You are on page 1of 68

NUMERICAL SIMULATION OF TWO-DIMENSIONAL COMPLEX FLOWS

OVER BLUFF BODIES USING THE IMMERSED BOUNDARY METHOD

A. L. F. LIMA E SILVAa,* , A. R. SILVAa AND A. SILVEIRA-NETOa

a
School of Mechanical Engineering, Federal University of Uberlndia - Minas Gerais,

Brazil;

Address: Joo Naves de vila Avenue, Campus Santa Mnica.

Telephone: 55(34) 32394148 R. 213, Fax: 55(34) 32394206

E-mail: alfsilva@mecanica.ufu.br, arsilva@mecanica.ufu.br, aristeus@mecanica.ufu.br

Abstract

This paper presents a two-dimensional numerical simulation of flows around bluff

bodies at Reynolds numbers Re = 100 and 200. The Immersed Boundary Method (IB),

developed by Peskin in 1977, requires the computation of the force density term, which

is added to the Navier Stokes equations. In the present work the Virtual Physical Model

(VPM), proposed by Lima e Silva et al. in 2003 is used. This methodology presents the

advantage of using a Cartesian Eulerian grid to represent the fluid domain and a set of

points that compose the Lagrangian grid, used to model the bluff body. Simulations

were performed for two circular cylinders of different diameter in tandem, two cylinders

of the same diameter in tandem and two cylinders placed in side by side arrangement.

The configurations of seven cylinders in a V arrangement, for angles of

* Corresponding author.
400 1800 , were also simulated. A configuration of 23 different bluff bodies,

obtained by a transverse cut in a central tower of an offshore structure, has been

simulated and the results were compared with a single compact square, of equivalent

size, immersed in the flow. The Strouhal number, the drag and the lift coefficients were

also calculated. The Strouhal number is calculated using the Fast Fourier Transform

(FFT) of the lift coefficient temporal distribution. Visualization of the vorticity and

pressure fields and the streamlines are presented for each simulation showing the flow

dynamics and patterns. It was possible to verify that the IB method with VPM is a

powerful methodology to simulate flows in the presence of complex geometries.

Keywords: Immersed Boundary Method, Virtual Physical Model, Bluff Bodies,

Numerical Simulation, Cartesian Grid.

1. Introduction

2
The Immersed Boundary Method was designed to solve problems such as the

biological fluid mechanics, which involve the interaction of a viscous fluid with an

elastic membrane (Peskin, 1977). This methodology has been extended to solve

complex engineering problems like those with immersed rigid bodies (Lima e Silva et

al., 2003). The main idea of this method is to use a regular, Cartesian-Eulerian grid, to

represent the total fluid domain, together with a Lagrangian grid to represent the

immersed boundary. The boundary applies a singular force to the fluid and, at the same

time, it reacts and moves inside the fluid at the local fluid velocity. The interaction

between the fluid and the boundary can be modeled by a well-chosen discrete delta

function, which is an approximation to the Dirac delta function. This methodology has

been applied successfully for studying the flow interference between bluff bodies and

the wake formed behind them. An important feature of this methodology is the

modeling and computation of the Lagrangian force.

Peskin (1977) proposed to model this force field using the generalized Hookes

law for immersed elastic boundaries. A constant of elasticity has been chosen to

simulate the elastic boundary. Goldstein et al. (1993) proposed to calculate the fluid-

solid interfacial force using a model that depends on two constants adjusted for each

simulation. The force field is chosen along a specified surface to have a magnitude and

direction opposing the local flow such that the flow is brought to the same velocity as

the immersed boundary. This is a kind of virtual boundary condition that is imposed

indirectly by the force field. For unsteady viscous flows the direct calculation of the

force is performed by a feedback scheme in which the velocity is used to interactively

determine the desired Lagrangian force field.

3
Saiki and Biringen (1996) employed the Goldstein model to simulate flows with

immersed solid bodies. The results presented spurious oscillations, which were caused

by the use of the virtual boundary method with a spectral discretization. The oscillations

were attenuated by the application of high-order finite differences.

Mohd-Yusof (1997) derived a discrete-time forcing scheme, which was used with

a spectral method to simulate flows around cylinders and spheres at moderate Reynolds

numbers. The force calculation was done in a way that the velocity value at the first grid

node near the interface was imposed to guarantee the no-slip condition.

Ye et al. (1999) proposed a different approach on a staggered grid, named a

Cartesian Grid Method, which does not use the concept of momentum forcing. In this

method a control volume near the boundary is reshaped into a body-fitted trapezium by

discarding the solid part of the cell and adding the neighboring cells. However, because

the stencil of the trapezoidal boundary cell is different from that of the regular cell, a

special numerical procedure is required in order to solve the discrete momentum

equations at each computational time step.

Fadlun et al. (2000) applied the Mohd-Yusof (1997) developments to a discrete

finite-difference method on a staggered grid. They adapted the Mohd-Yusof formulation

in relation to the velocity calculation at the first grid point. The velocity at the first grid

point external to the body is obtained by linearly interpolating the velocity at the second

grid point, which is determined directly by solving the Navier-Stokes equations and the

fluid velocity at the body surface.

Kim et al. (2001) introduced modifications to the Immersed Boundary Method by

adding a mass source/sink with the momentum forcing. The mass source and the

4
momentum forcing are applied on the body surface or in its neighborhood to satisfy the

no-slip velocity boundary condition.

