Download Internal Degrees of Freedom of Molecules ( + problem 33)

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Wave–particle duality wikipedia , lookup

Rigid rotor wikipedia , lookup

X-ray photoelectron spectroscopy wikipedia , lookup

Scalar field theory wikipedia , lookup

Renormalization group wikipedia , lookup

Particle in a box wikipedia , lookup

Hydrogen atom wikipedia , lookup

Canonical quantization wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Atomic theory wikipedia , lookup

Two-dimensional nuclear magnetic resonance spectroscopy wikipedia , lookup

Rutherford backscattering spectrometry wikipedia , lookup

T-symmetry wikipedia , lookup

Franck–Condon principle wikipedia , lookup

Molecular Hamiltonian wikipedia , lookup

Transcript
Internal Degrees of Freedom of Molecules
Adiabatic Theory of Molecule
For the internal dynamics of molecules it is of crucial importance that nuclei are much heavier than
electrons. This leads to the adiabatic picture of the nuclei motion with respect to the motion of
electrons. In the adiabatic picture, the coordinates of nuclei play the role of external parameters with
respect to the state of the electronic sub-system, that is they play the same role as the adiabatic
parameter λ in our consideration of the adiabatic process. Here, however, we are dealing with more
subtle situation where nuclei coordinates behave as external parameters with respect to the electrons
only, not to each other. Mathematically, the adiabatic picture implies the following structure of the
molecule wavefunction
Ψ(r, R) = ψ(R) ΦR (r) ,
(1)
where r and R stand for all the electron and nuclei coordinates, respectively. The function ψ(R) is
the wavefunction of nuclei. The function ΦR (r) is the wavefunction of electrons at fixed positions of
all nuclei. It obeys the Schrödinger equation
(R)
Hel
(R)
Φ = Eel Φ .
(2)
(R)
Here Hel is the electronic Hamiltonian in which the nuclei at fixed positions play the role of external
potentials. Both the wavefunction ψ and the eigenvalue Eel thus depend on R. The nuclei motion is
described by the effective Schrödinger equation
(kin)
Hnuc
ψ + Veff ψ = E ψ ,
(3)
(kin)
where Hnuc is the standard kinetic energy term and
(R)
Veff (R) = V (R) + Eel
,
(4)
is the effective potential energy equal to the sum of bare potential energy, V (R), (Coulomb repulsion
between nuclei) and the electronic energy as a function of R. Note that it is entirely due to the
(R)
attractive term Eel that nuclei bind into a molecule.
The effective adiabatic potential Veff (R) is very rigid in the sense that the typical inter-nuclei
separation turns out to be much smaller than its quantum uncertainty. Hence, the inter-nuclei
separations are good quantum numbers characterizing the structure of the molecule. Consider some
estimates proving this fact. First we estimate the electronic energy as
(R)
Eel
∼
h̄2
h̄2
,
∼
me R 2
me a2B
(5)
where me is the electron mass and aB is the Bohr radius. Then we estimate the strength of the
effective adiabatic potential Veff (R). To this end we replace it with a harmonic potential with respect
to the relative coordinates of the nuclei. This is a quite accurate approximation because relative
coordinates of nuvlei do not change significantly with respect to their equilibrium values and we can
expand Veff (R) in the Taylor series in the vicinity of the equilibrium configuration. The equivalent
harmonic potential with respect to one of the relative coordinates, R, then reads
Veff (R) = const − (k/2)(R − R0 )2 ,
1
(6)
where R0 is the equilibrium position, and k is the spring constant, which we estimate as
(R)
k ∼
∂ 2 Eel
∂R2
Eel
Eel
∼ 2 .
R02
aB
∼
(7)
Given the spring constant, we readily estimate the frequency of the vibrations
q
ωvib ∼
k/m∗
(8)
(m∗ is the typical nuclei mass) and the typical inter-level separation for the atomic vibrations
µ
Evib = h̄ωvib ∼ Eel
me
m∗
¶1/2
.
(9)
To estimate the quantum mechanical uncertainty, ∆R, of the coordinate R, we just need to recall the
result for the ground-state uncertainty of the coordinate of the harmonic oscillator of the mass m∗
and the spring constant k. Using the relation between ∆R and Evib ,
h̄2
∼ Evib ,
m∗ (∆R)2
we find
µ
∆R
R
¶2
µ
∼
me
m∗
(10)
¶1/2
¿ 1.
(11)
Let us compare inter-level separations of vibrational and rotational degrees of freedom. For the
