations by an expert. Moreover, comparisons with a relatedrecent active contour method showed that the proposed framework brings significant improvements inregard to accuracy and computational efficiency.

� 2011 Elsevier B.V. All rights reserved.

1. Introduction

The standardized left ventricle (LV) segmentation in Cerqueiraet al. (2002) prescribes the use of representative 2D cardiac slicesto generate 17 standardized LV segments relevant to the clinicalassessment of regional (localized) heart motion abnormalities.Such standard 2D segments are commonly used for regional anal-ysis and quantification of the LV function. An essential componentin the diagnosis of cardiovascular diseases related to localizedregions with movement abnormalities, obtaining these segmentsrequires accurate detection of the LV endo- and epicardium bound-aries in cardiac Magnetic Resonance (MR) sequences (Zhu et al.,2010; Ben Ayed et al., 2009a; Spottiswoode et al., 2009; Hautvastet al., 2006). The problem amounts to segmenting each frame intothree target regions: the LV cavity, myocardium, and background(refer to the examples in Fig. 4). Manual delineation in all theimages of a subject is prohibitively time-consuming and, as such,

ll rights reserved.

ON, Canada.ed).

automatic or semi-automatic algorithms are highly desired. Albeitan impressive research effort has been devoted to this problem,current algorithms are still not sufficiently fast and flexible for rou-tine clinical use, mainly because of the challenges inherent to MRcardiac images (Jolly, 2008). For instance, cardiac regions, such asthe papillary muscles and the myocardium, are connected andhave almost the same intensity profile (cf. the typical example inFig. 3). Furthermore, appropriate intensity and geometry modelsof the LV are hard to learn from a finite training set (Ben Ayedet al., 2009a; Hautvast et al., 2006; Jolly, 2008) given the substan-tial variations in the LV shape and intensity between subjects, par-ticularly those with pathological patterns.

Current algorithms are based, among others, on active con-tours (Ben Ayed et al., 2009a,c; Hautvast et al., 2006; Fradkinet al., 2008; Lynch et al., 2008; Pluempitiwiriyawej et al., 2005;Sun et al., 2005; El-Berbari et al., 2007; Kaus et al., 2004; Fritscheret al., 2005), active appearance and shape models (Andreopoulosand Tsotsos, 2008; Mitchell et al., 2002; Mitchell et al., 2001;Zambal et al., 2006; Van Assen et al., 2008), classification usingprobabilistic atlases (Lorenzo-Valdés et al., 2004), graph cuts(Zhu-Jacquot and Zabih, 2008; Ben Ayed et al., 2009), Bayesian fil-

88 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

tering (Punithakumar et al., 2010), and registration (Zhuang et al.,2008). Commonly, the problem is stated as the optimization of acost functional whose solution is obtained following the evolutionof an active contour toward the boundary of the target region.Optimization of active contour functionals is a prevalent and flex-ible choice in medical image analysis because it can introduce awide spectrum of intensity and geometry constraints1 on thesolution (Ben Ayed et al., 2009b; Rousson and Cremers, 2005; Liuet al., 2005). Generally, these constraints reference a sum overthe target region or its boundary of pixel-wise penalties fitting theimage to intensity and geometry models learned from a finite train-ing set. Pixel-wise intensity information cannot distinguish con-nected cardiac regions having almost the same intensity profile(Ben Ayed et al., 2009a; Hautvast et al., 2006), for instance the pap-illary muscles within the cavity and the myocardium. Therefore,most of existing methods bias the solution towards a model ofshapes learned a priori.

Although very effective in some cases, training-based algo-rithms may have difficulty in capturing the substantial subjectvariations in a clinical context (Ben Ayed et al., 2009a; Hautvastet al., 2006; Jolly, 2008). The ensuing results are bounded to thecharacteristics, variability, and mathematical description of thetraining set. For instance, a pathological case outside the trainingset of shapes may not be recovered, and intensity models have tobe updated for new acquisition protocols and sequences.

To relax the dependence on the choice of a training set, the re-cent active curve studies in Zhu et al. (2010), Ben Ayed et al.(2009a,c), Hautvast et al. (2006) build subject-specific modelsfrom a user-provided segmentation of one frame in the currentcardiac sequence. For instance, in Hautvast et al. (2006), theauthors propose to maintain a constant intensity environment inthe vicinity of the cavity boundary propagated over the sequence.Based on a global similarity measure between distributions, themethod in Ben Ayed et al. (2009a) maintains over a cardiacsequence a constant overlap between the intensity distributionsof the cavity and myocardium, which led to promising resultsfor mid-cavity images. In a closely related direction, the authorsin Ben Ayed et al. (2009c) maximize the similarities betweenthe intensity distributions of cardiac regions within consecutiveframes. Based only on the current data, these methods allow moreflexibility in clinical use, although at the price of a user initializa-tion. The LV segmentation methods in Ben Ayed et al. (2009a,c)follow on the effort of several recent studies in the context of gen-eral-purpose segmentation (Ben Ayed et al., 2009b, 2010; Ni et al.,2009; Zhang and Freedman, 2005; Freedman and Zhang, 2004;Aubert et al., 2003; Michailovich et al., 2007; Georgiou et al.,2007; Rother et al., 2006), which have shown that the use of globaldistribution measures outperforms standard techniques based onpixel-wise information, and is less sensitive to inaccuracies inestimating the models (Ben Ayed et al., 2009b; Michailovichet al., 2007). As such, it can relax the need of a large trainingset. Unfortunately, optimization of a global distribution measurewith respect to segmentation is NP-hard (Rother et al., 2006),and the problem has been commonly addressed with active con-tour optimization via partial differential equations (Ben Ayedet al., 2009b; Ni et al., 2009; Zhang and Freedman, 2005;Freedman and Zhang, 2004; Aubert et al., 2003; Michailovichet al., 2007; Georgiou et al., 2007). A gradient flow equation ofcontour evolution is derived in order to increase the similaritybetween the region within the contour and a given model (Zhangand Freedman, 2005; Freedman and Zhang, 2004; Aubert et al.,2003), or to decrease the similarity between the segmentation

1 Geometry constraints reference object shape, position, and size.

regions defined by the interior and exterior of the contour(Michailovich et al., 2007; Georgiou et al., 2007).

Several measures were studied within the active contour frame-work, for instance the Kullback–Leibler divergence (Aubert et al.,2003; Zhang and Freedman, 2005; Freedman and Zhang, 2004),the Earth Mover’s Distance (Adam et al., 2009), and the Bhattachar-yya coefficient (Zhang and Freedman, 2005; Freedman and Zhang,2004; Ben Ayed et al., 2009b; Michailovich et al., 2007). However,the latter has shown superior performances over other criteria(Zhang and Freedman, 2005; Michailovich et al., 2007). The Bhatta-charyya coefficient has a clear geometric interpretation (Comaniciuet al., 2003), outstanding theoretical properties which were wellstudied in information theory (Aherne et al., 1997), and a fixed(normalized) range which affords a conveniently practical apprai-sal of the similarity.

These methods based on distribution measures lead to compu-tationally intensive algorithms. Along with an incremental contourevolution, they require a large number of updates of computation-ally onerous integrals, namely, the distributions of the regions de-fined by the contour at each iteration and the correspondingmeasures. Active contour methods rely on stepwise gradient des-cent. As a result, the ensuing algorithms are notoriously slow, con-verge to a local minimum, and depend on the choice of anapproximating numerical scheme of contour evolution and the cor-responding parameters.

In this study, we state the detection of the LV endo- and epicar-dium boundaries in a cardiac MR sequence as the minimizationwith respect to a binary variable (labeling) of two original discretecost functions based on the Bhattacharyya measure, each contain-ing global intensity and geometry constraints. The global con-straints measure a similarity between the distributions of the LVcavity and myocardium regions in the current frame and subject-specific models learned from the first frame in the sequence. Aswe will see in the experiments, the global constraints are more rel-evant than pixel-wise ones in the context of cardiac regions. Usedin conjunction with regularization terms for smooth segmentationboundaries and a hard constraint to ensure the cavity is enclosedwithin the myocardium, the global constraints yield competitiveresults over 2280 cardiac images acquired from 20 subjects.However, they do not afford an analytical form amenable to fastmax-flow optimization because they do not reference pixel orpixel-neighborhood penalties. They evaluates a global similaritymeasure between distributions and, therefore, the ensuing optimi-zation problem is challenging and NP-hard. To solve efficiently theproblem, we first propose an original bound of the Bhattacharyyameasure by introducing an auxiliary labeling. From this bound,we reformulate the problem as the optimization of auxiliary func-tions by max-flow iterations, thereby obtaining a nearly real-timesegmentation algorithm. Then, we demonstrate that the proposedprocedure converges. The contributions of this study are not onlyin the application context but also in the scope of general-purposesegmentation:

1.1. Contributions in the application context

The proposed formulation removes the need of a large trainingset, handles intrinsically geometric variations of the LV withoutbiasing the solution towards a set of template shapes, relaxes opti-mization over geometric transformations, and prevents the papil-lary muscles from being included erroneously in the myocardium.

