Open Access
1 July 2003 In vivo endoscopic tissue diagnostics based on spectroscopic absorption, scattering, and phase function properties
Philippe Thueler, Igor Charvet, Frédéric P. Bevilacqua, Michel Saint Ghislain, G. Ory, Pierre Marquet, Paolo Meda, Ben Vermeulen, Christian D. Depeursinge
Author Affiliations +

1.

Introduction

The optical properties of a biological tissue depend on its biochemical composition and its cellular and subcellular structure. Therefore optical measurements have a strong potential for the development of noninvasive in vivo medical diagnostic tools, often called “optical biopsy.” Such techniques should significantly improve the efficiency of biopsies or help in determining the tumor margins in a surgical field. Several types of optical properties can be measured, such as fluorescence,1 2 Raman scattering,3 and absorption and scattering,2 4 5 6 7 8 each of them giving complementary information on the tissue.4

We focus in this paper on the absorption and scattering properties of tissue. In the visible and near-infrared range, the absorption spectrum is related to the concentration of chromophores, such as oxyhemoglobin and deoxyhemoglobin, fat and water.9 Such chromophores vary significantly with tissue metabolism.10 The scattering properties are related to the size distribution of cells and organelles, which are parameters used to differentiate normal from abnormal tissues in standard histopathology.5 6 Previous studies showed that the absorption and scattering of tumor tissues differ from their surrounding normal tissue.7 11 12

The absorption and scattering properties of a bulk tissue can be calculated from measurements of the diffuse reflectance using an appropriate model of the light propagation in tissue. Such techniques have been extensively described when interrogating a relatively large part of the tissue, typically cubic centimeters.10 11 12 In such cases, the light propagation can be modeled using diffusion theory.13 14 Nevertheless, potential medical applications, such as biopsy guidance, require the investigation of a small volume of tissue, on the order of few cubic millimeters. In such a case, only small source–detector separations must be used.

Promising results have been obtained using spectroscopic reflectance measurements at a short source–detector separation,2 4 5 6 7 15 16 demonstrating that important diagnostic information is contained in the reflectance spectrum. Nevertheless, most groups base their analysis on the wavelength dependence of the raw reflectance, without separating the absorption and scattering properties. In such a case, the interpretation of the reflectance spectrum in terms of tissue physiology is difficult, owing to coupling between the tissue absorption and scattering properties. Since absorption and scattering are fundamentally linked with different aspects of tissue physiology, and chromophore content and structure, respectively, the determination of both properties individually is expected to improve the discrimination among different tissues types and/or alterations.

We showed recently,7 17 along with other groups,4 8 18 that the local and superficial (average volume of 1 mm3) determination of the scattering and absorption properties of tissue was possible noninvasively. Our method relies on the measurements of reflectance spectra at different small source–detector separations, and on a model based on Monte Carlo simulations. The model takes into account the influence of the phase function and does not require any assumption about the optical properties of the medium, unlike other methods.4 8 18 Moreover, the use of spatially resolved reflectance could accommodate more complex models that take into account multilayer geometries.19

As shown recently by Kienle et al.;,20 the accuracy of the optical properties can be adversely affected if the effect of the phase function is not incorporated into the model. We consider the phase function parameter γ=(1−g2)/(1−g1), where g1 and g2 are the first two moments of the phase function, respectively. This parameter is the most important parameter of the phase function for reflectance measurements close to the source,17 and is related to the size distribution of scatterers inside the tissue.

In a previous paper we reported measurements performed in vivo on human brain.7 The measurements were performed only at few wavelengths and the data processing was not optimized. We present here further experimental validation and several significant improvements of the method, which make it routinely applicable in clinical situations. In particular, we built an instrument and an optical probe capable of measuring spectrally and spatially resolved reflectance through the working channel of an endoscope. We also developed an efficient and fast algorithm for the determination of broadband spectra of the absorption coefficient, reduced scattering coefficient, and phase function parameter γ, based on cubic B-spline interpolation of Monte Carlo simulations.

We first describe the methodology, including the experimental setup and the fitting algorithm. Second, we report results on the accuracy and precision of the method, based on simulations and tissue phantoms of known optical properties. Third, we report in vivo clinical measurements performed endoscopically on a human gastrointestinal tract.

2.

Materials and Methods

2.1.

Experimental Setup

The schematic of the experimental setup is shown in Fig. 1. The light source is a halogen lamp (Ocean Optics LS-1) that is stabilized (variations less than 1 after 40 min of warming up). Other light sources can be selected by an optical switch. The optical switch has one output connected to the illuminating fiber of the optical probe and another one to a reference channel that is connected directly to the spectrometer. Temporal variations of the source spectrum are corrected by the measurement made on the reference channel.

Figure 1

General scheme of the acquisition setup.

