Tải bản đầy đủ (.pdf) (10 trang)

Basic Theoretical Physics: A Concise Overview P36 pptx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (312.28 KB, 10 trang )

48.2 The Entropy of an Ideal Gas from the Microcanonical Ensemble 365
The Stirling approximation was used again, since
Ω(3N) ≈

2eπ
3N

3N
2
.
Using
N!

=

N
e

N
one thus obtains:
s

0
≡ ln


4πm
h
2

3


2
·e
5
2
·V
0
U
3
2
0

.
The entropy constant therefore only depends on the type of gas via the log-
arithm of the particle mass m.
The above result is reasonable because, with the factor N, it explicitly
expresses the additivity for the entropy of an ideal gas, which (as already
mentioned) depends crucially on the permutation factor
1
N!
for identical par-
ticles. Furthermore the law
1
T
=
∂S
∂U
yields the relation between U and T :
U =
3
2

Nk
B
T.
Similarly
p
T
=
∂S
∂V
yields pV = Nk
B
T . Only the expression for the chemical potential μ is
somewhat less apparent, but we shall require it later: From
μ
T
= −
∂S
∂N
it follows that
μ =
5
2
k
B
T − T ·
S
N

U + pV − T · S
N

.
Thus μ ≡
G
N
, i.e. the chemical potential μ is identical to the free enthalpy
G = U + pV − T · S
per particle. This identity is generally valid for a fluid system.
366 48 Canonical Ensembles in Phenomenological Thermodynamics
48.3 Systems in a Heat Bath: Canonical and Grand
Canonical Distributions
In the previous section we treated closed systems. Now we shall concentrate
on closed systems that consist of a large system, a so-called heat bath, to
which a small partial system (which is actually the system of interest) is
weakly coupled. For a canonical ensemble the coupling refers only to energy,
and only serves to fix the temperature. For a so-called grand canonical en-
semble on the other hand not only exchange of energy takes place with the
large system but also particles are exchanged. In this case not only the tem-
perature T of the small system is regulated by the heat bath but also the
chemical potential μ.
In the next section we shall consider how the transition from a micro-
canonical ensemble to a canonical and grand canonical ensemble is achieved
mathematically.
For a canonical ensemble the appropriate variable of state is the Helmholtz
free energy F (T,V,N). For a magnetic system it is the Gibbs free energy
F
g
(T,V,H,N). They can be obtained from the corresponding partition func-
tions
Z(T,V,N):=


j
e
−βE
j
(V,N)
and Z

(T,V,H,N):=

j
e
−βE
j
(V,H,N)
,
where
F (T,V,N)=−k
B
T · ln Z(T,V,N)and
F
g
(T,V,H,N)=−k
B
T · ln Z

(T,V,H,N) .
In a grand canonical ensemble not only the energy fluctuates but also the
number of particles of the “small system”
ˆ


(k)
j
= E
(k)
j
ψ
(k)
j
and
ˆ

(k)
j
= N
k
ψ
(k)
j
.
One therefore has, in addition to the energy index j, a particle-number index
k. Thus, in addition to the reciprocal temperature
β =
1
k
B
T
the chemical potential μ appears as a further distribution parameter. Both
parameters control the expectation values, i.e.,

ˆ

N
β,μ
=


j,k
N
k
p
(k)
j

β,μ
and
48.4 From Microcanonical to Canonical and Grand Canonical Ensembles 367
U = 
ˆ
H
β,μ
:=


j,k
E
(k)
j
p
(k)
j


β,μ
, where
p
(k)
j
(β,μ)=
e
−β

E
(k)
j
−μN
k

Z(β,μ)
,
with the grand canonical partition function
Z(β,μ):=

j,k
e
−β

E
(k)
j
−μN
k


.
The grand canonical Boltzmann-Gibbs distribution p
(k)
j
is therefore very simi-
lar to the canonical Boltzmann-Gibbs distribution. In particular the grand
canonical partition function Z is related to the grand canonical thermody-
namic potential Φ in a similar way as the free energy F (T,V,N) is related to
the usual partition function Z(T,V,N):
Φ(T,V,μ)=−k
B
T · ln Z(T,V,μ) .
The quantity Φ is the Gibbs grand canonical potential; phenomenologically it
is formed from the free energy by a Legendre transformation with respect to
N:
Φ(T,V,μ)=F(T,V,N(T,V,μ)) − μN , and dΦ = −pdV − N dμ − SdT.
48.4 From Microcanonical to Canonical
and Grand Canonical Ensembles
For an ergodic
2
system one can calculate the results for observables
ˆ
A
1
,which
only involve the degrees of freedom of the small system “1”, according to the
microcanonical distribution for E
I
≈ U − ε
i

