Sie sind auf Seite 1von 60

Chapter 3

The Equations of Change


for Isothermal Systems
In Chapter 2, velocity distributions were determined for
several simple flow systems by the shell momentum
balance method. The resulting velocity distributions were
then used to get other quantities, such as the average
velocity and drag force.
For more complex problems we need a general mass
balance and a general momentum balance that can be
applied to any problem, including problems with
nonrectilinear motion. That is the main point of this
chapter.
The two equations that we derive are called the equation
of continuity (for the mass balance) and the equation of
motion (for the momentum balance). These equations can
be used as the starting point for studying all problems
involving the isothermal flow of a pure fluid.
Those equations are called as equations of change
because they describe the change of velocity due to the
change of time and position in the fluid system.
3.1. THE EQUATION OF CONTINUITY
(MASS BALANCE)
This equation is developed by writing a mass balance over a
volume element Ax.Ay.Az , fixed in space, through which a
fluid is flowing (see Fig. 3.1-1):




The rate of mass entering the volume element through the
shaded face at x is (v
x
)|
x
Ay.Az, and the rate of mass leaving
through the shaded face at x + Ax is (v
x
)|
x+Ax
Ay.Az.


Similar expressions can be written for the other two pairs
of faces. The rate of increase of mass within the volume
element is Ax.Ay.Az(/t). The mass balance then
becomes



By dividing the entire equation by Ax.Ay.Az and taking
the limit as Ax, Ay, and Az go to zero, and then using the
definitions of the partial derivatives, we get

( ) ( )
( ) ( )
( ) ( )
x x x x x y y y y y
z z z z z
x y z y z v v x z v v
t
x y v v
A A
A

A A A A A A A
A A
+ +
+
c
(
(
= +
(

c
(
+

(3.1-2)
. (3.1-3)

This is the equation of continuity, which describes the
time rate of change of the fluid density at a fixed point in
space. This equation can be written more concisely by
using vector notation as follows:.




V.v = "divergence of v"
x y z
v v v
t x y z


| | c c c c
= + +
|
c c c c
\ .
The vector v is the mass flux, and its divergence has a
simple meaning: it is the net rate of mass efflux per unit
volume. A very important special form of the equation of
continuity is that for a fluid of constant density, for which
Eq. 3.1-4 assumes the particularly simple form
(incompressible fluid)
.

Of course, no fluid is truly incompressible, but frequently
in engineering and biological applications, the assumption
of constant density results in considerable simplification
and very little error.

Example 3.1-1. Normal Stresses at Solid
Surfaces for Incompressible Newtonian
Fluids
Show that for any kind of flow pattern, the normal stresses are
zero at fluid-solid boundaries, for Newtonian fluids with
constant density. This is an important result that we shall use
often.
SOLUTI ON
We visualize the flow of a fluid near some solid surface, which
may or may not be flat. The flow may be quite general, with all
three velocity components being functions of all three
coordinates and time.
At some point P on the surface we erect a Cartesian coordinate
system with the origin at P.
We now ask what the normal stress t
zz
is at P.
According to Table B.l or Eq. 1.2-6, t
zz
= -2(dv
z
/dz), because
for incompressible fluids. Then at point P on the surface
of the solid

( )
0 V = .v

.

First we replaced the derivative dv
z
/dz by using Eq. 3.1-3 with
constant. However, on the solid surface at z = 0, the velocity
v
x
is zero at any position of x by the no-slip condition (see
2.1), and therefore the derivative dv
x
/dx on the surface = 0.
The same is true of dv
y
/dy on the surface. Therefore t
zz
is zero.
It is also true that t
xx
and t
yy
are zero at the surface because of
the vanishing of the derivatives at z = 0. (Note: The vanishing
of the normal stresses on solid surfaces does not apply to
polymeric fluids, which are viscoelastic).
For compressible fluids, the normal stresses at solid surfaces
are zero if the density is not changing with time, as is shown in
Problem 3C.2.)
According to mass balance
for incompressible flow
3.2. THE EQUATION OF MOTION
(MOMENTUM BALANCE)
To get the equation of motion we write a momentum
balance over the volume element Ax.Ay.Az in Fig. 3.2-1
of the form.