018303j.1.jpg

The optical probe is made of eleven optical fibers, one for illumination and ten for detection [see Fig. 2(e)]. The fibers usedare multimode silica fibers (FVP-200 PF from Thorlabs Inc.) with a numerical aperture of 0.22 and a core diameter of 200 μm. They are coated with a 20-μm polyimide sheath. The ten detection fibers are placed at various distances (noted as ρi, i=1–10) from the illumination fiber, ranging approximately from 0.3 to 1.35 mm with a step of approximately 0.1 mm. They are held in place in the probe by two laser micromachined alignment pieces. This procedure leads to a high accuracy in fiber positioning (approximately 15 μm), which allows the construction of several identical probes. All the components have been chosen to with stand routine disinfections and/or sterilization without damage.

Figure 2

Diagram of the algorithm for the automatic spectral determination of μa, μs , and γ. (a) First, a set of reflectance curves simulated by the Monte Carlo method is produced for a wide range of discrete values for each optical coefficient (i.e., μa, μs , and γ). (b) Second, the simulated curves are interpolated by cubic B-splines to provide a reflectance curve (c) for any value of μa, μs , and γ. Finally, for each wavelength of the spectrum, a fit (d) of the measured reflectance curve (e) is performed onto the interpolated set of simulated curves, allowing for the determination of μa, μs , and γ. The fitting algorithm used is Levenberg-Marquardt.

018303j.2.jpg

The ten detecting fibers are connected to an imaging spectrometer (Jobin-Yvon CP 200) with a Peltier-cooled CCD camera (Hamamatsu C7041 and S7031-1008). A mechanical shutter sets the integrating time of the CCD camera (usually about 600 ms). The image acquired on the camera is digitized by a 12-bit analog/digital card (National Instrument PCI-MIO-16E-4). A single computer is used to control the instrumentation and to process the data. A typical measurement takes 4 s, including a first acquisition while the tissue is illuminated by the source fiber and a second acquisition made with the source fiber switched off. This allows the subtraction of the dark current noise and the background light.

In order to obtain absolute reflectance spectra, a two-step calibration procedure is performed. First, the effect of the spectral responses of the source, fibers, grating, and detector is corrected by performing a measurement on a spectrally flat reflectance standard. Typically, the probe is inserted inside an integrating sphere (made of Spectralon SRS-99, Labsphere Inc.). Second, the effective source intensity is obtained by performing a series of measurements on a solid turbid siloxane phantom whose optical properties a, μs , and γ) were previously determined at 675 nm by combining frequency-domain and spatially resolved measurements.7 The comparison of the phantom measurement with the expected absolute reflectance from a Monte Carlo simulation allows the calibration of the absolute intensity of the experimental reflectance. Note that this calibration factor is wavelength independent, since all the spectral features of the instrumentation have been corrected in the first step. This calibration routine is performed every time the probe is reconnected to the setup, before or after the measurement session. Repeated measurements on the calibration phantom over a six-month period showed a standard deviation of approximately 4, which was mostly due to the variability of the coupling between the probe and the phantom. Once the absolute reflectance spectra are calculated, the optical coefficients μa(λ), μs (λ) and γ(λ) are determined by the algorithm presented in Sec. 2.4.

2.2.

Optical Model at Short Source–Detector Distances: The γ Parameter

Consider a tissue illuminated by a fiber with a unit of incident power. The spatially resolved reflectance R(ρ) is defined as the power of the backscattered light per unit of area detected by a second fiber at the surface of the tissue at a distance ρ from the source. R(ρ) depends on the optical properties of the sample, i.e., the absorption coefficient μa, the scattering coefficient μs, and the phase function p(cos θ) (where θ is the scattering angle), the refractive index n, and the numerical aperture NA of the detecting fibers.

The phase function can be expanded into a series of Legendre polynomials Pn(cos θ):

Eq. (1)

p(cos θ)=n(2n+1)gnPn(cos θ).
In this expression, gn is the n’th order moment of the phase function:

Eq. (2)

gn=2π0πPn(cos θ)p(cos θ)d(cos θ),
where g1 is the conventional anisotropy factor, generally noted as g. In the diffusion approximation of the transport theory, g1 is the only parameter used to describe the phase function. In this case, it is useful to define the reduced scattering coefficient μs s(1−g1).

The diffusion approximation generally holds for μs ρ>5 and μs >100μa. For shorter source–detector separations, typically when 0.5<μs ρ<5, we showed17 that the second moment of the phase function g2 must also be taken into account. This case corresponds to the measurement performed with the probe described in Sec. 2.1.

Using the similarity relations introduced by Wyman et al.;,21 it is easy to show that besides the refractive index, only three parameters are needed to accurately describe the light propagation at such short source–detection distances17: μa, μs , and γ where

Eq. (3)

γ=1g21g1
and γ depends on the scatterer’s relative refractive index and on the ratio between the scatterer size and the wavelength; γ varies between 0.9 (Rayleigh-type scattering) and values larger than 2 (large scatterers compared with wavelength). We also showed that for a fractal distribution, γ is related to the fractal power of the size distribution of the scatterers.22 Note also that for a Henyey-Greenstein phase function, g2=g1 2 and, therefore,

Eq. (4)

g=g1=γ1.

