+ All Categories
Home > Documents > Estimating chromophore distributions from multiwavelength photoacoustic images

Estimating chromophore distributions from multiwavelength photoacoustic images

Date post: 01-Oct-2016
Category:
Upload: p-c
View: 214 times
Download: 0 times
Share this document with a friend
13
Estimating chromophore distributions from multiwavelength photoacoustic images B. T. Cox, 1, * S. R. Arridge, 2 and P. C. Beard 1 1 Department of Medical Physics and Bioengineering, University College London, London WC1E 6BT, UK 2 Department of Computer Science, University College London, London WC1E 6BT, UK * Corresponding author: [email protected] Received October 15, 2008; accepted December 11, 2008; posted December 15, 2008 (Doc. ID 102741); published January 30, 2009 Biomedical photoacoustic tomography (PAT) can provide qualitative images of biomedical soft tissue with high spatial resolution. However, whether it is possible to give accurate quantitative estimates of the spatially vary- ing concentrations of the sources of photoacoustic contrast—endogenous or exogenous chromophores—remains an open question. Even if the chromophores’ absorption spectra are known, the problem is nonlinear and ill- posed. We describe a framework for obtaining such quantitative estimates. When the optical scattering distri- bution is known, adjoint and gradient-based optimization techniques can be used to recover the concentration distributions of the individual chromophores that contribute to the overall tissue absorption. When the scat- tering distribution is unknown, prior knowledge of the wavelength dependence of the scattering is shown to be sufficient to overcome the absorption-scattering nonuniqueness and allow both distributions of chromophore concentrations and scattering to be recovered from multiwavelength photoacoustic images. © 2009 Optical Society of America OCIS codes: 170.5120, 100.3190. 1. INTRODUCTION The biomedical imaging modality photoacoustic tomogra- phy (PAT) has been developed over the past decade and shown to be able to provide images of soft biological tissue with high spatial resolution (100 m resolution at 5–10 mm depth) [15]. For contrast, it depends on the distribution of optical absorption in the imaged tissue, which in turn depends on the abundance and location of chromophores (light-absorbing molecules) within the tis- sue. The chromophores may be endogenous, such as he- moglobin or melanin, or exogeneous, such as dyes or nanoparticles that are introduced as contrast agents. Because of the close relationship between the photoa- coustic image and the tissue optical properties, much cur- rent research in PAT is concerned with the idea that spectroscopic methods could be applied to sets of photoa- coustic images obtained at multiple optical wavelengths to extract the distributions of the chromophore concentra- tions [68]. The prospect of being able to obtain accurate, quantitative, in vivo images of the distributions of endog- enous chromophores and tagged molecular markers to sub-mm resolution with nonionizing radiation is very en- ticing. However, extracting chromophore concentrations from PAT images is not trivial, and there have so far been few, if any, attempts to take into account the full nonlin- earity of the problem and tackle the nonunique depen- dence of PAT images on absorption and scattering. Photoacoustic amplitude spectra are not, in general, di- rectly proportional to the absorption coefficient spectra that gives rise to them. Were this the case, accurate esti- mates of chromophore concentration could be obtained straightforwardly by measuring photoacoustic images at a number of different wavelengths and using a simple, linear, pixel-by-pixel spectral best-fit. As is now increas- ingly being recognized, the spatially varying and wavelength-dependent distribution of light within the tis- sue must be taken into account if accurate results are to be achieved [7,916]. To do so, however, is complicated by the need to estimate the nonuniform light distribution, as it will depend on the distributions of both the optical ab- sorption and scattering coefficients of the tissue, neither of which is known in advance. This paper investigates the inversion that maps multi- wavelength photoacoustic images to chromophore distri- butions. The general framework involves the iterative ad- justment of the optical coefficients of a numerical model of light transport until the calculated absorbed energy den- sity at each wavelength matches the measured photo- aoustic images. Two approaches to this are (1) a two-step, wavelength-by-wavelength strategy that first recovers the absorption coefficient distribution from the photoacoustic image at each wavelength and then estimates the chro- mophore concentrations spectroscopically from knowledge of the chromophore spectra, or (2) a one-step, all- wavelengths-at-once approach in which the chromophore concentrations are recovered using a direct inversion without the intermediate step to the absorption coeffi- cients. Section 2 introduces the light transport models rel- evant to photoacoustic imaging, and Section 3 describes iterative algorithms based on these light models that are capable of separating chromophore distributions when the optical scattering distribution is known (wavelength- by-wavelength strategies). An adjoint model is introduced that can be used for very efficient calculation of the func- tional gradients [9,10]. Section 4 tackles the more general Cox et al. Vol. 26, No. 2/February 2009/J. Opt. Soc. Am. A 443 1084-7529/09/020443-13/$15.00 © 2009 Optical Society of America
Transcript
Page 1: Estimating chromophore distributions from multiwavelength photoacoustic images

1Tpsw5dwcsmn

crscttqestffed

rtmsa

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 443

Estimating chromophore distributions frommultiwavelength photoacoustic images

B. T. Cox,1,* S. R. Arridge,2 and P. C. Beard1

1Department of Medical Physics and Bioengineering, University College London, London WC1E 6BT, UK2Department of Computer Science, University College London, London WC1E 6BT, UK

*Corresponding author: [email protected]

Received October 15, 2008; accepted December 11, 2008;posted December 15, 2008 (Doc. ID 102741); published January 30, 2009

Biomedical photoacoustic tomography (PAT) can provide qualitative images of biomedical soft tissue with highspatial resolution. However, whether it is possible to give accurate quantitative estimates of the spatially vary-ing concentrations of the sources of photoacoustic contrast—endogenous or exogenous chromophores—remainsan open question. Even if the chromophores’ absorption spectra are known, the problem is nonlinear and ill-posed. We describe a framework for obtaining such quantitative estimates. When the optical scattering distri-bution is known, adjoint and gradient-based optimization techniques can be used to recover the concentrationdistributions of the individual chromophores that contribute to the overall tissue absorption. When the scat-tering distribution is unknown, prior knowledge of the wavelength dependence of the scattering is shown to besufficient to overcome the absorption-scattering nonuniqueness and allow both distributions of chromophoreconcentrations and scattering to be recovered from multiwavelength photoacoustic images. © 2009 OpticalSociety of America

OCIS codes: 170.5120, 100.3190.

liwsbtiso

wbjlsawaimowcwc

eictbtt

. INTRODUCTIONhe biomedical imaging modality photoacoustic tomogra-hy (PAT) has been developed over the past decade andhown to be able to provide images of soft biological tissueith high spatial resolution (�100 �m resolution at–10 mm depth) [1–5]. For contrast, it depends on theistribution of optical absorption in the imaged tissue,hich in turn depends on the abundance and location of

hromophores (light-absorbing molecules) within the tis-ue. The chromophores may be endogenous, such as he-oglobin or melanin, or exogeneous, such as dyes or

anoparticles that are introduced as contrast agents.Because of the close relationship between the photoa-

oustic image and the tissue optical properties, much cur-ent research in PAT is concerned with the idea thatpectroscopic methods could be applied to sets of photoa-oustic images obtained at multiple optical wavelengthso extract the distributions of the chromophore concentra-ions [6–8]. The prospect of being able to obtain accurate,uantitative, in vivo images of the distributions of endog-nous chromophores and tagged molecular markers toub-mm resolution with nonionizing radiation is very en-icing. However, extracting chromophore concentrationsrom PAT images is not trivial, and there have so far beenew, if any, attempts to take into account the full nonlin-arity of the problem and tackle the nonunique depen-ence of PAT images on absorption and scattering.Photoacoustic amplitude spectra are not, in general, di-

ectly proportional to the absorption coefficient spectrahat gives rise to them. Were this the case, accurate esti-ates of chromophore concentration could be obtained

traightforwardly by measuring photoacoustic images atnumber of different wavelengths and using a simple,

1084-7529/09/020443-13/$15.00 © 2

inear, pixel-by-pixel spectral best-fit. As is now increas-ngly being recognized, the spatially varying andavelength-dependent distribution of light within the tis-

ue must be taken into account if accurate results are toe achieved [7,9–16]. To do so, however, is complicated byhe need to estimate the nonuniform light distribution, ast will depend on the distributions of both the optical ab-orption and scattering coefficients of the tissue, neitherf which is known in advance.

