throbber
Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Optical properties of human skin
`
`Tom Lister
`Philip A. Wright
`Paul H. Chappell
`
`Petitioner Apple Inc. – Ex. 1033, cover p. 1
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Journal of Biomedical Optics 17(9), 090901 (September 2012)
`
`REVIEW
`
`Optical properties of human skin
`
`Tom Lister,a,b Philip A. Wright,a and Paul H. Chappellb
`aWessex Specialist Laser Centre, Salisbury District Hospital, Salisbury, SP2 8BJ, United Kingdom
`bUniversity of Southampton, School of Electronics and Computer Science, Southampton, SO17 1BJ, United Kingdom
`
`Abstract. A survey of the literature is presented that provides an analysis of the optical properties of human skin,
`with particular regard to their applications in medicine. Included is a description of the primary interactions of light
`with skin and how these are commonly estimated using radiative transfer theory (RTT). This is followed by analysis
`of measured RTT coefficients available in the literature. Orders of magnitude differences are found within published
`absorption and reduced-scattering coefficients. Causes for these discrepancies are discussed in detail, including
`contrasts between data acquired in vitro and in vivo. An analysis of the phase functions applied in skin optics,
`along with the remaining optical coefficients (anisotropy factors and refractive indices) is also included. The survey
`concludes that further work in the field is necessary to establish a definitive range of realistic coefficients for
`clinically normal skin. © 2012 Society of Photo-Optical Instrumentation Engineers (SPIE). [DOI: 10.1117/1.JBO.17.9.090901]
`
`Keywords: skin optics; radiative transfer theory; absorption; scatter; melanin; hemoglobin.
`Paper 12354V received Jun. 6, 2012; revised manuscript received Jul. 31, 2012; accepted for publication Jul. 31, 2012; published online
`Sep. 24, 2012.
`
`1 Introduction
`The color of human skin has long been used as a subjective
`adjunct to the detection and diagnosis of disease. More recently,
`the introduction of skin color measurements has extended this
`to include the potential for objective determination of skin fea-
`tures,1 including melanin and hemoglobin concentrations,2–6 the
`depth and diameter of blood vessels,7–9 the depth of pigmented
`skin lesions,10,11 the maturity and depth of bruises,12,13 and
`keratin fiber arrangements.14
`Such advances have proved invaluable for the advancement
`of skin laser treatments15 and photodynamic therapy,16–18 and
`have contributed to further advances in the diagnosis of cancer-
`ous and noncancerous skin lesions.10,19–21
`However, the success of these methods depends entirely
`upon adequate knowledge of the behavior of light as it impinges
`upon, and travels through, the skin. This article presents a
`description of the major interactions of visible light with skin
`and the principal skin features that contribute to these. This
`is followed by an analysis of published optical coefficients
`used in simulations of light transport through skin.
`
`2 Background
`
`2.1 Absorption
`
`Absorption describes a reduction in light energy. Within the
`visible region, there are two substances generally considered
`to dominate the absorption of light in skin: hemoglobin and
`melanin.
`Hemoglobin is the dominant absorber of light in the dermis.
`Normal adult hemoglobin (Hb A) is a protein consisting four
`polypeptide chains, each of which is bound to a heme.22 The
`heme in Hb A is named iron-photoporphyrin IX23,24 and is
`responsible for the majority of light absorption in blood. The
`free-electron molecular-orbital model describes this absorption
`
`as an excitation of loosely bound “unsaturation electrons” or
`“π-electrons” of the heme.25 Within the visible region, Hb A
`contains three distinctive peaks. The dominant peak is in the
`blue region of the spectrum and is referred to as the Soret
`peak or Soret band. Two further peaks can be distinguished
`in the green-yellow region, between 500 and 600 nm, that in
`combination with the Soret band cause Hb A to appear red.
`These are known as the α and β bands, or collectively as the
`Q-band, and have intensities of around 1% to 2% of the
`Soret band.26 The excitation levels of π-electrons vary, and
`therefore the positions and intensities of these bands vary
`with the ligand state of the heme (Fig. 1).
