13.07.2015 Views

modelling and measurements of laser-skin thermal interaction - F9

modelling and measurements of laser-skin thermal interaction - F9

modelling and measurements of laser-skin thermal interaction - F9

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

University <strong>of</strong> LjubljanaFaculty <strong>of</strong> mathematics <strong>and</strong> physicsDepartment <strong>of</strong> physicsMODELLING AND MEASUREMENTS OFLASER-SKIN THERMAL INTERACTIONGRADUATE SEMINARMartin GorjanADVISER: pr<strong>of</strong>. dr. Martin Čopič∗†CO-ADVISER: dr. Marko Marinček §†26th February 2008AbstractThermal <strong>laser</strong>-<strong>skin</strong> <strong>interaction</strong> has principal role in <strong>laser</strong> dermatology, <strong>and</strong> it can besuccessfully modelled <strong>and</strong> measured. For <strong>modelling</strong>, the seminar covers all the steps involved,which range from determination <strong>of</strong> <strong>laser</strong> beam <strong>and</strong> <strong>skin</strong> optical properties, treatment<strong>of</strong> light propagation in the tissue, <strong>thermal</strong> diffusion <strong>of</strong> heat to <strong>thermal</strong> effects prediction.As a simple <strong>and</strong> efficient mean for measuring the <strong>thermal</strong> <strong>interaction</strong>, <strong>thermal</strong>imaging is introduced. The seminar concludes with some recent results <strong>and</strong> comparisonbetween <strong>modelling</strong> <strong>and</strong> <strong>measurements</strong> obtained at Fotona.∗ Faculty <strong>of</strong> mathematics <strong>and</strong> physics† Jožef Stefan Institute§ Fotona d.d.


Contents1 Introduction 32 Laser 42.1 Laser beam characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 Skin 53.1 Skin optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 Modelling light propagation 74.1 Photon transport theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74.1.1 Monte-Carlo simulations . . . . . . . . . . . . . . . . . . . . . . . . . . 74.1.2 Calculated absorbed light distributions . . . . . . . . . . . . . . . . . . 85 Modelling <strong>thermal</strong> <strong>interaction</strong> 105.1 Estimation <strong>of</strong> <strong>thermal</strong> <strong>interaction</strong> effects . . . . . . . . . . . . . . . . . . . . . 105.2 Thermokinetic selectivity theory . . . . . . . . . . . . . . . . . . . . . . . . . 115.2.1 Calculated TRTs <strong>and</strong> temperature evolutions . . . . . . . . . . . . . . 126 Measurements <strong>of</strong> <strong>skin</strong> temperature 126.1 Measured temperature distributions . . . . . . . . . . . . . . . . . . . . . . . 137 Conclusion 152


1 IntroductionSoon after the first <strong>laser</strong> was conceived in early 1960s, its great potential for certain medicalapplications was recognized. The qualities <strong>of</strong> <strong>laser</strong> light make it possible to deliver big amounts<strong>of</strong> optical energy to small targets on the surface or below the surface. The energy that isabsorbed by the tissue can change its properties in a way that was found to be beneficial inmany fields <strong>of</strong> medicine. Today, the use <strong>of</strong> <strong>laser</strong>s is already an established practice, amongothers, in ophthalmology, dentistry <strong>and</strong> dermatology.The exact details <strong>of</strong> all the effects a <strong>laser</strong> beam can have on the tissue <strong>and</strong> their causingmechanisms are still being elusive. The complexity <strong>of</strong> living tissue structure <strong>and</strong> workings[1] reflects in the intricacy <strong>of</strong> <strong>laser</strong>-tissue <strong>interaction</strong>s. Nevertheless, much progress has beenmade in both directions: many physiological effects were observed <strong>and</strong> catalogued [2], whilefive major <strong>interaction</strong> mechanisms were identified <strong>and</strong> explained [3], some have even beensuccessfully modelled [4]. Overview <strong>of</strong> the <strong>thermal</strong> <strong>interaction</strong>, for its pivotal role in <strong>laser</strong>dermatology, which is an important part <strong>of</strong> Fotona’s market share, is the topic <strong>of</strong> the presentseminar.Thermal effects <strong>of</strong> a <strong>laser</strong> beam incident upon the <strong>skin</strong> can be predicted knowing full timedependence <strong>of</strong> temperature distribution within the <strong>skin</strong>. Calculating it requires <strong>modelling</strong><strong>of</strong> the two processes: light propagation through the tissue, as the absorbed light producesheat, <strong>and</strong> <strong>thermal</strong> diffusion <strong>of</strong> the heat from exposed to surrounding tissue. Presuming thatthe <strong>skin</strong>’s optical parameters don’t change during <strong>laser</strong> pulse duration, both processes can betreated separately. [5, 6] The first is effectively done with the so called photon transport theory[7], which is covered in fourth chapter. The latter is described by the heat equation. Its specialapplication, a theory known as thermokinetic selectivity [8], commonly used to account for<strong>thermal</strong> effects on <strong>skin</strong> small structures like vascular lesions, hair papilla [9], pigment clusters[10, 11], tattoo particles [12], <strong>and</strong> others [13], is covered in the fifth chapter. In order toaccount for the photon propagation in the <strong>skin</strong>, both <strong>laser</strong> beam <strong>and</strong> <strong>skin</strong> optical propertieshave to be known [14], so these are covered in the previous two chapters. Finally, <strong>thermal</strong>imaging as an effective technique for <strong>measurements</strong> <strong>of</strong> <strong>skin</strong> temperature [15, 16] <strong>and</strong> also forestimation <strong>of</strong> optical <strong>and</strong> <strong>thermal</strong> parameters [17, 18], is presented in the last chapter.Figure 1: A map <strong>of</strong> <strong>laser</strong>-<strong>skin</strong> <strong>thermal</strong> <strong>interaction</strong> study.3


