Introduction to Quantum Mechanics and Spectroscopy 3 Credits Fall Semester 2014 Laura Gagliardi. Lecture 27, December 5, 2014

Similar documents
Chem 4502 Introduction to Quantum Mechanics and Spectroscopy 3 Credits Fall Semester 2014 Laura Gagliardi. Lecture 28, December 08, 2014

A One-Slide Summary of Quantum Mechanics

Chemistry 3502 Physical Chemistry II (Quantum Mechanics) 3 Credits Fall Semester 2003 Christopher J. Cramer. Lecture 25, November 5, 2003

Semiempirical Molecular Orbital Theory

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 17, March 1, 2006

CHEM3023: Spins, Atoms and Molecules

Chem 4502 Introduction to Quantum Mechanics and Spectroscopy 3 Credits Fall Semester 2014 Laura Gagliardi. Lecture 21, November 12, 2014

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 22, March 20, 2006

The successful wavefunction can be written as a determinant: # 1 (2) # 2 (2) Electrons. This can be generalized to our 2N-electron wavefunction:

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester Christopher J. Cramer. Lecture 30, April 10, 2006

Computational Material Science Part II. Ito Chao ( ) Institute of Chemistry Academia Sinica

2 Electronic structure theory

Molecular Simulation I

Same idea for polyatomics, keep track of identical atom e.g. NH 3 consider only valence electrons F(2s,2p) H(1s)

SCF calculation on HeH +

One-Electron Hamiltonians

Electronic structure theory: Fundamentals to frontiers. 1. Hartree-Fock theory

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 31, April 12, 2006

This is a very succinct primer intended as supplementary material for an undergraduate course in physical chemistry.

APPENDIX 2. Semiempirical PM3 (Parametrization Method #3) When analytical solutions cannot be obtained, Schrödinger equation may be solved:

CHEM3023: Spins, Atoms and Molecules

Exercise 1: Structure and dipole moment of a small molecule

Session 1. Introduction to Computational Chemistry. Computational (chemistry education) and/or (Computational chemistry) education

Gustavus Adolphus College. Lab #5: Computational Chemistry

Yingwei Wang Computational Quantum Chemistry 1 Hartree energy 2. 2 Many-body system 2. 3 Born-Oppenheimer approximation 2

We also deduced that transformations between Slater determinants are always of the form

Pauli Deformation APPENDIX Y

Introduction to Hartree-Fock Molecular Orbital Theory

CHAPTER 11 MOLECULAR ORBITAL THEORY

Introduction to computational chemistry Exercise I: Structure and electronic energy of a small molecule. Vesa Hänninen

( ) + ( ) + ( ) = 0.00

Introduction to numerical projects

26 Group Theory Basics

Introduction to Electronic Structure Theory

Lecture 5: More about one- Final words about the Hartree-Fock theory. First step above it by the Møller-Plesset perturbation theory.

Quantum mechanics can be used to calculate any property of a molecule. The energy E of a wavefunction Ψ evaluated for the Hamiltonian H is,

Basis Sets and Basis Set Notation

Identical Particles in Quantum Mechanics

AN INTRODUCTION TO QUANTUM CHEMISTRY. Mark S. Gordon Iowa State University

Instructor background for the discussion points of Section 2

Quantum Mechanical Simulations

MODELING MATTER AT NANOSCALES

CHEMISTRY 4021/8021 MIDTERM EXAM 1 SPRING 2014

Jack Simons, Henry Eyring Scientist and Professor Chemistry Department University of Utah

Lecture 4: Hartree-Fock Theory

Self-consistent Field

CHEMISTRY Topic #1: Bonding What Holds Atoms Together? Spring 2012 Dr. Susan Lait

Section 3 Electronic Configurations, Term Symbols, and States

Lecture 6. Tight-binding model

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 7, February 1, 2006

Lecture 9. Hartree Fock Method and Koopman s Theorem

0 belonging to the unperturbed Hamiltonian H 0 are known

Density Functional Theory

Lecture 6. Fermion Pairing. WS2010/11: Introduction to Nuclear and Particle Physics

Chemistry 334 Part 2: Computational Quantum Chemistry

Electronic structure calculations: fundamentals George C. Schatz Northwestern University

Introduction to DFTB. Marcus Elstner. July 28, 2006

CHEM-UA 127: Advanced General Chemistry I

Quantum Chemical Simulations and Descriptors. Dr. Antonio Chana, Dr. Mosè Casalegno

I. CSFs Are Used to Express the Full N-Electron Wavefunction

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 9, February 8, 2006

Tentative content material to be covered for Exam 2 (Wednesday, November 2, 2005)

MO theory is better for spectroscopy (Exited State Properties; Ionization)

1 Rayleigh-Schrödinger Perturbation Theory

Consequently, the exact eigenfunctions of the Hamiltonian are also eigenfunctions of the two spin operators

The Hartree-Fock approximation

(Refer Slide Time: 5:21 min)

CHEMISTRY 333 Fall 2006 Exam 3, in class

Mean-field concept. (Ref: Isotope Science Facility at Michigan State University, MSUCL-1345, p. 41, Nov. 2006) 1/5/16 Volker Oberacker, Vanderbilt 1

3: Many electrons. Orbital symmetries. l =2 1. m l

Chemistry 3502/4502. Exam III. All Hallows Eve/Samhain, ) This is a multiple choice exam. Circle the correct answer.

