Вы находитесь на странице: 1из 24

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

net/publication/318039452

Forced Convection Heat Transfer

Chapter · May 2017


DOI: 10.1007/978-3-319-53829-7_9

CITATIONS READS
0 6,470

1 author:

Bahman Zohuri
University of New Mexico
414 PUBLICATIONS   799 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Business Resilience System (BRS): Driven Through Boolean, Fuzzy Logics and Cloud Computation: Real and Near Real Time Analysis and Decision Making System View
project

Our Daily life dependency driven by renewable and nonrenewable source of energy. View project

All content following this page was uploaded by Bahman Zohuri on 07 September 2018.

The user has requested enhancement of the downloaded file.


Chapter 9
Forced Convection Heat Transfer

Convection is the term used for heat transfer mechanism, which takes place in a
fluid because of a combination of conduction due to the molecular interactions and
energy transport due to the macroscopic (bulk) motion of the fluid itself. In the
above definition, the motion of the fluid is essential otherwise, the heat transfer
mechanism becomes a static conduction situation. When the term of convection is
used, usually a solid surface is present next to the fluid. There are also cases of
convection where only fluids are present, such as a hot jet entering into a cold
reservoir. However, the most of the industrial applications involve a hot or cold
surface transferring heat to the fluid or receiving heat from the fluid.

9.1 Introduction

Bear in your mind as we said in previous chapters, if the fluid motion is, sustained
by a difference of pressure created by an external device such as a pump or fan, the
term of “forced convection” is used. On the other hand, if the fluid motion is
predominantly sustained by the presence of a thermally induced density gradient,
then the term of “natural convection” is used.
Heat transfer by convection occurs as the result of a moving fluid encountering a
fixed surface. The moving fluid carries the heat and deposits it on the surface or
draws it out of the surface. There are two types of convection. In forced convection,
the fluid is being driven or forced along by some mechanism other than thermal
gradients at the surface. In free convection, the fluid is moved along by thermal
gradients or temperature differences at the surface. Convection obeys Newton’s law
of cooling given by

Q ¼ hAðT 1  T w Þ ð9:1aÞ

© Springer International Publishing AG 2017 323


B. Zohuri, Thermal-Hydraulic Analysis of Nuclear Reactors,
DOI 10.1007/978-3-319-53829-7_9

bahmanz@aol.com
324 9 Forced Convection Heat Transfer

q ¼ hð T 1  T w Þ ð9:1bÞ

q in this case is the heat flux per unit area at the wall. The symbol h is identified as
the film heat transfer coefficient. It has units of W/m2 K or Btu/h/ft2/R. Where k in
Eq. 3.30b, the thermal conductivity, is a function of only the material and its
temperature, h, the film heat transfer coefficient, depends on the properties of the
fluid, the temperature of the fluid, and the flow characteristics. Multiple correlations
have been determined for calculating an appropriate h for most materials and flow
situations. In Eqs. 9.1a and 9.1b, the wall temperature is designated by Tw and T1 is
the temperature of fluid far from the wall at free-stream condition.
To understand better the heat exchange between a solid and fluid, consider a
heated wall over which a fluid flows as sketched in Fig. 9.1. In this figure, U1 is the
velocity of the fluid under free-stream condition and far away from the wall as well.
For a given stream velocity, the velocity of the fluid decreases as we get closer to
the wall. This is due to the viscous effects of the flowing fluid. On the wall, because
of the adherence (nonslip) condition the velocity of the fluid is zero. The region in
which the velocity of the fluid varies from the free-stream value to zero is called
“velocity boundary layer.” Similarly, the region in which the fluid temperature
varies from its free-stream value to that on the wall is called the “thermal boundary
layer,” and both these boundary layers are defined in previous chapters. Since the
velocity of the fluid at the wall is zero, the heat must be transferred by conduction at
that point. Thus, we calculate the heat transfer by using the Fourier’s heat

t∞

3
1 2 y

y VELOCITY TEMPERATURE
U∞
PROFILES PROFILES

1
2
3

tw t
u
u

q'' HEATED WALL

Fig. 9.1 Convection heat transfer to a flow over a heated wall

bahmanz@aol.com
9.1 Introduction 325

conduction law (Eqs. 3.30a or 3.30b), with thermal conductivity of the fluid
corresponding to the wall temperature and the fluid temperature gradient at the wall.
The question at this point is that: since the heat flows by conduction in this layer,
why do we speak of convection heat transfer and need to consider the velocity of the
fluid. The short answer to this question is that the temperature gradient of the fluid
on the wall is highly dependent on the flow velocity of the free-stream. As this
velocity increases, the distance from the wall we travel to reach fret stream
temperature decreases. In other words, the thickness of velocity and thermal
boundary layers on the wall decreases. The consequence of this decrease is to
increase the temperature gradient of the fluid at the wall, i.e., an increase in the rate
of heat transferred from the wall to the fluid. The effect of increasing fret stream
velocity on the fluid velocity and temperature profiles close to the wall is illustrated
in Fig. 9.1. Note also that the temperature gradient of the fluid on the wall increases
with increasing free-stream velocity. Newton’s experiments end up finding the heat
flux on the wall is based on Eqs. 9.1a and 9.1b.
Table 9.1 gives the orders of magnitude of convective heat transfer coefficients.
We need to remind you that most flow that are occurring in practical applications
are turbulent. As we know so far, that turbulent flow characterization is based on
disorderly displacement of individual volumes of fluid within the flow.
From the above discussion, we conclude that the basic laws of heat conduction
must be coupled with those of fluid motion to describe, mathematically, the process
of convection. The mathematical treatment of the resulting system of differential
equations is very complex. Therefore, for engineering applications, the convection
will be treated by an ingenious combination of mathematical techniques, empirical
evidence, and experimentation.
Velocity, temperature, pressure, and other properties change continuously in
time at every point of turbulent flow and the governing equations of mass, energy,
and momentum that were describe in Chap. 5 applies here and are valid for
turbulent flow as well as laminar flow even in transient mode. Therefore, we have
to take the note of the fact, that all the quantities such as velocity, pressure, and
temperature in these equations are instantaneous values. In this chapter for time
being, we concentrate on problems of heat transfer related to laminar forced
convection flow in pipes and ducts.

