Download as pdf or txt
Download as pdf or txt
You are on page 1of 114

Chapter 2

Fluid Flow

Most aspects of fan engineering are concerned in some way with the flow
of fluids. Consequently, many fundamentals of flow will be discussed in this
chapter to provide a basis for the detailed discussions in subsequent chapters.
The first part of this chapter deals with flow under more or less ideal
conditions. While it does not ignore friction, it does not deal with it directly.
The second part examines the effects of viscosity and other factors that
determine the resistance to flow. The third part gives means for determining
the frictional losses in the various elements of a duct system. The last part
covers the measurement of pressure and flow.

Principles of Fluid Flow


The general principles that govern fluid flow are discussed in the
following sections. Only the integral equations have been developed.
Differential equations are more suitable for examining some flow-related
questions; if this is so, refer to a text on fluid mechanics.

Mathematical Models
A mathematical model is necessary if a flow situation is to be analyzed
quantitatively. The model must be capable of representing any variable that
changes significantly during the flow. (Flow variables include the position,
velocity, acceleration, pressure, temperature, density, enthalpy, and entropy of
the fluid.) The changes predicted by the model must be consistent with the
physical laws that govern fluid flow.
These laws are:
1) the general law regarding conservation of mass,
2) the first law of thermodynamics regarding conservation of energy,
3) Newton's second law of motion regarding force and momentum,
and
4) the second law of thermodynamics regarding entropy.
In most fluid mechanics models, the fluid is assumed to be a continuum.
That is, the velocities and other properties are assumed to vary continuously
throughout the fluid. Such an assumption could not be justified for highly
rarefied gas flows, but these usually do not occur in fan engineering.
In the mathematical formulation of the physical laws governing fluid flow,
.

© 1999 Howden Buffalo, Inc.


2-2 FAN ENGINEERING

coordinates must be assigned to points in space. In one method of modeling,


coordinates that are a function of time are also assigned to identifiable
particles (or portions) of the fluid. This Lagrangian approach requires that the
two sets of coordinates be distinguished from each other and leads to complex
equations. It is more usual in fluid mechanics to use the Eulerian approach,
which leads to simpler equations. In this method, the flow field is described;
that is, the fluid properties are specified at points in space by properly
constructed equations. The properties can vary with time at any point in
space, if that is a condition of modeling. And it is not necessary to assign
coordinates to individual portions of the fluid.
Any mathematical model that is capable of representing the most general
flow situation will be more complex than necessary for many fan-engineering
applications. Accordingly, various assumptions can be made to simplify the
model. However, caution is advisable in using simplified models until the
assumptions have been verified for the situation being modeled. The more
common assumptions are discussed below.
1) All real flows are three-dimensional to some extent, but it is not always
necessary to use a three-dimensional model to describe the flow. It may be
possible to analyze some flow situations with either a one-dimensional or a
two-dimensional flow model. In a one-dimensional flow model, the changes
in variables perpendicular to the main flow are taken into account by using
average values. Only one dimension or coordinate is required to establish
position along the flow path. Pipe flow, for example, usually can be treated as
one-dimensional by using the proper averages across the section for the
variables. In two-dimensional flow, just two independent coordinates are
required to describe the flow field. For instance, the flow across a wing can
be analyzed on a two-dimensional basis by using a correction factor to
account for the three-dimensional effects at the tips. A three-dimensional
model considers each variable to be a function of three space coordinates and
time.
2) All real flows probably have some unsteadiness. That is, the properties
at a point vary with time. However, the time variations may be small and
centered around a constant value for each property. In such cases, the
temporal average values can be used in the model, and the flow can be called
steady. However, in turbulence studies, certain effects of these variations
cannot be disregarded.
3) All flows are influenced by gravitational forces. However, in some
flows, particularly those involving gases, the effects of gravity are negligible
compared to the effects of other forces to which the fluid is exposed. If so,
the gravity terms can be omitted from the equations in the model.
4) All fluids are compressible, especially gases. The effects of
compressibility on liquids can be ignored, except for situations involving very
large or sudden changes in pressure. The effects can also be ignored in many
calculations involving gases when the pressure changes and Mach numbers
are small. Using a compressibility factor may be the most convenient way to
deal with compressibility when it cannot be ignored. Otherwise, the
appropriate compressible-flow equation must be used.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-3

5) All real flows involve friction. Often, however, it is convenient to


consider the flow frictionless or ideal. The analysis of some flows may be
divided into two parts: that in a boundary layer and that outside such a
boundary layer, with the latter considered ideal.
6) Other effects that are almost always ignored in fan engineering
applications are those due to surface tension, buoyancy, and Coriolis forces.
The number of equations required to model a flow situation must equal the
number of unknowns. Basically, a continuity equation and one or more
equations of motion will be needed. The latter will provide a mechanical
energy balance. If thermodynamic effects are significant, the general energy
equation will be needed. An equation of state will be required when an
explicit relationship among pressure, temperature, and density is necessary.
The equation of state for perfect gases is discussed in Chapter 1. Discussions
of the equations of continuity, motion, and energy are given in the next three
sections of this chapter. In these discussions, reference is made to a system,
as opposed to a control volume and its associated control surface. A system is
a definite mass of material that can be distinguished from its surroundings. A
control volume is a region in space and is signified by the abbreviation c.v.
This region is usually fixed, but it may be moving in space. The control
surface is the boundary of the control volume and is signified by the
abbreviation c.s. For the control volume equations, the flow enters the control
volume through the entrance area A1 on the control surface and leaves
through the exit area A2 . If there is more than one entrance or more than one
exit, the equations must be modified. Each discussion contains a system
equation in differential form to illustrate the underlying physical law and at
least one version of a control volume equation in integral form to provide the
basis for the simplified equations that are commonly used. Vector notation
(denoted by symbols with superior arrows) is used in the general equations to
convey the three-dimensional aspects without G necessitating
G three equations.
G
Vector quantities include velocity vector V , area vector A , force vector F ,
G G
surface force vector Fs , body force per unit volume vector B , radius or
G G
position vector r , and surface torque vector Ts . Most of the scalar quantities
ù
are defined as they appear; but time t, mass density ρ , and volume V should
be noted here.
In most fan engineering applications, the reader is not expected to use
either the general system equations or the general control volume equations.
These equations, however, do illustrate what has been omitted from the
simplified equations derived from them.

Continuity Equations
The general physical law regarding the conservation of mass requires that
the mass m of a particle or system of particles remain constant with time t .
Stated another way, the rate of change of mass dm / dt must be zero, which
leads to the system equation

© 1999 Howden Buffalo, Inc.


2-4 FAN ENGINEERING

dm
= 0. (2.1)
dt

The corresponding control volume equation states that the net rate of mass
flow through the control surface must equal the rate of change of mass in the
control volume, or

II
c .s .
G G
ρV ⋅ dA = −

∂t III
c.v .
ù
ρ dV . (2.2)

This equation, written in integral form with vector notation, is sufficiently


general to cover three-dimensional flow that is nonuniform, unsteady, and
compressible. In most fan engineering applications, the flow is steady and the
rate of change of mass in the control volume is zero, leading to

II
c .s .
G G
ρ V ⋅ dA = 0 . (2.3)

For one-dimensional compressible flow with only one entrance and one exit
in the control surface,

ρ 1V1 A1 = ρ 2V2 A2 . (2.4)

For one-dimensional incompressible flow with one entrance and one exit,

V1 A1 = V2 A2 . (2.5)

Although Equations 2.4 and 2.5 have been derived from one-dimensional
considerations, they can be used for nonuniform flow if the densities and
velocities across the section are represented by appropriate averages. For
Equation 2.5 (incompressible flow), it is appropriate to use the area-weighted
average of the velocities V across the section. For Equation 2.4
(compressible flow), it is also appropriate to use the area-weighted average of
the velocities provided that the volume-flow-weighted average of the densities
ρ is used. Alternatively, the area-weighted average of the products ρV can
be used in Equation 2.4. Remember that both equations are based on the
assumption of steady flow, which may not always be the case. Unsteady flow
occurs in fan engineering during start-up and shutdown, as well as in
situations involving pulsation. The equations must also be modified if there is
more than one entrance area A1 or exit area A2 .

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-5

Equations of Motion or Momentum Equations


Newton's second law of motion requires that the sum of the external
forces F acting on a particle equal the mass of the particle m times its
acceleration dV/dt. The corresponding system equation states that the rate of
change of momentum equals the sum of the external forces on the system, or

G G
d
dt
3 8
mV = F . (2.6)

Similarly, the control volume equation states that the sum of the surface
forces plus the sum of the body forces must equal the net rate of momentum
flow through the control surface plus the rate of change of momentum in the
control volume, or

G
Fs + III
c .v .
G ù
BdV = II
c .s .
G G G ∂
VρV ⋅ dA +
∂t III
c .v .
G ù
Vρ dV . (2.7)

This equation, written in integral form with vector notation, is sufficiently


general to cover three-dimensional flow that is nonuniform, unsteady, and
compressible, as well as the gravitational and frictional effects on the flow.
However, in most fan engineering applications, the flow is steady and
gravitational effects are negligible, so
G
Fs = II
c .s .
G G G
VdV ⋅ dA. (2.8)

This vector equation also can be written in component form. For instance, for
the component Fx and for uniform velocities at both A1 and A2 ,

Fx = mû (Vx 2 − Vx1 ) (2.9)

where mû is the mass flow rate and Vx is the velocity component on the x-axis.
Newton's second law also requires that the sum of the moments of the
external forces about a point acting on a particle equal the rate of change of
angular momentum of the particle. The corresponding system equation for
angular momentum is

d G G G G
(mr × V ) = r × F , (2.10)
dt

© 1999 Howden Buffalo, Inc.


2-6 FAN ENGINEERING

and the corresponding control volume equation is

II
c .s .
G G
r × dFs + III
c .v .
G G ù
r × BdV = II
c .s .
G G G G ∂
r × Vρ V ⋅ dA +
∂t III
c .v .
G G ù
r × Vρ dV . (2.11)

The latter, written in integral form with vector notation, is sufficiently general
to cover three-dimensional flow that is nonuniform, unsteady, and
compressible, as well as the gravitational and frictional effects on the flow.
Since, in most fan-engineering applications, the flow is steady and
gravitational effects are negligible,
G
Ts = II
c.s.
G G G G
r × Vρ V ⋅ dA (2.12)

where Ts is the net surface torque on the control volume. This vector equation
also can be written in component form. For instance, for the component Tz
and for uniform velocities at both A1 and A 2 ,

Tz = mû (r2Vt 2 − rV
1 t1 ) (2.13)

where mû is the mass flow rate, r is the radius, and Vt is the tangential
component of the velocity about the z-axis.
Equations 2.9 and 2.13 are simple and useful, but they do have limitations.
Both are based on uniform flow but can be used for non-uniform flow if the
velocities and radii across the section are represented by appropriate averages.
For Equation 2.9 (linear momentum), it is appropriate to use the mass-flow-
weighted average of the velocities V across the section. For Equation 2.13
(angular momentum), it is appropriate to use the mass-flow-weighted average
of the product of the radius and velocity rV . In either case, it would be
appropriate to use the area-weighted-average velocity only if an appropriate
momentum correction factor1 were applied to each average value. Both
equations also assume that the flow is steady, which is not always the case.
And both equations require that the right-hand side be divided by the
conversion factor gc if force and mass are assigned independent dimensions.
Equation 2.13 is useful particularly in the analysis of fans and other
turbomachines.
The equations of motion can be used to derive various equations that deal
with the mechanical energy balance in a system. These equations are also
called Euler or Bernoulli equations. However, since the general energy
equation, which contains the mechanical energy balance as well as a
thermodynamic balance, is discussed in the next section, no detailed
discussion is given here.
1
See the second section following.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-7

General Energy Equation


The first law of thermodynamics requires that the heat Q added to a
system less the work W done by the system must equal the change in energy
E of the system. Stated another way, the rate of heat transfer dQ dt to the
system minus the rate of work done dW dt by the system equals the rate of
change of energy dE dt in the system, which leads to the system equation

dE dQ dW
= − . (2.14)
dt dt dt

The directions of energy flow indicated in the above statement are considered
positive and dictate the sign convention. The work term is frequently given
the opposite sign in fan engineering, as noted later. Assuming that electrical,
magnetic, chemical, nuclear, and surface tension effects are negligible,

V2
E =U +m + gmZ (2.15)
2

where U is the internal energy, mV 2 2 is the kinetic energy, and gmZ is the
potential energy of position. The corresponding control volume equation
states that the rate of heat transfer to the system minus the rate of work done
by the system must equal the net rate of energy crossing the control surface
plus the rate of energy accumulation in the control volume, or

dQ dW
dt

dt II
c .s .
G G ∂
= eρ V ⋅ dA +
∂t III
c .v .
eρ dV
ù
(2.16)

where e is the energy per unit mass or specific energy and is equal to
u + V 2 2 + gZ . It is usual to separate the rate of work done into two parts, the

II
rate of shaft work and shear dWs dt and the rate of flow work at the control
G G
surface pV ⋅ dA where p is the pressure, resulting in
c .s .

dQ dWs
dt

dt c.s. II
G G
c .s .
II
G G ∂
− pV ⋅ dA = eρ V ⋅ dA +
∂t III
c .v .
ù
eρ V . (2.17)

This equation, written in integral form with some vector notation, is


sufficiently general to cover three-dimensional flow that is nonuniform,
unsteady, and compressible, as well as the gravitational and frictional effects
on the flow.

© 1999 Howden Buffalo, Inc.


2-8 FAN ENGINEERING

Assuming steady flow and substituting for e results in

çæ p + u + V + gZ äã ρ VG ⋅ dAG.
dQ dW
dt

dt
= II å ρ 2 â 2
(2.18)

Assuming one-dimensional steady flow and negligible variations in Z over


the entrance and exit areas yields

dQ dWs p çæV2 äã
p V2 çæ äã
dt

dt ρ2 å 2 â
ρ1 2 å
= 2 + u2 2 + gZ2 ρ 2V2 A2 − 1 + u1 + 1 + gZ1 ρ 1V1 A1 . (2.19)
â
Dividing by the mass flow rate and rearranging gives

p1 V12 p2 V22
+ u1 + + gZ1 + q = + u2 + + gZ2 + ws . (2.20)
ρ1 2 ρ2 2

Equation 2.20 is useful and relatively simple but does have restrictions.
First, it assumes that the flow is steady, which may not always be the case.
And second, it is based on one-dimensional flow, which means that the proper
average values across the section must be used for each term if the flow is
nonuniform. The appropriate averages for the flow work (sometimes called
pressure energy1), for the internal energy, and for the potential energy terms
are found by using volume-flow-weighted-average values of pressure and
density and mass-flow-weighted-average values of internal energy and
elevation. Since the area-weighted-average velocity is appropriate when
using the one-dimensional continuity equation for nonuniform flows, it would
be desirable to use the same average in the one-dimensional general energy
equation. However, the use of area-weighted average values for velocity does
not yield the proper average kinetic energy terms. Kinetic energy correction
factors2 must, therefore, be applied to each of the kinetic energy terms to
correct for this difference. The effect of these factors may or may not be
significant, depending on the extent to which the flow is uniform.
1
Grouping the flow work term with the energy terms, as is done to Equation 2.18, has led to
the concept of pressure energy. This concept is not theoretically proper because there is no
stored energy associated with pressure. Nevertheless, the expression is used and does not
cause any difficulties in practice.
2
See the following section.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-9

The equations in this section are dimensionally consistent only if the


dimensions used are force, length, and time or, alternatively, mass, length, and
time. If dimensions are used for both force and mass, each mass term must be
divided by the appropriate conversion factor gc . Mass is included in the
kinetic and potential energy terms. If heat is used as a dimension, each heat
term must be multiplied by the appropriate value of J , the mechanical
equivalent of heat. The internal energy and heat transfer terms are frequently
given dimensions of heat.
The general energy equation will be examined further in the sections for
incompressible and compressible flow situations.

Nonuniform-Velocity Correction Factors


When the velocity across a section is nonuniform, the area-weighted
average velocity can be used to determine kinetic energy or momentum only
if an appropriate correction factor is applied. The kinetic energy correction
factor is usually designated α , and the momentum correction factor is usually
designated β . Neither factor can be less than unity, but both factors do
approach unity in highly turbulent flows and, therefore, are frequently ignored
in fan engineering. The kinetic energy factor is

α=
II
A
V 3A
V 3dA
(2.21)

and the momentum factor is

β=
II
V 2A
V 2 dA
A
(2.22)

where V is the local velocity, V is the area-weighted average velocity, and A


is the area.

Specific Energy, Head, and Pressure


Equation 2.20, the general energy equation, is written so that each term
has dimensions of specific energy or energy per unit mass (or work per unit
mass). While these are the most convenient dimensions for deriving the
general energy equation, they are not the usual ones used in either hydraulics
or fan engineering. In hydraulics, head is usually used instead of specific
energy. Head can be obtained from specific energy by dividing by g , the
acceleration due to gravity. However, heads are seldom used in fan
engineering for the following reasons. First, the units of head are, in SI, the
meter of fluid or, in U.S. customary units, the foot of fluid; this leads to large
numerical values when air is the fluid. And second, it is difficult to directly
measure heads in air.

© 1999 Howden Buffalo, Inc.


2-10 FAN ENGINEERING

Therefore, the term usually used in fan engineering is pressure instead of


specific energy. Pressure can be obtained from specific energy by multiplying
by the mass density of the fluid. The usual units of pressure are the pascal in
SI and the inch water gage in U.S. customary units.
Each term in Equation 2.20 can be associated with a particular kind of
energy whether expressed as specific energy, head, or pressure. The total
energy at a section is the sum of the flow work and the internal, kinetic, and
potential energies at that point. The total head or total pressure at a section is
the sum of the corresponding heads or pressures. Not all of these terms are
named, as will be discussed in the following sections. Some of the terms may
be dropped or combined with others to make a new term.

Incompressible-Flow Energy Equations


The general energy equation can be rewritten for steady, one-dimensional
incompressible flow as

p2 − p1 α 2V22 − α 1V12 g ( Z2 − Z1 )
yF = + + + (u2 − u1 − q ) J (2.23)
ρ 2 gc gc

where all terms have dimensions of specific energy. The shaft work term in
Equation 2.20 represents the work that could be done by the fluid in a turbine,
and this turbine work is considered positive. In fan engineering it is
convenient to consider the shaft work done on the fluid (or fan work) as
positive, so the sign has been reversed and the symbol changed to yF. The
internal energy and heat transfer terms, when combined as shown, represent
the conversion of mechanical energy into thermal energy for incompressible
flow. This term can be considered the loss of mechanical energy since the
heat transfer and the increase in internal energy are not mechanically useful
under the incompressible assumption. For simplicity, the kinetic energy
factors are assumed to be equal to unity in the following discussion. This and
the assumption of steady one-dimensional flow inherent in Equation 2.23
limit the applicability of the resulting equations. Dividing Equation 2.23 by
g/gc changes the dimensions of each term to energy per unit weight, or head,
and gives

p2 − p1 V22 − V12
HF = + + ( Z2 − Z1 ) + H L1− 2 . (2.24)
w 2g

The flow work or pressure terms now contain w , the specific weight of the
fluid, and are called pressure heads. The kinetic energy terms are called
velocity heads. And the potential energy terms are called elevation heads.
The loss term is called the head loss term, and it contains all of the losses
.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-11

between 1 and 2 including the losses in any fans between those points. The
shaft work term is the gross fan head if the fan losses are left in the head loss
term. However, the shaft work term is the net fan head if the fan losses are
subtracted from the head losses. Heads are not normally used in fan
engineering, but there is some tendency to call pressures "heads," and vice
versa.
Multiplying Equation 2.23 by ρ changes the dimensions of each term to
energy per unit volume, or pressure, and gives

ρ (V22 − V12 ) ρg ( Z2 − Z1 )
pF = ( p2 − p1 ) + + + pL1− 2 . (2.25)
2 gc gc

The flow work or pressure terms are called pressures. The kinetic energy
terms are called velocity pressures, and the potential energy terms are called
elevation pressures. The loss term is called the pressure loss, and it contains
all of the losses between 1 and 2 including the losses in any fans between
those points. The shaft work term is the gross fan pressure if the fan losses
are left in the pressure loss term. But, the shaft work term is the net fan
pressure if the fan losses are subtracted from the pressure losses.
This last equation can be expressed in terms of gage static pressure pS ,
barometric pressure pB , velocity pressure pV , elevation pressure pZ , fan
pressure pF , and pressure losses p L1− 2 .

pF = ( pS 2 − pS 1 ) + ( pB 2 − pB1 ) + ( pV 2 − pV 1 ) + ( pZ 2 − pZ 1 ) + p L1− 2 . (2.26)

The sum of the static, barometric, velocity, and elevation pressures at a point
is the total pressure at that point. Often in fan engineering, the difference in
elevation between points 1 and 2 is very small, so the barometric and
elevation terms are dropped and the sum of the static pressure and the velocity
pressure is called the total pressure. Regardless of whether or not terms are
dropped, the fan pressure pF , the pressure losses p L1− 2 , and the total
pressures pT1 and pT 2 are related as follows:

pF = ( pT 2 − pT 1 ) + p L1− 2 . (2.27)

This equation will provide the basis for various discussions relating to fan
performance and system performance. Remember, however, that this
equation is based on incompressible, steady, one-dimensional flow with unity
kinetic energy factors. It is also based on negligible differences in elevation
whenever barometric and elevation terms are omitted. And there can be
exceptions to any of these conditions in fan engineering. Differences in
elevation are not negligible when there is a stack effect, for example.

