Fact-checked by Grok 2 weeks ago

Rule of mutual exclusion

The rule of mutual exclusion is a selection rule in molecular spectroscopy that governs the activity of vibrational modes in centrosymmetric molecules, dictating that no normal mode can be both infrared (IR)-active and Raman-active simultaneously. This principle arises from the symmetry properties analyzed through point group theory, specifically in molecules belonging to centrosymmetric point groups that include an inversion center (i). In such molecules, vibrational modes are classified by their behavior under inversion as either gerade (g, symmetric) or ungerade (u, antisymmetric). IR activity requires a change in the molecule's dipole moment during vibration, and the dipole moment transforms according to the ungerade representations (e.g., T1u in octahedral symmetry); conversely, Raman activity requires a change in polarizability, which transforms according to the gerade representations (e.g., A1g, Eg, T2g). These orthogonal symmetry requirements ensure that no single mode can satisfy both conditions, making IR and Raman spectroscopy complementary techniques for characterizing vibrations in centrosymmetric systems. The rule is exemplified in linear triatomic molecules like (CO2), which has D∞h symmetry: the symmetric stretch (σg) is Raman-active but IR-inactive, while the antisymmetric stretch (σu) and bending mode (πu) are IR-active but Raman-inactive. Similarly, in octahedral complexes such as SF6, the threefold degenerate T1u mode is IR-active, whereas the A1g, Eg, and T2g modes are Raman-active. This aids in assigning vibrational spectra, confirming , and distinguishing between symmetric and asymmetric structures, though it does not apply to non-centrosymmetric molecules where modes may be active in both techniques.

Introduction

Definition

The rule of mutual exclusion in molecular states that, for centrosymmetric molecules possessing a center of inversion, no fundamental vibrational mode can be both (IR) active and Raman active simultaneously. This principle arises from the properties of the and dictates the selection rules for spectroscopic transitions in such systems. A centrosymmetric molecule is one that has a center of symmetry, denoted by the inversion operation i, where every has a corresponding counterpart located at an equal distance on the opposite side of this central point. Examples include linear molecules like CO₂ or homonuclear diatomics like O₂, which exhibit this inversion symmetry. The underlying reason for the stems from the parity classification of vibrational modes under inversion: modes are either gerade (even , symmetric under inversion) or ungerade (odd , antisymmetric under inversion). activity requires a change in the , which transforms as an ungerade and thus couples only with ungerade modes, while Raman activity involves a change in , which is a gerade and couples only with gerade modes. This ensures that no single mode can satisfy both selection rules in centrosymmetric environments.

Historical Development

The application of group theory to molecular vibrations, which underpins the rule of mutual exclusion, originated in the early 1920s. In 1924, C.J. Brester first demonstrated how symmetry groups could classify vibrational modes in symmetric molecules, providing a foundational framework for predicting which modes would be spectroscopically active. The experimental discovery of the by in 1928 spurred rapid theoretical progress in selection rules for vibrational . Between 1930 and 1934, George Placzek developed a comprehensive quantum mechanical theory of , elucidating the changes required for Raman activity and contrasting them with the changes governing (IR) absorption. This work highlighted symmetry as a key determinant of activity in both techniques, setting the stage for the principle in centrosymmetric systems. During the 1930s, researchers including Gerhard Herzberg advanced the study of molecular spectra, articulating the rule of mutual exclusion as a direct consequence of parity selection rules in molecules with inversion symmetry: IR-active modes (ungerade, u) cannot be Raman-active (gerade, g), and vice versa. Herzberg formalized this for polyatomic molecules in his influential 1945 monograph Infrared and Raman Spectra of Polyatomic Molecules, integrating it with emerging understandings of point group symmetries and vibrational analysis. Post-World War II developments further refined the rule through systematic normal mode analysis. The 1955 text Molecular Vibrations by E. Bright Wilson Jr., J.C. Decius, and Paul C. Cross introduced the GF matrix method, enabling precise calculations of vibrational symmetries and activities to verify exclusion in complex systems. In the latter half of the 20th century, computational advances in —such as methods and —facilitated routine simulations of vibrational spectra, confirming the rule's predictions across diverse molecular structures and extending its verification beyond experimental limitations.

