ArticlePDF Available

Tomographic bioluminescence imaging by use of a combined optical-PET (OPET) system: A computer simulation feasibility study

IOP Publishing
Physics in Medicine & Biology
Authors:

Abstract and Figures

The feasibility and limits in performing tomographic bioluminescence imaging with a combined optical-PET (OPET) system were explored by simulating its image formation process. A micro-MRI based virtual mouse phantom was assigned appropriate tissue optical properties to each of its segmented internal organs at wavelengths spanning the emission spectrum of the firefly luciferase at 37 degrees C. The TOAST finite-element code was employed to simulate the diffuse transport of photons emitted from bioluminescence sources in the mouse. OPET measurements were simulated for single-point, two-point and distributed bioluminescence sources located in different organs such as the liver, the kidneys and the gut. An expectation maximization code was employed to recover the intensity and location of these simulated sources. It was found that spectrally resolved measurements were necessary in order to perform tomographic bioluminescence imaging. The true location of emission sources could be recovered if the mouse background optical properties were known a priori. The assumption of a homogeneous optical property background proved inadequate for describing photon transport in optically heterogeneous tissues and led to inaccurate source localization in the reconstructed images. The simulation results pointed out specific methodological challenges that need to be addressed before a practical implementation of OPET-based bioluminescence tomography is achieved.
Content may be subject to copyright.
Tomographic bioluminescence imaging by use of a combined
optical-PET (OPET) system: a computer simulation feasibility
study
George Alexandrakis
1
, Fernando R Rannou
2
, and Arion F Chatziioannou
1
1 David Geffen School of Medicine at UCLA, Crump Institute for Molecular Imaging, University of
California, 700 Westwood Plaza, Los Angeles, CA 90095, USA
2 Departamento de Ingenieria Informatica, Universidad de Santiago de Chile (USACH), Av. Ecuador
3659, Santiago, Chile
Abstract
The feasibility and limits in performing tomographic bioluminescence imaging with a combined
optical-PET (OPET) system was explored by simulating its image formation process. A micro-MRI
based virtual mouse phantom was assigned appropriate tissue optical properties to each of its
segmented internal organs at wavelengths spanning the emission spectrum of the firefly luciferase
at 37 °C. The TOAST finite-element code was employed to simulate the diffuse transport of photons
emitted from bioluminescence sources in the mouse. OPET measurements were simulated for single-
point, two-point and distributed bioluminescence sources located in different organs such as the liver,
the kidneys and the gut. An expectation maximization code was employed to recover the intensity
and location of these simulated sources. It was found that spectrally resolved measurements were
necessary in order to perform tomographic bioluminescence imaging. The true location of emission
sources could be recovered if the mouse background optical properties were known a priori.
Assumption of a homogeneous optical property background proved inadequate for describing photon
transport in optically heterogeneous tissues and lead to inaccurate source localization in the
reconstructed images. The simulation results pointed out specific methodological challenges that
need to be addressed before a practical implementation of OPET-based bioluminescence tomography
is achieved.
1. Introduction
The translation of putative cancer therapy agents to the clinic can greatly be facilitated by non-
invasive imaging methods that allow high throughput screening in animal models.
Additionally, time-lapse whole body imaging of animals bearing xenografts of appropriately
labelled cancer cells can provide new information on tumour growth dynamics and metastasis
patterns that is not possible to obtain by invasive experimental approaches. Bioluminescence
imaging has recently emerged as a modality that can meet these research needs by use of low
cost and easy to operate equipment (Contag et al 1998). The capacity to detect optical photons
with high efficiency and the intrinsically low autoluminescence background of mice make
bioluminescence detection a very sensitive technique if photon sources are located near the
tissue surface. However, most bioluminescence probes currently in use emit the bulk of their
optical signal at green wavelengths and are therefore attenuated very strongly if they are located
deep in tissues (Troy et al 2004). In addition, planar images of weak sources near the surface
E-mail: ArChatziioann@mednet.ucla.edu.
(1)
Author: Please provide journal title in ‘Arridge et al 1993’ and ‘Contag et al 1998’.
NIH Public Access
Author Manuscript
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
Published in final edited form as:
Phys Med Biol. 2005 September 7; 50(17): 4225–4241.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
could look identical to ones produced by stronger sources deeper in tissue (Kuo et al 2004).
Tomographic imaging is therefore required for source localization. The capacity of
bioluminescence imaging in producing accurate tomographic reconstructions is a topic of
ongoing investigation (Kuo et al 2004, Wang et al 2004, Chaudhari et al 2005).
Positron emission tomography (PET) is a well-established modality for small animal imaging
(Tai et al 2001, Chatziioannou 2002) and several commercial systems are now available. PET
can perform functional imaging of physiological processes with exquisite specificity if an
appropriate radiolabeled tracer analogue exists. New PET tracers that allow monitoring of
previously inaccessible physiological processes or simply ones that confer increased specificity
of detection in small animals can subsequently be tested on humans in the clinic without
modification (Ray et al 2004). Despite this direct translational advantage PET is hindered by
the need to use positron emitting tracers, which usually requires the presence of a nearby
cyclotron. This makes PET methodology a valuable but expensive and lower throughput tool.
A single device that will perform simultaneously both optical and PET (OPET) imaging on
the same animal is currently under development in our laboratory (Prout et al 2004, Rannou
et al 2004). The OPET radial field of view is only slightly larger than the mouse torso (figure
1). Both bioluminescence and scintillation photons from the converted annihilation gammas
are detected by photomultiplier tubes at the back end of the PET crystals and the two signal
types are distinguished by their very different pulse amplitudes (Prout et al 2004). Use of fusion
reporter gene probes co-expressing optical and PET signals under the same promoter (Ray et
al 2003) will produce images that are intrinsically co-registered in space. This unique advantage
of the OPET device will enable direct comparison of how well the optical and PET modalities
can localize and how sensitively they can detect emission sources existing at different tissue
depths.
The purpose of this work is to explore the feasibility and limits in performing bioluminescence
tomographic measurements with the proposed OPET system, by simulating its image formation
process. A mouse micro-MRI image volume (Segars et al 2004) provided the anatomical
features for a virtual phantom. Each of the segmented tissues was assigned appropriate optical
properties which were collected from the literature and were adjusted for blood content and
oxygenation levels (see the methods). OPET measurements were simulated for single-point,
two-point and distributed bioluminescence sources located in different organs such as the liver,
the kidneys and the gut. An expectation maximization (EM) code was employed to recover the
intensity and location of these simulated sources. The effects of spectrally-resolved
measurements, detector spatial sampling and the adequacy of approximating the mouse
background optical properties by a uniform medium were explored.
2. Methods
In order to assess the effect of background optical properties on the OPET image formation
process, photon propagation calculations were performed in an optically heterogeneous virtual
mouse phantom. The anatomical definitions of the different tissues in this phantom were
provided by a publicly available time-gated high resolution MRI image volume known as the
mouse phantom for molecular imaging research (Segars et al 2004) (MOBY). Different time
instances of the mouse image volume could be generated by use of a program utilizing the
natural spline definitions of its segmented organ volumes (Segars et al 2004). For our
simulations we used the first time gate of that image volume and we implemented some further
alterations in order to simplify computations: only the mouse torso was simulated, from the
neck to the base of the tail, the tissue surface was padded out to a perfect cylinder of 27 mm
diameter and the image volume was re-sampled to 2 mm size voxels (figure 2). The amount
of padding did not exceed 2 mm in the radial direction at any location near the torso surface.
Alexandrakis et al. Page 2
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
As the original MOBY image did not include the mouse skin, the latter was defined as the first
2 mm (one voxel) in from the cylinder’s surface. As a result, the padded volume voxels were
defined to be occupied by skin. The MOBY-defined organ volumes for the gall-bladder, vas
deferens, and testes were left out due to insufficient knowledge of their optical properties from
current literature data. The space occupied by these organs as well as any space between organs
in the modified mouse torso was defined as adipose tissue.
All of the organs included in the modified phantom were assigned optical properties for
wavelengths spanning the firefly luciferase emission spectrum (500–800 nm) reported by Kuo
et al. The absorption (μ
a
) and transport scattering (μ′) coefficients assigned to each organ, as
a function of photon wavelength (λ), were estimated on the basis of a compilation of relevant
tissue optical property data reported in the literature (tables 1 and 2). An empirical function
was employed to approximate μ′
s
(λ):
μ
s
(
λ
) =
a
×
λ
b
mm
−1
,
λ
in nm (1)
where a and b were constants controlling the spectral variation in each tissue (table 1).
