Sei sulla pagina 1di 18

The Babcock & Wilcox Company

Chapter 3
Fluid Dynamics

In the production and use of steam there are many


fluid dynamics considerations. Fluid dynamics addresses steam and water flow through pipes, fittings,
valves, tube bundles, nozzles, orifices, pumps and turbines, as well as entire circulating systems. It also considers air and gas flow through ducts, tube banks, fans,
compressors and turbines plus convection flow of gases
due to draft effect. The fluid may be a liquid or gas
but, regardless of its state, the essential property of a
fluid is that it yields under the slightest shear stress.
This chapter is limited to the discussion of Newtonian
liquids, gases and vapors where any shear stress is
directly proportional to a velocity gradient normal to
the shear force. The ratio of the shear stress to the velocity gradient is the property viscosity represented by
the symbol .
Liquids and gases are recognized as states of matter. In the liquid state, a fluid is relatively incompressible, having a definite volume. It is also capable of
forming a free surface interface between itself and its
vapor or any other fluid with which it does not mix. On
the other hand, a gas is highly compressible. It expands
or diffuses indefinitely and is subject only to the limitations of gravitational forces or an enclosing vessel.
The term vapor generally implies a gas near saturation conditions where the liquid and the gas phase
coexist at essentially the same temperature and pressure, during a process such as vaporization or boiling.
In a similar sense the term gas denotes a highly superheated steam. Sometimes steam may be treated as
an ideal gas and careful judgment is needed when
doing so.
Fluid dynamics principles normally consider the
fluid to be a continuous region of matter, a continuum,
and a molecular model is not required except for rare
instances. However, one property is noteworthy to consider due to the effect on steam generation fluid flow
and due to intermolecular forces. Surface tension, ,
is a liquid property of the vapor-liquid interface and
is the energy per unit area required to extend the interface. Surface tension is important in two-phase systems, such as a mixture flowing in a boiler tube, and
relates to the shape and flow regime of the bubble interface and also to the heat transfer area of droplets.
Vapor bubbles increase the resistance to fluid flow.
Steam 41 / Fluid Dynamics

The surface tension of water is dependent on temperature and its value goes to zero at the critical temperature (705.47 F, 374.15C). Supercritical water is considered single phase in fluid dynamic analysis due to
zero surface tension.
The recommended correlation1 for the surface tension of water and its vapor, , is:
(T T )
= 235.8 10 N / m c

1.256

Tc T
1 0.625 T

(1)

where Tc = 647.15K and T is the fluid temperature in K.


Water in steam generators operating at supercritical pressure (above 3200.1 psia, 22.1 MPa) will behave as a single phase fluid converting from liquid to
steam without creating bubbles. At the critical pressure and critical temperature, the density of water and
steam are identical and there is no distinguishable interface at equilibrium conditions. Surface tension is
also related to the latent heat of vaporization which
also decreases to zero at the critical temperature.2 This
chapter discusses single phase fluid flow. Chapter 5
pertains to two-phase fluid flow that occurs in boiling
tube circuits.

Fundamental relationships
Three fundamental laws of conservation apply to
fluid dynamic systems: conservation of mass, momentum and energy. With the exception of nuclear reactions where minute quantities of mass are converted
into energy, these laws must be satisfied in all flowing systems. Fundamental mathematical relationships
for these principles are presented in several different
forms that may be applied in particular fluid dynamic
situations to provide an appropriate solution method.
However, full analytical solutions are frequently too
complex without the use of a computer. Simplified
forms of the full equations can be derived by applying engineering judgment to drop negligible terms and
consider only terms of significant magnitude for cer3-1

The Babcock & Wilcox Company


tain classes of problems. Fluid dynamics problems can
be classified as compressible or incompressible, viscous
or inviscid. Engineering practice is based upon applying various assumptions and empirical relationships
in order to obtain a practical method of solution. A
more complete discussion of the derivation of these
conservation law relationships and vector notation
representing three dimensional spaces may be found
in References 3, 4, 5 and 6.

Conservation of mass
The law of conservation of mass simply states that
the rate of change in mass stored in a system must
equal the difference in the mass flowing into and out
of the system. The continuity equation of mass for one
dimensional single phase flow in a variable area channel or stream tube is:
A

V
A

= 0
+ A
+ V
+ AV
x
x
x
t

(2)

In its simplest form in x, y and z three dimensional


Cartesian coordinates, conservation of mass for a small
fixed control volume is:

u +
v +
w =
t
z
y
x

(3)

where u, v and w are the fluid velocities in the x, y


and z coordinate directions; t is time and is the fluid
density. An important form of this equation is derived
by assuming steady-state ( / t = 0) and incompressible (constant density) flow conditions:
u v w
= 0
+
+
x y z

(4)

Although no liquid is truly incompressible, the assumption of incompressibility simplifies problem solutions and is frequently acceptable for engineering
practice considering water and oils.
Another relationship useful in large scale pipe flow
systems involves the integration of Equation 3 around
the flow path for constant density, steady-state conditions. For only one inlet (subscript 1) and one outlet
(subscript 2):

The conservation of momentum for one dimensional


single phase flow in a variable area channel or stream
tube is:

Pf
1 G 1 G 2 A
+

+
gc t
A x
A
P
g
= 0
+
sin +
x
gc
where
P
G
A

Pf
g
gc

=
=
=
=
=
=
=
=
=

pressure, psia (MPa)


mass flux, G = V, lb/h ft2 (kg/s m2)
flow area of channel ft2 (m2)
density lb/ft3 (kg/m3)
wall shear stress, lb/ft2 (N/m2) (refer to Equation 26)
channel wetted perimeter, ft (m)
32.17 ft /s2 (9.8 m /s2)
32.17 lbm ft/lbf s2 (1 kg m /N s2)
angle of channel inclination for x distance

This relationship is useful in calculating steam generator tube circuit pressure drop.
The conservation of momentum is a vector equation and is direction dependent, resulting in one equation for each coordinate direction (x, y and z for Cartesian coordinates), providing three momentum equations for each scaler velocity component, u, v and w.
The full mathematical representation of the momentum equation is complex and is of limited direct use in
many engineering applications, except for numerical
computational models. As an example, in the x coordinate direction, the full momentum equation becomes:

u
u
u
u
+w
+v
+u

z
y
x
t
= fx
P
x
2 u v w
+

x 3 x y z

(5)

where is the average density, V is the average velocity, A is the cross-sectional area, and m is the mass
flow rate.

v u
+

y x y

w u
+

z
z x

 = 1 A1 V1 = 2 A2 V2
m

Conservation of momentum
The law of conservation of momentum is a representation of Newtons Second Law of Motion the
mass of a particle times its acceleration is equal to the
sum of all of the forces acting on the particle. In a flowing system, the equivalent relationship for a fixed (control) volume becomes: the rate of change in momentum entering and leaving the control volume is equal
to the sum of the forces acting on the control volume.
3-2

(6)

Term 1
Term 2
Term 3
Term 4
(7)

where x is the body force in the x direction, P is the


pressure, and is the viscosity. This equation and the
corresponding equations in the y and z Cartesian coordinates represent the Navier-Stokes equations
which are valid for all compressible Newtonian fluids
with variable viscosity. Term 1 is the rate of momentum change. Term 2 accounts for body force effects
such as gravity. Term 3 accounts for the pressure gradient. The balance of the equation accounts for moSteam 41 / Fluid Dynamics

