ArticlePDF Available

Effects of sampling strategy on image quality in noncontact panoramic fluorescence diffuse optical tomography for small animal imaging

Optica Publishing Group
Optics Express
Authors:

Abstract and Figures

Fluorescence diffuse optical tomography is an emerging technology for molecular imaging with recent technological advances in biomarkers and photonics. The introduction of noncontact imaging methods enables very large-scale data acquisition that is orders of magnitude larger than that from earlier systems. In this study, the effects of sampling strategy on image quality were investigated using an imaging phantom mimicking small animals and further analyzed using singular value analysis (SVA). The sampling strategy was represented in terms of a number of key acquisition parameters, namely the numbers of sources, detectors, and imaging angles. A number of metrics were defined to quantitatively evaluate image quality. The effects of acquisition parameters on image quality were subsequently studied by varying each of the parameters within a reasonable range while maintaining the other parameters constant, a method analogue to partial derivative in mathematical analysis. It was found that image quality improves at a much slower rate if the acquisition parameters are above certain critical values (approximately 5 sources, approximately 15 detectors, and approximately 20 angles for our system). These critical values remain virtually the same even if other acquisition parameters are doubled. It was also found that increasing different acquisition parameters improves image quality with different efficiencies in terms of the number of measurements: for a system characterized by a smaller threshold in SVA (less than 10(-5) in our study), the number of sources is the most efficient, followed by the number of detectors and subsequently the number of imaging angles. However, for systems characterized by a larger threshold, the numbers of sources and angles are equally more efficient than the number of detectors.
Content may be subject to copyright.
Effects of sampling strategy on image quality in noncontact
panoramic fluorescence diffuse optical tomography for small
animal imaging
Xiaofeng Zhang* and Cristian Badea
Department of Radiology, Center for In Vivo Microscopy, Duke University Medical Center, Durham,
NC, 27710, USA
Abstract
Fluorescence diffuse optical tomography is an emerging technology for molecular imaging with
recent technological advances in biomarkers and photonics. The introduction of noncontact imaging
methods enables very large-scale data acquisition that is orders of magnitude larger than that from
earlier systems. In this study, the effects of sampling strategy on image quality were investigated
using an imaging phantom mimicking small animals and further analyzed using singular value
analysis. The sampling strategy was represented in terms of a number of key acquisition parameters,
namely the numbers of sources, detectors, and imaging angles. A number of metrics were defined to
quantitatively evaluate image quality. The effects of acquisition parameters on image quality were
subsequently studied by varying each of the parameters within a reasonable range while maintaining
the other parameters constant, a method analogue to partial derivative in mathematical analysis.
Image quality improves at a much slower rate if the acquisition parameters are above certain critical
values (~5 sources, ~15 detectors, and ~20 angles for our system). These critical values remain
virtually the same even if other acquisition parameters are doubled. It is also found that increasing
different acquisition parameters improves image quality with different efficiencies in terms of the
number of measurements: for a system characterized by a smaller threshold in singular value analysis
(less than 105 in our study), the number of sources is the most efficient, followed by the number of
detectors and subsequently the number of imaging angles; however, for systems characterized by a
larger threshold, the numbers of sources and angles are equally more efficient than the number of
detectors.
1. Introduction
Fluorescence diffuse optical tomography (FDOT) in small animals is currently an active field
of research with promises for applications to basic science and medicine. This is because this
technology offers a potentially revolutionary means to noninvasive in vivo tomographic
imaging with unmatched sensitivity and molecular specificity for medical and preclinical
imaging, particularly in cancer studies [1–3].
In parallel with engineering advances in fluorescent probes that selectively bind to various
molecular targets, as reviewed in [4,5], researchers are also developing better imaging
apparatus and methods for FDOT to provide tomographic mapping of biomarker-driven
©2008 Optical Society of America
*Corresponding author: E-mail: steve.zhang@duke.edu.
OCIS codes: (170.3880) Medical and biological imaging; (110.3000) Image quality assessment; (110.6880) Three-dimensional image
acquisition; (110.6960) Tomography
NIH Public Access
Author Manuscript
Opt Express. Author manuscript; available in PMC 2010 March 30.
Published in final edited form as:
Opt Express. 2009 March 30; 17(7): 5125–5138.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
distributions of these probes, e.g., [6–12]. Compared to other leading biomedical imaging
modalities, such as x-ray computed tomography (CT) and magnetic resonance imaging (MRI),
achieving high-quality FDOT images is a challenging task mostly due to the diffuse nature of
light propagation in tissue [13,14]. Nonetheless, image quality can be improved from multiple
aspects such as advanced imaging apparatus, accurate photon migration models, sound
sampling strategy, and sophisticated reconstruction algorithms.
Rapid advances in photonics have greatly improved the variety and performance of available
instruments for FDOT, particularly laser diodes and charge-coupled devices (CCD). In terms
of system configuration, many FDOT systems for small animals use optical fibers to couple
light to and/or from the skin of the animal, e.g. [15,16]. As a result, these system configurations
typically lead to an underdetermined image reconstruction problem and are difficult to extend
to large-scale data acquisition because of the physical constraints posed by optical fibers. These
limitations can be eliminated by using noncontact acquisition techniques: a scanning laser as
the light source and a CCD camera as the optical sensor [4,6,9,16,17]. A recent publication has
shown that noncontact techniques produce better imaging results compared to fiber-based
methods [18]. In many small animal FDOT systems, imaging chambers filled with optical
matching fluid are used. In fiber-based systems, the main purpose of the imaging chamber is
to achieve proper optical contact; and in noncontact systems, computation can be significantly
simplified by using slab geometry due to optical matching, e.g., [4,16]. However, the imaging
chamber makes it very difficult to achieve panoramic data acquisition for in vivo studies, which
is critical to depth-resolution in FDOT. Studies show that FDOT could be performed without
the imaging chamber, e.g., [6,19,20]. With a fully noncontact and fluid-free method, a rotation
stage can be integrated to enable panoramic data acquisition on a vertically positioned animal,
as reported recently by Deliolanis et al. in [19].
The introduction of a fully noncontact imaging system is clearly an exciting advancement in
FDOT hardware development because it allows dramatically increased sampling density as
well as unprecedented flexibility in data sampling strategy compared to fiber-based systems.
With the typical number of pixels available on a modern CCD camera (~106), a reasonable
laser scanning density (~102–104 points within an area of ~25×25 mm), and the number of
angles a rotation stage can provide (~102), a noncontact FDOT system can potentially provide
1010–1012 measurements. Using this approach, the reconstruction problem in FDOT can easily
become overdetermined, given that the usual upper limit of field-of-view in small animal
imaging (30×30×60 mm) contains only 5.4×104 of 1 mm3 voxels, a typical resolution limit
due to scattering in tissue. Furthermore, with a scanning laser and a CCD camera, the points
of illumination and sensing can be virtually anywhere on the skin of the animal. Such flexibility
in sampling inevitably raises the interesting question – what would be the optimal data sampling
strategy in order to achieve the best image quality with the least amount of sampling? Extensive
studies regarding different imaging strategies have been conducted for FDOT and diffuse
optical tomography (DOT), which share similar physical and mathematical models, as
discussed in [21–26]. Nonetheless, the majority of these studies were performed in the context
of imaging in humans or using fiber-based systems. Studies for noncontact FDOT in small
animal imaging are still scarce in the literature. One such study on the optimal sampling
geometry for the FDOT was recently reported by Lasser et al. [27], in which an analytical
model based on the diffusion approximation was used to compute the forward problem and
singular value analysis (SVA) based on a previously determined threshold was used to optimize
source and detector arrangements and the field-of-view (FOV). The sampling strategies
determined by SVA were experimentally verified using an imaging phantom.