1.2. Contributions in the general-purpose context

Variables which are global over the segmentation regions havebeen generally avoided in the context of max-flow optimization(Boykov and Funka Lea, 2006; Rother et al., 2004; Kohli et al.,

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 89

2008; Blake et al., 2004). The proposed upper bound affords a po-tent global description at a very low computational cost. It requiresvery few max-flow iterations and KDEs to converge, unlike activecontours.

Although our cost functions contain constraints that are de-signed for cardiac image segmentation, the proposed bound opti-mization can be applied to any cost function in the form of theBhattacharyya similarity measure. Therefore, it can be used inother applications and can lead to segmentation algorithms whichshare the flexibility of active contours and computational advanta-ges of max-flow optimization.

The remainder of this paper is organized as follows. The nextsection details the proposed cost functions and the correspondingbound optimization. In Section 3, we report a quantitative andcomparative performance evaluation using several criteria alongwith the computation time/load, describe a typical example whichillustrates explicitly the relevance of global measures for cardiacimage segmentation, and give a representative sample of the re-sults for visual inspection. Section 4 contains a conclusion.

2. Formulation

Consider a MR cardiac sequence containing N image functions2

InðpÞ ¼ Inp : P � R2 ! I � R; n 2 ½1 . . . N�;

with P the image domain and I the set of intensity variables. Thepurpose of this study is to partition the domain P of each frame ninto three regions (cf. the examples in Fig. 4):

Cn : the heart cavityMn : the myocardium

P n ðCn [MnÞ : the background

We state the problem as the successive minimization with respectto a binary variable (labeling) of two original discrete cost functionsdesigned to address the problems related to cardiac MR images,each containing two kernel density matching terms, one intensity-based and the other distance-based. The cost functions are usedin conjunctions with regularization terms for smooth partitionboundaries and a hard constraint to ensure the cavity region is en-closed within the myocardium. Minimization of the first cost func-tion yields for each frame n an optimal labeling Ln

CðpÞ : P ! f0;1gwhich defines the heart cavity

Cn ¼ fp 2 P=LnCðpÞ ¼ 1g; ð1Þ

whereas the labeling minimizing the second cost function,Ln

M[CðpÞ : P ! f0;1g, defines the myocardium

Mn ¼ fp 2 P=LnM[CðpÞ ¼ 1 and Ln

CðpÞ ¼ 0g ð2Þ

Table 1Summary of the general notations for any labeling LðpÞ ¼ Lp : P ! f0;1g, any imageJ : P ! J , and any set of variables J .

2.1. General notations and definitions

We first consider several notations which will be used repeat-edly in the definition of the cost functions. For any labelingLðpÞ ¼ Lp : P ! f0;1g, any image J : P ! J , and any set of vari-ables J , we have:

� RL1 and RL0 are the complementary regions defined by

Notation Description

RL1 Region fp 2 P=LðpÞ ¼ 1gRL Region fp 2 P=LðpÞ ¼ 0g

2 The

RL1 ¼ fp 2 P=LðpÞ ¼ 1gRL0 ¼ fp 2 P=LðpÞ ¼ 0g ¼ P n RL1 ð3Þ

0

PL;J Distribution of J within RL1

number of frames N is typically equal to 20 or 25.

� PL;J is the kernel density estimate (KDE) of the distribution ofimage data J within region RL1 :

A(R)BJ ðf

8j 2 J ; PL;JðjÞ ¼P

p2RL1KjðJpÞ

AðRL1 Þ; ð4Þ

where A(R) denotes the number of pixels within a region R andKj the Gaussian kernel (r is the width of the kernel):

KjðJpÞ ¼1ffiffiffiffiffiffiffiffiffiffiffiffi

2pr2p exp�

ðj�JpÞ2

2r2 ; ð5Þ

The Gaussian kernel is commonly used in density estimation(Bishop, 2007). The value of r controls the smoothness of thedistribution estimate. For example, when r � 0, the Gaussiankernel in (4) approaches the Dirac function, which yields the dis-crete normalized histogram as estimate. r > 0 corresponds to asmooth, continuous distribution estimate (the higher sigma,the smoother the estimate).� BJ ðf ; gÞ is the Bhattacharyya coefficient measuring the amount

of overlap (similarity) between two distributions f and g:

BJ ðf ; gÞ ¼Xj2J

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffif ðjÞgðjÞ

qð6Þ

BJ ðf ; gÞ has a clear geometric interpretation (Comaniciu et al.,2003); it corresponds to the cosine of the angle between thevectors ðf ðjÞ; j 2 J ÞT and ðgðjÞ; j 2 J ÞT . Thus, it enforces the condi-tion

Pj2J f ðjÞ ¼ 1 and

Pj2J gðjÞ ¼ 1, thereby considering explic-

itly f and g as distributions. The range of the Bhattacharyyacoefficient is [0;1], 0 corresponding to no overlap between thedistributions and 1 to a perfect match. Such fixed (normalized)range affords a conveniently practical appraisal of the similarity.Table 1 summarizes the general notations for any labelingLðpÞ ¼ Lp : P ! f0;1g, any image J : P ! J , and any set ofvariables J .

2.2. The cavity-detection cost function

We assume that a detection of the cavity in first frame I1, i.e., alabeling L1

C defining a partition fC1;P n C1g, is given. Using priorinformation from frame I1 and labeling L1

C, the intensity and geom-etry model distributions of the cavity are learned and embedded inthe following distribution matching constraints to segment subse-quent frames.

2.2.1. Intensity-matching constraintGiven the learned model distribution of intensity, denotedMC;I ,

i.e.,

MC;I ¼ PL1C ;I

1 ; ð7Þ

the purpose of this term is to find, for each subsequent frame In,n 2 [2. . .N], a region Cn whose intensity distribution most closelymatches MC;I . To this end, we minimize the following intensitymatching function with respect to L:

The number of pixels within R; gÞ P

j2Jffiffiffiffiffiffiffiffiffiffiffiffiffiffiffif ðjÞgðjÞ

p2 ½0; 1�

90 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

LnC ¼ arg min

L:P!f0;1gBC;InðLÞ ¼ �BIðPL;In ;MC;I Þ

¼ �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPL;In ðiÞMC;I ðiÞ

qð8Þ

2.2.2. Distance-matching constraintThe purpose of this term is to constrain the segmentation with

prior geometric information (shape, scale, and position of the cav-ity) obtained from the learning frame. Let Oc be the centroid of cav-ity C1 in the learning frame and DcðpÞ ¼ Dc

p ¼kp�Ock

NDc: P ! D a

distance image measuring the normalized distance between p andOc, with D the space of distance variables and NDc a normalizationconstant which is computed systematically in order to restrict thevalues of image Dc within interval [0;1]. LetMC;D be the model dis-tribution of distances within the cavity in the learning frame:

MC;D ¼ PL1C ;D

c ð9Þ

We propose to find a region Cn whose distance distribution mostclosely matches MC;D by solving the following minimizationproblem:

LnC ¼ arg min

L:P!f0;1gBC;DðLÞ ¼ �BDðPL;Dc ;MC;DÞ

¼ �Xd2D

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPL;Dc ðdÞMC;DðdÞ

qð10Þ

Note that this geometric prior is invariant to rotation, and embedsimplicitly uncertainties with respect to scale via the kernel widthr in (5). The higher r, the more scale variations allowed. In ourexperiments, r = 2 was sufficient to handle effectively variationsin the scale of the cavity (cf. the examples in Fig. 4). Based on globalrather pixel-wise information, the proposed geometric prior relaxes(1) learning/modeling of geometric characteristics over a largetraining set and (2) explicit optimization with respect to geometrictransformations, unlike existing shape priors. It is worth noting thatsuch invariance can be achieved with other descriptions, e.g., invari-ant moments (Flusser et al., 2009). However, optimizing an invari-ant moment with respect to segmentation is not amenable to fastsolvers. It may result in difficult, computationally expensive optimi-zation problems.

2.2.3. Total cost functionWe propose to minimize a cost function containing the inten-

sity and distance matching terms as well as a regularization termfor smooth segmentation boundaries. For each n 2 [2. . .N], thealgorithm computes the optimal labeling Ln

C minimizing the fol-lowing discrete cost function over all L : P ! f0;1g:

LnC ¼ arg min

L:P!f0;1gF C;InðLÞ

¼ BC;InðLÞ|fflfflfflffl{zfflfflfflffl}Intensity Matching

þ cBC;DðLÞ|fflfflfflfflffl{zfflfflfflfflffl}Geometry Matching

þ kSðLÞ|fflffl{zfflffl}Smoothness

ð11Þ

where SðLÞ is a smoothness (regularization) term which ensures la-bel consistency of neighboring pixels (Boykov and Kolmogorov,2003)

SðLÞ ¼Xfp;qg2N

1kp� qk dLðpÞ–LðqÞ ð12Þ

with

dx–y ¼1 if x–y

0 if x ¼ y

�ð13Þ

and N is a 4-neighborhood system containing all unordered pairs{p,q} of neighboring elements of P. c and k are positive constantsbalancing the relative contribution of each term.

