Open Access
19 October 2012 In vivo spatial frequency domain spectroscopy of two layer media
Author Affiliations +
Abstract
Monitoring of tissue blood volume and local oxygen saturation can inform the assessment of tissue health, healing, and dysfunction. These quantities can be estimated from the contribution of oxyhemoglobin and deoxyhemoglobin to the absorption spectrum of the dermis. However, estimation of blood related absorption in skin can be confounded by the strong absorption of melanin in the epidermis and epidermal thickness and pigmentation varies with anatomic location, race, gender, and degree of disease progression. Therefore, a method is desired that decouples the effect of melanin absorption in the epidermis from blood absorption in the dermis for a large range of skin types and thicknesses. A previously developed inverse method based on a neural network forward model was applied to simulated spatial frequency domain reflectance of skin for multiple wavelengths in the near infrared. It is demonstrated that the optical thickness of the epidermis and absorption and reduced scattering coefficients of the dermis can be determined independently and with minimal coupling. Then, the same inverse method was applied to reflectance measurements from a tissue simulating phantom and in vivo human skin. Oxygen saturation and total hemoglobin concentrations were estimated from the volar forearms of weakly and strongly pigmented subjects using a standard homogeneous model and the present two layer model.

1.

Introduction

Diffuse optical spectroscopy (DOS) can be used to characterize the optical properties of skin by measuring the amount of radiative energy remitted from the skin.13 Spatial frequency domain imaging uses patterned illumination to determine the absorption and reduced scattering coefficients of a turbid medium based on the spatial frequency dependent reflectance function. The method has been described in detail in the literature.412 Instead of measuring the total reflectance, spatial frequency domain techniques measure the attenuation of specific spatial frequency components of the illuminating pattern as it propagates in a tissue. Typically, a one dimensional sine-wave pattern with controlled spatial frequency fx is used and set to have sensitivity to a particular physical depth since higher frequencies tend to probe a more superficial volume. Additionally, high spatial frequencies have been shown to be more sensitive to the tissue’s scattering coefficient. On the other hand, lower frequencies (i.e., fx=0mm1) have been shown to be sensitive to the tissue’s absorption coefficient.4,5,7,1315

Previously, we presented an accurate and efficient model of light transfer through optical media suitable for analyzing spatial frequency domain data for multilayered media. An artificial neural network was trained to map a set of optical properties of two layer media to a spatial frequency domain solution of the radiative transfer equation (RTE). Additionally, we developed an iterative inverse method with the neural network as a forward model that could be used to analyze reflectance of two layer media to determine the optical properties of each layer. In this study, we apply the neural network model to (i) simulated skin phantoms, (ii) two layer polydimethylsiloxane (PDMS) phantoms, and (iii) in vivo measurements from human volar forearms with varying pigmentation. The accuracy, application, and limitations of the present model are discussed in the context of biomedical spatial frequency domain spectroscopy of skin.

2.

Background

2.1.

Structure of The Skin

The human skin consists of the epidermis and dermis and is supported by a subcutaneous layer of fat. Epidermal thickness ranges between ten to hundreds of micrometer in thickness. The cells in the epidermis include the basal keratinocytes, melanocytes, and Langerhans.16 Melanocytes synthesize melanin, the skin protein mainly responsible for racial and seasonal variation in skin color. Melanin is contained in organelles knows as melanosomes.16,17 Melanosomes concentration in the epidermis varies between 1% and 43% of the epidermal volume corresponding to lightly or darkly pigmented skin, respectively.1619 However, the degree of pigmentation by melanin associated with genetic and seasonal variation was assessed in this study with the Fitzpatrick skin pigmentation scale which categorizes human skin into one of six categories.20 Skin Type I refers to white skin which burns easily when expose to the sun and never tans while Type VI refers to skin that is very dark, never burns due to solar exposure and tans easily. Melanosomes are concentrated near to the basement membrane.2123 In normal skin, melanin absorption dominates the total absorption in the epidermis in the visible range and NIR.14,1619

The dermis, located beneath the epidermis, is responsible for the skin's pliability, mechanical resistance and temperature control. The thickness of the dermis ranges between 450 μm and several millimeters.2427 The dermis is primarily composed of collagen fibers, nerves, capillaries, and blood vessels but also contains elastin, fibroblasts, and Schwann and endothelial cells.16,17 Collagen fibers make up 70% of the dermis, giving it strength and toughness. Elastin maintains normal elasticity and flexibility. The dermis consists of the papillary and reticular dermal layers. The papillary dermis connects the epidermis to underlying tissue and consists of long, thin randomly oriented collagen fibers. Thicker bundles of collagen run parallel to the skin surface in the deeper reticular layer. The volume of blood in the dermis ranges between 0.2 and 7% and varies with bodily location.19,28 Erythrocytes (red blood cells) occupy approximately half of the blood volume and are responsible for oxygen transfer from the lungs throughout the body. Erythrocytes are composed mainly of hemoglobin molecules which reversibly bind to oxygen molecules in the lungs to form oxyhemoglobin. Hemoglobin is known as deoxyhemoglobin once it has released its oxygen molecules. Hemoglobin absorption dominates the total absorption in the dermis in the visible range and near infrared (NIR).14,16,17

2.2.

Depth Sensitivity of Spatial Frequency Domain Reflectance

Spatial frequency domain spectroscopy involves illumination of media with a spatially varying intensity pattern of the form,

Eq. (1)

q(x)=P0(fx)cos(2πfxx),
where fx is the spatial frequency expressed in mm1, P0(fx) is the incident optical power at the spatial frequency fx and measured in mW, and x is the spatial coordinate parallel to the medium’s surface and measured in mm. The remitted energy differs from the illumination pattern due to the optical and geometric property characteristics of the sample.47 In fact, the spatial frequency dependent reflectance measured from a turbid medium encodes both depth and optical property information, enabling both quantitation and low resolution depth sectioning of the spatially varying optical properties of a medium.5,810

Several investigators have established the depth-sensitivity of spatial frequency domain imaging.5,810 Cuccia et al.,5 demonstrated that spatially modulated illumination facilitates quantitative wide-field optical property mapping and exhibits sensitivity to buried heterogeneities in turbid media. They performed spatial frequency domain imaging of tissue simulating phantoms with embedded heterogeneities using 42 spatial frequencies fx between 0 and 0.6mm1. They showed that changes in demodulated reflectance at low versus high spatial frequencies are sensitive to the lower versus upper embedded heterogeneities. This work showed that spatial frequency domain spectroscopy can detect contrast between background and heterogeneity, but did not provide a quantitative technique for determining the optical properties of the heterogeneity itself.