In this study, we investigate the effects of different sampling strategies on image quality using
a different approach and extend the number of parameters under investigation. Specifically,
three acquisition parameters that are fundamental to experimental designs were chosen for this
Zhang and Badea Page 2
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
study: the number of source positions, the number of detector positions, and the number of
imaging angles. The number of source positions is linearly related to the acquisition time in
scanning- or switching-source systems. The number of detector positions becomes less a
hardware limit in a noncontact optical imaging system due to the relatively large number of
pixels available in a CCD camera. Nonetheless, it is still a major limiting factor for the rate of
data acquisition. The number of imaging angles is linearly related to data acquisition time as
well. All of these acquisition parameters contribute linearly to the matrix size in image
reconstruction. We first constructed a noncontact liquid-free panoramic FDOT system and
performed experiments on an imaging phantom. A number of metrics of image quality were
adopted to quantitatively evaluate the effects of these acquisition parameters on imaging
results. Based on our experimental configuration, SVA was used to evaluate the overall
performance of the imaging system in terms of the number of singular values (NSV) above a
series of different thresholds to interpret and generalize our experimental results.
2. Methods
2.1. Hardware configuration and experimental design
Our FDOT system adopts the same architecture as reported in [19]. The optical sensor is an
intensified CCD (ICCD) camera (PicoStar HRI, LaVision, Germany), which has a 2-D array
of 1370×1040 pixels with 12-bit digitization. The light source is a 785 nm continuous-wave
(CW) laser diode (HL7851G, Hitachi, Japan), which is thermally and electrically stabilized
(LDC205C, TED200C, and TCLDM9, Thorlabs, Germany) and coupled to the surface of a
lab-built imaging phantom via a galvo-scanner, i.e., a dual-axis galvanometer equipped with
a pair of high-reflectance mirrors (6230H, Cambridge Technology, MA). The imaging
phantom is vertically positioned on a motorized rotation stage (Oriel 13049, Newport, CA).
The system configuration is schematically shown in Fig. 1(a) and (b). A filter wheel was used
to switch between the excitation and emission band-pass filters (785 and 830 nm central
wavelengths, respectively, and 10 nm pass-band).
The imaging phantom used in this study was constructed as a cylinder (Nylon 6/6) with 23 mm
outer diameter (1 mm wall thickness) and filled with approximately 1% v/v fat emulsion
(diluted from 10% Intralipid, Fresenius Kabi, Sweden) and India ink (off the shelf) in deionized
water to simulate the optical properties of small animals. The phantom enclosed two thin-wall
glass tubes (3 and 4 mm outer diameter, 0.3 and 0.4 mm wall thickness, respectively, NMR
tubes by Wilmad LabGlass, NJ) both filled with 10 µM indocyanine green (ICG or
Cardiogreen, Sigma-Aldrich, MO) aqueous solution (deionized water), see Fig. 1(c). The
optical properties of the components of the imaging phantom were listed in Table 1. The
absorption and scattering coefficients were measured at 785 nm, and the indices of refraction
were taken from typical values in the literature.
The laser and the camera were arranged in a transillumination configuration—180° apart from
each other with respect to the imaging phantom. The galvo-scanner moves the laser beam
horizontally across the surface of the phantom in small steps at a constant angular interval and
thus creates a discrete 1-D pattern of light source positions perpendicular to the symmetric axis
of the cylinder, as illustrated in Fig. 1(b) and (c). For each source position, the ICCD camera
was triggered (20 ms exposure time) to take a fluorescence image (at the emission wavelength,
using the 830 nm filter), in which a series of points were chosen in 1-D (at the same height as
the source positions) as the detector positions. For each detector position, its signal intensity
was an average over a small cluster of pixels (5×5 in this study) around it to increase the signal-
to-noise ratio (SNR), which corresponds to an area of 0.3×0.3 mm on the projected 2-D surface
of the imaging phantom. After the galvo-scanner completes a scan and resumes its initial state,
the imaging phantom was turned a small angle by the rotation stage, followed by another laser
scan by the galvo-scanner. This process repeats until the phantom returns to its initial angular
Zhang and Badea Page 3
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
position. Afterwards, the above imaging procedure was repeated at the excitation wavelength
(785 nm) to obtain the distribution of excitation light in order to normalize the previously
acquired fluorescence signal [10]. In this study, a total of 21 source positions, 30 detector
positions, and 180 imaging angles were used during data acquisition, after which subsets of
this full dataset were selected for each of the study cases (described later) to ensure consistent
experimental conditions. The source positions spanned an arc of 100°; and the detector
positions covered 140°on the cylindrical surface of the phantom, both of which had
approximately the same angular sampling density of 5°/sample on the surface of the imaging
phantom. In each subset, the sources, detectors, and imaging angles were evenly distributed
(rounded to the nearest available sampling points) in the full dataset. The triggering of the
ICCD camera and the actuation of the rotation stage were controlled using our lab-developed
LabVIEW (v8.5, National Instruments, TX) applications and auxiliary computer hardware.
Due to the large number of possible combinations of the acquisition parameters, an exhaustive
case study is impractical. We adopted an analyzing method analogue to partial derivative in
mathematical analysis: only one parameter is variable at a time while maintaining the other
two constant. In this study, image quality is analyzed for the following 9 cases, as listed in
Table 2: Case (1), variable number of sources and fixed numbers of detectors and imaging
angles; Case (2), repeat Case (1) but double the number of detectors; Case (3), repeat Case (1)
but double the number of angles; Cases (4)–(6), variable number of detectors and double the
numbers of sources and angles, respectively; Cases (7)–(9), variable number of imaging angles
and double the numbers of sources and detectors, respectively. The values of the constant
parameters used in Cases (1), (4), and (7) give reasonably good image quality, which were
chosen empirically based on initial trail experiments.
2.2. Forward problem and image reconstruction
The forward problem of FDOT was based on the Born approximation of the diffusion equation,
similar to the normalized Born method proposed by Ntziachristos and Weissleder in [10].
Fluorescence was modeled as a linear equation AX = b, where the measured fluorescence b is
a vector linearly related to the 3-D fluorophore distribution X via a sensitivity matrix A. The
fluorescence (optical signal measured at the emission wavelength) was normalized by the
signal at the excitation wavelength under otherwise identical experimental conditions; the
normalized sensitivity matrix was computed using a Monte Carlo (MC) simulation-based
method developed by Zhang et al. [28].
In MC simulations, the imaging medium was modeled to precisely represent our imaging
phantom as a two-layered cylinder: a homogeneous cylindrical scattering material (a liquid
mixture of Intralipid and India ink) enclosed by a thin layer of homogeneous material of
different optical properties (Nylon 6/6). 90 millions of photons were used in the simulations.
The voxel size of the discretized imaging medium was (0.1 mm)3 in MC simulations and was
down-sampled to (1 mm)3 while computing the sensitivity matrix. The exact geometry and
optical properties of the imaging phantom were set to match our experimental conditions
previously described in detail in section Hardware configuration and experimental design.
The exact positions of the sources and the detectors were obtained by geometrical calculation.
The orientations of the light sources (the incident angle of the laser) and the optical detectors
(the principle escaping angle of the fluorescent light) in the simulations are parallel to the axis
on which the galvo-scanner, the imaging phantom, and the camera are positioned. This parallel
configuration in MC simulations effectively avoids the type of distortion of signal intensity
due to geometrical effects arising from the curved surface when the sources and detectors are
modeled as being perpendicular to the surface of the phantom (as is the case in typical fiber-
based systems), which has to be corrected, for example, using the methods described by Schulz
et al. in [29]. Second-order geometrical distortion due to the lens and the scanning laser are
Zhang and Badea Page 4
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
ignored due to the relatively large distance from the phantom to the lens (~240 mm) and to the
galvo-scanner (~200 mm) compared to the representative size of the imaging phantom (~20
mm).