2.3.

Monte Carlo Simulation

We used Monte Carlo simulations as our basis for the model of light propagation. As already mentioned, standard diffusion theory is inadequate for the small source–detector separation we used in experiments17 The Monte Carlo simulations allow us to generate reflectance curves noted as R sim ias ,γ) [see Fig. 2(a)]. The code we used has been extensively tested and allows the use of any phase function given in discretized form,23 24 We used a modified form of the Henyey-Greenstein25 phase function p HG (θ,g HG ), noted as p MHG (θ,g HG ,ξ) (see Ref. 17):

Eq. (5)

pMHG(θ,gHG,ξ)=ξ×pHG(θ,gHG)+(1ξ) 34π cos2 θξ=01,
where ξ is a normalization factor. According to Eq. (1), adding the cos 2 θ term only affects the value of g2. Thus, choosing g1 and g2, i.e., setting γ, uniquely determines the values of g HG and ξ. The tissue was considered to be homogeneous and its refractive index was fixed at 1.4. Our simulations took into account the geometry of the probe. The fibers had a diameter of 200 μm; the index mismatch between the probe (n=1.5) and the tissue was taken into account by the Fresnel law. We showed in a previous paper7 that for the type of probe used in the experiment, the index mismatch outside the probe does not play a significant role.

In order to decrease the Monte Carlo noise and limit the number of photons (one million photons have been simulated for each reflectance curve), the numerical aperture was set at a higher value than in the experiments (0.279 instead of 0.157). This use of the higher numerical aperture was chosen as a tradeoff between simulation time, noise in the simulation, and accuracy degradation. Several tests showed that after renormalization of the intensity, this mismatch introduced a typical error smaller than 2 and a maximum error of 4. Note that in both the simulation and the experiments, the numerical aperture corresponds to a narrow solid angle (a half angle of 16° instead of 9°) and the angular distribution is close to a Lambertian distribution for such small-angle ranges.

Approximately forty discrete values were chosen for μa and μs , ranging respectively from 0.003 to 10 mm−1 and from 0.5 to 10 mm−1; twenty values were chosen for γ, ranging from 1.0 to 2.9. This resulted in a four-dimensional matrix of simulated reflectance curves, noted as R sim ias ,γ), each coefficient taking only discrete values. Using these parameters, 32,000 reflectance curves have been computed, each for a specific triplet of optical coefficients as ,γ). Since the path lengths and exit positions of each simulated photon were stored, a full Monte Carlo simulation was required only for each different value of γ (thus twenty simulations only). Assuming the tissue is homogeneous, scaling relationships17 allow us to derive any reflectance curve R sim ias ,γ) from a single simulation R sim ia=0,μs =1,γ), γ being kept constant.

2.4.

Algorithm for the Determination of μa, μs , and γ

We denote ρi, i=1–10, the distances between the ten detecting fibers and the illuminating fiber, and R mes i) as the measured reflectance at these distances for a given wavelength. The inverse problem corresponds to determining the optical coefficient triplet as ,γ) that is the closest to the reflectance curve R mes i).

The basic idea of the algorithm, as shown in Fig. 2, is to fit the measured reflectance curve R mes i) onto a set of previously simulated reflectance curves, forming a matrix noted as R sim ias ,γ). An algorithm based on a lookup table can be performed as a first step. However, in such a case, the number of simulations performed naturally limits the accuracy of the result. Since the reflectance curves are continuous in all their parameters, this problem could be overcome by interpolating the basic set of simulations. The result of the interpolation is thus a function sim ias ,γ), which is a numerical function defined for any value of the optical coefficient triplets as ,γ) within their definition ranges and for each discrete distance ρi, i=1–10.

The matrix R sim ias ,γ) was interpolated with cubic B-splines.26 Scho¨nberg et al.;27 have shown that any discrete function R(x), (x∈ℤ) may be interpolated by a unique combination c(k) of B-splines βn(x):

Eq. (6)

R(x)=kZc(k)βn(xk),
where c(k) is the weighing coefficient of the B-spline βn(x−k) centered on k.

Cubic B-splines β3(x) are defined like pieces of polynomials of degree 3, smoothly connected:

Eq. (7)

β3(x)={23|x|2+|x|3/2|x|1(2|x|)3/61<|x|20|x|>2
Interpolation of multivariable functions is generalized by replacing c(k) by c(k,l,…,m) and by replacing βn(x−k) by βn(x−k)×βn(y−l)…βn(z−m) in Eq. (6), the number of B-splines matching the number of variables. The derivative of the function R(x) may also easily be obtained by replacing the B-spline function by a combination of B-splines of the inferior degree but keeping the same coefficients c(k).