The Babcock & Wilcox Company


mentum change due to viscous transfer. Term 1 is
sometimes abbreviated as (Du /Dt) where Du /Dt is
defined as the substantial derivative of u. For a function (scaler or vector), D /Dt is the substantial derivative operator on function defined as:

D
+v
+u
=
y
x
t
Dt

+ v i
=
+w
t
z

(8)

or grad or del = i /x + j /y + k /z
For the special case of constant density and viscosity,
this equation reduces to (for the x coordinate direction):
(9)

The y and z coordinate equations can be developed


by substituting appropriate parameters for velocity u,
pressure gradient P / x, and body force x. Where viscosity effects are negligible ( = 0), the Euler equation
of momentum is produced (x direction only shown):
1 P
Du
= fx
Dt
x

(10)

Energy equation (first law of thermodynamics)


The law of conservation of energy for nonreacting
fluids states that the energy transferred into a system less the mechanical work done by the system must
be equal to the rate of change in stored energy, plus
the energy flowing out of the system with a fluid,
minus the energy flowing into the system with a fluid.
A single scaler equation results. The one dimensional
single phase flow energy equation for a variable area
channel or stream tube is:

P
H
H
1 P
= q H + q +
+G
A
J
x
t

(11)

where
P
G
A

PH
x
H
J

=
=
=
=
=
=
=
=
=

pressure, psia (MPa)


mass flux, lb/h ft2 (kg/s m2)
flow area of channel, ft2 (m2)
density, lb/ft3 (kg/m3)
wall shear stress, lb/ft2 (N/m2)
channel heated area, ft2 (m2)
channel distance, ft (m) for x distance
enthalpy, Btu/lb (kJ/kg)
mechanical equivalent of heat = 778.17 ft lbf/
Btu (1 N m/J)
q = heat flux at boundary, Btu/h ft2 (W/m2)
q = internal heat generation, Btu/h ft3 (W/m)
Steam 41 / Fluid Dynamics

DH
= q +
Dt

DP

+ ikT +

Dt
gc

Term 1 Term 2 Term 3 Term 4

where the vector gradient or grad or del operator on


function is defined as:

1 P
Du
2u 2u 2u
+ 2 + 2 + 2
= fx
z
y
Dt
x
x

A general form of the energy equation for a flowing system using an enthalpy based formulation and
vector notation is:

Term 5

(12)

where is the fluid density, H is the enthalpy per unit


mass of a fluid, T is the fluid temperature, q is the
internal heat generation, k is the thermal conductivity, and is the dissipation function for irreversible
work.6 Term 1 accounts for net energy convected into
the system, Term 2 accounts for internal heat generation, Term 3 accounts for work done by the system,
Term 4 addresses heat conduction, and Term 5 accounts for viscous dissipation.
As with the momentum equations, the full energy
equation is too complex for most direct engineering
applications except for use in numerical models. (See
Chapter 6.) As a result, specialized forms are based
upon various assumptions and engineering approximations. As discussed in Chapter 2, the most common
form of the energy equation for a simple, inviscid (i.e.,
frictionless) steady-state flow system with flow in at
location 1 and out at location 2 is:

JQ W = J ( u2 u1 ) + ( P2v2 P1v1 )
+

g
1
V22 V12 + ( Z2 Z1 )
gc
2 gc

(13a)

or
JQ W = J ( H 2 H1 )
+

g
1
V22 V12 + ( Z2 Z1 )
gc
2 gc

(13b)

where
Q = heat added to the system, Btu lbm (J/kg)
(See Note below)
W = work done by the system, ft-lbf/lbm (N m/kg)
J = mechanical equivalent of heat = 778.17 ft lbf/
Btu (1 N m/J)
u = internal energy, Btu/lbm (J/kg)
P = pressure, lbf/ft2 (N/m2)
= specific volume, ft3/lbm (m3/kg)
V = velocity, ft /s (m/s)
Z = elevation, ft (m)
H = enthalpy = u + P/J, Btu/lbm (J/kg)
g = 32.17 ft /s2 (9.8 m /s2)
g c = 32.17 lbm ft/lbf s2 (1 kg m /N s2)
Note: Where required for clarity, the abbreviation lb is augmented by f (lbf) to indicate pound force and by m (lbm) to
indicate pound mass. Otherwise lb is used with force or
mass indicated by the context.

3-3

The Babcock & Wilcox Company

Energy equation applied to fluid flow


(pressure loss without friction)
The conservation laws of mass and energy, when
simplified for steady, frictionless (i.e., inviscid) flow of
an incompressible fluid, result in the mechanical energy balance referred to as Bernoullis equation:
P1v + Z1

V2
V2
g
g
+ 1 = P2v + Z2
+ 2
gc
gc
2 gc
2 gc

(14)

The variables in Equation 14 are defined as follows


with the subscripts referring to location 1 and location 2 in the system:
P

Z
V

=
=
=
=

pressure, lbf/ft2 (N/m2)


specific volume of fluid, ft3/lbm (m3/kg)
elevation, ft (m)
fluid velocity, ft/s (m/s)

Briefly, Equation 14 states that the total mechanical energy present in a flowing fluid is made up of pressure energy, gravity energy and velocity or kinetic
energy; each is mutually convertible into the other
forms. Furthermore, the total mechanical energy is
constant along any stream-tube, provided there is no
friction, heat transfer or shaft work between the points
considered. This stream-tube may be an imaginary
closed surface bounded by stream lines or it may be
the wall of a flow channel, such as a pipe or duct, in
which fluid flows without a free surface.
Applications of Equation 14 are found in flow measurements using the velocity head conversion resulting from flow channel area changes. Examples are the
venturi, flow nozzle and various orifices. Also, pitot
tube flow measurements depend on being able to compare the total head, P + Z + (V2 /2 gc ), to the static
head, P + Z, at a specific point in the flow channel.
Descriptions of metering instruments are found in
Chapter 40. Bernoullis equation, developed from
strictly mechanical energy concepts some 50 years
before any precise statement of thermodynamic laws,
is a special case of the conservation of energy equation or first law of thermodynamics in Equations 13a
and b.
Applications of Equation 13 to fluid flow are given
in the examples on water and compressible fluid flow
through a nozzle under the Applications of the Energy Equation section in Chapter 2. Equation 18,
Chapter 2 is:
V2 =

2 gc J ( H1 H 2 ) = C H1 H 2

(15)

where
V2 =
gc =
J =
H1 =
H2 =
C =

downstream velocity, ft/s (m/s)


32.17 lbm ft/lbf s2 = 1 kg m/Ns2
778.26 ft lbf/Btu = 1 Nm/J
upstream enthalpy, Btu/lb (J/kg)
downstream enthalpy, Btu/lb (J/kg)
223.8 lbm/Btu ft/s (1.414 kg/J m/s)

This equation relates fluid velocity to a change in enthalpy under adiabatic (no heat transfer), steady, inviscid (no friction) flow where no work, local irrevers3-4

ible flow pressure losses, or change in elevation occurs.


