Open Access
1 September 2010 Optical characteristics of cartilage at a wavelength of 1560 nm and their dynamic behavior under laser heating conditions
Alexander P. Sviridov, Andrey V. Kondyurin
Author Affiliations +
Abstract
A double-integrating-sphere system was used to measure the diffuse transmittance, diffuse reflectance, and collimated transmittance of cartilage and polyacrylamide hydrogel samples as a function of temperature under 1560-nm laser heating conditions. The dynamic behavior of the absorption and scattering coefficients and scattering anisotropy of the biomaterials was calculated by the inverse Monte Carlo method. The absorption coefficient of the cartilage and hydrogel samples proved to be linear in temperature. Raising the temperature of the cartilage samples to 80°C caused their absorption coefficient to decrease by some 25%. The temperature-induced change of the absorption spectrum of the interstitial water was found to be responsible for the clarification of the cartilage tissue observed to occur under 1560-nm laser heating conditions. The temperature field produced in the tissue by the laser energy deposited therein was calculated using a bioheat transfer equation with temperature-dependent parameters. The calculation results demonstrated that the temperature-induced changes of the optical parameters of biological tissues should be taken into account to make their 1560-nm laser treatment effective and safe.

1.

Introduction

A number of treatment procedures have recently been developed on the basis of a moderate thermal heating of biological tissues with laser radiation.1 The principal merit of laser radiation is the possibility it offers to convey energy to a local region in the biotissue being treated as required by the given scenario. To provide for the necessary therapeutic effect and safe conduct of the medical procedures involved, one should strictly control the temperature field induced in the tissue by the laser. This can only be done if one knows the optical and thermophysical characteristics of the tissue and their variations during the course of laser treatment. The temperature field is determined by both the parameters of the laser radiation proper, namely, its wavelength, the energy distribution in the laser beam, and exposure time, and the optical characteristics of the tissue at the laser wavelength. The set of the three optical parameters—the absorption coefficient μa , scattering coefficient μs , and anisotropy factor g at a given scattering phase function—wholly determines the distribution of light in the bulk of the tissue, and hence, the heat-source function associated with light absorption. It characterizes the rate of temperature rise in the laser-exposed zone and controls the spatial and temporal configurations of the temperature field. These optical parameters are usually determined by indirect methods based on room-temperature measurements of diffuse reflection and transmission of low-intensity radiation,2, 3, 4, 5, 6, 7 taken with integrating spheres. The results obtained are subsequently used to describe the propagation of high-intensity laser radiation in the biological tissue of interest on the assumption that the radiation and the tissue both remain unchanged during the course of laser heating.

However, the thermal action of laser radiation on biological tissues gives rise therein to a great many physicochemical processes, at both the microscopic and macroscopic levels.8 As a result, changes in the structure and composition of the tissues take place that can lead to substantial alterations in their optical parameters at the laser wavelength. For example, the anisotropy factor and scattering coefficient usually change as a result of either conformation transitions of the macromolecules forming the biological tissue or its thermal coagulation.9, 10, 11, 12, 13, 14, 15 The absorption coefficient depends on the concentration of the chromophore in the laser-exposed zone. Where the chromophore is water, the reduction of the absorption coefficient is sometimes attributed to the thermal diffusion of water from this zone.16, 17, 18 In a number of cases, the absorption dynamics of biological tissues during laser action thereon is governed by the temperature dependence of the absorption spectrum of the intratissue water. Numerous facts are available today in the literature that evidence variations of the absorption spectrum of liquid water under the effect of various lasers emitting radiation in the wavelength range from the UV to the far-IR region. To illustrate, Ediger 19 demonstrated that when ablating hydrated collagen by the ArF laser radiation (λ=193nm) , its absorption of radiation increased materially even during a short pulse (25ns) . Later on it was found20 that this effect was due to a change in the absorption spectrum of water in the UV region upon heating. Jansen 21 investigated the transmission dynamics of a water layer for the Ho:YAG (λ=2.12μm) , Ho:YSGG (λ=2.09μm) , and Tm:YAG (λ=2.01μm) laser radiation at various temperatures. They showed that when water was heated to 70°C , the absorption of radiation by it lowered perceptibly as a result of the shift of its 2ν2 absorption band (1.94μm) toward the short-wavelength region. The clarification of pure water was also observed to occur under the effect of high-power Er:YAG (2.94μm) and Er:YSGG (2.79μm) laser pulses,22, 23, 24 when its temperature reached a few thousand degrees centigrade. The experimental results presented by Shori 22 and Walsh and Cummings,24 together with their respective coworkers, confirmed the shift of the ν1 (3.05μm) and ν3 (2.87μm) absorption bands of liquid water toward shorter wavelengths as a result of the weakening of the hydrogen bond between the water molecules on heating. They suggested models for crater formation in the laser ablation of biological tissues that allowed for this absorption band dynamics. A similar dynamics was also observed by Sobol 25 who studied the radiometric response of cartilage tissues and cornea to the action of a tunable free electron laser.