2.3. Efficient max-flow optimization via an upper bound of theBhattacharyya measure

The global terms BC;In ðLÞ and BC;DðLÞ in the cost function in Eq.(11) are not directly amenable to max-flow optimization becausethey do not reference pixel or pixel-neighborhood penalties. Theyevaluates a global similarity measure between distributions and,therefore, the ensuing optimization is a challenging and NP-hardproblem. To optimize efficiently these terms, we first propose anoriginal bound of the Bhattacharyya measure by introducing anauxiliary labeling. From this bound, we reformulate the problemas the optimization of auxiliary functions by max-flow iterations.Then, we formally demonstrate that the proposed proceduresconverge.

2.3.1. Upper boundsTo introduce our formulation, let us first introduce the follow-

ing proposition:

Proposition 1. Given a fixed (auxiliary) labeling La, for any labelingL verifying RL1 � RL

a

1 , i.e., the foreground region defined by L is withinthe foreground region defined by La, and "a 2 [0,1], we have thefollowing upper bound of BC;In ðLÞ

BC;In ðLÞ 6 J C;In ðL;La;aÞ ¼Xp2RL0

cp;Inð0Þ þ ð1� aÞXp2RL1

cp;In ð1Þ; ð14Þ

with cp;In ð0Þ and cp;In ð1Þ given for each p in P by

cp;Inð0Þ ¼ Lap

A RLa

1ð Þ BC;InðLaÞ þPi2I

Ki Inp

� � ffiffiffiffiffiffiffiffiffiffiffiffiMC;I ðiÞPLa ;In ðiÞ

r� �cp;Inð1Þ ¼ BC;In ðL

aÞAðRLa

1 Þ

8>><>>: ð15Þ

Similarly, we have the following upper bound of BC;DðLÞ

BC;DðLÞ 6 J C;DðL;La;aÞ ¼Xp2RL0

cp;Dð0Þ þ ð1� aÞXp2RL1

cp;Dð1Þ ð16Þ

with cp,D(0) and cp,D(1) given for each p in P by

cp;Dð0Þ ¼La

p

A RLa

1ð Þ BC;DðLaÞ þPd2D

KdðDcpÞ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiMC;DðdÞPLa ;Dc ðdÞ

r� �cp;Dð1Þ ¼ BC;DðLaÞ

AðRLa1 Þ

8>><>>: ð17Þ

Proof of proposition 1. We give a proof of Proposition 1 in Appen-dix A.

Definition 1. AðL; bLÞ is called auxiliary function of a given costfunction FðLÞ if it satisfies the following conditions:

FðLÞ 6 AðL; bLÞ ð18ÞAðL;LÞ ¼ FðLÞ ð19Þ

Auxiliary functions are commonly used in the Nonnegative MatrixFactorization (NMF) literature for optimization (Lee and Seung,2000). Rather than optimizing the cost function, one can optimizeiteratively an auxiliary function of the cost function. At each itera-tion t, this amounts at optimizing over the first variable

Lðtþ1Þ ¼ arg minLAðL;LðtÞÞ ð20Þ

Thus, by definition of auxiliary function and minimum, we obtainthe following monotonically decreasing sequence of the costfunction

FðLðtÞÞ ¼ AðLðtÞ;LðtÞÞP AðLðtþ1Þ;LðtÞÞP FðLðtþ1ÞÞ ð21Þ

Table 2a

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 91

Proposition 2. For a = 0, the following function is an auxiliary func-

Weights of the graph edges for optimizing auxiliary function AðL;L ;aÞ. The last termin wp;Tc is a hard constraint; it is a positive constant K used for each pixel outside RL

a

1

to ensure the solution L verifies RL1 � RLa

1 . A large K enforces the conditionwp;Tc > wp;Sc . This means that any pixel outside RL

a

1 remains connected to terminal

tion of F C;In ðLÞ

AC;InðL;La;aÞ ¼ J C;InðL;La;aÞ þ cJ C;DðL;La;aÞ þ kSðLÞ ð22Þ

Tc (refer to Fig. 1c) and, therefore, remains within the complement of the cavityregion.

Edge Weight (cost)

wp;Sc cp;In ð0Þ þ ccp;Dð0Þwp;Tc ð1� aÞðcp;In ð1Þ þ ccp;Dð1ÞÞ þ K 1� La

p

� �wp,q

kkp�qk

Proof. We give a proof of Proposition 2 in Appendix B. h

Proposition 2 instructs us to consider the following procedureMinimization of F C;In :

begin� Initialize the auxiliary labeling La

� Initialize a:a = a0 with 0 < a0 < 1repeat

1. Optimize the auxiliary function over L

Fig. 1. Illuthe other

LðtÞ ¼ arg minL:RL1�RL

a1

AC;InðL;La;aÞ

2. Update La by La ¼ LðtÞ3. Decrease a:a = aq with q > 1

until Conversgence;

The optimal labeling LnC is given by LðtÞ at convergence. This

optimal labeling defines the cavity in frame n (region Cn) accordingto Eq. (1).

Convergence proof: When a approaches zero, AC;In ðL;La;aÞ ap-proaches an auxiliary function of cost function F C;In and, therefore,the above procedure leads to a monotonically decreasing sequenceof F C;In . This comes directly from (21). Since the cost function islower bounded (because the Bhattacharyya measure is upperbounded by one), the above minimization procedure converges.

2.3.2. Max-flow optimizationNow notice that the auxiliary function AC;In in step 1 of the min-

imization of F C;In is the sum of unary and pairwise (sub-modular)penalties. In combinatorial optimization, a global optimum of suchsum can be computed efficiently in low-order polynomial time bysolving an equivalent max-flow problem (Boykov and Kolmogorov,2004). Furthermore, the condition that a solution L should verifyRL1 � RL

a

1 can be imposed easily by adding a hard constraint(Boykov and Funka Lea, 2006).

To optimize the auxiliary function, it suffices to build a weightedgraph (an illustration of the graph is depicted in Fig. 1a)

(a)stration of the min-cut/max-flow optimization: the minimum cut yields a partito the sink.

G ¼ hN;Ei; ð23Þ

where N is the set of vertices (nodes) and E the set of edges connect-ing these nodes. N contains a node for each pixel p 2 P and twoadditional nodes called terminals, one source Sc representing thecavity region Cn and one sink (Tc) representing the complement ofthe cavity, P n Cn. Let wp,q be the weight of the edge connectingneighboring pixels {p,q} in N , and wp;Sc and wp;Tc the weights ofthe edges connecting each pixel p to source Sc and sink Tc,respectively.

The edge weights (costs) in Table 2 correspond to the optimiza-tion of AC;In . Choosing these weights and following the max-flowalgorithm of Boykov and Kolmogorov (2004), we compute a mini-mum cut of G, i.e., a subset of edges in E whose sum of edgeweights is minimal and whose removal divides the graph intotwo disconnected subgraphs, each containing a terminal node (re-fer to Fig. 1b and c for an illustration). This minimum cut, whichassigns each node (pixel) p in P to one of the two terminals, in-duces at each iteration t an optimal labeling Lt which minimizesglobally the auxiliary function.

As illustrated in Fig. 1, we are using the Boykov–Kolmogorovmax-flow algorithm (Boykov and Kolmogorov, 2004) in the caseof 2D grids with a 4-neighborhood system. In this case, it is wellknown that the algorithm yields a state-of-the-art efficiency(Boykov and Kolmogorov, 2004). However, as noted in Boykovand Kolmogorov (2004), Juan and Boykov (2007), the efficiencyof the algorithm decreases when moving from 2D to 3D grids orwhen using denser (larger neighborhood) grids.

Interpretation of the minimum cut (min-cut/max-flow): Let uswrite the auxiliary function in terms of the graph weights in Table2 (we ignore the smoothness and hard constraints for a clearpresentation):

(b) (c)tion of the image domain into two regions, one connected to the source terminal and

92 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

AC;In ¼Xp2RL0

wp;Sc þXp2RL1

wp;Tc ð24Þ

For each pixel p, we have two cases:

Case 1: wp;Tc < wp;Sc (refer to the illustration in Fig. 1b)In this case, the minimum cut removes the edge with the lower

weight wp;Tc , i.e., the edge connecting p and Tc, thereby includingpixel p in the cavity region RL1 (i.e., foreground). This makes sensebecause including p in RL1 adds the lower weight to the expressionin (24) and, therefore, favors the minimization of function AC;In .

