Gaussian orbital

Last updated

In computational chemistry and molecular physics, Gaussian orbitals (also known as Gaussian type orbitals, GTOs or Gaussians) are functions used as atomic orbitals in the LCAO method for the representation of electron orbitals in molecules and numerous properties that depend on these. [1]

Contents

Rationale

The use of Gaussian orbitals in electronic structure theory (instead of the more physical Slater-type orbitals) was first proposed by Boys [2] in 1950. The principal reason for the use of Gaussian basis functions in molecular quantum chemical calculations is the 'Gaussian Product Theorem', which guarantees that the product of two GTOs centered on two different atoms is a finite sum of Gaussians centered on a point along the axis connecting them. In this manner, four-center integrals can be reduced to finite sums of two-center integrals, and in a next step to finite sums of one-center integrals. The speedup by 4-5 orders of magnitude compared to Slater orbitals outweighs the extra cost entailed by the larger number of basis functions generally required in a Gaussian calculation.

For reasons of convenience, many quantum chemistry programs work in a basis of Cartesian Gaussians even when spherical Gaussians are requested, as integral evaluation is much easier in the cartesian basis, and the spherical functions can be simply expressed using the cartesian functions. [3] [4]

Mathematical form

The Gaussian basis functions obey the usual radial-angular decomposition

,

where is a spherical harmonic, and are the angular momentum and its component, and are spherical coordinates.

While for Slater orbitals the radial part is

being a normalization constant, for Gaussian primitives the radial part is

where is the normalization constant corresponding to the Gaussian.

The normalization condition which determines or is

which in general does not impose orthogonality in .

Because an individual primitive Gaussian function gives a rather poor description for the electronic wave function near the nucleus, Gaussian basis sets are almost always contracted:

,

where is the contraction coefficient for the primitive with exponent . The coefficients are given with respect to normalized primitives, because coefficients for unnormalized primitives would differ by many orders of magnitude. The exponents are reported in atomic units. There is a large library of published Gaussian basis sets optimized for a variety of criteria available at the Basis Set Exchange portal.

Cartesian coordinates

In Cartesian coordinates, Gaussian-type orbitals can be written in terms of exponential factors in the , , and directions as well as an exponential factor controlling the width of the orbital. The expression for a Cartesian Gaussian-type orbital, with the appropriate normalization coefficient is

In the above expression, , , and must be integers. If , then the orbital has spherical symmetry and is considered an s-type GTO. If , the GTO possesses axial symmetry along one axis and is considered a p-type GTO. When , there are six possible GTOs that may be constructed; this is one more than the five canonical d orbital functions for a given angular quantum number. To address this, a linear combination of two d-type GTOs can be used to reproduce a canonical d function. Similarly, there exist 10 f-type GTOs, but only 7 canonical f orbital functions; this pattern continues for higher angular quantum numbers. [5]

Molecular integrals

Taketa et al. (1966) presented the necessary mathematical equations for obtaining matrix elements in the Gaussian basis. [6] Since then much work has been done to speed up the evaluation of these integrals which are the slowest part of many quantum chemical calculations. Živković and Maksić (1968) suggested using Hermite Gaussian functions, [7] as this simplifies the equations. McMurchie and Davidson (1978) introduced recursion relations, [8] which greatly reduces the amount of calculations. Pople and Hehre (1978) developed a local coordinate method. [9] Obara and Saika introduced efficient recursion relations in 1985, [10] which was followed by the development of other important recurrence relations. Gill and Pople (1990) introduced a 'PRISM' algorithm which allowed efficient use of 20 different calculation paths. [11]

The POLYATOM System

The POLYATOM System [12] was the first package for ab initio calculations using Gaussian orbitals that was applied to a wide variety of molecules. [13] It was developed in Slater's Solid State and Molecular Theory Group (SSMTG) at MIT using the resources of the Cooperative Computing Laboratory. The mathematical infrastructure and operational software were developed by Imre Csizmadia, [14] Malcolm Harrison, [15] Jules Moskowitz [16] and Brian Sutcliffe. [17]

See also

Related Research Articles

<span class="mw-page-title-main">Atomic orbital</span> Mathematical function describing the location and behavior of an electron within an atom

In atomic theory and quantum mechanics, an atomic orbital is a function describing the location and wave-like behavior of an electron in an atom. This function can be used to calculate the probability of finding any electron of an atom in any specific region around the atom's nucleus. The term atomic orbital may also refer to the physical region or space where the electron can be calculated to be present, as predicted by the particular mathematical form of the orbital.

<span class="mw-page-title-main">Wave function</span> Mathematical description of the quantum state of a system

In quantum physics, a wave function is a mathematical description of the quantum state of an isolated quantum system. The wave function is a complex-valued probability amplitude, and the probabilities for the possible results of measurements made on the system can be derived from it. The most common symbols for a wave function are the Greek letters ψ and Ψ.

Electron density or electronic density is the measure of the probability of an electron being present at an infinitesimal element of space surrounding any given point. It is a scalar quantity depending upon three spatial variables and is typically denoted as either or . The density is determined, through definition, by the normalised -electron wavefunction which itself depends upon variables. Conversely, the density determines the wave function modulo up to a phase factor, providing the formal foundation of density functional theory.