The amount of light absorption taking place in all tissues was assumed to depend only on the
resident oxy-haemoglobin (HbO
2
), deoxy-haemoglobin (Hb) and water (W) concentrations.
The spectral absorption coefficient (μ
a
(λ)) was then approximated as a weighted sum of the
three constituent absorption coefficients (μ
aHbO
s
(λ), μ
aHb
(λ) and μ
aW
(λ)), which were
calculated from the corresponding absorbance spectra reported by Prahl (2001):
μ
a
(
λ
) =
S
B
(
xμ
aHb
(
λ
) + (1
x
)
μ
aHbO
2
(
λ
)) +
S
W
μ
aW
(
λ
)
(2)
where x = (HbO
2
)/((HbO
2
)+(Hb)) and S
B
and S
W
were heuristic scaling factors adjusted to
match the absorption data currently available in the literature for each tissue (table 2). S
B
and
S
W
were consistent with the known blood and water volume fractions respectively in the
different mouse organs (Brown et al 1997) with the exception of the kidneys and the lungs.
The blood volume fractions in these two organs predicted higher μ
a
(λ) values than those
reported in the literature.
It is important to emphasize that the tissue optical property literature data are sparse and
incomplete. The publications compiled in table 1 report on measurements performed on tissues
from different species, in vitro or in vivo, and at a great variety of other experimental conditions.
In addition, the optical properties of the mouse stomach and bowel contents were unknown
and this material could possibly shift around and mix with air pockets during image acquisition.
In this work, the stomach and bowel contents were assumed to be homogeneous and static.
They were assigned the abdomen optical properties as inferred from bulk tissue measurements
(Torricelli et al 2001). Given the above shortcomings, it should be stated that the parametric
descriptions that we assumed for equations (1) and (2) could only be interpreted as reasonable
but rough estimates of the true mouse tissue optical properties.
The propagation of bioluminescence photons in the mouse torso phantom was first simulated
by Monte Carlo (MC) methods. An open source MC code, originally intended for computing
the time-resolved reflectance from photon sources located on the surface of optically
heterogeneous media (Boas et al 2002), was modified to simulate the migration of photons
emanating from sources at the mouse torso interior. The modified MC code required as inputs
the discretized bioluminescence emission spectra, the voxelized distributed photon sources and
the mouse torso optical property phantom. Although the background optical properties were a
continuous function of photon wavelength, we have approximated the firefly luciferase
emission spectrum (Kuo et al 2004) as five discrete intensity bins of 25 nm width each, centred
Alexandrakis et al. Page 3
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
at 600, 625, 650, 675 and 700 nm respectively. Photons emitted at longer wavelengths
comprised <1% of the total emission intensity and were ignored. Though photons from the
firefly luciferase spectrum could be emitted at wavelengths as low as 500 nm, we found that
these underwent extreme attenuation in all tissues and were therefore not included in the
simulations. The mouse torso optical property phantom was defined as a cylindrical volume
of 2 mm side cubic voxels. This rather coarse voxel size was employed in order to reduce
memory and computation time requirements. Though the phantom’s interior was voxelized,
its air–tissue boundary was modelled as a smooth cylindrical surface. This was done to ensure
that photons were refracting out of the phantom at the appropriate surface geometry, free of
voxelation artefacts. Photons were subsequently scored at virtual detectors, of 2 mm × 2 mm
size, which were located on the phantom’s surface.
As MC simulations were time consuming (15 minutes on a 1.6 GHz PC for 10
6
photon histories)
and their estimates for the photon flux reaching the phantom’s surface inevitably contained
statistical noise, we explored the alternative of using a deterministic code to perform these
calculations. The publicly available time-resolved optical absorption and scattering
tomography (TOAST) finite-element code (Arridge et al 1993, Schweiger et al 1995), which
solves the diffusion approximation to the radiative transfer equation, was employed to compute
the boundary fluxes resulting from point sources located in the mouse torso interior. The same
optical property mouse phantom was used to match the MC simulation conditions. Though not
used in this work, a smaller voxel size or even a non-segmented image volume of continuously
varying optical properties could have been used to assign optical properties to the finite-element
nodes defined in the TOAST code. A mesh generation tool which was part of the TOAST
package was employed to discretize the mouse torso volume into tetrahedrons of an average
0.65 mm side size. This was sufficiently small to confer numerical stability to the diffusion
computations. By comparing TOAST predictions with corresponding MC simulations we
verified that, with the exception of sources located at <1 mm from the torso surface, the
diffusion approximation was adequate. Figure 3 demonstrates the agreement between MC and
TOAST in the case of a point source in the mouse gut, located 3 mm from the torso surface
and emitting photons at 650 nm, for detector locations on the same transverse plane as the
source. Similar results were found for all other bioluminescence emission wavelengths, source
and detector locations considered in this work. However, in contrast to the MC code, TOAST
could not handle non-scattering regions located within the geometric tissue boundaries. This
presented a problem for modelling photon transport through the bladder. We verified by MC
simulation that, for the detector field-of-view (FOV) locations examined in this work, the
photon fluxes detected from point sources located at the edge of the FOV, and nearest to the
bladder, were independent of the bladder’s presence. Given this observation, the volume
occupied by the bladder was filled with a scattering medium, chosen to be adipose tissue.
TOAST was advantageous to use over MC because it could calculate boundary photon fluxes
with similar accuracy but in addition, its computations were free of statistical noise and were
completed faster by a factor of 7, or more, on the same computer. Therefore, the TOAST code
was employed to compute the photon fluxes from emission sources, located in internal organs,
reaching each of the boundary elements of the mouse torso surface. These computations were
the first step in simulating the virtual OPET detector measurements. A controlled amount of
Gaussian noise, having zero mean and variance amplitude equal to the square root of photons
crossing each boundary element, was then added to the noiseless flux calculations. In addition,
a wavelength-dependent autoluminescence background flux (Troy et al 2004) was added to
each of those photon counts as spatially invariant Gaussian noise. It was a reasonable
approximation to assume that the angular distribution of diffuse photons exiting the mouse
torso was isotropic (Martelli et al 1999). The noise-added photon counts from each torso
boundary element were then geometrically projected onto the virtual OPET detectors by use
of appropriate solid angle factors (Gotoh and Yagi 1971, Wielopolski 1984). Finally a
Alexandrakis et al. Page 4
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
wavelength independent detection efficiency and dark noise, estimated at 10% and 35 counts
s
1
respectively (Prout et al 2004), were also modelled into the virtual OPET measurements.
Source contributions from each of the five wavelength bins in which the firefly luciferase
spectrum was discretized were weighed by the fractional spectrum area that they subtended.
The intensity of source voxels was adjusted to achieve any desired overall signal-to-noise ratio
(SNR) at the detectors.
In addition to performing virtual measurements with the OPET detectors (2.2 mm × 2.2 mm,
384 in total) high resolution detectors spanning OPET’s geometrical FOV and having an area
equal to that of boundary finite elements (HiResOPET, 0.68 mm × 0.65 mm, 1740 in total)
were also utilized. These idealized HiResOPET virtual detectors were located on the mouse
torso surface and had 100% detection efficiency. Thus the performance of OPET detectors
could be compared to that of superior ones capable of higher spatial sampling and ideal
detection efficiency.
Bioluminescent source distributions were reconstructed by use of an EM algorithm for
emission tomography (Shepp and Vardi 1982). One of the required inputs for the EM code was
the P-matrix—the collection of probabilities that a photon emitted from any source location
in the image volume would reach any given detector. The P-matrix needs to be computed once
for any particular choice of tissue optical properties and organ topologies. We employed
TOAST to compute the P-matrix for point sources located at the centre of all 2 mm size voxels
of the virtual mouse torso phantom for all five wavelengths approximating the luciferase
emission spectrum. These were very time-consuming calculations that collectively took
approximately 3 weeks to complete on a cluster of 27 CPUs (2 GHz, 4 GB RAM each) which
were available at our institute. The EM algorithm utilized the P-matrix to reconstruct the most
likely distributed source that generated the simulated measurement data. A weakness of the
EM algorithm was that there was no official stopping criterion (Hebert and Leahy 1989) and
the latter had to be determined empirically. For our virtual mouse torso phantom and the OPET
detection geometry (Rannou et al 2004) we found that the reconstructed images did not change
significantly after 2500 iterations. All of the images presented in this work were reconstructed
with 5000 iterations, which were completed in approximately 1 h on a single CPU.
