You are on page 1of 10

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/254330059

Fluid Flow And Heat Transfer from a Cylinder


Between Parallel Planes

Article in Journal of Thermophysics and Heat Transfer · January 2004


DOI: 10.2514/1.6186

CITATIONS READS

36 102

3 authors, including:

Waqar Khan M. M. Yovanovich


Majmaah University University of Waterloo
159 PUBLICATIONS 3,147 CITATIONS 312 PUBLICATIONS 6,234 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Waqar Khan on 15 April 2014.

The user has requested enhancement of the downloaded file.


JOURNAL OF THERMOPHYSICS AND HEAT TRANSFER
Vol. 18, No. 3, July–September 2004

Fluid Flow and Heat Transfer from a Cylinder


Between Parallel Planes

W. A. Khan,∗ J. R. Culham,† and M. M. Yovanovich‡


University of Waterloo, Waterloo, Ontario N2L 3G1, Canada

An integral approach is employed to investigate the effects of blockage on fluid flow and heat transfer from a
circular cylinder confined between parallel planes. The integral form of the boundary-layer momentum equation
is solved using the modified von Kármán–Pohlhausen method, which uses a fourth-order velocity profile inside the
hydrodynamic boundary layer. The potential flow velocity, outside the boundary layer, is obtained by the method
of images. A third-order temperature profile is used in the thermal boundary layer to solve the energy integral
equation for isothermal and isoflux boundary conditions. Closed-form solutions are obtained for the fluid flow and
heat transfer from the cylinder with blockage ratio and Reynolds and Prandtl numbers as parameters. It is shown
that the blockage ratio controls the fluid flow and the transfer of heat from the cylinder and delays the separation.
The results for both thermal boundary conditions are found to be in a good agreement with experimental/numerical
data for a single circular cylinder in a channel.

Nomenclature α = thermal diffusivity, m2 /s


b = blockage ratio, identical to D/ST δ = hydrodynamic boundary-layer thickness, m
CD = total drag coefficient δT = thermal boundary-layer thickness, m
CDf = friction drag coefficient δ1 = displacement thickness, m
C Dp = pressure drag coefficient δ2 = momentum thickness, m
Cf = skin-friction coefficient ζ = dummy variable
Cp = pressure coefficient η = dimensionless normal distance
cp = specific heat of fluid, J/kg · K θ = angle measured, rad or deg
D = cylinder diameter, m λ = pressure gradient parameter
Df = shear force in flow direction, N µ = absolute viscosity of fluid, kg/m · s
Dp = pressure force in flow direction, N ν = kinematic viscosity of fluid, m2 /s
h̄ = average heat transfer coefficient, W/m2 · K ρ = density of the fluid, kg/m3
j = number of cylinders in transverse direction τ = shear stress, N/m2
k = thermal conductivity, W/m · K ψ, φ = stream and potential functions, m2 /s
NuD = average Nusselt number based on diameter
Subscripts
of cylinder, identical to h̄ D/k f
Pr = Prandtl number, identical to ν/α f = fluid or friction
p = pressure, N/m2 p = pressure
q = heat flux, W/m2 r, θ = plane polar coordinates
Re D = Reynolds number, identical to DU∞ /ν s = separation
r, θ = polar coordinates T = temperature
ST = vertical distance between parallel planes w = wall
s = distance along curved surface of circular cylinder ∞ = freestream conditions
measured from forward stagnation point, m
T = temperature, ◦ C Introduction
U (s) = potential flow velocity just outside boundary layer, m/s
u, v
w(z)
=
=
velocity components in boundary layer, m/s
complex potential in Cartesian coordinates, m2 /s T HE main objective of this study is to investigate analytically
the effects of blockage ratio b = D/ST on the fluid flow and
heat transfer from a cylinder under different thermal boundary con-
x, y = Cartesian coordinates
Y = distance normal to and measured from surface ditions. This parameter plays an important role in determining the
of circular cylinder, m fluid flow and heat transfer from a cylinder confined between parallel
z = complex variable in Cartesian coordinates, m planes. In practice, a cylinder is placed in flows restricted by walls.
This configuration is found in many applications, such as crossflow
heat exchangers, shrouded heat sinks, and electric heating elements
Received 31 October 2003; presented as Paper 2004-493 at the 42nd in boilers. It has been observed by Ž ukauskas1 and others that, as the
Aerospace Sciences Meeting and Exhibit, Reno, NV, 5–8 January 2004; re- blockage ratio increases, the velocity around the circular cylinder
vision received 30 January 2004; accepted for publication 31 January 2004. outside the boundary layer increases and the pressure and velocity
Copyright  c 2004 by the American Institute of Aeronautics and Astronau- distributions inside the boundary layer are changed accordingly. Po-
tics, Inc. All rights reserved. Copies of this paper may be made for personal tential flow velocity outside the boundary layer can be obtained by
or internal use, on condition that the copier pay the $10.00 per-copy fee to the method of images, which is a technique used in potential flow
the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA modeling to represent the presence of a channel wall by creating an
01923; include the code 0887-8722/04 $10.00 in correspondence with the
image of the cylinder.
CCC.
∗ Graduate Research Assistant, Department of Mechanical Engineering, The fluid flow and heat transfer from a cylinder (infinite
Microelectronics Heat Transfer Laboratory. flow conditions) have been studied analytically by Khan et al.,2
† Director, Microelectronics Heat Transfer Laboratory. and experimentally/numerically by many researchers including
‡ Distinguished Professor Emeritus, Department of Mechanical Engineer- Ž ukauskas,1 Roshko,3 Achenbach,4 Schlichting,5 Wieselsberger,6
ing, Microelectronics Heat Transfer Laboratory. Fellow AIAA. Churchill,7 Sucker and Brauer,8 Ž ukauskas and Ž iugžda,9 Eckert
395
396 KHAN, CULHAM, AND YOVANOVICH