Lima e Silva et al. (2003) presented a new proposal to model the force field over

the Lagrangian grid. This model does not require any constant to be adjusted nor a

special interpolation scheme near the Lagrangian cells. It has been named the virtual

physical model (VPM). The VPM is based on the momentum equation applied to a fluid

particle located over the fluid-solid interface. The authors presented results for

simulations of flows over a single two-dimensional cylinder at moderate Reynolds

numbers. The Strouhal number, the recirculation bubble length, the drag, the lift and the

pressure coefficient were compared with numerical and experimental results of other

authors. The model is free of modeling constants, as well as any special procedure to

recompose the discretization cells near the interface.

In the present work the Immersed Boundary Method with the Virtual Physical

Model (Lima e Silva et al., 2003) were applied to simulate and analyze flows over bluff

bodies. Simulations of two circular cylinders in tandem and in side by side arrangement

were done. Flows over other complex arrangements of cylinders such as a V

configuration and a composition of different geometries were simulated. The results

illustrate the potential of this methodology to simulate flows over any type of complex

configurations of immersed bodies.

2. Mathematical Model: Momentum Equations and the Virtual Physical Model

5
A mixed Eulerian-Lagrangian formulation is used to represent the flow and the

immersed boundary. A Cartesian grid describes the flow using a Finite Difference

method and a Lagrangian grid, composed by a finite number of points, describes the

immersed bodies. The Eulerian and the Lagrangian grids are coupled by a force field

calculated at the Lagrangian points and then distributed across the Eulerian nodes in the

body neighborhood.

2.1. Fluid Formulation

The momentum and continuity equations for an incompressible, Newtonian

viscous flow can be written as

u
uiu j p ui fi
2
i (1)
t x j xi x j x j

and

ui
0. (2)
xi

The Eulerian force field is given by


f x ,t F x ,t x x dx
k k k , (3)

6

where x xk is a Dirac delta function, xk are the Lagrangian points placed over the


immersed boundary (see Fig. 1), F xk is the Lagrangian force density and f x is the

Eulerian force, which is different from zero only over the immersed boundary. Equation

(3) models the interaction between the immersed boundary and the fluid flow, injecting

the force field into the fluid.

2.2. Fluid-Interface Coupling

The Dirac delta function which appears in Eq (3) is replaced by the distribution

function, Dij , proposed by Peskin (1977), in order to calculate the Eulerian force, in a

discrete form. Therefore Eq. (3) is replaced by


f x ,t D x x F x ,t s
k k
2
, (4)

with


D x xk

f xk - xi / h f yk - y j / h
,
(5)
h2

f 1 r if r 1

1
f r - f 1 2 r if 1 r 2 , (6)
2
0 if r 2

7
and

2
32 r 1 4 r 4 r
f1 r . (7)
8


The D x xk term is calculated from Eqs. (5)-(7) and s is the distance between the

Lagrangian points, as indicated in Fig. 1. The parameter r is x k xi / h or y k y j / h

, h is the Eulerian grid size and xi , y j are the coordinates of the Eulerian points.

2.3. Interface Force Field Formulation

The Virtual Physical Model developed by Lima e Silva et al. (2003) and used to

calculate the Lagrangian force field is based only upon the momentum equations. A

momentum balance is done over the fluid particle, illustrated in Fig. 2, taking into

account the Lagrangian force field F x k ,t . Isolating this Lagrangian force yields


F x k ,t Fa x k ,t Fi x k ,t F x k ,t F p x k ,t . (8)

8
The different terms that compose Eq. (8) are here named the acceleration force

Fa , the inertial force Fi , the viscous force F and the pressure force Fp . These

terms are given by


V
Fa x k ,t xk , (9)
t



Fi x k ,t VV , (10)


F x k ,t 2V , (11)


F p x k ,t p x k . (12)


2.4. Method to Calculate the F x k ,t Terms

The terms described by Eqs. (9)-(12) must be evaluated over the interface using

the Eulerian velocity field V x and pressure field p x . These calculations also take

into account that, at the interface, the fluid velocity must be equal to the interface

velocity, which guarantees the no-slip boundary condition. The velocity and pressure

spatial derivatives are computed using the fluid velocity and pressure obtained by means

of Eqs. (1) and (2).

The derivatives that compose the different terms are calculated using a second-

order polynomial of Lagrange. Equations (13)-(16) were established for a generalized

variable , which can represent the velocity components and the pressure.

9
The derivatives in the x direction are given by


xk , y k xk x2 1 xk x1 2 xk x1 xk x2 k (13)
x x1 x2 x1 xk x2 x1 x2 xk xk x1 xk x2

and

2 21 2 2 2 k
xk , y k
x 2
x1 x 2 x1 xk x2 x1 x2 x k xk x1 xk x2 , (14)

with similar terms representing the y direction derivatives.

The scheme used for these calculations was presented by Lima e Silva et al.


(2003) and is summarized in Fig. (3). The velocity components u xi ,t , xi ,t and


p xi ,t must be interpolated over the auxiliary points xi , with i = 1, 2, 3 and 4 (see

Fig. 3). The velocity and pressure interpolations are performed using the neighboring

Eulerian grid as illustrated by Figs. (4), (5) and (6). These interpolations do not account

for the velocities and pressure values inside the interface.

After the velocity components and the pressure on the auxiliary points 1, 2, 3 and

4 are known, they must be evaluated at point k over the interface. The velocities are

interpolated using all the Eulerian values around point k, as illustrated by Fig. 7(a). The

pressure at this point is estimated as shown by Fig. 7(b). The nearest Eulerian cell in the

normal direction is found and its pressure is transferred to point k.