Konecky, et al.,8 used spatial frequency domain imaging to detect tube heterogeneities buried in homogeneous tissue simulating phantoms. They measured spatial frequency dependent reflectance from the heterogeneous phantoms at 11 spatial frequencies. They then used an inverse method based on the diffusion approximation to reconstruct tomographic contrast images of the buried tubes. This model, however, may not be applicable if the highly absorbing heterogeneity is close to the surface (i.e., the epidermis of the skin).

2.3.

Artificial Neural Networks For Two Layer Spectroscopy

The present study couples the inherent depth sensitivity of spatial frequency domain imaging with an accurate and fast semi-empirical model of light transfer through tissue that does not suffer from the limitations of approximate methods such as the diffusion approximation. We performed multiple Monte Carlo simulations and then fit a machine learning algorithm—an artificial neural network—to the output data. An artificial neural network is a data structure that can accurately approximate a nonlinear relationship between a set of input and output parameters from multiple samples of the input-output pairs.29 Neural network based approaches to determining optical scattering and absorption coefficients of biological tissue and phantom reflectance measurements have been proposed by other investigators.3038 Wang, Shastri and Pfefer developed a neural network to detect the optical properties of two layer media using spatially resolved reflectance measurements.38 They then used their inverse model to detect the absorption and reduced scattering coefficients of two layer media in the ultra-violet and visible ranges, for which the tissues’ absorption coefficient is typically greater than its reduced scattering coefficient. However, they reported large prediction errors in all their parameters (absorption and scattering coefficients of the top and bottom layer). Additionally, their method required multiple spatially resolved measurements. This study uses a forward and inverse model designed for spatial frequency domain measurements of two layer media from a single detector.

3.

Methods

3.1.

Digital Skin Simulating Phantom Preparation

We first tested the artificial neural network on digital skin phantom data generated by Monte Carlo simulations. We modeled the skin as two layer system depicted in Fig. 1. The absorption coefficient of the epidermal layer μa,1 as a function of λ was defined as

Eq. (2)

μa,1(λ)=εmel(λ)fmel,
where εmel(λ) was the molar extinction coefficient of melanin measured in mm1/μM and fmel was the molarity of melanin in the epidermis measured in µM and was assumed to range between 0.01 and 0.06 µM.19,28 The absorption coefficient of the dermis μa,2 as a function of λ was calculated as

Eq. (3)

μa,2(λ)=[εoxy(λ)SO2+εdeoxy(λ)(1SO2)]fheme,
where εoxy and εdeoxy were the molar extinction coefficients of oxyhemoglobin and deoxyhemoglobin, respectively, SO2 was the percent oxygen saturation of hemoglobin, and fheme was the molarity of hemoglobin in the dermis measured in µM and was assumed to range between 20 and 200 µM. Values of fheme lower than 20 µM resulted in an essentially transparent epidermis while the spectral features of the dermal tissue became invisible for fheme larger than 200 µM. The scattering coefficients of the epidermis μs,1 and dermis μs,2 were modeled according to a power-law relationship with wavelength,

Eq. (4)

μs,1(λ)=A1λb1,

Eq. (5)

μs,2(λ)=A2λb2,
where Ai and bi are the so-called scattering amplitude and power, respectively.39

Fig. 1

Illustration of the two layer geometry, optical properties, and illumination considered.

JBO_17_10_107006_f001.png

The spatial frequency dependent reflectance R(n1,d1,μa,1,μs,1,μa,2,μs,2,fx), was determined with Monte Carlo simulations. Monte Carlo simulation software developed by Wang and Jacques40 was used to calculate the radial diffuse reflectance function R(n1,d1,μa,1,μs,1,μa,2,μs,2,ρ) where ρ was the radial distance from the simulation’s origin. Then, the spatial frequency domain diffuse reflectance function was calculated with the Hankel transform using the method suggested by Cuccia et al.,13

Eq. (6)

R(n1,d1,μa,1,μs,1,μa,2,μs,2,fx)=2π0ρJ0(2πfx)R(n1,d1,μa,1,μs,1,μa,2,μs,2,ρ)dρ,
where Jo is the 0’th order Bessel function of the first kind. Each simulation was run with 106 photon packets.

3.2.

Two Layer Silicone Phantom Preparation

A two layer skin simulating phantom was prepared by combining two homogeneous phantoms with distinct absorption and scattering characteristics.41,42 Polydimethylsiloxane (PDMS) served as a base for each phantom. Curing agent from a commercially available kit (Kit P-4, Eager Plastics) was added to bulk PDMS at a ratio of 10 parts PDMS to one part curing agent. PDMS is weakly absorbing and non-scattering in the NIR. Scattering was achieved by the addition of rutile TiO2 powder (Ti-602, Atlantic Equipment Engineers), which has a high index of refraction relative to PDMS, but negligible absorption in the NIR spectrum.43 The epidermis simulating phantom was colored with coffee.42 The absorption spectra of coffee and melanin both lack distinct spectral features and exhibit power law-like decay from the visible into the near infrared. For this reason, coffee has been used a cost-effective and readily available colorant for epidermis simulating phantoms in our study and by other investigators.44,45 India ink (PRO-4100, Pro Art) was added to the dermis simulating layer to achieve a desired level of absorption. Approximately 0.07 g India Ink (0.25g/l) was required to generate an absorption of roughly 0.02mm1, and approximately 0.2 g TiO2 (0.73g/l) was needed to generate reduced scattering coefficient of roughly 1mm1 at 650 nm. These amounts did not significantly change the volume of the phantom. The thicknesses of the top and bottom layer phantoms were 250 µm and 1 cm, respectively. The refractive index of both phantoms was approximately 1.43, which was a close approximation of tissue. After preparation and cleaning, the thin epidermal phantom was placed upon the dermal phantom. A roller was used to squeeze out any air bubbles that formed between the two phantoms during assembly.

3.3.