ATOMIC STRUCTURE. Atomic Structure. Atomic orbitals and their energies (a) Hydrogenic radial wavefunctions

Jack Simons Henry Eyring Scientist and Professor Chemistry Department University of Utah

Density Functional Theory. Martin Lüders Daresbury Laboratory

1 Introduction. 2 The hadronic many body problem

Chemistry 2000 Lecture 2: LCAO-MO theory for homonuclear diatomic molecules

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Fall Semester 2006 Christopher J. Cramer. Lecture 5, January 27, 2006

Introduction to Computational Chemistry Computational (chemistry education) and/or. (Computational chemistry) education

PRACTICE PROBLEMS Give the electronic configurations and term symbols of the first excited electronic states of the atoms up to Z = 10.

VALENCE Hilary Term 2018

PAPER:2, PHYSICAL CHEMISTRY-I QUANTUM CHEMISTRY. Module No. 34. Hückel Molecular orbital Theory Application PART IV

Lecture 08 Born Oppenheimer Approximation

Mechanics, Heat, Oscillations and Waves Prof. V. Balakrishnan Department of Physics Indian Institute of Technology, Madras

Lecture 9 Electronic Spectroscopy

Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

NMR and IR spectra & vibrational analysis

This is called a singlet or spin singlet, because the so called multiplicity, or number of possible orientations of the total spin, which is

Handbook of Computational Quantum Chemistry. DAVID B. COOK The Department of Chemistry, University of Sheffield

Answers Quantum Chemistry NWI-MOL406 G. C. Groenenboom and G. A. de Wijs, HG00.307, 8:30-11:30, 21 jan 2014

Introduction to Computational Chemistry

Hartree-Fock-Roothan Self-Consistent Field Method

On the Uniqueness of Molecular Orbitals and limitations of the MO-model.

MOLECULAR STRUCTURE. The general molecular Schrödinger equation, apart from electron spin effects, is. nn ee en

v(r i r j ) = h(r i )+ 1 N

We will use C1 symmetry and assume an RHF reference throughout. Index classes to be used include,

Multiplicity, Instability, and SCF Convergence Problems in Hartree Fock Solutions

Chemistry 431. Lecture 14. Wave functions as a basis Diatomic molecules Polyatomic molecules Huckel theory. NC State University

CHEM3023: Spins, Atoms and Molecules

Chem 442 Review for Exam 2. Exact separation of the Hamiltonian of a hydrogenic atom into center-of-mass (3D) and relative (3D) components.

Transcription:

Chem 4502 Introduction to Quantum Mechanics and Spectroscopy 3 Credits Fall Semester 2014 Laura Gagliardi Lecture 27, December 5, 2014 (Some material in this lecture has been adapted from Cramer, C. J. Essentials of Computational Chemistry, Wiley, Chichester: 2002; pp. 118-119, 154-158.) Solved Homework Consider the diagonal and off-diagonal kinetic energy matrix elements T µµ = µ 1 2 2 µ T µν = µ 1 2 2 ν Let's think about what the Laplacian operator does to the function on which it is operating. It reports back the sum of second derivatives in all coordinate directions. That is, it is a measure of how fast the slope of the function is changing in various directions. If two functions µ and ν are far apart, then since good basis functions go to zero at least exponentially fast with distance, ν is likely to be very flat where µ is large. The second derivative of a flat function is zero. So, every point in the integration will be roughly the amplitude of µ times zero, and not much will accumulate. For the diagonal element, on the other hand, the interesting second derivatives will occur where the function has maximum amplitude (amongst other places) so the accumulation should be much larger. Thus, we expect diagonal elements in this case to be larger than off-diagonal elements. Analogous arguments can be made for the nuclear attraction integrals. V µµ = µ nuclei Z k k r k µ V µν = µ nuclei k Z k r k ν The 1/r operator acting on ν will ensure that the largest contribution to the overall integral will come from the nucleus k on which basis function ν resides. Unless µ also has significant amplitude around that nucleus, it will multiply the result by roughly zero and the whole integral will be small. In this case, however, it is conceivable (i) that offdiagonal elements involving Z =100 might be larger than diagonal elements about Z =1 and (ii) that off-diagonal elements involving two basis functions on the same nuclear center might compete with diagonal elements involving a more diffuse basis function on that center (e.g., a 1s2s interaction on the same atom might be larger than a 5f interaction with itself, since the latter is so spread out that its amplitude isn't very large anywhere).