Table 9.1 Order of magnitude of convective heat transfer coefficients


Fluid and flow conduction h W/m2 K
Air, free convection 5–25
Water, free convection 15–100
Air or superheated steam, forced convection 30–300
Oil, forced convection 60–1800
Water, forced convection 300–15,000
Liquid sodium, forced convection 10,000–100,000
Boiling water 3000–60,000
Condensing steam 3000–100,000

bahmanz@aol.com
326 9 Forced Convection Heat Transfer

9.2 Heat Transfer in Laminar Tube Flow

In almost every operation of fluid transport phenomena where we need to


understand heat and mass transfer, we need to get involved with flow of a viscous
fluid in some form of closed conduit or tube. Then, understanding of heat transfer in
such internal flow requires complete knowledge of the behavior and mechanism of
the flow of fluids in pipes or tubes.
A common situation encountered by the chemical engineer is heat transfer to
fluid flowing through a tube. This can occur in heat exchangers, boilers, condensers,
evaporators, and a host of other process equipment. Therefore, it is useful to know
how to estimate heat transfer coefficients in this situation.
We can classify the flow of a fluid in a straight circular tube into either laminar or
turbulent flow. It is assumed from hereon that we assume fully developed incom-
pressible, Newtonian, steady flow conditions. Fully developed flow implies that the
tube is long compared with the entrance length in which the velocity distribution at
the inlet adjusts itself to the geometry and no longer changes with distance along
the tube.
If we consider the case of a constant wall heat flux, then the energy equation in
transient mode using cylindrical coordinates can be written as
"   #
2
∂T ∂T ∂T 1∂ ∂T ∂ T
þu þυ ¼α r þ 2 ð9:2Þ
∂t ∂r ∂z r ∂r ∂r ∂z

where u and υ are two components of velocity in r and z direction, respectively.


Previously, we have already solved for the velocity distribution in laminar tube
flow and know that the solution is given by

υ  r 2
¼1 for u ¼ 0 ð9:3Þ
υmax R

or with υaverage ¼ 0.5υmax, Eq. 9.3 can be written as follows:


  r 2 
υ
¼2 1 for u¼0 ð9:4Þ
υmax R

In two lateral equations, R is the internal radius of tube, and these velocities can
be combined with energy equation to yield
 " #
 r 2  ∂T α 1∂

∂T
 2
∂ T
2 1 ¼ r þ 2 ð9:5Þ
R ∂z υaverage r ∂r ∂r ∂z

bahmanz@aol.com
9.2 Heat Transfer in Laminar Tube Flow 327

Next, we assume that the heat transfer is fully developed. In this case, with
constant wall heat flux, we would expect the temperature to vary linearly in the
axial direction, so that

T ¼ C0 z þ Gðr=RÞ ð9:6Þ

where C0 is a constant to be determined, given the proper boundary conditions per


problem in hand.
The boundary conditions in the radial direction are

∂T
¼0 at r ¼ 0 ð9:7Þ
∂r
∂T
k ¼ q00 at r=R ¼ 1 ð9:8Þ
∂r

where q00 is the local heat flux. Inserting the expression for T in the energy equation
gives
  r 2    
α 1∂ ∂G
2C0 1  ¼ r ð9:9Þ
R υaverage r ∂r ∂r

Inserting twice,
    
1 r 4
r 2 α ∂G
R C02
 ¼ r þ C1 ð9:10Þ
R 2 R υaverage ∂r
  
1 2 r 2 1 r  4 α r 
R C0  ¼G þ C1 ln þ C2 ð9:11Þ
2 R 4 R υaverage R

The first boundary condition requires that C1 ¼ 0. The temperature is then


  
υaverage R 1  r 2 1 r 4
T ¼ C0 zþ  þ C02 ð9:12Þ
α 2 R 4 R

The second boundary condition requires that

2q00 α
C0 ¼ ð9:13Þ
kυaverage R

This gives the radial temperature distribution, where C2 is related to a reference


temperature T0,
 
2q00 α z q00 R  r 2 1 r 4
T¼ þ  þ T0 ð9:14Þ
kυaverage R k R 4 R

bahmanz@aol.com
328 9 Forced Convection Heat Transfer

The mixed mean temperature is given by


ð1
2
Tm ¼ ðr=υÞυTd ðr=υÞ
υaverage 0
ð9:15Þ
2q00 α z 7 q00 R
¼ þ þ T0
kυaverage R 24 k

but also the difference between the wall temperature and bulk temperature is

3 q00 R 7 q00 R 11 q00 R


Tw  Tm ¼  ¼ ð9:16Þ
4 k 24 k 24 k

Define a heat transfer coefficient or convection coefficient, h, and for a fully


developed pipe flow, the convection coefficient is a constant and is not varied along
the pipe length (as long as all thermal and flow properties are constant also).