© 1999 Howden Buffalo, Inc.


2-12 FAN ENGINEERING

Table 2.1 Velocities of Dry Air for Various Velocity Pressures


At 70ºF and 29.92 in. Hg Barometer
In fpm

in. wg 0 .1 .2 .3 .4 .5 .6 .7 .8 .9

0 0 1266 1791 2194 2533 2832 3102 3351 3582 3800


1 4005 4200 4387 4566 4739 4905 5066 5222 5373 5521
2 5664 5804 5940 6074 6204 6332 6458 6581 6702 6820
3 6937 7052 7164 7275 7385 7493 7599 7704 7807 7909
4 8010 8110 8208 8305 8401 8496 8590 8683 8775 8865
5 8955 9045 9133 9220 9307 9393 9478 9562 9645 9728
6 9810 9892 9972 10052 10132 10211 10289 10367 10444 10520
7 10596 10672 10747 10821 10895 10968 11041 11113 11185 11257
8 11328 11398 11469 11539 11608 11676 11745 11813 11881 11948
9 12015 12082 12148 12214 12279 12344 12409 12474 12538 12601

in. wg 0 1 2 3 4 5 6 7 8 9

10 12665 13283 13874 14440 14985 15511 16020 16513 16992 17457
20 17911 18353 18785 19207 19620 20025 20422 20811 21192 21568
30 21936 22299 22656 23007 23353 23694 24030 24361 24688 25011
40 25330 25644 25956 26263 26566 26866 27163 27457 27747 28035

Table 2.2 Velocities of Dry Air for Various Velocity Pressures


At Various Temperatures and 29.92 in. Hg Barometer
In fpm

in.wg 50ºF 60ºF 70ºF 80ºF 100ºF 150ºF 200ºF 300ºF 500ºF 650ºF

.1 1242 1254 1266 1278 1302 1359 1414 1517 1705 1833
.2 1757 1774 1791 1808 1841 1922 1999 2145 2411 2592
.3 2152 2173 2194 2214 2255 2354 2448 2627 2953 3175
.4 2485 2509 2533 2557 2604 2718 2827 3033 3410 3666
.5 2778 2805 2832 2859 2911 3038 3161 3391 3812 4099
.6 3043 3073 3102 3131 3189 3328 3462 3715 4176 4490
.7 3287 3319 3351 3382 3444 3595 3740 4013 4510 4850
.8 3514 3548 3582 3616 3682 3843 3998 4290 4822 5185
.9 3727 3763 3800 3835 3905 4076 4241 4550 5114 5500
1.00 3929 3967 4005 4043 4117 4297 4470 4796 5391 5797
1.25 4392 4435 4478 4520 4603 4804 4998 5362 6027 6481
1.50 4812 4859 4905 4951 5042 5263 5475 5874 6602 7100
1.75 5197 5248 5298 5348 5446 5684 5913 6345 7132 7669
2.00 5556 5609 5664 5717 5822 6077 6322 6783 7624 8198
2.25 5893 5951 6007 6064 6175 6446 6705 7194 8087 8696
2.50 6212 6272 6332 6392 6509 6794 7068 7583 8524 9166
2.75 6515 6577 6642 6704 6827 7126 7413 7953 8940 9613
3.00 6805 6871 6937 7002 7130 7443 7742 8307 9338 10040
4.00 7858 7934 8010 8085 8233 8594 8940 9592 10780 11590

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-13

For incompressible flow, the velocity pressure, density, and velocity are
related as indicated by
ρV 2 2 gc C p pV
pV = and V = . (2.28)
2 gc C p ρ

In SI units, pV is in Pa, ρ is in kg/m3, V is in m/s, gc is 1.0 m-kg/N-s2, and


1.0 Pa = 1.0 N/m2,

ρV 2 2 pV
pV = and V = . (2.29)
2 ρ

In U.S. customary units, pV is in in. wg, ρ is in lbm/ft3, V is in fpm, gc is


32.174 ft-lbm/lb-s2, and 1 in. wg = 5.193 lb/ft2, so

ç V äã
= ρæ
2
pV
pV
å 1097 â and V = 1097
ρ
. (2.30)

For standard air, ρ is 0.075 lbm/ft3, so

çæ V äã 2

pV =
å 4005â and V = 4005 pV . (2.31)

This indicates that for standard air a velocity pressure of one-inch water gage
corresponds to 4005 feet per minute, and vice versa. Numerous other
solutions of this equation are given in Table 2.1. Table 2.2 gives velocities
for various combinations of velocity pressure and temperature.
Another interesting value is 69.24, the number of ft of standard air
equivalent to 1 in. wg.
The useful power Po delivered by a fan to an incompressible fluid can be
defined using either the net fan work y F , the net fan head H F , or the net fan
pressure pF , by means of

y F mû H F wû p F Qû
Po = = = (2.32)
Cm Cw CQ

where mû is the mass flow rate, wû is the weight flow rate, and Qû is the
volume flow rate through the fan. The value of the constant Cm , Cw , or CQ
depends on the units used, as indicated in Table 2.3. In fan engineering, this
useful power has been called the air horsepower or, more generally, the air
power. It is the minimum power required to move air at the specified rate
against the specified resistance and can be considered the output power of a
fan for those conditions.

© 1999 Howden Buffalo, Inc.


2-14 FAN ENGINEERING

Note that density does not appear as a variable in any of the expressions
for air power.
The air power Po divided by the shaft power Ps gives the incompressible
flow efficiency η . See Equation 2.42 for comparisons with compressible
flow efficiencies.

Table 2.3 Units and Values for Equation 2.32

Po yF mû Cm HF wû Cw pF Qû CQ

W J/kg kg/s 1.0 m N/s 1.0 Pa m3/s 1.0


kW J/kg kg/s 1000 m N/s 1000 kPa m3/s 1.0
hp ft-lb/lbm lbm/s 550 ft lb/m 33000 in.wg cfm 6354
kW J/kg kg/s 1000 m kp/s 0.102 mm wg m3/s 102.1

Example 2.1 Incompressible Flow - Pressure, Power, and Efficiency

Given the diagram and information below, find the net fan pressure, the
velocity pressure at 4, the air power, and the efficiency.

FAN

5 3 1 2 4 6

pS 5 = pS 6 = 0 in. wg,
V1 = V2 = V3 = V4 = 2000 fpm, V5 = V6 = 0 fpm ,
Z1 = Z2 = Z3 = Z4 = Z5 = Z6 = 0 ft,
p L5−1 = 5 in. wg, p L2 − 6 = 5 in. wg,
ρ = 0.075 lbm/ft3,
Qû = 1000 cfm, and
Po = 2.0 hp

Using Equations 2.25, 2.30, and 2.32 with Table 2.3:


ρ (V62 − V52 ) ρg ( Z6 − Z5 )
pF = ( p6 − p5 ) + + + p L5−1 + p L 2 − 6 ,
2 gc gc
pF = 0 + 0 + 0 + 5 + 5 = 10 in. wg,
ç V äã
= ρæ
2
ç 2000äã
= 0.075 æ
2

å 1097 â å 1097 â = 0.25 in. wg,


4
pV 4

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-15

pF Qû 10 × 1000
Po = = = 1.57 hp, and
CQ 6354
Po 1.57
η= = = 0.786
Ps 2.0
Po can also be determined from mû and yF :
ρ Qû 0.075 × 1000
mû = = = 1.25 lbm/s
60 60
C p 5.193 × 10
yF = p F = = 692 ft ⋅ lb/lbm, and
ρ 0.075
y mû 692 × 1.25
Po = F = = 1.57 hp,
Cm 550
where C p is from Table 2.4.

Compressible-Flow Energy Equations


The general energy equation can be rewritten for steady, one-dimensional
compressible flow as

 p2   p1  α 2V22 − α1V12 g ( Z 2 − Z1 )
yF =  + u2 J  −  + u1 J  + qJ + + (2.33)
 ρ2   ρ1  2 gc gc

where all the terms have the dimensions of specific energy. As in the
previous section, the sign on the work term has been reversed to make fan
work positive. Shaft work can be determined from this equation if states 1
and 2 are defined and if the heat transfer from the surroundings can be
determined. If so, it might be appropriate to combine the pressure and
internal energy terms into enthalpy terms. The terms of Equation 2.33 cannot
be transformed into pressure terms simply by multiplying by the density, as in
the incompressible case of the preceding section, because the density varies.
Nevertheless, pressure terms are usually preferred in fan engineering. It will
be necessary to use approximations if pressures are to be used for
compressible flow. Suitable approximations can be obtained in various ways
including the use of a mean density and assumptions about the process. Some
idealized processes are examined below.
There are various thermodynamic processes that can be described by the
expression pv x = constant where p is the pressure, ν is the specific volume,
and x is an exponent that depends on the process.

© 1999 Howden Buffalo, Inc.


2-16 FAN ENGINEERING

For the isochoric or constant density process, x = ∞ . This process can


also be called isometric or constant volume. For an isobaric or constant
pressure process, x = 0 . A constant-temperature or isothermal process in a
perfect gas has x = 1. An isentropic process in a perfect gas has x = γ ,
where γ is the ratio of specific heats for the gas undergoing the process. This
constant entropy process is reversible and adiabatic; that is, there is no friction
and no heat transfer. A polytropic process can have any other positive
exponent n . Many actual processes can be approximated by one or more of
these mathematically simple processes. In addition, these processes can
provide standards of comparison for an actual process.
Although the derivation will not be given here, Equation 2.33 can be
rewritten as below for a reversible process that follows pv x = constant .

á
x çp ä
x −1
"# α V − α V
æ ã g ( Z2 − Z1 )
2 2
p
## 2 g
x

x −1 å p â
yF = 1 2
−1 + 2 2 1 1
+ . (2.34)
ρ1
! 1
$ c gc

This is reducible to the following for frictionless incompressible flow:

p2 − p1 α 2V22 − α 1V12 g ( Z2 − Z1 )
yF = + + . (2.35)
ρ 2 gc gc

The reversible compressible flow case can be expressed in a similar way by


incorporating a compressibility factor K p

yF =
1 p − p 6K
2 1 p
+
α 2V22 − α 1V12 g ( Z2 − Z1 )
+ (2.36)
ρ1 2 gc gc

where

çæ p äã
å p â for reversible isothermal processes,
2
ln
=
çæ p − 1äã
1
Kp (2.37)

åp â
2

γ çp ä
á "# γ −1

æ ã − 1#
γ

γ −1 å p â
2

! 1
#$ for isentropic processes,
Kp =
çæ p − 1äã (2.38)

åp â
2

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-17

áç p ä n −1
"#
n
æå p ãâ
n

n −1
2
−1
##
! 1
$
Kp =
çæ p − 1äã for reversible polytropic processes, (2.39)

åp â
2

K p = 1 for reversible isochoric processes, and

K p = 0 for reversible isobaric processes.

Note from Equation 2.36 that ρ 1 / K p can be considered a mean density.


The shaft work yF obtained from Equations 2.34 - 2.36 is the work
required to produce flow from 1 to 2 for an ideal reversible process. If the fan
inlet is at 1 and the fan outlet is at 2, yF can be considered the ideal work
output of the fan. This work output will vary with the process as illustrated in
Example 2.2. Losses in the fan are not considered since the process is
reversible in each case, but an appropriate amount of heat transfer, depending
on the process, is considered.
Multiplying Equation 2.34 by the mass flow rate mû yields the ideal power
for the reversible process pv x = constant:

p Qû x
áç p ä x −1
"# áα V − α V "#
g ( Z 2 − Z1 )
æå p ãâ
2 2

## ! 2 g
x
Po = 1 1 − 1 + mû +
$
2 2 2 1 1
. (2.40)
CQ x − 1
! 1
$ c gc

In terms of the pressure difference and K p ,

P =
1 p − p 6Qû K
2 1 1 p
+ mû
áα V − α V
2 2
+
1
g Z2 − Z1 6 "#
! 2g $
2 2 1 1
o . (2.41)
CQ c gc

Both Equations 2.40 and 2.41 contain the volumetric flow rate Qû , at section
1. Refer to Table 2.3 for values of CQ for various units. The efficiency η of
an actual process can be defined as the ideal power Po divided by the shaft
power Ps
P
η= o . (2.42)
Ps

Any process can be chosen as the standard of comparison. Using the


isothermal process leads to the isothermal efficiency. Similarly, the
polytropic

© 1999 Howden Buffalo, Inc.


2-18 FAN ENGINEERING

process yields the polytropic efficiency. When the isentropic process is used,
the efficiency can be called the isentropic or adiabatic efficiency. And using
the isochoric process leads to the isochoric efficiency, which has the same
value as the incompressible flow efficiency.

Example 2.2 Compressible Flow - Work, Power, and Efficiency

Given the diagram and information below, find the work output, ideal power,
and efficiency for reversible isochoric, isentropic, and reversible isothermal
processes in the fan. Assume that differences in kinetic and potential energies
are negligible.

FAN

5 3 1 2 4 6

p1 = 408 in. wg abs., p2 = 418 in. wg abs.,


∴ p2 − p1 = 10 in. wg or 51.9 lb/ft2,
ρ1 = 0.075 lbm/ft3, and
mû = 1.25 lbm/s ∴ Qû = 1000 cfm.

Using Equations 2.36, 2.41, 2.42, 2.37, and 2.38:

yF =
1 =
6
p2 − p1 K p 519 . Kp
= 692 K p ft-lb/lbm,
ρ1 0.075

Po =
1 û
p2 − p1 QK p6=
10 × 1000 K p
= 1.57 Kp hp,
CQ 6354
Po Po
η= = , and
Ps 2
á
γ çp ä
"# γ −1

ç p ä γ -1 æå p ãâ − 1#
lnæ ã
2
γ

å p â , or ! #$
2
1
K p = 1,
çæ p − 1äã çæ p − 1äã
1

åp â åp â
2 2

1 1

Process Kp yF Po η
reversible isochoric 1.0 692 ft-lb/lbm 1.57 hp 78.6 %
isentropic 0.9914 686 ft-lb/lbm 1.56 hp 77.9 %
reversible isothermal 0.9879 684 ft-lb/lbm 1.55 hp 77.6 %

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-19

Mach Number
The ratio of the stream velocity V to the sonic velocity c is called the
Mach number Ma.

V
Ma = (2.43)
c

The Mach number is also a measure of the ratio of the inertial force to the
elastic force and a measure of the kinetic energy to the internal energy, at a
point.
Dimensional analyses1 indicate and experimental data demonstrate that
Mach number is an important natural physical variable in compressible flow
situations. Refer to the chapter on fan laws for a discussion of dynamic
similarity and Mach number effects on fan law predictions.
The velocity of sound c in a perfect gas is a function of the ratio of
specific heats γ and the state of the gas, or

g cγ p
c= = g c γ RT (2.44)
ρ

where p is the pressure, ρ is the density, T is the absolute temperature, and


R is the gas constant. The speed of sound in air is tabulated for various
conditions in Table 1.2. While the sonic velocity is of the order of 1100 fps or
335 m/s for air, it is four times higher in hydrogen and about five times higher
in most common liquids. It is only about one-fourth of those values in very-
high-molecular-weight gases.

Temperature Rise Due to Compression


The temperature at any state can be calculated from other state properties
using the equation of state as discussed in Chapter 1. The temperature ratio
for any two states can be calculated from the pressure ratio using the equation
for the particular process if the process is known. The temperature rise
T2 − T1 experienced by a gas during compression can be calculated using the
absolute inlet temperature T1 , the absolute pressure ratio p2 / p1 , and

á% p ( x −1
"#
T −T = T & )
##
x
−1
!' p *
2
(2.45)
$
2 1 1
1

for any reversible process that follows pv x = constant .


The temperature rise can also be expressed in terms of the pressure ratio
p2 / p1 . This is facilitated by using the compressibility coefficient K p , which
.
1
See Appendix A for a discussion of dimensional analysis.

© 1999 Howden Buffalo, Inc.


2-20 FAN ENGINEERING

was introduced together with the exponent x in the discussion of the energy
equation for compressible flow, and

T2 − T1 =
T1
p1
1p2 − p1 K p6 çæå
x −1
x
.
äã
â (2.46)

For a reversible isothermal process, x = 1 and T2 − T1 = 0 .

For a reversible adiabatic or isentropic process, x = γ and

T1
1 6çæå
γ −1 äã 1
p2 − p1 K p C p 6
T2 − T1 =
p1
K p p2 − p1
γ
=
â
ρ 1c p J
. (2.47)

The second equality includes the inlet density ρ1 , the specific heat cp, and two
constants, Cp, and J, whose values can be determined from Table 2.4.

For a reversible polytropic process, x = n and

T2 − T1 =
T1
p1
1
K p p2 − p1
n
6çæå äãâ 1
n −1
=
p2 − p1 K p C p
ρ 1c p Jη p
6 (2.48)

where η p , the polytropic efficiency, is a function of γ and n and can be


calculated from

n γ −1
ηp = . (2.49)
n −1 γ

Note that as n approaches infinity (which corresponds to the constant density


or isochoric case) the polytropic efficiency approaches a constant value, e.g.
0.2857 for air. Experience indicates that the temperature rise in a fan is better
approximated by using a value for the polytropic efficiency that varies with
the degree of perfection in the design. This is still another illustration of the
need to examine carefully the effects of simplifying assumptions. The
assumption of a constant density process may serve well in some cases, but it
cannot be used when calculating temperature rise.
Temperature rise can also be calculated from the shaft power Ps the rate
of heat transfer Qû H to the fluid, and the mass flow rate mû as follows

Ps Cm + Qû H J
T2 − T1 = (2.50)
û
c p mJ

using units and values from Table 2.4.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-21

Table 2.4 Units and Values for Equations 2.46-2.50

t p ρ cp Cp J Ps QûH mû Cm
3
K kPa kg/m J/kg 1000 1.0 kW J/s kg/s 1000
°R in. wg lbm/ft3 Btu/lbm 5.193 778.2 hp Btu/s lbm/s 550

Example 2.3 Temperature Rise

Given the data below, find the temperature rise.

ρ 1 = 0.075 lbm/ft3 = 1.2 kg/m3,


p2 − p1 = 10 in. wg = 2.4836 kPa,
K p = 0.9923,
η p = 0.780 for a polytropic process,
P0 = 2.0 hp = 1.4914 kW,
mû = 1.25 lbm/s = 0.5670 kg/s, and
c p = 0.24 Btu/lbm-°R = 1005 J/kg-K.

Using Equation 2.48:

T2 − T1 =
1 6
p2 − p1 K p C p
=
10 × 0.9923 × 5193
.
= 4.72 °F
ρ 1c p Jη p 0.075 × 0.24 × 778.2 × 0.780
and

T2 − T1 =
1 6
p2 − p1 K p C p 2.4836 × 0.9923 × 1000
= = 2.62 °C.
ρ 1c p Jη p . × 1005
12 . × 10
. × 0.780

Using Equation 2.50 and assuming that Qû H = 0:


P C + Qû H J 2.0 × 550 + 0
T2 − T1 = s m = = 4.71 °F and
û
c p mJ 0.24 ×1.25 × 778.2
P C + Qû H J 1.4914 × 1000 + 0
T2 − T1 = s m = = 2.62 °C.
û
c p mJ 1005 × 0.5670 × 1.0

The results of using Equations 2.48 and 2.50 are substantially equivalent,
although there are small differences due to rounding off. However, the
assumption that Qû H = 0 may not be appropriate for other cases, e.g., when
calculating temperature rise for a fan at shut-off.

A common rule of thumb for fan temperature rise is 0.5°F per in. wg. A
corresponding figure is 1.1°C per kPa.

© 1999 Howden Buffalo, Inc.