After computing the Lagrangian force F x ,t at each point k over the interface,

this force is distributed using Eq. (4). This distribution process is illustrated in Fig. 8.

10
3. Numerical Algorithm

Equations (1) and (2) are solved by the finite difference method through a

fractional step pressure correction method (Armfield and Street, 1999). As follows, an

estimation of the velocity is explicitly calculated by

u in 1 u in 1 p n
t

xi

N u in L u in f i n , (15)

where u i is the estimated velocity component, t is the computational time step, n is

the substep index,

2ui
L ui (16)
x j x j

and

ui u j
N ui . (17)
x j

The pressure correction, n1 , is obtained by solving

11
2 n 1 u in 1
. (18)
x j x j t xi

The velocity field is updated using the solution of Eq. (13) by

n 1
u in 1 u in 1 t . (19)
xi

The previous pressure field and the correction pressure are used to calculate the

new pressure field using

p n 1 p n n 1 . (20)

The numerical method can be summarized as follows:



(a) The force field F x k ,t is calculated over the Lagrangian points x k , y k , by Eqs.

(8)-(12) using the velocity and pressure fields at the previous time;

(b) The force F x k ,t is distributed to the Eulerian grid using Eqs. (4)-(7);

(c) The fluid velocity field under the influence of the force field f is estimated using

Eq. (15);

(d) The linear system for the pressure correction, given by Eq. (18), is solved using an

iterative method;

(e) The fluid velocity is updated using Eq. (19);

(f) The pressure field is computed using Eq. (20);

(g) The divergence is verified using Eq. (2);

(h) The steps above are repeated for the new time step.

12
The linear system for the pressure correction, Eq. (18), is solved by the modified

strongly implicit procedure (MSI), Schneider and Zedan (1981). The interface force

field and the momentum equations are computed using an explicit method.

4. Numerical Results

4.1. Preliminary Aspects

The simulations were performed at Reynolds numbers ( Re U d / ) equal to

100 and 200, where U is the freestream velocity and d is the circular cylinder

diameter or the characteristic length for the other geometries and is the kinematic

viscosity of the fluid. The dimensionless time was defined as T tU / d , where t is the

physical time in seconds. Once the pressure and velocities fields are obtained, the drag

and lift coefficients and the Strouhal number are calculated using the Lagrangian force

field directly. Therefore, these coefficients are defined as

2 Fd
Cd , (21)
U 2 d

2 Fl
Cl , (22)
U 2 d

where Fd and Fl are the drag and the lift forces, respectively, given by

13

Fd f x dx Fx ds , (23)
0


Fl f y dx Fy ds (24)
0

where fx and fy are the Eulerian forces and Fx and Fy are the Lagrangian forces in the x

and y directions; is the domain around the interface in the Eulerian grid and is the

interface length.

The Strouhal number is defined as the dimensionless frequency of vortex

shedding behind the body, calculated by

fd
St , (25)
U

where f is the frequency of vortex shedding, which is determined by the FFT (Fast

Fourier Transform) of the lift coefficient time distribution.

The time step t 1.10 6 s is used at the first iteration and is increased gradually,

up to t 1.10 3 s , to ensure numerical stability. As the geometry of the interface is

established by the force field, the time step is calculated in such a way that the CFL <

1.0 criterion is attained.

A grid refinement was done with the objective of to verify its influence on the

drag and lift coefficients. It was observed the necessity of at least 30 grids inside the

body to guarantee the grid independence. At the present simulations 40 grids were used.

The aspect ratio of the domain is also important and therefore this study was also done

for each simulation.

14
This methodology was very well validated by Lima e Silva et al. (2003) for flows

over a single cylinder.

4.2. Cylinders of Different Diameters in Tandem

The IB simulations have been carried out in the domain shown in Fig. 9(a). This

figure also shows a global view of the domain with the main geometrical parameters.

Figure 9(b) shows a closer view of the arrangement with the Cartesian and the

Lagrangian grids. The upstream and the downstream cylinders diameters are d and d/2,

respectively. The simulations were done for the distances between the cylinder centers

of L = 1.5d, 2.0d, 2.5d, 2.7d, 3.0d and 4.0d. A grid of 500x250 points in the x and y

directions respectively was used. The flow direction is from the left to the right side of

the domain. The simulations were carried out at a Reynolds number Re = 200, based on

the diameter of the upstream cylinder, and the dimensionless time step, T tU / d , is

0.015. The indices 1 and 2 refer to the upstream and downstream cylinder, respectively.

The vorticity contours for four values of the time steps are shown in Fig. 10, for L

= 2.7d. It can be observed that the cylinder 2 (downwind cylinder) is embedded in the

vortex street generated by the cylinder 1 (Fig. 10(a)). As time advances, the vortex shed

by cylinder 1 retracts downward and begins to collide frontally against cylinder 2. This

phenomenon starts at T =150, as shown by Fig. 10 (c). At this time, the flow near

cylinder 2 changes its direction and this cylinder starts shedding vortices. These effects

(collision and vortex shedding) are directly observed in the temporal evolution of the

drag coefficients as shown in Fig. 11. There is a change in their behavior. The drag

coefficient of cylinder 2 (Cd2) presents negative values for T up to 150. For T >150, the

15
drag coefficient assumes a positive time distribution. This behavior has its origin in the

vortices transport dynamics explained above. The temporal sequence illustrated in Fig.

10 shows that as time passes the vortex shed from the upstream cylinder retracts and

collides frontally with the downstream cylinder. This behavior explains why Cd2 changes

its sign as a function of time. The Cd1 distribution also presents a sudden increase at the

same time T =150.

