INTERMATHS, VOL. 4, NO. 1 (2023), 25–47
https://doi.org/10.22481/intermaths.v4i1.12683
Article
cb licença creative commons
Simulation of the onset turbulent flow around a
Isothermal Complex Geometries: an analysis of thermofluid
dynamic flow
Rômulo D. C. Santos
a,
, Quétila G. Silva
a
, and Samia R. L. Tananta
a
a
Instituto Federal de Educação, Ciência e Tecnologia do Acre (IFAC) (IFAC), Rio Branco - AC, Brasil
* Correspondence: lrccarvalho@gmail.com
Abstract: In this work, in the area of Computational Fluid Dynamics (CFD), more specifically in
the area of thermofluid dynamics for two-dimensional flows (2D), and also considering, the fluid-body
interaction, allied to the phenomena of heat-transfer by mixed convection and the beginning of processes
of the turbulent flow phenomenon in the fluid-body interaction, a study is proposed that demonstrates
the efficiency in the analysis and simulation of these complex phenomena. We adopt an Eulerian
approach for a fixed mesh, which is intended to represent the thermofluid dynamic movement, working
together with a Lagrangian mesh, the latter being intended to discretize the immersed body. The
strategy, in this work, allows approaching complex isothermal geometries, which present a certain
aerodynamic degree on their surface, being popularly known as blunt body, where this, in turn, is
immersed in an incompressible Newtonian fluid. One of the contributions of this work is the introduction
of a simple but efficient method to calculate the Nusselt number. In relation to validating and modeling
the physical phenomena of interest, specifically the effectiveness of the Immersed Boundary Method, we
conducted an implementation with low computational cost. This implementation was used to transfer
mixed convection heat and model turbulence using the Spalart-Allmaras model within the framework
of the URANS (Unsteady Reynolds Average Navier-Stokes) methodology. Numerical results showed
good convergence with data available in the literature, which confirms the numerical precision and
reliability of the adopted model.
Keywords: Immersed Boundary Method; Mixed Convection; Onset Turbulence; Isothermal Bluff
Body.
Classification MSC: 76D05; 80A19
1 Introduction
The simulation of flows around complex geometries presents a challenging task
for computational fluid dynamics. Traditional body-fitted numerical methods, which
strongly couple the solution of governing equations and implementation of boundary
conditions, usually need mesh generation, in general, with high computational cost. Even
considering simple geometries, the generating a high quality mesh, is not a trivial job.
To overcome the disadvantage of a strong coupling between the solution of governing
equations and the implementation of the boundary condition in the immersed boundary,
Peskin [
1
] presented the idea of Immersed Boundary Method (IBM) in 1971 to simulate
blood flows in mitral valves of the human heart. The IBM is a numerical method that
Submitted 18 May 2023; Accepted 25 June 2023; Available online: 30 June 2023.
ISSN 2675-8318 Copyright ©2023 INTERMATHS. Published by Edições UESB. This is an Open Access article under the CC BY 4.0 license.
uses Cartesian Eulerian grid points to discretize the solution of governing equations
and Lagrangian points to represent the boundary of immersed body. Peskin’s idea
was to model the contour of the heart valve as a deformable elastic fiber with high
stiffness. A small deformation or movement of the boundary would produce a force
that tends to restore the boundary back to its original shape or position. The restoring
force at the Lagrangean point is then distributed in the surrounding fluid (Eulerian
grid point) as a body force, and the Navier-Stokes equations with body strength are
resolved throughout the domain. The IBM was reviewed in [
5
], presenting there an
extensive list of bibliographic references. The easy computational implementation of
IBM has attracted many reseachers. Various modifications and refinements have been
made. Among them, Lai and Peskin [
6
] presented a second-order accurate IBM and
applied it to simulate the flow past a circular cylinder. The interaction between the fluid
and immersed boundary was modeled by a suitable representative of the delta-function.
The work of Goldstein et al. [
7
] presents a method that uses control of the velocity on
the solid boundary to enforce the boundary conditions, and demonstrates its suitability
for a flow past a rigid cylinder. In the process of obtaining the forcing momentum
term iteratively, two negative semi-empirically constants are required in the numerical
scheme, creating severe stability problems in the computation. The forcing momentum
and the parameters are used into nearby Eulerian grids [
1
],[
2
]–[
7
]. Instead of adopting
user-specified parameters, the forcing term imposing a specific boundary condition on
the body was numerically extracted, so this scheme can be classified as a discrete forcing
approach.
The work of Yang [
8
] proposes the use of a methodology called Virtual Boundary
Method (VBM): a bilinear interpolation procedure is combined to realize the data
transmission between grid and boundary points. The selection of constant values in the
forcing term is discussed based on the error of the velocity at the boundary. Then, the
VBM method is used to simulate the flow around bluff bodies. The VBM feasibility is
verified by the good agreement between the simulation’s results and other reference’s
outcomes. Mohd-Yusof [
9
] introduced a momentum forcing term that does not affect
the stability of the discrete-time equation. The need to use a small computational
time step, which is regarded as an important advantage of this method over other
previous methods, is therefore avoided. Ye et al. [
10
] proposed the so-called Cartesian
grid method in non-staggered grids to simulate the unsteady and incompressible Navier-
Stokes equations in the physical domain with complex immersed boundary. In the
whole computational domain, a finite-volume method of second-order accuracy is applied
together with the two-step fractional-step procedure. Near the immersed boundary, an
interpolation procedure of second-order spatial accuracy is used.
The objectives of this study were to propose an immersed boundary method that
accounts for both momentum and heat transfer during mixed convection, taking into
consideration the onset of turbulence between a heated immersed body (referred to as
an isothermal immersed body) and the surrounding fluid. Furthermore, the study aimed
to analyze the effects resulting from the combination of these physical phenomena.. As
far as we know, heat-transfer by mixed convection combined with the phenomena of
turbulence around complex geometries is treated in a segregated way in many works,
some of them considering only forced convection for low Reynolds numbers. Santos
26 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
et al. [
11
] present the IBM for low and high Reynolds numbers coupled with the
Virtual Physical Model (VPM) proposed by Silva et al. [
12
] to calculate the Lagrangian
force field, based only on the momentum equation. To simulate incompressible two-
dimensional flows around heated square cylinders, a constant temperature was assumed
on the surface of the cylinders. A good numerical agreement was achieved, with a
margin of error of less than 3% compared to these previous works. The time evolution
of the drag and lift coefficients, as well as, the Nusselt number, were obtained with this
methodology, being the parameters obtained from the Eulerian fields. The geometry
used in this work has singularities, which were taken into account in the construction of
the algorithm/code. This fact holds significant importance as it enables its applicability
to various geometries. Across all simulations, the results consistently indicate that the
influence of the immersed heated body’s surface on the flow intensifies with higher
Reynolds numbers. A turbulence model was also used for transfer process between the
largest and smallest turbulent scales.