Case 2: wp;Sc < wp;Tc (refer to the illustration in Fig. 1c)

In this case, the minimum cut removes the edge connecting pand Sc, thereby excluding pixel p from the cavity region.

A statistical hypothesis testing interpretation of coefficientscp;In ð0Þ; cp;In ð1Þ; cp;Dð0Þcp;Dð1Þ:

In the following, we examine the link between the bound opti-mization and the classical theory of hypothesis testing (Lehmann,1986). To simplify the interpretation, let us omit the effect of thesmoothness, geometry-matching and hard constraints (i.e.,c = k = K = 0). At the first iteration (t = 0), assume that the cavity(foreground) region contains all the pixels in the image. Then,the algorithm performs at each iteration t the following statisticaltest for each pixel p within the current cavity regionðp 2 RL

t�1

1 ¼ RLa

1 Þ:

Hp ¼ wp;Sc �wp;Tc ¼ cp;In ð0Þ � ð1� aÞcp;Inð1Þ

¼ 1

A RLa

1

� �ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiMC;I In

p

� �PLa ;In In

p

� �vuuut þ a

BC;In ðLaÞAðRL

a

1 Þ

¼ 1

A RLa

1

� �ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiMC;I In

p

� �PLa ;In ðIn

vuut � aBIðPL;In ;MC;I Þ

A RLa

1

� � ð25Þ

The evaluation of the sign of Hp has a clear meaning, and amountsto a statistical hypothesis testing by an image likelihood ratio test. Itevaluates the hypotheses that the image at pixel p is drawn fromlearned modelMC;I or from the image distribution within the cur-rent foreground region. We have two cases:

Case 1: Hp < 0, i.e., the likelihood ratio is lower than the follow-ing critical value:

MC;I Inp

� �PLa ;In In

p

� � < ðaBIðPL;In ;MC;I ÞÞ2 ð26Þ

Fig. 2. Illustration of the hard constraint: minimization of the area of intersectionbetween Cn and RL0 ensures that the cavity region is enclosed within themyocardium.

In this case, we have cp;In ð0Þ < ð1� aÞcp;In ð1Þ, i.e., wp;Sc < wp;Tc .Therefore, the graph cut excludes pixel p from the current fore-ground region (refer to Fig. 1c) so as to decrease the auxiliary func-tion. This makes sense because it results in decreasing the imagedistribution within the current cavity region at value Ip, whichmeans a better match with the model at that value.

Case 2: Hp > 0In this case pixel p is kept within the cavity region becauseexcluding it would increase the discrepancy between PLa ;In

and model MC;I at value Inp .

Interpretation of a: Parameter a controls the critical value in theright-hand side of (26). The higher a, the more pixels excludedfrom the current foreground region and, therefore, the faster theevolution of the labeling towards the final segmentation. Thereforea can be viewed as an algorithmic time which controls the conver-gence speed.

2.4. The myocardium-detection cost function

Adopting a similar notation as previously and assuming that adetection of the myocardium in first frame I1, i.e., a labeling L1

M[C

defining a partition fM1 [ C1;P nM1 [ C1g, is given, we proposeto minimize the following cost function to detect the myocardium:

LnM[C ¼ arg min

L:P!f0;1gFM;InðLÞ

¼ BM;InðLÞ|fflfflfflfflffl{zfflfflfflfflffl}Intensity Matching

þ cBM[C;DðLÞ|fflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflffl}Geometry Matching

þ kSðLÞ|fflffl{zfflffl}Smoothness

þKA RL0 \ RLnC

1

� �|fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl}

Hard constraint

ð27Þ

The first term is an intensity matching term which measures thesimilarity between PL;In and a myocardium model of intensitylearned from the first frame:

BM;In ðLÞ ¼ �BIðPL;In ;MM;I Þ ¼ �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPL;In ðiÞMM;I ðiÞ

qð28Þ

where

MM;I ¼ PL1M ;I

1 with L1M ¼ LM[C1 1� L1

C

ð29Þ

The second term is a distance matching term:

BM[C;DðLÞ ¼ �BDðPL;Dmc ;MMC;DÞ ¼ �Xd2D

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPL;Dmc ðdÞMMC;DðdÞ

qð30Þ

where DmcðpÞ ¼ Dmcp ¼

kp�OmckNDmc

: P ! D is a distance image measuringthe normalized distance between p and Omc, which denotes the cen-troid of region M1 [ C1 in the learning frame.MMC;D is the model ofdistances learned from the first frame:

MMC;D ¼ PL1M[C ;D

mc ð31Þ

The last term is a hard constraint which measures the area of regionRL0 \ RL

nC

1 (K is a positive constant). Minimization of the area of inter-section between Cn and RL0 ensures that the cavity region is en-closed within the myocardium (refer to Fig. 2 for an illustration).The cost function in (27) is minimized by max-flow iterations usingprevious arguments. The obtained optimal labeling, Ln

M[C, definesthe myocardium region in frame n according to Eq. (2).

3. Experimental results and discussions

The evaluation was carried out over 120 short axis cardiac cineMR sequences acquired from 20 subjects: a total of 2280 imageswere automatically segmented, and the results were compared to

Table 3Details of the datasets used in evaluation of the proposed method.

Description Value

Number of subjects 20Scanner protocol FIESTAPatient ages 16–69 yearsShort-axis image resolution (256 � 256) pixelsTemporal resolution 20 volumesPixel spacing 1.17–1.56 mmSlice thickness 8–10 mm

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 93

independent manual segmentations by an expert. For each subject,the algorithm was applied to six slices including apical, mid-cavity,and basal slices. Each slice corresponds to a sequence of 20 frames.Table 3 summarizes the details of the datasets used in evaluation.Using the same datasets, we compared the accuracy andcomputational load/time of the proposed method, referred to asMFM (max-flow method), with the level-set segmentation of theLV in Ben Ayed et al. (2009a), referred to as LSM (level-set method).Similar to LSM (Ben Ayed et al., 2009a), MFM relaxes the need of atraining, and learns model distributions from a user-providedsegmentation of the first frame in each sequence. For a faircomparison, the same presegmented learning frame was used forboth algorithms.

In the following, we first describe a typical example which illus-trates explicitly the relevance of global measures for cardiac imagesegmentation, and give a representative sample of the results forvisual inspection. Then, we report the computation time/load,and describe a quantitative and comparative performance analysisusing several accuracy measures. In this analysis, the parameterswere unchanged for all the datasets, and were fixed as follows:c = 1; k = 0.0012; K = 103; r = 2 for the distance distributions,r = 10 for the intensity distributions of the cavity, and r = 10�8

for the intensity distributions of the myocardium; q = 5; a0 = 0.9for the cavity detection and a0 = 0.8 for the myocardium detection.Finally, we will describe a large number of experiments whichdemonstrate the effect of the weighting parameters (c and k) onthe results of the proposed algorithm.

3.1. Relevance of global measures: a typical example

Fig. 3 depicts a typical example, where the purpose is to find theboundary between the heart cavity and the myocardium. The ex-pert (ground truth) delineation, depicted by the green curve inFig. 3b, includes the papillary muscles within the cavity (region

(a) (b)Fig. 3. A typical example where the purpose is to find the boundary between the hearinterest. (b) The green curve depicts the expert (ground truth) delineation, and the red c(32). (c) MC: the distribution of image data within the ground-truth cavity (region insmyocardium (region outside the green curve). (For interpretation of the references to co

within the green curve). Recovering this expert segmentation isdifficult because the myocardium (region outside the green curve)and the papillary muscles within the cavity are connected and havethe same intensity profile.

The red curve in Fig. 3b depicts the segmentation obtained byminimizing the pixel-wise likelihood energy commonly used inmin-cut segmentation (Boykov and Funka Lea, 2006):

Lopt ¼ arg minL:P!f0;1g

�Xp2RL0

logMMðIpÞ �Xp2RL1

logMCðIpÞ; ð32Þ

where I is the image data;MC andMM are the assumed models ofimage data within the cavity and myocardium, respectively. Thesemodels were learned from the ground truth for illustration pur-poses:MC is the distribution of image data within the ground-truthcavity (region inside the green curve) whereas MM is the distribu-tion of image data within the ground-truth myocardium (regionoutside the green curve). Fig. 3c plots these models. The optimiza-tion in (32) performs the following test at each pixel p:

Case 1: �logMMðIpÞ < �logMCðIpÞ, i.e., MMðIpÞ >MCðIpÞIn this case, the pixel is included in the myocardium region (i.e.,

RL0 ) so as to obtain a lower value of the energy. Therefore, followingthe optimization in (32), all the cavity pixels corresponding to thegray area in Fig. 3c are erroneously included in the myocardium(refer to the red curve in Fig. 3b).

Case 2: �logMMðIpÞ > �logMCðIpÞ, i.e., MMðIpÞ <MCðIpÞIn this case, the pixel is included in the cavity region.

