10.06.2015 Views

o_19nf61j2o10399kdgfam962n4a.pdf

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Int J Thermophys (2013) 34:139–159<br />

DOI 10.1007/s10765-013-1396-0<br />

Analytical Solution of Thermal Wave Models<br />

on Skin Tissue Under Arbitrary Periodic<br />

Boundary Conditions<br />

R. Fazlali · H. Ahmadikia<br />

Received: 10 August 2012 / Accepted: 9 January 2013 / Published online: 26 January 2013<br />

© Springer Science+Business Media New York 2013<br />

Abstract Modeling and understanding the heat transfer in biological tissues is important<br />

in medical thermal therapeutic applications. The biothermomechanics of skin<br />

involves interdisciplinary features, such as bioheat transfer, biomechanics, and burn<br />

damage. The hyperbolic thermal wave model of bioheat transfer and the parabolic<br />

Pennes bioheat transfer equations with blood perfusion and metabolic heat generation<br />

are applied for the skin tissue as a finite and semi-infinite domain when the<br />

skin surface temperature is suddenly exposed to a source of an arbitrary periodic<br />

temperature. These equations are solved analytically by Laplace transform methods.<br />

The thermal wave model results indicate that a non-Fourier model has predicted the<br />

thermal behavior correctly, compared to that of previous experiments. The results of<br />

the thermal wave model show that when the first thermal wave moves from the first<br />

boundary, the temperature profiles for finite and semi-infinite domains of skin become<br />

separated for these phenomena; the discrepancy between these profiles is negligible.<br />

The accuracy of the obtained results is validated through comparisons with existing<br />

numerical results. The results demonstrate that the non-Fourier model is significant in<br />

describing the thermal behavior of skin tissue.<br />

Keywords Laplace transforms · Non-Fourier · Pennes equation · Skin tissue ·<br />

Thermal wave<br />

R. Fazlali (B)<br />

Young Researchers & Elites Club, Hamedan Branch, Islamic Azad University, Hamadan, Iran<br />

e-mail: rz_fazlali@yahoo.com; rz_fazlali@iauh.ac.ir<br />

H. Ahmadikia<br />

Department of Mechanical Engineering, University of Isfahan, 81746-73441 Isfahan, Iran<br />

e-mail: ahmadikia@eng.ui.ac.ir<br />

123


140 Int J Thermophys (2013) 34:139–159<br />

List of Symbols<br />

Variables<br />

A Frequency factor (1/s)<br />

C Thermal wave speed in the medium (m · s −1 )<br />

c b Specific heat of blood (J · kg −1 · ◦C −1 )<br />

c t Specific heat of tissue (J · kg −1 · ◦C −1 )<br />

E a Activation energy (J · kmol −1 )<br />

H Heaviside theta function<br />

I 0 Zero rank-modified Bessel function<br />

I 1 First rank-modified Bessel function<br />

k Thermal conductivity of tissue (W · m −1 · ◦C −1 )<br />

L Skin thickness (m)<br />

q ext Heat generated (W · m −3 )<br />

q met Metabolic heat generation (W · m −3 )<br />

R Universal gas constant (J · mol −1 · K −1 )<br />

T Tissue temperature ( ◦ C)<br />

T a Blood temperature ( ◦ C)<br />

T c Minimum temperature of skin surface boundary condition ( ◦ C)<br />

T s Maximum temperature of skin surface boundary condition ( ◦ C)<br />

t Time (s)<br />

t 0 Period time (s)<br />

x Distance along the surface (m)<br />

U Unit step function<br />

Greek Symbols<br />

α Thermal diffusivity (m 2 · s −1 )<br />

Δ Duration time of a period that the boundary condition of skin surface is<br />

equal to T s (s)<br />

δ Dirac delta function<br />

ρ b Density of blood (kg · m −3 )<br />

ρ t Density of skin tissue (kg · m −3 )<br />

τ q Thermal relaxation time (s)<br />

Ω Dimensionless thermal damage<br />

ϖ b Blood perfusion rate (ml · ml −1 · s −1 )<br />

Subscripts<br />

b<br />

P<br />

t<br />

Blood<br />

Pennes equation<br />

Tissue<br />

123


Int J Thermophys (2013) 34:139–159 141<br />

1 Introduction<br />

The skin is the most extensive living organ of the human body. The skin consists<br />

of three layers: epidermis, dermis, and hypodermis (fat subcutaneous tissue). Its<br />

contribution to the body is essential and includes sensory, thermoregulation, host<br />

defense, etc. Advances in laser, microwave, and similar technologies have led to<br />

recent developments in the thermal treatment of skin diseases and injured skin tissue<br />

such as skin cancer and skin burn. Understanding heat transfer and related thermomechanical<br />

properties in soft tissue such as skin is important in medical applications.<br />

The thermal behavior of heat transfer in skin tissue is a kind of heat conduction<br />

problem that is coupled with complicated physiological problems such as blood<br />

circulation, metabolic heat generation, and other processes. Skin is affected by<br />

many factors such as age, gender, etc. Furthermore, simple elastomeric constitutive<br />

models are not suitable to describe the complicated mechanical behavior of the<br />

skin [1].<br />

Solving the heat transfer problem in soft organs such as skin has attracted the attention<br />

of many researchers. The Fourier model (Pennes equation) introduced by Pennes<br />

[2] and some non-Fourier models such as the thermal wave model of bioheat transfer<br />

and the dual-phase-lag (DPL) model were applied in modeling bioheat transfer across<br />

the tissue.<br />

One of the most complicated problems in the heat transfer of skin tissue is blood<br />

perfusion. Arkin et al. [3] have investigated the effect of blood perfusion in heat transfer<br />

on the tissues. They argued that the Pennes interpretation of the vascular contribution<br />

to heat transfer in perfuse tissues fails to account for the actual thermal equilibration<br />

process between the flowing blood and the surrounding tissue. Lang et al. [4] used<br />

the nonlinear three-dimensional heat transfer model based on temperature-dependent<br />

blood perfusion in order to predict the temperature distribution. Xu et al. [5] used<br />

the Pennes equation for modeling heat transfer in skin tissue. They applied the Green<br />

function method for solving the Pennes equation.<br />

Zhao et al. [6] solved the one-dimensional Pennes equation with a two-level finite<br />

difference scheme. They compared the numerical and experimental results in order<br />

to validate the new numerical scheme. Shih et al. [7] adopted the Laplace transform<br />

method for solving the Pennes equation with an in-surface sinusoidal heating condition.<br />

They observed that the temperature oscillation in the initial period caused<br />

by the sinusoidal heating on the skin was unstable. Liu and Xu [8] performed the<br />

blood perfusion estimation by the phase shift method in the temperature response<br />

to sinusoidal heating for the Pennes equation. In their research, the sinusoidal heating<br />

flux was considered constant, while this parameter was changeable in Ref. [7].<br />

Erdmann et al. [9] studied the optimization of the temperature distribution for regional<br />

hyperthermia based on the Pennes model by the finite element method. Ng et al.<br />