`Melanins are ordinarily contained within the epidermis and
`produce an absorption spectrum that gradually decreases from
`the ultraviolet (UV) to the infrared (IR) regions. In contrast to
`hemoglobin, the variation and complexity of melanins means
`that their detailed structures are not yet fully understood, despite
`intense research over the last five decades, and this broadband
`absorbance spectrum is still a topic of scientific debate.5,28,29 At
`present, the scientific consensus appears to gravitate towards a
`chemical disorder model.5,28–33 This model proposes that mela-
`nins consist of a collection of oligomers or polymers in various
`forms arranged in a disordered manner. This results in a number
`of absorption peaks that combine to create a broadband absor-
`bance effect28,30 (Fig. 2).
`Further absorption of light may be attributed to chromophores,
`such as bilirubin and carotene,34 lipids,35 and other structures,
`including cell nuclei and filamentous proteins36,37 Although
`the individual contributions from these secondary chromophores
`may be considered separately,13,38,39 most simulations group them
`into a single overarching value.40,41
`Despite its abundance in all tissues, water is not a significant
`absorber of light in the visible region, although its contribution
`has been considered when simulating skin color.42
`
`Address all correspondence to: Tom Lister, Wessex Specialist Laser Centre, Salis-
`bury District Hospital, Salisbury, SP2 8BJ, United Kingdom. Tel: 01722780104;
`E-mail: tom.lister@soton.ac.uk
`
`0091-3286/2012/$25.00 © 2012 SPIE
`
`Journal of Biomedical Optics
`
`090901-1
`
`September 2012 (cid:127) Vol. 17(9)
`
`Petitioner Apple Inc. – Ex. 1033, p. 1
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Lister, Wright and Chappell: Optical properties of human skin
`
`collagen is the principal filamentous protein of the dermis and
`occupies approximately 18% to 30% of its volume.45 Further
`scatter is attributed to melanosomes in the epidermis, cell nuclei,
`cell walls and many other structures in the skin that occur in
`smaller numbers.46
`Scatter from filamentous proteins has been approximated
`using a Mie solution to Maxwell’s equations applied to data
`from in vitro skin samples.47,48 This approach provides an
`increase in simulated scattering probability with increasing
`fiber diameter and decreasing wavelength. The dependence of
`scatter on fiber diameter suggests that the protein structures
`of the dermis, which may be 10 times as large as those in
`the epidermis,45,49 possess a greater scattering cross-section.
`This in part compensates for the lower number densities of fila-
`mentous proteins in the dermis. The scattering events that occur
`are mainly in the forward direction, meaning that on average,
`light that returns to the surface undergoes a large number of scat-
`tering events.50 One implication of the wavelength dependence
`of scatter is that blue and green light that has returned to the
`surface of the skin will have, on average, travelled less deeply
`than red light. This is considered the primary reason why blood
`vessels and pigmented nevi that are situated deeper within the
`skin are only able to absorb light from the red end of the
`spectrum and therefore appear bluer than their superficial
`equivalents.51,52
`The volume fraction of melanosomes in the epidermis varies
`typically from 1% in pale skin to 5% in darker skin,53 although
`one group has suggested greater values.54,55 However, despite
`their
`low numbers relative to keratins, melanosomes are
`approximately 10 times the diameter of the largest keratin struc-
`tures in the epidermis56 and possess a greater refractive index57
`(and therefore a greater difference in refractive index at their
`interface with skin). Melanin has been shown to contribute sig-
`nificantly to the degree of scatter within the epidermis.58,59 As
`well as the volume fraction, the distribution and size of melanin
`structures in the epidermis also vary with skin type. Thus, the
`total amount of scatter that occurs as a result of melanin in the
`epidermis can vary substantially between individuals,60,61
`although this is not always taken into account when simulating
`the effects of varying melanin concentration on skin color,42,62
`or when simulating laser treatments,15,63 for example.
