Open Access
29 December 2015 Validation of quantitative attenuation and backscattering coefficient measurements by optical coherence tomography in the concentration-dependent and multiple scattering regime
Author Affiliations +
Abstract
Optical coherence tomography (OCT) has the potential to quantitatively measure optical properties of tissue such as the attenuation coefficient and backscattering coefficient. However, to obtain reliable values for strong scattering tissues, accurate consideration of the effects of multiple scattering and the nonlinear relation between the scattering coefficient and scatterer concentration (concentration-dependent scattering) is required. We present a comprehensive model for the OCT signal in which we quantitatively account for both effects, as well as our system parameters (confocal point spread function and sensitivity roll-off). We verify our model with experimental data from controlled phantoms of monodisperse silica beads (scattering coefficients between 1 and 30  mm−1 and scattering anisotropy between 0.4 and 0.9). The optical properties of the phantoms are calculated using Mie theory combined with the Percus–Yevick structure factor to account for concentration-dependent scattering. We demonstrate excellent agreement between the OCT attenuation and backscattering coefficient predicted by our model and experimentally derived values. We conclude that this model enables us to accurately model OCT-derived parameters (i.e., attenuation and backscattering coefficients) in the concentration-dependent and multiple scattering regime for spherical monodisperse samples.

1.

Introduction

Optical coherence tomography (OCT) is a light scattering-based imaging modality which allows micrometer-scale resolution volumetric imaging and characterization of tissue morphology. Several (preclinical) studies show a correlation between the pathological state of the tissue and an OCT-derived attenuation coefficient, μOCT, which quantifies the decay of the OCT signal in depth.14 Despite promising results, the underlying cause of this correlation is not theoretically explained. The hypothesis of a higher attenuation for cancerous versus noncancerous5 (relying on assumed structural differences between cancerous and noncancerous) does not seem to hold for all tissue types.68 For reliable clinical use, the validation of quantitative measurement of OCT-derived attenuation coefficient (μOCT) and amplitude [determined by the backscattering coefficient within the numerical aperture (NA) of the optics, μB,NA] is paramount, e.g., to determine cut-off values for various tissue types. To obtain tissue optical properties μOCT and μB,NA, system parameters [confocal point spread function (PSF) and sensitivity roll-off] should be calibrated for. Moreover, within the scattering regime of biological tissue effects as the concentration-dependent scattering and multiple scattering should be taken into account in order to understand the measured attenuation. Due to concentration-dependent scattering, the “bulk” scattering properties at high volume fractions cannot be calculated from super positioned scattered intensities of individual particles. Rather, scattered fields need to be added and their interference should be accounted for by incorporating a “structure factor” in the calculation of the scattering properties. The structure factor for discrete random media (DRM, i.e., suspensions of monodisperse silica beads) considered in this paper can be calculated using the Percus–Yevick equation. 9,10 Using Mie solutions to the Maxwell equations, one can calculate the scattering properties of a single spherical particle. At times, the Mie solutions for a single particle are linearly extrapolated to obtain the scattering properties of an ensemble of particles. For very dilute solutions, μS scales linearly with volume fraction. However, at higher concentrations, dependent scattering leads to lower values for the scattering coefficient (μS). At higher scattering coefficients and scattering anisotropy, the contribution of multiple scattered light to the OCT signal is increased, which leads to a decreased measured μOCT compared with the μS. The separation of the effects of concentration dependent and multiple scattering is challenging, as both lower the value of μOCT compared with predictions based on the linear extrapolation of Mie calculated scattering cross sections.

The aim of this study is to quantitatively derive and validate the values of μOCT and μB,NA from the OCT signal by correcting for system parameters and taking into account the effects of multiple and concentration-dependent scattering.

We first describe a generic model of the OCT signal in Sec. 2 in terms of system parameters (confocal PSF and sensitivity roll-off) and pursued properties μOCT and μB,NA. We discuss the calculation of the scattering properties of the phantom materials taking into account concentration-dependent scattering. Next, we summarize the extended Huygens–Fresnel (EHF) model11,12 to account for multiple scattering in OCT signals and combine it with the calculations of the DRM scattering properties to arrive at theoretical predictions for μOCT and μB,NA. Section 3 describes a practical way to account for system-induced attenuation, sample preparation, and data analysis. We propose to use DRM as validation phantoms, i.e., samples of spherical scatterers with known size distribution, for which the OCT-derived μOCT and μB,NA can be calculated. The values of μOCT and μB,NA of the DRM should match the range of values found in tissue as closely as possible (roughly 1 to 30  mm1 in our experiments). Therefore, samples of different size distributions and volume fractions need to be prepared. Finally, in Secs. 4 and 5, respectively, we present our results showing excellent agreement between experimental and predicted μOCT and μB,NA and discuss our results and their clinical implications.

2.

Theory

2.1.

Generic Model of the Optical Coherence Tomography Signal

The most common approach to date to quantify optical properties from OCT images is by the means of nonlinear least squares (NLLS) curve-fitting of an OCT signal model to the measured data. If the signal model is parameterized properly, such that no mutual dependence exists between the fit parameters, the covariance matrix obtained from the NLLS fit procedure yields uncertainty estimates on the fitted parameters such as standard deviation of multiple fits for which the region of interest (ROI) is changed by 10%. We adopt the model by Xu et al.,13 which assumes only single backscattered light contributes to the OCT signal. Light to and from the backscatter location at depth z is attenuated due to scattering and absorption. This is accounted for by the attenuation coefficient μt, such that μtdz is the probability of interaction in path length dz and 1μtdz is the transmission probability. The intensity incident at the backscattering site is thus reduced by a factor (1μtdz1)×(1μtdz2)××(1μtdzN) compared with the intensity impinged on the tissue. If the sample is homogeneous and the product μtdz is small, we write limn[1μt(z/n)]n=eμtz. Several preclinical and phantom studies demonstrated that, for most tissues, the OCT signal amplitude versus depth z is best described by this single exponential decay curve, combined with descriptions for the confocal PSF14,15 and sensitivity roll-off in depth for frequency domain OCT systems. The resulting expression for the mean-squared OCT depth signal after noise subtraction for zz0 is16,17

Eq. (1)

A2(z)=α·T(zzf)·H(z)·μB,NAexp[2μOCT(zz0)],
where zf and z0 are the depth positions of the focus and the tissue boundary with respect to zero-delay, respectively, and α is a system-dependent factor (which can be calibrated) that includes the power incident on the sample, the quantum efficiency of the detector, and the source coherence length.18,19 The function T(zzf) is the confocal PSF:

Eq. (2)

T(zzf)=1(zzf2nZR0)2+1,
where ZR0=πw02/λ0 is the Rayleigh length (half the depth-of-focus) of the sample arm optics measured in air, and n is the refractive index of the medium, λ0 is the center wavelength of the OCT source, and w0 is the beam waist at the focus (defined as the beam radius where the intensity drops to a factor 1/e2 of its maximum). The factor 2 accounts for the apparent doubling of the Rayleigh length for diffuse reflection.15,20 The PSF can be calibrated, for example, by measuring the depth-of-focus using a mirror or by knife-edge measurements of w0.