Note that Eq. 3.2-1 is an extension of Eq. 2.1-1 to
unsteady-state problems.
The fluid is allowed to move through all six faces of the
volume element.
Remember that Eq. 3.2-1 is a vector equation with
components in each of the three coordinate directions x, y,
and z.
We develop the x-component of each term in Eq. 3.2-1.
The y- and z-components may be treated analogously.
First, we consider the rates of flow of the x-component
of momentum into and out of the volume element shown
in Fig. 3.2-1. (second subscript x: direction of tangential
and normal stresses for molecular transport and direction
of momentum flux for convective transport)
Momentum enters and leaves Ax.Ay.Az by two
mechanisms: molecular transport (see 1.2) and
convective transport (see 1.7).


First subscripts x, y, and z (cause): directions of momentum
transfers due to the change of velocity for molecular stress or due
to convection represented by velocity in x, y and z directions for
convective stress. Cause can be from all directions.
Second subscripts x (effect): directions of tangential stress or
normal stress or direction of convective momentum flux
represented by direction of mass flux (in x direction). Effect is
only in one direction in a momentum balance
Second subscripts of all
components are the same,
i.e. x (direction of effect)
The rate at which the x-component of momentum enters
across the shaded face at x by all mechanisms-both
convective and molecular-is |
xx
|
x
AyAz and the rate at which
it leaves the shaded face at x + A x is |
xx
|
x+Ax
AyAz.
The rates at which x-momentum enters and leaves through
the faces at y and y + Ay are |
yx
|
y
AzAx and |
yx
|
y+Ay
AzAx
respectively.
Similarly, the rates at which x-momentum enters and leaves
through the faces at z and z + Az are |
zx
|
z
AxAy and |
zx
|
z+Az

AxAy
When these contributions are added we get for the net rate
of addition of x-momentum across all three pairs of faces.
.

Next there is the external force (typically the gravitational
force) acting on the fluid in the volume element. The x-
component of this force is

Equations 3.2-2 and 3.2-3 give the x-components of the
three terms on the right side of Eq. 3.2-1.
The sum of these terms must then be equated to the rate
of increase of x-momentum within the volume element:
Ax.Ay.Az (v
x
)/t. When this is done, we have the x-
component of the momentum balance. When this
equation is divided by Ax.Ay.Az and the limit is taken as
Ax, Ay and Az zero, the following equation results:
Here we have made use of the definitions of the partial
derivatives. Similar equations can be developed for the y-
and z-components of the momentum balance:
.



By using vector-tensor notation, these three equations can
be written as follows:
.

This is a vector equation (vector dot tensor = vector)
That is, by letting i be successively x, y, and z, Eqs. 3.2-4,5,
and 6 can be reproduced. The quantities v
i
are the Cartesian
components of the vector v, which is the momentum per
unit volume at a point in the fluid.
Similarly, the quantities g
i
are the components of the
vector g, which is the external force per unit volume. The
term -[.|]
i
is the ith component of the vector -[.|].
When the magnitude of ith component of Eq. 3.2-7 is
multiplied by the unit vector in the ith direction and the
three components are added together vectorially, we get

