Вы находитесь на странице: 1из 10

Available online at www.sciencedirect.

com

International Journal of Heat and Mass Transfer 51 (2008) 3190–3199


www.elsevier.com/locate/ijhmt

A general bioheat transfer model based on the theory of porous media


A. Nakayama *, F. Kuwahara
Department of Mechanical Engineering, Shizuoka University, 3-5-1 Johoku, Hamamatsu 432-8561, Japan

Received 24 October 2006; received in revised form 20 April 2007


Available online 12 September 2007

Abstract

A volume averaging theory (VAT) established in the field of fluid-saturated porous media has been successfully exploited to derive a
general set of bioheat transfer equations for blood flows and its surrounding biological tissue. A closed set of macroscopic governing
equations for both velocity and temperature fields in intra- and extravascular phases has been established, for the first time, using the
theory of anisotropic porous media. Firstly, two individual macroscopic energy equations are derived for the blood flow and its sur-
rounding tissue under the thermal non-equilibrium condition. The blood perfusion term is identified and modeled in consideration of
the transvascular flow in the extravascular region, while the dispersion and interfacial heat transfer terms are modeled according to con-
ventional porous media treatments. It is shown that the resulting two-energy equation model reduces to Pennes model, Wulff model and
their modifications, under appropriate conditions. Subsequently, the two-energy equation model has been extended to the three-energy
equation version, in order to account for the countercurrent heat transfer between closely spaced arteries and veins in the circulatory
system and its effect on the peripheral heat transfer. This general form of three-energy equation model naturally reduces to the energy
equations for the tissue, proposed by Chato, Keller and Seiler. Controversial issues on blood perfusion, dispersion and interfacial heat
transfer coefficient are discussed in a rigorous mathematical manner.
Ó 2007 Published by Elsevier Ltd.

Keywords: Bioheat transfer; Dispersion; Porous media; Volume averaging

1. Introduction only the venous blood stream as the fluid stream equili-
brated with the tissue.
A number of bioheat transfer equations for living tissue In order to overcome these shortcomings, a considerable
have been proposed since the landmark paper by Pennes [1] number of modifications have been proposed by various
appeared in 1948, in which the perfusion heat source was researchers. Wulff [2] and Klinger [3] considered the local
introduced. Although Pennes model is often adequate for blood mass flux to account the blood flow direction, while
roughly describing the effect of blood flow on the tissue Chen and Holmes [4] examined the effect of thermal equil-
temperature, some serious shortcomings exist in his model ibration length on the blood temperature and added the
due to its inherent simplicity, as pointed out by Wulff [2], dispersion and microcirculatory perfusion terms to the
namely, assuming uniform perfusion rate without account- Klinger equation.
ing for blood flow direction, neglecting the important ana- All foregoing papers concerned mainly with the cases of
tomical features of the circulatory network system such as isolated vessels and the surrounding tissue. The effect of
countercurrent arrangement of the system, and choosing countercurrent heat transfer between closely spaced arter-
ies and veins in the tissue must be taken into full consider-
ation when the anatomical configuration of the main
supply artery and vein in the limbs is treated. Following
*
Corresponding author. Tel./fax: +81 534781049. the experimental study conducted by Bazett and his col-
E-mail address: tmanaka@ipc.shizuoka.ac.jp (A. Nakayama). leagues [5,6], Scholander and Krog [7] and Mitchell and

0017-9310/$ - see front matter Ó 2007 Published by Elsevier Ltd.


doi:10.1016/j.ijheatmasstransfer.2007.05.030
A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199 3191

Nomenclature

A surface area e porosity


Aint interface between the fluid and solid m kinematic viscosity
af specific surface area q density
bij Forchheimer tensor x perfusion rate
cp specific heat at constant pressure x0 net filtration rate
hf interfacial heat transfer coefficient
k thermal conductivity Special symbols
Kij permeability tensor ~
/ deviation from intrinsic average
nj unit vector pointing outward from the fluid side h/i volume average
to solid side h/if,s,a,v intrinsic average
p pressure
Sm metabolic reaction rate Subscripts and superscripts
T temperature a artery
ui velocity vector dis dispersion
V representative elementary volume f fluid
x, y Cartesian coordinates s solid
a thermal diffusivity v vein