2 LaserLASER 1 light is produced in a <strong>laser</strong> device via selective amplification <strong>of</strong> light by stimulatedemission <strong>of</strong> radiation. The nature <strong>of</strong> light generation along with the presence <strong>of</strong> <strong>laser</strong> resonatorgives the <strong>laser</strong>light some unique properties: it has very well defined frequency (monochromacity),high spatial coherence <strong>and</strong> low divergence, meaning it can propagate in a well definedbeam. Consequently, its exceeding power per area per solid angle <strong>and</strong> per wavelength or spectralradiance dJdλ =d3 PdS dΩ dλmakes it capable <strong>of</strong> delivering great amounts <strong>of</strong> energy to smallspots <strong>and</strong> small parts <strong>of</strong> light spectra, which is very convenient for a range <strong>of</strong> medical <strong>and</strong>many other applications. [19]2.1 Laser beam characterizationLaser beams can be adequately characterized by a few parameters [19], which influence thebeam propagation in the tissue <strong>and</strong> also determine the type <strong>of</strong> <strong>interaction</strong> taking place (moreon <strong>interaction</strong>s follows in Sec. 5) [3].Beam wavelength λ is determined by the <strong>laser</strong> media, which usually also gives commonnames for <strong>laser</strong>s. The most important medical <strong>laser</strong>s are solid-state: neodymium yttriumaluminium-garnet(Nd:YAG) operating at λ = 1064 nm, frequency doubled Nd:YAG withλ = 532 nm, erbium YAG (Er:YAG) with λ = 2940 nm, alex<strong>and</strong>rite at λ = 755 nm <strong>and</strong>ruby at λ = 694 nm. Other important examples are gas CO 2 operating at λ = 10.6 µm <strong>and</strong>various diode <strong>laser</strong>s operating between λ = 800 − 1000 nm. [8]The spatial character <strong>of</strong> the <strong>laser</strong> beam is usually described by the transverse electromagnetic(or TEM) family <strong>of</strong> modes, with Gaussian beam (or TEM 00 , see Fig. 2a) being thediffraction limit for given wavelength. [19] However, because <strong>of</strong> heavy scattering in the tissue(more on it in Sec. 3.1), the beam angular characteristics are effectively r<strong>and</strong>omized immediatelyupon entering. Only the spatial energy distribution <strong>of</strong> the beam remains important - itcan be expressed as the spot size <strong>and</strong> spot shape. [7]Beam temporal character or pulse width is determined by the operating regime <strong>of</strong> the<strong>laser</strong> <strong>and</strong> with given size <strong>of</strong> <strong>laser</strong> media is closely connected with its peak power. A variety<strong>of</strong> regimes can produce pulses with vastly different widths <strong>and</strong> peak powers, as schematicallydepicted on Fig. 2a. Medical continuous wave (cw) <strong>laser</strong>s typically operate up to some watts.Quasi-cw <strong>laser</strong>s, which use pulsed pump sources, can produce pulses from microseconds tomillisecond widths, with typical peak powers <strong>of</strong> some kilowatts. Q-switched <strong>laser</strong>s modulatethe resonator losses so that all the energy stored in <strong>laser</strong> media is released in a few roundtriptimes, producing pulses <strong>of</strong> nanosecond lengths <strong>and</strong> peak powers in the megawatt range.Mode-locked <strong>laser</strong>s, where phases <strong>of</strong> many oscillating axial modes are matched, can produceextremely short pulses <strong>of</strong> a few femtoseconds in duration <strong>and</strong> corresponding peak power <strong>of</strong>gigawatts <strong>and</strong> more. [19]1 The word is an acronym for Light Amplification by Spontaneous Emission <strong>of</strong> Radiation4


