You are on page 1of 9

Mathematical Biosciences 269 (2015) 19

Contents lists available at ScienceDirect

Mathematical Biosciences
journal homepage: www.elsevier.com/locate/mbs

Bioheat transfer problem for one-dimensional spherical


biological tissues
Emmanuel Kengne, Ahmed Lakhssassi
Dpartement dinformatique et dingnierie, Universit du Qu bec en Outaouais, 101 St-Jean-Bosco, Succursale Hull, Gatineau (PQ) J8Y 3G5, Canada

a r t i c l e

i n f o

Article history:
Received 30 March 2015
Revised 6 July 2015
Accepted 20 August 2015
Available online 1 September 2015
Keywords:
Pennes bioheat transfer model
Point-heating source
Bioheat transfer problems
Tumors
Spherical living biological bodies

a b s t r a c t
Based on the Pennes bioheat transfer equation with constant blood perfusion, we set up a simplied onedimensional bioheat transfer model of the spherical living biological tissues for application in bioheat transfer
problems. Using the method of separation of variables, we present in a simple way the analytical solution of
the problem. The obtained exact solution is used to investigate the effects of tissue properties, the cooling
medium temperature, and the point-heating on the temperature distribution in living bodies. The obtained
analytical solution can be useful for investigating thermal behavior research of biological system, thermal
parameter measurements, temperature eld reconstruction and clinical treatment.

1. Introduction
Spatiotemporal temperature distribution in living biological tissues plays a vital role in many physiological processes. Investigation
of bioheat transfer problems requires the evaluation of temporal and
spatial distributions of temperature. This class of problems has been
traditionally addressed using the Pennes bioheat equation [1]. Scientic research in the bioheat transfer research eld has paved a key
foundation in hyperthermia cancer therapy, thermal diagnosis, cryogenic surgery etc. [28]. The quantitative, qualitative, and accurate
analysis of bioheat transfer is to effectively understand and model
the heat transfer mechanism of the biological system.
Heat transfer analysis on thermal medical problems, such as the
thermal diagnostics [9] and thermal comfort analysis [10,11], thermal
parameter estimation [1216], or burn injury evaluation [17], usually
has to simultaneously face the transient or spatial heating both on
skin surface and in interior of the biological bodies. The complexity
underlying in this class of medical problems remains not only for its
heterogeneity and anisotropy but also for conduction, convection,
and radiation heat ow, cells metabolism, and blood perfusion etc.
Therefore, to obtain a exible solution, which is capable of solving
any one of the above thermal medical problems, is very desirable.
Indeed, analytical solutions reect actual physical feature of the
models and can be used as standards to verify the corresponding
numerical results and as a proof to the reasonability of in-vitro mode

Corresponding author. Tel.: +18196430402.


E-mail address: kengem01@uqo.ca, ekengne6@yahoo.fr (E. Kengne).

http://dx.doi.org/10.1016/j.mbs.2015.08.012
0025-5564/ 2015 Elsevier Inc. All rights reserved.

2015 Elsevier Inc. All rights reserved.

analysis. Although people relied too much on numerical approaches


such as nite difference method (FDM), nite element method (FEM),
and boundary element method (BEM) for solving thermal medical
problems, the analytical solutions, if they can be obtained, are
often preferred [18,19]. Analytical solutions are very attractive since
their eciency depends weakly on the dimensions of the problem,
in contrast to the numerical methods. Knowing analytical solution,
temperature at a desired point at a given time can be performed independently from that of the other points within the domain, which can
be an asset when temperatures are needed at only some isolated sites
or times. It is also important to point out that analytical solutions
of bioheat transfer problems will save computational time greatly,
which is valuable in some hyperthermia practices. Motivated by the
importance of exact solutions of bioheat transfer problems in hyperthermia cancer therapy, thermal diagnosis, or cryogenic surgery,
we aimed in this paper to present analytical solutions to the Pennes
bioheat model in one-dimensional spherical coordinate system with
relatively complex boundary or volumetric heating conditions [20].
Various techniques have been performed to obtain exact solutions
of bioheat transfer problems in one dimensional Cartesian coordinate
[8,2127]. In our knowledge, a combination of the method of separation of variables with the Greens function method has not yet been
applied for explicitly solving bioheat transfer problems for spherical
symmetry. In this paper, we combine the Fourier method (method
of separation of variables) with the Greens function method to derive the exact solution of one-dimensional model of the spherical living tissue. Such a combination of the two methods has been used
by Durkee and Antich [28,29] when addressing the time-dependent
Pennes equation in 1-D multi-region Cartesian and spherical