The expression for the sensitivity roll-off in depth H(z) is given in terms of sampling and optical resolution (in wavenumbers) in the following equation:21

Eq. (3)

H(z)=[sin(0.5Δksamplingz)0.5Δksamplingz]2exp(Δkoptical2z28ln2).
The sampling resolution is calculated as (kmaxkmin)/NSAMP, where NSAMP is the number of pixels/samples taken per spectrum and kmax and kmin are the maximum and minimum wavenumbers supported by the spectrometer or swept source. For spectrometer-OCT systems, Δkoptical is determined by the dispersion line width of the spectrometer. For swept-source OCT systems, it is the instantaneous line width of the source. Both systems’ resolutions, the first- and second-term in Eq. (3), can be calibrated separately. In Sec. 3.1, we present a practical approach to calibrate T(zzf) and H(z) together from an OCT measurement of very weakly scattering media.

The last term of Eq. (1), μB,NAexp[2μOCT(zz0)], contains the optical properties of the sample, but is influenced by system properties as well. The backscattering coefficient within the NA (μB,NA) clearly depends on the confocal properties of the OCT system. Throughout this paper, we assume μB,NA only differs per sample, but is constant in depth for homogeneous phantoms for each individual sample. The decay constant μOCT depends on the scattering coefficient μS and absorption coefficient μA. For weakly scattering samples, with negligible contributions from multiple scattered light, μOCT=μS+μA. The experimentally obtained value may be influenced by the contribution of multiple scattering to the OCT signal, determined by the confocal properties of the system and on the angular scattering characteristics of the sample. Both a lower NA and samples with highly forward directed scattering combined with a high scattering coefficient, corresponding to values of the scattering anisotropy g close to unity, lead to larger contributions of multiple scattered light in the detected signals.12 Thus, for these situations, detection of multiple scattered light leads to a decreased decay of the OCT signal in depth, i.e., μOCT<μS+μA. In that case, the following equation is applicable:

Eq. (4)

μOCT=fNA,g(μS)+μA,
where the mapping function fNA,g depends on confocal system properties and sample angular scattering properties. A similar equation was proposed by Jacques et al.22 based on Monte Carlo simulations. The calculation of μS and μB,NA is detailed in Sec. 2.2. The multiple scattering mapping function fNA,g is discussed in Sec. 2.3. At the clinically applied wavelength of 1300 nm, scattering dominates over absorption (μSμA).

2.2.

Optical Properties of the Phantom

Consider an isotropic DRM consisting of identical hard spherical particles with known number density ρ [# particles/m3]. The particles cannot overlap in volume. For the individual particles, Mie theory23 yields the differential scattering cross section σS(θ) (m2sr1), where θ is the scattering angle. The total scattering cross section is found by integrating over the angular coordinates while taking into account the phase differences in the scattered fields from the different particles (which are determined by their distinct positions r)

Eq. (5)

σS=2π0πσS(θ)u=1Nv=1Neiq(θ)·rueiq(θ)·rvsinθdθ.
The term between chevrons defines the structure factor S(q), where q is the scattering vector q=koutkin, |q|=q=2ksin(θ/2). The ensemble average runs over a volume containing N particles with position ru (and rv) and thus depends on volume fraction fv. The structure factor describes the accumulative effect of interference between the fields scattered from the N contributing particles under the assumption that each of the particles is illuminated by the same incident field (first Born approximation). If no interference takes place, the terms where uv vanish and S=1. In general, interference takes place and the structure factor depends on number density ρ, or volume fraction fv=ρVP (“concentration-dependent scattering”), where VP=πD3/6 is the volume of a spherical particle with diameter D. For the DRM under consideration, the structure factor is calculated as the Fourier transform of the Percus–Yevick equation for the pair correlation function (PCF, which is interpreted as a normalized distribution of distances r between particle pairs).9,24 For dilute suspensions of hard spheres of diameter D, PCF(fV,r)=0 when r<D and unity otherwise. For higher volume fractions, PCF(fV,r) shows a damped oscillatory behavior, with increased probability of finding pair separations at multiples of D and decreased probability in between. In the limit r, PCF(fV,r) goes to unity. The structure factor that can be described as S(fV,θ) is approximately constant at unity for low volume fractions. Equation (5) shows that nonunity S(fV,θ) at higher values of fv causes an angular redistribution of scattered light by the medium compared with that of a single particle (it serves as a weighting function for the differential scattering cross section). Figure 1 shows a graph of the structure factor at volume fractions 0.001, 0.2, and 0.5 as function of Dq=(kD)*2sin(θ/2). The curve was calculated using the algorithm described in Ref. 24, as the numerical Fourier transform of the Percus–Yevick PCF.

Fig. 1

A graph of the structure factor at volume fractions fv=0.001, 0.2, and 0.5 as a function of Dq=(kD)*2sin(θ/2). The curve was calculated using the algorithm described by Tsang et al.,24 as the numerical Fourier transform of the Percus–Yevick pair correlation function (PCF).

JBO_20_12_121314_f001.png

The scattering coefficient (the product of particle density and scattering cross section, μS=ρσS) is then calculated as

Eq. (6)

μS=fVVp·  2π0πσS(θ)S(fV,θ)sinθdθ.
The backscattering coefficient within the detection NA is calculated similarly by adjusting the integration boundaries to [πarcsin(NA) to π]:

Eq. (7)

μB,NA=fVVp·2ππarcsin(NA)πσS(θ)S(fV,θ)sinθdθ.
Consequently, both μOCT and μB,NA depend both on the differential scattering cross section of the particles comprising the DRM (calculated by Mie theory) and on the volume fraction of the particles (accounted for by the Percus–Yevick structure factor).

2.3.

Multiple Scattering Model

For a high value of the sample’s scattering coefficient (μS) and a scattering anisotropy (g) close to 1, an increased contribution of multiple scattered light to the OCT signal will lead to a reduced measured μOCT. The most comprehensive model to date describing this effect is based on the EHF principle introduced by Schmitt and Knüttel11 and Thrane et al.12 Here, in the paraxial approximation, the mean squared OCT signal—excluding the effects on attenuation of the confocal PSF and sensitivity roll-off—is given by a contribution of three terms: the single-backscattered field [as in Eq. (1)], the multiple (forward) scattered field, and a coherent cross term between these two fields as

Eq. (8)

A2(z)exp(2μSz)+2exp(μSz)[1exp(μSz)]1+wS2(z)wH2(z)+[1exp(μSz)]2wH2(z)wS2(z).
To maintain readability of the equations, we substituted Δz=zz0 in the equations in this paragraph, so that Δz should be interpreted as the depth coordinate in tissue measured from the sample boundary at Z0.