Myers [8] investigated such an effect and successfully dem- comings in existing models will be overcome. We start with
onstrated that the countercurrent heat exchange reduces the case of isolated blood vessels and the surrounding tis-
heat loss from the extremity to the surroundings, which sue, to establish a two-energy equation model for the blood
could be quite significant due to a large surface to volume and tissue temperatures. We shall identify the terms
ratio. Keller and Seiler [9] established a bioheat transfer describing the blood perfusion and dispersion in the result-
model equation to include the countercurrent heat transfer, ing equation and revisit Pennes model, Wulff model and
using a one-dimensional configuration for the subcutane- their modifications.
ous tissue region with arteries, veins and capillaries. Wein- Subsequently, the two-energy equation model is
baum and Jiji [10] proposed a new model, which is based extended to the three-energy equation model, so as to
on some anatomical understanding, considering the coun- account for the effect of countercurrent heat transfer
tercurrent arterio–venous vessels. As pointed out by Roet- between closely spaced arteries and veins in the blood cir-
zel and Xuan [11], the model may be useful in describing a culatory system. In this model, three individual tempera-
temperature field in a single organ, but would not be con- tures are assigned for the arteries, veins and tissue. We
venient to apply to the whole thermoregulation system. shall examine the Keller and Seiler model [9] and Chato
Excellent reviews on these bioheat transfer equations may model [12] for the microcirculation as well as the model
be found in Chato [12] and Charny [13]. proposed by Xuan and Roetzel [16] for simulation of tran-
Khaled and Vafai [14] and Khanafer and Vafai [15] sient response of the human limb to an external stimulus.
stress that the theory of porous media is most appropriate Controversial issues on blood perfusion, dispersion and
for treating heat transfer in biological tissues since it con- heat transfer coefficient will be discussed in a rigorous
tains fewer assumptions as compared to different bioheat mathematical manner.
transfer equations. Roetzel and Xuan [11] and Xuan and
Roetzel [16] exploited the volume averaging theory
(VAT) previously established for the study of porous media 2. Volume averaging procedure
(e.g. Cheng [17], Nakayama [18]), to formulate a two-
energy equation model accounting for the thermal non- In an anatomical view, three compartments are identi-
equilibrium between the blood and peripheral tissue. In fied in the biological tissues, namely, blood vessels, cells
their model, the perfusion term is replaced by the interfa- and interstitium, as illustrated in Fig. 1. The interstitial
cial convective heat transfer term. This point should be space can be further divided into the extracellular matrix
examined since the interfacial convective heat transfer is and the interstitial fluid. However, for sake of simplicity,
different from perfusion heat transfer. Naturally, the for- we divide the biological tissue into two distinctive regions,
mer takes place even in the absence of the latter. namely, the vascular region and the extravascular region
In this study, we present a rigorous mathematical devel- (i.e. cells and the interstitium) and treat the whole anatom-
opment based on the volume averaging theory, so as to ical structure as a fluid-saturated porous medium, through
achieve a complete set of the volume averaged governing which the blood infiltrates. The extravascular region is
equations for bioheat transfer and blood flow. Most short- regarded as a solid matrix (although the extravascular fluid
3192 A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199

Blood vessels Cells


where Vf is the volume space which the fluid (blood) occu-
pies. Obviously, two averages are related as
f
h/i ¼ eh/i ; ð3Þ
where e  Vf/V is the local porosity, namely, the volume
fraction of the vascular space, which is generally less than
0.1. Following Cheng [17], Nakayama [18], Quintard and
Whitaker [19] and many others, we decompose a variable
into its intrinsic average and the spatial deviation from it:
f~
/ ¼ h/i þ /: ð4Þ
We shall exploit the following spatial average relationships:
Vascular region Extravascular region
f f f
h/1 /2 i ¼ h/1 i h/2 i þ h/ ~ 2 if ;
~ 1/ ð5Þ
Fig. 1. Schematic view of biological tissue.
  Z
o/ oh/i 1
¼ þ /ni dA or
oxi oxi V Aint
is present), and will be simply referred to the ‘‘tissue”  f f Z
o/ 1 oeh/i 1
region to differentiate it from the ‘‘blood” region. ¼ þ /ni dA; ð6a; bÞ
oxi e oxi V f Aint
Thus, we shall try to apply the principle of heat and fluid
flow in a fluid-saturated porous medium to derive a set of and
the volume averaged governing equations for the bioheat  
transfer and blood flow. In order for the volume averaging o/ oh/i
¼ ; ð7Þ
(smoothing process) to be meaningful, we consider a con- ot ot
trol volume V in a fluid-saturated porous medium, as
where Aint is the local interface between the blood and solid
shown in Fig. 2, whose length scale V1/3 is much smaller
matrix, while ni is the unit vector pointing outward from
than the macroscopic characteristic length V 1=3 c , but, at
the fluid side to solid side. The similarity between the vol-
the same time, much greater than the microscopic (anatom-
ume averaging and the Reynolds averaging used in the
ical structure) characteristic length (see e.g. Nakayama
study of turbulence is quite obvious. However, it should
[18]). Under this condition, the volume average of a certain
be noted that the present volume averaging procedure is
variable / is defined as
somewhat more complex than the Reynolds averaging pro-
Z
1 cedure, since it involves with surface integrals, as clearly
h/i  /dV : ð1Þ seen from (6).
V Vf
We subdivide the anatomic structure into the blood
phase (fluid phase) and the tissue and other solid tissue
Another average, namely, intrinsic average, is given by
phase (solid matrix phase), in which metabolic reactions
Z
f 1 may take place. We shall consider the microscopic govern-
h/i  /dV ; ð2Þ ing equations, namely, the continuity equation, Navier–
V f Vf
Stokes equation and energy equation for the blood phase
and the heat conduction equation for the solid matrix
phase.
For the blood phase:
ouj
¼ 0; ð8Þ
oxj
 
oui o 1 op o oui ouj
þ uj ui ¼  þ mf þ ; ð9Þ
ot oxj q oxi oxj oxj oxi
   
oT o o oT
qf cp f þ uj T ¼ kf : ð10Þ
ot oxj oxj oxj
For the solid matrix phase:
 