If the function R(x) is known on x values that are not integers—as is the case for the reflectance matrix R sim ias ,γ) —it must be made uniform. This step is performed by interpolating the function with nonuniform splines, resampling it on the integers, and finally interpolating it again with uniform B-splines. This process, fully described in Ref. 28, has been applied on R sim ias ,γ).

In terms of performance, this interpolation method is very efficient because the initial matrix R sim ias ,γ) is economically replaced by a matrix of coefficients c(ρi,k,l,m) of the same dimension, and any value of sim ias ,γ) is calculated by the summation of 3×4 B-splines weighted by their coefficients c(ρi,k,l,m). * The interpolated function sim ias ,γ) could then be properly used to fit any measured reflectance curve in order to determine its optical coefficients. To perform the fit, the Levenberg-Marquardt algorithm was chosen because of its efficiency and its robustness. Appropriately, this algorithm minimizes the quantity noted as χ2 and defined as

Eq. (8)

χ2=i=110(Rmes(ρi)R¯sim(ρi,μa,μs,γ)σi)2,
where σi is the standard deviation of each data point.

As required by the algorithm, the Jacobean of the sim ias ,γ) function can be calculated economically, as mentioned before, by the combination of square B-splines. For the lowest wavelength, a first set of parameters as ,γ) is determined using a lookup table based on the discrete set of simulations R sim ias ,γ). These values are then used as initial guesses for the Levenberg-Marquardt algorithm. For the next wavelength, the initial values are taken directly from the previous wavelengths (typically from an average of the four closest wavelengths). Nevertheless, we found the converged values to be insensitive to the initial guesses, which is related to the fact that only single minima were found in the χ2 function, as discussed in the next section. The determination of the optical coefficients for a whole spectrum, composed of 510 wavelengths (480 to 950 nm) takes about 50 s (on a 400-MHz Pentium personal computer).

3.

Results and Discussion

3.1.

Algorithm Robustness to Noise

Various assessments of the algorithm are reported in this section. First, in order for the algorithm to work properly, its solution must be unique. This is guaranteed if only a single minimum is present in the χ2 function. An inspection of the χ2 function over the whole range of values of μa, μs , and γ for which the algorithm is defined (i.e., μa from 0.003 to 10 mm−1, μs from 0.5 to 10 mm−1, and γ from 1.0 to 2.9) revealed only a single minimum for each set of parameters. Second, the robustness of the algorithm was studied by assessing the impact of noise on the accuracy of the optical coefficient determination. For this, noise (uniformly distributed) was added to simulated experimental reflectance curves. The relative errors between the result of the fit and the expected optical properties were estimated and are shown in Fig. 3 for noise of 2, 4, and 8 of the reflectance curve. Such noise figures are realistic for clinical tissue measurements, where tissue heterogeneity plays a significant role as noise in the reflectance curve. For each noise value, both the maximal and average errors of determination are plotted. Typically, the mean error on μs and γ is lower than 3, 5, and 10 for a noise of 2, 4, and 8, respectively. The maximum error on μa is lower than 6, 15, and 30 for a noise of 2, 4, and 8, respectively. The determination of μa is the most sensitive to noise because the sensitivity of the measurement to the absorption is not optimal owing to the relatively short path length at a short source–detector separation. Also, the error increases significantly at low absorption, especially for absorptions lower than 0.05 mm−1. Specifically, μa values smaller than 0.02 mm−1 cannot be reasonably estimated accurately with such a small source–detector separation. Therefore this method is best adapted to wavelengths in the visible (400 to 700 nm) or wavelengths higher than 900 nm, where μa is typically higher than 0.02 mm−1 in tissues. Overall, these results demonstrate that μa, μs , and γ can be determined simultaneously from the reflectance curve at short source–detector separations, which is a remarkable result.

Figure 3

Accuracy of the algorithm to determine μa, μs , and γ from noisy reflectance curves. The relative amplitudes of uniformly distributed noise added to the simulated reflectance curves were 2, 4, and 8, and both the maximal (thin lines) and average (bold lines) inaccuracies are plotted.

018303j.3.jpg

3.2.

Experiments With Phantoms

In order to complete the validation of the method, we performed a series of measurements on tissuelike phantoms made of suspended and dyed microspheres. Ge´le´bart et al.;29 have suggested that tissue phase functions could be advantageously simulated with a mixture of spheres with different diameters. Following this approach, we mixed polystyrene microspheres of five different diameters: 1, 0.5, 0.2, 0.1, and 0.05 μm. An identical concentration of 1.3 [g/liter] was set for each diameter, corresponding to a sphere density decreasing as d−3, where d is the sphere’s diameter. The scattering spectra μs (λ) and γ(λ) were calculated by Mie theory,30 assuming a refractive index of 1.60 for polystyrene and 1.33 for water.

The absorption properties of the solution were obtained by adding Nigrosin to the microsphere suspensions with a concentration of 59.09±1.65 μg/ml. The Nigrosin absorption spectrum μa(λ) was determined by spectrophotometry. Figure 4 shows a comparison of the predicted spectra of optical coefficients μa(λ), μs (λ), and γ(λ) and the corresponding spectra measured experimentally with the probe.