2-22 FAN ENGINEERING

Stagnation Properties
When a fluid moving along a streamline is brought to rest an increase in
temperature, pressure, and density results. The stagnation properties can be
determined by considering what happens at the stagnation point. Assuming
an adiabatic process, the stagnation temperature Tτ , can be determined from
the free-stream temperature T∞ , the free-stream velocity V∞ , and the specific
heat cp using

2
V∞
Tτ = T∞ + . (2.51)
2 gc c p J

Assuming a reversible adiabatic or isentropic process, the stagnation


pressure pτ , can be determined from the free-stream pressure p∞ , the free-
stream Mach number Ma ∞ , and the ratio of specific heats γ using absolute
pressures and

á "#
γ

2 γ −1
γ −1
= p 1 + Ma
! $
pτ ∞ ∞ . (2.52)
γ

This can also be expressed in terms of the free-stream velocity V∞ and the
free-stream density ρ ∞ using

pτ = p∞ +
ρ ∞V∞2 á
1+
Ma 2∞
+ 2 −γ 1
Ma 4∞
+... . 6 "#
! $
(2.53)
2 gc 4 24

Note that the stagnation pressure pτ is the sum of the static pressure p∞ and a
term that is reducible to the velocity pressure for incompressible flow when
the Mach number is very small.
The stagnation density can be calculated from the stagnation pressure and
stagnation temperature using


ρτ = (2.54)
RTτ

where R is the specific gas constant.

Radial and Vortex Flow


There are several flow situations that can be classified as radial flow or
vortex flow. These cases have practical applications in fan design and other
aspects of fluid flow.
Ideally, in the radial flow of air through a device such as that indicated in
.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-23

Figure 2.1, the radial velocity V , assuming uniform flow, varies inversely
with the radius R :

V2 R1
= . (2.55)
V1 R2

Disregarding friction, the conversion of velocity pressure into static pressure


in a radial diffuser of this type can be determined from

á1 − ç R ä "#.
2

∆ pS 2 −1 = pV 1 æå R ãâ #
1
(2.56)
! 2
$
This is equal to the difference in velocity pressures.
The flow at AA must be essentially radial to prevent separation from the
guiding surfaces. Equation 2.56 applies only to frictionless flow without
separation. The flow will be without separation only if the dimensions are
approximately as shown in Figure 2.1. When the discharge is into the
atmosphere, the average pressure within the radial diffuser will be less than
the atmospheric pressure, so that the confining walls will tend to approach
each other in spite of the impingement of the jet.

Figure 2.1 Confined Radial Flow

A free jet impinging against a flat plate has the velocity pattern indicated
in Figure 2.2. The central portion slows as it approaches the plate, producing
a corresponding increase in static pressure. This is followed by a
reconversion of static pressure to velocity pressure, which in the frictionless
case .

© 1999 Howden Buffalo, Inc.


2-24 FAN ENGINEERING

would reaccelerate the flow to the original velocity. This type of flow, when
bounded by walls approximating the natural flow boundaries, is an efficient
means of turning an air stream within a very short distance.

Figure 2.2 lmpingement of a Free Jet on a Flat Plate

The term vortex flow is used to describe the motion of a fluid whenever a
whirl exists. A circular vortex is one without a radial component; a spiral
vortex is one that has a radial component. The radial component of a spiral
vortex can be directed inward or outward.
In a free circular vortex, the velocity varies inversely with the radius
according to Equation 2.55, just as for radial flow. If there were no friction,
the velocity at the center would become infinitely great. In a free spiral
vortex, both the radial and tangential components of the velocity vary
inversely with the radius. Various types of vortex flow occur in nature as well
as in certain fan applications. Free spiral vortex conditions are approached in
the scroll-shaped housing of centrifugal fans and also in cyclone collectors.
The velocity-radius relation, as expressed in Equation 2.55, can be greatly
modified by friction and other forces. The extent of this modification can be
appreciated by noting that, in a frictionless straight-blade centrifugal fan that
produces one kind of forced vortex, the tangential velocity varies directly with
the radius, as indicated by

V2 R2
= . (2.57)
V1 R1

The change in static pressure due to such a change in tangential velocity can
be determined from

á1 − ç R ä "#.
2

∆ pS 2 −1 = pV 1 æå R ãâ #
2
(2.58)
! 1
$

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-25

Effects of Viscosity
The flow of any real fluid is resisted by friction forces that arise because
of the viscosity of the fluid. These forces are generated within the fluid
wherever there are velocity gradients. They are transmitted between the fluid
and any solid boundary by a layer of fluid that becomes attached to the
boundary. Forces are transmitted between this fluid and any other fluid by a
mixing layer of entrained fluid. The acceleration of a fluid is also resisted by
inertia forces. If flow is to take place, work must be done to overcome these
resistances. The energy required can be transferred to the fluid by a fan or by
some other means. In a fan system, this mechanical energy is delivered to the
air by the fan (or fans) and that portion which overcomes friction is gradually
converted into thermal energy as resistance is encountered throughout the
system.

Flow Regimes and Boundary Layers


Fluid flow can be laminar, turbulent, or transitional (in a transition state
between these two regimes). Laminar flow, as the name implies, is
considered to proceed in layers between which there is relative sliding motion.
It is characterized by the absence of local macroscopic velocity fluctuations.
Shearing stresses are transmitted essentially by intermolecular forces and
random molecular motion. Turbulent flow exhibits relatively large-scale local
velocity fluctuations. This results in eddying, mixing, and transport of
momentum, which becomes the main shearing-force-transmission mechanism
between adjacent portions of the flowing fluid. The presence of a boundary
surface in a flowing fluid produces various phenomena. As the fluid attaches
itself to the boundary, gradually increasing amounts are retarded forming a
boundary layer whose thickness increases in the direction of flow. The flow
in this layer can be laminar or turbulent, depending on the Reynolds number
of the main flow and on the disturbances present. If the boundary is a closed,
uniform conduit of sufficient length, the flow will eventually become fully
developed; that is, the velocity profile will be the same for all succeeding
sections. This condition will not occur if the conduit is nonuniform or
strongly curved. Similarly, with external flow around an immersed blunt
body, the boundary layer may not have sufficient kinetic energy to overcome
the resulting positive pressure gradients and other resistances. In this case, the
flow will separate, causing return flow and eddies that form and reform and
are swept out into the main stream.
Disturbed, redeveloping flow may persist for 50 diameters of straight
conduit after a disturbance. With a smooth entrance, flow may be fully
developed in 20 diameters or less, depending on Reynolds number.
Separation does not occur in accelerated flow. In contrast, decelerated
flow does promote separation. Since separation leads to eddy formation and,
.

© 1999 Howden Buffalo, Inc.


2-26 FAN ENGINEERING

ultimately, to energy dissipation, accelerated flow is usually less troublesome


and more efficient than decelerated flow.
Another phenomenon due to the presence of boundaries is secondary flow.
This can occur in either unestablished or established flow. It is the flow-
within-a-flow that occurs, for example, at bends or discontinuities

Dynamic and Kinematic Viscosity


Dynamic viscosity µ , also called absolute viscosity or coefficient of
viscosity, is that property of a fluid that resists the movement of one layer
over another. Mathematically, it is the proportionality factor relating shear
stress τ and fluid strain rate. In simple flows, strain rate is given by the
velocity gradient du dy (incremental velocity per incremental distance), or

du
τ =µ . (2.59)
dy

Kinematic viscosity ν is absolute viscosity divided by mass density ρ , or

µ
ν= . (2.60)
ρ

The dimensions1 of kinematic viscosity are L2 / T (the dimensions of


kinematics). The dimensions of dynamic viscosity are FT / L2 or M / TL
(the dimensions of dynamics). Various units of measurement2 are used for
viscosity; some with special names like poise and stokes. In addition, the
kinematic viscosities of liquids are sometimes reported as the measurement of
a particular viscosimeter, e.g. 100 SSU or 100 seconds efflux time on the
Saybolt Universal viscosimeter. Calibration data like that in Figure 2.7 are
needed to convert efflux time to kinematic viscosity.
As illustrated in Figures 2.3 through 2.6, the viscosity of a gas or a vapor
increases with increasing temperature due to increasing molecular agitation,
whereas the viscosity of a liquid decreases with increasing temperature due to
weakening intermolecular forces.

Reynolds Number
The Reynolds number Re at a point in a fluid stream is the ratio of the
inertia force to the viscous shearing force acting on an element of fluid at that
point. It is the dimensionless combination of some characteristic linear
dimension of the boundary surface D , the relative velocity V between the
element and that surface, and the physical properties of the fluid as
represented by the absolute viscosity µ and the mass density ρ or the
kinematic viscosity ν :
1
See Appendix A for a discussion of dimensions.
2
See Appendix C for conversion factors for units.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-27

Figure2.3 Dynamic Viscosities of Common Gases


Adapted from the data of G. A. Hawkins, H. L. Solberg, and A. A. Potter: "The Viscosity of
Superheated Steam," Trans. ASME, vol. 62, pp. 677-688, 1940, and that of SAE: Aeronautical
Information Report No. 24, 1952.

Figure 2.4 Dynamic Viscosities of Refrigerant Vapors


Adapted from the data of J. C. Reed and E. E. Ambrosius: "Viscosity of Refrigerants,"
Heating, Piping and Air Conditioning, June 1930, pp. 455-461, and that of A.E Benning and
W. H. Markwood, Jr.: "The Viscosities of' Freon Refrigerants,' " Journal of the ASRE in
Refrigerating Engineering, April 1939, pp. 243-247.

© 1999 Howden Buffalo, Inc.


2-28 FAN ENGINEERING

Figure 2.5 Dynamic Viscosities of Liquids


Adapted from the data of SAE: Aeronautical Information Report No. 24, 1952.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-29

Figure 2.6 Kinematic Viscosities of Gases and Liquids

Adapted from the data of SAE: Aeronautical Information Report No. 24, 1952.

© 1999 Howden Buffalo, Inc.


2-30 FAN ENGINEERING

Figure 2.7 Conversions from Efflux Time to Kinematic Viscosity


Adapted from the data of TEMA: Standards of the Tubular Exchanger Manufacturers
Association, TEMA, 1959.

DVρ DV
Re = = . (2.61)
µ ν

Some characteristic linear dimensions D are: the diameter of the opening for
orifices, the diameter of the exit opening for nozzles, the inside diameter for a
round conduit, the equivalent diameter based on the same hydraulic radius for
a rectangular duct and the chord length for an airfoil.
In all these devices, the local Reynolds number across a section varies
from point to point because of the velocity variation. Therefore, it is
convenient to use a Reynolds number based on the mean velocity Vm .
Dimensional analyses1 indicate, and experimental data demonstrate, that
Reynolds number is an important, natural physical variable in various flow
situations. Refer to the chapter on fan laws for a discussion of dynamic
similarity and Reynolds-number effects on fan law predictions. Refer also to
1
See Appendix A for a discussion of dimensional analysis.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-31

the particle dynamics chapter for information on the effects of Reynolds


numbers on particle behavior in an aerosol. The discussions that follow here
are concerned directly with flow in conduits, but the principles apply to many
flow situations.

Absolute and Relative Roughness


Any surface, no matter how polished, has peaks and valleys. The mean
distance between these high and low points is the absolute roughness ε .
Table 2.5 lists several roughness conditions with typical surfaces and
corresponding absolute roughnesses expressed in feet. The relative roughness
ε / D of the surface in a conduit is the absolute roughness divided by the
effective diameter. Any consistent set of units can be used.
Hydraulically, any value of relative roughness can represent either a
smooth or a rough condition depending on the Reynolds number. A brief
analysis of various flow phenomena will explain this. It has been established
that, for all but the most rarefied flow, there is a molecular layer of fluid
firmly anchored to the boundary surfaces. In laminar flow, there is a velocity
gradient all across the stream. In turbulent flow, there are velocity
fluctuations, but in the main stream, the average velocity profile in a pipe is
almost flat. However, a turbulent boundary layer is formed across which
there is a definite velocity gradient. There may also be a very thin laminar
sublayer if the absolute roughness is small compared to the boundary layer
thickness. If such a sublayer submerges the high points of the surface
sufficiently, that is, if sublayer thickness exceeds absolute roughness by a
factor of 4, the pipe surface can be described as hydraulically smooth.
In the wholly rough zone of turbulent flow, the condition of the surface
.

Table2.5 Roughness Information for Various Conditions

Condition Typical Surface Average ε Range ε n4 c4


Very smooth Drawn tubing .000005 ft - .20 0.1036
Medium smooth Aluminum duct1, 3 .00015 ft .00010'-.00020' .18 0.0807
Average Galvanized iron duct2 .0005ft .00045'-.00065' .16 0.0746
Medium rough Concrete pipe .003 ft .001'-.01' .14 0.0642
Very rough Riveted steel pipe .01 ft .003'-.03' .12 0.0560

1
Crimped slipjoint every 3'
2
Crimped slipjoint every 2-1/2'
3
New steel pipe also typical
4
n and c are for use in Equation 2.78.

Adapted from the data of F. W. Hutchinson: "Friction Losses in Round Aluminum Ducts,"
Trans. ASHVE, vol. 59, 1953, pp. 127-138.

© 1999 Howden Buffalo, Inc.


2-32 FAN ENGINEERING

can be described as hydraulically rough. The laminar sublayer is reduced to


about 1/6 or less of the absolute roughness and can be prevented from forming
at all as the roughness increases in comparison to the boundary layer
thickness.

Darcy Friction Factor


The Darcy friction factor f for flow in pipes is a dimensionless group
that relates two other dimensionless groups: the ratio of the loss of total head
between two points H L to the velocity head HV , and the ratio of the distance
between those two points L and some characteristic dimension D that
determines velocity.

HL L
= f . (2.62)
HV D

Pressures can be substituted for heads when the flow is considered


incompressible.
In the laminar zone,

64
f = (2.63)
Re

which appears as a straight sloping line on the Moody Chart in Figure 2.8.
There is a critical value of Re above which laminar flow can exist only if the
flow remains essentially undisturbed. Since this condition is quite easily
upset, a dotted line is shown for Reynolds numbers greater than about 2100.
This critical Reynolds number is for conduit flow, but for other types of flow,
the critical values will be different. For example, for flow between parallel
plates, the critical Reynolds number, based on the distance between plates, is
approximately 900.
In the wholly rough zone,

1
f =
çæ 2 log 3.7 äã , (2.64)

å ε D, â
which produces a series of straight horizontal lines in Figure 2.8. There is a
value of Re for each value of relative roughness, below which the flow cannot
be considered independent of Reynolds number. These values are indicated
by the dashed, curved line in Figure 2.8.
In the transition zone,

ç
ε D ä
1
f
= −2 log æå +
2.51
3.7 Re f
. ãâ (2.65)

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-33

Figure 2.8 Moody Chart for Darcy Friction Factor


Adapted from the data of L. F. Moody: "Friction Factors for Pipe Flow," Trans. ASME, vol.
66, 1944, pp. 671-684.

This is the Colebrook1 equation, which is based on the research of Nikuradse,


von Kármán, and others.
For hydraulically smooth pipes,

0.3164
f = . (2.66)
Re 0.25

This is the Blasius formula, which can be used for Reynolds numbers between
3000 and 105.
Also for smooth pipes,

ç ä
1
f
= −2 log æå
2.51
Re f
. ãâ (2.67)

This is the Prandtl universal resistance law, which can be used for Reynolds
numbers between 5 × 104 and 3.5 × 106.
An explicit function that gives satisfactory values over a wide range of
Reynolds numbers is
1
C. F. Colebrook: "Turbulent Flow in Pipes with Particular Reference to the Transition Points
Between Smooth and Rough Laws, "ICE Journal, vol. 11, 1938, pp. 133-156.

© 1999 Howden Buffalo, Inc.


2-34 FAN ENGINEERING

álogçæ ε D + 5.74 äã "# .


2

f = 0.25
! å 3.7 Re â $ 0.9
(2.67a)

Example 2.4 illustrates the use of the Darcy friction factor in a pressure-loss
calculation.

Example 2.4 Pressure Loss Using Friction Factor

Given 100 ft of 12-in. diameter galvanized iron duct with 1000 cfm of 70°F
air flowing, find the pressure loss.

Using Figure 2.6, Equation 2.61, and Table 2.5:

ν = 163
. × 10 −4 ft2/s,
A = 0.7854 ft2 for 12-in. pipe,
V = 1000/0.7854 = 1273 fpm = 21.22 fps,

DV 1 × 2122
.
Re = = = 130200,
ν . × 10−4
163

ε = 0.0005 ft, and


ε 0.0005
= = 0.0005.
D 1

Using Figure 2.8:

f = 0.020.

Alternatively, using Equation 2.65:

ç
ε D ä = −2 logç 0.0005 + 2.51 ä
1
f
= −2 log æå +
2.51
3.7 Re f ãâ æå 3.7 130200 f ãâ and

f = 0.020 (computation requires an iteration procedure).

Using Equations 2.30 and 2.62:

ç V äã
= ρæ
2
ç 1273äã
= 0.075 æ
2

pV
å 1097 â å 1097 â = 0.10 in. wg and

L 100
pL = f pV = 0.020 × × 010
. = 0.20 in. wg.
D 1

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-35

Pressure Losses for Duct Elements


In fan systems, the flow through each duct element (including straight
ducts, elbows, diffusers, hoods, etc.) can almost always be considered
incompressible. Accordingly, the incompressible-flow energy equation, as
expressed in Equation 2.23, can be used to examine the energy relationships
for a duct element. Excluding fans as duct elements, the fan work term y F
will be zero. The combined heat transfer and internal energy term
1 6
u2 − u1 − q J can be considered the loss of mechanical energy. Regardless of
whether this equation is expressed in terms of energy per unit mass, head, or
pressure, the loss is equal to the sum of the three remaining terms. One term
results from the change in static pressure, another from the change in velocity,
and the last from a change in elevation. In other words, the loss of
mechanical energy per unit mass y L1− 2 for a duct element is equal to the
change in total specific energy across that element. Similarly, the head loss
H L1− 2 is equal to the change in total head. Finally, the pressure loss p L1− 2 is
equal to the change in total pressure across the element.
For most duct elements, the change in elevation between 1 and 2 is
negligible. For many duct elements, the velocity is sufficiently uniform at
both 1 and 2 to assume that the kinetic energy correction factors α 1 and α 2 are
unity. For some duct elements, the kinetic energies at 1 and 2 are equal, so
the loss on a pressure basis equals the change in static pressure. However,
this is an exceptional case, and it is best to consider any losses based on
pressure to be equal to the change in total pressure.
There are many compilations that give data for calculating losses for duct
elements. Perhaps the most complete is that of Idel'chik.1 Since it is far too
extensive to repeat here, only data sufficient to cover most cases, at least
approximately, are given below. Both ASHRAE and SMACNA have taken
much of their data from this source.
The total pressure loss for a duct element is a function of the configuration
of the element and of the flow through the element. As noted in the
discussion on Darcy friction factor, certain dimensionless groups can be used
to conveniently express these relationships. These groups include relative
roughness, number of diameters, Reynolds number, and Mach number, all
previously discussed. A loss coefficient K L can be defined to relate the
pressure loss p L of a duct element and the velocity pressure pVx at some
location x in that element:

pL
KL = . (2.68)
pVx

1
I. E. Idel'chik, Handbook of Hydraulic Resistance Coefficients of Local resistance and of
Friction, Distributed by National Technical Information Service, U.S. Department of
Commerce, Springfield, Va., AEC-TR-6630, 1966.

© 1999 Howden Buffalo, Inc.


2-36 FAN ENGINEERING

The value of this coefficient for the different duct elements will vary
depending upon the dimensionless parameters listed above. For instance, for
straight ducts,

pL L
KL = = f (2.69)
pV D

where f is a function of Re and ε / D . Here, it is usually more convenient to


use f and L / D directly, but for other elements that are not so dependent on
L / D , K L is preferred.