Although the models were learned from the ground truth, opti-

mization of the pixel-wise energy in (32) excluded the papillarymuscles from the cavity, yielding a segmentation different fromthe ground truth. Pixel-wise information cannot distinguishbetween the papillary muscles and the myocardium because thepixels within these two connected regions have almost the sameintensity.

Table 4 illustrates explicitly the relevance of the proposedglobal measure over the pixelwise energy in (32). For each of thesegmentations in Fig. 3b, it reports the corresponding pixel-wiseenergy as well as the global intensity-matching constraint:

�BIðPL;I;MCÞ ¼ �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPL;IðiÞMCðiÞ

qð33Þ

The ground truth segmentation yields a pixel-wise energy higherthan the one corresponding to the red-curve segmentation (referto the second line in Table 4). This indicates that the desiredsegmentation does not correspond to the minimum of the

0 50 100 1500

0.005

0.01

0.015

0.02

0.025

0.03

0.035

Bins

Density

(c)t cavity and the myocardium. (a) A box limiting the segmentation to the region ofurve the segmentation obtained by minimizing the pixel-wise likelihood energy in

ide the green curve); MM: the distribution of image data within the ground-truthlour in this figure legend, the reader is referred to the web version of this article.)

Table 4Relevance of global measures: pixel-wise and global energies corresponding to thesegmentations in Fig. 3b.

Segmentation Green curve (ground truth) Red curve

Global energy �1 �0.86Pixel-wise energy 7.38 � 103 7.25 � 103

94 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

pixel-wise energy in (32). On the contrary, the global energy (33)corresponding to the ground truth is lower than the one correspond-ing to the red-curve segmentation (refer to the first line in Table 4).This demonstrates explicitly that minimization of the proposedintensity-matching constraint is more relevant for cardiac imagesegmentation.

3.2. Visual inspection: a representative sample of the results

In Fig. 4, we give a representative sample of the results for foursubjects, including mid-cavity, apical, and basal slices. The greenand red curves depict the automatic endo- and epicardium bound-aries, respectively. The yellow discontinuous curves depict theground-truth boundaries. The proposed method prevented the

Fig. 4. A representative sample of the results for four subjects. The yellow discontinuouautomatic endo- and epicardium boundaries, respectively. Weighting parameters c = 1 anreader is referred to the web version of this article.)

papillary muscles from being included erroneously in the myocar-dium. This task is challenging (El-Berbari et al., 2007) because thepapillary muscles and the myocardium are connected and have al-most the same intensity (refer to the green curves in the typicalexamples depicting mid-cavity frames in Fig. 4). The shown exam-ples also contain apical frames in the second and fourth row inFig. 4, where it is difficult to segment the cavity because of thesmall size of the structures and moving artifacts. These examplesshow how the method handles implicitly significant variations inthe scale of the cavity, although neither an additional optimizationover geometric transformations nor a large training set arerequired.

3.3. Computation time/load

Although based on global distribution measures, the proposedmethod (MFM) led to nearly real-time segmentation. Running ona 2 GHz machine, it needs 0.14 s to process a frame. Table 5 reportsthe computation time/load for NFM and the level-set method inBen Ayed et al. (2009a) (LSM). The proposed bound leads to a sig-nificant decrease in computation load because it requires only twokernel density estimations (KDEs) and max-flow iterations,

s curves depict the ground-truth boundaries. The green and red curves depict thed k = 0.0012. (For interpretation of the references to colour in this figure legend, the

Table 5Computation time (CPU) and number of iterations and kernel density estimations(KDEs) for the proposed method (MFM) and the level-set method in Ben Ayed et al.(2009a) (LSM). MFM led to nearly real-time segmentation and a significant decreasein the computation load.

Method MFM LSM

CPU/frame (secs) 0.14 4.33CPU/subject (secs) 15.96 494.45Nb of iterations and KDEs/frame 2 300

Table 6Quantitative performance evaluations over 20 subjects (2280 images) for theproposed method (MFM with c = 1 and k = 0.0012) and the method in Ben Ayedet al. (2009a) (LSM). The first two rows: The average RMSE. The second two rows:statistics of the DM expressed as mean ± standard deviation (the higher the DM, thebetter the performance). The last two rows: reliability of the DM ðRð0:80ÞÞ. The higherthe reliability, the better the performance.

Method MFM LSM

Endocardium RMSE 1.60 2.46Epicardium RMSE 1.99 1.89Cavity DM 0.92 ± 0.031 0.88 ± 0.090Myocardium DM 0.82 ± 0.061 0.81 ± 0.10Cavity reliability 1 0.89Myocardium reliability 0.79 0.75

3 DM is always in [0,1]. DM equal to 1 indicates a perfect match between manualand automatic segmentations.

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 95

whereas LSM requires approximately 300 KDEs and active-curveupdates. KDE allows an accurate and flexible description of modeldistributions, but is computationally onerous. Generally, the ensu-ing segmentation algorithms, for instance those based on level-setevolution, compute a large number of updates of kernel densitiesand, therefore, are computationally intensive. MFM requires veryfew updates of kernel densities, thereby allowing an accurate andflexible description with a very low computational load.

3.4. Quantitative and comparative performance evaluations

Table 6 summarizes the similarities between the ground truthand the segmentations obtained with MFM (c = 1 and k = 0.0012)and LSM using three measures: the Root Mean Squared Error(RMSE), the Dice metric (DM), and the correlation coefficient. RMSEis contour-based; it measures a distance between manual andautomatic boundaries. The Dice metric and correlation coefficientare region-based; they measure the similarity between the auto-matically detected and ground-truth regions.

3.4.1. The Root Mean Squared Error (RMSE)We evaluated the RMSE by computing the distances between

corresponding points on the manual and automatic boundaries.The RMSE over N points is given by

RMSE ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1N

XN

i¼1

ðxi � ~xiÞ2 þ ðyi � ~yiÞ2vuut ð34Þ

where ðxi; yiÞ is a point on the automatically detected boundary andð~xi;~yiÞ is the closest point to ðxi; yiÞ on the manually traced boundary.We used 240 points along the boundary, i.e., N = 240. RMSE mea-sures a distance between manual and automatic boundaries. Assuch, the lower RMSE, the better the conformity of the results tothe ground truth. The first two rows in Table 6 report the averageRMSE in pixels over all the data for MFM and LSM. For endocardiumdetection, MFM led to a significant improvement in the accuracyover LSM: MFM yielded an average RMSE equal to 1.60 pixels,whereas LSM yielded an average RMSE equal to 2.46 pixels.

Fig. 5a depicts the average RMSE of endocardium detection as afunction of the time, i.e., throughout the cardiac cycle, for MFM and

LSM. MFM led to a lower curve, which indicates a better confor-mity to the ground truth. For epicardium detection, MFM andLSM led approximately to the same accuracy: MFM yielded anaverage RMSE equal to 1.99, whereas LSM yielded an average RMSEequal to 1.89. Fig. 5b depicts the average RMSE of epicardiumdetection as a function of the time.

3.4.2. The Dice metricWe computed the Dice metric (DM) commonly used to measure

the similarity (overlap) between the automatically detected andground-truth regions (Ben Ayed et al., 2009a; Lynch et al., 2008;Pluempitiwiriyawej et al., 2005). The cavity and myocardium re-gions were evaluated for performance appraisal. Let Va,Vm, andVam be the volumes corresponding to an automatically segmentedregion, the corresponding hand-labeled region, and the intersec-tion between them, respectively. The volume is measured by thesum of areas of the considered region in six slices. Volume mea-surements are expressed as the number of pixels within the region.DM is given by3

DM ¼ 2Vam

Va þ Vmð35Þ

The higher the DM, the better the performance of the algorithm.The second two rows in Table 6 report the DM statistics over all

the data for MFM and LSM. For cavity detection, MFM led to a sig-nificant improvement in region accuracy over LSM: MFM yielded aDM equal to 0.92 ± 0.031 (DM is expressed as mean ± standarddeviation), whereas LSM yielded a DM equal to 0.88 ± 0.090. Notethat an average DM higher than 0.80 indicates an excellent agree-ment with manual segmentations (Pluempitiwiriyawej et al.,2005), and an average DM higher than 0.90 is, generally, difficultto obtain because the small structure of the cavity at the apex de-creases significantly the DM (Lynch et al., 2008). For myocardiumdetection, MFM led to a region accuracy slightly better than LSM:MFM yielded a DM equal to 0.82 ± 0.061, whereas LSM yielded aDM equal to 0.81 ± 0.10.

