Anda di halaman 1dari 25

4

BASIC LAWS FOR A CONTROL VOLUME IN INTEGRAL FORM

4.1 Introduction

The basic laws of nature, which apply to all systems, are:

● the law of conservation of mass;


● Newton’s second law of motion for (a) linear motion and (b) angular motion;
● the first law of thermodynamics, and
● the second law of thermodynamics.

The last two are used extensively in the analysis of thermal plant and would normally be covered
in a course on Thermodynamics. In analyzing fluid flow problems, they become important when
heat transfer and/or friction are present. In Fluid Mechanics we shall for the most part deal with
problems without heat transfer.

Our aim in this chapter is to derive the control volume forms of the law of mass conservation
and Newton’s second laws for linear and angular motion and to illustrate their use through
example problems.

4.2 System forms of the basic laws

Before we express the laws into forms suitable for the flow of a fluid through a control volume,
we shall write down the equations expressing them for a system. This will also serve the purpose
of introducing the notation required for each law.

4.2.1 Law of conservation of mass

If m is the total mass of a system, then the law of conservation of mass simply states that
m = constant (4.1)
In deriving the control volume form of the equation, we shall need to divide the system into two
or more subsystems. To this end, consider a system consisting of a number of elemental
subsystems such as that shown in Figure 4.1 below.

total volume, V
elemental subsystem
total mass, m
dm=ρdV

system boundary

Figure 4.1

1
The mass dm of an elemental subsystem is the density ρ of that subsystem multiplied by its
volume dV. Since mass is an extensive property1, the total mass m of the whole system is the
sum (integral) of the elemental masses:

m = ∫ dm = ∫ ρdV (4.2)
m V

The first integral is over the mass of the whole system and the second is over the volume of the
whole system. Note that the density may vary from point to point within the system boundary.

In order to express equation (4.1) for a control volume it is necessary first to write that equation
as a time-rate equation. Since the system mass is a constant, this is simply

 dm 
0=  (4.3)
 dt 
We shall find that the other laws of nature can also be expressed in a similar form.

4.2.2 Newton’s second law for linear motion

The linear momentum dΦ of an elemental subsystem is a vector quantity defined as its velocity
U multiplied by its mass dm, see Figure 4.2 which also shows the force dF acting on the
subsystem.

dm
U

dF

Figure 4.2

Since linear momentum is an extensive property, the total linear momentum Φ of the whole
system is given by

Φ = ∫ Udm = ∫ UρdV (4.4)


m V

Both the velocity and the density may vary from point to point within the system boundary.

Newton’s second law of motion states that the rate of change of linear momentum of the system
is equal to the resultant F of the forces acting on all the subsystems. In equation form the second
law states

1
An extensive property is one whose total value for a system is obtained by adding together the values of the
corresponding property for all of its subsystems. Two other examples of extensive properties are energy and
entropy. Temperature, pressure and density are not extensive properties.
2
 dΦ 
F =  (4.5)
 dt 
The resultant force consists of two types of force: (a) a surface force FS which includes forces
normal to the surface due to pressure and forces parallel to the surface due to shear stresses, and
(b) a body force FB due to gravity or some other force field (e.g. magnetic). Thus

F = FS + FB (4.6)
In most applications of Newton’s second law, we are interested in the force or the rate of change
of momentum in a particular direction, say the x or the y-direction. It is then more convenient to
express the equations in component form. Thus, for example, the x-component of equation (4.5)
would be
 dΦ x 
Fx =   (4.7)
 dt 
where Fx and Φx are the resultant force and momentum in the x-direction. The x-direction
momentum is obtained by replacing the vector velocity U in equation (4.4) with the x-direction
velocity u. Thus
Φ x = ∫ udm = ∫ uρ dV (4.8)
m V

and the resultant force is the sum of the surface and body forces in the x-direction

Fx = FSx + FBx (4.9)


Similar equations apply to other directions. In the y-direction, for example, the subscripts x in
the last three equations would be replaced by y and the x-direction velocity u in equation (4.8)
would be replaced by the y-direction velocity v.

4.2.3 Newton’s second law for angular motion

The corresponding form of the equation for angular motion is not a separate axiom. It is just a
corollary of the law for linear motion. To write it down as a rate equation, we first need to define
what we mean by ‘angular momentum’. For this purpose, consider an axis through some point O
which may be inside or outside the system boundary, see Figure 4.3. In general the system will
not be in a single plane and the axis through O may be at any angle. However, for ease of
presentation, it is helpful to think of the system as being a plane lamina and the axis through O
as being perpendicular to the plane of the figure. The position vector of the elemental mass with
respect to O as the origin is r.

O
r Usinθ
rsinθ U
θ
dm

Figure 4.3
3
The angular momentum dΨ of the element about the axis through O is defined as the vector
product r × U times the mass dm of the element. By definition of a vector product, r × U is
perpendicular to the plane containing r and U and its ‘sense’ is determined by the right hand
rule. For the case shown in the diagram above, it is represented by an arrow coming out of the
paper. Note that the order in which r and U appear in the vector product is important. The vector
U × r would have the same magnitude as that of r × U and would be parallel to it but its sense
would be into the paper.