q00 ¼ hðT wall  T bulk Þ ð9:17Þ

and finally, for constant heat addition, the Nusselt number is constant and equal to

hD 48
NuD ¼ ¼ ¼ 4:364 ð9:18Þ
k 11

Note that for the case of constant wall temperature, the Nusselt number
approaches a limit of Nu ¼ 3.658 asymptotically [1, p. 96].
Unlike turbulent flow, where the hydraulic diameter can be used for noncircular
duct shapes, care must be taken with laminar flows in noncircular ducts. Figure 9.2
shows Nusselt numbers for laminar heat transfer in a variety of duct shapes for both
constant heat flux and constant wall temperature boundary conditions.

Cross-sectional b/a Nu H Nu T Cross-sectional b/a Nu H Nu T


shape shape

4.364 3.66 a 4.0 5.33 4.44


b

a 1.0 3.61 2.98 a 8.0 6.49 5.60


b
b
a 1.43 3.73 3.08 8.235 7.54
b
a 2.0 4.12 3.39 5.385 4.86
b
a 3.0 4.79 3.96 3.00 2.35
b

Fig. 9.2 Nusselt numbers for fully developed velocity and temperature profile in tube of various
cross sections [1]

bahmanz@aol.com
9.3 Heat Transfer in Laminar Boundary Layers 329

9.3 Heat Transfer in Laminar Boundary Layers

Theory of boundary layer was fully discussed in Chap. 5, and development of a


boundary layer, its transition, the way a flow handles a pressure gradient, and a
possible separation were also considered.
Recall our discussion in Chap. 5 where assumed two parallel plates. Further, we
assumed the lower plate was stationary and the other one moving. We said that
there was a No Slip condition, which meant that the fluid does not slip past the solid
in contact. Needless to say that this is a typical effect of viscosity.
Considering Fig. 9.3 and a simple flat plate, we can follow the effects as flow
approaches a solid body, with a uniform (inviscid) flow in front of a flat plate at a
speed U1. As soon as the flow “hits” the plate No Slip Condition gets into action.
As a result, the velocity on the body becomes zero. Since the effect of viscosity is to
resist fluid motion, the velocity close to the solid surface continuously decreases as
the flow moves downstream. But away from the flat plate, the speed is equal to the
free-stream value of U1. Consequently, a velocity gradient is set up in the fluid in a
direction normal to flow. Thus, a layer establishes itself close to the wall with a
velocity gradient. This is what we call the boundary layer.
The boundary layer is not a static phenomenon. It is dynamic. The thickness of
boundary layer (the height from the solid surface where we first encounter 99% of
free-stream speed) continuously increases. A shear stress develops on the solid
wall. This shear stress causes a drag on the plate. Boundary layer has a pronounced
effect upon any object, which is immersed and moving in a fluid.
Drag on an aeroplane or a ship and friction in a pipe are some of the common
manifestations of boundary layer. A boundary layer may be laminar or turbulent
and Fig. 9.4, presents a typical velocity profile for laminar and turbulent boundary
layers where for turbulence there is an intense agitation.
A turbulent boundary layer forms only at larger Reynolds numbers. Those
calculating turbulent flow rely on what is called Turbulence Viscosity or Eddy
Viscosity, which has no exact expression. It has to be modeled. Several models
have been developed for this purpose and superimposing of Fig. 9.4 is depicted in
Fig. 9.5 below.

Inviscid Flow
Uniform Flow

Boundary
Layer

Fig. 9.3 Formation of a boundary layer

bahmanz@aol.com
330 9 Forced Convection Heat Transfer

Fig. 9.4 Typical velocity profiles for laminar and turbulent boundary layers

Fig. 9.5 Typical velocity


profiles for laminar and
turbulent boundary layers
superimposed

Base on above definitions, it is possible to do the same scaling for thermal


boundary layers as for momentum boundary layers and show that one of the
diffusion terms can be dropped, giving the steady-state parabolic form
2
∂θ ∂θ ∂ θ
u þυ ¼α 2 ð9:19Þ
∂x ∂y ∂y

where the boundary conditions are simplified by using

T  Tw
θ¼ ð9:20Þ
T1  Tw

For a flat plate with constant temperature, as shown in Fig. 9.6, the boundary
conditions are then

bahmanz@aol.com
9.3 Heat Transfer in Laminar Boundary Layers 331

Fig. 9.6 Thermal boundary layer on a flat plate

θ¼0 at y ¼ 0 ð9:21Þ
θ!1 at y ! 1 ð9:22Þ
θ¼1 at x ¼ 0 ð9:23Þ

We already have a solution for the simple case of v ¼ α or Pr ¼ 1 since the


solution is the same as for the momentum equation. We can also look for a
similarity solution for the case where Pr 6¼ 1, using the same similarity variable
as in Chap. 5,
y
η ¼ pffiffiffiffiffiffiffiffiffiffiffi ð9:24Þ
vx=U

We then look for a solution in the form of θ(η). Use of the stream function gives
simple forms for the velocities
u
¼ ζ 0 ðηÞ ð9:25Þ
U
∂ψ ∂ψ pffiffiffiffiffiffiffiffiffiffiffi
u¼ υ¼ ψ ¼ vxUζ ð9:26Þ
∂y ∂x