Figure 2: a Spatial parameters <strong>of</strong> <strong>laser</strong> beam are its angular (green) <strong>and</strong> radial (dark green) distribution<strong>of</strong> power; for Gaussian beams, the first depends on wavelength, waist size (w 0 ) <strong>and</strong> divergence (θ), thelatter follows a Gaussian distribution. b Temporal parameters <strong>of</strong> <strong>laser</strong> beams are chiefly determined by the<strong>laser</strong> operating mode: continuous wave (blue), free-run (red), Q-switched (green) <strong>and</strong> mode-locked (voilet)produce pulses <strong>of</strong> vastly differing durations <strong>and</strong> peak powers.3 SkinSkin is the human body’s largest organ, measuring ∼ 2 m 2 <strong>and</strong> weighting ∼ 5 kg. It is made <strong>of</strong>two distinct layers, the outer epidermis <strong>and</strong> inner dermis. Below the dermis lies subcutaneouslayer, composed <strong>of</strong> proteins <strong>and</strong> adipose tissue (fat), but is not part <strong>of</strong> the <strong>skin</strong> itself. Fig. 3aschematically shows the layered structure <strong>of</strong> the <strong>skin</strong>. [1]Epidermis is again made <strong>of</strong> four to five layers, together measuring around 100 µm. Theyare made <strong>of</strong> keratinocytes, the cells which produce fibrous protein keratin, <strong>and</strong> compose 90%<strong>of</strong> the epidermis. Deeper layers are made <strong>of</strong> younger cells, which gradually travel outwards,transforming into a hard protective layer along the way. Scattered between the keratinocytesare melanocytes, cells producing the pigment melanin, <strong>skin</strong>’s main light absorber. Packets <strong>of</strong>it are called melanosomes <strong>and</strong> their main function is to protect the keratinocytes from UVlight coming from the Sun. All people have roughly equal number <strong>of</strong> melanocytes, but theyproduce different amounts <strong>of</strong> melanin. Its distribution can also vary (e.g. forming freckles,patches <strong>and</strong> age spots), hence different <strong>skin</strong> colors. [1]A much thicker dermis (∼ 1 mm), unlike the epidermis, is richly supplied with blood<strong>and</strong> nerves. Most <strong>of</strong> the dermis, around 75%, is made <strong>of</strong> two proteins: collagen <strong>and</strong> elastin,providing <strong>skin</strong>’s strength <strong>and</strong> elasticity. There are many structures in the dermis: tiny loops<strong>of</strong> blood vessels with important role in regulation <strong>of</strong> body temperature, gl<strong>and</strong>s, receptors forvarious stimuli, <strong>and</strong> hair follicles. Hairs protruding from hair follicles are made <strong>of</strong> keratinwhile melanin gives them color. The heavy presence <strong>of</strong> blood provides yet another importantpigment, hemoglobin, which is by contrast extremely scarce in the epidermis. [1]5