Figure 4

Comparison of the predicted spectra of optical coefficients (dashed curve) and those measured (solid curve) on the phantoms. The bold line is the average and the thin solid ones are the standard deviations.

018303j.4.jpg

A preliminary analysis allowed us to make two adjustments in the experimental procedure that eliminated two sources of systematic errors. First, we found that the use of the two closest fibers (ρ≈0.3) did not fit well with the simulations and was causing a systematic error. This effect was most likely due to a mismatch in higher moments of the phase function between our theoretical and experimental models. Our theoretical model assumes that only the first two moments of the phase function affect the reflectance. We showed in Ref. 17 that such an assumption is valid for an optical distance ρμs >0.5. Shorter distances can be used only if higher moments are matched between the simulation and experiment. This is indeed generally not the case between the average phase function of microsphere mixtures made of only five sizes of spheres and the modified Henyey-Greenstein phase function used in the simulation. Thus, these two fibers were removed from the analysis of the phantom measurements. However, the first two fibers have been kept when analyzing tissue measurements, since in this case the high-order moments of the tissue phase function are expected to match significantly better with the ones of the modified Henyey-Greenstein phase function. Moreover, adding these two fibers improved the robustness of the fit. Second, a systematic difference of 5 was found in the reflectance intensity between experiments and simulations, which was mainly due to the refractive index difference between the phantom (n=1.33) and the refractive index assumed in the simulations (n=1.40, corresponding to tissue).

Figure 4 shows the comparison between the predicted and experimental μa(λ), μs (λ), and γ(λ) values. The agreement on μs (λ) and γ(λ) is excellent, with an error typically less than 5. A good agreement is also found for μa(λ) with an error less than 10. These results confirmed the algorithm test described in the previous section.

3.3.

In Vivo Measurements of Human Gastrointestinal Tracts

In this section we present measurements performed in vivo on the human stomach, in the antrum (lower part of the stomach) and fundus (upper part) during gastroscopic examination. These results represent an example of the clinical feasibility of the method. Other studies are currently being performed on the colon and the uterus. One of the medical interests of the method lies in the possibility of instantly discriminating gastritis from normal cases in vivo during gastroscopy. This necessitates the determination of normal values for the mucosa in the antrum and in the fundus, which are presented here. In a companion paper,31 we will report further studies on the stomach, showing that this technique is sensitive to different stages of gastritis and to subtypes as well.32

A set of 35 patients, 21 females and 14 males, aged from 23 to 87 with an average age of 50 were examined. For each patient, four sites were usually selected to be measured, two in the antrum and two in the fundus. The optical probe was inserted into the working channel of the gastroscope at the end of the examination and gently applied on the tissue. The pressure could not be monitored during acquisition, but other controlled measurements on mouse skin showed that pressure variations did not significantly alter the results.

The endoscope illumination was turned off after the probe placement because of its high intensity. During this time, four to six measurements were taken in approximately 20 s. At the end, the endoscope illumination was turned on again and measurements were accepted and averaged if the probe stayed at the same location. The accept rate was approximately 80. Immediately after that, a biopsy was taken on the measurement site, which could be precisely located by a light mark on the epithelium. The data were analyzed after the clinical session.

Figure 5 shows the average optical properties, μa(λ), μs (λ), and γ(λ), of twenty-two measurements performed in the antrum and sixty measurements performed in the fundus. All sites were found normal after biopsy. First, as expected, μa(λ) clearly shows the presence of hemoglobin, characterized by absorption peaks at 541 and 576 nm. The absorption decreases significantly in the near-infrared range. The slight increase for wavelengths larger than 850 nm is due to water and hemoglobin. The concentration of total oxyhemoglobin (HbO 2) and deoxyhemoglobin (Hb) were obtained by a least-square fit of μa(λ). The total hemoglobin concentration (i.e., [HbO 2]+[Hb]) was 20.0±2.6 μM in the antrum, with a saturation (sO 2) around 48. In the fundus, a similar total hemoglobin concentration was found, 21.0±2.6 μM, whereas the saturation drops to 21. This surprising very low saturation could nevertheless be explained by a high concentration of CO 2 in the mucosa. Further investigations are needed to validate such results, since no other data on the oxygenation and saturation of hemoglobin in the fundus mucosa have been published yet. The μs (λ) spectra measured in both the fundus and antrum are monotonically decreasing, corresponding approximately to a power function, as expected for tissue.9 29 33 The scattering spectrum μs (λ) is found to be significantly lower in the antrum than in the fundus over nearly the entire spectrum.

Figure 5

Average (bold lines) and standard deviations (thin lines) of optical coefficients measured for normal antrum (solid curve) and normal fundus (dashed curve).