V oT o oT
qs cs ¼ ks þ Sm; ð11Þ
ot oxj oxj
Vc where the subscripts f and s stand for the fluid and solid,
respectively. It is assumed that the fluid (blood) is incom-
Fig. 2. Control volume in a porous medium. pressible and Newtonian, and all properties are constant.
A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199 3193
Z   
3. Volume averaged continuity and momentum equations for 1 p oui ouj o f
 þ mf þ nj dA  h~uj ~ui i
blood flow Vf Aint qf oxj oxi oxj
mf
Let us integrate the continuity Eq. (8) over a local con- ¼  ehui if  be2 ðhuk if huk if Þ1=2 hui if ð15Þ
K
trol volume using the formula (6b) as
such that
f Z !
oehuj i 1 f f f f
þ uj nj dA ¼ 0; ð12Þ ohui i o f f 1 ohpi o ohui i ohuj i
oxj V Aint þ huj i hui i ¼  þ mf þ
ot oxj q oxi oxj oxj oxi
where Aint is the local interface between the blood and solid mf  1=2
f f f f
 ehuj i  bij e2 huk i huk i huj i ;
matrix within the control volume V, while nj is the unit vec- K ij
tor pointing outward from the fluid side to solid side. For ð16Þ
sake of simplicity, the porosity e is assumed to vary moder-
ately within a porous medium. where Kij and bij are the permeability and Forchheimer ten-
The second term describes the volume rate of the fluid sors, respectively. These tensors, which depend on the ana-
bleeding off to the solid matrix through the interfacial vas- tomical structure, can be determined following the
cular wall, as illustrated in Fig. 3. In most microcirculatory procedure established for anisotropic porous structure
systems of the body, there is a net filtration of fluid from (Nakayama et al. [21]), as sufficient information on the
the intravascular to the extravascular compartment, such anatomical structure and properties is provided For the
that capillary fluid filtration exceeds reabsorption. How- vessels of sufficiently small diameter, the foregoing equa-
ever, this would not cause fluid to accumulate within the tion reduces to Darcy’s law:
interstitium since the lymphatic system removes excess fluid
from the interstitium and returns it back to the intravascu- 1 ohpif mf
  huj i ¼ 0; ð17Þ
lar compartment, as indicated in the figure. Thus, the sec- q oxi K ij
ond term describing the net filtration is negligibly small,
such that Eq. (12) reduces to where huji = ehujif is the Darcian velocity (i.e. apparent
velocity). We may use the Darcy law for most tissue regions
ohuj i except for the regions where large arteries or veins are
¼ 0: ð13Þ
oxj located.
Accordingly, the Navier–Stokes equation (9) may be inte-
grated to give 4. Two-energy equation model for blood flow and tissue
 
ohui if o 1 ohpif o ohui i ohuj i Before actually integrating the energy Eq. (10), it may be
þ huj if hui if ¼  þ mf þ
ot oxj qf oxi oxj oxj oxi quite instructive to focus our attention on the volume aver-
Z   
1 p oui ouj age of the convection term. Using the Eqs. (5) and (6), it is
þ  þ mf þ nj dA
V f Aint q oxj oxi straightforward to show
o  f
 h~uj ~ui if : o o o
oxj e qf cp f uj T ¼ qf cp f huj ihT if þ uj T~ if
eq cp h~
oxj oxj oxj f f
ð14Þ Z
1 
þ qf cp f uj T nj dA; ð18Þ
In order to close the foregoing macroscopic momentum V Aint
Eq. (14), the terms associated with the surface integral
are modeled according to Vafai and Tien [20] as where the first term on the right hand-side describes the
macroscopic convection, while the second term on the
right hand-side takes account of the thermal dispersion
(Nakayama et al. [22]). It is the last term on the right
f
ρ f ωc p T
f
hand-side that corresponds to the blood ‘‘perfusion” heat
source. Thus, the blood perfusion heat source term is iden-
tified as an extra surface integral term resulting from
changing the sequence of integration and derivation, as
we obtain the macroscopic energy equation by integrating
the microscopic convection term over a local control
volume.
ρ f ωc p T
s
Having expanded the integrated convection term, we
f
may readily transform both the energy Eq. (10) for the
blood flow and the conduction Eq. (11) for the solid matrix
Fig. 3. Capillary blood flow and extravascular flow. into the corresponding volume averaged equations as
3194 A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199