and Soehngen,10 Churchill and Bernstein,11 Morgan,12 Hilpert,13 U∞ and the ambient temperature is assumed to be T∞ . The surface
Refai Ahmed and Yovanovich,14 Krall and Eckert,15 Giedt,16 and temperature of the cylinder wall is Tw in the case of the isothermal
Sarma and Sukhatme.17 However, the problem of fluid flow and cylinder and the heat flux is q for the isoflux boundary condition.
heat transfer from a cylinder placed inside a channel has not been The flow is assumed to be laminar, steady, and two dimensional.
studied analytically so far. Using an order-of-magnitude analysis, the reduced equations of
A review of existing literature reveals that few experimental/ continuity, momentum, and energy in the plane polar coordinates
numerical studies exist regarding the configuration of the present for an incompressible fluid can be written as follows:
problem. Perkins and Leppert18 investigated local heat transfer co- the continuity,
efficients from a uniformly heated cylinder with water in crossflow.
They used both potential flow and experimental pressure distribu- ∂u r ur 1 ∂u θ
+ + =0 (1)
tions to investigate and correlate the effects of blockage on the ve- ∂r r r ∂θ
locity and heat transfer distributions.
Ž ukauskas1 analyzed the work of Akilba’yev et al.19 about the the θ momentum,
influence of channel blockage on the flow and heat transfer of a  
∂u θ u θ ∂u θ 1 ∂p ∂ 2uθ 1 ∂u θ uθ
tube in a restricted channel. According to Ž ukauskas,1 they showed ur + =− +ν + − 2 (2)
that increasing channel blockage ratio from 0 to 0.8 caused the min- ∂r r ∂θ rρ ∂θ ∂r 2 r ∂r r
imum pressure point to be displaced from θ = 70 to 90 deg, and
the separation point to move downstream to θ = 100 deg. Their the r momentum,
theoretical calculations, by the method of Merk20 using the po- ∂p
tential flow velocity distribution, showed that the heat transfer on =0 (3)
∂r
the front portion of the tube increases with an increase in block-
age ratio. Ž ukauskas and Ž iugžda9 performed a series of experi- the energy,
ments with different freestream geometries to investigate the effects  
of channel blockage. They expressed the velocity distribution in ∂T uθ ∂ T ∂2T 1 ∂T
ur + =α + (4)
the outer boundary layer in terms of channel blockage and used ∂r r ∂θ ∂r 2 r ∂r
it to estimate the heat transfer behavior of a cylinder. Vaitiekünas
et al.21 investigated numerically the effects of the channel block- These equations can be rewritten by adopting a curvilinear system
age on the dimensionless shear stress, the location of Umax , point of of coordinates in which s denotes distance along the curved surface
boundary-layer detachment, and the local heat transfer coefficients. of the circular cylinder measured from the forward stagnation point
They approximated the velocity distribution outside the boundary A and Y is the distance normal to and measured from the surface
layer by the modified Hiemenz polynomial in which the coefficients as shown in Fig. 2. In this system of coordinates, the velocity com-
are functions of channel blockage. These functions were based on ponents u θ and u r are replaced by u and v in the local s and Y
the analysis of the experimental data of Ž ukauskas and Ž iugžda.9 directions, whereas r dθ and dr are replaced by ds and dY , respec-
They21 found satisfactory agreement with the experimental results tively. The potential flow velocity just outside the boundary layer is
of Ž ukauskas and Ž iugžda. denoted by U (s), which will be determined by the method of im-
Hattori and Takahashi22 performed experiments on forced con- ages. Therefore, the governing equations in this curvilinear system
vection heat transfer from a single row of circular cylinders in cross- will be as follows:
flow. They measured local and average Nusselt numbers for a cylin- the continuity,
der in the Reynolds number range from 80 to 6 × 103 and gave a
correlation for the average Nusselt number. Later Yamamoto and ∂u ∂v
+ =0 (5)
Hattori23 verified numerically their heat transfer values for the same ∂s ∂Y
arrangement. They found good agreement with those obtained from
experiments in water by Hattori and Takahashi.22 the s momentum,
It is obvious from the literature survey that no analytical study ex- ∂u ∂u 1 dp ∂ 2u
ists to give a closed-form solution for the fluid flow and heat transfer u +v =− +ν 2 (6)
from a circular cylinder in a channel for a wide range of blockage ra- ∂s ∂Y ρ ds ∂Y
tios, Reynolds numbers, and Prandtl numbers. In this study, a circu- the Y momentum,
lar cylinder in a channel is considered in crossflow with a Newtonian
fluid (Pr ≥ 0.71) to investigate the effects of the blockage ratio on dp
the fluid flow and heat transfer from the cylinder for a wide range =0 (7)
dY
of parameters, including blockage ratio, Reynolds numbers, and
Prandtl numbers. Closed-form solutions are obtained for the drag
coefficients and Nusselt numbers under different thermal boundary
conditions, which can be used for a wide range of parameters.