3.4.3. The reliabilityWe examined quantitatively and comparatively the reliability of

the algorithm by evaluating the reliability function–i.e., the comple-mentary cumulative distribution function (ccdf)–of the obtainedDice metrics, defined for each d 2 [0,1] as the probability of obtain-ing DM higher than d over all volumes:

RðdÞ ¼ PrðDM > dÞ

¼ Number of volumes segmented with DM > dTotal number of volumes

ð36Þ

RðdÞ measures how reliable the algorithm in yielding accuracy d,i.e., a DM higher than d. The higher R, the better the performance.The last two rows in Table 6 reports Rð0:80Þ for MFM and LSM,indicating the proposed algorithm brings 11% improvement in thereliability of cavity detection and 4% improvement in the reliability ofmyocardium detection. For cavity detection, MFM yieldedRð0:80Þ ¼ 1, i.e., an excellent agreement (DM > 0.80) in 100% ofthe cases, whereas LSM achieved 89% of the cases with a similaraccuracy. For myocardium detection, MFM yielded Rð0:80Þ ¼ 0:79,whereas LSM yielded Rð0:80Þ ¼ 0:75.

In Fig. 6a and b, we plotted R as a function of d for the cavityand myocardium, respectively. The proposed algorithm led to reli-ability curves higher than LSM. Fig. 6c and d depict the DM for arepresentative sample of the analyzed volumes for MFM and LSM.

0.4 0.6 0.8 10.4

0.6

0.8

1

1.2

Dice Metric

Rel

iabi

lity

(cav

ity)

MFMLSM

0.4 0.6 0.8 10.4

0.6

0.8

1

1.2

Dice metric

Rel

iabi

lity

(myo

card

ium

) MFMLSM

0 100 200 300

0.9

0.95

1

Volume number

Dic

e m

etric

(cav

ity) LSM

MFM

0 100 200 3000.6

0.7

0.8

0.9

1

Volume number

Dic

e m

etric

(myo

card

ium

)

MFMLSM

(a) (b)

(c) (d)Fig. 6. Region-based comparisons of manual and automatic segmentations of 2280 images (380 volumes) acquired from 20 subjects. (a) and (b): ReliabilityðRðdÞ ¼ PrðDM > dÞÞ for the proposed method (MFM) and the level set method in Ben Ayed et al. (2009a) (LSM). MFM led to reliability curves higher than LSM. (c) and(d): DM in a representative sample of the tested volumes for MFM and LSM. The first column corresponds to the cavity detection, and the second to the myocardiumdetection.

0 10 201

2

3

4

Image number

RM

SE (p

ixel

s)

Endocardium

MFMLSM

0 10 201

2

3

4

Image number

RM

SE (p

ixel

s)

Epicardium

MFMLSM

(a) (b)Fig. 5. Average RMSE (in pixels) over 20 subjects (2280 images) as a function of the time step for the proposed method (MFM) and the method in Ben Ayed et al. (2009a)(MFM): (a) endocardium detection; (b) epicardium detection. The average frame number corresponding to end diastole is approximately equal to 7.

0 2 4 6 8 10100

2

4

6

8

10

10−3Vm (manual)

10−3

Va (a

utom

atic

)

Cavity volumes

datay=x

0 2 4 6 8 100

2

4

6

8

10

10−3Vm (manual)

10−3

Va (a

utom

atic

)

Myocardium volumes

datay=x

(a) (b)Fig. 7. Automatic volumes versus manual volumes for the proposed method (MFM). (a) Cavity detection: the obtained correlation coefficient is 0.99. (b) Myocardiumdetection: the obtained correlation coefficient is 0.81.

96 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

Fig. 8. Effect of the geometry-matching term: The yellow discontinuous curvesdepict the ground truth boundary; (a) The red curve depicts the epicardiumboundary obtained without geometry-matching term (c = 0); (b) The red curvedepicts the epicardium boundary obtained with the geometry-matching term(c = 1); The smoothness weight is fixed for the two experiments (k = 0.00075). (Forinterpretation of the references to colour in this figure legend, the reader is referredto the web version of this article.)

Table 7Quantitative performance evaluations over 20 subjects (2280 images) for MFMwithout geometry-matching term (c = 0) and MFM without smoothness term (k = 0).The first two rows: The average RMSE. The second two rows: statistics of the Dicemetric (DM) expressed as mean ± standard deviation (the higher the DM, the betterthe performance). The last two rows: reliability of the DM ðRð0:80ÞÞ. The higher thereliability, the better the performance.

Method MFM (c = 0) MFM (k = 0)

Endocardium RMSE 2.36 3.69Epicardium RMSE 2.89 2.02Cavity DM 0.76 ± 0.2 0.88 ± 0.07Myocardium DM 0.77 ± 0.12 0.80 ± 0.06Cavity reliability 0.61 0.89Myocardium reliability 0.52 0.61

0 1 2 30.7

0.8

0.9

1

γ

DM

Cavity

(a)Fig. 9. The average Dice metric over all the data as a function

0 1 2

x 10−3

0.7

0.8

0.9

1

λ

DM

Cavity

(a)Fig. 10. The average Dice metric over all the data as a fun

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 97

3.4.4. The correlation coefficientThe proposed method yielded high correlation coefficients be-

tween manual and automatic volumes. The correlation coefficientis equal to 0.99 for cavity detection, and is equal to 0.81 for myo-cardium detection. The linear regression plots in Fig. 7 illustratethis high correlation. For all the data analyzed (380 volumes), it de-picts the volumes obtained with the proposed method versus man-ual volumes, along with the identity line. Most of the data pointsare very close to the identity line, which illustrates small differ-ences between manual and automatic segmentations.

3.5. Effect of the choice of the weighting parameters

In this section, we examine the role of the different terms in thecost function, and evaluate quantitatively the robustness of theproposed algorithm with respect to the choice of the weightingparameters. First, we start with a typical example which demon-strates explicitly the positive effect of the geometry-matchingterm. Then, we examine the performance of the algorithm withoutsmoothness (k = 0) or without geometry-matching (c = 0). Finally,we describe extensive experiments which evaluate the perfor-mance of the algorithm as a function of the parameters (Figs. 9and 10).

The example in Fig. 8 demonstrates explicitly the effect of thegeometry-matching term. The yellow discontinuous curve depictsthe ground truth epicardium boundary. The red curve in (a) depictsthe epicardium boundary obtained without geometry-matchingterm (c = 0). In this case, a part of myocardium region was ex-cluded erroneously from the solution. On the contrary, addingthe geometry-matching term (c = 1) biased the solution towardsthe ground-truth (refer to Fig. 8b).

Table 7 reports the performance of the proposed algorithm withc = 0. Removing the geometry-matching constraint affected signif-icantly all the performance measures. Table 7 also reports theperformance of the proposed algorithm without smoothness

0 1 2 30.7

0.8

0.9

1

γ

DM

Myocardium

(b)of c (the weight of the geometry-matching constraint).

0 1 2

x 10−3

0.7

0.8

0.9

1

λ

DM

Myocardium

(b)ction of k (the weight of the smoothness constraint).

98 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

constraint (k = 0). Removing the smoothness constraint did notchange significantly the average DM. However, it affected the RMSEand Reliability.

Fig. 9a and b plot the average Dice metric as a function of c (theweight of the geometry-matching constraint) for the cavity andmyocardium, respectively. We run the algorithm over 16 uni-formly-spaced values of c in the interval [0;3]. The values of c in[0.5;1.5] yielded a Dice metric above 0.9. Therefore, this parameterdoes not require a fine tuning. It is interesting to notice that thebest performance corresponds to c = 1. This makes sense becausethe intensity and geometry-matching constraints have the sameform of a Bhattacharyya measure and, therefore, the same rangeof values.

Fig. 10a and b plot the average Dice metric as a function of k(the weight of the smoothness constraint) for the cavity and myo-cardium, respectively. We run the algorithm over 11 uniformly-spaced values of k � 103 in the interval [0;2]. Most of the valuesof k within this interval yield a Dice metric above 0.9 for the cavityand above 0.8 for the myocardium. The Dice metric starts to dropsignificantly only at the neighborhood of k � 103 = 2. Fig. 10a andb demonstrate that the smoothness parameter does not requirefine tuning.

4. Conclusion

This study investigated nearly real-time detection of the LVendo- and epicardium boundaries in a cardiac magnetic resonance(MR) sequence. The solution is obtained following the optimizationof two original discrete cost functions, each containing globalgeometry and intensity constraints based on the Bhattacharyyasimilarity. Quantitative evaluations over 2280 images acquiredfrom 20 subjects demonstrated that the results correlate well withindependent manual segmentations by an expert. Compared to therecent active contour method in Ben Ayed et al. (2009a), the pro-posed formulation led to improvements in accuracy and a signifi-cant decrease in computation time. Built upon an original boundof the Bhattacharyya measure, the proposed formulation affordan important computational advantage over related active contourmethods: it does not require a large number of iterative updates ofthe segmentation and the corresponding kernel densities. Anotherinteresting aspect of the bound-optimization framework is that itcan be used for several other applications and, therefore, can leadto segmentation algorithms which share the flexibility of activecontours and computational advantages of max-flow optimization.Apart from these advantages over related general-purposesegmentation methods, the proposed formulation has severaldesirable properties in the application context. Based on globaldistribution information learned from the current data, it removesthe need of a large training set, handles intrinsically geometricvariations of the LV without biasing the solution towards a set oftemplate shapes, relaxes optimization over geometric transforma-tions, and prevents the papillary muscles from being includederroneously in the myocardium.