[10,11] predicted the skin burn injury and thermal profiles within heated human<br />

skin using the boundary element method. Durkee et al. [12] presented the exact<br />

analytical solution to the multiregional time-dependent bioheat equation (Pennes<br />

equation).<br />

123


142 Int J Thermophys (2013) 34:139–159<br />

Fourier’s law assumes that any thermal distribution on the body is instantaneously<br />

felt throughout the body. This means that the propagation speed of thermal distribution<br />

is infinite. This assumption is reasonable in the majority of engineering applications.<br />

This assumption is not true in the particular thermal condition, where heat conduction<br />

shows a non-Fourier feature or thermal wave phenomena, or hyperbolic heat conduction.<br />

The phase-lag behavior in the thermal wave may be justified if the layer has a<br />

very small thickness [13] or the time scale of the problem is very short [14]. Hader<br />

et al. [15] investigated the thermal behavior of a thin slab, under the effect of a fluctuating<br />

volumetric thermal disturbance described by the hyperbolic and dual-phase-lag<br />

heat conduction models. They found that the use of non-Fourier models is essential at<br />

high frequencies of the volumetric disturbance. The validity of using the microscopic<br />

hyperbolic heat conduction model under a harmonic fluctuating boundary heating<br />

source was investigated by Naji et al. [16]. They found that the use of the microscopic<br />

hyperbolic heat conduction model is essential when the value of the angular velocity<br />

of the fluctuating temperature is greater than 1 × 10 9 rad · s −1 for most metallic layers<br />

independent of their thickness.<br />

The modification of Fourier’s law presented by Cattaneo [17] and Vernotte [18] is<br />

a linear extension of the Fourier equation. The hyperbolic heat conduction model was<br />

extended to describe the thermal behavior of an anisotropic material by Al-Nimr and<br />

Naji [19]. In general, when materials (such as skin tissue) have a large relaxation time,<br />

the thermal wave phenomenon is observed in heat transfer conduction. Liu et al. [20]<br />

introduced a general form of the thermal wave model of bioheat transfer (TWMBT)<br />

in living tissues for the first time. Liu and Lu [21] and Lu et al. [22,23] reported<br />

that some thermal wave effects in bioheat transfer cannot be described by the Pennes<br />

equation. Mitra et al. [24] performed different experiments on processed meat with<br />

different boundary conditions. They observed wave-like phenomena in conduction<br />

heat transfer.<br />

Non-Fourier heat conduction of living tissue due to different kinds of heating methods<br />

are studied by employing a thermal wave model of bioheat transfer, such as heating<br />

by contact with hot material [25], laser heating [26–30], microwave heating [31,32],<br />

radio-frequency heating [33], and the dual-phase-lag model [34].<br />

In this study, the non-Fourier heat transport of skin biothermomechanics is solved<br />

analytically, where the skin surface temperature is exposed to a suddenly hot source<br />

of an arbitrary periodic temperature. The closed form of the temperature distribution<br />

function is obtained analytically for both the Pennes equation (Fourier) and thermal<br />

wave equation (non-Fourier) models for the skin as a finite and semi-infinite domain.<br />

The Pennes and thermal wave equations are solved analytically by the Laplace transform<br />

and the effect of metabolic heat. The thermal damage for skin for both Fourier<br />

and non-Fourier models are studied here.<br />

The finding of this study generates an exact analytical solution of the Pennes and<br />

thermal wave models of bioheat transfer for skin tissue as a finite and semi-infinite<br />

domain by considering an arbitrary periodic temperature for the skin surface boundary<br />

condition. Studying the thermal behaviors of skin tissue with this boundary condition<br />

by numerical solutions for Pennes, thermal wave, and dual-phase-lag bioheat transfer<br />

models is future work of the authors.<br />

123


Int J Thermophys (2013) 34:139–159 143<br />

2 Heat Transfer Models<br />

2.1 Pennes Bioheat Transfer Equation (PBHTE)<br />

For bioheat transfer, the Pennes equation is well established. This equation is based on<br />

the classical Fourier law. The Pennes bioheat transfer equation (PBHTE) is expressed<br />

as [2]<br />

ρ t c t<br />

∂T<br />

∂t<br />

+ ρ b ϖ b c b (T − T a ) = k ∂2 T<br />

∂x 2 + q met + q ext (1)<br />

Equation 1 is known as a parabolic bioheat equation. In this study, q ext is considered<br />

as zero.<br />

2.2 Thermal Wave Model of Bioheat Transfer (TWMBT)<br />

Cattaneo [17] and Vernotte [18] reported a modified unsteady heat conduction equation<br />

based on the concept of the finite heat propagation velocity as<br />

∂q(x, t)<br />

q(x, t) + τ q =−k∇T (x, t) (2)<br />

∂t<br />

BasedonEq.1, for a heat flux with the characteristic time τ q as well as the Pennes<br />

equation, a general form of the thermal wave model of bioheat transfer (TWMBT) in<br />

living tissues is expressed by Liu et al. [20]:<br />

τ q ρ t c t<br />

∂ 2 T<br />

∂t 2<br />

+ (ρ tc t + τ q ρ b ϖ b c b ) ∂T<br />

∂t<br />

+ ρ b ϖ b c b (T − T a )<br />

)<br />

(<br />

= k ∂2 T<br />

∂x 2 + ∂q met<br />

∂q ext<br />

q met + τ q + q ext + τ q<br />

∂t<br />

∂t<br />

(3)<br />

where τ q = α/C 2 is the thermal relaxation time [24,35].<br />

3 Analytical Solution of the Problem<br />

3.1 Arbitrary Periodic Surface Temperature Heating on a Finite-Domain Skin<br />

The skin surface temperature is exposed to a suddenly hot source with a temperature<br />

of T s (t). The skin tissue is considered as a perfect and infinitely wide and long layer.<br />

The adiabatic thermal condition is selected for the right boundary of the skin tissue.<br />

The geometrical problem and boundary conditions are shown in Fig. 1a.<br />

3.1.1 Solution of Thermal Wave Bioheat Transfer Equation (TWMBT)<br />

The bioheat transfer equation of a thermal wave as given in Eq. 3 is applied to describe<br />

the wave-like heat transfer process through the skin tissue. For simplicity, the new<br />

variable is defined as<br />

123


144 Int J Thermophys (2013) 34:139–159<br />

T(0,t)=T s<br />

Skin<br />

T x (L,t)=0<br />

T(0,t)=T s<br />

Skin<br />

T x (∞,t)=0<br />

x=L<br />

x<br />

x<br />

(a)<br />

skin as a finite domain<br />

(b)<br />

skin as a semi-infinite domain<br />

Fig. 1 Geometrical problem<br />

θ = T − T a (4)<br />

Equation 3, with respect to the constant q met and q ext = 0, can be rewritten in terms<br />

of a new variable and it becomes<br />

where<br />

τ q γ 1<br />

∂ 2 θ<br />

∂t 2 + (γ 1 + τ q γ 2 ) ∂θ<br />

∂t + γ 2θ = ∂2 θ<br />

∂x 2 + q m (5)<br />

γ 1 = ρ tc t<br />

k ,<br />

γ 2 = ρ bϖ b c b<br />

, q m = q met<br />

k<br />

k<br />

(6)<br />

The initial and boundary conditions can be written as<br />

θ(x, 0) = 0, θ t (x, 0) = 0 (7)<br />

θ(0, t) = T s (t) − T a = θ s (t), θ x (L, t) = 0 (8)<br />

If the boundary condition at the skin surface is periodic with an arbitrary function,<br />

