Turbulent Forced Convection

Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

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

net/publication/312751456

Turbulent Forced Convection

Chapter · January 2015


DOI: 10.1007/978-3-319-17434-1_6

CITATIONS READS
0 4,609

2 authors, including:

Bahman Zohuri
Galaxy Advanced Engineering
636 PUBLICATIONS 2,602 CITATIONS

SEE PROFILE

All content following this page was uploaded by Bahman Zohuri on 11 February 2019.

The user has requested enhancement of the downloaded file.


Chapter 6
Turbulent Forced Convection

Turbulence occurs nearly everywhere in nature. It is characterized by the efficient


dispersion and mixing of vorticity, heat, and contaminants. In flows over solid
bodies such as airplane wings or turbine blades, or in confined flows through ducts
and pipelines, turbulence is responsible for increased drag and heat transfer.
Turbulence is therefore a subject of great engineering interest. In this chapter, we
will look at the state of the fluid motion, which is independent of heat transfer and
this is where we speak of forced convection. Forced convection occurs when an
external force, such as a pump, fan, or a mixer, induces a fluid flow. On the other
hand, natural convection is caused by buoyancy forces due to density differences
caused by temperature variations in the fluid. At heating the density change in the
boundary layer will cause the fluid to rise and be replaced by cooler fluid that also
will heat and rise. These continuous phenomena are called free or natural
convection.

6.1 Introduction

Previously, we noted that the Navier–Stokes equations are nonlinear due to the
convective terms. This nonlinearity can give rise to multiple solutions, and in
particular, to the highly, transient phenomenon known as turbulence.
In case of forced convection, where the fluid motion is independent of heat
transfer, and yet it depends upon some external means in form of, for example, a
pressure difference. On the other hand, if the fluid motion is taking place via body
force within the fluid, such as density gradients near a solid surface due to the heat
transfer, then free (natural) convection presents itself.
A central role in determining the state of fluid motion is played by the Reynolds
number. In general, a given flow undergoes a succession of instabilities with
increasing Reynolds number and, at some point, turbulence appears more or less
abruptly. It has long been thought that the origin of turbulence can be understood by

© Springer International Publishing AG 2017 223


B. Zohuri, Thermal-Hydraulic Analysis of Nuclear Reactors,
DOI 10.1007/978-3-319-53829-7_6

[email protected]
224 6 Turbulent Forced Convection

sequentially examining the instabilities. This sequence depends on the particular


flow and, in many circumstances, is sensitive to a number of details. A careful
analysis of the perturbed equations of motion has resulted in a good understanding
of the first two instabilities in a variety of circumstances.
Given the Reynolds number that was established previously, we can define
different region of flow ranging from laminar to transition to turbulent as:
VDρ VD
Re ¼ ¼ ð6:1Þ
η ν

where
Re: Reynolds number (dimensionless number)
V: Velocity of flow (m/s)
D: Diameter of pipe (m)
ρ: Density of fluid (kg/m3)
η: Dynamic viscosity (kg/m s)
ν ¼ (η/ρ): Kinematic viscosity (m2/s)
Re < 2000: Laminar Flow
Re > 4000: Turbulent Flow
For Reynolds number being in range of 2000 < Re < 4000 transition region or
critical region flow can be either laminar or turbulent, which is very difficult to
pinpoint the condition of flow. Note that fluid with higher viscosity will tend to have
laminar flow or lower Reynolds number.
In addition, note that ordinary fluids, such as air, water, and oils, do behave as a
continuum fluid and they exhibit a linear relationship between the applied shear
stress and the rate of strain and such fluids are called Newtonian Fluids [1].
The mathematical expression relating the shear stress τs to the rate of strain
(velocity gradient ∂V/∂y in y-direction) for a Newtonian fluid in simple shear flow,
where only one velocity component is different from zero, is given by: [1]

∂V
τs ¼ μ ð6:2Þ
∂y

where μ is the dynamic viscosity of the fluid (kg/m s) or (N s/m2) and is a


measure of fluid resistance to flow as well as being strong function of temperature.
For each Newtonian fluid, the dynamic viscosity is constant for given temperature
and pressure and in contrast for non-Newtonian fluids, the dynamic viscosity at a
given temperature and pressure is a function of the velocity gradient. Note that the
Eq. 6.2 also is known as the Newton’s law of viscosity.
The surface shear stress can also be determined from:

ρU 21  
τ s ¼ cf N=m2 ð6:3Þ
2

[email protected]
6.1 Introduction 225

where cf is the friction coefficient or the drag coefficient, which is determined


experimentally in most cases. The drag force is calculated from:

ρU21
FD ¼ c f A ðNÞ ð6:4Þ
2

In summary, the flow of a fluid is such that the motion of particles at any point
varies rapidly in both magnitude and direction. Turbulent flow is characterized by
mixing of adjacent fluid layers.
In general, turbulent flows can be divided into three groups, and roughly
speaking they can be categorized as: [2]
1. Gridlike flow: Vortices generated by a grid structure interact, after some dis-
tance a homogeneous isotropic field of turbulence is generated. Used for funda-
mental studies.
2. Wall Shear Layers: The presence of a wall can dominate the generation of
turbulence, for both internal flows and boundary layers.
3. Free Shear Layers: Mixing layers, jets, wakes—downstream the extent of the
turbulent region always grows. Look for self-similar behavior.
Characteristics of turbulent flows are listed below: [2]
• Irregular fluctuations of velocity in all three directions, typically 10% or less of
the mean flow velocity.
• A gradient in the mean velocity must exist for turbulence to be self-sustaining
(grid turbulence is not self-sustaining). Turbulent fluctuations are the result of
vorticity in the fluid (potential flow is inherently stable).
• Time history of fluctuations looks random, but possesses structure.
• Velocity irregularities have spatial structures called eddies. (Shaped like a
vortex, imbedded jet, mushroom, etc.)
• Turbulent flows tend to entrain nonturbulent fluid, increasing the turbulent
region extent.
• Turbulent flows are diffusive—eddies transport momentum (heat, mass) from
regions of low to high velocity (temperature, concentration). Termed eddy
diffusivity. Note that the irregularities in the velocity field have certain spatial
structures known as eddies.
• Eddies exist at many scales; large eddies have internal smaller eddies which
have still smaller eddies. (Some ultimate limit is placed by viscous dissipation.)
Largest eddies are on scale of boundary layer, pipe diameter, or free jet.
• Energy can be transferred from large to small eddies (and sometimes vice versa).
Ultimately energy is dissipated by viscosity (dominantly in small eddies). Thus
viscosity is concerned with turbulent energy generation, transport of energy
from large to small eddies, and dissipation.

[email protected]
226 6 Turbulent Forced Convection

Characteristics of Transition (several ways): [2]


• Pipe flow—turbulent elements called puffs form in range 2000 < ReD < 2700.
Originate in region where the laminar profile is fully developed.
• Pipe flow—for ReD > 3200, dominant transition mechanism is slug generation
in hydrodynamic entrance region (core flow and boundary layers). Slug imme-
diately fills the pipe cross section and grows at ends as it is swept downstream.
• Boundary layers - several mechanisms - free-stream turbulence, wall roughness,
trip wires, acoustic noise, or Tollmien–Schlichting instability.
• Boundary layers—viscous stability limit is Rex < 60 , 000 though laminar
boundary layers can persist to 300,000 and higher (reentry vehicles are even
higher, negligible free-stream fluctuation and smooth surface). Favorable pres-
sure gradient extends laminar flow range (converging nozzles).
• For engineering, transition region is about as long as laminar region and wall
shear/heat transfer can be assumed to vary continuously from laminar to fully
turbulent values over this distance.
To provide a picture of turbulent phenomena, the best initial example is flow in a
tube. View of such turbulent tube flow is demonstrated in Fig. 6.1.

Laminar Sublayer

Buffer Zone
Fully Developed
Turbulent Flow
1

0.12

u,2
umax
LENT

TURBULENCE INTENSITY
TURBU

u
0.08
umax
R
INA
LAM

z,u

0.04
r,u
u,2
umax

0 0
1 0 1 0
r/R r/R

Fig. 6.1 Laminar and turbulent velocity distributions and turbulence intensity distributions for
pipe flow

[email protected]
6.1 Introduction 227

The primary points to note are that:


• Laminar and turbulent flow have different velocity profiles:
  r 2  
υ υ r 1=7
¼ 1 ¼ 1
υmax R υmax R
υave 1 υave 4 ð6:5Þ
¼ ¼
υmax 2 υmax 3
laminar ðfrom theoryÞ turbulent ðempiricalÞ

where the 1/7 power expression for turbulent flow is observed experimentally for
104 < ReD < 105 and is written in terms of average velocities, because with turbu-
lence the actual velocity fluctuates. Laminar flow has a parabolic dependence, and
the ratio of maximum to average velocity is 0.5; turbulent flow has a more
complicated velocity distribution, and the ratio of maximum to average velocity
is much closer to unity.
For turbulent flow, we can define a time-averaged velocity.
ð tþt0
1
υ¼ υdt ð6:6Þ
t0 t

Therefore, the instantaneous velocity is given by:

υ ¼ υ þ υ0 ð6:7Þ

where the time average of the fluctuating component υ0 ¼ 0, but υ0 2 is nonzero and
provides a measure of the magnitude of the local velocity fluctuation.
• In circular tubes, the transition from laminar to turbulent flow occurs at a
Reynolds number of about 2100. This value can vary, and depends on system
vibrations and tube wall smoothness.
• The size and type of fluctuation depends on the distance from the wall. Three
zones can be defined:
• Fully developed turbulent flow—transport occurs primarily by turbulent eddies.
• Buffer zone.
• Viscous sublayer—transport occurs primarily by molecular diffusion.
• Axial fluctuations tend to be larger than radial fluctuations, particularly at the
wall. Near the tube center, the fluctuations are more nearly isotropic
(see Fig. 6.1). This is quantified using the turbulence intensity,

 1=2 1=2
u0 u0 υ0 υ0
Ix ¼ Iy ¼ ð6:8Þ
υmax υmax

Also it is good to note that the turbulent kinetic energy shown below is of interest
in engineering calculation as well.

[email protected]
228 6 Turbulent Forced Convection

1  0 0   0 0 
K¼ uu þ υυ ð6:9Þ
2

6.2 Time-Averaged Conservation Equations for Turbulent


Flow in Duct

The analysis of turbulent duct flow and heat transfer is traditionally presented in
terms of time-averaged quantities, which are denoted by a bar superscript. For
example, the longitudinal velocity is decomposed as uð~ rÞ þ u0 ð~
r; tÞ ¼ uð~ r; tÞ, where
0
u is the time-averaged velocity and u is the fluctuation, or the time-dependent
difference between u and u. In the cylindrical coordinates (r, x) of the round tube
shown in Fig. 6.2, the time-averaged equations for the conservation of mass,
momentum, and energy are: [3]

∂
u 1 ∂
þ ðrυÞ ¼ 0 ðMassÞ ð6:10Þ
∂x r ∂r
∂
u ∂
u 1 d P 1 ∂ ∂T
u þυ ¼ þ r ð v þ εM Þ ðMomentumÞ ð6:11Þ
∂x ∂r ρ dx r ∂r ∂r

∂T ∂T 1 ∂ ∂T


u þυ ¼ r ð α þ εH Þ ðEnergyÞ ð6:12Þ
∂x ∂r r ∂r ∂r

Considering that the duct is a slender flow region, then the above sets of
Equations are simplified to the form that are written and the absence of second

derivatives in the longitudinal direction may be noted. In Eq. 6.11, the term dP=dx
 
means that d Pðr; xÞ ’ PðxÞ. The momentum eddy diffusivity εM and the thermal
eddy diffusivity εH are defined by:

∂
u ∂T
ρu0 υ0 ¼ ρεM and  ρcP υ0 T 0 ¼ ρcP εH ð6:13Þ
∂r ∂r

r u t0

q u y

x tapp
r0

Fig. 6.2 Distribution of apparent shear stress in fully developed turbulent flow

[email protected]
6.2 Time-Averaged Conservation Equations for Turbulent Flow in Duct 229

where u0 , υ0 , and T0 are the fluctuating parts of the longitudinal velocity, radial
velocity, and local temperature. The eddy diffusivities augment significantly the
transport effect that would occur in the presence of molecular diffusion alone, that
is, based on v for momentum and α for thermal diffusion.