For the blood phase: hf = Nu(kf/2R), such that afhf = Nu(ekf/R2), where Nu is
f the Nusselt number based on the local diameter of the vas-
ohT i o f cular tube. If, the local porosity e and specific surface area
eqf cp f þ qf cp f huj ihT i
ot oxj af are provided for the complex tissue–vascular structure,
Z !
o ohT i
f
kf we may estimate the interfacial heat transfer coefficient
f
¼ ek f þ Tnj dA  eqf cp f h~ uj T~ i using hf = Nu(kfaf/4e). Roetzel and Xuan [11] set Nu =
oxj oxj V Aint
Z Z 4.93 for both arterial and venous blood vessels. We may
1 oT 1  appeal to a numerical experiment proposed by Nakayama
þ kf nj dA  qf cp f uj T nj dA: ð19Þ
V Aint oxj V Aint et al. [23] for complex porous structures.
As for modeling the blood perfusion term, we may refer
For the solid matrix phase: back to Fig. 3, and note that the transcapillary fluid
s  s Z 
ohT i o ohT i ks exchange takes place between the blood and the surround-
ð1  eÞqs cs ¼ ð1  eÞk s  Tnj dA ing tissue. However, the fluid lost from the vascular space
ot oxj oxj V Aint
Z will be compensated by the flow of extravascular fluids
1 oT
 kf nj dA and lymph from the tissue to vascular space. It is quite rea-
V Aint oxj sonable to assume that extravascular fluids and all lymph
Z
1  in the tissue space have the same temperature as the tissue
þ qf cp f uj T nj dA þ ð1  eÞS m ;
V Aint itself. Thus, we assume that the transcapillary fluid
ð20Þ exchange takes place at the rate of x (m3/sm3) and model
the blood perfusion term as
s
where hTi is the intrinsic average of the solid matrix tem- Z
1 
perature. Note that the dispersion heat flux qf cp f h~uj T~ i ¼ f s
q cp uj T nj dA ¼ qf cp f xðhT i  hT i Þ: ð24Þ
uj T~ if appears in the volume averaged energy Eq.
eqf cp f h~ V Aint f f
(19) for the blood phase, which may well be modeled under
Note that the perfusion rate x, unlike that of Pennes, varies
the gradient diffusion hypothesis:
locally and we assume that its local value is provided every-
f
f ohT i where. Pennes found that his model fits the experimental
uj T~ i ¼ ek diskj
eqf cp f h~ ð21Þ data for x = 2  104–5  104 (m3/sm3). R
oxk
Furthermore,
R the surface integral terms kVf Aint Tnj dA and
A number of expressions have been proposed for the ther- ks
 V Aint Tnj dA present the tortuosity heat fluxes, which are
mal dispersion thermal conductivity k diskj . Nakayama et al. usually small, as convection dominates over conduction
[22] obtained a transport equation for the dispersion heat (see e.g. Nakayama et al. [24]). Therefore, their effects
flux vector, which naturally reduces to the foregoing gradi- may well be absorbed in effective thermal conductivities,
ent diffusion form. For a bundle of vessels of radius R, they as done by Xuan and Roetzel [16]. Having modeled the
obtained the following expression for the predominant ax- terms associated with dispersion, interfacial heat transfer,
ial component of k diskj : blood perfusion and tortuosity, the individual macroscopic
!2 energy equations may finally be written for the blood and
f f
1 qf cp f hui R q cp hui R
k disxx ¼ kf : f f <1 tissue phases as
48 kf kf For the blood phase:
ðcapillary blood vesselsÞ; ð22aÞ f
!7=8 ohT i o f
f f eqf cp f þ qf c p f huj ihT i
qf cp fhui R qf cp f hui R ot oxj
k disxx ¼ 2:55 Pr1=8 k f : >1 !
kf kf o ohT i
f
ohT i
f
¼ ek f þ ek disjk
ðlarge arteries and veinsÞ: ð22bÞ oxj oxj oxk
f s f s
In order to close the foregoing macroscopic energy Eqs.  af hf ðhT i  hT i Þ  qf cp f xðhT i  hT i Þ ð25Þ
(19) and (20), the terms associated with the surface integral,
describing the interfacial heat transfer and perfusion in which the left hand-side term denotes the macroscopic
between the fluid and solid, must be modeled. For the convection term, while the four terms on the right hand-
interfacial heat transfer, Newton’s cooling law may be side correspond to the macroscopic conduction, thermal
adopted as dispersion, interfacial convective heat transfer and blood
Z perfusion, respectively.
1 oT
kf nj dA ¼ af hf ðhT is  hT if Þ; ð23Þ For the solid tissue phase:
V Aint oxj 
s s
ohT i o ohT i
where af and hf are the specific surface area and interfacial ð1  eÞqs cs ¼ ð1  eÞk s þ af hf ðhT if  hT is Þ
ot oxj oxj
heat transfer coefficient, respectively. For the bundle of f s
vascular tubes of radius R, we have af = 2e/R and þ qf cp f xðhT i  hT i Þ þ ð1  eÞS m ð26Þ
A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199 3195