`Blood normally occupies around 0.2% to 0.6% of the physical
`volume of the dermis2,6,54,64–66 depending upon its anatomical
`location. The vessel walls surrounding this blood, in addition
`to the walls of vessels that remain vacant, may occupy a similar
`volume. Dermal vessels vary in thickness and structure from
`capillaries of around 10 to 12 μm diameter at
`the epider-
`mal junction to terminal arterioles and post-capillary venules
`(approximately 25 μm in diameter) in the papillary dermis and
`venules (approximately 30 μm) in the mid-dermis.67 Furthermore,
`blood vessels occur in higher densities at particular depths, giving
`rise to so-called blood vessel plexi.67 The contribution to light
`scatter by these structures, inclusive of refraction effects, may
`be significant 68–70 and varies with location and depth, as well
`as between individuals.* Larger, deeper vessels may also
`contribute to the color of skin.
`
`*Assuming a reduced scattering coefficient of 0.5 mm−1 for blood and 40 mm−1
`for vessel walls at 633 nm, a 0.5% volume fraction of each contributes approxi-
`mately 0.2 mm−1 to the dermal reduced scattering coefficient, measured at
`around 1-5 mm−1 (Fig. 9). The contribution will be larger within blood vessel
`plexi.
`
`Absorption Spectra for Four Ligand States of
`Human Adult Haemoglobin
`
`Hb
`
`Hb02
`
`HbCO
`
`MetHb
`
`18
`
`16
`
`14
`
`12
`
`10
`
`8
`
`6
`
`4
`
`2
`
`Millimolar Absorptivities (L.mmol-1.cm-1)
`
`0
`450
`
`500
`
`550
`
`650
`600
`Wavelength (nm)
`
`700
`
`750
`
`Fig. 1 Absorption spectra of deoxyhemoglobin (Hb), oxyhemoglobin
`(HbO2), carboxyhemoglobin (HbCO), and methemoglobin (MetHb) in
`the visible region, from Ref. 27.
`
`Fig. 2 Colored lines show individual absorption spectra of tetramer
`subunits within melanin extracted from human epidermis. The average
`absorption spectrum of these is shown by thick black line, shifted up
`1.5 units for clarity. Thin black lines shifted down by one unit represent
`absorption spectra from monomer subunits. a:u: ¼ arbitrary units.
`Reprinted figure with permission from Ref. 30.
`
`2.2 Scattering
`
`As well as absorption, scattering contributes significantly to the
`appearance of skin. Scattering describes a change in the direc-
`tion, polarization or phase of light and is commonly portrayed as
`either a surface effect (such as reflection or refraction) or as an
`interaction with a small region whose optical properties differ
`from its surroundings (particulate scatter).
`It has been estimated that 4% to 7% of visible light is
`reflected from the surface of the skin, independent of wave-
`length and skin color.43,44 The remaining light is refracted as
`it passes from air into the skin.
`The primary sources of particulate scatter within the skin are
`filamentous proteins. Keratins are the filamentous proteins of
`the epidermis and form this layer’s major constituent, whereas
`
`Journal of Biomedical Optics
`
`090901-2
`
`September 2012 (cid:127) Vol. 17(9)
`
`Petitioner Apple Inc. – Ex. 1033, p. 2
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Lister, Wright and Chappell: Optical properties of human skin
`
`Scattering from the remaining structures of the skin, includ-
`ing cell walls, nuclei, and organelles,36 hairs and glands, is rarely
`of central interest to a study of skin optics. As a result, the con-
`tributions from these structures to the total measured scattering
`coefficients are not routinely considered separately.71
`
`4 Optical Coefficients of Skin
`A considerable amount of work has been carried out to deter-
`mine appropriate values of the RTT coefficients. Cheong et al.74
`described both direct (in vitro) and indirect (in vivo) methods of
`measuring absorption and scatter. A comprehensive analysis of
`the literature involving each method is presented here.
