Mathematical Modeling and Numerical Simulation of Heat Transfer F
Mathematical Modeling and Numerical Simulation of Heat Transfer F
2021
Recommended Citation
Dehdashti, Esmaeil, "MATHEMATICAL MODELING AND NUMERICAL SIMULATION OF HEAT TRANSFER
FROM ISOLATED OBJECTS", Open Access Dissertation, Michigan Technological University, 2021.
https://doi.org/10.37099/mtu.dc.etdr/1231
By
Esmaeil Dehdashti
A DISSERTATION
DOCTOR OF PHILOSOPHY
2021
Mechanics.
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Péclet numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
iii
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
port . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5.1 Limit of Pe 1 . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5.2 Limit of Pe 1 . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
conductivity fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
dition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.3.3 Bridging results for limits of low and high Péclet numbers . . 65
4.4 Variation of Nusselt number for uniform heat flux boundary condition 66
iv
4.4.2 Limit of advection-dominated heat transport . . . . . . . . . . 69
4.4.3 Bridging results for limits of low and high Péclet numbers . . 70
4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
v
List of Figures
Stokes flow at (a) low and (b) high Péclet numbers. The black lines
and arrows show the flow streamlines and flow direction, respectively. 2
2.1 (a) Oblate spheroidal coordinates in a meridian plane. The arrows show
and η depicted in cyan and gray, respectively. The red curve in (a) and
2.2 (a) Bispherical coordinates in a meridian plane. The arrows show the
β depicted in cyan and gray, respectively. The red curves in (a) and
vi
whereas the right panel presents those obtained by setting m = 19,
considered here as exact results. (c) The Nusselt number Nu versus the
aspect ratio ε for oblate spheroids. Red, blue, and black lines represent
Nu of sphere and disk, and m = 19, respectively. The inset shows the
meridian planes) around two spheres whose centers are three radii
apart ( = 2/3). The left panel illustrates the results of the four-term
1 − , where the dashed and dotted lines represent the asymptotic ap-
inset shows the relative error of the approximations. (d) The Nusselt
for a pair of identical spheres. Red, blue, green, and black lines repre-
tively. The inset shows the relative difference between the approximate
vii
3.1 Schematic of a stationary particle of arbitrary shape, with surface Sp
and unit outward normal vector n, in a uniform fluid flow. The dashed
3.3 Numerically calculated plots of the Nusselt number versus Péclet num-
ber for forced convection heat transfer from spheroids of various aspect
Eqs. (3.53) and (3.54), respectively. The cut-off Péclet numbers for
the aspect ratios ε = 0.2, 0.5, 1, 2, 5 in (a) and (c) are, respectively,
viii
and the predictions of Eq. (4.71) for NuQ . . . . . . . . . . . . . . . . 72
ix
Preface
This dissertation includes contents that were previously published as journal articles,
• Conduction heat transfer from oblate spheroids and bispheres, S. Jafari Kang,
• Forced convection heat transfer from a particle at small and large Peclet num-
bers, E. Dehdashti and H. Masoud, Journal of Heat Transfer 142, 061803 (2020);
As evidenced from the list of authors, the first and last publications were collaborative
works involving, respectively, Saeed Jafari Kang and Meghdad Razizadeh. In the first
simulations, and writing of the manuscript. And, in the second one, the primary
contributors were the author of this dissertation (Esmaeil Dehdashti) and his advisor
(Dr. Hassan Masoud), with Meghdad Razizadeh being only involved in some early
analysis.
x
Abstract
In the area of heat transfer, like other fields of science and engineering, full- and
that can be used to identify relevant dimensionless parameters, to obtain basic insights
into the phenomena under consideration, to quickly quantify the effects of key factors,
and, ultimately, to pave the way for understanding more complex problems arising
in practice. These solutions can also serve as excellent benchmarks for calibrating
with the ultimate goal of deriving analytical or approximate expressions for the Nus-
selt number (denoted by Nu), which is a key dimensionless parameter that quantifies
the transfer of heat to and from a surface. First, we consider heat transfer by conduc-
tion from oblate spheroidal and bispherical surfaces into a stationary, infinite medium.
The surfaces are presumed to maintain a constant heat flux. Assuming steady-state
condition and uniform thermal conductivity, we analytically solve the Laplace equa-
tion for the temperature distribution and discuss the challenge of dealing with the
Neumann (uniform flux) versus more convenient Dirichlet (isothermal) boundary con-
dition. The solutions are obtained in boundary-fitting coordinate systems using the
for the average Nusselt number are presented along with their approximations.
Next, we examine forced convection heat transfer from a single particle in uniform
xi
laminar flows. Asymptotic limits of small and large Péclet numbers (denoted by Pe)
for the heat transfer coefficient that is valid for arbitrary particle shapes and Reynolds
with a constant heat flux surface condition in the limits of Pe 1 and small or
moderate Reynolds numbers. Specific results are given for the heat transfer from
Finally, we revisit the problem of steady-state heat transfer from a single particle
in a uniform laminar flow with the assumption that the thermal conductivity of the
and scaling analyses to derive approximate expressions for the Nusselt number of
arbitrarily shaped particles. The results cover the entire range of the Péclet number.
We find that, for a constant temperature boundary condition and fixed geometry, the
Nusselt number is essentially equal to the product of two terms, one of which is only a
function of Pe while the other one is nearly independent of Pe and mainly depends on
that, in contrast, when a uniform heat flux is imposed on the surface of the particle,
xii
Chapter 1: Background and significance
1.1 Introduction
Fundamental problems in any field of science and engineering are the foundational
blocks upon which the rest of the filed is formed. Often, complex phenomena and in-
tricate systems are interpreted and analyzed through the lens of these basic problems,
which highlights their importance in advancing the state-of-the-art and in making new
discoveries. This is certainly the case in thermal sciences, which among other topics,
deals with the transfer of thermal energy (i.e., heat) between objects and systems.
between small objects (i.e., particles) and their surrounding medium. The principal
heat transfer through theoretical examination of (i) conduction heat transfer from
oblate spheroids and bispheres, (ii) forced convection heat transfer from an arbitrarily-
shaped particle, and (iii) heat transfer from a single object in laminar flows of a
we put our work in context by reviewing some of the classical theoretical investigations
1
(a) (b)
on conduction and convection heat transfer from particles. In the next three chapters,
we discuss the details of our analysis for each of the above-mentioned problems. And,
in the last chapter, we summarize our study and give a few ideas for future directions.
parameter called the Péclet number (denoted by Pe), which measures the importance
conduction. When the Péclet number is very small (Pe 1), but finite, the transfer
of heat is still dominated by conduction (see, e.g., Fig. 1.1a). However, as Pe rises,
the influence of advection grows and eventually becomes comparable with that of
conduction when Pe ∼ O(1). With further increase of the Péclet number, especially
beyond O(102 ), advection becomes the dominant mode of heat transfer and conduction
2
will be limited to a narrow region (known as the boundary layer) close to the surface
very challenging, if not impossible. As a result, problems treated in the vast majority
Pe (i.e., Pe 1 and Pe 1) [1–16]. The latter choice makes the problem amenable
to the perturbation theory, using which very useful approximations can be derived
for the quantities of interest. One of such quantities is the dimensionless rate of total
heat transfer from a surface, which is known as the Nusselt number and is denoted
by Nu. In the following and beginning with the case of Pe = 0, we summarize several
regime under steady-state condition. We note that this review is not meant to be
comprehensive.
The literature on pure conduction is vast and most of the studies on this subject
center around the solution of the Laplace equation for the temperature distribution.
Once the temperature field is known, then the integral of the local heat flux on
the surface of the object is calculated to obtain the Nusselt number. Among many
other studies, Yovanovich [17] and Hahne and Grigull [18] analytically solved for and
catalogued Nu for a number of practical geometries. The Nusselt number for addi-
tional geometries are reported in classical heat transfer textbooks (see, e.g., [19]),
with references given to the original derivations. Furthermore, Brenner and Haber
3
[20] presented a general solution for the Laplace equation, using the symbolic oper-
ator method, that is valid for a general far-filed temperature distribution and is not
that the boundary value problems associated with pure conduction appear in other
fields such as electrostatics as well. And, they are often discussed in mathematical
physics textbooks.
In the limit of Pe 1, Brenner [4] showed, in his seminal work, that the leading-
order contribution of advection to the Nusselt number for the problem of uniform flow
Nusselt number of the same problem. This surprisingly general result was obtained by
using a singular perturbation expansion for the temperature and by the application of
pure conduction does not even depend on the details of the flow field. In the opposite
limit of Pe 1, Acrivos and Goddard [6, 7] generalized the work of Lighthill [1] in
their two-part study, where they developed a framework for calculating the Nusselt
number in cases of planar and axisymmetric laminar flows. These asymptotic results
for small and large Péclet number were later extended by [21] to problems with
Overall, our survey of the literature on conduction and convection heat trans-
fer from isolated objects reveals that while there exists a great body of theoretical
analyses, there are certain aspects that have received less attention. In particular, we
4
have noticed that the boundary conditions considered on the surface of the objects
have been limited, to a large extent, to the prescribed temperature condition. And,
therefore, there is a need for more extensive examinations of other boundary condi-
tion. The theoretical calculations detailed in the subsequent chapters are performed
5
Chapter 2: Conduction heat transfer from
oblate spheroids and bispheres
2.1 Introduction
In this chapter, we present analytical solutions for the problem of steady-state con-
duction heat transfer in an infinite medium due to the presence of hot/cold inclusions.
Specifically, two geometries for the heat source/sink are considered, namely an oblate
spheroid and a pair of spheres (see Figs. 2.1 and 2.2). For both cases, a uniform heat
flux is assumed to emanate from the surface of the inclusions. This boundary condi-
tion models many scenarios that appear frequently in engineering applications, such
as when a surface is covered by a thin layer of electric heater [19]. The assumption of
the Neumann boundary condition is the novel aspect of our work. In the following, we
first formulate the problem mathematically and then provide a detailed description
of the solutions for the two geometries of interest. Next, we discuss the results for
the temperature distribution and average Nusselt number and, finally, we give a brief
6
2.2 Problem formulation and solutions
Consider one or more objects surrounded by an infinite medium at rest. Suppose that
rate and that the temperature at infinity is maintained at a constant value. Of in-
assuming that the transport of heat is dominated by conduction and that the ther-
mal conductivity is constant. Given the above conditions, the boundary-value problem
∇2 T ? = 0 with
(2.1)
? ? ?
− k n · ∇T = qs for r ∈ Sp and T → T∞ as r → ∞,
medium, n is the unit vector outward normal to the surface of the objects denoted
?
by Sp , r is the position vector with magnitude r = |r|, and qs and T∞ are constants.