6.2.1 Time Averaging of the Equation of Motion

Quite often in engineering, the detailed motion is not of interest, but only the
longtime averages or means, such as the mean velocity in a boundary layer, the
mean drag of an airplane or pressure loss in a pipeline, or the mean spread rate of a
jet. It is therefore desirable to rewrite the Navier–Stokes equations for the mean
motion. The basis for doing this is the Reynolds decomposition, which splits the
overall motion into the time mean and fluctuations about the mean. These macro-
scopic fluctuations transport mass, momentum, and matter (in fact, by orders of
magnitude more efficiently than molecular motion), and their overall effect is thus
perceived to be in the form of additional transport or stress. This physical effect
manifests itself as an additional stress (called the Reynolds Stress) when the
Navier–Stokes equations are rewritten for the mean motion (the Reynolds equa-
tions). The problem then is one of prescribing the Reynolds stress, which contains
the unknown fluctuations in Quadratic Form.

Quadratic Form [1]


In mathematics, a Quadratic Form is a homogeneous polynomial of degree
two in a number of variables. For example,

qðx; yÞ ¼ ax2 þ 2xy  3y2

is a quadratic form in the variables x and y. Quadratic forms occupy a central


place in various branches of mathematics, including number theory, linear
algebra, group theory (orthogonal group), differential geometry (Riemannian
metric), differential topology (intersection forms of four-manifolds), and Lie
theory (the Killing form).
In general, quadratic forms are homogeneous quadratic polynomials in
n variables. In the cases of one, two, and three variables, they are called
Unary, Binary, and Ternary and have the following explicit form:

qðxÞ ¼ ax2 ðUnaryÞ


qðx; yÞ ¼ ax2 þ bxy þ cy2 ðBinaryÞ
qðx; y; zÞ ¼ ax2 þ by2 þ cz2 þ dxy þ exz þ fyz ðTernaryÞ

where a , b ,    , f are the coefficients.

(continued)

[email protected]
230 6 Turbulent Forced Convection

Note that quadratic functions, such as ax2 + bx + c in the one variable (x) case,
are not quadratic forms, as they are typically not homogeneous (unless b and c
are both 0, Unary condition).
The theory of quadratic forms and methods used in their study depend in a
large measure on the nature of the coefficients, which may be real or complex
numbers, rational numbers, or integers. In linear algebra, analytic geometry,
and in the majority of applications of quadratic forms, the coefficients are real
or complex numbers. In the algebraic theory of quadratic forms, the coeffi-
cients are elements of a certain field.1

A property of turbulence is that the Reynolds stress terms are comparable to the
other terms in the Reynolds equation, even when fluctuations are a small part of the
overall motion. An equation for the Reynolds stress itself can be obtained by
suitably manipulating the Navier–Stokes equations, but this contains third-order
terms involving fluctuations, and an equation for third-order terms involves fourth-
order quantities, and so forth. This is the closure problem in turbulence. The
Navier–Stokes equations are themselves closed, but the presence of nonlinearity
and the process of averaging result in nonclosure.
Note also that the mean flow may be two- or three-dimensional. However,
turbulence is three-dimensional and rotational and always occurs at high Reynolds
numbers. Vorticity plays a major role in turbulence; thus, the random vorticity
fluctuation that characterizes turbulent flows cannot maintain themselves without
the mechanism of vortex stretching which is absent in two-dimensional flows.
Turbulent flows are highly dissipative, and viscous shear stresses perform
deformation work at the expense of the kinetic energy of turbulence. Thus, without
the continuous supply of energy provided the mean flow, turbulence cannot be
sustained.
Finally, turbulence is not a property of the fluid; it is a property of the flow. Thus
turbulent flows are highly dependent on initial and boundary conditions.
Given this situation, much of the progress in the field has been due to [1]
exploratory experiments and numerical simulations of the Navier–Stokes equations
at low Reynolds numbers and [2] plausible hypotheses in conjunction with dimen-
sional reasoning, scaling arguments, and their experimental verification.
Some typical objectives of flow control are the reduction of drag of an object
such as an airplane wing, the suppression of combustion instabilities, and the
suppression of vortex shedding behind bluff bodies. Interest in flow control has
been stimulated by the discovery that some turbulent flows possess a certain degree
of spatial coherence at large scales. Successful control has also been achieved
through the reduction of the skin friction on a flat plate by making small longitu-
dinal grooves, the so-called riblets, on the plate surface, imitating shark skin.

1
Wikipedia on Answers.com

[email protected]
6.2 Time-Averaged Conservation Equations for Turbulent Flow in Duct 231

Turbulent flow is normally dealt with by considering the average flow velocity, and
lumping the small-scale velocity fluctuations into an effective eddy viscosity. The
transformation is obtained by substituting u ¼ u þ u0 , υ ¼ υ þ υ0 , and p ¼ p þ p0 into
the equations of motion for an incompressible fluid as developed previously,
∂ ∂
ðu þ u0 Þ þ ðυ þ υ0 Þ ¼ 0 ð6:14Þ
∂x ∂y
∂ ∂ ∂ ∂
ρ ðu þ u0 Þ þ ρ ðu þ u0 Þðu þ u0 Þ þ ρ ðu þ u0 Þðυ þ υ0 Þ ¼  ðp þ p0 Þ
∂t ∂x ∂y ∂x
þ μ∇2 ðu þ u0 Þ þ ρgx
ð6:15Þ

We then average these equations over some time-period, and obtain


∂
u ∂υ
þ ¼0 ð6:16Þ
∂x ∂y
 
∂ ∂ ∂ ∂
p ∂ 0 0 ∂
ρ u þ ρ uu þ ρ uυ ¼  ρ u u þ u0 υ0 þ μ∇2 u þ ρgx
∂t ∂x ∂y ∂x ∂x ∂y
ð6:17Þ

Note that the equation of continuity remains the same, except with time-
averaged quantities, and that the equation of motion is much the same, except for
the addition of two new terms. These are associated with the turbulent velocity
fluctuations, and are commonly called the Reynolds stresses. They can be written in
terms of the turbulent momentum flux τðtÞ .

τðxxtÞ ¼ ρu0 u0 τðxytÞ ¼ ρu0 υ0 ð6:18Þ

Note that the Reynolds stresses must be zero at the wall. Often it is desirable to
replace the Reynolds stress with an effective eddy viscosity,
∂ ∂ ∂ 1 ∂
p
u þ uu þ uυ ¼  þ ∇  ½ðν þ εM Þ∇  u þ gx ð6:19Þ
∂t ∂x ∂y ρ ∂x

We lose information by time averaging, and the Reynolds stresses are not
known. Now the number of equations is insufficient to solve the problem, giving
us a closure problem. Ad hoc, special assumptions must be made now to generate
solutions.

[email protected]
232 6 Turbulent Forced Convection

6.3 The Laminar Sublayer and Outer Turbulent Region

It is not easy to define turbulence. However, we can indicate its important features.
Probably, the most important characteristic of turbulence is its randomness. This
makes a deterministic approach extremely difficult and, in most cases, we rely on
statistical methods. The diffusivity of turbulence, which causes rapid mixing and
increased rates of momentum and heat transfer, is another feature.
Turbulent flows are significantly affected by the presence of walls. Obviously,
the mean velocity field is affected through the no-slip condition that has to be
satisfied at the wall. However, the turbulence is also changed by the presence of
the wall in nontrivial ways. Very close to the wall, viscous damping reduces the
tangential velocity fluctuations, while kinematic blocking reduces the normal
fluctuations. Toward the outer part of the near-wall region, however, the turbulence
is rapidly augmented by the production of turbulence kinetic energy due to the large
gradients in mean velocity.
In the thin, viscous region near the wall, the viscous sublayer, events leading to
the production of turbulence occur (Kays and Crawford, page 164) [4]. Simple
description: Sublayer is a continually developing laminar boundary layer, which
grows until its local Reynolds number becomes supercritical.
The sublayer becomes locally unstable and breaks down, creating a burst of
turbulence from the wall region.
The burst causes ejection from the wall region of a relatively large eddy of low
momentum fluid. Turbulent kinetic energy is created by collision of finite elements
of higher velocity fluid with low-velocity fluid.
A stable state is reached when the dissipation, plus some turbulent kinetic
energy that is convected and diffused outward, matches the generation of turbulent
kinetic energy.
Near the wall, the total apparent shear stress can be expressed as:
τ ∂
u
¼ ð ν þ εM Þ ð6:20Þ
ρ ∂y

with the assumption that u ¼ uðyÞ only, then v ¼ 0 if the wall velocity is zero, we
can write for steady state.
dτ dp
 þ ¼0 ð6:21Þ
dy dx

 
τ dp y
¼1þ ð6:22Þ
τ0 dx τ0

[email protected]
6.4 The Turbulent Boundary Layer 233

where τ0 is the wall shear. Nondimensionalize:


Shear velocity:
pffiffiffiffiffiffiffiffiffi
uτ ¼ τ0 =ρ ð6:23Þ

Nondimensional groups:
u u u=u1
uþ ¼ ¼ pffiffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffi ð6:24Þ
uτ τ0 =ρ cf =2
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
yuτ y τ0 =ρ yu1 cf =2
yþ ¼ ¼ ¼ ð6:25Þ
v v v

μðd p=dxÞ
pþ ¼ 3=2
ð6:26Þ
ρ1=2 τ0

when these are substituted into the preceding equations, we get,


τ
¼ 1 þ y þ pþ ð6:27Þ
τ0

6.4 The Turbulent Boundary Layer

The characteristics of the two-layer structure for flow in a smooth plane channel and
pipes are also applicable to turbulent boundary layers. Eq. 6.20 is representing the
total shear stress, where y is measured away from the wall, y ¼ r0  r, as indicated in
Fig. 6.2. The two contributions to total shear (or apparent τapp) stress τ, ρv∂ u=∂y
and ρεM ∂ u=∂y, are the molecular shear stress and the eddy shear stress, respec-
tively, as we mentioned previously. Note that τapp ¼ τ ¼ 0 at y ¼ 0. Using Adrian
Bejan [3], the momentum equation, of Eq. 6.9, reduces to
dP 1 ∂  
¼ rτapp ð6:28Þ
dx r ∂r

where both sides of the equation equal a constant. Integration of Eq. 6.28 on
interval of wall to the distance y in the fluid, and using the force balance on a flow
control volume of cross section A and length L as AΔP ¼ τwpL, one can show that
τ ¼ τapp decrease linearly from τ0 at the wall to zero on the centerline, [3]
 
y
τ ¼ τapp ¼ τ0 1  ð6:29Þ
r0

[email protected]
234 6 Turbulent Forced Convection

Sufficiently close to the wall, where y  r0, the apparent shear stress is nearly
constant, τapp ’ 0. The mixing-length analysis that produced the law of the wall for
the turbulent boundary layer applies near the tube wall and measurements confirm
that the time-averaged velocity profile fits the law of the wall,

uþ ¼ 2:5 ln yþ þ 5:5 ð6:30Þ

where 2.5 and 5.5 are curve-fitting constants (Reichardt [3] used these values),
and
 1=2
þ u þ u∗ y τ0
u ¼ y ¼ u∗ ¼ ð6:31Þ
u∗ v ρ

Figure 6.3 is presentation of typical velocity profile for flat plate turbulent
boundary layer and shows a comparison with experimental data.
In Fig. 6.3 where the log layer is shown with constant C1 ¼ 1κ and C2 ¼ C are
constant that per Eq. 6.30, they have the values of 2.5 and 5.5, respectively, which
are again these constants are for curve fitting purpose of experimental data. κ is Von
Karman’s constant and has been empirically determined to be 0.41. Therefore, has
value of (1/κ) ¼ (1/0.41) ’ 2.44. Although in Eq. 6.28 we have written the value of
(1/κ) ’ 2.5 and value of C ¼ 5.5, they are just for the purpose of curve fitting that is
presented in Fig. 6.3.

60
Viscous Log Degect
u+ = u Sublayer Layer Layer
u+

40

20
u+ = y+ 1
u+ = − lny+ +C
k

0
1 10 102 103 104
+
y = y u* / v

Fig. 6.3 Typical velocity profile for flat turbulent boundary layer

[email protected]
6.4 The Turbulent Boundary Layer 235

Numerical evaluation: Two empirical constants, κ and C, appear in Eq. 6.28.


Throughout the present study, the numerical values given to these constants
are:
κ ¼ 0:41
C ¼ 5:1

A great variety of other values, especially for k, can be found in the exper-
imental literature. However, in practically all cases where Eq. 6.28 is explic-
itly taken as a definition, kis found to lie between 0.39 and 0.41. Values
outside this range are usually the result of operations or assumptions which
change the definition of k and C.