The initial velocity is assumed to be zero and compressible flow is permitted. If the temperature (T ) and pressure (P ) of steam or water are known at points 1 and
2, Equation 15 provides the exit velocity using the enthalpy (H) values provided in Tables 1, 2 and 3 of Chapter 2. If the pressure and temperature at point 1 are
known but only the pressure at point 2 is known, the
outlet enthalpy (H2) can be evaluated by assuming constant entropy expansion from points 1 to 2, i.e., S1 = S2.

Ideal gas relationships


There is another method that can be used to determine velocity changes in a frictionless adiabatic expansion. This method uses the ideal gas equation of
state in combination with the pressure-volume relationship for constant entropy.
From the established gas laws, the relationship between pressure, volume and temperature of an ideal
gas is expressed by:
Pv = RT

or

Pv =

(16a)

R
T
M

(16b)

where
P

= absolute pressure, lb/ft2 (N/m2)


= specific volume, ft3/lb of gas (m3/kg)
= molecular weight of the gas, lb/lb-mole
(kg/kg-mole)
T = absolute temperature, R (K)
R = gas constant for specific gas, ft lbf/lbm R
(N m/kg K)
MR = R = the universal gas constant
= 1545 ft lb/lb-mole R (8.3143 kJ/kg-mole K)
The relationship between pressure and specific volume along an expansion path at constant entropy, i.e.,
isentropic expansion, is given by:

Pvk = constant

(17)

Because P1 and 1 in Equation 13 are known, the constant can be evaluated from P11k. The exponent k is
constant and is evaluated for an ideal gas as:
k = c p / cv = specific heat ratio

(18)

where
cp = specific heat at constant pressure, Btu/lb F (J/kg K)
cv = specific heat at constant volume, Btu/lb F (J/kg K)
= (u1 u2)/(T1 T2)
For a steady, adiabatic flow with no work or change
in elevation of an ideal gas, Equations 13, 16, 17 and
18 can be combined to provide the following relationship:
V22 V12

k 1

P2 k

k
= 2 gc
P1 v1 1 P
k 1
1

(19)

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company


When V1 is set to zero and using English units Equation 19 becomes:

V2 = 8.02

k 1

P2 k
k

P1 v1 1 , ft/s
k 1
P1

(20)

Equations 19 and 20 can be used for gases in pressure drop ranges where there is little change in k, provided values of k are known or can be calculated.
Equation 20 is widely used in evaluating gas flow
through orifices, nozzles and flow meters.
It is sufficiently accurate for most purposes to determine velocity differences caused by changes in flow
area by treating a compressible fluid as incompressible. This assumption only applies when the difference
in specific volumes at points 1 and 2 is small compared
to the final specific volume. The accepted practice is
to consider the fluid incompressible when:

(v2 v1 ) / v2 < 0.05

(21)

Because Equation 14 represents the incompressible


energy balance for frictionless adiabatic flow, it may
be rearranged to solve for the velocity difference as
follows:
V22 V12 = 2 gc ( Pv ) + Zg / gc

22)

where
(P) = pressure head difference between locations
1 and 2 = (P1 P2) , ft (m)
Z
= head (elevation) difference between locations 1 and 2, ft (m)
V
= velocity at locations 1 and 2, ft/s (m/s)
When the approach velocity is approximately zero,
Equation 22 in English units becomes:
V2 =

2 gh = 8.02 h , ft/s

most flow situations there are also bulk fluid interchanges known as eddy diffusion. The net result of
all inelastic momentum exchanges is exhibited in
shear stresses between adjacent layers of the fluid. If
the fluid is contained in a flow channel, these stresses
are eventually transmitted to the walls of the channel. To counterbalance this wall shear stress, a pressure gradient proportional to the bulk kinetic energy,
V 2 / 2 gc, is established in the fluid in the direction of
the bulk flow. The force balance is:

D = tube diameter or hydraulic diameter Dh ft (m)


Dh = 4 (flow area)/(wetted perimeter) for circular or noncircular cross-sections, ft (m)
dx = distance in direction of flow, ft (m)
w = shear stress at the tube wall, lb/ft2 (N/m2 )
Solving Equation 24 for the pressure gradient (dP /
dx):
4
dP
=
w
dx
D

Steam 41 / Fluid Dynamics

(25)

This pressure gradient along the length of the flow


channel can be expressed in terms of a certain number of velocity heads, , lost in a length of pipe equivalent to one tube diameter. The symbol is called the
friction factor, which has the following relationship to
the shear stress at the tube wall:

w =

f 1 V2
4 v 2 gc

(26)

Equation 25 can be rewritten, substituting for w from


Equation 26 as follows:

4 f 1 V2
dP
f 1 V2
=

=
dx
D 4 v 2 gc
D v 2 gc

(27)

The general energy equation, Equation 13, expressed


as a differential has the form:
VdV
+ d ( Pv ) = dQ dWk
gc

(28a)

VdV
+ Pdv + vdP = dQ dWk
gc

(28b)

du +

Pressure loss from fluid friction


So far, only pressure changes associated with the
kinetic energy term, V 2/2 gc, and static pressure term,
Z, have been discussed. These losses occur at constant
flow where there are variations in flow channel crosssectional area and where the inlet and outlet are at
different elevations. Fluid friction and, in some cases
heat transfer with the surroundings, also have important effects on pressure and velocity in a flowing fluid.
The following discussion applies to fluids flowing in
channels without a free surface.
When a fluid flows, molecular diffusion causes
momentum interchanges between layers of the fluid
that are moving at different velocities. These interchanges are not limited to individual molecules. In

(24)

where

(23)

In this equation, h, in ft head of the flowing fluid, replaces (P) + Z. If the pressure difference is measured in psi, it must be converted to lb/ft2 to obtain P
in ft.

D2
( dP ) = w D ( dx )
4

or

du +

Substituting Equation 26 of Chapter 2 (du = Tds


Pd) in Equation 28 yields:
VdV
Tds +
+ vdP = dQ dWk
(29)
gc
The term Tds represents heat transferred to or from
the surroundings, dQ, and any heat added internally
to the fluid as the result of irreversible processes.
These processes include fluid friction or any irrevers3-5

The Babcock & Wilcox Company


ible pressure losses resulting from fluid flow. (See
Equation 29 and explanation, Chapter 2.) Therefore:

Tds = dQ + dQF

(30)

where dQF is the heat equivalent of fluid friction and


any local irrecoverable pressure losses such as those
from pipe fittings, bends, expansions or contractions.
Substituting Equation 30 into Equation 29, canceling dQ on both sides of the equation, setting dWk equal
to 0 (no shaft work), and rearranging Equation 29
results in:
dP =

dQF
VdV

vgc
v

dQF
dx V 2
= f
D v 2 gc
v

(32)

Substitution of Equation 32 into Equation 31 yields:

VdV
f V2
dx

vgc
D v 2 gc

(33)

From Equation 5, the continuity equation permits


definition of the mass flux, G, or mass velocity or mass
flow rate per unit area [lb/h ft2 (kg/m2 s)] as:

V
= G = constant
(34)
v
Substituting Equation 34 into Equation 33 for a flow
channel of constant area:
dP = 2

G2 v
G2
dx
dv f
2 gc
2 gc D

(35)

Integrating Equation 35 between points 1 and 2, located at x = 0 and x = L, respectively:

P1 P2 = 2

3-6

G2 1
G2
(v2 v1 ) + f
2 gc
2 gc D

dx =
and

(31)

Three significant facts should be noted from Equation 31 and its derivation. First, the general energy
equation does not accommodate pressure losses due
to fluid friction or geometry changes. To accommodate
these losses Equation 31 must be altered based on the
first and second laws of thermodynamics (Chapter 2).
Second, Equation 31 does not account for heat transfer except as it may change the specific volume, ,
along the length of the flow channel. Third, there is
also a pressure loss as the result of a velocity change.
This loss is independent of any flow area change but
is dependent on specific volume changes. The pressure
loss is due to acceleration which is always present in
compressible fluids. It is generally negligible in incompressible flow without heat transfer because friction
heating has little effect on fluid temperature and the
accompanying specific volume change.
Equation 27 contains no acceleration term and
applies only to friction and local pressure losses. Therefore, dQF/ in Equation 31 is equivalent to dP of
Equation 27, or:

dP =

The second term on the right side of Equation 36 may


be integrated provided a functional relationship between and x can be established. For example, where
the heat absorption rate over the length of the flow
channel is constant, temperature T is approximately
linear in x, or:

L
0

vdx

(36)

L
0

vdx =

L
dT
T2 T1

L
T2 T1

2
1

vdT = Lvav

(37)

(38)

The term a is an average specific volume with respect to temperature, T.

vav = (v2 + v1 ) = v1 (vR + 1 )

(39)

where

R = 2 / 1
= averaging factor
In most engineering evaluations, is almost linear in T and l/2. Combining Equations 36 and
37, and rewriting 2 1 as 1 ( R 1):
P1 P2 = 2
+ f

G2
v1 ( vR 1 )
2 gc

L G2
v1 ( vR + 1 )
D 2 gc

(40)

Equation 40 is completely general. It is valid for compressible and incompressible flow in pipes of constant
cross-section as long as the function T = F(x) can be assigned. The only limitation is that dP/dx is negative at
every point along the pipe. Equation 33 can be solved
for dP/dx making use of Equation 34 and the fact that
P11 can be considered equal to P22 for adiabatic flow
over a short section of tube length. The result is:

dP
=
dx

Pf / 2 D
g Pv
1 c 2
V

(41)

At any point where V 2 = gcP, the flow becomes choked


because the pressure gradient is positive for velocities
greater than (gcP)0.5. The flow is essentially choked
by excessive stream expansion due to the drop in pressure. The minimum downstream pressure that is effective in producing flow in a channel is:

P2 = V 2 / v2 gc = v2 G 2 / gc

(42)

Dividing both sides of Equation 40 by G2 l / 2gc,


the pressure loss is expressed in terms of velocity
heads. One velocity head equals:
P (one velocity head) =

V2
V 2
=
2 gcCv
2 gcC

(43)

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company


where
P = pressure drop equal to one velocity head, lb/
in.2 (N/m2)
V = velocity, ft/s (m/s)
= specific volume, ft3/lb (m3/kg)
g c = 32.17 lbm ft/lbf s2 = 1 kg m/N s2
C = 144 in.2/ft2 (1 m2/m2)
= density, lb/ft3 (kg/m3)
In either case, represents the number of velocity
heads (Nvh) lost in each diameter length of pipe.
The dimensionless parameter defined by the pressure loss divided by twice Equation 43 is referred to
as the Euler number:

Eu = P / V 2 / gc

(44)

where is the density, or l/ .


Two other examples of integrating Equation 35
have wide applications in fluid flow. First, adiabatic
flow through a pipe is considered. Both H and D are
constant and Pl lm = P2 2m where m is the exponent
for constant enthalpy. Values of m for steam range
from 0.98 to 1.0. Therefore, the assumption P = constant = P1 1 is sufficiently accurate for pressure drop
calculations. This process is sometimes called isothermal pressure drop because a constant temperature ideal
gas expansion also requires a constant enthalpy. For P
= P1 1, the integration of Equation 35 reduces to:

v
G 2 2v1 v2
n 2
2 gc v1 + v2
v1
L G 2 2v1v2
+ f
D 2 gc v1 + v2

P1 P2 = 2

L G2
v
D 2 gc

L G
v
De 105

(47)

where
P =
=
L =
De =

fluid pressure drop, psi


friction factor from Fig. 1, dimensionless
length, ft
equivalent diameter of flow channel, in. (note
units)
= specific volume of fluid, ft3/lb
G = mass flux of fluid, lb/h ft2

Friction factor
The friction factor () introduced in Equation 26, is
defined as the dimensionless fluid friction loss in velocity heads per diameter length of pipe or equivalent
diameter length of flow channel. Earlier correlators in
this field, including Fanning, used a friction factor one
fourth the magnitude indicated by Equation 26. This
is because the shear stress at the wall is proportional
to one fourth the velocity head. All references to in
this book combine the factor 4 in Equation 25 with as
has been done by Darcy, Blasius, Moody and others.
The friction factor is plotted in Fig. 1 as a function of
the Reynolds number, a dimensionless group of variables defined as the ratio of inertial forces to viscous
forces. The Reynolds number (Re) can be written:
Re =

VDe
VDe
GDe
or
or

(48)

where

(46)

All terms in Equations 45 and 46 are expressed in


consistent units. However, it is general practice and
Steam 41 / Fluid Dynamics

P = f

(45)

Neither P2 nor 2 are known in most cases, therefore


Equation 45 is solved by iteration. Also, the term 21 2
/(1 + 2) can usually be replaced by the numerical average of the specific volumes av = 1/2 1(PR + 1) where
PR = P1 /P2 = 2/1. The maximum high side error at PR
= 1.10 is 0.22% and this increases to 1.3% at PR = 1.25.
It is common practice to use a numerical average for
the specific volume in most fluid friction pressure drop
calculations. However, where the lines are long, P2
should be checked by Equation 42. Also, where heat
transfer is taking place, P2 is seldom constant along the
flow channel and appropriate averaging factors should
be used. Computation using small zone subdivisions
along the length of the tube circuit is recommended to
limit errors in widely varying property values.
The second important example considering flow
under adiabatic conditions assumes an almost incompressible fluid, i.e., 1 is approximately equal to 2. (See
Equation 21.) Substituting for 1 and 2 in Equation 45, the result is:

P1 P2 = f

often more convenient to use mixed units. For example, a useful form of Equation 46 in English units
is:

=
=
=
V =
G =
De =

density of fluid, lbm/ft3 (kg/m3)


kinematic viscosity = /, ft2/h (m2/s)
viscosity of fluid, lbm/ft h (kg/m s)
velocity of fluid, ft/h (m/s)
mass flux of fluid, lb/h ft2 (kg/m2 s)
equivalent diameter of flow channel, ft (m)

Fluid flow inside a closed channel occurs in a viscous


or laminar manner at low velocity and in a turbulent
manner at high velocities. Many experiments on fluid
friction pressure drop, examined by dimensional
analysis and the laws of similarity, have shown that
the Reynolds number can be used to characterize a
flow pattern. Examination of Fig. 1 shows that flow
is laminar at Reynolds numbers less than 2000, generally turbulent at values exceeding 4000 and completely turbulent at higher values. Indeterminate conditions exist in the critical zone between Reynolds
numbers of 2000 and 4000.
Fluid flow can be described by a system of simultaneous partial differential equations. (See earlier Fundamental relationships section.) However, due to the
complexity of these equations, solutions are generally
only available for the case of laminar flow, where the
only momentum changes are on a molecular basis. For
laminar flow, integration of the Navier-Stokes equa3-7