Let ` be a characteristic length scale of the problem. Then, upon the change of
variables T = k (T ? − T∞
?
) /qs `, r̃ = r/`, and r̃ = r/`, Eq. (2.1) simplifies to the
∇2 T = 0 with
(2.2)
n · ∇T = −1 for r̃ ∈ Sp and T → 0 as r̃ → ∞.
7
Sp
Nu = , (2.3)
2πT
where Sp represents the dimensionless surface area of the object and T is the mean
value of T on Sp . Below, we solve Eq. (2.2) and calculate the average temperature for
cases in which Sp represents the surface of an oblate spheroid and a pair of spheres,
respectively. For each case, the solution is obtained via the method of separation
results against those obtained from the numerical solution of Eq. (2.2). A second-
order finite volume method as implemented in OpenFOAM (see, e.g., [22]) is used
Consider an oblate spheroid of equatorial radius ` and aspect ratio (ratio of polar to
nate system located at the center of the spheroid such that the z axis coincides with
solve Eq. (2.2), we adopt an oblate spheroidal coordinate system (ξ, η, ϕ) defined as
8
Figure 2.1: (a) Oblate spheroidal coordinates in a meridian plane. The
arrows show the direction of the unit vectors eξ and eη . (b) Surfaces of
constant ξ and η depicted in cyan and gray, respectively. The red curve in
(a) and its corresponding surface in (b) represent a hot/cold oblate spheroid
that releases/absorbs heat at a constant uniform rate.
[23, 24]
p
x / cos ϕ = y / sin ϕ = c (1 + ξ 2 ) (1 − η 2 ), z = c ξ η, (2.4)
√
where c = 1 − ε2 is the (dimensionless) radius of the focal circle and
As shown in Fig. 2.1, the surfaces of constant ξ and η are oblate spheroids and one-
√
sheet hyperboloids of revolution, respectively. In particular, ξ = ξ0 = ε / 1 − ε2
9
In the orthogonal curvilinear coordinate system (ξ, η, ϕ), Eq. (2.2) takes the form
of [23, 24]
2 1 ∂ ∂T
2 ∂ 2 ∂T
∇T = 2 2 1+ξ + 1−η
c (ξ + η 2 ) ∂ξ ∂ξ ∂η ∂η
1 ∂ 2T
+ = 0 with (2.7)
c2 (1 + ξ 2 ) (1 − η 2 ) ∂ϕ2
1 ∂T
= −1 and T → 0 as ξ → ∞.
hξ ∂ξ ξ=ξ0
∂ ∂T
2 ∂ 2 ∂T
1+ξ + 1−η = 0. (2.8)
∂ξ ∂ξ ∂η ∂η
Starting with the ansatz T (ξ, η) = X(ξ) H(η), where X and H are to-be-determined
functions, and following the steps involved in the separation of variables technique
[23, 24], it can be shown that the general solution of Eq. (2.8) is
∞
X 1
Am Pm (iξ) + A2m Qm (iξ) A3m Pm (η) + A4m Qm (η) ,
T = (2.9)
m=0
where m is an integer, A1m , . . . , A4m are constants, i2 = −1, and Pm and Qm are
Legendre functions of the first and second kind, respectively [25]. The latter function
is defined as
1 z+1 z z+1
Q0 (z) = ln , Q1 (z) = ln − 1,
2 z−1 2 z−1
(2.10)
(2n + 1) z Qn − n Qn−1
Qn+1 (z) = .
n+1
The function Pm (iξ) blows up as ξ → ∞ for m 6= 0 and so does Qm (η) at η =
10
±1 for all m. Therefore, to keep the solution finite, we need to set A1m = 0 for m 6= 0
and A4m = 0 for all m. By demanding T to vanish at infinity, we find that A10 is zero,
too, since P0 (iξ) = P0 (η) = 1 and Qm (iξ) decays to zero for large ξ. Thus, Eq. (2.9)
simplifies to
∞
X
T = Am Qm (iξ) Pm (η), (2.11)
m=0
where the constant coefficients Am are determined by applying the constant flux
boundary condition on Sp :
s
∞
1 ∂T i 1 + ξ02 X
= 2 2
Am Q0m (iξ0 ) Pm (η) = −1, (2.12)
hξ ∂ξ ξ=ξ0 c ξ0 + η m=0
with Q0m (x) = dQm (x)/dx. Note that although the gradient of T in the direction
normal to the boundary is constant, its derivative with respect to ξ (coordinate normal
to the boundary) is equal to the scale factor hξ , which varies along the boundary. Had
would have been also independent of η and Eq. (2.8) would have further reduced to
cot−1 ξ cot−1 ξ
T = = . (2.13)
cot−1 ξ0 cos−1 ε
(
Z 1 2 / (2m + 1) if n=m
Pm (x) Pn (x) dx = , (2.14)
−1
0 if n 6= m
11
where n is an integer. Using this feature, the unknown coefficients are obtained as
Z 1
i 2m + 1
q
Am = 0 Pm (η) ξ02 + η 2 dη. (2.15)
Qm (iξ0 ) 2(1 + ξ02 ) −1
The integral in Eq. (2.15) is zero for odd values of m and is otherwise calculated via
[27]
Z 1 q
Pm (η) ξ02 + η 2 dη
−1
m/2 (2.16)
X Γ(m/2 + n + 1/2) I2n (ξ0 )
= 2m ,
n=0
Γ(2n + 1) Γ(m − 2n + 1) Γ(n − m/2 + 1/2)
Z 1
2ξ0
q
2n −2
I2n (ξ0 ) = η ξ02 + η 2 dη = 2 F1 −1/2, n + 1/2; n + 3/2; −ξ0 . (2.17)
−1 2n + 1
where
!
ξ02
Z q
Sp = 1 dS = 2 π 1 + p coth−1 1 + ξ02 . (2.19)
Sp 1 + ξ02
It is worth noting that the results of this section for an oblate spheroid can be
converted to those for a prolate spheroid by allowing ε to be greater than one and by
12
2.2.2 Temperature field around two identical spheres
Consider a pair of identical spheres, with radius `, that are placed distance 2`/ apart,
where 0 < < 1. Recall the Cartesian coordinate system of §2.2.1. Only this time,
the origin is located in the middle of the line that connects the center of the spheres
and the z axis is oriented in the direction of that line. The natural coordinate system
for dealing with this geometry is a bispherical coordinate system (ζ, β, ϕ) defined as
[23, 24]
c sin β c sinh ζ
x / cos ϕ = y / sin ϕ = , z= , (2.20)
cosh ζ − cos β cosh ζ − cos β
√
where c = −2 − 1 is half the (dimensionless) focal distance and
c c sin β
hζ = hβ = , hϕ = ρ = . (2.22)
cosh ζ − cos β cosh ζ − cos β
Figure 2.2 depicts the surfaces of constant ζ and β that are non-intersecting spheres
surrounding the foci (points located at z = ±c) and intersecting tori passing through
the focal points, respectively. In this curvilinear coordinate system, the thermally
active spheres are represented by ζ = ±ζ0 = cosh−1 −1 and (ζ, β) → (0, 0) correspond
to r̃ → ∞.
13
Figure 2.2: (a) Bispherical coordinates in a meridian plane. The arrows
show the direction of the unit vectors eζ and eβ . (b) Surfaces of constant ζ
and β depicted in cyan and gray, respectively. The red curves in (a) and their
corresponding surfaces in (b) represent a pair of identical hot/cold spheres
that release/absorb heat at a constant uniform rate.
The boundary conditions are, again, independent of ϕ, which simplifies the Laplace
equation to
∂ sin β ∂T ∂ sin β ∂T
+ = 0. (2.24)
∂ζ cosh ζ − cos β ∂ζ ∂β cosh ζ − cos β ∂β
Unlike Eq. (2.8), this equation is not simply separable and belongs to a class of partial
differential equations called R-separable equations [24]. Therefore, to solve Eq. (2.24),
14
√
we start with the ansatz T (ζ, β) = cosh ζ − cos β Z(ζ) B(β), where Z and B are
the unknown functions we seek to determine. Substituting the proposed form for T
into Eq. (2.24), we arrive at a pair of ordinary differential equations for Z and B.
Upon solving those equations, the following general solution for T emerges [23, 24]:
∞
X
p 1
T = cosh ζ − cos β Bm sinh (m + 1/2)ζ
m=0 (2.25)
2
3 4
+ Bm cosh (m + 1/2)ζ Bm Pm (cos β) + Bm Qm (cos β) ,
1 4
where Bm , . . . , Bm are constant coefficients.
4
As indicated before, the function Qm (cos β) is singular at cos β = ±1. Thus, Bm
vanish in order to retain the regularity of the solution. One can infer from the geometry
of the problem and the boundary conditions that the temperature distribution is
symmetric about the x-y plane, which means ∂T /∂ζ = 0 at ζ = 0. To satisfy this
1
condition, we must set Bm = 0. Conveniently, the infinity boundary condition is
already satisfied thanks to the term outside the summation that approaches zero as
(ζ, β) → (0, 0). Incorporating these results, the solution for T reduces to
∞
X
p
T = cosh ζ − cos β Bm cosh (m + 1/2)ζ Pm (cos β), (2.26)
m=0
where the constant coefficients Bm are determined by enforcing the normal gradient
15
where
Yet, again, we see that the application of the constant flux boundary condition results
in a more convoluted equation for the unknown coefficients compared to the situation
In addition to the orthogonality property Eq. (2.14), the following integral rela-
Utilizing the above relations in conjunction with Eq. (2.14), we obtain a recursive
B1 = F0 + B0 ,
(2.30)
Bm+1 = Fm + Gm Bm + (1 − Gm ) Bm−1 ,
where
√
2 2 sinh ζ0 e−(m+1/2)ζ0
Fm = − , (2.31)
(m + 1) sinh (m + 3/2)ζ0
16
1 2m cosh ζ0 sinh (m + 1/2)ζ0
Gm = 1+ . (2.32)
m+1 sinh (m + 3/2)ζ0
We are now left with the task of determining B0 , accomplished by demanding the
terms in the solution that correspond to large eigenvalues to be finite. This condi-
manipulations, yields
B0 = − lim Hm , (2.33)
m→∞
where
H0 = F0 , H1 = F1 + G1 F0 ,
(2.34)
Hm = Fm + Gm Hm−1 + (1 − Gm ) Hm−2 .
With the coefficients known, the mean T over the surface of each sphere is calculated
via
2 π c2 π
Z Z
1 T sin β
T = T dS = 2 dβ
Sp Sp Sp 0 (cosh ζ0 − cos β)
∞ (2.35)
sinh ζ0 X
Bm 1 + e−(2m+1)ζ0 .