For distances smaller than 2.7d the wake formed by the cylinder 1 envelopes

cylinder 2 and near its surface there is a flow in the opposite direction (from right to

left). This behavior can be seen in Fig. 12 using the vorticity field, for the case where

the distance between the cylinders is L = 1.5d, Fig. 12(a) and L = 2.5d, Fig. 12 (b).

Figure 13 shows the velocity vectors, confirming the direction of the flow near the

smallest cylinder.

It is important to mention that in the IB methodology, the instantaneous and the

mean drag coefficient and the Strouhal number are directly obtained through the

Lagrangian force, calculated over the immersed body. We can follow the evolution of

the drag coefficient with the time as presented in Fig. 14 for L = 1.5d and for L = 2.5d.

The negative value of the drag coefficient is observed for both cases.

The drag coefficient for both cylinders as a function of the distance between their

centers is shown in Fig. 15. The drag coefficient sign of cylinder 2 switches from a

negative value to a positive value when the vortices of cylinder 1 start to collide with

cylinder 2, as presented before. The critical value of L/d = 2.7d is a little larger than the

value of 2.5d, presented by Surmas et al. (2003) and Flatschart et al. (2000).

Figure 16 shows the results of the Strouhal number. It was calculated with the

time distribution of Cl (lift coefficient) of the upstream cylinder. In the IB method the

16
Strouhal number can be easily obtained because the force distribution is already known.

The Strouhal number was calculated by taking the Fast Fourier Transformer (FFT) of

the lift coefficient distribution. The results have the same behavior when compared with

other methodologies (Surmas et al., 2003 and Flatschart et al., 2000).

4.3. Cylinders of Same Diameters in Tandem

These simulations were done for two cylinders of same diameter d separated from

each other by the following distances L = 1.5d, 2.0d, 3.0d and 4.0d. The simulations

were carried out on a 500x250 grid for Re = 200. The domain used in these simulations

and the geometrical parameters are shown in Fig. 17. As mentioned before, in this

arrangement, the index 1 is used for the upstream cylinder and the index 2 refers to the

downstream cylinder.

In Fig. 18, the vorticity field of the present work is compared with the result of

two other authors (Surmas et al., 2003 and Meneghini et al., 2001). These results

correspond to the time at which the vortices collide frontally with the second cylinder.

At this time the drag coefficient of cylinder 2 changes from negative to positive.

Figure 19 shows the drag coefficient of cylinder 1 (upstream) and cylinder 2

(downstream) as a function of the distance between their centers. This figure clearly

shows the point where the drag coefficient sign changes to positive. The transition

occurred at L = 4.0d in the present work, which is the same value presented by Surmas

et al. (2003) and Meneghini et al. (2001). When the cylinders have the same diameter,

the sign transition occurs in the larger distance (L = 4.0d) as compared with simulations

with cylinders of different diameters (L = 2.7d).

17
The Strouhal number, based on the diameter of cylinder 1, is presented in Fig. 20

as a function of the distance and between the cylinders centers. The results of the

present work are compared with results of other authors. The comparisons show that the

agreement among the three methods is very good. The present Strouhal value was 2

percent lower than the obtained by Meneghini et al. (2001) for L = 4.0d.

4.4. Cylinders of Same Diameter Side by Side

These simulations were done with two cylinders of same diameter disposed side

by side. The gap L between their centers was considered as 1.5d, 2.0d, 3.0d and 4.0d.

The simulation domain is shown in Fig. 21. A grid of 500x350 nodes was used and Re =

200, as in the last two cases.

Figure 22 shows the vorticity (left column) and pressure contours (right column)

for all the distances simulated. Similar results of Meneghini et al. (2001) and Surmas et

al. (2003) are presented in Fig. 23 for L = 2.0d and L = 4.0d. There is a great

interference between the wakes of the cylinders for small distances (a repulsive force

between the cylinders). This repulsive force is due to the movement of the stagnation

points of both cylinders where the pressure is high as observed by Meneghini et al.

(2001). The flow has a region of intense instabilities that cause changes in the drag and

lift coefficients. For L 2.0d only one wake is formed and it deflects in the direction

of one of the cylinders and later in the opposite direction (flopping phenomenon). This

behavior was also observed by Bearman and Wadcock (1973), Williamson (1985) and

Meneghini et al. (2001). The formation of two independent wakes can be observed as

18
the distance between the cylinders increases, and the wakes become independent of

each other as in isolated cylinders.

The mean drag coefficient is shown in Fig. 24. It has the same value for both

cylinders because they are symmetrically placed in the domain. The results of Surmas et

al., (2003) and Meneghini et al. (2001) are also presented in this figure. Again the

agreement among the three methodologies is not good and experimental results are

required in order to draw conclusions about these results.

The mean lift coefficients are presented in Fig. 25 for both cylinders. This

parameter has the same magnitude for both cylinders but with opposite signs. The index

1 indicates the upper cylinder and the index 2 indicates the lower cylinder. There is an

attraction force between the cylinders. As the gap between them increases, the mean lift

coefficient for both cylinders tends to zero. This behavior was expected as it was

observed for a single cylinder.

4.5. Cylinders of Same Diameter in V

The configuration shown in Fig. 26 was chosen to analyze and to verify the

influence between the wakes formed by seven cylinders of same diameter d. The

cylinders were arranged in different configurations by changing the angle . This


angle was assumed equal to 400, 600, 800, 1000, 1400 and 1800. The diagonal distance

between two consecutive cylinders, Fig. 26, was maintained equal to 2.5d for all the

simulations (different values of ). Cylinders 1 and 7 where chosen to be analyzed.