The group u∗ is known as the friction velocity. The major drawback of the τapp
approximation is that the velocity profile deduced from it, Eq. 6.28, has a finite
slope at the centerline. An empirical profile that has zero slope at the centerline and
matches Eq. 6.29 as y+ ! 0 is that of Reichardt [3]:
Eq. 6.32 is presentation of experimental data for the velocity outside the viscous
sublayer, wherey ¼ r0  r.
8 9
< 3ð1 þ r=r Þ =
uþ ¼ 2:5 ln h i yþ þ 5:5
0
ð6:32Þ
:2 1 þ 2ðr=r Þ2 ;
0

The friction factor is defined by Eq. (5.46a, b) and is related to the friction
velocity (τ0/ρ)1/2, designating V ¼ Vavg:
 1=2  1=2
τ0 f
¼V c ð6:33Þ
ρ 2

An analysis based on a velocity curve fit where, instead of Eq. 6.30, u+ is


proportional to (y+)1/7 (Prandtl [5]) leads to [3]:
1=4
f c ’ 0:078ReD ð6:34Þ

where ReD ¼ VD/v and D ¼ 2r0, using Fig. 6.2. Eq. 6.34 is in agreement with
measurements up to ReD ¼ 8  104. An empirical relation that holds at higher
Reynolds number in smooth tubes such as the one presented in Fig. 6.2 is:
1=5  
f c ’ 0:046ReD 2  104 < ReD < 106 ð6:35Þ

[email protected]
236 6 Turbulent Forced Convection

An alternative that has wider applicability is obtained by using the law of the
wall, that was presented in Eq. 6.30, instead of Prandtl’ 1/7 power law, u+  (y+)1/7.
This results in [5]

1  
¼ 1:737 ln ReD f 1=2
c  0:396 ð6:36Þ
f 1=2
c

which agrees with measurements for ReD values up to O(106). The heat transfer
literature refers to Eq. 6.36 as the Karman–Nikuradse relation [4, 6]; this relation is
displayed as the lowest curve in Fig. 6.3. This figure is known as the Moody chart
[7]. The laminar flow line in Fig. 6.3 is for a round tube. The figure shows that the
friction factor in turbulent flow is considerably greater than that in laminar flow in the
hypothetical case that the laminar regime can exist at such large Reynolds numbers.
For fully developed flow through ducts with cross sections other than round, the
Karman–Nikuradse relation of Eq. 6.34 still holds if ReD is replaced by the
Reynolds number based on hydraulic diameter, ReDh . Note that for a duct of
noncircular cross section, the time-averaged τ0 is not uniform around the periphery
of the cross section; hence, in the friction factor definition of Eq. 5.46, τ0 is the
perimeter-averaged wall shear stress.
Figure 6.4 also documents the effect of wall roughness. It is found experimen-
tally that the performance of commercial surfaces that do not feel rough to the touch
departs from the performance of well-polished surfaces. This effect is due to the
very small thickness acquired by the laminar sublayer in many applications [e.g.,
because UyvsL/v is of order 102 [8], where yvsL is the thickness of the viscous
sublayer]. In water flow through a pipe, with U ’ 10m/s and v  0.01cm2/s, yvsL is
approximately 0.01 mm. Consequently, even slight imperfections of the surface
may interfere with the natural formation of the laminar shear flow contact spots. If
the surface irregularities are taller than yvsL, they alone rule the friction process.
Nikuradse [9] measured the effect of surface roughness on the friction factor by
coating the inside surface of pipes with sand of a measured grain size glued as
tightly as possible to the wall. If ks is the grain size in Nikuradse’s sand roughness,
the friction factor fully rough limit is the constant given as Eq. 6.37.
 2
D
f c ’ 1:74 ln þ 2:28 ð6:37Þ
ks

The fully rough limit is that regime where the roughness size exceeds the order
of magnitude of what would have been the laminar sublayer in time-averaged
turbulent flow over a smooth surface,

ks ðτ0 =ρÞ1=2

s ¼ 10 ð6:38Þ
v

The roughness effect described by Nikuradse is illustrated by the upper curves in


Fig. 6.4.

[email protected]
6.4 The Turbulent Boundary Layer 237

Surface condition ks (mm)


Riveted steel 0.9−9
Concrete 0.3−3
Wood stave 0.18−0.9
Cast iron 0.26
Galvanized iron 0.15
Asphalted cast iron 0.12
Commercial steel or
Wrought iron 0.05
Drawn tubing 0.0015

0.1

0.05
0.04
0.04
0.05 0.02

Relative roughness, ks/D


0.015
0.04 0.01
0.008
0.03 0.006
0.004
4f

Laminar flow, 0.002


0.02 f = 16 0.001
0.0008
ReD 0.0006
0.0004
0.0002
Smooth pipes
0.0001
(the Karman- 0.000,05
0.01 Nikuradse relation)
0.000,01
103 104 105 106 107 0. 8
0.0000,010
00, 05
ReD = UD/u 001

Fig. 6.4 Friction factor for duct flow (from Bejan [8]; drawn after Moody [7])

Another way of dealing with the subject of turbulent boundary layer is the
approach taking by (Kays and Crawford, page 173) [4].
Consider a simple model, with a viscous sublayer and mixing length governed
turbulent region. Inside the viscous sublayer, assume that viscosity dominates
turbulence, v εM. With zero pressure gradient, and noting that the sheer force
is constant in y, then Eq. 6.18 can be integrated to give
ð u ð
τ0 y
du ¼ dy ð6:39Þ
0 μ 0

or
uþ ¼ y þ ð6:40Þ

The effective thickness of the sublayer is about y+ ¼ 10.8 from experiment.

[email protected]
238 6 Turbulent Forced Convection

Outside the viscous sublayer, assume that turbulence dominates viscosity,


v  εM. Then use the Prandtl mixing-length theory, where the maximum value
of the fluctuation of the local velocity is assumed to be proportional to a mixing
length as,
∂
u
u0max ¼ ‘0 ð6:41Þ
∂y

and in the y -direction, with some local constant k,

∂
u
v0max ¼ k‘0 ð6:42Þ
∂y

Then the product provides, using a new length ‘ which includes the phase
constant for the time when u0max and v0max occur,

u0max v0max
2
∂
u
u0 υ0 ¼ ¼ ‘2 ð6:43Þ
2 ∂y

where we assume that the mixing length is proportional to the distance from the
wall, ‘ ¼ ky
 2  2
τ0 ∂
u ∂
u
¼ ‘2 ¼ k 2 y2 ð6:44Þ
ρ ∂y ∂y

or in nondimensional form,

duþ 1
¼ ð6:45Þ
dyþ kyþ

Integrating from u+ ¼ y+ ¼ 10.8, and setting k ¼ 0.41, we obtain the law of the
wall,

uþ ¼ 2:44 ln yþ þ 5:0 ð6:46Þ

Eq. 6.46 is in agreement with previous derivation that is presented in Eq. 6.30
and Fig. 6.3 as well. Considering again that the constants in Eq. 6.30 are chosen for
purpose of curve fitting that is plotted in Fig. 6.3 from experimental data and Von
Karman constant value is very much empirical.
Fig. 6.5 shows a comparison with experimental data. Now that the Van Driest
model does a better job in the transition region.
From this velocity distribution, it is possible to obtain the local friction coeffi-
cient [Kays and Crawford, page 174 (Eq. 6.47a) and page 175 (Eq. 6.47b)] [4].

[email protected]
6.5 Fully Developed Turbulent Flow in a Pipe 239

Fig. 6.5 Turbulent boundary layer profiles in wall coordinate [4]

cf
¼ 0:0287Re0:2
x ð6:47aÞ
2
cf 1=4
¼ 0:0125Reδ2 ð6:47bÞ
2

6.5 Fully Developed Turbulent Flow in a Pipe

In real world of fluid mechanics, that we deal with in practice, on daily basis, we
encounter with fluid flow in circular and noncircular pipe. We even see this
situation at our house, using our hot or cold water when we take a shower or
wash our hand and do cooking in our kitchen, whenever we need to use our water,
which is pumped through our piping network around the house.
City water distribution through city piping network is another example of flow in
pipes. Blood circulation in our body and arteries as well as blood veins, are also
presentation fluid flow through piping network and these type of example can go
on. Therefore, fluid flow can be classified as internal and external, depending on the
circumstance of fluid flow and if it forced to flow over a surface or in a pipe or
conduit.
Fluid flow whether it is internal or external flow does exhibit very different
characteristics. So far we have discussed the time averaging of the fluid equations
such as motion, the laminar sublayer and outer turbulent region and finally the

[email protected]
240 6 Turbulent Forced Convection

turbulent boundary layer and now we analyze the case of fully developed turbulent
flow in the pipe.
The terms pipe, duct, and conduit are usually used interchangeably for flow
sections. In general, flow sections of circular cross section are referred to as pipes
(especially when the fluid is a liquid), and flow sections of noncircular cross section
as ducts, especially when the fluid is a gas. Small diameter pipes are usually referred
to as tubes. Given this uncertainty, we will use more descriptive phrases such as a
circular pipe or a rectangular duct whenever it is necessary to avoid any
misunderstandings [10].
The flow of fluid such as gas and liquid through pipes or ducts is usually forced
to flow by a fan or pump through a flow section. As results of such applications, we
need to have a particular attention to friction, which is related to the pressure drop
and head loss in a direct way, during fluid flow through pipes and ducts.
For fully developed flow, u is function of y alone and the radial velocity υ ¼ 0,
the momentum equation becomes
 
11 d u 1 d p
r ð v þ εM Þ ¼ ð6:48Þ
r dr dr ρ dz

The experimental eddy viscosity outside the viscous sublayer is matched by:
 "  2 #
εM κyþ r r
¼ ¼ 1þ 1þ2 ð6:49Þ
v 6 r0 r0

where y ¼ r0  r. The experimental data for the velocity outside the viscous
sublayer is then given by Eq. 6.30 previously and is modified to the following form:
8 9
< 1:5ð1 þ r=r Þ =
uþ ¼ 2:5 ln h i yþ þ 5:5
0
ð6:50Þ
: 1 þ 2ðr=r Þ2 ;
0

Again, the values of 2.5 and 5.5 are curve-fitting constants.


Note the similar form to that for an external boundary layer. This can be
approximated for moderate Reynolds (104 < ReD < 5  104) numbers as

uþ ¼ 8:6yþ1=7 ð6:51Þ

Then at the centerline

uþ þ1=7
c ¼ 8:6yc ð6:52Þ

and
u
¼ 8:6yþ1=7
c ð6:53Þ
uc

[email protected]
6.5 Fully Developed Turbulent Flow in a Pipe 241

The mean velocity Vavg as defined before is then


ð
2 r0
V avg ¼ 2 urdr ð6:54Þ
r0 0

V avg
¼ 0:817 ð6:55Þ
uc

Compare this to a value of 0.5 for laminar flow.


Furthermore, Eq. 6.53 above can be written in terms of dimensional variables,
 1=7
uc uτ R
¼ 8:6 ð6:56Þ
uτ v
qffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
τ0
But uτ ¼ =ρ ¼ V avg cf =2 , with friction coefficient cf ¼ fc/4 and fc being the
friction factor, then substitution of it into Eq. 6.56 yields
ffi!1=7
pffiffiffiffiffiffiffi
RV avg f =2
c
uc
pffiffiffiffiffiffiffi
ffi ¼ 8:2 ð6:57Þ
V avg f =2
c v

With Eq. 6.55 and Reynolds number ReD ¼ DVavg/v in order, the Eq. 6.57
reduces to the following form:
cf
¼ 0:039Re0:25
D ð6:58Þ
2

This expression fits experimental data very well for 104 < ReD < 5  104.
For a wider range of the Reynolds number, it is better to use the logarithmic
equation, giving the Karman–Nikuradse relation as defined in Eq. 6.34 in terms of
friction coefficient cf ¼ fc/4, where fc is friction factor.
 qffiffiffiffiffiffiffiffi
1
pffiffiffiffiffiffiffiffi ¼ 2:46 ln Re D
cf
=2 þ 0:30 ð6:59Þ
cf
=2

This is a bit of awkward situation, and thus the entire range of Reynolds number
is commonly fit by the following expression [11],
64

f ¼ ReD 0 < ReD < 2000 ð6:60Þ
.
Re0:33
f ¼ D 2000 < ReD < 4000 ð6:61Þ
381

f ¼ 0:316Re0:25
D 4000 < ReD < 30, 000 ð6:62Þ
f ¼ 0:184Re0:2
D 30, 000 < ReD < 10 6
ð6:63Þ

where again f ¼ 4cf.