we can write it in the Fourier series form. By considering the step function form at the<br />

skin surface that is shown in Fig. 2, the Fourier series form of the surface boundary<br />

condition can be written as<br />

123<br />

∞∑<br />

∞∑<br />

θ s (t) = T s (t) − T a = a 0 + a n cos(ω n t) + b n sin(ω n t) (9)<br />

n=1<br />

n=1


Int J Thermophys (2013) 34:139–159 145<br />

Fig. 2 Surface boundary<br />

condition<br />

where ω n = 2nπ/t 0 and the coefficients are obtained as<br />

a 0 =<br />

( )<br />

(Tb −T a )+(T c −T a )(t 0 −)<br />

t 0<br />

a n = 2 t 0<br />

[<br />

(Tb −T a ) sin(ω n )<br />

ω n<br />

b n =− 2 t 0<br />

[<br />

(Tb −T a )(cos(ω n )−1)<br />

ω n<br />

]<br />

+ (T c−T a )(sin(ω n t 0 )−sin(ω n ))<br />

ω n<br />

]<br />

+ (T c−T a )(cos(ω n t 0 )−cos(ω n ))<br />

ω n<br />

(10)<br />

By using the superposition theorem, this problem is divided into three sub-problems<br />

with the following boundary conditions:<br />

θ 1 (0, t) = a 0 , θ x (L, t) = 0 (11)<br />

θ 2 (0, t) = cos (ω n t) , θ x (L, t) = 0 (12)<br />

θ 3 (0, t) = sin (ω n t) , θ x (L, t) = 0 (13)<br />

By taking the Laplace transform of Eq. 5 and applying the initial conditions, the<br />

partial differential equation is changed into the following ordinary differential equation:<br />

d 2 θ<br />

dx 2 − P2 θ =− q m<br />

s ,<br />

P = √<br />

τ q γ 1 s 2 + (γ 1 + τ q γ 2 )s + γ 2 (14)<br />

The constant boundary condition problem (Eq. 11) is solved with q m and the other<br />

boundary conditions are solved with considering q m = 0, because Eq. 5 is nonhomogeneous<br />

due to term q m .<br />

The solution of Eq. 14 and applying the Laplace transform of boundary conditions<br />

(Eq. 11), can be obtained through<br />

¯θ 1 (x, s) = a 0 cosh (P(x − L))<br />

s cosh (PL)<br />

− q m cosh (P(x − L))<br />

sP 2 cosh(PL)<br />

+ q m<br />

sP 2 (15)<br />

The inverse theorem is applied for the inverse Laplace transform. For complementary<br />

information about the inverse theorem, refer to [36]. The first term of Eq. 15 has a pole<br />

123


146 Int J Thermophys (2013) 34:139–159<br />

at s = 0 and two other simple poles at s = S m + , and s = S− m , these poles are obtained<br />

as follows:<br />

⎡<br />

S m ± = 1<br />

( ) ]<br />

⎣−(γ 1 + τ q γ 2 ) ± √ (γ1 + τ q γ 2 )<br />

2τ q γ 2 − 4τ q γ 1<br />

[γ ⎤ 2 λm<br />

2 + ⎦ ,<br />

1 L<br />

( ) 2m − 1<br />

λ m =<br />

π, m = 1, 2, ... (16)<br />

2<br />

By calculating the residues at s = 0, s = S m + , and s = S− m , the inverse Laplace<br />

transform of the first term of Eq. 15 is obtained by<br />

θ 11 (x, t) = a 0 cosh ( √<br />

γ2 (x −L) ) ⎛<br />

( )<br />

∞<br />

cosh ( √<br />

γ2 L ) + ∑ 2a 0 e S+ m t λ m i cosh λm i<br />

L<br />

⎝<br />

(x −L) ⎞<br />

⎠<br />

L<br />

m=1<br />

2 S m + (2τ q γ 1 S m + +γ 1 +τ q γ 2 ) sinh (λ m i)<br />

⎛<br />

( )<br />

∞∑ 2a 0 e S− m t λ m i cosh λm i<br />

L<br />

+ ⎝<br />

(x − L) ⎞<br />

⎠<br />

L 2 Sm − (2τ q γ 1 Sm − (17)<br />

+ γ 1 + τ q γ 2 ) sinh (λ m i)<br />

m=1<br />

The poles of the second term of Eq. 15 are derived as follows:<br />

s = 0, s =− 1 τ q<br />

, s =− γ 2<br />

γ 1<br />

, s = S + m , s = S− m (18)<br />

By calculating the residues based on the above poles, the inverse Laplace transform<br />

of the second term of Eq. 15 is obtained as<br />

θ 12 (x, t) =− q m cosh ( √<br />

γ2 (x − L) )<br />

γ 2 cosh ( √<br />

γ2 L ) + q m τ q<br />

( )e − t q<br />

τq m γ 1<br />

+ ( ) e − γ 2<br />

γ1 t<br />

τq γ 2 − γ 1 γ1 − τ q γ 2 γ2<br />

(<br />

∞∑ 2q m e S+ m t cosh ( λ m i ( x<br />

−<br />

L<br />

− 1 )) )<br />

S + m=1 m λ m i(2τ q γ 1 S m + + γ 1 + τ q γ 2 ) sinh (λ m i)<br />

(<br />

∞∑ 2q m e S− m t cosh ( λ m i ( x<br />

−<br />

L<br />

− 1 )) )<br />

Sm − λ m i(2τ q γ 1 Sm − (19)<br />

+ γ 1 + τ q γ 2 ) sinh (λ m i)<br />

m=1<br />

The inverse Laplace transform of the third term of Eq. 15 is obtained by<br />

θ 13 (x, t) = q m<br />

⎛<br />

⎝ 1 γ 2<br />

−<br />

τ qe − t<br />

τq<br />

τ q γ 2 − γ 1<br />

+<br />

γ 1 e − γ 2 t<br />

γ 1<br />

( )<br />

γ 2 τq γ 2 − γ 1<br />

⎞<br />

⎠ (20)<br />

Finally, by adding Eqs. 17, 19, and 20, the closed form of the function θ 1 (x, t) is<br />

obtained by<br />

123<br />

θ 1 (x, t) = θ 11 (x, t) + θ 12 (x, t) + θ 13 (x, t) (21)


Int J Thermophys (2013) 34:139–159 147<br />

In this section, q m is assumed to be zero in Eq. 14. With a similar method, the inverse<br />

