Rigid Rotor#

What you need to know

  • The rigid rotator system is introduced as a prototype for the quantization of rotational degrees of freedom in molecules. The spherical coordinate system is introduced out of the necessity of taking advantage of the spherical symmetry of the problem, which leads to a reduction of dimensionality.

  • Solving the Schrodinger equation in spherical coordinates results in eigenfunctions in the form of spherical harmonics. Energy eigenvalues are found to be degenerate for one of the quantum numbers.

  • To compute rotational spectra we need to know moment of inertia of a molecules which for diatomics takes simple form \(I=\mu r^2\) where \(\mu\) is reduced mass and \(r\) is distance.

  • For polyatomic molecules may require more elaborate calculations to account for spatial mass distribution.

  • Connection with microwave spectroscopy is shown where spectral lines are predicted to occur in equal intervals equal \(2\tilde{B}\).

  • The selection rule is established via the recursion relation of spherical harmonics and is \(\Delta J\pm 1\) and \(M_J =0, \pm 1\).

  • Coupling of vibrational degrees leads to rovibronic transitions and necessitates the inclusion of vibrational quantum numbers for a more accurate account of transitions.

  • Rotatioal motion can lead to change in bond length, this is known as centrifugal effect and can be accounted for by adding an extra centrifgual correction term to rigid rotor model.

Classical picture: Rotating dumbbell#

DeD0

Fig. 30 Conservation of angular momentum: In the absence of torque angular momentum of conservative system remains constant. This has implications for rotational motion, for instance you can rotate faster if you decreese moment of inertia and vice versa.#

The rigid rotor is a model of a rotating dumbbell: two unequal masses held together via a rigid stick. The system is not acted upon by any external potential; hence the only energy is the kinetic energy of rotation:

\[ K=\frac{m_1 v_1^2}{2}+\frac{m_2 v_2^2}{2}=\frac{m_1 r_1^2+m_2 r^2_2}{2}\omega^2 \]

Where we have plugged in \(v_1=\omega r_1\) and \(v_2=\omega_2 r\) velocities of rotation of two masses rotating with frequency \(\omega\). The classical mechanical problem of two masses is once again reducible to a single reduced mass \(\mu\) rotating around constant radius \(r=r_1+r_2\) rotating around center of mass \(m_1 r_1=m_2 r_2\).

\[ K=\frac{I \omega^2}{2}=\frac{L^2}{2I} \]

Where \(L=I \omega\) is the angular momentum, the \(I=\mu r^2\) is moment of inertia and \(\mu=\frac{m_1 m_2}{m_1+m_2}\)

Quantum rigid rotor and angular momentum operator#

  • The hamiltonian for the rigid rotor model is the kinetic energy operator of an effective mass \(\mu\) wchih rotates around sphere of radius \(r=const\). To incorporate constraint \(r=const\) it is more convenient to adopt spherical coordinates \((x,y,z)\rightarrow (r,\theta,\phi)\). To the full laplacian in spherical coordinates is:

  • In spherical coordinates Hamitlonian is more conveniently expressed in terms of angular momentum operator as opposed to linear momentum operator:

\[ \hat{H}=-\frac{\hbar^2}{2\mu}\nabla_{x,y,z}^2 = -\frac{\hbar^2}{2\mu r^2}\nabla_{\theta,\phi}^2=\frac{\hat{L}^2}{2I} \]
  • Where \(I=\mu r^2\) is the moment of inertia and where identified the angular momentum operator as: $\( \hat{L}= -i\hbar \nabla_{\theta,\phi} \)$

Quantum numbers \((J,M_J)\) for quantizing \((\theta,\phi)\) coordiante pair.#

  • Having written down hamitlonain we now solve it anticipating two quantum numbers for two coordinates. The eigenfunctions turn out to be well known special functions called spherical harmonics \(Y(\theta,\phi)\):

\[ \hat{H}Y(\theta, \phi)=E_{J,m}Y(\theta,\phi) \]
  • We are once again able to separat two angular variables and solve the resulting ODE exactly.

  • We expect energy to depend on two quantum numbers \(J\) and \(M_J\) which quantize roational motion acorss \(\theta\) and \(\phi\) angles.

Rotational states of molecules are quantized#

  • Solving a rigid rotor problem, we find that eigenvalues depend only on the quantum number \(J\). This makes each energy level degenerate with respect to \(2J+1\) values assumed by \(M_J\) quantum number.

\[ \boxed{E_J = \frac{\hbar^2}{2I}J(J+1)=BJ(J+1)} \]
  • Where we have defined \(B=\frac{h^2}{8\pi^2 I}\) rotational constant with units of energy.

  • Quantization in this equation arises from the cyclic boundary condition rather than the potential energy, which is identically zero.

  • There is no rotational zero-point energy (\(J = 0\) is allowed).