`
`4.1 Absorption Coefficients
`
`4.1.1 In vitro absorption coefficients
`
`to produce repeat
`Direct measurements have the potential
`measurements of a predetermined volume or section of skin
`and, unlike in vivo measurements, can include transmission
`data. However, the processes necessary to extract and prepare
`a skin sample cannot be carried out without altering its optical
`properties.
`The in vitro studies presented in Fig. 3 vary significantly in
`tissue-processing methodologies, measurement setup and the
`interpretation of data. For example, Jacques et al.’s work75
`included three methods of tissue preparation. The epidermis was
`separated from the dermis using a micro-cryotome for one set
`of skin samples, after mild thermal treatment in a water bath in
`another set, and was not separated in a third set. The same mild
`thermal treatment was used to separate the dermis and epidermis
`in Prahl’s work55 and a micro-cryotome was also applied in
`Salomatina et al.’s study.59 No separation of the epidermis
`was reported by Chan et al.76 or Simpson et al.77 Although
`Salomatina’s work shows greater absorption from the in vitro
`epidermis when compared to the dermis, the studies analyzed
`here do not demonstrate a clear distinction in absorption coeffi-
`cients reported between the methods of separation described,
`nor between those that separated the epidermis and those that
`did not.
`The level of hydration is likely to have varied considerably
`between the studies analyzed. Prahl55 and Jacques et al.75 soaked
`samples in saline for at least 30 min before carrying out mea-
`surements, during which the samples were placed in a tank of
`saline. Salomatina et al.59 also soaked their skin samples prior to
`measurement and sealed them between glass slides to maintain
`hydration. Chan et al.76 and Simpson et al.77 did not soak their
`samples prior to or during measurement. Jacques et al. reported
`that soaking the sample increases back-scattered reflectance,
`although the effects on the calculated absorption are not
`described. Chan et al. commented that dehydration may elevate
`the measured absorption coefficient. However,
`the greatest
`reported absorption coefficients are those from rehydrated tissue
`samples. From the information available, the effect of tissue
`hydration on the measured absorption coefficients is not clear.
`Data was interpreted using Monte Carlo simulations by
`Salomatina et al.,59 Simpson et al.77 and Graaf et al.,78 an
`adding-doubling technique by Prahl et al.,55 and by direct inter-
`pretation in Chan et al.’s76 and Jacques et al.’s75 studies. Both the
`methods described in the Monte Carlo simulations and Prahl’s
`adding-doubling technique are based upon assumptions of opti-
`cally homogeneous tissue layers, uniform illumination and no
`time dependence, and both are essentially discrete solutions
`to the radiative transport equation. The methods described con-
`trast in their approach to internal reflection for beams exiting the
`skin model and the adding-doubling method relies upon accu-
`rate representation of the angular distribution of beams exiting
`the thin layer upon which the model is built. It is not directly
`
`3 Simulating Light Transport Through Skin
`Optical simulations involving mathematical models of healthy
`human skin generally approximate the surface as perfectly
`smooth, although some computer graphics models have applied
`calculations of directional reflectance from rough surfaces.72
`Surface scattering effects (reflection and refraction) can be
`calculated for smooth surfaces using Fresnel’s equations and
`Snell’s law, respectively:
`
`
`ða − cÞ2
`ða þ cÞ2
`
`1 þ ½cða þ cÞ − 1Š2
`½cða − cÞ þ 1Š2
`
`:
`
`(1)
`
`R ¼ 1
`
`2
`
`Fresnel reflection (R) of unpolarized light from air to skin,
`where c ¼ cosðθiÞ, θi is the angle of incidence, a ¼ n2 þ c2 − 1
`and n is the refractive index of skin.
`
`
`
`(2)
`
`sin θi
`
`:
`
`1 n
`
`θt ¼ arcsin
`
`Angle of refraction (θt) at the skin’s surface calculated using
`Snell’s law.