If n is a unit vector in the perpendicular direction, and if θ is the angle between the vectors r and
U, then by definition

dΨ = r × U dm = n rUsinθ dm (4.10)

where r and U are the magnitudes of the vectors r and U. It is seen that rsinθ is the perpendicular
distance from O to the line of action of the velocity vector U so that rsinθ times Udm is the
moment of the linear momentum of the elemental mass about O. Therefore, a more appropriate
name for ‘angular momentum’ is moment of momentum. Alternatively, since Usinθ is the
component of velocity perpendicular to the radius vector, then r times Usinθdm is again the
moment of momentum of the elemental mass about O. If we denote U sin θ by U ⊥ then
equation (4.10) becomes

dΨ = r × U dm = n rU ⊥ dm (4.11)

Like linear momentum angular momentum is an extensive property so that its total value Ψ for
the whole system is obtained by integration:

Ψ = ∫ r × Udm = ∫ r × UρdV (4.12)


m V

or
Ψ = ∫ nrU ⊥ dm = ∫ nrU ⊥ ρdV (4.13)
m V

According to Newton’s second law for angular motion, the time-rate of change of angular
momentum is equal to the resultant torque T about the axis through O. The torque is clearly a
vector quantity and its magnitude is the moment of all the forces acting on the system, about O:

 dΨ 
T =  (4.14)
 dt 
This is the vector form of the equation. However, as for the second law of linear motion, we
usually need to apply it about a particular axis, say an axis z through O in Figure 4.3. The z-
component of equation (4.14) is then

4
 dΨ z 
Tz =   (4.15)
 dt 

where Tz and Ψz are components of the torque and angular momentum about the axis Oz. Ψz is a
scalar quantity given by

Ψ z = ∫ rU ⊥ dm = ∫ rU ⊥ ρdV (4.16)
m V

In comparison with equation (4.13) the radius r in this integral is the radius of the elemental
mass from the axis Oz and U ⊥ is the velocity perpendicular to that radius.

The resultant torque in equation (4.14) or (4.15) can consist of contributions from surface and
body forces. The figure below shows that the resultant surface force FS acting through some
point S whose position vector with respect to O is RS such that the angle between the force and
the position vector is α.

O FS
RS
α
RSsinα
S

Figure 4.4

The magnitude of the torque TS due to this force about O is the magnitude of the force multiplied
by the perpendicular distance from O to the line of action of the force. It can also be expressed as
a vector product of RS and FS:

TS = RS × FS = n FS RS sin α (4.17)

where n is a unit vector perpendicular to the plane containing the vectors FS and RS. FS and RS
are the magnitudes of the respective vectors and RS sinα is the perpendicular distance from O to
the line of action of the force.

A similar equation applies to the torque due to gravity forces. The gravity force on an elemental
mass of the system is gdm where the vector g is the specific2 gravitational force. The torque due
to this force about the axis through O is r × gdm , see Figure 4.5. If β is the angle between r and
g, then
r × gdm = n gdm r sin β (4.18)

2
The term specific is used to describe a quantity divided by the mass of the system. Thus the specific
gravitational force is the gravitational force divided by the mass of the system.
5
O rsinβ
r
β

gdm
Figure 4.5

which means that the torque is the gravity force gdm times the perpendicular distance rsinβ
from O to the line of action of this force. Here g and r are the magnitudes of the respective
vectors. The total torque due to gravitation\al forces is TG which is obtained by integrating over
the mass or volume of the system

TG = ∫ r × gdm = ∫ r × gρdV (4.19)


m V

This expression is suitable when the system consists of a number of subsystems of different
masses at various positions. However, if the centre of mass of the system is at a point G whose
position vector with respect to O is RG, then the total torque due to gravity forces can be
simplified to

TG = RG × g m (4.20)

where m is the total mass of the system.

Some books include a third type of torque, Tshaft, which is the torque applied by a shaft
interacting across the system boundary. However, in essence this is really a torque arising from
surface shear forces constituting a couple and it should strictly be included in the surface force
component TS already discussed. Thus we shall write

T = TS + TG = RS × FS + RG × g m (4.21)

Again, in applications of the second law for angular motion, we are likely to need only one
component of the torques due to surface and body forces.

4.2.4 The first law of thermodynamics

Suppose that a system undergoes an infinitesimal process by which its total energy (an extensive
property) increases by an amount dE while an amount of heating δQ is done to it by the

6
surroundings and an amount of work δW is done by it on the surroundings3. The first law of
thermodynamics relates these three quantities by the equation

δQ − δW = dE (4.22)
Dividing by the time interval δt of the process and letting δt → 0 leads to the time-rate form of
the first law equation:
 dE 
Q& − W& =   (4.23)
 dt 
where Q& and W& are the instantaneous rates of heating and working respectively and (dE / dt ) is
the instantaneous rate of change of energy of the system. The total energy of the system is given
by
E = ∫ edm = ∫ eρdV (4.24)
m V

Here e is the specific energy (i.e. energy per mass). e consists of three components: the specific
internal energy4 (u), the specific kinetic energy (U2/2) and the specific potential energy (gz).
Thus
U2
e=u+ + gz (4.25)
2
Here U is the resultant velocity and z is the elevation of the centre of mass of the system above a
horizontal datum plane.