.
3
2
A classical system in a given energy range U −dU<E≤ U is called ergodic if for
almost all conformations in this energy region and almost all observables A(p, q)
the time average
1
t
0
t
0
R
0
dtA(p(t), q(t)) for t
0
→∞is almost identical with the
so-called ensemble average: A(p, q) =
R
U−dU<H(p,q)≤U
d
f
pd
f
qA(p,q)
R
U−dU<H(p,q)≤U
d
f
pd
f

q
. Most fluid
systems are ergodic, but important non-ergodic systems also exist, for example,
glasses and polymers, which at sufficiently low temperature often show unusual
behavior, e.g. ageing phenomena after weeks, months, years, decades or even
centuries, because the investigated conformations only pass through untypical
parts of phase space, so that for these systems application of the principles of
statistical physics becomes questionable.
3
Here and in the following we shall systematically use small letters for the small
system and large letters for the large system (or heat bath). For example, U ≈
E
I
+ ε
i
.
368 48 Canonical Ensembles in Phenomenological Thermodynamics
A
1
 =

U− dU−ε
i
<E
I
≤U−ε
i
ψ
i
|

ˆ
A
1
ψ
i
N
(2)
(U −ε
i
)

U− dU−ε
i
<E
I
≤U−ε
i
N
(2)
(U − ε
i
)
. (48.1)
In the following we shall omit the indices
1
and
(2)
.
We now introduce a Taylor expansion for the exponent of
N(U − ε

i
) ,
viz:
N(U − ε
i
) ≡ e
S(U−ε
i
)
k
B
=e
S(U)
k
B
·e

ε
i
k
B
·
dS
dU
·e

ε
2
i
d

2
S
2k
B
dU
2
· .
The first term is a non-trivial factor, the second term on the right-hand-side
of this equation, gives e

ε
i
k
B
T
; one can already neglect the next and following
terms, i.e., replace the factors by 1, as one sees, for example, for N →∞in
the term
ε
2
i
k
B
d
2
S
dU
2
, with U ≈
3

2
Nk
B
T, thus obtaining e

ε
2
i
d
3Nk
B
dT
(
1
k
B
T
)
→ 1 ,
i.e., if one uses a monatomic ideal gas as heat bath and replaces factors
e

const.
N
by unity. Inserting this Taylor expansion into the above formula one obtains
the Boltzmann-Gibbs distribution for a canonical system. One may proceed
similarly for the case of a grand canonical ensemble.
49 The Clausius-Clapeyron Equation
We shall now calculate the saturation vapor pressure p
s

(T ), which has already
been discussed in the context of van der Waals’ theory; p
s
(T ) is the equilib-
rium vapor pressure at the interface between the liquid and vapor phase of
a fluid. We require here the quantity
dp
s
dT
.
There are two ways of achieving this.
a) The first method is very simple but rather formal. Inversely to what is
usually done it consists of replacing differential quotients where necessary
by quotients involving differences, i.e. the non-differentiable transition
function from gas to liquid state from the Maxwell straight line is ap-
proximated by a gently rounded, almost constant transition function in
such a way that for the flat parts of a curve one may equate quotients of
differences with corresponding differential quotients.
Having dealt with mathematical aspects, we shall now consider the physics
of the situation: We have V = N
1
v
1
+ N
2
v
2
,wherev
1
and v

2
are the atomic
specific volumes in liquid and vapor phase respectively, and N
1
and N
2
are the
corresponding numbers of molecules. We therefore have ΔV = ΔN ·(v
2
−v
1
),
since N = N
1
+ N
2
is constant. For N
2
→ N
2
+ ΔN we also have N
1

N
1
− ΔN, and thus from the first law:
ΔU = δQ + δA = l · ΔN −pΔV ,
with the (molecular) specific latent heat of vaporization l(T )(≈ 530 cal/g
H
2

O
).
Using the Maxwell relation, which is essentially a consequence of the second
law, we obtain

∂U
∂V

T
=
ΔU
ΔV
= T
∂p
∂T
− p = T
dp
S
dT
− p
s
.
Thus
∂p
∂T
= ,or
dp
s
(T )
dT

=
l(T )
T · (v
2
− v
1
)
. (49.1)
370 49 The Clausius-Clapeyron Equation
This is known as the Clausius-Clapeyron equation. We shall now derive some
consequences from this equation, and in doing so we must take the sign into
account. In the case of water boiling, everything is “normal” provided one
is not in the close vicinity of the critical point: v
2
(vapor)  v
1
(liquid), and
thus from (49.1) we obtain
dp
s
dT

l
T · v
2
, and with v
2

k
B

T
p
s
:
dp
s
dT

l ·p
s
k
B
T
2
, i.e., p
s
(T ) ≈ p
0
· e

l
k
B
T
,
with constant p
0
. As a result there is a very fast drop in saturation vapor
pressure with increasing temperature.
There is no peculiarity here with regard to the sign; however, the Clausius-