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 1. Spherical model of organ showing the tumour either at the center of the model (a) or far from the center of the model (b).

geometry. In most of the existing analytical studies of bioheat transfer


in biological tissues based on either the method of separation of variables, or the Greens function method, or a combination of these two
methods, the solutions to the bioheat transfer problem are either for
a steady state, heat conduction equations, an innite domain, or for
a constant heating at skin surface or inside the tissue volume, which
may not be practical for some real bio-thermal situations (see for example Refs. [25,26,30]). The main aim of combining in the present
work the method of separation of variables with the Greens function method is to nd analytical exact solution that may incorporate
relatively complete situations such as the nite tissue domain, the
transient or space-dependent boundary conditions and volumetric
heating. The rest of the paper is organized as follows. In Section 2,
we describe the model under consideration and present its analytical
solution. The application of obtained solution to the tissues internal
temperature distribution and the thermal effect of tumors are discussed in Section 3, and the main results are summarized in Section 4.

2. Model and solution


An important number of theoretical analysis on bioheat transfer
problems is based on the Pennes equation [1], which describes the
inuence of blood ow on the temperature distribution in the tissue
in terms of volumetrically distributed heat sinks or sources. Thus
it will also be used in this paper. More precisely, we consider in
this paper one-dimensional (1-D) case of Pennes bioheat model
with constant thermal parameters, which is written in 1-D spherical
coordinate system as



1
T
2 T
=k 2
c
r
+ b b cb (Ta T ) + qm + Q (r, t ),
t
r
r r

(1)

where , c, and k are the density, the specic heat, and the thermal
conductivity of the tissue, respectively, b and cb denote density and
specic heat of blood, b stands for the blood perfusion, Ta and T
are the arterial temperatures which are treated as a constant and the
tissue temperature at a given position for a given time, respectively,
qm is the metabolic heat generation, and Q is the heat source due to
spatial heating. In Eq. (1), r2 = x2 + y2 + z2 , where (x, y, z) are the
rectangular coordinates of a point of the tissue, situated at a distance
r from the center of the model; here, 0 x L, where L is the distance
from the skin surface (x = L) to the body core (x = 0).
Following Hossain and Mohammadi [31], the initial temperature
eld for the basal state of biological bodies in the case of axially
symmetrical model can be obtained through solving the following

problem:




1 d

2 dT0

r
T0 + = 0
r2 dr

dr

dT0 
 = 0,

dr r=0

k dT0 

= ha (T0 Te )|r=R1 ,

dr 

(2)

r=R1

where = b b cb /k, = (b b cb Ta + qm )/k, and T0 (r) = T (r, 0)


is steady-state temperature elds prior to heating, R1 is the radius
of the concerned tissue, ha is the heat exchange coecient which
accounts for both the convection and radiation heat loss on the tissue
surface, and Te is the surrounding air temperature. Here, the skin surface is dened at r = R1 while the body core at r = 0. We assume that
the tumour is located either at the center of the spherical model (see
Fig. 1(a)) or far from the center of the spherical model (see Fig. 1(b)).
The solution to problem (2) is

2R21 ha ( Te ) sinh
T0 (r) =
+

Dr

(3)

where

D = R1 k + k ha R1 exp R1


+ ha R1 + R1 k k exp R1 .
During the practical thermal processes, the boundary condition
(BC) at the skin surface is often time-dependent. For axially symmetrical model the boundary conditions at tumor surface and at the skin
surface are described respectively as