`Within the skin, both absorption and scatter must be consid-
`ered simultaneously. These may be described in the classical
`approach by Maxwell’s equations, which consider the interac-
`tions between the electric and magnetic fields of light with mat-
`ter. However, an exact solution to Maxwell’s equations requires
`precise knowledge of each structure within the medium and
`becomes prohibitively complex for the case of human skin.
`The most commonly used approximation to Maxwell’s equa-
`tions in the field of skin optics is radiative transfer theory
`(RTT).73 This considers the transport of light in straight lines
`(beams),with absorption simulated as a reduction in the radiance
`of a beam and dependent upon the absorption coefficient (μa).
`The degree of scattering is described by the scattering coeffi-
`cient (μs), which considers both a loss of radiance in the direc-
`tion of the beam and a gain from beams in other directions, and
`the phase function (p), the probability that an individual beam
`will scatter in any particular direction. The reduced-scattering
`s ¼ μsð1 − gÞ,
`
`coefficient (μ 0s) combines these variables, i.e., μ 0
`where g is the anisotropy factor, the average cosine of the
`scattering angle θ:
`Z
`
`g ¼
`
`4π
`
`pðcos θÞ cos θdω;
`
`(3)
`
`where dω is a differential solid angle.
`In order for RTT to be valid, it must be assumed that any
`cause for increasing or decreasing the radiance of a beam other
`than that described by the absorption and scattering coefficients,
`including inelastic scatter (fluorescence or phosphorescence)
`and interactions between beams (interference), is negligible. The
`skin model must also consist of volumes that are homogeneous
`with regards to μs, μa and p, and that do not change over time.
`
`Journal of Biomedical Optics
`
`090901-3
`
`September 2012 (cid:127) Vol. 17(9)
`
`Petitioner Apple Inc. – Ex. 1033, p. 3
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Lister, Wright and Chappell: Optical properties of human skin
`
`Summary of Published Absorption Coefficients Measured In Vitro
`
`Summary of Published Absorption Coefficients Measured In Vivo
`
`10
`
`1
`
`0.1
`
`Dognitz 1998*
`
`Toricelli 2001* (mean)
`
`Meglinski 2002 epidermis†
`
`Meglinski 2002 dermis†*
`
`0.01
`
`Graaf 1993
`
`Zonios 2006
`
`Absorption coefficient (mm-1)
`
`Jacques 87
`Prahl 1988*
`Chan 1996*
`Salomatina 2006 epidermis*
`Salomatina 2006 dermis*
`Graaf 1993
`Simpson 1998*
`
`10
`
`1
`
`0.1
`
`0.01
`
`Absorption coefficient (mm-1)
`
`0.001
`350
`
`400
`
`450
`
`600
`550
`500
`Wavelength (nm)
`
`650
`
`700
`
`750
`
`0.001
`350
`
`Doornbos 1999† arm
`
`Doornbos 1999† forehead
`
`Svaasand 1995 epidermis
`
`Svaasand 1995 dermis
`
`Bosschaart 2011
`
`400
`
`450
`
`600
`550
`500
`Wavelength (nm)
`
`650
`
`700
`
`750
`
`Fig. 3 Summary of absorption coefficients available in the literature. In vitro data represents absorption coefficients from exsanguinated skin whereas
`dermal in vivo data is inclusive of blood absorption. *Data obtained from graphical presentation. †Data presented was not complete and required input
`of hemoglobin or water optical properties obtained from.35 The raw data is provided in the Appendix (Table 1).
`
`clear if or how these differences may have contributed to the
`higher absorption coefficients reported by Prahl et al.
`It is also of interest that Prahl et al.’s,55 Chan et al.’s76 and
`Salomatina et al.’s59 studies, whose samples varied in thickness
`between 60 and 780 μm, did not demonstrate a clear correlation
`between sample thickness and published absorption coefficients
`but Simpson et al.’s published absorption coefficients, which are
`an order of magnitude smaller than the other values analyzed
`involved much thicker samples (1500 to 2000 μm
`here,
`thick). Thus, differences between the published absorption
`coefficients across these studies may have resulted from varia-
`tions in the regions of skin investigated or the ability of the simu-
`lations to correctly account for boundary effects at the lower
`boundary.