Analysis
Consider a uniform flow of a Newtonian fluid (Pr ≥ 0.71) past
a fixed circular cylinder of diameter D, confined between parallel
planes, as shown in Fig. 1. The approaching velocity of the fluid is

Fig. 1 Physical model and coordinate system. Fig. 2 Flow over a circular cylinder.
KHAN, CULHAM, AND YOVANOVICH 397

the energy,
∂T ∂T ∂2T
u +v =α 2 (8)
∂s ∂Y ∂Y
the Bernoulli equation,
1 dp dU (s)
− = U (s) (9)
ρ ds ds

Hydrodynamic Boundary Conditions


At the cylinder surface, that is, at Y = 0, Fig. 3 Transverse row of doublets
or circular cylinders.
∂ u
2
1 ∂p
u = 0, = (10)
∂Y 2 µ ∂s
At the edge of the boundary layer, that is, at Y = δ(s),
∂u ∂ 2u
u = U (s), = 0, =0 (11)
∂Y ∂Y 2

Thermal Boundary Conditions


The boundary conditions for the uniform wall temperature (UWT)
and uniform wall flux (UWF) are

T = Tw for UWT
When this value is substituted into Eq. (17), the required complex
Y = 0, ∂T q potential will be
 =− for UWF
∂Y kf (12)
w(z) = φ + iψ = U∞ [z + C coth(π z/ST )] (20)
∂ T
2
Y = 0, =0 (13) where φ and ψ are the potential and stream functions and C is a
∂Y 2 constant given by
∂T C = (ST /π) sinh2 (π D/2ST ) (21)
Y = δT , T = T∞ , =0 (14)
∂Y
The stream function ψ in polar coordinates (r, θ) can be obtained
Velocity Distribution Inside Boundary Layer from Eq. (20) as follows:
Assuming a thin boundary layer around the cylinder, the velocity ψ = U∞ r sin θ
distribution in the boundary layer can be approximated by a fourth-

order polynomial as suggested by Pohlhausen,24 sin(2πr sin θ/ST )
− C U∞ (22)
u/U (s) = (2η − 2η3 + η4 ) + (λ/6)(η − 3η2 + 3η3 − η4 ) (15) cosh(2πr cos θ/ST ) − cos(2πr sin θ/ST )
The radial and transverse components of velocity at the surface of
where 0 ≤ η = Y/δ(s) ≤ 1 and λ is the pressure gradient parameter
the cylinder can be written as
given by
 
δ 2 dU (s) 1 ∂ψ  ∂ψ 
ur = − , uθ = (23)
λ=
ν ds
(16) r ∂θ r = D/2 ∂r r = D/2

With the help of velocity profiles, Schlichting5 showed that the pa- which gives
rameter λ is restricted to the range −12 ≤ λ ≤ 12.
u r = 0, u θ = U∞ f (θ) (24)
Velocity Distribution Outside Boundary Layer
where
When the method of images is used, a cylinder confined between
parallel planes (Fig. 1) can be modeled as a system of infinite trans- f (θ) = sin θ
verse row of doublets superimposed on a uniform flowfield (Fig. 3).  
Kochin et al.25 and Perkins and Leppert18 pointed out this case. For 2πC cos(π D cos θ/ST ) sin θ

this case, the complex potential can be written as ST cosh(π D cos θ/ST ) − cos(π D sin θ/ST )

∞  
µ π D sin θ
w(z) = U∞ z + + sin
j = −∞
2π(z − i j ST ) ST
 
µ πz sinh(π D cos θ/ST ) cos θ + sin θ sin(π D sin θ/ST )
= U∞ z + coth (17) × (25)
2ST ST [cosh(π D cos θ/ST ) − cos(π D sin θ/ST )]2
Setting C1 = π D/ST and substituting the value of C, we get
where j is the number of doublets or cylinders. Therefore, the com-  
plex velocity will be C1 cos(C1 cos θ) sin θ
 
 f (θ) = sin θ − 2 sinh 2
2 cosh(C1 cos θ) − cos(C1 sin θ )
w  (z) = U∞ − µπ 2ST 2
1 sinh2 (π z/ST ) (18)