typical inter-level separation of the rotational levels we have (only the kinetic part of the Hamiltonian
is relevant for the rotations)
Erot
h̄2
∼ Evib
∼
m∗ R 2
µ
me
m∗
¶1/2
∼ Eel
me
.
m∗
(12)
We see that
Erot ¿ Evib ¿ Eel .
(13)
Finally, we establish the orders of magnitudes of realistic values. As is known, Eel ∼ (1 ÷ 10) eV ,
where
1 eV = 1.16 × 104 K .
(14)
Note that for our purposes it is most convenient to measure inter-level separations in Kelvins, since
this immediately yields the ranges of the temperatures when corresponding degrees of freedom get
excited. So,
Eel ∼ (104 ÷ 105 ) K .
(15)
A reasonable estimate for the parameter
p
me /m∗ is ∼ 10−2 leads us to
Evib ∼ (102 ÷ 103 ) K ,
(16)
Erot ∼ (1 ÷ 10) K .
(17)
2
Spherical Quantum Rotator
To a very good approximation, the rotational degrees of freedom of a molecule decouple from the
other degrees of freedom, and thus can be treated as a separate subsystem—quantum rotator. Here
we consider the simplest case of a spherical quantum rotator. This case corresponds to an axially
symmetric molecule of (distinguishable) atoms and, in particular, to the case of a diatomic molecule
of two non-identical atoms.
The eigenstates of the spherical rotator are labelled by two quantum numbers, the angular momentum, l = 0, 1, 2, . . . , and the projections of the angular momentum (onto some given axis),
ml = −l, . . . , −1, 0, 1, . . . , l. For a given l there are (2l + 1) different values of ml . The energy
spectrum of the spherical rotator is
h̄2 l(l + 1)
El =
,
(18)
2I
where the parameter I is called moment of inertia. For example, in the case of a rigid diatomic
molecule
m1 m2
I=
R2
(rigid diatomic molecule) ,
(19)
m1 + m2 0
where m1 and m2 are the masses of the atoms, and R0 is the separation between the two nuclei. Note
that energies do not depend on ml , which means that each energy level is 2l + 1-fold degenerate.
It is convenient to introduce the parameter
h̄2
,
2I
θ=
(20)
which characterizes typical energy scale associated with rotations at l ∼ 1, and to write
El = θ l(l + 1) .
(21)
For the partition function we thus have
Z =
∞ X
l
X
e−(θ/T ) l(l+1) =
l=0 ml =−l
∞
X
(2l + 1) e−(θ/T ) l(l+1) .
(22)
l=0
There are two characteristic cases:
T Àθ
classical regime ,
(23)
T ¿θ
ultra-quantum regime .
(24)
and
Classical regime. In the classical regime the smallness of the parameter θ/T in Eq. (35) renders
the discreteness of l irrelevant and we may replace the summation with integration. With the same
2
accuracy, we
p do not distinguish between (2l + 1) and 2l, and also between l(l + 1) and l [because
typical l ∼ T /θ À 1].
Z=
Z ∞
0
−(θ/T ) l2
dl 2l e
=
Z ∞
0
dx e−(θ/T ) x = T /θ .
(25)
Then, by the standard thermodynamical procedure we find
F = −T ln(T /θ) ,
3
(26)
S = ln(T /θ) + 1 ,
(27)
E=T ,
(28)
C=1.
(29)
Ultra-quantum regime. In this regime the temperature is much smaller than the separation
between the ground state and the first excited state. Hence, with the exponential accuracy in the
parameter θ/T the molecule does not rotate—rotational degrees of freedom are frozen out.
We can easily take into account the main exponential correction. To this end we keep the second
term in the series (35), neglecting the rest of the terms [which are exponentially smaller]. We have
Z = 1 + 3e−2(θ/T ) ,
and then find
h
F = −T ln 1 + 3e−2(θ/T )
i
≈ −3T e−2(θ/T ) ,
(30)
(31)
S = [3 + 6(θ/T )] e−2(θ/T ) ,
(32)
E = 6θ e−2(θ/T ) ,
(33)
C = 12(θ/T )2 e−2(θ/T ) .
(34)
Gas of Molecules
The statistics of internal degrees of freedom of molecules is decoupled from the statistics of the
translational degrees of freedom, simply because the translational motion—the motion of the center
of mass of a molecule—is dynamically independent from the internal motions. Hence, for a dilute (=
non-interacting) gas of molecules the partition function factorizes into the product of the partition
functions of translational and internal motions:
Z(V, T ) = Ztr (V, T ) Zint (T ) .
(35)
Note that Zint is V -independent. Correspondingly, we have
F (V, T ) = Ftr (V, T ) + Fint (T ) ,
(36)
and see that the equation of state for the gas of molecules is the same as for the mono-atomic gas,
since Fint (T ) is V -independent and thus cannot change the expression for the pressure:
µ
∂F
P =−
∂V
¶
T
µ
∂Ftr
=−
∂V
¶
= N T /V .
(37)
T
Similarly, for the entropy we have
