Website by fretchen

WelcomeBlogAMOImageGen
Injected
WalletConnect
MetaMask

Lecture 17 - Rotation and Vibration of Molecules

Loading...

We will study the existance of vibrational and rotational levels in molecules. It allows us investigate the transitions of molecules and introduce the Franck-Condon principle. Finally, we will study how such intruiging molecules are used for the study of the permanent electric dipole moment of the electron.

We studied during the last two lectures the properties the electronic structure. For atoms the next step was the analysis of the transition rules to understand the spectrum. However, for molecules the electronic transition directly couple to the vibrational and rotational motion of the nuclei, which will have to study first.

A short reminder on nuclear motion

We discussed diatomic molecules, with NN electrons bound to the nuclei. The full Hamiltonian of the molecule could be written in the following fashion: H^=T^e+T^N+V(RA^,RB^,r^1,,r^N)\hat{H} = \hat{T}_e + \hat{T}_N + V(\hat{\mathbf{R}_\mathrm{A}}, \hat{\mathbf{R}_\mathrm{B}},\hat{\mathbf{r}}_1,\cdots, \hat{\mathbf{r}}_N) T^e\hat{T}_e describes the kinetic energy of the electrons, T^N\hat{T}_N the kinetic energy of the nuclei and VV the coupling between them. The we decomposed the full wavefunction over a nuclear part and an electronic part: Ψ(RA,RB,r1,,rN)=ψe(RA,RB,r1,,rN)ψn(RA,RB)\Psi(\mathbf{R}_\mathrm{A}, \mathbf{R}_\mathrm{B}, \mathbf{r}_1,\cdots, \mathbf{r}_N) = \psi_e(\mathbf{R}_\mathrm{A}, \mathbf{R}_\mathrm{B}, \mathbf{r}_1,\cdots, \mathbf{r}_N)\cdot \psi_n(\mathbf{R}_\mathrm{A}, \mathbf{R}_\mathrm{B}) This allowed us to decouple nicely the two motions and study the properties of the electronic potentials first. In the Born-Oppenheimer approximation we obtained:

T^NψN+Ee(Ra,Rb)ψN=EψN\hat{T}_N \psi_N + E_e (\vec{R}_\textrm{a}, \vec{R}_\textrm{b}) \psi_N = E \psi_N

with

T^N=12MaΔRa12MbΔRb,\hat{T}_N = - \frac{1}{2 M_\textrm{a}} \Delta_{\vec{R}_\textrm{a}} - \frac{1}{2 M_\textrm{b}} \Delta_{\vec{R}_\textrm{b}},

the total energy EE and the masses of the individual atoms MaM_\textrm{a} and MbM_\textrm{b}. For the electronic energy EeE_e, only R=RaRb\vec{R} = \vec{R}_\textrm{a} - \vec{R}_\textrm{b} matters. We then calculated the molecular potential curves Ee(R)E_e(R), which differ for each electronic configuration, discussed in the last lecture and sketched once more in Fig. 1.

The molecular potential curves obtained from the Born-Oppenheimer Approximation.

In the center of mass frame we can tranform and get:

(12MΔR+Ee(R))ψn(R)=Eψn(R),\left( - \frac{1}{2M} \Delta_{\vec{R}} + E_e (\vec{R}) \right) \psi_\textrm{n} (\vec{R}) = E \psi_\textrm{n} (\vec{R}),

R\vec{R} is spherically symmetric, and

M=MaMbMa+MbM = \frac{M_\textrm{a} \cdot M_\textrm{b}}{M_\textrm{a} + M_\textrm{b}}

is the reduced mass. This means that we can separate the angular and radial motion to obtain:

ψn(R,θ,φ)=1RS(R)Ylm(θ,φ)\psi_\textrm{n} (R,\theta,\varphi) = \frac{1}{R} S(R) \cdot Y_l^m (\theta,\varphi)

They describe the rotational and vibrational levels of the nucleus.

