+ All Categories
Home > Documents > Algebraic Decomposition of Fat and Water in MRI

Algebraic Decomposition of Fat and Water in MRI

Date post: 25-Sep-2016
Category:
Upload: bp
View: 216 times
Download: 0 times
Share this document with a friend
12

Click here to load reader

Transcript
Page 1: Algebraic Decomposition of Fat and Water in MRI

IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009 173

Algebraic Decomposition of Fat and Water in MRIMathews Jacob*, Member, IEEE, and Bradley P. Sutton, Member, IEEE

Abstract—The decomposition of magnetic resonance imaging(MRI) data to generate water and fat images has several ap-plications in medical imaging, including fat suppression andquantification of visceral fat. We introduce a novel algorithm toovercome some of the problems associated with current analyticaland iterative decomposition schemes. In contrast to traditionalanalytical schemes, our approach is general enough to accommo-date any uniform echo-shift pattern, any number of metabolitesand signal samples. In contrast to region-growing method that usea smooth field-map initialization to resolve the ambiguities withthe IDEAL algorithm, we propose to use an explicit smoothnessconstraint on the final fieldmap estimate. Towards this end, weestimate the number of feasible solutions at all the voxels, prior tothe evaluation of the roots. This approach enables the algorithmto evaluate all the feasible roots, thus avoiding the convergenceto the wrong solution. The estimation procedure is based on amodification of the harmonic retrieval (HR) framework to accountfor the chemical shift dependence in the frequencies. In contrastto the standard linear HR framework, we obtain the frequencyshift as the common root of a set of quadratic equations. On mostof the pixels with multiple feasible solutions, the correct solutioncan be identified by a simple sorting of the solutions. We use aregion-merging algorithm to resolve the remaining ambiguity andphase-wrapping. Experimental results indicate that the proposedalgebraic scheme eliminates most of the difficulties with thecurrent schemes, without compromising the noise performance.Moreover, the proposed algorithm is also computationally moreefficient.

Index Terms—Fat–water decomposition, harmonic retrieval,linear prediction, magnetic resonance imaging (MRI), Sylvestermatrix.

I. INTRODUCTION

I N MEDICAL imaging applications of magnetic resonanceimaging, the acquired image is dependent on various chem-

ical species as well as the magnetic field (B0) inhomogeneity.The decomposition of the images into chemical concentrationsand the B0 map is crucial for various applications. The mainutility of this scheme is the suppression of unwanted signalsfrom species such as fat, which often obscures the underlyingpathology [1]. This approach is also used to accurately estimatethe fat volume in obesity-induced illnesses [2], [3]. Since the B0induced frequency shift causes image distortions in many MRIacquisition schemes, a precise estimate of the field map is cru-cial in obtaining distortion-free reconstructions [4]–[7].

Manuscript received March 17, 2008; revised May 22, 2008. First publishedJuly 02, 2008; current version published January 30, 2009. Asterisk indicatescorresponding author.

*M. Jacob is with the Departments of Biomedical Engineering and ImagingSciences, University of Rochester, Rochester, NY 14642 USA.

B. P. Sutton is with the Department of Bioengineering and Beckman Institute,University of Illinois at Urbana–Champaign, Urbana, IL 61801 USA.

Color versions of one or more of the figures in this paper are available onlineat http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/TMI.2008.927344

Fig. 1. Water–fat–B0 model. The water and fat peaks are separated by a fre-quency indicated by� corresponding to the chemical shifts between them. Boththe peaks are shifted by the field map induced frequency shift �. The unknownsin the model are the water and fat concentrations (indicated by � and � )and the B0 induced frequency shift (denoted by �).

In recent years, there has been renewed interest in Dixon-like water–fat decomposition schemes due to their robustnessto B0 and B1 (radio-frequency field) inhomogeneities [8]–[16].The standard practice is to acquire two or three source MR im-ages, each with slightly different echo-times. These methodsare classified as two or three-point schemes depending onthe number of source images used. We focus on three pointacquisition in this paper. The source images are processedusing analytical or iterative algorithms to estimate the concen-trations and B0 field map. These algorithms assume a para-metric two-frequency model, where the difference betweenthe frequencies (chemical shift between water and fat) is aknown constant (see Fig. 1 for a graphical illustration). Thismodel-based approach enables the estimation of the parame-ters from fewer source images, thus providing the estimates ina reasonable scan time. However, this approach suffers fromtwo main sources of ambiguity.

1) When uniform echo-shifts are considered, the estimatedsolutions may suffer from phase wrapping. The range offield-map frequencies that can be uniquely estimated is

, where is echo-spacing. Theestimates of the field-map frequencies that are outside thisrange will be wrapped back to this fundamental range.

2) The algorithm leads to two feasible solutions on voxelswith only one metabolite. When one of the chemicals isabsent, either of the two frequencies in the model may fitthe single exponential due to the metabolite that is present.

Both of these uncertainties cannot be resolved from the mea-sured data; additional prior information has to be used to resolvethese ambiguities. For convenience, we will refer to the uniquesolution in the fundamental range as simply the unique solutionin the rest of the paper.

Analytical schemes offer an elegant means to estimate theparameters as simple nonlinear functions of the measuredsamples [8]–[12]. These methods evaluate two roots at everyvoxel (irrespective of whether both species are present or not).They then use a mixture of heuristic and prior information to

0278-0062/$25.00 © 2009 IEEE

Page 2: Algebraic Decomposition of Fat and Water in MRI

174 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

resolve the ambiguities [8]–[12]. Since these algorithms aredesigned for specific echo-shift patterns [8], [9], their extensionto arbitrary patterns is not obvious. The echo-shift patternsfor which these reconstruction schemes were designed oftenprovide redundant measurements, thus resulting in ambiguousreconstructions [13]. In addition, the extension of these methodsto estimate more than two metabolites is also nontrivial. More-over, the noise performance of these nonlinear schemes mightbe suboptimal.

The IDEAL algorithm formulates the estimation of themodel parameters as an optimization problem, where themaximum likelihood (ML) criterion is minimized [13], [14].This technique uses an iterative optimization scheme to min-imize the nonlinear criterion. In contrast to the analyticalmethods, IDEAL assumes a single solution at every voxel.When there are two feasible solutions, it converges to the one ofthem, depending on the initialization. Moreover, the nonlinearML criterion has several local minima, depending upon thefat–water concentrations and echo-shift parameters [17]. Aheuristic scheme to initialize IDEAL and thus avoid theseproblems has been proposed in [17]. This approach attemptsto resolve the uncertainties is by using the prior knowledgeof the field map smoothness [17]. This scheme estimates thevoxels sequentially along a rectangular spiral trajectory. Toderive the initialization of the current pixel on the trajectory,it performs a polynomial fit of the already solved pixels in asquare neighborhood. While it solves the convergence issuesin many cases, it often leads to propagation of errors along thetrajectory as explained in [17]. Moreover, there are often veryfew reliable pixels (in the square neighborhood) in regions closeto the edge of the object and noisy regions; this often leads toill-posed/poor polynomial fit leading to poor initializations asdemonstrated in the paper.