which is the differential statement of the law of conservation
of momentum. It is the translation of Eq. 3.2-1 into
mathematical symbols.
In Eq. 1.7-1 it was shown that the combined momentum
flux tensor | is the sum of the convective momentum flux
tensor vv and the molecular momentum flux tensor t,
and that the latter can be written as po +t . When we
insert | = po + vv + t into Eq. 3.2-8, we get the
following equation of motion
In this equation p is a vector (=vector times scalar)
called the "gradient of (the scalar) p" sometimes written
as "grad p ". The symbol [.t] is a vector (=vector dot
tensor) called the "divergence of (the tensor) t " and
[.vv] is a vector (=vector dot tensor) called the
"divergence of vv.
In the next two sections we give some formal results that
are based on the equation of motion. The equations of
change for mechanical energy and angular momentum are
not used for problem solving in this chapter, but will be
referred to in Chapter 7 (these sections are excluded
from the lecture material!).
3.5. THE EQUATIONS OF CHANGE IN
TERMS OF THE SUBSTANTIAL
DERIVATIVE
The Partial Time Derivative /t (derivative against one
variable)
Suppose we stand on a bridge and observe the concentration
of fish just below us as a function of time. We can then
record the time rate of change of the fish concentration at a
fixed location. The result is (c/t)|
x,y,z
the partial derivative
of c with respect to t, at constant x, y, and z.
The Total Time Derivative d/dt (derivative against all
variables)
Now suppose that we jump into a motor boat and speed
around on the river, sometimes going upstream, sometimes
downstream, and sometimes across the current as we wish.

All the time we are observing fish concentration. At any
instant, the time rate of change of the observed fish
concentration is
. (3.5-1)

in which dx/dt, dy/dt, and dz/dt are the components of the
velocity of the boat.
The Substantial Time Derivative D/Dt
Next we climb into a canoe and we just float along with
the current to observe the fish concentration.
x,y,z y,x,t x,y,t
x,z,t
dc c dx c dy c dz c
dt t dt x dt y dt z
| | c c c c
| | | | | |
= + + +
|
| | |
c c c c
\ . \ . \ .
\ .
In this situation the velocity of the observer = the velocity
v of the stream, which has components v
x
, v
y
, and v
z
.
If at any instant we report the time rate of change of fish
concentration, we are then giving


The special operator D/Dt = /t + v. is called the
substantial derivative (meaning that the time rate of
change is reported as one moves with the "substance").
The terms material derivative, hydrodynamic derivative,
and derivative following the motion are also used.
Now we need to know how to convert equations
expressed in terms of /t into equations written with
D/Dt. For any scalar function f(x,y,z,t) we can do the
following manipulations:
=0
According to mass balance
The quantity in the second parentheses in the second line =
0 according to the equation of continuity. Consequently
Eq. 3.5-3 can be written in vector form as


Similarly, for any vector function f(x,y,z,t),

These equations can be used to rewrite the equations of
change given in 3.1 to 3.4 in terms of the substantial
derivative as shown in Table 3.5-1.
: in Chapter 7, excluded in this lecture
D/Dt = /t + v.
Equation A in Table 3.5-1 tells how the density is
decreasing or increasing as one moves along with the fluid,
because of the compression [(.v) < 0] or expansion of the
fluid [(.v) > 0].
Equation B can be interpreted as (mass) x (acceleration) =
the sum of the pressure forces, viscous forces, and the
external force. In other words, Eq. 3.2-9 is equivalent to
Newton's second law of motion
(density x acceleration = summation of all
forces /volume)
Three most common simplifications of the equation of motion:
For constant and , insertion of the Newtonian expression
for t from Eq. 1.2-7 into the equation of motion leads to the
very famous Navier-Stokes equation, first developed from
molecular arguments by Navier, a French engineer, and from
continuum arguments by Stokes, an English mathematician:
.

When the acceleration terms in Navier-Stokes equation are
neglected-that is, when (Dv/Dt) = 0 -we get

which is called the Stokes flow equation. It is sometimes
called the creeping flow equation, because the term
(v.v] ~ 0 when the flow is extremely slow and can be
approached as steady flow.
D/Dt = /t + v.
temporal change
Spatial/positional change
When viscous forces in Navier-Stokes equation are
neglected - that is, .t =
2
v = 0 - the equation of
motion becomes (normal and shear stresses occur due
to viscosity)