Acknowledgments

This study is supported in part by the Natural Sciences andEngineering Research Council of Canada (NSERC), under the post-doctoral fellowship (PDF) awarded to Ismail Ben Ayed, and in partby GE Healthcare.

Appendix A. Proof of Proposition 1

Because RL1 and RL0 are complementary, we can rewrite RLa

1 asfollows:

RLa

1 ¼ ðRLa

1 \ RL1 Þ [ RLa

1 \ RL0� �

ðA:1Þ

For L verifying RL1 � RLa

1 , we have RLa

1 \ RL1 ¼ RL1 . Therefore, from Eq.(A.1), we can rewrite RL1 as follows:

RL1 ¼ RLa

1 n RLa

1 \ RL0� �

ðA:2Þ

Using this equation, we rewrite the kernel density estimate in (4) asfollows:

PL;In ðiÞ ¼P

p2RLa

1Ki In

p

� ��P

p2RLa

1 \RL0Ki In

p

� �A RL

a

1

� �� A RL

a

1 \ RL0� � ðA:3Þ

Now because A RLa

1 \ RL0� �

is nonnegative, we have the followinginequality

PL;In ðiÞPP

p2RLa

1Ki In

p

� ��P

p2RLa

1 \RL0Ki In

p

� �A RL

a

1

� �¼ PLa ;In ðiÞ �

Pp2RL

a1 \RL0

Ki Inp

� �A RL

a

1

� � ðA:4Þ

Using this lower bound in the Bhattacharyya measure, we obtainthe following upper bound of BC;In ðLÞ (i is omitted as argument ofthe distributions to simplify the equations)

BC;In ðLÞ 6 �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPLa ;In �

Pp2RL

a1 \RL0

Ki Inp

� �A RL

a

1

� �0@ 1AMC;I

vuuut

¼ �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPLa ;InMC;I

q ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1�

Pp2RL

a1 \RL0

Ki Inp

� �PLa ;In A RL

a

1

� �vuuut

¼ �X

i 2 IffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPLa ;InMC;I

q ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1�

Pp2RL

a1 \RL0

Ki Inp

� �P

p2RLa

1Ki In

p

� �vuuut ðA:5Þ

Now notice the following inequality for any 0 6 x 6 1ffiffiffiffiffiffiffiffiffiffiffiffi1� xp

P 1� x ðA:6Þ

Because RLa

1 \ RL0 � RLa

1 , we have

0 6

Pp2RL

a1 \RL0

Ki Inp

� �P

p2RLa

1Ki In

p

� � 6 1 ðA:7Þ

Thus, applying inequality (A.6) givesffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1�

Pp2RL

a1 \RL0

Ki Inp

� �P

p2RLa

1Ki In

p

� �vuuut P 1�

Pp2RL

a1 \RL0

Ki Inp

� �P

p2RLa

1Ki In

p

� � ðA:8Þ

Finally, combining this inequality with (A.5) gives the followingupper bound of BC;In ðLÞ

BC;In ðLÞ 6 �Xi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPLa ;InMC;I

q1�

Pp2RL

a1 \RL0

Ki Inp

� �P

p2RLa

1Ki In

p

� �0@ 1A

¼ BC;In ðLaÞ þXi2I

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPLa ;InMC;I

q Pp2RL

a1 \RL0

Ki Inp

� �PLa ;In AðRL

a

1 Þ

¼ BC;In ðLaÞ þXp2RL0

Lap

A RLa

1

� � Xi2I

Ki Inp

� � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiMC;I ðiÞPLa ;In ðiÞ

sðA:9Þ

Now because A RLa

1 \ RL0� �

þ A RL1

¼ A RLa

1

� �, we have:

I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100 99

BC;InðLaÞ ¼BC;In ðLaÞA RL

a

1 \ RL0� �

A RLa

1

� � þBC;In ðLaÞA RL1

A RL

a

1

� � ðA:10Þ

Now notice the following equalities:

A RLa

1 \ RL0� �

¼Xp2RL0

Lap

A RL1

¼Xp2RL1

1 ðA:11Þ

Embedding these equalities in (A.10) gives:

BC;InðLaÞ ¼Xp2RL0

LapBC;In ðLaÞ

A RLa

1

� � þXp2RL1

BC;InðLaÞA RL

a

1

� � ðA:12Þ

Now because the last term in (A.12) is positive, we have the follow-ing inequality "a 2 [0,1]:

BC;InðLaÞ 6Xp2RL0

LapBC;InðLaÞ

A RLa

1

� � þ ð1� aÞXp2RL1

BC;InðLaÞA RL

a

1

� � ðA:13Þ

Combining (A.9) and (A.13) proves the inequality in Eq. (14). Proofof the inequality in Eq. (16) follows the same steps.

Appendix B. Proof of Proposition 2

To prove Proposition 2, it suffices to verify conditions (18) and(19) for AC;In and F C;In . Condition (18) follows directly from Propo-sition 1. For Condition (19), it suffices to see that

When La ¼ L;Lap ¼ 0 8p 2 RL0 ; i:e:;

cp;In ð0Þ ¼ 0cp;Dð0Þ ¼ 0

�8p 2 RL0

In this case, we haveXp2RL0

cp;In ð0Þ ¼ 0 andXp2RL0

cp;Dð0Þ ¼ 0 ðB:1Þ

Therefore,

for a ¼ 0; AC;InðL;L;0Þ¼ J C;In ðL;L;0Þ þ cJ C;DðL;L; 0Þ þ kSðLÞ

¼Xp2RL1

cp;In ð1Þ þ cXp2RL1

cp;Dð1Þ þ kSðLÞ

¼Xp2RL1

BC;In ðLÞA RL1 þ c

Xp2RL1

BC;DðLÞA RL1 þ kSðLÞ

¼ BC;In ðLÞ þ cBC;DðLÞ þ kSðLÞ ¼ F C;In ðLÞ ðB:2Þ

which verifies condition (19).

References

Adam, A., Kimmel, R., Rivlin, E., 2009. On scene segmentation and histograms-basedcurve evolution. IEEE Transactions on Pattern Analysis and Machine Intelligence31, 1708–1714.

Aherne, F., Thacker, N., Rockett, P., 1997. The Bhattacharyya metric as an absolutesimilarity measure for frequency coded data. Kybernetika 32, 1–7.

Andreopoulos, A., Tsotsos, J.K., 2008. Efficient and generalizable statistical models ofshape and appearance for analysis of cardiac MRI. Medical Image Analysis 12,335–357.

Aubert, G., Barlaud, M., Faugeras, O., Jehan-Besson, S., 2003. Image segmentationusing active contours: calculus of variations or shape gradients? SIAM Journal ofApplied Mathematics 63, 2128–2154.

Ben Ayed, I., Chen, H.M., Punithakumar, K., Ross, I., Li, S., 2010. Graph cutsegmentation with a global constraint: recovering region distribution via abound of the Bhattacharyya measure. In: IEEE International Conference onComputer Vision and Pattern Recognition (CVPR 2010), San Francisco, CA, USA,pp. 1–7.

Ben Ayed, I., Li, S., Ross, I., 2009a. Embedding overlap priors in variational leftventricle tracking. IEEE Transactions on Medical Imaging 28, 1902–1913.

Ben Ayed, I., Li, S., Ross, I., 2009b. A statistical overlap prior for variational imagesegmentation. International Journal of Computer Vision 85, 115–132.

Ben Ayed, I., Li, S., Ross, I., Islam, A., 2009c. Myocardium tracking via matchingdistributions. International Journal of Computer Assisted Radiology and Surgery4, 37–44.

Ben Ayed, I., Punithakumar, K., Li, S., Islam, A., Chong, J., 2009. Left ventriclesegmentation via graph cut distribution matching. In: Medical ImageComputing and Computer-Assisted Intervention (MICCAI 2009), vol. 1,London, UK, pp. 901–909.

Bishop, C.M., 2007. Pattern Recognition and Machine Learning (Information Scienceand Statistics). Springer.

Blake, A., Rother, C., Brown, M., Pérez, P., Torr, P.H.S., 2004. Interactive imagesegmentation using an adaptive gmmrf model. In: European Conference onComputer Vision (ECCV 2004), vol. 1, Prague, Czech Republic, pp. 428–441.

Boykov, Y., Funka Lea, G., 2006. Graph cuts and efficient n-d image segmentation.International Journal of Computer Vision 70, 109–131.

Boykov, Y., Kolmogorov, V., 2003. Computing geodesics and minimal surfaces viagraph cuts. In: IEEE International Conference on Computer Vision (ICCV 2003),vol. 1, Nice, France, pp. 26–33.