For this work, we will consider Newtonian and incompressible fluids with constant
properties. The buoyancy term based on the Boussinesq approximation is presented,
considering the case of mixed convection. The momentum and heat transformations
between the Lagrangian and Eulerian coordinates were realized by using a smooth Dirac
delta-function, so that the nearby grid points are locally subject to momentum and
heat transfer. The stability regimes were validated for two empirical constants. The
governing equations were independently solved in each grid system coupled. The energy
generation term is disregarded/ignored in this context, because neither the effects of
internal heat (for example, absorption or emission radiation) nor the humidity (which
could be responsible for the latent heat exchange) are considered. Coriolis force or
rotation effects are also absent in this paper. The proposed methods were validated
for natural and forced (mixed) convection, with the isothermal and constant heat flux
boundary conditions. Finally, the heat-transfer and the onset of turbulence around
complex geometry with the surrounding fluid was examined. The numerical results
obtained are both accurate and stable, demonstrating a strong agreement with the
existing findings in the literature.
2 Mathematical methodology
2.1 Mathematical formulation for the fluid motion and temperature
Consider a viscous incompressible flow in a two-dimensional domain containing an
immersed boundary in the form of closed curve Γ, as shown in Fig. (1). Like in previous
works, the forcing term is also introduced in the momentum equations and the heat
source term is incorporated in the energy equation. As a result, the governing equations
describing the forced/natural convection can be written in the primitive variables form
as
ρ
0
u
t
+ (u ) u
!
= −∇p + µ
2
u + f , (1)
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 27
ρ
0
u
t
+ (u ) u
!
= −∇p + µ
2
u + ρ
0
g (1 β (T T
)) j + f , (2)
ρ
0
c
p
T
t
+ (u ) T
!
= k
2
T + q , (3)
u = 0 , (4)
Fig. 1. A two-dimensional domain containing an ishothermal bluff body (isothermal square
cylinder) Γ with non-uniform grid.
where Eqs. (1), (3) and (4) are for forced convection, while Eqs. (2), (3) and (4) are for
natural convection (in Eq. (2), the Boussinesq approximation is used). The symbols (i)
u,
p
,
T
and
T
denote velocity vector, pressure, temperature and reference temperature,
respectively, (ii)
ρ
0
,
µ
,
β
,
k
and
c
p
are fluid density at temperature
T
=
T
, viscosity,
thermal diffusivity, thermal expansion coeficient and specific heat at constant pressure,
respectively, (iii) g is a downward gravitational acceleration, (iv) the term
ρ
0
g
β (T T
)
accounts for the effects of the fluid temperature on the fluid flow, and (v) j is the unit
vector in the positive y-axis direction, respectively.
The governing equations are non-dimensionalized by using characteristic scales such
as the fluid density
ρ
0
for the characteristic density, the far-field velocity
U
for
the characteristic velocity, and the body diameter
D
for the characteristic length.
The following characteristic scales are introduced:
D/U
for the time,
ρ
0
U
2
for the
pressure, and
ρ
0
U
2
/D
for the Eulerian momentum force f or simply forcing term. The
dimenionless incompressible Navier-Stokes equations and energy equation are
28 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
u
t
+ (u · ) u = −∇p +
1
Re
2
u +
Gr
Re
2
gT + f , (5)
T
t
+ u · T =
1
Re · P r
2
T + q , (6)
where
Re
is the Reynolds number, defined mathematically by
Re
=
ρ
0
U
D
, and
Gr
is the Grashof number, defined here by
Gr
= g
β (T T
) D
3
2
.
P r
denotes the
Prandtl number, defined by
P r
=
ν
, where
α
and
ν
are the thermal diffusivity and
the kinematic viscosity, respectively. The non-dimensional continuity equation takes the
same form as the dimensional continuity equation.
The forcing term f in the momentum Eqs. (1) and (2) is the force density at the fluid
(Eulerian grid) point distributed from the force density at the boundary (Lagrangian)
point, which can be expressed as
f (x, t) =
Z
Γ
F (X (s) , t) δ (x X (s, t)) ds , (7)
where x is the Eulerian coordinate. X and F stand for Lagrangian coordinate and
the boundary force density respectively.
δ (x X (s, t))
is the Dirac delta function
representing the interaction between fluid and the immersed boundary. Similarly, the
heat source/sink term
q
in the energy, Eq. (3), is the heat density transferred to the
fluid from the heat flux at the immersed boundary, which can be written as
q (x, t) =
Z
Γ
Q (X (s) , t) δ (x X (s, t)) ds . (8)
Here, Q (X (s) , t) is the virtual boundary heat flux.
The boundary condition-enforced immersed boundary method developed in this paper
involves both velocity correction and temperature correction to satisfy the boundary
conditions. In the following, the velocity correction procedure will be describe first,
followed by the temperature correction procedure.
2.2 Velocity correction procedure
With the help of fractional step technique, the velocity field can be obtained in the
following predictor-corrector steps:
(1) Predictor step:
We start with the usual Navier-Stokes equations without the forcing term,
ρ
0
u
t
+ (u ) u
!
= −∇p + µ
2
u , (9)
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 29
ρ
0
u
t
+ (u ) u
!
= −∇p + µ
2
u + ρ
0
g (1 β (T T
)) j , (10)
By solving the above Eq. (9) or (10) together with Eq. (4), using the projection
method developed by Chorin [
13
], we obtain the predicted velocity u
. The basic
solution procedure is described as follows.
To advance the solution from time level
n
to
n
+ 1, the Crank-Nilson scheme is applied
for temporal discretization, and the immediate velocity
˜
u can be given by
ρ
0
˜u u
n
δt
=
ρ
0
2
[(
˜
u · )
˜
u + (u
n
· ) u
n
] +
µ
2
2
˜
u +
2
u
n
, (11)
ρ
0
˜u u
n
δt
=
ρ
0
2
[(
˜
u · )
˜
u + (u
n
· ) u
n
] +
µ
2
2
˜
u +
2
u
n
+
ρ
0
Gr
Re
2
(1 β (T
n
T
)) j . (12)
Then, the immediate velocity
˜
u
is corrected to the predicted velocity u
by the
following form
ρ
0
u
˜
u
δt
= −∇p
n+1
. (13)
The Richardson’s number (
Ri
=
Gr/Re
2
), present in Eq. (12), expresses the relationship
between the buoyancy of a fluid in relation to its viscosity. In this work, the Richardson’s
number is based on the distance between two isothermal plates,
Gr
= g
β (T T
) D
3
2
,
which represent the walls of a computing domain, separated by a characteristic distance
D
. Before using Eq. (11), the pressure field should be obtained first by solving the
following Poisson equation
2
p
n+1
= ρ
0
·
˜
u
δt
. (14)
In the implementation process, a staggered grid is used for the finite volume discretiza-
tion of Eqs. (9) and (10), i.e. a scalar variable pressure-like is defined at the center
of each cell, while velocity components are defined at the cell faces. For the spatial
derivatives, the well-know central difference scheme is employed.