4.2.5 The second law of thermodynamics

There are several equivalent ways of stating the second law of thermodynamics for a system.
Here we shall use the one which relates the change in an extensive property, called the entropy S,
of the system, Figure 4.6. When a system at thermodynamic (absolute) temperature T receives an
amount of heating δQ, the increase in its entropy dS is greater than or equal to δQ/T. Thus

δQ
T
dm
s

Figure 4.6

3
Note that we use δQ and δW for the amounts of heating and work which are interactions across the system
boundary. They depend upon the type of process and hence are not exact differentials. On the other hand, the
change in energy dE is an exact differential whose value depends only on the end states and not on the type of
process.
4
Do not confuse the specific internal energy u with the x-component of velocity which is also denoted by u.
Fortunately we shall not be making extensive use of the first law in this course so that the specific internal
energy and x-direction velocity will rarely appear together.
7
δQ
≤ dS (4.26)
T
The equality sign applies in the ideal case of a reversible process. As for the first law, we divide
by the time duration δt of the process and let δt → 0 to express this as a time-rate equation
Q&  dS 
≤  (4.27)
T  dt 
If s is the specific entropy (entropy/mass) of the elemental mass dm, the total system entropy is
evaluated by integration:

S = ∫ sdm = ∫ sρdV (4.28)


m V

4.2.6 Summary of system forms of the basic equations

Thus far we have obtained the equations for the basic laws in terms of changes in one or other of
the following extensive properties: m, Φ, Ψ, E or S. It is helpful to list them in a table as follows:

Law Time-rate equation Extensive property

 dm  m = ∫ dm
Conservation of mass 0= 
 dt  m

 dΦ  Φ = ∫ Udm
Second law of linear motion F = 
 dt  m

 dΨ  Ψ = ∫ r × U dm
Second law of angular motion T = 
 dt  m

 dE  E = ∫ edm
First law of thermodynamics Q& − W& =  
 dt  m

Q&  dS  S = ∫ s dm
Second law of thermodynamics ≤ 
T  dt  m

Before expressing the equations into forms suitable for the flow through a control volume, note
that all five equations can be expressed in the general form

 dN 
I =  (4.29)
 dt 

where the left hand side I is value of the particular interaction across the system boundary and is
the cause of a change in the generic extensive property N. For example, in the conservation of
mass equation, N is the mass m of the system and there is no cause for the mass to change so that
in this case I = 0. In the second law for linear motion, I is the force F which causes the linear
momentum Φ to change. Secondly, note that each of the extensive properties can be expressed
in the form

8
N = ∫ η dm (4.30)
m

where η is the appropriate extensive property divided by mass. For example if N = m then η = 1
whereas if N is the linear momentum Φ, then η is the vector velocity U (=mU/m). Here U is, in
fact, the momentum per mass i.e. the specific linear momentum.

4.3 Inertial and non-inertial control volumes

We shall consider a control volume of fixed dimensions. Before we adapt the system forms of
the laws to the flow through the control volume, we need to recognize that, in some applications,
the control volume may itself be moving. Figure 4.7(a) below shows a control volume that is at
rest and Figure 4.7(b) shows one that is moving with a uniform rectilinear velocity UC relative to
a fixed frame XYZ. The frame xyz is ‘attached’ to the control volume such that in case (a) it is at
rest whereas in case (b) it is moving at the same velocity as the control volume.

UC
y y
Y Y
CV CV
x x
z z

X X

Z Z
(a) stationary CV (b) moving CV

Figure 4.7 - Inertial control volumes

The word ‘uniform’ is used to imply ‘steady’ or ‘non-accelerating’ relative to the fixed frame,
and the word ‘rectilinear’ is used to mean ‘in a straight line’. The control volumes described here
are called inertial control volumes. There are numerous practical instances of stationary control
volumes. In a civil power station, for example, the turbine, the pump, the condenser, the boiler
and the superheater are all stationary and hence inertial control volumes. The jet engine of an
aircraft moving at a steady speed in a straight line is an example of a moving inertial control
volume.

In contrast, a non-inertial control volume will not only be moving at some velocity, it will also
have an acceleration aR relative to the fixed frame XYZ. One common example is an
accelerating car. Air is drawn into the engine and exhaust gases flow out of it while the car
speeds up. An example of a rotating non-inertial control volume is the domestic garden
sprinkler. When it starts from rest, the sprinkler has a radial acceleration towards the axis of
rotation as well as an angular acceleration. Even when it is rotating at steady speed, it still has a
radially inward acceleration.

9
Throughout this course, we shall be dealing with inertial control volumes only. For a discussion
of the method of applying the laws to non-inertial control volumes see, for example, Fox et al
(2009)5.

4.3.1 The general form of the equations for a control volume

Consider fluid flowing through an inertial control volume (CV) of fixed dimensions as shown in
Figure 4.8. The boundary of the control volume is the control surface (CS) shown as a closed
dashed line. The control volume and the coordinate frame xyz may be moving with a uniform
velocity relative to a fixed frame XYZ (not shown). We consider the fluid motion relative to the
control volume.

Next we mark a fluid system whose boundary at time t coincides exactly with the CS. After an
interval ∆t, the system moves on to a new position shown on the right of Figure 4.8 and, as it
does so, it may also change its shape.