= √
2 m=0
We begin with the results for oblate spheroids. First, we consider the limiting cases of
ε = 1 (sphere) and ε = 0 (disk). The results for the former have been already known
to be
T = 1/r̃, T = 1, Nu = 2. (2.36)
17
For the latter, however, we find
2
i (2m + 1) m!
if m ≡ 0 (mod2)
(m − 1) (m + 2) 2m [(m/2)!]2
Am = , (2.37)
m ≡ 1 (mod2)
0 if
Next, we examine other aspect ratios. Quite unexpectedly, our calculations indicate
that the series solution for the temperature (see Eq. (2.11)) is very well approximated
by its first two non-zero terms for the entire range of ε, i.e.,
q
Sp
T ≈ 1 + ξ0 2
cot−1 ξ
4π
5 2
2
Sp
+ 2 + 3ξ0 + 4 + 3ξ0 1 − (2.39)
64 2π
3ξ − (3ξ 2 + 1) cot−1 ξ
2
× 3η − 1 .
3 (1 + ξ02 ) ξ0 cot−1 ξ0 − (2 + 3ξ02 )
this result is provided in Fig. 2.3, where we present side-by-side contour plots of the
temperature field (in meridian planes) around an oblate spheroid of aspect ratio ε =
0.3. As you can see, the approximate results of Fig. 2.3a are barely distinguishable
Equally interesting, we also find that the average Nusselt number for oblate
spheroids varies almost linearly with the aspect ratio (see Fig. 2.3c). Therefore, the
18
(a) (b)
2
(c)
Two-term approximation
Linear interpolation
1.75 Exact
Nu
1.5
3
2
δ(%)
1.25
1
1
0.2 0.4 0.6 0.8 1
ε
Figure 2.3: (a)-(b) Contour plots of the (dimensionless) temperature field
θ (in meridian planes) around an oblate spheroid of aspect ratio ε = 0.3. The
left panel illustrates the results of the two-term approximation whereas the
right panel presents those obtained by setting m = 19, considered here as
exact results. (c) The Nusselt number Nu versus the aspect ratio ε for oblate
spheroids. Red, blue, and black lines represent the results corresponding to
m = 3, linear interpolation based on the Nu of sphere and disk, and m = 19,
respectively. The inset shows the relative difference between the approximate
and exact results.
19
Nu versus ε curve can be approximated by the line that passes through its end points,
i.e.,
3π 3π
Nu ≈ Nudisk + (Nusphere − Nudisk ) ε = + 2− ε. (2.40)
8 8
The maximum relative error of this approximation occurs at ε ≈ 0.5 and falls below
3%, which underscores its validity for all aspect ratios (see the inset of Fig. 2.3c).
Furthermore, from Fig. 2.3c and its inset, we learn that the Nusselt number calculated
We now analyze the results for bispheres. In the limit → 0, the spheres are very
far from each other and, therefore, the solution simplifies to Eq. (2.36) for a single
sphere. On the other hand, when → 1 (i.e., when the gap between the spheres
vanish), Eq. (2.26) for the temperature distribution does not immediately reduce to
a simple form. The average surface temperature and Nusselt number in this case are
Again, we observe that setting m = 3 in the series solution for T [Eq. (2.26)]
provides excellent results, accurate to within 5% of the converged solution for 0 <
< 4/5 (see, e.g., Figs. 2.4a and 2.4b). As the gap between the spheres narrows,
the error of the four-term approximation increases, which means that more terms are
needed to represent the temperature field accurately. For example, the error grows to
Given the recursive relation defined in Eq. (2.30), what carries the most weight in
20
(a) (b) 0
(c)
√
ln(1 − ) / 2 − 0.26085
√
−2 2
Exact
B0
−4
30
−6 20
δ(%)
10
0
−8
10−4 10−3 10−2 10−1 100
1−
6
(d)
4
1.75 δ(%) 2
0
Nu
21
calculating Bm is determining B0 . Figure 2.4c shows the variation of this coefficient as
√
a function of 1 − . As it can be seen, B0 approaches zero as 2 when the distance
√
between the spheres is large and it asymptotes to −∞ as ln(1 − ) / 2 − 0.26085
when the spheres almost touch each other. The asymptotic formulas can be used to
estimate, with reasonable accuracy, the value of B0 over a wide range of (see the
Lastly, our calculations indicate that the Nusselt number of the two-sphere system,
like that of the oblate spheroid changes almost linearly with (see Fig. 2.4d). Hence,
again, a linear interpolation using the Nu of a single sphere and a pair of touching
spheres can be employed to effectively approximate the curve of Nu versus (see the
inset of Fig. 2.4d). There exists another approach for approximating this curve based
on the asymptotic behavior of T in the limit → 0. When the spheres are widely
separated, they see each other as a point source/sink. Thus, to the leading order in
, the average surface temperature and consequently the Nusselt number take the
outperforms the linear interpolation for gap sizes larger than the radius of the spheres
2.4 Summary
We derived analytical solutions for the problems of conduction heat transfer from an
isolated oblate spheroid and a pair of spheres. The derivations were carried out in
22
curvilinear coordinate systems, befitting each geometry, using the method of separa-
tion of variables and eigenfunction expansion. While the solutions are in the form of
infinite series, we showed that considering only the first few terms provides excellent
results, often accurate to within a few percent of the exact values. We also found that
the Nusselt number in both cases considered varies rather linearly with respect to the
relevant dimensionless length scale of the problem. As we shall see in the following
chapters, our results can be used to develop perturbation solutions for the problems
of forced convection heat transfer from heated spheroids and bispheres in uniform
23
Chapter 3: Forced convection heat trans-
fer from a particle at small and
large Péclet numbers
3.1 Introduction
The transport of heat from a particle via an externally driven fluid flow is a phe-
nomenon commonly observed in natural and man-made systems. The ubiquity and
importance of forced convection heat transfer have led a large number of researchers
to study various aspects of this mode of heat transport. Among the investigations
surprisingly, however, the vast majority of theoretical studies in thhat area have been
limited to the case of an object with a known surface temperature distribution (see
e.g. [1, 4, 6, 7, 15, 30–36]), while little attention has been paid to the equally practical
problem of convection heat transfer from a particle with a prescribed surface heat
flux. Of course, in such a problem, the rate of heat transfer (i.e. the surface integral
of the imposed heat flux) is already known, but what is not known, and often sought
after, is the average surface temperature in response to the heat emanating from the
surface of the particle. For example, envisage a scenario where heat is dissipated from
24
an electronic element by blowing air over it. Assume that the rate at which the heat
is generated by the element is known. In this system, the goal is to set the flow speed
such that the average surface temperature stays well below a critical temperature,
It is not immediately obvious as to why cases with prescribed surface heat flux
condition has been overlooked by theoreticians. However, one might surmise that the
tor that has deterred them from considering this category of convection heat transfer
problems. To partly address this deficiency in the literature, in this chapter, we exam-
ine uniform laminar flows past a single hot/cold particle whose surface is presumed
expressions for the Nusselt number Nu (based on the average surface temperature)
in the limits of small and large Péclet numbers Pe. The accuracy of the formulas for
the specific case of a spheroidal particle in axisymmetric Stokes flow are tested via
In the rest of this chapter, we first pose the mathematical problem and describe
transport, respectively. Specific results are discussed next and a short summary is
25
3.2 Problem statement
stationary particle of arbitrary geometry and characteristic length scale ` (see Fig.
stant and e is a unit vector. Suppose that heat is released/absorbed from the surface
of the particle at a constant uniform rate qs and that the temperature vanishes at
infinity. Neglecting viscous dissipation and assuming that the fluid properties are con-
stant, the boundary-value problem that governs the steady-state distribution of the
Pe u · ∇T = ∇2 T with
(3.1)
n · ∇T = −1 for r ∈ Sp and lim T = 0,
r→∞
where the Péclet number is defined as Pe = ρU∞ cp `/k, with ρ, cp , and k being the
density, specific heat, and thermal conductivity of the fluid, respectively. Also, n is
the unit vector outward normal to the surface of the particle denoted by Sp and r
is the position vector with magnitude r = |r| (see Fig. 3.1). Here, the temperature,
We reiterate that the primary novelty of our study is the consideration of a Neumann
Sp
Nu = , (3.2)
2π T
26
Figure 3.1: Schematic of a stationary particle of arbitrary shape, with
surface Sp and unit outward normal vector n, in a uniform fluid flow. The
dashed line indicates an enclosing boundary in the far field.
where Sp represents the dimensionless surface area of the particle and T is the mean
variation of the Nusselt number (or equivalently T ) as a function of the Péclet number.