Clapeyron equation (49.1) is valid not only for a boiling transition but also
for melting. For water-ice transitions, v
2
, the atomic specific volume of the
liquid, is 10% smaller than v
1
, the atomic specific volume of the ice phase
1
.
As a result of this,
dp
s
dT
=
l
T · (v
2
− v
1
)
is now negative, in agreement with the anomalous behavior of the phase
diagram of H
2
O mentioned earlier.
We now come to a second derivation of the Clausius-Clapeyron equation:
b) The method is based on an ideal infinitesimal Carnot process,whichone
obtains by choosing for a given liquid or gas segment in the equation of
state the line p
s
(T ) corresponding to the Maxwell construction as the

lower Carnot path (i.e. T
2
≡ T ), whereas one chooses the saturation
pressure line p
s
(T + ΔT )astheupper Carnot path (i.e. T
1
≡ T + ΔT ).
We then find
ΔA =



1


2


p
s
dV = p
s
·(v
2
− v
1
)ΔN
!
=

ΔT
T
· Q
1
,
since
η =
ΔA
Q
1
=
ΔT
T
.
Using
Q
1
= l · ΔN and ΔA = Δp
s
· (v
2
− v
1
)ΔN =
ΔT
T
· Q
1
,
the Clausius-Clapeyron equation is obtained: (49.1).

These derivations imply – as we already know – that the Maxwell relations,
the second law, and the statement on the efficiency of a Carnot process are
all equivalent, and that in the coexistence region the straight-line Maxwell
section, e.g., p
s
(T ), is essential.
1
Chemists would again prefer to use the specific molar volume.
50 Production of Low
and Ultralow Temperatures;
Third Law of Thermodynamics
Low temperatures are usually obtained by a process called adiabatic demag-
netization. Ultralow temperatures are achieved (in Spring 2004 the record was
T
min
=0.45 ×10
−9
Kelvin) in multistage processes, e.g., firstly by adiabatic
demagnetization of electron-spin systems, then by adiabatic demagnetization
of nuclear spins, thirdly by laser cooling, and finally by evaporation methods.
(Many small steps prove to be effective.) Evaporation cooling is carried out on
atomic and molecular gas systems, mainly gases of alkali atoms, that are held
in an electromagnetic “ trap ”. The phenomenon of Bose-Einstein condensa-
tion is currently being investigated on such systems at extreme temperatures
(
<

10
−7
K and lower powers of ten). This will be discussed later. In 2001

the Nobel Prize was awarded for investigations of the Bose-Einstein conden-
sation of ultracold gases of alkali atoms (see below). These investigations
could only be performed after it had been discovered how to obtain ultralow
temperatures in a reproducible and controllable manner.
1
Next we shall consider the production of low temperatures in general. The
techniques usually depend on “ x-caloric effects ”, e.g., the magnetocaloric
effect. We shall consider the following examples:
a) Gay-Lussac’s experiment on the free expansion of a gas from a container
(see above). This occurs at a constant internal energy, such that

dT
dV

U
= −
∂U
∂V
∂U
∂T
.
With the Maxwell relation
∂U
∂V
= T
∂p
∂T
− p
and van der Waals’ equation of state
p = −

a
v
2
+
k
B
T
v −b
we obtain

dT
dV

U
= −
a
c
(0)
v
v
2
,
1
Nobel Prize winners: Cornell, Ketterle, Wiemann.
372 50 Production of Low and Ultralow Temperatures; Third Law
i.e. the desired negative value (see above). In this connection we should
recall that the exact differential
dU =
∂U
∂T

dT +
∂U
∂V
dV
!
=0.
b) The Joule-Thomson effect has also been discussed above. This involves
a pressure drop at constant internal enthalpy per particle.Oneobtains
the expression

dT
dp

I/N
= −
∂I/N
∂p
∂I/N
∂T
= =
2a
k
B
T
− b
c
(0)
p
·(1 − )
,

i.e., giving a negative value above and a positive value below the so-called
inversion temperature T
Inv.
.Thus,forT<T
Inv.
(this is the normal case)
a drop in temperature occurs for a reduction in pressure.
c) Thirdly we shall consider the magnetocaloric effect or the phenomenon
of temperature reduction by adiabatic demagnetization, i.e., dH<0for
constant entropy S(T,H). Here one obtains as above:

dT
dH

S
= −
∂S
∂H
∂S
∂T
,
so that one might imagine reducing the temperature indefinitely, if the
entropy S(T,H) behaved in such a way that for T = 0 at finite H also
S(0,H)werefinite,with
∂S
∂H
< 0 .
One would then only need to magnetize the magnetic sample in a first
stage (step 1) isothermally (e.g., H → 2H) and subsequently (step 2) to
demagnetize it adiabatically, i.e. at constant S, in order to reach absolute