which is known as the Euler equation for "inviscid"
fluid in unsteady flow. Of course, there are no truly
"inviscid" fluids, but there are many flows in which
the viscous forces are relatively unimportant (far from
solid surfaces or very high velocity). Examples are
the flow around airplane wings (except near the solid
boundary), flow of rivers around the upstream surfaces
of bridge supports, some problems in compressible gas
dynamics, and flow of ocean current.
Example 3.5-1. The Bernoulli Equation
for the Steady Flow of Inviscid Fluids
The Bernoulli equation for steady flow of inviscid,
incompressible fluids (conditions in Stokes flow and
Euler eqs) is one of the most famous equations in
classical fluid dynamics. Show how it is obtained from
the Euler equation of motion.
SOLUTI ON
Inviscid Fluids omit the time-derivative term in Eq. 3.5-
9, and then use the vector identity [.vvl = [v.vl =
(v.v) - [v x [ x v]] (Eq. A.4-23) to rewrite the Navier
Stokes equation as

Next we divide Eq. 3.5-10 by and then form the dot product
with the unit vector s = v/|v| in the flow direction. When the
fluid is inviscid, then there is no vorticity ( x v = 0) and
consequently v x ( x v) = 0, and (s.) can be replaced by
d/ds, where s is the distance along a streamline. Thus we get


When this is integrated along a streamline from point 1 to
point 2, we get


which is called the Bernoulli equation. It relates the velocity,
pressure, and elevation of two points along a streamline in a
fluid in steady-state flow of inviscid fluid.
3.6. USE OF THE EQUATIONS OF CHANGE
TO SOLVE FLOW PROBLEMS
To describe the flow of a Newtonian fluid at constant
temperature, we need in general
The equation of continuity Eq. 3.1-4
The equation of motion Eq. 3.2-9
The components of t Eq. 1.2-6
The equation of state = (p)
The equations for the viscosities = (p, T)
These equations, along with the necessary boundary
(related to positions) and initial (related to time)
conditions, determine completely the pressure, density, and
velocity distributions in the fluid.
They are seldom used in their complete form to solve fluid
dynamics problems. Usually restricted forms are used for
convenience, as in this chapter.
If it is appropriate to assume constant density and
viscosity, then we use
The equation of continuity Eq. 3.1-4 and Table B.4
The Navier-Stokes equation Eq. 3.5-6 and Tables B.5,
6, 7 along with initial and boundary conditions.
From these one determines the pressure and velocity
distributions.
Example 3.6-1. Steady Flow in a Long
Circular Tube
Rework the tube-flow problem of Example 2.3-1 using the
equations of continuity and motion. This illustrates the use
of the tabulated equations for constant viscosity and
density in cylindrical coordinates, given in Appendix B.5.
SOLUTI ON
We postulate that v = o
z
v
z
(r, z). This postulate implies that
there is no radial flow (v
r
= 0) and no tangential flow (v
u
=
0), and that v
z
f (u).
We assume that there is no change of velocity profile in z
direction.
Consequently, we can discard many terms from the
tabulated equations of change, leaving



The postulate and Eq. 3.6-1 indicates that v
z
depends only
on r; hence the partial derivatives in the second term on
the right side of Eq. 3.6-4 can be replaced by ordinary
derivatives.
By using the modified pressure P = p +gh (where h is
the height above some arbitrary datum plane and g is a
constant), we avoid the necessity of calculating the
components of g in r and u coordinates, and we obtain a
solution valid for any orientation of the axis of the tube.
Equations 3.6-2 and 3.6-3 show that P is a function of z
alone, and the partial derivative in the first term of Eq.
3.6-4 may be replaced by an ordinary derivative.
For constant change of P against z, by introducing a
constant C
0
, Eq. 3.6-4 reduces to


The P equation can be integrated at once. The v
z
-equation
can be integrated one operation after another on the left
side (do not "work out" the compound derivative there).
This gives
.


The four constants of integration can be found from the
boundary conditions:
.




ln 0 = indefinite, so C
2
= 0 to obtain definite v
z
.. The
resulting solutions are:
.