T 
k
= 0,
r r=0

(4a)



T 
k
= h f T f (t ) 
,

r=R1
r r=R

(4b)

where f(t) is the time-dependent temperature of the cooling medium,


hf is the transfer coecient due to convection and radiation (hf is the
heat transfer coecient in the direction of heat ow on the boundary
skin surface), and T/ r is the partial derivative of T along the outward direction normal to the skin surface. No heat loss is assumed
in remaining regions. The ux condition (4b) describes the exchange
between body and surrounding medium. The choice of BC (4b) is motivated by the fact that typical boundary conditions at the skin surface
for cancer hyperthermia or thermal comfort analysis are usually the

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 2. Steady-state temperature elds prior to heating for tissue properties given in the text. (a) Effect of the blood perfusion on the steady-state temperature distribution. (b)
Effect of the tissue thermal conductivity on the steady-state temperature distribution.

third BC. Through using the following transformation:

3. Results and discussion

T (r, t ) = u(r, t ) + f (t ),

(5)

Eq. (1) takes the following form:



c u 1 2 u
r
u + F,
= 2
k t
r
r r

(6)

where

F (r, t ) = (Ta f (t )) +

qm + Q
c df

.
k
k dt

(7)

The corresponding boundary and initial conditions are then expressed as


u 
k
= 0,
r r=0

u 
k
+ h f u|r=R1 = 0
r 

(8a)

(8b)

r=R1

u(r, 0) = T0 (r) f (0).

(8c)

Solving problems (6) and (8a)(8c) with the method of separation of


variables combined with the Greens function method [8] and using
Eq. (5) yield

T (r, t ) = f (t ) f (0) + exp


+

k
c

where

G(r, t; , ) = exp

R1
0

k
t T (r)
c 0

G(r, t; , )d ,

(9)
3.1. Temperature distribution under a point-heating source

+

k
F ( , ) sin [n ] sin [n r]
,
( t )
2
c
r
n=1 Un 

(10)

n are positive roots of equation


tan [R1 ]

kR1
k R1 h f

= 0,

(11)

and

Un 2 = n Si[2n R1 ]

In this section, we discuss the effect of major thermal parameters


on the temperature distribution of spherical living tissues. In our simulations, the typical tissue properties are applied as given in [9,32,33].
The apparent heat convection coecient due to natural convection and radiation is taken as ha = 10 W/(m2 C), while the forced
convection coecient is applied as h f = 100 W/(m2 C) and the
surrounding uid temperature is chosen as Te = 25 C. Further, as
demonstrated in many works [34,35], the interior tissue temperature
usually tends to a constant within a short distance such as 24 cm.
Therefore 0.02 m L 0.04 m will be used in our study. Curves of
Fig. 2 depict the steady-state temperature distributions of biological
bodies associated with the above typical tissue properties. Fig. 2(a)
and (b) shows the effect of the blood perfusion and the tissue thermal
conductivity on the steady-state temperature distribution. Obviously,
the steady-state temperature (initial temperature) of the tissue decreases from the body core to the skin surface. It is seen from Fig. 2(a)
that the steady-state temperature of the tissue increases as the blood
perfusion decreases; moreover, Fig. 2(a) shows that the blood perfusion has signicant effects on the temperature of the body core.
Curves of Fig. 2(b) show that the steady-state temperature near the
body core (near the skin surface) decreases as the tissue thermal conductivity increases (increases with the tissue thermal conductivity).
In our investigations, we study the temperature distribution when
the biological body is subjected either to a point-heating source or to
the most typical heat source that the heat ux decays exponentially
with the distance from the skin surface.

sin [n R1 ]
.
R1
2

In our analysis, we rst discuss the temperature response subject


to a point-heating. Practical examples for such heating can be found
in clinics where heat is deposited though inserting a conducting heating probe in the deep tumor site. Here, the point-heating source to be
used is [8]