Spatial Frequency Domain Imaging System

Spatial frequency domain imaging was achieved by illuminating a surface of interest with light having a spatially varying intensity and capturing the reflected image with a charge coupled device (CCD) camera. A schematic and photograph of the instrument used in this study is shown in Fig. 2. Broadband light was generated with a Newport Corporation power source (Newport Corporation, Irvine CA) and a 250 watt tungsten lamp. This light was used to illuminate a DLP Developers Kit 11024×768pixel digital micro-mirror device (Texas Instrument, Dallas TX), which spatially modulated the light to be projected onto the phantom. Frequency patterns for the spatial modulation were generated using customized software (Modulated Imaging Inc., Irvine CA). The patterned light was projected onto the surface of interest via a 45 deg mirror. Six spatial frequencies spaced evenly between from 0 to 0.25mm1 were used during the imaging.46 The diffusely reflected light images were acquired with a Nuance Multispectral Imaging System (CRi Inc., Woburn,MA), which consists of a liquid crystal tunable filter capable of acquiring at discrete wavelengths (10 nm bandwidth) between 650 nm and 1100 nm, and a 1040×1392pixel, front illuminated CCD. A 6×6cm field of view was acquired by the CCD using 4×4 binning in hardware and an exposure time of 100ms. Cross-polarizers were placed at the source and detector in order to eliminate specular reflection. Thus, the modeling component of this study only considered the diffuse reflectance because specular reflectance was removed by cross-polarization. Additionally, the modeling did not consider polarization of the diffusely reflected light because multiple scattering events of light in the tissue would effectively randomize the polarization angels of the reflected energy. Illumination by eleven (11) uniformly spaced wavelengths between 650 and 800 nm were used during two layer silicone phantom and in vivo measurements.

Fig. 2

Schematic (a) and photograph (b) of spatial frequency domain imager used in this study.

JBO_17_10_107006_f002.png

3.4.

In vivo Skin Measurements

The volar forearms of four subjects were imaged with the spatial frequency domain imaging (SFDI) system under UC Irvine IRB protocol HS# 2008-6307. Each subject of varying skin color was male, between 20 and 30 years of age, and in good health. Subjects were imaged in a comfortable, seated position. Subjects were allowed to rest for 5 min prior to imaging to reduce the effect of systemic vascular changes. Data post-processing and analysis for digital, silicone phantoms, and in vivo measurements were completed in MATLAB (Mathworks, Natick, MA).

For each patient, a resting baseline was measured for 3 min. A Hokanson AG101 Rapid Cuff Inflator was then used to immediately occlude venous blood flow into the volar forearm at a pressure of 200 mm Hg. This was maintained for 3 min before being released for a final 3 min of recovery.

During the entire 9 min period, continuous SFDI measurements were acquired using four equally spaced wavelengths between 650 and 749 nm at five spatial frequencies equally spaced from 0 to 0.5mm1.

4.

Analysis

4.1.

Two Layer Skin Model

Figure 1 shows the two layer geometry, optical properties, and illumination considered in this study. Human skin was approximated by two layers, the epidermis and the dermis. Note that we made the simplifying assumption that melanin and blood were uniformly distributed in the epidermis and dermis, respectively. The epidermis was illuminated by a collimated light source. The spatial frequency fx of the illumination pattern was considered to range between 0 and 0.25mm1.The index of refraction, absorption coefficient, reduced scattering coefficients, and thickness of the epidermis where denoted by n1, μa,1, μs,1, and d1, respectively. The index of refraction n1 was assumed to be 1.40 to represent biological tissue47 and was suitable for both the silicone phantom and human skin. The absorption coefficient μa,1 was assumed to range between 0.10 and 2.00mm1.4753 The reduced scattering coefficient μs,1 was assumed to be identically 1.00mm1 since it was shown previously that large changes in μs,1 resulted in minimal changes in the reflectance of the two layer medium(see Fig. 2 of Ref. 54). Thus setting μs,1 to a constant value of 1mm1 sufficiently captured the effect of scattering in the thin superficial layer. Finally, the thickness d1 was assumed to range between 15 and 150 μm.26,55,56 Additionally, it was shown that the individual values of μa,1 and d1 were not expected to be detectable from spatial frequency dependent reflectance (see Fig. 7 of 54). Hence, this study focused on determining the optical thickness τ1 of the epidermis defined as

Eq. (7)

τ1=d1(μa,1+μs,1).
The index of refraction, absorption coefficient, and reduced scattering coefficients of the dermal layer where denoted by n2, μa,2, and μs,2, respectively. For simplicity, a matched index boundary condition was assumed between the epidermis and dermis (i.e., n1=n2=1.40). The absorption coefficient of dermis, μa,2, was assumed to range between 0.01 and 0.20mm1 while reduced scattering coefficient μs,2 was assumed to range between 0.50 and 2.00mm1.4753

4.2.

Inverse Problem with the Least-Squares Fitting

The forward model NNf(τ1,μa,2,μs,2) defined in Eq. (5) of Ref. 54 was used along with an iterative inverse method to estimate parameters τ1, μ^a,2, and μ^s,2 from a spatial frequency dependent reflectance Rref(fx). This was done by choosing τ^1, μ^a,2, and μ^s,2 to minimize a quadratic cost function,

L=i=1N[NNf(τ1,μ^a,2,μ^s,2,fx,i)Rref(fx,i)]2,
where Rref(fx,i) was the spatial frequency dependent reflectance at spatial frequency fx,i and N was the total number of spatial frequencies. In this study, L was minimized with the Levenberg-Marquardt algorithm implemented in MATLAB. Six spatial frequencies fx between 0 and 0.25mm1 were used (N=6).

5.

Results and Discussion

5.1.