CS subregion 3
CV A of region III

subregion 1 III
of region I B
II

I D C
y
system boundary system boundary
time t time t+∆t

z x

Figure 4.8

By definition, the instantaneous time-rate of change of an extensive property N of the system is

 dN  N − Nt
  = lim t + ∆t (4.31)
 dt  ∆ t → 0 ∆t

where N t and N t + ∆t are the values of N at instants t and t+∆t. Referring to the sketch on the left,
it is seen that N t = ( N CV ) t since at time t the system occupies the same space as the control
volume. The sketch on the right shows that, at time t+∆t, the system consists of two subregions
II and III. Subregion II is still contained within the CV whereas subregion III is the part of the
system that has ‘gone out’ of the CV. It is also seen that subregion II is the whole of the control
volume minus subregion I which was inside the CV at time t. If we denote the value of N for
5
Introduction to Fluid Mechanics by Fox, R. W., Pritchard, P. J. and McDonald, A. T., Wiley, Seventh Edition,
2009.
10
subregions I and III by ∆N I and ∆N III respectively, then
N t + ∆t = ( N II + ∆N III ) t + ∆t = ( N CV − ∆N I + ∆N III ) t + ∆t . The above equation now becomes

 dN  ( N − ∆N I + ∆N III ) t + ∆t − ( N CV ) t
  = lim CV (4.32)
 dt  ∆t → 0 ∆t
which can be rearranged as

 dN  (N ) − ( N CV ) t (∆N III ) t + ∆t (∆N I ) t + ∆t


  = lim CV t + ∆t + lim − lim (4.33)
 dt  ∆t →0 ∆t ∆t →0 ∆t ∆t → 0 ∆t
1 2 3

The first term in this equation is, by definition, the time-rate of increase of the extensive property
N of the contents of the control volume. Thus

( N CV ) t + ∆t − ( N CV ) t0  ∂N 
1 = lim =  (4.34)
∆t → 0 ∆t  ∂t  CV

The second term is the time-rate at which the property N is going out of the control volume
through the area ABC of the control surface and the third is the rate at which it is coming into
the control volume through the area CDA of the control surface. It is appropriate to call these
terms N& out and N& in respectively:

(∆N III ) t + ∆t
2 = lim = N& out (4.35)
∆t → 0 ∆t

(∆N I ) t + ∆t
3 = lim = N& in (4.36)
∆t → 0 ∆t

Use of the last three definitions in equation (4.37) and of the general form of the system equation
(4.29) gives

 dN   ∂N 
I = =  + N& out − N& in (4.37)
 dt   ∂t  CV

Since the system boundary coincides with the control surface as ∆t → 0 , I is now the
instantaneous interaction across the control surface rather than across the system boundary.
Cleary N& out − N& in is the net rate at which the property N is flowing out of the control volume. In
words the equation states that: the amount of interaction across the control surface equals the
time-rate of increase of N of the mass inside the CV plus the net rate at which N is leaving the
CV.

11
In any particular problem we would need to evaluate the quantities (∂N / ∂t ) CV , N& out and N& in .
Consider the first of these. If dm=ρdV is the mass of an element inside the CV and η is the value
of N per mass over that element, then the value of N for the element is ηdm=ηρdV. The total
value for the whole control volume is obtained by integrating over the mass or the volume of the
control volume:

N CV = ∫ηdm = ∫ηρdV
mCV CV
(4.38)

where mCV is the mass and CV is the volume of the control volume. The time-rate of increase of
NCV becomes:

 ∂N  ∂
  = ∫ηρdV
 ∂t  CV ∂t CV
(4.39)

Next, in order to evaluate N& out , consider a subregion 3 of region III in Figure 4.8. An enlarged
view of this is shown below.
dA

α
U
α ∆l
∆h
dA

Figure 4.9 - Subregion 3 of region III

The value of the property N for this element is given by dN = ηdm = ηρdV where ρ is the
density and η is the specific value of N for the element with mass dm and volume dV. The
volume dV is the base area dA times the height ∆h perpendicular to the area. As seen in the
figure, ∆h = ∆l cos α . Here α is the angle between the velocity vector U and the area vector dA
and ∆l is the length of the element in the direction of flow, given by the scalar velocity U
multiplied by the time interval ∆t, i.e. ∆l = U∆t . Thus dV = dA∆h = dA∆l cos α = UdA cos α ∆t .
Now the product UdA cos α is just the scalar (dot) product of the vectors U and dA so that
dV = U ⋅ dA ∆t . Thus dN = ηρU ⋅ dA ∆t and the rate at which the property N is flowing out
through area dA is obtained by dividing by the time interval ∆t
∴ dN& out = ηρU ⋅ dA (4.40)
&
The total outflow rate N out is obtained by integration over the (exit) area AABC of the control
surface, see Figure 4.8:
N& out = ∫ηρU ⋅ dA
AABC
(4.41)

The inflow rate is obtained in a similar manner but in this case we must consider a subregion 1
of region I in Figure 4.8. An enlarged view of this is shown below. In this case the angle α

12
between the velocity vector and the area vector is obtuse6 so that the element height is
∆h = ∆l cos(180 o − α ) = −∆l cos α .