The Reynolds number based on the diameter of one cylinder is 100. The grid size used

for these simulations was 500x500 and the domain has dimensions 50d x 50d.

19
Figure 27 shows the vorticity field for all the angles simulated. It is interesting to

note that the vortex formation over individual cylinders is totally inhibited when the

angle is small. This flow is similar to that of a single immersed body like a delta.
The wake structure changes as the angle increases. The flows become more
complex and all the cylinders start to shed vortices, which strongly interact. There is a

great influence between their wakes, which changes the drag force coefficients.

The flow behavior, when the angle is equal to 1800, is compared qualitatively
with the experimental result of Fantasy of Flow (1993), Fig. 28, showing that both

results are very similar.

Figure 29 shows the drag coefficient as a function of the dimensionless time for

all the simulated angles . The drag coefficients for cylinders 1 and 7, numbered in Fig.

26, are compared with the drag coefficient of a single cylinder. For angles smaller than

800 cylinders 1 and 7 have drag coefficients smaller than that for a single cylinder.

Cylinder 1 has, for this range of , the smallest values of the drag coefficient, as can
be seen in Fig. 29 (a)-(c). As the angle decreases, the wake becomes more inhibited. The

amplitude of the wake oscillations increases with the increase of angle . Also, as

increases, the time variation of the drag coefficient becomes more and more complex.

This is associated with the complexity of the wakes, as can be seen in Figs. 27 (d) - (f).

The mean drag coefficient and the Strouhal number are shown in Table 1 for

cylinders 1 and 7. At = 400, a minimum value of Cd1 = 0.7323 is reached and St1 =

0.0851. These values are 52% and 44%, respectively, of the values found in an isolated

cylinder (Cd = 1.3997, St = 0.1955).

For 1000 the values of the mean drag coefficients for all cylinders increase

and reach values higher than that for an isolated cylinder. The amplitude of oscillation is

20
also higher as is increased, as can be seen in Fig. 29. The maximum value of Cd1 =
2.2272 is 60% higher than the value for an isolated cylinder.

4.6. Combination of Bluff Bodies of Several Shapes

The IB methodology with the VPM has the advantage of providing a flow

simulation around several geometries, without any type of additional complexity in the

grid generation process. The grid is always Cartesian irrespective of the complexity of

the configuration of immersed bodies. It is possible to obtain the force coefficients on

each immersed body or an overall value, because the force distribution is known over

each body individually at each time step.

In order to illustrate these characteristics, a transverse horizontal cut at the main

tower of an offshore structure was considered, which is illustrated in Fig. 30 (a). These

structures are subject to strong wind streams and maritime currents, which can cause

vibration, damaging the structural integrity. The schematic view of a composition of

bluff bodies, generated by the transverse cut, is shown in Fig. 30 (b). Each body was

numbered to facilitate the identification. For these simulations a non-uniform grid with

400x300 points was used, as illustrated in Fig. 31. The Reynolds number based on the

dimension H, Fig. 30(b), is Re = 200. The distance between the external bodies (bodies

1 to 16) was chosen equal to 0.1H, as indicated in Fig. 30(b).

The streamlines for the dimensionless times T = 10s, 20s, 30s and 44s are shown

in Fig. 32. The fluid flows through the immersed bodies but there is no vortex shedding

inside the structure because all bodies are very close. The system of bodies can be

considered as a porous media. The recirculation bubble generated behind the global

21
structure has the same order of H. The results of the mean drag coefficients of each

geometry are shown in Table 2. The value of the global drag coefficient calculated by

the sum of the forces from each body is equal to 1.9492.

We can observe that there are mass injections in the shear layer due to the gap

between the bodies. As a consequence the horizontal length of the recirculation bubble

is bigger than that of the flow over a square of the same dimensions. Figure 33 shows

the streamlines of the simulations with a square immersed in the flow for the non-

dimensional instants of time of T = 10s, 20s, 30s and 44s. For these simulations the

value of the mean drag coefficient was 1.6844. This value is 13% smaller than the drag

coefficient of the composition of bluff bodies.

The time evolution of the drag coefficient is represented in Fig. 34 for the

geometries numbered in Fig. 30. Figure 34(a) shows the time distribution of the drag

coefficient for bodies 3 and 11. These two bodies are placed symmetrically. The drag

distributions are out of phase but the amplitudes are the same for both bodies. The same

behavior is observed in Figs. 34(b)-(g). Figure 34(f) shows the drag coefficient for

bodies 21, 22 and 23 placed inside the configuration. As expected, the drag coefficient

of these bodies is small as compared with the downstream bodies (bodies 1 and 13-18).

Figure 34 (h) shows the drag coefficient for bodies 7 and 15 placed in the downwind

and upwind stagnation positions. The drag coefficient of body 15 is greater than body 7,

as expected. This results illustrate that it is possible, using the IB-VPM methodology to

easily calculate the drag and lift coefficients of each body that compose such a complex

geometry.