To overcome the problems associated with the analytic anditerative schemes, we propose a novel framework for fat–waterdecomposition. In contrast to the region growing approach ofinitializing the iterative algorithm with a smooth function toobtain a smooth solution, we propose to impose an explicitsmoothness constraint on the final field-map estimate. Due tothe nonlinear nature of the estimation process, the initializationwith a smooth function may not lead to a smooth field mapestimate. Besides, the polynomial fitting procedure to derivethe field map estimate [17] may be ill-posed at regions close tothe edge of the objects. The resulting errors may be propagatedalong the rectangular spiral trajectory used by the algorithm asexplained in [17]. To enforce the smoothness constraint on thefinal fieldmap estimate, we estimate all the feasible solutions ateach voxel. The number of feasible solutions (model order) ateach voxel is estimated from the noisy data, prior the estimationof the solutions. The correct solutions from these feasible setsare then chosen subject to the smoothness constraint. Thismethod is also much more computationally efficient than boththe standard IDEAL and region growing IDEAL algorithms.

The estimation of the feasible solutions is based on the har-monic retrieval (HR) frequency estimation theory. This well-established framework is used for the estimation of unknownfrequencies in a time domain signal [18], [19]; in the context ofMRI, it has been used to estimate the metabolite peak locations

[20]–[22], solvent suppression [20]–[23] in MR spectroscopicimaging, and field inhomogeneity estimation [24]. The maincontribution of this paper is the modification of the standard HRsetup, where the frequencies are assumed to be independent,to introduce the prior knowledge of difference in frequenciesof the chemical species. The introduction of this informationmakes the approach more robust than the standard HR model.Similar approaches (using chemical shift, amplitude constraints,and frequency range) to constrain harmonic retrieval have beeninvestigated in [22]. This approach was designed in the contextof MR spectroscopy applications, where large number of timesamples were available. Our main focus is to estimate the un-knowns from minimal number of samples, where this schemeis not readily applicable. Moreover, [22] assumes unique solu-tions, which is not true in our case.

We obtain the feasible solutions as the common roots of aset of quadratic equations. This is in contrast to the standardHR setup, where a set of linear equations are solved. Thanks tothe well established algebraic theory of polynomials [25]–[27],the model order evaluation and computation of the feasible so-lutions are efficiently implemented using simple matrix opera-tions. The ambiguity removal procedure picks the correct rootfrom the feasible sets (at each ambiguous voxel), such that asmoothness criterion is minimized. This criterion is only usedin selecting the correct component at the ambiguous voxels andhence does not smooth the measured field map as in regular-ization based schemes. We use a region merging heuristic to de-rive a computationally efficient scheme to remove the ambiguityand thus to derive the smooth field map estimate. Thanks to thesimple matrix manipulations at each voxel to derive the feasiblesolutions and the region merging step that consider large regionsrather than individual pixels, the proposed algorithm providessignificant computational savings over standard schemes.

The proposed scheme is general enough to be used with anyuniform echo-shift pattern. We also generalize it to accommo-date arbitrary number of metabolites and signal samples, thusenabling its use in other problems such as MR spectroscopy.Since the estimation procedure involves simple matrix opera-tions, the algorithm is computationally efficient. Using Monte-Carlo simulations, we show that the benefits associated with theproposed scheme come only at a marginal decrease in noise per-formance over iterative schemes.

The rest of the paper is organized as follows. In Section II,we review the Dixon decomposition scheme and the harmonicretrieval framework. In Section III, we introduce the novel algo-rithm. We address the uniqueness of the estimate in Section IV.In Section V, we consider the extension of the algorithm tomultiple chemical species (more than 2) and arbitrary numberof samples. We validate the fat–water decomposition algorithmand compare its noise performance with the iterative scheme inSection VI.

II. PRELIMINARIES

We will now review the three-point Dixon method and theharmonic retrieval framework on which our method is based.We will also define the notations followed in the paper.

Page 3: Algebraic Decomposition of Fat and Water in MRI

JACOB AND SUTTON: ALGEBRAIC DECOMPOSITION OF FAT AND WATER IN MRI 175

A. Dixon Decomposition: Mathematical Formulation

In the standard setup, the signal at each spatial location ismodeled as

(1)

where and are the concentration values ofwater and fat, respectively. Various system nonidealities suchas k-space sampling shifts, subject motion and RF effects canlead to spatially varying water and fat phases [28]. Hence, as-suming the concentrations to be real quantities may lead to ar-tifacts. The phases of and may be modeled to be thesame as in [28]. However, we assume them to be complex quan-tities for simplicity. denotes the frequency shift due to themagnetic field inhomogeneity. is the chemical shift betweenwater and fat peaks; see Fig. 1 for a graphical illustration. Theunknowns in (1) are the concentrations andand the field-inhomogeneity induced shift . Since the esti-mation procedure treats each voxel independently, we will omitthe dependence on the spatial variable for simplicity.

The standard practice is to estimate these unknowns fromthree uniformly spaced samples of the signal, specified by

. The sampling step is denoted by ,while is a shift parameter to make the echo-shift patternasymmetric. The Dixon decomposition scheme uses the sym-metric echo-shift scheme specified by ,while the IDEAL scheme assumes an asymmetric pattern

. Assuming this pattern, the signalsamples are given by

(2)

where and . andare the concentrations of water and fat, respectively. We

use the square brackets to distinguish between discrete and con-tinuous domain signals. Note that we model the fat and watersignals as undamped exponentials. Since the sampling step istypically much smaller than the and of the metabolites,this assumption does not affect the estimation.

As discussed before, fitting the two-frequency model, indi-cated in (1), to a single frequency signal (on voxels with onlyfat or water present) will result in two feasible solutions. Thesingle exponential may be represented by either of the two fre-quencies. This is a fundamental ambiguity and can only be re-solved by using additional prior information. We will discussour approach for removing this uncertainty in Section III-D.

B. Frequency Estimation Using HR

The HR framework deals with the estimation of unknown fre-quencies in a signal from its time samples [18]. The frameworkis rooted on the concept of the annihilating filter. Assume thatwe have the samples of an exponential given by

(3)

where . It is easy to see that the parametricfilter (termed as forward annihilating filter)

(4)

will annihilate the exponential (3); i.e., .Any two consecutive samples of will provide a linear equa-tion in . The unknown parameter is estimated from theselinear equations. This estimation of the frequencyis unique [18] and is also robust [19].

Since (exponential is undamped), the backward filterwill annihilate the sequence : the conjugate

of . This provides an additional set of equations that maybe used to improve the robustness of the estimate. We refer thereader to [18] for a more in-depth coverage of the harmonicretrieval framework.

If a two-frequency model is considered, the expression for thesignal is given by

(5)

The corresponding parametric filter is given by

(6)

The application of this filter on the time series will provide a setof linear equations in terms of the filter coefficients. In the stan-dard HR framework, the frequencies are obtained as the roots ofthis polynomial.

III. METHOD

We have seen that the number of feasible solutions is a vari-able, depending on the number of chemical species present inthe voxel. This makes it difficult to implement the estimationprocedure as a a one-step algorithm. We propose a sequentialapproach for the estimation of the field inhomogeneity and theconcentration maps. The basic steps involved are as follows.

1) Estimation of the number of feasible solutions (describedin Section III-C).

2) Evaluation of the feasible solutions (described in Sec-tion III-C).