(2) Corrector step:
The corrected velocity field can be obtained by solving the following equation with
the force term f ,
ρ
u
t
= f . (15)
30 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
In the conventional IBM, the force density f is determined in advance and then the
corrected velocity u can be obtained from Eq. (15). However, there is no guarantee
that the velocity at the Lagrangian point interpolated from this corrected velocity u
always satisfies the non-slip boundary condition. This drawback can be overcome by
the present method. The basic idea is that the force density f is considered as unknown,
which is evaluated implicitly in such a way that the velocity u
(X (s) , t)
at the boundary
(Lagrangian) point interpolated from the corrected velocity u at the Eulerian grids equals
the given boundary velocity U
B
.
By applying forward Euler method to approximated
the temporal derivative in Eq. (15), it is clear that adding a forcing term f is equivalent
to making a correction δu in the velocity field. By defining
u = u
+ δu . (16)
The Eq. (15) gives
ρ
δu
δt
= f , (17)
where
δt
is the time step size, u is the corrected velocity, u
is the intermediate velocity,
δ
u
is the velocity correction. Suppose that the immersed boundary is represented by a set of
Lagrangian points X
i
B
(i = 1, 2, · · · , M)
, and the fluid field is covered by a fixed uniform
Cartesian mesh x
j
(j = 1, 2, · · · , N)
with mesh spacing
x
= ∆
y
=
h
. Furthermore, let
δ (x X (s, t)) be smoothly approximated by a continous kernel distribution
D
ij
= D
x
j
X
i
B
=
1
h
2
δ
x
j
X
i
B
h
!
δ
y
j
Y
i
B
h
!
. (18)
Here,
h
is the mesh size of Eulerian mesh, and
δ (r)
is proposed by Lai and Peskin [
13
]
δ (r) =
1
8
3 2 |r| +
q
1 + 4 |r| 4r
2
, |r| 1
1
8
5 2 |r| +
q
7 + 12 |r| 4r
2
, 1 < |r| 2
0 , |r| > 2 .
(19)
Note from Eq. (7) that the force density f at the Eulerian point is distributed from
the boundary force F through the Dirac delta function
(δ (x X (s, t)))
interpolation.
Eq. (7) can be written in the discrete form as
f (x
j
, t) =
X
i
F
X
i
B
, t
D
x
j
X
i
B
s
i
, (j = 1, 2, · · · , N; i = 1, 2, · · · , M) .
(20)
Substituting Eq. (20) into Eq. (17), we have
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 31
δu (x
j
, t) =
X
i
F (X
i
B
, t) δt
ρ
D
x
j
X
i
B
s
i
, (i = 1, 2, · · · , M) , (21)
where
δ
u
(x
j
, t)
is the velocity correction at the
j
th
Eulerian point,
s
i
is the
i
th
boundary
segmente length. It is noted that the known quantities in Eq. (21) are the boundary
forces F
i
B
(abbreviation for F
(X
i
B
, t)
). To satisfy the non-slip boundary condition, the
velocity at the boundary point interpolated from the corrected fluid velocity field via
the smooth delta function
D
ij
must be equal to the boundary velocity U
(X
i
B
, t)
at the
same position
U
X
i
B
, t
=
X
j
u (x
j
, t) D
x
j
X
i
B
h
2
, (j = 1, 2, · · · , N) . (22)
Here, u
(x
j
, t)
is the corrected velocity at the
j
th
Eulerian grid point, expressed as
the sum of predicted velocity u
(x
j
, t) and velocity correction δu (x
j
, t)
u (x
j
, t) = u
(x
j
, t) + δu (x
j
, t) . (23)
Substituting Eqs. (23) and (21) into Eq. (22), gives
U
X
i
B
, t
=
X
j
u
(x
j
, t) D
x
j
X
i
B
h
2
+
X
j
X
k
F (X
i
B
, t) δt
ρ
D
x
j
X
k
B
s
k
D
x
j
X
i
B
h
2
, (24)
for i = 1, 2, · · · , M; j = 1, · · · , N; k = 1, · · · , M .
The system of equations for the boundary force can be written in the following matrix
form
A X = B , (25)
where
A =
δt
ρ
h
2
D
11
s
1
D
12
s
1
· · · D
1N
s
1
D
21
s
2
D
22
s
2
· · · D
2N
s
2
.
.
.
.
.
.
.
.
.
.
.
.
D
M1
s
M
D
M2
s
M
· · · D
MN
s
M
D
11
D
12
· · · D
1M
D
21
D
22
· · · D
2M
.
.
.
.
.
.
.
.
.
.
.
.
D
N1
D
N2
· · · D
NM
(26)
32 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
B =
U
1
B
U
2
B
.
.
.
U
M
B
h
2
D
11
D
12
· · · D
1M
D
21
D
22
· · · D
2M
.
.
.
.
.
.
.
.
.
.
.
.
D
N1
D
N2
· · · D
NM
u
1
u
2
.
.
.
u
N
(27)
X =
F
1
B
F
2
B
.
.
.
F
M
B
(28)
and U
i
B
(i = 1, · · · , M)
, u
j
(j = 1, · · · , N)
are the abbreviations for U
(X
i
B
, t)
and
u
(x
j
, t)
, respectively. In this way, the boundary force F
i
B
(i = 1, · · · , M)
is evaluated
directly and implicity so that the satisfying of no-slip condiction is accurately achieved. It
can be observed that the entries of matrix A depend only on the coordinate information
of the Lagrangian boundary points and their adjacent Eulerian points. By solving the
equation system (26), using a direct method or iteractive method, the unknow boundary
force F
i
B
(i = 1, · · · , M)
at all Lagrangian boundary points are obtained simultaneously,
which are then substituted into Eqs. (21) and (23) to calculate the velocity correction
δu
j
and the corrected velocity u
j
(j = 1, · · · , N).
A remarkable advantage of the present method is that the force on the Lagrangian
points can be directly derived, since it is a part of the solution. We know that
F
i
B
(i = 1, · · · , M)
is the boundary force density acting on the fluid, and due to the fact
that action and reaction are equal and opposite, it is obvious that the force exerceted
by fluid on the boundary should be F
i
B
(i = 1, · · · , M) at each Lagrangian point.
2.3 Temperature correction procedure
Similarly, using the fractional step approach, the solution of Eq. (3) can be also
obtained by predictor-corrector steps. The predictor step solves the following energy
equation without the heat source term q ,
ρ
0
c
p
T
t
+ (u ) T
!