in which the left hand-side term denotes the thermal inertia af hf


xPennes ’ x þ : ð29Þ
term, while the four terms on the right hand-side corre- q f cp f
spond to the macroscopic conduction, interfacial convec-
tive heat transfer, blood perfusion heat source and Thus, Pennes’ perfusion rate may be regarded as an effec-
metabolic heat source, respectively. tive one that includes interfacial convective heat transfer
The resulting Eqs. (25) and (26) appear to be a correct as well. Pennes assumes that blood enters the smallest
form for the case of thermal non-equilibrium, and are vessels of the microcirculation at Ta0, where all heat trans-
expected to clear up possible confusions associated with fer between the blood and tissue takes place. The assump-
the blood perfusion term. The continuity Eq. (13), Darcy’s tion of the complete thermal equilibration with the
law (17) and the two-energy equations (25) and (26) form a surrounding tissue is valid only when Peclet number is suf-
closed set of the macroscopic governing equations. The ficiently small.
present model in a multi-dimensional and anisotropic form
is quite general and can be applied to find both velocity and 5.2. Wulff model and Klinger model
temperature fields, as we prescribe the spatial distributions
of permeability tensor, porosity, interfacial heat transfer Wulff [2] criticized the Pennes model, pointing out that
coefficient, metabolic reaction rate and perfusion rate. It the moving blood through a tissue convects heat in any
is interesting to note that, when the velocity field, porosity direction, not just in the direction of the local tissue tem-
and metabolic reaction are prescribed, we only need to perature gradient. He assumed that the blood temperature
know the local value of the lumped convection–perfusion hTif is equivalent to the tissue temperature within a tissue
parameter, namely, ðaf hf þ qf cp f xÞ (in addition to appropri- control volume and proposed a new bioheat transfer equa-
ate thermal boundary conditions) to solve the two-energy tion. The equation later generalized by Klinger [3] runs in
Eqs. (25) and (26) for the blood and tissue temperatures, our notation as
hTif and hTis. ohT i
s
ð1  eÞqs cs
5. Comparison of present and existing bioheat transfer  ot s s
o ohT i ohuj ihT i
models ¼ ð1  eÞk s  qf c p f þ ð1  eÞS m :
oxj oxj oxj
It should be noted that most existing bioheat transfer ð30Þ
models already reside in the present model based on the We can obtain a similar equation by combining Eqs. (25)
theory of porous media. We shall revisit some of the exist- and (26) setting hTif = hTis as follows:
ing models and try to generate them from the present gen-
s
eral model. ohT i o
ðeqf cp f þ ð1  eÞqs cs Þ þ q f cp f huj ihT is
ot oxj
 s s
5.1. Pennes model o ohT i ohT i
¼ ðek f þ ð1  eÞk s Þ þ ek disjk þ ð1  eÞS m :
oxj oxj oxk
Pennes model [1] in our notation runs as
ð31Þ
s  s
ohT i o ohT i
ð1  eÞqs cs ¼ ð1  eÞk s We can easily see that the foregoing equation reduces to the
ot oxj oxj
s
Klinger equation when the ratio of vascular volume to total
þ qf cp f xPennes ðT a0  hT i Þ þ ð1  eÞS m ; volume (i.e. porosity e) is sufficiently small. Since the poros-
ð27Þ ity is generally less than 0.1, the foregoing two equations
are quite close to each other.
where xPennes is the mean blood perfusion rate, while Ta0 is Another interpretation on the directional effect on the
the mean brachial artery temperature. We compare the tissue temperature field is possible. When the blood flow
Pennes model against the energy Eq. (26) for the solid tis- is strong enough to neglect the macroscopic diffusion, the
sue phase and find the following relationship: energy Eq. (25) for the blood flow reduces to
qf cp f xPennes ðT a0  hT is Þ o f f s f s
f s f s
qf cp f huj ihT i ¼ af hf ðhT i  hT i Þ  qf cp f xðhT i  hT i Þ
¼ af hf ðhT i  hT i Þ þ qf cp f xðhT i  hT i Þ ð28Þ oxj
ð32Þ
Perhaps, Pennes considered that the blood perfusion is the
predominant heat source for the tissue, and did not bother Substitution of the foregoing equation into the energy
to describe the interfacial convective heat transfer between equation for the tissue (26) yields the Klinger Eq. (30).
the blood and tissue via the vascular wall. Instead, he intro- The assumption implicit here is that the blood flow
duced Ta0 to adjust the total heat transfer, which takes velocity is sufficiently high that the ratio of the bulk con-
place as the blood enters and leaves the tissue. We may as- vection heat transfer to conduction heat transfer, namely,
sume Ta0 ’ hTif for small vessels, and find the Peclet number, is much greater than unity. Thus, the
3196 A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199

Klinger model applies to the tissue with comparatively which is close to the equation of Chen and Holmes, except
large vessels. that qf cp f xj ðT a  T t Þ is missing, as in the models of Wulff
and Klinger, since it should vanish, as we add Eqs. (25) and
5.3. Cheng and Holmes model (26).