This paper investigates the inversion that maps multi-avelength photoacoustic images to chromophore distri-utions. The general framework involves the iterative ad-ustment of the optical coefficients of a numerical model ofight transport until the calculated absorbed energy den-ity at each wavelength matches the measured photo-oustic images. Two approaches to this are (1) a two-step,avelength-by-wavelength strategy that first recovers thebsorption coefficient distribution from the photoacousticmage at each wavelength and then estimates the chro-

ophore concentrations spectroscopically from knowledgef the chromophore spectra, or (2) a one-step, all-avelengths-at-once approach in which the chromophore

oncentrations are recovered using a direct inversionithout the intermediate step to the absorption coeffi-

ients.Section 2 introduces the light transport models rel-

vant to photoacoustic imaging, and Section 3 describesterative algorithms based on these light models that areapable of separating chromophore distributions whenhe optical scattering distribution is known (wavelength-y-wavelength strategies). An adjoint model is introducedhat can be used for very efficient calculation of the func-ional gradients [9,10]. Section 4 tackles the more general

009 Optical Society of America

Page 2: Estimating chromophore distributions from multiwavelength photoacoustic images

ckwittwduostcid

2MIfipoimtnptlfaf

AIrssagptaccst�fi

iwArthmdhti

partot

wTepd

Hcecse

tmwepotibtcetsf

Tsew

BTsiaoanicam

444 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

ase, when the scattering coefficient distribution is un-nown and potentially nonuniform, using an all-avelengths-at-once approach. In this case, the inversion

s ill-posed both because of the diffusive nature of lightransport and, more severely, because of the “photoacous-ic absorption-scattering nonuniqueness,” an example ofhich is given in Subsection 4.A. Section 4 describes andemonstrates that this nonuniqueness can be overcomesing prior information about the wavelength dependencef the scattering within [17]. The singular-value decompo-ition (SVD) of the Hessian matrix is used to demonstratehat the nonuniqueness has been removed, and a numeri-al example is given in which a chromophore distributions recovered simultaneously with the unknown scatteringistibution using Newton’s method.

. LIGHT PROPAGATION IN SCATTERINGEDIA

n a wavelength-by-wavelength inversion strategy, therst step is to extract optical absorption coefficients fromhotoacoustic images, and the second is to use knowledgef the chromophore absorption spectra in a spectroscopicnversion for chromophore concentrations. To see how it

ight be possible to extract optical absorption (and scat-ering) coefficients from a photoacoustic image, it is firstecessary to understand how a photoacoustic image de-ends on them. This section therefore briefly describeshe optical part of the photoacoustic forward problem:ight propagation and absorption. (The second part of theorward problem, the propagation and detection of thecoustic waves, is not covered here, but details can beound in the literature [4,18].)

. Absorbed Energy Density: Photoacoustic Imagen PAT, a short pulse of light, typically of nanoseconds du-ation, illuminates a region of soft tissue. The light iscattered and absorbed within the tissue, and can be de-cribed by a fluence rate ��x , t� in W/cm2, where x�� ispoint in the tissue and t is the time. For the examples

iven in this paper ��R2, but similar behavior is ex-ected in R3; see Section 5. The photons that are not scat-ered back out of the tissue are eventually absorbed by it,nd when the dominant de-excitation pathway of the ex-ited chromophores is via vibrational relaxation, the opti-al energy is converted to heat. The absorbed power den-ity, the rate at which the light energy is absorbed andherefore the rate at which the tissue is locally heated, isa� in W/cm3, where �a�x� is the optical absorption coef-cient of the tissue.The absorbed optical energy, or equivalently the depos-

ted heat energy, causes a temperature and pressure riseithin the tissue local to where the absorption took place.s soft tissue is an elastic medium, this local pressureise propagates as an acoustic (ultrasonic) wave. Becausehe optical propagation, absorption, and conversion toeat typically occur on a timescale much shorter than theechanical relaxation—i.e., the local tissue mass density

oes not change significantly until all the optical energyas been converted to heat—it is often assumed thereforehat, from the acoustic point of view, the heating occursnstantaneously. Under this assumption, the acoustic

ropagation can be modeled as an initial value problem,nd the temporal variation of the fluence rate � is not di-ectly relevant; the key quantity is its integral over time,he fluence, ��x�=���x , t�dt in J/cm2. The total amount ofptical energy deposited over the duration of the pulse,he absorbed energy density, H in J/cm3, is then given by

H�x,t� = �a�x���x���t� = h�x���t�, �1�

here the spatial part of H is written as h�x�=�a�x���x�.hermodynamic considerations [19] lead to the followingxpressions for the spatially varying temperature andressure rises T0 and p0, respectively, due to the energyeposited at time t=0:

T0 = h/��Cv�, p0 = ��vs2/Cp�h = h. �2�

ere � is the mass density, Cv and Cp are the specific heatapacities at constant volume and pressure, � is the lin-ar thermal expansivity, and vs is the sound speed. , theonversion factor between absorbed optical energy den-ity and acoustic pressure, is called the Grüneisen param-ter and is dimensionless.

The aim in photoacoustic image reconstruction is to es-imate the initial pressure distribution p0 accurately fromeasurements of the propagating acoustic pressureaves over a measurement surface surrounding p0. Sev-

ral exact and approximate algorithms have been pro-osed to solve this acoustic inversion [4,20]. As the focusf this paper is on recovering optical coefficients from pho-oacoustic images, and not on the reconstruction of themages themselves, it will be assumed that p0 has alreadyeen recovered accurately. It will also be assumed thathe Grüneisen coefficient is known, so that the photoa-oustic images can be scaled to be images of the absorbednergy density h=p0 /. The relationship between a pho-oacoustic image h and the optical absorption and reducedcattering coefficients �a and �s�, respectively, can there-ore be written as

h = �a���a,�s��. �3�

his is the fundamental equation regarding the relation-hip between the photoacoustic image and the optical co-fficients. With the dependence on spatial position andavelength x and shown explicitly, Eq. (3) is

h�x,� = �a�x,���x,,�a�x,�,�s��x,��. �4�

. Optical Absorption and Scatteringhe optical absorption of tissue arises from the optical ab-orption of its constituent molecules (potentially includ-ng naturally occurring chromophores, contrast agents,nd biomolecular probes). For some wavelength ranges,nly a few chromophores dominate the absorption. For ex-mple, in the near-infrared the absorption is predomi-antly due to oxy- and deoxy-haemoglobin, water, and lip-

ds [7]. If the concentrations of the K significanthromophores are written as ck�x�, k=1, . . . ,K, then thebsorption coefficient (over a given wavelength range)ay be written as the sum

Page 3: Estimating chromophore distributions from multiwavelength photoacoustic images

wotch

l��sdt

w[ttS

CIctiMeibTen

tfa“sivtsctSttrpl

eatr

waeflEa�wsfbtilpf

3CSApmtttlss

otthtadmtmows

ACWeic

wmsTat

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 445

�a�x,� = �k=1

K

ck�x��k��, �5�

here �k�� are the specific absorption coefficient spectraf the chromophores [7,21]. The main topic of this paper ishe inversion of Eqs. (4) and (5) to obtain the chromophoreoncentrations ck�x� from the multiwavelength images�x ,� when the spectra �k�� are known.The scattering in highly scattering media such as bio-

ogical tissue can be described by the scattering coefficients�x ,�, or equivalently the reduced scattering coefficient

s�=�s�1−g� where g is a parameter that accounts forome anisotropy in the scattering [22]. The wavelengthependence of �s� has been measured for many types ofissue and can often be approximated by

�s��x,� � a�x�−b, �6�

here the constant b�0 is known from experiment23–25]. Prior knowledge of the exponent b will be used inhe multiwavelength inversion in Section 4 to overcomehe absorption-scattering nonuniqueness described inubsection 4.A.