= k
2
T . (29)
The resultant solution from Eq. (29) is noted as predicted temperature
T
(x, t) ,
which can be obtained by solving the following equation obtained by the Crank-Nicolson
scheme
ρ
0
c
p
T
T
n
δt
=
ρ
0
c
p
2
h
u
n+1
·
T
+
u
n+1
·
T
n
i
+
k
2
2
T
+
2
T
n
. (30)
Similar to the velocity correction procedure, the corrected temperature
T (x, t)
is
obtained by solving the following equation,
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 33
ρ
0
c
p
T
t
= q . (31)
It can be clearly observed from Eq. (31) that adding a heat source/sink is equivalente
to making a temperature correction. Applying Euler explicit scheme to Eq. (31) gives
ρ
0
c
p
δT
δt
= q , (32)
with
δT (x, t) = T (x, t) T
(x, t) , (33)
where
δT (x, t)
is the temperature correction and
T (x, t)
is the corrected temperature.
Note from Eq. (8) that the heat source/sink,
q,
at the Eulerian grid point is distributed
from the boundary heat flux,
Q,
through Dirac delta function interpolation, which can
be expressed in the discret form as
q (x
j
, t) =
X
i
Q
X
i
B
, t
D
x
j
X
i
B
s
i
, (i = 1, · · · , M; j = 1, · · · , N) . (34)
Substituting Eq. (33) into Eq. (32) leads to
δT (x
j
, t) =
X
i
Q (X
i
B
, t) δt
ρc
p
D
x
j
X
i
B
s
i
, (i = 1, · · · , M; j = 1, · · · , N) .
(35)
It is noted that the unknows in Eq. (35) are the boundary heat fluxes
Q (X
i
B
, t)
. To
satisfy the physical boundary condiction, we have to make sure that the temperature
at the boundary point interpolated from the corrected temperature field by the delta
function D
ij
is equal to the specified temperature T
B
(X
i
B
, t), that is,
T
B
X
i
B
, t
=
X
j
T (x
j
, t) D
x
j
X
i
B
h
2
(i = 1, · · · , M; j = 1, · · · , N) . (36)
Substituting Eqs. (33) and (35) into Eq. (36) gives
T
B
X
i
B
, t
=
X
j
T
(x
j
, t) D
x
j
X
i
B
h
2
+
X
j
X
k
Q
X
k
B
, t
δt
ρc
p
D
x
j
X
k
B
s
k
D
x
j
X
i
B
h
2
, (37)
with
i
= 1
, · · · , M
;
j
= 1
, · · · , N .
The Eq. (37) can be put in the following matrix form
34 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
C Y = D , (38)
where
C =
δt
ρc
p
h
2
D
11
s
1
D
12
s
1
· · · D
1N
s
1
D
21
s
2
D
22
s
2
· · · D
2N
s
2
.
.
.
.
.
.
.
.
.
.
.
.
D
M1
s
M
D
M2
s
M
· · · D
MN
s
M
D
11
D
12
· · · D
1M
D
21
D
22
· · · D
2M
.
.
.
.
.
.
.
.
.
.
.
.
D
N1
D
N2
· · · D
NM
(39)
D =
T
1
B
T
2
B
.
.
.
T
M
B
h
2
D
11
s
1
D
12
s
1
· · · D
1N
s
1
D
21
s
2
D
22
s
2
· · · D
2N
s
2
.
.
.
.
.
.
.
.
.
.
.
.
D
M1
s
M
D
M2
s
M
· · · D
MN
s
M
T
B
T
2
.
.
.
T
N
(40)
Y =
Q
1
B
Q
2
B
.
.
.
Q
M
B
(41)
with
T
i
B
(i = 1, · · · , M)
,
T
j
(j = 1, · · · , N)
and
Q
i
B
(i = 1, · · · , M)
, representing
T (X
i
B
, t)
,
T
(x
j
, t) and Q (X
i
B
, t), respectively.
As shown above, the essence of temperature correction procedure is that the heat
source/sink is determined implicitly in such a way that the temperature
T (X (s) , t)
at
the boundary point interpolated from the corrected temperature field
T (x, t)
equals to
the specified boundary temperature
T
B
. In other words, the physical boundary condition
is enforced. The system of linear equations (38) can be solved by a direct method or
an interactive method. After
Q
i
B
at all Lagrangian points are obtained, they are then
substituted into Eq. (35) to obtain the temperature correction
δT
j
, which are further
substituted into Eq. (33) to get finally the corrected temperature T
j
(j = 1, · · · , N).
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 35
3 Evaluating the average number of Nusselt
Nusselt number is a key parameter in the heat-transfer problem. The local Nusselt
number Nu is defined as
Nu (X (s) , t) =
h (X (s) , t) L
c
k
, (42)
where
L
c
is a typical characteristic length. According to Newton’s cooling law and
Fourier’s law, the heat conducted away from the wall by the fluid is equal to the heat
convection from the wall, that is
k
T
n
(X (s) , t) = h (X (s) , t) (T
w
T
) . (43)
Substituting Eq. (43) into Eq. (42) one obtains
Nu (X (s) , t) =
1
T
w
T
k
T
n
(X (s) , t) L
c
k
=
L
c
T
w
T
T
n
(X (s) , t) . (44)
The average Nusselt number
Nu
is an important parameter to examine the rate of
heat-transfer from the heated surface and is defined by
Nu =
1
L
Z
Γ
Nu · ds =
1
L
Z
Γ
L
c
T
w
T
T
n
(X (s) , t) ds , (45)
where
L
=
R
Γ
ds
is the total length of the immersed boundary. As shown in Eq. (44) and
(45), the evaluation of the local and average Nusselt numbers involves the calculation of
the temperature gradient at the boundary points.
3.1 Estimation of the average Nusselt at Eulerian points
As shown in Eq. (8), the heat source at one Eulerian points is from the heat flux at
a few Lagrangian points. If we consider all the Eulerian points which receive the heat
from the immersed boundary, then from energy conservation law, we have
Z
q (x, t) dx =
Z
Γ
Q (X (s) , t) ds , (46)
in which the l.h.s. (left hand side) is the volume integral in the whole fluid domain, and
the r.h.s. (right hand side) is the surface integral along the cylinder wall. According to
Fourier’s law, Q (X (s) , t) can be written as
Q (X (s) , t) = k
T
n
(X (s) , t) . (47)
36 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
Substituting Eqs. (32) and (47) into Eq. (46) gives
Z
ρc
p
δT
δt
dx =
Z
Γ
k
T
n
(X (s) , t) ds . (48)
The r.h.s. of Eq. (48) can be also be written in terms of local Nusselt numbers
Nu (X (s) , t) as
Z
Γ
k
T
n
(X (s) , t) ds =
Z
Γ
k · Nu (X (s) t) (T
w
T
)
L
c
ds =
Nu (T
w
T
) kL
L
c
(49)
Eq. (48) can then be simplified to
Nu (T
w
T
)
L
c
kL =
Z
ρc
p
δT
δt
dx . (50)
Thanks to the approximation of the volume integral offered by the mean theorem, the
equality (50) can be finally written as
Nu =
P
j
ρc
p
δT
j
δt
x
j
y
j
kL (T
w
T
)
L
c
, (j = 1, · · · , N) . (51)
In Figs. (2) and (3), we present the behaviour of the flow around the neighborhood
of the ishotermal square cylinder (bluff body) at different values of
Re
(
Re
= 1
,
2 and
Re = 10) and Ri (Ri = 0 and Ri = 0, 5).