The Babcock & Wilcox Company


tion with velocity in the length direction only gives the
following equation for friction factor:
f = 64 / Re

(49)

The straight line in the laminar flow region of Fig. 1


is a plot of this equation.
It has been experimentally determined that the
friction factor is best evaluated by using the Reynolds
number to define the flow pattern. A factor /De is then
introduced to define the relative roughness of the
channel surface. The coefficient expresses the average height of roughness protrusions equivalent to the
sand grain roughness established by Nikuradse.6 The
friction factor values in Fig. 1 and the /De values in
Fig. 2 are taken from experimental data as correlated
by Moody.7

Laminar flow
Laminar flow is characterized by the parallel flowing of individual streams like layers sliding over each
other. There is no mixing between the streams except
for molecular diffusion from one layer to the other. A

small layer of fluid next to the boundary wall has zero


velocity as a result of molecular adhesion forces. This
establishes a velocity gradient normal to the main body
of flow. Because the only interchanges of momentum
in laminar flow are between the molecules of the fluid,
the condition of the surface has no effect on the velocity gradient and therefore no effect on the friction
factor. In commercial equipment, laminar flow is usually encountered only with more viscous liquids such
as the heavier oils.

Turbulent flow
When turbulence exists, there are momentum interchanges between masses of fluid. These interchanges are induced through secondary velocities,
irregular fluctuations or eddys, that are not parallel
to the axis of the mean flow velocity. In this case, the
condition of the boundary surface, roughness, does
have an effect on the velocity gradient near the wall,
which in turn affects the friction factor. Heat transfer is substantially greater with turbulent flow (Chapter 4) and, except for viscous liquids, it is common to
induce turbulent flow with steam and water without

Fig. 1 Friction factor/Reynolds number relationship for determining pressure drop of fluids flowing through closed circuits (pipes and ducts).

3-8

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company


excessive friction loss. Consequently, it is customary
to design for Reynolds numbers above 4000 in steam
generating units.
Turbulence fluctuations in the instantaneous velocity introduce additional terms to the momentum
conservation equation called Reynolds stresses. These
fluctuations influence the mean motion and increase
the flow resistance in a manner producing an increase
in the apparent viscosity. Analysis of turbulent flow
must consider the impact of the fluctuating velocity
component along with the mean flow velocity or resort to empirical methods that account for the additional momentum dissipation.4, 6, 8

Velocity ranges
Table 1 lists the velocity ranges generally encountered in the heat transfer equipment as well as in duct
and piping systems of steam generating units. These
values, plus the specific volumes from the ASME
Steam Tables (see Chapter 2) and the densities listed
in Tables 2 and 3 in this chapter, are used to establish
mass velocities for calculating Reynolds numbers and
fluid friction pressure drops. In addition, values of
viscosity, also required in calculating the Reynolds

number, are given in Figs. 3, 4 and 5 for selected liquids and gases. Table 4 lists the relationship between
various units of viscosity.

Resistance to flow in valves and fittings


Pipelines and duct systems contain many valves and
fittings. Unless the lines are used to transport fluids
over long distances, as in the distribution of process
steam at a factory or the cross country transmission
of oil or gas, the straight runs of pipe or duct are relatively short. Water, steam, air and gas lines in a power
plant have relatively short runs of straight pipe and
many valves and fittings. Consequently, the flow resistance due to valves and fittings is a substantial part
of the total resistance.
Methods for estimating the flow resistance in valves
and fittings are less exact than those used in establishing the friction factor for straight pipes and ducts.
In the latter, pressure drop is considered to be the result of the fluid shear stress at the boundary walls of
the flow channel; this leads to relatively simple boundary value evaluations. On the other hand, pressure
losses associated with valves, fittings and bends are
mainly the result of impacts and inelastic exchanges

Table 1
Velocities Common in Steam Generating Systems
Velocity
Nature of Service
Air:
Air heater
Coal and air lines,
pulverized coal
Compressed air lines
Forced draft air ducts
Forced draft air ducts,
entrance to burners
Ventilating ducts
Crude oil lines [6 to 30
in. (152 to 762 mm)]
Flue gas:
Air heater
Boiler gas passes
Induced draft flues
and breaching
Stacks and chimneys
Natural gas lines (large
interstate)
Steam:
Steam lines
High pressure
Low pressure
Vacuum
Superheater tubes

Fig. 2 Relative roughness of various conduit surfaces. (SI conversion: mm = 25.4 X in.)

Steam 41 / Fluid Dynamics

ft/min

m/s

1000 to 5000

5.1 to 25.4

3000 to 4500
1500 to 2000
1500 to 3600

15.2 to 22.9
7.6 to 10.2
7.6 to 18.3

1500 to 2000
1000 to 3000

7.6 to 10.2
5.1 to 15.2

60 to 3600

0.3 to 18.3

1000 to 5000
3000 to 6000

5.1 to 25.4
15.2 to 30.5

2000 to 3500
2000 to 5000

10.2 to 17.8
10.2 to 25.4

1000 to 1500

5.1 to 7.6

8000 to 12,000 40.6 to


12,000 to 15,000 61.0 to
20,000 to 40,000 101.6 to
2000 to 5000
10.2 to

Water:
Boiler circulation
Economizer tubes
Pressurized water
reactors
Fuel assembly channels
Reactor coolant piping
Water lines, general

61.0
76.2
203.2
25.4

70 to 700
150 to 300

0.4 to 3.6
0.8 to 1.5

400 to 1300
2400 to 3600
500 to 750

2.0 to 6.6
12.2 to 18.3
2.5 to 3.8

3-9

The Babcock & Wilcox Company

Table 2
Physical Properties of Liquids at 14.7 psi (0.101 MPa)
Liquid

Temperature F (C)

Water

70 (21)
212 (100)
70 (21)

Automotive oil
SAE 10
SAE 50
Mercury
Fuel oil, #6

70
70
180
70

Kerosene

55 to 57
57 to 59
846
60 to 65
60 to 65
50 to 51

(21)
(21)
(82)
(21)

Table 3
Physical Properties of Gases at 14.7 psi (0.101 MPa)**
Instantaneous
Specific Heat
Temperature Density, cp
cv
k,
F
lb/ft3 Btu/lb F Btu/lb F cp/cv

Air

70
200
500
1000

0.0749
0.0601
0.0413
0.0272

0.241
0.242
0.248
0.265

0.172
0.173
0.180
0.197

1.40
1.40
1.38
1.34

70
200
500
1000

0.1148
0.0922
0.0634
0.0417

0.202
0.216
0.247
0.280

0.155
0.170
0.202
0.235

1.30
1.27
1.22
1.19

H2

70
200
500
1000

0.0052
0.0042
0.0029
0.0019

3.440
3.480
3.500
3.540

2.440
2.490
2.515
2.560

1.41
1.40
1.39
1.38

Flue gas*

70
200
500
1000

0.0776
0.0623
0.0429
0.0282

0.253
0.255
0.265
0.283

0.187
0.189
0.199
0.217

1.35
1.35
1.33
1.30

70
200
500
1000

0.0416
0.0334
0.0230
0.0151

0.530
0.575
0.720
0.960

0.406
0.451
0.596
0.836

1.30
1.27
1.21
1.15

CO2

CH4

* From coal; 120% total air; flue gas molecular weight 30.
** SI conversions: T, C = 5/9 (F-32); , kg/m3 = 16.02 x lbm/
ft3; cp, kJ/kg K = 4.187 x Btu/lbm F.

