Anda di halaman 1dari 10

6 Chapter 0 The Subject of Transport Phenomena

center of mass and the position vector of the atom with respect to the center of mass, and
w e s o
we recognize that RA2 = -RAU ^ write the same relations for the velocity vectors.
Then we can rewrite Eq. 0.3-3 as

mArA + mBxB = mAxA + mBrB (0.3-4)

That is, the conservation statement can be written in terms of the molecular masses and
velocities, and the corresponding atomic quantities have been eliminated. In getting
Eq. 0.3-4 we have used Eq. 0.3-2 and the fact that for homonuclear diatomic molecules
mM = A2 = \ mA.
(c) According to the law of conservation of energy, the energy of the colliding pair of
molecules must be the same before and after the collision. The energy of an isolated mol-
ecule is the sum of the kinetic energies of the two atoms and the interatomic potential en-
ergy, , which describes the force of the chemical bond joining the two atoms 1 and 2 of
molecule A, and is a function of the interatomic distance \xA2 1|. Therefore, energy
conservation leads to

bnA2rA1 + ) + (1 + \2 + ) =
+ \mA1rA\ + ') + &'\\ + WB2rB\ + ) (0.3-5)

Note that we use the standard abbreviated notation that f\x = (1 f 1 ). We now write
the velocity of atom 1 of molecule A as the sum of the velocity of the center of mass of A
and the velocity of 1 with respect to the center of mass; that is, 1 = + 1. Then Eq.
0.3-5 becomes

(\mAr2A + uA) + (lmBr2B + uB) = %mArA2 + uA) + (lmBrB2 + uB) (0.3-6)

in which uA = \niMRAl + lmA2RA2 + is the sum of the kinetic energies of the atoms, re-
ferred to the center of mass of molecule , and the interatomic potential of molecule A.
That is, we split up the energy of each molecule into its kinetic energy with respect to
fixed coordinates, and the internal energy of the molecule (which includes its vibra-
tional, rotational, and potential energies). Equation 0.3-6 makes it clear that the kinetic
energies of the colliding molecules can be converted into internal energy or vice versa.
This idea of an interchange between kinetic and internal energy will arise again when
we discuss the energy relations at the microscopic and macroscopic levels.
(d) Finally, the law of conservation of angular momentum can be applied to a collision
to give

([1 X 1 1 ] + [2 X 2 2 ]) + ([rB1 X mmim] + [rB2 X mB2iB2]) =


([1 X 1 1 ] + [2 X 2 2 ]) + ([rB1 X 1 1 ] + [rB2 X 2 2 ]) (0.3-7)

in which X is used to indicate the cross product of two vectors. Next we introduce the
center-of-mass and relative position vectors and velocity vectors as before and obtain

([ x ] + 1) + ([rB X mBrB] + 1B) =


([ X ] + 1) + ([rB X ] + 1B) (0.3-8)

in which 1 = [1 X 1 1 ] + [RA2 x mA2RA2] is the sum of the angular momenta of the


atoms referred to an origin of coordinates at the center of mass of the moleculethat is,
the "internal angular momentum." The important point is that there is the possibility for
interchange between the angular momentum of the molecules (with respect to the origin
of coordinates) and their internal angular momentum (with respect to the center of mass
of the molecule). This will be referred to later in connection with the equation of change
for angular momentum.
0.4 Concluding Comments 7

The conservation laws as applied to collisions of monatomic molecules can be ob-


tained from the results above as follows: Eqs. 0.3-1, 0.3-2, and 0.3-4 are directly applica-
ble; Eq. 0.3-6 is applicable if the internal energy contributions are omitted; and Eq. 0.3-8
may be used if the internal angular momentum terms are discarded.
Much of this book will be concerned with setting up the conservation laws at the mi-
croscopic and macroscopic levels and applying them to problems of interest in engineer-
ing and science. The above discussion should provide a good background for this
adventure. For a glimpse of the conservation laws for species mass, momentum, and en-
ergy at the microscopic and macroscopic levels, see Tables 19.2-1 and 23.5-1.

0o4 CONCLUDING COMMENTS


To use the macroscopic balances intelligently, it is necessary to use information about in-
terphase transport that comes from the equations of change. To use the equations of
change, we need the transport properties, which are described by various molecular the-
ories. Therefore, from a teaching point of view, it seems best to start at the molecular
level and work upward toward the larger systems.
All the discussions of theory are accompanied by examples to illustrate how the the-
ory is applied to problem solving. Then at the end of each chapter there are problems to
provide extra experience in using the ideas given in the chapter. The problems are
grouped into four classes:
Class A: Numerical problems, which are designed to highlight important equa-
tions in the text and to give a feeling for the orders of magnitude.
Class B: Analytical problems that require doing elementary derivations using
ideas mainly from the chapter.
Class C: More advanced analytical problems that may bring ideas from other chap-
ters or from other books.
Class D: Problems in which intermediate mathematical skills are required.
Many of the problems and illustrative examples are rather elementary in that they in-
volve oversimplified systems or very idealized models. It is, however, necessary to start
with these elementary problems in order to understand how the theory works and to de-
velop confidence in using it. In addition, some of these elementary examples can be very
useful in making order-of-magnitude estimates in complex problems.
Here are a few suggestions for studying the subject of transport phenomena:
Always read the text with pencil and paper in hand; work through the details of
the mathematical developments and supply any missing steps.
Whenever necessary, go back to the mathematics textbooks to brush up on calculus,
differential equations, vectors, etc. This is an excellent time to review the mathemat-
ics that was learned earlier (but possibly not as carefully as it should have been).
Make it a point to give a physical interpretation of key results; that is, get in the
habit of relating the physical ideas to the equations.
Always ask whether the results seem reasonable. If the results do not agree with
intuition, it is important to find out which is incorrect.
Make it a habit to check the dimensions of all results. This is one very good way of
locating errors in derivations.
We hope that the reader will share our enthusiasm for the subject of transport phe-
nomena. It will take some effort to learn the material, but the rewards will be worth the
time and energy required.
8 Chapter 0 The Subject of Transport Phenomena

QUESTIONS FOR DISCUSSION


1. What are the definitions of momentum, angular momentum, and kinetic energy for a single
particle? What are the dimensions of these quantities?
2. What are the dimensions of velocity, angular velocity, pressure, density, force, work, and
torque? What are some common units used for these quantities?
3. Verify that it is possible to go from Eq. 0.3-3 to Eq. 0.3-4.
4. Go through all the details needed to get Eq. 0.3-6 from Eq. 0.3-5.
5. Suppose that the origin of coordinates is shifted to a new position. What effect would that
have on Eq. 0.3-7? Is the equation changed?
6. Compare and contrast angular velocity and angular momentum.
7. What is meant by internal energy? Potential energy?
8. Is the law of conservation of mass always valid? What are the limitations?
Part One
Momentum
Transport
Chapter 1

Viscosity and the Mechanisms


of Momentum Transport
1.1 Newton's law of viscosity (molecular momentum transport)
1.2 Generalization of Newton's law of viscosity
1.3 Pressure and temperature dependence of viscosity
1.4 Molecular theory of the viscosity of gases at low density
1.5 Molecular theory of the viscosity of liquids
1.6 Viscosity of suspensions and emulsions
1.7 Convective momentum transport

The first part of this book deals with the flow of viscous fluids. For fluids of low molecu-
lar weight, the physical property that characterizes the resistance to flow is the viscosity.
Anyone who has bought motor oil is aware of the fact that some oils are more "viscous"
than others and that viscosity is a function of the temperature.
We begin in 1.1 with the simple shear flow between parallel plates and discuss how
momentum is transferred through the fluid by viscous action. This is an elementary ex-
ample of molecular momentum transport and it serves to introduce "Newton's law of vis-
cosity" along with the definition of viscosity /. Next in 1.2 we show how Newton's law
can be generalized for arbitrary flow patterns. The effects of temperature and pressure
on the viscosities of gases and liquids are summarized in 1.3 by means of a dimension-
less plot. Then 1.4 tells how the viscosities of gases can be calculated from the kinetic
theory of gases, and in 1.5 a similar discussion is given for liquids. In 1.6 we make a
few comments about the viscosity of suspensions and emulsions.
Finally, we show in 1.7 that momentum can also be transferred by the bulk fluid
motion and that such convective momentum transport is proportional to the fluid density p.

1.1 NEWTON'S LAW OF VISCOSITY (MOLECULAR


TRANSPORT OF MOMENTUM)
In Fig. 1.1-1 we show a pair of large parallel plates, each one with area A, separated by a
distance . In the space between them is a fluideither a gas or a liquid. This system is
initially at rest, but at time t = 0 the lower plate is set in motion in the positive x direc-
tion at a constant velocity V. As time proceeds, the fluid gains momentum, and ulti-
mately the linear steady-state velocity profile shown in the figure is established. We
require that the flow be laminar ("laminar" flow is the orderly type of flow that one usu-
ally observes when syrup is poured, in contrast to "turbulent" flow, which is the irregu-
lar, chaotic flow one sees in a high-speed mixer). When the final state of steady motion

11
12 Chapter 1 Viscosity and the Mechanisms of Momentum Transport

Fig. 1.1-1 The buildup to


, Q Fluid initially the steady, laminar velocity
at rest profile for a fluid contained
between two plates. The
flow is called 'laminar" be-
cause the adjacent layers of
fluid ("laminae") slide past
Lower plate one another in an orderly
set in motion
fashion.

..
cb m a 1 1 f
Velocity buildup
vx(y, t) in unsteady flow

Final velocity
Large t distribution in
steady flow

has been attained, a constant force F is required to maintain the motion of the lower
plate. Common sense suggests that this force may be expressed as follows:

V (1.1-1)

That is, the force should be proportional to the area and to the velocity, and inversely
proportional to the distance between the plates. The constant of proportionality is a
property of the fluid, defined to be the viscosity.
We now switch to the notation that will be used throughout the book. First we re-
place F/A by the symbol ryx, which is the force in the x direction on a unit area perpen-
dicular to the direction. It is understood that this is the force exerted by the fluid of
lesser on the fluid of greater y. Furthermore, we replace V/Y by -dvx/dy. Then, in
terms of these symbols, Eq. 1.1-1 becomes
dvx
(1.1-2)1

This equation, which states that the shearing force per unit area is proportional to the
negative of the velocity gradient, is often called Newton's law of viscosity} Actually we

1
Some authors write Eq. 1.1-2 in the form
dvx
(1
"2)
in which [ = ] lty/ft2, vx [ = ] ft/s, [=] ft, and /JL [=] lbm/ft s; the quantity f is the "gravitational
conversion factor" with the value of 32.174 poundals/lty. In this book we will always use Eq. 1.1-2 rather
thanEq. l.l-2a.
2
Sir Isaac Newton (1643-1727), a professor at Cambridge University and later Master of the Mint,
was the founder of classical mechanics and contributed to other fields of physics as well. Actually Eq.
1.1-2 does not appear in Sir Isaac Newton's Philosophiae Naturalis Principia Mathematica, but the germ of
the idea is there. For illuminating comments, see D. J. Acheson, Elementary Fluid Dynamics, Oxford
University Press, 1990, 6.1.
1.1 Newton's Law of Viscosity (Molecular Transport of Momentum) 13

should not refer to Eq. 1.1-2 as a "law/' since Newton suggested it as an empiricism 3
the simplest proposal that could be made for relating the stress and the velocity gradi-
ent. However, it has been found that the resistance to flow of all gases and all liquids
with molecular weight of less than about 5000 is described by Eq. 1.1-2, and such fluids
are referred to as Newtonian fluids. Polymeric liquids, suspensions, pastes, slurries, and
other complex fluids are not described by Eq. 1.1-2 and are referred to as non-Newtonian
fluids. Polymeric liquids are discussed in Chapter 8.
Equation 1.1-2 may be interpreted in another fashion. In the neighborhood of the
moving solid surface at = 0 the fluid acquires a certain amount of x-momentum. This
fluid, in turn, imparts momentum to the adjacent layer of liquid, causing it to remain in
motion in the x direction. Hence x-momentum is being transmitted through the fluid in
the positive direction. Therefore ryx may also be interpreted as the flux of x-momentum
in the positive direction, where the term "flux" means "flow per unit area." This interpre-
tation is consistent with the molecular picture of momentum transport and the kinetic
theories of gases and liquids. It also is in harmony with the analogous treatment given
later for heat and mass transport.
The idea in the preceding paragraph may be paraphrased by saying that momentum
goes "downhill" from a region of high velocity to a region of low velocityjust as a sled
goes downhill from a region of high elevation to a region of low elevation, or the way
heat flows from a region of high temperature to a region of low temperature. The veloc-
ity gradient can therefore be thought of as a "driving force" for momentum transport.
In what follows we shall sometimes refer to Newton's law in Eq. 1.1-2 in terms of
forces (which emphasizes the mechanical nature of the subject) and sometimes in terms
of momentum transport (which emphasizes the analogies with heat and mass transport).
This dual viewpoint should prove helpful in physical interpretations.
Often fluid dynamicists use the symbol v to represent the viscosity divided by the
density (mass per unit volume) of the fluid, thus:
v = p/p (1.1-3)
This quantity is called the kinematic viscosity.
Next we make a few comments about the units of the quantities we have defined. If
we use the symbol [=] to mean "has units of," then in the SI system rXJX [=] N/m 2 = Pa,
vx [ = ] m/s, and [=] m, so that

l l
= ^ [ = ] ( P a ) [ ( m / s ) ( m )] = s (1.1-4)
\dy)
since the units on both sides of Eq. 1.1-2 must agree. We summarize the above and also
give the units for the c.g.s. system and the British system in Table 1.1-1. The conversion
tables in Appendix F will prove to be very useful for solving numerical problems involv-
ing diverse systems of units.
The viscosities of fluids vary over many orders of magnitude, with the viscosity of
5
air at 20C being 1.8 X 10~ Pa s and that of glycerol being about 1 Pa s, with some sili-
4
cone oils being even more viscous. In Tables 1.1-2,1.1-3, and 1.1-4 experimental data are

3
A relation of the form of Eq. 1.1-2 does come out of the simple kinetic theory of gases (Eq. 1.4-7).
However, a rigorous theory for gases sketched in Appendix D makes it clear that Eq. 1.1-2 arises as the
first term in an expansion, and that additional (higher-order) terms are to be expected. Also, even an
elementary kinetic theory of liquids predicts non-Newtonian behavior (Eq. 1.5-6).
4
A comprehensive presentation of experimental techniques for measuring transport properties can be
found in W. A. Wakeham, A. Nagashima, and J. V. Sengers, Measurement of the Transport Properties of Fluids,
CRC Press, Boca Raton, Fla. (1991). Sources for experimental data are: Landolt-Bornstein, Zahlemverte und
Funktionen, Vol. II, 5, Springer (1968-1969); International Critical Tables, McGraw-Hill, New York (1926);
Y. S. Touloukian, P. E. Liley, and S. Saxena, Thermophysical Properties of Matter, Plenum Press, New York
(1970); and also numerous handbooks of chemistry, physics, fluid dynamics, and heat transfer.
14 Chapter 1 Viscosity and the Mechanisms of Momentum Transport

Table 1.1-1 Summary of Units for Quantities


Related to Eq. 1.1-2

SI c.g.s. British
2 2
Pa dyn/cm poundals/ft
vx m/s cm/s ft/s
m cm ft
V> Pa-s gm/cm s = poise lb^/ft-s
2 2 2
V m /s cm /s ft /s
2
Note: The pascal, Pa, is the same as N/m , and the newton,
N, is the same as kg m/s2. The abbreviation for "centipoise"
is "cp."

Table 1.1-2 Viscosity of Water and Air at 1 atm Pressure

Water (liq.r Air"

Temperature Viscosity Kinematic viscosity Viscosity Kinematic viscosity


7TC) /JL (mPa s) v (cm2/s) / (mPa s) v (cm7s)
0 1.787 1.787 0.01716 13.27
20 1.0019 1.0037 0.01813 15.05
40 0.6530 0.6581 0.01908 16.92
60 0.4665 0.4744 0.01999 18.86
80 0.3548 0.3651 0.02087 20.88
100 0.2821 0.2944 0.02173 22.98
a
Calculated from the results of R. Hardy and R. L. Cottington, /. Research Nat. Bur. Standards, 42,
573-578 (1949); and J. F. Swidells, J. R. , Jr., and . . Godfrey, /. Research Nat. Bur. Standards, 48,1-31
(1952).
b
Calculated from "Tables of Thermal Properties of Gases," National Bureau of Standards Circular 464
(1955), Chapter 2.

Table 1.1-3 Viscosities of Some Gases and Liquids at Atmospheric Pressure"

Temperature Viscosity Temperature Viscosity


Gases T(C) / (mPa s) Liquids T(C) /x (mPa s)

i-QH 1 0 23 0.0076c (C 2 H 5 ) 2 O 0 0.283


SF 6 23 0.0153 25 0.224
CH4 20 0.0109* QH6 20 0.649
H2O 100 0.01211rf Br2 25 0.744
co
N 2
20
20
0.0146b
0.0175b
Hg
C 2 H 5 OH
20
0
1.552
1.786
2

o2 20
380
0.0204
0.0654'y
25
50
1.074
0.694
Hg
H 2 SO 4 25 25.54
Glycerol 25 934.
a
Values taken from N. A. Lange, Handbook of Chemistry, McGraw-Hill, New York, 15th edition
(1999), Tables 5.16 and 5.18.
b
H. L. Johnston and K. E. McKloskey, J. Phys. Chem., 44,1038-1058 (1940).
c
CRC Handbook of Chemistry and Physics, CRC Press, Boca Raton, Fla. (1999).
d
Landolt-Bornstein Zahlenwerte und Funktionen, Springer (1969).
1.1 Newton's Law of Viscosity (Molecular Transport of Momentum) 15

Table 1.1-4 Viscosities of Some Liquid Metals

Temperature Viscosity
Metal T(C) /x (mPa s)

Li 183.4 0.5918
216.0 0.5406
285.5 0.4548
Na 103.7 0.686
250 0.381
700 0.182
69.6 0.515
250 0.258
700 0.136
Hg -20 1.85
20 1.55
100 1.21
200 1.01
Pb 441 2.116
551 1.700
844 1.185

Data taken from The Reactor Handbook, Vol. 2, Atomic


Energy Commission AECD-3646, U.S. Government
Printing Office, Washington, D.C. (May 1955), pp. 258
et seq.

given for pure fluids at 1 atm pressure. Note that for gases at low density, the viscosity
increases with increasing temperature, whereas for liquids the viscosity usually decreases
with increasing temperature. In gases the momentum is transported by the molecules in
free flight between collisions, but in liquids the transport takes place predominantly by
virtue of the intermolecular forces that pairs of molecules experience as they wind their
way around among their neighbors. In 1.4 and 1.5 we give some elementary kinetic
theory arguments to explain the temperature dependence of viscosity.

2
EXAMPLE 1.1-1 Compute the steady-state momentum flux in lty/ft when the lower plate velocity V in Fig.
1.1-1 is 1 ft/s in the positive x direction, the plate separation is 0.001 ft, and the fluid viscos-
Calculation of ity ix is 0.7 cp.
Momentum Flux
SOLUTION
Since is desired in British units, we should convert the viscosity into that system of units.
Thus, making use of Appendix F, we find /x = (0.7 cp)(2.0886 X 10"5) = 1.46 X 10~5 lb, s/ft2.
The velocity profile is linear so that
dvx bvx = -1.0 ft/s
=
= -lOOOs"1 (1.1-5)
dy ~ 0.001 ft
Substitution into Eq. 1.1-2 gives

ryx = -fi^ = -(1.46 X 10~5)(-1000) = 1.46 X 10" 2 lb/ft 2 (1.1-6)


ay '

Anda mungkin juga menyukai