`
`4.1.2 In vivo absorption coefficients
`
`Indirect measurements do not suffer from such changes in the
`properties of the interrogated skin volume, although care must
`be taken to consider variations in blood perfusion, for example,
`which may result from sudden changes in ambient temperature,
`the use of some drugs and even contact between the skin and the
`measurement device.79
`In general, absorption coefficients measured in vivo may be
`expected to be higher than in vitro values where the highly
`absorbing pigments from blood are removed from the samples.
`This is particularly true in the blue-green regions of the visible
`spectrum. Assuming a value of 0.5% blood volume in the
`−1 at 410 nm
`dermis, this would contribute approximately 8 cm
`
`−1 at 500 nm and 1.4 cm−1 at 560 nm
`(Soret band), 0.6 cm
`−1 at 700 nm (values calcu-
`(Q-band), but only around 0.05 cm
`lated from35). This contribution is not reflected in the literature.
`Absorption coefficients obtained from in vivo work show greater
`variation, but are not consistently higher than those obtained
`from in vitro work (Fig. 3).
`Absorption coefficients from Svaasand et al.,80 Zonios
`et al.,81 and Meglinski and Matcher42 clearly demonstrate the
`effect of blood on the measured absorption coefficients. Each
`
`study shows an absorption peak between 400 and 450 nm
`corresponding to the Soret band and a double peak at approxi-
`mately 540 and 575 nm corresponding to the α and β bands
`of oxyhemoglobin (see Fig. 1). There are, however, notable
`differences between the absorption coefficients produced from
`the three studies. Meglinski and Matcher and Svaasand et al.
`considered epidermal absorption coefficients separately to der-
`mal values. The reported values from Svaasand et al. are greater,
`and show a different spectral curve to those from Meglinski and
`Matcher. This is a direct result of Svaasand et al.’s inclusion of
`0.2% blood by volume in the calculation of epidermal absorp-
`tion coefficients, representing blood infiltrating the modeled
`epidermal layer from the papillae. Compared to Meglinski and
`Matcher’s dermal values and Zonios et al.’s absorption coeffi-
`cients for their skin model consisting a single layer, both of
`which also included the influence of blood, Svaasand et al.’s
`reported dermal absorption coefficients were consistently high.
`This is despite using a dermal blood volume fraction of 2%,
`compared to an average of 12% from Meglinski and Matcher’s
`study and a value of 2.6% in Zonios et al.’s work. The cause of
`this discrepancy is the variation in magnitude of the blood
`absorption coefficients applied across the three studies (Fig. 6,
`Appendix). Bosschaart et al.82 employed a diffusion approxima-
`tion technique to their data collected from neonates, effectively
`applying a single value of absorption across the skin volume.
`Their data is in close agreement to Meglinski and Matcher’s
`dermal absorption coefficients in the 530- to 600-nm range,
`but the contribution of melanin produces a relative increase
`in Bosschaart et al.’s values at shorter wavelengths.