In recent times, there has been tremendous progress in the development of medical fiber and solid state erbium glass lasers emitting radiation in the wavelength range 1.54– 1.56μm . They are being extensively used in medical practice for correction of cartilage tissues,26, 27 skin rejuvenation by multiple local laser heating actions (fractional photothermolysis),28, 29 and spine disk regeneration.30 This has stimulated investigations into the mechanisms governing the interaction between the above-mentioned laser radiation and biological tissues. Bagratashvili 31 demonstrated earlier that cartilage tissues undergo reversible clarification upon laser heating at a wavelength of 1.56μm . Several mechanisms were suggested that could explain such a clarification. First of all, this can be due to the effect of the thermal diffusion of water from the laser-exposed zone and deformation of the tissue absorption spectrum upon heating (absorption band shift and reduction of the integral absorption intensity as a result of the deaggregation of water clusters upon heating). All the mechanisms suggested were adequately substantiated,32 but their individual contributions to the tissue clarification effect observed were not determined. It is important to note that the dynamics of the transmission of laser radiation through the cartilage tissue studied was measured in the above work31 with a receiver of 10mm in diameter located immediately behind the tissue sample, which provided for a sufficiently large angular aperture for the collection of the transmitted light and, hence, for the adequate reliability of the energy measurements. However, the experimental scheme used was incapable of measuring the optical parameters of the cartilage tissue and their dynamics during the course of laser treatment. At the same time, such measurements, on the one hand, are of practical importance in selecting optimal laser treatment conditions, and on the other hand, they can help elucidate the mechanism behind the clarification of cartilage tissues at a laser wavelength of 1.56μm .

In essence, the present work was a continuation of the work by Bagratashvili 31 Its objective it was to measure the optical parameters of cartilage tissue and polyacrylamide hydrogel samples at a wavelength of 1.56μm and study their dynamics in the course of laser heating. Subject to consideration here is the problem of safety in medical laser procedures associated with the tissue clarification effects. The results obtained can be extended to other biological materials as well.

2.

Materials and Methods

The samples studied were prepared from calf nasal septal hyaline cartilage. The tissue was freed from perichondrium and mucous coat. The samples were prepared the day the experiment was conducted and stored in saline at a temperature of 4°C . Such a sample storage method minimizes the tissue decomposition as a result of natural decay processes.33

Polyacrylamide (PAA) hydrogel was selected to serve as a model system. This synthetic material is being frequently used as an equivalent of biological tissues.34 Its merits are good elasticity and thermal stability at temperatures up to 90°C . The PAA hydrogel is convenient to use because its samples can be synthesized to differ in water content and degree of cross-linking. The PAA hydrogels were synthesized through radical copolymerization of acrylamide (Lancaster Synthesis, Windham, New Hampshire) and N,N -methylenebisacrylamide (Amresco, Solon, Ohio) in the presence of catalytic amounts of ammonium persulfate (Amresco) and N,N,N,N -tetramethylenediamine (Lancaster Synthesis).35 Using this method, we synthesized a PAA hydrogel 1:19 in cross-linking degree and 70% by mass in water content that possessed optical and thermophysical characteristics similar to those of the cartilage.34

To find the temperature dependence of the optical parameters, we used a measuring system including a radiation source, two integrating spheres 8 cm diam, and a sample temperature measuring unit (Fig. 1 ). The inside surface of the spheres was coated with fine-dispersed barium sulfate (Kodak, Los Angeles, California), ensuring high reflectance in the optical and near-IR regions. The radiation intensity on the inside surface of each sphere was measured with InGaAs/InP photodiode (Telam, Moscow, Russia) operating in the spectral range from 1100 to 1620 nm. The diameters of detector ports were equal to 5 mm. Investigations into the integrating capacity of the spheres showed that variations in the direction of the 1560-nm radiation entering the sphere had practically no effect on the sensor signal being measured. Each sphere had two ports opposite to each other. The diameters of both ports were equal to 1 cm. Both spheres were mounted on movable supports so that their ports could be aligned. One more photodiode sensor was installed on the same axis at a distance of 120cm from the exit port of the second sphere. The data from all the sensors were collected with a Model USB-6009 multichannel analog-to-digital converter (National Instruments, Austin, Texas). The dark noise level amounted to 0.002 V at a characteristic signal level of 1–1.5 V. On the basis of these data and using calibration coefficients, we calculated the diffuse reflection (Rd) and diffuse transmission (Td) intensities, and also the intensity of light passing through the sample without deflection from the original direction (Tc) . The measurement accuracy of these quantities came to 0.2, 0.3, and 6×105mW , respectively.

Fig. 1

Schematic diagram of the double-integrating sphere setup for measuring diffuse transmittance, diffuse reflectance, and collimated transmittance.

055003_1_014005jbo1.jpg

The radiation source used was a fiber erbium glass laser (IRE-Polus, Moscow, Russia) 1.56μm in radiation wavelength and up to 5 W in output power. The transmission of radiation was by means of a quartz optical fiber 0.6 mm diam. We measured the spatial light intensity distribution at the exit from the fiber at different distances from its end face by scanning the light beam with a sensor whose sensing element was 25μm across. These measurements showed that the intensity distribution in the near and far fields could be well described by a Gauss formula. The exit face of the optical fiber was fixed inside the first integrating sphere at a distance of 7mm from the sample surface. The intensity distribution in the sample plane was described by

Eq. 1

I(r)=I0exp(r2WL2),
where WL=0.8mm is the radius of the laser beam and r is the distance from its center.