Laplace transforms of the solutions of this equation can be obtained by applying the<br />

boundary conditions (Eqs. 12, 13) in the Laplace domain.<br />

(√<br />

)<br />

cosh −τ q γ 1 ωn 2 + (γ 1 + τ q γ 2 )i ⌢ ω n + γ 2 (x − L) e iω nt<br />

θ 2 (x, t) =<br />

(√<br />

)<br />

2 cosh −τ q γ 1 ωn 2 + (γ 1 + τ q γ 2 )i ⌢ ω n + γ 2 L<br />

(√<br />

)<br />

cosh −τ q γ 1 ωn 2 + (γ 1 + τ q γ 2 )i ⌢ ω n + γ 2 (x − L) e −iω nt<br />

+<br />

(√<br />

)<br />

2 cosh −τ q γ 1 ωn 2 − (γ 1 + τ q γ 2 )i ⌢ ω n + γ 2 L<br />

⎡<br />

∞∑ 2S<br />

+ ⎣<br />

m + λ mi cosh ( λ m i ( x<br />

L<br />

− 1 )) ⎤<br />

e S+ m t<br />

(<br />

m=1 L 2 S m +2 (2τq<br />

+ ωn) 2 γ 1 S m + )<br />

⎦<br />

+ γ 1 + τ q γ 2 sinh (λm i)<br />

⎡<br />

∞∑ 2S<br />

+ ⎣<br />

m −λ mi cosh ( λ m i ( x<br />

L<br />

− 1 )) ⎤<br />

e S− m t<br />

(<br />

m=1 L 2 Sm −2 (2τq<br />

+ ω n<br />

2)<br />

γ 1 Sm − )<br />

⎦ (22)<br />

+ γ 1 + τ q γ 2 sinh (λm i)<br />

(√<br />

)<br />

cosh −τ q γ 1 ωn 2 + (γ 1 + τ q γ 2 )iω n + γ 2 (x − L) e iω nt<br />

θ 3 (x, t) =<br />

(√<br />

)<br />

2i cosh −τ q γ 1 ωn 2 + (γ 1 + τ q γ 2 )iω n + γ 2 L<br />

(√<br />

)<br />

cosh −τ q γ 1 ωn 2 − (γ 1 + τ q γ 2 )iω n + γ 2 (x − L) e −iω nt<br />

−<br />

(√<br />

)<br />

2i cosh −τ q γ 1 ωn 2 − (γ 1 + τ q γ 2 )iω n + γ 2 L<br />

⎡<br />

∞∑<br />

2ω<br />

+ ⎣<br />

n λ m i cosh ( λ m i ( x<br />

L<br />

− 1 )) ⎤<br />

e S+ m t<br />

(<br />

m=1 L 2 S m +2 (2τq<br />

+ ω n<br />

2)<br />

γ 1 S m + )<br />

⎦<br />

+ γ 1 + τ q γ 2 sinh (λm i)<br />

⎡<br />

∞∑ 2ω<br />

+ ⎣<br />

n λ m i cosh ( λ m i ( x<br />

L<br />

− 1 )) ⎤<br />

e S− m t<br />

( )<br />

⎦ (23)<br />

m=1 L 2 Sm<br />

−2 + ωn<br />

2 (2τ q γ 1 Sm − + γ 1 + τ q γ 2 ) sinh (λ m i)<br />

Finally, the closed form of function θ(x, t) is obtained by<br />

∞∑<br />

∞∑<br />

θ(x, t) = θ 1 (x, t) + a n θ 2 (x, t) + b n θ 3 (x, t) (24)<br />

n=1<br />

n=1<br />

3.1.2 Solution of the Pennes Bioheat Transfer Equation (PBHTE)<br />

The Pennes bioheat transfer equation as given in Eq. 1 is applied here to describe the<br />

parabolic heat transfer process through the skin tissue. Substituting Eq. 4 by Eq. 1, it<br />

123


148 Int J Thermophys (2013) 34:139–159<br />

gives<br />

γ 1<br />

∂T<br />

∂t<br />

+ γ 2 θ = ∂2 θ<br />

∂x 2 + q m (25)<br />

The boundary and initial conditions are similar to that of the thermal wave model.<br />

Taking the Laplace transform of Eq. 25 and considering the initial condition, we have<br />

the following:<br />

d 2 θ<br />

dx 2 − (γ 1s + γ 2 ) θ =− q m<br />

s<br />

(26)<br />

Substituting the boundary conditions of Eq. 11 for the general solution of Eq. 26,the<br />

function ¯θ P1 (x, s) on the Laplace domain is expressed as<br />

¯θ P1 (x, s) =<br />

[ ] [√<br />

a0<br />

s − q m cosh γ1 s + γ 2 (x − L) ]<br />

s (γ 1 s + γ 2 ) cosh ( √<br />

γ1 s + γ 2 L ) + q m<br />

s (γ 1 s + γ 2 )<br />

By employing the inverse theorem, the inverse Laplace transform of Eq. 27 is obtained<br />

by<br />

(<br />

θ P1 (x, t) = a 0 − q ) [√<br />

m cosh γ2 (x − L) ]<br />

γ 2 cosh ( √<br />

γ2 L ) + q m<br />

γ 2<br />

(( [√ ∞∑ 2 cosh γ1 S P + γ 2 (x − L) ] )(<br />

))<br />

e S P t<br />

+<br />

S P Lγ 1 sinh (√ γ 1 S P + γ 2 L ) √ q m<br />

a 0 γ1 S P + γ 2 − √<br />

γ1 S p + γ 2<br />

n=1<br />

(27)<br />

(28)<br />

where<br />

S P =− γ 2<br />

− 1 ( ) 2 λm<br />

, λ m =<br />

γ 1 γ 1 L<br />

( 2m − 1<br />

2<br />

)<br />

π, m = 1, 2, ... (29)<br />

In this section, q m is assumed to be zero in Eq. 26. With a similar method, the inverse<br />

Laplace transform of the solutions of this equation by applying the boundary conditions<br />

(Eqs. 12, 13), can be obtained as<br />

θ P2 (x, t) =<br />

(√<br />

)<br />

cosh γ 1 i ⌢ ω n + γ 2 (x − L) e iωnt<br />

(√<br />

) +<br />

2cosh γ 1 i ⌢ ω n + γ 2 L<br />

+<br />

(√<br />

)<br />

cosh −γ 1 i ⌢ ω n + γ 2 (x − L)<br />

)<br />

2cosh( √−γ1<br />

i ⌢ ω n + γ 2 L<br />

[ √ ∞∑ 2SP γ1 S P + γ 2 cosh (√ γ 1 S P + γ 2 (x − L) ) ]<br />

e S P t<br />

(<br />

S<br />

2<br />

P<br />

+ ωn<br />

2 )<br />

Lγ1 sinh (√ γ 1 S P + γ 2 L )<br />

m=1<br />

e −iωnt<br />

(30)<br />

123


Int J Thermophys (2013) 34:139–159 149<br />

θ P3 (x, t) =<br />

(√<br />

)<br />

cosh γ 1 i ⌢ ω n + γ 2 (x − L) e iωnt<br />

2i cosh (√ γ 1 iω n + γ 2 L )<br />

+<br />

m=1<br />

(√<br />

)<br />

cosh −γ 1 i ⌢ ω n + γ 2 (x − L) e −iωnt<br />

− 2i cosh (√ −γ 1 iω n + γ 2 L )<br />

[ √ ∞∑ 2ωn γ1 S P + γ 2 cosh (√ γ 1 S P + γ 2 (x − L) ) ]<br />

e S P t<br />

(<br />

S<br />

2<br />

P<br />

+ ωn<br />

2 )<br />

Lγ1 sinh (√ γ 1 S P + γ 2 L )<br />

(31)<br />

Finally, the closed form of the function θ P (x, t) is obtained by<br />

∞∑<br />

∞∑<br />

θ P (x, t) = θ P1 (x, t) + a n θ P2 (x, t) + b n θ P3 (x, t) (32)<br />

n=1<br />

n=1<br />

3.2 Arbitrary Periodic Surface Temperature Heating on a Semi-infinite Domain Skin<br />

In this case, the skin is assumed as a semi-infinite domain. Thus, the new boundary<br />