Rotations

If we assume a "rigid" molecule where the distance between the atoms is fixed, the rotational energy is simply given by:

Erot(R)=J(J+1)2MR2(a.u.)E_\text{rot} (R) = \frac{J\cdot (J+1)}{2M R^2}\, (\text{a.u.})

where MM is the reduced mass of the nuclei in atomic units and JJ is the angular momentum quantum number. The factor MR2MR^2 represents the moment of inertia. For more complex atoms the relationship is not quite as simple and the rotational energy is typically described by the moment of interia IijI_{ij}. The Hamiltonian for this rotation reads then:

H^rot=Jx22Ixx+Jy22Iyy+Jz22Izz\hat{H}_{rot} =\frac{J_x^2}{2I_{xx}}+\frac{J_y^2}{2I_{yy}}+\frac{J_z^2}{2I_{zz}}

The molecule H2H_2 has then a rotational frequency ω/2π=31012J(J+1)Hz\omega/2\pi =3\cdot 10^{12}\cdot J(J+1) \text{Hz}.

Vibrations

As already known from the hydrogen atom we can use the angular solutions to discuss the radial solutions. We have to solve now:

(12Md2dR2+Ee(R)+12MJ(J+1)R2)S(r)=EvibS(R)\left( \frac{1}{2M} \frac{d^2}{dR^2} + E_e (R) + \frac{1}{2M} \frac{J(J+1)}{R^2} \right) S(r) = E_\text{vib} S(R)

If the extension from the minimum (see 2) is small, we can approximate it by a harmonic potential. We can then find a vibrational energy Evib=ω0(ν+12)ν=0,1,E_\text{vib} = \omega_0 (\nu+\frac{1}{2})\, \nu=0,1,\cdots The harmonic expansion around the minimum reads:

EeE0+12Mω02(RR0)2E_e \approx E_0 + \frac{1}{2} M \omega_0^2 (R-R_0 )^2

For the example of H2, we get ω/2π1014\omega/2\pi \sim 10^{14} Hz.

Internuclear potential. In the limit of a harmonic expansion around the minimum, the vibrationally excited states are equidistant.

A better approximation of the vibrational level structure than the simple harmonic oscillator is the Morse potential.

The Morse potential

In this case we approximate the molecular potential curves by:

Ee(R)Vmorse(R)Vmorse(R)=hcDe(1eax)2 with a=k2hcDex=RR0E_e(R) \approx V_\text{morse}(R)\\ V_\text{morse}(R) =hcD_e(1-e^{-ax})^2\text{ with }a =\sqrt{\frac{k}{2hcD_e}}\\ x = R-R_0

Its particular usefulness stems from the fact that it is still analyitically solvable and we obtain:

Evib=(ν+12)ω(ν+12)2ωxeωxe=a22ME_{vib}=(\nu+\frac{1}{2})\hbar \omega-(\nu+\frac{1}{2})^2 \hbar \omega x_e\\ \omega x_e = \frac{a^2\hbar}{2M}

xex_e is then called the anharmonicity parameter.

Molecular transitions

We are now ready to discuss the different transitions that might appear in the spectrum. And we will work our way through the different levels of energy as we will see that they are all coupled.

Rotational transitions

We will start out with the transitions of the lowest frequency, the rotational transitions. So, we would like to know if it is possible to transition from a state ϵ,J,MJ\left|\epsilon, J, M_J\right\rangle to another state ϵ,J,MJ\left|\epsilon, J', M_J'\right\rangle, where ϵ\epsilon describes the electronic and vibrational degree of freedom. This means that we have to calculate as usual the the electric dipole moment:

ϵ,J,MJDϵ,J,MJ\left\langle\epsilon, J', M_J'\right| \vec{D}\left|\epsilon, J, M_J\right\rangle

Within the Born-Oppenheimer approximation electronic and rotational degree of freedom decouple and we can write:

ϵ,J,MJDϵ,J,MJ=J,MJϵDϵJ,MJ\left\langle\epsilon, J', M_J'\right| \vec{D}\left|\epsilon, J, M_J\right\rangle = \left\langle J', M_J'\right| \left\langle\epsilon\right|\vec{D}\left|\epsilon\right\rangle\left| J, M_J\right\rangle

This electric dipole transitions were forbidden in atoms as they do not have a permanent electric dipole moment. However, hetero-nuclear atoms can have such a permanent electric dipole moment, they are called polar molecules. Examples are alkali-alkali molecules like NaK, NaCs, KRb whose permenanent electric dipole moment can be up 3ea03 ea_0. It follows that:

  • Pure rotational transitions exist in polar molecules.