The cartilage tissue or PAA hydrogel sample of specified thickness (1.5mm) was placed between two 1-mm-thick glass plates. The sample temperature was measured with a K-type thermocouple (Omega Engineering, Stamford, Connecticut) 125μm in thickness that was inserted into the sample in order to locate its junction at the center of the radiation-exposed zone. Thanks to its small thickness, the thermocouple had practically no effect on optical measurements. Furthermore, under laser heating conditions, it was practically always in thermal equilibrium with the surroundings. Our detailed investigation of temperature dynamics at the beginning and end of laser pulse showed that the temperature jump caused by direct heating of thermocouple by laser light at 1.56μm , does not exceed 1.5°C . We consider this value to be rather small and acceptable for our experiments. The data from the thermocouple were collected with a model no. cDAQ-9172, 9211 analog-to-digital converter (National Instruments). The random error of temperature measurement was 0.1°C .

Thus, the experimental setup described allowed us to synchronously measure, in real time, the diffuse reflectance Rd , diffuse transmittance Td , collimated transmittance Tc , and sample temperature in the radiation-exposed zone.

2.1.

Calculation of Optical Parameters

The diffuse reflectance Rd , diffuse transmittance Td , and collimated transmittance Tc were calculated on the basis of a Monte Carlo algorithm similar to that used by Wang 36 and Pope and Wang.37 The photon flux reaching the sensitive element of the third sensor was calculated with due consideration for the angular distribution of the incident beam and the intensity distribution on the sample surface. The Henyey–Greenstein function generally accepted in biomedical practice was used to serve as a scattering phase function; 3×105 iterations of the Monte Carlo program were needed for calculation of the Rd , Td , and Tc parameters.

The determination of the unknown parameters μs , μa , and g reduced to the search for their values at which the sum of the squared deviations of the theoretical quantities Rdtheor , Tdtheor , and Tctheor from their experimental counterparts Rdexp , Tdexp , and Tcexp was minimal. This problem was solved by the method of linear regression using the modified Levenberg–Marquardt algorithm.38 The quantities Rdexp , Tdexp , and Tcexp for each sample were measured five times, which made it possible to calculate five statistically independent groups of the parameters μs , μa , and g and the corresponding standard deviations σμs , σμa , and σg that came to 0.4, 0.1, and 0.01cm1 , respectively. In these calculations, the values of μs , μa , and g were taken to be the same throughout the bulk of the sample.

3.

Experimental Results

Using the above-described measuring system, we found the temperature dependences of Rd , Td , and Tc for our cartilage tissue and PAA hydrogel samples under laser heating conditions. Characteristic results for the cartilage tissue are presented in Fig. 2 . As the laser heated the cartilage samples to 80°C , their Rd , Td , and Tc rose monotonically by approximately 5, 65, and 60%, respectively. The Rd , Td , and Tc values measured for the PAA hydrogel samples proved close to those for their cartilage tissue counterparts, as were their temperature dependences. The Rd , Td , and Tc values measured following the spontaneous cooling of the samples to room temperature and their repeated laser heating under the same conditions practically coincided with the values measured previously for both the cartilage tissue and PAA hydrogel samples. It should be noted that the slope of the Rd temperature-dependence curve for the cartilage tissue somewhat decreased in the temperature range from 45 to 60°C , this being true of all the samples and during the course of the repeated laser heating cycle as well. At the same time, no such deviations from linearity were observed in the PAA hydrogel samples. No Rd measurements were taken for water. Doubling the laser radiation intensity increased the heating rate of the samples by two times, but had practically no effect on the character of the temperature dependence of Rd , Td , and Tc .

Fig. 2

Temperature dependences of (a) Rd , (b) Td , and (c) Tc for the cartilage tissue (● — first laser heating cycle; ○ — repeated laser heating cycle).

055003_1_014005jbo2.jpg

The experimental results obtained for Rd , Td , and Tc allowed us to calculate the corresponding optical parameters: the scattering coefficient μs , absorption coefficient μa , and anisotropy factor g of the cartilage and PAA hydrogel. The temperature dependences of these parameters are presented in Fig. 3 . It can be seen that the coefficient μs for both materials varies within 5%, its increase being certainly monotonic. The anisotropy factor also increases monotonically and practically linearly from 0.83 to 0.9 as the cartilage sample is heated from 20 to 80°C . The absorption coefficient of the cartilage at room temperature is 5cm1 , and during the course of laser heating, it decreases linearly, the slope of its temperature curve amounting to some 0.025cm1°C . Thus, it decreases to 3.6cm1 upon laser heating to 80°C , which comes to some 75% of its original value. Approximately similar values of the coefficient μa and the same slope of its temperature curve are also observed for the PAA hydrogel. For comparison, we measured the temperature dependence of the absorption coefficient of liquid water during the course of laser heating at a wavelength of 1.56μm by a similar scheme using a liquid cell with a built-in thermocouple. In that case, the exit port of the second integrating sphere was shut with a standard diffuse reflector. The results obtained are presented in Fig. 4 . The absorption coefficient of water measured at room temperature is 12cm1 , which agrees with the available literature data.39 One can see that the absorption coefficient of water also suffers a linear drop; its relative change agrees well with similar variations for the cartilage and PAA samples under the same heating conditions, with due regard for their water content, and amounts to some 0.005°C1 .

Fig. 3

Temperature dependences of the coefficients (a) μs , (b) μa , and (c) g for the cartilage (●) and PAA hydrogel (○).