Here, wH(z) is the local beam waist in the absence of forward scattering (e.g., of the single-backscattered beam):

Eq. (9)

wH2(Δz)=w02[(zzf2nZRO)2+1],
where w0 is the beam waist at the focus, measured in air as before (Sec. 2.1). The factor 2 in the denominator of the term between square brackets [] accounts for the apparent doubling of the Rayleigh length for diffuse reflection.15,20 The expression for the local beam waist in the presence of multiple forward scattering wS(z) can be found in Ref. 12 as

Eq. (10)

wS2(Δz)=wH2(Δz)+13(μSΔz)θrms2(Δz/n).
The broadening of the beam thus depends on the average number of scattering events μSz and on the angular distribution of the scattered light. In the EHF model, this distribution is described through the root-mean-square (rms) scattering angle θrms (a small value for small scattering angles; compared with the scattering anisotropy g which yields values close to unity for small scattering angles). The physical interpretation is that for highly forward directed scattering (small θrms, g close to unity), the shape of the forwarded scattered beam is only marginally altered on subsequent scattering events. Thus, it remains possible for this beam to be coupled back in to the OCT system, leading a large contribution of multiple forward scattered light. The contribution of multiple scattered light will show as an overall signal decay that is slower than the single backscattered light, e.g., μOCT<μS. If the scattering is more isotropic (larger θrms, g1), each subsequent scattering event considerably broadens the forward scattered beam. This reduces the coupling efficiency of that beam and subsequently the contribution of multiple scattered light will be less; leading to a measured μOCT much closer to μS.

3.

Methods

The OCT data were recorded with the Santec IVS 2000 swept-source OCT system, at a center wavelength of λ0=1309  nm, 140-nm sweep range, with a sweep rate of 50 kHz. Optical and sampling resolutions were approximately equal at ΔkOPTICALΔkSAMPLING0.0003  μm1. From a knife-edge measurement, the diameter of the collimated beam was measured at 2.54±0.01  mm. The NA of the system was 0.02 with a focal length of 65 mm, and the Rayleigh length measured in air was 960  μm. The mode field diameter of the fiber was 2.9  μm. The position of the focus was coaligned with the position of the reference mirror (zero-delay). The combined sensitivity roll-off, expressed as the depth where the signal drops to 3  dB of its initial value, was 3.8  mm. The measured axial and lateral resolutions were 12 and 25.5  μm in air, respectively. A cross-section image (B-scan) was built from 1000 adjacent A-lines. The samples were measured in a 1-mm quartz cuvette. The inner boundary of the cuvette was placed at z0=650  μm ±34  μm from zero delay. The cuvette was placed under an angle of 10  deg to avoid specular reflection from the boundary.

3.1.

System Parameter Calibration

All recorded data were corrected for system-induced attenuation, i.e., caused by the confocal PSF and sensitivity roll-off as described in Sec. 2. Due to the low NA and high spectral resolution of our system, both are slowly decaying functions of depth and can in practice be approximated by a single exponential decay with attenuation constant μCAL. Thus, Eq. (1) can be modified to

Eq. (11)

A2(z)α·exp(2μCALz)·μB,NAexp[2μOCT(zz0)].
The μCAL can straightforwardly be obtained from an OCT measurement of a very weakly scattering sample. We used a 1000-fold dilution of Intralipid 20% in water, for which the scattering coefficient at our center wavelength is estimated to be μS,IL0.01  mm1. The fitted signal attenuation corresponding to μCAL was found to be 1.1  mm1 for this system. Note that the absorption of water (from literature: 0.135  mm1)25 is also accounted for in this manner.

3.2.

Sample Preparation

Monodisperse silica beads (Kisker Biotech, Steinfurt, Germany) with measured mean diameters of 0.47, 0.70, 0.91, and 1.60  μm were suspended in distilled water for a monodisperse concentration series. For an accurate measurement of the size distribution, the silica beads were imaged using transmission electron microscopy (TEM) (Philips CM-10) (Fig. 2). The TEM images were acquired and analyzed following the protocol described in previous work.10 The obtained values of the mean diameter and standard deviation of the bead sizes are shown in Table 1. To prevent aggregation of the beads in suspension, 0.03 mM of sodium dodecyl sulfate was added to all samples.26 The 0.47- and 0.70-μm sized beads were diluted to 10 samples with volume fractions ranging from 0.02 to 0.2. The 0.91- and 1.60-μm sized beads were diluted to 5 samples with volume fractions from 0.02 to 0.1. Suspensions with a higher volume fraction of silica were not studied due to aggregation of the beads. All samples were 4 to 6 times alternatingly vortexed and sonicated for 15 min.

Table 1

Mean and standard deviation of the size distribution of monodisperse silica beads measured with transmission electron microscopy.

Monodisperse silica beads (Kisker Biotech)Mean diameter (μm)Standard deviation (μm)
PSI 0.50.470.03
PSI 0.80.700.03
PSI 1.00.910.02
PSI 1.51.600.03

Fig. 2

Histogram of the size distribution of monodisperse silica beads (PSI 1.0) measured with transmission electron microscopy.

JBO_20_12_121314_f002.png

3.3.

Optical Coherence Tomography Data Analysis

We used a custom-written code (LABVIEW 2013, National Instruments) for the analysis of the acquired OCT data. The attenuation (μOCT) was obtained from the data by NLLS fit of a single exponential decay model y=AMP*exp(μz)+y0 to the OCT signal in depth. The data were first averaged over at least 1000 A-lines. The selection of the fit was done manually where the start region was chosen to approximately be 50  μm behind the cuvette boundary. The end point of the fit was chosen such that within the ROI, upon visual inspection, the signal appeared to follow a single exponential decay curve. An offset added to the fit model was fixed at the average noise level directly at the backside of the cuvette. The amplitude and decay constant μ were independently varied parameters. The fit was repeated 100 times, during which the boundaries of the ROI were varied within 10% of their operator-selected values. The mean value of the 100 measurements was taken as the final μ value, with uncertainty estimated as the standard deviation of the fits. The obtained attenuation μ was corrected for system parameters, as described in Sec. 3.1, yielding μOCT=μμCAL.

The backscattering coefficient, μB,NA, was determined from the amplitude of the backscattered OCT signal18,19 in a superficial ROI starting at zROI=0.03  mm below the glass–sample interface. The ROI is chosen as such to exclude contribution of multiple scattered light. The region extends only 0.026 mm (4 pixels) in depth to minimize the effects of attenuation, which was verified by ensuring that the OCT amplitude speckle contrast in the ROI was 0.52,27 corresponding to fully developed speckle, which can only be achieved if no appreciable attenuation in depth takes place in the ROI.28 By Eq. (11), in this ROI, we have A2(zROI)=αexp(μCALzROI)μB,NA which features only system-dependent parameters. This allows us to define a scaling factor between measured amplitude and backscatter coefficient μB,NA=sfØA2(z0), where the scaling factor is determined separately for each particle diameter using the highest volume fraction fvMAX: sfØ=[μB,NA(MIEPY)/A2(z0)]Ø,fvMAX.

3.4.