At the stagnation points, z = ±D/2 and w (±D/2) = 0. Therefore, sinh(C1 cos θ) cos θ + sin θ sin(C1 sin θ )
+ sin(C1 sin θ) ×
Eq. (18) gives [cosh(C1 cos θ) − cos(C1 sin θ)]2
µ/2ST = (U∞ ST /π) sinh2 (π D/2ST ) (19) (26)
398 KHAN, CULHAM, AND YOVANOVICH

The resultant potential flow velocity is When Eq. (36) is solved with Eq. (37), the local dimensionless
momentum thickness can be written as
U = U∞ f (θ) (27) 
 θ
δ2 0.485 1
Temperature Distribution =√ f 5 (ζ ) dζ (41)
D Re D f 6 (θ) 0
Assuming a thin thermal boundary layer around the cylinder,
the temperature distribution in the thermal boundary layer can be The local dimensionless boundary-layer thickness can be obtained
approximated by a third-order polynomial, from Eq. (16),
(T − T∞ )/(Tw − T∞ ) = A + BηT + CηT2 + DηT3 (28) 
δ/D = λ/2 Re D g(θ) (42)
where ηT = Y /δT (s). When the aforementioned thermal boundary where g(θ) is the first derivative of f (θ) with respect to θ obtained
conditions are used, the temperature distribution will be from Eq. (26). When Eqs. (35) and (42) are solved and the results
are compared with Eq. (41), the values of the pressure gradient pa-
(T − T∞ )/(Tw − T∞ ) = 1 − 32 ηT + 12 ηT3 (29)
rameter λ are obtained corresponding to each position along the
cylinder surface. These values are positive from 0 ≤ θ ≤ θ1 = π/2,
for the isothermal boundary condition and for example, region 1, and negative from θ1 = π/2 ≤ θ ≤ θs , for ex-
 ample, region 2, as shown in Fig. 2. Thus, the whole range of interest
T − T∞ = (2qδT /3k f ) 1 − 32 ηT + 12 ηT3 (30) 0 ≤ θ ≤ θs can be divided into two regions, and the λ values can be
fitted separately for each blockage ratio by the least-squares method
for the isoflux boundary condition. into two polynomials. These polynomials can be used to determine
the drag and the local heat transfer coefficients in both regions.
Boundary-Layer Parameters
In dimensionless form, the momentum integral equation can be Fluid Flow
written as The first parameter of interest is fluid friction, which manifests
  
dU δ2 ∂u 
itself in the form of the drag force FD , where FD is the sum of the
U δ2 dδ2 δ1 δ22
+ 2+ = (31) skin-friction drag D f and pressure drag D p . Skin-friction drag is
ν ds δ2 ν ds U ∂Y Y = 0 due to viscous shear forces produced at the cylinder surface pre-
dominantely in those regions where the boundary layer is attached.
where The component of shear force in the flow direction is given by
 1 
u D
δ1 = δ 1− dη (32) Df = τw sin θ dθ (43)
0 U A 2

1
u u where τw is the shear stress along the cylinder wall, which can be
δ2 = δ 1− dη (33) determined from Newton’s law of viscosity,
U U
0

∂u 
When velocity distribution from Eq. (15) is used, Eqs. (32) and (33) τw = µ  (44)
can be written as ∂y y =0

δ1 = (δ/10)(3 − λ/12) (34) In dimensionless form, it can be written as



√ 
δ2 = (δ/63)(37/5 − λ/15 − λ2 /144) (35) C f = τw 12 ρU∞2
= 13 (λ + 12) Re D f (θ) 2g(θ)/λ (45)

assuming The angle of separation depends only on the velocity distribution


outside the boundary layer. Khan et al.2 have shown that, for infinite
δ22 dU flow conditions, separation is calculated to occur at θs = 107.71 deg.
Z= , K=Z From Eq. (45) it can be shown that the angle of separation depends
ν ds
on the blockage ratio. The friction drag coefficient is defined as
Equation (31) can be reduced to a nonlinear differential equation of  π
the first order for Z , which is given by CDf = C f sin θ dθ
0
dZ H (K )
= (36)  θs  π
ds U
= C f sin θ dθ + C f sin θ dθ (46)
where H (K ) = 2 f 2 (K ) − 2K [2 + f 1 (K )] is a universal function 0 θs
and is approximated by Walz26 using a straight line,
Because no shear stress acts on the cylinder surface after boundary-
H (K ) = 0.47 − 6K (37) layer separation, the second integral will be zero and the friction
drag coefficient can be written as
with  θs
CDf = C f sin θ dθ
63(3 − λ/12)
f 1 (K ) = (38) 0
10(37/5 − λ/15 − λ2 /144)  
θ1
   1 2g(θ)
1 λ 37 λ λ 2 = √ (λ1 + 12) f (θ) sin θ dθ
f 2 (K ) = 2+ − − (39) 3 Re D 0 λ1
63 6 5 15 144
  
