Fact-checked by Grok 2 weeks ago

Slater-type orbital

A Slater-type orbital (STO), also referred to as a Slater-type function (STF), is an analytical expression used in to approximate the wave function of an in an , serving as a in methods like the (LCAO) for building molecular orbitals. Introduced by American physicist John C. Slater in 1930, STOs extend simple hydrogen-like orbitals by incorporating shielding effects via an effective nuclear charge, providing a simplified yet physically motivated model for multi-electron s. The mathematical form of an STO consists of a radial component proportional to r^{n-1} e^{-\zeta r}, where r is the radial distance from the nucleus, n is the principal quantum number, and \zeta is the orbital exponent that controls the decay rate, multiplied by the appropriate spherical harmonic for the angular part. This exponential decay ensures STOs satisfy Kato's cusp condition at the nucleus—describing the sharp behavior of the wave function near the atomic core—and exhibit the correct long-range asymptotic falloff, mimicking the exact solutions for hydrogenic atoms. Despite these desirable properties, STOs present computational challenges in quantum chemistry calculations, particularly in evaluating multicenter integrals required for molecular systems, which has historically limited their direct use in large-scale methods. To address this, STOs are often approximated by sums of Gaussian-type orbitals (), as in basis sets like STO-nG, enabling efficient computation while retaining much of the accuracy for atomic and molecular properties such as energies, geometries, and spectra. Ongoing research continues to refine techniques for direct STO implementations, enhancing precision in areas like electron correlation and relativistic effects.

Definition and Motivation

Historical Development

Slater-type orbitals were introduced by John C. Slater in 1930 to provide a simplified yet effective approximation for the wavefunctions of multi-electron atoms in the context of the emerging -Fock self-consistent field method. In his paper "Note on 's Method," published in , Slater proposed representing the radial portion of atomic orbitals with a single , which captured the essential screening effects of inner s on outer ones, thereby making numerical solutions more tractable compared to the full procedure. This innovation addressed the computational challenges posed by electron correlations in atoms beyond the hydrogen-like cases, enabling approximate analytical treatments that advanced early quantum mechanical calculations. The motivation for Slater-type orbitals stemmed from the limitations of exact hydrogenic functions in multi-electron systems, where electron shielding required adjustments that were difficult to handle numerically; Slater's approach offered a parameterized form that balanced physical with ease of use in atomic structure computations. During the 1930s and 1940s, STOs were widely adopted in for atomic and studies, as they facilitated the estimation of energies and properties with reasonable accuracy while avoiding the intensive numerical integrations of the time. A pivotal advancement came in the 1950s with Clemens C. J. Roothaan's development of the of atomic orbitals-molecular orbital (LCAO-MO) method, which incorporated STOs as basis functions to expand molecular wavefunctions, laying the foundation for molecular calculations as outlined in his influential 1951 review. Slater provided further refinements to STO parameters in his 1960 book Quantum Theory of Atomic Structure, where he systematically derived screening constants and orbital exponents through a combination of variational principles and empirical fits to spectroscopic data, solidifying their role in . However, by the , the computational advantages of Gaussian-type orbitals began to overshadow STOs, as the latter's multi-center integrals proved analytically challenging despite their superior cusp and asymptotic behavior; this shift was accelerated by the 1969 work of Hehre, Stewart, and Pople, which demonstrated efficient Gaussian expansions of STOs, culminating in the release of the Gaussian 70 program that popularized GTO-based methods.

Basic Concept

Slater-type orbitals (STOs) are exponentially decaying functions designed to approximate the radial distribution of electrons in multi-electron atoms, offering a practical balance between physical accuracy and computational feasibility in quantum chemistry calculations. Introduced by John C. Slater in 1930 as simple analytic forms for atomic wave functions based on shielding concepts, STOs provide an effective means to model electron behavior without resorting to fully numerical solutions. The primary motivation for STOs stems from the need to approximate solutions to the for atoms beyond hydrogen, where electron-electron interactions complicate exact treatments. By incorporating an that accounts for screening by inner electrons, STOs enable efficient estimation of orbital shapes and energies, bypassing the intensive required in full Hartree-Fock methods for multi-electron systems. This approach facilitates broader applications in atomic and molecular modeling while maintaining reasonable fidelity to observed atomic properties. A key advantage of STOs lies in their ability to accurately capture the cusp-like behavior of near the and the at large distances, features that align closely with quantum mechanical solutions and provide superior physical realism compared to basis functions with or other decay forms. However, their evaluation in multi-center integrals for molecular systems poses significant computational challenges, limiting their direct use in large-scale calculations without approximations. Conceptually, STOs function as single-zeta basis sets, each representing a single effective orbital per atomic through adjustment of the screening parameter to simulate the reduced attraction experienced by electrons. This streamlined form underscores their role as versatile tools for initial approximations in , emphasizing electron penetration and shielding effects in a computationally tractable manner.

Mathematical Formulation

Functional Form

The Slater-type orbital (STO) is expressed in spherical coordinates as \psi_{nlm}(r, \theta, \phi) = N r^{n-1} e^{-\zeta r} Y_{lm}(\theta, \phi), where r is the radial distance from the nucleus in atomic units (a_0 = 1), N is the normalization constant, \zeta is the orbital exponent, and Y_{lm}(\theta, \phi) are the spherical harmonics. This form provides a simple analytical approximation to atomic orbitals, capturing the exponential decay characteristic of bound states while incorporating screening effects through \zeta. A common normalized form for the radial component of the STO is R_{nl}(r) = \sqrt{ \left( \frac{2\zeta}{n} \right)^{2l+1} \frac{(n-l-1)!}{2n [(n+l)!]} } \, e^{-\zeta r} \, r^{l} \, L_{n-l-1}^{2l+1}(2\zeta r), but in the simplest STO approximation, the associated Laguerre polynomial L is omitted or approximated by a constant, reducing to R_{nl}(r) \propto r^{n-1} e^{-\zeta r}. The normalization ensures \int_0^\infty R_{nl}^2(r) r^2 \, dr = 1. For the nodeless approximation often used, the power r^{n-1} provides the correct number of radial nodes without polynomial complexity. The angular part is specified by the Y_{lm}(\theta, \phi), which depend on the m and determine the orbital's symmetry (e.g., Y_{00} = 1/\sqrt{4\pi} for s orbitals). These functions are orthonormal over the unit sphere, \int Y_{lm}^* Y_{l'm'} \, d\Omega = \delta_{ll'} \delta_{mm'}, ensuring the total orbital integrates to unity when combined with the normalized radial part. This functional form derives from the exact hydrogen-like orbitals by replacing the nuclear charge Z with an effective charge Z_\mathrm{eff} = Z - \sigma to account for electron screening, while simplifying the radial part to a power-law term for computational tractability. The resulting STOs mimic the cusp at the nucleus and asymptotic decay of true atomic orbitals.

Parameters and Screening

In Slater-type orbitals (STOs), the orbital exponent, denoted \zeta, serves as a measure of the effective nuclear charge Z_\mathrm{eff} experienced by an electron, expressed as Z_\mathrm{eff} = Z - \sigma, where Z is the atomic number and \sigma quantifies the shielding from electron-electron repulsion. The exponent is parameterized as \zeta = Z_\mathrm{eff}/n = (Z - \sigma)/n, where n is the principal quantum number. This adjusts the radial decay of the orbital to account for the reduced attraction felt by outer electrons due to the intervening charge cloud of inner electrons. The physical basis of screening lies in the electrostatic within multi-electron atoms: inner-shell electrons effectively reduce the charge perceived by electrons, leading to a slower tuned by \zeta rather than the full Z. This shielding arises from the cumulative repulsion among electrons, modeled approximately as a constant \sigma. As a result, STOs capture the qualitative behavior of atomic orbitals by incorporating this effective charge, mimicking the cusp and asymptotic decay near the seen in exact solutions. The value of \sigma is obtained through , variational minimization of the total energy in self-consistent field calculations, or empirical adjustments to match spectroscopic or Hartree-Fock results. For the 1s orbital, \sigma = 0 yields \zeta = 1. For first-row atoms, valence orbital \zeta values increase with Z, such as approximately 1.69 for 1s, 1.06 for 2s, and 1.83 for 2p in minimal basis sets using . These values reflect the growing nuclear charge outpacing the additional screening from more electrons. Sensitivity to \zeta is pronounced, as small changes alter orbital extent and stability: larger \zeta contracts the orbital, decreasing the expectation value \langle r \rangle roughly as n / (2\zeta) and stabilizing the energy by making it more negative (scaling approximately as -\zeta^2 / 2 for 1s-like functions in variational treatments). Conversely, smaller \zeta expands the orbital, raising the energy and increasing spatial delocalization, which can significantly impact predicted atomic radii and ionization potentials if \zeta is not accurately tuned. This dependence highlights \zeta's role in balancing computational simplicity with physical fidelity in STO approximations.

Approximations and Properties

Relation to Hydrogen-like Orbitals

Hydrogen-like orbitals represent the exact solutions to the for one-electron atoms with nuclear charge Z, characterized by quantum numbers n, l, and m. The wave function is expressed as \psi_{nlm}(r, \theta, \phi) = R_{nl}(r) Y_{lm}(\theta, \phi), where Y_{lm} are and the radial function is R_{nl}(r) = \sqrt{\left( \frac{2Z}{n a_0} \right)^3 \frac{(n-l-1)!}{2n [(n+l)!]}} \, e^{-\rho/2} \rho^l L_{n-l-1}^{2l+1}(\rho), with \rho = 2 Z r / (n a_0) and L denoting associated . This form incorporates radial nodes determined by the Laguerre polynomials and ensures orthogonality among orbitals with different quantum numbers. Slater-type orbitals (STOs) approximate these hydrogenic orbitals by simplifying the radial part to a single-exponential form without radial nodes or polynomial variations: R(r) \propto (2 \zeta r)^l e^{-\zeta r}, multiplied by the angular spherical harmonics Y_{lm}, where the exponent \zeta is an effective orbital parameter. This approximation collapses the multi-nodal structure of the exact hydrogenic radial functions into a nodeless, analytically tractable expression, motivated by the need for computational simplicity in multi-electron systems. STOs accurately reproduce key physical features of hydrogenic orbitals, including the cusp at the nucleus—where the wave function's logarithmic derivative satisfies Kato's condition \frac{\partial \ln |\psi|}{\partial r} \big|_{r=0} = -Z for s-orbitals—and the correct exponential decay at long range, \exp(-2 \sqrt{2 I} r), with I the ionization potential. However, the absence of radial nodes leads to underestimated orthogonality between orbitals of the same l but different n, as the overlap integrals \int R_{n l}^{\text{STO}}(r) R_{n' l}^{\text{H}}(r) r^2 dr (where superscript H denotes hydrogenic) are non-zero, unlike in the exact case. Mathematical comparisons via overlap integrals between STOs and hydrogenic functions demonstrate reasonable fidelity, particularly for inner-shell orbitals, with explicit formulas available for evaluation. Error metrics for radial densities, such as shifts in the position of maxima r_{\max}, reveal discrepancies; for example, in lithium, the 2s STO places r_{\max} at 4.57 a_0 compared to 3.55 a_0 in Hartree-Fock (approximating hydrogenic for light atoms), indicating overestimation of orbital extent in valence regions. Screening effects further adjust the effective Z in STO exponents to account for electron-electron interactions, improving overall approximation quality.

Slater's Rules for Exponents

Slater introduced empirical rules in for determining the screening constants σ used in the exponents ζ = Z - σ of Slater-type orbitals, providing a simple way to account for the shielding effects of inner electrons on outer orbitals in multi-electron atoms. These rules were derived from fitting approximate wave functions to known atomic properties, such as potentials and sizes, without solving the full many-electron . The rules involve grouping the electrons into shells: (1s), (2s,2p), (,3p), (), (4s,4p), (), (), etc. For an in a given group, the screening constant σ is the sum of contributions from other electrons: 0 from electrons in outer groups (higher n), 0.35 from each other electron in the same group (except 0.30 for the other electron in the 1s group), 0.85 from each electron in the (n-1) group, and 1.00 from each electron in all deeper groups. For or electrons, the d or f groups are treated separately from s,p, with the same contribution values. The is then ζ = Z - σ, which serves as the orbital exponent (sometimes scaled by 1/n in early formulations). For example, in neutral lithium (1s² 2s¹), for the 2s electron: no other electrons in (2s,2p) group, so 0; two 1s electrons contribute 2 × 0.85 = 1.70; thus σ = 1.70, ζ = 3 - 1.70 = 1.30. These values ensure the orbital's radial extent roughly matches that of hydrogen-like atoms with reduced nuclear charge, facilitating quick estimates in early quantum chemical calculations. Subsequent refinements improved upon these empirical choices by optimizing exponents through numerical Hartree-Fock calculations. In 1974, Clementi and Roetti compiled tables of STO exponents derived from minimizing the energy expectation value for atomic ground states, offering more accurate values than Slater's rules, particularly for core orbitals and heavier elements. These optimized exponents, often denoted as minimal basis set parameters, reduce errors in computed atomic energies and properties compared to the original empirical approximations. Despite their utility, exhibit limitations, as they rely on group-averaged screening that oversimplifies interactions in transition metals, where d-electron contributions lead to variable shielding not captured by the fixed adjustments. This results in less reliable exponents for systems involving partially filled d or f shells, prompting the use of optimized tables for precise applications.

Computational Aspects

Evaluation of Integrals

In quantum chemical calculations using Slater-type orbitals (STOs), one-electron form the foundation for constructing the core elements. The overlap S_{ab} = \int \phi_a(\mathbf{r}) \phi_b(\mathbf{r}) \, d\mathbf{r} between two STOs \phi_a and \phi_b is essential for assessing orbital non-orthogonality. For s-type STOs, the radial component is [ \zeta_a \zeta_b / (\zeta_a + \zeta_b) ]^{3/2}, multiplied by explicit angular factors that account for the relative orientation and separation of the orbital centers; for normalized 1s STOs on the same center, this simplifies to S_{ab} = 8 (\zeta_a \zeta_b)^{3/2} / (\zeta_a + \zeta_b)^3. The T_{aa} = -\frac{1}{2} \int \phi_a(\mathbf{r}) \nabla^2 \phi_a(\mathbf{r}) \, d\mathbf{r} for an s-type STO on a single center evaluates to \zeta_a^2 / 2, reflecting the orbital's effective screening through the exponent \zeta_a. For off-diagonal elements T_{ab} between different centers, analytical expressions involve derivatives of the overlap , incorporating inter-center distance and dependencies. The V_{ab} = -\sum_A Z_A \int \phi_a(\mathbf{[r](/page/R)}) \phi_b(\mathbf{[r](/page/R)}) / |\mathbf{[r](/page/R)} - \mathbf{R}_A| \, d\mathbf{[r](/page/R)}, where the sum is over centers A with charges Z_A, captures electron- interactions. For a 1s STO centered on A, the one-center term is Z_A \zeta_a, derived from the expectation value \langle 1/[r](/page/R) \rangle = \zeta_a; two-center variants require expansion in auxiliary functions or for closed-form evaluation. Two-electron integrals, particularly the repulsion type (ab|cd) = \iint \phi_a(\mathbf{r}_1) \phi_b(\mathbf{r}_1) (1/r_{12}) \phi_c(\mathbf{r}_2) \phi_d(\mathbf{r}_2) \, d\mathbf{r}_1 d\mathbf{r}_2, are central to electron correlation in methods like Hartree-Fock. These contribute to Coulomb (J) and exchange (K) terms, with analytical forms often expressed via incomplete gamma functions or Boys function analogs adapted for STOs. For the simplest ss|ss case (all s-type STOs), the one-center expression involves $2\pi^{5/2} / (\zeta_a + \zeta_b + \zeta_c + \zeta_d)^3 scaled by additional factors from angular integration; multi-center variants, such as two-center Coulomb integrals, use prolate spheroidal coordinates for exact evaluation. Evaluating four-center two-electron over STOs presents significant challenges due to their analytical complexity and the O(N^4) scaling with basis size N in molecular systems, where each requires handling multiple inter-center distances and exponents. This computational bottleneck historically motivated approximations, such as fitting STOs to sums of Gaussian-type orbitals to facilitate .

Numerical and Analytical Methods

Analytical methods for evaluating integrals over Slater-type orbitals (STOs) often rely on closed-form expressions involving incomplete gamma functions, particularly for electron repulsion integrals with non-integer principal quantum numbers. These expressions facilitate exact computation by transforming radial parts into hypergeometric series or finite sums via recurrence relations, ensuring across a range of parameters. For overlap and one-electron integrals, the Löwdin alpha function provides closed-form solutions based on exponential integrals akin to gamma functions, enabling arbitrary precision without . Analogs to the Boys function, such as the generalized function G_m(T, U), extend these techniques to multicenter STO integrals by incorporating incomplete gamma forms that mirror the Gaussian Boys function F_m(T), allowing analytical treatment of interactions. relations adapted for higher momenta, similar to Obara-Saika schemes but tailored for in STOs, reduce complex integrals to lower-order overlap-like terms using functions and vertical transfers, improving efficiency for d- and f-type orbitals. These , often involving auxiliary functions for buildup, have been detailed in early works on multicenter expansions. Numerical methods address cases where analytical forms are intractable, employing techniques like Gauss-Legendre for two-center overlap integrals over STOs, which approximate the radial integrals with high accuracy using weighted sums over finite points. The McMurchie-Davidson scheme, while originally for Gaussians, has been adapted in hybrid approaches for STOs via Gaussian expansions, facilitating numerical evaluation of difficult two-electron integrals by decomposing products into Hermite-like intermediates. For relativistic STOs, grid-based on radial meshes uses finite differences to compute integrals numerically, accommodating components and Dirac-Coulomb interactions with controlled accuracy. Efficiency improvements in STO computations stem from contracting multiple primitive STOs into fewer basis functions, reducing the number of integrals by fixing linear combinations that preserve accuracy while minimizing computational cost. This contraction is particularly beneficial in large basis sets, where parallelization on GPUs accelerates numerical of Slater integrals through OpenACC directives, achieving speedups for high-angular-momentum cases without loss of precision. Historically, approximate evaluation in the leveraged Rys polynomials for integrals over Gaussian approximations to STOs, providing a quadrature-based that efficiently handles the $1/r_{12} via orthogonal expansions, paving the way for modern hybrid techniques.

Applications and Implementations

Role in

Slater-type orbitals (STOs) are fundamental basis functions in , particularly within the Hartree-Fock , where they expand molecular orbitals according to the Roothaan-Hall equations. This approach approximates the multi-electron wavefunction, facilitating calculations of ground-state energies and properties for molecules and atoms with high fidelity when using optimized STO exponents. In post-Hartree-Fock methods, such as configuration interaction and the for many-body , STOs enable correlated treatments that capture electron interactions beyond mean-field level, yielding accurate ionization potentials and band gaps. STOs excel in applications requiring precise atomic property descriptions, as they closely mimic numerical Hartree-Fock radial functions for elements across the periodic table, often achieving errors below 0.1% in energies and expectation values. Their exponential form accurately models through the nuclear cusp condition, outperforming Gaussian-type orbitals () in regions near the where peaks sharply. In relativistic contexts, STOs integrated with the Douglas-Kroll-Hess provide an effective two-component for heavy-element systems, decoupling positive and negative energy states to compute spin-orbit effects and scalar relativistic corrections efficiently. For modeling, STOs underpin Slater's method, which approximates core-level electron binding energies by averaging initial and final states, enabling simulations of and emission spectra in complexes with chemical accuracy. Compared to , STOs offer advantages in capturing the correct short-range cusp for descriptions and long-range for anions, though they introduce challenges like basis set superposition error in intermolecular interactions due to overcomplete basis sets in molecules. In modern embedded cluster methods, STOs describe active sites in extended materials, such as metal surfaces or zeolites, by combining high-accuracy quantum treatment of clusters with electrostatic embedding of the environment, improving predictions of adsorption and reactivity. Furthermore, Slater-type orbitals facilitate quantum electrodynamic (QED) corrections in relativistic calculations for few-electron atoms, with non-integer variants used in some approaches; these incorporate and effects to achieve sub-microhartree precision in energy levels. While STO integrals pose evaluation difficulties compared to , analytical techniques mitigate this for practical use.

Software and Basis Sets

The STO-nG family of basis sets approximates each Slater-type orbital (STO) through a least-squares fit of n to match the radial behavior of the STO, prioritizing the outer for chemical relevance. Developed by Hehre, Stewart, and Pople in 1969, the STO-3G variant—using three GTOs per —provides a minimal basis suitable for large molecules, with parameters optimized for atoms from to ; the fitting minimizes the weighted squared ∫ [R_STO(r) - Σ c_i exp(-α_i r^2)]^2 r^2 dr, where weights emphasize valence regions, yielding orbital errors typically under 0.01 and bond length predictions accurate to 0.01 for first- and second-row hydrides. This approximation facilitates efficient evaluation while retaining much of the STO's cusp and asymptotic decay properties, though it slightly overestimates core densities. Double-zeta STO basis sets, pioneered by Clementi and Raimondi in 1963, use separate exponents for core and valence shells to better capture radial nodal structure, offering improved flexibility over minimal sets; these were optimized via Hartree-Fock calculations for first-row atoms (Li-Ne) and later extended, providing energy accuracies within 0.1 hartree for atomic properties compared to numerical Hartree-Fock results. Polarized extensions, such as STO-3G*, augment the STO-3G approximation with d-functions on second-row atoms and p-functions on hydrogen, enhancing description of polarization effects in molecular environments; this set maintains the minimal size while reducing basis set superposition errors in intermolecular interactions by up to 50% relative to unpolarized STO-3G. Major quantum chemistry software packages incorporate STO-based basis sets, often via approximations for computational efficiency. The Gaussian program supports the full STO-nG family (n=1 to 6) as built-in minimal bases, enabling Hartree-Fock and DFT calculations with automatic integral generation over the fitted . Similarly, GAMESS-US includes STO-3G and related sets for methods, leveraging libraries for rapid evaluation of multicenter integrals in molecular systems up to hundreds of atoms. For exact STO handling, the package in the Modeling Suite employs unapproximated STOs as its core basis functions, optimized for with frozen-core techniques to handle heavy elements efficiently. integrates STO approximations, such as in the r^2SCAN-3c composite method, for hybrid DFT applications requiring quick, balanced accuracy across the periodic table. Recent advancements in the have introduced high-precision STO basis sets for broader applicability. For instance, the GW100 TZ and QZ sets, developed for many-body , provide triple- and quadruple-zeta quality with diffuse and functions, yielding energies within 1 kcal/mol of complete basis set limits for molecules. Optimized STO sets covering elements 1–118, refined for all-electron calculations, support high-accuracy benchmarks in transition metals and s, with radial exponents tuned to reproduce experimental potentials to 0.1 . As of 2025, STO basis sets continue to be applied in specialized studies, such as all-electron TZ2P sets for electronic structures revealing trends in ligand asymmetry, and STO-based Kohn–Sham densities for accurate descriptions. These updates, building on earlier fitting strategies, emphasize segmented contractions and even-tempered exponent sequences to minimize linear dependence while enhancing convergence in post-Hartree-Fock methods.

References

  1. [1]
    [PDF] mi 19960103 042 - DTIC
    Slater-type orbitals (STOs), i.e., exponential type orbitals, would be desirable for basis sets in quantum chemistry and molecular physics because they can ...
  2. [2]
  3. [3]
    (PDF) Molecular Integrals over Slater-type Orbitals. From pioneers to ...
    Slater orbitals are the natural basis functions in quantum molecular calculations. Their use was hindered over the last four decades by integration problems.
  4. [4]
  5. [5]
    Atomic Shielding Constants | Phys. Rev.
    Abstract. In analogy with the method of Zener for the atoms from Li to F, simple rules are set up giving approximate analytic atomic wave functions for all the ...Missing: concept | Show results with:concept
  6. [6]
    [PDF] Analytical Methods Using Slater-Type Orbitals in Quantum Chemistry
    Our main concern is with the development of a Slater-type orbital (STO) multicenter molecular integral package for use with standard ab initio quantum.
  7. [7]
    [PDF] slater screening constants in atomic and molecular orbitals
    Thus the concept of a screening constant for each type of orbital represents an approximation. ... In 1930, Slater made the use of screening constants simple.Missing: original | Show results with:original
  8. [8]
  9. [9]
    Alternative Approach to Calculate Two-Center Overlap Matrix ...
    In this work, we propose an alternative approach to evaluate two-center overlap integrals. It is computationally more efficient than the standard procedure ...
  10. [10]
    Calculation of two-center integrals between Slater-type orbitals
    Two-center overlap, nuclear-attraction and kinetic energy integrals between Slater-type orbitals (STOs) centered on two different points are calculated.
  11. [11]
    [PDF] Analytical Evaluation of Two-Center Kinetic Energy Integral for ...
    A new way of computing kinetic energy integral over Slater type orbitals is proposed. This article is concerned with the construction of the general formulae ...
  12. [12]
    Analytical evaluation for two-center nuclear attraction integrals over ...
    Jul 15, 2008 · In this study, we shall suggest analytical expressions for two-center nuclear attraction integrals over STO's with a one-center charge
  13. [13]
    An analytic method for three‐center nuclear attraction integrals: A ...
    A completely analytic method for evaluating three-center nuclear attraction integrals for STOS is presented. The method exploits a separation of the STO ...
  14. [14]
    Calculation of two-centre two-electron integrals over Slater-type ...
    Oct 22, 2014 · In this paper, which constitutes the first part of the series, we consider calculation of two-centre Coulomb and hybrid integrals over Slater-type orbitals ( ...
  15. [15]
    [PDF] arXiv:2212.13911v3 [quant-ph] 4 Jan 2023
    Jan 4, 2023 · The electron repulsion integrals over the Slater−type orbitals with non−integer principal quantum numbers are considered.
  16. [16]
    Evaluation of Slater-type geminal integrals using tailored Gaussian ...
    In his pioneering work, he formulated them in terms of the special function called G m ( T , U ) , in analogy with the Boys function F m ( T ) in the ERI ...Missing: analogs | Show results with:analogs
  17. [17]
    Unified kernel function approach to two-center integrations in ...
    As an example, two-center overlap integrals over Slater- type orbitals are calculated recursively for higher angular momentum quantum numbers and compared ...
  18. [18]
    [PDF] UCRL 13 791 '•<* - OSTI.GOV
    The one electron integrals can be further classified as. (a) those that can be done in closed form and (b) those that require the same numerically approximated ...
  19. [19]
    Gauss Quadrature Calculation for Two-Centre Overlap Integrals of ...
    Abstract. Abstract: The Guass-Legendre numerical quadrature method was introduced into the calculation of STO two-centre overlap integrals.
  20. [20]
    Efficient GPU Implementation of the McMurchie–Davidson Method ...
    Oct 10, 2025 · Quantum chemistry offers the formal machinery to derive molecular and physical properties arising from (sub)atomic interactions.
  21. [21]
    Numerical Procedures for Relativistic Atomic Structure Calculations
    Aug 4, 2004 · ... orbitals are represented by numerical values on a grid using finite differences for integration and differentiation. The numerical accuracy ...1. Introduction · 4.3. The Fermi Nucleus · 8. Slater Integrals
  22. [22]
    Automatic Generation of Accurate and Cost-Efficient Auxiliary Basis ...
    Sep 4, 2023 · ... primitive functions is key to making the approach computationally efficient. But, when the present contraction scheme is used with NAOs, it ...
  23. [23]
    [PDF] The Numerical Evaluation of Slater Integrals on GPU
    This article presents SlaterGPU, a graphics processing unit (GPU) accelerated library that uses OpenACC to numerically compute Slater-type orbital (STO) ...<|control11|><|separator|>
  24. [24]
    Evaluation of molecular integrals over Gaussian basis functions
    This paper is concerned with the efficient computation of the ubiquitous electron repulsion integral in molecular quantum mechanics.
  25. [25]
    GW100: A Slater-Type Orbital Perspective - ACS Publications
    To this end, we present two new Slater-type orbital (STO) basis sets of triple-(TZ) and quadruple-ζ (QZ) quality, whose polarization is adequate for correlated ...
  26. [26]
    Scalar relativistic all-electron density functional calculations on ...
    Feb 16, 2005 · This approach is based on the third-order Douglas–Kroll–Hess approximation, allowing the treatment of all electrons on an equal footing.
  27. [27]
  28. [28]
    The moderately-large-embedded-cluster method for metal surfaces
    The electronic structure of the substrate slab model is described using a linear-combination-of-Slater-type-orbitals method.
  29. [29]
    Chapter 3: Accurate Relativistic Calculations Including QED ...
    In our present approach, we have applied Slater-type spinor orbitals with non-integer principal quantum numbers (NISTOs) [29, 20,21] centered on the nuclei.
  30. [30]
    Basis Sets | Gaussian.com
    May 17, 2021 · Most methods require a basis set be specified; if no basis set keyword is included in the route section, then the STO-3G basis will be used. ...
  31. [31]
    Basis sets and atomic fragments — ADF 2025.1 documentation - SCM
    The basis functions used in ADF are commonly known as Slater Type Orbitals (STOs). The ADF package comes with a database of STO basis sets. The basis sets are ...
  32. [32]
    Optimization of the r 2 SCAN-3c Composite Electronic-Structure ...
    Jun 2, 2022 · In this work, we present an optimized Slater-type atomic orbital (STO) variant of the composite r2SCAN-3c DFT method with a customized all- ...