055003_1_014005jbo3.jpg

Fig. 4

Temperature dependence of the coefficient μa of liquid water (● — literature data;40 ○ — experimental data).

055003_1_014005jbo4.jpg

4.

Discussions

Our measurements showed that the laser heating of the cartilage tissue and PAA hydrogel from 25 to 80°C caused their optical parameters to undergo reversible changes. The reversible character of these changes is evidence that no irreversible processes, such as coagulation10, 13 or dehydration,17 took place in the samples under our laser heating conditions. We consider the following physicochemical processes as possible causes of the effects observed:

  • 1. the formation of a thermal lens in the sample

  • 2. mass transfer of water contained in the biological tissue or hydrogel as a result of the laser-induced temperature gradient therein

  • 3. reversible conformation transformations of the cartilage and hydrogel macromolecules

  • 4. alteration of the absorption spectrum (shape and integral intensity) of water upon heating

Note that these processes can occur simultaneously, each being capable of contributing to the alteration of the optical properties of biological tissues.

The laser heating of a biological tissue sample induces therein an inhomogeneous temperature field that can generally give rise to two phenomena: formation of a thermal lens and mass transfer of water contained in the biomaterial from the center of the laser-exposed zone to the periphery. On the one hand, the laser-induced temperature gradient produces a refractive index gradient along the laser beam path in the biomaterial, which can affect the readings of the sensors, especially those of the sensor used to measure Tc .13, 16 On the other hand, the presence of a temperature gradient causes some redistribution of water—the principal 1.56-μm chromophore of biomaterials—in the laser-exposed zone, which can also affect the readings of the sensors used to measure Td and Tc . But in our case, these effects are apparently unimportant. Both of these effects depend on the magnitude of the temperature gradient in the laser-exposed zone and ought, as a first approximation, to be linear in radiation power. Doubling the laser power in our experiments caused no perceptible changes in Td and Tc , provided that the sample temperature remained the same. This means that thermal lens formation and water redistribution in the laser-exposed zone have no significant effect on the Rd , Td , and Tc values being measured, hence on the optical parameters determined on their basis.

The similarity between the temperature dependences of the coefficients μs and g for the cartilage and polyacrylamide hydrogel bears witness to the fact that their changes are primarily caused by the alteration of the state of water contained in the samples. However, the slight difference between the μs(T) curves for the cartilage and PAA hydrogel in the temperature range 4560°C points to individual specificities of these systems. It is well known that hydrogels as well as biological tissues contain kinetically and thermodynamically unequal water molecules.40 The nuclear magnetic resonance spectra of such systems show two lines41, 42 instead of a single one, as is the case with liquid water. This fact points to the existence of water specifically linked to the polymer and also water not linked to the polymer (“free water”). There is also information that water changes from one state to the other upon heating. Therefore, the changes observed in μs and g are possibly due to a gradual change of the “linked” water to the “free” state under the effect of laser heating, as well as conformation changes of macromolecules consequent on changes in temperature and water state.

Worthy of note is the hypothesis that the laser heating of biological tissues and PAA hydrogels causes their dissolved gas to liberate in the form of bubbles, which increases light scattering. Indeed, the solubility of gases in water decreases with rising temperature. If their concentration is close to saturation, some dissolved molecules must change to the gaseous state upon heating. Obviously this process can take place without any temperature-controlled threshold, and in a reversible manner and rather rapidly. But to confirm this hypothesis requires additional investigation.

The substantial reduction of the absorption coefficient μa of the cartilage and PAA hydrogel with rising temperature is apparently due to the change of the absorption coefficient of liquid water. That this is so, is supported by both our measurements of the total transmission of liquid water upon laser heating (Fig. 4) and the well-known investigations into the deformation of the 1470-nm absorption band (ν1+ν3) of liquid water upon ordinary heating under equilibrium conditions.43 Indeed, the relative changes of the dependence μa(T) for the cartilage, PAA hydrogel, and water practically coincided at equal temperatures, which points to a common nature of the phenomenon. As follows from the work by Segtnan, 43 as temperature grows higher, the absorption band maximum of liquid water gradually shifts toward the short-wavelength region. These authors believe that such a deformation of the absorption band (1470 nm; ν1+ν3 ) of water upon ordinary heating under equilibrium conditions is due to the equilibrium change of one type of water clusters formed by several water molecules to another type of cluster consisting of a fewer number of molecules. This mechanism is probably also applicable to the description of the temperature dependence of the coefficient μa of any biomaterials, including cartilage tissues and PAA hydrogels. Our analysis of the above work43 showed that the temperature dependence of μa of water at a wavelength of λ=1.56μm proved practically linear, the slope of its curve being in good agreement with our measurements (Fig. 4).

When normalized to water concentration, the values of the coefficient μa measured for the PAA hydrogel and cartilage samples at room temperature turned out to be somewhat lower than the expected 78cm1 , namely, 55.5cm1 . This difference can be due to various factors.

First, account should be taken of the fact that the optical parameters determined are the result of solution of a complex inverse problem involving a great many parameters and a number of assumptions. For example, the use of the Henyey–Greenstein scattering phase function in computing these parameters is, in principle, unfounded. This function is of a phenomenological character but offers a number of advantages and conveniences. In many cases, it yields reasonable results, provided that the experimental data obtained are used to calculate other quantities, for example, the radiation power density distribution inside the scattering medium.