zero T = 0 immediately in this second step. This supposed behavior of
S(T,H) is suggested by the high temperature behavior:
S(T,H) ∝ a(T ) −
b
H
2
.
However, it would be wrong to extrapolate this behavior to low temper-
atures.
In fact, about 100 years ago the third law of thermodynamics was proposed by
the physico-chemist Walter Nernst. This is known as Nernst’s heat theorem,
which can be formulated, as follows:
Let S(T,X) be the entropy of a thermodynamic system, where X repre-
sents one or more of the variables of state, e.g., X = V , p , m
j
or H. Then,
for X>0, the limit as S(T → 0,X) is zero; and the convergence to zero is
such that the absolute zero of temperature in Kelvin, T =0, is unattainable
50 Production of Low and Ultralow Temperatures; Third Law 373
Fig. 50.1. Third Law of Thermodynam-
ics (schematically). The low temperature
behavior of the entropy S between 0 and
4 units is presented vs. the absolute tem-
perature T (here between 0 and 1.2 units);
only the 2nd and 4th curve from above
(i.e., with S(T =0)≡ 0) are realis-
tic, whereas the 1st and 3rd lines repre-
sent false extrapolations suggested by the
high-T asymptotes
in a finite number of steps. In the case of, for example, adiabatic demagneti-

zation this results in a countably-infinite number of increasingly small steps
(see Fig. 50.1):
Figure 50.1 shows the qualitative behavior of the entropy S(T,H)of
a paramagnetic system as a function of T for two magnetic field strengths.
The first and third curves from the top correspond to extrapolations sug-
gested by high-temperature behavior; but they do not give the true behavior
for T → 0. This is instead represented by the second and fourth curves, from
which, for the same high-temperature behavior, we have for all H =0,as
postulated by Nernst:
S(T =0,H) ≡ 0 .
The third law, unattainability of absolute zero in a finite number of steps
–whichisnot a consequence of the second law – can be relatively easily
proved using statistical physics and basic quantum mechanics, as follows.
Consider the general case with degeneracy, where, without loss of gener-
ality, E
0
= 0. Let the ground state of the system be g
0
-fold, and the first
excited state g
1
-fold; let the energy difference (= E
1
− E
0
)beΔ(X). Then
we obtain for the x-caloric effect:

dT
dX


S
= −
∂S
∂X
∂S
∂T
, where S = −
∂F
∂T
and F = −k
B
T · ln Z,
with the following result for the partition function:
Z = g
0
+ g
1
· e
−βΔ
+ .
Elementary calculation gives
S(T,X)
k
B
=lng
0
+
g
1

g
0
·
Δ
k
B
T
· e

Δ
k
B
T
+ ,
where the dots describe terms which for k
B
T  Δ can be neglected. If one
assumes that only Δ, but not the degeneracy factors g
0
and g
1
, depend on
374 50 Production of Low and Ultralow Temperatures; Third Law
X, it follows strictly that

dT
dX

S
≡ T ·

∂Δ
∂X
Δ
+ ,
since the exponentially small factors
∝ e
−βΔ
in the numerator and denominator of this expression cancel each other out. In
any case, for T → 0 we arrive at the assertion of unattainability of absolute
zero. Furthermore, we find that the assumption, S → 0, in Nernst’s heat
theorem is unnecessary. In fact, with g
0
≡ 2 for spin degeneracy of the ground
state, one obtains:
S(T =0,H ≡ 0) = k
B
ln 2(=0).
In spite of this exception for H ≡ 0, the principle of unattainability of abso-
lute zero still holds, since one always starts from H =0,whereS(0,H)=0.
In this respect one needs to be clear how the ultralow temperatures men-
tioned in connection with Bose-Einstein condensation of an alkali atom gas
are achieved in a reasonable number of steps. The deciding factor here is that
ultimately only the translational kinetic energy of the atoms is involved, and
not energetically much higher degrees of freedom. Since we have
M

v
2

T

2
=
3k
B
T
2
,
the relevant temperature is defined by the mean square velocity of the atoms,
k
B
T =
M ·

v
2

T
3
,
where we must additionally take into account that the relevant mass M is
not that of an electron, but that of a Na atom, which is of the order of
0.5 × 10
5
larger. One can compare this behavior with that of He
4
,whereat
normal pressure superfluidity (which can be considered as some type of Bose-
Einstein condensation for strong interaction) sets in at 2.17 K, i.e., O(1) K.
The mass of a Na atom is an order of magnitude larger than that of a He
4

atom, and the interparticle distance δr in the Na gas considered is three
to four orders of magnitude larger than in the He
4
liquid, so that from the
formula
k
B
T
c


2
2M(δr)
2
one expects a factor of ∼ 10
−7
to ∼ 10
−9
, i.e. temperatures of
O

10
−7

to O

10
−9

K

are accessible.

×