Digital Skin Phantom Spectroscopy

Figure 3(a) shows estimates of the reduced optical thickness τ1(λ) for d1 between 40 and 150 μm, fmel between 0.01 and 0.10 µM, fheme equaling to 75 µM, and SO2 equaling to 50%, A1 and A2 both equaling 103, b1 equaling to 1.30, and b2 equaling to 1.35. Relative percent error between the true and estimated value are shown as a function of wavelength in the upper left corner of each panel. For these biometric parameters and for λ between 650 and 800 nm, μa,1(λ) ranged between 0.13 and 1.66mm1, μa,2(λ) ranged between 0.013 and 0.142mm1, μs,1 ranged between 1.77 and 2.32mm1, and μs,1(λ) ranged between 1.35 and 1.79mm1. Figure 3(a) illustrates that the reduced optical thickness τ1 was well estimated in the NIR by the present inverse method from spatial frequency dependent reflectance with six evenly spaced spatial frequencies fx between 0.00 and 0.25mm1. Furthermore, it was apparent that the estimation error was less for thick and highly pigmented epidermis (i.e., d1=140μm and fmel=0.06μM) than for thin and weakly pigmented epidermis (i.e., d1=50μm and fmel=0.01μM). This is due to the diminished sensitivity of the spatial frequency dependent reflectance function to τ1 when τ1 was small as depicted in Fig. 7 of Ref. 54. Figure 3(b) shows estimates of the reduced optical thickness τ1 retrieved from spatial frequency dependent reflectance functions with 2% additive noise. True noise levels depend on the (i) instrumentation used, (ii) integration time, (iii) filtering, and (iv) oversampling. A 2% noise level was added for illustrative purposes. As expected, noise corruption has a larger effect on estimation of optically thin epidermal layers than on optically thick layers.

Fig. 3

Estimation of τ1 for tissue simulating digital phantom (a) without noise and (b) with 2% noise. Relative percent error between the true and estimated value are shown as a function of wavelength in the upper left corner of each panel.

JBO_17_10_107006_f003.png

Figure 4(a) depicts estimates of the dermal absorption coefficient μa,2(λ) for d1 between 40 and 150 μm, fmel between 0.01 and 0.10 µM, fheme equaling to 75 µM, and SO2 equaling to 50%, A1 and A2 both equaling 103, b1 equaling to 1.30, and b2 equaling to 1.35. The melanin concentrations fmel equaling to 0.01and 0.10 µM corresponded to an epidermal absorption coefficient of 0.20 and 2.0mm1 at 694 nm and were typical of lightly and moderately pigmented subject, respectively. Accurate determination of μa,2(λ) is essential for transcutaneous tissue oximetry which attempts to determine the dermal blood volume and oxygen saturation by decomposing μa,2(λ) into absorption contribution by oxyhemoglobin and deoxyhemoglobin according to Eq. (3). It is desired that estimation of μa,2(λ) be independent of epidermal thickness and pigmentation when, for example, imaging tissue with large changes in epidermal thickness or melanin due to natural or pathological factors. Figure 4(a) illustrates that the present method accurately detects μa,2(λ) in the NIR for a large range of epidermal thicknesses and melanin concentration. In addition, Fig. 4(b) shows that the method was robust to 2% additive noise. Figure 4 illustrates that analysis of spatial frequency dependent reflectance can probe beneath optically thin and thick superficial layers.

Fig. 4

Estimation of μa,2 for tissue simulating digital phantom (a) without noise and (b) with 2% noise. Relative percent error between the true and estimated value are shown as a function of wavelength in the upper left corner of each panel.

JBO_17_10_107006_f004.png

Figure 5(a) illustrates the effectiveness of the present method in estimating the reduced scattering coefficient of the dermal layer μs,2 for d1 between 40 and 150 μm, fmel between 0.01 and 0.10 µM, fheme equaling to 75 µM, and SO2 equaling to 50%, A1 and A2 both equaling 103, b1 equaling to 1.30, and b2 equaling to 1.35. Relative percent error between the true and estimated value are shown as a function of wavelength in the upper left corner of each panel. Additionally, Fig. 5(b) shows these results in the presence of 2% additive noise. The method performed well in retrieving μs,2(λ) in the NIR despite large variations in epidermal thickness and pigmentation and μs,1(λ) which varied with λ.

Fig. 5

Estimation of μs,2 for tissue simulating digital phantom (a) without noise and with (b) 2% noise. Relative percent error between the true and estimated value are shown as a function of wavelength in the upper left corner of each panel.

JBO_17_10_107006_f005.png

5.2.

Two-Layer Silicone Phantom Spectroscopy

Figure 6 shows the estimated and true optical thickness τ1 of the two layer silicone phantom as a function of wavelength λ between 650 and 980 nm. The standard deviations of the estimates at each wavelength are also shown. The true optical thickness τ1 of the simulated epidermal layer was a superposition of the monotonically decreasing absorption coefficient of coffee onto the relatively featureless absorption characteristics of pure PDMS. The small absorption peaks of PDMS were observed at 910 nm. The absorption peak at 980 nm was attributed to coffee. The estimated absorption coefficient of the simulated epidermal layer predicted by the present model was similar in trend and magnitude to its true value.

Fig. 6

Estimated and true optical product of the top layer thickness and absorption coefficient of the two layer silicone phantom. The standard deviations of the estimates at each wavelength are also shown.

JBO_17_10_107006_f006.png

Figure 7 shows estimated and true absorption coefficient μa,2 of the supporting layer of silicone phantom representing dermis in skin. The true absorption coefficient μa,2 was close to 0.02mm1 for nearly the entire wavelength range between 650 and 980 nm. It exhibits a strong peak at 910 nm from PDMS. The estimated absorption coefficient μa,2 did not deviate from the true absorption coefficient by more than 10% for the entire range displayed. Additionally, the strong absorption peak near 910 nm was well captured by the two layer model. Similarly, Fig. 8 depicts estimated and true reduced scattering coefficient of the bottom layer μs,2 as a function of wavelength λ between 650 and 980 nm. The true reduced scattering coefficient exhibits the typical exponential decay with wavelength and monotonically decreases from 1.1 and 0.6mm1. The estimated reduced scattering coefficient closely followed the true value of μs,2 for the entire range shown. In fact, the two values were never deviated from one another by more than 10%.

Fig. 7

Estimated and true absorption coefficient μa,2 of the supporting layer of silicone phantom representing dermis in skin. The standard deviations of the estimates at each wavelength are also shown.

JBO_17_10_107006_f007.png

Fig. 8

Estimated and true reduced scattering coefficient μs,2 of the supporting layer of silicone phantom representing dermis in skin. The standard deviations of the estimates at each wavelength are also shown.

JBO_17_10_107006_f008.png

5.3.