`Data selected for analysis in this work involved “Caucasian”
`skin types only. Where stated, these studies involved skin types
`described as Northern European. Where not stated,
`it was
`assumed that such skin types were used except for the studies
`carried out by Zonios et al.81 and Torricelli et al.83 that were
`conducted in Southern Europe. However, the latter two studies
`did not report higher absorption coefficients, as may be expected
`from measurements on darker skin types. In Zonios et al.’s
`
`Journal of Biomedical Optics
`
`090901-4
`
`September 2012 (cid:127) Vol. 17(9)
`
`Petitioner Apple Inc. – Ex. 1033, p. 4
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Lister, Wright and Chappell: Optical properties of human skin
`
`work, this is primarily a result of the low values of blood absorp-
`tion coefficient applied. Torricelli et al.83 was the only group to
`apply time-resolved reflectance spectroscopy. This involves a
`prediction of the temporal spread of a laser pulse using a diffu-
`sion model. The values presented can only be as good as the
`diffusion model, and rely upon a wavelength dependence deter-
`mined from phantom measurements.84
`The absorption coefficients from both Graaf et al.’s78 and
`Doornbos et al.’s85 studies were lower than those from the
`remaining studies. Both studies involved an integrating sphere
`and multifiber probe, respectively, as did the higher values from
`Svaasand et al.40 and Meglinski and Matcher.42 Graaf et al. and
`Doornbos et al. applied a Monte Carlo Simulation and diffusion
`approximation respectively, as did Meglinski and Matcher and
`Svaasand et al. The multiple layered mathematical skin models
`that Svaasand et al. and Meglinski and Matcher applied when
`separately considering the effects of the epidermis and dermis
`may be a more accurate approach than the single homogeneous
`layer used in Graaf et al.’s and Doornbos et al.’s work. Although
`the cause of lower values is not clear, Graaf et al. commented
`that their absorption coefficient at 633 nm was “much smaller
`than expected from in vivo” results. Doornbos et al. did not com-
`ment directly on the cause of their low values, but mentioned
`that their “results resemble those of Graaf et al.”
`
`4.2 Scattering Coefficients
`
`4.2.1 In vitro scattering coefficients
`Of the studies analyzed here, both Prahl’s55 and Jacques et al.’s75
`studies describe a number of processes between tissue extraction
`and measurement that are likely to have had an effect on the
`measured reduced-scattering coefficient, including: exposure
`to a 55°C water bath for 2 min to aid with separating the epi-
`dermis from the dermis; freezing, cutting and stacking of 20-μm
`thick slices of the dermis; and soaking in saline to rehydrate and
`wash away any blood. The bloodless samples were then held
`
`between glass slides in a saline-filled tank and illuminated
`using a 633-nm laser. Freezing and drying, heating to remove
`the epidermis and deformation of skin samples have all been
`reported to change the measured values of scattering and absorp-
`tion coefficients.75,77,78 In particular, experimental work by Pick-
`ering et al.86 suggested that heating tissue to 55°C may increase
`the value of μs. Also, Jacques et al.75 commented that soaking
`the dermis (in saline) will increase the backscattered reflectance,
`and thus may increase the calculated scattering coefficient. In
`contrast, Chan et al.76 and Simpson et al.,77 whose reduced-scat-
`tering coefficients were substantially lower, reported minimal
`tissue processing (although Chan et al.’s specimens had pre-
`viously been frozen).
`A further source of disparity between in vivo data from ear-
`lier studies,55,75,78 which involved more tissue processing than
`the more recent in vivo data presented in Fig. 4,59,76,77 may
`have come from the choice of measurement setup. For example,
`Graaf et al.78 reported that discrepancies may arise when internal
`reflectance is not taken into consideration. Due to a larger dif-
`ference in refractive indices, this will have a greater effect for
`samples in air compared to samples in water or saline solution.
`All samples were placed between glass slides. However, only
`the earlier studies analyzed here, those which produced higher
`values of reduced-scattering coefficient, submerged the sample
`in water or saline.55,75
`Salomatina et al.’s
`study59 determined separately the
`reduced-scattering coefficients of the epidermis and dermis.
`Their data show that the epidermal reduced-scattering coeffi-
`−1 higher than the dermal
`cient was consistently 2 to 3 mm
`reduced-scattering coefficient over the visible spectrum. This
`suggests that studies that excluded the epidermis, such Chan
`et al.’s76 and Simpson et al.’s,77 should provide lower values
`of
`reduced-scattering coefficient
`than data obtained from
`studies in which the epidermis remained, such as Graaf
`et al.’s78 and Prahl’s.55 However, most probably due to a prevail-
`ing effect from the aforementioned influences, this is not the
`case.