3. Results
The need for tomographic bioluminescence measurements was first demonstrated by
simulating the OPET detector data for the case of a point source located at the mouse torso
centre (figure 4(a)). Even for noise-free detection, there is no intuitive way to infer the true
depth of the point source, or even that it is indeed a point source, by simply looking at the
detector data image (figure 4(b)). It is even harder to infer any conclusions about the true
localization of the source when its intensity is not much larger than the detected background
noise (figure 4(c)). An additional difficulty is that tissue optical properties vary with photon
wavelength. Thus both photon propagation and detection need to be considered in a spectrally-
resolved fashion.
It was first investigated whether the proposed OPET system could attain tomographic images
from simulated monochromatic detector measurements. In these test cases the tissue optical
properties were assumed to be known and the simulated detector data were noise free. For a
point source located at the mouse torso centre emitting at 650 nm (figure 4(a)) the reconstructed
source was placed near the torso surface (figure 5(a)), which was clearly not its true location.
The apparent near-surface reconstructed source placement was also observed for simulated
sources located at different tissue depths emitting monochromatically at any wavelength in the
firefly luciferase spectrum. An attempt to improve point source localization by employing the
higher resolution virtual detectors (see the methods) was unsuccessful. In fact, the total photon
Alexandrakis et al. Page 5
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
counts reaching each detector from the reconstructed source intensity distribution were always
nearly identical to those reaching these same detectors from the true point source location. The
similarity in detected intensity profiles from two different monochromatic emission source
distributions demonstrated the ill-posed nature of the problem which has been previously
reported (Kuo et al 2004).
The ability to reconstruct the true location of point emission sources by use of measurements
at two wavelengths was then investigated. When noiseless detector data were simulated for the
point source in figure 4(a), now emitting both at 600 nm and 650 nm, the EM reconstruction
recovered its true location (figure 5(b)). Similarly, the true locations of point sources anywhere
in the mouse torso could be recovered when detector data from any two of the five wavelengths
used in this work (see the methods) were used in the reconstructions. Inclusion of noiseless
detector data at additional wavelengths did not substantially improve the quality of
reconstructed images. Furthermore, it was examined whether noisy detector data at two
wavelengths would still be sufficient for the EM reconstructions to recover the true locations
of point sources. The same point source as in figure 4(a) was simulated to have an emission
intensity sufficient to produce a total of OPET detector counts which was three standard
deviations above the sum of the overall autoluminescence background and detector dark noise.
Measurements at any two wavelengths yielded reconstructed images of a diffuse source
distribution which encompassed the true source location (e.g. figure 5(c)). Measurements at
three wavelengths improved the point source localization (figure 5(d)). Incremental inclusion
of additional measurements at up to five wavelengths (figure 5(e)) further improved
reconstructions though ever less so. Similar results were found for most point source locations
within the detectors’ geometrical FOV. Exceptions to these observations were any sources
located deep within highly attenuating tissues, e.g. at the liver centre. In the latter case and for
physiologically relevant source intensities (Troy et al 2004), photon count contributions from
wavelengths below 650 nm were lower than background noise.
The above simulations were performed under the assumption that the mouse optical properties
were known exactly. This is difficult to achieve by prior non-invasive optical measurements
even when additional anatomical information is provided by some other high resolution
imaging modality (Barnett et al 2003). In addition, time-consuming P-matrix calculations
would need to be repeated for every mouse, which is impractical. We have examined the
possibility of assuming a uniform optical property background for the mouse torso while
reconstructing the measurements obtained from the heterogeneous torso phantom. We found
that assigning the volume average optical properties of the whole mouse torso to the
homogeneous phantom did not describe photon transport well in all regions of the optically
heterogeneous torso. The reason was that the volume average properties were close to those
of liver, which were too attenuating to accurately describe the photon transport in less
attenuating tissues such as the gut. We decided to examine more closely the favourable case
of having the OPET geometrical FOV encompass the less heterogeneous middle part of the
torso which mostly contained adipose tissue and the gut (μ
a
(λ) and μ′
s
(λ) volume averages
shown in table 3). Noiseless detector data from three separate point source locations, all located
on an axial plane which was centred in the OPET FOV, were reconstructed. In the radial
direction one source was located at the cylinder’s centre, one at half-radius and one at the edge
voxel (figure 6(a)). The reconstruction of the source at the centre was poor (figure 6(b)). The
reconstructed source distribution at half-radius (figure 6(c)) was slightly mis-positioned
relative to the true point source location and was somewhat diffuse. The point source at the
cylinder’s edge (figure 6(d)) was reconstructed at its correct location. These findings
demonstrated that assumption of a uniform optical property background was appropriate only
when point sources were superficial.
Alexandrakis et al. Page 6
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
The capacity of OPET measurements to tomographically resolve two neighbouring-point
sources at different tissue depths and in different organs was then explored for the case where
the mouse optical properties were known exactly. The arbitrary criterion for considering the
two sources resolved was that in a line profile through the image a trough, dipping at >15% of
the two peak amplitudes’ mean, would exist between the two source peak intensities. First, it
was examined whether two point sources separated by a distance of 2 voxels (or 4 mm) could
be resolved. When two such sources of equal intensity were placed on either side of the mouse
torso centre in either the gut or the liver, they could not be resolved in the reconstructed images
(figure 7(a), SNR = 5). In addition, these sources could also not be resolved for unequal relative
intensity ratios (data not shown). When those source pairs were moved on either side of the
torso half-radius, they could still not be resolved if they were of equal intensities (figure 7(b)).
However, they could be resolved in the liver if their relative intensities were in the vicinity of
2:1 (figure 7(b)) but did not exceed 3:1. Finally, the two sources were moved near the torso
edge so that the one nearest to the surface was located at 1 mm depth. Even at these superficial
tissue depths two equal intensity sources in the gut could not be resolved but they could be
resolved in the liver (figure 7(c)). Nevertheless, these two sources could actually be resolved
for unequal intensity ratios ranging from 2:1 to 5:1 in the gut and for up to 10:1 in the liver
(data not shown). When the two sources were separated by a centre-to-centre distance of 3
voxels, or 6 mm, they were always resolved at all tissue locations if the true tissue optical
properties were assumed to be known. For example, two equal intensity point sources located
in the gut and the spleen respectively could be resolved under these circumstances (figure 7
(d)). However, when a uniform optical property background was assumed the two point sources
were reconstructed as a single source located half-way between the two true source locations
(figure 7(d)). In all of the above cases where the two equal intensity point sources were resolved
they were typically not reconstructed at equal intensities though they were within a factor of
4 of each other.
Interestingly, image reconstruction artefacts in the form of additional ‘phantom’ emission
sources appeared at mouse torso locations from where emitted photons had a small probability
of reaching the OPET detectors. For example in the case of two point sources, separated by 4
mm, in the gut (figure 8(a)) a phantom source appeared at the edge of the FOV and near the
liver centre (figure 8(b), SNR = 5). The artefact persisted even for noiseless detector data (figure
8(c)) though its intensity was much weaker. A similar image artefact was observed in the heart
blood pool area when the OPET FOV was shifted to the upper half of the mouse torso (data
not shown). As photons emitted from these artefactual sources were highly attenuated, they
did not contribute significantly to the total detected signal. Interestingly, these artefacts
disappeared and the two point sources were better localized when virtual measurements were
performed by the more numerous HiResOPET detectors located on the torso surface (figure 8
(d), SNR = 5). Nevertheless, even with measurements performed by these higher resolution
detectors the two equal intensity sources at the torso centre were neither resolved nor were they
reconstructed at equal intensities.
Finally the capacity to reconstruct distributed emission sources from OPET detector data was
investigated. In two test cases the segmented volumes occupied by the kidneys (encircled green
voxels in figure 9(a)) and the gut (green–yellow voxels in figure 9(d)) were each defined as
distributed sources. The same emission intensity was assigned to all source voxels with each
one contributing an overall SNR 5 at the OPET detectors. When the true torso optical
properties were utilized in the reconstructions, the true spatial extents of the distributed sources
defined by the kidneys (figure 9(b)) and the gut (figure 9(e)) were both recovered. As found
previously, assumption of a uniform optical property background resulted in artefacts. The
reconstructed kidneys volumes were shifted relative to their true locations and an apparent
exclusion of source emission activity, similar to that seen in figure 6(b), could also be observed
(figure 9(c)). The problem of emission source exclusion from the torso centre was seen more
Alexandrakis et al. Page 7
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
clearly when the gut was reconstructed while assuming a uniform optical property background
(figure 9(f )).