Multiple Scattering

A practical approach for assessing the influence of multiple scattering for confocal reflectance microscopy and OCT was proposed by Jacques et al.22 Based on Monte Carlo simulation, a reference grid is constructed, allowing for the mapping of scattering the coefficient and scattering anisotropy to the amplitude and decay rate and vice versa.22,29 We adopt this approach, but rather than relying on Monte Carlo simulations, we use the analytical EHF model proposed in Sec. 2.3 to simulate the OCT signal at a range of values of the scattering coefficient and anisotropy and, subsequently, fit the single exponential decay to the simulated signal to retrieve the corresponding predicted value of μOCT. This procedure is illustrated in Fig. 3.

Fig. 3

Simulated extended Huygens–Fresnel (EHF) model for optical coherence tomography (OCT) (green dashed line), experimental OCT curve (black solid line) fitted with a single exponential decay within the ROI for samples of 0.1 volume fraction: (a) 0.47-μm beads and (b) 1.60-μm beads. The simulated single scattering (Mie+PY) curve (blue dotted line) is plotted for reference. The exponential decay of the fits on the experimental curves is (a) μexp=5.0  mm1 and (b) μexp=10.8  mm1. After correction of system-induced attenuation (μCAL=1.1  mm1), the decay rates match closely with the decay of the simulated curves: (a) μsim=3.6  mm1 and (b) μsim=9.7  mm1.

JBO_20_12_121314_f003.png

3.5.

Calculations

To calculate predictions for μOCT and μB,NA, we have written a Labview code (Labview 2013, National Instruments). We implemented Mie calculations from Ref. 30 and calculated the differential scattering cross section for each particle by integrating over the size distribution measured with TEM (Table 1 and Fig. 2). The refractive index of the silica beads was fixed at 1.425 (within the range provided by the manufacturer) and converted to 1309 nm using the dispersion curve from Malitson31 (n=1.419 to 1.449). The refractive index of the medium was assumed to be equal to that of water (1.324).32 The Percus–Yevick structure factor S(θ) was calculated in Ref. 24, using volume fraction and mean particle diameter (Table 1) as input parameters. By combining the Mie-calculated σS(θ) with this structure factor, we calculate μS [Eq. (6)], μB,NA [Eq. (7)], and rms scattering angle θrms for each sample (Mie+PY). The EHF calculations were based from Ref. 12 and took the system properties center wavelength, focal length, and Rayleigh length as input parameters, as well as the medium refractive index and the μS and θrms (MIE+PY) for each sample. To quantify the accuracy of the prediction, for each model, the standard error of the estimate (s) was calculated as (μOCT,EXPERIMENTALμOCT,PREDICTION)2/N, where N is the number of data points (volume fractions). The lower s corresponds to a better predictive value. An uncertainty estimate on s was calculated using a bootstrapping method,33 we calculate N values of s from an (N1) sized dataset by leaving out each value of the original dataset once. The standard deviation of the resulting sequence of s-values is taken as an uncertainty estimate (Table 2). A similar analysis was performed for the backscattering coefficient (Table 3).

Table 2

The standard error of estimates is calculated to evaluate the match between the experimental data and the theoretical predictions for μOCT by the three models (Mie, Mie–PY, and Mie–PY–EHF). Uncertainties are estimated using a bootstrapping method.

ØMIEMIE–PYMIE–PY–EHF
0.472.7 (±0.2)0.29 (±0.02)0.24 (±0.02)
0.704.7 (±042)0.41 (±0.02)0.51 (±0.02)
0.912.7 (±0.5)0.7 (±0.1)0.35 (±0.06)
1.6011 (±2)7 (±1)0.36 (±0.06)

Table 3

The standard error of estimates is calculated to evaluate the match between the experimental data and the theoretical predictions for μB,NA by the two models (Mie and Mie–PY). Uncertainties are estimated using a bootstrapping method.

ØMIEMIE–PY
0.472.1(±0.2)×1055.9(±0.6)×106
0.706.3(±0.3)×1072.7(±0.1)×107
0.913.2(±0.1)×1063.1(±0.1)×106
1.604.4(±0.3)×1074.4(±0.3)×107

4.

Results

The experimental data points and predicted calculated curves of μOCT as a function of volume fraction of silica beads are plotted in Fig. 4 (ranging from 1 to 10  mm1). In every graph, the calculated curves of μOCT using Mie theory (single, concentration-independent scattering), Mie+PY (single, concentration-dependent scattering), and Mie+PY+EHF (concentration dependent and multiple scattering) models are plotted for reference. To quantitatively evaluate the match between experimental data and the theoretical predictions for μOCT, the standard error of the estimate is calculated (Table 2). For the beads with a mean diameter of 0.47 and 0.70  μm, the experimental values of μOCT are well described by the calculated single scattering curve, in which concentration-dependent scattering properties of the phantoms are accounted for by combining Mie theory with the Percus–Yevick radial distribution (Mie+PY). No contribution of multiple scattered light is observed for the smallest beads (0.47  μm). For the 0.70-μm beads, multiple scattering contributes only at high volume fractions for which μS is high. For the beads with a diameter of 0.91  μm and corresponding intermediate forward scatter (scattering anisotropy g0.8), we see a significant contribution of multiple scattering for the higher volume fractions of scatterers. For the beads with a mean diameter of 1.60  μm and corresponding larger forward scatter (high scattering anisotropy g0.9), we see a significant contribution of multiple scattered light. By combining Mie+PY with the EHF model, we are able to accurately describe the contribution of multiple scattering as the experimental data points and calculated curves (Mie+PY+EHF) are in good agreement. To illustrate the influence of concentration-dependent scattering on the scattering anisotropy, we plotted the scattering anisotropy (g) of the samples, as calculated with Mie theory and Mie+PY as a function of volume fraction in Fig. 5. This figure shows the influence of the concentration of scatters on the average cosine of the scattering angle. The weighting by the concentration-dependent structure factor leads to relatively more backscatter and thus a decrease in scattering anisotropy with increasing volume fraction. This effect is more pronounced for smaller sized beads. Figure 6 shows the experimental data points and calculated curves of the backscattering coefficient, μB,NA, as a function of volume fraction calculated with Mie+PY. Per bead size, a scaling factor is used to scale experimental amplitudes to μB,NA (Sec. 3.3), which takes into account the system-dependent parameters. The curves calculated with only Mie scattering (no concentration-dependent scattering) are plotted as a reference. The values of the standard error of the estimate are shown in Table 3. The experimental data points and calculated Mie+PY curves are in good agreement.

Fig. 4

The volume-fraction dependence of the OCT attenuation coefficient, μOCT, of four different sized silica beads (mean diameter: 0.47, 0.70, 0.91, and 1.60  μm). The black dots depict the experimental data point and the error bars depict the uncertainty estimated as the standard deviation of 100 fits. The plotted curves show the calculated data. The blue dotted lines are the Mie calculations, the red dashed lines are the Mie calculations combined with Percus–Yevick formalism (Mie+PY) to include concentration-dependent scattering, and the green lines are the calculations combining Mie+PY with the EHF model to account for multiple scattering.

JBO_20_12_121314_f004.png

Fig. 5

The calculated scattering anisotropy (g) of the silica beads as a function of their volume fractions. The blue dotted lines are the single, concentration-independent scattering, Mie, calculations. The red dashed lines are the single, concentration-dependent scattering calculation, where Mie theory is combined with the Percus–Yevick PCF (Mie+PY). A change in scattering anisotropy as a function of volume fraction is predicted when comparing Mie+PY (single, concentration-dependent scattering) to Mie (single, concentration-independent scattering) calculations.