Image reconstruction was performed on the normalized Born equation to resolve the
fluorophore distribution in 3-D, using a conjugate gradient-based iterative method. Starting
from an all-zero initial guess, the algorithm iteratively updates the solution according to the
conjugate gradient method [30] until the value of the cost function H is less than a
predetermined limit of 106. This cost function was defined as a weighted sum of three penalty
terms: square of L-2 norm of the residual error, image intensity, and its gradient
(1)
where k2 is the L-2 norm of vector k, Ω is the region of interest, and r is position. Using this
definition for the cost function, the reconstructed image is a regularized result of image fidelity,
sparsity, and smoothness, respectively. The optimal regularization parameters α and β were
determined by inspection (103 and 104, respectively) and were maintained the same
throughout this study.
As a result of the relatively long data acquisition time (~45 minutes for each 360° acquisition
consisting of 21 source positions and 180 imaging angles with 0.5 and 4 s delays between
consecutive laser scanning positions and phantom rotations, respectively) and the relatively
high energy density of the laser (~50 mW/mm2 at the surface of the phantom), the effect of
photo-bleaching [31] on signal intensity has to be taken into account in image reconstruction:
a fluorescence image taken immediately after regular data acquisition (i.e., the 181st image at
the emission wavelength) produced an average intensity of about 40% less than the 1st image
at all the detector positions. This effect was corrected for each measurement by modeling photo-
bleaching as a global effect on the entire imaging phantom using the following formula:
(2)
where pm,n(i) is the fractional signal intensity measured for the m-th source and the n-th detector
at the i-th imaging angle (0 i N), and N is the total number of imaging angles. It is worth
noting that the index of imaging angle was used to model photo-bleaching instead of using the
exposure time of the fluorophore to the laser because the laser was scanning at a constant time
interval. The effect of photo-bleaching on measurements at the excitation wavelength was
ignored because of the relatively small fraction of absorption by the fluorophore at this
wavelength compared to its surrounding medium in the imaging phantom.
The raw images produced by the image reconstruction algorithm had a matrix size of 23×23×15
with a voxel size of (1 mm)3. They were further processed for later quantitative analysis: in a
raw 3-D image, only the center slice where the sources and detectors were located was retained
due to the symmetry of light propagation about this plane; the background noise was removed
using a threshold such that only two disconnected features remain; and furthermore, the
reconstructed fluorescent structures were defined as the clusters of pixels having values greater
than half of the local maxima. This image processing procedure was programmed using
MATLAB (R2008a, The MathWork, MA) and applied to all raw reconstructed images without
further human intervention. It is worth noting that although the particular choice of the image
processing method is less than ideal and that it has incorporated prior information (we assumed
that two distinct fluorescent structures exist as a prior), which is not generally available in
typical biomedical imaging, the use of this automated procedure greatly reduces the
Zhang and Badea Page 5
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
dependence of the reconstruction results on a particular operator and thus makes image quality
comparison at the later stage more objective.
2.3 Image quality metrics and singular value analysis
The definition of image quality is a key to its evaluation. From a theoretical point of view, in
the context of biomedical imaging the reconstructed image should truthfully represent the
actual shape of the internal structures under investigation, correctly localize the structures, and
also accurately reconstruct intensities of different structures. In accordance with these criteria,
image quality was defined in terms of the following metrics: root mean square (RMS) error,
localization error, and relative intensity. The RMS error describes the global difference
between the reconstructed and the true images
(3)
where xi and x0,i are the i-th elements in the reconstructed and the true images respectively,
and N is the total number of voxels within the FOV. The localization error was defined as the
displacement of the centroid of the smaller fluorescent structure in the reconstructed image
from its true location. The relative intensity was defined as the ratio of the average intensities
of the two reconstructed structures, which is particularly relevant in FDOT because it is an
indicator of imaging artifact arising from the spatially variant sensitivity function associated
with diffuse optical imaging [17,32,33].
From a practical point of view, the efficiency of measurements is another important factor in
biomedical imaging because of the constraints in time and resources due to various reasons.
In this study, the measurement efficiency is quantified by plotting the values of image quality
metrics against the total number of measurements. The number of measurements dictates the
required imaging and computing resources, which in most cases are the limiting factors for
high-resolution or high-speed imaging.
SVA was used to interpret the overall performance of the FDOT system in terms of the number
of truncated singular values, which represents the number of usable modes in the measurement
space that can be mapped onto the image space given by singular value decomposition. More
detailed descriptions on SVA in the context of diffuse optical imaging were given in [27,34].
Specifically, the normalized sensitivity matrix underwent singular value decomposition (SVD)
using built-in routines in MATLAB. The resulting vector of singular values was normalized
by its largest element. The normalized vector of singular values was then truncated at a series
of thresholds (103, 104, 105, 106, and 107) respectively. Subsequently, the remaining
numbers of singular values after truncation were plotted against different acquisition
parameters (numbers of sources, detectors, and imaging angles) and the total number of
measurements, i.e., the product of the acquisition parameters, respectively.
3. Results and Discussion
3.1. Experimental data
Representative reconstruction results from different number of sources (from 1 to 20) and
different number of detectors (from 2 to 10) with fixed 16 imaging angles are shown in Fig. 2.
Image quality shows strong dependence (by visual inspection) on the acquisition parameters
(variable numbers of sources and detectors, with a fixed number of imaging angles). Fig. 2
also clearly demonstrates that image quality improves in general with increased acquisition
parameters; and that abnormalities (“exceptions”) exist when the numbers of source and
Zhang and Badea Page 6
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
detectors are small, as shown in the lower-left quadrant of Fig. 2. Such abnormality in image
quality can mislead towards false interpretations of the images.
In Fig. 3, image quality metrics are plotted against the number of sources, the number of
detectors, and the number of imaging angles (from left to right) for the study cases listed in
Table 2. For RMS and localization errors, as shown in Figs. 3(a) and (b), respectively, a smaller
value indicates better image quality. For relative intensity, shown in Fig. 3(c), however, a value
closer to unity represents better image quality. In these plots, the curves of image quality metrics
either monotonically improve or severely oscillate before they eventually stabilize to a certain
level. Although the total number of measurements can differ significantly in different cases,
for example, comparing Case 1 with Case 2 or 3, the values of all image quality metrics stabilize
after the acquisition parameters are greater than certain critical values: ~5 for the number of
sources, ~15 for detectors, and ~20 for imaging angles, as marked by broken lines in Figs. 3
(a)–(c). Noted that these critical values remain virtually the same even though other parameters
have changed, as shown in Fig. 3(a1) in particular. This might seem to contradict the
conventional belief that more measurement would result in better image quality. In fact, as
further proven using SVA, this phenomenon reflects the differences in measurement
efficiencies of different acquisition parameters.
To better illustrate the differences in measurement efficiency, the same image quality metrics
are plotted against the total number of measurements, which is the product of the numbers of
sources, detectors, and angles, for all study cases, as shown in Figs. 3(d)–(f). We define the
criteria of finding higher measurement efficiency for different acquisition parameters as
“obtaining the same image quality with a smaller number of measurements,” or equivalently
as “obtaining a better image quality with the same number of measurements.” Using these
criteria, Fig. 3(d1) suggests that the increase in the number of sources is the most efficient way
to lower RMS errors. Indeed, doubling either the number of detectors or angles does not reduce
RMS error, although the number of measurements has been doubled for the same number of
sources (Cases 2 and 3 compared to Case 1). From Fig. 3(d2), one can draw the conclusion
that increasing the number of sources is more efficient that the number of detectors, which is
subsequently more efficient than the number of angles, because for the same number of
measurements, doubling the number of sources (Case 5) gives a smaller RMS error, while
doubling the number of angles (Case 6) results in a larger RMS error. Fig. 3(d3) shows that
doubling the number of sources (Case 8) reduces the RMS error. Overall, Fig. 3(d) consistently
demonstrate that for our experimental configuration and reasonably close to the chosen values
of acquisition parameters, increasing the number of sources is more efficient than the number
of detectors, which is in turn is more efficient than increasing the number of imaging angles.