condition in the Laplace domain is defined as<br />

¯θ x (∞, s) = 0 (33)<br />

The geometrical problem and boundary conditions are shown in Fig. 1b. The boundary<br />

conditions at the skin surface are similar to that of Eqs. 11–13.<br />

3.2.1 Solution of Thermal Wave Bioheat Transfer Equation (TWMBT)<br />

By using the Laplace transform of the boundary conditions of Eq. 11 at the skin surface<br />

and replacing the boundary condition at x = L with Eq. 33, the solution of Eq. 14 is<br />

obtained as<br />

¯θ 1 (x, s) =<br />

The inverse Laplace transform of Eq. 34 is calculated by<br />

( a0<br />

s − q m<br />

sP 2 )<br />

e −Px + q m<br />

sP 2 (34)<br />

θ 1 (x, t) = £ −1 [ ¯θ 1 (x, s) ] =<br />

∫ t<br />

0<br />

G 1 (t − v) Q (x,v) dv + £ −1 ( q m<br />

sP 2 )<br />

(35)<br />

where<br />

G 1 (t) = £ −1 ( a 0<br />

s − q m<br />

sP 2 )<br />

, Q (x, t) = £ −1 ( e −Px) (36)<br />

123


150 Int J Thermophys (2013) 34:139–159<br />

The function Q (x, t) is simplified as follows:<br />

( (<br />

£ −1 e −Px) =£<br />

[exp<br />

−1 −x<br />

√τ q γ 1<br />

(s+ 1 )(<br />

s+ γ ) )]<br />

2<br />

τ q γ 1<br />

⎡ ⎛<br />

√<br />

(<br />

=exp − γ )<br />

√√√√ ⎛<br />

⎞⎞⎤<br />

2<br />

£ −1 ⎢ ⎜<br />

⎣exp ⎝−x √ ( )<br />

τq γ 1<br />

1<br />

γ 1 − τ q γ 2 s ⎝s+ ( ) ⎠⎟⎥<br />

γ 1 γ 1 − τ q γ 2 τq<br />

⎠⎦ ,<br />

γ 1<br />

γ 1 −τ q γ 2<br />

γ 1<br />

>τ q (37)<br />

γ 2<br />

In order to obtain the inversion of Eq. 37, the following equation is applied [37]:<br />

(<br />

£<br />

[exp<br />

−1 −x √ )] ⎡ ⎛ √<br />

κs 2 + s<br />

√ = £ −1 ⎣exp ⎝ −x κs ( s + 1 ) ⎞⎤<br />

κ<br />

√ ⎠⎦<br />

a a<br />

= δ (t − xb) s 1 (x, t) + H (t − xb) s 2 (x, t) (38)<br />

where<br />

( )<br />

s 1 (x, t) = e − 2κ t 1<br />

I0<br />

√t<br />

2κ<br />

2 −x 2 b 2 ,<br />

xbe − 2κ<br />

t<br />

s 2 (x, t) =<br />

2κ √ t 2 − b 2 x I 2 1<br />

( )<br />

1 √t<br />

2κ<br />

2 − x 2 b 2<br />

(39)<br />

where b = √ κ/a. δ and H represent the Dirac delta and Heaviside theta functions,<br />

respectively. With a similar method and using the Laplace transform of the boundary<br />

conditions of Eqs. 12 and 13 at the skin surface and assuming the skin as a semi-infinite<br />

domain, the closed form of functions θ(x, t) is obtained by<br />

∫ t<br />

θ(x, t) = D 1 +<br />

123<br />

+<br />

+<br />

∞∑<br />

n=1<br />

∞∑<br />

n=1<br />

0<br />

a n<br />

⎛<br />

⎝<br />

b n<br />

⎛<br />

⎝<br />

∫ t<br />

(G 1 (t − v) D 2 ) dv + H (t − xb)<br />

∫ t<br />

0<br />

∫ t<br />

0<br />

xb<br />

(G 2 (t − v) D 2 ) dv + H (t − xb)<br />

(G 3 (t − v) D 2 ) dv + H (t − xb)<br />

(G 1 (t − v) D 3 ) dv<br />

∫ t<br />

xb<br />

∫ t<br />

xb<br />

⎞<br />

(G 2 (t − v) D 3 ) dv⎠<br />

⎞<br />

(G 3 (t − v) D 3 ) dv⎠<br />

(40)


Int J Thermophys (2013) 34:139–159 151<br />

where<br />

⎛<br />

⎞<br />

D 1 = q m<br />

⎝ 1 −<br />

τ qe − τq<br />

t<br />

γ 1 e − γ 2 t<br />

γ 1<br />

+ ( ) ⎠ (41)<br />

γ 2 τ q γ 2 − γ 1 γ 2 τq γ 2 − γ 1<br />

⎛<br />

D 2 =δ (v − xb) exp ⎝−<br />

(<br />

v γ 1 − τ q γ 2 + γ 2<br />

2τ q γ 1<br />

γ 1<br />

)<br />

⎞<br />

(( )<br />

γ1 − τ<br />

⎠ q γ 2 √ )<br />

I 0 v<br />

2τ q γ 2 − x 2 b 2 (42)<br />

1<br />

xb ( ) (<br />

) (( )<br />

γ 1 − τ q γ 2 exp − v(γ 1−τ q γ 2 +1) γ1 −τ<br />

2τ q γ 1<br />

I q γ 2<br />

√ )<br />

1 2τ q γ 1 v 2 − x 2 b 2<br />

D 3 =<br />

√ (43)<br />

2τ q γ 1 v 2 − x 2 b 2<br />

⎛<br />

( )<br />

G 1 (t − v) = a 0 − q m<br />

⎝ 1 − τ qe − t−v<br />

⎞<br />

τq<br />

+<br />

γ 1e − γ 2 (t−v)<br />

γ 1<br />

( ) ⎠ (44)<br />

γ 2 τ q γ 2 − γ 1 γ 2 τq γ 2 − γ 1<br />

G 2 (t − v) = cos (ω n (t − v)) , G 3 (t − v) = sin (ω n (t − v)) , b = √ τ q γ 1 (45)<br />

3.2.2 Solution of the Pennes Bioheat Transfer Equation (PBHTE)<br />

By substituting the boundary conditions of Eq. 11 at the skin surface and Eq. 33 for<br />