180-α
dA ∆h U

α
∆l

Figure 4.10 Subregion 1 of region I

Following a procedure similar to that above, the volume of this subregion is dV = −U ⋅ dA ∆t and
the value of the property N for this subregion is dN = −ηρU ⋅ dA ∆t . The rate at which property
N flows into the CV through the area dA is again obtained by dividing this by ∆t:

∴ dN& in = −ηρU ⋅ dA (4.42)

The total inflow rate N& in is obtained by integrating with respect to the area over the (inlet) area
ACDA of the control surface, see Figure 4.8:

N& in = − ∫ ηρU ⋅ dA (4.43)


ACDA

Finally substituting the expressions for ∂N CV / ∂t , N& out and N& in into equation (4.37), we obtain

 dN  ∂
I = = ∫ηρdV + ABC
 dt  ∂t CV
∫ηρU ⋅ dA + CDA
∫ηρU ⋅ dA (4.44)

Note that the sign between the last two integrals on the right is positive because the expression
for N& has a negative sign in front of the integral.
in

Referring again to the right of Figure 4.8, it is seen that the areas AABC and ACDA in the second
and third integrals on the right of (4.44) make up the total area of the control surface so that
these integrals can be combined into a single integral over the entire control surface:

 dN  ∂
I = = ∫ηρdV + CS∫ηρU ⋅ dA
 dt  ∂t CV
(4.45)

where CS denotes the total area of the control surface.

While this vector equation is elegant in form and was the object of this section, the earlier form
of the equation (4.37) is more useful in appreciating the physical meanings of the terms
involved.

6
The positive direction of the area vector is such that it points out of the control surface.
13
4.3.2 An important reminder about the velocity U in equation (4.45)

Armed with equation (4.45) we are ready to obtain the equations for the basic laws for a control
volume. Before doing that, however, we should remind ourselves that in arriving at the equation,
we considered fluid motion relative to the control volume. The velocity U which accounts for the
mass flux ρU ⋅ dA through an element of area is thus the fluid velocity relative to the control
volume.

In many problems we shall indeed consider a control volume that is fixed so that the velocity U
is the absolute velocity. When we occasionally need to deal with problems in which the control
volume is moving at constant velocity UC, then the fluid velocity relative to the control volume
is the absolute fluid velocity U minus UC. If we denote the latter by Urel, then

U rel = U − U C (4.46)

In this case, the velocity which accounts for the mass flux term is Urel so that equation (4.43)
becomes

 dN  ∂
I = = ∫ηρdV + CS∫ηρU rel ⋅ dA
 dt  ∂t CV
(4.47)

While this equation is equally applicable to a control volume that is at rest, it is cumbersome to
carry the subscript “rel” on U. Therefore we shall only use the last equation to deal with
problems in which the control volume is actually moving.

4.3.3 Conservation of mass for flow through a control volume

The system form of the equation for this is equation (4.3) where the interaction I is zero. In this
case the extensive property N in equation (4.37) is the mass m and the specific property η is
simply the number 1(=m/m). Equation (4.37) becomes

 dm   ∂m 
0= =  + m& out − m& in (4.48)
 dt   ∂t  CV
which can also be written

 ∂m 
  = m& in − m& out . (4.49)
 ∂t  CV

The latter form of the equation is easily remembered since it makes the common sense statement
that the rate of increase of mass within the CV is the net rate at which mass is coming into the
CV.

An alternative integral form of equation (4.48) can be obtained by putting N = m and η = 1 in


equation (4.45). Thus
14
 dm  ∂
0= = ∫ ρdV + CS∫ ρU ⋅ dA
 dt  ∂t CV
(4.50)

Equation (4.48) or (4.50) are alternative ways of expressing the conservation of mass for the
flow through a control volume. They are called the equation of continuity.

The second integral on the right of (4.50) is the net mass outflow rate through the entire control
surface. When the integral is taken over a particular section of area A, it represents the mass flow
rate through that area and is denoted by m& .

m& = ∫ ρU ⋅ dA (4.51)
A

Equations (4.48) and (4.50) are in a general form in which the flow may be steady or unsteady
and the velocity and density may vary with position. In the following we consider particular
cases where simplifications can be made.

Steady flow

If the flow is steady, then by definition, there is no time dependence ( ∂ / ∂t ≡ 0 ) so that equations
(4.48) and (4.50.) simplify to
0 = ∫ ρU ⋅ dA = m& out − m& in (4.52)
CS

As expected in this case, the mass flow rate into the CV equals the mass flow rate out of it.

Uniform flow

If the density and velocity are uniform across a flow cross-section, then ρ and U can be ‘rolled
across’ the integral in equation (4.51) to obtain

m& = ρU ⋅ A (4.53)
If the area is chosen such that it is perpendicular to the velocity vector, then the expression for
the mass flow rate is simplified to

m& = ρUA (4.54)

where U and A are the magnitudes of the velocity and the area vectors. This equation is used in
most thermodynamics problems where the above assumptions are valid to a reasonable
approximation.