Theoretical Foundations

Symmetry Operations and Point Groups

Symmetry operations are the transformations that leave a indistinguishable from its original configuration, forming the basis for classifying in . These operations include the (E), which leaves the unchanged; proper s (C_n), which rotate the by 360°/n around an axis, such as a C_2 of 180°; reflections (σ), which mirror the molecule across a , categorized as horizontal (σ_h) to the principal axis, vertical (σ_v) containing the principal axis, or dihedral (σ_d) bisecting axes; and inversion (i), which maps each point through a central point to its opposite. These operations are essential in molecular as they determine how interacts with molecular wavefunctions. Point groups are collections of these symmetry operations that share a common point, typically the molecular center, and are denoted by symbols like C_{nv}, D_{nh}, or O_h, with centrosymmetric point groups distinguished by the presence of an inversion center. Centrosymmetric groups, which possess the inversion operation (i), include D_{\infty h} for linear molecules like CO_2, featuring infinite rotation axes (C_{\infty}) and vertical reflection planes (σ_v); O_h for octahedral molecules like SF_6, with multiple high-order rotations (e.g., 3C_4, 4C_3) and reflection planes; and D_{6h} for planar molecules like benzene, incorporating a sixfold rotation axis (C_6), perpendicular C_2 axes, and both horizontal and vertical reflection planes. These groups are particularly relevant to the rule of mutual exclusion, as their inversion symmetry leads to parity classifications that affect spectroscopic activity./02%3A_Symmetry_and_Group_Theory/2.02%3A_Point_Groups) The inversion operation (i) is mathematically represented as \mathbf{r} \to -\mathbf{r}, where \mathbf{r} denotes the position vector of a point relative to the inversion center, transforming coordinates (x, y, z) to (-x, -y, -z). This operation is absent in non-centrosymmetric molecules, such as chiral ones, but defines the in centrosymmetric point groups. Character tables for point groups systematically organize these operations and their effects on irreducible representations, classifying vibrational modes (or other basis sets) as gerade (g, even parity) or ungerade (u, odd parity) based on their behavior under inversion. In these tables, a g designation corresponds to a character of +1 for the i operation, indicating the mode remains unchanged (symmetric), while u corresponds to -1, signifying a sign change (antisymmetric). This g/u classification is crucial for distinguishing modes in centrosymmetric molecules, though detailed application to vibrations is addressed elsewhere.

Vibrational Modes and Symmetry

In molecular spectroscopy, normal modes of vibration represent the independent oscillatory motions of atoms around their equilibrium positions in a . These modes are characterized by collective displacements where all atoms oscillate at the same frequency and in phase, though with varying amplitudes proportional to the eigenvectors derived from the of the . For a nonlinear with N atoms, there are $3N - 6 such vibrational , excluding translations and rotations, while linear s have $3N - 5. This decoupling into normal modes simplifies the analysis of vibrational spectra by treating each as a . The of these modes is classified according to the irreducible s (irreps) of the molecule's , which describes how the vibrational coordinates transform under the group's operations. Each normal mode serves as a basis for one or more irreps, allowing vibrations to be labeled by symmetry species such as A_1, B_2, or \Sigma_g^+ depending on the ./04%3A_Symmetry_and_Group_Theory/4.04%3A_Examples_and_Applications_of_Symmetry/4.4.02%3A_Molecular_Vibrations) To determine this classification, the total representation of all atomic displacements is first constructed from the character table of the , then reduced by subtracting the irreps corresponding to translations and rotations; the remaining representation is further decomposed into irreps that correspond to the vibrational modes. This reduction process often employs projection operators conceptually to project the total representation onto the irreducible subspaces, yielding the symmetry-adapted combinations of vibrational coordinates without requiring explicit . operators act on basis functions (such as stretches or bends) to generate linear combinations that transform purely as specific irreps, facilitating the identification of mode symmetries. For instance, in ( C_{2v}), the three normal modes transform as A_1, A_1, and B_2, reflecting their behavior under and operations. In centrosymmetric molecules belonging to point groups like D_{\infty h} or O_h, vibrational modes are additionally assigned parity labels: gerade (g) for modes unchanged under spatial inversion (even parity) and ungerade (u) for those that change sign (odd parity). This parity arises from the inversion operation i, where the character is +1 for g modes and -1 for u modes, distinguishing symmetric and antisymmetric vibrations relative to the molecular center. Such classification is essential for understanding mode degeneracies and activity in centrosymmetric systems./Spectroscopy/Fundamentals_of_Spectroscopy/Selection_rules_and_transition_moment_integral)