4. Discussion and conclusions
Simulations were performed to assess the capacity of a proposed OPET system to perform
tomographic bioluminescence imaging. Measurements at a single wavelength were found to
be inadequate for achieving three-dimensional source localization. The latter was not even
obtained when the idealized HiResOPET detectors were employed, even though these were
located on the phantom surface. Therefore the need for spectrally resolved detection to achieve
tomographic bioluminescence imaging was independent of the imaging system specifics.
Though OPET measurements at two wavelengths did produce tomographic images, source
localization improved when additional spectral measurements were included. With the
exception of sources located on the skin surface, photons emitted at wavelengths <600 nm
suffered extreme attenuation in all tissues. In addition sources located deep in highly
attenuating tissues, such as the liver centre, did not have significant signal contributions to the
detectors for wavelengths <650 nm. Therefore, image reconstructions for experimental data
from these source locations are expected to suffer not only from fewer photon counts but also
from having a smaller fraction of the firefly luciferase spectrum available for spectrally
resolved measurements.
The capacity of the OPET system to resolve two neighbouring point sources was a function of
their locations in different tissues and decreased with tissue depth. It was also observed that
two neighbouring sources were better resolved in the liver than in the gut even though the
former was a more absorbing tissue. It is hypothesized that this was a result of the higher spatial
gradient of spectral attenuation in the more absorbing tissues which would favour depth
resolution. When two equal intensity point sources were resolved, they were typically not
reconstructed at equal intensities. In addition, the range of relative intensity ratios for which
two point sources could be resolved was a function of their locations in tissues. The overall
dynamic range for resolving two neighbouring sources separated by 2 voxels (4 mm) was rather
limited, not exceeding 4:1, unless these were located within a few mm of the tissue surface. It
should nevertheless be noted that the limits in resolving neighbouring sources will not have
been fully explored until image reconstructions utilizing P-matrices of 1 mm voxel size, or
less, are performed. Given that the computation of the 2 mm voxel P-matrices used in this work
was time consuming, use of higher resolution P-matrices is presently impractical.
The spatial distribution of photon fluence reaching the tissue surface depends both on the tissue
optical properties and the size and shape of the mouse organs in which bioluminescence sources
are embedded. Exact knowledge of the tissue optical properties in every mouse being imaged
may not be practical or even possible. In addition the optical property map in the mouse changes
with breathing and food digestion. Even if that dynamic optical property map was known it
would be impractical to compute a new P-matrix for every mouse. On the other hand, the
attempted simplification of assuming a uniform optical property background was shown to be
inadequate unless emission sources were superficial. The above findings imply that
development of computationally efficient methods taking into account the differences between
individual mouse optical properties would be the key to any practical implementation of
bioluminescence tomography. In future work we plan to explore image artefacts arising from
assigning inaccurate optical properties to tissues whose anatomical boundaries are obtained by
an independent high resolution imaging modality.
The detector spatial resolution was shown to play an important role in the quality of the resulting
reconstructed images. The OPET detector spatial resolution was restricted to 2.2 mm. The latter
was equal to the centre-to-centre distance between neighbouring PET crystals (block of 8 × 8,
Alexandrakis et al. Page 8
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
2 mm × 2 mm each, 0.2 mm inter-crystal gap) which were designed to match the physical size
of the channels on a 64-channel photomultiplier tube. Unfortunately, bioluminescence
detection at 2.2 mm resolution resulted in reconstructed image artefacts in areas of high photon
attenuation in the mouse. These artifacts were eliminated when the OPET detectors were
replaced by the HiResOPET ones, which were simulated to have a width of approximately 650
μm—a resolution easily achievable by CCD cameras. These considerations indicate some of
the compromises one is faced with when designing a hybrid imaging modality system.
The proposed OPET system presents a unique opportunity to determine the relative
bioluminescence and PET tracer detection sensitivities as a function of source depth in the
mouse in a spatially co-registered setting. Optical photons undergo both strong and spatially
variable attenuation as they travel through tissues. Based on calculations performed in this
work, typical OPET detection probabilities per bioluminescence photon emitted are 0.15% for
a point source located at the mouse torso centre in the gut and 0.01% for one at the liver centre.
Near the torso surface, bioluminescence photon detection probabilities are about 2%. In
contrast, annihilation gamma photons undergo little (~20%) attenuation as they travel through
the mouse but they are also hard to stop even by high-density detector materials. As a result
small animal PET systems have an absolute sensitivity of the order of 4% (Chatziioannou
2002). It is important to remember though that the detected emission intensities from both PET
and bioluminescence sources are also the functions of their respective tracer kinetics in vivo
as well as the underlying cell biology of each tracer. Future comparisons between simultaneous
optical and PET images will enable exploring the detection sensitivity of bioluminescence
tomography under the design constraints of a dual modality system.
Acknowledgements
We would like to thank Dr Richard Taschereau for providing us with the schematic shown in figure 1. We would also
like to thank Dr Martin Schweiger and Jason Riley of University College, London, for their advice on the use of the
TOAST code. We would like to acknowledge Dr Richard Leahy, Dr Felix Darvas, and Abhijit Chaudhari of the
University of Southern California for fruitful discussions and advice on finite-element bioluminescence computations.
Finally, we would like to thank Dr Michael Patterson of McMaster University, Canada, for his comments on the
manuscript. This work was supported by the NIBIB R01-EB001458, NIH/NCI R25 CA098010:01 and DE-
FC02-02ER63520 grants.
References
Arridge SR, Schweiger M, Hiraoka M, Delpy DT. A finite element approach for modeling photon
transport in tissue. 1993;20:299–309.
Barnett AH, Culver JP, Sorensen AG, Dale A, Boas DA. Robust inference of baseline optical properties
of the human head with three-dimensional segmentation from magnetic resonance imaging. Appl Opt
2003;42:3095–108. [PubMed: 12790461]
Beek JF, Blokland P, Posthumus P, Aalders M, Pickering JW, Sterenborg HJCM, vanGemert MJC. In
vitro double-integrating-sphere optical properties of tissues between 630 and 1064 nm. Phys Med Biol
1997a;42:2255–61. [PubMed: 9394410]
Beek JF, vanStaveren HJ, Posthumus P, Sterenborg HJCM, vanGemert MJC. The optical properties of
lung as a function of respiration. Phys Med Biol 1997b;42:2263–72. [PubMed: 9394411]
Boas DA, Culver JP, Stott JJ, Dunn AK. Three dimensional Monte Carlo code for photon migration
through complex heterogeneous media including the adult human head. Opt Express 2002;10:159–
70.
Brown RP, Delp MD, Lindstedt SL, Rhomberg LR, Beliles RP. Physiological parameter values for
physiologically based pharmacokinetic models. Toxicol Ind Health 1997;13:407–84. [PubMed:
9249929]
Chatziioannou AF. Molecular imaging of small animals with dedicated PET tomographs. Eur J Nucl Med
Mol Imaging 2002;29:98–114. [PubMed: 11807613]
Alexandrakis et al. Page 9
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Chaudhari A J, Darvas F, Cherry S R and Leahy R M 2005 Resolution studies in bioluminescence and
fluorescence optical tomography using hyperspectral data acquisition Mol. Imaging Biol. (Orlando)
vol 7 (New York: Springer) p 125
Cheong WF, Prahl SA, Welch AJ. A review of the optical-properties of biological tissues. IEEE J
Quantum Elect 1990;26:2166–85.
Contag PR, Olomu IN, Stevenson DK, Contag CH. Bioluminescent indicators in living mammals.
1998;4:245–7.
Doornbos RMP, Lang R, Aalders MC, Cross FW, Sterenborg HJCM. The determination of in vivo human
tissue optical properties and absolute chromophore concentrations using spatially resolved steady-
state diffuse reflectance spectroscopy. Phys Med Biol 1999;44:967–81. [PubMed: 10232809]
Enejder AMK, Swartling J, Aruna P, Andersson-Engels S. Influence of cell shape and aggregate
formation on the optical properties of flowing whole blood. Appl Opt 2003;42:1384–94. [PubMed:
12638895]
Firbank M, Hiraoka M, Essenpreis M, Delpy DT. Measurement of the optical-properties of the skull in
the wavelength range 650–950 nm. Phys Med Biol 1993;38:503–10. [PubMed: 8488176]
Gotoh H, Yagi H. Solid angle subtended by a rectangular slit. Nucl Inst Methods 1971;96:485–6.
Hebert T, Leahy R. A generalized Em algorithm for 3-D Bayesian reconstruction from poisson data using
Gibbs priors. IEEE Trans Med Imaging 1989;8:194–202.