The same conclusion can be drawn using the localization error and relative intensity, as shown
in Figs. 3(e) and (f), respectively. It is relatively less definitive to use localization error to
determine the order of efficiency because its stabilized value reaches sub-millimeter levels,
which are less than the voxel size in image reconstruction, and thus makes comparison
relatively difficult.
3.2. Singular value analysis
The SVA results for different sampling strategies (Cases 1–9 listed in Table 2) are shown in
Fig. 4. In the top row, the NSV above a series of thresholds (103, 104, till 107) are plotted
against the number of sources, the number of detectors, and the number of imaging angles, as
shown in Figs. 4(a), which corresponds to Cases 1, 4, and 7, respectively. In these plots, the
behavior of each curve is noticeably different from each other with respect to different
acquisition parameters: the amplitudes that the curves reach and their linearity, for example.
The difference is more prominent when the threshold is relatively small (starting from 105,
but 106 and 107 in particular).
Zhang and Badea Page 7
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
We compared the critical values observed from the experimental data and SVA, at which the
improvement of image quality and the increment rate of the NSV curves change significantly,
and found that the threshold of 105 in SVA best characterize our experimental data, as shown
in Fig. 4(b). In each of these plots, a NSV curve initially increases rapidly and eventually
reaches a plateau. The transition between these two stages is the “critical value” of the
corresponding acquisition parameter in our FDOT system because it is around this point when
the increment rate of NSV starts to decrease dramatically. It is interesting to notice the similarity
between the NSV curves with different acquisition constants, i.e., comparing curves for Cases
1 through 3, 4 through 6, and 7 through 9. The critical values remain virtually the same even
though other constant parameters have been doubled—the same conclusion drawn from the
experimental data (~5 sources, ~15 detectors, and ~20 angles, as marked by broken lines in
Fig. 3). The shapes of the NSV curves are also consistent with our experimental results. For
example, the transition of the NSV curves using variable number of detectors is more gradual
than those of the other two parameters (Fig. 4[b2] compared to [b1] and [b3]), which results
in less significant changes in image quality (from experimental data) before and after the critical
values with variable number of detectors then the other two acquisition parameters (Fig. 3[a2]–
[c2] compared to [a1]–[c1] and [a3]–[c3]). As expected, when the constant acquisition
parameters were doubled, the NSV curves (green and blue) are in general higher than the
original curves (red).
When the NSV are plotted against the number of measurements, the red curve can be on top
of, in between, or beneath the green and blue curves, as shown in Fig. 4(c), which clearly
indicates the relative measurement efficiencies of different acquisition parameters. More
specifically, because doubling the number of detectors and angles does not increase the NSV
for variable number of sources, as shown in Fig. 4(c1); because doubling the number of sources
increases the NSV, while doubling the number of angles not, for variable number of detectors,
Fig. 4(c2); and because doubling the number of sources and detectors both increase the NSV
for variable number of angles, Fig. 4(c3), it can be concluded that with a threshold of 105,
increasing the number of sources is more efficient in improving NSV than the number of
detectors, which in turn is more efficient than the number of angles. This conclusion is identical
to the one drawn from our experimental data. We note that an increase in the NSV is always
desirable.
However, if using a larger threshold, e.g., 104, the relative order of measurement efficiency
changes to: increasing the numbers of sources or angles is equally more efficient than the
number of detectors, as shown in Fig. 4(d). Using a threshold of 103 gives the same result
(data not shown). Using a smaller threshold, e.g. 106 (Fig. 4(e)) and 107 (data not shown),
the relative efficiencies of acquisition parameters are the same as using the threshold of 105,
except that the differences between the acquisition parameters are less significant and the
overall trend of the NSV curves resembles closer to a straight line, which occurs when the
threshold is zero.
In this study, the overall performance of our FDOT system was characterized by a threshold
of 105 used in SVA. The effect of sampling strategies on image quality was compared in terms
of a number of image quality metrics for reconstruction results and in terms of the NSV in a
more generalized SVA. The results are representative for similar noncontact FDOT systems,
such as [19], and can be used to assist the choice of acquisition parameters in experimental
designs—a tradeoff between image quality and resources.
The NSV curves are in general consistent with the image quality metrics defined in this study.
Nonetheless, a larger NSV does not necessarily translate into a higher score in every metric.
This is because a particular image quality metric represents only one figure of merit of a
complex imaging system, which consists of imaging apparatus and reconstruction methods.
Zhang and Badea Page 8
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
On the other hand, differences due to specific imaging hardware, experimental configurations,
and reconstruction methods can alter the outcome of the analysis. Nonetheless, this study
adopted a way to characterize the overall performance of an imaging system in terms of the
threshold in SVA. With the same imaging apparatus and similar experimental configurations,
this threshold can be used to predict outcomes of FDOT imaging using other sampling
strategies. With this method, a more generalized analysis of sampling strategies can be
conducted with more acquisition parameters, for example, the 2-D distribution of sampling
points for sources and detectors, as well as angular coverage and distribution of imaging angles.
The effect of experimental configuration, e.g., relative angular positions of the scanning laser
and the camera, on the characteristic SVA threshold and further on image quality can be
investigated using this approach as well, which is beyond the scope of this study.
It is important to point out that this study has been purposefully limited to 2-D on a limited
imaging parameters, although indeed FDOT is fundamentally a 3-D imaging method. This is
primarily because the image quality is affected by so many parameters that it would be
impractical to cover the full effects of all possible imaging parameters and experimental
configurations. Instead, we are focusing only on the most fundamental imaging parameters
that are pertinent to the majority of FDOT experiments. For the same reason, the phantom
design and image analysis were limited to 2-D as well in order to avoid being too specific on
particular image processing algorithms and to avoid the complexity of image analysis in 3-D.
4. Conclusions
We found that critical values for acquisition parameters exist, beyond which image quality
improves at a much slower rate. These critical values largely remain the same despite
significant change of other acquisition parameters. In terms of the number of measurements,
increasing the number of sources is more efficient in improving image quality than the number
of detectors and subsequently the number of angles, if the SVA threshold is small (less than
or equal to 105 using our imaging system); or the number of sources and the number of imaging
angles are almost equally more efficient than the number of detectors, if the SVA threshold is
large (greater or equal to 104 in our study). Based on this study, we believe a generalized
optimal imaging strategy can be developed for FDOT if we conduct larger scaled studies based
on more sophisticated modeling techniques that include more acquisition parameters.
Acknowledgment
This study was funded by NIH/NCRR P41 RR005959 and NCI SAIRP U24 CA092656. The authors are grateful to
Professor G. Allan Johnson for his mentorship and to Sally Zimney for her assistance in editing.
References and links
1. Wessels JT, Busse AC, Mahrt J, Dullin C, Grabbe E, Mueller GA. In vivo imaging in experimental
preclinical tumor research--a review. Cytometry A 2007;71:542–549. [PubMed: 17598185]
2. Ntziachristos V, Ripoll J, Wang LV, Weissleder R. Looking and listening to light: the evolution of
whole-body photonic imaging. Nat. Biotechnol 2005;23:313–320. [PubMed: 15765087]
3. Grimm J, Kirsch DG, Windsor SD, Kim CF, Santiago PM, Ntziachristos V, Jacks T, Weissleder R.
Use of gene expression profiling to direct in vivo molecular imaging of lung cancer. Proc Natl. Acad.