Differentiating and substituting into the energy equation,


rffiffiffiffiffiffi
∂θ ∂θ 1 vU ∂θ ∂η
u þυ ¼ ζ ð9:27Þ
∂x ∂y 2 x ∂η ∂y
  !
2 2
∂ θ ∂ ∂θ ∂η ∂ 1 ∂θ α ∂ θ ∂η
α 2¼α ¼α pffiffiffiffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffi 2 ð9:28Þ
∂y ∂y ∂η ∂y ∂y vx=U ∂η vx=U ∂η ∂y

bahmanz@aol.com
332 9 Forced Convection Heat Transfer

Pr 0
θ00 þ ζθ ¼ 0 ð9:29Þ
2

where the new boundary conditions become

θ ð 0Þ ¼ 0 θ ð1Þ ¼ 1 ð9:30Þ

This can be integrated as follows:

dθ0 Pr 0 dθ0 Pr
þ ζθ ¼ 0 þ ζdη ¼ 0 ð9:31Þ
dη 2 θ0 2
 
0 Pr Ð η
θ ¼ C1 exp  0 ζdζ
  2 
Ðη ð9:32Þ
Pr Ð η
θ ¼ C1 0 exp  ζdζ dη þ C2
2 0

The boundary condition that θ(0) ¼ 0 requires that C2 ¼ 0. The boundary condi-
tion at infinity then gives
Ð η Pr Ð η 
exp  2 0 ζdζ dη
θ ¼ Ð 1 Pr Ð η
0  ð9:33Þ
0 exp  2 0 ζdζ dη

since we already know, this integral can be evaluated with the help numerical
analysis methods. In terms of the heat transfer coefficient h and heat flux q00 ,
   
00 ∂T ∂θ
q ¼ hðT w  T 1 Þ ¼ k ¼ kðT 1  T w Þ
∂y o ∂y o
k ðT w  T 1 Þ 0
¼ pffiffiffiffiffiffiffiffiffiffiffi θ ð0Þ ð9:34Þ
vx=U

This can be related to the Nusselt number

hx x
Nux ¼ ¼ pffiffiffiffiffiffiffiffiffiffiffi θ0 ð0Þ ¼ Rex1=2 θ0 ð0Þ ð9:35Þ
k vx=U

The function can be approximated for moderate values of the Prandtl number,
giving

Nux ¼ 0:332Pr 1=3 Re1=2


x ð9:36Þ

Note the comparison between the results of this equation for Pr ¼ 1 and our
previous results for the friction coefficient,

bahmanz@aol.com
9.3 Heat Transfer in Laminar Boundary Layers 333

Table 9.2 Values of Pr


Nux Rex1=2 for several Prandtl m 0.7 0.8 1.0 5.0 10.0
numbers [1]
0.0753 0.242 0.253 0.272 0.457 0.570
0 0.292 0.307 0.332 0.585 0.730
0.111 0.331 0.348 0.378 0.669 0.851
0.333 0.384 0.403 0.440 0.792 1.013
1.0 0.496 0.523 0.570 1.043 1.344
4.0 0.813 0.858 0.938 1.736 2.236

Fig. 9.7 Values of β for various wedge angles

0:664 0:664
cf ¼ pffiffiffiffiffiffiffiffiffiffiffi ¼ 1=2 ð9:37Þ
vx=U Rex

It is also possible to generate similarity solutions for stagnation point flows and
wedge flows, where the free-stream velocity varies as u1 ¼ Cxm. Table 9.2 gives
values of the Nusselt number for a variety of wedge flows and Prandtl numbers,
where the wedge angles are illustrated in Fig. 9.7, where m ¼ (β/π)(2  β/π).
Heat transfer is significantly higher in the entrance region of a tube than in the
fully developed region. Table 9.3 gives Nusselt number values for laminar flow in
the entrance region of a uniform temperature pipe.

bahmanz@aol.com
334 9 Forced Convection Heat Transfer

Table 9.3 Nusselt number values for the entry region of circular tubes with constant surface
temperature combined thermal and hydrodynamic entry length x [1]
x=R Nux (local) Nun (mean over length x)
RePr Pr ¼ 0.7 Pr ¼ 2 Pr ¼ 5 Pr ¼ 0.7 Pr ¼ 2 Pr ¼ 5
0.001 16.8 14.8 13.5 30.6 25.2 22.1
0.002 12.6 11.4 10.6 22.1 19.1 16.8
0.004 9.6 8.8 8.2 16.7 14.4 12.9
0.006 8.25 7.5 7.1 14.1 12.4 11.0
0.01 6.8 6.2 5.9 11.3 10.2 9.2
0.02 5.3 5.0 4.7 8.7 7.8 7.1
0.05 4.2 4.1 3.9 6.1 5.6 5.1
1 3.66 3.66 3.66 3.66 3.66 3.66

9.4 Heat Transfer in Turbulent Tube Flow

A common situation encountered by the nuclear, mechanical and chemical engineer


is heat transfer to fluid flowing through a tube. This can occur in heat exchangers,
boilers, condensers, evaporators, and a host of other process equipment such as
internal core of nuclear power plant with fuel bundles in it. Therefore, it is useful to
know how to estimate heat transfer coefficients in this situation. We can classify the
flow of a fluid in a straight circular tube into either laminar or turbulent flow. It is
assumed from here on that we assume fully developed incompressible, Newtonian,
steady flow conditions. Fully developed flow implies that the tube is long compared
with the entrance length in which the velocity distribution at the inlet adjusts itself
to the geometry and no longer changes with distance along the tube.
Todreas and Kazimi [2, p. 442] indicates heat transfer correlations in turbulent
flow for empirical theoretical aspect of heat transfer coefficients for turbulent
pipe flow.
For turbulent tube flows, the time-averaged energy transport equation can be
written in cylindrical coordinates in analogy to the time-averaged momentum
transport equation
   