The transition rules are di-atomic molecules: ΔJ=±1\Delta J = \pm 1 and ΔMJ=0,±1\Delta M_J= 0, \pm1. For more complex molecules these transition rules can vary quite substantially as the rotational degree of freedom might have to be described by an additional quantum number.

Vibrational transitions

In the next step, we would like to understand the transitions between different vibrational levels. Hence, we are investigating the electric dipole moment

ϵ,νDϵ,ν=νDϵν\left\langle\epsilon, \nu'\right| \vec{D}\left|\epsilon, \nu\right\rangle= \left\langle\nu'\right| \vec{D}_\epsilon\left|\nu\right\rangle

The evaluation is now not quite as simple as for the rotational degree of freedom as both ν\nu and ϵ\epsilon will influence the length of the molecule, they both directly depend on RR. We can develop the electric dipole moment as a function of distance from the equilibrium and write then:

νDϵν=ν(Dϵ(0)+dDϵdxx+)ν=dDϵdxνxν+\left\langle\nu'\right| \vec{D}_\epsilon\left|\nu\right\rangle= \left\langle\nu'\right| \left(\vec{D}_\epsilon(0)+ \frac{d\vec{D}_\epsilon}{dx}x+\cdots\right)\left|\nu\right\rangle\\ = \frac{d\vec{D}_\epsilon}{dx}\left\langle\nu'\right|x\left|\nu\right\rangle+\cdots

So vibrational transistions will only happen in molecules for which the permanent electric dipole changes as a function of distance. Once again they are non-existant in homo-nuclear molecules.

Vibronic transitions

At this stage, we are ready to discuss electronic transitions. If we are performing an electronic transition this also implies a change on the molecular potential curve as indicated in Fig. 3. Imagine now the transition of the ground state molecular branch (called the X branch) to a higher electronic shell (called A, B, C, ...). Such a transition will happen at constant internuclear radius as they are much faster than the nuclei motion. This implies that an electronic transition will typically excite the molecule into a high vibrational branch. The dipole moment is then proportional too:

ϵ,νDϵ,νDϵ,ϵνν\left\langle\epsilon', \nu'\right| \vec{D}\left|\epsilon, \nu\right\rangle\approx \vec{D}_{\epsilon, \epsilon'} \left\langle\nu'\right| \left|\nu\right\rangle
The Franck-Condon principle for a simple toy model.

The factor S(ν,ν)=νν2S(\nu, \nu')=|\left\langle\nu\right|\left|\nu'\right\rangle|^2 is then called the Franck-Condon factor and it describes the strength of the transitions.

It is exactly this coupling of different hierarchies that makes the molecular spectra so rich and also extremely tough to control.

Can we get into the groundstate ?

Given all the complexities of molecules it seems non-trivial to find a scheme that gets them into the ground state. For atoms laser cooling has proven very efficient as we will discuss later. However, it mainly adresses the cooling of external degrees of freedom. In molecules a significant amount of energy its in the rotational and vibrational levels. In this connection, a beautiful solution has been demonstrated in Ni 2008.

The scheme is visualized in Fig. 4.

TProduction of groundstate molecules of K + Rb. Figure is taken from Ni 2008

In a first step the atoms are cooled and then associated to a highly excited molecule in the a^3^Σ\Sigma state. From there the atom has to be transferred down in to the ground state g\left|g\right\rangle. A direct thransfer is not possible as the Franck-Condon factors do not allow for it. Another path is to go through an intermediate level (here the 23Σ^3 \Sigma level), which has overlap with both of them. However, this level has typically overlap with plenty of other levels and a finite lifetime. How can we then optimize the transfer ? The idea is to use the concept of dark states in the triplet of {i,e,g}\{i, e, g\}.