Eq. 26, the function ¯θ P1 (x, s) is obtained as<br />

¯θ P1 (x, s) =<br />

(<br />

a0<br />

s −<br />

)<br />

q m<br />

e −√ γ 1 s+γ 2 x +<br />

s (γ 1 s + γ 2 )<br />

q m<br />

s (γ 1 s + γ 2 )<br />

(46)<br />

The inverse Laplace transform of Eq. 46 is calculated by<br />

θ P1 (x, t) = 1 2<br />

√<br />

1<br />

(<br />

a 0 − q )<br />

m<br />

γ 2<br />

γ<br />

( ) (<br />

qm<br />

− e −γ t £ −1<br />

γ 2<br />

((<br />

) )<br />

e −γ t £ −1 1<br />

1<br />

(√ √ ) − (√ √ ) e −√ √<br />

γ 1 sx<br />

s− γ s+ γ<br />

e −√ γ 1<br />

√ sx<br />

s<br />

)<br />

( ) ( )<br />

qm 1<br />

+ £ −1<br />

γ 2 s − 1<br />

(s + γ )<br />

(47)<br />

where γ = γ 2 /γ 1 . In order to obtain the inversion of Eq. 47, the following equations<br />

are applied [37]:<br />

£ −1 (<br />

£ −1 (<br />

)<br />

e −z√ s<br />

√ s + a<br />

e −z√ s<br />

s<br />

)<br />

= √ 1<br />

(<br />

e −z2 /4t − ae za e a2t erfc a √ t +<br />

πt<br />

( ) z<br />

= erfc<br />

2 √ t<br />

z<br />

2 √ t<br />

)<br />

z ≥ 0 (48)<br />

(49)<br />

123


152 Int J Thermophys (2013) 34:139–159<br />

Finally, the closed form of functionθ P1 (x, t)is obtained by<br />

θ P1 (x, t) = 1 (<br />

2 √ a 0 − q ) (<br />

m √γ<br />

e −γ t e<br />

− √ γ √ (<br />

γ 1 x e γ t erfc − √ √ )<br />

γ1 x<br />

γ t +<br />

γ γ 2 2 √ t<br />

+ √ γ e √ γ √ ( √ ))<br />

γ 1 √γ x e γ t γ1 x<br />

erfc t +<br />

2 √ t<br />

( ) (√ ) ( )<br />

qm<br />

+ e −γ t γ1 x<br />

erfc<br />

γ 2 2 √ qm (1<br />

+ − e<br />

−γ t ) (50)<br />

t γ 2<br />

By substituting the boundary conditions of Eq. 12 at the skin surface and Eq. 33 for<br />

Eq. 26 and taking q m = 0, the function ¯θ P2 (x, s) is obtained as<br />

(<br />

s<br />

¯θ P2 (x, t) =<br />

s 2 + ωn<br />

2<br />

)<br />

e −√ γ 1 s+γ 2 x<br />

(51)<br />

The inverse Laplace transform of Eq. 51 is calculated as follows:<br />

( ) ∫t<br />

θ P2 (x, t) = e −γ t 1<br />

2 √ (θ P21 (t − v) θ P22 (x, t)) dv (52)<br />

γ + iω n<br />

where<br />

0<br />

θ P21 (t − v) =−iω n e (t−v)(γ −iωn) + δ (t − v) (53)<br />

θ P22 (x, t) = √ (<br />

γ + iω n e −√ √<br />

γ +iω n γ1 x e (γ +iωn)t erfc − √ √ )<br />

√ γ1 x<br />

γ + iω n t +<br />

2 √ t<br />

+ √ γ + iω n e √ (<br />

√<br />

γ +iω n γ1 √γ<br />

√ )<br />

x e (γ +iωn)t √ γ1 x<br />

erfc + iωn t +<br />

2 √ t<br />

(54)<br />

By substituting the boundary conditions of Eq. 13 at the skin surface and Eq. 33 for<br />

Eq. 26 and taking q m = 0, the function ¯θ P3 (x, s) is<br />

( )<br />

ωn<br />

¯θ P3 (x, t) =<br />

s 2 + ωn<br />

2 e −√ γ 1 s+γ 2 x<br />

(55)<br />

The inverse Laplace transform of Eq. 55 is obtained by<br />

123<br />

θ P3 (x, t) = e−γ t<br />

4i<br />

(( ) ( ) )<br />

1<br />

1<br />

√ Y 1 (x, t) − √ Y 2 (x, t)<br />

γ + iωn γ − iωn<br />

(56)


Int J Thermophys (2013) 34:139–159 153<br />

where<br />

(<br />

Y 1 (x, t)=a 1 e −za 1e a2 1 t erfc<br />

Y 2 (x, t)=a 2 e −za 2e a2 2 t erfc<br />

√<br />

−a 1 t +<br />

z<br />

2 √ t<br />

( √<br />

−a 2 t +<br />

z<br />

2 √ t<br />

)<br />

(<br />

+a 1 e za 1e a2 1 t erfc<br />

)<br />

+a 2 e za 2e a2 2 t erfc<br />

)<br />

√<br />

a 1 t +<br />

z<br />

2 √ t<br />

( √<br />

a 2 t +<br />

z<br />

2 √ t<br />

)<br />

(57)<br />

(58)<br />

and<br />

a 1 = √ γ + iω n , a 2 = √ γ − iω n , z = √ γ 1 x (59)<br />

Finally, the closed form of the equation θ P (x, t) is obtained by<br />

∞∑<br />

∞∑<br />

θ P (x, t) = θ P1 (x, t) + a n θ P2 (x, t) + b n θ P3 (x, t) (60)<br />

n=1<br />

n=1<br />

3.3 Thermal Damage<br />

The burn evaluation is one of the most essential characteristics in the bioengineering<br />

science of skin tissue. The thermal damage begins when the basal layer temperature<br />

rises to 44 ◦ C[38]. The basal layer is located between the epidermis and dermis layers<br />

of the skin. A quantitative analysis of thermal damage was first proposed by Moritz<br />

and Henriques [39,40] based on the fact that the tissue damage could be represented<br />

as an integral of a chemical process rate:<br />

=<br />

∫ t<br />

0<br />

A exp (−E a /RT) dt (61)<br />

where A is a material parameter equivalent to a frequency factor, E a is the activation<br />

energy, and R is the universal gas constant. The constantsA and E a are obtained experimentally.<br />

By fitting these experimental data, a linear relation is observed betweenE a<br />

and ln(A), expressed as [41]<br />

E a = 21149.324 + 2688.367ln(A) (62)<br />

4 Results and Discussion<br />

In this study, the arterial blood and maximum temperature of the skin surface boundary<br />

condition are considered as T a = 376 ◦ C and T s = 100 ◦ C, respectively. The blood<br />

perfusion rate is considered as W b = ρ b ϖ b = 0.5kg · m −3 · s −1 here [7]. The<br />

other blood properties and one-layer skin properties are given in Table 1. Itisworth<br />

mentioning that the blood and tissue properties in all the results are based on data<br />

given in Table 1.<br />

123


154 Int J Thermophys (2013) 34:139–159<br />

Table 1 Thermophysical properties of blood and skin tissue [42,43]<br />

Parameters<br />

Skin density<br />

Skin specific heat<br />

Thermal conductivity of skin<br />

Metabolic heat generation<br />

Skin depth<br />

Blood density<br />

Blood specific heat<br />

Value<br />

1190 kg · m −3<br />

3600 J · kg −1 · K −1<br />

0.235 W · m −1 · K −1<br />

368.1 W · m −3<br />

0.006 m<br />

1060 kg · m −3<br />

3770 J · kg −1 · K −1<br />

Fig. 3 Temperature distribution at x = 0.0001 m for both the Pennes and thermal wave models of bioheat<br />

transfer<br />

The accuracy of the obtained analytical solutions is validated through comparison<br />

with the existing numerical results. Therefore, the temperature profiles for both the<br />