. Light Transport Modelsn order to study the inversion of Eq. (4), it is necessary tohoose a form or a model to describe the light fluence dis-ribution ��x�. Light propagation in a scattering mediums often modeled using the random walk approach of

onte Carlo simulations [26,27], which is widely consid-red the most accurate technique, but is computationallynefficient as the paths of many millions of photons muste calculated to obtain a good estimate of the fluence.his inefficiency makes it an unsuitable candidate for it-rative inversions in which the fluence must be calculatedumerous times.Alternative models are usually based on Boltzmann’s

ransport equation (sometimes called the radiative trans-er equation) in which the tissue is characterized by thebsorption and scattering coefficients �a and �s and aphase function” that describes the directionality of thecattering process [22]. This integrodifferential equationn the radiance—a description of the light as a time-arying function of direction at every point—expresseshe conservation of energy during the scattering and ab-orption processes. Because it is difficult to solve analyti-ally, in practice approximations are used and in all buthe simplest cases are solved numerically [22,28,29].ome approximations, including the diffusion approxima-ion used in this paper, have the additional advantagehat the equations are simple enough to manipulate di-ectly, which can be helpful when tackling the inverseroblem, e.g., by allowing gradients to be calculated ana-ytically rather than numerically.

The “diffusion approximation” to the radiative transferquation has been used widely in biomedical optical im-ging, particularly in diffuse optical tomography [28]. Theime-independent case, relevant in this case, takes on theelatively simple form of a diffusion equation

��a − � · � �� = q0, �7�

here = �3��a+�s���−1 is the optical diffusion coefficient,

nd q0 is an isotropic source term. To obtain the diffusionquation from the radiative transport equation, the lightuence, �, is assumed to be almost isotropic everywhere.quation (7) is therefore usually considered an accuratepproximation to the radiative transport equation when

s���a [28]. However, in tissue, light is quite strongly for-ard scattered (typically g�0.9), so for a nondiffuse

ource exterior to the domain, this model is accurate onlyor distances greater than a scattering length inside theoundary, 1/�s�, where the fluence has become diffuse. Forhis reason, a collimated beam incident on the boundarys often modeled by a point source placed one scatteringength inside it [30]. This is the approach taken in this pa-er, using a finite-element (FE) implementation of the dif-usion equation [29,30].

. MULTIWAVELENGTH INVERSIONS FORHROMOPHORE CONCENTRATIONS:CATTERING KNOWNs stated previously, the main aim of this paper is to ex-lore ways in which chromophore concentrations ck�x�ight be recovered from photoacoustic images. This sec-

ion and Section 4 describe inversion techniques for whenhe optical scattering is known and unknown, respec-ively. These cases are fundamentally different in that theatter is much more ill-posed due to the absorption-cattering nonuniqueness (Subsection 4.A), althoughimilar optimization tools can be employed to solve both.

When the scattering coefficient distribution is known,ne way to estimate the chromphore distributions is firsto estimate the absorption coefficient distributions fromhe images of absorbed energy one wavelength at a time,�x ,0�→�a�x ,0�, h�x ,1�→�a�x ,1�, etc, and then usehese recovered absorption coefficient spectra �a�x ,l� inlinear inversion of Eq. (5) to estimate the chromophore

istributions ck�x�. An alternative approach is to mini-ize the difference between the measured images and

hose generated using a model by adjusting the chro-ophores, thus estimating ck�x� directly using nonlinear

ptimization, without first obtaining the single-avelength absorption coefficients. Both methods are de-

cribed below.

. Fixed-Point Iterative Inversion for Absorptionoefficientshen the scattering is known, there is a simple way to

stimate the absorption coefficient from a photoacousticmage [9,10]. Given the image h�x�, the absorption coeffi-ient can be recovered using the fixed-point iteration

�a�n+1��x� = h�x�/���n��x� + ��, �8�

here ��n�=���a�n� ,�s�� is the fluence calculated from a

odel of light transport using the nth estimate of the ab-orption coefficient, and � is a regularization parameter.his approach was applied to experimental data by Yuannd Jiang [31]. If the specific absorption coefficient spec-ra of the chromophores within the tissue � �� are

k
Page 4: Estimating chromophore distributions from multiwavelength photoacoustic images

kma

BBvswr[ssaaflraaatmiocutm

sditacgsbmse

wc

watcit

CSAea

cqa

(c

eae�

Te

wae

e(g(tmtBtci

DEet

wmlwi

446 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

nown, then the absorption coefficient images can beapped to chromophore distributions by inverting Eq. (5),single matrix inversion.

. Other Inversions for Absorption Coefficientsanerjee et al. [32] have recently proposed a noniterativeersion similar to the fixed-point iteration above by as-uming the diffusion coefficient in Eq. (7) �1/ �3�s�� (truehen �s���a), which allows the fluence to be obtained di-

ectly by solving �� · � ��= h−q0. Ripoll and Ntziachristos33] also describe an inversion scheme based on the diffu-ion model that can recover small perturbations in the ab-orption coefficient distribution when both the scatteringnd background absorption coefficients are known. Yin etl. [34] suggest making additional measurments of theuence leaving the tissue and using diffuse optical tomog-aphy (DOT) to estimate the interior fluence distribution,lthough this will suffer from the poor spatial resolutionchieveable with DOT. Yuan et al. [35] propose the use ofpriori structural information as a means of regulariza-

ion, where “the PAT image (absorbed energy densityap) is used both as input data and as prior structural

nformation” [35], p. 18078. The difficulty with this formf regularization is that the PAT image may not give ac-urate structural information as it is distorted by the non-niform fluence. Indeed, the desire to obtain an imagehat is physiologically accurate structurally is one of theotivations for inverting for the optical coefficients.The difference between two images can be used to as-

ess changes before and after a contrast agent is intro-uced, or to compare two images at different wavelengthsf one of the chromophores contributes negligible absorp-ion at one of the wavelengths and the other has similarbsorption at both. A “difference” or “subtraction” imagean provide useful qualitative images that highlight re-ions where the absorption has changed [16]. However,imple considerations show that this approach is of noenefit to quantitative imaging. If in the first measure-ent there is only one absorber present �a1, but in the

econd there is also a second �a2, then the absorbed en-rgy images in the two cases will be

h1 = �a1�1, �9�

h2 = ��a1 + �a2��2, �10�

here �1 and �2 are the fluence distributions in the twoases. The difference image is given by

h2 − h1 = �a1��2 − �1� + �a2�2, �11�

hich, if �2=�1, would be the same as a photoacoustic im-ge taken with only the second absorber present. Whilehis may be a useful qualitative tool in cases where thehange in the fluence is minimal, it does nothing to assistn our attempt to extract chromophore concentrations, ashe fluence distribution is still unknown.