Explanation of the Rule

Conditions for Applicability

The rule of mutual exclusion applies specifically to molecules possessing a center of inversion, which belong to centrosymmetric point groups such as D_{\infty h} (e.g., CO_2), O_h (e.g., SF_6), or C_{2h} (e.g., trans-N_2F_2). In these cases, the presence of the inversion symmetry element i ensures that no vibrational mode can change both the and the tensor simultaneously, leading to mutually exclusive activity in (IR) and Raman spectra. The rule does not apply to non-centrosymmetric molecules, which lack an inversion center and belong to point groups such as C_{2v} (e.g., H_2O). In such systems, vibrational modes can be active in both and , as symmetry constraints do not prohibit simultaneous changes in and . This principle holds rigorously for isolated molecules in the gas phase, where intermolecular interactions are negligible and the remains intact. However, in condensed phases such as liquids or solids, the rule may weaken or break down due to molecular collisions or interactions that temporarily distort the inversion , allowing weak activity in the otherwise forbidden spectrum. For instance, in liquid CS_2, modes expected to be Raman-inactive appear weakly in the spectrum. Theoretically, the rule relies on the quantum mechanical treatment of vibrations under the harmonic approximation, where normal modes are defined as independent harmonic oscillators without anharmonic perturbations that could couple modes or alter symmetry selections. This assumption simplifies the analysis of symmetry properties but may not fully capture effects in real molecules with .

IR and Raman Selection Rules

In infrared (IR) , a vibrational mode is active if it leads to a change in the molecular \mu. This requires the representation of the vibrational mode \Gamma_\text{vib} to transform as the representation of the operator \Gamma_\mu, such that \Gamma_\text{vib} \subset \Gamma_\mu (or \Gamma_\text{vib} = \Gamma_\mu for irreducible representations). In centrosymmetric molecules, which possess a center of inversion, the components transform as ungerade (u) representations (e.g., x, y, z coordinates have odd under inversion). Thus, only ungerade vibrational modes can be IR-active in such systems. For , a vibrational mode is active if it induces a change in the molecular tensor \alpha. The selection rule is that \Gamma_\text{vib} must be contained in \Gamma_\alpha, the of the , so \Gamma_\text{vib} \subset \Gamma_\alpha. The , being a second-rank tensor derived from terms like x^2, [xy](/page/XY), xz, transforms as gerade (g) representations in centrosymmetric molecules (even under inversion). More precisely, the Raman transition is allowed if the \Gamma_\text{vib} \otimes \Gamma_\alpha contains the totally symmetric of the point group (e.g., A_{1g}). Consequently, only gerade vibrational modes are Raman-active in centrosymmetric systems. The rule of mutual exclusion arises directly from this distinction in centrosymmetric molecules. Ungerade (u) modes, which are IR-active due to matching \Gamma_\mu (u), cannot be Raman-active because they do not overlap with \Gamma_\alpha (g). Conversely, gerade (g) modes, active in via \Gamma_\alpha (g), are IR-inactive as they lack u . This parity mismatch ensures no vibrational mode can be simultaneously active in both IR and Raman spectra, providing a powerful tool for analysis.