In vivo Skin Measurements

Previous sections established the ability of the present model in analyzing simulated and measured diffuse reflectance phantoms from two layer media. The present section compares results from a previously developed homogeneous model6 and the present two layer model in analysis of spatial frequency domain spectroscopy measurements from human skin on the volar forearm. It aims to justify the complexity of two layer spectroscopy in the context of tissue oximetry. For example, Fig. 9 shows estimates of oxyhemoglobin and deoxyhemoglobin concentrations denoted by foxy and fdeoxy, respectively, for four subjects with increasingly darker skin type predicted by a homogeneous model6 and by the present two layer model. The absorption coefficient estimated by the homogeneous model6 denoted by μa,skin(λ)–was spectrally decomposed to determine the constituent chromophore concentrations. Results are presented for spectral decomposition with and without melanin as a contributing chromophore. Furthermore, Fig. 9 shows the estimation of oxyhemoglobin and deoxyhemoglobin by spectral decomposition of the dermal absorption coefficient μa,2(λ) determined from analysis of the same data with the present model. We assumed that μa,2(λ) was completely determined by absorption due to blood and that all melanin absorption was segregated to μa,1(λ). The cross-talk between epidermal melanin and dermal chromophore concentrations is most apparent in Fig. 9(b) that shows the apparent deoxyhemoglobin concentration increasing artificially for subjects with darker skin tone. This effect was apparent for both melanin compensated and non-compensated decomposition of μa,skin(λ). On the other hand, the estimates oxyhemoglobin foxy and deoxyhemoglobin fdeoxy by the present two layer model did not exhibit coupling with melanin. In fact, estimates of deoxyhemoglobin–which were apparently more sensitive to melanin concentration–stay within 5 µM of each other regardless of skin type. Since all subjects were age-matched, healthy males, and rested prior to imaging, similar oxyhemoglobin and deoxyhemoglobin measurements were expected.

Fig. 9

Estimates of (a) oxyhemoglobin foxy and (b) deoxyhemoglobin fdeoxy concentration for four subjects with increasingly darker skin type predicted by a homogeneous model (with and without melanin compensation) and by the present two layer model.

JBO_17_10_107006_f009.png

Figure 10 shows the effectiveness of the two layer model in predicting transcutaneous oximetry information in the presence of variable melanin content in the epidermis. It depicts estimates of total hemoglobin concentration and oxygen saturation denoted by fheme and SO2, respectively, for four subjects with increasingly darker skin type from Type I to Type IV on the Fitzpatrick skin pigmentation scale.20 The homogeneous model6 over predicts total hemoglobin concentration for darker subjects because melanin absorption artificially increase μa,skin(λ). Furthermore, the homogeneous model predicts artificially low oxygen saturations since both melanin and deoxyhemoglobin absorb more strongly in the low NIR. In fact, for the Type IV subject, the predicted oxygen saturation was 44% which was unrealistically low value. The SO2 values predicted by the homogeneous model ranged between 44% and 65% and decreased as skin pigmentation increased. On the other hand, the two layer model predicts realistic and similar values of SO2 between 69% and 75% for all four subjects regardless of pigmentation. Additionally, the SO2 values predicted by the present model shows no coupling to epidermal melanin concentration.

Fig. 10

Estimates of (a) total hemoglobin fheme and (b) oxygen saturation SO2 for four subjects with increasingly darker skin type predicted by a homogeneous model (with and without melanin compensation) and by the present two layer model.

JBO_17_10_107006_f010.png

A practical tissue oximetry system would both remove the effects of melanin from the transcutaneous oximetry signal and quantify pigmentation in the epidermis. Figure 11 shows estimates of epidermal melanin content for subjects of varying skin type predicted by spectral decomposition of the homogeneous absorption coefficient and the present two layer model. Figure 11(a) shows the wavelength dependent product of epidermal thickness d1 (assumed to be 70 µm) and calculated epidermal absorption coefficient μa,1 with the epidermal reduced scattering coefficient assumed to be μs,1. The epidermal absorption coefficient μa,1 exhibits monotonically decreasing behavior with wavelength, typical of melanin absorption. A melanin index can be defined as,

Eq. (8)

MI1=10,000fmel,
where fmel was estimated by spectral decomposition of μa,skin(λ). Additionally, we defined a melanin index based on results from the two layer model defined as

Eq. (9)

MI2=10,000μa,1(650nm)d1.
Figure 11(b) shows MI1 and MI2 for four subjects with increasing skin pigmentation. It indicates that the melanin index predicted by the homogeneous model did not correlate with skin type. In fact, MI1 Type I and IV subjects—which should represent extremes of MI1–were 423 and 460, respectively. The maximum melanin index of 569 was calculated for the subject with Type III skin. On the other hand, values of MI2 for Type I, II, III and IV subjects was 141, 258, 330, and 875, respectively. As expected, the melanin index increased with increasing skin type. In fact, Type IV skin exhibited a melanin index 6.2 times higher than Type I skin.

Fig. 11

(a) The product of epidermal thickness and absorption coefficient as a function of wavelength estimated by the present model. An epidermal thickness d1 of 70 µm and epidermal scattering coefficient μs,1 of 1mm1 were assumed. (b) Melanin index MIi predicted by a homogeneous model and the present two layer model.

JBO_17_10_107006_f011.png

Figure 12 illustrates estimates of the reduced scattering coefficient for four subjects with increasingly darker skin type predicted by a homogeneous model6 and the present two layer model as a function of wavelength. The reduced scattering coefficients predicted by both models have similar shape. However, the reduced scattering coefficient predicted by the two layer model was approximately 0.4mm1 lower than the prediction by the homogeneous model. This may be due to a coupling between melanin absorption and the tissue’s reduced scattering coefficient which both exhibit an exponential decay with wavelength. When considered homogeneous, the tissue’s bulk absorption coefficient is a combination between melanin and blood absorption. Subsequently, the bulk reduced scattering coefficient μs may be exaggerated to compensate for the artificially large absorption coefficient. However, the present two layer model segregates melanin absorption to the epidermal absorption coefficient μa,1 and so reduces coupling between melanin absorption and the dermal reduced scattering coefficient μs,1.

Fig. 12

Estimates of the reduced scattering coefficient for four subjects with increasingly darker skin type predicted by a (a) homogeneous model and (b) the present two layer model.

JBO_17_10_107006_f012.png

5.4.