Holboke MJ, Tromberg BJ, Li X, Shah N, Fishkin J, Kidney D, Butler J, Chance B, Yodh AG. Three-
dimensional diffuse optical mammography with ultrasound localization in a human subject. J Biomed
Opt 2000;5:237–47. [PubMed: 10938789]
Jacques SL, Prahl SA. Modeling optical and thermal distributions in tissue during laser irradiation. Laser
Surg Med 1987;6:494–503.
Karagiannes JL, Zhang Z, Grossweiner B, Grossweiner LI. Applications of the 1-D diffusion-
approximation to the optics of tissues and tissue phantoms. Appl Opt 1989;28:2311–7.
Kienle A, Lilge L, Patterson MS, Hibst R, Steiner R, Wilson BC. Spatially resolved absolute diffuse
reflectance measurements for noninvasive determination of the optical scattering and absorption
coefficients of biological tissue. Appl Opt 1996;35:2304–14.
Kuo C, Coquoz O, Stearns D G and Rice B 2004 Diffuse luminescence tomography of in vivo
bioluminescent markers using multi-spectral data Society for Molecular Imaging 3rd Annual Meeting
(St. Louis) vol 3 (Cambridge: MIT Press) p 227
Lualdi M, Colombo A, Farina B, Tomatis S, Marchesini R. A phantom with tissue-like optical properties
in the visible and near infrared for use in photomedicine. Laser Surg Med 2001;28:237–43.
Marchesini R, Bertoni A, Andreola S, Melloni E, Sichirollo AE. Extinction and absorption-coefficients
and scattering phase functions of human-tissues in vitro. Appl Opt 1989;28:2318–24.
Martelli F, Sassaroli A, Zaccanti G, Yamada Y. Properties of the light emerging from a diffusive medium:
angular dependence and flux at the external boundary. Phys Med Biol 1999;44:1257–75. [PubMed:
10368017]
Matern U, Haberstroh J, el Saman A, Pauly E, Salm R, Farthmann EH. Emergency laparoscopy: technical
support for the laparoscopic diagnosis of intestinal ischemia. Surg Endosc 1996;10:883–7. [PubMed:
8703143]
Mitic G, Kolzer J, Otto J, Plies E, Solkner G, Zinth W. Time-gated transillumination of biological tissues
and tissuelike phantoms. Appl Opt 1994;33:6699–710.
Parsa P, Jacques SL, Nishioka NS. Optical-properties of rat-liver between 350 and 2200 nm. Appl Opt
1989;28:2325–30.
Pifferi A, Torricelli A, Taroni P, Bassi A, Chikoidze E, Giambattistelli E, Cubeddu R. Optical biopsy of
bone tissue: a step toward the diagnosis of bone pathologies. J Biomed Opt 2004;9:474–80. [PubMed:
15189084]
Prahl S A 2001 http://omlc.ogi.edu/spectra/index.html Oregon Medical Laser Clinic
Prout DL, Silverman RW, Chatziioannou A. Detector concept for OPET—a combined PET and optical
Imaging system. IEEE Trans Nucl Sci 2004;51:752–6. [PubMed: 16429601]
Rannou FR, Kohli V, Prout DL, Chatziioannou AF. Investigation of OPET performance using GATE, a
Geant4-based simulation software. IEEE Trans Nucl Sci 2004;51:2713–7. [PubMed: 16429604]
Alexandrakis et al. Page 10
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Ray P, De A, Min JJ, Tsien RY, Gambhir SS. Imaging tri-fusion multimodality reporter gene expression
in living subjects. Cancer Res 2004;64:1323–30. [PubMed: 14973078]
Ray P, Wu AM, Gambhir SS. Optical bioluminescence and positron emission tomography imaging of a
novel fusion reporter gene in tumor xenografts of living mice. Cancer Res 2003;63:1160–5. [PubMed:
12649169]
Ritz JP, Roggan A, Germer CT, Isbert C, Muller G, Buhr HJ. Continuous changes in the optical properties
of liver tissue during laser-induced interstitial thermotherapy. Laser Surg Med 2001;28:307–12.
Schweiger M, Arridge SR, Hiraoka M, Delpy DT. The finite-element method for the propagation of light
in scattering media—boundary and source conditions. Med Phys 1995;22:1779–92. [PubMed:
8587533]
Segars WP, Tsui BM, Frey EC, Johnson GA, Berr SS. Development of a 4-D digital mouse phantom for
molecular imaging research. Mol Imaging Biol 2004;6:149–59. [PubMed: 15193249]
Shepp LA, Vardi Y. Maximum likelihood reconstruction for emission tomography. IEEE Trans Med
Imaging 1982;1:113–22.
Solonenko M, Cheung R, Busch TM, Kachur A, Griffin GM, Vulcan T, Zhu TC, Wang HW, Hahn SM,
Yodh AG. In vivo reflectance measurement of optical properties, blood oxygenation and motexafin
lutetium uptake in canine large bowels, kidneys and prostates. Phys Med Biol 2002;47:857–73.
[PubMed: 11936174]
Srinivasan S, Pogue BW, Jiang SD, Dehghani H, Kogel C, Soho S, Gibson JJ, Tosteson TD, Poplack SP,
Paulsen KD. Interpreting hemoglobin and water concentration, oxygen saturation, and scattering
measured in vivo by near-infrared breast tomography. Proc Natl Acad Sci USA 2003;100:12349–54.
[PubMed: 14514888]
Swartling J, Palsson S, Platonov P, Olsson SB, Andersson-Engels S. Changes in tissue optical properties
due to radio-frequency ablation of myocardium. Med Biol Eng Comput 2003;41:403–9. [PubMed:
12892362]
Tai C, Chatziioannou A, Siegel S, Young J, Newport D, Goble RN, Nutt RE, Cherry SR. Performance
evaluation of the microPET P4: a PET system dedicated to animal imaging. Phys Med Biol
2001;46:1845–62. [PubMed: 11474929]
Thueler P, Charvet I, Bevilacqua F, St Ghislain M, Ory G, Marquet P, Meda P, Vermeulen B, Depeursinge
C. In vivo endoscopic tissue diagnostics based on spectroscopic absorption, scattering, and phase
function properties. J Biomed Opt 2003;8:495–503. [PubMed: 12880356]
Torricelli A, Pifferi A, Taroni P, Giambattistelli E, Cubeddu R. In vivo optical characterization of human
tissues from 610 to 1010 nm by time-resolved reflectance spectroscopy. Phys Med Biol
2001;46:2227–37. [PubMed: 11512621]
Torricelli A, Quaresima V, Pifferi A, Biscotti G, Spinelli L, Taroni P, Ferrari M, Cubeddu R. Mapping
of calf muscle oxygenation and haemoglobin content during dynamic plantar flexion exercise by
multi-channel time-resolved near-infrared spectroscopy. Phys Med Biol 2004;49:685–99. [PubMed:
15070196]
Troy T, Jekic-McMullen D, Sambucetti L, Rice B. Quantitative comparison of the sensitivity of detection
of fluorescent and bioluminescent reporters in animal models. Mol Imaging 2004;3:9–23. [PubMed:
15142408]
Ugryumova N, Matcher SJ, Attenburrow DP. Measurement of bone mineral density via light scattering.
Phys Med Biol 2004;49:469–83. [PubMed: 15012014]
Wang G, Li Y, Jiang M. Uniqueness theorems in bioluminescence tomography. Med Phys 2004;31:2289–
99. [PubMed: 15377096]
Wei HJ, Xing D, Wu GY, Jin Y, Gu HM. Optical properties of human normal small intestine tissue
determined by Kubelka–Munk method in vitro. World J Gastroentero 2003;9:2068–72.
Wielopolski L. Monte Carlo calculation of the average solid angle subtended by a parallellepiped detector
from a distributed source. Nucl Instum Methods Phys Res 1984;226:436–48.
Zijp JR, ten Bosch JJ. Optical properties of bovine muscle tissue in vitro: a comparison of methods. Phys
Med Biol 1998;43:3065–81. [PubMed: 9814535]
Alexandrakis et al. Page 11
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1.
A schematic of the proposed OPET system. Its gantry size is only slightly larger than the mouse
torso diameter.
Alexandrakis et al. Page 12
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2.
(a) Sagittal view of the MOBY mouse. (b) Corresponding sagittal view for the modified mouse
torso phantom. Different colours correspond to different tissue types (dark blue: adipose tissue,
light blue: liver, turquoise: lungs, yellow: bone, red: whole blood, orange: heart wall, light
green: gut, and purple: skin).
Alexandrakis et al. Page 13
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 3.
Comparison between MC and TOAST boundary flux predictions for a 3 mm deep point source
in the mouse gut emitting at 650 nm. Results are shown for a single ring of HiResOPET
detectors centred on the axial plane containing the point source.