Incompressible flow

If density changes are negligible (as in most liquids and gases for which changes in temperature
and pressure are not too large), then the density in equation (4.50) can be cancelled through to
obtain

15
 ∂V 
0=  + ∫ U ⋅ dA = ∫ U ⋅ dA
 ∂t  CV CS CS

Note that the time derivative term is zero since, at the outset, we have restricted consideration to
a control volume of fixed dimensions. The integral on the right, without the density ρ, is now the
net volume flow rate out of the control surface V&out − V&in instead of the net mass flow rate
m& out − m& in . The above equation may now be written as


0 = U ⋅ dA = V& − V&
CS
out in (4.55)

so that the volume flow rate V&in into the CV equals the volume flow rate V&out out of it, an
expected result.

When the integral in (4.55) is taken over a particular area A of the control surface, it represents
the volume flow rate through that area and is denoted by the symbol V&

∫ U ⋅ dA = V& .
A
(4.56)

4.3.4 Choice of control surface, mean velocity and uniform velocity

We shall find that it is usually best to choose the flow cross-sections of a control surface such
that the velocity across any section is perpendicular to the area, as shown in Figure 4.11. In this
case the area vector and the velocity vector are parallel to each other so that the mass flow rate
equation (4.51) becomes

m& = ∫ ρU ⋅ dA = ∫ ρUdA (4.57)


A A

where U and dA are now the magnitudes of the vectors U and dA.

In general, the density and velocity will vary across the cross-section. The variation in density
across the section is usually negligible. As for the velocity, it is often convenient and also
accurate enough to work in terms of an average (mean) velocity Um across the section, see
Figure 4.11. In that case, equation (4.57) becomes

m& = ρU m A (4.58)

CS A CS A U
Um

Figure 4.11 – Definition of mean velocity Figure 4.12 – Uniform velocity


16
If the velocity U is indeed uniform across a section, as in Figure 4.12, then clearly U m = U .

4.3.5 Significance of the sign of the integral ∫ ρ U ⋅ dA


A

The sign of the numerical value of this integral will depend upon whether the angle α between
the velocity vector U and the area vector dA is acute or obtuse, see Figure 4.13.

U
dA

α U

dA

(a) Inflow : ∫ ρU ⋅ dA < 0


A
(b) Outflow : ∫ ρU ⋅ dA > 0
A

Figure 4.13

In case (a) where α is obtuse, cos α < 0 so that the integral will be negative, indicating flow into
the control volume. In case (b) with an acute angle α the flow is out of the control volume.

At this stage it would be helpful to illustrate the use of the equations derived so far with some
examples.

Example 4.1

Fluid of constant density flows through a device with three flow ports, 1, 2 and 3 as shown in the
diagram below.

A2
y
A1
A3
50o

The cross-sectional areas are A1 = 0.05 m2, A2 = 0.01 m2 and A3 = 0.06 m2. The uniform
velocities at ports 1 and 2 are U1 = - 4i m/s and U2 = -8j m/s where i and j are the unit vectors in
the coordinate directions shown.

Calculate the volume flow rate and the vector velocity at port 3.
_______________________________________

17
Example 4.2

A porous plate of length 2 m is held parallel to a uniform stream of water as shown below. The
plate width perpendicular to the plane of the diagram is 1.5 m.

A B
4 m/s 4 m/s
4 m/s 1.2 mm
δ

0 x
0.3 m/s

The steady water velocity parallel to the plate is 4 m/s. At the surface of the plate, water is
sucked away through the plate such that the uniform velocity across the plate surface is 0.3 m/s.
At the leading edge of the plate (x = 0) the x-wise velocity is uniform at 4 m/s.

Owing to the effect of viscosity a ‘boundary layer’ develops over the plate such that, at any
transverse section, the x-wise velocity u varies with distance y from the plate according to the
equation
πy
u = u ∞ sin (0 ≤ y ≤ δ )

where δ is the ‘thickness’ of the boundary layer and u ∞ = 4 m/s is the x-wise velocity at y=δ.
Measurements showed that at the end of the plate δ = 1.2 mm.

Choose a suitable control volume and calculate the mass flow rate through the area represented
by the line AB in the diagram. State whether this flow is upward or downward.
_______________________________________

Example 4.3 – unsteady flow

A circular tank of internal diameter 0.8 m discharges water through a pipe of internal diameter
25 mm at its lower end. Calculate the rate at which the height of the water level is changing at
the instant when the water velocity at exit from the pipe is 2 m/s. Assume that the velocity
distribution across the pipe is uniform.
_______________________________________

4.3.6 Newton’s second law of linear motion – inertial control volume

The system form of the equation for this law in the x-direction is equation (4.7). Here the
interaction I is the resultant force Fx on the system and the extensive property N is the linear
momentum Φx. For this case equation (4.37) becomes

18
 dΦ x   ∂Φ x 
Fx =  =  + Φ& x ,out − Φ& x,in (4.59)
 d t   ∂t  CV

In this equation Fx represents the instantaneous net x-wise force on the contents of the control
volume rather than on any system. In words, the equation states that, at any instant, the net x-
wise force on the contents of the CV equals the time-rate of increase of x-wise momentum of the
contents of the CV (= ∂Φ x / ∂t) CV plus the net rate of x-momentum flowing out of the CV
(= Φ& − Φ& ) . We shall refer to this equation as the momentum equation in the x-direction.
x ,out x ,in

Similar equations can be written for other directions.