To this end, we use the ideas of the reciprocal theorem in conjunction with the method
that are valid in the limits of Pe 1 and Pe 1. Details of the calculations are
27
3.3 Perturbation solution in the limit of conduction-
dominated heat transport
Suppose that the Péclet number is small, but finite. In this limit, we seek to deter-
mine the O(Pe) contribution to the Nusselt number. It is well-known that a regular
perturbation expansion in terms of Pe is only valid in the vicinity of the particle, i.e.,
regardless of the magnitude of Pe, there exists a domain (r/` & O(Pe−1 )) where the
lar perturbation expansion is used that involves separate expansions covering regions
close to and far from the particle, i.e., the inner and outer regions, respectively (see,
e.g., [4, 32, 34]). The inner and outer expansions are matched asymptotically in an
intermediate region where both expansions are valid and, together, constitute a per-
Specifically, the inner expansion of the temperature field takes the form of
28
which results in
˜ T̃ (1) = ∇
e·∇ ˜ 2 T̃ (1) with lim T̃ (1) = 0. (3.6)
r̃→∞
The ∼ overebar in Eqs. (3.5) and (3.6) denotes that the temperature field and the
with r̃ = |r̃|. The remaining boundary conditions of Eqs. (3.4) and (3.6) are furnished
by enforcing
We first consider the solution for T (0) . To satisfy the matching requirement at the
0) solution of the original problem described in Eq. (3.1). Thus, far from the particle,
Sp Sp
T (0) = + O(r−3 ) = Pe + O(Pe−3 ). (3.8)
4πr 4πr̃
This means that the far-field temperature distribution, to the leading order, may
the “heat” center of the particle. Next, given Eq. (3.8) and applying the matching
29
(1) Sp 1
T̃ = exp − (r̃ − e · r̃) . (3.10)
4πr̃ 2
We now implement the ideas of the reciprocal theorem (see, e.g., [37, 38]). We
multiply the Laplace equation (3.4) by T and Eq. (3.1) by T (0) and then subtract the
Integrating this equation over the fluid domain V and using the divergence theorem,
we arrive at
Z Z Z
(0) (0)
T n · ∇T dS = T n · ∇T dS + Pe T (0) u · ∇T dV, (3.12)
Sp Sp V
where the integrands decay sufficiently fast for contributions from surfaces at infinity
Z Z (0)
(0) (1)
Sp T = T dS − Pe T (0) u · ∇T dV = Sp T + Pe T + o(Pe), (3.13)
Sp V
(0)
where T is the zeroth-order contribution to the average surface temperatures and
(1)
T is the O(Pe) correction to it. Given Eqs. (3.13), (3.3), and (3.5), we deduce
Z Z
(1)
− Sp T = T (0)
u · ∇T (0)
dV + e · T̃ (0) ˜ (1)
∇ T̃ − T̃ (0)
dr̃, (3.14)
V R3
where the second integral on the right-hand side is over the entire three-dimensional
real space R3 and T̃ (0) = Sp /4πr̃. The first integral on the right-hand side can be
30
written as
Z Z h i
(0) (0) (0) 2
2 T u · ∇T dV = ∇· T u dV
V V
Z Z (3.15)
(0) 2 (0) 2
=− T n · u dS − T n · u dS,
Sp S∞
where both surface integrals are zero because of the no penetration condition on Sp
Z
e· ˜ T̃ (0) dr̃
T̃ (0) ∇
R3
2 Z ∞ Z π Z 2π
Sp 1 2
=− sin θ cos ϕ dϕ dθ dr̃ = 0. (3.16)
4π 0 r̃ 0 0
e·r
Z
(1) Sp 1
T =− exp − (r̃ − e · r̃) dr̃. (3.17)
16π 2 R3 r̃4 2
Let ex , ey , and ez be the unit vectors of the Cartesian coordinate system (x, y, z)
located at the center of the particle, see Fig. 3.1. Then, expressing the position vector
and setting the arbitrary unit vector e to ez (for convenience), the above relation
simplifies to
Z ∞ Z π Z 2π
(1) Sp cos θ sin θ
T =−
16π 2 0 0 0 r̃
(3.18)
r̃ Sp
× exp − (1 − cos θ) dϕ dθ dr̃ = − ,
2 8π
31
which, in turn, yields
!
Nu(0)
(0) Sp (0)
T =T 1 − Pe (0)
+ o(Pe) = T 1 − Pe + o(Pe). (3.19)
8πT 4
Substitution of Eq. (3.19) into Eq. (3.2) gives the final expression for the Nusselt
number: 2
Sp Nu(0)
Nu = = Nu(0) + Pe + o(Pe). (3.20)
2πT 4
Of course, this result is consistent with the one obtained by Leal [34] for the special
There are a couple of important points to make here. First, thanks to the reciprocal
theorem-inspired approach that was adopted, the derivation of Eqs. (3.19) and (3.20)
did not require a detailed knowledge of the velocity field. All we utilized were the
facts that the flow is divergence-free and that it does not penetrate into the particle.
Even no-slip condition was not essential and we assumed no restriction on the flow
Reynolds number defined as Re = ρU∞ `/µ, where µ is the fluid viscosity. Second, and
perhaps equally notable, Eq. (3.20) for the dependence of the Nusselt number on the
Péclet number is identical to the formula obtained by Brenner [4] for an isothermal
particle. Equation (3.20) is also the same as the expression derived by Gupalo et
al. [39] for the Sherwood number (analogue of Nu for mass transfer problems) as
surface. A natural question to ask at this point is that over what range of Pe does
Eq. (3.20) produce accurate results? We will answer this question for the special case
32
of axisymmetric Stokes flow past a spheroids in §2.3.
Suppose that the Péclet number is large (i.e., Pe 1) and the Reynolds number
is small or moderate. In this limit, the temperature distribution outside the particle
is mainly restricted to a thin layer around the particle (see, e.g., Fig. 3.2), whose
thickness is proportional to Pe−1/3 (see e.g. [6]). This scaling can be deduced by
equating the order of magnitude of the advective and conductive terms in Eq. (3.1),
that, unlike the temperature field, the velocity field surrounding the particle is not
confined to a narrow region. Similar to the previous section, here also we wish to
Following the standard boundary layer theory (see, e.g., [6]), we assume that the
already known velocity field and the to-be-solved-for temperature distribution are
boundary layer coordinate system (x, y, ϕ), where the first two components measure,
respectively, the distance along and perpendicular to the surface of the particle in
the plane of flow (see Fig. 3.2) and the third component is the azimuthal angle for
axisymmetric and the distance from the x-y plane for two-dimensional problems. The
33
U∞
T
y
x
q = Constant
x=0
Th
erma
l bounda
ry layer
cosine of the angle between the axis of rotation and the tangent to Sp . Thus, κ =
In the vicinity of Sp , Eq. (3.1) can be expressed in terms of the boundary layer
coordinates as
ux ∂T uy ∂T
Pe +
hx ∂x hy ∂y
1 ∂ hy hϕ ∂T ∂ hx hϕ ∂T
= + with (3.22)
hx hy hϕ ∂x hx ∂x ∂y hy ∂y
∂T
= −1 and lim T = 0,
∂y y=0
y→∞
where
34
1 ∂ψ 1 ∂ψ
ux = and uy = − (3.23)
hϕ hy ∂y hϕ hx ∂x
are the velocity components in the x and y directions, respectively, with ψ being the
stream function. Remember that ∂T /∂ϕ and uϕ are zero. Since we are interested in
where
∂ux
2ψ2 = % = % τ0 (3.25)
∂y y=0
with τ0 being the dimensionless shear stress at the surface of the particle. Note that
Given how the thickness of the boundary layer scales with the Péclet number, we
consider
ỹ = Pe1/3 y
where
∂ dψ2 ∂ ∂2
L (0) = 2ỹ ψ2 − ỹ2 − % 2, (3.27a)
∂x dx ∂ỹ ∂ỹ
35
(1) 2 ∂ dψ3 ∂ ∂ ∂
L = 3ỹ ψ3 − ỹ3 − (α + %κ) ỹ . (3.27b)
∂x dx ∂ỹ ∂ỹ ∂ỹ
Taking the form of Eq. (4.38) into account, it is natural to expand the temperature
T = Pe−1/3 T (0) + Pe−1/3 T (1) + O(Pe−1 ), (3.28a)
(0) (1)
T = Pe−1/3 T + Pe−1/3 T + O(Pe−1 ), (3.28b)
where
Z xm −1 Z xm
T = % dx T |ỹ=0 % dx, (3.29a)
0 0
Sp
Nu(0) = (0)
, (3.29b)
2πT
2 2π T (1)
Nu(1) = − Nu (0)
, (3.29c)
Sp
maximum value of x, respectively. We note that, for two dimensional problems, the
quantities described by Eqs. (3.28) and (3.29) belong to the temperature boundary
layer that forms over one side of the stagnation point (see Fig. 3.2), as the energy
equation in the two layers can be treated separately, but in the same manner. Fur-
thermore, as clearly articulated by Acrivos and Goddard [6], the above expansions
are not expected to be fully applicable to all scenarios, such as where there exists
36
these expansions to insure the accuracy of the results.
(0) (1)
Below, we derive T and T , and, by extension, Nu(0) and Nu(1) . Substituting
for T in Eq. (3.27) and requiring the energy equation and its boundary conditions to
hold for all orders of Pe, we arrive at the following parabolic equations for T (0) and
T (1) :
We first consider Eq. (3.30a), which is indeed the standard boundary layer approxi-
mation of Eq. (3.27). One might be tempted to develop a similarity solution for this
equation, as done traditionally when Sp is considered isothermal (see, e.g., [40]). How-
ever, such an approach would fail here because of the Neumann boundary condition
Z x p p
t= 2ψ2 (s) %(s) ds and z = 2ψ2 (x) ỹ, (3.31)
0
∂T (0) ∂ 2 T (0)
2ψ2 % z − = 0 with (3.32)
∂t ∂z2
∂T (0) 1
= −√ , lim T (0) = 0, and lim T (0) = 0.
∂z z=0 2ψ2 z→∞ t→0
37
According to Sutton [41], the solution of the above problem is
t
∂T (0)
Z
(0)
T =− G(t, z; t̂, 0) dt̂, (3.33)
0 ∂z z=0
where
∂G ∂ 2 G
z − = δ(t̂, ẑ) with
∂t ∂z2
(3.36)
∂G
= lim G = lim G = 0.
∂z z=0 z→∞ t→t̂
Here, I−1/3 is the modified Bessel function of the first kind and of order −1/3 and
δ is the Dirac delta function. Replacing for G and the heat flux distribution in Eq.
(3.33), we find
Z t −1/2
1 −2/3
e−z / 9 (t−t̂) dt̂,
3
(0)
T = 1/3 t − t̂ 2ψ̂2 (3.37)
3 Γ(2/3) 0
where Γ is the gamma function and ψ̂2 = ψ2 (x̂). Hence, the leading-order contribution
to the average surface temperature can be obtained via (see Eqs. (3.28b) and (3.29a))
Z xm Z x
(0) 1 %̂
T = 1/3 Rx 2/3 dx̂ % dx, (3.38)
3 Γ(2/3) 0 m % dx 0 0 t − t̂
where %̂ = %(x̂) and t̂ = t(x̂). Remember that all needed to calculate t(x) is the
38
With the Green’s function known, the solution of Eq. (3.30b) can be written
formally as
Z t Z ∞
(1)
T = f (t̂, ẑ) G(t, z; t̂, ẑ) dẑ dt̂, (3.39)
0 0
where
L (1) T (0)
f (t, z) = −
2ψ2 %
(3.40a)
∂ 2 T (0) ∂T (0)
1 3
= −√ Az − Bz + C ,
2ψ2 % ∂z2 ∂z
3ψ3
A(x) = % − (α + %κ) , (3.40b)
2ψ2
!
d ψ3
B(x) = , (3.40c)
dx (2ψ2 )3/2
Substituting for T (0) in Eq. (3.40a), multiplying the result by Eq. (3.34), and carrying
out the ẑ integration (see, e.g., [6, 27]), we reach, after some simplifications,
2 Z x
(1) 1 1/3
T =− 1/3
t − t̂
3 Γ(2/3) 0
!
x̂
z3
Z
1
× exp − 2 2/3
9 t − t̂
0 ˆ
t − t̂ ˆ
t̂ − t̂
n h i
ˆ
× t − t̂ 3A + 9 t̂ − t̂ B̂ˆ (3.41)
ˆ
!
2 t̂ − t̂ z3
ˆ Ĉ − 2Aˆ
× 1 F1 2; ; + t − t̂
3 t − t̂ˆ 9 t − t̂
ˆ
!)