3-10

Specific Heat
Btu/lb F (kJ/kg C)

62.4 (999.4)
59.9 (959.3)

of momentum. These losses are frequently referred to


as local losses or local nonrecoverable pressure losses.
Even though momentum is conserved, kinetic energies are dissipated as heat. This means that pressure
losses are influenced mainly by the geometries of
valves, fittings and bends. As with turbulent friction
factors, pressure losses are determined from empirical correlations of test data. These correlations may

Gas

Density
lb/ft3 (kg/m3 )

1.000 (4.187)
1.000 (4.187)

(881 to 913)
(913 to 945)
(13,549)
(961 to 1041)
(961 to 1041)
(801 to 817)

0.435
0.425
0.033
0.40
0.46
0.47

(1.821)
(1.779)
(0.138)
(1.67)
(1.93)
(1.97)

be based on equivalent pipe lengths, but are preferably defined by a multiple of velocity heads based on
the connecting pipe or tube sizes. Equivalent pipe
length calculations have the disadvantage of being
dependent on the relative roughness (/D) used in the
correlation. Because there are many geometries of
valves and fittings, it is customary to rely on manufacturers for pressure drop coefficients.
It is also customary for manufacturers to supply
valve flow coefficients (CV) for 60F (16C) water. These
are expressed as ratios of weight or volume flow in the
fully open position to the square root of the pressure
drop. These coefficients can be used to relate velocity
head losses to a connecting pipe size by the following
expression:
(50)

N v = kD 4 / CV 2

Table 4
Relationship Between Various Units of Viscosity
Part A: Dynamic (or Absolute) Viscosity,
Pa s
Ns
m2

Centipoise

kg
ms

0.01 g
cm s

1.0
0.001
1.49
413 x 106
47.90

1000
1.0
1488
0.413
47,900

lbm
ft s
672 x 103
672 x 106
1.0
278 x 106
32.2

lbm
ft h

lbf s
ft2
20.9 x 103
20.9 x 106
0.0311
8.6 x 106
1.0

2420
2.42
3600
1.0
115,900

Part B: Kinematic Viscosity, = /


Centistoke
m2
s
1.0
106
92.9 x 103
25.8 x 106

0.01 cm2
s

ft2
s

ft2
h

106
1.0
92,900
25.8

10.8
10.8 x 106
1.0
278 x 106

38,800
0.0389
3600
1.0

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company

Fig. 4 Absolute viscosities of some common gases at atmospheric


pressure.

Fig. 3 Absolute viscosities of some common liquids (Pa s =


0.000413 X lbm/ft h).

where
N = number of velocity heads, dimensionless
k = units conversion factor: for CV based upon
gal/min/()1/2, k = 891
D = internal diameter of connecting pipe, in.
(mm)
CV = flow coefficient in units compatible with k and
D: for k = 891, CV = gal/min/()1/2
CV and corresponding values of N for valves apply
only to incompressible flow. However, they may be extrapolated for compressible condition using an average
specific volume between P1 and P2 for P values as high
as 20% of P1. This corresponds to a maximum pressure
ratio of 1.25. The P process for valves, bends and fittings is approximately isothermal and does not require
the most stringent limits set by Equation 21.
When pressure drop can be expressed as an equivalent number of velocity heads, it can be calculated by
the following formula in English units:
P = N v

v
12

G
105

(51)

where
P = pressure drop, lb/in.2
N = number of equivalent velocity heads, dimensionless
= specific volume, ft3/lb
G = mass flux, lb/ft2 h
Another convenient expression, in English units only,
for pressure drop in air (or gas) flow evaluations is:
Steam 41 / Fluid Dynamics

P = N v

30 T + 460 G
B 1.73 105 103

(52)

where
P = pressure drop, in. wg
B = barometric pressure, in. Hg
T = air (or gas) temperature, F
Equation 52 is based on air, which has a specific
volume of 25.2 ft3/lb at 1000R and a pressure equivalent to 30 in. Hg. This equation can be used for other
gases by correcting for specific volume.
The range in pressure drop through an assortment
of commercial fittings is given in Table 5. This resistance
to flow is presented in equivalent velocity heads based
on the internal diameter of the connecting pipe. As noted,
pressure drop through fittings may also be expressed as
the loss in equivalent lengths of straight pipe.

Contraction and enlargement irreversible


pressure loss
The simplest sectional changes in a conduit are converging or diverging boundaries. Converging boundaries can stabilize flow during the change from pressure energy to kinetic energy, and local irrecoverable
flow losses (inelastic momentum exchanges) can be
practically eliminated with proper design. If the included angle of the converging boundaries is 30 deg
(0.52 rad) or less and the terminal junctions are
smooth and tangent, any losses in mechanical energy
are largely due to fluid friction. It is necessary to consider this loss as 0.05 times the velocity head, based
on the smaller downstream flow area.
3-11

The Babcock & Wilcox Company


When the elevation change (Z2 Z1) is zero, the
mechanical energy balance for converging boundaries
becomes:
P1v +

V12
2 gc

= P2v +

V22
V2
+ Nc 2
2 gc
2 gc

(53)

Subscripts 1 and 2 identify the upstream and downstream sections. Nc, the contraction loss factor, is the
number of velocity heads lost by friction and local nonrecoverable pressure loss in contraction. Fig. 6 shows
values of this factor.
When there is an enlargement of the conduit section in the direction of flow, the expansion of the flow
stream is proportional to the kinetic energy of the
flowing fluid and is subject to a pressure loss depending on the geometry. Just as in the case of the contraction loss, this is an irreversible energy conversion
to heat resulting from inelastic momentum exchanges. Because it is customary to show these losses
as coefficients of the higher kinetic energy term, the
mechanical energy balance for enlargement loss is:
P1v +

V12
V2
V2
= P2v + 2 + N e 1
2 gc
2 gc
2 gc

(54)

The case of sudden enlargement [angle of divergence


= 180 deg ( rad)] yields an energy loss of (V1 - V2)2/
2gc. This can also be expressed as:
Ne

A
= 1 1
A2

(55)

where A1 and A2 are the upstream and downstream


cross-sectional flow areas, respectively and (A1 < A2).
Even this solution, based on the conservation laws,
depends on qualifying assumptions regarding static
Fig. 5 Absolute viscosities of saturated and superheated steam.