(a) (b)
Fig. 2. Streamlines for flow around ishotermal square cylinder at Re = 1 and Re = 2 for different
Richardson numbers (Ri).
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 37
(c) (d)
Fig. 3. Streamlines for flow around ishotermal square cylinder at
Re
= 10 for different Richardson
numbers (Ri).
The fluid begins the process of separation of bluff body when
Re
2, for different
Ri
values. A detachment process from the circulation bubble is also observed in Fig. (2-(c))
for Re = 10 with Ri = 0.
Other works using different numerical methodologies have presented a similar scenario;
for example, see Chatterjee [
14
], where it is conducted a study for a pair of square
cylinders in a tandem vertical channel configuration, with low Reynolds numbers. Here,
the author concluded that the beginning of separation of the flow occurs for
Re
= 1
2,
Re
= 2
3 and
Re
= 3
4, considering the same values for the Richardson number
presented in our work. Therefore, in this work, it was observed that, as the Reynolds
number is (gradually) increased, the separation of the flow occurs at the trailing edge of
the immersed ishotermal body and the two (symmetrical) vortices of the recirculation
bubble begin to be formed downstream of the obstacle, even under the permanent
outflow regime. This can be attributed to the increase in buoyancy, which leads to a
higher velocity gradient on the surface of the cylinder. Consequently, the pressure on
the immersed body’s surface decreases. Other significant studies, including Singh [
15
]
and Moulay [
16
], in addition to presenting similar scenarios, there is a critical value to
be studied for the adimensionless number
Ri
, i.e. the authors take into account the
onset of the vortex detachment and analyze the influence of this parameter with the
increase of Re.
Table 1. Comparison of drag coefficiet (
C
d
) and average Nusselt (
N u
) numbers for isothermal
boundary condition, for different Reynolds (Re) numbers.
Re Cd Nu
Present 1 4,8500 0,6917
Anjaiah et. al. [17] 1 4,8600 0,6916
Dhiman et. al. [18] 1 4,8714 0,6916
Sharma et. al. [19] 1 4,8714 0,6928
Present 10 0,9300 1,5532
Anjaiah et. al. [17] 10 0,9437 1,5624
Continued on next page
38 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
Table 1 continued from previous page
Dhiman et. al. [18] 10 0,9437 1,5624
Sharma et. al. [19] 10 0,9343 1,5573
Present 40 0,2497 2,6802
Anjaiah et. al. [17] 40 0,2538 2,6969
Dhiman et. al. [18] 40 0,2538 2,6969
Sharma et. al. [19] 40 0,2493 2,6012
Note that the r.h.s. of Eq. (51) are dimensionless. The actual form of Eq. (51) for
dimensionless variables depend on the reference characteristics taken for the problem
considered. In the numerical studies reported below, the specific forms of Eq. (51) will
be given.
The heat flux at Lagrangian points obtained from equation system (38) can be directly
used to compute the average Nusselt number, that is,
Nu =
1
L
Z
Γ
L
c
k (T
w
T
)
Q (X (s) , t) ds . (52)
The the discretized form of Eq. (52) is
Nu =
P
i
Q
i
B
s
i
kL (T
w
T
)
L
c
, (i = 1, · · · , M) . (53)
Note that Eq. (53) is also based on dimensionless, variables. Its specific form for
dimensionless variables will be provided in Section 5.
4 Onset of turbulence
In addition to laminar regime flow, we have also simulated (onset of turbulence) flows
for Reynolds numbers going up to 5
×
10
2
. Thus, it is necessary to use turbulence
models to close the Navier-Stokes equations. A model has been implemented, namely
the Spalart-Allmaras model, with URANS model (Unsteady Reynolds Averaged Navier-
Stokes).
The URANS model is used to refer to RANS (Reynolds Averaged Navier-Stokes), where
in this model, the dependent variables of the Navier-Stokes equation are decomposed
into filtered components and floating components, and then all terms are filtered.
4.1 Spalart-Allmaras model
The Spalart-Allmaras model is a one-equation model that solves the transport equation
for the kinematic eddy turbulent viscosity [
29
], without involving turbulent energy,
dissipation or vorticity calculations, available in other models. Thus, the Spalart-
Allmaras model concentrates into a single parameter the behavior of turbulence, being
classified as a closed model.
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 39
The equation for the turbulent viscosity is constructed using mainly flow empirical
considerations, dimensional analysis and Galileo’s principle of relativity. The model uses
a working variable,
e
ν, given by the following transport equation
e
ν
t
+
x
j
(u
j
e
ν) = c
b
1
(1 f
t
2
)
e
S
e
ν +
1
σ
"
x
j
(ν +
e
ν)
e
ν
x
j
!
+ c
b
2
e
ν
x
j
e
ν
x
j
#
c
w
f
w
c
b
1
k
2
f
t
2
e
ν
d
w
2
+ f
t
1
U
2
,
(54)
where the terms on the right side are, respectively: the production of turbulent viscosity,
the molecular and turbulent diffusion of
e
ν
, the dissipation of
e
ν
, the destruction of
e
ν
,
which reduces the turbulent viscosity near the wall, and, finally, the terms that model
the transition effects, indicated by sub-index
t
. The turbulence viscosity,
ν
t
, is calculated
from the auxiliary variable of the Spalart-Allmaras model and damped by the function
f
v1
along the wall, and is given by:
ν
t
=
e
νf
v
1
, (55)
where
f
v
1
=
χ
3
χ
3
+ C
3
v
1
, (56)
with
χ =
e
ν
ν
. (57)
5 Results
5.1 Mixed convection around a isothermal cylinder
Forced convection over an isothermal cylinder has been extensively studied and used
as a benchmark to examine the capability of numerical methods. The experimental [
27
]
and numerical [
17
]-[
26
] results are available for square cylinder. The fluid and heat flow
are characterized by Reynolds number
Re
=
ρU
D
/µ
and Prandtl number
P r
=
µc
p
/k
,
where
ρ
is the fluid density,
U
is the free stream velocity,
D
is the cylinder diameter,
µ
is the dynamic viscosity,
c
p
is heat at constant pressure and
k
the thermal diffusivity.
In this work, numerical simulations are conducted for different low Reynolds numbers
Re
= 1
500, while keeping the Prandtl number fixed at
P r
= 0
.