2 t̂ − t̂ z3
×1 F1 1; ; %̂ˆ dx̂
ˆ dx̂,
ˆ 9 t − t̂
3 t − t̂
ˆ = t(x̂),
where t̂ ˆ Aˆ = A(x̂), B̂ = B(x̂), Ĉ = C(x̂), %̂ˆ = %(x̂),
ˆ and 1 F1 is the confluent
39
hypergeometric function of the first kind [25]. The first correction to the mean surface
2 Z xm Z x
(1) 1 1 1/3
T = − R xm 1/3
t − t̂
0
% dx 3 Γ(2/3) 0 0
Z x̂ (
1 h
ˆ
× 2 2/3 t − t̂ 3A+
0 ˆ
t − t̂ ˆ
t̂ − t̂
)
i
ˆ B̂ + t − t̂
9 t̂ − t̂ ˆ Ĉ − 2Aˆ %̂ˆ dx̂
ˆ dx̂ % dx. (3.42)
We choose Stokes (zero Reynolds number) flow to exemplify the general results of the
biology, engineering, and material science. And, we choose spheroids because of their
practical significance. In what follows, we first present the results for the limits of
small and large Péclet numbers and then show the comparison with the full numerical
solution of Eq. (3.1). We acknowledge that the choice of Stokes flow does not affect
the low Pe analytical result. For the sake of completeness, we also provide the results
for spheroids with isothermal surface condition following the works of Brenner [4] and
40
3.5.1 Limit of Pe 1
Consider a spheroid of equatorial radius ` and aspect ratio (ratio of polar to equatorial
radius) ε. As discussed in §3.3, when the Péclet number is small, the details of the
flow field are irrelevant for calculating the Nusselt number to the leading order in Pe.
3/2
(0)
[− (1 + ξ02 )] S2p
Nu = ∞ , (3.43)
X Q2m (iξ0 ) 2
8π 2 ξ02 16m (4m + 1) 0 D
m=0
Q2m (iξ0 ) m
where
m
X Γ(m + n + 1/2)
Dm =
n=0
Γ(2n + 1) Γ(2m − 2n + 1) Γ(n − m + 1/2)
−2
2 F1 −1/2, n + 1/2; n + 3/2; −ξ0
× , (3.44)
2n + 1 !
ξ2
q
Sp = 2 π 1 + p 0 2 coth−1 1 + ξ02 , (3.45)
1 + ξ0
ε
ξ0 = √ . (3.46)
1 − ε2
Here, m and n are integers, Q2m are Legendre functions of the second kind [25] with
Q02m (x) = dQ2m (x)/dx, and 2 F1 is the hypergeometric function [25]. The summation in
the denominator of Eq. (3.43) converges very quickly, to the extent that taking only
two terms of the series produces results accurate to within 0.25% of the exact values.
41
It is interesting to contrast Eq. (3.43) with its counterpart for the problem of
Allowing for imaginary values of the square root and inverse cosine functions, this
expression and also the forthcoming Eq. (3.50) are valid for the entire range of ε. Here,
its analog for the constant heat flux surface condition. Clearly, Eq. (3.47) is far less
cumbersome than Eq. (3.43). That aside, the substitution of either Eq. (3.43) or (3.47)
in Eq. (3.20) gives the Nusselt number correct to the order of Pe for incompressible
3.5.2 Limit of Pe 1
Consider the spheroid of §3.5.1 with its axis of revolution coinciding with the z axis
O(Pe1/3 ) Nusselt number and its first correction, we need to calculate the integrals in
Eqs. (3.38) and (3.42), which involve the functions %(x), t(x), A(x), B(x), and C(x).
The first and last functions depend only on the geometry of the spheroid whereas the
remaining three are additionally dependent on the flow field. It is more convenient to
42
s
dx ε2 + (1 − ε2 ) η 2
= −hη = − and − 1 ≤ η ≤ 1, (3.48)
dη 1 − η2
stream function for axisymmetric Stokes flow past a spheroid is known (see, e.g.,
[28]). Granted this, we find, following the definitions given in §3.4, that
p
%(η) = 1 − η2, (3.49a)
r
F ε −1 p
t(η) = cos η − η 1 − η 2 , (3.49b)
16π
5ε2
1 2
A(η) = 1−ε − 2 (3.49c)
2hη ε η + ε2 (1 − η 2 )
r 3/2
η π 1 3ε2
B(η) =
3hη F (1 − η 2 ) ε [η 2 + ε2 (1 − η 2 )]2
× 1 + 3 ε2 − 1 1 − η 2 − 1 + ε2
(3.49d)
ε 1
C(η) = 1+ 2 , (3.49e)
hη η + ε2 (1 − η 2 )
where
3/2
8π (1 − ε2 )
F = √ (3.50)
(1 − 2ε2 ) cos−1 ε + ε 1 − ε2
is the magnitude of the Stokes drag experienced by the spheroid. This quantity is
made dimensionless by µU∞ `. The average surface temperature and its first correction
can now be calculated by replacing the foregoing relations in Eqs. (3.38) and (3.42).
(0) (1)
Substitution of T and T in Eqs. (3.29b) and (3.29c) then yields Nu(0) and Nu(1) .
Equations (3.49) can also be used to determine the O(Pe1/3 ) and O(1) contribu-
tions to the Nusselt number for high-Péclet number heat transfer from an isothermal
43
spheroid in axisymmetric Stokes flow. Availing ourselves of the general results of
Acrivos and Goddard [6], it can be shown, after much reduction, that
1
Nu(0) = (12πF ε)1/3 , (3.51a)
8 Γ(4/3)
(1) 4ε2 + 1
Nu(1) = Nusphere , (3.51b)
5ε
where
( −2/3
4 Γ(2/3) π γ − 12 sin 2γ
Z
(1) 5
Nusphere = 0.92301 = 1−
3 [Γ(1/3)]2 0 π
1/3 )
γ − 12 sin 2γ
1 − cos γ
× 1− sin2 γ dγ (3.52)
π π
Again, simplicity-wise, the contrast between the above formulas and those for Nu(0)
and Nu(1) is quite remarkable. Equation (3.51a) was also reported by Sehlin [9], though
in a different form. However, to the best of our knowledge, Eq. (3.51b) has not been
reported elsewhere, and is, therefore, another original contribution of this work.
Lastly, we note that the existence of a rear stagnation point on the spheroid in
Stokes flow renders the perturbation expansion described by Eq. (3.28a) invalid in
the vicinity of η = −1. Fortunately, however, the contribution of this singular region
to the Nusselt number is beyond O(1), and, hence, has no effect on Nu(1) and Nu(1)
(see [6]).
44
3.5.3 Comparison with direct numerical solutions
To find out the true limits for which the perturbation calculations of §3.5.1 and §3.5.2
for the Nusselt number are valid, we compare our theoretical results with those ob-
tained from the full numerical solution of Eq. (3.1). A second-order finite volume
method as implemented in OpenFOAM (see, e.g., [22]) is used to perform the numer-
ical calculations. The Stokes equations for the velocity field u is solved first using the
SIMPLE algorithm, and the advection-diffusion equation for the temperature distri-
bution T is treated next. The outer boundary at infinity is modeled as a large cylinder,
whose center coincides with the center of the spheroid. The diameter of the cylinder
is equal to its length, which is 200 times the semi-major axis of the spheroid. 2D
domain and grid-independence tests are performed by refining the mesh in the en-
tire domain and repeating the simulations. In all cases considered, the computational
grid is chosen such that the change in the results due to the refinement is marginal.
Figure 3.3 shows the results of the numerical calculations for the spheroids of vari-
ous aspect ratios. Interestingly, the plots of Nu versus Pe for constant heat flux and
isothermal boundary conditions are very much alike, not only qualitatively but also
Perhaps, the simplest way to construct a global approximation for the Nusselt
45
102
ε = 0.2
ε = 0.2
ε = 0.5
ε = 0.5
ε=1
Nusselt number
ε=1
ε=2
ε=2
101
ε=5
ε=5
100
10−3 10−2 10−1 100 101 102 103 104 105
Péclet number
Figure 3.3: Numerically calculated plots of the Nusselt number versus
Péclet number for forced convection heat transfer from spheroids of various
aspect ratios in an axisymmetric uniform Stokes flow subject to constant
heat flux (solid lines) and isothermal (dashed lines) boundary conditions on
the surface of the spheroid.
2
(0)
Nul
(0)
Nul + Pe if Pe < Pec
Nu = 4 , (3.53)
Pe1/3 Nu(0) + Nu(1)
if Pe ≥ Pec
h h
where Pec is the cut-off Péclet number and the subscripts l and h indicate that
the coefficients belong to the low and high Pe limits, respectively. Alternatively, the
Nusselt number may be approximated over the entire range of the Péclet number by
46
which only incorporates the leading order terms in the perturbation expansions of
Nu in the asymptotic limits of Pe. Figure 3.4 presents the percent difference between
the results of Fig. 3.3 and those of Eqs. (3.53) and (3.54). Overall, we see that both
approximations are quite accurate. Specifically, Eq. (3.53) provides more precise pre-
dictions when Pe 1 and Pe 1 whereas better estimates are given by Eq. (3.54)
at intermediate values of Pe, where the approximations deviate the most from the
numerical results. Finally, we note that the kinks in Figs. 3.4a and 3.4c are associated
with transition from low- to high-Péclet-number solutions at Pec , which vary from
0.2 to 1 for the range of aspect ratios considered. In comparison with the plots for
in these cases, both asymptotic formulas overestimate the Nusselt number (i.e., ∆ is
positive) in the neighborhood of the cut-off point and no discontinuity exist at Pec .