[email protected]
242 6 Turbulent Forced Convection

6.6 Turbulent Flow in Other Cross-Sectional Shape

Because turbulent flow velocity change is confined near the wall, the shape of the
tube cross section has little effect on shear stress at the wall (except in sharp
corners). Thus, Eqs. Eq. 6.62 andEq. 6.63 can be used in noncircular tubes, using
the hydraulic diameter,
flow cross sectional area Ac
Dh ¼ 4r h ¼ 4 ¼4 ð6:64Þ
wetted perimeter P

For shapes such as squares, rectangular or annular ducts where the height and
width are comparable, the characteristical dimension for internal flow situations is
taken to be the hydraulic diameter, Dh, defined as Eq. 6.64.
In Eq. 6.64, Ac is the pipe cross-sectional area, while P is the wetted perimeter
and by definition, is perimeter of the cross-sectional area that is “wet,” as shown in
Fig. 6.6.
The wetted perimeter can be defined mathematically as
X
1
P¼ li ð6:65Þ
i¼0

where li is the length of each surface in contact with the aqueous body. Note that
in open channel flow, the wetted perimeter is defined as the surface of the channel
bottom and sides in direct contact with the aqueous body.
Laminar flow cannot be treated using the hydraulic diameter. For flow in tube
banks, special correlations are available as will be discussed later.
In case of flow in pipe or tube, the Reynolds number is generally defined as:
ρV avg Dh V avg Dh QDh
Re ¼ ¼ ¼ ð6:66Þ
μ v vAc

where:
Dh: is the hydraulic diameter of the pipe; its characteristic travelled length, L, (m)
Q: is the volumetric or discharge (Q ¼ AcVavg) flow rate (m3/s)
Ac: is the pipe cross-sectional area (m2)
Vavg: is the mean velocity of the fluid (m/s)
μ: is the dynamic viscosity of the fluid (Pa s ¼ N s/m2 ¼ kg/(m s))
v: is the kinematic viscosity (v ¼ μ/ρ)(m2/s)
ρ: is the density of the fluid (kg/m3)

Fig. 6.6 Cross-sectional


area of a trapezoidal open
channel, red highlights
wetted perimeter

[email protected]
6.6 Turbulent Flow in Other Cross-Sectional Shape 243

Note that for uniform flow (no changes in depth for selected length or cross-
section of the channel), Q is referred to as Normal Discharge. The value of Q for
different open channel type with different materials can be determined by the
following equation:
1:0 2=3
Q¼ Ac r h S1=2 ð6:67Þ
n

where S is slope drop of channel and n is the material type that channel made of.
Example 6.1: Using Fig. 6.7, determine normal discharge (uniform flow) for a
200 mm inside diameter common clay (n ¼ 0.013) drainage tile running half-full if
the slope drops 1 m over 1000 meters.
Solution: The slope of channel can be determined as follows:

S ¼ ð1=1000Þ ¼ 0:001
  n o
and Ac ¼ ð1=2Þ  πD2 =4 ¼ 0:5  π  ð0:2Þ2 =4 ¼ 0:0157 m2
wet perimeter P ¼ ð1=2Þ  ðπDÞ ¼ 0:5  π  0:2 ¼ 0:3141 m
hydraulic radius r h ¼ 0:05m

and for common clay (n ¼ 0.013), substituting all these values in Eq. 6.65, we
get the following answer:

Q ¼ ð1:0=0:013Þ  ð0:0157Þ  ð0:05Þ2=3 ð0:001Þ1=2


Q ¼ 5:18  103 m3 =s

Fig. 6.7 Circular drain tile


running half-full

D/2

[email protected]
244 6 Turbulent Forced Convection

Further study of Eq. 6.65 reveals that, conveyance and most efficient channel
shapes come from by looking at the RHS of this equation. Other than the S term, all
other terms are related to channel cross-sectional and its features.
These terms together are referred to as the Conveyance (K ) of the channel.

1:0 2=3
K¼ Ac r h ð6:68Þ
n

With this definition Eq. 6.67 reduces to Eq. 6.69 as follows:

Q ¼ KS1=2 ð6:69Þ

K is maximum when wet perimeter P is at a minimum for a given area. This is


also the most efficient cross section for conveying flow. For circular section, half-
full is the most efficient and for other shapes, see Table 6.1 below.
All types of open channel shapes are illustrated in Fig. 6.8.
In addition, all types of open channel flows are defined as follows and some
scenarios are depicted in Fig. 6.9.
1. Steady Flow: when discharge Q does not change with time
2. Uniform Flow: when depth of fluid does change for a selected length or section
of channel
3. Uniform Steady Flow: when discharge does change with time and depth
remains constant for a selected section. Note that cross section should remain
unchanged—referred to as a prismatic channel
4. Varied Steady Flow: when depth changes but discharge remains the same
5. Varied Unsteady Flow: when both depth and discharge change along a channel
length of interest
6. Rapidly Varying Flow: depth is rapid
7. Gradually Varying Flow: depth change is gradual

Section 1—Rapidly varying flow


Section 2—Gradually varying flow
Section 3—Hydraulic jump
Section 4—Weir and waterfall
Section 5—Gradually varying
Section 6—Hydraulic drop due to change in channel slope

[email protected]
6.7 Effects of Surface Roughness 245

Table 6.1 Other shapes


Section Area A Wetted Perimeter WP Hydraulic Radius R
Rectangle 2.0y2 4y y/2
(half of a square)
b = 2y = T

Triangle y2 2.83y 0.354y


(half of a square)
T = 2y

1 1 y
z= 1 z= 1
45° 45°

Trapezoid 1.73y2 3.46y y/2


(half of a hexagon)
T = 2.309y

1 1
b
L=

= 0.577 y z = 0.577

60° 60°

b= 1.155y

Semicircle 2 πy
1 2 πy y/2

T = 2y

D = 2y

6.7 Effects of Surface Roughness

Many researchers using surface roughness factors such as the Wenzel roughness
factor and so on have studied the effects of surface roughness on wettability.
However, these factors depend on each other, which make a precise discussion
difficult.

[email protected]
246 6 Turbulent Forced Convection

Fig. 6.8 Different type of open channels shape

The effect of roughness of a solid surface on its wettability by a liquid has


been studied theoretically using mechanistic arguments. Nikuradse [9] did the
original work, using uniform sand paper grains to roughen pipes. From the
equivalent sand grain, roughness size ks a nondimensional roughness Reynolds
number is obtained:
pffiffiffiffiffiffiffi
uτ ks Re f =8
Rek ¼ ¼ ð6:70Þ
v D=ks

[email protected]
6.7 Effects of Surface Roughness 247

gate
Hydraulic jump

Reservoir Weir

Hydraulic drop

1 2 3

4 5

Fig. 6.9 Various flows configuration

0.07
Critical Transition
zone zone
Complete turbulence, rough pipes
0.03
0.05 0.02
0.015
0.04
0.010

Relative Roughness D
0.006
0.03
Friction factor f

f = 64/Re 0.004
Laminar-
circular 0.002
pipes only
0.02 0.001
0.0006

λ
0.0004
0.015
λ, cm 0.0002
Concrete 0.03-0.3
Cast iron 0.026 0.0001
Galvanized iron 0.015 Smooth
Commercial steel 0.0045 pipes 0.000 05
0.010 Drawn tubing 0.00015

0.008 0.000 01
3 2 4 6 4 2 4 6 5 2 4 6 6 2 4 6 7 2 4 6 8
10 10 10 10 10 10
Re = UD/ν
 
p0 p
Fig. 6.10 Moody chart for friction factors f. Here f ¼ 4cf ¼ DL ρU2 =2

Rek < 5.0 Perfectly smooth surface (all above boundary layer defined before and pipe
flow equations will work.
5.0 < Rek < 70 Transitional roughness takes place and we can use Moody Chart (1944)
depicted in Fig. 6.10.
Rek > 70 Fully rough case is the one, where friction coefficient cf becomes indepen-
dent of Reynolds number and viscosity is no longer a significant variable.

[email protected]
248 6 Turbulent Forced Convection

8
fc ¼ ð6:71Þ
½2:46 ln ðD=ks Þ þ 3:222

Normally, friction factors are extracted from Moody chart for any turbulent flow
in pipes, which includes the effects of surface roughness as well.
In summary, we can mention few bullet points here as follows:
• Surface topography includes roughness, waviness, and form.
• Surface roughness is measured by profilometry, optical interferometers, and
AFM (Atomic Force Microscope, 1986) [12].
• Surface roughness has influence on the real contact area and type of surface
deformation (elastic/plastic—plasticity index).
• Surface roughness influences all of the main causes of friction: adhesion,
abrasion, deformation of asperities.

6.8 Numerical Modeling of Turbulence

What we have covered and learned in previous section so far is adequate for plane
or pipe flows, but it is not enough for large free stream turbulence, large surface
curvature, separation regions, and three-dimensional flows. In order to deal with
circumstance as such, we need to develop higher-order turbulence theories, using
one or more additional partial differential equations. Therefore, finding a numerical
modeling approach for turbulent flow has been ongoing goal for most fluid dyna-
mists for nearly a century.
It is believed that the Navier–Stokes equations can be used to fully describe
turbulent flows, but current limitations in computational horsepower have made the
direct solution of the Navier–Stokes equations impractical for all but very simple
flows at low Reynolds numbers. This is because current computers do not allow for
the resolution of the wide range of length and time scales associated with turbu-
lence. Many complex fluid dynamic applications are directed at determining time-
averaged quantities, and hence it is desirable to find a means to obtain these mean
quantities short of solving the full unsteady Navier–Stokes equations for all of the
length and time scales associated with the turbulence [13].
Some basic knowledge of turbulence and an understanding of how turbulence
models are developed can help provide insight into choosing and applying these
models to obtain reasonable engineering simulations of turbulent flows. This effort
is directed at production users of Computational Fluid Dynamics (CFD). The
attempts provide the basic information required to choose and use currently avail-
able turbulence modeling techniques [13].
As computers developed and numerical simulation evolved differential equation
based transport type turbulence models became the turbulence simulation method-
ology of choice. It should always be remembered that transport models are

[email protected]
6.9 Friction Factors 249

empirically calibrated. The use of transport type turbulence models has become
standard practice for most engineering applications. Many current researchers are
now solving the unsteady Navier–Stokes equations for large-scale, or grid realized,
turbulence and modeling the smaller, or sub-grid, turbulent scales that cannot be
captured on the computational grid [13].
A turbulent kinetic energy equation can be applied to describe the production,
convection, diffusion, and dissipation of kinetic energy associated with fluctuating
quantities u’, υ’. (The equation has a form similar to the energy conservation
equation.) The eddy diffusivity is then determined from an assumed relationship
between eddy diffusivity and turbulent energy. The models can be listed in order of
complexity:
1. One Equation Models use the turbulent kinetic energy equation with an empir-
ical relation for length scale similar to the mixing length relation as it was
covered in previous sections.
2. Two Equation Models also having another partial differential equation for local
values of the length scale can be derived from Navier Stokes.
3. Stress Equation Models require solutions of Partial Differential Equations
(PDEs) for all the components of the turbulent stress tensor.
4. Largely Eddy Simulation attempts to capture large scale eddies, using a lower-
level model for small-scale turbulence.
Although the ultimate general-purpose turbulence model has yet to be
established and developed, turbulence modeling has matured to the point that
reasonably accurate results can be obtained for a wide range of engineering
applications with the current class of computers. As computer, technology con-
tinues to improve both the role and the form of turbulence models will continue to
evolve.

6.9 Friction Factors

If we consider a steady driven flow with constant density, there are two systems of
kind that can be studied [14].
(a) The fluid flows in a straight conduit of uniform cross section.
(b) The fluid flows around a submerged object that has a symmetry axis, or two
planes of symmetry parallel to the direction of the approaching fluid.
For channel flows, one often wants to calculate the pressure drop ΔP for given
volume flow rate. For submerged bodies, predictions of the drag force for a given
velocity are often needed. One can express the force on body Fk or the pressure drop
in a channel Fk/A as the product of a characteristic area A, characteristic kinetic
energy per unit volume K, and dimensionless friction factor fc,

[email protected]
250 6 Turbulent Forced Convection

AKf
Fk ¼ ð6:72Þ
4

This choice often yields convenient empirical formulations as it is commonly


found from experimental data that fc is a simple function of geometry and Reynolds
number.
For both types of systems (a) and (b), Eq. 6.70 can be used, and clearly, for any
given flow system, fc is not defined until A and K are specified.