Applications in Spectroscopy

Examples in Diatomic and Linear Molecules

In homonuclear diatomic molecules such as N₂, which belong to the D∞h point group and possess a center of inversion, the rule of mutual exclusion manifests clearly due to the absence of a permanent . The single vibrational mode is a symmetric stretch of Σ_g^+ , which does not change the and thus renders the molecule . However, this mode alters the , making it Raman-active, with the fundamental transition observed at approximately 2331 cm⁻¹. This separation of activities exemplifies the rule's application in simple systems, where the vibrational stretch can be conceptually sketched as the two nuclei moving symmetrically toward and away from each other along the bond axis, preserving even parity (g) under inversion and thus excluding IR activity while allowing . Homonuclear diatomics like O₂ follow analogous behavior, with their Raman-active stretches confirming the exclusion without IR counterparts. For linear triatomic molecules like CO₂, also in the D∞h point group with a center of symmetry, the four vibrational (3N-5=4) decompose into three distinct modes: a symmetric stretch (ν₁, Σ_g^+), an asymmetric stretch (ν₃, Σ_u^+), and a doubly degenerate bend (ν₂, Π_u). The symmetric stretch, where both C-O bonds elongate and contract in phase, maintains even parity and changes but not , so it is Raman-active only, appearing at ~1333 cm⁻¹. In contrast, the asymmetric stretch, involving out-of-phase bond motions creating a temporary , and the bending mode, which also induces dipole changes, are both IR-active only, at ~2349 cm⁻¹ and ~667 cm⁻¹, respectively. Conceptually, these modes can be sketched as follows: the symmetric stretch shows parallel arrows on both oxygen atoms moving away from and toward the central carbon, with g parity; the asymmetric stretch has opposing arrows (one oxygen approaching while the other recedes), yielding u parity; and the bend depicts perpendicular displacements of oxygens in the plane (degenerate in two directions), also u parity. This clear dichotomy in CO₂'s spectra—Raman detecting only the symmetric mode and detecting the others—directly illustrates the , as no mode is active in both techniques due to the inversion center.

Examples in Polyatomic Molecules

Polyatomic non-linear molecules with a center of symmetry provide clear illustrations of the rule of mutual exclusion, where the total number of vibrational modes is given by the formula 3N-6, with N being the number of atoms, and these modes decompose into symmetry species that determine their spectroscopic activity. For instance, benzene (C₆H₆), belonging to the D₆ₕ point group, has 12 atoms and thus 30 vibrational modes (3×12 - 6 = 30), ten of which are doubly degenerate, resulting in 20 fundamental frequencies. These active modes span the irreducible representations: 2A₁g + E₁g + 3E₂g (Raman-active, gerade) and A₂u + 4E₁u (IR-active, ungerade), ensuring no overlap between IR and Raman activities due to the centrosymmetric structure. A simpler example is trans-1,2-dichloroethene (C₂H₂Cl₂), which has and 4 atoms, yielding 6 vibrational modes (3×4 - 6 = 6). In this , the modes classify as 3A_g + B_g (Raman-active) and A_u + B_u (IR-active), with all modes strictly either gerade or ungerade, prohibiting any vibration from being active in both spectroscopies. This separation highlights the rule's effect in lower-symmetry centrosymmetric systems, where degeneracy is absent but the exclusion remains absolute. Experimental spectra confirm these predictions: for , IR bands such as the E₁u mode at approximately 1030 cm⁻¹ appear prominently in infrared but are absent in Raman, while Raman-active A₁g modes like the ring breathing at 992 cm⁻¹ show no corresponding IR intensity, with silent modes (e.g., B₂g) unobserved in both. Similarly, in trans-1,2-dichloroethene, gas-phase IR spectra reveal A_u and B_u fundamentals (e.g., C-Cl stretch at ~700 cm⁻¹), while Raman spectra display only A_g and B_g bands (e.g., C-H bend at ~900 cm⁻¹), with no shared frequencies, verifying the mutual exclusion and the silence of forbidden modes.

Implications and Limitations

Spectroscopic Consequences

The rule of mutual exclusion has profound implications for experimental vibrational spectroscopy, particularly in the complementary use of infrared (IR) and Raman techniques to characterize molecular vibrations in centrosymmetric systems. In such molecules, IR spectroscopy selectively detects ungerade (u) vibrational modes that involve a change in the dipole moment, such as asymmetric stretches often associated with functional groups like C=O or O-H. Conversely, Raman spectroscopy identifies gerade (g) modes that alter molecular polarizability, including symmetric stretches like those in C=C bonds or homonuclear diatomics. This separation ensures that the two methods probe orthogonal subsets of the vibrational spectrum, providing a more complete picture when used together; for instance, in benzene (D_{6h} symmetry), the symmetric ring-breathing mode at approximately 992 cm^{-1} appears strongly in Raman but is absent in IR, while the out-of-plane C-H bends are IR-active only. A key consequence is the enhanced assignment of spectral features, where the rule facilitates precise labeling of vibrational modes and confirmation of . By comparing and Raman spectra, researchers can assign modes based on their activity: the presence of a band in only one spectrum supports centrosymmetric , while mutual activity may suggest lower , aiding in structural elucidation. This symmetry-based assignment is particularly valuable in complex spectra, as it reduces ambiguity; depolarization ratios in Raman further refine classifications, with polarized bands indicating totally symmetric () modes. In polyatomic molecules, this approach streamlines the correlation of observed peaks to specific symmetry species, improving the accuracy of vibrational analysis without relying solely on computational predictions. In practical applications, the rule underpins analytical advantages in fields like forensics and , where combined IR-Raman spectroscopy yields a comprehensive vibrational for and . For example, in forensic analysis of , IR excels at detecting polar functional groups in residues, while Raman reveals non-polar backbone structures, enabling unambiguous material matching even in mixtures. Similarly, in , such as microplastic , simultaneous IR-Raman acquisition overcomes limitations of individual techniques, providing high-resolution spectra that distinguish types and additives with submicron . Quantitatively, intensity ratios between corresponding modes in IR and Raman spectra can highlight subtle perturbations, such as those from environmental interactions, offering insights into structural integrity without altering the core exclusion principle.

Exceptions and Special Cases

The rule of applies strictly only to molecules possessing a of inversion in their ground electronic state; thus, in non-centrosymmetric molecules such as cis-isomers (e.g., cis-1,2-dichloroethene, which belongs to the C_{2v} ) or chiral molecules (e.g., helicenes), vibrational modes can exhibit activity in both () and Raman spectra without violation, as the absence of an inversion allows ungerade modes to contribute to changes. Similarly, in systems where electronic resonance effects perturb the ground-state , such as through vibronic , the of vibrational wavefunctions can mix, enabling forbidden modes to gain intensity in the "inactive" technique; for instance, in layered , resonant exciton-phonon in the near- leads to a breakdown of , with Raman spectra showing IR-active modes due to valley-specific interactions. The Jahn-Teller effect provides another resonance-induced exception, where degenerate electronic states drive geometric distortions that lower the and eliminate the inversion ; in octahedral Cu(II) complexes like [Cu(H2O)6]^{2+}, the axial elongation reduces the from O_h to D_{4h}, allowing originally Raman-only e_g modes to become partially IR-active through vibronic mixing. In condensed phases, intermolecular interactions often perturb the ideal of isolated molecules, inducing weak activity in otherwise forbidden modes. For (CS_2), which is strictly centrosymmetric (D_{\infty h}) in the gas —exhibiting no overlap between (ν_2 and ν_3 modes) and Raman (ν_1 mode) bands—the liquid shows violations, with the Raman-inactive ν_1 appearing weakly in (~1-5% ) and IR-active modes gaining faint Raman signals, attributed to transient during molecular collisions. This effect arises from short-lived perturbations in the local environment, as described in semiclassical models of collisional dynamics. In solids or liquids more generally, lattice vibrations or hydrogen bonding can couple to intramolecular modes, activating silent vibrations; for example, in crystalline , intermolecular forces induce low-intensity Raman activity in u modes that are IR-forbidden in the free . Isotopomeric substitution or structural defects can break exact in otherwise symmetric molecules, leading to partial of forbidden modes. In (CO_2), the symmetric ^{12}C^{16}O_2 isotopomer obeys , with the symmetric stretch (ν_1) Raman-active but IR-inactive; however, the asymmetric ^{13}C^{16}O^{18}O isotopomer lacks an inversion center (C_s symmetry), rendering ν_1 both IR- and Raman-active due to altered and derivatives. Defects, such as vacancies or impurities in like , similarly relax selection rules by local symmetry reduction; in defective C_{60} fullerenes, isotopic substitution or stone-wales defects activate IR-silent modes in Raman spectra through at symmetry-broken sites. In modern contexts, such as nanomaterials and high-pressure environments, density functional theory (DFT) calculations reveal symmetry reductions that violate the exclusion rule. For polyyne chains in solution or nanoscale assemblies, end-group capping or solvent interactions lower the linear D_{\infty h} symmetry to C_{\infty v}, enabling mutual IR-Raman activity in stretching modes, as confirmed by DFT-optimized geometries showing bent configurations. Under high pressure, molecules like CrSBr undergo phase transitions with symmetry lowering (e.g., from P2_1/c to lower groups), resulting in emergent Raman peaks from originally IR-only modes, as observed in in situ spectroscopy and reproduced by DFT simulations of pressure-induced distortions up to 10 GPa. In nanomaterials such as carbon nanotubes, radial breathing modes exhibit partial IR activity due to curvature-induced symmetry breaking, with DFT predicting intensity borrowings for symmetric stretches becoming dipole-allowed.

References

  1. [1]
    Raman active modes
    Therefore it obeys the rule of mutual exclusion: IR active modes are Raman inactive, and vice versa. through the equilibrium position.
  2. [2]
    None
    ### Summary of Rule of Mutual Exclusion
  3. [3]
    [PDF] Raman Spectroscopy
    MUTUAL EXCLUSION RULE. For centro-symmetric molecules (those with a center of inversion), transitions are allowed either for Raman scattering or by infrared ...<|control11|><|separator|>
  4. [4]
    1.13: Selection Rules for IR and Raman Spectroscopy
    Feb 23, 2025 · Rule of Mutual Exclusion. A molecule is centrosymmetric if it has a center of inversion and their corresponding point group contains the class ...
  5. [5]
    Complementary vibrational spectroscopy - PMC - NIH
    The group theory states that fundamental vibrational modes of molecules with the center of symmetry cannot be both IR and Raman active (which is known as the ...Missing: definition | Show results with:definition
  6. [6]
    [PDF] Rotational and Vibrational Spectroscopy 1 Chapter 27 Problems
    The molecules with a center of symmetry are N2, O2, and CO2. No normal mode of a centrosymmetric molecule is both Raman and IR active. Since the homonuclear ...
  7. [7]
    [PDF] "Resonance Raman Spectroscopy" in
    If a molecule or an ion possesses a center of inversion, there is a rule of mutual exclusion; no fundamental vibration that is active in the IR absorption ...
  8. [8]
    Instrumental Analysis Lectures
    Rule of Mutual Exclusion. from chemlibre link "A molecule is centrosymmetric if it has a center of inversion and their corresponding point group contains the ...
  9. [9]
    Practical Group Theory and Raman Spectroscopy, Part I: Normal ...
    Centrosymmetric molecules follow the rule of mutual exclusion, affecting IR and Raman spectral activity. ... gerade and ungerade, which here mean even and odd, ...
  10. [10]
    The Analysis of Vibrational Spectra: Past, Present and Future
    ### Historical Development of Selection Rules in Vibrational Spectroscopy
  11. [11]
    [PDF] Theory of Normal Vibrations - COPYRIGHTED MATERIAL
    “mutual exclusion rule.” It should be noted, however, that in polyatomic ... Herzberg, Molecular Spectra and Molecular Structure, Vol. II: Infrared and ...
  12. [12]
  13. [13]
    Group Theory and Symmetry, Part I: Symmetry Elements
    Dec 1, 2009 · The next symmetry element is the center of inversion, symbolized by i. This symmetry operation is a reflection through a point at the center ...
  14. [14]
    [PDF] Physical Chemistry 5 (Spectroscopy) A short introduction to Group ...
    Some common point groups: D∞h group symmetry elements example. D∞h. E, C ... Some common point groups: D6h group symmetry elements example. D6h. E, C6 ...
  15. [15]
  16. [16]
    [PDF] Molecular Vibrations - - Sherrill Group
    The eigenvectors ak are the normal modes of vibration. For each normal mode, all the atoms move with the same frequency and phase, but with different amplitudes ...
  17. [17]
    [PDF] 4. symmetry of molecules and molecular vibrations
    Character Tables​​ The character table is used to classify molecular vibrations to the irreducible representation of the symmetry group.
  18. [18]
    [PDF] Assigning Symmetries of Vibrational Modes 1 Introduction 2 O Has ...
    The 1's and -1's in the table indicate whether the irreducible representation (or irrep, for short) is symmetric or antisymmetric for that symmetry operation. ...
  19. [19]
    [PDF] Chem 104A, UC, Berkeley
    Important relationship between symmetry & Vibrations. Each normal mode of vibration forms a basis for an irreducible representation of the point group of the ...
  20. [20]
    [PDF] Group theory
    4.1.3 Permutation-inversion operations and CNPI groups. As mentioned already, the point groups are well suited to describe rigid molecules. However, for ...
  21. [21]
    [PDF] Raman Spectroscopy
    The general rule: symmetric vibrations give intense Raman bands and weak IR bands. The fact that H. 2. O does not obey the rule of mutual exclusion indicates ...
  22. [22]
    None
    ### Summary of Rule of Mutual Exclusion Discussion
  23. [23]
    Molecular Vibration - an overview | ScienceDirect Topics
    We recognize that the concept of normal mode is based on the harmonic approximation. Under harmonic approximation, there are no couplings among the normal modes ...
  24. [24]
    [PDF] SYMMETRY AND VIBRATIONAL SPECTROSCOPY 5.1 Potential ...
    For the bending vibrations, both eu and a2u should be observed in the IR spectrum. One should be polarized in the molecular plane, and the other polarized ...
  25. [25]
    [PDF] Comparison between IR and Raman
    IR Selection Rule: For any IR absorption, the vibrational excitation must change the dipole moment of the molecule in order for radiation absorption to occur.Missing: Γ_vib Γ_μ Γ_α mutual exclusion
  26. [26]
    Raman Techniques: Fundamentals and Frontiers - PMC
    This rule is general and for nonlinear molecules, mutual exclusion is relaxed. In materials without inversion symmetry, (ro-)vibrational mode transition can ...
  27. [27]
    [PDF] Shift in the Raman symmetric stretching band of N2, CO2, and CH4 ...
    At constant temperature, Raman peak position, including the more intense CO2 peak (ν+) decreases (shifts to lower wavenumber) with increasing pressure for all ...
  28. [28]
    Nitrogen - the NIST WebBook
    Nitrogen (N2) has a molecular weight of 28.0134 and is also known as Nitrogen gas, Dinitrogen, or Molecular nitrogen.
  29. [29]
    18.3: Applications of Raman Spectroscopy - Chemistry LibreTexts
    Apr 27, 2019 · Homonuclear diatomic molecules are all IR inactive, fortunately, the ... Mutual Exclusion rule [5], which says that for a molecule with ...
  30. [30]
    IR spectroscopic characterization of the co-adsorption of CO2 and ...
    Free CO2 has three fundamental modes: the symmetric and anti-symmetric stretch vibrations ν1 and ν3, and the bending vibration ν2. Due to the symmetry of CO2, ...
  31. [31]
    [PDF] N2 Gas Mixtures by Raman Spectroscopy - HAL
    Nov 20, 2019 · Only the ν1 mode (~ 1333 cm–1) is Raman- active. However, the experimental spectrum of CO2 presents two strong bands because of Fermi ...
  32. [32]
    [PDF] Vibrational Modes and Symmetry
    CO2. 3x3-5 = 4 vibrations: Σg. + + Σu. + + Πu. Page 4. D3h. PCl5. 6x3-6 = 12 ... 7x3-6 =15 vibrations: A1g + Eg + 2T1u + T2g + T2u. Octahedral Symmetry.
  33. [33]
    [PDF] Chapter 5. IR Spectroscopy and Raman Scattering
    The symmetric stretch is an easier deformation than the asymmetric stretch so the asymmetric stretch occurs at a higher wavenumber. The bending vibration is ...
  34. [34]
    Intrinsic molecular vibration and rigorous vibrational assignment of ...
    Oct 21, 2020 · Benzene has thirty vibrational modes and ten of them are double degenerated which leads to twenty independent vibrational modes and fundamental ...
  35. [35]
    [PDF] C6H6.mcd
    The symmetry of the vibrational modes and their IR and Raman activity are given below: IR active: A2u and 3E1u. Raman active: 2A1g, E1g, and 4E2g.
  36. [36]
    Infrared and Raman Spectra of the Isomers of 1,2-Dichloroethylene ...
    Origin of the liquid-liquid phase transition for trans -1,2-dichloroethylene observed by IR spectroscopy. The Journal of Chemical Physics 2008, 129 (7) ...<|control11|><|separator|>
  37. [37]
    [PDF] Raman Spectroscopy: Theory
    We have seen that due to the rule of mutual exclusion, asymmetric vibrations with wavenum- bers Qn2 and Qn3 (normal modes Q2 and Q3) are inactive in the ...Missing: original | Show results with:original
  38. [38]
  39. [39]
    Vibrational Spectroscopy of Hexahalo Complexes - ACS Publications
    Apr 5, 2022 · This highlights the complementarity of the three forms of spectroscopy: symmetry assignments are unambiguous from the infrared and Raman spectra ...
  40. [40]
    Optical photothermal infrared spectroscopy with simultaneously ...
    Nov 5, 2022 · IR and Raman spectroscopy are complementary methods to some extent. IR spectroscopy targets the vibrations of molecular bonds upon a change in ...
  41. [41]
    Teller effect III. The rotational and vibrational spectra of symmetric ...
    We have developed the physical theory required for interpreting the infra-red, Raman and microwave spectra of molecules in orbitally degenerate electronic ...
  42. [42]
    Motions of Molecules in Condensed Systems. XII. Infrared Spectrum ...
    Induced infrared absorption by modes which are forbidden in the vapor can yield much information on the crystal structure. In addition, the spectra of the ...
  43. [43]
    Correlation of the Vibrational Spectra of Isotopomers: Theory and ...
    Using the concept of a mass reaction path, we developed an algorithm that leads to the correct correlation of the vibrational modes of two isotopomers.<|separator|>
  44. [44]
    Effect of isotopic substitution on the Raman spectrum of | Phys. Rev. B
    Apr 1, 1995 · Alternate symmetry-breaking mechanisms are suggested to explain the observed weak modes. References (15). Zheng-Hong Dong, Phys. Rev.Missing: activity centrosymmetric
  45. [45]
    Pressure-induced structural phase transitions in CrSBr - Nature
    Jun 11, 2025 · Since various peaks appear and disappear, we know that CrSBr has both symmetry restorations and reductions under pressure. ... mutual exclusion is ...