Alexandrakis et al. Page 14
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 4.
(a) Transverse view of the mouse torso gut area with a point source located at its centre (star).
(b) Noise-free and spectrally blind OPET detector measurements for the point source described
in (a), emitting at wavelengths and spectral intensities defined by the luciferase spectrum. (c)
As in (b) but with the point source emitting at an intensity that is three times higher than the
standard deviation of background noise as measured at the mouse torso surface. The six OPET
detector blocks can be discerned.
Alexandrakis et al. Page 15
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 5.
(a) Reconstructed source distribution based on noiseless boundary flux data generated by the
point source in figure 4(a) emitting at 650 nm. (b) As in (a) but for spectrally distinct detection
of photons emitted simultaneously at 625 nm and 650 nm. (c) Reconstructed source distribution
for spectrally distinct detection of noise-added boundary flux data generated by the point source
in figure 4(a) emitting at 625 nm and 650 nm. (d) As in (c) but with the source emitting at 600,
625 and 650 nm. (e) As in (c) but with the source emitting at 600, 625, 650, 675 and 700 nm.
Alexandrakis et al. Page 16
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 6.
(a) Transverse view of the homogeneous mouse torso with three individually simulated point
sources sequentially placed at its centre, at half-radius and at 1 mm from the surface. (b)
Reconstructed source distribution based on noiseless boundary flux data generated by the point
source at the torso centre emitting at all five wavelengths. (c) As in (a) but for a source located
at half-radius; the red cross indicates the correct position of the point source. (d) As in (a) but
for a source located at 1 mm from the surface.
Alexandrakis et al. Page 17
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 7.
Line profiles through the two source peak intensities in reconstructed images of source pairs
in different anatomical locations. (a) Equal intensity sources separated by 4 mm in the gut
(solid curve) and the liver (short dashed curve). (b) As in (a) but with the two sources located
symmetrically around the torso half-radius. The two sources could be resolved in the liver for
a relative intensity ratio of 2:1 (grey long dashed curve). (c) As in (a) but with the two sources
located at 1 mm and 5 mm from the torso surface. (d) Two equal intensity point sources, 6.3
mm apart, when the true tissue optical properties (solid curve) and a uniform optical property
background (short dashed curve) were utilized in the image reconstructions.
Alexandrakis et al. Page 18
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 8.
(a) Coronal view of the mouse torso phantom with two equal intensity point sources (red stars)
located on either side of the torso centre in the gut. The artefactual source (orange oval) was
located at the liver centre and near the OPET FOV (red dashed lines) edge. (b) Reconstructed
image of the two point sources based on the OPET detector measurements at SNR = 5. (c) As
in (b) but for noiseless detector data. (d) Reconstructed image based on the virtual HiResOPET
detector measurements at SNR = 5.
Alexandrakis et al. Page 19
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 9.
(a) Transverse view of the mouse torso for an axial slice through the kidneys (encircled green
voxels). (b) Transverse view of the kidneys reconstructed as a uniform distributed source. The
red crosses indicate the peak intensity value in each kidney. (c) As in (b) but assuming a uniform
optical property background. The red crosses indicate a shift in the reconstructed kidney source
locations. (d) Transverse view of the mouse torso for an axial slice through the gut (yellow–
green voxels). (e) Transverse view of the gut reconstructed as a uniform distributed source.
(f ) As in (e) but assuming a uniform optical property background.
Alexandrakis et al. Page 20
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Alexandrakis et al. Page 21
Table 1
Values for the parameters a and b used to estimate μ′
s
(λ) for each tissue by use of equation (1). Definitions of
these parameters are provided in the text.
Tissue type a (mm
1
) b (no units) References
Adipose 38.0 0.53 (Mitic et al 1994, Kienle et al 1996, Holboke et al 2000, Srinivasan et al
2003)
Bone 35600 1.47 (Firbank et al 1993, Beek et al 1997a, 1997b, Pifferi et al 2004, Ugryumova
et al 2004)
Bowel 3670 1.24 (Matern et al 1996, Torricelli et al 2001, Solonenko et al 2002, Wei et al
2003)
Heart wall 10600 1.43 (Swartling et al 2003)
Kidneys 41700 1.51 (Solonenko et al 2002)
Liver and spleen 629 1.05 (Karagiannes et al 1989, Marchesini et al 1989, Parsa et al 1989, Kienle et
al 1996, Beek et al 1997a, 1997b, Ritz et al 2001, Srinivasan et al 2003)
Lung 68.4 0.53 (Beek et al 1997a, Beek et al 1997b, Srinivasan et al 2003)
Muscle 4.e7 2.82 (Cheong et al 1990, Kienle et al 1996, Beek et al 1997a, 1997b, Zijp and ten
Bosch 1998, Torricelli et al 2004)
Skin 2850 1.1 (Jacques and Prahl 1987, Marchesini et al 1989, Beek et al 1997a, 1997b,
Doornbos et al 1999, Lualdi et al 2001)
Stomach wall 792 0.97 (Thueler et al 2003)
Whole blood 133 0.66 (Enejder et al 2003)
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Alexandrakis et al. Page 22
Table 2
Values for the parameters S, x, and y used to estimate μ
a
(λ) for each tissue by use of equation (2). Definitions of
these parameters are provided in the text. The literature references employed for these estimates are identical to
those in table 1.
Tissue type S
B
x S
W
Adipose 0.0033 0.7 0.5
Bone 0.049 0.8 0.15
Bowel 0.0093 0.8 0.5
Heart wall 0.05 0.75 0.5
Kidneys 0.056 0.75 0.8
Liver and spleen 0.30 0.75 0.7
Lung 0.15 0.85 0.85
Muscle 0.07 0.8 0.5
Skin 0.06 0.75 0.5
Stomach wall 0.01 0.7 0.8
Whole blood 1.0 0.75 0.0
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Alexandrakis et al. Page 23
Table 3
The uniform mouse torso optical property values at the five wavelengths used in this work.
Emission wavelength μ
a
(λ) (mm
1
) μ′
s
(λ) (mm
1
)
600 0.19 1.66
625 0.062 1.59
650 0.038 1.53
675 0.028 1.47
700 0.022 1.41
Phys Med Biol. Author manuscript; available in PMC 2005 December 20.
... For biological applications with visible and infrared light, the spectral optical parameters for target tissues can be approximated as a weighted sum of the three constituent optical parameters for oxyhemoglobin, deoxyhemoglobin, and water. [26][27][28] In addition, the choice between reflection and transparent configurations should also be determined by the form and environment of the objects of interest, as well as whether mild depolarization is acceptable, such as the X ¼ 12 and X ¼ 22 in reflection and X ≅ 5 in transparent geometry. ...
Article
Full-text available
Significance: The depolarization of circularly polarized light (CPL) caused by scattering in turbid media reveals structural information about the dispersed particles, such as their size, density, and distribution, which is useful for investigating the state of biological tissue. However, the correlation between depolarization strength and tissue parameters is unclear. Aim: We aimed to examine the generalized correlations of depolarization strength with the particle size and wavelength, yielding depolarization diagrams. Approach: The correlation between depolarization intensity and size parameter was examined for single and multiple scattering using the Monte Carlo simulation method. Expanding the wavelength width allows us to obtain depolarization distribution diagrams as functions of wavelength and particle diameter for reflection and transparent geometries. Results: CPL suffers intensive depolarization in a single scattering against particles of various specific sizes for its wavelength, which becomes more noticeable in the multiple scattering regime. Conclusions: The depolarization diagrams with particle size and wavelength as independent variables were obtained, which are particularly helpful for investigating the feasibility of various particle-monitoring methods. Based on the obtained diagrams , several applications have been proposed, including blood cell monitoring, early embryogenesis, and antigen-antibody interactions.
... Therefore, a light with a wavelength of 1240 nm is unsuitable because of its strong optical absorption by water. For biological applications with visible and infrared light, the spectral optical parameters for target tissues can be approximated as a weighted sum of the three constituent optical parameters for oxy-hemoglobin, deoxy-hemoglobin, and water [26][27][28]. ...