  • The ground state rotational wavefunction has equal probability amplitudes for each orientation.

Example

What are the reduced mass and moment of inertia of \(H^{35}Cl\)? The equilibrium internuclear distance \(R_e\) is 127.5 pm. What are the values of \(L, L_z\) and \(E\) for the state with \(J = 1\)? The atomic masses are: \(m_H = 1.673470 \cdot 10^{-27}\) kg and \(m_{Cl} = 5.806496 \cdot 10^{-26}\) kg.

Rotational spectra of diatomic molecules#

  • We assume that the molecule is a rigid rotor, which means that the molecular geometry does not change during molecular rotation. We have solved this problem already

  • Energies are typically expressed in wavenumber units (\(cm^{-1}\) although the basic SI unit is \(m^{-1}\)) by dividing \(E\) by \(hc\). The use of wavenumber units is denoted by including a tilde sign above the variable (e.g., \(\tilde{\nu}\)). The rotational energies expressed in wavenumbers are given by:

\[\boxed{\tilde{E}_r(J) = \frac{E_r}{hc} = \tilde{B} J(J+1)}\]
  • Where the rotational constant is usually expressed in \(cm^{-1}\) units and is given by:

\[{\tilde{B} = \frac{h}{8\pi^2Ic}}\]

Selection rules#

The rotational constant defines the rotational energy levels for a rigid diatomic molecule. When the molecule is in electronic state \(e\) and vibrational state \(v\), the total wavefunction is written as \(\psi = \psi_e\psi_v\psi_{J,M}\). The transition moment between two different rotational levels \(J,M\) and \(J',M'\):

\[{\int\int\int\psi_e^*\psi_v^*\psi_{J',M'}\hat{\mu}\psi_e\psi_v\psi_{J,M}d\tau_ed\tau_{rot}d\tau_{vib}}\]

where \(\hat{\mu}\) is the transition dipole operator and only the rotational wavefunction has change. The electronic part gives the permanent dipole moment:

\[{\mu_0^{(e)} = \int\psi_e^*\hat{\mu}\psi_ed\tau_e}\]

Therefore we can reduce the last equation to:

\[{\int\int\psi_v^*\psi^*_{J',M'}\mu_0^{(e)}\psi_v\psi_{J,M}d\tau_{rot}d\tau_{vib}}\]

The vibrational part just gives the dipole moment for the molecule in vibrational state \(v\) and we can write:

\[{\int\psi_{J',M'}^*\mu_0\psi_{J,M}d\tau_{rot}}\]
  • The rotational transition can only occur if this integral has a non-zero value. Clearly \(\mu_0\) must be non-zero for the transition to occur, which means that the molecule must have a permanent dipole moment for the rotational transition to occur.

  • For example, homonuclear diatomic molecules like \(H_2\) and \(O_2\) will not show pure rotational spectra. Heteronuclear molecules show pure rotational spectra.

Using the known properties for spherical harmonics, one can show the following selection rule holds for the rigid rotor model:

\[\boxed{\Delta J = J' - J = \pm 1}\]

Since photons have one unit of angular momentum, the above rule can be understood in terms of angular momentum transfer. The transition frequencies between the rotational levels are given by (\(J = 0,1,2,...\)):

\[{\tilde{\nu} = \tilde{E}_r(J + 1) - \tilde{E}_r(J) = \left((J+1)(J+2) - J(J+1)\right)\tilde{B} = 2\tilde{B}(J+1)}\]

The successive line positions in the rotational spectrum are given by \(2\tilde{B}, 4\tilde{B}, 6\tilde{B},...\). Note that molecules with different atomic isotopes have different moments of inertia and hence different positions for the rotational lines.

DeD0

Fig. 31 Rigid rotor model predicts evenlys spaced spectral lines#

Population of states#

Another factor that affects the line intensities in a rotational spectrum is related to the thermal population of the rotational levels. Thermal populations of the rotational levels is given by the Boltzmann distribution (for a collection of molecules):

\[{f_J = \frac{g_Je^{-hc\tilde{E}_r(J) / (k_B T)}}{\sum\limits_{J'}g_{J'}e^{-hc\tilde{E}_r(J') / (k_B T)}} = \frac{g_Je^{-hc\tilde{E}_r(J) / (k_B T)}}{q}}\]
  • The \(q\) is called the partition function and \(g_J = 2J + 1\) corresponds to the degeneracy count of state \(J\). A useful comparison of thermal energy is given by \(kT\) and if the energy of a state is much higher than this, it will not be thermally populated.

  • one expects the intensities to first increase as a function of the initial state \(J\), reach a maximum, and then decrease because the thermal populations decrease. In an absorption experiment, one can see the thermal populations of the initial rotational levels.