3) Resolution of the ambiguities (described in Section III-D).4) Computation of the concentration maps (described in Sec-

tion III-E).The resulting algorithm is computationally efficient. Thanksto the algebraic theory of polynomials, the first two steps areperformed efficiently using matrix operations. We use a fast,greedy, algorithm to pick the correct root from the feasiblesolution sets at each ambiguous voxel. Once the field-inhomo-geneity is determined unambiguously, the computation of theconcentration maps involves the inversion of a system of linearequations.

A. Field Map Estimation: Formulation Using HR

Comparing (5) with the fat–water model specified by (1), wefind that and . Unlike the standardHR setting, the frequencies in our model (specified by (1)) arerelated to each other through the relation

(7)

where is a known quantity. This is the central dif-ference between the standard HR framework and the proposed

Page 4: Algebraic Decomposition of Fat and Water in MRI

176 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

scheme. We make use of the above relation to derive a filter, pa-rameterized in terms of

(8)

Since and are on the unit circle, we can also use the back-ward relation. The backward filter is given by . Ap-plying these filters to the time sequences and , we ob-tain the following set of quadratic equations in

(9)

(10)

Note that these quadratic equations are valid for any value of .The actual number of equations at our disposal depends on thenumber of available signal samples. In three-point Dixon setup,we have three signal samples and hence two quadratic equations.This is the minimum number of samples required to estimate theparameters using this framework. Similarly, we will have fourquadratic equations, when four samples are available.

Given the samples , (9) and (10) are the necessary con-ditions to be satisfied by the solution. We term a solution asfeasible, if it satisfies all the equations specified by (9) and(10). Thus, the roots of the greatest common divisor (GCD) ofthe polynomials will indicate the set of feasible solutions. Thenumber of feasible solutions is obtained as the degree of theGCD polynomial. The GCD thus provides the complete solu-tion to the problem at hand. We will now focus on the evaluationof the GCD.

B. Evaluation of the Polynomial GCD

In this section, we restrict ourselves to the three point Dixonscheme for simplicity. We will generalize this scheme to arbi-trary number of metabolites and signal samples in Section V.The evaluation of the GCD can be performed in terms of theassociated Sylvester matrix [29]. Assume two quadratic poly-nomials specified by and

. The 4 4 Sylvester matrix associated withthese polynomials is given by

(11)

The 2 4 submatrix , corresponding to the polynomial, is defined as

(12)

while is defined likewise. Postmultiplying the Sylvester ma-trix by the vector provides a vector of polyno-mials. These polynomials serve as a convenient basis for theGCD; premultiplication of this vector by cor-responds to evaluating the polynomial ,where and . The Sylvestermatrix has some nice properties that are very useful for our ap-plication.

1) If the GCD of and is a polynomial of degree ,then will be of rank [30].

2) If is triangulated to row-echelon form using only rowoperations, then the th row gives the coefficients ofthe polynomial GCD [26].

When the signal samples are not corrupted by noise, these prop-erties enable us to estimate 1) the number of feasible solutionsand 2) the solutions themselves. We will now focus on the noisycase.

C. GCD Estimation When the Samples Are Noisy

When the coefficients of and are corrupted bynoise, they may not have any common roots; will be a fullrank matrix in this case. Hence, we look for the approximateGCD of the two polynomials [29]. This involves the estimationof the degree of the GCD and the evaluation of the feasible so-lutions (steps 1 and 2 of our sequential algorithm).

We use the singular value decomposition (SVD) to estimatethe rank (and hence the degree of the GCD) of in the pres-ence of noise. Once the rank is estimated, the denoised matrix(denoted by ) is obtained by truncating the lower singularvalues. This is a standard procedure in signal processing for rankreduction and matrix denoising.1

We denote the SVD of the Sylvester matrix as. is a diagonal matrix, whose singular values are

arranged in the descending order; i.e.,. Similarly, the SVD of the denoised matrix is given by

, where is derived from . We then derivethe approximate GCD (whose degree is indicated by ) usingthe LU factorization of [26]. The GCD estimation procedureinvolves the following basic steps.

• Perform the singular value decomposition of :. Set and .

• Threshold the singular value with an appropriate limitparameter. If , then . Else .

• Perform LU decomposition of ; the throw provides the approximate GCD of the two polyno-mials.

We set , because we expect at least one feasible solu-tion to the system of quadratic equations. The threshold foris determined experimentally on a large number of acquired MRimages (we use a threshold of ). The worst-case sce-nario is when a pixel with a nonunique solution (two zero sin-gular values) is classified as a unique one (single zero singularvalue). If the single solution in this case is the wrong one, itis not possible to correct it in the subsequent processing steps.In contrast, if a unique pixel is classified as nonunique and oneof the solutions is the correct one, the resulting ambiguity canbe resolved using the region merging algorithm. Hence, we setthe value of the threshold high so as to avoid nonunique regionsbeing labeled as unique. Alternatively, the optimal value ofmay be evaluated using the minimum descriptor length crite-rion [32]. Intuitively, the threshold that is used to determine the

1This approach does not preserve the Sylvester structure of the matrix. How-ever, it has been shown to be a robust approach and provides reasonably accurateestimates. The quality of the algorithm may be improved by using structure pre-serving algorithms [31].

Page 5: Algebraic Decomposition of Fat and Water in MRI

JACOB AND SUTTON: ALGEBRAIC DECOMPOSITION OF FAT AND WATER IN MRI 177

number of metabolites in a voxel is a function of the noise vari-ance. Its value is nonlinearly dependent on the minimum con-centration of the second metabolite (the one with lower concen-tration), which is required to provide a unique solution at thespecified noise variance.

D. Selecting the Solutions From the Feasible Sets

Performing steps 1 and 2 of the sequential algorithm,described in the previous subsection, provides a feasible solu-tion-set at every pixel. We have seen that there is a fundamentalambiguity when any of the metabolites are absent. For pixelscontaining only water, the correct solution is the one with thehighest field-map value. Similarly, the correct solution is theone with lowest field-map value for pixels containing onlyfat. In in vivo applications, it is found that fat often coexistswith water. Hence the solution is unique at the fat voxels. Incontrast, many pixels will only have water present in it. Thus,the correct solution (on the nonunique pixels) is almost alwaysthe one with the highest field-map value. This can be seenfrom the illustration shown in Fig. 8(a). A major reason forthis assumption to be wrong in in vivo applications is due tophase-wrapping. If the field map is too low at a pixel with onlywater present, the wrong solution may wrap and thus have ahigher field map value. Similarly, if the field-map is too high,the correct solution may wrap and end up being lower in valuethan the wrong solution.

We use the above assumption (solution with highest field mapvalue is the correct one) to only derive a good initial guess.Starting with this guess, we propose to use the prior knowledgeof the smoothness of the field map to resolve both of these un-certainties (nonunique solutions and phase-wrapping). Thus ouralgorithm is also able to derive the correct solution even if thereare pixels with only fat present. We formulate the ambiguity re-moval problem as the minimization of the global cost function

(13)

Here indicates the gradient of . is the total number ofpixels. indicates the index of the selected root (

on ambiguous pixels, while on the uniquepixels). Similarly, the integer determines the phase offset atthe th pixel. The vectors and are essentially the unknowns.