7. Both heat and
fluid flow characteristics like the drag coefficient
C
d
, recirculation behind the cylinder,
streamline and isotherm patterns, average Nusselt number on the cylinder surface are
presented and compared with previous results in the literature.
A schematic view of two-dimensional heat and fluid flow across a heated square
cylinder is shown in Fig. (1). In the simulation, governing equations in dimensionless
form are used. The free-stream velocity
U
and the cylinder diameter
D
are chosen as
the reference velocity and reference length, respectively. The temperature is normalized
by
T
=
T T
T
c
T
, (58)
40 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
where
T
c
is the temperature on the cylinder surface and
T
is the free-stream temperature.
The specific dimensionless forms to compute the average Nusselt number by the methods
1 and 2 abovementioned, are
Nu =
Re · P r
π
X
j
δT
j
δt
x
j
y
i
, (j = 1, · · · , N) , (59)
and
Nu =
P
i
Q
i
B
s
i
π
, (i = 1, · · · , M) , (60)
respectively. Note that in Eq. (59), the dimensionless form of
Q
i
B
is actually
T
n
(X
i
B
, t)
,
which can be directly obtained from the solution of equation system (38). The drag
coefficient is defined as
C
d
=
F
d
1
/2 ρU
2
D
, (61)
where
F
d
is the drag force that, in the present work, it is computed from the solution of
equation system (25), which makes the drag force calculation very easy. In the order
to validade our code, simulations with a single heated square cylinder were compared
with the results found in the literature. To mimic free boundary conditions, it was
considered
L
x
= 55
d
and width
L
y
= 30
d
. The heated isothermal square cylinder was
placed at x = 16.5d and y = 15d, to minimize boundary influence. The grid resolution
for these simulations was 318
×
164, 840
×
600 and 1120
×
800 points, a non-uniform
grid was used to better capture the effects with a total of 202 Lagrangian grid points
inside the immersed body. The grid is uniform in the region of the cylinder, maintaining
a minimum number of 30 grid inside. The time step used in the calculation process
is in the range 1
.
0
×
10
6
s
to 1
.
0
×
10
4
s
, which is dynamically calculated by the
Courant-Friedrichs-Lewy (CFL) condition:
CF L =
max (|u| + c)
[x
j
0.5, x
j
+0.5]
× t
x
j
. (62)
To avoid numerical problems, the maximum ratio of 3% grid expansion was employed
in this region in both directions. In the present work, the ratio between Lagrangian
and Eulerian grid size was maintained at 0
.
98. The boundary condition were prescribed
as follow: 1) uniform and nonuniform flow with velocity
U
, pressure
p
= 0, and
temperature
T
input
= 0 at the entrance of the domain (left side); 2) flow fully developed
in the output
(u/∂x = 0, T/∂x = 0)
; 3) condition of symmetries at the upper and
lower boundaries
(u/∂y = v = 0)
,
(T/∂y = 0)
. Concerning the initial condition, we
have considered
u
=
v
= 0, at
t
= 0 (time), for the entire computational domain. The
square cylinder is maintained at a constant dimensionless temperature equal to 1, i. e.,
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 41
T
c
= 1, while the fluid has an initial temperature equal to zero (
T
f
= 0). Although the
boundary condition at the output is not a reflexive condition, the simulation results
successfully corroborate the formation of large (vortex) structure outside the domain
without any reflection. The current pressure boundary conditions at the inlet and outlet
are imposed to ensure consistency with the equations for velocities, which is achieved
through the arrangement of the staggered grid.
5.2 Mixed convection around isothermal cylinders in tandem
In this section, the flow around a pair of heated circular cylinders with different
configurations is studied, where the heat transfer process around obstacles has its
importance and relevance in engineering problems. Here, two cylinder configurations
will be considered, where in both cases the two cylinders have equal diameters and the
same center-to-center distance (
L
cc
). In the first case, the angle formed by the segment
joining the centers of the two cylinders and the axis of the abscissa is zero.
5.2.1 Description of the problem and boundary conditions
In the Fig. (4) the two cylinders are identical and fixed with the same diameter,
maintained in tandem configuration, the cylinder B, downstream of the cylinder A.
Fig. 4. Ilustration of the computational domain with two cylinders in tandem configuration.
In this case, the cylinders is confined to a channel with free flow, with uniform velocity
(
U
) and constant temperature (
T
c
(> T
)
). The horizontal and vertical spacing between
the cylinders are fixed in
L
u
= 16
.
5
d
and
L
d
= 19
.
5
d
, repectively. These values are
chosen to reduce the effect of boundary condictions on the inlet and outlet relative to
the flow pattern at the cylinder boundary. Moreover, these choices are also consistent
with other contemporary works available in the literature, such as, for example, Mahir
and Altaç [
30
], Sohankar and Etminan [
26
], Santos [
32
], Laidoudi and Bouzit [
33
] and
Santos et. al. [
11
]. The drag (
C
d
) and lift (
C
l
) coefficients for the calculation of the each
cylinder are performed as follows:
42 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
C
d
= C
dp
+ C
dv
=
2F
d
ρU
2
d
, (63)
C
l
= C
lp
+ C
lv
=
2F
l
ρU
2
D
, (64)
where
C
lp
and
C
lv
represent the lift coeficients due to pressure and viscous forces,
respectively.
In a similar way,
C
dp
and
C
dv
represent the drag coeficients due to the pressure and
viscous forces.
F
d
and
F
l
are forces of drag and lift, respectively, acting on the surface of
the cylinder. Thus, the drag and lift coefficients can be obtained from the expressions:
C
dp
= 2
Z
1
0
(p
f
p
r
) dy ,
C
dv
=
2
Re
Z
1
0
(
u
y
!
s
+
u
y
!
i
)
dx +
u
x
!
f
+
u
x
!
r
dy
,
(65)
C
lp
= 2
Z
1
0
(p
i
p
s
) dy ,
C
lv
=
2
Re
Z
1
0
v
x
!
f
+
v
x
!
b
dx +
(
v
y
!
t
+
v
y
!
i
)
dy
.
(66)
The subscripts
f
,
b
,
i
and
s
refer to "front", "back", "bottom" and "top" surfaces of the
cylinders, respectively.
The main results for the simulations can be summarized as follows:
A wake is formed upstream of the second cylinder, and it is necessary to investigate
whether it can be reduced or eliminated by increasing the distance between the
cylinders;
The isothermal lines reflect the same behavior of the pattern of the streamlines
(current lines);
The average Nusselt number increase for
Re
= 500 for different values of
Ri
, even
keeping the distance between the cylinders;
The thermal buoyancy is suppressed in the recirculation zones of the tandem
cylinders, even with a mounting angle;
The thermal buoyancy tends to increase the coefficient of drag and the average
Nusselt number of the cylinder more than the second;
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 43
5.2.2 Variations of the Nusselt number
One of the main objectives of heat-transfer calculations involving cylinders is to
determine the local and overall heat-transfer around isothermal cylinders. The effect of
the flow, especially regarding the heat-transfer, can be better observed by analyzing the
local heat-transfer coeffient, also known as the Nusselt local number.