 2 θs
2g(θ)
λ 37 λ λ2 + (λ2 + 12) f (θ) sin θ dθ (47)
K= − − (40) λ2
3963 5 15 144 θ1
KHAN, CULHAM, AND YOVANOVICH 399

The drag coefficient C D f is calculated for different blockage ratios for region 1, and
and correlated into a single expression:
d 45D 2
[45.72 − 39.9 exp(−0.95b3.44 )] δT [ f (θ)δT ζ (λ2 + 12)] = (58)
CDf = √ (48) dθ Pr Re D
Re D
for region 2. These two equations can be solved separately in the
Pressure drag is due to the unbalanced pressures that exist between two regions for the local thermal boundary layer thicknesses,
the relatively high pressures on the upstream surfaces and the lower 
pressures on the downstream surfaces. The component of pressure 
force in the flow direction is given by δT1 1 3 90 f 1 (θ) λ1
= (59)
 D 1
2
1 (λ1 + 12)2 f 2 (θ) 2g(θ)
D Re D Pr 3

Dp = p cos θ dθ (49) 
A 2 
δT2 1 3 90 f 3 (θ) λ2
In dimensionless form, it can be written as = (60)
D 1
2
1 f 2 (θ) 2g(θ)
 Re D Pr 3
π
C Dp = C p cos θ dθ (50) where functions f 1 (θ) and f 3 (θ) are given by
0
 θ
where C p is the pressure coefficient and is defined as
f 1 (θ) = f (θ)(λ1 + 12) dθ (61)
1 0
C p = p ρU∞
2
(51)
2  θ
The pressure difference p is obtained by integrating Eq. (2) with f 2 (θ) = f (θ)(λ2 + 12) dθ (62)
respect to θ. Then the pressure drag coefficient is calculated, using θ1
Eq. (50), for different blockage ratios and correlated into a single
expression: f 1 (θ) f 2 (θ)
f 3 (θ) = + (63)
λ1 + 12 λ2 + 12
C Dp = {6.1 − 4.95 exp(−0.76b 2.63
)}
For the isothermal boundary condition, the local heat transfer coef-
1.49 − 0.23 exp(−5.81b2.15 ) ficient is defined as follows:
+ (52)
Re D k f (∂ T /∂Y )|Y = 0 3k f
h(θ) = − = (64)
The total drag coefficient C D can be written as the sum of both drag Tw − T∞ 2δT
coefficients:
Thus, local heat transfer coefficients for both regions are written as
[45.72 − 39.9 exp(−0.95b3.44 )]
CD = √  
Re D h 1 (θ) = 3k f 2δT1 , h 2 (θ) = 3k f 2δT2 (65)
+ [6.1 − 4.95 exp(−0.76b 2.63
)] which give the local Nusselt numbers for isothermal boundary
condition:
[1.49 − 0.23 exp(−5.81b 2.15
)] 
+ (53) 
Re D
3 3 (λ1 + 12)2 f 2 (θ) 2g(θ) 12 1
N u D1 (θ) = Re D Pr 3 (66)
Heat Transfer 2 90 f 1 (θ) λ1
The second parameter of interest in this study is the dimension-  
less average heat transfer coefficient N u D for large Prandtl numbers. 3 f 2 (θ) 2g(θ) 12 1
3
This parameter is determined by integrating Eq. (8) from the cylin- N u D2 (θ) = Re D Pr 3 (67)
der surface to the thermal boundary-layer edge. When the presence 2 90 f 3 (θ) λ2
of a thin thermal boundary layer δT along the cylinder surface is
assumed, the energy integral equation for the isothermal boundary The average heat transfer coefficient is defined as
condition can be written as  π
  h̄ =
1
∂ T 
δT h(θ) dθ
d π
(T − T∞ )u dY = −α (54)
∂Y 
0
ds 0 Y =0  
θs π
1
When velocity and temperature profiles are used, and ζ = δT / = h(θ) dθ + h(θ) dθ (68)
π θs
δ < 1 is assumed, Eq. (54) can be simplified to 0

d It has been observed experimentally by many researchers that, at low


δT [U (s)δT ζ (λ + 12)] = 90α (55) Reynolds numbers (up to Re D = 5000, according to Ž ukauskas and
ds
Ž iugžda9 ), there is no appreciable increase in the local heat transfer
Because s = (D/2)θ ⇒ ds = (D/2) dθ and U (s) = U∞ f (θ), after the separation point. However, at high Reynolds numbers, the
therefore, local heat transfer increases from the separation point to the rear
stagnation point. Hence, the average heat transfer coefficient can be
d 45D 2 written as
δT [ f (θ)δT ζ (λ + 12)] = (56)
dθ Pr Re D  θs
1
This equation can be rewritten separately for the two regions h̄ = h(θ) dθ
π 0
mentioned earlier, that is,
 θ1  θs
d 45D 2 1
δT [ f (θ)δT ζ (λ1 + 12)] = (57) = h 1 (θ) dθ + h 2 (θ) dθ (69)
dθ Pr Re D π 0 θ1
400 KHAN, CULHAM, AND YOVANOVICH