Table 5
Resistance to Flow of Fluids Through
Commercial Fittings*
Fitting

Loss in Velocity Heads

L-shaped, 90 deg (1.57 rad)


standard sweep elbow
L-shaped, 90 deg (1.57 rad)
long sweep elbow
T-shaped, flow through run
T-shaped, flow through 90 deg
(1.57 rad) branch
Return bend, close
Gate valve, open
Check valve, open
Globe valve, open
Angle valve, 90 deg (1.57 rad) open
Boiler nonreturn valve, open

0.3 to 0.7
0.2 to 0.5
0.15 to 0.5
0.6
0.6
0.1
2.0
5.0
3.0
1.0

to
to
to
to
to
to
to

1.6
1.7
0.2
10.0
16.0
7.0
3.0

* See Fig. 9 for loss in velocity heads for flow of fluids


through pipe bends.

3-12

Fig. 6 Contraction loss factor for >30 deg (Nc = 0.05 for 30 deg).

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company


pressures at the upstream and downstream faces of
the enlargement.
Experimental values of the enlargement loss factor, based on different area ratios and angles of divergence, are given in Fig. 7. The differences in static
pressures caused by sudden and gradual changes in
section are shown graphically in Fig. 8. The pressure
differences are shown in terms of the velocity head at
the smaller area plotted against section area ratios.

Flow through bends


Bends in a pipeline or duct system produce pressure
losses caused by both fluid friction and momentum
exchanges which result from a change in flow direction. Because the axial length of the bend is normally
included in the straight length friction loss of the pipeline or duct system, it is convenient to subtract a calculated equivalent straight length friction loss from
experimentally determined bend pressure loss factors.
These corrected data form the basis of the empirical
bend loss factor, Nb.
The pressure losses for bends in round pipe in excess of straight pipe friction vary slightly with Reynolds numbers below 150,000. For Reynolds numbers
above this value, they are reasonably constant and
depend solely on the dimensionless ratio r/D, the ratio of the centerline radius of the bend to the internal
diameter of the pipe. For commercial pipe, the effect
of Reynolds number is negligible. The combined effect of radius ratio and bend angle, in terms of velocity heads, is shown in Fig. 9.
Flow through rectangular ducts
The loss of pressure caused by a direction change
in a rectangular duct system is similar to that for cylindrical pipe. However, an additional factor, the shape
Fig. 8 Static pressure difference resulting from sudden and gradual
changes in section.

of the duct in relation to the direction of bend, must


be taken into account. This is called the aspect ratio,
which is defined as the ratio of the width to the depth
of the duct, i.e., the ratio b/d in Fig. 10. The bend loss
for the same radius ratio decreases as the aspect ratio
increases, because of the smaller proportionate influence
of secondary flows on the stream. The combined effect
of radius and aspect ratios on 90 deg (1.57 rad) duct
bends is given in terms of velocity heads in Fig. 10.
The loss factors shown in Fig. 10 are average values of test results on ducts. For the given range of
aspect ratios, the losses are relatively independent of
the Reynolds number. Outside this range, the variation with Reynolds number is erratic. It is therefore
recommended that Nb values for b/d = 0.5 be used for
all aspect ratios less than b/d = 0.5, and values for b/
d = 2.0 be used for ratios greater than b/d = 2.0. Losses
for bends other than 90 deg (1.57 rad) are customarily considered to be proportional to the bend angle.

Fig. 7 Enlargement loss factor for various included angles.

Steam 41 / Fluid Dynamics

Turning vanes
The losses in a rectangular elbow duct can be reduced by rounding or beveling its corners and by in3-13

The Babcock & Wilcox Company


For applications requiring a uniform velocity distribution directly after the turn, a full complement or
normal arrangement of turning vanes (see Fig. 12b)
is required. However, for many applications, it is sufficient to use a reduced number of vanes, as shown in
Fig. 12c.
For nonuniform flow fields, the arrangement of
turning vanes is more difficult to determine. Many
times, numerical modeling (see Chapter 6) and flow
testing of the duct system must be done to determine
the proper vane locations.

Fig. 9 Bend loss for round pipe, in terms of velocity heads.

stalling turning vanes. With rounding or beveling, the


overall size of the duct can become large; however,
with turning vanes, the compact form of the duct is
preserved.
A number of turning vane shapes can be used in a
duct. Fig. 11 shows four different arrangements. Segmented shaped vanes are shown in Fig. 11a, simple
curved thin vanes are shown in Fig. 11b, and concentric splitter vanes are shown in Fig. 11c. In Fig. 11c,
the vanes are concentric with the radius of the duct.
Fig. 11d illustrates simple vanes used to minimize flow
separation from a square edged duct.
The turning vanes of identical shape and dimension, Fig. 11b, are usually mounted within the bend
of an elbow. They are generally installed along a line
or section of the duct and are placed from the inner
corner to the outside corner of the bend. Concentric
turning vanes, Fig. 11c, typically installed within the
bend of the turn, are located from one end of the turn
to the other end.
The purpose of the turning vanes in an elbow or turn
is to deflect the flow around the bend to the inner wall
of the duct. When the turning vanes are appropriately
designed, the flow distribution is improved by reducing flow separation from the walls and reducing the
formation of eddy zones in the downstream section of
the bend. The velocity distribution over the downstream
cross-section of the turn is improved (see Fig. 12), and
the pressure loss of the turn or elbow is decreased.
The main factor in decreasing the pressure losses
and obtaining equalization of the velocity field is the
elimination of an eddy zone at the inner wall of the
turn. For a uniform incoming flow field, the largest
effect of decreasing the pressure losses and establishing a uniform outlet flow field for a turn or elbow is
achieved by locating the turning vanes closer to the
inner curvature of the bend. (See Figs. 11d and 12c.)
3-14

Fig. 10 Loss for 90 deg (1.57 rad) bends in rectangular ducts.

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company

Fig. 12 Velocity profiles downstream of an elbow: a) without


vanes, b) with typical vanes, and c) with optimum vanes (adapted
from Idelchik, Reference 12).

Fig. 11 Turning vanes in elbows and turns: a) segmented, b) thin


concentric, c) concentric splitters, and d) slotted (adapted from
Idelchik, Reference 12).

Pressure loss
A convenient chart for calculating the pressure loss
resulting from impact losses in duct systems conveying air (or flue gas) is shown in Fig. 13. When mass
flux and temperature are known, a base velocity head
in inches of water at sea level can be obtained.
Flow over tube banks
Bare tube The transverse flow of gases across tube
banks is an example of flow over repeated major crosssectional changes. When the tubes are staggered, sectional and directional changes affect the resistance.
Experimental results and the analytical conclusions
of extensive research by The Babcock & Wilcox Company (B&W) indicate that three principal variables
other than mass flux affect this resistance. The primary variable is the number of major restrictions, i.e.,
the number of tube rows crossed, N. The second variable is the friction factor which is related to the
Reynolds number (based on tube diameter), the tube
spacing diameter ratios, and the arrangement pattern
(in-line or staggered). The third variable is the depth
factor, Fd (Fig. 14), which is applicable to banks less
than ten rows deep. The friction factors for various
in-line tube patterns are given in Fig. 15.
The product of the friction factor, the number of
major restrictions (tube rows) and the depth factor is,
in effect, the summation of velocity head losses
through the tube bank.
N v = f N Fd

Flow through stacks or chimneys


The flow of gases through stacks or chimneys is established by the natural draft effect of the stack and/
or the mechanical draft produced by a fan. The resistance to this flow, or the loss in mechanical energy be-

(56)

The N value established by Equation 56 may be


used in Equations 51 or 52 to find the tube bank pressure loss. Some test correlations indicate values
higher than the isothermal case for cooling gas and
lower for heating gas.
Finned tube In some convective boiler design applications, extended surface tube banks are used.
Many types of extended surface exist, i.e., solid helical fin, serrated helical fin, longitudinal fin, square fin
and different types of pin studs. For furnace applications,
the cleanliness of the gas or heat transfer medium dictates whether an extended surface tube bank can be used
and also defines the type of extended surface.
Steam 41 / Fluid Dynamics

Several different tube bank calculation methods


exist for extended surface, and many are directly related to the type of extended surface that is used.
Various correlations for extended surface pressure loss
can be found in References 9 through 15. In all cases,
a larger pressure loss per row of bank exists with an
extended surface tube compared to a bare tube. For
in-line tube bundles, the finned tube resistance per
row of tubes is approximately 1.5 times that of the bare
tube row. However, due to the increased heat transfer absorption of the extended surface, a smaller number of tube rows is required. This results in an overall
bank pressure loss that can be equivalent to a larger
but equally absorptive bare tube bank.