As pointed out in Example 2.3-1, Eq. 3.6-13 is valid only in
the laminar-flow regime, and at locations not too near the
tube entrance and exit. For Re > about 2100, a turbulent-flow
regime exists downstream of the entrance region, and Eq.
3.6-13 is no longer valid.
Example 3.6-2. Falling Film with Variable
Viscosity
Set up the problem in Example 2.2-2 by using the
equations of Appendix B.5. This illustrates the use of the
equation of motion in terms of t.
SOLUTI ON
As in Example 2.2-2 we postulate a steady-state flow
with constant density, but with viscosity depending on x.
We postulate, as before, that the x- and y-components of
the velocity are zero (v
x
and v
y
= 0) and that v
z
= v
z
(x).
With these postulates, the equation of continuity is zero.
According to Table B.l, the only nonzero components of t
are t
xz
= t
zx
= -(dv
z
/dx). The components of the equation
of motion in terms of t are, from Table B.5,

.



Integration of Eq. 3.6-14 gives
.
in which f(y, z) is an arbitrary function. Equation 3.6-15
shows that f cannot be a function of y.


We next recognize that the pressure in the gas phase is very
nearly constant at the prevailing atmospheric pressure p
atm
.
Therefore, at the gas-liquid interface x = 0, the pressure is
also constant at the value p
atm
. Consequently, f can be set
equal to p
atm
, and we obtain finally from 3.6-14.
.
(p is function of x only). Equation 3.5-16 then becomes
.
which is the same as Eq. 2.2-10. The remainder of the
solution is the same as in 2.2.
Example 3.6-3. Operation of a Couette
Viscometer
The viscosity may also be determined by measuring the
torque required to turn a solid object in contact with a
fluid. The forerunner of all rotational viscometers is the
Couette instrument, which is sketched in Fig. 3.6-1.
Determine velocity distribution and shear stress for the
laminar, tangential flow of an incompressible fluid
between 2 co-axial vertical cylinders. Outer cylinder
rotates with angular velocity O
o
(see Figure 3.6-1). End-
effects is negligible.
Solution
In steady-state laminar flow, fluid moves in circular
direction with velocity components v
r
= 0 and v
z
= 0.
There is no pressure gradient in u direction (p = p(r,z)). It
is expected that p depends on z due to gravity and on r
due to centrifugal force.
For these postulates all the terms in the equation of
continuity are zero, and the components of the equation of
motion simplify to




The first equation tells how the centrifugal force affects the
pressure.
The second equation gives the velocity distribution.
The third equation gives the effect of gravity on the
pressure (the hydrostatic effect)
For the problem at hand we need only the u-component of
the equation of motion for velocity distribution
0
z
p
g
z

c
= +
c

Integration of Eq. 3.6-21 results in
.





The boundary conditions are that the fluid does not slip at
the two cylindrical surfaces:
.

These boundary conditions can be used to get the
constants of integration, which are then inserted in Eq.
3.6-26. This gives
.


From the velocity distribution we can find the momentum
flux by using Table B.1:
The torque acting on the inner cylinder is then given by
the product of the inward momentum flux (-t
ru
), the
surface of the cylinder, and the lever arm, as follows:
.


Therefore, measurement of the angular velocity of the cup
makes it possible to determine the viscosity. The same
kind of analysis is available for other rotational
viscometers.
For any viscometer it is essential to know when
turbulence will occur. The critical Reynolds number
(O
o
R
2
/), above which the system becomes turbulent, is
shown in Fig. 3.6-2 as a function of the radius ratio k.
One might ask what happens if we hold the outer cylinder
fixed and cause the inner cylinder to rotate with an
angular velocity O
i
(the subscript "i" stands for inner).
Then the velocity distribution is



This is obtained by making the same postulates (see
before Eq. 3.6-20) and solving the same differential
equation (Eq. 3.6-21), but with a different set of boundary
conditions.

Das könnte Ihnen auch gefallen