. Gradient-Based Inversion for Absorption orcattering: Adjoint Modeln alternative approach to estimating the absorption co-fficient images �a�x� from h�x�, and one which has thedvantage that it can be used to estimate the scattering

oefficient if �a is known, is to minimize a functionaluantifying the difference between the model output hnd the measurements h by adjusting �a�x�:

argmin�a�x�

E�a=

1

2 � �h��a� − h�2d�. �12�

The 12 is included here so that the derivatives are not

luttered by factors of 2.)One way to find the minimum is to calculate the gradi-

nts of the functional E�awith respect to �a at each point

nd perform a directed search for the minimum. Thequivalent problem for scattering is obtained by replacinga with �s� in Eq. (12):

argmin�s��x�

E�s�=

1

2 � �h��s�� − h�2d�. �13�

he functional gradients for both problems can be foundfficiently by using the adjoint equation

��a − � · � ��* = �a�h − h�, �14�

ith adjoint solution �*. The functional gradients for thebsorption and scattering can then be calculated via thequations [36]

�E�a

��a�x�= ��x��h�x� − h�x�� − �*�x���x�, �15�

�E�s�

��s��x�= 3 �x�2 � �*�x� · ���x�. �16�

In other words, by solving Eqs. (7) and (14) just onceach, the functional gradients can be calculated from Eqs.15) and (16)—a much more efficient way to calculate theradients than the common method of finite differences.Note that because both the forward and adjoint equa-ions are in the form of diffusion equations, the same nu-erical model can be used for both.) It has been shown

hat by using a gradient descent algorithm, such as theroyden–Fletcher–Goldfarb–Shanno (BFGS) minimiza-

ion routine, either the absorption or reduced scatteringoefficient distributions can be recovered when the others known in advance [36].

. Inversion for Chromophore Concentrationsxtending the minimization approach from absorption co-fficients to chromophores is straightforward. In this casehe problem becomes

argminck�x�

Ec =1

2 �� �h�ck� − h�2d�d, �17�

here h and h now represent a set of images obtained atultiple wavelengths and a second integral over wave-

ength is included. Using Eq. (5) the functional gradientith respect to the chromophores ck can be calculated us-

ng

�Ec

�ck=� �k��

�E�a

��a��d. �18�

Page 5: Estimating chromophore distributions from multiwavelength photoacoustic images

actwpts2s

ipftdt

Ert

4CSTiestafortmawE

AWhcrtfitmsa

Ffbd

Fbt

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 447

In the following example, the absorbed energy densityt four wavelengths (650, 750, 850, and 950 nm) was cal-ulated assuming that only two chromophores contributeo the absorption. The reduced scattering coefficient �s�as set constant at 2 mm−1. A point source of light wasositioned 0.25 mm below the boundary; see Fig. 1. Thewo chromophores were chosen to have absorption spectraimilar to deoxyhemoglobin and oxyhemoglobin; see Fig.. Their concentrations, c1�x� and c2�x�, respectively, arehown in Figs. 3A and 3B.

In the simulated photoacoustic images shown in Fig. 1,t is immediately apparent that it is not possible to see theatterns of both underlying chromophore distributionsrom these plots. However, the least-squares minimiza-ion, Eq. (17), can recover the two separate chromophoreistributions accurately. The estimates after 500 itera-ions are shown in Figs. 3C, 3D, and 4. The adjoint model,

A

C

ig. 1. (Color online) Images of absorbed optical energy densityour wavelengths: (A) 650, (B) 750, (C) 850, (D) 950 nm. The imageneath the upper surface, and the anisotropy factor is 0.9. Eaynamic range.

650 750 850 9500

0.002

0.004

0.006

0.008

0.010

0.012

0.014

wavelength (nm)

specificab

sorptio

ncoeffi

cien

t(mm

/l/g

)

ig. 2. Specific absorption coefficients (mm/l/g) of oxyhemoglo-in and deoxyhemoglobin, the two chromophores whose concen-rations are shown in Figs. 3A and 3B.

q. (14), provided the functional gradients, and the BFGSoutine as encoded in Matlab’s fminunc function was usedo perform the minimization.

. MULTIWAVELENGTH INVERSION FORHROMOPHORE CONCENTRATIONS:CATTERING UNKNOWNhe assumption in Section 3 was that the optical scatter-

ng was known. In some situations it may be possible tostimate the scattering coefficient accurately, for in-tance, in tissues that are fairly homogeneous. In general,hough, this will not be the case, and the photoacousticbsorption-scattering nonuniqueness makes the inversionor the chromophore concentrations ill-posed. An examplef this nonuniqueness is given in Subsection 4.A. In theemainder of this section, the notation and framework forackling the inversion using a well-known nonlinear opti-ization (Newton’s method) will be given, and an ex-

mple will be used to show that prior knowledge of theavelength dependence of the scattering, exponent b inq. (6), is sufficient to remove the nonuniqueness.

. Absorption-Scattering Nonuniquenesshile the gradient-based approach described in Section 3

as the two advantages that (a) the gradients can be cal-ulated efficiently using the adjoint model, and (b) it canecover either the absorption or scattering coefficient dis-ributions when the other is known, there remains a dif-culty when trying to recover both absorption and scat-ering coefficient distributions togther. Given aeasurement of the absorbed energy density h�x� at a

ingle wavelength, it is not possible in general to recoverbsorption and scattering distributions simultaneously

B

D

o two chromophores with different absorpion spectra, shown atnsions are 3.75 mm�8 mm, a point source is positioned 0.25 mmge is normalized by its maximum value to optimize the visible

due te dimech ima

Page 6: Estimating chromophore distributions from multiwavelength photoacoustic images

atwmwct

ctSdchoabi

c�rdst

Fw�aanen

Fsaa

Fpe

448 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

nd uniquely. The reason is that both the absorption andhe scattering can affect the fluence distribution, and sohen the calculated absorbed energy h�x� differs from theeasured absorbed energy h�x� it is not possible to sayhether this is due to an error in the absorption coeffi-

ient distribution or in the scattering coefficient distribu-ion.

A numerical example of this nonuniqueness will help tolarify the difficulty it poses. By using a FE implementa-ion of the diffusion model of light transport described inubsection 2.C, the absorption and scattering coefficientistributions shown in Fig. 5A and 5B, �a1 and �s1, werealculated to give rise to the absorbed energy distribution1 shown in Fig. 5C. (The model was encoded in Matlabn a 25�50 mesh representing a 4 mm�8 mm rectangle,

point source was placed 0.25 mm inside the upperoundary, and the boundary condition set such that thencoming photon current is zero.) The scattering coeffi-

A

B

C

D

ig. 3. (Color online) True concentration distributions (g/l) of twuccessfully obtained by minimizing Eq. (17) using a gradient-bages from Fig. 1 as input data. The functional gradients were calcpriori. The image dimensions are 3.75 mm�8 mm.

−4 −3 −2 −1 0 1 2 3

10

20

30

40

50

x (mm)

chromph

oreconc

entration(g/l) A

ig. 4. (Color online) Profiles through the chromophore concenrofiles at 1.6 mm through Figs. 3A and 3C, respectively. (B) Prxact (dashed), estimated (dotted–dashed) through Figs. 3B and

ient was then set to the distribution shown in Fig. 5E,s2, and a gradient-based minimization [with the Matlaboutine fminunc using the BFGS algorithm, and the gra-ients calculated from Eq. (15)] was used to find the ab-orption coefficient �a2 that would minimize the func-ional Enonunique

Enonunique =1

2 � �h2��a2,�s2� − h1��a1,�s1��2d�. �19�

ig. 5D shows �a2 when the differences between h2 and h1ere negligible. h2 is shown in Fig. 5F and the differencesa1−�a2, �s1−�s2, and h1−h2 are shown in Figs. 5G, 5H,nd 5J, respectively. The two pairs ��a1 ,�s1� and ��a2 ,�s2�re an example of the absorption-scattering nonunique-ess, in the sense that both result in the same absorbednergy distribution. A standard response to a nonunique-ess in an inverse problem like this is to try and incorpo-

10

20

30

40

50

0

4

8

12

16

200

ophores (images A and B) and their estimates (images C and D)gorithm (BFGS) with the multiwavelength absorbed energy im-efficiently using an adjoint model, and the scattering was known

−4 −3 −2 −1 0 1 2 3 4−5

0

5

10

15

20

x (mm)

chromph

oreconc

entration(g/l) B

s (g/l) shown in Fig. 3. (A) Exact (solid) and estimated (dotted)at 1.6 mm, exact (solid) and estimated (dotted), and at 2.3 mm,spectively.

o chromsed al

ulated

4

trationofiles3D, re

Page 7: Estimating chromophore distributions from multiwavelength photoacoustic images

ressdrfrntcbo

BTtcr

t(k

Tmtg

Ta

Frdtdws factor

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 449

ate some additional information, such as prior knowl-dge of the type of solution, into the problem to reduce theize of the solution space. Unfortunately, simple con-traints on the smoothness of the coefficientistributions—such as may be provided by Tikhonov-styleegularization—will not be sufficient in this case. Indeed,rom this example it is clear that both sets of optical pa-ameters have similar degrees of smoothness, and it willot be possible to separate them on that basis. However,he problem posed by the nonuniqueness can be over-ome, in the sense that the chromphores can be recoveredy using prior knowledge of the wavelength dependencef the scattering as described below [17,21].