JBO_20_12_121314_f005.png

Fig. 6

The volume-fraction dependence of the backscattering coefficient (μB,NA). The black dots show the experimental data points and the error bars depict the standard deviation from five independent measurements. The blue dotted lines are the single, independent scattering, Mie calculations. The red dashed lines are the single, concentration-dependent scattering, Mie+PY calculations. Here, we do not include any multiple scattering, as the experimental data are taken in a thin superficial layer to minimize the contribution of multiple scattered light. A scaling factor (sf) [sfØ=[μB,NA  (MIEPY)/A2(z0)]Ø,fvMAX] is used to obtain μB,NA from A. Per particle size, we used the measurement at highest volume fraction to calculate the scaling factor.

JBO_20_12_121314_f006.png

5.

Discussion

We describe a method that validates measurement of OCT-derived attenuation coefficient μOCT which is correlated to tissue disease state in various clinical studies, ranging from ophthalmology,34 cardiology,2 dermatology,6 to urology.35,36 Paramount to clinical relevance is understanding how μOCT depends on properties of the OCT system itself, as well as the influence of multiple scattering by tissue properties: only then can reproducible results be obtained either with the same OCT system or by different systems. Therefore, we described the influence of the confocal PSF and sensitivity roll-off in depth [Eq. (1)] and introduced a practical method to calibrate their influence (Sec. 3.1) for commonly applied, low-NA clinical systems. Furthermore, we describe DRM for calibration of μOCT with tissue-mimicking scattering properties (μOCT between 1 to 10  mm1 and g between 0.4 and 0.9).

5.1.

Concentration-Dependent Scattering Versus Multiple Scattering

For higher scattering coefficients and low values of θrms (g close to 1), multiple forward scattered light has a large contribution to the OCT signal. Corresponding DRM for calibration purposes requires high volume fraction of the scatterers. However, for high volume fractions, the bulk optical properties of the DRM cannot be extrapolated from the optical properties of a single scatterer simply through scaling by volume fraction. Rather than adding scattered intensities, scattered fields (amplitude and phase) from the individual particles in the DRM should be added (taking into account their interference) to calculate the optical properties. This situation is commonly referred to as “concentration-dependent scattering” as the bulk optical properties now depend nonlinearly on volume fraction. The effect of concentration-dependent scattering is that the scattering coefficient μS scales as a lower-than-linear rate with volume fraction [μS<(fV/VP)σSCAT]; Sec. 2.2), leading to a lower μOCT compared with the “concentration-independent scattering” case. Increased detection of multiple-forward scattering leads to further reduction of μOCT. Therefore, care must be taken to separate both effects in data analysis. Here, we predict μOCT by first calculating μS and θrms for our DRM samples by using Mie theory (scattering properties of individual spheres) and the Percus–Yevick structure factor (concentration-dependent scattering). These values are then input to the EHF equation that describes the OCT signal in the multiple scattering regime. μOCT is derived by fitting a single exponent to the simulated data. The measured values for μOCT correspond closely to these predictions (Fig. 4). For samples with a low amount of forward scattering (low g-value, Fig. 5), no contribution of multiple scattered light is observed. Although the EHF model is limited by the paraxial approximation, its prediction does not deviate from the concentration-dependent scattering curve and is in good agreement with the experimental data, which suggests a broad applicability of the EHF model. For the case of weakly scattering media with moderate scattering anisotropies, the simpler Mie–Percus–Yevick model may be used for accurate prediction of μOCT. The backscattering coefficient, μB,NA, is best predicted by Mie–Percus–Yevick theory for all particle sizes. Even though no statistical difference is found between the prediction based on Mie theory or Mie–Percus–Yevick theory, only the latter model yields accurate predictions for both parameters. The physical explanation is that, as the volume fraction increases, the increment in scattering coefficient decreases [e.g., the integral part of Eq. (6) is a decreasing function of volume fraction] while at the same time increasingly more light is scattered to the backward direction [e.g., the integral part of Eq. (7) remains almost constant]. Effectively, the backscattering coefficient scales linearly with volume fraction in both models.

We conclude that our approach correctly separates the contributions of concentration dependent and multiple scattering, substantiated by the correct prediction of the backscattering coefficient μB,NA from the same calculations (Fig. 6). We measure μB,NA at depths in the order of 1 mean free path to minimize attenuation of the signal within the ROI (minimum mean free path is μS1=30  μm for Ø=1.60-μm particles, volume fraction=0.1). Moreover, we have recently demonstrated correct prediction of scattering coefficients μS by Mie–Percus–Yevick theory, using transmission-based low-coherence interferometry.10 In that experiment, the contribution of small-angle multiple forward scattering is greatly reduced by probing only the ballistic light transmitted through the sample.

The weighting of the differential scattering cross section with the structure factor (concentration-dependent scattering) also leads to more light being scattered into the backward direction. Ultimately, these two contrary effects result in an almost linear increase of the backscattering coefficient μB,NA with volume fraction. Clearly, this result is not general but depends on the exact scattering phase function of the used sample/tissue and the confocal configuration of the OCT system as predicted by the presented theory.

To eliminate contributions of multiple scattering, Xu et al.13 propose to confine the ROI of the fit between Imax and Imax/e. As can be seen from the EHF equations [Eqs. (8)–(10)], the multiple scattering starts to contribute directly from z=z0, or Δz=(zz0)=0. The nonscattered beam attenuates with exp(μsΔz) and the (small angle, forward) scattered beam grows in intensity with [1exp(μsΔz)]. This latter beam, however, broadens upon propagation [Eq. (10)], so that coupling back into the single-mode fiber of the OCT system will become more difficult with increasing Δz. The contribution of multiple scattering, therefore, is a balance between these two factors. At small values of Δz, the power in the scattered beam is small but, since the spread (increased divergence of the scattered beam) is small, it is more effectively coupled back. In contrast, for larger values of Δz, the power in the scattered beam is increased, but due to the increased spread, the coupling efficiency is worse. At very large depths, the highly increased magnitude of the scattered beam starts to dominate the single-backscattered contribution.

Two solutions proposed in literature are commonly adopted to take into account the multiple scattered light contribution to the OCT signal: the EHF model described here and a Monte Carlo derived model introduced by Jacques et al.22 We have adopted the EHF model in this work, but adjusted for the confocal PSF for a diffuse reflection, where we assume that the lateral phase of the probing beam is lost upon reflection. In practice, this is accounted for by setting the Rayleigh length in the medium to 2 times the value measured for specular reflection (Sec. 2.3).15,20 In the other approach, an empirical equation is derived by Monte Carlo simulations to approximate the contribution of multiple scattering.22,29 Levitz et al.29 use this approach and translate μOCT and amplitude to μS and g based on calibration of the peak signal of a known interface (scattering phantom with known optical properties). The strength of our approach is that the variation of any parameter in the model may be changed easily to study its influence on the retrieved optical properties (of course, within the range of validity of the model itself, e.g., dominant forward scattering). The empirical model by Jacques et al.22 would require a new set of simulations if, for example, the detection NA changes. In concept, however, both methods are equivalent in that the mapping function f(μS) in Eq. (1) is studied either using an analytical model or Monte Carlo simulations.