Slater-type orbitals (STOs) are functions used as atomic orbitals in the linear combination of atomic orbitals molecular orbital method. They are named after the physicist John C. Slater, who introduced them in 1930.

In computational physics and chemistry, the Hartree–Fock (HF) method is a method of approximation for the determination of the wave function and the energy of a quantum many-body system in a stationary state.

<span class="mw-page-title-main">Gaussian integral</span> Integral of the Gaussian function, equal to sqrt(π)

The Gaussian integral, also known as the Euler–Poisson integral, is the integral of the Gaussian function over the entire real line. Named after the German mathematician Carl Friedrich Gauss, the integral is

Q-Chem is a general-purpose electronic structure package featuring a variety of established and new methods implemented using innovative algorithms that enable fast calculations of large systems on various computer architectures, from laptops and regular lab workstations to midsize clusters and HPCC, using density functional and wave-function based approaches. It offers an integrated graphical interface and input generator; a large selection of functionals and correlation methods, including methods for electronically excited states and open-shell systems; solvation models; and wave-function analysis tools. In addition to serving the computational chemistry community, Q-Chem also provides a versatile code development platform.

In rotordynamics, the rigid rotor is a mechanical model of rotating systems. An arbitrary rigid rotor is a 3-dimensional rigid object, such as a top. To orient such an object in space requires three angles, known as Euler angles. A special rigid rotor is the linear rotor requiring only two angles to describe, for example of a diatomic molecule. More general molecules are 3-dimensional, such as water, ammonia, or methane.

A multipole expansion is a mathematical series representing a function that depends on angles—usually the two angles used in the spherical coordinate system for three-dimensional Euclidean space, . Similarly to Taylor series, multipole expansions are useful because oftentimes only the first few terms are needed to provide a good approximation of the original function. The function being expanded may be real- or complex-valued and is defined either on , or less often on for some other .

In theoretical and computational chemistry, a basis set is a set of functions that is used to represent the electronic wave function in the Hartree–Fock method or density-functional theory in order to turn the partial differential equations of the model into algebraic equations suitable for efficient implementation on a computer.

In quantum chemistry, a configuration state function (CSF), is a symmetry-adapted linear combination of Slater determinants. A CSF must not be confused with a configuration. In general, one configuration gives rise to several CSFs; all have the same total quantum numbers for spin and spatial parts but differ in their intermediate couplings.

<span class="mw-page-title-main">Spartan (chemistry software)</span>

Spartan is a molecular modelling and computational chemistry application from Wavefunction. It contains code for molecular mechanics, semi-empirical methods, ab initio models, density functional models, post-Hartree–Fock models, and thermochemical recipes including G3(MP2) and T1. Quantum chemistry calculations in Spartan are powered by Q-Chem.

In atomic, molecular, and optical physics and quantum chemistry, the molecular Hamiltonian is the Hamiltonian operator representing the energy of the electrons and nuclei in a molecule. This operator and the associated Schrödinger equation play a central role in computational chemistry and physics for computing properties of molecules and aggregates of molecules, such as thermal conductivity, specific heat, electrical conductivity, optical, and magnetic properties, and reactivity.

STO-nG basis sets are minimal basis sets, where primitive Gaussian orbitals are fitted to a single Slater-type orbital (STO). originally took the values 2 – 6. They were first proposed by John Pople. A minimum basis set is where only sufficient orbitals are used to contain all the electrons in the neutral atom. Thus for the hydrogen atom, only a single 1s orbital is needed, while for a carbon atom, 1s, 2s and three 2p orbitals are needed. The core and valence orbitals are represented by the same number of primitive Gaussian functions . For example, an STO-3G basis set for the 1s, 2s and 2p orbital of the carbon atom are all linear combination of 3 primitive Gaussian functions. For example, a STO-3G s orbital is given by:

A hydrogen-like atom (or hydrogenic atom) is any atom or ion with a single valence electron. These atoms are isoelectronic with hydrogen. Examples of hydrogen-like atoms include, but are not limited to, hydrogen itself, all alkali metals such as Rb and Cs, singly ionized alkaline earth metals such as Ca+ and Sr+ and other ions such as He+, Li2+, and Be3+ and isotopes of any of the above. A hydrogen-like atom includes a positively charged core consisting of the atomic nucleus and any core electrons as well as a single valence electron. Because helium is common in the universe, the spectroscopy of singly ionized helium is important in EUV astronomy, for example, of DO white dwarf stars.

The Yoshimine sort is an algorithm that is used in quantum chemistry to order lists of two electron repulsion integrals. It is implemented in the IBM Alchemy program suite and in the UK R-matrix package for electron and positron scattering by molecules which is based on the early versions of the IBM Alchemy program suite.

<span class="mw-page-title-main">Cubic harmonic</span>

In fields like computational chemistry and solid-state and condensed matter physics the so-called atomic orbitals, or spin-orbitals, as they appear in textbooks on quantum physics, are often partially replaced by cubic harmonics for a number of reasons. These harmonics are usually named tesseral harmonics in the field of condensed matter physics in which the name kubic harmonics rather refers to the irreducible representations in the cubic point-group.

<span class="mw-page-title-main">Riemann–Silberstein vector</span>

In mathematical physics, in particular electromagnetism, the Riemann–Silberstein vector or Weber vector named after Bernhard Riemann, Heinrich Martin Weber and Ludwik Silberstein, is a complex vector that combines the electric field E and the magnetic field B.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

The linearized augmented-plane-wave method (LAPW) is an implementation of Kohn-Sham density functional theory (DFT) adapted to periodic materials. It typically goes along with the treatment of both valence and core electrons on the same footing in the context of DFT and the treatment of the full potential and charge density without any shape approximation. This is often referred to as the all-electron full-potential linearized augmented-plane-wave method (FLAPW). It does not rely on the pseudopotential approximation and employs a systematically extendable basis set. These features make it one of the most precise implementations of DFT, applicable to all crystalline materials, regardless of their chemical composition. It can be used as a reference for evaluating other approaches.

References

  1. Gill, Peter M.W. (1994). "Molecular integrals Over Gaussian Basis Functions" (PDF). Advances in Quantum Chemistry. 25: 141–205. Bibcode:1994AdQC...25..141G. doi:10.1016/S0065-3276(08)60019-2. ISBN   9780120348251 . Retrieved 17 June 2011.
  2. Boys, S. F. (1950). "Electronic Wave Functions. I. A General Method of Calculation for the Stationary States of Any Molecular System". Proc. R. Soc. Lond. A. 200 (1063): 542–554. Bibcode:1950RSPSA.200..542B. doi:10.1098/rspa.1950.0036. JSTOR   98423. S2CID   122709395.
  3. Schlegel, H.; Frisch, M. (1990). "Transformation between Cartesian and pure spherical harmonic Gaussians". International Journal of Quantum Chemistry. 54 (2): 83–87. doi:10.1002/qua.560540202. S2CID   94417974.
  4. Mathar, Richard J. (2002). "Mutual Conversion of Three Flavors of Gaussian Type Orbitals". International Journal of Quantum Chemistry. 90 (1): 227–243. arXiv: physics/9907051 . Bibcode:2002IJQC...90..227M. doi:10.1002/qua.10085. S2CID   119100125.
  5. Cramer, Christopher J. (2004). Essentials of computational chemistry : theories and models (2nd ed.). Chichester, West Sussex, England: Wiley. p. 167. ISBN   9780470091821.
  6. Taketa, Hiroshi; Huzinaga, Sigeru; O-ohata, Kiyosi (1966). "Gaussian-Expansion Methods for Molecular Integrals". Journal of the Physical Society of Japan. 21 (11): 2313–2324. Bibcode:1966JPSJ...21.2313T. doi:10.1143/JPSJ.21.2313.
  7. Živković, T.; Maksić, Z. B. (1968). "Explicit Formulas for Molecular Integrals over Hermite-Gaussian Functions". Journal of Chemical Physics. 49 (7): 3083–3087. Bibcode:1968JChPh..49.3083Z. doi:10.1063/1.1670551.
  8. McMurchie, Larry E.; Davidson, Ernest R. (1978). "One- and two-electron integrals over Cartesian Gaussian functions". Journal of Computational Physics. 26 (2): 218–31. Bibcode:1978JCoPh..26..218M. doi:10.1016/0021-9991(78)90092-X.
  9. Pople, J. A.; Hehre, W. J. (1978). "Computation of electron repulsion integrals involving contracted Gaussian basis functions". J. Comput. Phys. 27 (2): 161–168. Bibcode:1978JCoPh..27..161P. doi:10.1016/0021-9991(78)90001-3.
  10. Obara, S.; Saika, A. (1986). "Efficient recursive computation of molecular integrals over Cartesian Gaussian functions". J. Chem. Phys. 84 (7): 3963–74. Bibcode:1986JChPh..84.3963O. doi:10.1063/1.450106.
  11. Gill, Peter M. W.; Pople, John A. (December 1991). "The Prism Algorithm for Two-Electron Integrals" (PDF). International Journal of Quantum Chemistry. 40 (6): 753–772. doi:10.1002/qua.560400605 . Retrieved 17 June 2011.
  12. Csizmadia, I.G.; Harrison, M.C.; Moskowitz, J.W.; Sutcliffe, B.T. (1966). "Nonempirical LCAO-MO-SCF-CI calculations on organic molecules with gaussian-type functions. Introductory review and mathematical formalism". Theoretica Chimica Acta. 6 (3): 191. doi:10.1007/BF02394698. S2CID   198176437.
  13. A.C. Wahl, Chemistry by computer, Scientific American, pages 54-70, April, 1970.
  14. Imre Csizmadia, Professor Emeritus of Chemistry, University of Toronto, in Reviews in Computational Chemistry, vol.15, p.248
  15. Malcolm C. Harrison, Professor of Computer Science, New York University
  16. Jules W. Moskowitz, Professor Emeritus of Chemistry, New York University
  17. Brian T. Sutcliffe, Professor of Chemistry, York University