22
5. Conclusions

The Immersed Boundary Method with the Virtual Physical Model was used to

simulate flows over different compositions of bluff bodies. The results of the mean drag

coefficient and the Strouhal number for cylinders in tandem and side-by-side were

compared with other authors results. The flow dynamics changes with the distance

between the cylinders. It was observed the presence of a repulsion force when the

distance between the bodies is small. Consequently the drag coefficient is affected and

can change from negative to positive. For the simulations with different cylinders and

with two identical cylinders in tandem, it was observed that the distance where the Cd

sign changes was 2.7d and 4.0d, respectively. In the side-by-side arrangement, a

repulsive force between the cylinders has been observed for gaps L 2.0d . The

simulations with seven cylinders in a V configuration show the potential of the

methodology to study flows over complex geometries. There is no restriction on the

number of bodies and the computational cost is not significantly increased when a great

number of bodies are immersed in the flow. In other words the main cost is due to the

solution of the linear system for the pressure correction. With these simulations, it is

possible to choose the optimal configuration for which the drag coefficients are

minimized. The last case simulated was the flow in the presence of different bluff

bodies, illustrating a horizontal section on an offshore structure. The global drag

coefficient was compared with that of a square with the same dimensions. The drag

coefficient value obtained for the compact square was approximately 13% smaller than

that for the composed geometry. There was a change in the shape of the recirculation

bubble and the vortex shedding starts earlier than it is for the simulations with a square.

23
It is also important to emphasize that in the IB method a Cartesian grid is used to

simulate flows over any geometric shape. Another advantage is the ease of calculating

the force coefficients for an isolated body or the global value for the configuration.

6. Acknowledgements

The authors acknowledge CNPq (National Council for Scientific and Technological

Development, from Brazilian Ministry of Education) and Fapemig (Foundation for the

Support of Research in Minas Gerais, Brazil) for the financial support and the

Mechanical Engineering College of the Federal University of Uberlndia.

7. References

Armfield, S. and Street, R., 1999. The Fractional-Step Method for the Navier-Stokes

Equations on Staggered Grids: The Accuracy of three Variations. Journal of

Computational Physics, 153, 660.

Bearman, P. W. and Wadcock, A. J., 1973. The interaction between a pair of circular

cylinders normal to a stream. Journal of Fluid Mechanics, 61, 499-511.

24
Fadlun, E. A., Verzicco, R., Orlandi, P. e Mohd-Yusof, J., 2000. Combined Immersed-

Boundary finite-Difference Methods for Three-dimensional Complex Flow Simulations.

Journal of Computational Physics, 161, 35.

Fantasy of Flow: The World of Fluid Flow Captured in Photographs, Vol. 1, 1993, The

Visualization Society of Japan (Editor), IOS Press, Hardcover, ISBN 90-5199-138-x.

Flatschart, R. B., Saltara, F., Ferrari Jr., J. A., Meneghini, J. R., Siqueira, C. R., 2000.

Numerical Simulation of Flow Interference Between two Circular Cylinders in Tandem.

19th International Conference on Offshore Mechanics and Artic Engineering, February

14-17, New Orleans, LA, USA.

Goldstein, D., Adachi, T. e Sakata, H., 1993. Modeling a No-Slip Flow with an External

Force Field. Journal of Computational Physics, 105, 354.

Juric, D., 1996. Computation of phase change. Ph. D. Thesis, Mech. Eng. Univ. of

Michigan, USA.

Kim, J., Kim, D. and Choi, H., 2001. An Immersed-Boundary Finite-Volume Method

for Simulations of Flow in Complex Geometries. Journal of Computational Physics,

171, 132.

Lima e Silva, A. L. F., Silveira-Neto, A. and Damasceno, J. J. R., 2003. Numerical

Simulation of Two Dimensional Flows over a Circular Cylinder using the Immersed

Boundary Method. Journal of Computational Physics, 189, 351-370.

25
Meneghini, J.R., Saltara, F., Siqueira, C.L.R., Ferrari Jr, J. A., 2001. Numerical

Simulation of Flow Interference Between Two Circular Cylinders in Tandem and Side-

by-Side Arrangements. Journal of Fluids and Structures, 15, 327350

Mohd-Yusof, J., 1997, Combined Immersed boundaries/B-Splines Methods for

Simulations of Flows in Complex Geometries. CTR Annual Research Briefs, NASA

Ames/Stanford university.

Peskin, C. S., 1977. Numerical Analysis of Blood flow in the Heart. Journal of

Computational Physics, 25, 220.

Saiki, E. M. e Biringen, S., 1996. Numerical Simulation of a Cylinder in Uniform Flow:

Application of a Virtual Boundary Method. Journal of Computational Physics, 123,

450.

Schneider, G. E. and Zedan, M., 1981. A Modified Strongly Implicit Procedure for the

Numerical Solution of Field Problems. Numerical Heat Transfer, 4, 1.

Surmas, R., Santos, L. O. E. and Philippi, P. C., 2003. Caractersticas da Formao e

Desprendimento de Vrtices em Grupos de Cilindros. Segundo Congresso Brasileiro de

Pesquisa e Desenvolvimento em Petrleo e Gs, Rio de Janeiro.

26
Ye, T., Mittal, R., Udaykumar, H. S. and Shyy, W., 1999. An Accurate Cartesian Grid

Method for Viscous Incompressible Flows with Complex immersed Boundaries. Journal

of Computational Physics, 156, 209.

Williamson, C. H. K, 1985. Evolution of a single wake behind a pair of bluff bodies.

Journal of Fluid Mechanics, 159, 1-18.

Figure Captions

Figure 1 Immersed boundary illustration: Lagrangian grid for the interface and

Eulerian grid for the domain.

Figure 2 Illustration of a particle of fluid located over an immersed interface point.

Figure 3 Scheme for the velocity components and pressure interpolations.

Figure 4 Scheme for the interpolation of the horizontal velocity component u at

point 3.