Preprint
Full-text available
Significance: The depolarization of circularly polarized light caused by scattering in turbid media reveals structural information about the dispersed particles, such as their size, density, and distribution, which is useful for investigating the state of biological tissue. However, the correlation between depolarization strength and tissue parameters is unclear. Aim: We aimed to examine the generalized correlations of depolarization strength with the particle size and wavelength, yielding depolarization diagrams. Approach: The correlation between depolarization intensity and size parameter were examined for single and multiple scattering using the Monte Carlo simulation method. Expanding the wavelength width allows us to obtain depolarization distribution diagrams as functions of wavelength and particle diameter for reflection and transparent geometries. Results: Circularly polarized light suffers intensive depolarization in a single scattering against particles of various specific sizes for its wavelength, which becomes more noticeable in the multiple scattering regime. Conclusions: The depolarization diagrams with particle size and wavelength as independent variables were obtained, which are particularly helpful for investigating the feasibility of various particle-monitoring methods. Based on the obtained diagrams, several applications have been proposed, including blood cell monitoring, early embryogenesis, and antigen-antibody interactions.
... To assess the impact of depth of inflammation on contrast using the proposed normalised FFT feature, a simple 20 × 30 mm two-dimensional (2-D) model representing the anatomy of the PIP joint was constructed, as shown in Fig. 4, with the skin and bone thickness assigned as 1 and 10 mm, respectively, dimensions approximately estimated from an available MRI image of a PIP joint. Physiological values assigned to each of these regions are shown in Table 1, with total haemoglobin (tHb) and water (H 2 O) chosen according to previous literature, 35 all tissue oxygenation (StO 2 ) values assumed to be 70% and values for the synovia region based on previous joint imaging studies, 36 while scattering amplitude and scattering power were both assumed to be homogeneous with values of 0.77 mm −1 and 2.145, respectively. A single source and 27 virtual detectors separated by 1 mm along the opposing boundary were assigned to this model, to reflect the point source illumination and a pixel array from camera-based detection used in this study. ...
... Additionally, two cylindrical tumors, measuring 3 mm in height and 4 mm in diameter, were inserted into the liver of the mouse (Figure 2a). The scattering and absorption coefficients were allocated to the organs based on the study conducted by [34]. ...
Article
Full-text available
Single-view cone-beam X-ray luminescence computed tomography (CB-XLCT) has recently gained attention as a highly promising imaging technique that allows for the efficient and rapid three-dimensional visualization of nanophosphor (NP) distributions in small animals. However, the reconstruction performance is hindered by the ill-posed nature of the inverse problem and the effects of depth variation as only a single view is acquired. To tackle this issue, we present a methodology that integrates an automated restarting strategy with depth compensation to achieve reconstruction. The present study employs a fast proximal gradient descent (FPGD) method, incorporating L0 norm regularization, to achieve efficient reconstruction with accelerated convergence. The proposed approach offers the benefit of retrieving neighboring multitarget distributions without the need for CT priors. Additionally, the automated restarting strategy ensures reliable reconstructions without the need for manual intervention. Numerical simulations and physical phantom experiments were conducted using a custom CB-XLCT system to demonstrate the accuracy of the proposed method in resolving adjacent NPs. The results showed that this method had the lowest relative error compared to other few-view techniques. This study signifies a significant progression in the development of practical single-view CB-XLCT for high-resolution 3−D biomedical imaging.
Article
Significance The depolarization of circularly polarized light (CPL) caused by scattering in turbid media reveals structural information about the dispersed particles, such as their size, density, and distribution, which is useful for investigating the state of biological tissue. However, the correlation between depolarization strength and tissue parameters is unclear. Aim We aimed to examine the generalized correlations of depolarization strength with the particle size and wavelength, yielding depolarization diagrams. Approach The correlation between depolarization intensity and size parameter was examined for single and multiple scattering using the Monte Carlo simulation method. Expanding the wavelength width allows us to obtain depolarization distribution diagrams as functions of wavelength and particle diameter for reflection and transparent geometries. Results CPL suffers intensive depolarization in a single scattering against particles of various specific sizes for its wavelength, which becomes more noticeable in the multiple scattering regime. Conclusions The depolarization diagrams with particle size and wavelength as independent variables were obtained, which are particularly helpful for investigating the feasibility of various particle-monitoring methods. Based on the obtained diagrams, several applications have been proposed, including blood cell monitoring, early embryogenesis, and antigen–antibody interactions.
Preprint
Full-text available
Significance: Pulsatile blood Oxygen Saturation (SpO2 ) via pulse oximetry is a valuable clinical metric for assessing oxygen delivery. Individual anatomical features, including skin tone, may affect current optical pulse oximetry methods. Aim: Develop an optical pulse oximetry method based on Dual-Ratio (DR) measurements to suppress individual anatomical features on SpO2. Approach: Design a DR-based finger pulse oximeter, hypothesizing that DR would suppress confounds from optical coupling and superficial tissue-absorption. This method is tested using Monte Carlo (MC) simulations and in vivo experiments. Results: Different melanosome volume fraction in the epidermis, a surrogate for skin tone, cause changes in the recovered SpO2 on the order of 1%. Different heterogeneous pulsatile hemodynamics cause greater changes on the order of 10%. SpO2 recovered with DR measurements showed less variability than the traditional Single-Distance (SD) transmission method. Conclusions: For the models and methods considered here, SpO2 measurements are more strongly impacted by heterogeneous pulsatile hemodynamics than by melanosome volume fraction. This is consistent with previous reports that, the skin tone bias is smaller than the observed variation in recovered SpO 2 across individual people. The partial suppression of variability in the SpO2 recovered by DR suggests promise of DR for pulse oximetry.
Article
Background and Objectives A threshold fluence for melanosome disruption has the potential to provide a robust numerical indicator for establishing clinical endpoints for pigmented lesion treatment using a picosecond laser. Although the thresholds for a 755‐nm picosecond laser were previously reported, the wavelength dependence has not been investigated. In this study, wavelength‐dependent threshold fluences for melanosome disruption were determined. Using a mathematical model based on the thresholds, irradiation parameters for 532‐, 730‐, 755‐, 785‐, and 1064‐nm picosecond laser treatments were evaluated quantitatively. Study Design/Materials and Methods A suspension of melanosomes extracted from porcine eyes was irradiated using picosecond lasers with varying fluence. The mean particle size of the irradiated melanosomes was measured by dynamic light scattering, and their disruption was observed by scanning electron microscopy to determine the disruption thresholds. A mathematical model was developed, combined with the threshold obtained and Monte Carlo light transport to calculate irradiation parameters required to disrupt melanosomes within the skin tissue. Results The threshold fluences were determined to be 0.95, 2.25, 2.75, and 6.50 J/cm² for 532‐, 730‐, 785‐, and 1064‐nm picosecond lasers, respectively. The numerical results quantitatively revealed the relationship between irradiation wavelength, incident fluence, and spot size required to disrupt melanosomes distributed at different depths in the skin tissue. The calculated irradiation parameters were consistent with clinical parameters that showed high efficacy with a low incidence of complications. Conclusion The wavelength‐dependent thresholds for melanosome disruption were determined. The results of the evaluation of irradiation parameters from the threshold‐based analysis provided numerical indicators for setting the clinical endpoints for 532‐, 730‐, 755‐, 785‐, and 1064‐nm picosecond lasers.
Article
Objective Regulation of blood flow to bone is critical but poorly understood, particularly in humans. This study aims to determine whether nitric oxide (NO), a major regulator of vascular tone to other tissues, contributes also to the regulation of blood flow to bone. Methods In young healthy adults ( n = 16, 8F, 8M), we characterized NO‐mediated vasodilation in the tibia in response to sublingual nitroglycerin and contrasted it to lower leg. Blood flow responses were assessed in supine individuals by continuously measuring tibial total hemoglobin (tHb) via near‐infrared spectroscopy and lower leg blood flow (LBF) as popliteal flow velocity via Doppler ultrasound in the same leg. Results LBF increased by Δ9.73 ± 0.66 cm/s and peaked 4.4 min after NO administration and declined slowly but remained elevated (Δ3.63 ± 0.60 cm/s) at 10 min. In contrast, time to peak response was longer and smaller in magnitude in the tibia as tHb increased Δ2.08 ± 0.22 μM and peaked 5.3 min after NO administration and declined quickly but remained elevated (Δ0.87±0.22 μM) at 10 min ( p = .01). Conclusions In young adults, the tibial vasculature demonstrates robust NO‐mediated vasodilation, but tHb is delayed and diminishes faster compared to LBF, predominately reflective of skeletal muscle responses. Thus, NO‐mediated vasodilation in bone may be characteristically different from other vascular beds.