Second, some systematic error in determining the optical parameters can arise in measuring and modeling Tc . In our model, we tried to take account of the actual geometrical factors of the experimental scheme, the angular and spatial light distribution in the laser beam incident on the sample. However, our model assumed a geometrical character of light propagation and failed to allow for possible coherence effects affecting the wavefront in the near and far fields.

Third, the fact cannot be ruled out that the spectral properties of pure water can actually differ perceptibly from those of water contained in biological tissues and PAA hydrogels. Therefore, evaluating the value of the coefficient μa as the product of the value of this coefficient for liquid water by the water concentration corresponding to the biomaterial in hand is conditional enough. Calculations by the Monte Carlo method of the total diffuse transmission Td as a function of the sample thickness h (measured in centimeters) at measured optical parameters and laser beam parameters corresponding to our experiment yielded the exponential dependence Td=K exp(μeffh) , where μeff=11.6cm1 . The diffusion length μd=3μa[μa+(1g)μs] of photons wandering in the cartilage at room temperature amounted to 11.8cm1 at the parameters μa=5cm1 , μs=28cm1 , and g=0.85 , measured by us at room temperature. The coefficients μa and μeff agree well with the effective absorption coefficient of the cartilage tissue, μeff=10.8±0.5cm1 , measured independently by us34 by the method of remote laser IR radiometry. These facts indirectly bear witness to the reliability of the values of the optical parameters obtained.

When calculating the optical parameters, we assumed that the values of μs , μa , and g were the same throughout the bulk of the sample. But in actual fact, with the temperature field inside the sample being inhomogeneous and the quantities μs , μa , and g , temperature dependent, one cannot strictly state that they remain the same for all points of the sample upon laser heating. In the ideal case, to determine μs , μa , and g , one ought to solve a complex inverse problem, with due consideration given to the mutual influence of the light and temperature fields under varying optical parameter conditions. The problem is solvable, given specified a priori temperature dependences of μs , μa , and g , but requires substantial complication of the light and heat propagation models available at the given moment. In this work, we did not attempt such a task, restricting ourselves to the approximation presented above. Apparently such an approach was justified, for the temperature dependences of μs , μa , and g proved rather gradual and the temperature gradients proper in the region wherethrough the bulk of photons passed were not very great.

The variation of the optical parameters observed during the course of laser heating to 80°C at a wavelength of λ=1.56μm leads to a 30% increase in the depth of penetration of laser radiation, which is important in selecting irradiation regimes as regards their efficiency and safety. To evaluate the effect of the dynamics of μs , μa , and g on the temperature field in biological tissues, use was made of the model presented in our work.34 The calculation of the laser-induced three-dimensional temperature field was based on the solution of the classical heat conduction equation expressed in cylindrical coordinates,

Eq. 2

Tt=χ1rr(rTr)+χ2Tz2+f(r,z,t),0rR,0zZ,
where T(r,z,t) is the temperature at the point with the coordinates (r,z) at the instant t,f(r,z,t) is the laser heating rate (in Kelvin per second) of the medium, and χ=0.13 is its thermal diffusivity (in millimeters squared per second). The radiation intensity inside the sample was assumed to be described by the Beer law; hence, the function f(r,z,t) for light beams with a Gaussian energy distribution was found to be

Eq. 3

f(r,z,t)=(1Rd)μeffPρCpπWL2exp[(rWL)2]exp(μeffz)θ(t),θ(t)={1,0<ttimp0,t> timp},
where μeff is the effective laser radiation attenuation index, ρCp=3.5 is the specific heat capacity (in Joules per cubic centimeters per degree kelvin), P is the laser power (in watts), WL=0.1 is the laser beam radius (in millimeters), and Rd is the dimensionless diffuse reflectance measured with an integrating sphere and timp , is the time that the laser source is turned on. The spatial distribution of the laser power density in an absorbing optically inhomogeneous medium depends on μs , μa , and g and, strictly speaking, is not amenable to the Beer law. Nevertheless, the use of an exponential law with some generalized index μeff to describe the distribution of heat sources in an optically inhomogeneous absorbing medium seemed quite justified. We also assumed that μeff(T)=μd(T)=3μa(T){μa(T)+[1g(T)]μs(T)} , where the dependences μa(T) , μs(T) , and g(T) corresponded to their measured counterparts. It turned out that the dependence μeff(T) could be well described by the straight line μeff(T) =0.065T+13.8 , where μeff is measured in inverse centimeters and temperature (in degrees Celsius).

The temperature of the sample at the initial instant of time was taken to be homogeneously distributed throughout its bulk and equal to the ambient temperature T0 . Boundary conditions of third kind were specified at the boundary of the sample. The problem formulation provided for the independence of thermal diffusivity and specific heat capacity from temperature, coordinates, and time.

Within the scope of these notions, the solution of the thermal problem by the finite difference method showed a substantial difference between the temperature fields produced upon laser heating in cartilage tissues with the coefficient μeff remaining steadily at 11.8cm1 and with that varying linearly with temperature. The temperature was computed at each finite difference step of the thermal model, and then the optical properties adjusted according to the new temperatures to determine a new light distribution based on Eq. 3 with the process repeated for each time subsequent step. Figure 5 presents the distribution of this temperature difference as a function of sample depth and distance from the laser beam axis, calculated at a radiation power of 1 W, laser spot radius of WL=0.8mm , and heating time of 6 s, which approximately corresponds to conditions typical of laser nasal septum reshaping procedures. As one can see, the temperature difference on the front surface of the sample reaches some 12°C , this on the opposite surface amounting to 8°C . Such a temperature difference is important and should be taken into account when selecting laser treatment conditions for biological tissues.