5.2.

Limitations and Clinical Implications

Our model calculations assume applicability of Mie–Percus–Yevick calculations to account for concentration-dependent scattering and the EHF model to account for multiple small-angle forward scattering. Input for the Mie–Percus–Yevick calculations is size and refractive index of the silica spheres making up the DRM. In our present calculations, we use the mean diameter of the size distribution as measured by TEM. Alternatively, an effective “scattering diameter” can be computed by weighting the size distribution with the corresponding scattering cross sections which yield only small deviations from our present calculations (<1% in μOCT). By averaging over all particle sizes, we can draft a rule of thumb expression for the concentration-dependent scattering coefficient μS as a function of volume fraction fV and the concentration-independent scattering cross section σS,MIE of particles with volume VP:

μS(12.4{±0.2}fv)fvVPσS,MIE,
which is intriguingly similar to the result obtained for blood37 (see below).

Since the specified uncertainty of the refractive index of silica beads is 2%, we have selected the refractive index (n=1.425) that describes all measured data most accurately. To study how the refractive index affects the calculations, we have varied the optimal refractive index with 1% (not shown here). This resulted in a maximum uncertainty in μOCT of 20% for 0.70  μm spheres at 0.2 volume fraction (μOCT8.9±1.8  mm1).

An alternative approach to determining μOCT would be direct fitting of the EHF model [Eq. (8)] to the OCT signal in depth to obtain μS and θrms. However, this is not always possible because μS and θrms (via wS) are competing parameters: change in one of the parameters can often be compensated by a change in the other leading to different sets of (μS, θrms) with equivalent statistical goodness of fit. The resulting curve fit may, therefore, not always converge to realistic values.15 A promising venue to be explored in future studies is the combination of OCT with other (fiber-based) technologies such as single-fiber reflectance spectroscopy38 that can provide independent parameterization of the scattering phase function—and thus in principle of θrms.

We calculate the amplitude directly from the OCT data and do not use the amplitude from the fitted A-scans. A practical concern is that in many cases of interest, such as epithelial tissues, the number of data points is limited for an NLLS curve fit. With a common OCT resolution of 10  μm, and for example a layer thickness of 50  μm, only five unique data points are available, which may lead to inconclusive fitting results on μOCT and μB,NA. We scaled the measured OCT signal amplitude to the backscattering coefficient, μB,NA, within the system’s NA using a scaling factor that was calculated at the highest volume fraction. Ideally, a single calibration factor could be used for scaling.10 In (clinical) practice, however, the scaling is influenced by parameters that may change between measurements, for example, when switching disposable probes with slightly different coupling efficiencies. In practice, it is, therefore, desirable to calibrate each probe/measurement individually using phantoms with known scattering properties. In this work, it required calibration for each volume-fraction series, per particle diameter.

To obtain quantitative measurement of OCT-derived parameters μOCT and μB,NA, it is crucial to calibrate for system parameters (confocal PSF and sensitivity roll-off). However, the contribution of multiple scattering is challenging to quantify. From Fig. 4, we conclude that multiple scattering starts to have a large (>10%) effect on the measured μOCT for samples with a high anisotropy, g>0.8, and high scattering coefficient, μS>10  mm1. In this case, it is advisable to consider multiple scattering to map μOCT to μS. For samples with lower anisotropy factor and scattering cross-section, values of μS are within 10% variation of μOCT.

The role of concentration-dependent scattering is absent for tissue, since this only relates to upscaling optical properties of individual scatterers to those of a solution of scatterers at a given volume fraction. In general, both “an individual scatterer” and “volume fraction” cannot be defined for biological tissue. Instead of a DRM, tissue is arguably better described as a continuous random medium,39 where statistical properties, such as variance, correlation length, and fractal dimension of the spatial refractive index fluctuation, take the role of refractive index contrast, correlation length of Percus–Yevick PCF, and particle size that describes the DRM. Linking these different sets of parameters to OCT measured optical properties will be subject of further research, deploying increasingly hybrid samples (duo and polydisperse bead mixtures of known size distribution).40

The exceptional biological specimen for which scattering properties may indeed be described by the Mie–Percus–Yevick formalism, is whole blood. The main scatterers (red blood cells) can be regarded as spherical when sufficient averaging over orientation is possible (validating the use of Mie theory).41 The physiological volume fraction (fV0.45) is sufficiently high to warrant the use of the Percus–Yevick structure factor. In a recent theoretical study annex literature review, we found that μs(1fV)2(fV/VRBC)σSCAT as opposed to μs=(fV/VRBC)σSCAT for concentration-independent scattering.37 Specifically, the predicted scattering coefficient and scattering anisotropy g at 800 nm are μs=71  mm1, g=0.9812; μA=0.38  mm1 and μs=63  mm1, g=0.9820; μA=0.47  mm1 for oxygenated and deoxygenated blood, respectively. The major contribution of multiple forward scattered light in this case resulted in the highly decreased contribution of scattering properties to the total OCT attenuation coefficient, e.g., measured μOCT of 5.5 and 5.8  mm1 for oxygenated and deoxygenated blood, respectively.42 Accurate separation of scattering and absorption contributions to μOCT may enable quantitative OCT-based measurement of μA (spectra) leading to quantification of localized oxygen saturation. Previously, we raised the question of whether quantitative measurements of attenuation by blood coefficients by OCT are feasible43 and concluded that better modeling of the concentration-dependent OCT signal was needed. With the theoretical results presented in Ref. 37 and the experimental validation of the influence of light scattering described in this contribution, accurate determination of attenuation by blood, and subsequent extraction of oxygen saturation, seems feasible and warrants further experimental validation studies.

6.

Conclusions

We propose a comprehensive model of the OCT signal accounting for the influence of the confocal PSF and sensitivity roll-off together with an experimental procedure from which the amplitude μB,NA and decay constant μOCT can be quantified. Both experimental quantities have been investigated as markers of disease (progression) in various medical fields. We validate their quantification using OCT measurements of DRM consisting of silica spheres within a wide range of scattering properties. We show that such validation measurements over the range of scattering properties found in tissue require DRM of high volume fractions, for which concentration-dependent scattering and multiple scattering effects cannot be neglected. We, therefore, include the Percus–Yevick structure factor and the EHF formalism in our model and demonstrate excellent agreement between predicted and measured values of μOCT and μB,NA versus volume fraction.

Acknowledgments

This research was funded as part of the IOP Photonic Devices, Project number IPD 12020, managed by the Netherlands Enterprise Agency, and was performed within the framework of Institute Quantivision, a collaboration initiative of AMC, NKI, UvA, VU, and VUmc. We thank E. van de Pol, PhD, for the size determination of the silica beads.

References

1. 