Cheng and Holmes [4] assumed that all tissue-arterial


6. A general three-energy equation model for countercurrent
blood heat exchange occurs along the circulatory network
bioheat transfer
after the blood flows through the terminal arteries and
before it reaches the level of the arterioles, which
Bazett and his colleagues [6] found that the axial temper-
prompted them to propose the following bioheat transfer
ature gradient in the limb artery of human, under condi-
model:
tions of very low ambient temperature, is an order of
oT t o magnitude higher than under normal ambient conditions.
qc þ qf c p f huj iT t Their experimental finding brought attention to the role
ot oxj
  of countercurrent heat exchange in bioheat transfer. A
o oT t oT t schematic view of the tissue layer close to the skin surface
¼ ðek f þ ð1  eÞk s Þ þ kp
oxj oxj oxj is shown in Fig. 4, in which the arteries and veins are
paired, such that the countercurrent heat transfer takes
þ qf cp f xj ðT a  T t Þ þ ð1  eÞS m ; ð33Þ
place. Mitchell and Myers [8] mathematically modeled this
important role in a more general manner than that pre-
where
sented by Scholander and Krog [7] and demonstrated that
q ¼ eqf þ ð1  eÞqs ; ð34aÞ the countercurrent heat exchange reduces heat loss from
the extremity to the surroundings. However, their one-
c ¼ ðeqf cp f þ ð1  eÞqs cs Þ=q ð34bÞ dimensional model was not able to take account of either
metabolic reaction or perfusion bleed-off from the artery
and to vein. The foregoing survey prompts us to establish a
f s multi-dimensional model, which can be applied to the
T t ¼ ðeqf cp f hT i þ ð1  eÞqs cs hT i Þ=qc ð34cÞ
regions of extremity, where the countercurrent heat trans-
fer between closely spaced arteries and veins in the blood
is the temperature of the continuum based on a volume
circulatory system plays an important role in the peripheral
average. Moreover, xj is the perfusion bleed-off to the tis-
heat transfer from the extremity to the surroundings.
sue only from the micro-vessels past the jth generation of
Thus, we assign individual temperatures hTiahTiv and hTis
branching, while T a is the blood temperature at the jth gen-
to the arterial blood, venous blood and tissue, respectively,
eration of branching. Both xj and T a require the anatom-
to propose a general three-energy equation model as
ical data. Chen and Holmes also take account of the
follows:
‘‘eddy” conduction due to the random flow of blood, by
For the arterial blood phase:
introducing the thermal conductivity kp, which corre-
a
sponds to our dispersion thermal conductivity kdis. The en- ohT i o a a
ea qf cp f þ q f cp f ea huj i hT i
ergy equation similar to their Eq. (33) may be obtained by ot oxj
combining the two-energy equations (25) and (26) in the  a a
present model as o ohT i ohT i a s
¼ ea k a þ ea k disajk  aa ha ðhT i  hT i Þ
oxj oxj oxk
oT t o f
qc þ qf cp f huj ihT i a
ot oxj  qf cp f x0a hT i : ð37aÞ
!
f s f
o ohT i ohT i ohT i
¼ ek f þ ð1  eÞk s þ ek disjk þ ð1  eÞS m
oxj oxj oxj oxk
Cutaneous layer
ð35Þ

When the three temperature gradients on the right hand-


side are close and ek disjk ¼ k p djk , the foregoing equation re-
duces to Venous vessel
Arterial vessel
oT t o f
qc þ qf c p f huj ihT i
ot oxj
 
o oT t oT t Deep tissue layer
¼ ðek f þ ð1  eÞk s Þ þ kp þ ð1  eÞS m
oxj oxj oxj
Fig. 4. Schematic view of countercurrent heat exchange near the skin
ð36Þ surface.
A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199 3197

For the venous blood phase: Veins


v
ohT i o v v
ev qf cp f þ qf cp f ev huj i hT i
ot oxj
 v v
o ohT i ohT i
¼ ev k v þ ev k disvjk  av hv ðhT iv  hT is Þ Capillaries
oxj oxj oxk
v
 qf cp f x0v hT i : ð37bÞ Tissue space
For the solid tissue phase:
 