Q (r, t ) = P1 (t )(r r0 ),

(12)

where P1 (t) is the point-heating power, (r r0 ) is the Dirac function,


and r0 is the position of the point-heating source. Following Anderson
and Burnside [36], a sinusoidal point-heating power

P1 (t ) = q0 + qw cos [0 t]

(13)

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

will be used. Here, q0 and qw are the constant term and the oscillation amplitude of sinusoidal point-heating power, respectively, and
0 is heating frequency. It is important to point out that the spatial
sinusoidal heating was also proposed [8] to measure the blood perfusion where the heat was deposited to the biological body using ultrasound, and the temperature response was monitored at the skin
surface. As the time-dependent temperature of the cooling medium,
we will consider a sinusoidal temperature as follows:

f (t ) = q0 f + qw f cos

T (r, t ) = f (t ) f (0) + exp

0 f t ,

R qw sin [n r0 ]

r0 k2 2 + 02 2 c

+

k
1
P1 ( ) sin [n r0 ]
( t )
2
c
k
r0
U


n
n=1



sin [n ]
qm
c df
sin [n r]
+
( )

.
Ta f ( ) +

k
k dt
r

(14)

 gI (t ) sin [n r]
k
n
t T0 (r) +
,
2
c
r
U

n
n=1
(15)

where

gIn (t ) =

where q0f and qwf are the constant term and the oscillation amplitude,
respectively, of sinusoidal temperature of the cooling medium, and
0f is the temperature frequency.
Under the above considerations, Eq. (10) becomes

G(r, t; , ) = exp

Therefore, the temperature response is obtained from Eqs. (9) and


(14):

k
t
k exp
c


k cos [0t] + 0 c sin [0t]
2



R1 q0 sin [n r0 ]
+ Si[n R1 ]
kr0




kTa + qm k q0 f
k
t
1 exp

k
c





k
t cos 0 f t .
+ qw f Si[n R1 ] exp
c

Fig. 3 depicts temporal temperature distributions in the tissues


under a surface point-heating with the sinusoidal heating power
P1 (t ) = 2500 + 2450 cos [0.02t] at r0 0. The sinusoidal temperature

Fig. 3. Temperature distribution at three positions when a surface point-heating with the sinusoidal point-heating power and a sinusoidal temperature of the cooling medium are
adopted.

Fig. 4. Temperature distribution at different times when a surface point-heating with the sinusoidal point-heating power P1 (t ) = 1500 + 1450 cos [0.02t] W/m3 is applied. A
sinusoidal temperature of the cooling medium f (t ) = 25 + cos [0.02t] is used.

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 5. Effect of some biological parameter on the temperature distribution for typical tissue parameters shown in Table 1. (a) Inuence of the tissue density . (b) Inuence of the
tissue thermal conductivity k.

Table 1
Typical material properties of the tissue.
Parameters

Value and unit

, b
c, cb
Ta
k

1000 [kg/m3 ]
4200 [J/(kg C)]
37 [C]
0.5 [W/(m C)]
0.0005 [ml/(s ml)]
33,800 [W/m3 ]

qm

of the cooling medium f (t ) = 25 + cos [0.02t] C is used. Due to the


surface cooling by the owing medium, the lowest tissue temperature occurs at the skin surface (x = L = 0.03) while the highest tissue
temperature reaches at the body core.
Fig. 4 depicts the temperature distributions at time t = 100 s of
biological bodies subject to a surface sinusoidal point-heating and
sinusoidal temperature of the cooling medium. The surface pointheating is applied at r0 = 0.01. Due to the surface cooling by the owing medium, the tissue temperature decreases as both the time and
the depth increase. At a given time, the position for the local highest temperature is just stayed at the site of the point source. This is
very benecial for hyperthermia therapy since one can then selectively heat the deep regional tumor.
Fig. 5 illustrates the temperature distribution at time
t = 100 s for the tissues subjected to point source P1 (t ) =
2500 + 2450 cos [0.02t] W/m3 , r0 = 0.015 m under the sinusoidal
cooling medium temperature f (t ) = 25 + cos [0.02t] C for the