∂T ∂T ∂T 1 ∂ ∂T ∂ ∂T
þ u þυ ¼ r ð α þ εT Þ þ r ð α þ εT Þ ð9:38Þ
∂t ∂z ∂r r ∂r ∂r ∂z ∂z

where εT is the turbulent thermal eddy diffusivity. In laminar flow, the entry region
can be long, but in turbulent flow the entry region is typically quite short and the
velocity profiles become fully developed quite rapidly, so it is valid to approximate
for fully developed, steady tube flow,
 
dT 1 ∂ ∂Tm
u ¼ r ð α þ εT Þ ð9:39Þ
dz r ∂r ∂r

bahmanz@aol.com
9.4 Heat Transfer in Turbulent Tube Flow 335

where the velocity profile is assumed to be independent of z. Fully developed


temperature profiles are normally obtained after 10–15 tube diameters. We can
introduce a new variable as

y¼Rr ð9:40Þ

so
 
d Tm 1 ∂ ∂T
u ¼ R  y ð v þ εT Þ ð9:41Þ
dz ðR  yÞ ∂y ∂y

With boundary conditions

∂T
¼0 at y ¼ R ð9:42Þ
∂y
∂T ∂T
keff ¼ ρcðα þ εT Þ ¼ q00 at y ¼ 0 ð9:43Þ
∂y ∂y

where q00 is heat transfer from the fluid. For turbulent flow, to first order we can
assume slug flow, or u ¼ U ¼ constant. Integrating,
 
y2 d Tm ∂T
Ry  U ¼ ð R  y Þ ð α þ εT Þ þ C1 ð9:44Þ
2 dz ∂y

Applying the first boundary condition gives

R2 dTm
C1 ¼ U ð9:45Þ
2 dz

Inserting and solving for the derivative, we get


 2 
∂T y R2 dTm
ðR  yÞðα þ εT Þ ¼  Ry þ U ð9:46Þ
∂y 2 2 dz
 
∂T R  y U q00
¼ ð9:47Þ
∂y α þ εT 2 dz

The second boundary condition allows us to quantify the mean temperature


gradient, giving
 
∂T 1  y=R q00
¼ ð9:48Þ
∂y α þ εT ρc

bahmanz@aol.com
336 9 Forced Convection Heat Transfer

Integrating, it yields
00 ðy
q 1  y=R
T ¼ dy þ T w ð9:49Þ
ρc 0 α þ εT

where the constant of integration is set to give the wall temperature. Inserting our
nondimensional temperature and length variables
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
þ τ0 =ρ
y þ ðT w  TÞ τ0 =ρ
y ¼ T ¼ ð9:50Þ
v q00 =ðρcÞ

we obtain
ð yþ
þ 1  y=R
T ¼ dyþ ð9:51Þ
0 1=Pr þ εT =v

For an analytic solution, we can use the same two-region model used for solution
of the momentum equation. Inside the viscous sublayer, (y+ < 13.2), eddy diffusion
is neglected and since the layer is thin, y/R  0. Outside the viscous sublayer,
molecular diffusion is neglected, and the eddy diffusion is assumed to be given
by the Reichardt equation, used in Chap. 6,
  r 2  ε ε
εM Kyþ  r M T εT
¼ ¼ 1þ 1þ2 ¼ ¼ Pr t ð9:52Þ
v 6 R R εT v v

Then, for y+ > 13.2


!
Pr 1:5 1  r=R
T þ ¼ 13:2Pr þ ln yþ
t
 5:8 ð9:53Þ
K 1 þ 2ðr=RÞ2

We can approximate Prt ¼ 0.9 and K ¼ 0.4, giving


!
þ 1:5ð1  r=RÞ
T þ ¼ 13:2Pr þ 2:25 ln y  5:8 ð9:54Þ
1 þ 2ðr=RÞ2

This temperature profile is plotted in the Fig. 9.8 for Re ¼ 30,000. Note the effect
of Prandtl number, and also note that this expression should not be used for Pr
outside the range from 0.5 to 1.0 (thus use it for gases).
When the 1/7 power law is used to determine the mean temperature
distribution [1].

T  T w  r 1=7
¼ 1  ð9:55Þ
Tc  T w R

bahmanz@aol.com
9.4 Heat Transfer in Turbulent Tube Flow 337

Fig. 9.8 Effect of Prandtl number on the turbulent flow temperature distribution [1]

where Tc is the centerline temperature, and previous relationships for the wall
shears as a function of Re are applied, then the Nusselt number for 0.5 < Pr < 1.0 is
found to be

0:152Re0:9 Pr
Nu ¼  ð9:56Þ
0:833 2:25 ln 0:11Re0:9 þ 13:2Pr  5:8

Figure 9.9 provides Nusselt number values for a wider range of Prandtl and
Reynolds number values.
Dittus and Boelter [3] gave a more accurate empirical form for smooth tubes that
is commonly applied when the viscosity evaluated at the wall temperature is close
to that at the bulk temperature. Valid in the range of 0.7 < Pr < 100, ReD > 10,000
and