When Eqs. (59–65) are used, Eq. (69) can be solved for the average Results and Discussion
heat transfer coefficient that gives the average Nusselt number for Flow Characteristics
an isothermal cylinder: The effects of the blockage ratio b on the velocity distribution
1 1
outside the boundary layer are shown in Fig. 4. It shows that as
N u D |isothermal = [0.843 − 0.25 exp(−2.65b2.5 )] Re D Pr 2 3 (70) the blockage ratio decreases the velocity outside the boundary layer
decreases. These results are compared with the experimental data
For the isoflux boundary condition, the energy integral equation can of Akilba’yev et al.19 (reported by Ž ukauskas1 ) for two blockage
be written as ratios. A good agreement between potential theory and experiment
 δT
is observed for the front part of the cylinder where laminar boundary
d q layer exists.
(T − T∞ )u dY = (71) √
ds 0 ρc p The dimensionless local shear stress, C f (Re D ), is plotted in
Fig. 5 for b = 2.5. It shows that C f is zero at the stagnation point
For constant heat flux and thermophysical properties, Eq. (71) can and reaches a maximum at θ ≈ 60 deg. The increase in shear stress
be simplified to is caused by the deformation of the velocity profiles in the boundary
layer, a higher velocity gradient at the wall, and a thicker boundary
d
45D 2 layer. In the region of decreasing C f preceding the separation point,
f (θ)δT2 ζ (λ + 12) = (72) the pressure gradient decreases further and finally C f falls to zero,
dθ Pr Re D
where boundary-layer separation occurs. Beyond this point, C f re-
Rewriting Eq. (72) for the two regions in the same way as Eq. (56), mains close to zero up to the rear stagnation point. These results
one obtains the dimensionless local thermal boundary layer thick- are compared with the numerical results of Vaitiekünas et al.21 for
nesses δT1 /D and δT2 /D under isoflux boundary condition, the same blockage ratio. The results are again in good agreement
for the front part of the cylinder. In Fig. 6, it can be seen that the
  angle of separation depends on the blockage ratio. As the blockage
δT1 1 3 45θ λ1 ratio increases, the location of the boundary-layer separation moves
= (73)
D 1
2
1 (λ1 + 12) f (θ) 2g(θ) backward. This movement is due to the change in the velocity distri-
Re D Pr 3
bution outside the boundary layer. In Fig. 7, the effects of blockage
  ratio on the drag coefficient can be seen for different Reynolds num-
δT2 1 3 45 f 4 (θ) λ2 bers. It is clear that the drag coefficient decreases with the blockage
= (74)
D 1
2
1 f (θ) 2g(θ)
Re D Pr 3

with

f 4 (θ ) = θ/(λ1 + 12) + (θ − π/2)/(λ2 + 12) (75)

The local surface temperatures for the two regions is obtained from
Eq. (30),

T1 (θ) = 2qδT1 3k f (76)

T2 (θ) = 2qδT2 3k f (77)

The local heat transfer coefficient is obtained from its definition as

h 1 (θ ) = q/T1 (θ), h 2 (θ) = q/T2 (θ) (78)

which will give local Nusselt numbers for isoflux boundary condi-
tion as follows:
  Fig. 4 Effect of blockage ratio on velocity distribution.
3 3 (λ1 + 12) f (θ) 2g(θ) 12 1
N u D1 (θ ) = Re D Pr 3 (79)
2 45θ λ1
 
3 3 f (θ) 2g(θ) 12 1
N u D2 (θ ) = Re D Pr 3 (80)
2 45 f 4 (θ) λ2

Following the same procedure for the average heat transfer coeffi-
cient as described earlier, one obtains the average Nusselt number
for an isoflux cylinder as
1 1
N u D |isoflux = [1.104 − 0.47 exp(−1.54b2.77 )]Re D2 Pr 3 (81)

Combining the results for both thermal boundary conditions, we


have

NuD 0.843 − 0.25 exp(−2.65b2.5 ) for UWT
=
1
2
Re D Pr
1
3
1.104 − 0.47 exp(−1.54b 2.77
) for UWF
Fig. 5 Distribution of dimensionless shear stress for given blockage
(82) ratio.
KHAN, CULHAM, AND YOVANOVICH 401

Fig. 8 Drag coefficient as a function of Reynolds number ReD for dif-


ferent blockage ratios.

Fig. 6 Effect of blockage ratio on angle of separation.

Fig. 9 Local Nusselt number for two thermal boundary conditions.