. Error Functional and Sensitivity Equationshe unknown quantities from now on are the spatial dis-ributions of the concentrations of the K chromophoresk�x� and the spatial dependence of the scattering a�x�,ather than the absoprtion and scattering coefficient dis-

absorptioncoefficient 1 (mm )

abcoeffici

0

0.3

scatteringcoefficient 1 (mm )

scacoefficie

10

30

photoacoustic image 1

0

1

-1

-1

photoaco

A

B E

C F

D

ig. 5. (Color online) Optical absorption-scattering nonuniqueneise to the absorbed energy distribution (to which the photoacouistributions D and E also give rise to the same absorbed energyion, scattering, and absorbed energy images are shown on the rensities are indistinguishable (J is virtually zero everywhere) davelength photoacoustic image and the underlying optical coeffi

itioned 0.25 mm beneath the upper surface, and the anisotropy

ributions �a�x� and �s��x�. The problem is similar to Eq.17), except the scattering amplitude a�x� is also un-nown:

argminck�x�,a�x�

E =1

2 �� �h�ck,a� − h�2d�d. �20�

he sensitivity of E to changes in ck and a guide the mini-ization by indicating the local shape of the error func-

ional. Differentiating Eq. (20) with respect to ck and aives

�E�ck

=�� �h

�ck�h�ck,a� − h�d�d, �21�

�E�a

=�� �h

�a�h�ck,a� − h�d�d. �22�

he derivatives �h /�ck and �h /�a are related to �h /��and �h /���, respectively, by

ion(mm )

0

0.08

g(mm )

−12

15

−1 x 10

1 x 10−6

−6

-1

-1

image 2

scattering coefficient 1- scattering coefficient 2

absorption coefficient 1- absorption coefficient 2

photoacoustic image 1- photoacoustic image 2

H

G

J

e absorption and scattering coefficient distributions A and B giveage is proportional) shown in C. The absorption and scatteringution, which is shown in F. The differences between the absorp-G, H, and J, respectively. The fact that these absorbed energy

strates the nonuniqueness of the relationship between a single-. The image dimensions are 4 mm�8 mm, a point source is po-is 0.9.

sorptent 2

tterinnt 2

ustic

ss. Thstic imdistribight inemoncients

s

Page 8: Estimating chromophore distributions from multiwavelength photoacoustic images

a

w�tEcb

S

Tfigebmard

CIsctbo(stt

tv

DWdtImettk

Tb

wwf

i[nka

wTrHEr

A+pH

450 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

�h

�ck= �k

�h

��a, �23�

�h

�a= −b

�h

��s�, �24�

nd �h /��a and �h /��s� may be calculated from

�h�x�

��a�x��= ��x���x − x�� + �a�x�

���x�

��a�x��, �25�

�h�x�

��s��x��= �a�x�

���x�

��s��x��, �26�

hich come from differentiating Eq. (3). Equations for� /��a and �� /��s�, the sensitivity of fluence with respecto the optical parameters, may in turn be derived fromq. (7). The sensitivity of the fluence ��x� at point x to a

hange in the absorption coefficient �a�x�� at point x� maye calculated from

��a − � · � ����x�

��a�x��= − ��x���x − x��. �27�

imilarly for the diffusion coefficient

��a − � · � ����x�

� �x��= � · ���x − x�� � ��x��. �28�

he reduced scattering coefficient and the diffusion coef-cient are related by = �3��a+�s���

−1, so differentiatingives the sensitivity relation � /��s�=−3 2. All the gradi-nts and sensitivities required for a minimization coulde calculated from these equations. However, as the lightodel was encoded using a FE model, the gradients could

lso be obtained by differentiating the basis functions di-ectly. This avoids the numerical difficulties posed by theiscretization of terms such as � · ���x−x�����x��.

. Discrete Notationmages are not continuous functions of x but discretized,o it is helpful to have a notation to describe the discretease. For the remainder of this paper, the coordinates ofhe pixels (or voxels) of the photoacoustic image of h wille denoted by xm, m=1, . . . ,M, and the optical coefficientsr chromophores will be defined at points xn, n=1, . . . ,N.The meshes defined by these points may of course be theame.) Also, the subscripts k and l will be used to indicatehe different chromophores and wavelengths, respec-ively. The following column vectors will be useful:

• the absorbed energy distribution at wavelength l,hl= �h1

1 , . . . ,hM1�T= �h�x1 ,l� , . . . ,h�xM ,l��T,

• the concentration distribution of chromophore k,ck= �ck1 , . . . ,ckN�T= �ck�x1� , . . . ,ck�xN��T,

• and the spatial variation of the scattering,a= �a1 , . . . ,aN�T= �a�x1� , . . . ,a�xN��T.

For the multiwavelength inversions, these column vec-ors are concatenated into long, multiwavelength columnectors:

h = h1

]

hL =

h11

]

hM1

]

h1L

]

hML

, c = c1

]

cK =

c11

]

c1N

]

cK1

]

cKN

.

. Gradient, Hessian, and Jacobianith continuous variables, the problem was to find the

istributions ck�x� and a�x� that minimized the error func-ional in Eq. (20) given the continuous measured data h.n the discrete case, the problem is still to find the chro-ophore concentrations and scattering that minimize an

rror functional, but they are no longer continuous func-ions but finite-length vectors c and a. For succinctness,hey will be combined together into a single vector of un-nowns:

u = �c

a� . �29�

he error functional E is no longer defined as an integralut as a sum over image pixels and wavelengths,

E�u� =1

2�l=1

L

�m=1

M

�hml�u� − hm

l�2 =1

2eTe, �30�

here e=h− h is the vector of residuals. From now on Eill be used to refer to this discrete version of the error

unctional.One way to search for the minimum of E is to use the

terative inversion scheme known as Newton’s method37]. A brief description of it is given here for complete-ess. First note that E is a continuous function of the un-nown parameter vector u, and so its Taylor expansionbout u0 exists as

E�u0 + �� � E�u0� + gT� +1

2�TH� + ¯ , �31�

here � represents a perturbation to the unknowns u0.he first-order derivative vector g and second-order de-ivative matrix H are called the (functional) gradient andessian, respectively. Differentiating the Taylor series inq. (31) and setting it to zero gives g=−H�, which can be

earranged into an expression for an “update” vector

� = − H−1g � − �JTJ�−1Je. �32�

t each step, the latest estimate of u is updated, u←u�, until the value of u that minimizes E, or a good ap-roximation to it, is reached. The gradient vector g andessian matrix H are given by

Page 9: Estimating chromophore distributions from multiwavelength photoacoustic images

tbgfi3tFl

Tcts

Tnwt

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 451

g = �E�c11

, . . . ,�E

�cKN� �E

�a1, . . . ,

�E�aN

�T

, �33�

H = ��2E

�c112

. . .�2E

�c11�cKN

�2E�c11�a1

. . .�2E

�c11�aN

] � ] ] � ]

�2E�cKN�c11

. . .�2E

�cKN2

�2E�cKN�a1

. . .�2E

�cKN�aN

�2E�a1�c11

. . .�2E

�a1�cKN

�2E

�a12

. . .�2E

�a1�aN

] � ] ] � ]