Entrance Conditions
On entering a duct system, the air accelerates from zero velocity in the
surrounding atmosphere to the duct velocity. This process involves the
transformation of pressure energy into kinetic energy, which means that, as
the velocity pressure goes up, the static pressure must go down. In addition,
there will be a loss whose magnitude will depend on the configuration at the
entrance, and the total pressure will also decrease. If the surrounding
atmospheric pressure is taken to be zero gage pressure, the velocity pressure
in the duct will be positive, the total pressure will be negative, and the static
pressure will be even more negative. This last quantity is usually referred to
as the static suction in the duct near the entrance, or static suction, or hood
suction for short.
The loss coefficient K L for an entrance condition can be determined from
the data of Figure 2.9 and Figure 2.10. These data are based on the duct
velocity pressure pV 1 , so the total pressure loss p L is

p L = K L pV 1 . (2.70)

Note that the notched entry approximates a bell-mouth; apparently, the vortex
in the notch promotes smooth flow into the duct. The curves in Figure 2.10
are based on an area ratio of 5 to 1. The peak values need not be reduced
more than 4 or 5%, even for an area ratio of 2 to 1. For area ratios in these
ranges, the loss around the outer edge of the hood is quite small compared to
that due to contraction in the pipe, so the addition of a flange will not
materially change the loss. Nevertheless, flanges can control the pattern of
airflow and thereby perform a very useful function.
The static suction pS can be used to determine the flow rate entering a
duct, by means of

p
Qû = Cv K e A S (2.71)
ρ

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-37

Figure 2.9 Entrance Conditions

Figure 2.10 Coefficients for Converging Entrances (Hoods)


Adapted from data of A. D. Brandt: "Energy Losses at Suction Hoods," Trans. ASHVE, vol.
2, 1946, pp. 205-236.

© 1999 Howden Buffalo, Inc.


2-38 FAN ENGINEERING

where Cv is 1097 for U.S. customary units1 or 2 for SI units2, Ke is the


coefficient of entry from Figure 2.9 or Figure 2.10, ρ is the air density, and
A is the area of the duct.
If the fan is located at the entrance to a duct system, the entrance loss
becomes part of the fan loss and need not be calculated as part of the system
resistance. The exception is a fan that is rated for conditions with an inlet
duct and that duct is omitted.

Example 2.5 Entrance and Exit Losses

Given a flanged entry to a 12-in. diameter duct with 1000 cfm of standard air,
find the entry loss, the static suction, the static pressure in the duct, and the
total pressure in the duct.

pV = 0.10 in. wg (from Example 2.4)

Using Figure 2.9 and Equations 2.70 and 2.71:

K L = 0.50, Ke = 0.82,
pL = K L pV = 0.50 × 010
. = 0.05 = 0.05 in. wg, and
ρQû 2
0.075 × 10002
pS = 2 =
1
Cv Ke A 6 1 1097 × 0.82 × 0.785462
= 0.15 in. wg.

The static suction is 0.15 in. wg; therefore, the static pressure in the duct is -
0.15 in. wg. The pressure loss is 0.05 in. wg; therefore, the total pressure in
the duct is 0 - 0.05 = -0.05 in. wg.

Given a 12-in. diameter duct with 1000 cfm of standard air, find the exit loss.

pV = 0.10 in. wg (from Example 2.4)

Using Equation 2.72:

pL = pV = 0.10 in. wg.

Exit Conditions
On exiting from a duct system, the air decelerates from the duct velocity
to zero velocity in the surrounding atmosphere. This process involves
dissipation of the kinetic energy of the air stream. The static pressure in the
issuing stream will be equal to the surrounding atmospheric pressure. If the
latter pressure is taken to be zero gage pressure, the total pressure at the exit
will equal the velocity pressure pV at the exit, and the exit loss p L (assuming
that the kinetic energy factor α is unity) will also equal the velocity pressure.
1 2
See Equation 2.30. See Equation 2.29.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-39

This is the same as saying that the loss coefficient K L is unity, or

p L = K L pV . (2.72)

The kinetic energy in the issuing stream can be increased by using a noz-
zle at the end of the duct. (This may be necessary to increase the distance the
stream is projected, to promote dispersion, etc.) Energy can be conserved by
using an evasé at the end of the duct. While the evasé will introduce, an addi-
tional loss of its own, it is more than compensated by the reduction in exit
loss.
Even if a fan is located at the exit from a duct system, the exit loss should
be calculated as part of the system resistance. The exception is a fan that is
rated in terms of fan static pressure. Energy can always be conserved by us-
ing an evasé when the exit velocity is not needed.

Figure 2.11 Exit Conditions

Straight Ducts
The resistance to flow through a straight duct, expressed as a pressure loss
pL , can be determined using the velocity pressure pV , the length L , and the
equivalent diameter D .
For any straight round duct,

L
pL = f pV . (2.73)
D

The proportionality factor f , known as the Darcy friction factor, was previ-
ously discussed and can be determined from Figure 2.8. As noted there, it is a
function of Reynolds number and relative roughness. Since there are so many
significant factors that influence duct resistance, it is common practice to
draw duct friction charts for an average relative roughness and to apply
roughness corrections if necessary. A roughness correction chart is drawn in
Figure 2.12. The correction, as read at the right or left, should be applied as a
multiplying factor to the values from the appropriate chart.

© 1999 Howden Buffalo, Inc.


2-40 FAN ENGINEERING

Figure 2.12 Roughness Corrections for Ducts

For any straight rectangular duct,

L L
pL = f ′ pV = 4 f ′ pV . (2.74)
M D

The proportionality factor f ′ is known as the Fanning friction factor. It is


equal to the Darcy friction factor divided by 4.
The equivalent diameter D for the same mean hydraulic radius M is
4 M . Since the mean hydraulic radius is the cross-sectional area divided by
the wetted perimeter, the equivalent diameter in terms of duct dimensions
x and y is

4 xy 2 xy
D = 4M =
1
2 x+ y 6 1
=
x+ y
.
6 (2.75)

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-41

The Darcy friction factor f can be considered to be the number of veloc-


ity pressures lost divided by the number of diameters. Its reciprocal N is the
number of diameters for a loss of one velocity pressure. After rewriting,
Equation 2.65 becomes

ç ε D + 2.51 N ä .
N = −2 log æå 3.7 Re ãâ (2.76)

Figure 2.13 is a graphical representation of Equation 2.76 for standard air


and average roughness. Corrections for other degrees of roughness can be
made by using the appropriate factor from Figure 2.12 based on the descrip-
tions in Table 2.5. Corrections can also be made for kinematic viscosity, if
different from that of standard air. Since kinematic viscosity appears in the
denominator of the Reynolds number and velocity appears in its numerator, an
equivalent velocity equal to the actual velocity multiplied by the ratio of stan-
dard kinematic viscosity to actual kinematic viscosity can be used together
with Figure 2.13.
The total pressure losses in terms of N for round and rectangular ducts
are

1 L çæ
L x+ y äã
pL =
N D
pV =
å
N 2x y
pV .
â (2.77)

Figure 2.14 can be used to determine the equivalent diameter (to be used
with constant velocity) for any rectangular dimensions. This chart gives the
round-duct diameter, which has the same mean hydraulic radius as the rectan-
gular duct. The two do not have the same cross-sectional areas. This equiva-
lent diameter can be used directly in Figure 2.13 together with the actual duct
velocity to determine the number of diameters for one velocity pressure loss.
Figure 2.15 is another type of duct friction chart based on standard air and
average roughness. The appropriate factor from Figure 2.12 can be used to
correct for other degrees of roughness. However, there is no simple way of
applying a viscosity correction. This limitation is unimportant in air condi-
tioning but can cause significant error at high temperatures or for gases other
than air. Nevertheless, Figure 2.15 has many useful features including a ca-
pacity-velocity-diameter conversion.
Figure 2.16 can be used to determine the equivalent diameter (to be used
with constant capacity) for any rectangular dimensions. This chart gives the
round-duct diameter, which has the same friction per foot of length as the
rectangular duct, but the two do not have the same cross-sectional areas. This
equivalent diameter can be used directly in Figure 2.15 together with the ac-
tual capacity, to determine the friction loss per 100 feet.
Equation 2.73 can be combined with the equations for Darcy friction factor

© 1999 Howden Buffalo, Inc.


2-42 FAN ENGINEERING

Figure 2.13
Duct Friction Chart in Diameters per Velocity Pressure
Adapted from the data of R. D. Madison and W. R. Elliot: "Friction Charts for Gases Includ-
ing Correction for Temperature, Viscosity and Pipe Roughness," ASHVE Journal Section of
Heating, Piping and Air Conditioning, October 1946, pp. 107-112.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-43

Figure 2.14
Equivalent Diameters for Use With Constant Velocity

© 1999 Howden Buffalo, Inc.


2-44 FAN ENGINEERING

Figure 2.15
Duct Friction Chart in Inches per 100 Feet
Adapted from the data of D. K. Wright, Jr.: "A New Friction Chart for Round Ducts:" Trans
ASHVE, vol. 51,1945, pp. 303-316.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-45

Figure 2.16
Equivalent Diameters for Use With Constant Capacity

© 1999 Howden Buffalo, Inc.


2-46 FAN ENGINEERING

to produce pressure-drop expressions for each zone of the Moody chart. The
resulting expressions in the wholly rough and transition zones are rather com-
plex. A compromise expression for turbulent flow is given here. An exact
expression for laminar flow is given later.
For turbulent flow,

cLV 2 .0− n ρ 1.0− n µ n


pL = . (2.78)
gc D1.4 − n

Any set of units can be used with this expression provided that they are di-
mensionally consistent. The coefficient c and the exponent n can be deter-
mined from Table 2.5 opposite the appropriate condition of roughness. The
effect of changes in the exponent n more than offsets the effect of changes in
coefficient c so that pressure drop pL increases with roughness. The rougher
the pipe or duct, the more nearly the pressure drop p L varies as the square of
the velocity V , the first power of the density ρ , the 1.4 power of the diameter
D , and the zero power of the viscosity µ . It is common practice to deter-
mine pressure drops from charts and tables drawn up for standard conditions,
to correct for density on the basis of a first power relationship, and to ignore
the effect of viscosity. A somewhat more accurate approach is to use the re-
lationships indicated in Equation 2.78. The most accurate results can be ob-
tained by using equivalent velocity in Figure 2.13 as outlined in the discussion
of that chart.
For laminar flow,

32 LVµ
pL = . (2.79)
gc D 2

This expression is correct when used with any set of units that are dimension-
ally consistent. The difference between turbulent and laminar flow, as regards
losses, is easily ascertained by comparing Equations 2.78 and 2.79. In lami-
nar flow, the loss pL is independent of fluid density ρ . The actual value of
roughness has no effect on loss. As in any kind of flow, the loss is propor-
tionate to length L , but the relationships with velocity V and diameter D for
laminar flow are different from those for turbulent flow.

Example 2.6 Round-Duct Loss

Given 100 ft of 12-in. round galvanized duct, 1000 cfm of air at 0.075 lbm/ft3
density and 1.22 × 10-5 lbm/ft-s viscosity, find the pressure loss.

Qû 1000
V= = = 1273 fpm = 21.22 fps and
A 0.7854

pV = 0.10 in. wg from Example 2.4.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-47

Using Figure 2.13 and Equation 2.77:

1 L çæ äã çæ äã
1 100
N = 51 and p L =
N D
pV =
51å âå â
1
01
. = 0.20 in. wg.

Alternatively, using Figure 2.15:

L 100
p L ′ = 0.2 in. wg/100 ft and p L = p L ′ = 0.20 = 0.20 in. wg.
100 100

Alternatively, using Table 2.5 and Equation 2.78:

n = 0. 16, c = 0.0746,

pL =
cLV 2 .0− n ρ 1.0− n µ n 0.0746 × 100 × 2122
=
2 . × 10−5
. 1.84 × 0.0750.84 × 122 7 0.16

,
gc D1.4 − n 32.17 × 10. 1.24

p L = 1.19 lb/ft2 = 0.23 in. wg.

Refer to Example 2.4 for use of the Moody Chart and the Colebrook Equa-
tion.

Example 2.7 Rectangular-Duct Loss

Given 100 ft of 10-in. × 15-in. rectangular galvanized duct and 1326 cfm of
standard air, find the pressure loss.

Qû 1326
V= = = 1273 fpm and pV = 0.10 in. wg (from Example 2.4).
A 104
.

Using Figure 2.14, Figure 2.13, and Equation 2.77:

D = 12 in. (for use with constant velocity),

2 xy 2 × 10 × 15
(Also, D =
1 x+ y6 1
=
10 + 156 = 12 in. from Equation 2.75.)

1 L 1 100çæ äã
N = 51, and p L =
N D
pV =
51 1 å â
. = 0.20 in. wg.
01

Alternatively, using Figure 2.14 and Figure 2.15:

© 1999 Howden Buffalo, Inc.


2-48 FAN ENGINEERING

D = 12 in. (for use with constant velocity),

p L ′ = 0.20 in. wg/100 ft based on 1273 fpm and 12 in., and

L 100
pL = pL ′ = 0.20 = 0.20 in. wg.
100 100

Alternatively, using Figure 2.16 and Figure 2.15:

D = 13.2 in. (for use with constant capacity),

p L ′ = 0.20 in. wg/100 ft based on 1326 cfm and 13.2 in., and

L 100
pL = pL ′ = 0.20 = 0.20 in. wg.
100 100

All methods should, and do, give the same results.

Elbows
Elbows, bends, and miters are used to guide a fluid through a change in
direction of flow. Both shock losses and friction losses are caused by such
devices. The relative amounts of each, as well as the total resistance, will de-
pend on the abruptness of the change in direction, the Reynolds number, and
the roughness.
Elbow losses can be expressed in terms of a loss coefficient, an equivalent
length of straight duct, or an extra equivalent length of straight duct. Using
the loss coefficient K L , the pressure loss p L is obtained from

p L = K L pV (2.80)

where the velocity pressure pV is the average at any section, provided that
there is no change of section throughout the elbow. The size of the elbow will
have very little effect compared with certain other geometrical considerations.
The most significant factors are: the shape of the elbow, that is, whether
round, square, or rectangular; the aspect ratio, if rectangular; the angle of
bend; and the radius ratio or curve ratio.
The curve ratio CR of an elbow is its inside radius Ra divided by its out-
side radius Rb , assuming concentricity as indicated in Figure 2.17.

Ra
CR = (2.81)
Rb

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-49

Figure 2.17 Hard and Easy Bends

The radius ratio RR of an elbow is its centerline radius R divided by its


width W in the plane of the bend, also assuming concentricity as indicated in
Figure 2.17.

R
RR = . (2.82)
W

The aspect ratio AR of an elbow is its depth D along the axis of the bend
divided by the width W in the plane of the bend, as indicated by

D
AR = . (2.83)
W

For square elbows the depth equals the width W , while for round elbows
both depth and width equal the diameter D . For both elbows, AR = 1.0, and
CR and RR are related according to:

RR − 0.5
CR = and (2.84)
RR + 0.5

çæ 1 + CR äã .
RR = 0.55
å 1 − CR â (2.85)

Figure 2.18 can be used to predict the losses for round miters and for 90º
round elbows and bends of various curve or radius ratios. This figure is based
on averages similar to those of Locklin, and includes both shock and friction
losses.
The amount of the loss is not exactly proportional to the angle of bend, as
indicated by Figure 2.19. Determine the factor based on the appropriate

© 1999 Howden Buffalo, Inc.


2-50 FAN ENGINEERING

Figure 2.18 Loss Coefficients for 90º Round Elbows and Miters
Adapted from the data of D. W. Locklin: "Energy Losses in 90 Degree Duct Elbows:" Trans.
ASHVE, vol. 56, 1950, pp. 479-502.

description and angle of bend. Obtain the loss for the corresponding 90º el-
bow from Figure 2.18. Then multiply by this angle-of-bend factor.
Figure 2.20 can be used to determine the losses for square miters and for
90º square elbows of various curve or radius ratios. It is based on tests con-
ducted in the Buffalo Forge Company laboratory. (There is notable agree-
ment between this data and the average data of several investigators, as com-
piled by Locklin.1) From this figure, it is apparent that size effect is negligible
except at high-curve ratios where shock losses decrease to the point of insig-
nificance compared with friction losses.
1
D. W. Locklin, "Energy Losses in 90 Degree Duct Elbows," Trans. ASHVE, vol. 56, 1950,
pp. 479-502.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-51

Figure 2.19 Angle of Bend Factors for Elbows


Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows," Trans. ASME, vol. 58, 1936, pp. 167-176.

© 1999 Howden Buffalo, Inc.


2-52 FAN ENGINEERING

Figure 2.20 Loss Coefficients for 90º Square Elbows and Miters
Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows": Trans. ASME, vol. 11, 1916, pp. 161-116

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-53

Figure 2.21 Aspect Ratio Factors for Rectangular Elbows


Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows," Trans. ASME, vol. 58, 1936, pp. 167-176.

Figure 2.21 gives aspect ratio factors for rectangular elbows. Determine
the factor based on the appropriate description and aspect ratio. Obtain the
loss for the corresponding square elbow from Figure 2.20. Then multiply by
the aspect ratio factor. Note that there is a broad range of aspect ratios for
which no correction is necessary.
The loss in an elbow located at the end of a duct is much higher than that
for a similar elbow followed by a short run of straight pipe. This is similar to
the phenomenon observed with orifices in that there is a contraction of the
stream, and unless expansion is allowed to take place before exiting, a consid-
erable amount of kinetic energy is wasted. The curves of Figure 2.22 illus-
trate the effect of both aspect ratio and curve or radius ratio on the pressure
loss in a 90º elbow discharging directly into the atmosphere.
The combined loss of a compound elbow may be quite different from the
sum of the individual losses, as indicated by Figures 2.23 and 2.24. The col-
umns marked "actual" are the loss as measured. The columns marked "esti-
mated sum" are the sums of the individual loss coefficients, estimating the
loss coefficient for the upstream elbow as if it were followed by a duct and
also estimating the loss coefficient for the downstream elbow according to
conditions as marked. The ratio of actual to estimated loss is given as "% of
est." The estimated combined loss coefficient with splitters is usually quite
close to its actual loss.

© 1999 Howden Buffalo, Inc.


2-54 FAN ENGINEERING

Figure 2.22 Loss Coefficients for 90º Elbows and Miters


Discharging to Atmosphere
Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows," Trans. ASME, vol. 58, 1936, pp. 167-176.

Conventional elbows are those having concentric inner and outer radii and
constant areas of cross section throughout the bend. From time to time,
various special elbows have been proposed incorporating either a change in
area or a change in the shape of that area throughout the bend. For
comparable curve and aspect ratios, conventional elbows are superior to
special elbows. For the lowest possible loss, conventional elbows with
splitters or miters with turning vanes should be used.
Splitters are curved vanes placed in an elbow concentric with both the
inside and outside radii and extending the full angle of bend from face to face.
Splitters, in effect, divide the flow into parallel channels, each having a larger
curve ratio and a larger aspect ratio than that of the original elbow. Usually,
the change in aspect ratio has much less effect upon elbow loss than the
change in curve ratio. Disregarding the change in aspect ratio then, the ideal
locations for splitters are those that divide the elbow into components, each
with the same curve ratio. For any number of splitters n , the new curve ratio
CR′ of each component elbow formed by the splitters can be determined
using the curve ratio CR of the original elbow without splitters in

1 61
CR ′ = CR
6.
1/ n +1
(2.86)

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-55

Figure 2.23 Actual and Estimated Losses for Compound Elbows


(3" x 12")

Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows," Trans. ASME, vol. 58, 1936, pp. 167-176.

© 1999 Howden Buffalo, Inc.


2-56 FAN ENGINEERING

Figure 2.24 Actual and Estimated Losses for Compound Elbows


(12" x 12")

Adapted from the data of R. D. Madison and J. R. Parker: "Pressure Losses in Rectangular
Elbows," Trans. ASME, vol. 58, 1936, pp. 167-176.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-57

Figure 2.25
Locations of Splitters in Rectangular Elbows

© 1999 Howden Buffalo, Inc.


2-58 FAN ENGINEERING

Figure 2.25 gives the radius of each splitter at the intersection of a straight
line drawn between points that represent the inside and outside radii of the
elbow without splitters. The example drawn on the chart shows that, for an
inside radius of 2 in. and for an outside radius of 20 in., the radius of the inner
splitter should be approximately 4.25 in., and the radius of the outer splitter
should be about 9.25 in.
The loss due to a 90º square-section elbow with one, two, or three splitters
can be determined from the curves in Figure 2.20. Splitters produce apprecia-
ble reduction in pressure loss when the original curve ratio is low, but there is
no material reduction when the original curve ratio is high. Although this data
is strictly applicable only to square-section elbows, corrections for aspect ratio
can be made for rectangular-section elbows.
Turning vanes can be used to reduce the loss through a miter of either
round or rectangular section. As indicated in Figure 2.26, the loss depends
upon the type of miter and the type of turning vane. The estimated values
shown are based on a vane depth of approximately six times the spacing and,
where thickened vanes are indicated, a value of 0.5 curve ratio between the
back of one and the front of the next. The very considerable reduction in loss
coefficient from a value of 1.15 (see Figures 2.18 and 2.20) to the approxi-
mate values indicated in Figure 2.26 results from the conversion of the miter
into a series of high-aspect-ratio (or easy-bend) elbows.