R. A. McLaughlin et al., “Parametric imaging of cancer with optical coherence tomography,” J. Biomed. Opt., 15 (4), 046029 (2010). http://dx.doi.org/10.1117/1.3479931 JBOPFO 1083-3668 Google Scholar

2. 

G. van Soest et al., “Atherosclerotic tissue characterization in vivo by optical coherence tomography attenuation imaging,” J. Biomed. Opt., 15 (1), 011105 (2010). http://dx.doi.org/10.1117/1.3280271 JBOPFO 1083-3668 Google Scholar

3. 

R. Wessels et al., “Optical coherence tomography in vulvar intraepithelial neoplasia,” J. Biomed. Opt., 17 116022 (2012). http://dx.doi.org/10.1117/1.JBO.17.11.116022 JBOPFO 1083-3668 Google Scholar

4. 

L. Scolaro et al., “Parametric imaging of the local attenuation coefficient in human axillary lymph nodes assessed using optical coherence tomography,” Biomed. Opt. Express, 3 (2), 366 (2012). http://dx.doi.org/10.1364/BOE.3.000366 BOEICL 2156-7085 Google Scholar

5. 

T. Xie, M. Zeidel and Y. Pan, “Detection of tumorigenesis in urinary bladder with optical coherence tomography: optical characterization of morphological changes,” Opt. Express, 10 (24), 1431 –1443 (2002). http://dx.doi.org/10.1364/OE.10.001431 OPEXFF 1094-4087 Google Scholar

6. 

R. Wessels et al., “Functional optical coherence tomography of pigmented lesions,” J. Eur. Acad. Dermatol. Venereol., 29 (4), 738 –744 (2015). http://dx.doi.org/10.1111/jdv.12673 Google Scholar

7. 

Y. Yang et al., “Optical scattering coefficient estimated by optical coherence tomography correlates with collagen content in ovarian tissue,” J. Biomed. Opt., 16 (9), 090504 (2011). http://dx.doi.org/10.1117/1.3625247 JBOPFO 1083-3668 Google Scholar

8. 

E. C. C. Cauberg et al., “Quantitative measurement of attenuation coefficients of bladder biopsies using optical coherence tomography for grading urothelial carcinoma of the bladder,” J. Biomed. Opt., 15 (6), 066013 (2010). http://dx.doi.org/10.1117/1.3512206 JBOPFO 1083-3668 Google Scholar

9. 

J. K. Percus and G. J. Yevick, “Analysis of classical statistical mechanics by means of collective coordinates,” Phys. Rev., 110 (1), 1 –13 (1958). http://dx.doi.org/10.1103/PhysRev.110.1 PHRVAO 0031-899X Google Scholar

10. 

V. D. Nguyen et al., “Dependent and multiple scattering in transmission and backscattering optical coherence tomography,” Opt. Express, 21 (24), 29145 –29156 (2013). http://dx.doi.org/10.1364/OE.21.029145 OPEXFF 1094-4087 Google Scholar

11. 

J. M. Schmitt and A. Knüttel, “Model of optical coherence tomography of heterogeneous tissue,” J. Opt. Soc. Am. A, 14 (6), 1231 (1997). http://dx.doi.org/10.1364/JOSAA.14.001231 JOAOD6 0740-3232 Google Scholar

12. 

L. Thrane, H. T. Yura and P. E. Andersen, “Analysis of optical coherence tomography systems based on the extended Huygens-Fresnel principle,” J. Opt. Soc. Am. A, 17 (3), 484 (2000). http://dx.doi.org/10.1364/JOSAA.17.000484 JOAOD6 0740-3232 Google Scholar

13. 

C. Xu et al., “Characterization of atherosclerosis plaques by measuring both backscattering and attenuation coefficients in optical coherence tomography,” J. Biomed. Opt., 13 (3), 034003 (2008). http://dx.doi.org/10.1117/1.2927464 JBOPFO 1083-3668 Google Scholar

14. 

T. G. Van Leeuwen, D. J. Faber and M. C. Aalders, “Measurement of the axial point spread function in scattering media using single-mode fiber-based optical coherence tomography,” IEEE J. Sel. Top. Quantum Electron., 9 (2), 227 –233 (2003). http://dx.doi.org/10.1109/JSTQE.2003.813299 IJSQEN 1077-260X Google Scholar

15. 

D. J. Faber et al., “Quantitative measurement of attenuation coefficients of weakly scattering media using optical coherence tomography,” Opt. Express, 12 (19), 4353 –4365 (2004). http://dx.doi.org/10.1364/OPEX.12.004353 OPEXFF 1094-4087 Google Scholar

16. 

R. Leitgeb, C. Hitzenberger and A. Fercher, “Performance of Fourier domain vs. time domain optical coherence tomography,” Opt. Express, 11 (8), 889 –894 (2003). http://dx.doi.org/10.1364/OE.11.000889 OPEXFF 1094-4087 Google Scholar

17. 

N. Nassif et al., “In vivo high-resolution video-rate spectral-domain optical coherence tomography of the human retina and optic nerve,” Opt. Express, 12 (3), 367 –376 (2004). http://dx.doi.org/10.1364/OPEX.12.000367 OPEXFF 1094-4087 Google Scholar

18. 

V. M. Kodach et al., “Determination of the scattering anisotropy with optical coherence tomography,” Opt. Express, 19 (7), 6131 –6140 (2011). http://dx.doi.org/10.1364/OE.19.006131 OPEXFF 1094-4087 Google Scholar

19. 

N. Bosschaart et al., “Measurements of wavelength dependent scattering and backscattering coefficients by low-coherence spectroscopy,” J. Biomed. Opt., 16 (3), 030503 (2011). http://dx.doi.org/10.1117/1.3553005 JBOPFO 1083-3668 Google Scholar

20. 

H. R. G. W. Verstraete et al., “Towards model-based adaptive optics optical coherence tomography,” Opt. Express, 22 (26), 32406 –32418 (2014). http://dx.doi.org/10.1364/OE.22.032406 OPEXFF 1094-4087 Google Scholar

21. 

B. Cense et al., “Ultrahigh-resolution high-speed retinal imaging using spectral-domain optical coherence tomography,” Opt. Express, 12 (11), 2435 –2447 (2004). http://dx.doi.org/10.1364/OPEX.12.002435 OPEXFF 1094-4087 Google Scholar

22. 

S. L. Jacques et al., “Measuring tissue optical properties in vivo using reflectance-mode confocal microscopy and OCT,” Proc. SPIE, 6864 68640B (2008). http://dx.doi.org/10.1117/12.761803 PSISDG 0277-786X Google Scholar

23. 

H. C. van der Hulst, Light Scattering by Small Particles, Dover Publications, New York (1957). Google Scholar

24. 

L. Tsang et al., Scattering of Electromagnetic Waves, Numerical Simulations, John Wiley & Sons, Inc., Hoboken, New Jersey (2001). Google Scholar

25. 

G. M. Hale and M. R. Querry, “Optical constants of water in the 200-nm to 200-μm wavelength region,” Appl. Opt., 12 (3), 555 –563 (1973). http://dx.doi.org/10.1364/AO.12.000555 APOPAI 0003-6935 Google Scholar

26. 