typical biological parameters shown in Table 1. The effect of the


tissue density and the effect of the tissue thermal conductivity k on
the temperature distribution are illustrated in Fig. 5(a) and Fig. 5(b),
respectively. For each value of either the tissue density or the tissue
thermal conductivity k, the position for the local highest temperature
is just stayed at the site of the point source. Plots of Fig. 5(a) and (b)
show that the position for the local highest temperature depends
both on the tissue density and the tissue thermal conductivity. It
is seen from Fig. 5(a) that the higher the tissue density, the higher
the temperature near the skin surface. The effect of the tissue
thermal conductivity on the temperature distribution is illustrated
in Fig. 5(b); this gure indicates that the thermal conductivity has
insignicant effect on the temperature distribution at the body core
as it varied between 0.48 W/(m C) and 0.54W/(m C). The curves
of Fig. 5(b) indicate that the higher the tissue thermal conductivity,
the lower the temperature near the body core and the higher the
temperature near the skin surface. Such a phenomenon in thermal
distribution of biological body produces a distinguishable elevation
in either the skin surface or the body core temperature and is a tool
for analyzing benign stage tumour.
Fig. 6 depicts the inuence of tissue specic heat c on the temperature distribution at time t = 100 s when tissue was subjected
to point-heating source Q (r, t ) = P1 (t )(r 0.015), and the temperature of cooling medium is f (t ) = 25 C. It shows that, the larger tissue
specic heat, the higher temperature increases.
Fig. 7 shows the temperature distributions at time t = 100 s for
the tissues subjected to constant point source P1 (t ) = 2500 W/m3 ,

Fig. 6. Effect of the tissue specic heat c on the temperature distribution in biological tissues subject to a constant point-heating under a constant temperature of the cooling
medium.

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 7. Inuence of cooling medium temperature to tissue temperature distribution.

Fig. 8. Inuence of heating power on tissue temperature distribution.

r0 = 0.018 m and under different cooling medium temperatures. As


we can see from plots of this gure, the larger the temperature of the
cooling medium, the lower the body core temperature decreases. It is
also seen from this gure that the magnitude and position of the local
highest temperature are changeless on the whole. This means that
an optimum heating can be obtained through regulating the surface
cooling.
We show in Fig. 8 the inuence of heating power on the temperature distribution when tissue was subjected to a constant pointheating source Q (r, t ) = P1 (t )(r 0.018), and the temperature of
cooling medium is f (t ) = 25 C. It shows that although the effect
of the heating power on the temperature distribution is negligible,
the larger the heating power, the higher the temperature increases.
Moreover, the position for the local highest temperature is xed for
all these heatings.
Fig. 9 gives out the effects of the cooling medium temperature (top
plot) and the effects of the heating power (bottom plot) on the temperature transients at skin surface. Curves of the top panel are obtained when the biological tissue was subjected to a constant pointheating source Q (r, t ) = 2500(r 0.018). To generate the plots of
the bottom panel, we used the constant cooling medium temperature f (t ) = 25 C. As it is seen from the curves of the top panel,
the skin surface temperature decreases when the cooling medium
temperature increases. The bottom panel of Fig. 9 indicates that, the

larger the heating power, the higher the skin surface temperature
increases.