Figure 2.26 Loss Coefficients for 90º Μiters


Μ with Turning Vanes

Adapted from the data of L. Wirt: "New Data for the Design of Elbows in Duct Systems,"
General Electric Review, June 1927, pp. 286-296.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-59

As a rule of thumb in the region of 0.5 curve ratio, the equivalent length of
straight duct for a 90º round elbow is about 9 to 10 diameters. Similarly, the
equivalent length of a 90º rectangular elbow is about 6 to 8 equivalent
diameters.
When estimating duct friction, it is often convenient to measure straight
lengths as if all elbows were miters. To compensate, elbow losses are figured
on the extra equivalent length basis. Values were formerly reported on this
basis in the Guide.1

Example 2.8 Round-Elbow Loss

Given a 4-piece 90ƒ round elbow with a 12-in. diameter, an 18-in. centerline
radius, and 1000 cfm of standard air, find the loss.

pV = 0.10 in. wg (from Example 2.4)

Using Equation 2.82, Figure 2.18, and Equation 2.80:


R 18
RR = = = 15
.
W 12
K L = 0.33, and
p L = K L pV = 0.33 × 010
. = 0.033 in. wg.

Example 2.9 Rectangular-Elbow Loss

Given a rectangular elbow with a 12-in. width, a 15-in. depth, a 6-in. inside
radius, and 1326 cfm of standard air, find the loss with and without splitters.

pV = 0.10 in. wg (from Example 2.7)

Using Equation 2.81, Figure 2.20, and Equation 2.80:


R 6
CR = a = = 0.33,
Rb 18
K L = 0.22 for no splitters, and
p L = K L pV = 0.22 × 010
. = 0.02 in. wg ;

K L = 0.10 for 2 splitters, and


p L = K L pV = 010
. × 0.10 = 0.01 in. wg.

Using Equation 2.83 and Figure 2.21:


D 15
AR = = = 125
. , and
W 12
aspect ratio factor is 1.01 from curve A and can be ignored.
1
ASHRAE Guide and Data Book - Equipment, ASHRAE, New York, 1969, p. 32.

© 1999 Howden Buffalo, Inc.


2-60 FAN ENGINEERING

Changes in Duct Section


Various fittings can be used to connect two ducts that are of different sizes
or shapes.
If the downstream duct is of the same shape but smaller than the upstream
duct, a converging cone or a converging pyramid (with total included angles
of convergence less than 15º) can be used. The total pressure loss pL , which
will be a small fraction of the leaving velocity pressure pV 2 , can be estimated
by using the mean values of D and f in the equation for straight ducts:

L
pL = f pV 2 . (2.87)
D

A considerably larger loss would result if an abrupt contraction were used.


This will be discussed in a subsequent section
If the downstream duct is of similar shape but larger than the upstream
duct, a diverging cone or a diverging pyramid can be used. Conical and plane
diffusers of this kind are discussed in a later section. An abrupt enlargement
can also be used. This, too, is discussed below.
If the downstream duct is not of the same shape as the upstream duct, a
transformation piece can be used. But, even if the areas of the two ducts are
the same, the transformation will have elements with different slopes. If all
the elements are converging, the total pressure loss can be estimated as out-
lined above for converging cones or pyramids. If all the elements are diverg-
ing, the total pressure loss can be estimated from the data for conical and
plane diffusers. However, should some of the elements converge and others
diverge, the total pressure loss can be determined as if the transformation
piece were a straight duct, but only if the slopes of the diverging elements are
very small. Otherwise, there will be some additional diffuser loss.
If the downstream duct is not coaxial with the upstream duct, the data on
bends and miters must be used to calculate total pressure loss.

Abrupt Enlargement
In an abrupt enlargement, the fluid flows into a conduit without contrac-
tion but at less than full bore and then expands to full bore. Not every sudden
enlargement of duct section fulfills these conditions. The conduit must be
sufficiently long (at least 3 or 4 diameters) for the necessary expansion to take
place. On the other hand, if a fluid does enter a conduit with subsequent con-
traction and re-expansion, the condition between the plane of the vena con-
tracta and the plane where the expansion becomes complete can be consid-
ered an abrupt enlargement
The total pressure loss pL for an abrupt enlargement, or Borda-Carnot
loss, is a function of the velocity V1 in the upstream duct and the velocity V2
in the downstream duct, as indicated by

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-61

Figure 2.27 K L and K R for Abrupt Enlargements

ç V − V äã
= ρæ
2

å C â
1 2
pL (2.88)
v

where ρ is the gas density and Cv is 2 for SI units or 1097 for U.S. cus-
tomary units. This can also be expressed in terms of the velocity pressure pV 1
in the upstream duct and the area ratio A1 A2 :

ç Aä
= æ1 − ã
2

å Aâ pV 1 = K L pV 1. .
1
pL (2.89)
2

Obviously, the bracketed term can be considered a loss coefficient K L . Val-


ues of K L versus A1 A2 are plotted in Figure 2.27.
Another important quantity for an abrupt, or even a gradual, enlargement
is the change in static pressure. Static pressure will increase in the direction
of flow and is, therefore, called static pressure regain pSR , or regain for short.
The regain is the difference between the change in velocity pressure and the
total pressure loss, and can also be expressed in terms of pV 1 and A1 A2 :

çæ A äã çæ1 − A äã p
pSR = 2
å A âå A â = K R pV 1 .
1 1
V1 (2.90)
2 2

© 1999 Howden Buffalo, Inc.


2-62 FAN ENGINEERING

The coefficient of pressure recovery K R is not equal to one minus the total
pressure loss coefficient K L as is often, but erroneously, stated. Values of
K R are plotted in Figure 2.27.
The loss and regain can be expressed as functions of the change in veloc-
ity pressure ∆ pV by using an effectiveness factor µ . This is done in the sec-
tion on diffusers.
The ideal case considered above assumes that duct friction is negligible,
that the velocity distribution is uniform in the upstream duct, and that flow is
completely turbulent. Corrections for low Reynolds number, nonuniform ki-
netic energy, and nonuniform momentum are given by ldel'chik1 for various
configurations involving abrupt enlargement.

Example 2.10 Abrupt Enlargement Loss and Regain

Given an abrupt enlargement from an 8-in. duct to a 12-in. duct and 1000 cfm
of standard air, find the loss and the regain.

A1 0.3491 Qû 1000
= = 0.4445 and V1 = = = 2865 fpm
A2 0.7854 A1 0.3491

Using Equation 2.30:

ç V äã
= ρæ
2
ç 2865äã
= 0.075æ
2

å 1097 â å 1097 â = 0.51 in. wg.


1
pV 1

Using Figure 2.27 and Equations 2.89 and 2.90:

K L = 0.309,
K R = 0.494,

p L = K L pV 1 = 0.309 × 0.51 = 016


. in. wg, and

pSR = K R pV 1 = 0.494 × 0.51 = 0.25 in. wg.

Abrupt Contraction
In an abrupt contraction, the fluid flows into a conduit at full bore but im-
mediately contracts and then re-expands to full bore. Not every sudden re-
duction of duct section fulfills these conditions. The conduit must be suffi-
ciently long for the necessary contraction and re-expansion to take place. The
conditions up to the vena contracta are the same as for a square-edged orifice
with an opening equal to the cross section of the conduit. The conditions be-
yond the vena contracta are the same as for an abrupt enlargement.
1
Idel'chik, pp. 128-132.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-63

Figure 2.28 K L for Abrupt Contractions

The total pressure loss pL of an abrupt contraction is a function of the


shape of that contraction, the upstream area A1 the downstream area A2 and
the downstream velocity pressure pV 2 , as indicated by

çæ
pL = K L ′ 1 −
A1äã
å A2 â
pV 2 = K L pV 2 . (2.91)

In the square-edged case, K L ′ is approximately 0.5. Values of the loss coef-


ficient K L versus A2 A1 , are plotted in Figure 2.28 for high Reynolds num-
ber and K L ′ = 0.5 .

Example 2.11 Abrupt Contraction Loss

Given an abrupt contraction from a 12-in. duct to an 8-in. duct and 1000 cfm
of standard air, find the loss.

pV 2 = 0.51 in. wg (from Example 2.8)


Using Figure 2.28 and Equation 2.91:
K L = 0.278 and
p L = K L pV 2 = 0.278 × 0.51 = 014
. in.wg.

Idel'chik1 gives data for calculating K L for rounded and beveled edges, as
well as for low Reynolds numbers, in the square-edged case.
1
Idel'chik, pp. 98-99.

© 1999 Howden Buffalo, Inc.


2-64 FAN ENGINEERING

Diffusers and Evasés (Diverging Tapers)


A diffuser is a flow passage in which kinetic energy is converted into
pressure energy. For subsonic flow, the passage must diverge in the direction
of flow. A diffuser is distinguished from an abrupt enlargement by gradual
divergence. When a diffuser is located at the exit end of a duct, it is known as
an evasé. A diffuser on the outlet of a fan sometimes is called an evasé even
if there is additional ductwork. And frequently the fan is said to be coned
even if the cross sections of the diffuser are not circular. A conical diffuser,
however, does have circular cross sections. Two-dimensional, or plane, dif-
fusers have rectangular cross sections and two parallel sides. Any other di-
verging passage with rectangular sections is called a three-dimensional dif-
fuser.
The performance of a diffuser, like that of an abrupt enlargement, will
vary with inlet velocity pressure pV 1 and the ratio of outlet to inlet area
A2 / A1 . In addition, performance will vary with the centerline length L and
the total included angle of divergence 2 θ .
In an ideal diffuser, the regain pSR would equal the change in velocity
pressure ∆pV , and the loss in total pressure pl would be zero. However, in a
real diffuser, there will always be a loss in total pressure, and the effectiveness
η , which is equal to pSR / ∆pV , will be less than unity. Regain is frequently
given in percent of inlet velocity pressure. The dimensionless fraction
pSR / pV 1 is also known as the coefficient of pressure recovery K R .
The flow in a diffuser may follow the walls of the passage, or it may stall
and be deflected away from the walls by reverse flow. Figure 2.29, which is
drawn for plane diffusers with straight walls, shows the various flow regimes
that might be encountered. It is generally applicable for entrance Reynolds
numbers greater than 5 × 104 and for Mach numbers less than 0.2. The coor-
dinates are area ratio A2 / A1 , and the ratio of length to inlet width L / W1 , but
the angle of divergence is also shown. The flow in each regime is described
in detail in the reference cited; only brief explanations follow here.
In jet flow, separation from both walls begins near the throat and covers
both walls. In two-dimensional stall, separation begins near the throat and
covers only one wall. This fixed stall will remain there unless it is diverted to
the other wall by a large disturbance at the inlet or outlet. There is a region
between the two zones where either jet flow or two-dimensional stall can ex-
ist. Transitory stall varies with diffuser geometry but generally begins in the
corners, builds up, and is swept away repeatedly. Note that the lines of
maximum pressure recovery and maximum effectiveness fall into the region
of "some stall."
Figures 2.30 and 2.31 are also drawn for plane diffusers. Figure 2.30
shows the coefficients of pressure recovery and Figure 2.31 the effectiveness
values, for various geometrys. However, these charts are drawn for only one
.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-65

Figure 2.29
Flow Regimes in Plane Diffusers
Adapted from the data of L. R. Reneau, J. P. Johnston and S. J. Kline: "Performance and
Design of Straight Two-Dimensional Diffusers," ASME Paper No. 66-FE-10, 1966.

© 1999 Howden Buffalo, Inc.


2-66 FAN ENGINEERING

Figure 2.30
Coefficients of Pressure Recovery for Plane Diffusers
Adapted from the data of L. R. Reneau, J. P. Johnston and S. J. Kline: "Performance and
Design of Straight Two-Dimensional Diffusers," ASME Paper No. 66-FE-10, 1966.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-67

Figure 2.31
Effectiveness Values for Plane Diffusers
Adapted from the data of L. R. Reneau, J. P. Johnston and S. J. Kline: "Performance and
Design of Straight Two-Dimensional Diffusers," ASME Paper No. 66-FE-10, 1966.

© 1999 Howden Buffalo, Inc.


2-68 FAN ENGINEERING

Figure 2.32
Coefficients of Pressure Recovery for Conical Diffusers

Adapted from the data of A. T. McDonald and R. W. Fox: "An Experimental Investigation of
Incompressible Flow in Conical Diffusers," ASME Paper No. 65-FE-25, 1965.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-69

Figure 2.33
Effectiveness Values for Conical Diffusers

Adapted from the data of A. T. McDonald and R. W. Fox: "An Experimental Investigation of
Incompressible Flow in Conical Diffusers," ASME Paper No. 65-FE-25, 1965.

© 1999 Howden Buffalo, Inc.


2-70 FAN ENGINEERING

inlet boundary layer condition: a turbulent inlet boundary layer such as might
be generated if the diffuser were preceded by several diameters of straight
duct and if flow were uniform. Chart values should be reduced as much as
25% when inlet flow is nonuniform or boundary layers are thicker. Plots for
other conditions are given in the reference.
To determine static pressure regain, use either

pSR = K R pV 1 or (2.92)

pSR = η∆pV . (2.93)

To determine total pressure loss, use

1 6
p L = 1 − η ∆pV . (2.94)

Peak pressure recovery for a given length or a given area ratio can be
determined from Figure 2.30. A line connecting all such points is drawn on
the chart. It is also drawn on Figure 2.29 to illustrate that peak recovery
occurs in the zone of appreciable stall. On the other hand, maximum
effectiveness generally occurs in the "some stall" zone. Figure 2.31 shows
that maximum effectiveness at constant area ratio will be produced with a
plane diffuser angle of approximately 7ƒ1
Figures 2.32 and 2.33 are drawn for conical diffusers using area ratio
A2 / A1 and length to inlet radius ratio L / R1 as coordinates. These charts can
be used in the same way as Figures 2.30 and 2.31 to determine performance
for any geometry or to optimize performance. These conical diffuser data are
for free-jet exit conditions so that some improvement could be expected, at
least for large angles of divergence, if a discharge duct were used. In contrast
to the situation for plane diffusers, conical diffusers attain peak recovery for a
given length or area ratio before the onset of stall. This probably is due to the
absence of side-wall corners.
The performance of any diffuser can be improved by fairing the entrance
corners. There is some evidence1 that the presence of a resistance at the exit
of a diffuser allows the use of greater angles of divergence for a given total
pressure loss and may even prevent the onset of stall. Wide-angle diffusers
can be improved by using splitter vanes.2 Diffusers with curved centerlines
have lower performance than straight-walled diffusers.3
1
C. H. McLellan and M. R. Nichols, "An Investigation of Diffuser-Resistance Combinations
in Duct System," NACA Wartime Report L-329, 1942.
2
O. G. Fail, "Vane Systems for Very-Wide-Angle Subsonic Diffusers," ASME Paper No. 64-
FE-4, 1964.
3
C. J. Sagi and J. P. Johnston, "The Design and Performance of Two-Dimensional Curved
Diffusers," ASME Paper, No. 67-FE-6, 1967.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-71

According to Fox,1 the coefficients of pressure recovery for conical dif-


fusers with low angles of divergence increase as inlet Mach number is in-
creased to about 0.9. Effectiveness appears to decrease rapidly as the Mach
number is increased to about 0.2 to 0.3. However, effectiveness remains rela-
tively constant over the 0.3 to 0.9 Mach-number range. At Mach numbers
above 0.9, both effectiveness and coefficient of pressure recovery decline
rapidly.
ldel'chik2 gives data for calculating coefficient of loss K L for conical, py-
ramidal, plane, transformation, curved, stepped, and annular diffusers and for
the effects of baffles and screens.

Example 2.12 Diffuser Loss and Regain

Given a 36-in. long conical diffuser, an 8-in. to 12-in. diameter, and 1000 cfm
of standard air, find the loss and the regain.

A2 0.7854
= = 2.25,
A1 0.3491

pV 1 = 0.51 in. wg (from Example 2.8),

pV 2 = 0.10 in. wg (from Example 2.4), and

L 36
= = 9.
R1 4

Using Figures 2.32 and 2.33 and Equations 2.92, 2.93, and 2.94:

K R = 0.64,

η = 0.80,

pSR = K R pV 1 = 0.64 × 0.51 = 0.33 in. wg,

1 6 1
pSR = η pV 1 − pV 2 = 0.80 0.51 − 010 6
. = 0.33 in. wg, and

1 61 6 1
p L = 1 − η pV 1 − pV 2 = 0.20 0.51 − 010 6
. = 0.08 in. wg.

Note that this loss is considerably less than that for an abrupt enlargement as
illustrated in Example 2.10.
1
R. W. Fox, "Subsonic Flow in Conical Diffusers," Technical Report FMTR-67-1, Purdue
Research Foundation, Lafayette, Ind., 1967.
2
Idel'chik, pp. 166-188.

© 1999 Howden Buffalo, Inc.


2-72 FAN ENGINEERING

Takeoffs and Junctions


The total pressure losses of takeoffs and junctions are affected by the ratio
of downstream to upstream velocity, by the branching angle, and by other
geometrical relations. Figure 2.34 shows the average loss coefficient for a
divided-flow fitting of round cross section for both the run of the main K L1− 2
and the takeoff K L1− 3 . The total pressure loss p L1− 2 for the run of the main is

pL1− 2 = K L1− 2 pV 2 . (2.95)

The total pressure loss p L1− 3 for the takeoff is

p L1− 3 = K L1− 3 pV 3 . (2.96)

Three different branching angles are shown for the takeoff. Its loss can be
reduced considerably by using a converging taper between the main and the
branch pipe. The dotted curve shows the approximate effect of a generous
taper, one side of which makes a branching angle of approximately 45ƒ.
The above data apply only to takeoffs where one stream is divided into
two. Similar fittings are used at junctions where two streams are combined
into one. The recommended procedure for such a junction is to join the up-
stream main with the downstream main by means of a taper at least two up-
stream diameters long and to join the branch to the taper at an angle of 30ƒ
with the upstream main. Figure 2.35 gives the coefficients of loss for both the
upstream main to downstream main K L1− 3 and the branch to downstream
main K L2 − 3 when the sum of the areas of the upstream main and the branch
equals the downstream main area. Note that the loss can be negative depend-
ing on the ratio of the branch flow to the total flow Qû 2 / Qû 3 . This is due to the
transfer of momentum from one stream to the other.
The total pressure loss p L1− 3 for the upstream main to the downstream
main is

p L1− 3 = K L1− 3 pV 3 . (2.97)

The total pressure loss p L2 − 3 for the branch to the downstream main is

p L 2 − 3 = K L 2 − 3 pV 2 . (2.98)

Idel'chik1 presents data on calculating the coefficient of loss for numerous


takeoffs and junctions, including wyes, double wyes, tees, crosses, and head-
ers.
1
Idel'chik, pp. 260-304.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-73

Figure 2.34 Pressure Losses for Divided Flow Fittings


Adapted from the data of S. E Gilman: "Pressure Losses of Divided-Flow Fittings," Trans.
ASHAE, vol. 61,1955, pp. 281-296.

Example 2.13 Takeoff Losses

Given a divided-flow fitting with a 12-in. diameter main, an 8-in. diameter


60ƒ takeoff, and 1000 cfm of standard air divided equally, find the losses.

A1 = A2 = 0.7854 ft2,
A3 = 0.3491 ft2,
V1 = 1000 / 0.7854 = 1273 fpm,
V2 = 500 / 0.7854 = 500/0.7854 = 637 fpm,
V3 = 500 / 03491 = 1432 fpm,

© 1999 Howden Buffalo, Inc.


2-74 FAN ENGINEERING

V2 637
= = 0.05,
V1 1273
K L1− 2 = 150/100= 1.5 from Figure 2.34,
V3 1432
= = 112
. ,
V1 1273
K L1− 3 = 70/100= 0.7 from Figure 2.34.

Using Equations 2.30, 2.95, and 2.96,


çæ äã
V2
2
637 çæ
2
äã
pV 2 = ρ
å â
1097
=.075
1097 å= 0.03,
â
ç V äã
= ρæ
2
çæ 1432 äã 2

å 1097 â = 0.075
å 1097 â = 0.13,
3
pV 3

p L1− 22 = K L1− 2 pV = 15. × 0.03 = 0.05 in. wg for the run of the main, and
p L1− 3 = K L1− 3 pV 3 = 0.7 × 013
. = 0.09 in. wg for the takeoff.