In vivo Skin Measurements During Forearm Occlusion

Finally, Fig. 13 shows estimates of the oxygen saturation measured from lightly and strongly pigment subjects during a forearm occlusion experiment predicted by the standard homogeneous model and the present two layer model. Figure 13(a) indicates that estimates of SO2 measured from the strongly pigmented subject exhibit a baseline around 52%. On the other hand, the weekly pigmented subject exhibits a baseline of approximately 65%. The reasons for this discrepancy is described above. On the other hand, Fig. 13(b) indicates that the two layer model predicts a baseline value around 70%. Furthermore, both subjects exhibit a drop of approximately 10% during occlusion and a the typical hyperemic response after pressure was released. Figure 13 indicates that the present two layer model can effectively remove the effects of strong pigmentation from the SO2 signal during dynamic perfusion events.

Fig. 13

Oxygen saturation measured from lightly and strongly pigment subjects during a forearm occlusion experiment predicted by a (a) homogeneous model and (b) the present two layer model.

JBO_17_10_107006_f013.png

6.

Conclusion

This study describes a technique for analyzing spatial frequency dependent reflectance of two layer turbid media. The present model can be used to determine the optical thickness of a strongly absorbing superficial layer and the absorption and reduced scattering coefficient of a supporting semi-infinite supporting layer. An artificial neural network was used to map input optical properties to a spatial frequency dependent reflectance function of two layer media. The accurate and efficient forward method was used in conjunction with a gradient descent method to determine the optical properties of the two layer medium.

The neural network inverse method was applied to (i) simulated spatial frequency dependent reflectance of skin, (ii) a two layer silicone tissue simulating phantom, and (iii) in vivo human skin from subjects of varying skin type. Analysis of the simulated spatial frequency dependent reflectance of skin showed that the optical thickness of the epidermis and absorption and reduced scattering coefficients of the dermis could be determined independently and with minimal coupling. Additionally, the optical properties of the two layer tissue simulated phantom could be determined from spatial frequency dependent reflectance measurements. Finally, in vivo experiments showed that the present two layer model can be used to quantify melanin pigmentation and estimate hemoglobin concentration and saturation in humans of varying skin pigmentation.

Current work involves implementing a neural network approach which directly maps a measured spatial frequency dependent reflectance spectrum to optical properties as described in Eq. (7) of Ref. 54. This would allow for rapid generation of high resolution two dimensional chromophore maps similar to those reported in Ref. 6 but presenting information about two distinct tissue layers. However, present experiments with this approach indicate that the direct method may have been over-trained on sample data and is thus highly sensitive to experimental noise and offset. We are also pursuing human and animal model occlusion studies to show the utility of this approach over the full range of realistic vascular conditions associated with ischemia, hyperemia, inflammation, and, for example, necrosis. Once characterized in a controlled lab setting, the present algorithm can be deployed to analysis of clinical spatial frequency domain imaging data.

Acknowledgments

The authors gratefully acknowledge funding provided by the NIH SBIRs1R43RR030696-01A1, the NIH NIBIB Biomedical Technology Research Center (LAMMP: 5P-41RR01192), the Military Photomedicine Program, AFOSR Grant No. FA9550-08-1-0384, and the Beckman Foundation.

References

1. 

J. C. Hirschet al., “Near-infrared spectroscopy: what we know and what we need to know—a systematic review of the congenital heart disease literature,” J. Thorac. Cardiovasc.Surg., 137 (1), 154 –159 (2009). http://dx.doi.org/10.1016/j.jtcvs.2008.08.005 JTCSAQ 0022-5223 Google Scholar

2. 

S. J. EricksonA. Godavarty, “Hand-held based near-infrared optical imaging devices: a review,” Med. Eng. Phys., 31 (5), 495 –509 (2009). http://dx.doi.org/10.1016/j.medengphy.2008.10.004 MEPHEO 1350-4533 Google Scholar

3. 

S. J. Matcheret al., “Performance comparison of several published tissue near-infrared spectroscopy algorithms,” Anal. Biochem., 227 (1), 54 –68 (1995). http://dx.doi.org/10.1006/abio.1995.1252 ANBCA2 0003-2697 Google Scholar

4. 

BevilacquaF.et al., “Method and apparatus for performing quantitative analysis and imaging surfaces and subsurfaces of turbid media using spatially structured illumination,” U.S. patent 6,958,815 B2 (2003).

5. 

D. J. Cucciaet al., “Modulated imaging: quantitative analysis and tomography of turbid media in the spatial-frequency domain,” Opt. Lett., 30 (11), 1354 –1356 (2005). http://dx.doi.org/10.1364/OL.30.001354 OPLEDP 0146-9592 Google Scholar

6. 

D. J. Cucciaet al., “Quantitation and mapping of tissue optical properties using modulated imaging,” J. Biomed. Opt., 14 (2), 024012 (2009). http://dx.doi.org/10.1117/1.3088140 JBOPFO 1083-3668 Google Scholar

7. 

J. R. Weberet al., “Noncontact imaging of absorption and scattering in layered tissue using spatially modulated structured light,” J. Appl. Phys., 105 (10), 102028 (2009). http://dx.doi.org/10.1063/1.3116135 JAPIAU 0021-8979 Google Scholar

8. 

S. D. Koneckyet al., “Quantitative optical tomography of sub-surface heterogeneities using spatially modulated structured light,” Opt. Express, 17 (17), 14780 –14790 (2009). http://dx.doi.org/10.1364/OE.17.014780 OPEXFF 1094-4087 Google Scholar

9. 

J. R. WeberD. J. CucciaB. J. Tromberg, “Modulated imaging in layered media,” in Proc. IEEE Engineering in Medicine and Biology Society (EMBS'06), 6674 –6676 (2006). Google Scholar

10. 

S. Belangeret al., “Real-time diffuse optical tomography based on structured illumination,” J. Biomed. Opt., 15 (1), 016006 (2010). http://dx.doi.org/10.1117/1.3290818 JBOPFO 1083-3668 Google Scholar

11. 

N. RajaramT. H. NguyenJ. W. Tunnell, “Lookup table-based inverse model for determining optical properties of turbid media,” J. Biomed. Opt., 13 (5), 050501 (2008). http://dx.doi.org/10.1117/1.2981797 JBOPFO 1083-3668 Google Scholar

12. 

A. Bassiet al., “Spatial shift of spatially modulated light projected on turbid media,” J. Opt. Soc. Am. A Opt. Image Sci. Vis., 25 (11), 2833 –2839 (2008). http://dx.doi.org/10.1364/JOSAA.25.002833 Google Scholar

13. 