The derivation of the optimal and that minimize (13) isessentially a quadratic integer optimization problem. The evalu-ation of the exact solution has a prohibitive computational com-plexity. We have seen that the solution is well defined in most re-gions of the image, except for small connected subregions wherethe solutions are swapped or phase wrapped. Hence, we proposeto use a region-merging heuristic to derive a computationallytractable algorithm. This greedy technique is a straightforwardextension of the approach pursued in [33] for phase unwrap-ping. While more sophisticated optimization schemes such as[34] may provide better results than our approach, we used thegreedy strategy due to its ease of implementation. Moreover,this approach worked well in all our experiments.

The criterion (13) depends on the relative indices (differ-ence in indices) between the neighboring pixels. We start by

assuming an initial set of regions, where the indices of theroots as well as the phase offsets of all the pixels in the sameregion are the same. One could start by assuming each voxel asan independent region. However, to reduce the computationalcomplexity, we follow the approach in [33]. We split the imageinto different connected regions, inside which the phase valuesremain in a certain interval. This is achieved by partitioning theimage into different regions such that the phase values fall inspecified ranges (e.g., ).The connected subregions of each of these regions are identifiedand are assigned a different region number.

With this partitioning and the assumption that the the indicesand the phase offsets are the same for all pixels in the

same region, we rewrite (13) as

(14)

where

(15)

Here denote the different subregions andindicates the index of the solutions on the th

pixel. is the integer phase offset of the region . Heredenotes the set of four neighboring pixels of the pixel .is the measure of the discrepancy between the boundary

pixels of regions and . is the sum of the discrete differ-ences of the pixels inside the regions. Since does not playa part in the optimization process, we ignore it in the optimiza-tion process. With this criterion, we follow the approach taken in[33]. We sequentially merge pairs of regions, until all the regionsare merged. At each step, we compare all the unmerged regionpairs and compare the possible relative indices and integer off-sets as in [33]. We combine the pair, whose relative indices andoffsets if chosen correctly, will provide the maximal decrease inthe criterion. This choice ensures the maximal decrease of costat each region-merging step. The merged region is then consid-ered as a single entity. Since this is a greedy approach, no guar-antees regarding the convergence of the criterion to the globalminimum can be made. However, this approach works well inall the experiments considered and is computationally efficient.The different steps of the region merging process is illustratedin Fig. 8.

E. Concentration Estimation (Step 4)

Having derived the field map, the signal samples are linearlyrelated to fat and water concentrations. This is similar to theprocedure followed in [13]. From (2), we have

(16)

Page 6: Algebraic Decomposition of Fat and Water in MRI

178 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

Fig. 2. The (a) magnitude and (b) phase of the nonlinear function ��� � ���� � � as a function of � . Curves corresponding to four different shift parameters,ranging from ������ to �����, at a regular interval of ���� are shown. We assume the optimal sampling step � � �����. Note that �� � � when � � ;the symmetric echo-shift scheme always yields two feasible solutions. The second solution will be lagging or leading the correct root depending on the value of� �. The solution is unique for all the asymmetric patterns, except when � � ��. These cases correspond to the voxel containing only water or fat, where thesolution is fundamentally ambiguous. When the second solution is close to the unit circle, (when � � ��), it may be misclassified to be on the unit circle inthe presence of noise. In these worst case scenarios, �� � is farthest from the �� � � � line, when � � ����. (a) Magnitude1. (b) Phase.

We estimate the concentrations as

(17)

IV. UNIQUENESS OF THE SOLUTION

In this section, we analyze the uniqueness of the algorithm inthe context of three-point fat–water decomposition. Numericalevaluation of the Cramer–Rao bounds was used in [14] to derivethe optimal shift. In contrast, we use the quadratic relationship(9) to analytically derive the uniqueness condition. We inject thediscrete fat–water model, specified by (2), into the forward (9)and identify its roots as

(18)

where . The first root is the expected solution, while thesecond one is the expected solution multiplied by a functionthat depends on the echo-shift parameters and , the chemicalshift , and the fat–water ratio . The nonlinearfunction is given by

(19)

Due to the relation between the coefficients of (9) and (10), theroots of these equations are also related to each other. If isa root of (9), then will be a root of (10). Thus, the firstroot in (18) is also a root of (10), and hence a feasible solution.The second root will also be a feasible solution if and only if

. This in turn implies that the feasible solution is uniquewhen . To study the uniqueness of the solution derivedby the proposed approach, we plot the magnitude and phase ofas a function of (for different values of ) in Fig. 2(a) and (b)respectively. We assumed that , the optimal valuederived in [13].

Note from Fig. 2 that the symmetric case leadsto , irrespective of the value of . This implies that boththe roots will be on the unit circle. There will always be two

feasible solutions. The second feasible solution will be givenby , where is a function of the ratio of fat–waterconcentrations: . Note that this solution will be lagging theactual solution when and will be leading when

. It is also seen from Fig. 2 that for assymetricecho-shift patterns only whenor . This implies that these patterns lead to uniquesolutions, unless either fat or water is absent in the voxel (or ). This corresponds to the fundamental ambiguitydiscussed earlier. Note that in these cases . It can alsobe seen from Fig. 2(b) that the second solution will lead theoriginal one by , when fat is absent in the voxel (i.e.,

). Similarly, converges to when water is absent.Thus, if the wrong solution is chosen, the estimates of water andfat will be swapped. We have thus shown that when both fat andwater concentrations are nonzero, all the asymmetric echo-shiftschemes provide unique solutions.

If is close to unity, the second solution will be close tothe unit circle. It may be misclassified as a valid solution in thepresence of noise. Thus, the regions close to indicatethe worst case scenarios. We see from the plots that is mostdifferent from unity when . It is easy to see that theoptimal shift parameter is invariant to shifts by . This shiftis also the one that provides the most robust solution as shownnumerically in [13].

V. GENERALIZATION TO MULTIMETABOLITE

AND MULTIPOINT CASES

Our main focus, so far, was on the Dixon-three point decom-position of water and fat. The generalization of the proposed al-gorithm to arbitrary number of metabolites and number of sam-ples is straightforward due to the well developed algebraic for-malism [25]. These schemes can be used for the MRI exam ofthe breast with silicone implants as well as MR spectroscopicimaging schemes, where the concentration maps of multiplemetabolites have to be derived from the MR data. We will nowbriefly describe these generalizations.

Page 7: Algebraic Decomposition of Fat and Water in MRI

JACOB AND SUTTON: ALGEBRAIC DECOMPOSITION OF FAT AND WATER IN MRI 179

A. Arbitrary Number of Metabolites

Assuming the number of metabolite resonances to be , thediscrete model is given by

(20)

where , and is the chemical shiftof the th metabolite. , where is theconcentration of the th metabolite and is a fieldand chemical-shift induced phase. Note that the observed signalis the sum of exponentials. The corresponding forward filteris

(21)

It is an -degree polynomial filter, parametrized in terms of theunknown term . Applying this filter to the time series , weget the th degree polynomial equation

(22)

where are the coefficients of the filter. Similarly, the backward equation provides another

th degree polynomial, denoted by . Similar to the Dixonthree point estimation scheme, the feasible solutions are givenby the GCD of these polynomials. The GCD is computed usingthe Sylvester matrix, specified by

(23)

where the submatrix is given by

......