�2E�aN�c11

. . .�2E

�aN�cKN

�2E�aN�a1

. . .�2E

�aN2

� �34�

EAdniosccstetmw(t

uf“iGrlsw

sostTdateiTtn

Note that the gradient, unlike the Hessian, depends onhe measured data h. Both the Hessian and gradient cane calculated from the Jacobian matrix, as H�JTJ and=JTe, although the gradient can be calculated more ef-ciently using the adjoint model described in Subsection.C. The elements of the Jacobian matrix are the sensi-ivities of the model output h to changes in the unknowns.or instance, the Jacobian matrices for c and a at wave-

ength l are

Jcl = �

�h1l

�c1

. . .�h1

l

�cKN

] � ]

�hMl

�c1

. . .�hM

l

�cKN

� , �35�

Jal = �

�h1l

�a1

. . .�h1

l

�aN

] � ]

�hMl

�a1

. . .�hM

l

�aN

� . �36�

he elements of the single-wavelength Jacobians can bealculated column by column using Eqs. (23)–(28), andhe multiwavelength Jacobian matrix can then be con-tructed as

J = �Jc

1 Ja1

Jc2 Ja

2

]

JcL Ja

L� . �37�

his potentially huge multiwavelength Jacobian does notecessarily need to be stored in full, because the multi-avelength Hessian and gradient could be calculated as

he sum of single-wavelength Hessians and gradients.

. Exampleproof-of-principle numerical example will be used to

emonstrate that the absorption-scattering nonunique-ess is not a problem for multiwavelength chromophore

nversions when using prior knowledge of the dependencef the scattering on wavelength. Figures 8A and 8B belowhow the spatial distributions of a single chromophoreoncentration c�x� and the spatial part of the scatteringoefficient a�x�, respectively. A small 25�50 mesh repre-enting 3.6 mm�7.5 mm was deliberately chosen to keephe size of the inversion reasonable. Even with this smallxample consisting of only two unknown parameter dis-ributions, the number of unknowns is 2500, the Hessianatrix has 25002=6.25�106 elements and the multi-avelength Jacobian 5000�2500=12.5�106 elements.

The large scale of this type of inversion is discussed fur-her in Section 5 below.)

The FE model of light transport described above wassed both to simulate the “measured” data h and as the

orward model in the inversion scheme. To mitigate thisinverse crime” the former was calculated on and linearlynterpolated from a larger, noncoincident mesh, andaussian noise was added to give a mean signal-to-noise

atio in the “measured” images of �30 dB. The wave-ength dependence of the chromophore was chosen to beimilar to that of oxyhemoglobin, and the scatteringavelength dependence was set to b=1.3, see Fig. 6.The ranges of the resulting absorption and reduced

cattering coefficients are shown in Table 1 as a functionf wavelength. (A scaling factor was introduced to thecattering, �s��x�=a�x�a0−b mm−1, where a0=500, so thathe unknowns c�x� and a�x� were of similar magnitude.)hese coefficients were chosen to be sufficiently small toemonstrate clearly the principle that the scattering-bsorption nonuniqueness could be overcome using mul-iwavelength data. When the absorption or scattering co-fficient is large, the fluence may be small at some pointsn the image, resulting in a low signal-to-noise ratio there.he practical question of the range of signal-to-noise ra-

ios for which this inversion is achievable in practice isot tackled directly in this paper.

Page 10: Estimating chromophore distributions from multiwavelength photoacoustic images

FAsJifiubHetam�ttsi

aadtttvwsT

las

isTFdwtniIl

muccsst

Fdttntrttn

452 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

. Ill-Conditioning and Regularizationsolution for the update � in the matrix Eq. (32) will

trictly exist only when the Hessian, or its approximationTJ, has an inverse. However, the existence of the inverse

s not sufficient to ensure that the updates � calculatedrom it are stable, in the sense that a small perturbationn the measured data leads to a small perturbation in thepdate �. This will be the case only if the condition num-er of the Hessian is not too large. The conditioning of theessian can be obtained, for small problems such as the

xample here, by calculating the SVD H=U�VT, wherehe columns of the full matrices U and V contain the “left”nd “right” singular vectors ui and vi, and � is a diagonalatrix containing the corresponding singular values,

1 ,�2 , . . . . The singular values in � appear in order fromhe largest at the top to the smallest at the bottom, andhe condition number is the ratio of the largest to themallest [38]. Substituting the SVD representation of Hnto Eq. (32) gives a way to calculate the update [39]

� = − V�−1UTg = − �i

uiTg

�ivi. �38�

From this it is clear that if the smallest singular valuesre very small then their reciprocals will be very largend will make the solution overly sensitive to noise in theata. Measures taken to prevent this are genericallyermed regularization. Two popular ways are theruncated-SVD, in which the sum over i in Eq. (38) isruncated to include only some of the singular values andectors in the reconstructed image, and Tikhonov, inhich a filter �i

2 / ��i2+�2� is used to weight the inverted

ingular values. � is a variable regularization parameter.he value of i at which to truncate the SVD, or the regu-

1

2

3

4

5x 10

−3

650 700 750

wavele

absorption

specificab

sorptio

ncoeffi

cien

t(m

m-1(g

/l)-1

)

Fig. 6. Wavelength dependence of the chromophore abs

Table 1. Ranges of the Absorption and ReducedScattering Coefficients Used in the

Multiwavelength Inversion Example as aFunction of Wavelength

(nm) �a �mm−1� �s� �mm−1�

650 0.01–0.02 0.5–1.1750 0.01–0.03 0.4–0.9850 0.02–0.06 0.4–0.8950 0.02–0.07 0.3–0.7

arization parameter �, can be chosen automatically usingmethod such as the L-curve, or in order to maximize the

ubjective quality of the image, as was done here.The Hessian was calculated for the example above us-

ng one, two, and four wavelengths. Figure 7 shows theingular value spectra for these Hessians. The effect ofikhonov regularization on the spectrum is also shown.irst, it is clear that in the single wavelenth case the con-ition number is huge, �1023, and so the single-avelength Hessian is very ill-conditioned–evidence of

he absorption-scattering nonuniqueness. Increasing theumber of wavelengths in the Hessian from one to two

mproves the conditioning considerably, to perhaps 1011.nterestingly, further increases in the number of wave-engths do not improve the conditioning more.

Intuitively, if the number of (independent) measure-ent samples, here L�M, is greater than the number ofnknown parameters, �K+1��N, then there is a goodhance the nonuniqueness in the inversion will be over-ome. Here, the inversion is for two parameters, and theignificant reduction in the condition number of the Hes-ian when two wavelengths are included is indicative ofhis removal of the nonuniqueness.

850 900 9500.06

0.08

0.10

0.12

(nm)

scattering

scattering,a0 λ -b

and scattering used in the example in Subsection 4.E.

0 500 1000 1500 2000 2500

10−15

10−10

10−5

100

105

Tikhonov

single λ

index, i

sing

ular

valuesσ i

ig. 7. Singular value spectrum of the Hessian matrix whenata at one, two, and four wavelengths are used in its construc-ion. The nonuniqueness in the single wavelength case gives riseo a gap in the singular value spectrum of several orders of mag-itude. The nonuniqueness, and therefore the gap in the spec-rum, disappears when two or more wavelengths are used in theeconstruction. However, the condition number is still large dueo a second type of ill-posedness caused by the diffusive nature ofhe light propagation. This can be treated using standard tech-iques such as Tikhonov regularization, as shown.