3.2. Temperature distribution under the most typical heating source


In this section, we study the temperature distribution when the
tissue is subjected to the typical heating source for which the heat
ux decays exponentially with the distance from the skin surface. In
other words, we work with the spatial heating [37]

Q (r, t ) = P0 (t ) exp [r],

(16)

where P0 (t) is the time-dependent heating power on skin surface


and is the scattering coecient. We focus on the constant heating power (P0 (t ) = q p = constant), reecting the situation where the
human skin was heated by a laser.
In the present situation, the temperature response is obtained
from Eqs. (9) and (14):

 gII (t ) sin [n r]
k
n
t T (r) +
,
T (r, t ) = f (t ) f (0) + exp
2
c 0
r
U

n
n=1
(17)

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 9. Inuence of cooling medium temperature (top) and inuence of heating power (bottom) on the skin surface temperature response.

where

gIIn (t ) =

k Ta + qm k q0 f Si[n R1 ]

k



1 exp

k
t
c





k
t
qw f Si[n R1 ] cos 0 f t exp
c


  R
1
qp
k
sin [n ]
t
+
1 exp
exp [ ]
d .
k
c

0
Fig. 10 shows the effect of the heating power P0 (t) (top) and the
effect of the scattering coecient to the skin surface temperature
response. To generate the curves of this gure, we use the constant
cooling medium temperature f (t ) = 25 C and the tissue properties
shown in Table 1. The top panel of Fig. 10 shows the transient temperatures of tissues subject to four constant heating with = 200 m1
[37]. Obviously, the larger the heating power, the higher the temperature increases. Such information is valuable for thermal comfort
evaluation. The bottom panel of Fig. 10 gives out the effects of the
scattering coecient on the temperature transients at skin surface
for P0 (t ) = 250 W/m2 . The plots of this gure show that the larger
the coecient, the higher the temperature decreases. Although the
scattering coecient seems to have insignicant effect on the
temperature distribution on the skin surface, the bottom panel of
Fig. 10 reveals that the tissue temperature increases as the scattering

coecient decreases. Because different heating apparatus such as


laser or microwave may have different power P0 (t) and scattering
coecient , we conclude that the above results are expected to
be useful for the heating-dose planning during the hyperthermia
treatment or parameter estimation.
Fig. 11 is the transient temperature at three positions of biological
bodies subject to a constant spatial heating with P0 (t ) = 250 W/m2
and = 200 m1 [37]. Curves, A, B and C are the transient temperatures of tissues at positions r = 0.01 m, r = 0.015 m, and r = 0.025 m.
Plots of Fig. 11 indicate that the tissue temperature decreases as the
tissue depth increases; this means that the highest temperature occurs at the body core, while the lowest temperature occurs at the skin
surface. This situation is due to the surface cooling by the owing
medium. It is also seen from this gure that the temperature at any
given depth decreases as a function of time.
4. Conclusion
In this work, the bioheat transfer problems for the spherical living biological tissues have been investigated based on 1-D Pennes
model, and the analytical solution of the corresponding equation
has been found with the help of the method of separation of variables. The obtained exact solution of the problem has been used to
study the tissue temperature distribution in radial direction. Our results show that the scattering coecient and the heating powers

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

Fig. 10. Effects of the heating power (top) and scattering coecient (bottom) on temperature response at skin surface for f (t ) = 25 C. Top: = 200 m1 ; Bottom: P0 (t ) =
250 W/m2 .

Fig. 11. Transient temperatures at three positions when a constant spatial heating with P0 (t ) = 250 W/m2 and = 200 m1 was applied. The cooling medium temperature is taken
to be f (t ) = 25 C.