`
`100
`
`Summary of Published Reduced Scattering Coefficients Measured
`in Vitro
`
`100
`
`Summary of Published Reduced Scattering Coefficients Measured
`in Vivo
`
`10
`
`1
`
`Dognitz 1998*
`
`Toricelli 2001* (mean)
`
`Reduced Scattering Coefficient (mm-1)
`
`600
`550
`500
`Wavelength (nm)
`
`650
`
`700
`
`750
`
`0.1
`350
`
`Graaf 1993
`
`Doornbos 1999 arm
`
`Doornbos 1999 forehead
`
`Svaasand 1995
`
`Bosschaart 2011
`
`Zonios 2006
`
`400
`
`450
`
`600
`550
`500
`Wavelength (nm)
`
`650
`
`700
`
`750
`
`10
`
`1
`
`Jacques 87
`
`Prahl 1988*
`
`Chan 1996*
`
`Salomatina 2006 epidermis*
`
`Salomatina 2006 dermis*
`
`Graaf 1993
`
`Simpson 1998*
`
`0.1
`350
`
`400
`
`450
`
`Reduced Scattering Coefficient (mm-1)
`
`Fig. 4 Summary of reduced- scattering coefficients available in the literature. In vitro data represents absorption coefficients from exsanguinated skin
`whereas dermal in vivo data is inclusive of blood absorption. *Data obtained from graphical presentation. †Data presented was not complete and
`required input of hemoglobin or water optical properties obtained from Ref. 35. Raw data is provided in the Appendix (Table 2).
`
`Journal of Biomedical Optics
`
`090901-5
`
`September 2012 (cid:127) Vol. 17(9)
`
`Petitioner Apple Inc. – Ex. 1033, p. 5
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Journal-of-Biomedical-Optics on 11/1/2018
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Lister, Wright and Chappell: Optical properties of human skin
`
`4.2.2 In vivo scattering coefficients
`In addition to their interpretation of Prahl’s in vitro data,55 Graaf
`et al.78 performed measurements of reflection on five male sub-
`jects with “white” skin at 660 nm using an LED source. Despite
`using a similar wavelength light source to Prahl’s 633 nm,
`reduced-scattering coefficients from Graaf et al.’s in vivo
`measurements were appreciably lower than their interpretation
`of in vitro data (Fig. 4). This is likely to be a result of the
`posthumous tissue processing performed in Prahl’s study as pre-
`viously described. However, this effect is not reflected across
`the literature, as in general reduced-scattering coefficients from
`in vivo studies were not substantially lower than those evaluated
`from in vitro studies, nor did they demonstrate an appreciable
`difference when considering the variation in reduced-scattering
`coefficients across the visible spectrum.
`Any solution involving two independent variables (such as
`μ 0
`s and μa) can suffer from nonuniqueness, where equivalent
`results can be obtained from two or more sets of input values
`(local minima). When applying RTT to skin, a simulated
`increase or reduction in reflection can be attributed to a change
`in either μa or μ0
`s. The reduced- scattering coefficients from
`Svaasand et al.’s study40 were considerably higher than any
`of the other in vivo studies assessed. This is in addition to
`their high values of absorption coefficient discussed in the pre-
`vious section. The paper stated that “the fact that the calculated
`[skin reflectance] values tend to be higher than the measured
`ones might indicate that the used values for the epidermal and
`dermal (reduced) scattering coefficients are somewhat too high.”
`The reduced-scattering coefficient from Svaasand et al.’s work
`was derived from a single data point at 577 nm measured by
`Wan et al.87 fitted to a simple μ0
`s ∝ wavelength
`−1 relationship
`and therefore may not be as reliable as data derived from a series
`of direct measurements. It should also be noted that the remain-
`ing studies that produced the highest reduced-scattering coeffi-
`cients analyzed here also provided the highest absorption
`coefficients, including both in vivo and in vitro data. Similarly,
`those studies presenting the lowest reduced-scattering coeffi-
`cients produced the lowest absorption coefficients (Fig. 4).
`Furthermore, when applying high values of dermal scattering
`to a minimization procedure, Verkruysse et al. demonstrated
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket