Você está na página 1de 34

APPENDIX 1 SOME RESULTS FROM THE KINETIC THEORY OF GASES A1.

1 DISTRIBUTION OF MOLECULAR VELOCITIES IN A GAS


A1.1.1 THE DISTRIBUTION DERIVED FROM THE BAROMETRIC FORMULA

In Chapter 2 the variation of gas density in the atmosphere was derived. In the upper atmosphere where the gas temperature is approximately constant
"gh -------2 a0

!(z + h) ------------------- = e !(z) Rewrite (A1.1) in terms of molecular parameters.


n(z + h) ------------------- = e n(z)
mgh ---------kT

(A1.1)

(A1.2)

where n ( z ) is the number of molecules per unit volume at the height z , m = M w N is the mass of each molecule (or the average mass in the case of a mixture like air). The Boltzmanns constant, k = 1.38 10
23

Joules K

(A1.3)

is related to the gas constant and Avagadros number N by


(A1.4) k = Ru N . Equation (A1.2) expresses the distribution of molecules in any system at constant temperature subject to a body force. The fraction of molecules per unit area above a given altitude z + h is also given by the distribution (A1.2). That is

$ n ( z ) dz- = e ---------------------------------(z + h) #

mgh ---------kT

(A1.5)

n ( z ) dz
z

which is a basic property of an exponential distribution.

bjc

A1.1

7/12/11

Distribution of molecular velocities in a gas

The distribution of molecules under the downward pull of gravity given by (A1.5) can be used to infer how the velocities of gas molecules are distributed. Consider molecules that pass through some reference height z on their way to height z + h . z+h

uh m

The upward velocity component u h of a molecule whose kinetic energy at z is equal to its potential energy when it arrives at z + h satises mgh = mu h 2 .
2 (A1.6)

This is the minimum positive velocity needed for a molecule starting at z to reach the height z + h . Let J u > u ( z ) be the ux of molecules ( number of molecules/area-sec ) passing h through the plane z that possess a vertical velocity component greater than u h and let J u > 0 ( z ) be the ux of molecules passing through the plane z with any positive upward velocity. The fraction of molecules passing upward through z with the requisite velocity is J u > u ( z ) J u > 0 ( z ) . This ratio should be the same
h

as the fraction of molecules above the height z + h . Therefore one can infer the equality n ( z ) dz Ju > u (z) z+h h ----------------------- = -----------------------------# J u > 0(z) n ( z ) dz

# (A1.7)

Using (A1.5) we deduce that

7/12/11

A1.2

bjc

Distribution of molecular velocities in a gas

Ju > u (z) h ----------------------- = e J u > 0(z) which is plotted in Figure A.1. Ju > u (z) h ----------------------J u > 0(z)
1 0.8 0.6 0.4 0.2

mgh ---------kT

= e

mu h -----------2kT

(A1.8)

0.5

1.5

2.5

m --------- u h 2kT

Figure A.1 Fraction of molecules at any reference height with positive vertical velocity component exceeding u h .

Now use (A1.7) to generate a probability density function (pdf) g ( u h ) for the velocities. The pdf is dened such that, at any height z , g ( u h )du h equals the fraction of molecules per unit volume with velocity between u h and u h + du h . The pdf we are seeking would look something like Figure A.2. g ( uh )
duh

uh
Figure A.2 Probability density function of one velocity component. The shaded area divided by the whole area (which is one) is the fraction of molecules with velocity between u h and u h + du h .

bjc

A1.3

7/12/11

Distribution of molecular velocities in a gas

The desired pdf is normalized so that

$
Note that

# #

g ( u h ) du h = 1 .

(A1.9)

$#

exp ( x ) dx =

%.

The number of molecules that pass through a unit area of the plane z in unit time (the ux of molecules) with vertical velocity component u h is u h g ( u h )du h . In terms of the velocity pdf Ju > u (z) h ----------------------- = J u > 0(z)
# mu h -----------2kT
2

uh

u h g ( u h )du h = e

(A1.10)

Differentiate both sides of (A1.10).


mu h -----------2kT
2

m u h g ( u h ) = ------u h e kT Thus far

(A1.11)

g ( u h ) = ------e . kT Apply the normalization condition (A1.9) to (A1.12).


mu h -----------# m 2kT
2

mu h -----------m 2kT

(A1.12)

------e kT #

2%m 1 2 du h = ( ---------- ) . & kT '

(A1.13)

Finally, the normalized 1-dimensional velocity pdf is


mu h 1 2 -----------2kT m )
2

g ( u h ) = ( ------------& 2 % kT '

(A1.14)

7/12/11

A1.4

bjc

Distribution of molecular velocities in a gas

We arrived at this result using a model of particles with random vertical velocity components moving in a gravitational eld. In a gas in thermodynamic equilibrium at a temperature T , the molecules constantly undergo collisions leading to chaotic motion with a wide range of molecular velocities in all three directions. The randomness is so strong that at any point no one direction is actually preferred over another and one can expect that the probability density function for a velocity component in any direction at a given height z will have the same form as (A1.14).
A1.1.2 THE MAXWELL VELOCITY DISTRIBUTION FUNCTION

We can use a slightly different, more general, argument to arrive at the velocity distribution function in three dimensions. The randomness produced by multiple collisions leads to a molecular motion that is completely independent in all three coordinate directions and so the probability of occurrence of a particular triad of velocities ( u x, u y, u z ) is equal to the product of three one-dimensional probability distributions. f ( u x, u y, u z ) = g ( u x )g ( u y )g ( u z ) .
(A1.15)

In addition, since no direction in velocity space ( u x, u y, u z ) is preferred over any other, then the three dimensional velocity probability distribution function must be invariant under any rotation of axes in velocity space. This means f ( u x, u y, u z ) can only depend on the radius in velocity space f ( u x, u y, u z ) = f ( u x + u y + u z ) (A1.16) The only smooth function that satises both (A1.115) and (A1.116) and the requirement that the probability goes to zero at innity is the decaying exponential f ( u x, u y, u z ) = Ae
B ( u x + u y + uz )
2 2 2

(A1.17)

In a homogeneous gas the number density of molecules with a given velocity is accurately described by the Maxwellian velocity distribution function
2 2 2 m 32 m exp ( --------- ( u 1 + u 2 + u 3 )) f ( u 1, u 2, u 3 ) = ( -------------) & 2kT ' & 2 % kT ' (A1.18)

bjc

A1.5

7/12/11

Distribution of molecular velocities in a gas

where the correspondence ( u x, u y, u z ) * ( u 1, u 2, u 3 ) is used. The result (A1.118) is consistent with the use of (A1.114) in all three directions. The dimensions of f are sec M . The pdf (A1.18) is shown in Figure A.3.
1 0.8

f m ( -------------) & 2%kT '


3 -2
0.6 0.4 0.2

0.5

1.5

2.5

m ( --------- ( u 2 + u 2 + u 2 )) - 1 2 3 ' 3 & 2kT

1 -2

Figure A.3 Maxwellian probability density function of molecular velocities.

As with the 1-D pdf, the 3-D pdf is normalized so that the total probability over all velocity components is one.

$ # $ # $ # f du du du
1 2

3 = 1.

(A1.19)

To understand the distribution (A1.18) imagine a volume of gas molecules at some temperature T .

At a certain instant a snapshot of the motion is made and the velocity vector of every molecule in the gas sample is measured as indicated in the sketch above.

7/12/11

A1.6

bjc

Distribution of molecular velocities in a gas

The velocity components of each molecule dene a point in the space of velocity coordinates. When data for all the molecules is plotted, the result is a scatter plot with the densest distribution of points occurring near the origin in velocity space as shown in Figure A.4. u2

u1

u3
Figure A.4 Scatter plot of gas molecules in velocity space. Each point corresponds to the velocity vector of a molecule in the sample of gas.

The probability density function (A1.18) can be thought of as the density of points in Figure A.4. The highest density occurs near the origin and falls off exponentially at large distances from the origin. If the gas is at rest or moving with a uniform velocity the distribution is spherically symmetric since no one velocity direction is preferred over another. The one-dimensional probability distribution is generated by integrating the Maxwellian in two of the three velocities.
# #
2 2 2 m 32 m ( -------------) exp ( --------- ( u 1 + u 2 + u 3 )) du 2 du 3 & 2kT ' & 2 % kT '

f 1D =

# #

$$

(A1.20)

which can be rearranged to read


( mu 1 ) m 32 exp + ------------, f 1D = ( -------------) & 2 % kT ' & 2kT '
2 #

$
#

( mu 2 ) exp + ------------, du 2 & 2kT '

$
#

( mu 3 ) exp + ------------, du 3 . & 2kT '

(A1.21)

Carrying out the integration in (A1.21) the 1-D pdf is

bjc

A1.7

7/12/11

Mean molecular velocity

( mu 1 ) m 12 f 1D = ( -------------) exp + ------------ , . & 2 % kT ' 2kT ' & which is identical to (A1.14).

(A1.22)

A1.2 MEAN MOLECULAR VELOCITY


Any statistical property of the gas can be determined by taking the appropriate moment of the Maxwellian distribution. The mean velocity of molecules moving in the plus x 1 direction is,
# # #

u1 =

$ $

# # 0

m 32 2 2 2 m u 1 ( -------------) $ & 2 % kT-' exp ( ---------- ( u 1 + u 2 + u 3 )) du1 du 2 du 3 & 2kT '

(A1.23)

Note that the integration extends only over velocities in the plus x direction. When the integral is evaluated the result is, kT 1 2 u 1 = ( ---------- ) & 2 % m'
(A1.24)

Lets work out the ux in the 1-direction of molecules crossing an imaginary surface in the uid. This is the number of molecules passing through a surface of unit area in unit time.

The ux is simply kT 1 2 J 1 = nu 1 = n ( ---------- ) & 2 % m'


(A1.25)

At equilibrium, the number of molecules crossing per second in either direction is the same. Note that the molecular ux through a surface is the same regardless of the orientation of the surface.
7/12/11 A1.8 bjc

Distribution of molecular speeds

The mean square molecular speed is dened as the average of the squared speed over all molecules 2 2 2 2 u = u1 + u1 + u1 From the Maxwellian
2 u =
#

(A1.26)

$$
#

m 32 2 2 2 2 2 2 m ( -------------) $ & 2 % kT-' ( u 1 + u 2 + u 3 ) exp ( ---------- ( u 1 + u 2 + u 3 )) du1 du 2 du 3 & 2kT ' 3kT = --------m

(A1.27)

where the integration is from minus innity to plus innity in all three directions. The root-mean-square speed is. 3kT --------(A1.28) m This result is consistent with the famous law of equipartition which states that, for a gas of monatomic molecules, the mean kinetic energy per molecule (the energy of the translational degrees of freedom) is u rms = 1 2 3 E = -- mu = -- kT . 2 2
(A1.29)

2 u =

A1.3 DISTRIBUTION OF MOLECULAR SPEEDS


The Maxwellian distribution of molecular speed is generated from the full distribution by determining the number of molecules with speed in a differential spherical shell of thickness du and radius u in velocity space. The volume of the shell is dV u = 4 % u du and so the number of molecules in the shell is = u 1 + u 2 + u 3 . The Maxwellian probability distribution for the molecular speed is therefore, m 32 2 mu u exp ( --------- ) f u = 4 % u f = 4 % ( -------------) & 2 % kT ' & 2kT '
2 2 (A1.30) fdV u = 4 % u fdu where u
2 2

plotted below.
bjc A1.9 7/12/11

Pressure

0.35

fu ----------------------------------m 12 ( -------------) 4% & 2%kT '

0.3 0.25 0.2 0.15 0.1 0.05 0.5 1 1.5 2 2.5 3

m 12 ( --------- ) u & 2kT ' Figure A.5 Maxwell distribution of molecular speeds.

The mean molecular speed is


# #

u =

$
0

u f u du =

$
0

m 32 3 mu 4 % ( -------------) u exp ( --------- ) du = & 2kT ' & 2 % kT '

8kT --------- . %m

(A1.31)

In summary, the three relevant molecular speeds are, u1 = kT ---------- ; 2%m u rms = 3kT --------- ; m u = 8kT --------- . %m
(A1.32)

A1.4 PRESSURE
A1.4.1 KINETIC MODEL OF PRESSURE

Before we use the Maxwellian distribution to relate the gas pressure to temperature it is instructive to derive this relation using intuitive arguments. Consider a gas molecule conned to a perfectly elastic box. The molecule is moving randomly and collides and rebounds from the wall of the container preserving its momentum in the direction normal to the wall. In doing so the molecule undergoes the change in momentum

- p = 2 mu 1
7/12/11 A1.10

(A1.33)

bjc

Pressure

and imparts to the wall A 1 the momentum 2mu 1 . Suppose the molecule reaches the opposite wall without colliding with any other molecules. The time it takes for the molecule to bounce off A 2 and return to collide again with A 1 is 2L u 1 and so the number of collisions per second with A 1 is u 1 2L . x2 L L m u1 -u1 A1 L x1

A2

x3
Figure A.6 An elastic box containing an ideal gas

Thus the momentum per unit time that the molecule transfers to A 1 is Force on wall by one molecule = mu 1 L .
2 (A1.34)

Let N p be the number of molecules in the box. The total force on A 1 is the pressure of the gas times the area and is equal to the sum of forces by the individual molecules
Np 2

PL

1 = -L

i=1

mi ui

2 1

(A1.35)

where the index i refer to the ith molecule. If the molecules all have the same mass
p ( ) 2, Nm) + 1 ( -------- -----P = ui & V ' +N 1, & pi = 1 '

(A1.36)

bjc

A1.11

7/12/11

Pressure

The rst factor in (A1.36) is the gas density ! = N p m V . No one direction is preferred over another and so we would expect the average of the squares in all three directions to be the same.
Np Np Np

1 -----Np

i=1

ui

2 1 = -----1 Np

i=1

ui

2 1 = -----2 Np

i=1

ui

2 3

(A1.37)

The pressure is now written P = ! u rms 3 where


p p ( p ) 2 2 2 2, 1 + ui + ui (A1.39) u rms = -----ui + 1 1 1, N p+ &i = 1 ' i=1 i=1 The root-mean-square velocity dened by a discrete sum in (A1.39) is the same as that derived by integrating the Maxwellian pdf in (A1.28). Using (A1.28) and (A1.38) the pressure and temperature are related by

(A1.38)

! 3kT P = -- ( ---------) = ! RT 3& m '

(A1.40)

We arrived at this result using a model that ignored collisions between molecules. The model is really just a convenience for calculation. It works for two reasons; the time spent during collisions is negligible compared to the time spent between collisions and, in a purely elastic exchange of momentum between a very large number of molecules in statistical equilibrium, there will always be a molecule colliding with A 2 with momentum mu 1 while a molecule leaves A 1 with the same momentum.
A1.4.2 PRESSURE DIRECTLY FROM THE MAXWELLIAN PDF

The momentum ux at any point is


#

/ ij =

$$
#

m 32 2 2 2 m ( -------------) $ & 2 % kT-' ( nm )u i u j exp ( ---------- ( u 1 + u 2 + u 3 )) du1 du 2 du 3 & 2kT '

(A1.41)

7/12/11

A1.12

bjc

Pressure

For i 0 j the integral vanishes since the integrand is an odd function. For i = j the integral becomes
#

mn 2kT / 11 = ----------- ( ---------) 3 2& m ' %


#

( mu 1 ) 2 u 1 exp + ------------ , du 1 & 2kT '


2

( mu 1 ) exp + ------------ , du 1 & 2kT ' = nkT

( mu 1 ) exp + ------------ , du 1 & 2kT '

(A1.42)

similarly / 22 = nkT and / 33 = nkT . The uid density is

! = nm
and the gas constant is M w Ru Ru - k = ----- = -------- -------- = mR N Mw N where N is Avogadros number, N = 6.023 10 Thus
26

(A1.43)

(A1.44)

molecules kmole .

(A1.45)

! / 11 = / 22 = / 33 = nkT = ( --- ) mRT = ! RT = P & m'

(A1.46)

The normal stress derived by integrating the Maxwellian distribution over the molecular ux of momentum is just the thermodynamic pressure. P 0 0 0 P 0 0 0 P

/ ij = P 1 ij =

(A1.47)

bjc

A1.13

7/12/11

The mean free path

Recall the mean molecular speed u =


2

8kT % m . The speed of sound for a gas

is a = " P ! . Thus the speed of sound is directly proportional to the mean molecular speed.
12 12 12 ( " P) ( "%) ( " nkT ) --------- a = = = ----u & !' & 8' & nm ' (A1.48)

A1.5 THE MEAN FREE PATH


A molecule of effective diameter 2 sweeps out a collision path which is of diameter 2 2 . The collision volume swept out per second is %2 u . With the number of molecules per unit volume equal to n , the number of collisions per second experienced by a given molecule is 3 c = n ( %2 ) 8kT % m . The average distance that a molecule moves between collisions is, u rel u rel 4 = -------- = ---------------(A1.49) 2 3c n %2 u where the mean relative velocity between molecules is used. This accounts for the fact that the other molecules in the volume are not static; they are in motion. When this motion is taken into account, the mean free path is estimated as, 1 4 = -------------------(A1.50) 2 2n %2 Typical values for the mean free path of a gas at room temperature and one atmosphere are on the order of 50 nanometers (roughly 10 times the mean molecular spacing).
2 2

7/12/11

A1.14

bjc

Viscosity

A1.6 VISCOSITY
Consider the ux of momentum associated with the net particle ux across the plane y = y 0 .

y U(y) y
0
m

4 x 54 J

Figure A.7 Momentum exchange in a shear ow.

On the average, a particle passing through y 0 in the positive or negative direction had its last collision at y = y 0 4 or y = y 0 + 4 , where 4 is the mean free path. Through the collision process, the particle tends to acquire the mean velocity at that position. The particle ux across y 0 in either direction is J = n ( kT 2 % m )
12

. The net ux of x 1 - momentum in the x 2 -direction is


(A1.51)

dU / 12 = mJU ( y 0 4 ) mJU ( y 0 + 4 ) 6 2mJ 4 ------dy The viscosity derived from this model is kT 1 2 = 2mJ 4 = 2 !4 ( ---------- ) & 2 % m' In terms of the mean molecular speed this can be written as 1 8kT 1 2 1 - = -- !4 ( ---------) = -- ! u 4 2 & %m ' 2 The viscosity can be expressed in terms of the speed of sound as follows.

(A1.52)

(A1.53)

bjc

A1.15

7/12/11

Heat conductivity

1 8nkT 1 2 2 12 = -- !4 ( ------------ ) = ( ----- ) !a4 & %" ' 2 & % nm '

(A1.54)

This result can be used to relate the Reynolds number and Mach number of a ow. L ! UL UL R e = ---------- 6 ------- = M ( -- ) & 4' a4
12 (A1.55)

where L is a characteristic length of the ow and the factor ( 2 %" ) has been dropped. The ratio of mean free path to characteristic length is called the Knudsen number of the ow.

4 K n = -L

(A1.56)

A1.7 HEAT CONDUCTIVITY


The same heuristic model can be used to crudely estimate the net ux of internal energy (per molecular mass) across y 0 thus, dT Q 2 = mJ C v T ( y 0 4 ) mJ C v T ( y 0 + 4 ) 6 2mJ C v 4 -----dy The heat conductivity derived from this model is 1 - 7 = -- ! u 4 C v (A1.58) 2 The model indicates that the Prandtl number for a gas should be proportional to the ratio of specic heats. Cp C p P r = ---------- = ------ = " 7 Cv A more precise theory gives 4" P r = -------------- . 9" 5 the number of degrees of freedom of the molecular system.
(A1.59) (A1.57)

(A1.60)

which puts the Prandtl number for gases in the range 2 3 8 P r < 1 depending on

7/12/11

A1.16

bjc

Specific Heats, the law of equipartition

Notice that we got to these results using an imprecise model and without using the Maxwellian distribution function. In fact, a rigorous theory of the transport coefcients in a gas must consider small deviations from the Maxwellian distribution that occur when gradients of temperature or velocity are present. This is the so-called Chapman-Enskog theory of transport coefcients in gases.

1.8 SPECIFIC HEATS, THE LAW OF EQUIPARTITION


Classical statistical mechanics leads to a simple expression for C p and C v in terms of 9 , the number of degrees of freedom of the appropriate molecular model,

9+2 -R C p = ------------ ; 2

9 C v = -- R ; 2

9+2 " = ------------ . 9

(A1.61)

For a mass point, m , with three translational degrees of freedom, 9 = 3 , the energy of the particle is
2 1 2 1 2 1 E = -- mu + -- mv + -- mw 2 2 2 (A1.62)

where ( u, v, w ) are the velocities in the three coordinate directions. The law of equipartition of energy says that any term in E which is quadratic (proportional to a square) in either the position or velocity contributes ( 1 2 )kT to the thermal energy of a large collection of such mass points. Thus the thermal energy (internal energy) per molecule of a gas composed of a large collection of mass points is e = ( 3 2 )kT Over one mole of gas, N e = ( 3 2 )R u T internal energy is, e = ( 3 2 )RT where, recall R = R u molecular weight .
(A1.66) (A1.65) (A1.64) (A1.63)

where R u = Nk is the universal gas constant. On a per unit mass of gas basis the

bjc

A1.17

7/12/11

Diatomic gases

This is a good model of monatomic gases such as Helium, Argon, etc. Over a very wide range of temperatures, 5 C p = -- R ; 2 from near condensation to ionization. 3 C v = -- R 2
(A1.67)

A1.9 DIATOMIC GASES


A1.9.1 ROTATIONAL DEGREES OF FREEDOM

The simplest classical model for a diatomic molecule is a rigid dumb bell such as that shown in Figure A.8. In the classical theory for a solid object of this shape there is one rotational degree of freedom about each coordinate axis as shown in the gure. :2 , ;2

d1

m1

m2

d2

: 1 , ;1

:3 , ;3
Figure A.8 Rigid dumb bell model of a diatomic molecule

The energy of the rotating body is


2 1 2 1 2 1 E solid dumbell = -- I 1 : 1 + -- I 2 : 2 + -- I 3 : 3 2 2 2 The moment of inertia about the 2 or the 3 axes is (A1.68)

I 2 = I 3 = mr D . where
7/12/11 A1.18

(A1.69)

bjc

Diatomic gases

m1 m2 m r = -------------------m1 + m2

(A1.70)

is the reduced mass. The moment of inertia of the connecting shaft is neglected in (A1.69) where it is assumed that d 1, 2 << D so that the masses of the spheres can be assumed to be concentrated at the center of each sphere. In classical theory the energy (A1.68) is a continuous function of the three angular frequencies. But in the quantum world of a diatomic molecule the energies in each rotational degree of freedom are discrete. As in the classical case the energy varies inversely with the moment of inertia of the molecule about a given axis. When the Schroedinger equation is solved for a rotating diatomic molecule the energy in the two coordinate directions perpendicular to the internuclear axis is h 2K (K + 1) h 2K (K + 1) E 2, 3diatomic molecule = ( ----- ) ---------------------- = ( ----- ) ---------------------& 2%' & 2%' 2I 2 2, 3 2 ( mr D ) where h is Plancks constant h = 6.626 10 J-sec (A1.72) The atomic masses are assumed to be concentrated at the two atomic centers. The rotational quantum number K is a positive integer greater than or equal to zero. ( K = 0, 1, 2, 3, ). The law of equipartition can be used with (A1.71) to roughly equate the rotational energy with an associated gas temperature. Let 1 -- k < r 6 E 2, 3diatomic molecule . 2 Equation (A1.73) denes the characteristic temperature h 2 1 < r = ( ----- ) ------------------------& 2 %' 2 2k ( m r D )
(A1.73) 34 (A1.71)

(A1.74)

for the onset of rotational excitation in the two degrees of freedom associated with the 2 and 3 coordinate axes. In terms of the characteristic temperature, the rotational energy is E 2, 3diatomic molecule = K ( K + 1 )k < r .
(A1.75)

bjc

A1.19

7/12/11

Diatomic gases

Rotational parameters for several common diatomic species are given in Table 1.1. m r 10 kg
H2 N2 O2 CO 0.8393 11.629 13.284 11.392 27

D 10 M
0.74166 1.094 1.207 1.13 Table 1.1

10

< r K
87.28 2.89 2.08 2.77

The characteristic rotational excitation temperatures for common diatomic molecules are all cryogenic and except for hydrogen fall far below the temperatures at which the materials would liquefy. Thus for all practical purposes, with the exception of hydrogen at low temperature, diatomic molecules in the gas phase have both rotational degrees of freedom fully excited.
A1.9.2 WHY ARE ONLY TWO ROTATIONAL DEGREES OF FREEDOM EXCITED?

The discretized energy for the degree of freedom associated with rotation about the internuclear axis is h 2 ( ----- ) K ( K + 1 ) - ---------------------E 1diatomic molecule = & 2 %' 2I
1 (A1.76)

and the characteristic excitation temperature is h 2 1 < r1 = ( ----- ) ---------- . & 2 %' 2kI
1 (A1.77)

Almost all the mass of an atom is concentrated in the nucleus. This would suggest that the appropriate length scale, d , to use in (A1.76) might be some measure of the nucleus diameter. Consider a solid sphere model for the two nuclear masses. Then for a homonuclear molecule the characteristic rotational excitation temperature for the 1-axis degree of freedom is

7/12/11

A1.20

bjc

Diatomic gases

< r1

solid nuclear sphere

h 2 1 = ( ----- ) ---------------------------& 2 %' 2 2k ( md 5 )


14

(A1.78)

The diameter of the nucleus is on the order of 10 M four orders of magnitude small than the diameter of an atom. Thus one would expect

< r1

6 10 < r solid nuclear sphere

(A1.79)

an astronomically high temperature that would never be reached in a practical situation other than at the center of stars. Another approach is to ignore the nuclear contribution to the moment of inertia along the 1-axis but include the moment of inertia of the electron cloud. If we take all of the mass of the electron cloud and concentrate it in a thin shell at the atomic diameter the characteristic rotational temperature is

< r1

electron cloud hollow sphere

h 2 1 ( ----- ) ---------------------------------------------------------------------= & 2 %' 2 2k ( m electron cloud d atom 3 )


10

(A1.80)

Most atoms are on the order of 10 M in diameter which is comparable to the internuclear distances shown in Table 1.1. But the mass of the proton is 1835 times larger than the electron. So the mass of the nucleus of a typical atom with the same number of neutrons and protons is roughly 3670 times the mass of the electron cloud. In this model we would expect

< r1

electron cloud hollow sphere

6 3670 < r

(A1.81)

This puts the rotational excitation temperature along the 1-axis in the 8 to10, 000K range, well into the range where dissociation occurs. The upshot of all this is that because I 1 <<< I 2 or I 3 the energy spacing of this degree of freedom is so large that no excitations can occur unless the temperature is extraordinarily high; so high that the molecule is likely to be dissociated. Thus only two of the rotational degrees of freedom are excited in the range of temperatures one is likely to encounter with diatomic molecules. The energy of a diatomic molecule at modest temperatures is

bjc

A1.21

7/12/11

Diatomic gases

E = ET + ER
2 1 2 1 2 1 E T = -- mu + -- mv + -- mw . 2 2 2 2 1 2 1 E R = -- I 2 : 2 + -- I 3 : 3 2 2 At room temperature, (A1.82)

7 5 (A1.83) C p = -- R ; C v = -- R 2 2 From the quantum mechanical description of a diatomic molecule presented above we learn that C p can show some decrease below ( 7 2 )R at very low temperatures. This is because of the tendency of the rotational degrees of freedom to freeze out as the molecular kinetic energy becomes comparable to the rst excited rotational mode. More complex molecules with three-dimensional structure can have all three rotational degrees of freedom excited and so one might expect n = 6 for a complex molecule at room temperature.
A1.9.3 VIBRATIONAL DEGREES OF FREEDOM

At high temperatures C p can increase above ( 7 2 )R because the atoms are not rigidly bound but can vibrate around the mean internuclear distance much like two masses held together by a spring as indicated below. 7 m1 m2
x

The energy of a classical harmonic oscillator is


2 1 2 1 - E spring-mass = -- m r x + -- 7 x 2 2 (A1.84)

7/12/11

A1.22

bjc

Diatomic gases

where m r is the reduced mass, 7 is the spring constant and x is the distance between the mass centers. In the microscopic world of a molecular oscillator the vibrational energy is quantized according to h 1 E V = ( ----- ) : 0 ( j + --) & 2 %' & 2'
(A1.85)

where : 0 is the natural frequency of the oscillator and the quantum number j = 0, 1, 2, 3, . The natural frequency is related to the masses and effective spring constant by

:0 =

7 ----- . mr

(A1.86)

For a molecule the spring constant is called the bond strength. Notice that the vibrational energy of a diatomic molecule can never be zero even at absolute zero temperature. A characteristic vibrational temperature can be dened using the law of equipartition just as was done earlier for rotation. Let 1 -- k < v 6 E V . 2 Dene h :0 h 1 7 - < v = ( ----- ) ------ = ( ----- ) -- ----& 2 %' k & 2 %' k m r
(A1.88) (A1.87)

Vibrational parameters for several common diatomic species are given in Table 1.2

bjc

A1.23

7/12/11

Diatomic gases

. m r 10 kg
H2 N2 O2 CO 0.8393 11.629 13.284 11.392 27

7( N M )
570 2240 1140 1860 Table 1.2

< v K
6297 3354 2238 3087

The characteristic vibrational excitation temperatures for common diatomic molecules are all at high combustion temperatures. At high temperature seven degrees of freedom are excited and the energy of a diatomic molecule is E = ET + ER + EV
2 1 2 1 2 1 E T = -- mu + -- mv + -- mw 2 2 2 2 1 2 1 E R = -- I 1 : 1 + -- I 2 : 2 2 2 (A1.89)

1 2 1 2 - E V = -- mx + -- 7 x 2 2 where the various energies are quantized as discussed above. At high temperatures the heat capacities approach 9 7 (A1.90) C p = -- R ; C v = -- R 2 2 Quantum statistical mechanics can be used to develop a useful theory for the onset of vibrational excitation. The specic heat of a diatomic gas over a wide range of temperatures is accurately predicted by
2 Cp 7 A ( < v 2T ) B ------ = -- + ? --------------------------------- @ R 2 = Sinh ( < v 2T ) > (A1.91)

7/12/11

A1.24

bjc

Energy levels in a box

The enthalpy change of a diatomic gas at and above temperatures where the rotational degrees of freedom are fully excited is
h(T ) h(T 1) =

$T 1 C p dT

= R$

T T1

2 7 A ( < v 2T ) B -- + ? ---------------------------------- @ dT 2 = Sinh ( < v 2T ) >

(A1.92)

With T 1 set (somewhat articially) to zero, this integrates to ( <v T ) h(T ) 7 ----------- = -- + -------------------------------RT 2 ( <v T ) (e 1) & ' plotted below.
4.2

(A1.93)

h/RT
0.5 3.8 1 1.5 2

T/<v

3.6

Figure A.9 Enthalpy of a diatomic gas.

A1.10 ENERGY LEVELS IN A BOX


We discussed the quantization of rotational and vibrational energy levels but what about translation? In fact the translational energies for particle motion in the three coordinate directions are quantized when the particle is conned to a box. The wave function for a single atom of mass m contained inside a box with sides L x, L y, L z satises the Schroedinger equation
2 ih CD ( x, t ) h ----- -------------------- + ------------- E D ( x, t ) V ( x )D ( x, t ) = 0 . 2 2% Ct 8% m 2 (A1.94)

bjc

A1.25

7/12/11

Energy levels in a box

where h = 6.626068 10

34

m kg/sec is Plancks constant. The potential is

A F 0, 0<x<L x, 0 < y < L y, 0 < z < L z V ( x) = ? F #, on and outside the box =

(A1.95)

The solution must satisfy the condition D = 0 on and outside the walls of the box. For this potential (A1.94) is solved by a system of standing waves 8 G ( x, t ) = ( ----------------&L L L ' x y z
2%E k i ( ---------------) t 12 & h ' )

( n x %x) ( n y %y) ( n z %z) Sin + ----------- , Sin + ----------- , Sin + ---------- , (A1.96) & Lx ' & Ly ' & Lz '

where the quantized energy corresponding to the wave vector ( n x % n y % n z %) k = + -------- , -------- , --------, & L x L y Lz ' is
2 2) 2 2( 2 2 2 2 h + n x n y nz , h - E k = ------------- ( k x + k y + k z ) = ------ ----- + ----- + ----2 2 2, 8m + 2 8% m & L x L y Lz ' (A1.97)

(A1.98)

The corresponding frequency is Ek 3 = -------- . h


(A1.99)

The volume of the box is V = L x L y L z . The wave function is zero everywhere outside the box and must satisfy the normalization condition

$ $ $
0 0

L x L y Lz 0

G ( x, t )G* ( x, t ) dx dy dz = 1

(A1.100)

on the total probability that the particle is inside the box.

7/12/11

A1.26

bjc

Energy levels in a box

The implication of (A1.100) and the condition D = 0 on the walls of the box, is that the quantum numbers ( n x, n y, n z ) must be integers greater than zero. The energy of the particle cannot be zero similar to the case of the harmonic oscillator. For a box of reasonable size the energy levels are very closely spaced. A characteristic excitation temperature can be dened for translational motion at the lowest possible energy. Let 1 -- k < T 6 E T 2 where 3h E T = -------------------23 8mV The translational excitation temperature is
2 2 (A1.102) (A1.101)

3h < T = ----------------------- . (A1.103) 23 4mkV Take the monatomic gas helium for example. The characteristic translational excitation temperature is
<T
3 ( 6.626 10 ) 3.57 10 3h = ------------------------ = --------------------------------------------------------------------------------- = ------------------------------ K . (A1.104) 4m He kV
2 -3 2 34 2 14

He

4 ( 6.643 10

27

)1.38 10

2 - 23 3

2 -3

In general, the characteristic translational temperature is far below any temperature at which a material will solidify. Clearly translational degrees of freedom are fully excited for essentially all temperatures of a gas in a box of reasonable size.
A1.10.1 COUNTING ENERGY STATES

For simplicity let L x = L y = L z = L . In this case, the energy of a quantum state is


2 2 2 h E k = -------------------- ( n x + n y + n z ) 23 8mV 2 (A1.105)

bjc

A1.27

7/12/11

Energy levels in a box

Working out the number of possible states with translational energy of the monatomic gas less than some value E amounts to counting all possible values of the quantum numbers, n x , n y and n z that generate E k < E . Imagine a set of coordinate axes in the n x , n y and n z directions. According to (A1.98) a surface of constant energy is approximately a sphere (with a stair-stepped surface) with its center at the origin of the n x , n y and n z coordinates. The number of possible states with energy less than the radius of the sphere is directly related to the volume of the sphere. Actually only 1 8 th of the sphere is involved corresponding to the positive values of the quantum numbers. Figure A.10 illustrates the idea. ny
4 16

ny
32

ny

nz 4

nx

nz

nx
16 16

nz

nx
32 32

Figure A.10 Points represent quantum states in a box for energies that satisfy h 8 n x + n y + n z . Three cases are shown where the maximum value of the quantum numbers are 4, 16 and 32. 8EmV
23 2 2 2 2

Following Boltzmanns notation let W be the number of states.


2 2 32 2 2 32 1 4% 2 % 2 - W one atom = -- ( ----- ( n x + n y + n z ) ) = -- ( n x + n y + n z ) (A1.106) ' 6 8& 3

In terms of the volume of the box and the gas kinetic energy the number of states with energy E k < E is, according to (A1.105) and (A1.106)

% 8m 3 2 3 2 W one atom ( E, V ) = -- ( ------ ) VE 6 & h2 '

(A1.107)

7/12/11

A1.28

bjc

Energy levels in a box

Equation (A1.107) gives the number of possible energy states with energy less than E for a single atom. If we add a second atom to the box then the number of possible states lies within the positive region of a six dimensional sphere of radius R = 8mV
23

Eh .
(A1.108)

2 2 2 2 2 2 8mV E ------------------------ = ( n x1 + n y1 + n z1 + n x2 + n y2 + n z2 ) 2 h 3 6

23

The volume of a sphere in six dimensions is ( % 6 )R . For a box with N molecules the total number of possible energy states of the system lies within a sphere of 3N dimensions with volume
3N % --------------- R ( 3N ) ! ------& 2' 3N 2 (A1.109)

The volume of the positive region of this 3N dimensional sphere is % R 3N ( -- ) --------------- & ' ( 3N ) ! 2 ------& 2'
3N 2 (A1.110)

In terms of the energy, the number of distinct states with energy less than E is % 2m 3N 2 N 3N 2 V E W ( E, V , N ) = --------------------- ( ------ ) 3N ) & h 2 ' N! ( ------- ! & 2'
3N 2 (A1.111)

States with the same energy must not be counted more than once. Since the atoms are indistinguishable, the additional factor of N! in the denominator is needed to account for states that are repeated or degenerate. If the number of molecules in the volume V is more than a few tens or so, the number of states can be represented using the Stirling approximation for the factorial H! I H e
H H (A1.112)

bjc

A1.29

7/12/11

Energy levels in a box

Now
3N 2 5N 2 ( V ) N ( 4mE ) 3N 2 --------------W ( E, V , N ) = % e & N' & 2 ' (A1.113)

3h N The result (A1.113) is presented in Beckers Theory of Heat, page 135, equation 35.4. A more accurate version of the Stirling approximation is H! I ( 2%H )
1 2 H H

H e

which would give.


3N 2 5N 2 (A1.114)

% e V N 4mE 3N 2 W ( E, V , N ) = -------------------------------- ( --- ) ( -------------) & N' & 2 ' 6%N 3h N

Note that for values of N greater than a few hundred the number of states with energy less than E is virtually the same as the number of states with energy equal to E because of the extremely steep dependence of (A1.114) on N . All of the energy states are concentrated within a membrane thin spherical shell at the energy E .
A1.10.2 ENTROPY OF A MONATOMIC GAS IN TERMS OF THE NUMBER OF STATES

According to the Law of Equipartition and the development in the early partof this appendix, the internal energy of a monatomic gas is 2 1 3 E = -- mN u = -- kNT 2 2 The number of states can be expressed in terms of the temperature as W (T, V, N ) = %
3N 2 5N 2 ( V ) N ( 2mkT ) 3N 2 ---------------e & N' & 2 ' (A1.115)

(A1.116)

Take the logarithm of (A1.116). ( V 3 3 2%mk 5) Log ( W ( T , V , N ) ) = N + Log ( --- ) + -- Log ( T ) + -- Log ( ------------- ) + --, (A1.117) & N' 2 & 2 ' 2' 2 & h Differentiate (A1.117) and multiply both sides by Boltzmanns constant, k .

7/12/11

A1.30

bjc

Energy levels in a box

dW 3kN dT d(V N ) k -------- = ---------- ------ + kN -------------------W 2 T VN

(A1.118)

Recall R u = k N a where N a is Avogadros number. Equation (A1.118) now becomes 3R u dT dW d(V n) -------- = n --------- ------ + nR u ------------------k W Vn 2 T where the number of moles of gas in the box is n = N p N . The model we are considering is that of a monatomic gas with only three degrees of freedom. For such a gas the constant volume molar heat capacity is C v = 3R u 2 . Equation (A1.119) now reads dW dT d(V n) k -------- = nC v ------ + nR u ------------------W T Vn Recall the Gibbs equation dE P dS = ------ + ---dV T T
(A1.120) (A1.119)

(A1.121)

For an ideal gas with the equation of state PV = nR u T the Gibbs equation becomes dT dV dS = nC v ------ + nR u -----T V Comparing (A1.122) with (A1.120) leads to the conclusion that
(A1.122)

dW (A1.123) dS = k -------- . W Integrating (A1.123) leads to the famous Boltzmann relation for the entropy. S = kLog ( W ) + H Substutute (A1.113) into (A1.124). The entropy of the gas is ( V E 3 2) 5 3 4%m ) ( ( --- ) ( --- ) + -- + -- Log ( ---------- ) , S = kN + Log - - & 2 '' & & N' & N' ' 2 2 & 3h
(A1.124)

(A1.125)

bjc

A1.31

7/12/11

Energy levels in a box

The relation (A1.125) is known as the Sackur-Tetrode equation after its discoverers in the early 1900s. It can be used to derive the various equations that govern a gas. In terms of the temperature the entropy expression is equation (A1.117) multiplied by k . ( 32 V 3 2%mk 5) S = kN + Log ( ( --- ) ( T ) ) + -- Log ( ------------- ) + --, & 2 ' 2' & & N' ' 2 & h which agrees with equation 35.8 in Beckers Theory of Heat, page 136. The additive constant in (A1.126) is not in disagreement with the third law nor is the fact that (A1.126) becomes singular at T = 0 . The theory just developed is for a Boltzmann gas that necessarily implies temperatures well above absolute zero where the material is not in a condensed state and Boltzmann statistics apply. The constant is needed for the theory of condensation. The Sackur-Tetrode equation provides a remarkably accurate value of the entropy for monatomic gases. As an example, select helium at a pressure of 10 N m and temperature of 298.15K . At these conditions V N = 4.11641 10 The mass of the helium atom is m = 6.64648 10 entropy evaluated from (A1.26) is
27 5 2 26 3 (A1.126)

m .

kg using these values the

( 5 2 ( V ) ( 2%mkT ) 3 2) S --- ---------------------- = Log + e , = 15.1727 & N' & 2 ' kN & ' h

(A1.127)

The value provided in tabulations of helium properties is 15.17271576 . Similar results can be determined for other monatomic gases. According to the third law H = 0 and S = kLog ( W )
(A1.128)

is the absolute entropy of the system. Consider the state of a single atom contained in a box in the limit T * 0 . From the solution to the Schroedinger equation, (A1.98), the least energy that the atom can have is

7/12/11

A1.32

bjc

Energy levels in a box

3h E 111 = -------------------- . (A1.129) 23 8mV This is a tiny but nite number. In this model of a single atom gas, as the temperature goes to zero the atom in the box falls into the lowest quantum state so the entropy is S = kLog ( 1 ) = 0 . We derived (A1.128) using the energy states in an ideal monatomic gas rather than follow the more general proof of Boltzmann. The remarkable thing about (A1.128) is that it applies to a wide range of systems other than ideal gases including solids and liquids.

bjc

A1.33

7/12/11

Energy levels in a box

7/12/11

A1.34

bjc

Você também pode gostar