Figure 2.35 Pressure Losses for Junctions

Adapted from the data of I. E. Idel'chik, Handbook of Hydraulic, Resistance - Coefficients of


Local Resistance and of Friction, Distributed by National Technical Information Service,
U.S. Department of Commerce, Springfield, Va., AEC-TR-6630, 1966, p. 268.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-75

Obstructions
Various objects can partially obstruct the flow through a duct. Idel'chik1
gives data for many isolated obstructions like pipes, beams, and trusses; for
orifice plates; and for many distributed resistances like grids, screens, packed
beds, and tube bundles. Only a few will be examined here.
A simplified formula for the loss pL due to an isolated obstruction is

A3
A1
p L = CD
çæ1 − τ A äã 3
k1k 2 pV 1 (2.99)

å Aâ
3

where A3 is the projected area of the obstruction, A1 is the area of the duct,
CD is the drag coefficient for the obstruction (obtained from data like that in
Table 2.6), and pV 1 is the velocity pressure based on the free area of the duct.
Idel'chik includes means for determining the effects of velocity distribution
k1 and offset k2 . The value of k1 is approximately 1.15 for struts running
completely across the duct and 1.30 for objects with three-dimensional flow
around them. The value of k2 is unity for zero offset, and it decreases as the
object nears the wall (in the extreme, 0.3 to 0.6). The value of τ is unity for
smooth objects and approximately 1.5 for objects with sharp edges.
The general formula for the loss p L due to an orifice plate is

á ç Aä ç Aä
= k æ1 − ã + æ1 − ã
2
A0 çæ A äã
L "#
! å Aâ å Aâ å â
+ τ 1− 1 − 0 + f 0 pV 0 (2.100)
#$
0 0
pL
1 2 A1 A2 D0

where A1 is the area of the upstream duct, A2 is the area of the downstream
duct, f is the Darcy friction factor, and pV 0 is the velocity pressure based on
the area A0 of the orifice. The factor k depends on the shape of the inlet
edge of the orifice, ranging from nearly zero for a well-rounded entry to unity
for a re-entrant condition. The factor τ depends on both the shape of the inlet
edge and the ratio of the length L0 of the orifice (thickness of the orifice
plate) to its hydraulic diameter D0 , ranging from nearly zero for long pas-
sages to 1.4, or so, for thin orifice plates. This equation is reducible to that for
an abrupt contraction when A2 = A0 and to that for an abrupt enlargement
when A1 = A0 .
1
Idel'chik, pp. 388-399.

© 1999 Howden Buffalo, Inc.


2-76 FAN ENGINEERING

Table 2.6 Drag Coefficients

A perforated plate can be considered a parallel arrangement of orifices.


For a thin plate with square-edged openings, k = 0.5 , f L0 / D0 is negligible,
and τ = 1414
. . For a uniform duct, A2 = A1 and Equation 2.100 is reducible
to

á A çæ A äã "# çæ A äã
2 2

= 0.707
å
1− 0 + 1− 0
â$ å A â
1

!
pL pV 1 (2.101)
A1 A1 0

where A0 is the combined area of the openings, A1 is the duct area, and pV 1
is the velocity pressure based on the free area of the duct.
A woven wire screen can also be considered a parallel arrangement of ori-
fices. For round wires, τ = 2 k 1/ 2 and k = 13. . For a uniform duct, A2 = A1
and Equation 2.100 is reducible to

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-77

  A0   A1  
2

pL = 1.30 1 −  +  − 1  pV 1 (2.102)
  A1   A0  

where symbols are the same as for Equation 2.101. The equations given
above for plates and screens are for relatively high Reynolds numbers and low
Mach numbers. For low Reynolds number and high Mach number correc-
tions, consult Idel'chik.1

Example 2.14 Obstruction Losses

Given a 3-in. I-beam running through a 12-in. round duct with 1000 cfm of
standard air, find the obstruction loss when the web is parallel to the flow and
the beam is centered in the duct.

2.33 × 12
A3 = = 0.1942 ft2,
144
A3 01942.
= = 0.2472,
A1 0.7854
k1 = 1.15 for struts running completely across
k 2 = 1.0 for zero offset,
τ = 1.5 for sharp objects, and
pV 1 = 0.10 (from Example 2.4).

Using Table 2.6 and Equation 2.99:

CD = 1.2 for I-beams,


A3
A1
p L = CD
çæ A3 äã
k k p , and
3 1 2 V1

å
1− τ
A1 â
0.2472
p L = 12 × 115
. × 10
. × 010
. = 0.14 in. wg.
2 1 67
. 3
1 − 15
. 0.2472

Given a woven wire screen with 40% open area in a 12-in. round duct with
1000 cfm of standard air, find the screen loss.

A0
= 0.4 and
A1
pV 1 = 0.10 (from Example 2.4)
1
Idel'chik, pp. 321-349.

© 1999 Howden Buffalo, Inc.


2-78 FAN ENGINEERING

Using Equation 2.102:

á . ç1 − A ä + ç A − 1ä "# p and
2

æ ã æ ã
! å A â å A â #$
pL = 130 0 1
V1
1 0

. 11 − 0.46 + 12.5 − 16 9010


= 4130 . = 0.3 in. wg.
2
pL

Compressible Flow
The flow of any gas is compressible, but as noted previously, the flow
through a fan system can almost always be considered incompressible. The
error in ignoring compressibility effects in the computation of pressure varia-
tions will be less than one percent for Mach numbers up to about 0.2 (based
on Equation 2.53, using this value to evaluate the bracketed terms). Satisfac-
tory calculations can be made at even higher subsonic Mach numbers if a
mean density is used.
The coefficients of friction and loss given in the preceding sections proba-
bly can be used for Mach numbers even higher than 0.2. Shapiro1 states that
incompressible friction formulae are applicable up to Mach numbers of 1.0.
Although Benedict, et al2 have shown that, as Mach numbers increase, coeffi-
cients of loss for abrupt enlargements decrease and those for abrupt contrac-
tions increase, the changes are significant only for large steps. (Benedict also
suggests that the ratio of exit to inlet total pressure is not affected by com-
pressibility in subsonic flow.)
For uniform ducts, the properties of the gas will vary along the flow path
because of friction. In subsonic adiabatic flow, the pressure, temperature, and
density will decrease while the velocity and Mach number will increase. In
supersonic flow, the opposite variations take place, and they can be very sig-
nificant depending on the length of the duct and the initial Mach number. In
fact, there is a maximum length of duct for a given mass flow rate, or con-
versely, there is a maximum mass flow rate for a given length of duct. This
choking effect occurs when the Mach number at the exit becomes unity.
In order to obtain a supersonic velocity in a duct system, there must be a
converging section followed by a diverging section, with Ma = 1.0 at the
throat. The diverging section becomes a supersonic nozzle just as the con-
verging section becomes a subsonic nozzle, with velocity increasing and pres-
sure decreasing in both. In a supersonic wind tunnel, such a nozzle combina-
tion is located upstream of the test section. Downstream of the test section, a
similar combination serves as a diffuser. The converging section is the super-
sonic diffuser, and the subsequent diverging section is the subsonic diffuser.
1
A. H. Shapiro, The Dynamics, and Thermodynamics of Compressible Fluid Flow, The
Ronald Press Co., New York, vol. II, Chapter 28, p. 1131.
2
R. P. Benedict and N. A. Carlucci, Handbook of Losses in Flow Systems, Plenum Press, New
York, 1966, pp. 23-26.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-79

Liquid Flow
The data for the incompressible flow of air and gases usually can be ap-
plied to the flow of liquids. However, since water at ordinary temperatures
weighs approximately 800 times more than air, elevation differences are more
significant in liquid flow than in airflow. In addition, under certain conditions
of reduced pressure or increased temperature, the vapor pressure of the liquid
may approach the pressure of the liquid. When it does, the vapor and the liq-
uid phases may exist simultaneously. Then cavitation will result from the al-
ternate formation and collapse of vapor bubbles in the liquid, producing noise,
wear on parts, and pressure surges. The friction charts for the flow of air
through conduits, Figures 2.13 and 2.15, can be used for the flow of water
provided that the appropriate corrections are made (continued on page 2-82)

Figure 2.36 Pipe Friction Chart in Diameters per Velocity Head

Adapted from the data of R. D. Madison and W. R. Elliot: "Pressure Losses for Liquid Flow
in Pipes," Chemical Engineering Progress, vol. 44, September 1948, pp. 703-706.

© 1999 Howden Buffalo, Inc.


2-80 FAN ENGINEERING

Figure 2.37 Pipe Friction Chart in Feet per 100 Feet

Adapted from the data of R. D. Madison and W. R. Elliot: "Pressure Losses for Liquid Flow
in Pipes," Chemical Engineering Progress, vol. 44, September 1948, pp. 703-706.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-81

Figure 2.38 Roughness Corrections for Pipes


Adapted from. the data of R. D. Madison and W. R. Elliot: "Pressure Losses for Liquid Flow
in Pipes," Chemical Engineering Progress, Vol. 44, September 1948, pp. 703-706.

Table 2.7 Resistances of Standard Pipe Fittings


to Flow of Liquids
Equivalent Lengths in Feet

Velocity Elbows Close 90° Valves(open)


Size Head Std Std Long Re- Pipe
in. Factor 90° 45° 90° turn Miter Globe Angle Gate Check
1/2 .490 1.6 .86 1.1 2.7 3.1 18.0 7.0 .70 7.0
3/4 .340 2.1 1.1 1.4 3.7 4.0 24.0 8.5 .90 9.0
1 .270 2.6 1.4 1.8 4.2 5.0 30.0 13.0 1.10 11.0
1-1/4 .180 3.6 1.9 2.4 5.8 7.0 40.0 16.0 1.50 15.0
1-1/2 .145 4.1 2.2 2.8 6.8 8.0 46.0 18.0 1.80 18.0
2 .110 5.3 2.9 3.5 8.5 10.0 60.0 25.0 2.30 23.0
2-1/2 .090 6.2 3.5 4.1 11.0 13.0 70.0 29.0 2.60 26.0
3 .067 7.8 4.0 5.3 13.0 15.0 90.0 35.0 3.50 33.0
4 .049 10.2 5.6 7.0 16.0 20.0 120.0 48.0 4.40 44.0
5 .036 13.0 7.0 9.0 21.0 25.0 145.0 60.0 5.50 55.0
6 .029 16.0 8.0 11.0 25.0 30.0 170.0 73.0 6.50 65.0
8 .021 20.0 11.0 14.0 34.0 41.0 220.0 98.0 9.00 85.0

Adapted from the data of Crane Co.: "Flow of Fluids Through Valves, Fittings, and Pipe,"
Tech. Paper No. 110, 1957, p. A-31.

© 1999 Howden Buffalo, Inc.


2-82 FAN ENGINEERING

for differences in kinematic viscosity and roughness. For convenience, two


similar charts, Figures 2.36 and 2.37, have been drawn for 70ºF water flowing
through clean steel or wrought iron pipes. Figure 2.38 can be used to correct
for very rough pipe. Figure 2.36, like Figure 2.13, is particularly useful when
differences in kinematic viscosity must be taken into account. The method, as
previously described, is simply one of maintaining the same Reynolds number
by calculating an equivalent velocity corresponding to the new kinematic
viscosity.
Much of the data on elbows and other fittings given for air and gases was
actually determined from tests conducted with water and other liquids.
Additional data on the resistance, in equivalent length, of certain standard
pipe fittings are given in Table 2.7. The second column in this table gives a
factor for converting the loss in equivalent length to the loss in velocity heads.
For instance, the equivalent length of a standard 3"-diameter 90º elbow is 7.8
feet, and the number of velocity heads corresponding to the loss is equal to
7.8 × 0.067, or 0.52 velocity heads.

Pressure and Flow Measurements


The instruments and methods that are generally used in fan engineering to
measure pressure and flow are discussed in this section. Since the
measurement of barometric pressure was examined in Chapter 1, only gage
pressures will be considered here. Because all the flow measurement methods
described include the use of pressure measurements, the discussions of
pressure and flow are interspersed.
There are two stages of pressure measurement: sensing and indicating.
Different sensors must be used depending upon whether they are intended to
respond to total, static, or velocity pressure. These sensors may be connected
to separate indicating gages or transducers. Alternatively, the indicator may
be integrated with the sensor.

Pressure Taps and Probes


A pressure tap is basically a small opening in the wall of a container or
duct. It may be equipped with a fitting for the connection of a gage or
transducer via a hose or other conduit as shown in Figure 2.39. Alternatively,
a transducer may be exposed to the pressure to be measured by being mounted
flush with the hole. Pressure taps, or wall taps as they are frequently called,
are used to sense static pressure. If one leg of a manometer is connected to
the wall tap and the other leg is open to the atmosphere, the gage static
pressure will be indicated. Wall taps can be used to sense the static pressure
in either a gas at rest or a moving gas stream. No special precautions are
necessary for gases at rest, but impact and aspiration effects must be avoided
for moving gas streams. Wall taps in a duct must be small, free from burrs,
and sufficiently far from disturbances such as elbows and internal obstructions
so that disturbance effects are negligible. Theoretically, the holes should be

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-83

Figure 2.39 Pressure Tap

infinitely small with square edges, but for Reynolds numbers less than 107 and
for hole sizes less than 0.01 duct diameters, the error will be less than 0.01
duct velocity pressures.1 Rayle2 has shown that failure to remove burrs can
result in a negative error of up to fifteen percent of the velocity pressure.
Wall taps can sense the pressure only in the vicinity of the hole, but the same
pressure should prevail across the duct section if the flow is reasonably
uniform.

Figure 2.40 Round Nose Static Pressure Probe

The static pressures across a section can be sensed with a static pressure
probe. This is usually a bent tube with a rounded nose pointed into the flow
and having a series of static taps drilled about 8 diameters from the nose along
the 24-diameter stem as shown in Figure 2.40. At this point, the nose effect,
which increases tap pressure, compensates for the decrease due to stem
effects. Other forms of static pressure probes require careful calibration to
determine the effects of flow on the reading. Directional probes, discussed
below, can also be used as static pressure probes.
1
R. P. Benedict, Fundamentals of Temperature, Pressure, and Flow Measurements, Second
Edition, John. Wiley and Sons, New York, 1977, p 348.
2
R. E Rayle, "Influence of Orifice Geometry on Static Pressure Measurements:" ASME Paper
No. 59-A-234, 1959.

© 1999 Howden Buffalo, Inc.


2-84 FAN ENGINEERING

Figure 2.41 Impact Probe

The total pressure at a point in a moving gas stream can be measured with
an impact probe pointed into the flow. The Pitot tube, which is a bent tube
having a hole in the nose as shown in Figure 2.41, is usually used, but direc-
tional probes may have impact taps, too. The shape of the nose is not impor-
tant except when the Pitot tube is combined with a static probe for velocity
pressure measurements.

Figure 2.42 Pitot-Static Tube

The velocity pressure at a point can be measured with a variety of probes


that sense both total and static pressure. The Pitot-static tube, illustrated in
Figure 2.42, is the standard for laboratory measurements. In practice, the total
pressure tap is connected to one side of a manometer, and the static pressure
tap is connected to the other side. Either static pressure and velocity pressure
or total pressure and velocity pressure can be measured simultaneously by
using a tee connection in the appropriate line and a second manometer.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-85

Figure 2.43 Stauscheibe Probe

A properly constructed and maintained Pitot-static probe is generally con-


sidered a prime instrument that does not require calibration. However, errors
will occur if the probe is not properly aligned with the flow, so a straightener
should be located upstream of the measuring section. Errors due to turbu-
lence, velocity gradient, and blockage are generally negligible.1 Nevertheless,
field measurement with a Pitot-static tube can be troublesome. A bent tube is
difficult to insert through a pipe coupling connection, and the small holes are
susceptible to plugging. These difficulties can often be avoided by using a
Stauscheibe, or type-S, probe. This probe, illustrated in Figure 2.43, has a
tube facing forward and also a reverse tube. The differential pressure across
the combined forward-reverse tubes is higher than the velocity pressure and
requires a calibration correction. Both the Pitot- static and the forward-
reverse probes are substantially non-directional and should not be used if the
flow directions are unknown.
Directional probes are generally patterned after the Fechheimer tube,2
which is a cylindrical probe having two independent holes located as shown in
Figure 2.44. To determine yaw, the probe is inserted into the stream; the two
holes are differentially connected to a manometer; and the probe is rotated
until the gage nulls. A line bisecting the angle between the holes indicates the
direction of the oncoming flow, and the angle between this line and the duct
axis is the angle of yaw. The holes are spaced so that a manometer connected
to only one tap would register the static pressure under ideal conditions. In
practice, however, a calibration is required for different Reynolds numbers.
Although a three-hole probe may have a different shape, it is operated in the
same way. In addition, the third hole, which is located to face into the flow, is
used to measure the total pressure. A five-hole probe has two additional holes
.
1
R. P. Benedict, Fundamentals of Temperature, Pressure, and Flow, Measurements, pp.- 350-
370.
2
C. J. Fechheimer, "Measurement of Static Pressure," Trans. ASME, vol. 48, 1926, pp. 965-
977.

© 1999 Howden Buffalo, Inc.


2-86 FAN ENGINEERING

Figure 2.44 Fechheimer Tube

located so that a differential pressure obtained by connecting them to a ma-


nometer, together with the differential from the total and static taps, can be
used to determine pitch from a calibration.

Pressure Gages and Transducers


The pressure taps and probes described in the previous section can serve
only as sensors. To measure pressure, an indicating device is also needed.
Various principles can be used to make different pressure gages and transduc-
ers.
For instance, the U-tube manometer can be made from 3/16 to 1/4-inch
bore glass tubing and a suitable linear scale. In operation, the gage is partially
filled with water or other fluid of known specific gravity. One leg is then
connected to a pressure tap or probe, and the other leg is connected to some
reference pressure. The liquid will be displaced by the higher pressure and
will rise in the other leg as shown in Figure 2.45. This illustrates the most
general case: one in which the fluid for the reference pressure is different
from the fluid for the pressure being measured. The densities and heights of
these fluids affect the displacement of the manometer fluid in accordance with

á ç ä ç h ä − ç ρ ä ç h + 1ä "#h .
ρ
! æå ãâ æå h ãâ æå ρ ãâ æå h ãâ $
g
p1 − p2 = ρ m 1+ 2 2 1 1
(2.103)
ρm
m
gc m m m

The pressure difference p2 − p1 is the gage pressure relative to the refer-


ence pressure p2 . In fan engineering, the reference leg of the manometer is
usually left open to the atmosphere so that the reference pressure is atmos-
pheric pressure. The units of the gage pressures obtained from Equation
. .

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-87

Figure 2.45 U-Tube Manometer

2.103 will depend on the units used for the density ρ m of the manometric
fluid, the height hm of the manometric column, the acceleration due to gravity
g , and the gravitational correction factor gc . Any consistent units can be
used for the terms in the bracketed factor. If ρ is in lbm/ft3, h in ft, g in ft/s2,
and gc in ft-lbm/lb-s2 then p1 − p2 will be in lb/ft2. . If ρ is in kg/m3, h in m,
g in m/s2, and gc in m-kg/s2-N, then p1 − p2 will be in Pa. In fan engineer-
ing, it is customary to drop the ρ m g / g c term and to directly report gage pres-
sures in terms of the height of the manometric fluid. Actually the ρ m g / g c
term should not be dropped but, rather, replaced by ρ m g / ρ m0 g0 where ρ m0 is
the density of the manometric fluid at standard temperature and g0 is standard
gravitational acceleration. The bracketed term is reducible to 1 − ρ 1/ ρ m if
ρ 1 = ρ 2 and h1 = h2 , which usually is so in fan engineering. This gas-
column-balancing effect is only about 0.1% and, therefore, is frequently ig-
nored. Rewriting Equation 2.103 and taking into account the above consid-
erations gives

ρm g
p1 − p2 = hm . (2.104)
ρ m 0 g0

Values for the densities of water and mercury at various temperatures are
listed in Table 2.8. Standard temperature for water gages is usually 68ºF or
20ºC, while standard temperature for mercury gages (particularly barometers)
is usually 32ºF or 0ºC.
In simple U-tube manometers, the height of the manometric column is
determined by reading the level of the fluid in the two legs and then subtract-
ing the lower from the higher. The scale used for these measurements should
be calibrated and the readings adjusted for any change in temperature. This

© 1999 Howden Buffalo, Inc.