ohT is o ohT is a s
ð1  eÞqs cs ¼ ð1  eÞk s þ aa ha ðhT i  hT i Þ
ot oxj oxj Arteries
þ qf cp f x0a hT ia þ av hv ðhT iv  hT is Þ
Fig. 5. One-dimensional model for countercurrent heat exchange.
v
þ qf cp f x0v hT i þ ð1  eÞS m ; ð38Þ
where s
d2 hT i a s v s
e ¼ ea þ ev ð39Þ ð1  eÞk s þ aa ha ðhT i  hT i Þ þ av hv ðhT i  hT i Þ
dx2
a s
Since the arterial–venous anastomoses provide direct paths þ qf cp f x0 ðhT i  hT i Þ þ ð1  eÞS m ¼ 0 ð44Þ
from terminal arteries to veins, the net volume filtration
rates of the arterial and venous vessels, x0a and x0v , are which is almost identical to what we would get for the one-
no longer negligible for
R the
 peripheral
heat transfer of dimensional case from our multi-dimensional expression
this kind, such that Aint qf cp f uj T nj dA=V ¼ qf cp f x0 hT if . (43), except that the temperature difference in the perfusion
Accordingly, the velocity fields should be determined term somewhat differs from ours. Keller and Seiler ob-
from tained solutions assuming that the arterial blood enters
f a the peripheral region at the isothermal core temperature
oea huj i 1 ohpi m
þ x0a ¼ 0 and   ea huj ia ¼ 0 and that the venous blood is completely equilibrated with
oxj q oxi K aij the tissue at the cutaneous layer.
ð40a; bÞ
oev huj i
v
1 ohpi m
v 6.2. Chato model
v
þ x0v ¼ 0 and   ev huj i ¼ 0;
oxj q oxi K vij
Chato’s countercurrent heat transfer model [12] differs
ð41a; bÞ from Keller and Seiler [9] in its neglect of heat transfer
where ea and ev are the volume fractions of the arterial between the blood and tissue. In this way, he was able to
blood and that of the venous blood, respectively. For the concentrate on the two temperatures instead of three as
microcirculation of peripheral tissue in which capillaries in Keller and Seiler. Chato’s one-dimensional model can
provide a continuous connection between the terminal ar- easily be generated from our general expressions (37a)
tery and vein (i.e. arterial–venous anastomoses), we may and (37b) along with (40a) and (41a), dropping the thermal
readily set inertia and conduction terms as
x0a ¼ x0v ð42Þ d a a a v a
qf cp f ea hui hT i ¼ aa ha ðhT i  hT i Þ  qf cp f x0a hT i ;
dx
such that the present energy equation for the solid tissue
ð45aÞ
phase reduces to
 s d
ohT i
s
o ohT i qf cp f ev huiv hT iv ¼ av hv ðhT iv  hT ia Þ þ qf cp f x0a hT ia ;
ð1  eÞqs cs ¼ ð1  eÞk s þ aa ha ðhT ia  hT is Þ dx
ot oxj oxj ð45bÞ
v s
þ av hv ðhT i  hT i Þ
þ qf cp f x0a ðhT ia  hT iv Þ þ ð1  eÞS m : where
a a
ð43Þ ea hui ¼ u0  x0a x and ea hui ¼ u0 þ x0a x: ð46a; bÞ
Note that u0 is the apparent velocity at x = 0 and that the
6.1. Keller and Seiler model right hand-side terms in the two Eqs. (45a) and (45b) can-
cels out each other, as they should for this ‘‘perfect” heat
Keller and Seiler [9] noted that the axial temperature exchange system. Chato obtained arterial and venous tem-
gradient in the limb is much higher than the transverse perature profiles along the length of the vessels and demon-
one and considered an energy balance within a control vol- strated that the effect of perfusion bleed-off is to increase
ume for the idealized one-dimensional steady case, as illus- the heat transfer between the vessels as compared with
trated in Fig. 5, for which they proposed the case of constant mass flow rate.
3198 A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199