E. Kengne, A. Lakhssassi / Mathematical Biosciences 269 (2015) 19

P0 (t) and P1 (t) of the spatial heating are neglected for further analysis of the temperature distribution, though the tissue density, the
tissue specic heat and the cooling medium temperature are to be
take into account when investigating temperature distribution in the
biological bodies. The solution of the bioheat problem presented in
this paper is found to be very useful for a variety of bio-thermal
studies.
Acknowledgment
This work was supported by the Natural Sciences and Engineering
Research Council of Canada.
References
[1] H.H. Pennes, Analysis of tissue and arterial blood temperature in the resting human forearm, J. Appl. Physiol. 1 (1948) 93122.
[2] J. Chato, Measurement of thermal properties of biological materials, in: A. Shitzer,
R.C. Eberhart (Eds.), Heat transfer in Medicine and Biology, Vol. 1, Plenum Press,
NY, 1985, pp. 167173.
[3] H.F. Bowman, Estimation of tissue blood ow, in: A. Shitzer, R.C. Eberhart (Eds.),
Heat transfer in Medicine and Biology, Vol. 1, Plenum Press, NY, 1985, pp. 193
230.
[4] T.R. Gowrishankar, D.A. Stewart, G.T. Martin, J.C. Weaver, Transport lattice models
of heat transport in skin with spatially heterogeneous, temperature-dependent
perfusion, BioMed. Eng. OnLine 3 (42) (2004) 117.
[5] E. Kengne, A. Lakhssassi, R. Vaillancourt, W.-M. Liu, Monitoring of temperature
distribution in living biological tissues via blood perfusion, Eur. Phys. J. Plus 127
(89) (2012) 15.
[6] J. Lang, B. Erdmann, M. Seebass, Impact of nonlinear heat transfer on temperature
control in regional hyperthermia, IEEE Trans. Biomed. Eng. 46 (1999) 11291138.
[7] E. Kengne, A. Lakhssassi, R. Vaillancourt, Temperature distributions for regional
hypothermia based on nonlinear bioheat equation of pennes type: dermis and
subcutaneous tissues, Appl. Math. 3 (2012) 217224.
[8] Z.-S. Deng, J. Liu, Analytical study on bioheat transfer problems with spatial or
transient heating on skin surface or inside biological bodies, Trans. ASME 124
(2002) 638649.
[9] J. Liu, L.X. Xu, Boundary information based diagnostics on the thermal states of
biological bodies, Int. J. Heat Mass Transf. 43 (2000) 28272839.
[10] S.D. Burch, S. Ramadhyani, J.T. Pearson, Analysis of passenger thermal comfort in
an automobile under severe winter conditions, in: Proceedings of the ASHRAE
Transactions Winter Meeting, Atlanta, 1991, pp. 247257.
[11] E. Arens, P. Bosselmann, Wind, sun and temperature. predicting the thermal comfort of people in outdoor spaces, Build. Environ. 24 (1989) 315320.
[12] J.C. Chato, Measurement of thermal properties of growing tumors, Ann. N.Y. Acad.
Sci. 335 (1980) 6785.
[13] S. Weinbaum, L.M. Jiji, D.E. Lemons, Theory and experiment for the effect of vascular microstructure on surface tissue heat transfer. Part I. Anatomical foundation
and model conceptualization, ASME J. Biomech. Eng. 106 (1984) 321330.
[14] G.T. Martin, H.F. Bowman, W.H. Newman, Basic element method for computing
the temperature eld during hyperthermia therapy planning, Adv. Bioheat Mass
Transf. 231 (1992) 7580.
[15] J.W. Valvano, A.F. Badeau, In Vivo measurement of intrinsic and effective thermal conductivity using sinusoidally heated thermistors, in: Proceedings of the
6th Southern Biomedical Engineering Conference, 1987, pp. 14.
[16] J.W. Valvano, J.T. Allen, H.F. Bowman, The simultaneous measurement of thermal
conductivity, thermal diffusivity, and perfusion in small volumes of tissue, ASME
J. Biomech. Eng. 106 (1984) 192197.