800

ngthorption

Page 11: Estimating chromophore distributions from multiwavelength photoacoustic images

phoolb

ide8ftw

5IofPdpaowtatp

Fp(s

Fcecde

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 453

However, because of the diffusive nature of light trans-ort in scattering media, and the subsequent blurring ofigh-spatial-frequency information, there remains a sec-nd type of ill-posedness, indicated by the gradual decayf the singular values. This type of ill-posedness is muchess severe than the nonuniqueness and can be overcomey applying Tikhonov (or other) regularization.The results of the inversion for c�x� and a�x�, with an

nitial guess of 5 everywhere for both parameters, usingata at four wavelengths, and following three Newton it-rations with Tikhonov regularization are shown in Fig.. Profiles are shown in Fig. 9. Although the noise has af-ected the estimate of a more than that for c, it is clearhat both parameters have successfully been recoveredithout any “crosstalk” between them.

−3 −2 −1 0 1 2 3

5

7

9

11

13

15

mm

chromop

hore

concen

tration(g/l)

ig. 9. (Color online) Profiles for the multiwavelength inversionrofile through the concentration distribution, and a central versolid), initial guess (dashed), estimate after one Newton iteratiopond to slices through Figs. 8C and 8D.

A

B

ig. 8. (Color online) Results from Newton inversion using threntration distribution c�x� in (g/l). (B) True scattering distributired scattering distribution estimate. The initial distributions wentration and the scattering. The image dimensions are 3.6 mistributions have been recovered, without crosstalk between thexample.

. DISCUSSIONn this paper, several approximations have been used inrder to find a way to extract chromophore concentrationsrom photoacoustic images. First, it was assumed that aAT image gives a measurement of the absorbed energyensity distribution h�x�. This is true only if (a) the initialressure distribution p0�x� has been recovered exactly,nd (b) the Grüneisen parameter �x� is known. Neitherf these conditions will be quite true in practice, althoughith the use of calibrated broadband ultrasound detec-

ors, a complete set of acoustic pressure measurements onsurface surrounding p0, and an exact image reconstruc-

ion algorithm, a good quantitative estimate of p0�x� isossible.

0 0.5 1 1.5 2 2.5 3 3.54

5

6

7

8

9

10

scatterin

gpa

rametera(x)

mmple described in Subsection 4.E and Fig. 8. A central horizontalrofile through the scattering distribution show the true valuesed–dashed) and after three iterations (dotted). The latter corre-

5

10

15

4

6

8

10

tions with Tikhonov regularization. (A) True chromophore con-. (C) Recovered chromophore concentration estimate. (D) Recov-sen to be uniform and equal to 5 for both the chromophore con-mm. Profiles through these images are shown in Fig. 9. Bothough the scattering is clearly more sensitive to the noise in this

examtical pn (dott

C

D

ee iteraon a�x�ere chom�7.5m, alth

Page 12: Estimating chromophore distributions from multiwavelength photoacoustic images

tt(rots

slltbpssmimiPawaatdns(humtecowm

sTtdwaspd

6Tafistiic(

psttont

ATTS

R

1

1

1

1

1

454 J. Opt. Soc. Am. A/Vol. 26, No. 2 /February 2009 Cox et al.

Second, an approximate light model was used to modelhe light transport in the tissue. The diffusion approxima-ion will be accurate for depths greater than �1/�s��0.5−1 mm in tissue), and if a more accurate model isequired for shallower depths, either one of the higher-rder Pn approximations to the radiative transport equa-ion [28] or the delta-Eddington adjustment to the diffu-ion equation formulation [6,40,41] could be used.

Third, the simulations in this paper are in two dimen-ions in order to keep the number of unknown parametersow and therefore the inversion tractable. In reality, theight will propagate in three dimensions. However, solu-ions to the diffusion equation do not differ fundamentallyetween two and three dimensions, and so the inverseroblem is expected to behave similarly in three dimen-ions. Even for the very simple 2D example given in Sub-ection 4.E, the multiwavelength Jacobian required al-ost 100 MB storage. Clearly for more realistic problems,

n three dimensions and with more chromophores, theatrices could easily become very large indeed. A 1 cm3

mage at 100 �m resolution—achievable with currentAT technology—has 1�106 voxels. If four chromophoresnd scattering are included in the inversion, the Hessianill have 25�1012 elements, requiring hundreds of ter-bytes of storage. This is a large-scale inverse problem,nd while Newton’s method was used here to demonstratehe principle that knowledge of the wavelength-ependence of scatter can be used to overcome the nonu-iqueness, for a large scale problem it would not be fea-ible to store the Hessian, let alone calculate its inverse.This might be feasible up to a point with state-of-the-artigh-performance computing, but such facilities are notniversally available.) In this case alternative approachesust be used. One key step is the adjoint model, Eq. (14),

hat can be used to calculate the gradients efficiently,ven for large-scale problems. If the gradients can be cal-ulated, then conjugate-gradient or quasi-Newton meth-ds such as BFGS could be used to tackle the inversion,hich would obviate the need to calculate the Hessianatrix directly.Another practical issue of interest is the range of

ignal-to-noise ratios over which this inversion will work.his is not investigated in this paper, but is seems likelyhat when the target tissue is illuminated from just oneirection, there will be a trade-off between the depth tohich this inversion is accurate and the degree of attenu-tion of the light (the magnitude of the absorption andcattering coefficients). In some circumstances it may beossible to design illumination geometries to mitigate thisifficulty.

. CONCLUSIONShe nonlinear optical inversion of photoacoustic (PAT) im-ges for chromophore concentrations and scattering coef-cients was described, and a framework given for theirolution. The principle contributions of this paper are (1)o show that different chromophores can be separated us-ng a multiwavelength approach when the optical scatter-ng is known (an adjoint model was provided for the effi-ient calculation of the functional gradients in this case),2) to show that a scattering-absorption nonuniqueness

revents inversions for the absorption coefficient from aingle-wavelength photoacoustic image unless the scat-ering is known a priori, and (3) the demonstration thathe use of prior knowledge of the wavelength dependencef the scattering is sufficient to overcome this nonunique-ess and allow the recovery of the concentration distribu-ions of the constituent chromphores.

CKNOWLEDGMENTShe authors thank Jan Laufer for helpful discussions.his work was funded by the Engineering and Physicalciences Research Council (UK) (EPSRC).

EFERENCES1. R. A. Kruger, P. Liu, Y. R. Fang, and C. R. Appledorn,

“Photoacoustic ultrasound (PAUS)-reconstructiontomography,” Med. Phys. 22, 1605–1609 (1995).

2. C. G. A. Hoelen, F. F. M. de Mul, R. Pongers, and A.Dekker, “Three-dimensional photoacoustic imaging of bloodvessels in tissue,” Opt. Lett. 23, 648–650 (1998).

3. X. Wang, Y. Pang, G. Ku, X. Xie, G. Stoica, and L. V. Wang,“Noninvasive laser-induced photoacoustic tomography forstructural and functional in vivo imaging of the brain,”Nat. Biotechnol. 21, 803–806 (2003).

4. M. Xu and L. V. Wang, “Photoacoustic imaging inbiomedicine,” Rev. Sci. Instrum. 77, 041101 (2006).

5. E. Z. Zhang, J. G. Laufer, and P. C. Beard, “Backward-mode multiwavelength photoacoustic scanner using aplanar Fabry-Perot polymer film ultrasound sensor forhigh-resolution three-dimensional imaging of biologicaltissues,” Appl. Opt. 47, 561–577 (2008).

6. J. G. Laufer, C. Elwell, D. Delpy, and P. Beard, “In vitromeasurements of absolute blood oxygen saturation usingpulsed near-infrared photoacoustic spectroscopy: accuracyand resolution,” Phys. Med. Biol. 50, 4409–4428 (2005).

7. J. G. Laufer, C. Elwell, D. Delpy, and P. Beard,“Quantitative spatially resolved measurement of tissuechromophore concentrations using photoacousticspectroscopy: application to the measurement of bloodoxygenation and haemoglobin concentration,” Phys. Med.Biol. 52, 141–168 (2007).

8. M.-L. Li, J.-T. Oh, X. Xie, G. Ku, W. Wang, C. Li, G. Lungu,G. Stoica, and L. V. Wang, “Simultaneous molecular andhypoxia imaging of brain tumours in vivo usingspectroscopic photoacoustic tomography,” Proc. IEEE 96,481–489 (2008).