Note: For systems, where the rotational degrees of freedom may exchange identical nuclei, an additional complication arises from the symmetry requirement for the nuclear wavefunction. Recall that bosons must have symmetric wavefunctions and fermions antisymmetric. We will not discuss this in more detail here.

Example

Calculate the relative populations of the first five rotational levels of the ground vibrational state of \(H^{35}\)Cl at 300 K. The ground vibrational state rotational constant \(B_0 = 10.44\) cm\(^{-1}\).

Ro-vibrational spectra, R, P and Q branches**#

DeD0

Fig. 32 A cartoon depiction of a ideal rovibrational spectrum.#

  • Often times we are interested in transitions among rotational levels that accompany excitation from ground vibrational state \(v=0\rightarrow v=1 \). This can be described by combing rigid rotor and harmonic oscillator models:

\[ E_{v,J} = \tilde{\omega}(v+1/2)+\tilde{B}J(J+1) \]
  • Since at room temperatures molecules mostly occupy vibrational ground state we are interested in rotational transitions taking place between ground (v=0) and the first excited (v=1) vibrational states.

  • Rigid rotor model predicts different frequencies for absorption and emission transitions between any two rotational states
    $\( \tilde{\nu}_{\Delta J=\pm1} = E_{v+1,J\pm1} - E_{v,J} = \tilde{\omega} \pm 2\tilde{B}(J+1) \)$

  • The transitions with \(\Delta J=+1\) are called R branch: \(\tilde{\nu}_{\Delta J=+}=\tilde{\omega} + 2\tilde{B}(J+1)\)

  • The transitions with \(\Delta J=-1\) are called P branch: \(\tilde{\nu}_{\Delta J=+}=\tilde{\omega} - 2\tilde{B}(J+1)\)

  • The Q-branch \(\Delta J =0\) is predicted to be absent because it is forbidden by the selection rule of the rigid rotor model.

Beyond rigid rotor 1#

DeD0

Fig. 33 A cartoon depiction of a ideal rovibrational spectrum.#

DeD0

Fig. 34 A cartoon depiction of a real rovibrational spectrum.#

Centrifugal distortion#

In reality molecules are not rigid rotors and one must consider the coupling between \(H_{rot}\) and \(H_{vib}\). Classically thinking, with increasing rotational motion, the chemical bond stretches due to centrifugal forces, which increases the moment of inertia, and consequently, the rotational energy levels come closer together. It can be shown that this can be accounted for by including an additional term the energy expresison for rigid rotor:

\[{\tilde{E}_r(J) = \tilde{B}J(J+1) - \tilde{D}J^2(J+1)^2}\]
  • The \(\tilde{D}\) is the centrifugal distortion constant (\(cm^{-1}\)). Note that both \(\tilde{B}\) and \(\tilde{D}\) are positive. When the centrifugal distortion is taken into account, the rotational transition frequencies are given by:

\[{\tilde{\nu} = \tilde{E}_r(J+1) - \tilde{E}_r(J) = 2\tilde{B}(J+1) - 4\tilde{D}(J+1)^3\textnormal{ where }J=0,1,2,...}\]

Example

Measurement of pure rotational spectrum of H\(^{35}\)Cl molecule gave the following positions for the absorption lines:

\[\tilde{\nu} = \left(20.794\textnormal{cm}^{-1}\right)\left(J+1\right) - \left(0.000164\textnormal{cm}^{-1}\right)\left(J+1\right)^3\]

What is the equilibrium bond length and what is the value of the centrifugal distortion constant?

Rovibronic coupling#

  • As a diatomic molecule vibrates, its bond length changes. Since the moment of inertia is dependent on the bond length, it too changes and, in turn, changes the rotational constant B. We assumed above that B of R(0) and B of P(1) were equal, however they differ because of this phenomenon.

  • The v dependence is captured via the following expression showing that rotational constant is a linearly decreasing function of v!

\[ B_v = B_e-\alpha_e(v+1/2) \]
  • Where \(B_e\) is the rotational constant for a rigid rotor and \(\alpha_e\) is the rotational-vibrational coupling constant. The information in the band can be used to determine \(B_0\) and \(B_1\) of the two different energy states as well as the rotational-vibrational coupling constant, which can be found by the method of combination differences.

The R branch with rovibronic coupling:

\[ \tilde{\nu}_{\Delta J=+1} = \tilde{\omega} + 2\tilde{B_1}+(3\tilde{B_1}-\tilde{B}_0)J+(\tilde{B_1}-\tilde{B}_0)J^2 \]

The P brnach with rovibronic coupling: $\( \tilde{\nu}_{\Delta J=-1} = \tilde{\omega} - (\tilde{B_1}+\tilde{B}_0)J+(\tilde{B_1}-\tilde{B}_0)J^2 \)$

  • When \(B_0=B_1\) we recover rigid-rotor harmonic oscillator predictions.