018303j.5.jpg

The γ(λ) spectra are approximately constant. This is consistent in the case of a large distribution of scatterer sizes, as found in tissue. As a matter of fact, similar findings were displayed in the experiment and simulation using the microsphere suspensions. Computations with Mie scatterers showed that the γ(λ) value is related to the scatterer size distribution.22 32 Therefore the absolute γ value might be important in differentiating tissue types.

For the antrum and the fundus, the γ values are 1.9±0.1 and 2.0±0.1, respectively. Assuming a Henyey-Greenstein phase function, these values would correspond to a g of 0.9 and higher. Such values are basically in agreement with g values commonly accepted for tissue. However, the γ values are found in some cases to be larger than 2, which could not be produced by a simple Henyey-Greenstein phase function. This seems to confirm previous results showing that a Henyey-Greenstein function alone does not approximate the tissue phase function adequately at large angles, but needs to be completed by either a second Henyey-Greenstein function with a negative g value or a cos 2 θ function.17 20

The differences between the fundus and antrum found in the scattering properties μs (λ) and γ(λ) can be correlated with the histological analysis of the biopsy samples. Figure 6 shows typical histological images of tissue resected after measurements. These images clearly show a difference in the thickness of the foveolar and glandular regions between the antrum and the fundus. These differences are likely to be the source of the differences observed in the μs (λ) and γ(λ) coefficients. We recently showed34 correlations between spectra of optical coefficients measured with this setup and histological analysis of a mouse skin alteration model.

Figure 6

Histological slices of antrum (a) and fundus (b). Whereas the layered structure is clearly visible in the fundus with the glandular region (G) clearly distinct from the foveolar (F), the structure of the epithelium is thinner and less contrasted in the antrum. Moreover, the mucosa is two times thicker in the fundus than in the antrum.

018303j.6.jpg

Globally, these results demonstrate that the scattering, μs (λ) and γ(λ), and absorption, μa(λ), properties of the tissue can be separated quantitatively. Only a fairly small coupling effect with the μa(λ) spectrum can be observed in both μs (λ) and γ(λ) spectra. Such a coupling is unavoidable considering that actual tissue is not homogeneous, which is the opposite of the theoretical model. As already mentioned, the layer structure is obvious in Fig. 6.

Two approaches are possible to further diminish such a coupling and improve accuracy. First, the μa(λ), μs (λ), and γ(λ) spectral shapes could be constrained. The μa(λ) spectrum can be assumed to result from known chromophores such as Hb, HbO 2, water and fat. As discussed earlier, in tissue, μs (λ) approximately follows a power law and γ(λ) is approximately a linear function. Incorporating such conditions in the inverse problem would significantly decrease coupling effects. Second, a layer model could be used to more closely model the tissue structure. Spatially resolved measurements seem appropriate to determine some of the properties of a two-layer geometry.19

4.

Conclusions

We have presented a method to determine the scattering and absorption spectra of small volumes of tissue in vivo. This method is based on the measurement of the spatially and spectrally resolved reflectance with a small optical probe (2.3-mm diameter), that can be used in the working channel of endoscopes. Three coefficients are determined simultaneously from the measurements, μa(λ), μs (λ), and the phase function parameter γ(λ). The computation of these parameters was obtained from a Levenberg-Marquardt fit of the measured reflectance to cubic B-spline interpolation of Monte Carlo simulations.

The robustness of the algorithm has been assessed using simulated reflectance curves, adding various degrees of noise. These results showed that μa(λ) is the coefficient most sensitive to noise, especially for low absorption. The accuracy of the method was also tested on tissue phantoms with known optical properties. A good agreement between the measurement and prediction was found. Typically, μa(λ), μs (λ), and γ(λ) can be determined with an error smaller than 10.

The ability to perform clinical measurements was demonstrated. Measurements on 35 patients were performed on the stomach epithelium through the working channel of a gastroscope. The average was found to be significantly different in the antrum and in the fundus. To our knowledge, these measurements are the first report of the simultaneous in vivo determination of broadband spectra of the absorption coefficient μa(λ), reduced scattering coefficient μs (λ), and phase function parameter γ(λ). We believe that this method has promise for providing important clinical information to guide biopsies efficiently.

Acknowledgments

We are grateful to M. A. Brundler for the histological analysis. We thank P. Marquet for his help in the project and A. Munoz, P. Thevenaz, and M. Unser for the spline interpolation. We thank the Common Program in Biomedical Engineering of the HUG-UNIL-EPFL-UNIGE-HCV, project MICRODIAG for their financial support. F. Bevilacqua gratefully acknowledges support from the Swiss National Science Foundation and from the Laser Microbeam and Medical Program of the U.S. National Institutes of Health (grant RR01192).

REFERENCES

1. 

N. Ramanujam , “Fluorescence spectroscopy of neoplastic and non-neoplastic tissues,” Neoplasia , 2 89 –117 (2000). Google Scholar

2. 

I. Bigio and J. R. Mourant , “Ultraviolet and visible spectroscopies for tissue diagnostics: Fluorescence spectroscopy and elastic-scattering spectroscopy,” Phys. Med. Biol. , 42 803 –814 (1997). Google Scholar

3. 