G. Singh and L. Song, “Influence of sodium dodecyl sulfate on colloidal fouling potential during ultrafiltration,” Colloids Surf. A, 281 138 –146 (2006). http://dx.doi.org/10.1016/j.colsurfa.2006.02.036 Google Scholar

27. 

T. R. Hillman et al., “Correlation of static speckle with sample properties in optical coherence tomography,” Opt. Lett., 31 (2), 190 (2006). http://dx.doi.org/10.1364/OL.31.000190 OPLEDP 0146-9592 Google Scholar

28. 

J. W. Goodman, Speckle Phenomena in Optics: Theory and Applications, Roberts and Company Publishers, Greenwood Village, Colorado (2007). Google Scholar

29. 

D. Levitz et al., “Quantitative characterization of developing collagen gels using optical coherence tomography,” J. Biomed. Opt., 15 (2), 026019 (2010). http://dx.doi.org/10.1117/1.3377961 JBOPFO 1083-3668 Google Scholar

30. 

J. J. ten Bosch, “Pascal program to perform Mie calculations,” Opt. Eng., 32 (7), 1691 –1695 (1993). http://dx.doi.org/10.1117/12.138819 Google Scholar

31. 

I. H. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am., 55 (10), 1205 (1965). http://dx.doi.org/10.1364/JOSA.55.001205 JOSAAH 0030-3941 Google Scholar

32. 

K. F. Palmer and D. Williams, “Optical properties of water in the near infrared,” J. Opt. Soc. Am., 64 (8), 1107 (1974). http://dx.doi.org/10.1364/JOSA.64.001107 JOSAAH 0030-3941 Google Scholar

33. 

R. Wessels et al., “Learning curve and inter-observer variance in quantification of the optical coherence tomography attenuation coefficient,” J. Biomed. Opt., 20 (12), 121313 (2015). http://dx.doi.org/10.1117/1.JBO.20.12.121313 Google Scholar

34. 

K. A. Vermeer et al., “Depth-resolved model-based reconstruction of attenuation coefficients in optical coherence tomography,” Biomed. Opt. Express, 5 (1), 322 –337 (2014). http://dx.doi.org/10.1364/BOE.5.000322 BOEICL 2156-7085 Google Scholar

35. 

M. T. J. Bus et al., “Volumetric in vivo visualization of upper urinary tract tumors using optical coherence tomography: a pilot study,” J. Urol., 190 (6), 2236 –2342 (2013). http://dx.doi.org/10.1016/j.juro.2013.08.006 Google Scholar

36. 

K. Barwari et al., “Advanced diagnostics in renal mass using optical coherence tomography: a preliminary report,” J. Endourol., 25 (2), 311 –315 (2011). http://dx.doi.org/10.1089/end.2010.0408 Google Scholar

37. 

N. Bosschaart et al., “A literature review and novel theoretical approach on the optical properties of whole blood,” Lasers Med. Sci., 29 (2), 453 –479 (2014). http://dx.doi.org/10.1007/s10103-013-1446-7 Google Scholar

38. 

U. A. Gamm et al., “Measurement of tissue scattering properties using multi-diameter single fiber reflectance spectroscopy: in silico sensitivity analysis,” Biomed. Opt. Express, 2 (11), 3150 (2011). http://dx.doi.org/10.1364/BOE.2.003150 BOEICL 2156-7085 Google Scholar

39. 

J. M. Schmitt and G. Kumar, “Turbulent nature of refractive-index variations in biological tissue,” Opt. Lett., 21 (16), 1310 –1312 (1996). http://dx.doi.org/10.1364/OL.21.001310 OPLEDP 0146-9592 Google Scholar

40. 

J. M. Schmitt and G. Kumar, “Optical scattering properties of soft tissue: a discrete particle model,” Appl. Opt., 37 (13), 2788 –2797 (1998). http://dx.doi.org/10.1364/AO.37.002788 APOPAI 0003-6935 Google Scholar

41. 

J. M. Steinke and A. P. Shepherd, “Comparison of Mie theory and the light scattering of red blood cells,” Appl. Opt., 27 (19), 4027 –4033 (1988). http://dx.doi.org/10.1364/AO.27.004027 APOPAI 0003-6935 Google Scholar

42. 

D. J. Faber et al., “Toward assessment of blood oxygen saturation by spectroscopic optical coherence tomography,” Opt. Lett., 30 (9), 1015 –1017 (2005). http://dx.doi.org/10.1364/OL.30.001015 OPLEDP 0146-9592 Google Scholar

43. 

D. J. Faber and T. G. van Leeuwen, “Are quantitative attenuation measurements of blood by optical coherence tomography feasible?,” Opt. Lett., 34 (9), 1435 –1437 (2009). http://dx.doi.org/10.1364/OL.34.001435 OPLEDP 0146-9592 Google Scholar

Biography

Mitra Almasian is a PhD candidate in the Biomedical Engineering and Physics Department of the Academic Medical Center (AMC), University of Amsterdam, Amsterdam, The Netherlands. Her research is focused on optical coherence tomography. After an interdisciplinary bachelor with a major in chemistry, she received her master’s degree in physical sciences from the University of Amsterdam in 2012.

Nienke Bosschaart is an assistant professor at the Biomedical Photonic Imaging Group (BMPI) of the MIRA Institute at the University of Twente, Enschede, The Netherlands. She received her PhD at the AMC of the University of Amsterdam in 2012. Her research interests are the noninvasive quantification of tissue (optical) properties for medical diagnostics, in particular using low-coherence spectroscopy and spectroscopic optical coherence tomography.

Ton G. van Leeuwen is full professor in biomedical physics and in 2008 was appointed as head of the Biomedical Engineering and Physics Department at the Academic Medical Center of the University of Amsterdam. His current research focuses on the physics of the interaction of light with tissue, and using that knowledge for the development, introduction, and clinical evaluation of newly developed optical imaging techniques for gathering quantitative functional and molecular information of tissue.

Dirk J. Faber is an assistant professor (Universitair Docent) of biomedical optics at the Department of Biomedical Engineering and Physics of the AMC, University of Amsterdam. He studied applied physics at the University of Twente in 2000 and graduated in “Functional Optical Coherence Tomography” at AMC in 2005. His research interests are the physics of light–tissue interaction with specific application to cancer diagnosis.

© 2015 Society of Photo-Optical Instrumentation Engineers (SPIE) 1083-3668/2015/$25.00 © 2015 SPIE
Mitra Almasian, Nienke Bosschaart, Ton G. van Leeuwen, and Dirk J. Faber "Validation of quantitative attenuation and backscattering coefficient measurements by optical coherence tomography in the concentration-dependent and multiple scattering regime," Journal of Biomedical Optics 20(12), 121314 (29 December 2015). https://doi.org/10.1117/1.JBO.20.12.121314
Published: 29 December 2015
Lens.org Logo
CITATIONS
Cited by 57 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Scattering

Optical coherence tomography

Mie scattering

Light scattering

Multiple scattering

Signal attenuation

Backscatter

Back to Top