Sci. U S A 2005;102:14404–14409. [PubMed: 16183744]
4. Patwardhan S, Bloch S, Achilefu S, Culver J. Time-dependent whole-body fluorescence tomography
of probe bio-distributions in mice. Opt. Express 2005;13:2564–2577.
5. Cherry SR. In vivo molecular and genomic imaging: new challenges for imaging physics. Phys. Med.
Biol 2004;49:R13–R48. [PubMed: 15012005]
Zhang and Badea Page 9
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
6. Herve L, Koenig A, Da Silva A, Berger M, Boutet J, Dinten JM, Peltie P, Rizo P. Noncontact
fluorescence diffuse optical tomography of heterogeneous media. Appl. Opt 2007;46:4896–4906.
[PubMed: 17676093]
7. Patwardhan SV, Culver JP. Quantitative diffuse optical tomography for small animals using an ultrafast
gated image intensifier. J. Biomed. Opt 2008;13:011009. [PubMed: 18315358]
8. Lin Y, Gao H, Nalcioglu O, Gulsen G. Fluorescence diffuse optical tomography with functional and
anatomical a priori information: feasibility study. Phys. Med. Biol 2007;52:5569–5585. [PubMed:
17804882]
9. Meyer H, Garofalakis A, Zacharakis G, Psycharakis S, Mamalaki C, Kioussis D, Economou EN,
Ntziachristos V, Ripoll J. Noncontact optical imaging in mice with full angular coverage and automatic
surface extraction. Appl. Opt 2007;46:3617–3627. [PubMed: 17514324]
10. Ntziachristos V, Weissleder R. Experimental three-dimensional fluorescence reconstruction of
diffuse media by use of a normalized Born approximation. Opt. Lett 2001;26:893–895. [PubMed:
18040483]
11. Turner G, Zacharakis G, Soubret A, Ripoll J, Ntziachristos V. Complete-angle projection diffuse
optical tomography by use of early photons. Opt. Lett 2005;30:409–411. [PubMed: 15762444]
12. Milstein AB, Stott JJ, Oh S, Boas DA, Millane RP, Bouman CA, Webb KJ. Fluorescence optical
diffusion tomography using multiple-frequency data. J. Opt. Soc. Am. A Opt. Image Sci. Vis
2004;21:1035–1049. [PubMed: 15191186]
13. Gibson AP, Hebden JC, Arridge SR. Recent advances in diffuse optical imaging. Phys. Med. Biol
2005;50:R1–R43. [PubMed: 15773619]
14. Jorge R, Vasilis N. Imaging scattering media from a distance: theory and applications of noncontact
optical tomogarphy. Modern Physics Letters, B 2004;18:1403–1431.
15. Ntziachristos V, Tung CH, Bremer C, Weissleder R. Fluorescence molecular tomography resolves
protease activity in vivo. Nat. Med 2002;8:757–760. [PubMed: 12091907]
16. Schulz RB, Ripoll J, Ntziachristos V. Noncontact optical tomography of turbid media. Opt. Lett
2003;28:1701–1703. [PubMed: 13677542]
17. Kepshire DS, Davis SC, Dehghani H, Paulsen KD, Pogue BW. Subsurface diffuse optical tomography
can localize absorber and fluorescent objects but recovered image sensitivity is nonlinear with depth.
Appl. Opt 2007;46:1669–1678. [PubMed: 17356609]
18. Schulz RB, Peter J, Semmler W, D'Andrea C, Valentini G, Cubeddu R. Comparison of noncontact
and fiber-based fluorescence-mediated tomography. Opt. Lett 2006;31:769–771. [PubMed:
16544618]
19. Deliolanis N, Lasser T, Hyde D, Soubret A, Ripoll J, Ntziachristos V. Free-space fluorescence
molecular tomography utilizing 360 degrees geometry projections. Opt. Lett 2007;32:382–384.
[PubMed: 17356660]
20. Bloch S, Lesage F, McIntosh L, Gandjbakhche A, Liang K, Achilefu S. Whole-body fluorescence
lifetime imaging of a tumor-targeted near-infrared molecular probe in mice. J. Biomed. Opt
2005;10:054003. [PubMed: 16292963]
21. Pogue BW, McBride TO, Osterberg UL, Paulsen KD. Comparison of imaging geometries for diffuse
optical tomography of tissue. Optics Express 1999;4:270–286. [PubMed: 19396284]
22. Culver JP, Ntziachristos V, Holboke MJ, Yodh AG. Optimization of optode arrangements for diffuse
optical tomography: A singular-value analysis. Opt. Lett 2001;26:701–703. [PubMed: 18040425]
23. Sun J, Fu K, Wang A, Lin AW, Utzinger U, Drezek R. Influence of fiber optic probe geometry on
the applicability of inverse models of tissue reflectance spectroscopy: computational models and
experimental measurements. Appl. Opt 2006;45:8152–8162. [PubMed: 17068558]
24. Boas DA, Dale AM, Franceschini MA. Diffuse optical imaging of brain activation: approaches to
optimizing image sensitivity, resolution, and accuracy. Neuroimage 2004;23:S275–S288. [PubMed:
15501097]
25. Xu H, Dehghani H, Pogue BW, Springett R, Paulsen KD, Dunn JF. Near-infrared imaging in the
small animal brain: optimization of fiber positions. J. Biomed. Opt 2003;8:102–110. [PubMed:
12542386]
Zhang and Badea Page 10
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
26. Toronov V, D'Amico E, Hueber D, Gratton E, Barbieri B, Webb A. Optimization of the signal-to-
noise ratio of frequency-domain instrumentation for nearinfrared spectro-imaging of the human brain.
Optics Express 2003;11:2717–2729.
27. Lasser T, Ntziachristos V. Optimization of 360 projection fluorescence molecular tomography.
Medical Image Analysis 2007;11:389–399. [PubMed: 17524701]
28. Zhang X, Toronov V, Webb AG. Simultaneous integrated diffuse optical tomography and functional
magnetic resonance imaging of the human brain. Optics Express 2005;13:5513–5521.
29. Schulz, R.; Joerg, P.; Wolfhard, S.; Cosimo, D.; Gianluca, V.; Rinaldo, C. Quantifiability and image
quality in noncontact fluorescence tomography. In: Kai, L.; Rinaldo, C., editors. SPIE. 2005. p.
58590Z
30. Fletcher R, Reeves CM. Function minimization by conjugate gradients. The Computer Journal
1964;7:149–154.
31. Nakayama A, Bianco AC, Zhang CY, Lowell BB, Frangioni JV. Quantitation of brown adipose tissue
perfusion in transgenic mice using near-infrared fluorescence imaging. Mol. Imaging 2003;2:37–49.
[PubMed: 12926236]
32. Jacob M, Bresler Y, Toronov V, Zhang X, Webb A. Level-set algorithm for the reconstruction of
functional activation in near-infrared spectroscopic imaging. J. Biomed. Opt 2006;11:064029.
[PubMed: 17212552]
33. Pogue BW, McBride TO, Prewitt J, Osterberg UL, Paulsen KD. Spatially variant regularization
improves diffuse optical tomography. Appl. Opt 1999;38:2950–2961. [PubMed: 18319877]
34. Graves EE, Culver JP, Ripoll J, Weissleder R, Ntziachristos V. Singular-value analysis and
optimization of experimental parameters in fluorescence molecular tomography. J. Opt. Soc. Am. A
Opt. Image Sci. Vis 2004;21:231–241. [PubMed: 14763766]
Zhang and Badea Page 11
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Fig. 1.
Schematics of (a) the FDOT system setup, (b) experimental configuration and sampling
strategy, and (c) geometry of the imaging phantom.
Zhang and Badea Page 12
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Fig. 2.
Comparison of reconstructed images (23×23 pixels FOV with 1 mm/pixel) using different
numbers of sources and detectors (fixed 16 imaging angles). In the true image, the upper area
is the smaller fluorescent structure (3 mm diameter); and the lower one is the larger structure
(4 mm).
Zhang and Badea Page 13
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Fig. 3.
Image quality metrics derived from experimental data in terms of (a) RMS error, (b)
localization error, and (c) relative intensity in different study cases, from left to right: variable
number of sources (Cases 1 to 3), detectors (Cases 4 to 6), and imaging angles (Cases 7 to 9).
The bottom rows (d–f) are derived from the same data as in (a–c) but are plotted against the
number of measurements.
Zhang and Badea Page 14
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Fig. 4.
(a) and (b) Comparison of the NSV against variable acquisition parameters and (c)–(e) the
number of measurements using different thresholds in SVA in each of the study cases, from
left to right: variable number of sources (Cases 1 to 3), number of detectors (Cases 4 to 6), and
number of angles (Cases 7 to 9). All of the vertical axes represent the NSV, although their
scales are different among different rows. The horizontal axes in (a) and (b) are the number of
source, the number of detectors, and the number of angles (left to right, respectively); and those
in (c)–(e) are the number of measurements.
Zhang and Badea Page 15
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Zhang and Badea Page 16
Table 1
Optical properties of the imaging phantom at 785 nm.
Component Intralipid-ink Nylon 6/6
Absorption coefficient µa (1/cm) 0.1 <0.001
Reduced scattering coefficient µ’s (1/cm) 10 3.9
Index of refraction n1.4 1.5
Opt Express. Author manuscript; available in PMC 2010 March 30.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Zhang and Badea Page 17
Table 2
Sampling strategies used in experiment (var are variable acquisition parameters.)
Case #1 #2 #3 #4 #5 #6 #7 #8 #9
Number of sources var var var 5 10 5 5 10 5
Number of detectors 15 30 15 var var var 15 15 30
Number of angles 20 20 40 20 20 40 var var var
Opt Express. Author manuscript; available in PMC 2010 March 30.
... Diffuse optical tomography (DOT) has increasingly been applied in many clinical areas, such as breast imaging [1][2][3], joint imaging [4][5][6], brain imaging [7,8], vascular imaging [9,10], and small animal imaging [11][12][13], through model-based image reconstruction algorithms [14][15][16], due to its non-ionizing radiation, portability, real-time imaging, and low instrumentation cost. ...
... Matrix Q is also called the measurement operator, a mandatory component in optimization problems. This model provides a way to derive Q with (19), (20), (10) and (11). We employ backward ray tracing [37] for the construction of Q. ...
Article
Full-text available
This work presents a generalized angle-dependent ray transfer function that can accurately map the angular and spatial distribution of light intensities on the tissue surface onto a camera image plane in a non-contact camera-based imaging system. The method developed here goes beyond existing ray transfer models that apply to angle-averaged tomographic data alone. The angle-dependent ray transfer operator was constructed using backward ray tracing based on radiation surface theory. The proposed method was validated using numerical phantoms and experimental data from an actual non-contact imaging system.
... This is especially true in biomedical optical imaging, in which direct detection is often the approach for signal acquisition. As a result, the performance of imaging instrument and methodology were directly associated to the specific hardware used (or, assumed) in the studies, e.g., the results demonstrated in [1][2][3][4][5]. Under many circumstances, one wishes to incorporate desired features of advanced detection technologies in new applications to create novel capabilities. ...
... We applied our previously developed regularized iterative inversion algorithm based on a conjugate gradient method [4], which minimizes the following cost function ...
Article
Full-text available
Dense spatial sampling is required in high-resolution optical imaging and many other biomedical optical imaging methods, such as diffuse optical imaging. Arrayed photodetectors, in particular charge coupled device cameras are commonly used mainly because of their high pixel count. Nonetheless, discrete-element photodetectors, such as photomultiplier tubes, are often desirable in many performance-demanding imaging applications. However, utilization of the discrete-element photodetectors typically requires raster scan to achieve arbitrary retrospective sampling with high density. Care must be taken in using the relatively large sensitive areas of discrete-element photodetectors to densely sample the image plane. In addition, off-line data analysis and image reconstruction often require full-field sampling. Pixel-by-pixel scanning is not only slow but also unnecessary in diffusion-limited imaging. We propose a super-resolution method that can recover the finer features of an image sampled with a coarse-scale sensor. This generalpurpose method was established on the spatial transfer function of the photodetector-lens system, and achieved super-resolution by inversion of this linear transfer function. Regularized optimization algorithms were used to achieve optimized deconvolution. Compared to the uncorrected blurred image, the proposed super-resolution method significantly improved image quality in terms of resolution and quantitation. Using this reconstruction method, the acquisition speed with a scanning photodetector can be dramatically improved without significantly sacrificing sampling density or flexibility.
... These parameters were applied to the Jacobian matrix. [1][2][3][4][5][6] These methods were so time-consuming due to the large dimensions of the Jacobian matrix. For solving the time-consuming problem, the number of nodes must be reduced as much as possible, or the simple two-or three-dimensional geometries were used. ...
Article
Full-text available
Background: In diffuse optical tomography, determining the optimal angle between the source and detector is an effective method to reduce the number of projections while maintaining the quality of the reconstructed images. In this study, a new parameter is introduced to evaluate the source-detector geometries. Methods: A two-dimensional mesh with the radius of 20 mm and 7987 nodes were built. In each reconstruction, 0.5 mm heterogeneity with the absorption coefficient of 0.06 mm-1 and the dispersion coefficient of 0.6 mm-1 was added in different parts of the sample randomly. The relationship between the mean square error (MSE), sensitivity Laplacian ratio (SLR), and sensitivity standard deviation ratio (SSR) was evaluated based on their correlation coefficients. The quality of the images achieved using the optimized projections were compared with that of the full projections for the same depths. Results: MSE decreases by increasing the SLR magnitudes which indicate that the parameter could be used to evaluate the scanning geometries. There was a negative correlation coefficient (R = -0.76) with the inverse relationship between the SLR and MSE indices. SSR does not have a significant relationship with the quality of the reconstructed images. For each scanning depth, the comparison of the images obtained using the full and optimized-selective projections did not show any considerable difference despite the decrease of the projection numbers in scanning geometry with the optimized-selective projections. Conclusion: The unnecessary projections could be eliminated by placing the detectors at the specific angles, which were determined using the SLR. Thus, a proper compromise between the quality of the reconstructed images and reconstruction time might establish.
... [4,5] The optimization of the source-detector geometry is one of the most practical and effective methods that have been used in recent studies. [6,7] Various methods, such as singular-value analysis [8,9] and orthogonality of the Jacobian matrix, were used to evaluate the source-detector geometries and to optimize sampling frequency, the field of view, etc. [10,11] However, the singular-value decomposition and orthogonal methods could not assess different geometries. They did not change steadily with the changes of the source-detector geometry. ...
Article
Full-text available
Background Molecular fluorescence imaging is widely used as a noninvasive method to study the cellular and molecular mechanisms. In the optical imaging system, the sensitivity is the change of the intensity received by the detector for the changed optical characteristics (fluorescence) in each sample point. Sensitivity could be considered as a function of imaging geometry. A favor imaging system has a uniform and high-sensitivity coefficient for each point of the sample. In this study, a new parameter was proposed which the optimal angle between the source and detector could be determined based on this parameter. Methods For evaluation of the new method, a two-dimensional mesh with a radius of 20 mm and 5133 nodes was built. In each reconstruction, 0.5-mm fluorescence heterogeneity with a contrast-to-purpose ratio of fluorescence yield of 10 was randomly added at different points of the sample. The source and the detector were simulated in different geometric conditions. The calculations were performed using the NIRFAST and MATLAB software. The relationship between mean squared error (MSE) and sensitivity uniformity ratio (SUR) was evaluated using the correlation coefficient. Finally, based on the new index, an optimal geometrical strategy was introduced. Results There was a negative correlation coefficient (R = −0.78) with the inverse relationship between the SUR and MSE indices. The reconstructed images showed that the better image quality achieved using the optimal geometry for all scanning depths. For the conventional geometry, there is an artifact in the opposite side of the inhomogeneity at the shallow depths, which has been eliminated in the reconstructed images achieved using the optimal geometry. Conclusion The SUR is a powerful computational tool which could be used to determine the optimal angles between the source and detector for each scanning depth.