6.10 Flow in Conduits

Further analysis of Eq. 6.70 for the flow in conduits indicates that characteristic area
A is wetted surface, while K is taken to be 12 ρhυi2 . In particular, for circular tubes of
radius R and length L, the Eq. 6.70 yields a new form for friction factor fc as follows:
 
1
Fk ¼ ð2πRLÞ ρhυi2 f c ð6:73Þ
2

Furthermore, a balance force on the fluid between 0 and L in direction of a flow


that is fully developed is measured based on quantity of the pressure difference
p0  pL and the elevation h0  hL. In this case, a new form of Eq. 6.71 can be
produced in the following format:

Fk ¼ ½ðp0  pL Þ þ ρgðh0  hL ÞπR2


ð6:74Þ
¼ ðP0  PL ÞπR2

If we eliminate F between Eq. 6.71 and Eq. 6.74, then we obtain [14]
  !
1 D P0  PL
fc ¼ ð6:75Þ
2ρ < υ>
4 L 1 2

In this equation form, the symbol fc is sometimes called the Fanning friction
factor2 and D ¼ 2R is the tube diameter.
Friction factor for flow in tube defined by Eqs. Eq. 6.71 and Eq. 6.72 is depicted
in Fig. 6.11 as follows:

2
This friction factor definition is due to J. T. Fanning, A Practical Treatise on Hydraulic and
Water-Supply Engineering, Van Nostrand, New York, 1st edition (1877), 16th edition (1906); the
name “Fanning” is used to avoid confusion with the “Moody friction factor,” which is larger by a
factor of 4 than the f used here [L. F. Moody, Trans. ASME, 66, 671–684 (1944)

[email protected]
6.11 Flow Around Submerged Objects 251

1.0

0.5

0.2

0.1
Friction factor f

La
0.05 m
in
f = 16 ar
Re
0.02
Turbulent
0.01 k/D = 0.004
U f = 0.07 0.001
0.005 su 91
al Re 1/4 0.0004
ly
un "Hydrau 0.0001
sta lically sm
0.002 bl ooth"
e
0.001
102 103 104 105 106 107
Reynolds number Re = D<u> r/m

Fig. 6.11 Curves of L. F. Moody, Trans. ASME, 66,671–684 (1944) as presented in W. L.


McCabe and J. C. Smith, Unit Operations of Chemical Engineering, McGraw-Hill, New York
(1954)

6.11 Flow Around Submerged Objects

In this case, the characteristic area A is obtained by projecting the solid onto a plane
perpendicular to the velocity of the fluid and form of Eq. 6.70 will be established as
 
  1 2
Fk ¼ πR2 ρυ1 f c ð6:76Þ
2

In this case, K is considered to be 12 ρυ21 for flow around a sphere of radius R,


where υ1 is the approach velocity of the fluid at a large distance from the
submerged objects and fc is the friction factor that is presented by Eq. 6.76.
For the steady state of falling sphere through the flow, the approach velocity of the
fluid υ1 reaches to the terminal velocity υ of the sphere and therefore Fk is equal to
the counter-balanced for the gravitational and buoyant force and can be written as:
4 4
Fk ¼ πR3 gρsphere  πR3 g ρfluid ð6:77Þ
3 3

Elimination of Fk between Eqs. Eq. 6.76 and Eq. 6.77 results in [14]
 
4 gD ρsphere  ρfluid
fc ¼ 2 ð6:78Þ
3 υ1 ρfluid

[email protected]
252 6 Turbulent Forced Convection

Equation Eq. 6.78 allows one to obtain friction force f from terminal velocity
data, which also sometimes is called the drag coefficient and is symbolized ascD
Therefore, from the experiment perspective, and the Eq. 6.76, the drag force
resulting from fluid flow can be predicted by
 
  1
Fk ¼ cD πR2 ρh υ i 2 f c ð6:79Þ
2

where the drag coefficient is equal to cD ¼ fc/4, and yet our previous scaling
showed that fc ¼ fc(Re) for a smooth sphere.
Our previous analysis for Stokes flow, or creeping flow, showed that
Fs ¼ 6πμRυ1 ð6:80Þ

Now if we substitute Eq. 6.80 into Eq. 6.79, then we have


 
24  2  1
Fs ¼   πR ρhυ1 i2 ð6:81Þ
Dυ1 ρ 2
μ

or
24
cD ¼ ð6:82Þ
ReD

At higher ReD the friction factor has the behavior shown in Fig. 6.12.

SPHERE DRAG COEFFICIENT


104
SEPARATION

TRANSITION
TURBULENT

103

102
NEWTON'S
CD CD=24/ReD LAW
10

STOKE'S 3/5
1 LAW CD=18.5/ReD
CD=0.44
0.1
10-3 10-2 10-1 1 10 102 103 104 105 106
ReD

Fig. 6.12 Drag coefficients for sphere

[email protected]
6.13 Flow in Pipes and Ducts 253

6.12 Turbulent Flow in Noncircular Tubes

It is common to use the following empirical procedure: First we define a “mean


hydraulic radius” Rh as follows:

S
Rh ¼ ð6:83Þ
Z

In this equation, S is the cross section of the conduit and Z is the wetted
perimeter. Then we can use Eq. 6.75 along with Fig. 6.11 and replacing diameter
D of circular pipe with 4Rh we can obtain the following form of friction factor
equation:
  !
Rk P0  P L
fc ¼ 2
ð6:84Þ
2 ρhυz i
L 1

And obtaining f from Fig. 6.11 with a Reynolds number defined as

4Rh < υz > ρ


Reh ¼ ð6:85Þ
μ

For laminar flows in noncircular passages, this method is less satisfactory [14].

6.13 Flow in Pipes and Ducts

Consider a control volume around a pipe or a section of duct, as shown in Fig. 6.13.
We can apply the equation for the conservation of mechanical energy to this control
volume. This is often referred to as the Bernoulli Equation, and it can be written as
following forms at the starting point for flows in pipes or ducts.

p1 V 21
þ þ gz1 ¼ constant ð6:86aÞ
ρ 2

Fig. 6.13 Control volume


for flow in a pipe or duct

[email protected]
254 6 Turbulent Forced Convection

p1 V 21 p V2
þ þ gz1 ¼ 2 þ 2 þ gz2 ð6:86bÞ
ρ 2 ρ 2

In general form, we can write the two above equations in the following format
and the factors a1 and a2 relate the actual kinetic energy to the average velocity Vavg
ð p1 , υ2
a2 V 22  a1 V 21 dp
þ gð z 2  z 1 Þ þ þWþF¼0 ð6:87cÞ
2 p2 , υ1 ρ

Each term in the Bernoulli equation has the dimensions of energy per unit mass.
The terms on the left-hand side (LHS) are the pressure heads, the velocity head, and
the elevation head. The term “head” arises because, for a static flow, V ¼ 0, and the
equation can be arranged, under certain circumstances, to give

p ¼ ρgz ð6:88Þ

This gives the pressure at a point z below the surface of a liquid. The equation
states that the sum of these terms is constant along a streamline. The right-hand side
(RHS) of the equation says that the sum of the terms is constant at any two points
(points 1 and 2) along the streamline. The Bernoulli equation may be derived quite
simply from the steady flow energy equation (SFEE) used in thermodynamics and
Q being the heat energy, while the u is presentation of enthalpy (internal energy)
and m is mass flow.

1  
mQ  W ¼ mðu2  u1 Þ þ mgðz2  z1 Þ þ m V 22  V 21 þ mðp2 V 2  p1 V 1 Þ
2
ð6:89Þ

By simply knocking out the thermodynamic properties, or the terms associated


with the non-flow energy equation,

W ¼ mðu2  u1  QÞ ð6:90Þ

So another way of establishing the Bernoulli equation very similar to Eq. 6.86a is
by knocking out the time-dependent term and the shear stress term from the Navier–
Stokes equation, in the tensor form

~
∂V h i
∇  τ  ∇ðPÞ þ ρ~
g¼ρ þρ V~  ∇V
~ ð6:91Þ
∂t

Knocking out the shear stress and viscous terms gives


h i
0¼ρ V~  ∇~
V þ ∇ðPÞ  ρ~
g ð6:92Þ

[email protected]
6.13 Flow in Pipes and Ducts 255

And integrating with respect to z, taking into account the proper physical
direction for the gravity vector ~g, gives out Eq. 6.86a
In order to determine the pressure differential across a pipe work or ducting
system, we use the Bernoulli equation derived from the steady flow energy equation
but retain terms for the internal energy and some small amount of heat dissipation.
The Bernoulli equation now alters to

p1 V 21 p V2
þ þ gz1 ¼ 2 þ 2 þ gz2 þ u2  u1  Q ð6:93Þ
ρ 2 ρ 2

In other words, the viscous shear stresses in the fluid reduce the pressure and
increase the fluids internal energy, u2  u1. Some of this heat energy is dissipated to
the surroundings, Q. If we link in with this last equation the work done per unit
mass, w, by pump driving the system then straight away we are back to the SFEE.
However, we are not particularly interested in the thermodynamics of the problem;
we are more interested in determining the pressure losses in the pipe.
The “Head loss” due to friction in the pipe, or system, is usually determined by
simply measuring the velocity, elevation, and pressure differences across the
system. Hence, head loss, Δh, is given by:
   
p1 V 21 p2 V 22
Δh ¼ þ þ z1  þ þ z2 ð6:94Þ
ρg 2g ρg 2g

And the “Pressure loss”, by


   
V 21 V2
Δp ¼ p1 þ ρ þ ρgz1  p1 þ ρ 2 þ ρgz2 ð6:95Þ
2 2

Note that the more turbulent the flow then the greater the head loss. Turbulence
increases with increasing velocity, c. Head losses in turbulent flows are proportional
to velocity. A “Loss coefficient” for a pipe, duct, or component is defined as

Δh Δp
K¼ V2
¼ ρV 2
ð6:96Þ
2g 2

This is a dimensionless parameter. For valves and angle fittings the velocity, V,
is taken as the inlet velocity. For combining and dividing fittings, such as “T” and
“Y” pieces, the velocity is taken as the velocity in that leg where the flow is
combined. For all fittings, the loss coefficient, K, is obtained from charts or tables,
which have been compiled by the manufacturers. These charts and tables are made
up from the results of simple experimental measurements on the fittings. They are
therefore derived “empirically.” These tables and charts are available from the book
by D. S. Millar “Internal Flow Systems,” of the British Hydraulic Research
Association (BHRA), and the ASHRAE (American Society of Heating,

[email protected]
256 6 Turbulent Forced Convection

Dh
V2
2g
friction gradient, f = Dh

non-dimensional head
V2 L
2g D

non-dimensional length (L / D)

Fig. 6.14 Nondimensional head versus lengths

Refrigeration and Air-conditioning Engineers) catalogues. Examples of these charts


and tables are given presently.
Clearly for pipes it would be impractical to tabulate the loss coefficient, K, for
every length, diameter, and roughness category. In this case, it is easier to define a
“Friction coefficient”,

L V2 L V2
Δh ¼ f or Δp ¼ f ρ ð6:97Þ
D 2g D 2

The friction coefficient is the ratio of the head loss, in a one-diameter length of
pipe, to the mean velocity head. (The velocity in a pipe actually varies as a function
of the radius, r.) This idea is illustrated below as Fig. 6.14.
The friction factor for pipes, f, may be determined either from equations or,
again, from a chart such as Moody in Figs. 6.10 and 6.11. For Reynolds number we
can write

VDρ VD
ReD ¼ ¼ ð6:98Þ
μ v

In addition, Henry Darcy (1803–1858) determined that for laminar flow in pipes,
friction factor fc is given by:

64
f ¼ ð6:99Þ
ReD

Furthermore, in 1939, C. F. Colebrook determined that for turbulent flow in


pipes, the friction factfc is given by:
!
1 ε=D 2:51
pffiffiffi ¼ 2:0log10 þ pffiffiffiffi ð6:100Þ
f 3:7 ReD f c

[email protected]
6.13 Flow in Pipes and Ducts 257

In these equations D is pipe diameter, while ε is the pipe roughness in mm in


Moody chart.
This equation is cumbersome to evaluate as it contains fc on both the left- and
right-hand sides and can therefore only be solved by iterative means. An alternative,
although slightly less accurate, formula was given by S. E. Haaland in 1983
  !
1 ε=D 1:11 6:9
pffiffiffi ¼ 1:8log10 þ ð6:101Þ
f 3:7 ReD

To explain the flow in pipes and ducts, the Bernoulli equation provides a balance
between kinetic energy, potential energy, flow work, shaft work, and friction work.
Often the density of the fluid does not vary significantly over the control volume,
and it is valid to replace the density with an average value, making the third term
( p2  p1)/ρavg.
The kinetic energy entering and leaving the control volume depends on the
velocity distribution across the inflow or outflow area. Again, the factors a1 and a2
relate the actual kinetic energy to the average velocityVavg.
ð
1 υ
aV 2avg ¼ v~
~ ndA ð6:102Þ
AV avg A 2gc
ð
1
V avg ¼ v~
~ ndA ð6:103Þ
A A

Table 6.2 provides typical values of the factor a for different flow conditions.
For ducts, the friction force F is related to the wall shear, and is given by the
friction factor f ¼ 4fc (note that f ¼ fTODERAS/KAZIMI ¼ 4fBSL ¼ 4cf kays/crawford).
       
f PL 1 2 f L 1 2
F¼ ρV ¼ ρV ð6:104Þ
4 4Ac 2 avg 4 Dh 2 avg

where Ac is the cross-sectional area of the duct, P is the perimeter of the duct, and
Dh ¼ 4Ac/P is the hydraulic diameter.