3.6 Summary
We examined the problem of heat transfer from a stationary hot (or cold) particle im-
the perturbation theory to derive two-term approximations for the average Nusselt
number in the asymptotic limits of the Péclet number. At small Pe, Nu was approxi-
mated as the summation of the conduction Nu and the O(Pe) correction. We showed
that, for arbitrary particle shapes and flow Reynolds numbers, the correction term is
equal to the square of the zeroth-order term divided by four (see Eq. (3.20)). At high
47
ε = 0.2 ε = 0.5 ε=1 ε=2 ε=5
15 5
(a) (b)
10
5 0
0
−5
(%)
(%)
−5
∆
∆
−10 −10
−15
−20 −15
10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe
15 5
(c) (d)
10
5 0
0
−5
(%)
(%)
−5
∆
∆
−10 −10
−15
−20 −15
10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe
Pe, the boundary layer theory was employed to analytically solve for the temperature
distribution within the thermal boundary layer up to O(Pe−2/3 ). The solutions were
used to calculate the O(Pe1/3 ) and O(1) contributions to the Nusselt number. These
both low and high Péclet number results are due to the assumption of constant heat
48
flux condition on the surface of the particle.
isymmetric Stokes flow past a spheroid. The specific results were then compared
against those obtained from the full numerical solution of the underlying conser-
vation of thermal energy equation. The comparisons confirmed the accuracy of the
49
Chapter 4: Heat transfer from a particle in
laminar flows of a variable ther-
mal conductivity fluid
4.1 Introduction
Heat transfer from a hot or cold object exposed to an external fluid flow is arguably
the most basic form of forced convection heat transfer encountered in industrial pro-
cesses and technological applications. When analyzing this category of heat transfer
problems, in many cases, it is well justified to assume that the fluid properties such
general, this assumption renders the energy equation (the partial differential equation
that governs the distribution of the fluid temperature) linear and decoupled from the
There exist, however, practical cases where at least one of the fluid properties
cannot realistically be considered constant. For instance, it has been shown that the
temperature (see, e.g., [45]) or, for liquid metals, k has been found to vary roughly
linearly with the temperature in a wide range of operating conditions (see, e.g., [46]).
50
The energy equation in these situations is no longer linear and, therefore, becomes
more challenging to solve, which is the cost of adding realism to the mathematical
model of the underlying transport phenomenon. Perhaps for this reason, the majority
of textbook examples and classical problems in convective heat transfer from objects
are solved under the assumption that the fluid properties are constant. Hence, it is
of both academic and practical interest to revisit those problems with the goal of
extending their solutions to cases with variable fluid properties. To this end, in this
chapter, we examine the steady-state transfer of heat from a particle via an externally
driven laminar flow, with the premise that the thermal conductivity of the fluid is a
Building on previous theoretical efforts on the subject (see, e.g., [21, 44, 47–58]),
geometry. The derivations are based on asymptotic and scaling analyses. For com-
pleteness, we consider both constant temperature and uniform heat flux boundary
conditions on the surface of the particle. The results are presented for the full range
In the rest of this chapter, we first describe the problem we wish to solve. Then, we
present the solutions for the above-mentioned surface conditions. The validity ranges
of the solutions are discussed next and a brief summary is given in the end.
51
4.2 Problem statement
object of arbitrary geometry and characteristic length `. Suppose that the free-stream
?
unit vector, and that the temperature approaches a constant value, denoted T∞ , at
large distances from the particle. Also, let the thermal conductivity of the fluid vary
dissipation, the equation that governs the steady-state distribution of T outside the
particle is
Pe u · ∇T = ∇ · [(1 + β T ) ∇T ] , (4.1)
where the Péclet number is defined as Pe = ρU∞ cp `/k0 . The boundary conditions
for the case in which the surface of the particle Sp is held at a constant temperature
Ts , and
for the one where a uniform heat flux qs is applied on Sp . As before, r is the position
vector with magnitude r = |r| and n is the unit vector outward normal to the surface
52
of the particle. Furthermore, the length and fluid velocity are non-dimensionalized,
T ? − T∞
?
T ? − T∞?
or , (4.4)
Ts? − T∞
? qs `/k0
consistent with the boundary conditions in Eqs. (4.2) and (4.3). The star superscript
is used to denote the dimensional temperature. Also, the average Nusselt numberis
defined as
Z
1+β
NuT = − n · ∇T dS (4.5)
2π Sp
for the problem with a prescribed temperature on the surface of the particle (see Eq.
(4.2)) and as
Sp
NuQ = (4.6)
2πT
for the one with a prescribed heat flux on Sp (see Eq. (4.3)), where Sp represents the
dimensionless surface area of the particle and T is the mean value of T on Sp . In the
following sections, we seek to develop approximate formulas for the variations of NuT
We begin the derivation of NuT by first considering the asymptotic limits of small and
large Péclet numbers and then bridging the gap between the two limits by introducing
53
a smooth transition function. Our approach builds on ideas presented in the classical
works on the topic of heat and mass transfer from an isolated particle in uniform
Suppose conduction is the dominant mode of heat transport, i.e., Pe 1, but finite.
In this limit, an effective approach for dealing with the nonlinearity of Eq. (4.1) is to
Z T Z T
1 β 2
θ= k(T) dT = (1 + βT) dT = T + T. (4.7)
k0 0 0 2
This is known as the Kirchhoff transformation (see, e.g., [21, 52]), with the reference
Z
1
NuT = − n · ∇θ dS, (4.8)
2π Sp
where
∇θ
Pe u · √ = ∇2 θ, with
1 + 2β θ
Our goal is to obtain an asymptotic expression for NuT , and, to that end, we
54
θ = θ(0) + Pe θ(1) + o(Pe), (4.10)
which is known as the inner expansion of θ. Upon substitution of Eq. (4.10) into Eqs.
NuT = Nu(0)
T
+ Pe Nu(1)
T
+ o(Pe)
"Z Z #
1
=− n · ∇θ(0) dS + Pe n · ∇θ(1) dS + o(Pe), (4.11)
2π Sp Sp
Far from the particle, on the other hand, we consider θ to take the form of
where the tilde overbar denotes that the transformed temperature field is written as
Rewriting Eq. (4.9) in terms of r̃ and replacing Eq. (4.13) for θ̃, we find
˜ θ̃(1) = ∇
e·∇ ˜ 2 θ̃(1) with lim θ̃(1) = 0, (4.15)
r̃→∞
˜ and ∇
with ∇ ˜ 2 operators representing derivatives with respect to the stretched co-
ordinates. The inner and outer expansions are required to match asymptotically, i.e.,
55
Enforcing the above equation at every order of Pe, furnishes the missing boundary
conditions of Eqs. (4.12) and (4.15). Following Brenner [4], the zeroth-order inner
solution away from the particle and the first-order outer solution can be written,
respectively, as
Nu(0) Nu(0)
θ(0) = T
+ O(r−3 ) = Pe T
+ O(Pe−3 ), (4.17)
2r 2r̃
(0)
NuT 1
θ̃(1) = exp − (r̃ − e · r̃) , (4.18)
2r̃ 2
where r is measured from a proper origin located at the particle’s “heat center”.
θ(0) ∇θ
Z Z Z
(0) (0)
θ n · ∇θ dS = θ n · ∇θ dS + Pe u· √ dV, (4.19)
Sp Sp V 1 + 2β θ
where, the contributions from surfaces at infinity (denoted by S∞ ) vanish because the
integrands decay faster than dS grows at large distances from the particle. Equation
θ(0) ∇θ
Z
(0) Pe
NuT = NuT + u· √ dV. (4.20)
2 π (1 + β/2) V 1 + 2β θ
1
Nu(1) =
T
2 π (1 + β/2)
"Z #
θ(0) ∇θ(0)
Z
× u· p dV + e · ˜ θ̃(1) − θ̃(0) dr̃ ,
θ̃(0) ∇ (4.21)
V 1 + 2β θ (0) R3
56
where the second integral on the right-hand side is over the entire three-dimensional
first on the right-hand side of Eq. (4.21). To that effect, we convert this volume
θ(0) ∇θ(0)
Z
u· p dV
V 1 + 2 β θ(0)
Z
1 hp i
(0) βθ (0) − 1 u dV
= ∇ · 1 + 2 β θ
3 β2 V
"Z
1 p
1 + 2 β θ(0) β θ(0) − 1 n · u dS
=− 2
3β Sp
Z p
(0)
+ 1 + 2β θ (0) β θ − 1 n · u dS . (4.22)
S∞
Z
n · u dS = 0 (4.23)
S∞
because the flow is incompressible (i.e, ∇ · u = 0). Therefore, the values of both
surface integrals in Eq. (4.22) amount to zero, which means that the volume integral
Z
e· ˜ θ̃(0) dr̃
θ̃(0) ∇
R3
2
Nu(0)
T
Z ∞
1
Z π Z 2π
2
=− sin θ cos ϕ dϕ dθ dr̃ = 0. (4.24)
4 0 r̃ 0 0
57
2
Nu(0)
T
Z
e · r̃
1
(1)
NuT = exp − (r̃ − e · r̃) dr̃
8 π (1 + β/2) R3 r̃4 2
2
(0)
NuT
= . (4.25)
4 (1 + β/2)
Nusselt number and the latter is the constant conductivity pure conduction Nusselt
number. Remember that Eq. (4.12) describes a linear boundary value problem and,
Equation (4.26) indicates that, to the leading order in Pe, NuT is the product of a
Pe and the geometry of the particle. It is worth emphasizing that this relation was
derived with no restriction on the flow Reynolds number, defined as Re = ρU∞ `/µ.
58
Even no-slip condition was not necessary. The only conditions enforced were the flow
incompressibility and no flow penetration into the particle. Lastly, we note that Eq.
(4.26) recovers the classical result of Brenner [4] for β = 0 and generalizes the work
of Polyanin [21] for the special case of Stokes flow (i.e., Re = 0).
other words, the temperature variations are confined to a narrow layer next to the
boundary of the particle. Here, similar to §4.3.1, we aim for developing an asymptotic
boundary layer coordinate system (x, y, z), with the corresponding unit vectors ex ,
The direction of the x coordinate is chosen to be the same as the component of the
fluid velocity vector projected onto the plane normal to ey . The direction of the third
Equation (4.1), written in terms of the (x, y, z) coordinates takes the form of
ux ∂T uy ∂T 1 ∂ hy hz ∂ β 2
Pe + = T+ T
hx ∂x hy ∂y hx hy hz ∂x hx ∂x 2
59
∂ hx hz ∂ β 2 ∂ hx hy ∂ β 2
+ T+ T + T+ T
∂y hy ∂y 2 ∂z hz ∂z 2
1 ∂Ψ 1 ∂Ψ
ux = and uy = − (4.28)
hy hz ∂y hx hz ∂x
are the velocity components in the x and y directions, respectively, where Ψ is a pseudo
stream function. Note that uz is zero per the definition of the generalized boundary
thin region close to Sp , it is convenient to expand the metric coefficients and Ψ about
y = 0 as
hy = 1, (4.30)
1
Ψ = Ψ2 (x, z) y2 + · · · , (4.32)
2
where
∂ux
Ψ2 = hz = hz,0 τ0 , (4.33)
∂y y=0
with τ0 being the dimensionless shear stress at the surface of the particle. To ensure
the validity of Eq. (4.32), the Prandtl number Pr = cp µ/k0 is assumed to be large.
60
Needless to say, both Ψ and ∂Ψ/∂y are zero at y = 0, due to the no-slip condition.