Fig. 7 Effect of blockage ratio and Reynolds number on drag


coefficient.

ratio. The variation of the drag coefficient C D with Reynolds num-


ber Re D for different blockage ratios is shown in Fig. 8. In this case,
the drag coefficient decreases with the increase in Reynolds number
Re D for a specific blockage ratio. The present results are compared
with the experimental results of Wieselsberger6 for infinite flow con-
ditions. They are in good agreement except at Re D = 2 × 103 , where
a downward deviation (23.75%) in the experimental results was no-
ticed. No physical explanation could be found in the literature for
this deviation.

Heat Transfer Characteristics


The comparison of local Nusselt numbers for the isothermal and Fig. 10 Average Nusselt number for isoflux cylinder.
isoflux boundary conditions for a given blockage ratio is shown in
Fig. 9. The isoflux boundary condition gives a higher heat transfer
coefficient over the larger part of the circumference. On the front part of Hattori and Takahashi22 and Yamamoto and Hattori.23 It is clear
of the cylinder (up to θ ≈ 40 deg), there is no appreciable effect of that the present results are in very good agreement with the previous
boundary condition. Higher heat transfer coefficients have also been work for a given range of Reynolds numbers.
observed experimentally by Perkins and Leppert18 for a blockage It is clear from Fig. 11 that the heat transfer values are higher
ratio of 0.41 with the isoflux boundary condition. for the smaller blockage ratios, they decrease as the blockage ra-
The results for average heat transfer from a single isoflux cylinder tio decreases, and finally they approach to the values for an infi-
are shown in Fig. 10 for a given blockage ratio and Prandtl number, nite cylinder. The heat transfer values increase with the Reynolds
where they are compared with the experimental and numerical data number. The effects of the blockage ratios on the heat transfer
402 KHAN, CULHAM, AND YOVANOVICH

1/2
parameter N u D /Re D Pr 1/3 for the two thermal boundary condi-
tions are shown in Fig. 12. Figure 12 shows that the heat transfer
rates are higher for the isoflux boundary condition. Heat transfer
rates, for both thermal boundary conditions, decrease first with the
blockage ratio and then become constant.
The average Nusselt numbers for the isothermal cylinder for a
given blockage ratio are compared in Fig. 13 with the experimental
results of Niggeschmidt27 (reported by Hausen28 ) and Hausen.28 The
average Nusselt number N u D values are found to be in a good agree-
ment with both results. However, both previous results are found to
be higher at high Reynolds number due to freestream turbulence.

Summary
The influence of blockage ratio on the fluid flow and heat trans-
fer from a circular cylinder, placed between two parallel planes,
has been investigated. The method of images is used to obtain the
velocity distribution outside the boundary layer. Three correlations
are obtained, Eq. (53) for total drag coefficient, Eq. (70) for heat
transfer from an isothermal cylinder, and Eq. (81) for heat trans-
fer from a cylinder under the isoflux boundary condition. These
correlations can be used to determine the drag coefficient and the
dimensionless heat transfer coefficient from a cylinder confined in
Fig. 11 Effect of blockage ratio and Reynolds number on average
Nusselt numbers.
a channel with different blockage ratios. The present results indi-
cate good agreement with the experimental/numerical results for
a wide range of blockage ratio, Reynolds numbers, and Prandtl
numbers.
Acknowledgments
The authors gratefully acknowledge the financial support of Nat-
ural Sciences and Engineering Research Council of Canada and the
Centre for Microelectronics Assembly and Packaging.
References
1 Ž ukauskas, A., Advances in Heat Transfer, Academic Press, New York,
1972, pp. 93–160.
2 Khan, W. A., Culham, J. R., and Yovanovich, M. M., “Fluid Flow and
Heat Transfer from a Pin-Fin: Analytical Approach,” AIAA Paper 2003-163,
Jan. 2003.
3 Roshko, A., “Experiments on the Flow Past Circular Cylinders at Very
High Reynolds Number,” Journal of Fluid Mechanics, Vol. 10, No. 3, 1961,
pp. 345–356.
4 Achenbach, E., “Total and Local Heat Transfer From a Smooth Circular
Cylinder in Cross Flow at High Reynolds Number,” International Journal
of Heat and Mass Transfer, Vol. 18, No. 12, 1975, pp. 1387–1396.
5 Schlichting, H., Boundary Layer Theory, 7th ed., McGraw–Hill, New
York, 1979, Chap. 10.
6 Wieselsberger, C., “New Data on The Laws of Fluid Resistance,” NACA
TN 84, March 1922.
7 Churchill, S. W., Viscous Flows: The Practical Use of Theory,
Fig. 12 Effect of blockage ratio and thermal boundary condition on Butterworths Ser. in Chemical Engineering, Butterworths, Boston, 1988,
heat transfer. pp. 317–358.
8 Sucker, D., and Brauer, H., “Investigation of the Flow Around Trans-
verse Cylinders,” Wärme- und Stoffübertragung, Vol. 8, No. 3, 1975,
pp. 149–158.
9 Ž ukauskas, A., and Ž iugžda, J., Heat Transfer of a Cylinder in Crossflow,
Hemisphere, New York, 1985.
10 Eckert, E. R. G., and Soehngen, E., “Distribution of Heat-Transfer
Coefficients Around Circular Cylinders in Cross Flow at Reynolds Num-
bers from 20 to 500,” Transactions of the ASME, Vol. 74, April 1952,
pp. 343–347.
11 Churchill, S. W., and Bernstein, M., “A Correlating Equation for Forced
Convection from Gases and Liquids to a Circular Cylinder in Cross Flow,”
Journal of Heat Transfer, Vol. 99, No. 2, 1977, pp. 300–306.
12 Morgan, V. T., “The Overall Convective Heat Transfer from Smooth
Circular Cylinders,” Advances in Heat Transfer, Vol. 11, Academic Press,
New York, 1975, pp. 199–264.
13 Hilpert, R., “ Ẅ armeabgabe von geheizten Drahten und Rohren,”
Forsch. Geb. Ingenieurwes, Vol. 4, 1933, pp. 215–224.
14 Refai-Ahmed, G., and Yovanovich, M. M., “Analytical Method for
Forced Convection from Flat Plates, Circular Cylinders, and Spheres,”
Journal of Thermophysics and Heat Transfer, Vol. 9, No. 3, 1995,
pp. 516–523.
15 Krall, K. M., and Eckert, E. R. G., “Heat Transfer to a Transverse
Circular Cylinder at Low Reynolds Number Including Refraction Effects,”
Fig. 13 Average Nusselt number for isothermal boundary condition. Heat Transfer, Vol. 3, 1970, pp. 225–232.
KHAN, CULHAM, AND YOVANOVICH 403