...(24)

Similar to the fat–water case considered in Section III, we esti-mate the degree of the GCD (indicated by ) and the GCD itselffrom the Sylvester matrix. The degree of the GCD is given by

[25]. The GCD polynomial is obtainedby the row echelon decomposition of the matrix, followed bypicking the th row. Note that the minimum number ofsignal samples required for the estimation is . As in thetwo metabolite case, the number of feasible solutions is upperbounded by , depending on the number of metabolites that arepresent in the voxel. The adaptation of the rank estimation, de-noising, and the ambiguity removal algorithm is straightforwardand hence will not be discussed. As the number of metabolitesincrease, the number of feasible solutions may also increase (de-pending on the location of the metabolite peaks). The extensionof the region merging algorithm to this case is straightforward.The computational complexity of the region merging algorithmincreases as , where is the number of metabolites and

hence the upper bound on the number of solutions; there aredifferent possibilities of solution indices between any two

nonunique regions.

B. Arbitrary Number of Samples

We will now generalize the multimetabolite decompositionalgorithm to accommodate more samples. As explained in Sec-tion III, the number of available equations also increase with thenumber of signal samples. This approach can be used to improvethe robustness of the estimate. The increased number of sampleswill also improve the accuracy of the model order estimate. Letus assume that we have polynomials of degree indicatedby . The feasible solutions will satisfy allthe polynomial relations . Hence,the solution set is obtained as the roots of the GCD polynomial.The evaluation of the GCD of polynomials can be performedby using the generalized Sylvester matrix [25], spec-ified by

...(25)

where each of the submatrices are defined as in (24). Thanksto the generalized theory introduced in [25], the GCD and itsdegree in this case is obtained as in the previous cases. Thedegree of the GCD (denoted by ) is given by

, while the GCD is obtained by reducingto row-echelon form and picking the th

row. Note that the generalized Sylvester matrix is rectangular asopposed to the square matrix in the critically sampled cases con-sidered above. The number of feasible solutions, when the datais nonnoisy, is independent on the number of samples. Hence,the computational complexity of the region growing algorithmin this case would be exactly the same as the three point case.

VI. RESULTS

In this section, we validate the fat–water decomposition algo-rithm using numerical simulations as well as experimental data.

A. Simulation Study of Noise Performance

The IDEAL algorithm minimizes the maximum likelihoodcriterion and attains the Cramer–Rao lower bounds on the vari-ances of the estimates [14]. In contrast, we resort to a multistepalgorithm to avoid the convergence to wrong solutions and localminima. It may be argued that the proposed scheme may lead toa decrease in noise performance (i.e., increase in variance of theestimates). Hence, in this subsection, we analyze the noise-per-formance of the algorithm using Monte-Carlo experiments.

For comparisons, we use the effective number of signal aver-ages (NSA) metric, discussed in [9], [14]. This is the traditionalmeasure of noise efficiency in the context of fat–water decom-position [9]; it is defined as

(26)

Page 8: Algebraic Decomposition of Fat and Water in MRI

180 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

Fig. 3. Noise robustness of the proposed algorithm, in comparison to that of IDEAL. The effective number of signal averages (NSA) versus the fat–water ratio�� �, are plotted. Monte-Carlo simulation with 1000 trials is used to generate the plots. We assumed �� � ���� and the echo-shift pattern specified by � ������� � � ����� to generate the signal samples. Gaussian white noise of a specified variance, such that the � � ���, was added to the signal samples.Note that the NSA figures of the magnitudes, estimated with the proposed algorithm are exactly the same as IDEAL. The NSA figures corresponding to the phasesand the field map is marginally lower than the IDEAL scheme. In short, the results indicates that the noise performance of the algorithm is almost comparable tothe IDEAL scheme that is shown to attain the Cramer–Rao lower bound. (a) Water magnitude. (b) Fat magnitude. (c) Water phase. (d) Fat phase. (e) Field map.

Fig. 4. Source MR images of the slices studied in Figs. 5–9. Three slices of a water–fat-inhomogeneity phantom, acquired using the GRE sequence, are shownin (a)–(c). (a) A slice of the phantom that is the furthest from the air-filled ball and hence has a small field-map dynamic range. It shows the water filled containerand the two jars with oil. (b) and (c) Slices of the same phantom that are closer to the ball. The field-inhomogeneity variations are strong in these slices, especiallyclose to the ball. One of the GRE source images used for the illustration in Fig. 8 is shown in (d). (e) Source image of the human brain, acquired using a spin echosequence used to generate the images in Figs. 9. (a) Used in Fig. 5. (b) Used in Fig. 6. (c) Used in Fig. 7. (d) Used in Fig. 8. (e) Used in Fig. 9.

where is the magnitude estimate of one of the metabolites,its variance, and the variance of one of Dixon images. Thevalue of the NSA varies from 0 to , where is the number ofmeasurements (3 in the three-point Dixon case). The traditionaldefinition is valid only for the magnitude estimates. It has beenextended to the other parameters in [14]. For a detailed descrip-tion, we refer the readers to [14].

We plot the NSA indices for the parameters of interest inFig. 3. We used 1000 trials to generate the plots. We assumed

, and in-teger. The experiments were repeated at different water–fat ra-tios. Gaussian white noise of a specified variance was added tothe samples, such that the SNR of the samples is 100, a reason-able value for SNR of our 3-T images. Note that the NSA indicesfor the water and fat magnitude for both the algorithms are es-sentially the same. For the phases as well as the field map, theNSA indices of the proposed scheme is marginally lower thanthe IDEAL algorithm. Therefore, the benefits of the proposedscheme come only at a negligible loss in robustness.

B. Validation Using Phantom Data

To test the ability of the algorithm to converge to the correctroot under controlled conditions, we created a water–fat-inho-mogeneity phantom. It consists of an outer container filled withwater. Small sealed jars of corn-oil were attached to the inte-rior of the container. We also fixed an air-filled ball to gen-erate a large field map. This phantom was scanned using gra-dient echo sequence with three different echo times. We used

ms and assumed the optimal echo shift parame-ters: ( and ). This translates

to ms, 4.2 ms, and 5 ms, respectively. Three slicesof the phantom are shown in Fig. 4(a)–(c). The slice indicatedin Fig. 4(a) is the furthest from the ball and hence has a smalldynamic range for the field map. The slices in Fig. 4(b) and (c)are close to the ball and hence has a large dynamic range.

The data was processed using the IDEAL algorithm, theregion growing IDEAL algorithm and the proposed algo-rithm. Since the original implementations of IDEAL andregion-growing IDEAL were not available, we reimplementedthem in MATLAB. For the region growing IDEAL algorithm,we assumed the same parameters2 (block size of 41 41 andfirst order polynomial fit) and the same initialization schemeas reported in [17]. The decompositions corresponding to theslices are shown in Figs. 5, 6, and 7, respectively. The dynamicrange of the field map is small in the Fig. 5 and the iterativealgorithm converged to the correct solution. As expected, allthe algorithms gave good decompositions of fat and water inthis case. In contrast, in Fig. 6, the IDEAL algorithm convergedto the wrong solution in regions with large frequency shift,while the algebraic and the region growing methods gave goodestimates. In Fig. 7, both the IDEAL and the region growingmethods converged to the wrong solution. The proximity of theregions with large field map to the edge of the object led to theregion growing algorithm performing poorly. Note that in com-parison to the IDEAL, the region growing method propagatedthe errors to large regions of the image along the rectangular

2The results of the region growing IDEAL algorithm may be improved withthe optimization of the parameters. We varied the block-size without observingmuch change in the results. However, a rigorous analysis of the performance ofthis algorithm is beyond the scope of this paper.