A. Mahadevan-Jansen and R. Richards-Kortum , “Raman spectroscopy for the detection of cancers and precancers,” J. Biomed. Opt. , 1 31 –70 (1996). Google Scholar

4. 

I. Georgakoudi , B. C. Jacobson , J. Van Dam , V. Backman , M. B. Wallace , M. G. Mu¨ller , Q. Zhang , K. Badizadegan , D. Sun , G. A. Thomas , L. T. Perelman , and M. S. Feld , “Fluorescence, reflectance, and light-scattering spectroscopy for evaluating dysplasia in patients with Barrett’s esophagus,” Gastroenterology, 120 1620 –1629 (2001). Google Scholar

5. 

L. T. Perelman , V. Backman , M. Wallace , G. Zonios , R. Manoharan , A. Nuserat , S. Shields , M. Seiler , C. Lima , T. Hamano , I. Itzkan , J. Van Dam , J. M. Crawford , and M. S. Feld , “Observation of periodic fine structure in reflectance from biological tissue: a new technique for measuring nuclear size distribution,” Phys. Rev. Lett. , 80 (3), 627 –630 (1998). Google Scholar

6. 

V. Backman , M. B. Wallace , L. T. Perelman , J. T. Arendt , R. Gurjar , M. G. Mu¨ller , Q. Zhang , G. Zonios , E. Kline , T. McGillican , S. Shapshay , T. Valdez , K. Badizadegan , J. M. Crawford , M. Fitzmaurice , S. Kabani , H. S. Levin , M. Seiler , R. R. Dasari , I. Itzkan , J. Van Dam , and M. S. Feld , “Detection of preinvasive cancer cells,” Nature (London) , 406 (6791), 35 –62 (2000). Google Scholar

7. 

F. Bevilacqua , D. Piguet , P. Marquet , J. D. Gross , B. J. Tromberg , and C. Depeursinge , “In vivo local determination of tissue optical properties: applications to human brain,” Appl. Opt. , 38 (22), 4939 –4950 (1999). Google Scholar

8. 

G. Zonios , L. T. Perelman , V. Backman , R. Manoharan , M. Fitzmaurice , J. Van Dam , and M. S. Feld , “Diffuse reflectance spectroscopy of human adenomatous colon polyps in vivo,” Appl. Opt. , 38 6628 –6637 (1999). Google Scholar

9. 

F. Bevilacqua , A. J. Berger , A. E. Cerussi , D. Jakubowski , and B. J. Tromberg , “Broadband absorption spectroscopy in turbid media by combining frequency-domain and steady-state methods,” Appl. Opt. , 39 6498 –6507 (2000). Google Scholar

10. 

A. E. Cerussi , A. J. Berger , F. Bevilacqua , N. Shah , D. Jakubowski , J. Butler , R. F. Holcombe , and B. J. Tromberg , “Sources of contrast for quantitative non-invasive optical spectroscopy of breast tissue physiology,” Acad. Radiol. , 8 211 –218 (2001). Google Scholar

11. 

J. B. Fishkin , O. Coquoz , and B. J. Tromberg , “Frequency-domain photon migration measurements of normal and malignant tissue optical properties in a human subject,” Appl. Opt. , 36 (1), 10 –20 (1997). Google Scholar

12. 

B. J. Tromberg , O. Coquoz , J. B. Fishkin , T. Pham , E. R. Anderson , J. Butler , M. Cahn , J. D. Gross , V. Venugopalan , and D. Pham , “Non-invasive measurements of breast tissue optical properties using frequency-domain photon migration,” Philos. Trans. R. Soc. London, Ser. B , 352 661 –668 (1997). Google Scholar

13. 

A. Ishimaru , “Diffusion of a pulse in densely distrubuted scatterers,” J. Opt. Soc. Am. , 68 1045 (1978). Google Scholar

14. 

R. A. J. Groenhuis , J. J. Ten Bosch , and H. A. Ferwerda , “Scattering and absorption of turbid materials determined from reflection measurements. 2. Measuring method and calibration,” Appl. Opt. , 22 2463 –2467 (1983). Google Scholar

15. 

R. Marchesini , N. Cascinelli , D. R. Ventrali et al.;, “In vivo spectrophotometric evaluation of neoplastic and non-neoplastic skin pigmented lesions. II: Discriminant analysis between nevus and melanoma,” Photochem. Photobiol. , 55 515 –522 (1992). Google Scholar

16. 

V. P. Wallace , D. C. Crawford , and J. C. Bamber , “Spectrophotometric assessment of pigmented skin lesions: methods and feature selection for evaluation of diagnostic performance,” Phys. Med. Biol. , 45 735 –751 (2000). Google Scholar

17. 

F. Bevilacqua and C. Depeursinge , “Monte Carlo study of diffuse reflectance at source–detector separations close to one transport mean free path,” J. Opt. Soc. Am. A , 16 (12), 2935 –2945 (1999). Google Scholar

18. 