An alternative form of the momentum equation is obtained by using the generic equation (4.45)
with the specific property η replaced by the momentum per mass in the x-direction. The latter is
just the x-wise velocity u. Thus

 dΦ x  ∂
Fx = FSx + FBx =  = ∫ uρdV + CS∫ u ρU ⋅ dA .
 dt  ∂t CV
(4.60)

Here we have used equation (4.9) to write the force Fx as the sum of surface and body forces FSx
and FBx. Note that the quantity ρU ⋅ dA in the second integral is the same as in equation (4.45)
and is the mass flow rate through the area element.

We should remind ourselves that the fluid velocity in the above equation is the velocity relative
to the inertial frame xyz which is attached to the control volume. If the latter is at rest relative to
the fixed reference system XYZ, then it is simply the absolute velocity of the fluid. If it is
moving, then it is advisable to use a subscript “rel” on the symbol for velocity to emphasize that
it is the fluid velocity relative to the moving frame. Equation (4.60) then becomes

 dΦ x  ∂
Fx = FSx + FBx =  = ∫ u rel ρdV + CS∫ u rel ρU rel ⋅ dA
 dt  ∂t CV
(4.61)

The momentum equation derived in this section applies to the flow through an inertial control
volume. The extension to non-inertial control volumes is not difficult. The only point to
remember is that in such cases, the velocity in the definition of linear momentum [equation
(4.4)] must be the velocity relative to the fixed frame, i.e. the absolute velocity. Thus, if Urel is
the fluid velocity relative to the moving CV and UC is the velocity of the CV relative to the fixed
frame then U = U rel + U C . Since the velocity of a non-inertial control volume is not constant,
then an additional integral taking account of the acceleration of the control volume will appear
on the right hand side of equation (4.61)7.

7
For further details, see Introduction to Fluid Mechanics by Fox, R. W., Pritchard, P. J. and McDonald, A. T.,
Wiley, Seventh Edition, 2009.
19
The term ∫ u ρU ⋅ dA in equation (4.60) is the net rate at which x-wise momentum is flowing out
CS

through the entire control surface. If the integral is taken over a particular area, then the integral
represents the rate of flow of x-wise momentum through that area

Φ& x = ∫ u ρU ⋅ dA . (4.62)
A

If it can be assumed that the x-wise velocity is uniform over the cross-section, then u can be
rolled across the integral sign so that

Φ& x = u ∫ ρU ⋅ dA = um& (4.63)


A

since ∫ ρU ⋅ dA is just the mass flow rate through the area, m& [see equation (4.51)].
A

4.3.7 Some advice on the application of the momentum equation

A clearly labelled diagram showing the control surface is absolutely essential in any application
of the momentum equation. A dashed line is suggested for this purpose to distinguish the
contents of the control volume from the surroundings. At each flow cross-section, it is best to
draw the control surface such that it is perpendicular to the flow velocity in which case the mass
flow rate through that section is just the product of density, area and velocity. Other parts of the
control surface are usually dictated by the force that is to be calculated. In some cases more than
one choice will lead to the required answer but a little prior thought can lead to a simpler and
quicker analysis. These ideas will now be elaborated in the following examples.

Example 4.4
V
b
A A a

g section AA

The diagram above shows a vertically downward stream of water at uniform velocity V and
having a rectangular cross-section a×b.

The stream is deflected through an angle θ by a circular cylinder with its axis horizontal. The
cross-section of the stream is unchanged.

20
Calculate the horizontal component of the force exerted by the water on the cylinder if V=3 m/s,
a = 12.5 mm and b = 2.5 mm and θ = 20o.
______________________________________________

Example 4.5
The diagram below shows a vertical section through a horizontal channel in which water flows
through a sluice gate. The water depth and velocity upstream of the gate are 3 m and 1 m/s
respectively. Downstream of the gate, the water depth is 0.429 m. Assume that the velocities are
uniformly distributed across each section and consider unit width perpendicular to the plane of
the diagram.

Calculate

(a) the velocity at the downstream section,


(b) the horizontal force exerted by the water on the gate.

Compare the force in part (b) with the force exerted when the water depth on the upstream face
is 3 m and the gate is closed.

sluice gate

g
1 m/s 3m 0.429 m

Since it is required to calculate the force exerted by the water on the gate, it would seem best to
exclude the gate from the control surface. However, using the fact that the force required to keep
the gate in place must be equal and opposite to the force that is exerted by the water on the gate,
it is possible to define a simpler control surface which includes the gate.
______________________________________________

Example 4.6

A horizontal jet of water issues from a nozzle at a steady velocity of 30 m/s which is uniform
across the cross-section of area 0.01 m2. The nozzle is placed coaxially in a larger circular pipe
of area 0.075 m2. In the annular area between the nozzle and the pipe there is a secondary water
stream of uniform velocity 3 m/s. At some distance downstream, the jet stream and the
secondary stream are thoroughly mixed and exit the pipe at a uniform velocity across the whole

21
pipe cross-section. The pressures at the jet inlet section and at pipe exit are uniform across the
cross-sections.

Calculate
(a) the water velocity at pipe exit and
(b) the change in pressure from inlet to exit.

Suggest a possible application for such a device.