Fig. 5

Difference between temperature fields calculated with and without allowance for the optical parameter dynamics.

055003_1_014005jbo5.jpg

5.

Conclusions

Short-term 1.56-μm laser heating of cartilage tissue to 80°C causes its optical parameters to undergo reversible changes. The values of the coefficients μa , μs , and g at room temperature come to 5.1±0.1 , 27.5±0.4 , and 0.83±0.01cm1 , respectively. As temperature rises to 80°C , the coefficients μs and g increase monotonically by 6–8%, while μa decreases by some 25%. The dependence μa(T) is almost linear, with the slope of its curve amounting to 0.025cm1°C .

The reduction of the coefficient μa observed to occur with rising temperature is due to the laser-induced change of the absorption spectrum of water contained in the biomaterials. The temperature dependences of the coefficient μa of the cartilage and PAA hydrogel, normalized to their water content, are characterized by one and the same slope, similar to that of liquid water, which is evidence of the same nature of the phenomenon in these materials.

Modeling the temperature field dynamics with and without allowance for the temperature dependences of the coefficients μa , μs , and g showed that the maximum difference between the respective temperature fields produced in the cartilage tissue upon its laser heating to 80°C in conditions typical of laser reshaping procedures reached 12°C . Allowing for the tissue clarification process increased the penetration depth of laser radiation. From these considerations, it is recommended that one should take account of the temperature dependence of the optical parameters of the cartilage tissue when modeling the temperature fields induced therein by 1.56-μm laser radiation and selecting laser treatment conditions for biological tissues in practice.

Acknowledgments

The authors thank A. A. Nikulchin, V. P. Minaev, and F. D Lepeshkin for their assistance in taking measurements with integrating spheres and also V. N. Bagratashvili for useful discussions and his interest in this work. The work was supported by the Russian Foundation for Basic Research (Grants No. 07-08-00448, No. 09-02-01408, and No. 09-02-00714).

References

1. 

M. H. Niemz, Laser-Tissue Interactions: Fundamentals and Applications, Springer, New York (2002). Google Scholar

2. 

W. Cheong, S. A. Prahl, and A. J. Welch, “A review of the optical properties of biological tissues,” IEEE J. Quantum Electron., 26 (12), 2166 –2185 (1990). https://doi.org/10.1109/3.64354 0018-9197 Google Scholar

3. 

C. Darling, G. Huynh, and D. Fried, “Light scattering properties of natural and artificially demineralized dental enamel at 1310nm,” J. Biomed. Opt., 11 (3), 034023 (2006). https://doi.org/10.1117/1.2204603 1083-3668 Google Scholar

4. 

M. Friebel, A. Roggan, G. Muller, and M. Meinke, “Determination of optical properties of human blood in spectral range 250to1100nm using Monte Carlo simulations with hematocrit-dependent effective phase functions,” J. Biomed. Opt., 11 (3), 034021 (2006). https://doi.org/10.1117/1.2203659 1083-3668 Google Scholar

5. 

S. J. Madsen, E. A. Chu, and B. J. F. Wong, “The optical properties of Nasal Cartilage,” IEEE JST Quantum Electron., 5 (4), 1127 –1133 (1999). https://doi.org/10.1109/2944.796339 Google Scholar

6. 

P. Parsa, S. L. Jacques, and N. S. Nishioka, “Optical properties of rat liver between 350 and 2200nm,” Appl. Opt., 28 (12), 2325 –2330 (1989). https://doi.org/10.1364/AO.28.002325 0003-6935 Google Scholar

7. 

Y. Du, M. Cariveau, X. Ma, G. W. Kalmus, and J. Q. Lu, “Optical properties of porcine skin dermis between 900nm and 1500nm,” Phys. Med. Biol., 46 (1), 167 –181 (2001). https://doi.org/10.1088/0031-9155/46/1/312 0031-9155 Google Scholar

8. 

A. J. Welch and M. J. C. van Gemert, Optical-Thermal Response of Laser-Irradiated Tissue, Plenum Press, New York (1995). Google Scholar

9. 

R. Agah, A. H. Gandjbakhche, M. Motamedi, R. Nossal, and R. F. Bonner, “Dynamics of temperature dependent optical properties of tissue: dependence on thermally induced alteration,” IEEE Trans. Biomed. Eng., 43 (8), 839 –846 (1996). https://doi.org/10.1109/10.508546 0018-9294 Google Scholar

10. 

R. Basu, B. J. F. Wong, and S. J. Madsen, “Wavelength dependent scattering of during Nd:YAG Laser heating of porcine septal cartilage,” Proc. SPIE, 4257 221 –230 (2001). https://doi.org/10.1117/12.434728 0277-786X Google Scholar

11. 

J. P. Cummings and J. T. Walsh Jr., “Erbium laser ablation: the effect of dynamic optical properties,” Appl. Phys. Lett., 62 (16), 1988 –1990 (1993). https://doi.org/10.1063/1.109512 0003-6951 Google Scholar

12. 