S(V, T ) = Str (V, T ) + Sint (T ) ,
4
(38)
and from the fact that Sint depends only on T conclude that for any heat capacity Cξ [ξ is either V ,
or P , or any other fixed quantity] we get
(tr)
Cξ = Cξ
+ C (int) ,
(39)
where
dSint
.
dT
The contribution of the internal degrees of freedom to Cξ is one and the same for any ξ.
C (int) = T
(40)
Classical limits for internal motions. The vibrational degrees of freedom are just harmonic
oscillators. Hence, we conclude that each vibrational degree contributes at T À Evib the amount
T = (T /2 + T /2) to the energy and, correspondingly, 1 = (1/2 + 1/2) to the heat capacity (recall that
each half comes independently from kinetic and potential energies). So, if νvib is the total number of
vibrational degrees of freedom, we have
Evib = νvib T
(T À Evib ) ,
cvib = νvib
(41)
(T À Evib ) .
(42)
(Here we use a small letter c to emphasize that we are dealing with the heat capacity of a single
molecule rather than that of the whole gas.)
Let us find the total contribution to the energy of all the internal degrees of freedom in the classical
limit. The total number of degrees of freedom is 3ν, where ν is the total number of nuclei. As we
remember, in the classical statistics the kinetic and potential energies are statistically independent.
Moreover, the total kinetic energy is always equal to (3/2)νT . In our case, the potential energy
consists only of νvib harmonic potentials (in terms of relative distances), each of these potentials
contributing T /2 to the energy [1/2 to the heat capacity]. Subtracting (3/2)T associated with the
translational motion of the molecule, we arrive at the energy and heat capacity of the internal degrees
of freedom:
νvib
3(ν − 1)
T +
T
(T À Evib , Erot ) ,
(43)
Eint =
2
2
3(ν − 1)
νvib
cint =
+
(T À Evib , Erot ) .
(44)
2
2
The number νvib can be related to ν by subtracting from 3ν [the total number of degrees of freedom]
the 3 translational degrees and the rotational degrees. In a general case, there are 3 rotational degrees
of freedom (3 independent angles), while in the axially symmetric case there are only 2 rotational
degrees, since the rotation around the symmetry axis does not change the configuration of particles
and thus is not a degree of freedom. We thus have
(
νvib =
3ν − 5
3ν − 6
(axial symmetry) ,
(no axial symmetry) .
(45)
In the temperature regime Erot ¿ T ¿ Evib the vibrational degrees of freedom are still frozen
out, while the rotational degrees of freedom are already in their classical limit. Hence, it is reasonable
to separate in the equations (43) and (44) vibrational contributions from the rotational ones, so
that we could use their classical limits independently. Subtracting (41) and (42) from (43) and (44),
respectively, and taking into account (45), we get
(
Erot =
T
(3/2) T
(axial symmetry)
(no axial symmetry)
5
(T À Erot ) ,
(46)
(
crot =
1
3/2
(axial symmetry)
(no axial symmetry)
(T À Erot ) .
(47)
The results for the axially symmetric case are already known to us from the classical limit of the
spherical quantum rotator.
The fact that in the classical limit the heat capacity of internal degrees of freedom is just a
constant number has a simple implication in terms of entropy, and, correspondingly adiabats. Indeed,
if C (int) = N cint is just a temperature-independent number then
Sint = N cint ln T + const = N ln T cint + const .
(48)
This means that the exponents of the adiabats will get an extra term equal to the heat capacity
of the internal degrees of freedom with the elegant final results in terms of the total (translational
plus internal) heat capacities per molecule:
V T cV = V0 T0cV
(T, V )−adiabat ,
T cP
T0cP
=
P
P0
(T, P )−adiabat .
(49)
(50)
Problem 33. Typical vibrational frequency of the H2 molecule is ∼ 6000 K. Typical rotational level separation
for H2 is less than 100 K. Estimate CV and CP per molecule for the H2 gas at room temperature. [Note that I
use a convenient Stat. Mech. jargon: First, I measure energy in Kelvins, which causes no problem since Kelvins
are unambiguously related to Joules by the Boltzmann constant. Secondly, I also use Kelvins for measuring
frequencies which basically means that I do not distinguish between ω and h̄ω. Hence, when I say that the
frequency ω = 6000 K, I actually mean that h̄ω = 6000 K = 6000 kB J.]
6