Table 6.2 Correction factor a, Eq. 6.99, relating kinetic energy to the average velocity in a
channel, from Bonilla [15]
Annulus R2/R1 Wide Flat Square
Pipe 3 2 1.2 Channel Channel
Laminar flow 2 1.568 1.552 1.543 1.543 2.2
Turbulent flow f
0.04 1.135
0.02 1.073 1.055
0.01 1.040 1.022

[email protected]
258 6 Turbulent Forced Convection

Dimensional analysis shows that [16]


f ¼ f ðRe; L=Dh ; ks =Dh Þ ð6:105Þ

for long pipes, the velocity distribution becomes fully developed, and fc is
independent of L/Dh.
Tabular data on friction pressure drop for air and water is commonly used in
design calculations by engineers and can be found in handbooks such as Cameron
Hydraulic Data [17] and Crane Technical Paper No. 410 [18].
Pressure drop in entrances, exits, valves, and fittings can be calculated in terms
of either a form loss factorK.
 
1 2
Δp ¼ K ρV avg ð6:106Þ
2

or as an equivalent length of pipe, with values of Lequivalent/D being specified for


a given fitting.
Figure 6.15 gives approximate values for a variety of fittings. It is often difficult
to specify accurate loss factor K values, as they depend strongly on the specific
valve or fitting geometry, and on whether the fittings are closely spaced or widely
separated.

All pipe sizes


Fitting Description K value Fitting L/D
Pipe exit projecting 1.0 Close Return Bend 50
sharp edged
rounded
Pipe entrance Inward 0.78
Standard Tee thru flow 20
projecting
0.50 thru
Pipe entrance flush sharp edged 60
branch
r r/d = 0.02 0.28 30
90° Standard elbow
r/d = 0.04 0.24
d
r/d = 0.06 0.15
0.09 45° Standard Elbow 16
r/d = 0.10
r/d = 0.15 and up 0.04

Ball valve Gate valve Angle valve 90°Bends, pipe bends, r/d = 1 20
flanged/butt welded r/d = 2 12
L/D−3 L/D−8 L/D = 150
elbows r/d = 3 12
Butterfly valve r/d = 4 14
L/D = 45 (2 to 8") r/d = 6 17
L/D = 35 (10 to 14") r/d = 8 24
L/D = 25 (16 to 24") r/d = 10 30
r/d = 12 34
Globe valve Swing check valve r/d = 14 38
L/D = 340 L/D = 100 L/D = 50 r/d = 16 42
r/d = 18 46
r/d = 20 50

From Crane Co. Technical Paper 410

Fig. 6.15 Loss coefficient and equivalent lengths for several fittings [17, 18]

[email protected]
6.13 Flow in Pipes and Ducts 259

With the expressions for friction work, the mechanical energy balance takes its
most useful form,

a2 V 22  a1 V 21 1 X  L V2
þgðz2  z1 Þ þ ðp2  p1 Þ þ W þ f
2 ρ i
D 2 i
|fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl}
ðsum on all straight sectionsÞ
X  V2
þ K ¼0 Eq:6:104
i
2 i
|fflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflffl}
ðsum on all fitting;valves and elbowsÞ

With the Bernoulli equation and friction factors and form loss coefficients, the
flow and pressure drop through complex piping systems can be evaluated.
Pipe Networks
• The total flow into any junction must equal the flow out.
• The flow in each pipe must satisfy the Bernoulli equation.
• The algebraic sum of the heat loss around any closed circuit must be zero.
Example 6.2: Pumping Head Requirement for a Simple System What is the
required pump capacity to maintain 300 gpm flow in a system with a cooling
tower and heat exchanger. Consider a system with 200 ft of 4 in. Schedule
40 pipe, 15 90 elbows, 4 gate valves, and a heat exchanger with 15–10 ft long
parallel 1 in. Schedule 40 pipes, with Δh ¼ 8 ft. Use Fig. 6.16 below.
Solution: Pressure drop data can be found in tabulated form for the flow of water
through standard diameter pipe, from sources such as the Cameron Hydraulic Data
[17] and Crane [18] handbooks.

Fig. 6.16 Diagram for Example 6.2

[email protected]
260 6 Turbulent Forced Convection

Here we show the calculation for the pressure drops. First the Reynolds numbers
are determined using fluid properties and pipe diameters from the appendix,
  
4ð300 gal=minÞ 0:1337 ft3 =gal 144 in:2 =ft2
V 400 ¼ ¼ 7:56 ft=s
π ð4:02 in:Þ2 ð60 s=minÞ
  
4ð20 gal=minÞ 0:1337 ft3 =gal 144 in:2 =ft2
V 100 ¼ ¼ 7:44 ft=s
π ð1:049 in:Þ2 ð60 s=minÞ

ρVD ð998 kg=m3 Þð7:56 ft=sÞð4:026 in:Þð12 in:=ftÞð0:0254 m=in:Þ2


Re400 ¼ ¼  ¼ 234500
μ 1:002  103 Ns=m2

Re100 ¼ 60000

Next friction factors are selected from the Moody chart, Fig. 6.10, for relative
wall roughnesses of λ=D400 ¼ (0.0045 cm)/(4.026 in.)(2.54 cm/in.) ¼ 0.00044 and
λ=D100 ¼ 0:00011,

f c400 ¼ 0:0185 f c100 ¼ 0:0205

Then the respective pressure drops, per 100 ft of pipe, are

ð100 ftÞð12 in:=ftÞ ð7:56 ft=sÞ2


ΔH400 =100 ft ¼ ð0:0185Þ ¼ 4:89
ð4:026 in:Þ 2ð32:2 ft=sÞ2
ð100 ftÞð12 in:=ftÞ ð7:44 ft=sÞ2
ΔH 100 =100 ft ¼ ð0:021Þ ¼ 20:65
ð1:049 in:Þ 2ð32:2 ft=sÞ2

Next, we set up a table using information from Fig. 6.15:

ΔP/length Pressure
Quantity Component L/D Equivalent length (ft) (ft/100 ft) drop (ft)
400 Sch. 40 pipe 2000
15 90 400 elbows 14 15  0.330  14 ¼ 700
4 Gate valves 8 4  0.330  8 ¼ 110
Total 400 pipe 2810 4.89 13.70
100 Sch. 40 pipe 100 20.65 2.10
Total elevation change 8.00
Entrance and exit losses 2.60
Total head loss 26.40

where entrance and exit losses are found from tabular information for the fluid
velocity.

[email protected]
6.13 Flow in Pipes and Ducts 261

400 pipe entrance and exit:

1 ð7:56 ft=sÞ2
ð0:5 þ 1:0Þ ¼ 1:330
2 32:2 ft=s2

100 pipe entrance and exit:

1 ð7:44 ft=sÞ2
ð0:5 þ 1:0Þ ¼ 1:280
2 32:2 ft=s

Example 6.3: Transient Emptying of a Tank Derive an expression for the emptying
rate of a tank and the time required to empty the tank from a level zs(0). Assume
Rt < < Rp (velocity in the tank is small) and that the pipe remains primed
(no reverse flow of air). Use Fig. 6.17 below.
Solution: Starting with the Bernoulli equation and noting that the pressure is equal
at the top and bottom, and then we have
   
1 2   zp 1 2 1 2
ρV þ ρg zp þ zs þ f ρV þ ðK in þ K out Þ ρV p ¼ 0
2 ρ 2Rp 2 p 2
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  ffi
u
u 2g z þ z
Vp ¼ t
p s
z
1 þ f 2Rpp þ K in þ K out

Fig. 6.17 Diagram for


Example 6.3

[email protected]
262 6 Turbulent Forced Convection

Mass Balance:

dzs
q ¼ πR2t ¼ πR2p V p
dt
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  ffi
u
dzs R2p u 2g z þ z
¼ 2t
p s
z
dt Rt 1 þ f 2Rpp þ K in þ K out

ð0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð
1 R2p 2g ts
pffiffiffiffiffiffiffiffiffiffiffiffiffi dzs ¼  2 zp dt
zs zp þ zs Rt 1 þ f 2Rp þ K in þ K out 0
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi R2p 2g
2 zp  zp þ zs ¼ ts 2 zp
Rt 1 þ f 2Rp þ K in þ K out
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
z
pffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiR2p 1 þ f 2Rpp þ K in þ K out
ts ¼ 2 zp þ zs  zp 2
Rt 2g

6.14 Flow in Rod Bundles

The fuel element geometry most frequently used in nuclear reactors is the rod bundle.
Most nuclear fuel rods are arranged into arrays in either square or triangular pitched
patterns with coolant flowing axially through the subchannels formed between the
rods. The mixing of cooling fluid in a rod bundle through the gaps between the rods
reduces the temperature differences in the coolant as well as along the perimeter of
the rods. The prediction of the temperature distribution in the coolant and along the
rod perimeter is of major importance in nuclear reactor safety and design.
Flow in rod bundles has many similarities with flow in porous media, discussed
in Chap. 5. However, flows in rod bundles are commonly turbulent, and thus the
pressure drop varies with the square of the velocity, rather than linearly with
velocity as occurs for laminar porous media flows.
To ensure good thermal performance of a nuclear reactor, a detailed knowledge
of the heat transfer and fluid flow phenomena taking place within the core is
required. Coolant flow rates in different parts of the complex rod bundles and the
manner in which the single-phase and the two-phase flows are distributed in the
subchannels are very important for evaluating enthalpy distribution and perfor-
mance parameters, such as the onset of boiling and critical heat flux.
The subchannel analysis method is an important tool for predicting the thermal
hydraulic performance of rod bundle nuclear fuel element. It considers a rod bundle
to be a continuously interconnected set of parallel flow subchannels which are
assumed to contain one-dimensional flow and are coupled to each other by cross
flow mixing; the axial length is divided into a number of increments such that the
whole flow space of a rod bundle is divided into a number of nodes.

[email protected]
6.14 Flow in Rod Bundles 263

The principle of subchannel analysis is the application of continuity and conser-


vation equations to the flow between these nodes. The conservation equations relate
the local variations of velocity and enthalpy of each node to those of its neighboring
ones. The relation between subchannel flow rate, which is the mass flow rate in an
axial direction through subchannel area, and diversion cross-flow, which is the mass
flow in a transverse direction resulting from local pressure differences between two
subchannels, is strongly governed by momentum balance in a transverse direction.
For most nuclear reactors, rod diameters vary from 0.5 to 1.3 cm, with spacing
from 0.5 mm to 0.5 cm and lengths from 30 to 500 cm. The rods can be grouped into
triangular or rectangular arrays, where Fig. 6.18 shows the characteristic dimen-
sions are the rod diameter D and pitch P.
Spacers, shown in Fig. 6.19, are commonly used to separate the rods uniformly
and prevent bowing. These spacers can consist of wire spiraled around individual
rods, or honeycomb grid spacers.