G. J. Derbyshire, D. K. Bogen, and M. Unger, “Thermally induced optical property changes in myocardium at 1.06μm,” Lasers Surg. Med., 10 (1), 28 –34 (1990). https://doi.org/10.1002/lsm.1900100108 0196-8092 Google Scholar

13. 

M. Ith, M. Frenz, and H. P. Weber, “scattering and thermal lensing of 2.12μm laser radiation in biological tissue,” Appl. Opt., 40 (13), 2216 –2223 (2001). https://doi.org/10.1364/AO.40.002216 0003-6935 Google Scholar

14. 

J. Laufer, R. Simpson, M. Kohl, M. Essenpreis, and M. Cope, “Effect of temperature on the optical properties of ex vivo human dermis and subdermis,” Phys. Med. Biol., 43 (9), 2479 –2489 (1998). https://doi.org/10.1088/0031-9155/43/9/004 0031-9155 Google Scholar

15. 

A. M. K. Nilsson, C. Sturesson, D. L. Liu, and S. Andersson-Engels, “Changes in spectral shape of tissue optical properties in conjuction with laser-induced thermotherapy,” Appl. Opt., 37 (7), 1256 –1267 (1998). https://doi.org/10.1364/AO.37.001256 0003-6935 Google Scholar

16. 

W.-C. Lin, M. Motamedi, and A. J. Welch, “Dynamics of tissue optics during laser heating of turbid media,” Appl. Opt., 35 (19), 3413 –3419 (1996). https://doi.org/10.1364/AO.35.003413 0003-6935 Google Scholar

17. 

F. Chambettaz, F. Marquis-Weible, and R. P. Salathe, “Effect of dehydration on optical properties of tissue,” Proc. SPIE, 1646 383 –390 (1992). https://doi.org/10.1117/12.137483 0277-786X Google Scholar

18. 

D. Zhu, Q. Luo, and J. Cen, “Effects of dehydration on the optical properties of in vitro porcine liver,” Lasers Surg. Med., 33 (4), 226 –231 (2003). https://doi.org/10.1002/lsm.10215 0196-8092 Google Scholar

19. 

M. N. Ediger, G. H. Pettit, and D. W. Hahn, “Enhanced ArF laser absorption in collagen target under ablative conditions,” Lasers Surg. Med., 15 (1), 107 –111 (1994). https://doi.org/10.1002/lsm.1900150114 0196-8092 Google Scholar

20. 

P. T. Staveteig and J. T. Walsh, “Dynamic 193-nm optical properties of water,” Appl. Opt., 35 (19), 3392 –3403 (1996). https://doi.org/10.1364/AO.35.003392 0003-6935 Google Scholar

21. 

E. D. Jansen, T. G. van Leeuwen, M. Motamedi, C. Borst, and A. J. Welch, “Temperature dependence of the absorption coefficient of water for midinfrared laser radiation,” Lasers Surg. Med., 14 (2), 258 –268 (1994). https://doi.org/10.1002/lsm.1900140308 0196-8092 Google Scholar

22. 

R. K. Shori, A. A. Walston, O. M. Stafsudd, D. Fried, and J. T. Walsh Jr., “Quantification and modeling of the dynamic changes in the absorption coefficient of water at λ=2.94μm,” IEEE JST Quantum Electron., 7 (6), 959 –970 (2001). https://doi.org/10.1109/2944.983300 Google Scholar

23. 

K. L. Vodopyanov, “Saturation study of H2O and HDO near 3400cm1 using intense picosecond laser pulses,” J. Phys. Chem., 94 (8), 5389 –5393 (1991). https://doi.org/10.1063/1.460500 0022-3654 Google Scholar

24. 

J. P. Cummings, “Effect of the dynamic optical properties of water on the midinfrared laser ablation,” Lasers Surg. Med., 15 (3), 295 –305 (1994). https://doi.org/10.1002/lsm.1900150310 0196-8092 Google Scholar

25. 

E. N. Sobol, A. P. Sviridov, M. S. Kitai, and G. S. Edwards, “Temperature alterations of the light absorption by cartilage and cornea under free electron laser radiation,” Appl. Opt., 42 (13), 2443 –2449 (2003). https://doi.org/10.1364/AO.42.002443 0003-6935 Google Scholar

26. 

A. B. Ovchinnikov, E. N. Sobol, V. N. Svistushkin, A. B. Shekhter, V. N. Bagratashvili, and A. P. Sviridov, “Laser septochondrocorrection,” Arch. Facial Plast. Surg., 4 (3), 180 –185 (2002). https://doi.org/10.1001/archfaci.4.3.180 Google Scholar

27. 

E. N. Sobol, A. V. Baskov, O. L. Zakharkina, and A. P. Sviridov, “Technology and equipement for laser reconstruction intervertebral disks,” Almanakh Clinic. Med., 57 (2), 242 –245 (2008). Google Scholar

28. 

D. Manstein, G. S. Herron, K. Sink, H. Tanner, and R. R. Anderson, “Fractional photothermolysis: a new concept for cutaneous remodeling using microscopic patterns of thermal injury,” Lasers Surg. Med., 34 (5), 426 –438 (2004). https://doi.org/10.1002/lsm.20048 0196-8092 Google Scholar

29. 