4 Modelling light propagationThe most fundamental approach on calculating light propagation is based on physics <strong>of</strong>Maxwell’s equations. However, because <strong>of</strong> the complexities involved, such approach can be <strong>of</strong>limited applicability. [7] Due to its simplicity, universality <strong>and</strong> good practical results, photontransport theory is frequently used instead. [3]4.1 Photon transport theoryA beam <strong>of</strong> light is described with its radiance J(r, s), expressing the power at the position rflowing in the direction s. Radiance is governed by the radiative transport equation as followsdJ(r, s)ds= −µ t J(r, s) + µ ∫sp(s, s’)J(r, s’)dΩ ′ . (1)4π 4πRadiance J(r, s) is being diminished proportionally to the total attenuation coefficient µ t =µ a + µ s which is a sum <strong>of</strong> absorption <strong>and</strong> scattering attenuation coefficients. On the otherh<strong>and</strong>, radiance J(r, s) is being augmented by the light scattered from direction s’ to directions, magnitude <strong>of</strong> which is described the phase function p(s, s’). Commonly used is a Henyey-Greenstein phase function, which can be expressed as [7]p(s, s’) = p(θ) = µ s 1 − g 2, (2)µ t (1 + g 2 − 2gcosθ) 3 2depending on the average cosine or anisotropy factor g = cos θ: g = 1 denotes purely forwardscattering, g = −1 purely backward <strong>and</strong> g = 0 isotropic scattering. Most tissues are mainlyforward scattering; for <strong>skin</strong> g was found to be around from 0.8 to 0.9. [3, 4]Radiative transport equation is difficult to solve directly <strong>and</strong> many approximations exist,among others the Kubelka-Munk, inverse adding-doubling <strong>and</strong> the diffusion approximation.Since the mean free path <strong>of</strong> the light propagating in <strong>skin</strong> is much smaller than the typicaldimensions involved, propagation quickly becomes effectively r<strong>and</strong>om. [7] The Monte-Carlosimulation is then the preferred method for numerically solving the radiative transport problem.[3, 4]4.1.1 Monte-Carlo simulationsIn Monte-Carlo simulation the <strong>laser</strong> beam is represented as a stream <strong>of</strong> a large number N<strong>of</strong> ”photons”, each having the coordinate, direction <strong>and</strong> <strong>and</strong> energy weight. Each photon isstatistically ray-traced through tissue following five steps (depicted on Fig. 4):1. Photon generation: Initial location <strong>and</strong> direction <strong>of</strong> propagation are r<strong>and</strong>omly determinedaccording the original beam.2. Pathway generation: Path to the next event is determined: the direction <strong>of</strong> propagationis determined according to the phase function p(θ). The traveled distance ∆r7


is determined by choosing a r<strong>and</strong>om number ξ between 0 <strong>and</strong> 1 <strong>and</strong> using logarithmicdistribution∆r = − log(ξ)µ t. (3)3. Absorption: The photon energy weight E is decreased after each event according toabsorption <strong>and</strong> scattering coefficient∆E = − µ aµ tE. (4)4. Elimination: When the photon energy weight goes below some predetermined threshold,its propagation is terminated <strong>and</strong> a new photon is launched.5. Detection: The absorbed energy ∆E is registered at the event coordinates after eachevent.At the end <strong>of</strong> procedure a three-dimensional distribution <strong>of</strong> absorbed energy is produced.More details <strong>of</strong> Monte-Carlo algorithm implementation for light propagation through <strong>skin</strong> canbe found in [6, 20].Figure 4: a Schematic representation <strong>of</strong> Monte-Carlo simulation steps <strong>of</strong> <strong>laser</strong> beam propagation in <strong>skin</strong>.It consists <strong>of</strong> five steps (see text above): 1 - photon generation, 2 - pathway generation, 3 - absorption,4 - elimination <strong>and</strong> detector count. Black dots represents scattering events <strong>and</strong> blue squares detector cellcounts. Grid size is exaggerated for clarity. b Side view <strong>of</strong> actual calculation for N = 100 photons.4.1.2 Calculated absorbed light distributionsWe used optical ray-tracing program Zemax, which features a highly efficient Monte-Carloimplementation, to obtain the distributions <strong>of</strong> absorbed light. Simulation consisted <strong>of</strong> threelayers <strong>of</strong> <strong>skin</strong> (epidermis, dermis <strong>and</strong> subcutaneous tissue) hit by Nd:YAG, λ = 1064 nm <strong>laser</strong>beams. Distributions <strong>of</strong> absorbed light were calculated for various spot sizes <strong>of</strong> square (or”top-hat”) collimated beams. We observed the absorbed power densities at the surface <strong>and</strong>in the depth for each spot size <strong>and</strong> also intraspot variations on the surface. The calculated8


distributions are presented for 4 mm spot size. On the surface distribution (Fig. 5a) the scattering”tails” can be observed, while the depth distribution reveals penetration depth (Fig.5b). Interestingly, aplot <strong>of</strong> maximum power densities versus spot size reveals strong dependence<strong>and</strong> saturation effect (Fig. 5c), which is explained by the fact that the scattered ”tail”is somewhate invariant to the spot size, making ratio <strong>of</strong> scattered energy inversely proportionalto spot size. Similar reasoning can be applied for spot-size dependence <strong>of</strong> penetrationdepth (Fig. 5d). Furthermore, because optical parameters <strong>of</strong> <strong>skin</strong> are strongly wavelengthdependant (see Sec. 3.1), <strong>laser</strong> wavelength as well as the <strong>skin</strong> type, i.e. its absorption <strong>and</strong>scattering coefficients, are also <strong>of</strong> essential importance in determining the beam penetrationdepth.Figure 5: Calculated absorbed light distribution cross-sections for 4 mm diameter beam viewed from a top,where scattering ”tails” can be observed <strong>and</strong> b side, where strong absorption in epidermis is evident. Bothc average superficial absorbed power density (horizontal line in a) <strong>and</strong> d penetration depth (vertical line inb) significantly depend on beam diameter.9