The thickness of the temperature boundary layer is known to scale with Pe−1/3 ,
which results from equating the order of magnitude of the advective and conductive
terms within the boundary layer (see, e.g., [6, 34]). Considering this scaling and the
Y = Pe1/3 y, (4.36)
Z x
X= hx,0 hz,0 dx̂. (4.37)
x0
d2 (0)
(0) β (0) 2 2 dT
T + T + 3 η = 0 with
dη 2 2 dη
61
where the new independent variable η is defined, according to the von Mises trans-
formation, via
p Z X p −1/3
η = Ψ2 9 Ψ2 dX̂ Y. (4.40)
X0
Taking everything into account, the leading-order term in the Nusselt number expan-
∂T (0)
Z
(0) 1+β
NuT = − dS
2π Sp ∂ỹ ỹ=0
dT (0)
4
= −Γ (1 + β) Nu(0) , (4.41)
3 dη η=0
T,0
where Nu(0)
T,0
represents the value of Nu(0)
T
corresponding to β = 0. Of course, Eq.
(4.41) reproduces the classical results of Lighthill [1] and Acrivos [40].
dT (0)
4
c = −Γ (1 + β) , (4.42)
3 dη η=0
where T (0) satisfies Eq. (4.39). We first consider the limit of β 1 and proceed with
(0) (0)
T (0) = T0 + β T1 + O(β 2 ). (4.43)
62
Substituting Eq. (4.43) in Eq. (4.39), we find that
(0) (0)
d2 T0 dT
2
+ 3 η 2 0 = 0 with
dη dη
(0) (0)
T0 = 1 at η = 0 and lim T0 =0 (4.44)
η→∞
and
(0) (0)
d2 T1 2 dT1 1 d2 (0) 2
+ 3 η + T = 0 with
dη 2 dη 2 dη 2 0
(0) (0)
T1 = 0 at η = 0 and lim T1 = 0. (4.45)
η→∞
Z η
(0) 1
exp −η̂ 3 dη̂
T0 =1− (4.46)
Γ(4/3) 0
which is derived using the integrating factor method. Applying the same approach to
where the first term on the right-hand side is numerically calculated (by enforcing
63
(0)
dT1
= 0.667. (4.48)
dη
η=0
Given Eqs. (4.43), (4.46), and (4.48), for small β, we can write
(0)
4 dT1 β + O(β 2 )
c = 1 + 1 − Γ
3 dη
η=0
Eq. (4.39), that dT (0) /dη scales with β −1/3 in this limit. Informed by this scaling, we
expand T (0) as
(0)
T (0) = T0 + O(β −1/3 ) (4.50)
and introduce
η̃ = β −1/3 η. (4.51)
(0)
d2 (0) 2
2 dT0
T0 + 6 η̃ = 0 with
dη̃ 2 dη̃
(0) (0)
T0 = 1 at η̃ = 0 and lim T1 = 0. (4.52)
η̃→∞
We solve the above nonlinear ordinary differential equation numerically and determine
that
(0)
4 dT0
c = −Γ β 2/3 + O(β 1/3 )
3 dη̃
η̃=0
64
= 0.710 β 2/3 + O(β 1/3 ). (4.53)
What is truly surprising, based on the results of Eqs. (4.49) and (4.53), is that the
prefactor c is very well approximated by a single formula (for the entire range of β)
as
2/3
3β
c≈ 1+ . (4.54)
5
Note that the above expression captures both low- and high-β asymptotes remarkably
2/3
3β
NuT ≈ 1 + NuT,0 + O(1), (4.55)
5
4.3.3 Bridging results for limits of low and high Péclet numbers
where 1/2 / a / 3/5 and 2/3 / b / 1. The functional dependency of these two
65
√
(2/3) Pe + b0
b≈ √ , (4.58)
Pe + b0
where
The above relations are motivated by the full numerical solution of Eq. (4.1), subject
to the boundary conditions in Eq. (4.2), for several basic particle shapes. The accuracy
of the predictions made by Eqs. (4.56)-(4.59) and the details of the numerical solutions
will be discussed in §4.5. Note that, for some simple geometries, approximate formulas
Overall, our asymptotic analyses in this section (summarized by Eq. (4.56)) sug-
gest that the ratio between the Nusselt number and its corresponding value for the
emphasize that this result was derived for particles of arbitrary shape under a fairly
only demanded the Prandtl number to be large when the Péclet number is high.
In §4.3, we considered Eq. (4.1) assuming that the surface of the particle is held at a
constant temperature (see Eq. (4.2)). There, the assumption of a Dirichlet boundary
66
condition on Sp allowed us to effectively employ the Kirchhoff and von Mises trans-
formula for NuT . Unfortunately, neither techniques can be directly applied to find the
general form of the Nusselt number when a uniform heat flux is imposed on Sp . That
a Neumann boundary condition is more challenging to deal with analytically than its
Dirichlet counterpart is a known matter (see, e.g., §2 and §3). Despite this difficulty,
we derive an estimate for NuQ with the aid of the following scaling arguments.
and
Applying the Kirchhoff transformation (see Eq. (4.7)), it can be shown that
67
p
(0) 1 + 2 β θ(0) − 1
T = , (4.63)
β
where
∇2 θ(0) = 0 with
Additionally, consistent with Eq. (4.60), the Nusselt number can be expressed as
(1)
!
Sp T
NuQ = (0)
1 − Pe (0)
+ ···
2πT T
= Nu(0)
Q
+ Pe Nu(1)
Q
+ ··· , (4.65)
where the overbar indicates average over Sp (see also Eq. (4.71)). The zeroth-order
Nusselt number is determined by the solution of Eq. (4.64), with its dependency on
Nu(1)
Q
= Nu(1)
Q,0
+ β Nu(1)
Q,1
+ ··· . (4.66)
Since both Pe and β are small, then, we need to only retain Nu(1)
Q,0
in the above
1 1
T (0) ∼ √ and T (1) ∼ . (4.67)
β β
68
Hence, we conclude that, again, to the leading order, Nu(1)
Q
is independent of β. Given
for small Péclet numbers as a summation of two terms: one that only depends on β
Consider the limit of Pe 1, where the thermal boundary layer shrinks at a rate
proportional to Pe−1/3 . In this case, the temperature and Nusselt number can be
where, upon substituting for T in Eqs. (4.1) and (4.3), we have (in generalized bound-
Equation (4.70) does not depend on β and, in fact, is identical to the one for β = 0.
(0)
Recognizing that T is not a function of β, we deduce from Eq. (4.69) that NuQ for
large Péclet numbers is approximately equal to the sum of a Pe-dependent term and
69
4.4.3 Bridging results for limits of low and high Péclet numbers
The main takeaway point of our scaling analyses in §4.4.1 and §4.4.2 is that NuQ may
NuQ ≈ NuQ,0 + NucQ − NucQ,0 , (4.71)
where NuQ,0 represents the value of NuQ for β = 0 (i.e., for constant conductivity) and
Nusselt number). The above approximation possesses the required form and has zero
error for the cases of Pe = 0 and β = 0. We will test the validity of Eq. (4.71) in §4.5.
Before we conclude this section, it is worth emphasizing that, for certain geome-
tries, fairly accurate estimates of NuQ,0 exist in the literature (see, e.g., §2). Further-
Sp β p −1
c (0)
NuQ = 1 + 2β θ − 1
2π
q −1 s !−1
Sp β Sp β Sp β
≈ 1 + 2 β θ(0) − 1 = 1+ −1 . (4.72)
2π 2π π NucQ,0
Equation (4.72) is exact for a spherical particle and found to be unexpectedly precise
for other particle shapes such as ellipsoids, cylinders, cones, and cubes.
70
β = 0.1 β=1 β = 10
Re = 0.1 Re = 1 Re = 10
(a) (b) (c)
20 20 20
∆ (%)
∆ (%)
∆ (%)
10 10 10
0 0 0
10−3 10−1 101 103 105 10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe Pe
(d) (e) (f )
20 20 20
∆ (%)
∆ (%)
∆ (%)
10 10 10
0 0 0
10−3 10−1 101 103 105 10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe Pe
(g) (h) (i)
20 20 20
∆ (%)
∆ (%)
∆ (%)
10 10 10
0 0 0
10−3 10−1 101 103 105 10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe Pe
(j) (k) (l)
20 20 20
∆ (%)
∆ (%)
∆ (%)
10 10 10
0 0 0
10−3 10−1 101 103 105 10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe Pe
(m) (n) (o)
20 20 20
∆ (%)
∆ (%)
∆ (%)
10 10 10
0 0 0
10−3 10−1 101 103 105 10−3 10−1 101 103 105 10−3 10−1 101 103 105
Pe Pe Pe
Figure 4.1: Percent difference ∆ between the results of full numerical sim-
ulations and the predictions of Eq. (4.56) for NuT .
71
β = 0.1 β=1 β = 10
Re = 0.1 Re = 1 Re = 10
(a) (b) (c)
0 0 0
∆ (%)
∆ (%)
∆ (%)
-10 -10 -10
0 0 0
∆ (%)
∆ (%)
∆ (%)
-10 -10 -10
0 0 0
∆ (%)
∆ (%)
∆ (%)
-10 -10 -10
0 0 0
∆ (%)
∆ (%)
∆ (%)
0 0 0
∆ (%)
∆ (%)
∆ (%)
Figure 4.2: Percent difference ∆ between the results of full numerical sim-
ulations and the predictions of Eq. (4.71) for NuQ .
72
4.5 Comparison with direct numerical solutions
In §4.3 and §4.4, we presented approximate formulations for NuT and NuQ via per-
turbation analyses in the asymptotic limits of Pe and β. To give an idea about the
estimation error of the proposed formulas, here, we compare their predictions for
spherical, cubic, and ellipsoidal particles against the results obtained from the full
implemented in COMSOL Multiphysics [60, 61], is employed to carry out the compu-
tations. We first solve the steady-state incompressible Navier-Stokes equations for the
fluid flow, and, then, use the calculated velocity field u to compute the solution of Eq.
(4.1) for the temperature distribution T . The flow and advection-diffusion equations
are solved iteratively, while discretized by P2+P1 and quadratic Lagrange schemes,
respectively. The outer boundary at infinity is modeled as a large sphere, whose center
coincides with the center of the particle. Specifically, the diameter of the sphere is
set to 200 times the characteristic length of the particle. We use tetrahedral elements
to mesh the computational domain such that the grid density is the highest near the
particle. Grid-independence studies are performed to ensure that the results change
Figures 4.1 and 4.2 show the outcome of our calculations for NuT and NuQ , respec-
tively. The results are presented in the form of ∆ versus Pe plots for β = 0.1, 1, 10
and Re = 0.1, 1, 10, where the parameter ∆ is defined as the percent difference be-
tween the directly computed and predicted values of the Nusselt number. Each figure
73
consists of fifteen sub-figures that are organized into five rows and three columns.