R. G. Geronemus, “Fractional photothermolysis: current and future applications,” Lasers Surg. Med., 38 (3), 169 –176 (2006). https://doi.org/10.1002/lsm.20310 0196-8092 Google Scholar

30. 

E. Sobol, O. Zakharkina, A. Baskov, A. Shekhter, I. Borschenko, A. Guller, V. Baskov, A. Omelchenko, and A. Sviridov, “Laser engineering of spine discs,” Laser Phys., 19 (4), 825 –835 (2009). https://doi.org/10.1134/S1054660X09040422 1054-660X Google Scholar

31. 

V. N. Bagratashvili, N. V. Bagratashvili, V. P. Gapontsev, G. S. Makhmutova, V. P. Minaev, A. I. Omel’chenko, I. E. Samartsev, A. P. Sviridov, E. N. Sobol’, and S. I. Tsypina, “Change in the optical properties of hyaline cartilage heated by the near-ir laser radiation,” Quantum Electron., 31 (6), 534 –538 (2001). https://doi.org/10.1070/QE2001v031n06ABEH001996 1063-7818 Google Scholar

32. 

V. N. Bagratashvili, E. N. Sobol, A. P. Sviridov, V. K. Popov, A. I. Omel’chenko, and S. M. Howdle, “Thermal and diffusion processes in laser-induced stress relaxation and reshaping cartilage,” J. Biomech., 30 (8), 813 –817 (1997). https://doi.org/10.1016/S0021-9290(97)00028-6 0021-9290 Google Scholar

33. 

J. I. Youn, S. A. Telenkov, E. Kim, N. C. Bhavaraju, B. J. F. Wong, J. W. Valvano, and T. E. Milner, “Optical and thermal properties of nasal septal cartilage,” Lasers Surg. Med., 27 (2), 119 –128 (2000). https://doi.org/10.1002/1096-9101(2000)27:2<119::AID-LSM3>3.0.CO;2-V 0196-8092 Google Scholar

34. 

A. V. Kondyurin and A. P. Sviridov, “Equivalent of a cartilage tissue for simulations of laser-induced temperature fields,” Quantum Electron., 38 (7), 641 –646 (2008). https://doi.org/10.1070/QE2008v038n07ABEH013845 1063-7818 Google Scholar

35. 

T. Tanaka, “Gels,” Sci. Am., 244 (31), 110 –138 (1981). https://doi.org/10.1038/scientificamerican0181-124 0036-8733 Google Scholar

36. 

L. Wang, S. L. Jacques, and L. Zheng, “MCML—Monte Carlo modeling of light transport in multi-layered tissues,” Comput. Methods Programs Biomed., 47 (2), 131 –146 (1995). https://doi.org/10.1016/0169-2607(95)01640-F 0169-2607 Google Scholar

37. 

K. Pope and L. Wang, “Deriving optical properties in the near infrared using an inverse Monte carlo program,” Proc. SPIE, 3914 300 –304 (2000). https://doi.org/10.1117/12.388055 0277-786X Google Scholar

38. 

J. J. Moré, “The Levenberg-Marquardt algorithm: implementation and theory,” Numer. Anal. Lect. Notes Math., 630 105 –116 (1977). https://doi.org/10.1007/BFb0067700 Google Scholar

39. 

L. Kou, D. Labrie, and P. Chylek, “Refractive indices of water and ice in the 0.65 to 2.5μm spectral range,” Appl. Opt., 32 (19), 3531 –3540 (1993). https://doi.org/10.1364/AO.32.003531 0003-6935 Google Scholar

40. 

V. J. McBrierty, F. X. Quinn, C. Keely, A. C. Wilson, and G. D. Friends, “Water in hydrogels. 4. poly(N-vinyl-2-pyrrolidone-methylmethacrylate)/saline systems,” Macromolecules, 25 4281 –4284 (1992). https://doi.org/10.1021/ma00043a006 0024-9297 Google Scholar

41. 

S. Katayama and S. Fujiwara, “Study of the freezing/thawing mechanism of water in polyacrylamide gel,” J. Phys. Chem., 84 (18), 2320 –2325 (1980). https://doi.org/10.1021/j100455a022 0022-3654 Google Scholar

42. 

P. McConville and J. M. Pope, “A comparision of water binding and mobility in contact lens hydrogels from NMR measurements of water self-diffusion coefficient,” Polymer, 41 9081 –9088 (2000). https://doi.org/10.1016/S0032-3861(00)00295-0 0032-3861 Google Scholar

43. 

V. H. Segtnan, S. Sasic, T. Isaksson, and Y. Ozaki, “Studies on the structure of water using two-dimensional near-infrared correlation spectroscopy and principal component analysis,” Anal. Chem., 73 (13), 3153 –3161 (2001). https://doi.org/10.1021/ac010102n 0003-2700 Google Scholar
©(2010) Society of Photo-Optical Instrumentation Engineers (SPIE)
Alexander P. Sviridov and Andrey V. Kondyurin "Optical characteristics of cartilage at a wavelength of 1560 nm and their dynamic behavior under laser heating conditions," Journal of Biomedical Optics 15(5), 055003 (1 September 2010). https://doi.org/10.1117/1.3484749
Published: 1 September 2010
Lens.org Logo
CITATIONS
Cited by 4 scholarly publications and 2 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
Back to Top