16 Giedt, W. H., “Investigation of Variation of Point Unit Heat-Transfer 22 Hattori, N., and Takahashi, T., “Heat Transfer from a Single Row of Cir-
Coefficient Around a Cylinder Normal to an Air Stream,” Transactions of cular Cylinders Placed in the Transverse Direction of Water Flow,” Transac-
the ASME, Vol. 71, May 1949, pp. 375–381. tions of the Japan Society of Mechanical Engineers, Pt. B, Vol. 59, No. 568,
17 Sarma, T. S., and Sukhatme, S. P., “Local Heat Transfer from a Hor- 1993, pp. 4064–4068.
izontal Cylinder to Air in Cross Flow: Influence of Free Convection and 23 Yamamoto, H., and Hattori, N., “Flow and Heat Transfer Around a Sin-
Free Stream Turbulence,” International Journal of Heat and Mass Transfer, gle Row of Circular Cylinders,” Heat Transfer—Japanese Research, Vol. 25,
Vol. 20, No. 1, 1977, pp. 51–56. No. 3, 1996, pp. 192–200.
18 Perkins, H. C., and Leppert, G., “Local Heat Transfer Coefficients on 24 Pohlhausen, K., “Zur Näherungsweise Integration der Differential
a Uniformly Heated Cylinder,” International Journal of Heat and Mass Gleichung der Laminaren Reibungschicht,” Zeitschrift für angewandte
Transfer, Vol. 7, No. 2, 1964, pp. 143–158. Mathematic und Mechanic, Vol. 1, 1921, pp. 252–268.
19 Akilba’yev, Z. S., Isata’yev, S. I., Krashtalev, P. A., and Masle’yeva, 25 Kochin, N. E., Kibel, I. A., and Roze, N. V., Theoretical Hydromechan-
N. V., “The Effect of Channel Blockage on the Local Heat Transfer Coef- ics (translated from 5th Russian edition), Interscience, New York, 1964,
ficient of a Uniformly Heated Cylinder,” Problemy Teploenergetyki i Prik- Chap. 5.
ladnoi Teplofiziki, Vol. 3, 1966, pp. 179–198. 26 Walz, A., “Ein neuer Ansatz für das Greschwindligkeitsprofil der lam-
20 Merk, H. J., “Rapid Calculations for Boundary Layer Transfer Using inaren Reibungsschicht,” Lilienthal-Bericht, Vol. 141, 1941, p. 8.
Wedge Solutions and Asymptotic Expansions,” Journal of Fluid Mechanics, 27 Niggeschmidt, W., “Druckverlust und Warmeubergang bei fluchtenden,
Vol. 5, 1959, pp. 460–480. versetzten und teilversetzten querangestromten Rohrbundeln,” Dissertation,
21 Vaitiekünas, P. P., Bulota, A. J., and J. J., “Analysis of the Effect of Duct Darmstadt, Germany, 1975.
Blocking on Crossflow and Heat Transfer of a Cylinder,” Heat Transfer— 28 Hausen, H., Heat Transfer in Counterflow, Parallel Flow and Cross
Soviet Research, Vol. 17, No. 4, 1985, pp. 79–86. Flow, McGraw–Hill, New York, 1983, Chap. 2.

View publication stats

You might also like