Fig. 6.18 Rod bundle configuration

Fig. 6.19 Typical rod bundle spacer design

[email protected]
264 6 Turbulent Forced Convection

6.15 Flow Parallel to Rod Bundles

The pressure drop occurring with single-phase flow parallel to rod bundles can be
determined by using appropriate hydraulic diameters and friction factors, and by
accounting for the effects of spacers. The channel cross-sectional area can be
broken into individual, geometrically similar subchannels, as seen in Fig. 6.18.
Reduced flow occurs in subchannel types II and III next to shrouds, due to the larger
ratio of wetted perimeter to flow area. The enthalpy rise in these shroud subchannels
can be larger than the average enthalpy rise in the bundle. However, interchannel
mixing can reduce this effect and help maintain uniform temperature distribution
along the rod bundle.
For laminar flow parallel to rod bundles, consult Todreas and Kazimi,
pg. 382 [19].
The friction factor values for rod bundles are similar to those obtained for flow in
circular pipes, but friction factors for the specific geometry should be used when
available. Particularly when rods are closely spaced, stagnant zones can occur in the
narrow gaps, reducing the shear stress. These results in lower friction factors than
predicted for circular tubes, 63% lower when the rods are in contact (Subbotin et al.,
[20]). For triangular arrays, Rehme [21] proposed the following corrections to the
circular tube friction factors,
ForReDe ¼ 104 :
 
f P
¼ 1:045 þ 0:071 1 ð6:107Þ
f c:t: D

ForReDe ¼ 105 :
 
f P
¼ 1:036 þ 0:054 1 ð6:108Þ
f c:t: D

Todreas and Kazimi [19] provide additional information on page 386.

6.16 Pressure Drop Across Spacers

For flow parallel to rod bundles, spacers create form drag, which can result in
pressure drop comparable to the friction along bare rods. For flow over honeycomb
spacers (Fig. 6.19) and wire grid spacers, DeStrodeur [22] correlated the pressure
drop across an individual spacer as
 
Δps ¼ Cs ρV 2s =2 ðAs =Av Þ2 ð6:109Þ

[email protected]
6.16 Pressure Drop Across Spacers 265

where Vs is the velocity in the spacer region, As the projected frontal area of the
spacer, and Av the unrestricted flow area away from the grid or spacer. For spiral
wire wrapped rods, the calculation for the bare-rod pressure drop must be modified
if the DeStrodeur spacer correlation is used. The projected frontal area of the helical
wire is taken as the projected annular area occupied by the wire, which for the entire
rod length L is As ¼ π(D + ds)ds(L/h), where ds is the wire diameter and h is the wire
pitch (length per complete revolution). Example Problem 6.5 shows how to apply
the DeStrodeur correlation to wire wrapped rods.
Example 6.4: Pressure Drop through a Fuel Assembly Determine the pressure
drop due to flow of water 6.31 x 103 m3/s at 90 C through a rod bundle with spiral
wire spacers, given the following parameters (Rust Page 152) [23].
• 11 by 11 rod bundle with a square pitch
• Rod diameter D ¼ 0.49 cm
• Spiral wire diameter ds ¼ 0.64 mm
• Pin Pitch P ¼ 0.56
• Rod Length L ¼ 61.0 cm
• Helical spacer wire pitch h ¼ 15.3 cm
Solution: In this calculation, we determine the pressure drop due to flow over bare
rods and flow over the spacers separately, and then obtain the total pressure drop by
summing the two contributions. To begin the calculation we determine the number
of subchannels of the three types shown in Fig. 6.16. The number of central coolant
subchannels is.
nI ¼ (11  11)(11  1) ¼ 100 Central subchannels
The number of side and corner subchannels is, respectively.
nII ¼ 2  (11  1) + 2  (11  1) ¼ 40 Side subchannels
nIII ¼ 4 Corner subchannels
The areas of the three subchannel types are
π
AI ¼ P2  D2 ¼ 0:125  104 m2
4
AI
AII ¼ ¼ 0:063  104 m2
2
AII
AIII ¼ ¼ 0:031  104 m2
2

To apply the DeStrodeur correlation for helically wrapped rods, it is necessary to


calculate the bare tube pressure drop based on a modified subchannel flow area. For
bundles with honeycomb or wire grid spacers, the subchannel areas calculated
above would be used directly.

[email protected]
266 6 Turbulent Forced Convection

π
A0I ¼ AI  d 2s ¼ 0:122  104 m2
4
A0
A0II ¼ I ¼ 0:061  104 m2
2
A0
A0III ¼ II ¼ 0:030  104 m2
2

The actual and modified hydraulic diameters of the subchannels are then

4AI 4A0I
DeI ¼ ¼ 0:325  102 m D0eI ¼ ¼ 0:280  102 m
πD π ðD þ ds Þ
4AII 4A0II
DeII ¼ ¼ 0:188  102 m D0eII ¼ ¼ 0:171  102 m
P þ πD=2 P þ π2 ðD þ d s Þ
4AIII 4A0III
DeIII ¼ ¼ 0:131  102 m D0eIII ¼ ¼ 0:121  102 m
P þ πD=4 P þ π4 ðD þ d s Þ

Because the subchannels are linked, the bare tube pressure drop through each
channel is approximately equal (neglecting the effects of spacers), and is given by:
 
L ρu2i
ΔpR ¼ f w, i
D e, i 2

The total volume flow QT is given by the sum of the flows through individual
subchannels i, which assuming that the friction factor is equal for all channels is
given by:
P  1=2 P  
QT ¼ ui A0i ¼ 2gLfc ΔpR Ai D0eI
1=2
w
"  0 1=2  0 1=2 #
 1=2   A D A D
2gc ΔpR II III
Ai D0eI
1=2
¼ Lf 100 þ 40  1=2 þ 4  eIII1=2
eII
0
w
AI DeI AI D0eII
 1=2  1=2
¼ 2gLfc ΔpR 123Ai D0eI
w

The friction coefficient for the subchannels comes from the Reynolds number
and the Moody chart,
QI QT 6:3  103 m3 =s
uI ¼ ¼ ¼   ¼ 4:20 m=s
AI 0 123AI 0 123 0:122  104 m2
 
uDeI ð4:20 m=sÞ 0:280  102 m
Re ¼ ¼ ¼ 36, 000
v 0:327  106 m2 =s
f w ¼ 0:021

[email protected]
6.16 Pressure Drop Across Spacers 267

Finally, the pressure drop for flow over bare rods comes from the expression
above for QT (note that now DeI is used instead of DeI, following DeStordeur [22])
 
L ρQ2T
ΔpR ¼ f w
DeI 2gc ð123AI Þ2
 2
ð0:61 mÞ ð965 kh=m3 Þ 6:31  106 m2 =s
¼ 0:021     
0:325  106 m 2 ð123Þ 0:122  104 m2 =s 2
¼ 33, 600 Pa

The pressure drop due to the helical wire spacer is given by:

ρu2I S ρQ2T S
Δps ¼ Cs 0 ¼ Cs  2
2gc AI 2gc A0 123A0
I I

The projected frontal area of the spacers is

L
S ¼ π ðD þ ds Þds
h  
 2 2
 2
 61:0cm
¼ π 0:49  10 þ 0:064  10 m 0:064  10 m ¼ 0:444  104 m2
15:3cm

The drag coefficient Cs is found from Fig. 6.20, using a Reynolds number based
on the wire diameter,
 
uI ds ð4:20 m=sÞ 0:064  102 m
Res ¼ ¼ ¼ 8200
v 0:327  102 m2 =s

Cs ¼ 0.49 From Fig. 6.20


The pressure drop due to the helical wire spacer is then
 2  
ð965 kg=m3 Þ 6:31  103 m2 =s 0:444  104 m2
Δps ¼ ð0:49Þ  3 ¼ 15, 200 Pa
2ð123Þ2 0:122  104 m2

The total pressure drop is the sum of the bare rod and spacer pressure drops,

Δptotal, calculated ¼ Δps þ ΔpR ¼ 15, 200 Pa þ 33, 600 Pa ¼ 48, 800 Pa

This calculated value for the total pressure drop can be compared with the
experimental value given by Rust [23] (pg. 157),

Δptotal, calculated ¼ 48, 800 Pa

[email protected]
268 6 Turbulent Forced Convection

Spiral Wire Spacers Transverse Grid Spacers


1.0 2.4

0.9 2.2
0.8 2.0
Drag Coefficient Cs

Drag Coefficient Cs
Honeycomb grids
0.7 1.8
0.6 1.6
Circular wires
0.5 1.4

0.4 1.2 Lenticular wires


0.3 1.0

0.2 0.8
3 2 3 4 5 6 7 89 4 2 3 4 5 6 7 89 5 3 2 3 4 5 6 7 89 4 2 3 4 5 6 7 89 5
10 10 10 10 10 10
Reynolds number, Re=dsUsρ/μ Reynolds number, Re=dsUsρ/μ

Fig. 6.20 Drag coefficient for rod bundle space [22]

Discussion: Rust [23] provides comparisons of this technique for several flow
rates through this bundle geometry. He found that the calculated pressure losses
agree with experimental values within 12%.
Now getting back to discussion of our textbook here, Fig. 6.20 shows values of
the drag coefficient for honeycomb grids, circular wire grids, and lenticular wire
grids. At high Reynolds numbers, approaching 105, the drag coefficient approaches
the value Cs  1.7, while circular wires give friction factors about 10% lower.
Rehme [24] suggests an alternative form for the spacer pressure drop correlation,
which is easier to apply, and improves correlation of the data,
 
Δps ¼ Cv ρV 2v =2 ðAs =Av Þ2 ð6:110Þ

where Vv is the average bundle fluid velocity. For square arrays, Rehme [24]
recommends Cv ¼ 9.5 at
Re ¼ 104and Cv ¼ 6.5 at Re ¼ 105. Figure 6.21 summarizes his experimental
results for triangular and square arrays.

6.17 Flow Across Rod Bundles

For lateral flow across bare rod arrays, the pressure drop can be given by:

NG2max
Δp ¼ f Z ð6:111Þ

where Gmax is the maximum mass flux between the rods, N the number of rod rows
in the direction of flow, Z a correction factor for the array arrangement, and f a friction
factor given by Fig. 6.22 for inline rods or by Fig. 6.23 for staggered rods. In these
figures the Reynolds number is given by Re ¼ GmaxD/u, where D is the rod diameter.

[email protected]
6.17 Flow Across Rod Bundles 269

20

10
Cv
6

3
3x103 104 105
Re
Triangular array
Spacer coils Triangular-type spacer
Honey-comb-type spacer, n=1 Rhombus-type spacer
Honey-comb-type spacer, n=2 Ring-type spacer

Spengos
20

10
Cv
6

3
3x103 104 105
Re
Square array
Tube spacer transversally connected
Rehme Tube spacer axially connected
Honey-comb-type spacer

Spengos

Fig. 6.21 Modified drag coefficient for rod bundle spacer [24]

[email protected]
270 6 Turbulent Forced Convection

4 For xT ≠ xL
xT= PT/D 101
PL 103
2 xL= PL/D 4 10
4

10 2 5
8 PT Re = 10
6 Z 101 10
6
106
4 4
2
2
6 6 6
f 10-1 100 101
1.0 (xT-1)/(x -L1)
8
6 xL = 1.25
4 1.50
2.0
2
2.5
0.1
8

101 2 4 6 8102 2 4 6 8 103 2 4 6 8 104 2 4 6 8 105 2 4 6 8 106 2

Re

Fig. 6.22 Friction factor f and correction factor Z for in-line tubes for use with Eq. 6.108 [25]

100
8 xT = PT/D
6 For xT ≠ xD
4 xL = PL/D
2
xD = PD/D 1.6 Re = 10
5
2 PL 10
1.4 10
3
10 Z
4
8 PT 1.2 10
104
6 103
4 1.0 102 105
PD
f 2 0.4 0.6 0.8 1 2 4
xT= 1.25 xT/xL
1.50
1.0
8
6
4
2.0
2
2.5
0.1
8
6
4

101 2 4 68
102 2 4 68
103 2 4 68
104 2 4 68
105 2 4 68
106 2

Re
Fig. 6.23 Friction factor f and correction factor Z for staggered tubes for use with Eq. 6.108 [25]

[email protected]
6.18 Problems 271

6.18 Problems

Problem 6.1: A fluid flows over a plane surface 1 m by 1 m with a bulk temperature
of 50 C. The temperature of the surface is 20 C. The convective heat transfer
coefficient is 2000 W/m2 C. Use the following equation for natural or free
convection.
The equation for convection can be expressed as:

Q ¼ hc AΔt ð1Þ

where
Q ¼ heat transferred per unit time (W).
A ¼ heat transfer area of the surface (m2).
hc ¼ convective heat transfer coefficient of the process (W/(m2K) or W/(m2 C)).
Δt ¼ temperature difference between the surface and the bulk fluid (K or 0C)
Problem 6.2: A flat wall is exposed to the environment. The wall is covered with a
layer of insulation 1 in. thick whose thermal conductivity is 0.8 Btu/h ft F.
The temperature of the wall on the inside of the insulation is 600 F. The wall
loses heat to the environment by convection on the surface of the insulation. The
average value of the convection heat transfer coefficient on the insulation
surface is 950 Btu/h ft2 F. Compute the bulk temperature of the environment
(Tb) if the outer surface of the insulation does not exceed 105 F.
Problem 6.3: Engine oil at 60 C flows over a 5 m long flat plate whose temperature
is 20 C with a velocity of 2 m/s. Determine the total drag force and the rate of
heat transfer per unit width of the entire plate. Use the following figure to solve
this problem, assume the critical Reynolds number is 5  105 and use Eq. (1) of
Problem 6.1. Use Fig. P6.3.

oil

T¥ = 60°C

V¥ = 2 m/s Q°
Ts = 20°C

L=5m

Fig. P6.3 Velocity profile

Problem 6.4: Using Fig. P6.4, calculate slope of channel in that figure if Q ¼ 50 ft3/s
and assume for concrete n ¼ 0.017.