The sub-figures in the left, middle, and right columns correspond to Re = 0.1, 1,
and 10, respectively. Also, those in the first and second rows belong to spherical and
cubic particles, whereas the rest are for an ellipsoidal particle of semi-axis lengths a,
b, and c, where b/a = 2/3 and c/a = 1/3. Within the bottom three rows, the plots in
the first, second, and third rows are for flows along the principal axes of the ellipsoid
associated with a, b, and c, respectively. Note that, in calculating the Nusselt and
Péclet numbers the characteristic length ` is set to the radius for the sphere, to half
the side length for the cube, and to the largest semi-axis for the ellipsoid.
Overall, we see that the predictions of Eqs. (4.56) and (4.71) are quite accurate,
with the absolute value of ∆ being less than 16.5% for all the cases considered. The
approximations are more precise for particles with more streamlined shapes. Also, as
expected, the estimations deviate the most from the numerical results when β and Re
are large and Pe is in the intermediate range. Perhaps surprisingly, however, ∆ is very
small for β . O(1), irrespective of its corresponding Reynolds and Péclet numbers.
Another observation that can be made is that ∆ is mostly positive in the plots of
Fig. 4.1 whereas it is mainly negative in those of Fig. 4.2. Note that ∆ is defined such
that it is positive when the predicted Nusselt number overestimates the numerically
calculated one. Lastly, for the same shape, Re, and β, the Péclet number at which
the approximation error is maximum is generally higher for NuQ than it is for NuT .
74
4.6 Summary
We studied the problem of forced convection heat transfer from a particle of ar-
asymptotic as well as scaling analyses to develop approximate relations for the varia-
tions of the Nusselt number with the Péclet number and the slope of the (normalized)
form heat flux boundary conditions on the surface of the particle, and discovered
that, for the former, NuT can be estimated as a product of a Pe-dependent term and
one that primarily changes with β. We also found that, for the latter, NuQ may be
our derivations offer a straightforward way to estimate the Nusselt number for any
β by just knowing the the Nusselt number corresponding to β = 0, i.e., the constant
predictions for NuT and NuQ with those calculated based on direct numerical solutions
of the governing equations. The comparisons confirmed that the proposed approxima-
tions are valid over a wide range of parameters. More specifically, they demonstrated
75
Chapter 5: Summary and future directions
fer from isolated objects, namely (i) conduction heat transfer from oblate spheroids
and bispheres, (ii) forced convection heat transfer from an arbitrarily-shaped parti-
cle, and (iii) heat transfer from a single object in laminar flows of a variable thermal
conductivity fluid. We had noticed that although there have been many studies on
mathematical modeling of heat transfer from objects of various shapes (or analogous
have focused on the isothermal (Dirichlet) boundary condition and a relatively small
number have dealt with the uniform flux (Neumann) boundary condition. In an at-
tempt to partially fill this gap in the literature, in this dissertation, we particularly
considered this former condition in our analyses, and demonstrated that mathemat-
ical derivation under this choice becomes often more challenging compared to the
sion were used to solve the Laplace equation in curvilinear coordinate systems for
76
the conduction problems. The convection problems, on the other hand, were treated
via a combination of the perturbation and boundary layer theories, reciprocal theo-
rem, and scaling arguments. In all cases, we developed precise, but straightforward,
approximations for the Nusselt number that can be readily used by engineers to de-
sign new thermal systems or better understand the function of the existing ones.
The accuracy of the calculated estimations were confirmed through comparison with
worth emphasizing that the formulas derived for the Nusselt number in this work
are all equally applicable for approximating the Sherwood number in equivalent mass
transfer problems.
nish a sound basis for future theoretical investigations on heat transfer from multiple
particles, which can be considered as a natural extension of our study. We can envi-
sion using the method of reflections in addition to the approaches discussed here to
Furthermore, our work in §4 can be extended to cases where the thermal conductivity
is a non-linear function of the temperature and where other fluid properties such as
77
Bibliography
boundary layer,” Proc. Roy. Soc. (London), vol. 202, no. 1070, pp. 359–377, 1950.
[2] A. Acrivos, “Solution of the laminar boundary layer energy equation at high
[3] A. Acrivos and T. D. Taylor, “Heat and mass transfer from single spheres in
[4] H. Brenner, “Forced convection heat and mass transfer at small Peclet numbers
from a particle of arbitrary shape,” Chem. Eng. Sci., vol. 18, no. 2, pp. 109–122,
1963.
[5] T. D. Taylor, “Heat transfer from single spheres in a low reynolds number slip
convection heat and mass transfer,” J. Fluid Mech., vol. 23, no. 2, pp. 273–291,
1965.
78
[7] J. D. Goddard and A. Acrivos, “Asymptotic expansions for laminar forced-
convection heat and mass transfer Part 2. Boundary-layer flows,” J. Fluid Mech.,
[8] N. A. Frankel and A. Acrivos, “Heat and mass transfer from small spheres and
cylinders freely suspended in shear flow,” Phys. Fluids, vol. 11, no. 9, pp. 1913–
1918, 1968.
[9] R. C. Sehlin, “Forced-convection heat and mass transfer at large Péclet num-
bers from an axisymmetric body in laminar flow: Prolate and oblate spheroids,”
[10] F. Morrison Jr and S. Griffiths, “On the transient convective transport from a
lems of convective mass and heat transfer of reacting particles with the flow,”
Int. J. Heat Mass Transfer, vol. 27, no. 2, pp. 163–189, 1984.
M. Tain, “Simplified analytical models for forced convection heat transfer from
cuboids of arbitrary shape,” J. Electron. Packag., vol. 123, no. 3, pp. 182–188,
2001.
79
[14] Z.-G. Feng, “Forced heat and mass transfer from a slightly deformed sphere at
small but finite peclet numbers in stokes flow,” J. Heat Transfer, vol. 135, no. 8,
2013.
[15] ——, “Forced heat and mass transfer from a slightly deformed sphere at small
but finite peclet numbers in stokes flow,” J. Heat Trans., vol. 135, no. 8, p.
081702, 2013.
[16] R. Alassar and B. Alminshawy, “Heat conduction from two spheres,” AlChE J.,
[18] E. Hahne and U. Grigull, “Shape factor and shape resistance for steady multi-
dimensional heat conduction,” Int. J. Heat Mass Transf., vol. 18, pp. 751–767,
1975.
[20] H. Brenner and S. Haber, “Symbolic operator solutions of Laplace’s and Stokes’
equations: Part I Laplace’s equation,” Chem. Eng. Commun., vol. 27, no. 5-6,
80
[21] A. D. Polyanin, “An asymptotic analysis of some nonlinear boundary-value prob-
lems of convective mass and heat transfer of reacting particles with the flow,”
[22] F. Moukalled, L. Mangani, and M. Darwish, The Finite Volume Method in Com-
McGraw-Hill, 1953.
[24] P. Moon and D. E. Spencer, Field Theory Handbook: Including Coordinate Sys-
1988.
[26] R. S. Alassar, “Heat conduction from spheroids,” J. Heat Trans., vol. 121, no. 2,
[28] J. Happel and H. Brenner, Low Reynolds Number Hydrodynamics, with Special
81
Applications to Particulate Media. The Hague, The Netherlands: Martinus
Nijhoff, 1983.
[29] R. S. Alassar and B. J. Alminshawy, “Heat conduction from two spheres,” AIChE
[30] G. W. Morgan and W. H. Warner, “On heat transfer in laminar boundary layers
at high Prandtl number,” J. Aero. Sci., vol. 23, no. 10, pp. 937–948, 1956.
[32] A. Acrivos and T. D. Taylor, “Heat and mass transfer from single spheres in
[33] C. R. Robertson and A. Acrivos, “Low Reynolds number shear flow past a rotat-
ing circular cylinder. Part 2. Heat transfer,” J. Fluid Mech., vol. 40, no. 4, pp.
705–718, 1970.
[35] L. M. Relyea and A. S. Khair, “Forced convection heat and mass transfer from
a slender particle,” Chem. Eng. Sci., vol. 174, pp. 285–289, 2017.
82
[36] R. Shah and T. Li, “The thermal and laminar boundary layer flow over prolate
and oblate spheroids,” Int. J. Heat Mass Transf., vol. 121, pp. 607–619, 2018.
[37] V. Vandadi, S. Jafari Kang, and H. Masoud, “Reciprocal theorem for convective
heat and mass transfer from a particle in Stokes and potential flows,” Phys. Rev.
[38] H. Masoud and H. A. Stone, “The reciprocal theorem in fluid dynamics and
particle of arbitrary shape having a liquid flowing around it,” Fluid Dyn., vol. 10,
[40] A. Acrivos, “Solution of the laminar boundary layer energy equation at high
[42] L. A. Romero, “Low or high Péclet number flow past a prolate spheroid in a
saturated porous medium,” SIAM J. Appl. Math., vol. 55, no. 4, pp. 952–974,
1995.
[43] R. Clift, J. R. Grace, and M. E. Weber, Bubbles, Drops, and Particles. New
83
[44] A. D. Polyanin and V. V. Dil’man, “The method of asymptotic analogies in the
mass and heat transfer theory and chemical engineering science,” Int. J. Heat
thermal conductivity enhancement for nanofluids,” vol. 125, no. 4, pp. 567–574,
2003.
[46] W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer. New
variable thermal conductivity and capacity,” vol. 31, no. 1-2, pp. 25–31, 1978.
[48] ——, “Thermal boundary layer in liquid metals with variable thermal conduc-
[49] A. D. Polyanin, “Method for solution of some non-linear boundary value prob-
[50] T. C. Chiam, “Heat transfer in a fluid with variable thermal conductivity over a
linearly stretching sheet,” vol. 129, no. 1-2, pp. 63–72, 1998.
84
[52] S. Kakaç, Y. Yener, and C. P. Naveira-Cotta, Heat Conduction. Boca Raton,
[53] M. A. Ezzat, “State space approach to solids and fluids,” vol. 86, no. 11, pp.
1241–1250, 2008.
with variable thermal conductivity and fractional-order heat transfer,” vol. 36,
fluid with fractional order of heat transfer,” vol. 100, pp. 305–315, 2016.
[58] ——, “The effects of thermal and mechanical material properties on tumorous
[59] E. Dehdashti and H. Masoud, “Forced convection heat transfer from a particle
85
[60] W. B. J. Zimmerman, Multiphysics Modeling with Finite Element Methods. Sin-
[61] D. W. Pepper and J. C. Heinrich, The Finite Element Method: Basic Concepts
and Applications with MATLAB, MAPLE, and COMSOL. New York: CRC
Press, 2017.
86