27
Figure 5 Scheme for the interpolation of the vertical velocity component at point

3.

Figure 6 Scheme for the interpolation of the pressure field at point 3.

Figure 7 Interpolation for the velocity components (a) and pressure (b), at point k.

Figure 8 Illustration of the Lagrangian force distribution process to the Eulerian grid.

Figure 9 Geometric parameters of the simulation domain with two cylinders of

different diameters disposed in tandem; global view (a) and closer view (b)

Figure 10 Arrangement in tandem for different cylinders. Sequence of Vorticity

Contours for L = 2.7d; 2.5s (a), 5.0s (b), 10.0s (c) and 21s (d).

Figure 11 Arrangement in tandem for different cylinders. Time evolution of Cd1 and

Cd2 for L = 2.7d

Figure 12 Arrangement in tandem for different cylinders. Vorticity Contours for L =

1.5d (a) and L = 2.5d (b).

Figure 13 The velocity vectors for the flows around two cylinders of different

diameter for L = 1.5d.

28
Figure 14 Arrangement in tandem for different cylinders. Evolution of the drag

coefficient with time for L = 1.5d and for L = 2.5d.

Figure 15 - Arrangement in tandem for different cylinders. Mean drag coefficient as a

function of the gap between cylinders centers. 1 upstream cylinder and 2 downstream

cylinder.

Figure 16 - Arrangement in tandem for different cylinders. Strouhal number as a

function of the gap between the cylinders centers.

Figure 17 Geometric parameters of the simulation domain with two cylinders of same

diameter disposed in tandem; global view (a) and closer view with the Eulerian and

Lagrangian grids (b).

Figure 18 - Arrangement in tandem for cylinders of the same diameter. Vorticity

Contours for L = 4.0d; present work (a), Surmas et al. (2003) (b) and Meneghini et al.

(2001) (c).

Figure 19 Arrangement in tandem for cylinders of the same diameter. Drag coefficient

for both cylinders as a function of the gap.

Figure 20 Arrangement in tandem for cylinders of same diameter. Strouhal Number as

a function of the gap.

29
Figure 21 Geometric parameters of the simulation domain with two cylinders of same

diameter disposed side by side.

Figure 22 - Vorticity Contours and pressure field for L = 1.5d (a), L = 2.0d (b), L =

3.0d (c) and L = 4.0d (d); present work.

Figure 23 - Vorticity Contours for Side by side arrangement and L = 2.0d and 4.0d,

respectively. Surmas et al. (2003) (a) and Meneguini et al. (2001) (b).

Figure 24 Drag coefficient as a function of the gap between the cylinders for the side

by side arrangement.

Figure 25 Lift coefficient as a function of the gap between the cylinders for the side

by side arrangement.

Figure 26 Schematic view of the V Configuration.

Figure 27 Vorticity Field for the V Configuration. = 400 (a), = 600 (b), = 800

(c), = 1000 (d), = 1400 (e) and = 1800 (f).

Figure 28 Flow Visualization for the V Configuration at = 1800. Present work (a)

and experimental result of Fantasy of Flow book, (b).

30
Figure 29 Drag coefficient as a function of time for cylinders 1 and 7 for the V

Configuration. = 400 (a), = 600 (b), = 800 (c), = 1000 (d), = 1400 (e) and =

1800 (f).

Figure 30 Picture of an offshore structure (a) and schematic view of a horizontal

plane of the central tower (b).

Figure 31 Non-uniform grid used for the simulations over a group of bluff bodies.

Figure 32 Streamlines formed by the flow over group of bluff bodies; 10s (a), 20s (b),

30s (c) and 45s (d).

Figure 33 Streamlines formed by the flow over a square; 10s (a), 20s (b), 30s (c), and

45s (d).

Figure 34 Time evolution of the drag coefficient for the geometries represented in

Fig. 30.

31
Tables Legend

Table 1 Mean drag coefficient and Strouhal number for the cylinders 1 and 7 for

different angles.

Table 2 Mean drag coefficients for the bluff bodies of Figure 30(b).

32
Eulerian grid

Lagrangian grid

Fig. 1

33
Fig. 2

34
4

3
1 2


xk

Fig. 3

35
u i-3, j+2 u i-2, j+2 u i-1, j+2 u i, j+2 u i+1, j+2

u i-3, j+1 u i-2, j+1 u i-1, j+1 4 u i, j+1 u i+1, j+1

u i-1, j 3 u i, j u i+1, j

u i, j-1 u i+1, j-1

k 1 2

u i+1, j-2

Fig. 4

36
i-2, j+2 i-1, j+2 i, j+2 i+1, j+2 i+2, j+2

i-2, j+1 i-1, j+1 i, j+1 i+1, j+1 i+2, j+1

i, j i+1, j i+2, j

3
i+1, j-1 i+2, j-1
k 1 2
i+1, j-2 i+2, j-2

Fig. 5

37
pi-2,j+2 pi-1,j+2 pi,j+2 pi+1,j+2 pi+2,j+2

pi-2,j+1 pi-1,j+1 pi,j+1 4 pi+1,j+1 pi+2,j+1

pi,j 3 pi+1,j pi+2,j

pi+1,j-1 pi+2,j-1

k 1 2

pi+1,j-2 pi+2,j-2

Fig. 6

38
vfk

ufk
k

(a)

k x

(b)

Fig. 7

39
Fig. 8

40
(a)

(b)

Fig. 9

41
(a) (b)

(c) (d)