The dark states in three level systems

We can visualize the idea of the dark state transfer through the following Hamiltonian:

H^=Ω1(ie+ei)+Ω2(ge+eg)\hat{H}= \Omega_1\left(\left|i\right\rangle\left\langle e\right|+\left|e\right\rangle\left\langle i\right|\right)+\Omega_2\left(\left|g\right\rangle\left\langle e\right|+\left|e\right\rangle\left\langle g\right|\right)

We can rewrite it as:

H^=(Ω1i+Ω2g)e+e(Ω1i+Ω2g)Be+eBB=Ω1i+Ω2gΩ12+Ω22\hat{H}= (\Omega_1\left|i\right\rangle+\Omega_2\left|g\right\rangle)\left\langle e\right|+\left|e\right\rangle(\Omega_1\left\langle i\right|+\Omega_2\left\langle g\right|)\\ \propto\left|B\right\rangle\left\langle e\right|+\left|e\right\rangle\left\langle B\right|\\ \left|B\right\rangle= \frac{\Omega_1\left|i\right\rangle+\Omega_2\left|g\right\rangle}{\sqrt{\Omega_1^2+\Omega_2^2}}

So in the three level scheme the excited state is always could to the so-called bright state, which is a coherent superposition of g\left|g\right\rangle and i\left|i\right\rangle. The orthogonal state is the dark state:

D=Ω1gΩ2iΩ12+Ω22BD=0\left|D\right\rangle= \frac{\Omega_1\left|g\right\rangle-\Omega_2\left|i\right\rangle}{\sqrt{\Omega_1^2+\Omega_2^2}}\\ \langle B| D\rangle = 0

Now we can also discuss the transfer sequence non as STIRAP (stimulated Raman adiabatic passage).

STIRAP

STIRAP transfers the loosely bound molecules coherently into the groundstate without ever passing through the lossy excited level. It has the following steps:

  1. The dressing laser Ω2\Omega_2 is ramped on. The initial i\left|i\right\rangle is now the dark state.

  2. The coupling laser Ω1\Omega_1 is ramped on, while the laser Ω2\Omega_2 is ramped down. This transfers the i\left|i\right\rangle adiabatically into the state g\left|g\right\rangle, which is the dark state for fully switched of Ω2\Omega_2.

The molecules are now in the groundstate with a transfer efficiency of roughly 50%50\%.

Measurement of the electron electric dipole moment

Despite their complexity, molecules can be an enormously powerful tool for precision measurements as you might find in this or this review. The test of the existance of a permanent electric dipole moment (electron edm) of the electron is one of these tests.

What does does the existance of electron edm actually mean ? We have already discussed quite heavily the existance of a permanent edm for polar molecules. The amplitude of their dipole moment is in the order of a few ea0ea_0, which is also the natural unit for the induced edm of atoms. One could now also imagine that the electron itself has an edm, which is aligned with its spin De=dese\vec{D}_e = d_e \vec{s}_e. The standard model actually predicts such a permanent electron edm, but only of the amplitude de1030ea0d_e \approx 10^{-30}ea_0, which is fantastically small . However, the search continues as most extensions of the standard model actually predict substantially higher values as summarized in Fig. 5. As we can see the most precise measurments are actually performed in very heavy di-atomic molecules.

In these molecules the electron 'feels' enormous effective electric fields, which can reach the several GV/cm regime.

Search for the permanent electric dipole moment. Figure is taken from here

The search for the dipole moment is then testing the dependence of the electron energy:

E±=±(μB0+deE)E_\pm = \pm(\mu B_0 + d_e E)

This energy difference can be read out through Ramsey spectroscopy. Switching the electric field allows then to switch the frequency difference by δω=4deE\hbar \delta \omega = 4d_e E. Only an upper limit is known up to now de<8.7ecm|d_e|< 8.7 e cm.