Skip to main content
Chemistry LibreTexts

2.7: The Carbonyl Group

  • Page ID
    364136
  • \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \)

    \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)

    \( \newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\)

    ( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\)

    \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\)

    \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\)

    \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\)

    \( \newcommand{\Span}{\mathrm{span}}\)

    \( \newcommand{\id}{\mathrm{id}}\)

    \( \newcommand{\Span}{\mathrm{span}}\)

    \( \newcommand{\kernel}{\mathrm{null}\,}\)

    \( \newcommand{\range}{\mathrm{range}\,}\)

    \( \newcommand{\RealPart}{\mathrm{Re}}\)

    \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\)

    \( \newcommand{\Argument}{\mathrm{Arg}}\)

    \( \newcommand{\norm}[1]{\| #1 \|}\)

    \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\)

    \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    \( \newcommand{\vectorA}[1]{\vec{#1}}      % arrow\)

    \( \newcommand{\vectorAt}[1]{\vec{\text{#1}}}      % arrow\)

    \( \newcommand{\vectorB}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \)

    \( \newcommand{\vectorC}[1]{\textbf{#1}} \)

    \( \newcommand{\vectorD}[1]{\overrightarrow{#1}} \)

    \( \newcommand{\vectorDt}[1]{\overrightarrow{\text{#1}}} \)

    \( \newcommand{\vectE}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash{\mathbf {#1}}}} \)

    \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \)

    \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)

    Let’s look at the electron transitions of H2CO, formaldyhyde, for a typical description of the MO of an organic molecule.

    Start.png

    Pretty valence-bond theory

    (first year description of formaldehyde): https://www.chemtube3d.com/orbitalsformaldehyde

    View 1: Before (but after hybridization)

    Carbon and the oxygen are both \(\mathrm{sp}^{2}\) hybridized.

    formaldehyde_n.png

    (CC BY-SA-NC; Nick Graves via ChemTube3D)

    View 2: Moved together to demonstrate overlap

    When moved together, there is overlap of the \(\pi\) and \(\mathrm{sp}^{2}\) orbitals.

    Q.png

    (CC BY-SA-NC; Nick Graves via ChemTube3D)

    View 3: Generation of new bonds to make the carbonyl bond)

    Formation of bonding interactions to form the sigma bond (from the \(\mathrm{sp}^{2}\) orbitals on \(\ce{C}\) and \(\ce{O}\) AND an \(\pi\) bond from the two non-hybridized \(p\) orbitals on \(\ce{C}\) and \(\ce{O}\).

    formaldehyde_n - Copy (2).png

    (CC BY-SA-NC; Nick Graves via ChemTube3D)

    H2CO is a planar molecule with \(C_{2v}\) symmetry. We ignore the n(1s) core electrons on C and O as well as the \(σ\) electrons in the C-H bonds. Very low energy excitations will not be involved in the UV-VIS region. The ground state valence electronic configuration of interest is:

    \[n_{a}^{2} \sigma^{2} \pi^{2} n_{b}^{2}\left(\pi^{*}\right)^{0}\left(\sigma^{*}\right)^0 \nonumber \]

    This is the lowest energy MO’s that originate from the valence atomic orbitals of the \(\ce{C}\) and \(\ce{O}\).

    What do they look like? http://pubs.acs.org/doi/pdf/10.1021/ed050p400

    clipboard_e4bb322245e2e6af1bc2934ca87716cbd.png

    Figure XX: http://pubs.acs.org/doi/pdf/10.1021/ed050p400. All Rights Reserved ACS

    The MO energies are

    \[n_{a}<\sigma<\pi<n_{b}<\pi^{*}<\sigma^{*} \nonumber \]

    from a simple MO calculation. \(n < \sigma\) because \(2 s \ll 2p\). Let's ignore the lowest non-bodning state and focus on the "frontier orbitals" - the higher lying MOs.

    Possible Electronic Transitions

    There are six outer valence e- are involved in major electronic transitions. The lowest energy transition is a \(n \rightarrow \pi^{*}\) transition, while \(\pi \rightarrow \pi^{*}\) and \(n_a \rightarrow \sigma^{*}\) occur at higher \(ΔE\) and shorter \(λ\).

    clipboard_ebdc8020134134eaf2f0c81bdeb20fb76.png

    Energy levels of molecular orbitals in formaldehyde.

    Molecular orbitals of formaldehyde emphasizing the transitions (after Guillet [19]).

    clipboard_e725875b83dd53a0184f5ae2e13fdd092.png

    Triplet States

    In addition to change in the orbital type, the excited electron may keep the same \(m_s\), spin quantum number producing a singlet excited state. If \(m_s\) changes during the transition, the excited state had two ↑↑ electrons and is a triplet state. For instance, the \(n \rightarrow \pi^{*}\) transition can lead to a \(^1\left(n_{1} \pi^{*}\right)\) if the \(m_s\) does not change, but to \(^3\left(n_{1} \pi^{*}\right)\) if it does change (flip). Both the singlet and triplet states have the same configuration, but lies lower in energy because ↑↑ electrons are on average further apart than ↓↑ electrons (reducing e-e correlation).

    This MO description is oversimplified. e-e interactions are different in an excited state than in the ground state. So A single configuration description is not very good; however this descriptions can be improved by allowing configurations to exit together to minimize the energy of the excited state (e.g. CI-doubles, CI-triples). We will discuss this later.


    2.7: The Carbonyl Group is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

    • Was this article helpful?