5 Modelling <strong>thermal</strong> <strong>interaction</strong>Light that is absorbed in tissue can induce many light-tissue <strong>interaction</strong>s. Five such <strong>interaction</strong>sare empirically classified today. Two governing parameters, which determine the<strong>interaction</strong>, were identified: the peak power density <strong>and</strong> exposure time, which can simply beequated with the pulse duration. [3] The five <strong>interaction</strong> zones are depicted on Fig. 6a.Figure 6: a Circles roughly represent regions <strong>of</strong> five <strong>laser</strong>-tissue <strong>interaction</strong> mechanisms that are classifiedtoday. Main determining parameter is the pulse duration (with interdependent power density), while totaldeposited energy is <strong>of</strong> lesser importance <strong>and</strong> mostly lies between 1 J <strong>and</strong> 1000 J. b For irreversible <strong>thermal</strong>effects to take place, the duration <strong>of</strong> the elevated temperature is important as well. Red line is empiricallyestimated critical temperature.5.1 Estimation <strong>of</strong> <strong>thermal</strong> <strong>interaction</strong> effectsThermal <strong>interaction</strong> is itself a family <strong>of</strong> biological effects following tissue exposure to increasedlocal temperatures. The irreversible cell effects range from hyperthermia (at 45 ◦ C), reductionin enzyme activity <strong>and</strong> cell immobility (at 50 ◦ C), denaturation <strong>of</strong> proteins <strong>and</strong> coagulation(at 60 ◦ C), vaporization <strong>and</strong> <strong>thermal</strong> ablation (at 100 ◦ C) to carbonization (> 150 ◦ C) <strong>and</strong>final melting (> 300 ◦ C). [3]The duration <strong>of</strong> the elevated temperature is important as well. Correlation <strong>of</strong> criticaltemperature <strong>and</strong> exposure time τ can be estimated by Arrhenius type damage integral [4]ln( ) ∫ C(0)τ= ΓC(τ) 0e −EaRT (r,t)dt, (5)using ratio <strong>of</strong> initial <strong>and</strong> final concentrations <strong>of</strong> native cells as the measure <strong>of</strong> damage. The twoparameters, frequency factor Γ <strong>and</strong> activation energy E a , must be experimentally obtained byobserving optical or histological changes <strong>of</strong> the irradiated tissue. A plot <strong>of</strong> critical temperatureversus duration <strong>of</strong> exposition is shown on Fig. 6b. [3]10


5.2 Thermokinetic selectivity theoryOnce distribution <strong>of</strong> absorbed light is obtained as presented in Sec. 4, the full time evolution<strong>of</strong> temperature in irradiated <strong>and</strong> surrounding tissue can be modeled by the 3D heat equation∂T (r, t)∂t= k ∇ 2 j(r, t)T (r, t) + , (6)ρc p ρc pusing calculated absorbed power density j(r,t) as the heat source; its temporal part is determinedby the <strong>laser</strong> pulse evolution with time. Skin <strong>thermal</strong> parameters are contained in the<strong>thermal</strong> conductivity k <strong>and</strong> volumetric heat capacity ρc p <strong>of</strong> the <strong>skin</strong>.Temperature increase in <strong>skin</strong>’s stronger absorbing structures can be much higher thanthe bulk temperature raise, depending on the particle absorption coefficient, its depth, size<strong>and</strong> <strong>laser</strong> pulse duration. [8, 4] A theory designated as thermokinetic selectivity exposes thestrong correlation between particle size <strong>and</strong> its <strong>thermal</strong> relaxation time (T RT ). [8] A onedimensionalmodel <strong>of</strong> spherical heat source (depicted on Fig. 7b) is used to calculated theT RT <strong>and</strong> temperature evolution with time depending on <strong>laser</strong> pulse duration. The governingequation for difference between particle <strong>and</strong> surrounding tissue temperature ∆T is then∂T (r, t)∂t= k (∂ρc p r 2 r∂r2 ∂T (r, t)∂r)a j(x, t)+ , (7)ρc pwhere a is the absorption ratio <strong>of</strong> a small structure versus bulk <strong>skin</strong> <strong>and</strong> j(x, t) the calculated<strong>laser</strong> intensity at depth x multiplied with the pulse temporal function. Thermal relaxationtime (T RT ) depends only on <strong>skin</strong> <strong>thermal</strong> diffusivitykρc p<strong>and</strong> particle radius r 0 , while obtainingactual temperatures also requires knowing a j(x, t).Figure 7: a Skin is an inhomogenous tissue, having embedded many small structures like melanosomes (1),hair follicles (2) <strong>and</strong> vascular lesions (3). b Their temperature increase due to <strong>laser</strong> irradiation is treatedwith thermokinetic selectivity theory, where small structures are represented as small source <strong>of</strong> radius r 0embedded in sphere <strong>of</strong> <strong>skin</strong> with radius R >> r 0 .11