Nu ¼ 0:023Re0:8 Pr 0:4 when fluid is heated ð9:57Þ


Nu ¼ 0:023Re0:8 Pr 0:3 when fluid is cooled ð9:58Þ

where Re and Pr are evaluated at the arithmetic mean bulk temperature (average of
the inlet and outlet bulk temperature).

bahmanz@aol.com
338 9 Forced Convection Heat Transfer

103
8
6
5 Pr = 1000
4
3

2 100

30
102
8
10
6 3
Nu 5
4 1
3 0.7 Pr = 0.03
0.5
2
0.01
0.003
101
8
0.001
6
5 0.000
4 laminar flow
3

2
2 3 4 5 6 8 104 2 3 4 5 6 8 105 2 3 4 5 6 8 106

Fig. 9.9 Nusselt numbers for fully developed velocity and temperature profile in circular tubes
with constant heat rates [1]

The usual recommendation is to use this correlation for ReD > 10,000, but in
practice it is used even when the flow is in transition between laminar and turbulent
flow for lack of better correlations. A modern correlation that is slightly more
accurate is recommended in the textbooks for your use.

ðf c =8ÞðRe  1000ÞPr
Nu ¼ ð9:59Þ
1 þ 12:7ðf c =8Þ1=2 Pr 2=3  1

where fc is friction coefficient as before. Another researcher Mills suggested using


the above correlation for Reynolds number between 3000 and 106. This will not
work when Pr ¼ 1, but there are no fluids with that precise value of Prandtl number.
Physical properties to be used in these correlations are evaluated at the average
of the inlet and exit temperatures of the fluid. The friction factor fc is the Darcy
friction factor, and you can use Petukhov’s formula for evaluating it and is shown
here.

1
fc ¼ ð9:60Þ
½0:790 ln ðReÞ  1:642

This result is good for turbulent flow in smooth pipes for Re  5  106.

bahmanz@aol.com
9.5 Heat Transfer in High-Speed Laminar Boundary-Layer Flow along a Flat Plate 339

When the temperature difference between the wall and the bulk fluid is large, the
difference between the viscosity evaluated at the wall temperature μw and that at the
bulk temperature μb becomes significant and the Seider and Tate equation [4] is
useful, in the range of 0.7 < Pr < 120, Re > 10,000 and L/D > 60.
 0:14
μb
Nu ¼ 0:023Re Pr 0:8 0:3
ð9:61aÞ
μw

Another version Seider–Tate correlation is as follows:


   
Dh 1=3 μf 0:14
Nu ¼ 1:86 RePr ð9:61bÞ
L μw

where Re and Pr are evaluated at the arithmetic mean bulk temperature.


Todreas and Kazimi [4, p., 442] provide more detailed information on heat
transfer correlations and entrance effects.

9.5 Heat Transfer in High-Speed Laminar


Boundary-Layer Flow along a Flat Plate

Previously, we studied heat transfer for laminar flow along flat plate when the effect
of viscosity energy dissipation is negligible along the boundary layer. But under the
conditions where the free-stream velocity is high, the viscous energy dissipation
effect cannot be ignored and the temperature gradients in the boundary layer
becomes so large as well to the point that the properties of the fluid vary with
temperature in most significant way. A number of special studies and techniques
have been developed to deal with solution of such complex problem and we refer
the reader to most common heat transfer books written by different authors. For
most practical purpose, we use the brief following analysis done by Pohlhausen [5],
where he considers, the heat transfer rate in high-speed flow along a flat plate at
uniform temperature by considering a low-speed incompressible flow with heat
transfer coefficient h and a temperature difference Tw  Taw, where Tw is the actual
wall temperature and is the adiabatic wall temperature. Under these circumstances,
we are interested to see how the analysis was done to determine the adiabatic wall
temperature Tw.
The problem was set up by considering the high-speed flow of an incompress-
ible, constant-property fluid at a temperature T1 with a velocity u1 along a flat
plate using Fig. 9.10 and utilizing the system equations of Continuity, Momentum,
and Energy as below, in a dimensional form to determine the heat transfer solution
of the boundary-layer equations.

bahmanz@aol.com
340 9 Forced Convection Heat Transfer

q* = T — Tw
Velocity Temperature Velocity
profile profile boundary
u• layer
u'• Thermal
q*• boundary
q*
• u (x,y) layer
y q* (x,y)
δ(x) δ t (x)
0 x
q*w = 0 Wall

Fig. 9.10 Velocity and thermal boundary layers for laminar flow over a flat plate [6]

And all the governing equation are as given below

∂u ∂υ
þ ¼0 Continuity ð9:62aÞ
∂x ∂y
2
∂u ∂u ∂ u
u þυ ¼v 2 x-direction momentum ð9:62bÞ
∂x ∂y ∂y
2  2
∂T ∂T ∂ T μ ∂u
u þυ ¼α 2þ Energy ð9:62cÞ
∂x ∂y ∂y ρcp ∂y

where the last term on the right-hand side of the energy equation is for the viscous-
dissipation effects. Now taking under consideration solution of the governing
Eqs. 9.62 for special case of an adiabatic plate, which dictates that the first derivate
of temperature in y component to vanish for y ¼ 0 (i.e., (∂T/∂y)|y ¼ 0 ¼ 0 at y ¼ 0).
With this situation in mind, the appropriate boundary conditions of Eqs. 9.62a–
9.62c are illustrated as:

∂T
u¼0 υ¼0 ¼ 0 at y ¼ 0 ð9:63aÞ
∂y
u ! u1 T ! T1 as y ! 1 ð9:63bÞ

Set of governing Eqs. 9.62a–9.62c, given the boundary conditions provided by


set of Eqs. 9.63a and 9.63b, was solved by Pohlhausen [5] and it was revealing that
the difference between the adiabatic wall temperature Taw and the external-flow
temperature T1 can be expressed in the form of:

u21
T aw  T 1 ¼ r ð9:64Þ
2cp gc J

bahmanz@aol.com
9.5 Heat Transfer in High-Speed Laminar Boundary-Layer Flow along a Flat Plate 341

1.9 Pr1/3
101

Recovery factor, r
Pr1/2
4

100

4
6 100 2 4 6101 2 4 6 102 2 4 6 103
Prandil number, Pr

Fig. 9.11 Recovery factor for laminar flow along an adiabatic flat plate [6]

In this equation, r is called the recovery factor, which is a function of the Prandtl
number, where gc and J are the conversion factors. Figure 9.11 above shows a plot
of the computed numerical values of the recovery factor versus the Prandtl numbers
(i.e., gas, water) following relation relates recovery factor r to the Prandtl number

r ffi Pr 1=2 for 0:6 < Pr < 15 ð9:65Þ

For limiting case when Pr ! 1, Eq. 9.65 converts to the new form as Eq. 9.66.

r ffi Pr 1=3 ðvery large PrÞ ð9:66Þ

Utilization of Eq. 9.66 is more applicable for fluids such as oils as oils, which
have a very large Prandtl number.
The most significant aspect of recovery factor becomes more visible in case of
fluid such an ideal gas at a temperature T1 with velocity u1 that is slowed down
adiabatically to zero velocity.
As a result, the conversion of kinetic energy in the gas into internal energy will
produce at a gas temperature T0 given by the solution as

u21
T0  T1 ¼ ð9:67Þ
2cp gc J

where T0 is called the stagnation temperature.


Comparing Eqs. 9.64 with 9.67 shows an interesting conclusion that for recovery
factor r ¼ 1, the adiabatic wall temperature Taw is equivalent to the stagnation

bahmanz@aol.com
342 9 Forced Convection Heat Transfer

temperature T0, which also can be seen from Fig. 9.11 for Pr ¼ 1 and velocity u1
for the gas as well.
If the Pr > 1 then, the recovery factor r is greater than unity and the adiabatic
wall temperature exceeds the stagnation temperature. In contrast for Pr < 1, the
recovery factor r is less than unity and the adiabatic wall temperature is less than the
stagnation temperature.
Now returning our attention to the discussion of problem for high-speed flow of
an incompressible, with constant fluid property at a temperature T1 with a velocity
u1 along a flat plate that is maintained at a uniform temperature Tw, it shows that
the local heat flux qx using Eq. 9.36 is:

k
qx ¼ 0:332Pr 1=3 Re1=2
x ðT w  T aw Þ ð9:68Þ
x

If a local heat transfer coefficient h(x) is now presented as

qx ¼ hðxÞðT w  T aw Þ ð9:69Þ

Then, by forming these two recent equations (Eqs. 9.68 and 9.69), we obtain the
following:

hðxÞx
Nux  ¼ 0:332Pr 1=3 Re1=2
x ð9:70Þ
k

If we compare the Eq. 9.70 with Eq. 9.36, we conclude that the heat transfer
coefficient h(x) based on the temperature difference (Tw  Taw) for the high-speed
flow considered above is exactly, the same as the heat transfer coefficient for the
low-speed flow. The average value of the heat transfer coefficient hm over the
length 0  x  L of the plate is calculated by:

hm ¼ 2hðxÞx¼L ð9:71Þ

Generally, in case of high-speed flow, temperature gradients in the boundary


layer are high; therefore, the properties of the fluid vary most likely with
temperature.
Eckert [7] recommended that the variation of properties could be approximately
included in the heat transfer coefficients, which is calculated by Eq. 9.69 providing
that the properties of the fluid are evaluated at the following reference temperature
Tr as

T r ¼ T 1 þ 0:5ðT w  T 1 Þ  0:22ðT aw  T 1 Þ ð9:72Þ

bahmanz@aol.com
9.6 Problems 343

9.6 Problems

Problem 9.1: Consider water as a fluid that its velocity profile under fully devel-
oped laminar flow condition in a tube is given as u ¼ umax[1  (r/R)2]. For this
problem, assume steady flow and consider for water μ ¼ 2:1  105 lbf s ft2
,
umax ¼ 10 s , and R ¼ 3 in. By using Fig. P9.1, we
ft

(a) Obtain and expression for the shear force per unit volume in the r-direction
(b) Evaluate its maximum value at these conditions

dr
r
(rt)2pdz
r

Fig. P9.1 Illustration of Problem 9.1

Problem 9.2: Assuming the relationship for the maximum temperature in a fluid
occurs at the midpoint between the plates is given by

μu21
T max  T 0 ¼
8k

where μ and k are fluid dynamic viscosity and thermal conductivity, respec-
tively, and u1 is the fluid velocity.
Determine the maximum temperature rise in the fluid with a velocity
of u1 ¼ 6 m/s for a given heavy lubricating oil {μ ¼ 0.25 kg/(m s),
k ¼ 0.125 W/m
C)} at room temperature that flows in the clearance between
a journal and its bearing. Assume both the bearing and the journal are at the same
temperature.