Article
Full-text available
The absorption and transport scattering coefficients of biological tissues determine the radial dependence of the diffuse reflectance that is due to a point source. A system is described for making remote measurements of spatially resolved absolute diffuse reflectance and hence noninvasive, noncontact estimates of the tissue optical properties. The system incorporated a laser source and a CCD camera. Deflection of the incident beam into the camera allowed characterization of the source for absolute reflectance measurements. It is shown that an often used solution of the diffusion equation cannot be applied for these measurements. Instead, a neural network, trained on the results of Monte Carlo simulations, was used to estimate the absorption and scattering coefficients from the reflectance data. Tests on tissue-simulating phantoms with transport scattering coefficients between 0.5 and 2.0 mm(-1) and absorption coefficients between 0.002 and 0.1 mm(-1) showed the rms errors of this technique to be 2.6% for the transport scattering coefficient and 14% for the absorption coefficients. The optical properties of bovine muscle, adipose, and liver tissue, as well as chicken muscle (breast), were also measured ex vivo at 633 and 751 nm. For muscle tissue it was found that the Monte Carlo simulation did not agree with experimental measurements of reflectance at distances less than 2 mm from the incident beam.
Article
Full-text available
The applicability and limits of time-resolved transillumination to determine the internal details of biological tissues are investigated by phantom experiments. By means of line scans across a sharp edge, the spatial resolution (Δx) and its dependence on the time-gate width (Δt) can be determined. Additionally, measurements of completely absorbing bead pairs embedded in a turbid medium demonstrate the physical resolution in a more realistic case. The benefit of time resolution is especially high for a turbid medium with a comparatively small reduced scattering coefficient of approximately µ(s)' = 0.12 mm(-1). Investigations with partially absorbing beads and filled plastic tubes demonstrate the high sensitivity of time-resolving techniques with respect to spatial variations in scattering or absorption coefficients that are due to the embedded disturber. In particular, it is shown that time gating is sensitive to variations in scattering coefficients.
Article
Full-text available
We describe a novel Monte Carlo code for photon migration through 3D media with spatially varying optical properties. The code is validated against analytic solutions of the photon diffusion equation for semi-infinite homogeneous media. The code is also cross-validated for photon migration through a slab with an absorbing heterogeneity. A demonstration of the utility of the code is provided by showing time-resolved photon migration through a human head. This code, known as 'tMCimg', is available on the web and can serve as a resource for solving the forward problem for complex 3D structural data obtained by MRI or CT.
Article
AIM: To study the optical properties of human normal small intestine tissue at 476.5 nm, 488 nm, 496.5 nm, 514.5 nm, 532 nm, 808 nm wavelengths of laser irradiation. METHODS: A double-integrating-sphere system, the basic principle of measuring technology of light radiation, and an optical model of biological tissues were used in the study. RESULTS: The results of measurement showed that there were no significant differences in the absorption coefficients of human normal small intestine tissue at 476.5 nm, 488 nm, 496.5 nm laser in the Kubelka-Munk two-flux model (P > 0.05). The absorption coefficients of the tissue at 514.5 nm, 532 nm, 808 nm laser irradiation were obviously increased with the decrease of these wavelengths. The scattering coefficients of the tissue at 476.5 nm, 488 nm, 496.5 nm laser irradiation were increased with the decrease of these wavelengths. The scattering coefficients at 496.5 nm, 514.5 nm, 532 nm laser irradiation were obviously increased with the increase of these wavelengths. The scattering coefficient of the tissue at 532 nm laser irradiation was bigger than that at 808 nm. There were no significant differences in the total attenuation coefficient of the tissue at 476.5 nm and 488 nm laser irradiation (P > 0.05). The total attenuation coefficient of the tissue at 488 nm, 496.5 nm, 514.5 nm, 532 nm, 808 nm laser irradiation was obviously increased with the decrease of these wavelengths, and their effective attenuation coefficient revealed the same trend. There were no significant differences among the forward scattered photon fluxe, backward scattered photon fluxe, and total scattered photon fluxe of the tissue at 476.5 nm, 488 nm, 496.5 nm laser irradiation. They were all obviously increased with attenuation of tissue thickness. The attenuations of forward and backward scattered photon fluxes, and the total scattered photon fluxe of the tissue at 514.5 nm laser irradiation were slower than those at 476.5 nm, 488 nm, 496.5 nm laser irradiation respectively. The attenuations of forward and backward scattered photon fluxes, and total scattered photon fluxes at 532 nm laser irradiation were obviously slower than those at 476.5 nm, 488 nm, 496.5 nm, 514.5 nm laser irradiation. The attenuations of forward and backward scattered photon fluxes, and total scattered photon fluxe at 808 nm laser irradiation were all obviously slower than those at 476.5 nm, 488 nm, 496.5 nm, 514.5 nm, 532 nm laser irradiation respectively. CONCLUSION: There are significant differences in optical parameters of human normal small intestine tissue in the Kubelka-Munk two-flux model at six different wavelengths of laser radiation. The results would provide a new method of information analysis for clinical diagnosis. Keywords: $[Keywords] Citation: Wei HJ, Xing D, Wu GY, Jin Y, Gu HM. Optical properties of human normal small intestine tissue determined by Kubelka-Munk method in vitro . World J Gastroenterol 2003; 9(9): 2068-2072
Article
A solid angle subtended by a rectangular slit with an arbitrary thickness at an arbitrary point is represented by inverse trigonometric functions.
Article
A Monte Carlo method utilizing total variance reduction is applied to the calculation of the solid angle subtended by a parallelepiped detector from an arbitrarily positioned point. Although there is an analytical expression for the solid angle subtended by a rectangle from an arbitrary point, the method should prove valuable when a mean solid angle from a distributed source is required. In addition, the algorithm provides an efficient way to sample the directions of gamma rays which intercept a parallelepiped detector, required in the Monte Carlo calculations of the efficiency of square detectors.
Article
The propagation of light energy in tissues is an important problem in phototherapy, especially with the increased use of lasers as light sources. Often a slight difference in delivered energy separates a useless, efficacious, or disastrous treatment. Methods are presented for experimental characterization of the optical properties of a tissue and computational prediction of the distribution of light energy within a tissue. A standard integrating sphere spectrophototneter measured the total transmission, Tt, total reflectance, Rt, and the on-axis transmission, Ta, for incident collimated light that propagated through the dermis of albino mouse skin, over the visible spectrum. The diffusion approximation solution to the one-dimensional (1-D) optical transport equation computed the expected Tt, and Rt, for different combinations of absorbance, k, scattering, s, and anisotropy, g, and by iterative comparison of the measured and computed Tt, and Rt, values converged to the intrinsic tissue parameters. For example, mouse dermis presented optical parameters of 2.8 cm−1, 239 cm−1, and 0.74 for k,s, and g, respectively, at 488 nm wavelength. These values were used in the model to simulate the optical propagation of the 488-nm line of an argon laser through mouse skin in vivo. A 1-D Green's function thermal diffusion model computed the temperature distribution within the tissue at different times during laser irradiation. In vitro experiments showed that the threshold temperature range for coagulation was 60°–70° C, and the kinetics were first order, with a temperature-dependent rate constant that obeyed an Arrhenius relation (molar entropy 276 cal/mol-°K, molar enthalpy 102 kcal/mol). The model simulation agreed with the corresponding in vivo experiment that a 2-s pulse at 55 W/cm2 irradiance will achieve coagulation of the skin.
Article
Previous investigators have determined the optical properties of the liver at isolated wavelengths using various models of light propagation. In this study we derived coefficients for absorption, scattering, and anisotropy for rat liver between 350 and 2200 nm using measurements of integrated reflectance, transmittance, and on-axis attenuation and a diffusion approximation algorithm to solve the 1-D transport equation. These results were used to estimate the effective attenuation coefficient, penetration depth, and temperature rise induced by laser irradiation. The data should be of value to investigators in the field of tissue optics and developers of photomedical applications involving the liver.
Article
Optical properties of different human tissues in vitro have been evaluated by measuring extinction and absorption coefficients at 635- and 515-nm wavelengths and a scattering angular dependence at 635 nm. Extinction was determined by the on-axis attenuation of light transmitted through sliced specimens of various thicknesses. The absorption coefficient was determined by placing samples into an integrating sphere. The Henyey-Greenstein function was used for fitting experimental data of the scattering pattern. The purpose of this work was to contribute to the study of light propagation in mammalian tissues. The results show that, for the investigated tissues, extinction coefficients range from ~200 to 500 cm(-1) whereas absorption coefficients, depending on wavelength, vary from 0.2 to 25 cm(-1). Scattering is forward peaked with an average cosine of ~0.7.
Article
Optical constants of animal and plant tissues were measured over a wide spectral range using an integrating sphere spectrophotometer. The data were analyzed with two formulations of the 1-D diffusion approximation differing in the phase function. The accuracy of the similarity transformation was examined with tissue phantoms incorporating known chromophores in light scattering media.