2-88 FAN ENGINEERING

Table 2.8 Densities of Water and Mercury


Temperature Water Mercury

ºF ºC lbm/ft3 kg/m3 ρm / ρm0 lbm/ft3 kg/m3 ρm / ρm0

32 0 62.4229 996.511 1.00164 848.714 13595.1 1.00000


35.6 2 62.4291 1000.018 1.00174 848.402 13590.1 0.99963
39.2 4 62.4310 1000.048 1.00177 848.096 13585.2 0.99927
42.8 6 62.4291 1000.018 1.00174 847.784 13580.2 0.99890
46.4 8 62.4235 999.928 1.00165 847.478 13575.3 0.99854
50 10 62.4141 999.778 1.00150 847.172 13570.4 0.99818
53.6 12 62.4010 999.568 1.00129 846.860 13565.4 0.99782
57.2 14 62.3854 999.318 1.00104 846.554 13560.5 0.99746
60.8 16 62.3667 999.018 1.00074 846.248 13555.6 0.99709
64.4 18 62.3448 998.668 1.00039 845.942 13550.7 0.99673
68 20 62.3205 998.278 1.00000 845.630 13545.7 0.99637
71.6 22 62.2937 997.848 0.99957 845.324 13540.8 0.99601
75.2 24 62.2637 997.368 0.99909 845.018 13535.9 0.99565
78.8 26 62.2318 996.858 0.99858 844.712 13531.0 0.99529
82.4 28 62.1975 996.308 0.99803 844.407 13526.1 0.99492
86 30 62.0507 993.955 0.99567 844.101 13521.2 0.99456

Adapted from the data of J. A. Dean, Ed., Lange’s Handbook of Chemistry, 11th Edition,
McGraw- Hill Book Company, New York, 1973, p. 10-125.

adjustment is usually negligible if the gage is used near the calibration tem-
perature. The manometric fluid will form a meniscus at the surface. For wa-
ter in glass, the meniscus will be concave upward, while for mercury in glass
the meniscus will be concave downward. The level of the fluid should be de-
termined at the same point in the meniscus for each leg. Various adaptations
of the U-tube manometer are used to increase accuracy, to provide the con-
venience of a single reading, or to do both.
Precision U-tube manometers are called micromanometers. In one type of
micromanometer, the tubes are both enlarged at the levels where the readings
will be made. As shown in Figure 2.46, the level is determined by moving a
sharp, pointed index into contact with the surface of the manometric fluid.
Contact may be sensed visually or electrically. The difference between a ref-
erence level and a liquid level is determined by individual micromanometers
as shown. Other micromanometers use a precision lead screw with a mi-
crometer head. This lead screw may be motorized to move either leg, one of
which has a small inclined portion in which the meniscus is located precisely.
An inclined manometer is usually constructed as shown in Figure 2.47. In
operation, the gage is zeroed, the differential pressure is applied, and only the
deflection of the inclined leg is measured along the inclined scale. Since the
vertical leg deflection is not measured, that deflection either must be negligi-
ble or else the inclined scale must be adjusted to take that deflection into

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-89

account. The inclined scale, in effect, magnifies the reading of the


manometric deflection. The degree of magnification depends on the angle of
inclination α . The magnification factor would equal cot α except for the
deflection of the vertical leg. Inclined manometers should always be
calibrated against a micromanometer.

Figure 2.46 Micromanometer (Micrometer Type)

Figure 2.47 Inclined Manometer

Numerous other pressure transducers can be used for pressure


measurements. There are other manometers such as the McLeod gage and the
Zimmerli type for low absolute pressures. Other mechanical types use
Bourdon tubes, bellows, or diaphragms, together with a mechanical linkage,
for a wide range of pressure measurements. Piezoelectric elements are used
to measure sound pressures and other rapidly fluctuating pressures. Other
electrical pressure transducers use strain gages, potentiometers, capacitors,
transformers, and reluctance elements.

© 1999 Howden Buffalo, Inc.


2-90 FAN ENGINEERING

Flow Rate by Velocity Traverses


The flow rate at a particular cross section in a duct can be determined by
measuring the local velocities at a sufficient number of points to establish the
distribution and then integrating over the area. The local velocities can be
measured with various anemometers as discussed in the next section. Alter-
natively, they can be determined by measuring the local velocity pressures
with a velocity pressure probe as discussed in an earlier section. In either
case, the number and locations of the measuring points must be established.
Various means of establishing points are discussed below.
The velocity profile in a duct depends on Reynolds number, relative
roughness, and upstream disturbances. However, in general, the flow will be
retarded near the walls. Since it is virtually impossible to measure the flow
very close to the walls, assumptions about the velocity distribution in this re-
gion have to be made. The various methods described below differ in their
assumptions about the distribution in the main flow as well as in the boundary
flow.
One of the most common ways to establish the locations of the traversing
stations is to divide the duct into a number of equal areas and then take the
measurement at the centroid of each. Figure 2.48 shows these locations for
both rectangular and circular ducts. The sketches are for a particular number
of equal areas, but the associated tables give values for various other numbers
of measuring points. This method does not take into account the retardation
of the flow near the wall, so a positive error nearly always results. All meas-
uring points are given equal weight in the integrating procedure, which is de-
scribed below. In a variation of this method that was used in AMCA 210-67,
the outermost points for the circular duct traverse were relocated according to
the Prandtl 1/7-power law. For a traverse based on five points per radius, the
location of the last point was moved from 0.474 to 0.480.
Another technique, introduced by Winternitz and Fischl1 and known as the
log-linear method, is based on the Nikuradse formula for fully developed
flow. Figure 2.49 shows the locations of the traverse points for both rectan-
gular and circular ducts using this method. Once again, the sketches are for a
particular number of measuring stations, and the table for circular ducts offers
locations for various numbers of points per radius. The 26-point technique
shown for the rectangular duct was presented by Miles, Whitaker, and Jones2
for adoption as a British standard. The individual points are not weighted
equally in the integration procedure; rather, they are given the weightings
.
1
F. A L. Winternitz and C. F. Fischl, "A Simplified Integration Technique for Pipe-Flow
Measurement," Water Power, vol. 9, no. 6, June 1957, pp. 225-234.
2
D. J. Myles, J. Whitaker, and M. R. Jones, "A Simplified Integration Technique for Measur-
ing Volume Flow in Rectangular Ducts," NEL Report No. 251, National Engineering Labo-
ratory, East Kilbride, Glasgow, 1966.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-91

Figure 2.48
Centroids of Equal Areas for Rectangular and Circular Ducts

© 1999 Howden Buffalo, Inc.


2-92 FAN ENGINEERING

Figure 2.49
Log-Linear Traverse Points for Rectangular and Circular Ducts

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-93

Figure 2.50
Log-Tchebycheff Traverse Points for Rectangular and Circular Ducts

© 1999 Howden Buffalo, Inc.


2-94 FAN ENGINEERING

shown in the table for the 26-point method. The individual readings are,
however, equally weighted for circular ducts. Various profiles were explored
by Brown,1 which led to the adoption of the 24-point technique (6 radii, each
with four points) in AMCA 210-74/ASHRAE 51-75.
The log-Tchebycheff method postulates a logarithmic distribution near the
wall and a polynomial distribution of velocity elsewhere. Figure 2.50 shows
the locations of the measuring stations for both rectangular and circular ducts
with varying numbers of points per side or radius. All measurements are
equally weighted in the integration procedure. This method is referenced in
various international standards.
Regardless of whether the traversing method is based on the centroids of
equal areas, on the log-linear distribution, or on the log-Tchebycheff
distribution of measuring points, it is essential that the probe be properly
aligned with the flow. Most laboratory standards call for a calming length and
a straightener to ensure that the direction of the flow is parallel with the duct
walls across the measuring plane. This facilitates the use of any non-
directional probe provided that it is properly aligned with the duct axis. Such
alignment, however, is not necessarily used in non-laboratory situations. For
on-site testing, a directional probe is recommended. Such a probe should be
used properly aligned with the flow as determined by the nulling feature of
such probes, and only the component of velocity along the duct axis should be
used for the calculation of flow rate. It is also important to check that the duct
is clear of obstructions and to determine the inside area accurately. Whenever
it is expected that a traverse will take considerable time to execute, monitor
the flow to ensure against variations with time. This can be done by fixing a
second probe in a suitable location and monitoring its readings.
The integration procedure to find flow rate from centroid-of-equal-area
measurements, log-linear measurements, or log-Tchebycheff measurements is
really an arithmetic procedure used to estimate the integral. First, compute
the individual local normal velocities Vnj apply weighting factors Fj as
required, and then average arithmetically. This average normal velocity,
when multiplied by the cross-sectional area A , gives the volume flow rate Qû :

3 8
n
1
Qû = A ∑ FjVnj . (2.105)
n j =1

The individual velocities can be calculated using Equations 2.28 through 2.31
as appropriate. Note that it is not proper to simply average the velocity
pressures even when the density is constant across the section. The cited
equations can be used to determine the average velocity when the density is
1
N. Brown, "A Mathematical Evaluation of Pitot Tube Traverse Methods," ASHRAE Paper
No. 2335, presented at Atlantic City, 1975.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-95

constant only by using the average square root of the velocity pressure. The
mass flow rate mû can also be determined from these measurements using

mû = A
1 n
3
∑ ρ j FjVnj
n j =1
8 (2.106)

where ρ j is the local mass density at the various measuring points.


Graphical integration, too, can be used to find flow rate from traverse
measurements. This method offers much more flexibility in selecting the
number and locations of the measuring points. For a rectangular duct, the
measurements for each traverse of the probe should be plotted as ordinate
with their normalized locations as abscissa. The normalized locations are the
actual distances from the duct wall divided by the total distance across the
duct. Additional measurements should be made whenever the distribution
appears to be questionable. Some assumption will be required regarding the
flow at the boundaries. The area under the curve can be determined by using
a planimeter. This area is the mean velocity for the traverse. A similar pro-
cedure is used for all the other traverses, and the resulting mean values are
then plotted against their normalized locations in the other dimension of the
duct. The area under this new curve is the mean for the entire cross section of
the duct. For a circular duct, the measurements for each traverse are plotted
against the square of their normalized radial locations, and instead of meas-
uring the area under each curve, the values on selected circumferences are ar-
ithmetically averaged and the averages are then plotted against the square of
their normalized locations. The area under this curve is the mean for the en-
tire cross section of the duct
Even when arithmetical methods are used to find the flow rate, it may be
wise to check the measured distribution graphically. Unusually high or low
values should be investigated.
A single reading can sometimes be used to estimate the flow rate. When
the flow is turbulent and the test section is at least forty diameters downstream
from any disturbance, the approximate mean velocity Vm can be determined
from the centerline velocity Vc and the friction factor f using

1
Vm = Vc . (2.107)
1 + 1439
. f

Average Pressures from Traverses


In the early portions of this chapter, the fundamental equations of fluid
flow were examined in their one-dimensional form, and appropriate average
.

© 1999 Howden Buffalo, Inc.


2-96 FAN ENGINEERING

values were suggested for use in those equations when the flow was not one-
dimensional. These same average values are used here, but other types of av-
erages are possible and may even be desirable in some situations.
In the preceding section, volume flow rate was determined from an area-
weighted average velocity. Similarly, mass flow rate was determined from an
area-weighted average product of density and velocity. These quantities are
useful in calculating the average pressures at the traverse planes. Another
useful quantity is the average density ρ at the traverse plane, which can be
defined as the mass flow rate mû divided by the volume flow rate Qû , or

1 n
3 8 ∑ 3ρ 8
n


∑ ρ j FjVnj
n j =1
A
j =1
j FjVnj
ρ= = = (2.108)
Qû 1 n
3 8 ∑ 3F V 8
n
A ∑ FjVnj j nj
n j =1 j =1

where ρ j is the density and Vnj is the normal velocity at the measuring point.
Note that this equation is entirely consistent with the continuity equation and
that the average density is a volume-flow-weighted value.
The average static pressure pS is also a volume-flow-weighted value:

1 n
3 8 ∑3p 8
n
A ∑ p FV
n j =1 S j j nj Sj FjVnj
j =1
pS = = (2.109)
1 n
3 8 ∑ 3F V 8
n
A ∑ FjVnj j nj
n j =1 j =1

Note that, since the static pressure is joined with the density in the flow work
term of the general energy equation, it is consistent that averages for both
static pressure and density are volume-flow-weighted.
The average specific kinetic energy eK and average specific potential en-
ergy e P are both mass-flow-weighted values. The specific potential energy is
usually negligible in fan engineering, but the average specific kinetic energy
can be determined from

1 1 n
3 8 ∑ 3ρ 8
n
A ∑ ρ j Fj Vnj
3 3 3 3
j Fj Vnj
2 n j =1 1 j =1
eK = = . (2.110)
3 8 ∑ 3ρ F V 8
n n
1
gc A ∑ ρ j FjVnj
2 gc
j j nj
n j =1 j =1

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-97

The average velocity pressure pV can be obtained from the average spe-
cific kinetic energy and the average density using

ρ eK
pV = (2.111)
Cp

where C p is 5.193 in U.S. customary units and 1.0 in SI units.


The average total pressure is simply the sum of the average static pressure
and the average velocity pressure as indicated by:

pT = pS + pV . (2.112)

Anemometers
The term anemometer is used here to describe those instruments that can
be used to determine local velocities but utilize principles other than the rela-
tionship between pressure and velocity. Included in this classification are ro-
tating-vane anemometers, swinging-vane anemometers (velometers), turbine
meters, hot-wire anemometers, Kata thermometers, and laser-Doppler meters.
The vane anemometer is a rotating-vane type of instrument that can be
designed either to register velocity or to give a reading over a timed interval
that can then be converted to velocity. Such an anemometer requires frequent
calibration because readings are greatly affected by the condition of the bear-
ings. This instrument is useful in measuring low velocities at supply-register
and exhaust-grill openings when in-duct measurements are not convenient and
high accuracy is not required.
The operator should move the anemometer slowly and uniformly over the
whole flow area in order to arrive at an average determination. Exhaust ca-
pacity Qû can be calculated by using the average velocity V (obtained with
the dial of the anemometer facing the grill), the gross area of the grill Ag and
a correction factor K E that can be determined from Table 2.9 in

Qû = K EVAg . (2.113)

Supply capacity Qû can be determined by using the average velocity V (ob-


tained with the dial facing the operator), the gross area Ag plus the net free
area A f , and a correction factor KS determined from Table 2.9 in

3 8
Qû = 0.5KSV Ag + A f . (2.114)

The turbine meter can be designed for radial flow, axial flow, or tangential

© 1999 Howden Buffalo, Inc.


2-98 FAN ENGINEERING

Table 2.9 Values of KS and K E in Anemometer Formulae

Average Indicated Velocity, fpm


Grilles 150 200 300 400 500 600 700 800
KS Supply .952 .957 .967 .977 .985 .992 .998 1.00
KE Exhaust .762 .772 .789 .806 .820 .828 .832 ----

Adapted from the data of L. E. Davies, "The Measurement of Flow of Air through Registers
and Grilles," Trans. ASHVE, vol. 36, 1930, pp. 201-224.

flow, but the axial-flow type is probably most used for gas metering. It is
usually mounted on the end of a probe for traversing the duct. The velometer
is a direct-reading, air-velocity meter that is operated by the impact of the
flowing air against a swinging vane. By suitably designing the inlet and outlet
jets, the instrument can be adapted to read either high or low velocities. The
velocity can be taken at any point, or an averaging jet can be used to obtain
the average velocity over an area. These instruments have the advantages that
they can be read directly, are easy to use, and require no additional equipment.
Occasionally, the meter itself is placed in the airflow being investigated. If
the flow is appreciably different at the two ends of the meter, such as in front
of exhaust hoods, a material error may result.
A hot-wire anemometer consists of a resistance wire placed in the air
stream and heated by an electric current. The temperature of a current-
carrying wire in an air stream depends on the current and the rate of heat loss
to the air. Since this heat loss varies with velocity, airflow can be determined
from the relation between the current and the temperature of the wire, or be-
tween the current and the temperature rise of the air over the wire. Wire tem-
perature can be established in terms of its resistance.
The time required for the reading of a heated Kata thermometer to fall
through a specified interval (usually 100º to 95º) is a measure of the non-
directional air velocity. Heated-thermocouple and heated-thermometer ane-
mometers each use a pair of temperature-sensitive devices, one heated and the
other not. The difference in reading for a given heating rate is a measure of
the air velocity over the elements. All of the above devices (when properly
calibrated) can be used to measure low velocities with good accuracy. Refer
to the earlier section on flow rate from traverses for traversing patterns and
calculation procedures.

Differential Pressure Meters


Differential pressure meters can be of the orifice, nozzle, or venturi- meter
types. Each of these devices produces a pressure difference that can be cor-
related with the flow rate. These kinds of meters differ from both anemo-

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-99

meters and pressure probes in that the whole flow must pass through the me-
ter, whereas both pressure probes and anemometers must be traversed across
the flow and sample only a portion at a time.
Given a plate with a hole in it and a pressure difference, fluid will flow
from all directions on the higher-pressure side and issue as a jet on the lower
pressure side. The jet becomes substantially unidirectional at a point some-
what downstream from the opening. At this point, called the vena contracta,
the contraction of the jet (which occurs because of the multi-directional ap-
proach and corresponding directional momentum on the high-pressure side)
becomes maximal. That is, the area at the vena contracta is the minimum that
can be achieved with such a free jet. When the edge of the opening is per-
fectly square, the area of the stream at the vena contracta will be very close to
60% of the area of the opening. The location of the vena contracta will be
one half of an opening diameter downstream of the face of the opening. Con-
siderable static pressure is transformed into velocity pressure in producing
acceleration by means of an orifice. This conversion is highly efficient, and
the loss is amazingly low. However, the opposite conversion, from velocity
to static pressure, requires careful design in order to obtain high efficiencies
or avoid high losses. The pressure loss between the upstream face and the
vena contracta affects the velocity profile at the vena contracta.
The flow coefficients, as defined below, can be used to determine flow
rates and pressure losses for any orifice or nozzle. The subscript 1 will be
used to denote the upstream location, whether it be pipe or plenum, and the
subscript 3 to denote the plane of the vena contracta. The subscript 2 will be
used to denote the plane of area A2 . For an orifice, A2 is the area of the
opening at entrance. For a nozzle, A2 is the area of the opening at exit.
Based on Equation 2.26, the energy balance for incompressible flow for
any orifice or nozzle can be written:

p S 1 + pV 1 = pS 3 + pV 3 + p L1− 3 (2.115)

in which the pressure loss is denoted by p L1− 3


The coefficient of contraction CC , defined as the ratio of the area of the
vena contracta to that of the area at entrance to an orifice or at exit from a
nozzle, can be written:

A3
CC = (2.116)
A2

The coefficient of velocity CV , defined as the ratio of the actual average


velocity at the vena contracta to the velocity that would be obtained if there
were no loss, can be expressed by

pV 3
CV = . (2.117)
p L1− 3 + pV 3

© 1999 Howden Buffalo, Inc.


2-100 FAN ENGINEERING

The coefficient of resistance CR , defined as the ratio of the loss of total


pressure to the velocity pressure at the vena contracta, can be written:

p L1− 3 1
CR = = 2 − 1. (2.118)
pV 3 CV

The coefficient of discharge CD of an orifice or nozzle can be defined as


the product of the coefficients of contraction and velocity, or

CD = CC × CV . (2.119)

Equation 2.5, which is for incompressible flow, can be rewritten:

Qû = V3 A3 = V2 A2 = V1 A1 (2.120)

where volume flow rate Qû is in cfm when areas A are in ft2 and velocities V
are in fpm. The corresponding SI units are m3/s, m2, and m/s. Combining this
with Equation 2.28 produces

p
Qû = Cv A3 V 3 (2.121)
ρ

where Cv is 1097 if velocity pressure pV is in in. wg and density ρ is in


lbm/ft3, or Cv is 2 if SI units are used. Incorporating the definition of coef-
ficient of contraction yields

pV 3
Qû = Cv CC A2 . (2.122)
ρ

Similarly, using the definition of coefficient of velocity results in

pV 3 + pL1− 3
Qû = Cv CC CV A2 . (2.123)
ρ

Substituting the coefficient of discharge for its equivalent produces

pV 3 + p L1− 3
Qû = Cv CD A2 . (2.124)
ρ

Manipulating Equation 2.115 and making the appropriate substitutions in


Equation 2.124 gives

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-101

pT 1 − pS 3
Qû = Cv CD A2 or (2.125)
ρ

pV 1 + pS 1 − pS 3
Qû = Cv CD A2 . (2.126)
ρ

After removing the approach velocity pressure pV 1 from under the radical,

pS 1 − pS 3
Qû = Cv CD A2φ i . (2.127)
ρ

The true velocity of approach factor for incompressible flow φ i can be


found by equating Equations 2.126 and 2.127, or 2.125 and 2.127:

pV 3 + pS 1 − pS 3 pT 1 − pS 3
φi = = . (2.128)
pS 1 − pS 3 pS 1 − pS 3

Using the relationships embodied in Equations 2.115 and 2.117,

1 1
φi =
1
1 − pV 1 / pV 3 + p L1− 3
=
6
1 − CV pV 1 / pV 3
2
.
1 6 (2.129)

The true velocity of approach factor can also be written as a function of area
ratio and coefficient of velocity or discharge:

1 1
φi = =
1 − CV
2
1A / A 6
3 1
2 2
1
1 − CD A2 / A1 6.
2
(2.130)

A pseudo velocity of approach factor, i.e. φ i ′ = 1 / 1 − A2 / A1 1 6 2


or one
based only on area ratio, is frequently used. In such cases a pseudo coeffi-
cient of discharge, i.e. C ′ ≠ C × C or one different from that defined, must
D C V
be used in Equation 2.127 (and in Equation 2.136 which is given in the sec-
tion on nozzles). Note that φ ′C ′ = φ C .
D D
For a nozzle or orifice discharging into the atmosphere, the static pressure
at the vena contracta p3 is equal to zero. Thus, it is necessary only to meas-
ure either the average total pressure or the average static pressure ahead of the
nozzle or orifice in order to find the flow rate. Since it is usually much more
convenient to measure the average static pressure rather than the average total
pressure, Equation 2.127 is used even though it does involve the velocity of
.