27-2 The 2J K terms are somewhat more subtle to evaluate, but it will still generally be the case that diagonal terms will generate larger magnitudes than off-diagonal terms. We will not explore the point closely. The Hartree-Fock Self-Consistent Field Procedure To find the HF MOs we need to solve the HF secular determinant F 11 ES 11 F 12 ES 12! F 1N ES 1N F 21 ES 21 F 22 ES 22! F 2N ES 2N " " # " = 0 (27-1) F N1 ES N1 F N2 ES N2! F NN ES NN and find its various roots. We know that we can compute overlap integrals and that Fock matrix elements are defined by nuclei F µν = µ 1 2 2 ν Z k µ 1 ν k r k + P λσ λσ & ( µν λσ) 1 µλ νσ ' 2 ( ) ( ) (27-2) So, now for the tricky part. Remember that the density matrix elements appearing in eq. 27-2 are defined as occupied P λσ = 2 a λi a σi (27-3) where the a values are the coefficients of the basis functions in the occupied molecular orbitals. But, the whole point of solving the secular equation is to find those coefficients. So, if we don't know them already, how can we compute the matrix elements, and if we do know them already, why are we doing this in the first place? This is, of course, the same paradox we saw in the Hartree formalism. And, just as in the Hartree method, the HF method follows a self-consistent field (SCF) procedure, where first we guess the orbital coefficients (e.g., from an effective Hamiltonian method, like Hückel theory or a modification thereof) and then we iterate to convergence. The full process is described schematically by the flow-chart on the next page. The flow chart includes consideration of optimizing the geometry, but we'll ignore that for now. i

27-3 Choose a basis set Choose a molecular geometry q (0) Compute and store all overlap, one-electron, and two-electron integrals Guess initial density matrix P (0) Construct and solve Hartree- Fock secular equation Replace P (n 1) with P (n) Construct density matrix from occupied MOs Choose new geometry according to optimization algorithm no Is new density matrix P (n) sufficiently similar to old density matrix P (n 1)? yes no Optimize molecular geometry? yes no Does the current geometry satisfy the optimization criteria? Output data for unoptimized geometry yes Output data for optimized geometry After convergence of the MOs, one can compute the HF energy by evaluating the Hamiltonian operator for the HF determinant (the determinant will be formed in the usual

27-4 way given all the occupied orbitals to enforce antisymmetry). There is, however, a slightly easier way, too. If we think about the Fock operator for a particular electron, it is, remember f i = 1 2 i 2 Z k HF + V k r i j ik nuclei { } (27-4) It only operates on an electron in the ith orbital, so if we take the expectation value of the Fock operator over that orbital call it ε i it represents the total energy of an electron in that orbital. eq. 27-4 includes the attraction of an electron in the orbital to all the nuclei, its kinetic energy, and its repulsion with all the other electrons. So, if we add up all the ε values for all of the occupied orbitals and double that (because in restricted HF we have two electrons in every orbital), we'll have included all of the physical contributions to the energy. However, we'll have overcounted electronelectron repulsion. That's because for an electron in orbital 1, we evaluate the repulsion that electron feels from an electron in orbital 2 (and 3, and 4, etc.) But, when we go and consider the electron in orbital 2 (and 3, and 4, etc.) we account for that interaction again in dictating the energy of electron 2 (etc.) So, we double count all of the electron repulsions. Thus, an alternative way of computing the total energy is to add up all of the individual orbital energies and deduct half of the total electron repulsion energy (cf. eq. 26-8). That is occupied E HF = Ψ HF H Ψ HF = 2 ε i 2J ij K ij (27-5) i occupied Hartree-Fock theory as constructed using the Roothaan approach is quite beautiful in the abstract. This is not to say, however, that it does not suffer from certain chemical and practical limitations. It s chief chemical limitation is the one-electron nature of the Fock operators. Other than exchange, all electron correlation is ignored. It is, of course, an interesting question to ask just how important such correlation is for various molecular properties, and we will examine that later on. Furthermore, from a practical standpoint, HF theory posed some very challenging technical problems to early quantum chemists. One problem was choice of a basis set. The LCAO approach using hydrogenic orbitals remains attractive in principle, however, this basis set requires numerical solution of the four-index integrals appearing in the Fock matrix elements, and that is a very tedious process (no analytic solution has yet been found). Moreover, the number of four-index integrals is daunting. Since each index runs over the total number of basis functions, there are in principle N 4 total integrals to be evaluated, and this quartic scaling behavior with respect to basis-set size proves to be the bottleneck in HF theory applied to essentially any molecule. Historically, two philosophies began to emerge at this stage with respect to how best to make further progress. The first philosophy might be summed up as follows: The i, j