J. C. Davidet al., “Quantitation and mapping of tissue optical properties using modulated imaging,” J. Biomed. opt., 14 (2), 024012 (2009). http://dx.doi.org/10.1117/1.3088140 JBOPFO 1083-3668 Google Scholar

14. 

R. B. SaagerD. J. CucciaA. J. Durkin, “Determination of optical properties of turbid media spanning visible and near-infrared regimes via spatially modulated quantitative spectroscopy,” J. Biomed. Opt., 15 (1), 017012 (2010). http://dx.doi.org/10.1117/1.3299322 JBOPFO 1083-3668 Google Scholar

15. 

A. Yafiet al., “Postoperative quantitative assessment of reconstructive tissue status in a cutaneous flap model using spatial frequency domain imaging,” Plast. Reconstruct. Surg., 127 (1), 117 –130 http://dx.doi.org/10.1097/PRS.0b013e3181f959cc PRSUAS 0032-1052 Google Scholar

16. 

A. R. Young, “Chromophores in human skin,” Phys. Med. Biol., 42 (5), 789 (1997). http://dx.doi.org/10.1088/0031-9155/42/5/004 PHMBA7 0031-9155 Google Scholar

17. 

R. R. AndersonJ. A. Parrish, “The optics of human skin,” J. Investigat. Dermatol., 77 (1), 13 –19 (1981). http://dx.doi.org/10.1111/jid.1981.77.issue-1 JIDEAE 0022-202X Google Scholar

18. 

T. DwyerG. ProtaL. BlizzardR. AshboltM. R. Vincensi, “Melanin density and melanin type predict melanocytic naevi in 19–20 year olds of northern European ancestry,” Melanoma Res., 10 (4), 387 –394 (2000). http://dx.doi.org/10.1097/00008390-200008000-00011 MREEEH 0960-8931 Google Scholar

19. 

S. L. Jacques, “Origins of tissue optical properties in the UVA, visible, and NIR regions,” in OSA TOPS on advances in optical imaging and photon migration 2, 364 –371 (1996). Google Scholar

20. 

T. B. Fitzpatrick, “The validity and practicality of sun-reactive skin types I through VI,” Arch. Dermatol., 124 (6), 869 –871 (1988). http://dx.doi.org/10.1001/archderm.1988.01670060015008 ARDEAC 0003-987X Google Scholar

21. 

P. J. MattsP. J. DykesR. Marks, “The distribution of melanin in skin determined in vivo,” Brit. J. Dermatol., 156 (4), 620 –628 (2007). http://dx.doi.org/10.1111/bjd.2007.156.issue-4 BJDEAZ 1365-2133 Google Scholar

22. 

K. P. Nielsenet al., “The importance of the depth distribution of melanin in skin for DNA protection and other photobiological processes,” J. Photochem. Photobiol. B: Biol., 82 (3), 194 –198 (2006). http://dx.doi.org/10.1016/j.jphotobiol.2005.11.008 JPPBEG 1011-1344 Google Scholar

23. 

T. Tadokoroet al., “UV-induced DNA damage and melanin content in human skin differing in racial/ethnic origin,” FASEB J., 17 (9), 1177 –1179 (2003). http://dx.doi.org/10.1096/fj.02-0865fje FAJOEC 0892-6638 Google Scholar

24. 

W. F. W. Southwood, “The thickness of the skin,” Plastic Reconstruct. Surg., 15 423 (1955). http://dx.doi.org/10.1097/00006534-195505000-00006 PRSUAS 0032-1052 Google Scholar

25. 

Y. LeeK. Hwang, “Skin thickness of Korean adults,” Surgical and Radiologic Anatomy, 24 (3–4), 183 –189 (2002). http://dx.doi.org/10.1007/s00276-002-0034-5 0930-1038 Google Scholar

26. 

T. Gambichleret al., “In vivo data of epidermal thickness evaluated by optical coherence tomography: Effects of age, gender, skin type, and anatomic site,” J. Dermatol. Sci., 44 (3), 145 –152 (2006). http://dx.doi.org/10.1016/j.jdermsci.2006.09.008 JDSCEI 0923-1811 Google Scholar

27. 

J. Sandby-MollerT. PoulsenH. C. Wulf, “Epidermal thickness at different body sites: relationship to age, gender, pigmentation, blood content, skin type and smoking habits,” Acta DermatoVenereol., 83 (6), 410 –413 (2003). http://dx.doi.org/10.1080/00015550310015419 ADVEA4 0001-5555 Google Scholar

28. 

A. KrishnaswamyG. V. G. Baranoski, “A Biophysically Based Spectral Model of Light Interaction with Human Skin,” Comput. Graph. Forum, 23 (3), 331 –340 (2004). http://dx.doi.org/10.1111/j.1467-8659.2004.00764.x Google Scholar

29. 

R. O. Dudaet al., And a Scene Pattern Classification, Wiley, New York, Chichester (2001). Google Scholar

30. 

D. Warnckeet al., “A neural network based approach for determination of optical scattering and absorption coefficients of biological tissue,” J. Phys.: Conf. Ser., 178 (1), 012047 (2009). Google Scholar

31. 

M. C. Panet al., “Artificial neural networks-based diffuse optical tomography,” in Biomedical Optics, OSA Technical Digest (CD), Google Scholar

32. 

L. ZhangZ. WangM. Zhou, “Determination of the optical coefficients of biological tissue by neural network,” J. Mod. Opt., 57 (13), 1163 –1170 http://dx.doi.org/10.1080/09500340.2010.500106 JMOPEW 0950-0340 Google Scholar

33. 

T. J. Farrellet al., “A CCD and neural network based instrument for the non-invasive determination of tissue optical properties in-vivo,” Proc. SPIE, 2135 117 –128 (1994). http://dx.doi.org/10.1117/12.175987 Google Scholar

34. 

T. J. FarrellB. C. WilsonM. S. Patterson, “The use of a neural network to determine tissue optical properties from spatially resolved diffuse reflectance measurements,” Phys. Med. Biol., 37 (12), 2281 (1992). http://dx.doi.org/10.1088/0031-9155/37/12/009 PHMBA7 0031-9155 Google Scholar

35. 

T. J. Pfeferet al., “Reflectance-based determination of optical properties in highly attenuating tissue,” J. Biomed. Opt., 8 (2), 206 (2003). http://dx.doi.org/10.1117/1.1559487 JBOPFO 1083-3668 Google Scholar

36. 