In the Fig. (5), for different Richardson numbers, the distributions of Nusselt numbers
along the perimeter of the upstream and downstream cylinders are provided. For
L
cc
/d
= 3,
Re
= 100,
Re
= 200 and
Re
= 500, for different Richardson numbers, the
local distributions of the Nusselt number along the perimeter of the upstream and
downstream cylinders are provided. For
L
cc
/d
= 3, the heat transfer characteristics
between two closely spaced cylinders differ significantly. While the local Nusselt number
profile of the upstream cylinder resembles that of an isolated cylinder, the downstream
cylinder exhibits distinct behavior. This discrepancy arises due to the strong connection
between heat transfer and fluid flow. Specifically, we observed localized regions of
minimal heat transfer at the front and back stagnation points of the downstream
cylinder. These points experience relatively low flow velocities, which impact the heat
transfer process.
(a) (b)
Fig. 5. Local variation of the Nusselt number to the same dimensionless instant. (a) -
Re
= 100,
Re
200 and
Re
= 500 for
Ri
= 0 (forced convection) and (b) -
Re
= 500 for
Ri
= 1
.
0,
Ri
= 2
.
0
and Ri = 5.0 (natural convection).
Thus, in Fig. (5-(a)), the maximum heat-transfer from the downstream cylinder is
exhibited with a double protuberance occuring in
θ
57
and
θ
265
from the
cylinder wall, where thermal layers (also known as thermal plumes) and hydrodynamics
becomes thinner. The formation of vortices in the downstream region of the cylinder
coincides with the oscillation of the average Nusselt number from large amplitude to
low amplitude during a vortex release period for
L
cc
/d
= 3 and
Re
= 500 for different
values of Ri as see in Fig. (5-(b)).
It is important to note that although the Nusselt’s local distribution of the downstream
44 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
cylinder resembles that of the upstream cylinder, typified as a large protuberance, its
magnitude is smaller than of the upstream cylinder, indicating smaller heat-to-cylinder
convection to downstream.
6 Conclusions
In this work, a boundary condition-enforced immersed boundary method is developed
for simulations of heat and mass transfer problems. The effect of thermal boundaries
in the flow and temperature fields is considered through the velocity and temperature
corrections. The temperature correction is evaluated implicit in such a way that the
temperature at the immersed boundary, interpolated from the corrected temperature
field, satisfies the physical boundary conditions.
For the momentum transfer between the immersed body and the surrounding fluid,
the additional momentum forcing obtained by using the forcing term is added to the
fluid-body equations. To model the onset of turbulence, the Spalart-Allmaras model is
used. This model uses URANS concept, with only one transport equation for turbulence
viscosity, being calibrated in pressure gradient layers. A computational code was
developed to implement the mentioned methodology and analyze the combined heat-
transfer phenomena and the onset of turbulence for thermofluid dynamics interactions
around complex isothermal geometries. The agreement of the results with the available
data in the literature validates the numerical method.
Abbreviations
IBM = Immersed Boundary Method
l.h.s. = left hand side
r.h.s. = right hand side
RANS = Reynolds Averaged Navier-Stokes
Disclosure statement. The authors declare no conflict of interest in the writing of the
manuscript, or in the decision to publish the results.
ORCID
Rômulo D. C. Santos https://orcid.org/0000-0002-9482-1998
Quétila G. Silva https://orcid.org/0009-0000-5928-507X
Samia R. L. Tananta https://orcid.org/0009-0004-6842-5221
References
1.
Peskin, C. S. (1972). Flow Patterns Around Heart Valves: A Numerical Method. J. Com-
put. Phys., 10(2), 252–271. https://doi.org/10.1016/0021-9991(72)90065-4
2.
Peskin, C. S. (1977). Numerical Analysis of Blood Flow in the Heart. J. Comput. Phys.,
25, 220–252.
3.
Park, S. G., Chang, C. B., Kim, B. and Sung, H. J. (2017). Simulation of Fluid-
Flexible Body Interaction with Heat Transfer. Int. J. Heat Mass Transfer, 110, 20–33.
https://doi.org/10.1016/j.ijheatmasstransfer.2017.03.012
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 45
4.
Ashrafizadeh, A. and Hosseinjani, A. A. (2017). A Phenomenological Study
on the Convection Heat Transfer around Two Enclosed Rotating Cylinders via
an Immersed Boundary Method. Int. J. Heat Mass Transfer, 107, 667–685.
https://doi.org/10.1016/j.ijheatmasstransfer.2016.11.078
5.
Mittal, R. and Iaccarino, G. (2005). Immersed Boundary Method. Annu. Rev. Fluid Mech.,
37, 239–261. https://doi.org/10.1146/annurev.fluid.37.061903.175743
6.
Lai, M. C., & Peskin, C. S. (2000). An Immersed Boundary Method with Formal Second-
Order Accuracy and Reduced Numerical Viscosity. J. Comput. Phys., 160(2), 705–719.
https://doi.org/10.1006/jcph.2000.6483
7.
Goldstein, D., Handler, R., & Sirovich, L. (1993). Modeling a No-slip Flow
Boundary with an External Force Field. J. Comput. Phys., 105(2), 354–366.
https://doi.org/10.1006/jcph.1993.1081
8.
Yang, Q. and Cao, S. (2013). Numerical Simulation of Flow around Bluff Bodies Based on
Virtual Boundary Method. The 8th Asia–Pacific Conference on Wind Engineering, Chennai,
pp. 10-14, 582-591. https://doi.org/10.3850/978-981-07-8012-8_154
9.
Mohd-Yusof, J. (1997). For simulations of flow in complex geometries. Annual Research
Briefs, 317.
10.
Ye, T., Mittal, R., Udaykumar, H. S., & Shyy, W. (1999). An Accurate Cartesian Grid
Method for Viscous Incompressible Flows with Complex Immersed Boundaries. J. Com-
put. Phys., 156(2), 209–240. https://doi.org/10.1006/jcph.1999.6356
11.
Santos, R. D., Gama, S. M., & Camacho, R. G. (2018). Two-Dimensional Simu-
lation of the Navier-Stokes Equations for Laminar and Turbulent Flow around a
Heated Square Cylinder with Forced Convection. Applied Mathematics, 9(03), 291–312.
https://doi.org/10.4236/am.2018.93023
12.
Silva, A. L. E., Silveira-Neto, A. and Damasceno, J. J. R. (2003). Numerical Simulation of
Two-Dimensional Flows over a Circular Cylinder using the Immersed Boundary Method.