Problem 9.3: Assume the pressure drop is given by ΔP ¼ f ðL=DÞ ρu2m =2 mN2
(Eq. 6.94) where the variables in it are defined as follows:
ΔP ¼ Pressure drop across tube
f ¼ friction factor
um ¼ Mean flow velocity
D ¼ Tube inside diameter
ρ ¼ Fluid density
L ¼ Length of tube
Also, assume pumping power is given by
  
m3 N N m
Pumping PowerðWÞ ¼ V M ΔP 2 ¼ V M ΔP or W
s m s

bahmanz@aol.com
344 9 Forced Convection Heat Transfer

where VM is the flow rate in cubic meters per second through the pipe and it is
given by
π 
VM ¼ n D 2 um
4

Now given engine oil that is pumped with a mean velocity of um ¼ 0.6 m/s
through a bundle of n ¼ 80 tubes each of inside diameter D ¼ 2.5 cm and length
of L ¼ 10 m. The physical properties of the oil are kinematic viscosity
v ¼ 0.75  104 m2/s and density ρ ¼ 868 kg/m3. Calculate the pressure drop
across each tube and the total power required for pumping the oil through
80 tubes to overcome the fluid friction to flow.
Problem 9.4: Assume that heat transfer coefficient h for heating by condensing
steam or for laminar forced convection inside a circular tube in the hydrody-
namically and thermally developed region under constant wall temperature
boundary condition is given by

hD hD
Nu  ¼ 3:657 and Nu  ¼ 4:364
k k

Now, consider the heating of atmospheric air flow with a mean velocity of
um ¼ 0.5 m/s inside a thin-walled tube 2.5 cm in diameter in the hydrodynam-
ically and thermally developed region. Heating can be done either by condensing
steam on the outer surface of the tube, thus maintaining a uniform surface
temperature, or by electric resistance heating, thus maintaining a uniform sur-
face heat flux. Calculate the heat transfer for both of these heating conditions by
assuming air properties can be evaluated at 350
K. The air properties at 350
K
are kinematic viscosity v ¼ 20.76  106 m2/s and thermal conductivity
k ¼ 0.03 W/(m
C)
Problem 9.5: Air at atmospheric pressure and at a temperature 150
F (56.6
C)
flows with a velocity of 3 ft/s (0.915 m/s) along a flat plate which is kept at a
uniform temperature 250
F (121.1
C). Determine the local heat transfer
coefficient h(x) at a distance x ¼ 2 ft (0.61 m) from the leading edge of the
plate and the average heat transfer coefficient hm over the length x ¼ 0 to 2 ft
(0.61 m). Calculate the local heat transfer rate from the plate to the air over the
region x ¼ 0 to 2 ft per foot width of the plate. The physical properties of air at
200
F (i.e., arithmetic mean of Tw and T1) and at atmospheric pressure are

v ¼ 0:24  103 ft2 =s 0:223  104 m2 =s

k ¼ 0:0181 Btu=h ft Fð0:0313 W=m



Pr ¼ 0:692

bahmanz@aol.com
References 345

Problem 9.6: Air at a temperature T1 ¼ 460


R(255.6 K) and pressure
P ¼ 30
1
atm flows with a velocity u1 ¼ 2000 ft/s(609.6 m/s) along a 1 ft
(0.305 m) long flat plate maintained at a uniform temperature Tw ¼ 560
R
(311.1 K). Determine the heat transfer rate to the plate over 1 f. length per unit
width. As a hint, assume this at high-speed laminar flow case and assume the
Prandtl number for air at 460
R is Pr ¼ 0.711. The specific heat at constant
pressure for the air is cp ¼ 0.24.
Problem 9.7: Crude oil is heated with water vapor in a vertical pipe with internal
diameter d ¼ 10 mm. The inlet temperature of oil is 20
C and the wall temper-
ature is Tw ¼ 100
C. The oil mass flow rate in the pipe is W ¼ 0.04 kg/s.
Calculate the length of the pipe L that is required to heat the oil with 20
C.
Physical properties of oil are as follows: specific heat cp ¼ 2.09 kJ/kg K,
density ρ ¼ 1000 kg/m3, thermal conductivity λ ¼ 0.15 W/m K. The dynamic
viscosity is a function of temperature, as given in the table below

Temp
C 20 30 40 100
μ, Pas 2.5 x 102 1.8 x 102 1.2 x 102 0.3 x 102

References

1. W.M. Kays, M.E. Crawford, Convective Heat and Mass Transfer, 2nd edn. (McGraw-Hill,
New York, 1980)
2. N.E. Todreas, M.S. Kazimi, Nuclear Systems I: Thermal Hydraulic Fundamentals.
3. F.W. Dittus, L.M.K. Boelter, University of California, Berkeley, Publ. Eng. 2, 443 (1930)
4. E.N. Sieder, G.E. Tate, Ind. Eng. Chem., 28, 1429 (1936)
5. E. Pohlhausen, Z. Angew, Math. Mech., 1, 115 (1921)
6. M. Necati Ozisk, Basic Heat Transfer (McGraw-Hill, New York, 1997)
7. E.G. Eckert, Engineering relations for heat transfer and friction in high-velocity laminar and
turbulent boundary layer flow over surface with constant pressure and temperature. Trans.
ASME. 78, 1273–1284 (1956)

bahmanz@aol.com
View publication stats

Вам также может понравиться