D. Sharmaet al., “Evaluation of a fiberoptic-based system for measurement of optical properties in highly attenuating turbid media,” Biomed. Eng. Online, 5 49 (2006). http://dx.doi.org/10.1186/1475-925X-5-49 1475-925X Google Scholar

37. 

J. T. Bruulsemaet al., “Correlation between blood glucose concentration in diabetics and noninvasively measured tissue optical scattering coefficient,” Optics letters, 22 (3), 190 –192 (1997). http://dx.doi.org/10.1364/OL.22.000190 OPLEDP 0146-9592 Google Scholar

38. 

Q. WangK. ShastriT. J. Pfefer, “Experimental and theoretical evaluation of a fiber-optic approach for optical property measurement in layered epithelial tissue,” Appl. Opt., 49 (28), 5309 –5320 (2010). http://dx.doi.org/10.1364/AO.49.005309 APOPAI 0003-6935 Google Scholar

39. 

A. Cerussiet al., “In vivo absorption, scattering, and physiologic properties of 58 malignant breast tumors determined by broadband diffuse optical spectroscopy,” J. Biomed. Opt., 11 (4), 044005 (2006). http://dx.doi.org/10.1117/1.2337546 JBOPFO 1083-3668 Google Scholar

40. 

L. WangS. L. JacquesL. Zheng, “MCML—Monte Carlo modeling of light transport in multi-layered tissues,” Computer methods and programs in biomedicine, 47 (2), 131 –146 (1995). http://dx.doi.org/10.1016/0169-2607(95)01640-F CMPBEK 0169-2607 Google Scholar

41. 

F. Ayerset al., “Fabrication and characterization of silicone-based tissue phantoms with tunable optical properties in the visible and near infrared domain,” Proc. SPIE, 6870 687007 (2008). http://dx.doi.org/10.1117/12.764969 Google Scholar

42. 

R. Saageret al., “Multilayer silicone phantoms for the evaluation of quantitative optical techniques in skin imaging,” Proc. SPIE, 7567 756706 (2010). http://dx.doi.org/10.1117/12.842249 Google Scholar

43. 

M. FirbankM. OdaD. T. Delpy, “An improved design for a stable and reproducible phantom material for use in near-infrared spectroscopy and imaging,” Phys. Med. Biol., 40 (5), 955 (1995). http://dx.doi.org/10.1088/0031-9155/40/5/016 PHMBA7 0031-9155 Google Scholar

44. 

I. NishidateY. AizuH. Mishina, “Estimation of melanin and hemoglobin in skin tissue using multiple regression analysis aided by Monte Carlo simulation,” J. Biomed. Opt., 9 (4), 700 (2004). http://dx.doi.org/10.1117/1.1756918 JBOPFO 1083-3668 Google Scholar

45. 

M. Amourouxet al., “Diffuse reflectance spectroscopy’s sensitivity to a melanic layer thickness variations: an in vitro study on skin phantoms,” Proc. SPIE, 6628 66280Y (2007). http://dx.doi.org/10.1117/12.727851 JOSAAH 0030-3941 Google Scholar

46. 

J. R. Weberet al., “Noncontact imaging of absorption and scattering in layered tissue using spatially modulated structured light,” J. Appl. Phys., 105 (10), 102028 (2009). http://dx.doi.org/10.1063/1.3116135 JAPIAU 0021-8979 Google Scholar

47. 

V. V. Tuchin, Tissue Optics: Light Scattering Methods and Instruments for Medical Diagnosis, SPIE press, Bellingham, WA (2000). Google Scholar

48. 

I. V. MeglinskiS. J. Matcher, “Quantitative assessment of skin layers absorption and skin reflectance spectra simulation in the visible and near-infrared spectral regions,” Physiol. Measur., 23 (4), 741 –753 (2002). http://dx.doi.org/10.1088/0967-3334/23/4/312 PMEAE3 0967-3334 Google Scholar

49. 

L. DouvenG. Lucassen, “Retrieval of optical properties of skin from measurement and modeling the diffuse reflectance,” Proc. SPIE, 3914 312 –323 (2000). http://dx.doi.org/10.1117/12.388058 Google Scholar

50. 

T. Ryan, “Cutaneous circulation,” in Biochemistry and Physiology of the Skin, 817 –877 (1983). Google Scholar

51. 

K. Stenn, “The skin,” in Cell Tissue Biology, 541 –572 (1988). Google Scholar

52. 

M. J. van Gemertet al., “Skin optics,” IEEE Trans. Biomed. Eng., 36 (12), 1146 –1154 (1989). http://dx.doi.org/10.1109/10.42108 IEBEAX 0018-9294 Google Scholar

53. 

R. R. AndersonJ. A. Parrish, “The optics of human skin,” J. Invest. Dermatol., 77 (1), 13 –19 (1981). http://dx.doi.org/10.1111/jid.1981.77.issue-1 JIDEAE 0022-202X Google Scholar

54. 

D. YudovskyA. J. Durkin, “Spatial frequency domain spectroscopy of two layer media,” J. Biomed. Opt., 16 (10), 107005 (2011). http://dx.doi.org/10.1117/1.3640814 JBOPFO 1083-3668 Google Scholar

55. 

J. Sandby-MollerT. PoulsenH. C. Wulf, “Epidermal thickness at different body sites: relationship to age, gender, pigmentation, blood content, skin type and smoking habits,” ActaDerm-Venereol., 83 (6), 410 –413 (2003). http://dx.doi.org/10.1080/00015550310015419 ADVEA4 0001-5555 Google Scholar

56. 

Y. LeeK. Hwang, “Skin thickness of Korean adults,” Surg. Radiol. Anat., 24 (3–4), 183 –189 (2002). http://dx.doi.org/10.1007/s00276-002-0034-5 0930-1038 Google Scholar
© 2012 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2012/$25.00 © 2012 SPIE
Dmitry Yudovsky, John Quan M. Nguyen, and Anthony J. Durkin "In vivo spatial frequency domain spectroscopy of two layer media," Journal of Biomedical Optics 17(10), 107006 (19 October 2012). https://doi.org/10.1117/1.JBO.17.10.107006
Published: 19 October 2012
Lens.org Logo
CITATIONS
Cited by 25 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Absorption

Skin

Scattering

Spatial frequencies

Reflectivity

In vivo imaging

Spectroscopy

Back to Top