Pennes and thermal wave models for skin as a finite domain are obtained by applying<br />

the parameters introduced by Liu et al. [34], shown in Fig. 3. They assumed that the<br />

skin surface is exposed to a sudden heat source of constant temperature of 100 ◦ C and<br />

after contacting for 15 s, the heat source is removed and skin is cooled by a coolant<br />

of 0 ◦ C for 30 s. In this case, T c = 0 ◦ C, Δ = 15 s, t 0 = 45 s, and τ q = 10 s are<br />

considered. In fact, the results of the first period are compared with the numerical<br />

results of Liu et al. [34]inFig.3. By comparing the obtained results here with that of<br />

Liu et al. [34]atx = 0.0001 m, good agreement is observed. Of course, the results of<br />

this study have a little difference with the numerical results of Liu et al. [34], especially,<br />

at the locations where the thermal shocks (instantaneous jumping) have occurred. In<br />

the numerical solution of the thermal wave model due to the thermal shock, a large<br />

123


Int J Thermophys (2013) 34:139–159 155<br />

Fig. 4 Comparison between temperature response for the skin as a finite and semi-infinite domain at t =<br />

100 s and 250 s for Pennes bioheat transfer equation<br />

amount of oscillations was observed. The oscillations were decreased by increasing<br />

the accuracy of the numerical solution. In the thermal wave bioheat transfer model,<br />

the thermal wave propagation speed is finite, so instantaneous jumping is observed<br />

in the skin tissue temperature (see Fig. 3). The magnitude of the relaxation time τ q<br />

is an important characteristic of the thermal modeling of biomedical tissue and has<br />

an important effect on the temperature prediction. Mitra et al. [24] observed similar<br />

experimental wave-like results. These results show that the Pennes bioheat transfer<br />

model could not predict the instantaneous jump in the skin temperature, while the<br />

thermal wave model could predict this phenomenon correctly.<br />

In the thermal wave model of bioheat transfer, the thermal relaxation time generally<br />

is considered as τ q = 16 s for the skin tissue in previous research [24]; thus, τ q = 16 s is<br />

used here. The temperature profiles along the skin depth for both skins, having finite<br />

and semi-infinite domains at two different times for the Pennes and thermal wave<br />

models of bioheat transfer are shown in Figs. 4 and 5, respectively. Here, T c = 37 ◦ C,<br />

Δ = 45 s, and t 0 = 90 s are applied. The Pennes equation is based on an infinite speed<br />

of thermal wave propagation and due to this assumption, the type of bottom surface<br />

boundary condition gains an important effect on the temperature distribution even<br />

during the initial stages of heating. Figure 4 illustrates that the discrepancy between<br />

the results of the finite and semi-infinite skin domains is increased with increasing<br />

time. As shown is Fig. 5, in the results of the thermal wave model, there is good<br />

agreement in the finite and semi-infinite skin domains up to about 100 s, but after<br />

this time reaches about t = 102.47 s, the temperature profiles related to the skin as<br />

finite and semi-infinite domains become separated since the first thermal wave reaches<br />

x = L. This discrepancy is related to the right surface boundary condition. As for the<br />

finite domain, at the thermal wave reaches x = L, the adiabatic boundary condition<br />

123


156 Int J Thermophys (2013) 34:139–159<br />

Fig. 5 Comparison between temperature response for the skin as a finite and semi-infinite domain at t =<br />

100 s and 250 s for thermal wave model of bioheat transfer<br />

Fig. 6 Temperature response for the skin as a finite domain at x = 0.0001 m for both the Pennes and<br />

thermal wave models of bioheat transfer<br />

at x = L is not a suitable assumption because it means that heat cannot transfer into<br />

the depth more than x = L. Therefore, the results of Figs. 4 and 5 are not reality at<br />

large times and only illustrate that the boundary condition of the end of the skin is<br />

not important in the results of the thermal wave model while the first thermal wave is<br />

reached at x = L.<br />

123


Int J Thermophys (2013) 34:139–159 157<br />

Fig. 7 Temperature response for the skin as a finite domain at x = 0.0016 m for both the Pennes and<br />

thermal wave models of bioheat transfer<br />

The temperature response of the Pennes and thermal wave models for two selected<br />

blood perfusion rates at two different spatial locations, x = 0.0001 m and x =<br />

0.0016 m, are shown in Figs. 6 and 7, respectively. Here, T c =37 ◦ C, Δ = 5 s, and t 0 =<br />

10 s are applied. In these figures, for the temperature profile of the thermal wave model,<br />

a wave-like behavior is observed. As shown in Figs. 6 and 7, the first instantaneous<br />

temperature jump profile at the locations of x = 0.0001 m and 0.0016 m occurred at<br />

about t = 1.7 s and t = 27.3 s, respectively. The effect of the blood perfusion rate<br />

in tissue heat transfer is investigated. In general, the skin temperature decreases with<br />

an increase in the blood perfusion rate, where large amounts of heat can be carried<br />

away through the rate of blood perfusion. Figure 6 shows that the blood perfusion rate<br />

has little effect on the temperature distribution near the skin surface (x = 0.0001 m).<br />

On the other hand, Fig. 7 shows that the blood perfusion rate has a relatively large<br />

influence on heat transfer at the skin depth (x = 0.0016 m).<br />

The variation of thermal damage with time for both the Pennes and thermal wave<br />

models with τ q = 16 s, T c = 37 ◦ C,Δ = 5 s, and t 0 = 10 s at the locations of<br />

x = 0.0001 m and 0.0016 m is shown in Fig. 8. In this study, the thermal damage<br />

of skin is calculated by using the burn integration of Eq. 61, with the frequency<br />

factor A = 3.1 × 10 98 and the ratio of activation energy to universal gas constant<br />

E a /R = 75 000 [40]. A wave-like behavior can also be observed in the results of<br />

the thermal damage of the thermal wave model. The results in Fig. 8 illustrate that,<br />

in general, the thermal damage predicted by the Pennes and thermal wave models<br />

are different. This difference is due to the lack of consideration with respect to the<br />

non-Fourier effects on the materials that have the great thermal relaxation times when<br />

exposed to sudden temperature variations. Thus, it can be deduced that on neglecting<br />

these conditions, large deviations are observed in the predictions of the temperature<br />

distribution, thermal damage, and thermal stress.<br />

123


158 Int J Thermophys (2013) 34:139–159<br />

Fig. 8 Thermal damage profile for the skin as a finite domain at x = 0.0001 m and 0.0016 m for both the<br />

thermal wave model and Pennes equation<br />

5 Conclusions<br />

The thermal behavior of skin tissue was investigated here. The thermal wave and<br />

Pennes bioheat transfer models were applied for the energy equation in skin as finite<br />

and semi-infinite domains. The governing equations were solved analytically by means<br />

of the Laplace transform. The inverse theorem is applied to calculate the inverse<br />