Boykov, Y., Kolmogorov, V., 2004. An experimental comparison of min-cut/max-flow algorithms for energy minimization in vision. IEEE Transactions on PatternAnalysis and Machine Intelligence 26, 1124–1137.

Cerqueira, M.D., Weissman, N.J., D, V., Jacobs, A.K., Kaul, S., Laskey, W.K., Pennell,D.J., Rumberger, J.A., Ryan, T., Verani, M.S., 2002. Standardized myocardialsegmentation and nomenclature for tomographic imaging of the heart: astatement for healthcare professionals from cardiac imaging committee of thecouncil on clinical cardiology of the american heart association. Circulation 105,539–542.

Comaniciu, D., Ramesh, V., Meer, P., 2003. Kernel-based object tracking. IEEETransactions on Pattern Analysis and Machine Intelligence 25, 564–575.

El-Berbari, R., Bloch, I., Redheuil, A., Angelini, E.D., Mousseaux, É., Frouin, F.,Herment, A., 2007. Automated segmentation of the left ventricle includingpapillary muscles in cardiac magnetic resonance images. In: FunctionalImaging and Modeling of the Heart (FIMH 2007), Salt Lake City, UT, USA,pp. 453–462.

Flusser, J., Zitova, B., Suk, T., 2009. Moments and Moment Invariants in PatternRecognition. Wiley.

Fradkin, M., Ciofolo, C., Mory, B., Hautvast, G., Breeuwer, M., 2008. Comprehensivesegmentation of cine cardiac mr images. In: Medical Image Computing andComputer-Assisted Intervention (MICCAI 2008), vol. 1, New York, NY, USA, pp.178–185.

Freedman, D., Zhang, T., 2004. Active contours for tracking distributions. IEEETransactions on Image Processing 13, 518–526.

Fritscher, K.D., Pilgram, R., Schubert, R., 2005. Automatic cardiac 4d segmentationusing level sets. In: Functional Imaging and Modeling of the Heart (FIMH 2005),Barcelona, Spain, pp. 113–122.

Georgiou, T., Michailovich, O., Rathi, Y., Malcolm, J., Tannenbaum, A., 2007.Distribution metrics and image segmentation. Linear Algebra and itsApplications 425, 663–672.

Hautvast, G., Lobregt, S., Breeuwer, M., Gerritsen, F., 2006. Automatic contourpropagation in cine cardiac magnetic resonance images. IEEE Transactions onMedical Imaging 25, 1472–1482.

Jolly, M.P., 2008. Automatic recovery of the left ventricular blood pool in cardiaccine mr images. In: Medical Image Computing and Computer-AssistedIntervention (MICCAI 2008), vol. 1, New York, NY, USA, pp. 110–118.

Juan, O., Boykov, Y., 2007. Capacity scaling for graph cuts in vision. In: IEEEInternational Conference on Computer Vision (ICCV 2007).

Kaus, M.R., von Berg, J., Weese, J., Niessen, W., Pekar, V., 2004. Automatedsegmentation of the left ventricle in cardiac mri. Medical Image Analysis 8,245–254.

Kohli, P., Rihan, J., Bray, M., Torr, P.H.S., 2008. Simultaneous segmentation and poseestimation of humans using dynamic graph cuts. International Journal ofComputer Vision 79, 285–298.

Lee, D.D., Seung, H., 2000. Algorithms for nonnegative matrix factorization. In:Neural Information Processing Systems (NIPS 2000), vol. 13, Denver, CO, USA,pp. 556–562.

Lehmann, E.L., 1986. Testing Statistical Hypotheses. Wiley, New York.Liu, H., Chen, Y., Ho, H.P., Shi, P., 2005. Geodesic active contours with adaptive

neighboring influence. In: Medical Image Computing and Computer-AssistedIntervention (MICCAI 2005), vol. 2, Palm Springs, CA, USA, pp. 741–748.

Lorenzo-Valdés, M., Sanchez-Ortiz, G.I., Elkington, A.G., Mohiaddin, R.H., Rueckert,D., 2004. Segmentation of 4d cardiac mr images using a probabilistic atlas andthe em algorithm. Medical Image Analysis 8, 255–265.

Lynch, M., Ghita, O., Whelan, P., 2008. Segmentation of the left ventricle of the heartin 3-D+T MRI data using an optimized nonrigid temporal model. IEEETransactions on Medical Imaging 27, 195–203.

Michailovich, O.V., Rathi, Y., Tannenbaum, A., 2007. Image segmentation usingactive contours driven by the Bhattacharyya gradient flow. IEEE Transactions onImage Processing 16, 2787–2801.

Mitchell, S.C., Bosch, J.G., Lelieveldt, B.P.F., van der Geest, R.J., Reiber, J.H.C., Sonka,M., 2002. 3-D active appearance models: segmentation of cardiac mr andultrasound images. IEEE Transactions on Medical Imaging 21, 1167–1178.

Mitchell, S.C., Lelieveldt, B.P.F., van der Geest, R.J., Bosch, J.G., Reiber, J.H.C., Sonka,M., 2001. Multistage hybrid active appearance model matching: segmentationof left and right ventricles in cardiac mr images. IEEE Transactions on MedicalImaging 20, 415–423.

100 I. Ben Ayed et al. / Medical Image Analysis 16 (2012) 87–100

Ni, K., Bresson, X., Chan, T.F., Esedoglu, S., 2009. Local histogram based segmentationusing the wasserstein distance. International Journal of Computer Vision 84,97–111.

Pluempitiwiriyawej, C., Moura, J.M.F., Wu, Y.J.L., Ho, C., 2005. Stacs: new activecontour scheme for cardiac mr image segmentation. IEEE Transactions onMedical Imaging 24, 593–603.

Punithakumar, K., Ben Ayed, I., Islam, A., Ross, I., Li, S., 2010. Tracking endocardialmotion via multiple model filtering. IEEE Transactions on BiomedicalEngineering 57, 2001–2010.

Rother, C., Kolmogorov, V., Blake, A., 2004. Grabcut: interactive foregroundextraction using iterated graph cuts. ACM Transactions on Graphics 23, 309–314.

Rother, C., Minka, T.P., Blake, A., Kolmogorov, V., 2006. Cosegmentation of imagepairs by histogram matching - incorporating a global constraint into mrfs. In:IEEE International Conference on Computer Vision and Pattern Recognition(CVPR 2006), vol. 1, New York, NY, USA, pp. 993–1000.

Rousson, M., Cremers, D., 2005. Efficient kernel density estimation of shape andintensity priors for level set segmentation. In: Medical Image Computing andComputer-Assisted Intervention (MICCAI 2005), vol. 2, Palm Springs, CA, USA,pp. 757–764.

Spottiswoode, B.S., Zhong, X., Lorenz, C.H., Mayosi, B.M., Meintjes, E.M., Epstein, F.H.,2009. Motion-guided segmentation for cine dense mri. Medical Image Analysis13, 105–115.

Sun, W., Cetin, M., Chan, R., Reddy, V., Holmvang, G., Chandar, V., Willsky, A.S., 2005.Segmenting and tracking the left ventricle by learning the dynamics in cardiac

images. In: Information Processing in Medical Imaging (IPMI 2005), GlenwoodSprings, CO, USA, pp. 553–565.

Van Assen, H.C., Danilouchkine, M.G., Dirksen, M.S., Reiber, J.H.C., Lelieveldt, B.P.F.,2008. A 3-d active shape model driven by fuzzy inference: application to cardiacct and mr. IEEE Transactions on Information Technology in Biomedicine 12,595–605.

Zambal, S., Hladuvka, J., Bühler, K., 2006. Improving segmentation of the leftventricle using a two-component statistical model. In: Medical ImageComputing and Computer-Assisted Intervention (MICCAI 2006), vol. 1,Copenhagen, Denmark, pp. 151–158.

Zhang, T., Freedman, D., 2005. Improving performance of distribution trackingthrough background mismatch. IEEE Transactions on Pattern Analysis andMachine Intelligence 27, 282–287.

Zhu, Y., Papademetris, X., Sinusas, A.J., Duncan, J.S., 2010. Segmentation of the leftventricle from cardiac mr images using a subject-specific dynamical model.IEEE Transactions on Medical Imaging 29, 669–687.

Zhu-Jacquot, J., Zabih, R., 2008. Segmentation of the left ventricle in cardiac mrimages using graph cuts with parametric shape priors. In: IEEE InternationalConference on Acoustics, Speech, and Signal Processing (ICASSP 2008), CaesarsPalace, Las Vegas, Nevada, USA, pp. 521–524.

Zhuang, X., Rhode, K.S., Arridge, S.R., Razavi, R., Hill, D.L.G., Hawkes, D.J., Ourselin, S.,2008. An atlas-based segmentation propagation framework using locally affineregistration - application to automatic whole heart segmentation. In: MedicalImage Computing and Computer-Assisted Intervention (MICCAI 2008), vol. 2,New York, NY, USA, pp. 425–433.