© 1999 Howden Buffalo, Inc.


2-102 FAN ENGINEERING

approach factor. Also note that for a plenum approach the velocity pressure
pV 1 equals zero, and therefore, the velocity of approach factor φ i equals 1.0.
Numerous values for the various flow coefficients have been established
empirically. Where similar conditions of flow and measurement exist, these
data can be used to predict flow rates from pressure readings. However, if
there is any doubt whatsoever about the uniformity of the upstream flow or
the conditions of measurement, or both, an in-place calibration against a Pitot
tube should be made.
There are numerous differential pressure devices that can be used to
measure flow. The flow coefficients for some of these devices, together with
the method of pressure measurement and brief details of construction, are
given in the paragraphs below.

The Square-Edged Orifice


The square-edged orifice is a flat plate with a hole in it. The hole may be
of any shape. In the extreme, it may even be a slot with a very large aspect
ratio. The coefficient of discharge with a plenum approach is about 0.6 re-
gardless of shape. Orifice plates should be perfectly flat and the holes accu-
rately made for precise area determinations. All burrs must be removed from
the inlet edge so that the air flows over a sharp 90º corner. The thickness of
the edge should not exceed 1/50 of the orifice diameter. Recommended
thicknesses for various diameters are 1/16 in. for up to 6 in., 3/32 in. for up to
12 in., 1/8 in. for up to 24 in., and 3/16 in. for up to 48 in. Thicker plates,
when required for rigidity as in high-pressure work, should be beveled away
on the downstream side of the orifice. Table 2.10 lists flow coefficients for
square-edged orifices with various ratios of hole diameter to pipe diameter.
The combined velocity-of-approach factors and coefficients of discharge
φ i CD represent the averages for pipe sizes ranging from 1-1/2 in. to 16 in.
These data are based on vena contracta taps and pipe Reynolds numbers of
106 and will give reasonably accurate results whether flange, radius, or vena
contracta taps are used. Figure 2.51 shows the location of the various types
of pressure taps. Of the remaining coefficients of flow listed in Table 2.10,
the coefficient of velocity CV for orifices in pipes was assumed to be 0.975,
which is the average of values generally listed for orifices discharging from
plenums. The coefficient of contraction CC was calculated using the listed
values of φ i CD the appropriate definitions, and the assumed value for CV .
The coefficient of discharge CD was then calculated using the coefficients of
velocity and contraction, and the velocity-of-approach factor determined ac-
cordingly.
The pseudo velocity-of-approach factor φ ′ and the pseudo coefficient of
i

discharge CD ′ are also given. All data are based on the difference in pressure
from the upstream location to the vena contracta location. The static pressure
at the vena contracta will be zero only when discharge is to the atmosphere.
Otherwise, the discharge pressure must be measured.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-103

Figure 2.51 Square-Edged Orifice and Pressure Tap Locations

Adapted from the data of ASME: ASME Power Test Codes, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5; 4-1959, p. 9.

© 1999 Howden Buffalo, Inc.


2-104 FAN ENGINEERING

Table 2.10 Flow Coefficients for Square-Edged Orifices


(Also Abrupt Contractions)

D2 / D 1 .7 .65 .6 .55 .5 .45 .4 .35 .3 From


A2 / A1 .490 .422 .360 .302 .250 .202 .160 .122 .090 Plenum

φ i ′ CD .699 .670 .650 .635 .623 .614 .608 .603 .600 .60-.61
CV .975 .975 .975 .975 .975 .975 .975 .975 .975 .97-.98
CC .673 .660 .648 .638 .631 .625 .621 .617 .615 .62
CD .660 .644 .632 .623 .615 .609 .605 .601 .599 .60-.61
CR .052 .052 .052 .052 .052 .052 .052 .052 .052 .06-.04
φi 1.059 1.041 1.028 1.018 1.012 1.007 1.005 1.003 1.001 1.00

φi′ 1.147 1.104 1.072 1.049 1.033 1.022 1.013 1.007 1.004 1.00

CD ′ .609 .607 .606 .605 .604 .602 .600 .598 .597 .60-.61

Adapted from the data of ASME: ASME Power Test Codes, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5;4-1959, pp. 20-39.

Figure 2.52 Pressure-Loss Coefficients for Flow Meters


Adapted from the data of ASME: ASME Power Test Codes, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5; 4-1959, p. 12.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-105

Table 2.11 Equivalent Flow Coefficients for Square-Edged Orifices


Formed by a Uniform Downstream Pipe

D2 / D1 .7 .65 .6 .55 .5 .45 .4 .35 .3 From


A2 / A1 .490 .422 .360 .302 .250 .202 .160 .122 .090 Plenum

CD * .861 .850 .839 .831 .823 .818 .813 .811 .81 .81-.83
Cp * .912 .885 .863 .849 .835 .824 .817 .813 .81 .81-.83
CR * .35 .39 .42 .45 .47 .49 .51 .52 .53 .53-.48-

An average coefficient of resistance is shown. This value is calculated


from the assumed coefficient of velocity. It is nearly equal to the average
usually shown for orifices discharging from plenums. The contraction loss
p L1− 3 can be determined from the coefficient of resistance CR and the velocity
pressure at the vena contracta pV 3 using

p L1− 3 = CR pV 3 . (2.131)

When discharge is to the atmosphere, the pressure corresponding to the ve-


locity at the vena contracta represents an additional loss.
When the orifice is in a uniform pipe, the overall contraction and re-
expansion loss p L1− 4 can be determined from the meter differential pS 1 − pS 2
and the data of Figure 2.52 using

1
p L1− 4 = K L pS 1 − pS 2 .6 (2.132)

When the square-edged orifice is formed by a uniform downstream pipe,


some of the velocity pressure at the vena contracta is converted to, or re-
gained as, static pressure, provided that the discharge pipe is long enough for
complete expansion to take place. The minimum length of pipe usually nec-
essary for this to occur is three opening diameters. When such a configuration
is used, the coefficients listed in Table 2.10 still apply provided that the pres-
sure at the vena contracta is used as indicated. Alternatively, the equivalent
coefficients of discharge CD * and resistance CR * given in Table 2.11 can be
used in the following equations:

pT 1 − pS 4
Qû = Cv CD * A2 , (2.133)
ρ

pS 1 − pS 4
Qû = Cvφ i CD * A2 , and (2.134)
ρ

p L1− 4 = CR * pV 4 (2.135)

© 1999 Howden Buffalo, Inc.


2-106 FAN ENGINEERING

where the subscript 4 indicates the location at which full flow is established in
the downstream pipe. With a short pipe, pS 4 will be zero.

Rounded-Entry Nozzles
If the edge of an orifice opening is rounded rather than square, the coeffi-
cient of contraction is increased. In the extreme, when the opening is rounded
to the degree indicated in Figure 2.53, the coefficient of contraction reaches
unity. There is no contraction beyond the nozzle, and the stream issues from
the nozzle full bore.
In all real nozzles, there is some small loss of energy due to fluid friction.
For accurately made contours, this loss will be less than the corresponding
loss in a free jet issuing from a square-edged orifice. It increases with rough-
ness but decreases with Reynolds number. The ASME Power Test Code PTC
19.5; 4-1959 shows a coefficient of discharge of 0.994 for long-radius, low-
ratio nozzles handling air when the Reynolds number is above 5 × 105. The
coefficient of discharge drops to a value of 0.942 at a Reynolds number of 104
for long-radius, high-ratio nozzles. Calculations from a later reference1 yield
slightly lower values at high Reynolds numbers and slightly higher values at
low Reynolds numbers.
As distinguished from the reference area of an orifice, which is the open-
ing area or entrance area, the reference area for a nozzle is the discharge or
exit area. When the coefficient of contraction is 1.0, A2 is equal to A3 . Ac-
cordingly, the capacity equations are often rewritten using the area of the vena
contracta A3 :

pS 1 − pS 3
Qû = Cv CD A3φ i and (2.136)
ρ

p − pS 3
Qû = Cv CD A3 T 1 . (2.137)
ρ

Velocity-of-approach factors, together with combined coefficients using a co-


efficient of discharge of either 0.99 or 0.98, are listed in Table 2.12 for vari-
ous ratios of nozzle discharge area A3 to upstream pipe area A1 . The coeffi-
cient of velocity can be taken to be equal to the coefficient of discharge when-
ever full-bore flow is assumed, that is, whenever the coefficient of contraction
equals 1.0. Accordingly, the coefficient of resistance can be taken to be 0.02
for coefficients of discharge of 0.99, and 0.04 for coefficients of discharge of
0.98. The loss can be calculated from either Equation 2.131 or 2.132 using
CR as listed above or K L from Figure 2.52.
A short-radius nozzle known as the I.S.A. (International Standards Asso-
ciation) nozzle is illustrated in Figure 2.54. According to ASME Power Test
.
1
Measurement of Fluid Flow in Pipes Using Orifices, Nozzles, and Venturi, ASME MFC-
3M-1989, ASME, New York.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-107

Table 2.12 Flow Coefficients for Long-Radius-Flow Nozzles

D3 / D1 .7 .65 .6 .55 .5 .45 .4 .35 .3 From


A3 / A1 .490 .422 .360 .302 250 .202 .160 .122 .090 Plenum

φ i 1.147 1.104 1.072 1.049 1.033 1.022 1.013 1.007 1.004 1.00
.99φ i 1.136 1.093 1.061 1.039 1.023 1.012 1.003 .997 .994 .99
.98φ i 1.124 1.082 1.051 1.028 1.012 1.002 .993 .987 .984 .98

Figure 2.53 Long-Radius-Flow Nozzle (ASME)


Adapted from the data of ASME: ASME Power Test Codes, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5; 4-1959, p. 13.

Figure 2.54 Short-Radius-Flow Nozzle (ISA)


Adapted from the data of ASME: ASME Power Test Code, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5; 4-1959, p. 26.

© 1999 Howden Buffalo, Inc.


2-108 FAN ENGINEERING

Table 2.13 Flow Coefficients for ISA-Flow Nozzles

A3 / A1 .5 .45 .4 .35 .3 .25 .2 .15 .1 .05

φ i CD 1.081 1.059 1.041 1.028 1.016 1.0061 .999 .993 .989 .987

Adapted from. the data of ASME: ASME Power Test Codes, Instruments and Apparatus, Part
5, Measurement of Quantity of Materials, PTC 19.5; 4-1959, p. 26.

Code PTC 19.5; 4-1940, the combined coefficient of discharge and velocity-
of-approach factor is as listed for various areas in Table 2.13 for Reynolds
numbers above approximately 105. Data on the I.S.A. nozzle was omitted
from PTC 19.5; 4-1959 presumably because results obtained with it are not
always accurate.

Conical-Entry Nozzles (Converging Tapers)


If the edge of an orifice opening is beveled, performance will fall between
that of a square-edged orifice and a rounded-entry nozzle, approaching the
former as the included angle between sides nears 180º. The effect of the de-
gree of tapering on coefficient of discharge is illustrated in Figure 2.55 for
discharge into the atmosphere. The three different curves correspond to the
three different approach conditions as illustrated in the accompanying
sketches. For all three the jet will continue to contract beyond the discharge
of the nozzle producing a vena contracta at a point slightly downstream. The
vena contracta location can be determined experimentally by means of an
impact tube in the center of the stream since it is at that point where the im-
pact reading is greatest.

Figure 2.55 Discharge Coefficients for Converging Tapers


Discharging into the Atmosphere

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-109

If a converging taper is used as a nozzle for flow-measuring purposes,


Equation 2.125 can be used to calculate flow rate Qû . The coefficient of
discharge can be obtained from Figure 2.55. The area of the nozzle exit A2 ,
the density ρ of the air or gas, and the upstream total pressure pT1 can all be
measured. The static pressure at the vena contracta is zero.

Re-entrant Pipes
The flow nozzle, or rounded-entry orifice, represents one extreme in
orifice construction. The opposite extreme is the re-entrant pipe. A perfect
re-entrant pipe would have an infinitely thin pipe wall and would be just long
enough to ensure completely reversed flow along the outer surface but would
not be so long that the flow would reattach itself to the pipe wall. Re-
attachment is generally assumed to occur between three and four diameters
downstream from the opening. Table 2.14 values are all approximate since
they are based on assumed values for CC and CV Flow rates can be
determined from measurements of the upstream plenum total pressure pT1 and
static pressure at the vena contracta pS 3 using Equation 2.125 and CD .
Alternatively, Equation 2.133 and CD * can be used with measurements of the
static pressure at the point where the flow re-attaches pS 4 . Both involve the
area of the pipe opening A2 . Similarly, total pressure losses can be
determined from either Equation 2.131 or 2.135 and the appropriate
coefficients and velocity pressures.

Table 2.14 Flow Coefficients for Re-entrant Pipes

Flow CC CV CD CR CD * CR *

Separated .545 .975 .531 .052


Re-attached .545 .975 .531 .052 .799 .872

Venturi Meter
The venturi meter provides a very convenient method of producing a
pressure difference suitable for measurement and convertible to flow rate. As
indicated in Figure 2.56, the venturi meter consists of a combination of
converging and diverging tapers usually connected by a short, straight pipe
known as the throat. For minimum loss, the included angle in the convergent
section should be 30º or less, and the included angle in the divergent section
should be 7º to 8º. Suitable pressure differences are obtained with diameter
ratios of 1/2 to 1/3 or with area ratios of 1/4 to 1/9. For highest accuracy, the
meter should be calibrated in place; however, for the proportions listed, a
coefficient of discharge equal to about 0.98 can be used in Equation 2.136
where subscript 1 refers to the upstream pipe and subscript 3 refers to the
throat. The pressure loss can be estimated from Figure 2.52 and Equation
2.132.

© 1999 Howden Buffalo, Inc.


2-110 FAN ENGINEERING

Figure 2.56 Venturi Meter and Pressure Graph

Two datum lines are shown in Figure 2.56. The lower datum line shows
positive gage pressures throughout the meter. The higher datum line
illustrates that a negative gage pressure can be developed in the throat,
depending on the upstream gage pressure level.

Compressible-Flow Measurements
The discussions and equations in the preceding sections on pressure and
flow measurement are generally based on the assumption of incompressible
flow. As noted in an earlier section on stagnation properties, when a moving
fluid is brought to rest there is an increase in temperature, pressure, and
density. If a probe is inserted into a stream, the fluid will come to rest at some
point on the probe. A total pressure, or impact, probe is designed to measure
this increased pressure and does so whether the flow is compressible or
incompressible. A static pressure tap is designed to avoid stagnation whether
the flow is compressible or incompressible. The differential pressure, or
velocity pressure, can be used directly in Equation 2.28 to determine velocity
for incompressible flow. However, as indicated by Equation 2.53, a
somewhat more complicated expression must be used for compressible flow.
Such an expression is
á 1 ç ∆ p ä + γ − 1 ç ∆ p ä "#
2

æ ã æ ã
! 2γ å p â 6γ å p â #$
pV = ∆ p 1− 2
(2.138)

where ∆ p is the differential pressure, p is the absolute static pressure, and γ


is the ratio of specific heats.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-111

When a temperature probe is inserted into a moving stream, the indicated


temperature will be between the static temperature and the stagnation
temperature as given by Equation 2.51. To determine the absolute static
temperature TS from the indicated temperature T1 and flow measurements, an
expression such as

T1
TS =
γ −1 ∆ p çæ äã (2.139)
1 + FR
γ p å â
can be used. The recovery factor FR will depend on the design of the
temperature probe.
The flow through a well-designed nozzle can be considered isentropic for
all practical purposes. The subscript 1 will be used to denote the inlet or
admission plane, the subscript 2 to denote the plane of the throat, the subscript
3 to denote the plane just beyond discharge or the area into which the nozzle
discharges. The weight rate of flow wû can be determined from the specific
weight w1 and the absolute static pressures p1 and p2 using any consistent
set of units in

ç γ ä
áç p ä ç p ä "# 2 γ +1

2 gæ
å γ − 1ãâ p w !æå p ãâ − æå p ãâ ##$ .
γ γ
wû = φ c CD A2 1 1
2 2
(2.140)
1 1

The volume rate of flow at inlet conditions Qû , can be determined from the
gage pressures pS1 and pS 2 in in. wg, area A in ft2, and density ρ in lbm/ft3
using 1097 for Cv in

pS 1 − pS 2
Qû = Cvφ c CD A2ψ . (2.141)
ρ

Cv is 2 for SI units. The velocity of approach factor φ c for compressible


flow can be determined from Figure 2.57, which is based on

1
φc = . (2.142)
çæ A äã çæ p äã
2
2
γ
1−
åAâ åpâ
2 2

1 1

The expansion or compression factor ψ can be determined from Figure 2.58,


which is based on

© 1999 Howden Buffalo, Inc.


2-112 FAN ENGINEERING

Figure 2.57 Velocity of Approach Factors

á ç p ä "# γ −1

1− æ ã
γ

γ çp ä å p â ## .
2 2

æ ã
γ

γ −1å p â
ψ= 1

##
2
(2.143)
p
1 1− 2

!
p 1
#$
The coordinate scales of Figure 2.58 were chosen so that data could be read
directly for expansion processes in which the absolute static pressure ratios
p1 / p2 exceed unity. For compression processes in which the reciprocal
pressure ratios p2 / p1 exceed unity, the compression factor ψ c is the
reciprocal of the expansion factor ψ e indicated on the chart. Similarly, the
temperature at the downstream location T2 will be higher than the upstream
temperature T1 after compression, whereas the reverse is true after expansion.
The indicated temperature ratio is for expansion, so its reciprocal should be
used for compression. The product of φ c and ψ is usually called Y in
compressor test codes.
Because pressure variations cannot be transmitted through a fluid at a
velocity greater than that of sound through the fluid, a limiting condition
develops when the fluid velocity reaches the acoustic velocity.

© 1999 Howden Buffalo, Inc.


CHAPTER 2 – FLUID FLOW 2-113

Figure 2.58 Expansion Factors

In the equations presented above, it cannot be assumed that the pressure at


the throat p2 (even with a steadily converging shape) is always equal to the
backpressure p3 into which the nozzle discharges. There is a critical pressure
1
ratio p2 / p1 6 cr
that defines the lowest throat pressure that can exist for any
admission pressure p1 . The numerical value of this critical pressure ratio is
about 0.53 for normal air, 0.55 for highly superheated steam, and about 0.58
for saturated steam. It can be determined from the ratio of specific heats γ
using

çæ p äã = çæ 2 äã
γ
γ −1

å p â å γ + 1â
2
. (2.144)
1 cr

The critical or acoustical velocity c2 cr can be determined from

gcγ p2
c2 cr = = gcγ RT2 (2.145)
ρ2

© 1999 Howden Buffalo, Inc.


2-114 FAN ENGINEERING

using the mass density ρ and the absolute pressure p , or the gas constant R
and the absolute temperature T .
The critical value of the throat pressure p2 , is the lowest value that can
exist for any given admission pressure p1 . The pressure p2 will equal p3
whenever p3 is equal to or greater than p2 cr . For lower p3 values, p2 will
equal p2 cr and, therefore, will be higher than p3 . In the first case, the
maximum possible velocity can be achieved with a purely converging nozzle.
In the second case, an increase in velocity can be achieved beyond the throat
by adding a diverging section. In fact, this is the only type of nozzle that will
produce supersonic velocities.

© 1999 Howden Buffalo, Inc.

You might also like