______________________________________________

Example 4.7

Water flows steadily in a horizontal supply duct with a square cross-section of side 75 mm. The
flow is turned vertically upward by a right angled bend as shown in the diagram below.

g y
vmax
vmin
supply duct B
x

A
At the inlet of the bend (A), the gauge pressure is 84 kPa and the horizontal velocity is 8 m/s,
both uniformly distributed across the cross-section. At the exit (B), the velocity varies linearly
with the coordinate x shown from a maximum value vmax to a minimum vmin such that vmax=2vmin.
It does not vary in the direction perpendicular to the plane of the diagram. The pressure at exit is
atmospheric. The mass of the bend is 2 kg and its internal volume is 0.004 m3.

Calculate the horizontal and vertical components of the force exerted by the bend on the supply
duct.
______________________________________________

The next two examples deal with inertial control volumes that are moving at uniform velocity.

Example 4.8

The diagram shows a horizontal plane jet of water with a rectangular cross-section of 15.5 mm ×
5 mm striking a curved splitter plate moving horizontally at a steady speed of 10 m/s.The splitter
divides the jet into two streams of rectangular cross-section. One stream, at volume flow rate Q2

22
exits vertically upward and the other at volume flow rate Q3 exits at an angle of 30o to the
horizontal.

Q2

g A 10 m/s
25 m/s
5 mm A 30o Q3

15.5 mm

section AA

Neglect gravity forces and assume that there is no friction on the splitter surface so that the
magnitude of the water velocity relative to the splitter remains unchanged. Calculate the ratio
Q2/Q3 for which there is no force on the splitter in the vertical direction. For this flow ratio,
calculate the horizontal force acting on the splitter which ensures that its speed is constant.
______________________________________________

Example 4.9

A continuous jet of water issues at a steady velocity of 86.6 m/s from a nozzle of diameter 50
mm and strikes a series of identical curved vanes moving at a steady velocity of 50 m/s, as
shown in the diagram below.

The geometry of the vanes is as shown. Assume that (i) all the flow that leaves the nozzle
crosses the vanes, (ii) there is no friction on the vane so that the magnitude of the water velocity
relative to the vane at its exit is the same as it is at inlet and (iii) gravity forces are negligible.

45o

50 m/s

86.6 m/s
30o

23
Calculate:
(a) the nozzle angle α required to ensure that the jet strikes the vane tangentially at its leading
edge;
(b) the force that must be applied to the vane to maintain a constant speed, and
(c) the power output from the vane/nozzle system.
_____________________________________________

4.7 Newton’s second law of angular motion for an inertial control volume

The z-component of the system form of the equation for angular motion is equation (4.15) where
the interaction I is the torque Tz on the system about the axis Oz and Ψz is the angular
momentum about the same axis. Using equation (4.37) with I and N replaced by Tz and Ψz
respectively gives the control volume form of the angular momentum equation:

 dψ z   ∂Ψ z  &
Tz =  =  +Ψ z ,out −Ψ&z ,in (4.64)
 dt   ∂t 

or, using equation (4.45) with I and N replaced as above and with η replaced by rU ⊥ which is
the specific angular momentum about the z-axis, we obtain the alternative form

 dΨ z  ∂
Tz =  = ∫ rU ⊥ ρdV + CS∫ rU ⊥ ρU ⋅ dA
 dt  ∂t CV
(4.65)

In the above equations Tz is the torque on the contents of the control volume. In words, the
equations state that, about the axis Oz, the net torque on the control volume is the time-rate of
change of angular momentum of the contents of the CV ( = (∂Ψ z / ∂t) CV plus the net rate of flow
of angular momentum out of the CV (= Ψ& −Ψ& ) .
z , out z ,in

The angular momentum equation is essential for the analysis of rotary machines handling fluids,
e.g. centrifugal pumps and turbines. Rotodynamic machines are not studied in this course. We
shall nevertheless illustrate the use of the equation in two examples. The first is one in which it is
necessary to apply both the linear and the angular momentum equations.

Example 4.10

Crude oil of density 950 kg/m3 flows at the rate of 0.58 m3/s through a pipeline of internal
diameter 0.25 m having two right angled bends as shown in the diagram below. The pipeline is
in a horizontal plane(i.e. gravity acts perpendicular to the plane of the diagram). The distance
between the inlet (A) and the exit (B) is 20 m. The gauge pressures at A and B are 345 kPa and
332 kPa respectively.

24
Calculate the force exerted by the supply pipe and the delivery pipe on the pipe with bends.

supply pipe
A
3
0.58 m /s

20 m
0.25 m id

B delivery pipe
______________________________________________

Example 4.11

The diagram below shows a small lawn sprinkler from which the water exits via two
diametrically opposite nozzles of diameter 4 mm. The jets from the nozzle are inclined at an
angle of 30o to the horizontal. At an inlet gauge pressure of 20 kPa, the total volume of flow rate
of water through the sprinkler is 7.5 litres per minute and the sprinkler rotates steadily at 30
r/min8.

30o
gauge pressure = 20 kPa
150 mm

Calculate
(a) the water speed relative to the sprinkler nozzle at exit, and
(b) the frictional torque about the sprinkler axis.
______________________________________________

8
The SI system committee recommends that the abbreviation ‘rpm’ or ‘RPM’ for revolutions per minute’ should
be replaced by ‘r/min’.
25

Anda mungkin juga menyukai