5.2.1 Calculated TRTs <strong>and</strong> temperature evolutionsBy definition <strong>thermal</strong> relaxation time is the time it takes a particle at initial high temperatureT 0 to cool down to T 0 /e. I’s proportional to the second power <strong>of</strong> particle’s radiusT RT [µs] ∼ r 2 [µm 2 ]. [8] Laser pulses with durations shorter than particle’s T RT cause maximumtemperature raise, on the contrary, pulses with durations much longer than T RT don’tchange its temperature significantly, as can be seen on Fig. 8. The <strong>laser</strong> pulse duration isthus a crucial selectivity parameter when targeting small <strong>skin</strong> ”imperfections”.Figure 8: Time evolution <strong>of</strong> temperature difference in 10 µm particle with T RT ∼ 100 µs, irradiated withpulses <strong>of</strong> different durations <strong>and</strong> same energy density 1 J/mm 3 , leads to significantly different temperaturedifferences.6 Measurements <strong>of</strong> <strong>skin</strong> temperatureInfrared (or <strong>thermal</strong>) imaging is the method <strong>of</strong> choice for fairly accurate, fast <strong>and</strong> noncontact<strong>measurements</strong> <strong>of</strong> temperature distributions. It can be used to obtain in vivo timedependent temperature distributions from <strong>skin</strong> surface. [15, 16] Moreover, using the modelbasedparameter estimation by setting such model parameters for solution to fit the measuredtemperature spatial distributions or temporal course, some structural <strong>and</strong> optical parameterstogether with their uncertainties can be estimated. [17]In order to get quality images with well-defined timings, <strong>laser</strong> pulse <strong>and</strong> camera frametakingmust be exactly synchronized. Two general techniques <strong>of</strong> measurement are possible asdepicted on Fig. 9a:1. Immediately after <strong>laser</strong> pulse an usual recording is triggered. Each frame is taken sothe frame-rate determines the temporal resolution ∆t f = 1F P S2. Only the first frame after the <strong>laser</strong> pulse is taken. Repeating the measurement by varyingthe delay between the <strong>laser</strong> pulse <strong>and</strong> camera frame taking time, temporal resolutionsmuch lower than ∆t f can be reached.12