Fig. 13 Mass flux/velocity head relationship for air.

3-15

The Babcock & Wilcox Company

Fig. 14 Draft loss depth factor for number of tube rows crossed in
convection banks.

tween the bottom and the top of the stack, is a result of


the friction and stack exit losses. Application examples
of these losses are given in Chapter 25.

example of inelastic momentum exchanges occurring


within the fluid streams.
The injector is a jet pump that uses condensing steam
as the driving fluid to entrain low pressure water for
delivery against a back pressure higher than the pressure of the steam supplied. The ejector, similar to the
injector, is designed to entrain gases, liquids, or mixtures of solids and liquids for delivery against a pressure less than that of the primary fluid. In a waterjet aspirator, water is used to entrain air to obtain a
partial vacuum. In the Bunsen type burner, a jet of
gas entrains air for combustion. In several instances,
entrainment may be detrimental to the operation of
steam boilers. Particles of ash entrained by the products of combustion, when deposited on heating surfaces, reduce thermal conductance, erode fan blades,
and add to pollution when discharged into the atmosphere. Moisture carrying solids, either in suspension
or in solution, are entrained in the stream. The solids
may be carried through to the turbine and deposited
on the blades, decreasing turbine capacity and efficiency. In downcomers or supply tubes, steam bubbles
are entrained in the water when the drag on the
bubbles is greater than the buoyant force. This reduces the density in the pumping column of natural
circulation boilers.

Pressure loss in two-phase flow


Evaluation of two-phase steam-water flows is much
more complex. As with single-phase flow, pressure loss
occurs from wall friction, acceleration, and change in
elevation. However, the relationships are more complicated. The evaluation of friction requires the assessment of the interaction of the steam and water phases.
Acceleration is much more important because of the
large changes in specific volume of the mixture as
water is converted to steam. Finally, large changes in
average mixture density at different locations significantly impact the static head. These factors are presented in detail in Chapter 5.

Entrainment by fluid flow


Collecting or transporting solid particles or a second fluid by the flow of a primary fluid at high velocity is known as entrainment. This is usually accomplished with jets using a small quantity of high pressure fluid to carry large quantities of another fluid or
solid particles. The pressure energy of the high pressure fluid is converted into kinetic energy by nozzles,
with a consequent reduction of pressure. The material to be transported is drawn in at the low pressure
zone, where it meets and mixes with the high velocity jet. The jet is usually followed by a parallel throat
section to equalize the velocity profile. The mixture
then enters a diverging section where kinetic energy
is partially reconverted into pressure energy. In this
case, major fluid flow mechanical energy losses are an

3-16

Fig. 15 Friction factor (f ) as affected by Reynolds number for


various in-line tube patterns; crossflow gas or air.

Steam 41 / Fluid Dynamics

The Babcock & Wilcox Company

Boiler circulation
An adequate flow of water and steam-water mixture is necessary for steam generation and control of
tube metal temperatures in all circuits of a steam generating unit. At supercritical pressures this flow is
produced mechanically by pumps. At subcritical pressures, circulation is produced by the force of gravity

or pumps, or a combination of the two. The elements


of single-phase flow discussed in this chapter, twophase flow discussed in Chapter 5, heat input rates,
and selected limiting design criteria are combined to
evaluate the circulation in fossil-fired steam generators. The evaluation procedures and key criteria are
presented in Chapter 5.

References
1.
2.
3.
4.
5.
6.
7.
8.

Meyer, C.A., et al., ASME Steam Tables, Sixth Ed.,


American Society of Mechanical Engineers, New York,
New York, 1993.
Tabor, D., Gases, Liquids and Solids: and Other
States of Matter, First Ed., Penguin Books, Ltd.,
Harmondsworth, England, United Kingdom, 1969.
Lahey, Jr., R.T., and Moody, F.J., The Thermal-Hydraulics of a Boiling Water Nuclear Reactor, American Nuclear Society, Hinsdale, Ilinois, 1993.
Rohsenow, W., Hartnett, J., and Ganic, E., Handbook
of Heat Transfer Fundamentals, McGraw-Hill Company, New York, 1985.
Burmeister, L.C., Convective Heat Transfer, Second
Ed., Wiley-Interscience, New York, New York, 1993.
Schlichting, H.T. Gersten, K., and Krause, E.,
Boundary-Layer Theory, Eighth Ed., Springer-Verlag,
New York, New York, 2000.
Moody, L.F., Friction Factors for Pipe Flow, Transactions of the American Society of Mechanical Engineers (ASME), Vol. 66, 8, pp. 671-684, November, 1944.
Hinze, J.O., Turbulence: An Introduction to Its
Mechanism and Theory, Second Ed., McGraw-Hill
Company, New York, New York, 1975.

Steam 41 / Fluid Dynamics

9. Briggs, D.E., and Young, E.H., Convective heat


transfer and pressure drop of air flowing across triangular pitch banks of finned tubes, Chemical Engineering Progress Symposium Series (Heat Transfer),
AIChE, Vol. 41, No. 41, pp. l-10, Houston, Texas, 1963.
10. Grimison, E.D., Correlation and utilization of new
data on flow resistance and heat transfer for crossflow
of gases over tube banks, Transactions of ASME,
Process Industries Division, Vol. 59, pp. 583-594, New
York, New York, 1937.
11. Gunter, A.Y., and Shan, W.A., A general correlation
of friction factors for various types of surfaces in crossflow, Transactions of ASME, Vol. 67, pp. 643-660,
1945.
12. Idelchik, I.E., Handbook of Hydraulic Resistance,
Third Ed., Interpharm/CRC, New York, New York,
November, 1993.
13. Jakob, M., Discussion appearing in Transactions of
ASME, Vol. 60, pp. 384-386, 1938.
14. Kern, D.Q., Process Heat Transfer, p. 555, McGrawHill Company, New York, New York, December, 1950.
15. Wimpress, R.N., Hydrocarbon Processing and Petroleum Refiner, Vol. 42, No. 10, pp. 115-126, Gulf Publishing Company, Houston, Texas, 1963.

3-17

The Babcock & Wilcox Company

Laser velocity measurements in a steam generator flow model.

3-18

Steam 41 / Fluid Dynamics

Potrebbero piacerti anche