Page 9: Algebraic Decomposition of Fat and Water in MRI

JACOB AND SUTTON: ALGEBRAIC DECOMPOSITION OF FAT AND WATER IN MRI 181

Fig. 5. Fat water decomposition on a water phantom using gradient echo ac-quisition. One of the source images of this slice is shown in Fig. 4(a). This sliceincludes the jars containing corn-oil enclosed in the outer container. The crosssection of two plastic combs can also be seen in the image. Since this slice is thefarthest from the ball, the magnetic field is relatively homogeneous. In this case,all the algorithms gave good estimates. (a) IDEAL: water. (b) IDEAL: fat. (c)IDEAL: field-map. (d) RG: water. (e) RG: fat. (f) RG: field-map. (g) Algebraic:water. (h) Algebraic: fat. (i) Algebraic: field-map.

Fig. 6. Fat water decomposition on a water phantom using gradient echo acqui-sition. One of the source images of this slice is shown in Fig. 4(b). This slice onlycontains water. Close to the ball, the magnetic field is very in-homogeneous.IDEAL (a) Initialized using zero field map; it converged to the wrong solution.The region growing (b) and the algebraic decomposition algorithms were ableto converge to the correct solution. The water and fat images were omitted sinceno noticeable swapping of intensity is present in any of the images. The valueof the signal is pretty small at the voxels where IDEAL converged to the wrongsolution. (a) IDEAL: Field-map. (b) RG: Field-map. (c) Algebraic: Field-map.

spiral trajectory. The algebraic method worked well in this caseas well.

The MATLAB implementation of the algebraic method tookaround 4 s of computation time for a 128 128 image ona 2.33-GHz Intel Core2Duo processor, without any parallelthreads. Note that this also included the time taken for theregion merging algorithm, which took around 1.8 s of the total4 s computation time. In comparison, the standard IDEALalgorithm with 45 iterations on each pixel took around 32 s,while our MATLAB implementation of the region growingIDEAL algorithm took 118 s. The IDEAL algorithm requiresthe evaluation and least squares inversion of a 6 4 matrix

Fig. 7. Fat–water decomposition on a fat–water phantom using gradient echoacquisition. This slice only contains water. One of the source images of thisslice is shown in Fig. 4(c). Close to the ball, the magnetic field is very in-ho-mogeneous. The IDEAL algorithm (a)–(c) was initialized using zero field map;it converged to the wrong solution, resulting in the water and fat signals beingswapped in the decomposition. In regions close to the boundary of the object, theregion growing algorithm only has information from few reliable pixels to per-form polynomial fitting; the initialization is wrong thus leading the convergenceof the algorithm to wrong results (d)–(f). Note the error propagating nature ofthe region growing algorithm; the wrong estimates are spread to a larger region,along the rectangular spiral trajectory, than the IDEAL algorithm. In contrastto both the standard algorithms, the algebraic algorithm gave pretty good re-sults, correctly identifying the fat and the water signals (g)–(h). Note that therewere a few isolated pixels close to the boundary, where the region merging al-gorithm failed. (a) IDEAL: Water. (b) IDEAL: Fat. (c) IDEAL: Field-map. (d)RG: Water. (e) RG: Fat. (f) RG: Field-map. (g) Algebraic: Water. (h) Algebraic:Fat. (i) Algebraic: Field-map.

(with real entries) at every pixel, at each iteration. On the otherhand, the algebraic algorithm requires only one evaluation ofthe SVD of a 4 4 matrix to evaluate the phase value, followedby the least squares inversion of a 3 2 matrix with complexentries. Since our region-merging algorithm starts from a fewregions rather than individual pixels, it is very fast.

C. Experimental Results With Brain MR Data

We consider two brain data-sets, one acquired using the gra-dient echo scheme and the other using spin echo method. Weused for the spin echo andand gradient echo acquisitions. The repetition times (TR) werechosen as 1200 and 500 ms, respectively. The study was ap-proved by the Institutional Review Board of UIUC and writteninformed consent was obtained from the volunteers before thestudy began. The source images are shown in Fig. 4(d) and (e),respectively.

We used the gradient echo acquisition to demonstrate the re-gion merging algorithm in Fig. 8. We have adjusted the shim

Page 10: Algebraic Decomposition of Fat and Water in MRI

182 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

Fig. 8. Illustration of the region merging process. One of the three GRE images of the slice is shown in Fig. 4(d). (a) Phase map before region merging. Thepixels on which the solutions are unique are shown in (b). Note that most of the fat pixels are unique. This phase-map image is generated by combining the uniquesolutions and the solution with a higher field map value (on the nonunique pixels). Note that this map is close to the final solution [shown in (d)], except for someregions. The region merging algorithm splits the image into simply connected regions with almost the same phase-map value as shown in (c). The different regionsare assigned different colors. These regions are merged using the approach discussed above so as to minimize the criterion (13). The solution obtained using theIDEAL algorithm is shown in (e), while the one using region growing IDEAL is shown in (f). Note that the region growing IDEAL gave almost the same resultas the algebraic method in this case. The residual error and the solutions obtained using both algorithm at the pixels indicated by the arrows in (e) are shown in(g)–(h). Pixel 1 is identified as a unique pixel by the algebraic method. The corresponding solution is indicated by a red cross in (g). The IDEAL algorithm alsoconverged to the same solution, marked by the black circle. Pixel 2 is identified by the algebraic algorithm as a non-unique pixel. The corresponding solutions aremarked by red crosses in (h). Here the low value of the field map led to the wrong solution to wrap and appear as the one with the higher field-map value. However,the region-merging algorithm resolved this problem by picking the one with the lower field-map. In contrast, IDEAL converged to the wrong solution at this pixel,although the correct solution is closer to the initialization. This is because IDEAL is not a gradient based method; the value of � in IDEAL at each iteration is notupdated depending on the gradient of (27). Hence, the gradient of (27) pointing in the correct direction at the initialization is not a sufficient condition for IDEALto converge to the correct solution. IDEAL converges to the correct solution on this pixel, only if the initialization is below�����. (a) Before region merging. (b)Unique pixels. (c) Regions. (d) After region merging. (e) IDEAL. (f) Region growing IDEAL. (g) Solutions at pixel 1. (h) Solutions at pixel 2.

settings to introduce a phase ramp in one direction. This re-sults in the phase map varying from approximately 100 Hz to

Hz over the image. The phase map derived by sorting thesolutions are shown in Fig. 8(a). Note that this is a good ini-tial guess, with the solutions different from the original valuesonly when there is a phase wrap on the wrong solution. The re-gion merging algorithm divided the image into multiple regions[see Fig. 8(c)] and then merged them to obtain Fig. 8(d) so asto minimize the smoothness criterion. Note that the IDEAL al-gorithm converged to the wrong solution in regions with largephase values [Fig. 8(e)], while the region growing IDEAL wasable to avoid this problem [Fig. 8(f)]. In Fig. 8(g)–(h), we com-pare the solutions derived by IDEAL and the algebraic methodat two different voxels. The blue curve indicates the residualerror [17]. It is obtained by plotting

(27)

as a function of the phase value . The matrix is the onespecified in (16). The IDEAL algorithm converges to the local-minimum of this curve. It is seen that in Fig. 8(h), the algebraicalgorithm picked both the feasible solutions. The correct onefrom this set was selected by the region merging algorithm. In

contrast, IDEAL converged to the wrong feasible solution at thispixel.