Figure 9: a Every frame (black high) can be taken immediately after <strong>laser</strong> pulse (green high) is triggered<strong>and</strong> the frame-rate determines resolution ∆t f . Conversely, only the first frame after the <strong>laser</strong> pulse istaken. The delay between them roughly determines the resolution ∆t p instead. b Thermovision camera(yellow) with IR optics (blue) was used to record distributions <strong>of</strong> <strong>skin</strong> surface temperature after <strong>laser</strong> pulseirradiation (green) emerging from a h<strong>and</strong>piece (grey).6.1 Measured temperature distributionsWe employed both techniques using a thermovision camera with spatial resolution <strong>of</strong> 320 by240 pixels, frame-rate <strong>of</strong> 50 Hz interlaced <strong>and</strong> detector integration time <strong>of</strong> 1.5 ms in a set-upas depicted schematically on Fig. 9b. The composite video output from the camera was usedfor recording while its sync part was used for triggering the Nd:YAG <strong>laser</strong>. Pulse durationsranged from 100 µs to 2 ms. By overlapping the camera integration time with the <strong>laser</strong> pulse<strong>and</strong> varying the delay as shown on lower part <strong>of</strong> Fig. 9a, temporal resolution ∆t p <strong>of</strong> less than100 µs was reached. Laser light was delivered to <strong>skin</strong> on arms <strong>of</strong> three different people by anoptical fiber ended with h<strong>and</strong>piece, which determined the beam shape <strong>and</strong> spot size.Surface temperature evolution following a single pulse (shown on Fig. 10) reveals timescale<strong>of</strong> <strong>skin</strong> surface cooling. The presence <strong>of</strong> hot-spots due to <strong>skin</strong> inhomogenity is clearly visibleon the first frame. Fig. 11 shows <strong>skin</strong> temperatures obtained with different spot sizes <strong>and</strong>the same fluence <strong>of</strong> 40 J/cm 2 where spot dependence can be noted. Fig. 12 shows <strong>skin</strong>temperatures obtained with same spot size <strong>and</strong> fluence <strong>of</strong> 40 J/cm 2 , but on different people.The role <strong>of</strong> local <strong>skin</strong> type is apparent.Additional data can be extracted with some computer image analysis. Assuming circularsymmetry <strong>and</strong> rolling out hot-spots, the angular average <strong>of</strong> surface temperature (shownFig. 13a) reveals scattering ”tails”, which can be used to validate the scattering parameters:coefficient µ s <strong>and</strong> anisotropy factor g. The evolution with time (not shown) further revealsthe cooling direction, which is mainly vertical. Average spot temperature versus spot size(Fig. 13b) show strong dependence with saturation around spot diameter <strong>of</strong> 6 mm, which isin decent agreement with calculations (Fig. 5c). Measured temporal evolution <strong>of</strong> the averagespot temperatures agree very well with the calculated ones (Fig. 13c), found by solvingplanar heat equation with initial temperature distribution obtained from the Monte-Carlo13


simulation. By using model-based parameter estimation, it appears that the relaxation time<strong>of</strong> the fast part (∼ 20 ms) is mainly determined by epidermis thickness, the relaxation time<strong>of</strong> the slow part (∼ 10 s, not shown) by penetration depth <strong>and</strong> the ratio <strong>of</strong> amplitudes by theratio <strong>of</strong> epidermis/dermis absorption coefficients. Finally, probing for relaxation time <strong>of</strong> <strong>skin</strong>hot-spots (Fig. 13d) gives a rough estimate <strong>of</strong> less than 100 µs, meaning their size is in therange <strong>of</strong> 10 µm.Figure 10: Skin surface temperature evolution after irradiation with a pulse <strong>of</strong> collimated Nd:YAG <strong>laser</strong>beam with 4 mm spot size.Figure 11: Skin surface temperature immediately after irradiation with a pulse <strong>of</strong> collimated Nd:YAG <strong>laser</strong>beam with different spot sizes.Figure 12: Skin surface temperature immediately after irradiation with a pulse <strong>of</strong> collimated Nd:YAG <strong>laser</strong>beam on different people.14


Figure 13: a Angular averaged temperature <strong>of</strong> <strong>skin</strong> surface immediately after <strong>laser</strong> pulse. b Spot averagedtemperatures <strong>of</strong> <strong>skin</strong> surface for different spot-sizes <strong>and</strong> same fluences. c Temporal evolutions <strong>of</strong> <strong>skin</strong>temperature measured with direct recording agree well with calculated ones. d Temporal evolutions <strong>of</strong>hot-spots measured with pump-probe technique gives estimation <strong>of</strong> relaxation time.7 ConclusionA basic overview <strong>of</strong> studying the <strong>thermal</strong> <strong>laser</strong>-<strong>skin</strong> <strong>interaction</strong>, which is an essential part <strong>of</strong><strong>laser</strong> dermatology, was given in the seminar. All the steps from <strong>laser</strong> beam characterization,<strong>skin</strong> optical properties determination, light propagation in the tissue, <strong>thermal</strong> relaxation time<strong>and</strong> damage <strong>modelling</strong> were concisely covered. Thermal imaging was presented as a measuringmethod, which can be used to validate the models, determine the <strong>skin</strong> properties or tune<strong>laser</strong> application parameters. Additionally, some calculated as well as measured results werepresented with good agreement obtained between them.15


