Previous Up

Description of Electronically Excited States

Excited States of Molecules

Most molecules have bound higher energy excited electronic states. These states may be described by promotion of one or more electrons to higher energy orbitals. Electromagnetic energy of appropriate wavelength can cause the ground state electronic structure of a molecule to change into such state. As the rearrangement of electrons is much faster than the motion of nuclei, the nuclear configuration does not change significantly during the energy absorption process. Thus, the absorption spectrum of molecules is characterized by vertical excitation energies. Upon (usually rapid) relaxation, the nuclei adopt a new optimum geometry that is at equilibrium with the excited state electronic wavefunction. The energy difference between the relaxed excited state energy and the ground state energy is called the adiabatic excitation energy (which is measure in emission spectra). The actual absorption spectra of molecules in the gas phase are complex due to non-vertical transitions from the lowest vibrational level of the ground elecronic state into several vibrational levels of the excited electronic state. The spectra of molecules in solvents that interact strongly with the chromophore are broad and often featureless.

Computational Methods for Excited States

Meaningful excited state calculations can be difficult to carry out. There are special cases where a simple method like Configuration Interaction Singles (CIS) will give useful answers (see below). In general, however, one must be aware of pitfalls such as:

Configuration Interaction Singles (CIS)

The simple CIS approach is accurate in certain special cases, in particular for so-called charge transfer transitions as discussed in class. In the CIS approach we use orbitals of the Hartree-Fock solution to generate all singly excited determinants of the configuration interaction expansion. This treatment can be thought of as the Hartree-Fock method for excited states. It allows one to simultaneously solve for a large number of excited states and to optimize the geometry of any (desired) selected state. Both spin singlet and spin triplet states can be generated. The CIS method has some appealing features:

The main problem with the CIS method is that it gives accurate excitation energies only for transitions that are dominated by single excitations. Otherwise, typical errors are 1 eV, which makes it difficult to assign observed spectral lines in the absence of symmetry. For low-lying electronic states, that are not charge transfer transitions, a form of density functional theory, known as time-dependent density functional theory (TDDFT), is preferred. For small molecules higher level wavefunction-based methods currently offer the best choice.

The table below compares the performance of different methods in predicting the UV spectrum of formaldehyde:

                EXP     CIS    CIS-MP2     TDHF    TDDFT     CASSCF  CASPT2     EOM-CCSD
1A2 (n -> Pi*)  4.07    4.48    4.58       4.35	   3.92      4.62    3.91       4.04
1B2 (n -> Sg*)  7.11    8.63    6.85       8.59	   6.87      6.88    7.30       7.04

Here CIS-MP2 is the CIS method with a perturbation correction for double excitations. TDHF stands for time-dependent Hartree-Fock, which is a modification of CIS to include ground state electron correlation. The last three columns refer to higher level methods. Note that the n -> pi** transition is given fairly well by CIS and the MP2 correction (CIS-MP2) improves the n -> sg* result considerably. In fact,the results of CIS-MP2 are very close to those of one particular high level method (CASSCF). For these specific transitions the EOM-CCSD treatment discussed briefly in class is the best.

Running and Analyzing CIS calculations

The CIS method is implemented in the program Gaussian. To run the calculation, specify the keyword CIS. The number of desired excited states can be specifed as an option to the CIS keyword: CIS(NStates=8) requests the 8 lowest excited states. The CIS calculation is more resource-consuming than the Hartree-Fock calculation and calculations with large basis sets, such as aug-cc-pVTZ, may not be possible for larger molecules. Smaller basis such as 6-31+G(d) may be appropriate for larger molecules. Below is a sample output from the UV spectrum of a nucleobase uracil in Cs geometry with 6-31+G(2d,p) basis. The experimental spectrum of uracil in water shows two intense bands, centered around 257 nm and 220 nm. Based on the calculated intensities (f values are the oscillator strengths, which are the measure of intensity), these can be identified as excited state 2 and excited state 8. Notice that excited state 1 has very small intensity: this transition is nearly forbidden by orbital symmetry considerations. Symmetry considerations are especially useful in identifying transitions in highly symmetric molecules. In this case the CIS transition energies are significantly in error. The HOMO is orbital 43; the LUMO is orbital 44. Thus, the main contribution to state 2 (as determined by squaring the given coefficient) arises from excitation of an electron from the HOMO to LUMO + 5. You can look at the shape of orbitals involved using MOLDEN.

 Excited State   1:   Singlet-A"     5.7786 eV  214.56 nm  f=0.0008
      43 -> 44         0.57421
      43 -> 45        -0.29806
      43 -> 46         0.13312
      43 -> 50         0.12401
 This state for optimization and/or second-order correction.
 Copying the Cisingles density for this state as the 1-particle RhoCI density.

 Excited State   2:   Singlet-A'     5.8742 eV  211.06 nm  f=0.3761
      43 -> 47        -0.10128
      43 -> 49         0.66751

 Excited State   3:   Singlet-A"     6.5816 eV  188.38 nm  f=0.0023
      43 -> 44         0.35589
      43 -> 45         0.48737
      43 -> 46        -0.24746
      43 -> 48        -0.11210

 Excited State   4:   Singlet-A"     6.6584 eV  186.21 nm  f=0.0002
      37 -> 49         0.23353
      41 -> 49         0.54352
      41 -> 57        -0.20644
      41 -> 72         0.10528
      41 -> 78        -0.13593

 Excited State   5:   Singlet-A'     7.0599 eV  175.62 nm  f=0.0028
      43 -> 47         0.66177
      43 -> 52        -0.15451

 Excited State   6:   Singlet-A"     7.0916 eV  174.83 nm  f=0.0037
      39 -> 44         0.10592
      43 -> 45         0.19574
      43 -> 46         0.53360
      43 -> 48        -0.12071
      43 -> 50        -0.23089
      43 -> 51         0.16163
      43 -> 55         0.13265

 Excited State   7:   Singlet-A"     7.5672 eV  163.84 nm  f=0.0108
      39 -> 44         0.10410
      43 -> 45         0.15085
      43 -> 46         0.15656
      43 -> 48        -0.23182
      43 -> 50         0.54108
      43 -> 56         0.16348
      43 -> 62        -0.11541

 Excited State   8:   Singlet-A'     7.6361 eV  162.36 nm  f=0.5017
      39 -> 57        -0.10651
      43 -> 52         0.10904
      43 -> 54        -0.28671
      43 -> 57         0.50535
      43 -> 59        -0.19194
      43 -> 61        -0.19798

Previous Up

Material by Dr. Kalju Kahn and Bernie Kirtman, Department of Chemistry and Biochemistry, UC Santa Barbara. ©2007.