Fig. 10

42
1.5

1.0
Cd1, Cd2

0.5

0.0
Cd1
Cd2
-0.5
50 100 150 200 250 300 350
T
Fig. 11

43
(a)

(b)

Fig. 12

44
Fig. 13

45
0.2

0
L=1.5d
-0.2
Cd2

-0.4
L=2.5d

-0.6

50 100 150 200


T
Fig. 14

46
2

1
Cd1 ,Cd2

Cd1 present work


-1 Cd1 Surmas et al. (2003)
Cd1 Flatschart et al. (2000)
Cd2 present work
Cd2 Surmas et al. (2003)
-2 Cd2 Flatschart et al. (2000)

1.5 2 2.5 3 3.5 4 4.5


L/d
Fig. 15

47
0.25

0.2

0.15
St

0.1
present work
Surmas et al. (2003)
0.05
Flatschart et al. (2000)

0
1.5 2 2.5 3 3.5 4
L/d
Fig. 16

48
(a)

10.5 d

10 d

9.5 d

16 d 16.5 d 17 d 16.5 d + L

(b)

Fig. 17

49
(a)

(b)

(c)

Fig. 18

50
2

0
Cd1 ,Cd2

-1 Cd1 present work


Cd1 Surmas et al. (2003)
Cd1 Meneghini et al. (2001)
Cd2 present work
-2 Cd2 Surmas et al. (2003)
Cd2 Meneghini et al. (2001)

-3
1 2 3 4
L/d
Fig. 19

51
0.25

0.2

0.15
St

0.1

present work
Surmas et al. (2003)
0.05
Meneghini et al. (2001)

0
1 2 3 4
L/d
Fig. 20

52
d

Inlet
L Ly = 25d

16.5 d

Lx = 50 d

Fig. 21

53
(a) (b)

(c) (d)

(e) (f)

(g) (h)

Fig. 22

54
(a)

(b)

Fig. 23

55
1.8

1.6

1.4
Cd

1.2
Cd1 present work
Cd1 Surmas et al. (2003)
Cd1 Meneghini et al. (2001)
1

0.8
1 2 3 4
L/d
Fig. 24

56
1

Cl 0

-1 Cd1 present work


Cd1 Surmas et al. (2003)
Cd1 Meneghini et al. (2001)
Cd2 present work
Cd2 Surmas et al. (2003)
-2 Cd2 Meneghini et al. (2001)

1 2 3 4
L/d

Fig. 25

57
Fig. 26

58
(a) (b) (c)

(d) (e) (f)

Fig. 27

59
(a) (b)

Fig. 28

60
1.6 1.6
= 60
o

single cylinder
1.4 single cylinder
1.4
1.2
7
Cd

Cd
1.2
1

7
0.8 1
1 1
0.6
0.8
100 200 300 400 500 0 100 200 300 400 500 600
tU/d tU/d
(a) (b)
1.8


2

1.6 single cylinder 7 7


1.8

1.6
Cd

1.4
Cd

1.4

1.2 single cylinder 1


1 1.2

1 1
0 100 200 300 400 500 600 0 100 200 300 400 500
tU/d tU/d
(c) (d)
2.6
2.6 1

2.4 2.4
1
2.2 2.2

2 2
Cd

1.8
Cd

1.8
1.6 1.6 7
7
1.4 1.4
single cylinder single cylinder
1.2 1.2
1 1
0 100 200 300 400 500 0 100 200 300 400 500
tU/d tU/d
(e) (f)

61
Fig. 29
0
.1H

(a) (b)

Fig. 30

62
Fig. 31

63
(a) (b)

(c) (d)

Fig. 32

64
(a) (b)

(c) (d)

Fig. 33

65
1.20 2.4
Cd3 Cd1
1.00 Cd11 2.2 Cd13
2.0
0.80
Cd

Cd
1.8
0.60
1.6
0.40
1.4
1000 2000 3000 500 1000 1500
T T
(a) (b)
0.8
Cd5 0.8 Cd6
Cd9 Cd8
0.6
0.6
Cd
Cd

0.4
0.4

0.2
0.2
500 1000 1500 500 1000 1500
T T
(c) (d)
1.1 1.0
Cd17
Cd21
1.0 Cd18
0.8 Cd22
0.9 Cd23
Cd

Cd

0.6
0.8

0.7 0.4

0.6 0.2
500 1000 1500 500 1000 1500
T T
(e) (f)
0.70 2.5
Cd2 Cd7
0.60 Cd12 2.0 Cd15

1.5
0.50
Cd
Cd

1.0
0.40
0.5

0.30 0.0
1000 2000 3000 200 400 600 800
T T
(g) (h)

Fig. 34

66
Table 1

Cd,1 Cd,7 St1 St7

400 0.7323 0.8017 0.0851 0.0849

600 0.9709 1.1173 0.1128 0.1132

800 1.1767 1.3975 0.1194 0.1198

1000 1.4556 1.5358 0.1286 0.1280

1400 1.9752 1.8249 0.2074 0.1857

1800 2.2272 1.9320 0.2294 0.1892

Table 2

67
Body Number Body Number
Cd Cd

1 1.9520 13 1.9560

2 0.5000 14 1.9340

3 0.7600 15 1.7080

4 0.4880 16 1.9280

5 0.3740 17 0.8440

6 0.4780 18 0.8460

7 0.4370 19 0.4340

8 0.4840 20 0.4340

9 0.3840 21 0.5500

10 0.4960 22 0.4240

11 0.7680 23 0.4240

12 0.5000

68

You might also like