The spin echo data-set was processed using all the three al-gorithms (IDEAL, region growing IDEAL, and algebraic) andthe results are shown in Fig. 9. Note that both IDEAL and theregion growing IDEAL resulted in wrong phase values close tothe sinus regions, indicated by the arrows. Also note that the fatand the water regions are swapped. This region corresponds tolow signal as well as highly inhomogeneous magnetic field. Incontrast, the algebraic algorithm converged to the correct so-lution, resulting in a smoother field map. The isolated pixelswith discontinuous phase map values, derived using the alge-braic method, correspond to pixels with low signal. This is con-firmed by the fact that water and fat signals are not swapped atthese pixels [see Fig. 9(g) and (h)].

VII. DISCUSSION

In MR brain imaging, the subcutaneous lipids from theextra-cranial regions generates large signals, which result inmany artifacts. It is therefore a common practice to suppressthem using preparatory fat-suppression pulses. Unfortunately,this also leads to the suppression of the useful signal, thusaffecting the performance of SNR-challenged schemes such

Page 11: Algebraic Decomposition of Fat and Water in MRI

JACOB AND SUTTON: ALGEBRAIC DECOMPOSITION OF FAT AND WATER IN MRI 183

Fig. 9. Fat–water decomposition on SE acquired brain data. One of the sourceimages of this slice is shown in Fig. 4(e). Top row (a)–(c) indicates the resultsobtained using the region growing IDEAL algorithm. The second row denotesthe decomposition using the IDEAL algorithm. The last row indicates the de-composition using the proposed algebraic method. Note that both the regiongrowing and the IDEAL algorithm converged to the wrong solution close to thesinus regions. This regults in fat and water signals being swapped. In contrast,the algebraic method converges to the correct result. We used the optimal shiftsderived in [13]; i.e., (� � ����� and � � ��� � �����) for gradient echoimages and� � ����� and � � ����� for spin echo images. (a) RG: Water.(b) RG: Fat. (c) RG:.Field map. (d) IDEAL: Water. (e) IDEAL: Fat. (f) DEAL:Field map. (g) Algebraic: Water. (h) Algebraic: Fat (i) Algebraic: Field map.

as MR spectroscopic imaging. An alternate approach is toconstrain the reconstructions using spatial and the field mapinformation, estimated using the Dixon scheme [4]. This isin-fact the application that motivated the development of thisapproach. However, the main utility of this scheme may be inbody imaging, where the field map variations will be consid-erable [35], [36].

A limitation of the current implementation of the regionmerging algorithm is its inability to resolve the ambiguityon isolated voxels (separated from the main object). It doesnot pose a major concern in practical applications sincethese pixels are usually due to noise [see Fig. 7(i)]. Thislimitation is due to the finite difference implementationof the gradient, where only the difference between adja-cent pixels is considered. This problem may be minimizedby using a criterion that uses smoothed gradients such as

, where is asmooth, finitely supported filter (e.g., truncated Gaussian).This modification ensures that the evaluation of the smoothedgradient involves more neighboring pixels and hence couldpropagate the correct solution to isolated voxels.

Since the algebraic method relies on andbeing on the unit circle, it is not possible to extend

this algorithm to damped exponentials to account for oreffects. However, since the sampling step that is chosen inpractice is often much smaller than or , this is not neededfor most applications.

VIII. CONCLUSION

We introduced a general algorithm for the decomposition ofwater, fat, and field map from Dixon MRI data. The proposedalgorithm is based on a modification of the harmonic retrievalframework to accommodate for the frequency shift betweenthe chemical species. We estimated the field inhomogeneityinduced frequency shift as the common root of two quadraticpolynomials. Using the algebraic framework for the evaluationof the greatest common divisor of polynomials, we developeda computationally efficient algorithm. In contrast to traditionalanalytical schemes, the algebraic scheme is general enough toaccommodate arbitrary equispaced echo-shift patterns, numberof metabolites and signal samples. Since the algebraic methodestimates all the feasible solutions as opposed to iterativeschemes that assume a single solution, it is not affected by theconvergence to the wrong solution. Experimental results showthat the algebraic scheme eliminated most of the problemsassociated with traditional schemes without compromising thenoise performance.

ACKNOWLEDGMENT

This work was supported by the Beckman Foundation and thestartup funds at the University of Rochester. The authors wouldlike to thank Prof. J. Fessler and H. Wonseok at the Univer-sity of Michigan at Ann Arbor for providing the region-mergingIDEAL code, which was modified for the comparisons. The au-thors would also like to thank the anonymous reviewers, whosecomments significantly improved the paper.

REFERENCES

[1] E. de Kerviler, A. Leroy-Willig, O. Clement, and J. Frija, “Fat suppres-sion techniques in MRI: An update,” Biomed. Pharmacotherapy, vol.52, no. 2, pp. 69–75, 1998.

[2] W. Gong, H. Ren, H. Tong, X. Shen, J. Luo, S. Chen, J. Lai, X. Chen,H. Chen, and W. Yu, “A comparison of ultrasound and magnetic reso-nance imaging to assess visceral fat in the metabolic syndrome,” AsiaPacific J. Clin. Nutrition, vol. 16, pp. 339–345, 2007, Suppl 1.

[3] D. Armao, J.-P. Guyon, Z. Firat, M. A. Brown, and R. C. Semelka, “Ac-curate quantification of visceral adipose tissue (vat) using water-satu-ration MRI and computer segmentation: Preliminary results,” J. Magn.Reson. Imag., vol. 23, no. 5, pp. 736–741, 2006.

[4] M. Jacob, X. Zhu, A. Ebel, N. Schuff, and Z.-P. Liang, “Improvedmodel-based magnetic resonance spectroscopic imaging,” IEEE Trans.Med. Imag., vol. 26, no. 10, pp. 1305–1318, Oct. 2007.

[5] I. Khalidov, D. Vandeville, M. Jacob, F. Lazeyras, and M. Unser,“BSLIM: Spectral localization by imaging with explicit B0 fieldinhomogeneity compensation,” IEEE Trans. Med. Imag., vol. 26, no.7, pp. 990–1000, Jul. 2007.

[6] B. Sutton, D. Noll, and J. Fessler, “Fast, iterative image reconstructionfor MRI in the presence of field inhomogeneities,” IEEE Trans. Med.Imag., vol. 22, no. 2, pp. 178–188, Jan. 2003.

[7] K. S. Nayak and D. Nishimura, “Automatic field map generation andoff-resonance correction for projection reconstruction imaging,” Magn.Reson. Med., vol. 43, pp. 151–4, 2000.