6.3. Roetzel and Xuan model transfer from the extremity to the surroundings. The
three-energy equations coupled with the continuity and
Roetzel and Xuan [11] used the theory of porous media Darcy’s laws may be solved to find both velocity and tem-
to simulate a transient response of the limb to external perature fields, as we prescribe the spatial distributions of
stimulus, in which the effect of the countercurrent heat permeabilities, volume fractions, interfacial heat transfer
exchange on the temperature response is expected to be sig- coefficients and perfusion rates.
nificant. Their energy equation for the tissue in our nota- As pointed out by Roetzel and Xuan [11], some physio-
tion runs as logical parameters such as porosity and specific surface
s  s area depend on such factors as the body temperature and
ohT i o ohT i
ð1  eÞqs cs ¼ ð1  eÞk s þ aa ha ðhT ia  hT is Þ interaction with the environment, as well as vasoconstric-
ot oxj oxj
tor and vasodilator mechanisms. These physiological
v s
þ av hv ðhT i  hT i Þ þ ð1  eÞS m : ð47Þ parameters and model constants, which should be deter-
Comparison of the foregoing equation against our expres- mined experimentally, are urgently in need. Shortage of
sion (43) for the tissue reveals that the perfusion term these experimental data, however, should not hinder us
a v
qf cp f x0a ðhT i  hT i Þ is missing. Obviously, they did from applying the present bioheat transfer model to certain
not retain the term describing the transcapillary fluid cases using some estimated values. It is believed that even
exchange via the applications with the estimated values do not affect
R  arterial–venous 0 f
anastomoses, namely,
explanation of the applicability of the present bioheat
Aint
qf c p f u j T n j dA=V ¼ q f c p f x hT i . If they did, they
would have obtained our expression (43), which may be transfer model. Such attempts are underway.
rearranged in their form as
References
s
ohT i
ð1  eÞqs cs
[1] H.H. Pennes, Analysis of tissue and arterial blood temperature in the
 ot s resting human forearm, J. Appl. Physiol. 1 (1948) 93–122.
o ohT i a s
¼ ð1  eÞk s þ ðaa ha þ qf cp f x0a ÞðhT i  hT i Þ [2] W. Wulff, The energy conservation equation for living tissue, IEEE
oxj oxj Trans. Biomed., Eng. BME-21 (1974) 494–495.
v s
þ ðav hv  qf cp f x0a ÞðhT i  hT i Þ þ ð1  eÞS m : ð48Þ [3] H.G. Klinger, Heat transfer in perfused biological tissue. II. The
‘macroscopic’ temperature distribution in tissue, Bull. Math. Biol. 40
In their model, the convection–perfusion parameters, (1978) 183–199.
namely, ðaf hf  qf cp f x0 Þ, are replaced by the interfacial [4] M.M. Chen, K.R. Holmes, Microvascular contributions in tissue heat
transfer, Ann. N.Y. Acad. Sci. 335 (1980) 137–150.
convective heat transfer coefficients, afhf. This difference [5] H.C. Bazett, L. Love, L. Eisenberg, R. Day, R.E. Forster, Temper-
should not be overlooked since the perfusion heat sources ature change in blood flowing in arteries and veins in man, J. Appl.
could be quite significant for the bioheat transfer in the Physiol. 1 (1948) 3–19.
extremities, as Chato [12] demonstrated using his model. [6] H.C. Bazett, E.S. Mendelson, L. Love, B. Libet, Precooling of blood
in the arteries, effective heat capacity and evaporative cooling as
factors modifying cooling of the extremities, J. Appl. Physiol. 1 (1948)
6.4. Concluding remarks 169–182.
[7] P.F. Scholander, J. Krog, Countercurrent heat exchange and vascular
A rigorous mathematical development based on the vol- bundles in sloths, J. Appl. Physiol. 10 (1957) 405–411.
ume averaging theory has been presented to give a correct [8] J.W. Mitchell, G.E. Myers, An analytical model of the counter-
current heat exchange phenomena, Biophys. J. 8 (1968) 897–911.
set of bioheat transfer equations for the blood flow and its
[9] K.H. Keller, L. Seiler, An analysis of peripheral heat transfer in man,
surrounding biological tissue. The blood perfusion heat J. Appl. Physiol. 30 (1971) 779–789.
source term is identified as an extra surface integral term [10] S. Weinbaum, L.M. Jiji, A two phase theory for the influence of
resulting from changing the sequence of integration and circulation on the heat transfer in surface tissue, in: M.K. Wells (Ed.),
derivation, as we obtain the macroscopic energy equation Advances in Boiengineering, ASME, New York, 1979, pp. 179–
by integrating the microscopic convection term within a 182.
[11] W. Roetzel, Y. Xuan, Transient response of the human limb to an
local control volume. Using this general bioheat transfer external stimulus, Int. J. Heat Mass Transfer 41 (1998) 229–239.
model, we have revisited existing bioheat transfer models, [12] J.C. Chato, Heat transfer to blood vessels, ASME J. Biomech. Eng.
such as Pennes, Wulff, Klinger, Cheng and Holmes, so as 102 (1980) 110–118.
to point out possible shortcomings in the models. [13] C.K. Charny, Mathematical models of bioheat transfer, Advances in
The two-energy equation model has been further Heat Transfer, vol. 22, Academic Press, New York, 1992, pp. 19–155.
[14] A.-R.A. Khaled, K. Vafai, The role of porous media in modeling flow
extended to the three-energy equation model, to investigate and heat transfer in biological tissues, Int. J. Heat Mass Transfer 46
the countercurrent heat exchange between the arterial and (2003) 4989–5003.
venous blood vessels in the circulatory system. Keller and [15] K. Khanafer, K. Vafai, The role of porous media in biomedical
Seiler model and Chato model may easily be generated, engineering as related to magnetic resonance imaging and drug
delivery, Heat Mass Transfer 42 (2006) 939–953.
writing the present model for the idealized one-dimensional
[16] Y. Xuan, W. Roetzel, Bioheat equation of the human thermal system,
case. The present three-energy equation model in a multi- Chem. Eng. Technol. 20 (1997) 268–276.
dimensional and anisotropic form is found to be quite [17] P. Cheng, Heat transfer in geothermal systems, Advances in Heat
general and can be applied to all regions peripheral heat Transfer, vol. 14, Academic Press, New York, 1978, pp. 1–105.
A. Nakayama, F. Kuwahara / International Journal of Heat and Mass Transfer 51 (2008) 3190–3199 3199

[18] A. Nakayama, PC-aided numerical heat transfer and convective flow, [22] A. Nakayama, F. Kuwahara, Y. Kodama, A thermal dispersion flux
CRC Press, Boca Raton, 1995. transport equation and its mathematical modelling for heat and fluid
[19] M. Quintard, S. Whitaker, One and two equation models for flow in a porous medium, J. Fluid Mech. 563 (2006) 81–96.
transient diffusion processes in two-phase systems, Adv. Heat [23] A. Nakayama, F. Kuwahara, T. Hayashi, Numerical modelling for
Transfer 23 (1993) 369–465. three-dimensional heat and fluid flow through a bank of cylinders in
[20] K. Vafai, C.L. Tien, Boundary and inertia effects on flow and heat yaw, J. Fluid Mech. 498 (2004) 139–159.
transfer in porous media, Int. J. Heat Mass Transfer 24 (1981) 195– [24] A. Nakayama, F. Kuwahara, M. Sugiyama, G. Xu, A two-energy
203. equation model for conduction and convection in porous media, Int.
[21] A. Nakayama, F. Kuwahara, T. Umemoto, T. Hayashi, Heat and J. Heat Mass Transfer 44 (2001) 4375–4379.
fluid flow within anisotropic porous medium, J. Heat Transfer 124
(2002) 746–753.

Вам также может понравиться