[email protected]
272 6 Turbulent Forced Convection

8 ft

2 ft

4 ft

Fig. P6.4 Trapezoidal channel

Problem 6.5: Design rectangular channel if form of unfinished concrete assuming


Q ¼ 6.75.m3/s, slope S ¼ 1.2%, normal depth of ½ of the width of the channel
and assume for concrete n ¼ 0.017.
Problem 6.6: In Problem 6.4, the final width was set at 2 m and the maximum
Q ¼ 12 m3/s; find the normal depth for this maximum discharge. Assume for
concrete n ¼ 0.017.
Problem 6.7: What pressure gradient is required to cause diethyl aniline, C6H5N
(C2H5)2, to flow in a horizontal, smooth, circular tube of inside diameter
D ¼ 3 cm at a mass rate of 1028 g/s at.20 C? At this temperature, the density
of diethyl aniline is ρ ¼ 0.935 g/cm3 and its viscosity is μ ¼ 1.95 cp
Problem 6.8: Determine the flow rate, in pounds per hour, of water at 68 F through
a 1000 ft length of horizontal 8-in. Schedule 40 steel pipe of internal diameter
7.981 in., under a pressure difference of 3.00 psi. For such a pipe use Fig. 6.11
and assume that (k/D) ¼ 2.3  104.
Problem 6.9: Glass spheres of density ρsphere ¼ 2.62 g/cm" are to be allowed to fall
through liquid CCl4 at 20 C in an experiment for studying human reaction times
in making time observations with stopwatches and more elaborate devices. At
this temperature, the relevant properties of CCl4 are ρ ¼ 1.59 g/cm3 and μ ¼ 9.58
millipoises. What diameter should the spheres be to have a terminal velocity of
about 65 cm/s? Use Fig. P6.9 below.

[email protected]
6.18 Problems 273

Fig. P6.9 Graphical 2.0 1.86


procedure used in
Problem 6.9
Line of slope 1.0
f = 1.86 × 10-5 Re
1.0

0.8

0.6
0.5
Portion of f versus Re
f 0.4
curve from Fig. 6.3-1
0.3

0.2

0.1
104 2 3 4 5 6 8 105
2.4 × 104
Re

Problem 6.10: The velocity profile for hydrodynamically developed laminar flow
inside a circular tube is given by:

 r 2
uðr Þ ¼ 2um 1 
R

where R is the inside radius of the tube and um is the mean flow velocity.
Develop an expression for the friction factor f for flow inside the tube. Assume
the following expression is obtained for definition of the friction factor:

8μ ∂u
f ¼ ð1Þ
ρu2m ∂r wall

where μ is dynamic viscosity of fluid and ρ is its density.


Hint: The result should be analogous to Eq. 6.58
Problem 6.11: Engine oil (ρ ¼ 868 kg/m3, v ¼ 0.75  104m2/s) flows with a mean
velocity of um ¼ 0.15 m/s inside a circular tube having an inside diameter
D ¼ 2.5 cm. Calculate the friction factor f and the pressure drop over the length
L ¼ 100 m of the tube.

[email protected]
274 6 Turbulent Forced Convection

Problem 6.12: Compute the loss of head and pressure drop in a horizontal 60 m
pipe of asphalted cast iron which is 150 mm in diameter which is carrying water
at 2 m/s. For water at 20 C the density ρ ¼ 998 kg/m3 and the viscosity
μ ¼ 0.001 Ns/m2 (Pas). Assume for asphalted cast pipe roughness is
ε ¼ 0.12 mm.
Problem 6.13: Oil, with ρ ¼ 900 kg/m3 and v ¼ 0.00001 m2/s, flows at v/Ac ¼ 0.2 m
3
/s through 500 m of 200 mm diameter cast-iron pipe. Determine (a) the head
loss and (b) the pressure drop if the pipe slopes down at 10 in the flow direction.
Assume for cast pipe roughness isε ¼ 0.26 mm.
Problem 6.14: By solving both Problems 6.12 and 6.13, summarize what have you
learned for solutions of these two problems.
Problem 6.15: Mineral slurry with ρ ¼ 950 kg/m3 and V_ ¼ vAc ¼ 2  105 m2 =s
flows through a 30 cm diameter wrought iron pipe 100 m long with a head loss of
8 m. The roughness ratio is ε/D ¼ 0.0002. Find the average velocity and
flow rate.
Problem 6.16: Calculate the velocity of water in 60 m of horizontal 150 mm
diameter asphalted cast iron pipe where the head loss is 1.63 m. (Note that this
is a “work-back” of the Problems 6.10 and 6.11).
Problem 6.17: Mineral slurry of density ρ ¼ 950 kg/m3 and kinematic viscosity
v ¼ 2  105 m2/s flows through a 100 m long horizontal wrought iron pipe of
roughness ε ¼ 0.06 mm at a flow rate of v_ ¼ 0:342 m3 =s. The head loss is
Δh ¼ 8 m. Determine what diameter the pipe ought to be.
Problem 6.18: Assume frictionless flow in a long, horizontal, conical pipe. The
diameter is 2.0 ft at one end and 4.0 ft at the other. The pressure head at the
smaller end is 16 ft of water. If water flows through this cone at a rate of 125.6 ft3
/s, find the velocities at the two ends and the pressure head at the larger end. Try
to modify the Bernoulli’s Eq. 6.86 to a new form in terms of potential, kinetic
energy and momentum for water. You need to derive the Bernoulli’s equation
results from the application of the general energy equation and the first law of
thermodynamics, for a steady flow system. Assume in which no work is done on
or by the fluid, no heat is transferred to or from the fluid, and no change occurs in
the internal energy (i.e., no temperature change) of the fluid.
Problem 6.19: This problem is given for Thin-Walled Elastic Conduit
Three different cases for the anchoring of the conduit against longitudinal
movement are:
(a) Conduit anchored against longitudinal movement throughout its length

D 
ψ¼ 1  v2 ð1Þ
e

[email protected]
6.18 Problems 275

(b) Conduit anchored against longitudinal movement at the upper end [Wylie
and Streeter, 1983]

D
ψ¼ ð1  0:50vÞ ð2Þ
e

(c) Conduit with frequent expansion joints

D
ψ¼ ð3Þ
e

If the general expression for the wave velocity defined by Halliwell3 as:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K
a¼  ð4Þ
ρ 1 þ ðK=EÞψ

The parameters used in the above equations are defined below:


ψ ¼ A nondimensional parameter that depends on the elastic properties of conduit
D ¼ Conduit diameter
e ¼ Wall thickness
v ¼ The Poisson ratio
K ¼ Bulk modulus of elasticity of conduit walls
ρ ¼ Density of the fluid
E ¼ Young’s module of elasticity of the conduit walls

Then Compute the wave velocity in the steel Penstock of the Kootenay Canal
hydroelectric power plant, BC, Canada, The data for different segments of the
penstock are listed in Table below. The values of for steel, for concrete, and and
for water are 207 GPa, 2.19 GPa, and 999 kg/m3, respectively. Assume the Poisson
ratio v is 0.3.

Pipe Length (m) Diameter (m) Wall thickness (mm) Remarks


1 244.0 6.771 19 Expansion coupling at one end
2 36.5 5.55 22 Encased in concrete

3
Halliwell, A. R., “Velocity of a Waterhammer Wave in an Elastic Pipe,” Jour., Hydraulics Div.,
Amer. Soc. Civil Engrs., vol. 89, No. HY4, July 1963, pp. 1–21.

[email protected]
276 6 Turbulent Forced Convection

Data for Penstock Power Plant in British Columbia, Canada


In addition for steel-lined tunnel, the ψ can be written as:

DE
ψ¼ ð5Þ
GD þ Ee

in which e ¼ thickness of the steel liner and E ¼ module of elasticity of steel.

References

1. S. Kakac, Y. Yener, Convective heat transfer, 2nd edn. (CRC Press, Boca Raton, FL)
2. R.L. Panton, Incompressible Flow, 4th edn. (John Wiley, New York, 2013)
3. A. Bejan, Force Convection: Internal Flows, Chapter 5. (Department of Mechanical E Duck
University, Durham, North Carolina, 1963)
4. W.M. Kays, M.E. Crawford, Convective and Mass Transfer, 2nd edn. (McGraw-Hill Book
Company, New York, 1980)
5. Prandtl, J. Fluid Mech. 37(04), 785–798 (1969), Cambridge University Press.
6. W.M. Kays, H.C. Perkins, Forced convection, internal flow in ducts, in Handbook of Heat
Transfer, ed. By W.M. Rohsenow, J.P. Hartnett. (McGraw-Hill, New York, 1973, Sec. 7)
7. L.F. Moody, Friction factors for pipe flow. Trans ASME 66(8), 671–684 (1944)
8. A. Bejan, Convection Heat Transfer, 2nd edn. (Wiley, New York, 1985)
9. J Nikuradse, Laws of flow in rough pipes. Translation of Stromungsgesetze in rauhen Rohren,
Nikuradse, Forschung auf dem Gebiete des Ingenieurwesens. NACA Technical Memorandum
1292 (1933)
10. Y.A. Cengel, J.M. Cimbala, Fluid Mechanics: Fundamentals and Applications, 2nd edn.
(McGraw-Hill, New York, 2013)
11. R.A. Seban, J.A. Hodgson, Laminar film condensation in a tube with upward vapor flow. Int.
J. Heat Mass Transf. 25, 1291–1299 (1982)
12. http://www.qudev.ethz.ch/phys4/studentspresentations/afm/Atomic%20force%20micro
scope.pdf
13. R.H. Nichols, Turbulence Models and Their Application to Complex Flows, University of
Alabama at Birmingham, Revision 4.01.
14. R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena (John Wiley and Sons,
New York, 1960)
15. C.F. Bonilla, Nuclear Engineering, 1st edn. (McGraw-Hill, New York, 1957)
16. B. Zohuri, Dimensional Analysis and Self-Similarity Methods for Engineers and Scientist
(Springer, New York, 2015)
17. C.R. Westaway, A.W. Loomis, Cameron Hydraulic Data, 16nd edn. (Ingersoll-Rand,
Woodcliff Lake, NJ, 1981)
18. Flow of Fluids Through Valves, Fittings, and Pipe, Crane Technical Paper No. 410, 1981.
19. N.E. Todreas, M.S. Kazimi, Nuclear Systems, vol I (CRC Publications, Boca raton, FL,
1990)
20. V.I. Subbotin, Heat Removal for the Reactor Fuel Elements Coold by Liquid Metals, in
Proceedings of the 3rd United Nations International Conference on the Peaceful Uses of
Atomic Energy (1964) p. 328
21. K. Rehme, Pressure drop performance of rod bundles in hexagonal arrangements. Int. J. Heat
Mass Transf. 15, 2499 (1972)

[email protected]
References 277

22. A.M. DeStordeur, Drag coefficients for fuel elements spacers. Nucleonics 19, 74 (1961)
23. J.H. Rust, Nuclear Power Plant Engineering (Haralson Publishing Company, Atlanta, GA,
1979)
24. K. Rehme, Pressure drop correlations for fuel elements spacers. Nucl. Technol. 17,
15 (1973)
25. A. Zukauskas, Heat transfer from tubes in crossflow. Adv. Heat Transfer 8, 93 (1972)

[email protected]

View publication stats

You might also like