9. B. T. Cox, S. Arridge, K. Köstli, and P. Beard, “Quantitativephotoacoustic imaging: fitting a model of light transport tothe initial pressure distribution,” Proc. SPIE 5697, 49–55(2005).

0. B. T. Cox, S. R. Arridge, K. Köstli, and P. C. Beard, “2Dquantitative photoacoustic image reconstruction ofabsorption distributions in scattering media using a simpleiterative method,” Appl. Opt. 45, 1866–1874 (2006).

1. K. Maslov, M. Sivaramakrishnan, H. F. Zhang, G. Stoica,and L. V. Wang, “Technical considerations in quantitativeblood oxygenation measurement using photoacousticmicroscopy in vivo,” Proc. SPIE 6086, 60860R (2006).

2. K. M. Stantz, B. Liu, M. Cao, D. Reinecke, K. Miller, and R.Kruger, “Photoacoustic spectroscopic imaging of intra-tumour heterogeneity and molecular identification,” Proc.SPIE 6086, 608605 (2006).

3. X. Wang, X. Xie, G. Ku, and L. V. Wang, “Noninvasiveimaging of hemoglobin concentration and oxygenation inthe rat brain using high-resolution photoacoustic imaging,”J. Biomed. Opt. 11, 024015 (2006).

4. K. Maslov, H. F. Zhang, and L. V. Wang, “Effects ofwavelength-dependent fluence attenuation on thenoninvasive photoacoustic imaging of hemoglobin oxygen

Page 13: Estimating chromophore distributions from multiwavelength photoacoustic images

1

1

1

1

1

2

2

2

2

2

2

2

2

2

2

3

3

3

3

3

3

3

3

3

3

4

4

Cox et al. Vol. 26, No. 2 /February 2009 /J. Opt. Soc. Am. A 455

saturation in subcutaneous vasculature in vivo,” InverseProbl. 23, S113–S122 (2007).

5. M. Sivaramakrishnan, K. Maslov, H. F. Zhang, G. Stoica,and L. V. Wang, “Limitations of quantitative photoacousticmeasurements of blood oxygenation in small vessels,” Phys.Med. Biol. 52, 1349–1361 (2008).

6. A. de la Zerda, C. Zavaleta, S. Kere, S. Vaithilingam, S.Bodapati, Z. Liu, J. Levi, B. R. Smith, T.-J. Ma, O. Oralkan,Z. Cheng, X. Chen, H. Dai, B. T. Khuri-Yakub, and S. S.Gambhir, “Carbon nanotubes as photoacoustic molecularimaging agents in living mice,” Nat. Nanotechnol. 3,557–561 (2008).

7. B. T. Cox, S. R. Arridge, and P. C. Beard, “Simultaneousestimation of chromophore concentration and scatteringdistributions from multiwavelength photoacoustic images,”Proc. SPIE 6856, 68560Y (2008).

8. B. T. Cox, S. Kara, S. R. Arridge, and P. C. Beard, “K-spacepropagation models for acoustically heterogeneous media:application to biomedical photoacoustics,” J. Acoust. Soc.Am. 121, 3453–3464 (2007).

9. B. T. Cox and P. C. Beard, “Fast calculation of pulsedphotoacoustic fields in fluids using k-space methods,” J.Acoust. Soc. Am. 117, 3616–3627 (2005).

0. M. Agranovsky, P. Kuchment, and L. Kunyansky, “Onreconstruction formulas and algorithms for thermoacousticand photoacoustic tomography,” arXiv:0706.1303v1 (2007).

1. A. Corlu, R. Choe, T. Durduran, K. Lee, M. Schweiger, S.Arridge, E. Hillman, and A. Yodh, “Diffuse opticaltomography with spectral constraints and wavelengthoptimization,” Appl. Opt. 44, 2082–2093 (2005).

2. A. J. Welch and M. van Gemert, Optical-Thermal Responseof Laser-Irradiated Tissue (Plenum, 1995).

3. T. L. Troy and S. N. Thenadil, “Optical properties of humanskin in the near infrared wavelength range of1000 to 2200 nm,” J. Biomed. Opt. 6, 167–176 (2001).

4. A. N. Bashkatov, E. A. Genina, V. I. Kochubey, and V. V.Tuchin, “Optical properties of human skin, subcutaneousand mucous tissues in the wavelength range from400 to 2000 nm,” J. Phys. D: Appl. Phys. 38, 2543–2555(2005).

5. S.-H. Tseng, A. Grant, and A. J. Durkin, “In vivodetermination of skin near-infrared optical propertiesusing diffuse optical spectroscopy,” J. Biomed. Opt. 13,014016 (2008).

6. S. L. Jacques and L. Wang, “Monte Carlo Modeling of LightTransport in Tissues,” in Optical-Thermal Response ofLaser-Irradiated Tissue, A. J. Welch and M. J. C. VanGemert, eds. (Plenum 1995).

7. L. Wang, S. L. Jacques, and L. Zheng, “MCML -Monte

Carlo modeling of light transport in multi-layered tissues,”Comput. Methods Programs Biomed. 47, 131–146 (1995).

8. S. R. Arridge, “Optical tomography in medical imaging,”Inverse Probl. 15, R41–R93 (1999).

9. S. Arridge, M. Schweiger, M. Hiraoka, and D. Delpy, “Afinite element approach for modelling photon transport intissue,” Med. Phys. 20, 299–309 (1993).

0. M. Schweiger, S. Arridge, M. Hiraoka, and D. Delpy, “Thefinite element method for the propagation of light inscattering media: Boundary and source conditions,” Med.Phys. 22, 1779–1792 (1995).

1. Z. Yuan and H. Jiang, “Quantitative photoacoustictomography: Recovery of optical absorption coefficientmaps of heterogeneous media,” Appl. Phys. Lett. 88,231101 (2006).

2. B. Banerjee, S. Bagchi, R. M. Vasu, and D. Roy,“Quantitative photoacoustic tomography from boundarypressure measurements: noniterative recovery of opticalabsorption coefficient from the reconstructed absorbedenergy map,” J. Opt. Soc. Am. A 25, 2347–2356 (2008).

3. J. Ripoll and V. Ntziachristos, “Quantitative point sourcephotoacoustic inversion formulas for scattering andabsorbing media,” Phys. Rev. E 71, 031912 (2005).

4. L. Yin, Q. Wang, Q. Zhang, and H. Jiang, “Tomographicimaging of absolute optical absorption coefficient in turbidmedia using combined photoacoustic and diffusing lightmeasurements,” Opt. Lett. 32, 2556–2558 (2007).

5. Z. Yuan, Q. Wang, and H. Jiang, “Reconstruction of opticalabsorption coefficient maps of heterogeneous media byphotoacoustic tomography coupled with diffusion equationbased regularized Newton method,” Opt. Express 15,18076–18081 (2007).

6. B. T. Cox, S. R. Arridge, and P. C. Beard, “Gradient-basedquantitative photoacoustic image reconstruction formolecular imaging,” Proc. SPIE 6437, 64371T (2007).

7. W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P.Flannery, Numerical Recipes in C��, The Art of ScientificComputing (Cambridge U. Press, 2005).

8. L. N. Trefethen and D. Bau, Numerical Linear Algebra(SIAM, 1997).

9. P. C. Hansen, “Deconvolution and regularization withToeplitz matrices,” Numer. Algorithms 29, 323–378 (2002).

0. T. Spott and L. O. Svaasand, “Collimated light sources inthe diffusion approximation,” Appl. Opt. 39, 6453–6465(2000).

1. B. T. Cox, S. R. Arridge, and P. C. Beard, “Quantitativephotoacoustic image reconstruction for molecular imaging,”Proc. SPIE 6086, 60861M (2006).


Recommended