Laplace transform. The thermal damage was studied at x = 0.0001 m and 0.0016 m<br />

locations. The obtained results were compared with those obtained by Liu et al. [34]<br />

to validate the accuracy of the analytical solution here. This comparison shows good<br />

agreement between both results. The results show that wave-like behavior is observed<br />

for the thermal wave model and the thermal relaxation time τ q has an important effect<br />

on the temperature distribution. The results indicate that the thermal wave model<br />

could predict the instantaneous jump in the temperature profile, consistent with the<br />

experimental results obtained by Mitra et al. [24], while the Pennes equation does not<br />

show a prediction of this kind of jump in the temperature profile.<br />

The results of the Pennes equation for skin as either a finite or semi-infinite domain<br />

are compared. There is a discrepancy between the temperature responses from the<br />

finite and semi-infinite domains even during the initial stage heating. On the other<br />

hand, the relaxation time of skin is great and the speed of the thermal wave is finite.<br />

The results of the thermal wave model show that two graphs of finite and semi-infinite<br />

models of skin tissue are quite identical before the first thermal wave is propagated<br />

at the right boundary of the tissue. The temperature profiles become separated. This<br />

phenomenon show that the boundary condition at the end of the skin is not important<br />

in the results of the thermal wave model when the first thermal wave reached x = L.<br />

123


Int J Thermophys (2013) 34:139–159 159<br />

The thermal damage results show that the obtained burn times for the Pennes and<br />

thermal wave models are different in general. The thermal wave behavior is observed<br />

in the thermal damage profile.<br />

References<br />

1. Y.C. Fung, Biomechanics: Motion, Flow, Stress, and Growth (Springer, New York, 1990)<br />

2. H.H. Pennes, J. Appl. Physiol. 1, 93 (1948)<br />

3. H. Arkin, L.X. Xu, K.R. Holmes, IEEE Trans. Biomed. Eng. 41, 97 (1994)<br />

4. L. Lang, B. Erdmann, M. Seebass, IEEE Trans. Biomed. Eng. 46, 1129 (1999)<br />

5. F. Xu, T.J. Lu, K.A. Seffen, J. Mech. Phys. Solids 56, 1852 (2008)<br />

6. J.J. Zhao, J. Zhang, N. Kang, F. Yang, Appl. Math. Comput. 171, 320 (2005)<br />

7. T.C. Shih, P. Yuan, W.L. Lin, H.S. Kou, Med. Eng. Phys. 29, 946 (2007)<br />

8. J. Liu, L.X. Xu, IEEE Trans. Biomed. Eng. 46, 1037 (1999)<br />

9. B. Erdmann, J. Lang, M. Seebass, Ann. N.Y. Acad. Sci. 858, 36 (1998)<br />

10. E.Y.K. Ng, H.M. Tan, E.H. Ooi, Burns 35, 987 (2009)<br />

11. E.Y.K. Ng, H.M. Tan, E.H. Ooi, Phil. Trans. R. Soc. Lond., Ser. A 368, 655 (2010)<br />

12. J.W. Durkee Jr, P.P. Antich, C.E. Lee, Phys. Med. Biol. 35, 847 (1990)<br />

13. M.A. Hader, M.A. Al-Nimr, Heat Transf. Eng. 23, 35 (2002)<br />

14. A.F. Khadrawi, M.A. Al-Nimr, M. Hammad, Int. J. Thermophys. 23, 581 (2002)<br />

15. M.A. Hader, M.A. Al-Nimr, B.A. Abu Nabah, Int. J. Thermophys. 23, 1669 (2002)<br />

16. M. Naji, M.A. Al-Nimr, M. Hader, Int. J. Thermophys. 24, 545 (2003)<br />

17. C. Cattaneo, Comp. Rend. 247, 431 (1958)<br />

18. P. Vernotte, Comp. Rend. 246, 3154 (1958)<br />

19. M.A. Al-Nimr, M. Naji, Int. J. Thermophys. 21, 281 (2000)<br />

20. J. Liu, Z. Ren, C. Wang, Chin. Sci. Bull. 40, 1493 (1995)<br />

21. J. Liu, W.Q. Lu, Space Med. Med. Eng. 10, 391 (1997)<br />

22. W.-Q. Lu, J. Liu, Y. Zeng, Eng. Anal. Bound. Elem. 22, 167 (1998)<br />

23. W.-Q. Lu, J. Liu, Y. Zeng, J. Eng. Thermophys. 19, 728 (1998)<br />

24. K. Mitra, S. Kumar, A. Vedavarz, M.K. Moallemi, J. Heat Transfer, Trans. ASME 117, 568 (1995)<br />

25. K.C. Liu, Int. J. Therm. Sci. 47, 507 (2008)<br />

26. M. Jaunich, S. Raje, K. Kim, K. Mitra, Z. Guo, Int. J. Heat Mass Transf. 51, 5511 (2008)<br />

27. F. Xu, T.J. Lu, K.A. Seffen, E.Y.K. Ng, Appl. Mech. Rev. 62, 50801 (2009)<br />

28. M.M. Tung, M. Trujillo, J.A. Lopez Molina, M.J. Rivera, E.J. Berjano, Math. Comput. Modell. 50,<br />

665 (2009)<br />

29. K. Ting, K.T. Chen, S.F. Cheng, W.S. Lin, C.R. Chang, Jpn. J. Appl. Phys. 47, 361 (2008)<br />

30. J. Zhou, J.K. Chen, Y. Zhang, Comput. Biol. Med. 39, 286 (2009)<br />

31. S. Ozen, S. Helhel, S. Bilgin, Radiat. Environ. Biophys. 50, 483 (2011)<br />

32. S. Ozen, S. Helhel, O. Cerezci, Burns 34, 45 (2008)<br />

33. J.A. Lopez Molina, M.J. Rivera, M. Trujillo, F. Bordio, J.L. Lequerica, F. Hornero, E.J. Berjano, Open<br />

Biomed. Eng. J. 2, 22 (2008)<br />

34. K.C. Liu, P.J. Cheng, Y.N. Wang, J. Therm. Sci. 15, S61 (2011)<br />

35. D.Y. Tzou, Macro to Micro-scale Heat Transfer: The Lagging Behavior (Taylor and Francis, Washington,<br />

DC, 1997)<br />

36. V.S. Arpaci, Conduction Heat Transfer (Addison-Wesley, New York, 1966)<br />

37. A. Erdely, M.F. Oberhettinger, F.G. Tricomi, Tables of Integral Transforms (McGraw-Hill, New York,<br />

1954)<br />

38. A.M. Stoll, L.C. Greene, J. Appl. Physiol. 14, 373 (1959)<br />

39. A.R. Moritz, F.C. Henriques, Am. J. Pathol. 23, 695 (1947)<br />

40. F.C. Henriques, A.R. Moritz, Am. J. Pathol. 23, 531 (1947)<br />

41. F. Xu, K.A. Seffen, T.J. Lu, IAENG Int. J. Comput. Sci. 35, 1 (2008)<br />

42. F. Xu, K.A. Seffen, T.J. Lu, Int. J. Heat Mass Transf. 51, 2237 (2008)<br />

43. F. Xu, T. Lu, Introduction to Skin Biothermomechanics and Thermal Pain (Science Press, Beijing and<br />

Springer, Berlin, 2011)<br />

123

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

Saved successfully!

Ooh no, something went wrong!