Fact-checked by Grok 2 weeks ago

Koopmans' theorem

Koopmans' theorem is a fundamental approximation in quantum chemistry that equates the ionization potential of an N-electron system to the negative of the energy of its highest occupied molecular orbital (HOMO) as computed in the Hartree–Fock self-consistent field method, assuming no relaxation of the remaining orbitals upon electron removal. Formulated by Dutch physicist Tjalling C. Koopmans in his 1934 paper "Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den einzelnen Elektronen eines Atoms" published in Physica, the theorem provides a computationally efficient way to estimate vertical ionization energies from a single reference calculation on the neutral system, without needing to optimize the geometry or electronic structure of the cation. Similarly, it approximates the electron affinity of the N-electron system as the negative of the lowest unoccupied molecular orbital (LUMO) energy of the neutral system, enabling predictions of both electron removal and addition processes. The theorem's validity relies on the frozen-orbital approximation inherent to the Hartree–Fock framework, where the molecular orbitals and their occupation remain unchanged during the ionization or attachment process, neglecting electron correlation and orbital relaxation effects that become significant for deeper core orbitals or highly correlated systems. Despite these limitations, Koopmans' theorem remains widely used in for interpreting data and screening molecular candidates in fields like and , often serving as a starting point for more advanced methods such as the extended Koopmans' theorem or approximations. Its simplicity has made it a cornerstone of education and practical applications in estimating frontier orbital energies for reactivity predictions.

Introduction

Statement of the theorem

Koopmans' theorem provides an approximation within the Hartree-Fock framework for relating the energies of s to experimentally observable quantities such as ionization potentials. For a closed-shell system with N s, the theorem states that the vertical ionization energy to form the (N-1)-electron cation by removing an from the highest occupied (HOMO) is given by \mathrm{IE} \approx -\varepsilon_{\mathrm{HOMO}}, where \varepsilon_{\mathrm{HOMO}} is the eigenvalue of the HOMO from the Hartree-Fock equations. More generally, for ionization from any occupied orbital i, the ionization energy is approximated as \mathrm{IE}_i \approx -\varepsilon_i, where \varepsilon_i is the corresponding orbital energy.80011-X) Physically, this approximation interprets the single-particle orbital energies \varepsilon_i as the negative of the energy required to remove an from orbital i, thereby connecting the eigenvalues of the effective one-electron in the Hartree-Fock method to many-body observables like ionization potentials. This linkage arises under the assumption that the orbitals of the neutral system remain unchanged ("frozen") upon ionization, allowing the orbital energies to serve as direct estimates of removal energies without recalculating the cation's wavefunction.80011-X) A simple illustration of the theorem's accuracy is seen in the helium atom within the Hartree-Fock approximation. The HOMO energy is \varepsilon = -0.8965 hartree, yielding a predicted ionization energy of $0.8965 hartree, which compares favorably to the experimental value of $0.9037 hartree, demonstrating reasonable agreement for this basic system.

Historical background

Koopmans' theorem originated with the work of Dutch physicist Tjalling C. Koopmans, who published his seminal paper in 1934 while studying in the Netherlands.90013-X) In this publication, titled "Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den einzelnen Elektronen eines Atoms," Koopmans addressed the challenge of assigning specific wavefunctions and energy eigenvalues to individual electrons in multi-electron atoms.90013-X) His motivation stemmed from the need to improve approximations for calculating ionization energies, particularly in the context of atomic spectra, by leveraging self-consistent field solutions to interpret electron removal processes more accurately. This development was deeply connected to prior advancements in . Koopmans built directly on Douglas Hartree's 1928 introduction of the self-consistent field method, which iteratively solved for electron orbitals in the mean field of others. Vladimir Fock extended this in 1930 by incorporating quantum mechanical effects into the equations, providing a more rigorous antisymmetric wavefunction treatment. John C. Slater's contemporaneous 1930 work on the determinant form of multi-electron wavefunctions further facilitated practical implementations of these ideas. Together, these contributions formed the foundation of what would later be known as Hartree-Fock theory, within which Koopmans' approximation interprets orbital energies as ionization potentials. Initially applied to atomic systems for estimating ionization potentials, Koopmans' approach gained broader traction in the post-World War II era as shifted toward molecular applications. With the advent of electronic computers in the late 1940s and early , computational methods enabled extensions to molecules, where the theorem provided a simple means to approximate affinities and potentials without full reconfiguration calculations. By the mid-, as digital computing revolutionized the field, Koopmans' theorem was formally recognized as a cornerstone approximation in , influencing the development of semi-empirical and techniques.

Theoretical Basis

Derivation in Hartree-Fock theory

In –Fock theory for closed-shell systems, the total energy E_N of an N-electron system (with N even) is expressed using the as the expectation value of the over a single built from doubly occupied spatial orbitals \{\phi_i\}: E_N = 2 \sum_{i=1}^{m} \langle \phi_i | \hat{h} | \phi_i \rangle + \sum_{i=1}^{m} \sum_{j=1}^{m} (2 J_{ij} - K_{ij}), where m = N/2 is the number of occupied spatial orbitals, \hat{h} is the one-electron core , J_{ij} = \iint |\phi_i(\mathbf{r}_1)|^2 \frac{1}{r_{12}} |\phi_j(\mathbf{r}_2)|^2 \, d\mathbf{r}_1 d\mathbf{r}_2 is the Coulomb integral, and K_{ij} = \iint \phi_i^*(\mathbf{r}_1) \phi_j(\mathbf{r}_1) \frac{1}{r_{12}} \phi_j^*(\mathbf{r}_2) \phi_i(\mathbf{r}_2) \, d\mathbf{r}_1 d\mathbf{r}_2 is the exchange integral. The canonical orbital energies \varepsilon_i, obtained from the Hartree–Fock equations \hat{F} \phi_i = \varepsilon_i \phi_i with Fock operator \hat{F} = \hat{h} + \sum_{j=1}^m (2 \hat{J}_j - \hat{K}_j), are \varepsilon_i = \langle \phi_i | \hat{h} | \phi_i \rangle + \sum_{j=1}^m (2 J_{ij} - K_{ij}). Substituting this into the energy expression yields the equivalent form E_N = 2 \sum_{i=1}^m \varepsilon_i - \sum_{i=1}^m \sum_{j=1}^m (2 J_{ij} - K_{ij}), which highlights the double counting of interactions in the sum of orbital energies. The ionization energy is the difference between the ground-state energies of the (N-1)-electron cation and the N-electron neutral system: \mathrm{IE} = E_{N-1} - E_N. Koopmans' theorem approximates this difference using the frozen-orbital approximation, in which the orbitals \{\phi_i\} of the cation are identical to those of the neutral system, neglecting relaxation effects upon ionization. Assuming ionization from the highest occupied molecular orbital (HOMO) with index h = m, the frozen-orbital energy of the cation is obtained by evaluating the expectation value of the Hamiltonian over the corresponding restricted open-shell Hartree–Fock Slater determinant (with orbitals 1 to h-1 doubly occupied and h singly occupied). Within this approximation, the difference simplifies exactly to E_N - E_{N-1}^{\rm frozen} = \varepsilon_h, because the contribution of the removed electron to the total energy is given by its canonical orbital energy \varepsilon_h under the neutral system's Fock operator. Thus, \mathrm{IE} \approx -\varepsilon_{\mathrm{HOMO}}. This equality holds exactly within the frozen-orbital approximation in Hartree–Fock theory, as originally formulated for atomic systems and extended to molecules.

Key assumptions and limitations

Koopmans' theorem in Hartree-Fock theory rests on the assumption that the molecular orbitals remain unchanged upon removal of an , thereby neglecting orbital relaxation effects in the resulting cation. This frozen-orbital implies that the Fock and orbital coefficients for the N-electron system are directly applicable to the (N-1)-electron system. Additionally, the theorem operates within the mean-field Hartree-Fock framework, which excludes electron correlation beyond the average interaction, assuming the single provides an exact description of both the neutral and ionized states. These assumptions introduce significant limitations, particularly in the accuracy of predicted ionization potentials. The neglect of orbital relaxation leads to an overestimation of ionization energies because the actual relaxation of the remaining electrons stabilizes the cation, reducing the energy difference relative to the neutral species; this relaxation error (\Delta_{\rm relax}) typically contributes 1–2 eV for valence orbitals and up to 5 eV or more for core orbitals, with larger values in highly polarizable systems where electron reorganization is pronounced. The absence of electron correlation further contributes an error (\Delta_{\rm corr}), generally in the opposite direction, as correlation stabilizes the neutral system more than the ion (due to greater correlation energy in the N-electron system), leading to an underestimation of the ionization energy that partially cancels the relaxation effect but still results in net errors of 0.5–1 eV on average for valence ionization in small molecules. The total error can be qualitatively decomposed as \mathrm{IP}_{\rm exact} \approx -\varepsilon_i - \Delta_{\rm relax} + \Delta_{\rm corr}, where \varepsilon_i is the Hartree-Fock orbital energy, \Delta_{\rm relax} > 0 (\sim1–5 from valence to core transitions), and \Delta_{\rm corr} > 0 (\sim0.5–3 , depending on system size and electron count). This partial cancellation explains why the theorem yields reasonable estimates despite its approximations, with mean absolute errors around 0.8 for valence orbitals across diverse small organic molecules. However, performance degrades markedly for delocalized systems like metals, where collective screening and band-like orbital delocalization amplify relaxation and correlation discrepancies, often exceeding 2–5 errors. In contrast, the theorem holds best for systems with localized orbitals, such as insulators or rigid small molecules, where relaxation is minimal and errors can stay below 0.5 for valence predictions with adequate basis sets.

Applications

Calculation of ionization potentials

Koopmans' theorem provides a practical means to estimate vertical ionization potentials (IPs) through a single Hartree-Fock calculation on the neutral closed-shell system, where the IP for removing an from the k-th orbital is approximated as IP_k = -\epsilon_k, with \epsilon_k denoting the corresponding orbital energy. This vertical approximation assumes no nuclear relaxation or electron reorganization upon ionization, distinguishing it from adiabatic IPs that account for such changes in . The procedure involves solving the Hartree-Fock equations for the N-electron system to obtain the canonical orbital energies, with the negative of the highest occupied (HOMO) energy serving as the primary IP estimate for the first ionization. A classic example is the atom, where the Hartree-Fock energy yields a Koopmans' IP of 24.6 , compared to the experimental vertical IP of 21.6 , resulting in an overestimate of approximately 3 due to neglected relaxation and effects. This discrepancy highlights the theorem's utility as a rapid predictive tool rather than a highly precise one, particularly for screening purposes in larger systems. In early applications, Koopmans' orbital energies were instrumental in constructing diagrams to predict and assign peaks in photoelectron spectroscopy (PES) spectra, correlating and other orbital energies directly to observed ionization bands for molecules like and . For validation, the ΔSCF approach—computing the difference between separate Hartree-Fock calculations on the N- and (N-1)-electron systems—yields more accurate results, such as 22.4 for , which overestimates the experimental value by about 0.8 but is closer than the Koopmans' value. This comparison underscores Koopmans' theorem as an efficient screening method, avoiding the computational cost of multiple self-consistent field calculations while providing reasonable initial estimates for IP hierarchies in molecular systems.

Estimation of electron affinities

Koopmans' theorem provides an analogous approximation for the (EA) of a neutral system by relating it to the energy of the lowest unoccupied (LUMO) obtained from a Hartree-Fock calculation on the neutral species, such that EA ≈ −εLUMO, where εLUMO is the LUMO orbital energy. This formulation assumes that the anion wave function is formed by attaching an electron to the LUMO without relaxation of the core orbitals or inclusion of correlation effects. The practical procedure involves computing the Hartree-Fock orbitals for the neutral molecule or atom and directly taking the negative of the LUMO energy as the EA estimate; if this value is negative, it suggests the corresponding anion is unbound relative to the neutral plus free electron. This approach is computationally efficient, as it avoids separate calculations on the anion, but its accuracy is generally lower than for ionization potentials due to greater orbital relaxation upon electron attachment and Hartree-Fock self-interaction errors, which artificially lower the LUMO energy. A representative example is the fluorine atom, where the Koopmans' approximation yields an EA of approximately 4.5 , compared to the experimental value of 3.4 ; this overestimation arises primarily from neglected electron correlation in the anion. Such discrepancies highlight the theorem's limitations for electron attachment, where correlation stabilizes the anion more than the neutral. The approximation performs best for systems exhibiting positive EAs, such as , where the added occupies a compact orbital with minimal ; it is less reliable for non-polarizable molecules, where relaxation effects dominate.

Extensions and Generalizations

For excited-state and open-shell systems

Koopmans' theorem can be generalized to excited-state ions through the extended Koopmans' theorem (EKT), which provides a framework for approximating ionization potentials (IPs) from excited configurations of the neutral system. In this approach, the IP for ionizing an electron from an excited neutral state to reach a specific ionic state is approximately the negative of the orbital energy corresponding to the relevant occupied orbital in that excited configuration, assuming frozen orbitals and neglecting relaxation effects. This extension is exact for the lowest-energy states of a given and can be applied to higher excited states under certain conditions, such as when the removal orbital is properly chosen to match the spin and of the target . A notable application of this generalization arises in the study of core-hole states for (), where the sudden approximation underpins the interpretation of core-level ionization spectra. Here, Koopmans' theorem relates the core to the negative of the core orbital energy in the ground-state Hartree-Fock calculation, providing an initial unrelaxed estimate that aids in understanding the sudden creation of the core hole and subsequent electron shake-up processes. This approximation is particularly useful for estimating binding energies in molecules, though it requires corrections for relaxation and correlation to match experimental XPS shifts accurately. For open-shell systems, such as radicals, is adapted using restricted open-shell Hartree-Fock (ROHF) methods, where ionization potentials are approximated as the negative of the orbital energies of the singly occupied molecular orbitals () for the appropriate components ( or ). In ROHF, orbitals are defined such that the Fock operator satisfies conditions analogous to those in closed-shell cases, enabling direct use of orbital eigenvalues ε_i to estimate vertical IPs from the open-shell , with ε_singly occupied serving as a proxy for removing the . However, for unrestricted Hartree-Fock (UHF) treatments of open shells, challenges emerge due to increased contamination, which distorts the orbital energies and leads to less reliable Koopmans-like approximations. Additionally, relaxation effects are more pronounced in open-shell ions, necessitating corrections beyond the frozen-orbital assumption. In open-shell contexts, the ionization energy for removing an electron from orbital i can be refined as ≈ -ε_i + Δ, where Δ accounts for self-interaction errors particularly affecting unpaired s, though this correction is often estimated variationally rather than explicitly in pure Hartree-Fock frameworks. These adaptations maintain the utility of Koopmans' theorem for qualitative predictions in radicals and transition metal complexes, but quantitative accuracy typically requires hybrid approaches to mitigate spin and relaxation issues.

Adaptations in correlated methods

In post-Hartree-Fock methods, Koopmans' theorem is adapted through the extended Koopmans' theorem (EKT), which employs Dyson orbitals to compute ionization potentials (IPs) using correlation-renormalized orbitals, yielding IP ≈ -ε_eff where ε_eff are effective orbital energies derived from correlated wave functions. Dyson orbitals, defined as the overlap between the (N-1)-electron ionized state Ψ_k and the N-electron Ψ_0 upon removal of an from orbital k, φ_d^k = ⟨Ψ_k | a_k | Ψ_0 ⟩, capture and orbital relaxation effects neglected in the original theorem. A generalized form of the theorem expresses the vertical IP as IP_k = E_0 - E_k + corrections involving the orbital overlap, where the term |⟨Ψ_k | a_k | Ψ_0 ⟩|^2 (the strength or residue) quantifies the contribution of the specific channel, approaching 1 for minimal relaxation and enabling accurate energies in correlated frameworks like configuration interaction () or coupled cluster (). In second-order Møller-Plesset (MP2) or CC with single and double excitations (CCSD), orbital energies are further refined using Brueckner orbitals, which are optimized to satisfy the Brillouin condition and minimize relaxation errors by incorporating double excitations into the reference determinant. For the water molecule, these adaptations significantly improve IP predictions: Hartree-Fock Koopmans' theorem yields 13.5 for the first (1b_1) ionization, while correlated methods like CCSD achieve 12.6 , matching the experimental value of 12.6 and reducing errors from and relaxation. While these correlated adaptations enhance accuracy to ~0.1 for valence IPs, they increase computational cost substantially due to the need for multi-reference wave functions and evaluations, limiting applicability to small systems compared to mean-field approaches.

Counterpart in density functional theory

In (DFT), an analogous result to Koopmans' theorem is the ionization potential (IP) theorem, which holds for the exact exchange-correlation functional and states that the negative of the highest occupied Kohn-Sham (KS) orbital energy equals the ionization potential of the N-electron system: IP = -ε_HOMO. This equality arises as an extension of Janak's theorem, which establishes that the total energy E varies with the occupation number n_i of each KS orbital i according to \frac{\partial E}{\partial n_i} = \varepsilon_i. For integer occupations, integrating over the HOMO occupation from n_HOMO = 1 to 0 directly yields IP = -ε_HOMO. The underlying reason for this theorem is the piecewise linearity of the exact KS total energy E(N) with respect to the number of electrons N between integer values, which ensures a constant μ within each interval and a discontinuity in the derivative (and thus in the KS potential) at integer N. This linearity condition, established in ensemble DFT, guarantees the exact matching of orbital energies to removal/addition energies at integer points. In practice, approximate functionals such as the local density approximation (LDA) or generalized gradient approximations (GGAs) violate this theorem because they lack the proper derivative discontinuity of the exchange-correlation potential, leading to curvature in E(N) and inaccurate orbital energies. Despite these shortcomings, approximate DFT often yields ionization potentials in better agreement with experiment than Hartree-Fock theory; for benzene, a hybrid GGA like B3LYP gives -ε_HOMO ≈ 9.2 eV, matching the experimental value of 9.24 eV, whereas Hartree-Fock overestimates it at ≈ 10.5 eV. Additionally, the partial self-interaction cancellation inherent in LDA and GGA functionals tends to improve estimates relative to Hartree-Fock, where self-interaction effects are uncorrected and often lead to underestimated (or negative) affinities for neutral molecules.

Orbital energies in many-body perturbation theory

In (MBPT), Koopmans' theorem emerges as a zeroth-order approximation to the quasiparticle energies, with Hartree-Fock () orbital energies serving as the starting point, and higher-order captured through the self-energy \Sigma. The , derived from Hedin's equations, provides a systematic improvement by treating the self-energy perturbatively, yielding quasiparticle energies \epsilon_{qp} via the equation in the diagonal approximation: \epsilon_{qp}^i = \epsilon_{HF}^i + \langle \phi_i | \Sigma(\epsilon_{qp}^i) - V_{xc} | \phi_i \rangle, where \phi_i are the HF orbitals, and V_{xc} accounts for the starting-point exchange-correlation potential (often from DFT in practice). This formulation refines the orbital picture by incorporating dynamic screening and correlation effects beyond the static HF mean field. In the GW method, the self-energy is approximated as \Sigma = i G W \Gamma, where G is the one-particle Green's function, W is the screened Coulomb interaction, and the vertex function \Gamma \approx 1 in the basic GW scheme, simplifying the full Hedin coupling to a tractable form. This leads to the negative GW highest occupied molecular orbital (HOMO) energy -\epsilon_{HOMO}^{GW} directly approximating the ionization potential (IP), achieving mean absolute errors of approximately 0.1 eV for frontier orbitals in molecules across benchmark sets like GW100. The approach extends effectively to solid-state systems for predicting band gaps, where GW quasiparticle corrections dramatically outperform HF estimates. For example, in silicon, GW yields an indirect band gap of 1.2 eV, closely matching the experimental value of 1.1 eV, whereas HF overestimates it to approximately 3.3 eV (about three times the experimental value) due to neglect of screening. This orbital-based quasiparticle framework, rooted in Hedin's equations, naturally connects to optical excitations via the Bethe-Salpeter equation (BSE), where GW energies form the basis for electron-hole interactions in the four-point response function, enabling accurate spectra beyond single-particle approximations.

References

  1. [1]
  2. [2]
    Accurate Electron Affinities from the Extended Koopmans' Theorem ...
    The extended Koopmans' theorem (EKT) provides a systematic way to compute electron affinities (EAs) from any level of theory. Although, it is widely applied ...Introduction · Theoretical Approach · Results and Discussion · References
  3. [3]
    Physical Interpretation of Koopmans' Theorem: A Criticism of the ...
    In this paper we present a criticism of current didactic presentation of Koopmans' theorem and we propose a new didactic approach.
  4. [4]
    Koopmans' Theorem in the Restricted Open-Shell Hartree−Fock ...
    A general formulation of Koopmans' theorem is derived for high-spin half-filled open shells in the restricted open-shell Hartree−Fock (ROHF) method based on ...
  5. [5]
    Validation of Koopmans' theorem for density functional theory ...
    For Hartree–Fock, HF, wavefunctions, Koopmans' theorem, KT, which states that the initial state BE = −ε ιs rigorous. However, the KT relationship is commonly ...
  6. [6]
  7. [7]
  8. [8]
    2. Hartree-Fock methods — Advanced Topics in Computational ...
    Koopman's theorem states that for example the ionization energy of a closed-shell system is given by the energy of the highest occupied single-particle state.
  9. [9]
    None
    ### Summary of Koopmans' Theorem Derivation in Hartree-Fock Theory for Closed-Shell Systems
  10. [10]
    Validity of the Extended Koopmans' Theorem - ACS Publications
    Mar 24, 2009 · Koopmans' theorem (1) is widely used to estimate ionization energies on the basis of Hartree−Fock calculations. In ...
  11. [11]
    [PDF] On the accuracy of density functional theory and wave ... - OSTI
    the errors for the HF ionization energies were found to be significantly ... Koopmans' theorem: Calculation of correlation and relaxation energies,” J.<|control11|><|separator|>
  12. [12]
    Extension of Koopmans' theorem. II. Accurate ionization energies ...
    Jan 1, 1975 · Koopmans' theorem does not yield a 2p ionization energy (1s2 2s2 → 1s2 2p), but the extended Koopmans' theorem yields an ionization energy ...<|control11|><|separator|>
  13. [13]
    Koopmans-Compliant Spectral Functionals for Extended Systems
    May 23, 2018 · A new study demonstrates how functional theory can be used to compute the spectral properties of semiconductors and insulators, ...
  14. [14]
    Koopmans' theorem in the Hartree-Fock method. General formulation
    This work presents a general formulation of Koopmans' theorem (KT) in the Hartree-Fock (HF) method which is applicable to molecular and atomic systems with ...
  15. [15]
    Ionization Potential, Electron Affinity, Electronegativity, Hardness ...
    The directly calculated ionization potential (IP), electron affinity (EA), electronegativity (χ), hardness (η), and first electron excitation energy (τ) are all ...<|control11|><|separator|>
  16. [16]
    Calculated Ionization Energy for Ne (Neon atom)
    Calculated Ionization Energy for Ne (Neon atom) ; 16.503 · 21.440 · 21.440 ; 15.546 · 20.262 · 20.262 ...
  17. [17]
    The Electron Affinity as the Highest Occupied Anion Orbital Energy ...
    ... approximation (Koopmans) to minus the ionization energy (IP) and the LUMO orbital energy is a Koopmans approximation to minus the electron affinity (EA).
  18. [18]
    LUMO (Molecular Orbital) - an overview | ScienceDirect Topics
    Electron affinities calculated via Koopmans' theorem are usually quite poor. The derivation of Koopmans' theorem assumes that the electronic wave function of ...
  19. [19]
    Theoretical study of stable negative ions of polar molecules
    The Koopmans' theorem approximation equates the electron affinity with - €(LUMO), where €(LUMO) is the orbital energy of the lowest unoccupied molecu- lar ...
  20. [20]
    Calculated Electron Affininty for F (Fluorine atom) - CCCBDB
    Calculated Electron Affininty for F (Fluorine atom). 18 06 04 13 53. Experimental Electron Affinity is 3.40129 ± 0.000003 eV. differences original data ...
  21. [21]
    Koopmans' theorem for inner-shell ionization - ScienceDirect
    Inner-shell ionization energies calculated from Koopmans' theorem are shown, by the sudden approximation, to relate to a weighted average energy for singly ...Missing: potential | Show results with:potential<|control11|><|separator|>
  22. [22]
    Koopmans' theorem in the ROHF method: Canonical form for the ...
    Nov 30, 2006 · Based on this analysis, we derive the new (canonical) form for the Hamiltonian of the Hartree-Fock equation, the eigenvalues of which obey ...
  23. [23]
    Orbital energies and Koopmans' theorem in open-shell Hartree-Fock ...
    Cite. https://doi.org/10.1016/0009-2614(72)80030-7 Get rights and content. Full text access. Abstract. In (restricted) open-shell Hartree-Fock theory, the ...Missing: formulation | Show results with:formulation
  24. [24]
    The extended Koopmans' theorem and its exactness - AIP Publishing
    The extended Koopmans' theorem (EKT) is shown to give accurate and potentially exact values for the lowest ionization potential (IP).Missing: seminal | Show results with:seminal
  25. [25]
    Exploring Dyson's Orbitals and Their Electron Binding Energies for ...
    Oct 7, 2021 · We argue that Dyson's orbitals enable description of the response states compatible with the concepts of molecular orbital theory.
  26. [26]
    Brueckner orbitals, Dyson orbitals, and correlation potentials
    Aug 6, 2025 · Computation of the ionization energies and of the respective Dyson's orbitals based on the use of the extended Koopmans' theorem (EKT) is ...
  27. [27]
    Near Hartree‐Fock Calculations on the Ground State of the Water ...
    The Hartree‐Fock vertical ionization potentials (in electron volts), 11.1(2B1), 13.3(2A1), and 17.6(2B2), are too low by 1–1.5 eV as expected.
  28. [28]
    Ab initio determination of the ionization potentials of water clusters ...
    Jun 26, 2012 · Both MP2 and CASPT2 give rise to similar values, whereas CCSD gives the best result, as expected from a single reference system as the neutral ...
  29. [29]
    The extended Koopmans' theorem for orbital-optimized methods
    Oct 17, 2013 · ... error of 0.11 eV. There is almost a 5-fold reduction in MAE compared to the Koopmans' theorem. Errors of the EKT-MP2.5 (MAE = 0.21 eV) , EKT ...
  30. [30]
    Density-Functional Theory for Fractional Particle Number: Derivative ...
    Dec 6, 1982 · Density-Functional Theory for Fractional Particle Number: Derivative Discontinuities of the Energy. John P. Perdew · Robert G. Parr · Mel Levy.Missing: linearity | Show results with:linearity
  31. [31]
    Proof that in density-functional theory | Phys. Rev. B
    Dec 15, 1978 · It is shown that the variation of the total energy, as constructed in density-functional theory, with respect to an orbital occupation is equal to the ...
  32. [32]
    On Koopmans' theorem in density functional theory - AIP Publishing
    Nov 1, 2010 · This paper clarifies why long-range corrected (LC) density functional theory gives orbital energies quantitatively.Missing: original | Show results with:original
  33. [33]
    Piecewise Linearity of Approximate Density Functionals Revisited
    Mar 19, 2013 · Here, we show that, contrary to conventional wisdom, most of the required piecewise linearity of an arbitrary approximate density functional can be restored.
  34. [34]
    Koopmans' condition for density-functional theory | Phys. Rev. B
    Sep 23, 2010 · Here, we first examine self-interaction in terms of the discrepancy between total and partial electron removal energies, and then highlight the ...Missing: paper | Show results with:paper
  35. [35]
    Accurate GW frontier orbital energies of 134 kilo molecules - Nature
    Sep 5, 2023 · According to recent reports, GW accuracy on various test sets reaches 0.1(0.2) eV, a factor of 2(4) larger than the chemical accuracy. Here we ...
  36. [36]
    The GW Miracle in Many-Body Perturbation Theory for the Ionization ...
    We use the GW100 benchmark set to systematically judge the quality of several perturbation theories against high-level quantum chemistry methods.
  37. [37]
    [PDF] The GW Approximation - Many-Body Perturbation Theory - CORE
    In this lecture we present many-body perturbation theory as a method to determine quasiparticle excitations in solids, especially electronic band structures, ...<|control11|><|separator|>