References[1] Jorgen Serup <strong>and</strong> G. B. E. Jemec. H<strong>and</strong>book <strong>of</strong> Non-Invasive Methods <strong>and</strong> the Skin.CRC Press, 1995.[2] David H. Sliney <strong>and</strong> Stephen L. Trockel. Medical Lasers <strong>and</strong> Their Safe Use. Springer-Verlag, 1993.[3] Markolf H. Niemz. Laser-Tissue Interactions. Springer-Verlag, 1996.[4] Ashley J. Welch <strong>and</strong> Martin J. C. van Gemert. Optical-Thermal Response <strong>of</strong> Laser-Irradiated Tissue. Plenum Press, 1995.[5] Derek A. Affonce <strong>and</strong> Alex J. Fowler. The effect <strong>of</strong> <strong>thermal</strong> lensing during selectivephotothermolysis. Journal <strong>of</strong> Quantitative Spectroscopy <strong>and</strong> Radiative Transfer, 73(2-5):473–479, 2002.[6] J. J. Crochet, S. C. Gnyawali, Y. C. E. C. Lemley, <strong>and</strong> L. V. Wang anf W. R. Chen. Temperaturedistribution in selective <strong>laser</strong>-tissue <strong>interaction</strong>. Journal <strong>of</strong> Biomedical Optics,11(3), 2006.[7] A. Ishimaru. Wave propagation <strong>and</strong> scattering in r<strong>and</strong>om media. Academic Press, 1978.[8] David J. Goldberg. Laser <strong>and</strong> lights, 2005.[9] R. M. Williams, M. M. Christian, <strong>and</strong> R. L. Moy. Hair removal using the long-pulsedruby <strong>laser</strong>. Dermatologic Clinics, 17(2), 1999.[10] Gereon Huttmann, Cuiping Yao, <strong>and</strong> Elmar Endl. New Concepts in Laser Medicine:Towards A Laser Surgery With Cellular Precision. Medical Laser Application, 20(2):135–139, 2005.[11] Vincent KM Poon, Lin Huang, <strong>and</strong> Andrew Burd. Biostimulation <strong>of</strong> dermal fibroblastby sublethal Q-switched Nd:YAG 532 nm <strong>laser</strong>: collagen remodeling <strong>and</strong> pigmentation.Journal <strong>of</strong> Photochemistry <strong>and</strong> Photobiology B: Biology, 81(1), 2005.[12] E. F. Bernstein. Laser treatment <strong>of</strong> tattoos. Clinics in Dermatology, 24(1), 2006.[13] P. M. Becker-Wegerich, A. Kuhn, L. Malek, P. Lehman, M. Megahed, <strong>and</strong> T. Ruzicka.Treatment <strong>of</strong> nonmelanotic hyperpigmentation with the q-switched ruby <strong>laser</strong>. Journal<strong>of</strong> American Academy <strong>of</strong> Dermatology, 43(2), 2000.[14] R. R. Anderson <strong>and</strong> J. A. Parrish. The optics <strong>of</strong> human <strong>skin</strong>. Journal <strong>of</strong> InvestigativeDermatology, 77(1), 1981.[15] M. Le<strong>and</strong>ri, M. Saturno, L. Spodavecchia, G. D. Iannetti, <strong>and</strong> G. Crucco A. Truni. Measurements<strong>of</strong> <strong>skin</strong> temperature after infrared <strong>laser</strong> stimulation. Clinical Neurophysiology,36(4):207–218, 2006.16


[16] W. Verkryusse, W. Jia, W. Franco, T. E. Milner, <strong>and</strong> J. S. Nelson. Infrared Measurement<strong>of</strong> Human Skin Temperature to Predict the Individual Maximum Safe Radiant Exposure(IMSRE). Lasers in Surgery <strong>and</strong> Medicine, 39:757–766, 2007.[17] J. V. Beck <strong>and</strong> K. J. Arnold. Parameter estimation in engineering <strong>and</strong> science. JohnWiley <strong>and</strong> Sons, 1977.[18] S. A. Prahl, I. A. Vitkin, U. Bruggemann, B.C. Wilson, <strong>and</strong> R. R. Anderson. Determination<strong>of</strong> optical properties <strong>of</strong> turbid media using pulsed photo<strong>thermal</strong> radiometry. Physicsin Medicine <strong>and</strong> Biology, 37:1203–1217, 1992.[19] Orazio Svelto <strong>and</strong> David C. Hanna. Principles <strong>of</strong> Lasers. Springer, 4th edition, 1998.[20] Lihong Wang <strong>and</strong> Steven L. Jacques. Monte Carlo Modeling <strong>of</strong> Light Transport inMulti-layered Tissues in St<strong>and</strong>ard C, 1998.17

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!