... In the diffuse optics community, theoretical performance prediction for system design or optimization is usually accomplished with Monte Carlo simulation [6,9,10] or with singular value decomposition (SVD) [11][12][13][14]. The former can be very time-and resource-intensive, and the results are valid only for the particular inversion algorithm simulated. ...
Article
Full-text available
Spatial frequency domain imaging (SFDI) is a wide-field diffuse optical imaging modality that has attracted considerable interest in recent years. Typically, diffuse reflectance measurements of spatially modulated light are used to quantify the optical absorption and reduced scattering coefficients of tissue, and with these, chromophore concentrations are extracted. However, uncertainties in estimated absorption and reduced scattering coefficients are rarely reported, and we know of no method capable of providing these when look-up table (LUT) algorithms are used to recover the optical properties. We present a method to generate optical property uncertainty estimates from knowledge of diffuse reflectance measurement errors. By employing the Cramér-Rao bound, we can quickly and efficiently explore theoretical SFDI performance as a function of spatial frequencies and sample optical properties, allowing us to optimize spatial frequency selection for a given application. In practice, we can also obtain useful uncertainty estimates for optical properties recovered with a two-frequency LUT algorithm, as we demonstrate with tissue-simulating phantom and in vivo experiments. Finally, we illustrate how absorption coefficient uncertainties can be propagated forward to yield uncertainties for chromophore concentrations, which could significantly impact the interpretation of experimental results.
... These parameters were applied to the Jacobian matrix. [1][2][3][4][5][6] These methods were so time-consuming due to the large dimensions of the Jacobian matrix. For solving the time-consuming problem, the number of nodes must be reduced as much as possible, or the simple two-or three-dimensional geometries were used. ...
Article
Full-text available
Background Cervical cancer brachytherapy has an effective role on the tumor control probability as a boost and/or single treatment option. High-dose-rate 60Co brachytherapy units are used in many radiation oncology centers in Iran. Rectum and bladder tissues are considered as organs at risk in radiation treatment of cervical cancers, and they should be spared from unwanted radiation risk. Purpose In the present study, the effect of additional tungsten shield was investigated in a new colpostat design for bladder protection against radiation. Methods and Materials Monte Carlo (MC) simulation has been performed using the MC N-Particle eXtended version 2.4.0 transport code. The HDR GZP6 brachytherapy source applicator was simulated along with its colpostats. The f6 tally was used for absorbed dose calculation; and for model validation, we used from dosimetric features of the GZP6 treatment planning system. Results Results calculated by MC simulation method showed that dose reduction at the end of the colpostat was 2.44% for the medium colpostat and a dose increase of 1.35% was obtained for the small colpostat. In the reference point of the bladder (at the distance 1 cm from the end point of the colpostat), the percentages of dose reductions were also 25% and 15% for medium and small colpostats, respectively. Conclusions Results show that the absorbed dose in the bladder tissue can be reduced significantly using a shielded colpostat.
... One such approach that has proven popular is singular value analysis (SVA), which involves computing the singular value spectrum of the Jacobian (sensitivity matrix of the forward problem) associated with a particular system configuration and counting the number of singular values above a given threshold. [3][4][5][6][7][8][9] While SVA is conceptually straightforward, it does not make use of the information contained in the singular vectors of the Jacobian, which can lead to errors in performance predictions. 10,11 It also does not exploit knowledge of the covariance of the measurements. ...
Article
We evaluated the potential of the Cramér-Rao lower bound (CRLB) to serve as a design metric for diffuse optical imaging systems. The CRLB defines the best achievable precision of any estimator for a given data model; it is often used in the statistical signal processing community for feasibility studies and system design. Computing the CRLB requires inverting the Fisher information matrix (FIM), however, which is usually ill-conditioned (and often underdetermined) in the case of diffuse optical tomography (DOT). We regularized the FIM by assuming that the inhomogeneity to be imaged was a point target and assessed the ability of point-target CRLBs to predict system performance in a typical DOT setting in silico. Our reconstructions, obtained with a common iterative algebraic technique, revealed that these bounds are not good predictors of imaging performance across different system configurations, even in a relative sense. This study demonstrates that agreement between the trends predicted by the CRLBs and imaging performance obtained with reconstruction algorithms that rely on a different regularization approach cannot be assumed a priori. Moreover, it underscores the importance of taking into account the intended regularization method when attempting to optimize source-detector configurations.
... Using simulations, we have confirmed that an increase in the number of pinholes from 5 to 20, would increase the condition number from 12.95 to26.32 and will result in more inaccuracies in the reconstruction. A singular value analysis of the optimal number of x-ray pencil beams, their optimal translation and the number of angles is not the focus of this study but could be performed as described before in FDOT imaging [9]. ...
Article
Full-text available
X-ray Luminescence CT (XLCT) is a hybrid imaging modality combining x-ray and optical imaging in which x-ray luminescent nanophosphors (NPs) are used as emissive imaging probes. NPs are easily excited using common CT energy x-ray beams, and the NP luminescence is efficiently collected using sensitive light based detection systems. XLCT can be recognized as a close analog to fluorescence diffuse optical tomography (FDOT). However, XLCT has remarkable advantages over FDOT due to the substantial excitation penetration depths provided by x-rays relative to laser light sources, long term photo-stability of NPs, and the ability to tune NP emission within the NIR spectral window. Since XCLT uses an x-ray pencil beam excitation, the emitted light can be measured and back-projected along the x-ray path during reconstruction, where the size of the X-ray pencil beam determines the resolution for XLCT. In addition, no background signal competes with NP luminescence (i.e., no auto fluorescence) in XLCT. Currently, no small animal XLCT system has been proposed or tested. This paper investigates an XLCT system built and integrated with a dual source micro-CT system. Two novel sampling paradigms that result in more efficient scanning are proposed and tested via simulations. Our preliminary experimental results in phantoms indicate that a basic CT-like reconstruction is able to recover a map of the NP locations and differences in NP concentrations. With the proposed dual source system and faster scanning approaches, XLCT has the potential to revolutionize molecular imaging in preclinical studies.
Conference Paper
Based on light propagation theory, the measurements of a contact-free imaging system with generalized optical components can be obtained from a linear transformation of the light intensity distribution on the surface of the imaging object. In this work, we derived the linear measurement operator needed to perform this transformation. Numerical experiments were designed and conducted for validation.
Article
Full-text available
Even though endoscopic fluorescence diffuse optical tomography (eFDOT) exhibits significant potential, currently its application is limited due to the existence of strong autofluorescence and the imaging inaccuracy caused by a very short source-detector distance. Motivated by the emerging X-ray luminescence computed tomography (XLCT) technology, we presented an endoscopic XLCT (eXLCT) methodology. In the methodology, the aperture angle of the objective was incorporated into the forward model, providing a more accurate description of light propagation. Numerical simulation with a heterogeneous geometry and an imaging experiment with a physical phantom were illustrated to demonstrate the feasibility of the presented eXLCT methodology.