[8] W. Dixon, “Simple proton spectroscopic imaging,” Radiology, vol.153, pp. 189–194, 1984.

[9] G. Glover and E. Schneider, “Three point Dixon technique for truewater/fat decomposition with B0 inhomogeneity correction,” Magn.Reson. Med., vol. 18, pp. 371–383, 1991.

Page 12: Algebraic Decomposition of Fat and Water in MRI

184 IEEE TRANSACTIONS ON MEDICAL IMAGING, VOL. 28, NO. 2, FEBRUARY 2009

[10] Q. Xiang and L. An, “Water-fat imaging with direct phase encoding,”J. Mag. Res. Imag., vol. 7, pp. 1002–1015, 1997.

[11] G. H. Glover, “Multipoint Dixon technique for water and fat protonand susceptibility imaging,” J. Magn. Reson. Imag., vol. 1, no. 5, pp.521–530, 1991.

[12] F. J. Rybicki, R. V. Mulkern, R. L. Robertson, C. D. Robson, T. Chung,and J. Ma, “Fast three-point Dixon MR imaging of the retrobulbarspace with low-resolution images for phase correction: Comparisonwith fast spin-echo inversion recovery imaging,” AJNR Am. J. Neuro-radiol., vol. 22, no. 9, pp. 1798–1802, 2001.

[13] S. B. Reeder, A. R. Pineda, Z. Wen, A. Shimakawa, H. Yu, J. H. Brit-tain, G. E. Gold, C. H. Beaulieu, and N. J. Pelc, “Iterative decompo-sition of water and fat with echo asymmetry and least-squares estima-tion (IDEAL): Application with fast spin-echo imaging,” Magn. Reson.Med., vol. 54, no. 3, pp. 636–644, Sept. 2005.

[14] A. R. Pineda, S. B. Reeder, Z. Wen, and N. J. Pelc, “Cramér-Raobounds for three-point decomposition of water and fat,” Magn. Reson.Med., vol. 54, no. 3, pp. 625–635, Sept. 2005.

[15] Q.-S. Xiang, “Two-point water-fat imaging with partially-op-posed-phase (POP) acquisition: An asymmetric Dixon method,”Magn. Reson. Med., vol. 56, no. 3, pp. 572–584, 2006.

[16] T. E. Skinner and G. H. Glover, “An extended two-point Dixon algo-rithm for calculating separate water, fat, and B0 images,” Magn. Reson.Med. Magn. Reson. Med., vol. 37, no. 4, pp. 628–630, 1997.

[17] H. Yu, S. B. Reeder, A. Shimakawa, J. H. Brittain, and N. J. Pelc,“Field map estimation with a region growing scheme for iterative3-point water-fat decomposition,” Magn. Reson. Med., vol. 54, no. 4,pp. 1032–1039, 2005.

[18] P. Stoica and R. Moses, Introduction to Spectral Analysis. UpperSaddle River, NJ: Prentice Hall, 1997.

[19] D. Tufts and R. Kumaresan, “Estimation of frequencies of multiple si-nusoids: Making linear prediction perform like maximum likelihood,”Proc. IEEE, vol. 70, pp. 975–989, 1982.

[20] H. Barkhuijsen, R. de Beer, and D. van Ormondt, “Improved algorithmfor noniterative time-domain model fitting to exponentially dampedmagnetic resonance signals,” J. Magn. Reson., vol. 73, pp. 553–557,1987.

[21] T. Laudadio, Y. Selen, L. Vanhamme, P. Stoica, P. VanHecke, and S.VanHuffel, “Subspace-based MRS data quantitation of multiplets usingprior knowledge,” J. Magn. Reson., vol. 168, pp. 53–65, 2004.

[22] P. Stoica, Y. Selén, N. Sandgren, and S. V. Huffel, “Using prior knowl-edge in SVD-based parameter estimation for magnetic resonance spec-troscopy-the ATP example,” IEEE Trans. Biomed. Eng., vol. 51, no. 9,pp. 1568–1578, Sep. 2004.

[23] T. Sundin, L. Vanhamme, P. Van Hecke, I. Dologlou, and S. Van Huffel,“Accurate quantification of H spectra: From finite impulse responsefilter design for solvent suppression to parameter estimation,” J. Magn.Reson., Aug. 1999.

[24] H. M. Nguyen, J. R. L. Morrison, B. P. Sutton, and M. N. Do, “Jointestimation in MRI using harmonic retrieval methods,” in Proc. IEEEInt. Symp. Biomed. Imag., 2006, pp. 1004–1007.

[25] S. Fatouros and N. Karcanias, “Resultant properties of GCD of manypolynomials and a factorization representation of GCD,” Int. J. Control,vol. 76, pp. 1666–1683, 2003.

[26] K. Geddes, S. Czapor, and G. Labahn, “Another theorem relatingsylvester’s matrix and the greatest common divisor,” Math. Mag., vol.42, no. 3, pp. 126–128, 1969.

[27] E. Kaltofen and Z. Yang, “Approximate greatest common divisors ofseveral polynomials with linearly constrained coefficients and singularpolynomials,” in Proc. 2006 Int. Symp. Symbolic Algebraic Computa-tion, 2006, pp. 169–176.

[28] H. Yu, S. B. Reeder, C. A. Mckenzie, A. C. S. Brau, A. Shimakawa,J. H. Brittain, and N. J. Pelc, “Single acquisition water-fat separation:Feasibility study for dynamic imaging,” Magn. Reson. Med., vol. 55,no. 2, pp. 413–422, Feb. 2006.

[29] R. Corless, P. Gianni, B. Trager, and S. Watt, “The singular value de-composition for polynomial systems,” in Proc. Int. Symp. Symbolic Al-gebraic Computation, 1995, pp. 195–207.

[30] I. Emiris, A. Galligo, and H. Lombardi, “Certified approximate uni-variate gcds,” J. Pure Appl. Algebra, vol. 117, pp. 229–251, 1997.

[31] E. Kaltofen, Z. Yang, and L. Zhi, “Structured low rank approximationof a sylvester matrix,” in Proc. Int. Workshop Symbolic-Numeric Com-putation, 2005, pp. 188–201.

[32] J. Rissanen, “Modelling by shortest data description,” Automatica, vol.14, pp. 465–471, 1978.

[33] M. Jenkinson, “Fast, automated, N-dimensional phase unwrapping al-gorithm,” Magn. Reson. Med., vol. 49, pp. 193–197, 2003.

[34] V. Kolmogorov and R. Zabih, “What energy functions can be mini-mized via graph cuts?,” IEEE Trans. Pattern Anal. Mach. Intell., vol.26, no. 2, pp. 147–159, Feb. 2004.

[35] A. V. Barger, D. R. DeLone, M. A. Bernstein, and K. M. Welker, “Fatsignal suppression in head and neck imaging using fast spin-echo-idealtechnique,” AJNR Am. J. Neuroradiol., vol. 27, no. 6, pp. 1292–1294,2006.

[36] S. B. Reeder, M. Markl, H. Yu, J. C. Hellinger, R. J. Herfkens, andN. J. Pelc, “Cardiac cine imaging with ideal water-fat separation andsteady-state free precession,” J. Magn. Reson. Imag., vol. 22, no. 1, pp.44–52, 2005.


Recommended