[17] D.A. Torvi, J.D. Dale, A nite element model of skin subjected to a ash re, ASME
J. Biomech. Eng. 116 (1994) 250255.
[18] C.L. Chan, Boundary element method analysis for the bioheat transfer equation,
ASME J. Biomech. Eng. 114 (1992) 358365.
[19] B. Mochnacki, E. Majchrzak, Sensitivity of the skin tissue on the activity of external heat sources, Comput. Model. Eng. Sci. 4 (2003) 431438.
[20] K. Yue, X. Zhang, F. Yu, An analytic solution of one-dimensional steady-state
Pennes bioheat transfer equation in cylindrical coordinates, J. Thermal Sci. 13 (3)
(2004) 255258.
[21] M. Zhou, Q. Chen, Estimation of temperature distribution in biological tissue
by analytic solutions to Pennes equation, in: Proceedings of the 2nd International Conference on Biomedical Engineering and Informatics, BMEI, IEEE,
Nanjing, China, 2009.
[22] M. Zhou, Q. Chen, Study of the surface temperature distribution of the tissue affected by the point heat source, in: Proceedings of the IEEE International Conference on Bioinformatics and Biomedical Engineering and Informatics, 2007.
[23] E. Kengne, F.B. Hamouda, A. Lakhssassi, Extended generalized Riccati equation
mapping for thermal traveling-wave distribution in biological tissues through a
bio-heat transfer model with linear/quadratic temperature-dependent blood perfusion, Appl. Math. 4 (2013) 14711484.
[24] X. Liang, X. Ge, Y. Zhang, G. Wang, A convenient method of measuring the thermal
conductivity of biological tissue, Phys. Med. Biol. 36 (1991) 1599.
[25] R. Vyas, M.L. Rustgi, Greens function solution to the tissue bioheat equation, Med.
Phys. 19 (1992) 13191324.
[26] B. Gao, S. Langer, P.M. Corry, I.J. Hyperthermia, Application of the time-dependent
Greens function and fourier transforms to the solution of the bioheat equation,
Int. J. Hyperth. 11 (1995) 267285.
[27] J.W. Durkee, P.P. Antich, C.E. Lee, Exact-solutions to the multiregion timedependent bioheat equation. I: solution development, Phys. Med. Biol. 35 (1990)
847867.
[28] J.W. Durkee, P.P. Antich, Exact-solution to the multiregion time-dependent bioheat equation with transient heat-sources and boundary-conditions, Phys. Med.
Biol. 36 (1991) 345368.
[29] J.W. Durkee, P.P. Antich, Characterization of bioheat transport using exact solution
to the cylindrical geometry, multiregion, time-dependent bioheat equation, Phys.
Med. Biol. 36 (1991) 13771406.
[30] W.H. Newman, P.P. Lele, H.F. Bowman, Limitations and signicance of thermal
washout data obtained during microwave and ultrasound hyperthermia, Int. J.
Hyperth. 6 (1990) 771784.
[31] S. Hossain, F.A. Mohammadi, Development of an estimation method for interior
temperature distribution in live biological tissue of different organs, Int. J. Eng.
Appl. Sci. 3 (2) (2013) 4558.
[32] K.R. Holmes, Biological structures and heat transfer, in: Proceedings of Allerton
Workshop on The future of Biothermal Engineering, University of Illinois, 1997,
pp. 1437.
[33] B. Erdmann, J. Lang, M. Seebass, Optimization of temperature distributions for
regional hyperthermia based on a nonlinear heat transfer model, Ann. N. Y. Acad.
Sci. 858 (1998) 3646.
[34] S. Weinbaum, L.M. Jiji, D.E. Lemons, Theory and experiment for the effect of vascular microstructure on surface tissue heat transfer. Part I: anatomical foundation
and model conceptualization, ASME J. Biomech. Eng. 106 (1984) 321330.
[35] J. Liu, L.X. Xu, Estimation of blood perfusion using phase shift in temperature
response to sinusoidal heating at the skin surface, IEEE Trans. Biomed. Eng. 46
(1999) 10371043.
[36] G.T. Anderson, G. Burnside, A noninvasive technique to measure perfusion using
a focused ultrasound heating source and a tissue surface temperature measurement, in: R.G. Roemer, J.W. Valvano, L. Hayes, G.T. Anderson (Eds.), Advances in
measuring and computing temperatures in biomedicine: Thermal tomography
techniques, bio-heat transfer models, ASME, New York, 1990, pp. 3135.
[37] J.H. Li, H. Liang, Laser MedicineApplications of Laser in Biology and Medicine,
Science Press, Beijing, 1989.

You might also like