J. R. Mourant , I. J. Bigio , D. A. Jack , T. M. Johnson , and H. D. Miller , “Measuring absorption coefficients in small volumes of highly scattering media: source–detector separations for which path lengths do not depend on scattering properties,” Appl. Opt. , 36 5655 –5761 (1997). Google Scholar

19. 

C. K. Hayakawa , J. Spanier , F. Bevilacqua , A. K. Dunn , J. S. You , B. J. Tromberg , and V. Venugopalan , “Perturbation Monte Carlo methods to solve inverse photon migration problems in heterogeneous tissues,” Opt. Lett. , 26 1335 –1337 (2001). Google Scholar

20. 

A. Kienle , F. K. Forster , and R. Hibst , “Influence of the phase function on determination of the optical properties of biological tissue by spatially resolved reflectance,” Opt. Lett. , 26 1571 –1573 (2001). Google Scholar

21. 

D. R. Wyman , M. S. Patterson , and B. C. Wilson , “Similarity relations for anisotropic scattering in Monte Carlo simulation of deeply penetrating neutral particles,” J. Comput. Phys. , 81 137 –150 (1989). Google Scholar

22. 

P. Thueler, F. Bevilacqua, and C. Depeursinge, “Relationship between the phase function parameter γ and the dimension of a fractal distribution of scatterers,” (submitted).

23. 

P. Marquet , F. Bevilacqua , C. Depeursinge , and E. B. de Haller , “Determination of the reduced scattering and absorption coefficient by a single charge-coupled-device array measurement. Part I: Comparison between experiments and simulations,” Opt. Eng. , 34 2055 –2063 (1995). Google Scholar

24. 

F. Bevilacqua, “Local optical characterization of biological tissues in vitro and in vivo,” PhD Thesis No. 1789, Swiss Federal Institute of Technology, Lausanne (1998).

25. 

L. G. Henyey and J. L. Greenstein , “Diffuse radiation of the galaxy,” Astrophys. J. , 93 70 –83 (1941). Google Scholar

26. 

M. Unser , “Splines: A perfect fit for signal and image processing,” IEEE Signal Process. Mag. , 16 (6), 22 –38 (1999). Google Scholar

27. 

I. J. Scho¨nberg , “Contribution to the problem of approximation of equidistant data by analytic functions,” Quart. Appl. Math. , 4 45 –99 (1946). Google Scholar

28. 

A. Mun˜oz Barrutia, T. Blu, and M. Unser, “Non-uniform to uniform grid conversion using least-squares splines,” Proceedings of the Tenth European Signal Processing Conference (EUSIPCO’00), Tampere, Finland, September 4–8, 2000, Vol. IV, pp. 1997–2000.

29. 

B. Ge´le´bart , E. Tinet , J. M. Tualle , and S. Avrillier , “Phase function simulation in tissue phantoms: A fractal approach,” Pure Appl. Opt. , 5 377 –388 (1996). Google Scholar

30. 

C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles, Wiley, New York (1983).

31. 

I. Charvet, P. Thueler, G. Ory, M.-A. Brundler, M. Saint-Ghislain, A. Hadengue, C. Depeursinge, P. Meda, and B. Vermeulen, “Noninvasive diagnosis of gastritis alterations,” (submitted).

32. 

P. Thueler, “Optical spectral probing for epithelial tissue characterization,” PhD, Dissertation No. 2241, Swiss Federal Institute of Technology, Lausanne (2001).

33. 

J. R. Mourant , J. P. Freyer , A. H. Hielscher , A. A. Eick , D. Shen , and T. M. Johnson , “Mechanisms of light scattering from biological cells relevant to non-invasive optical tissue diagnostics,” Appl. Opt. , 37 (16), 3586 –3593 (1998). Google Scholar

34. 

I. Charvet , P. Thueler , B. Vermeulen , M. Saint-Ghislain , C. Biton , J. Jacquet , F. Bevilacqua , C. Depeursinge , and P. Meda , “A new optical method for the noninvasive detection of minimal tissue alterations,” Phys. Med. Biol. , 47 2095 –2108 (2002). Google Scholar
©(2003) Society of Photo-Optical Instrumentation Engineers (SPIE)
Philippe Thueler, Igor Charvet, Frédéric P. Bevilacqua, Michel Saint Ghislain, G. Ory, Pierre Marquet, Paolo Meda, Ben Vermeulen, and Christian D. Depeursinge "In vivo endoscopic tissue diagnostics based on spectroscopic absorption, scattering, and phase function properties," Journal of Biomedical Optics 8(3), (1 July 2003). https://doi.org/10.1117/1.1578494
Published: 1 July 2003
Lens.org Logo
CITATIONS
Cited by 120 scholarly publications and 3 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Reflectivity

Tissue optics

Tissues

Monte Carlo methods

Absorption

Scattering

In vivo imaging

Back to Top