Article
Full-text available
A complete methodology has been developed to integrate simultaneous diffuse optical tomography (DOT) and functional magnetic resonance imaging (MRI) measurements. This includes development of an MRI-compatible optical probe and a method for accurate estimation of the positions of the source and detector optodes in the presence of subject-specific geometric deformations of the optical probe. Subject-specific head models are generated by segmentation of structural MR images. DOT image reconstruction involves solution of the forward problem of light transport in the head using Monte Carlo simulations, and inversion of the linearized problem for small perturbations of the absorption coefficient. Initial results show good co-localization between the DOT images of changes in oxy- and deoxyhemoglobin concentration and functional MRI data.
Article
Full-text available
We present a fast scanning fluorescence optical tomography system for imaging the kinetics of probe distributions through out the whole body of small animals. Configured in a plane parallel geometry, the system scans a source laser using a galvanometer mirror pair (tauswitch~1ms) over flexible source patterns, and detects excitation and emission light using a high frame rate low noise, 5 MHz electron multiplied charge-coupled device (EMCCD) camera. Phantom studies were used to evaluate resolution, linearity, and sensitivity. Time dependent (deltat=2.2 min.) in vivo imaging of mice was performed following injections of a fluorescing probe (indocyanine green). The capability to detect differences in probe delivery route was demonstrated by comparing an intravenous injection, versus an injection into a fat pocket (retro orbital injection). Feasibility of imaging the distribution of tumor-targeted molecular probes was demonstrated by imaging a breast tumor-specific near infrared polypeptide in MDA MB 361 tumor bearing nude mice. A tomography scan, at 24 hour post injection, revealed preferential uptake in the tumor relative to surrounding tissue.
Article
Optical tomography of turbid media has been largely limited to systems that require fixed geometries or measurements employing fibers. Noncontact optical measurements from diffuse media could facilitate the use of large detector arrays at multiple angles that are well-suited for tomography applications. Such imaging strategies eliminate the need for individual fibers in contact with the highly scattering volume, the use of restricted geometries and the need for matching fluids. Here we review the different approaches and systems developed for noncontact optical measurements and concurrent registration of the three-dimensional surface information of the diffuse medium. We present the basic theoretical formulation and its experimental validation, finally applying it to the specific case of fluorescence tomography of small animals. We discuss how these new technologies can considerably simplify experimental procedures and improve our ability to visualize functional and molecular processes in vivo. Future perspectives and work are also outlined.
Article
Non-contact detection schemes for optical or fluorescence tomography offer several advantages compared to classic approaches, most importantly the ability to obtain images with a CCD in the absence of a matching fluid or fiber optics. This allows the acquisition of high density datasets, as well as simplified experimental procedures. Herein we create a unified framework for contact and non-contact detection procedures and present experimental results that show the ability of the non-contact method to quantify the concentration of fluorochromes hidden in turbid media as well as the improvement in image quality between conventional and non-contact detection.
Article
A quadratically convergent gradient method for locating an unconstrained local minimum of a function of several variables is described. Particular advantages are its simplicity and its modest demands on storage, space for only three vectors being required. An ALGOL procedure is presented, and the paper includes a discussion of results obtained by its used on various test functions.
Article
Frequency-domain near-infrared spectro-imaging offers significant advantages over the continuous-wave method in human brain applications. However, the drawback of existing instruments is a low signal-to-noise ratio for measured phase and modulation depth changes caused by cerebral activation. In this paper we show that in the case of the geometry specific for the activated area in the human brain, the SNR can be significantly improved by increasing the modulation frequency. We present the results of two studies: one performed experimentally using a subnanosecond pulsed light source and a spherical absorbing inhomogeneity immersed in a highly scattering solution, and the other performed numerically using Monte Carlo simulations of light transport in an MRI based digital phantom of the adult human head. We show that changes caused by the absorbing inhomogeneity in both phase and modulation depth increase with frequency and reach maximum values at frequencies between 400 and 1400 MHz, depending on the particular source-detector distance. We also show that for the human head geometry an increase of the modulation frequency from 100 to 500 MHz can increase the phase SNR 2-3 times, and the modulation depth SNR up to 10 times.
Article
Images produced in six different geometries with diffuse optical tomography simulations of tissue have been compared using a finite element-based algorithm with iterative refinement provided by the Newton-Raphson approach. The source-detector arrangements studied include (i) fan-beam tomography, (ii) full reflectance and transmittance tomography, as well as (iii) sub-surface imaging, where each of these three were examined in a circular and a flat slab geometry. The algorithm can provide quantitatively accurate results for all of the tomographic geometries investigated under certain circumstances. For example, quantitatively accurate results occur with sub-surface imaging only when the object to be imaged is fully contained within the diffuse projections. In general the diffuse projections must sample all regions around the target to be characterized in order for the algorithm to recover quantitatively accurate results. Not only is it important to sample the whole space, but maximal angular sampling is required for optimal image reconstruction. Geometries which do not maximize the possible sampling angles cause more noise artifact in the reconstructed images. Preliminary simulations using a mesh of the human brain confirm that optimal images are produced from circularly symmetric source-detector distributions, but that quantitatively accurate images can be reconstructed even with a sub-surface imaging, although spatial resolution is modest.
Article
Systematic efforts are under way to develop novel technologies that would allow molecular sensing in intact organisms in vivo. Using near-infrared fluorescent molecular beacons and inversion techniques that take into account the diffuse nature of photon propagation in tissue, we were able to obtain three-dimensional in vivo images of a protease in orthopic gliomas. We demonstrate that enzyme-activatable fluorochromes can be detected with high positional accuracy in deep tissues, that molecular specificities of different beacons towards enzymes can be resolved and that tomography of beacon activation is linearly related to enzyme concentration. The tomographic imaging method offers a range of new capabilities for studying biological function; for example, identifying molecular-expression patterns by multispectral imaging or continuously monitoring the efficacy of therapeutic drugs.
Article
We investigate fiber placement issues associated with a hybrid magnetic resonance imaging (MRI) near-infrared (NIR) imaging technique for small animal brain studies. Location of the optical fibers on the cranium is examined, with an emphasis on maximizing the recovered resolution and contrast in the region of interest, which in this case is the murine brain. In a series of simulation studies, singular value decomposition of the Jacobian is used in order to determine the measurement sites that provide the most information about the region of interest. The modeling results indicate that data collected using fibers arranged on one side of the head near the brain contain as much information about optical changes within the brain as those positioned equally spaced around the entire periphery of the head. Practical space limitation considerations favor the one-sided fiber array geometry in the case where the NIR acquisition is expected to occur simultaneously with MRI.
Article
Brown adipose tissue (BAT; brown fat) is the principal site of adaptive thermogenesis in the human newborn and other small mammals. Of paramount importance for thermogenesis is vascular perfusion, which controls the flow of cool blood in, and warmed blood out, of BAT. We have developed an optical method for the quantitative imaging of BAT perfusion in the living, intact animal using the heptamethine indocyanine IR-786 and near-infrared (NIR) fluorescent light. We present a detailed analysis of the physical, chemical, and cellular properties of IR-786, its biodistribution and pharmacokinetics, and its uptake into BAT. Using transgenic animals with homozygous deletion of Type II iodiothyronine deiodinase, or homozygous deletion of uncoupling proteins (UCPs) 1 and 2, we demonstrate that BAT perfusion can be measured noninvasively, accurately, and reproducibly. Using these techniques, we show that UCP -1/-2 knockout animals, when compared to wild-type animals, have a higher baseline perfusion of BAT but a similar maximal response to beta 3-receptor agonist. These results suggest that compensation for UCP deletion is mediated, in part, by the control of BAT perfusion. Taken together, BAT perfusion can now be measured noninvasively using NIR fluorescent light, and pharmacological modulators of thermogenesis can be screened at relatively high throughput in living animals.