J. Comput. Phys., 189, 351–370. https://doi.org/10.1016/s0021-9991(03)00214-6
13.
Chorin A. J. (1968). Numerical Solution of the Navier-Stokes Equations. Mathematics of
computation;22(104):745–762. https://doi.org/10.1090/s0025-5718-1968-0242392-2
14.
Chatterjee, D. (2010). Mixed Convection Heat Transfer from Tandem Square Cylinder in
Vertical Channel at Low Reynolds Numbers. Numerical Heat Transfer, Part A: Applications,
58(9), 740–755. https://doi.org/10.1080/10407782.2010.516703
15.
Singh, S., Panigrahi, P. & Muralidhar, K. (2007). Effect of Bouyancy on the Wakes of
Circular and Square Cylinders: A Schlieren-Interferometric Study. Experiments in fluids,
43(1):101–123. https://doi.org/10.1007/s00348-007-0329-8
16.
Moulay, M. A., Belkady, M., Aounallah, A. et.al. (2017). Effect of Opposing Bouyancy on
Two-dimensional Laminar Flow and Heat Transfer Across a Confined Circular Cylinder.
Mechanics, 23(6):859–865. https://doi.org/10.5755/j01.mech.23.6.17291
17.
Anjaiah, N., Dhiman, A. & Chhabra, R. (2006). Mixed Convection Heat Transfer from
a Square Cylinder to Power-Law Fluids in Cross-Flow. In AMSE 2006 2nd Joint US-
Europen Fluids Engineering Summer Meeting Collocated with the 14
t
h
International Con-
ference on Nuclear Engineering, pp. 1435–1444. American Society of Mechanical Engineers.
https://doi.org/10.1115/fedsm2006-98072
18.
Dhiman, A., Anjaiah, N., Chhabra, R. & Eswaran, V. (2007). Mixed Convection from
Heated Square Cylinder to Newtonian and Power-Fluids. Journal of Fluids Engineering,
129(4):506-513. https://doi.org/10.1115/1.2436586
19.
Sharma, N., Dhiman, A. K., & Kumar, S. (2012). Mixed Convection Flow and
Heat Transfer Across a Square Cylinder Under the Influence of Aiding Buoy-
ancy at Low Reynolds Numbers. Int. J. Heat Mass Transfer, 55(9-10), 2601–2614.
https://doi.org/10.1016/j.ijheatmasstransfer.2011.12.034
20.
Breuer, M., Bernsdorf, J., Zeiser, T. & Durst. F. (2000). Accurate Computations of
46 | https://doi.org/10.22481/intermaths.v4i1.12683 R D. C. Santos; Q. G. Silva; S. R. L. Tananta
the Laminar Flow Past a Square Cylinder Based on Two Different Methods: Lattice-
Boltzmann and Finite Volume. International Journal of Heat and Fluid Flow, 21(2):186–196.
https://doi.org/10.1016/s0142-727x(99)00081-8
21.
Kim, J., Kim, D. & Cho, H. (2001). An Immersed-Boundary Finite-Volume Method
for Simulations of Flow in Complex Geometries. J. Comput. Phys., 171(1):132–150.
https://doi.org/10.1006/jcph.2001.6778
22.
Kim, J. & Cho, H. (2004). An Immersed-Boundary Finite-Volume Method for Simulation
of Heat Transfer in Complex Geometries. KSME International Journal, 18(6):1026–1035.
https://doi.org/10.1007/bf02990875
23.
Perumal, D. A., Kumar, G. V., & Dass, A. K. (2012). Numerical Simulation of Viscous Flow
over a Square Cylinder Using Lattice Boltzmann Method. ISRN Mathematical Physics,2012.
https://doi.org/10.5402/2012/630801
24.
Sharma, A. (2003). Numerical Investigation of Unconfined and Channel-Confined Flow
Across a Square Cylinder with Forced and Mixed Convection Heat Transfer. Ph.D. thesis,
Indian Institute of Technology Kanpur, India.
25.
Sharma, A. & Eswaran, V. (2004). Effect of Aiding and Opposing Buoyancy on the Heat
and Fluid Flow Across a Square Cylinder at
Re
= 100. Numerical Heat Transfer, Part A:
Applications, 45(6):601–624. https://doi.org/10.1080/10407780490277798
26.
Sohankar, A. & Etminan, A. (2009). Forced-Convection Heat Transfer from Tandem Square
Cylinders in Cross Flow at low Reynolds Numbers. International Journal for Numerical
Methods in Fluids, 60(7):733–751. https://doi.org/10.1002/fld.1909
27.
Yen, S., San, K., & Chuang, T. (2008). Interactions of Tandem Square Cylinders
at Low Reynolds Numbers. Experimental Thermal and Fluid Science, 32(4):927–938.
https://doi.org/10.1016/j.expthermflusci.2007.07.001
28. Smagorinsky, J. (1963) General Circulation Experiments with the Primitive Equations: I.
The Basic Experiment. Monthly Weather Review, 91, 99–164. https://doi.org/10.1175/1520-
0493(1963)091%3C0099:gcewtp%3E2.3.co;2
29.
Spalart, P. & Allmaras, S. (1992). A One-Equation Turbulence Model for Aerodynamics
Flows. Recherce Aeropatiale, No. 1, 5–21.
30.
Mahir, N. & Altaç, Z.(2008). Numerical Investigation of Convective Heat Transfer in
Unteady Flow Past Two Cylinders in Tandem Arragements. International Journal of Heat
and Fluid Flow, 29(5):1309–1318. https://doi.org/10.1016/j.ijheatfluidflow.2008.05.001
31.
Sohankar, A., Norberg, C. & Davidson, L. (1999). Simulation of Three-Dimensional Flow
Around a Square Cylinder at Moderate Reynolds Numbers. Physics of Fluids, 11(2):
288–306. https://doi.org/10.1063/1.869879
32.
Santos, R. D. C. d. (2014). Análise Bidimensional Termo-Fluido Dinâmica de Cilindros
Rotativos com o Método da Fronteira Imersa/Modelo Físico Virtual. Master’s thesis.
Universidade Federal de Itajubá - Minas Gerais, Brazil.
33.
Laidoudi, H. & Bouzit, M. (2016). Mixed Convection in Poiseuille Fluid from an Asymmet-
rically Confined Heated Circular Cylinder. Thermal Science, (00): 172–172.
34.
Zdravkovich, M. (1985). Flow Induced Oscillations of Two Interfering Circular Cylin-
ders. Journal of Sound and Vibrations, 101(4): 511–521. https://doi.org/10.1016/s0022-
460x(85)80068-7
© INTERMATHS
CC BY-NC 4.0
R D. C. Santos; Q. G. Silva; S. R. L. Tananta INTERMATHS, 4(1), 2547, June 2023 | 47