27-5 HF equations are very powerful but still, after all, chemically flawed since they ignore electron-electron correlation. Thus, other approximations that may be introduced to simplify their solution, and possibly at the same time improve their accuracy (by some sort of parameterization to reproduce key experimental quantities), are well justified. Many quantum chemists continue to be guided by this philosophy today, and it underlies the motivation for so-called semiempirical MO theories, which, alas, we lack time to touch upon. The second philosophy essentially views HF theory as a stepping stone on the way to exact solution of the Schrödinger equation. HF theory provides a very well defined energy, one which can be converged in the limit of an infinite basis set, and the difference between that converged energy and reality is the electron correlation energy (ignoring relativity, spin-orbit coupling, etc.). It was anticipated that developing the technology to achieve the HF limit with no further approximations would not only permit the evaluation of the chemical utility of the HF limit, but also probably facilitate moving on from that low-altitude base camp to the Schrödinger equation summit. Such was the foundation for further research on ab initio (Latin for "from the beginning") HF theory. Along the way it became clear that, perhaps surprisingly, HF energies could be chemically useful. Typically their utility was manifest for situations where the error associated with ignoring the correlation energy could be made unimportant by virtue of comparing two or more systems for which the errors could be made to cancel. In any case, we will examine ab initio HF theory next. Gaussian Orbitals as Basis Functions The basis set is the set of mathematical functions from which the wave function is constructed. As detailed in this and earlier lectures, each MO in HF theory is expressed as a linear combination of basis functions, the coefficients for which are determined from the iterative solution of the HF SCF equations (as flowcharted above). The full HF wave function is expressed as a Slater determinant formed from the individual occupied MOs. In the abstract, the HF limit is achieved by use of an infinite basis set, which necessarily permits an optimal description of the electron probability density. In practice, however, one cannot make use of an infinite basis set. Thus, much work has gone into identifying mathematical functions that allow wave functions to approach the HF limit arbitrarily closely in as efficient a manner as possible. Efficiency in this case involves three considerations. As noted above, in the absence of additional simplifying approximations like those present in semiempirical theory, the number of two-electron integrals increases as N 4 where N is the number of basis functions. So, keeping the total number of basis functions to a minimum is computationally attractive. In addition, however, it can be useful to choose basis set functional forms that permit the various integrals appearing in the HF equations to be evaluated in a computationally efficient fashion. Thus, a larger basis set can still represent a computational improvement over a smaller basis set if evaluation of the greater number of integrals for the former can be carried out faster than for the latter.

27-6 Finally, the basis functions must be chosen to have a form that is useful in a chemical sense. That is, the functions should have large amplitude in regions of space where the electron probability density (the wave function) is also large, and small amplitudes where the probability density is small. The simultaneous optimization of these three considerations is at the heart of basis set development. We previously derived, by solution of the one-electron-atom Schrödinger equation, hydrogenic orbitals; recall that they are characterized by exponential decay that goes as the first power of the radial distance from the atom on which they are centered. In quantum calculations, such orbitals are usually called Slater -type orbitals (STOs, named for the same Slater who lent his name to the determinantal wave function). While STOs have the attractive feature that they are based on hydrogenic atomic orbitals, in ab initio HF theory they suffer from a fairly significant limitation. There is no analytical solution available for the general four-index integral (eq. 27-10) when the basis functions are STOs. The requirement that such integrals be solved by numerical methods severely limits their utility in molecular systems of any significant size. In 1950, Boys proposed an alternative to the use of STOs. All that is required for there to be an analytical solution of the general four-index integral formed from such functions is that the radial decay of the STOs be changed from e r to e r2. That is, the AO-like functions are chosen to have the form of a gaussian function. The general functional form of a normalized gaussian-type orbital (GTO) in atom-centered cartesian coordinates is ( ) = 2α % π φ x, y, z;α,i, j, k $ & ' ( ) (27-6) 3 / 4 ( 8α) i+ j+k 1/ 2 ( i! j!k! + * ) ( 2i)! ( 2 j)! ( 2k)! - x i y j z k e α x2 + y 2 +z 2, where α is an exponent controlling the width of the GTO, and i, j, and k are non-negative integers that dictate the nature of the orbital in a cartesian sense. In particular, when all three of these indices are zero, the GTO has spherical symmetry, and is called an s-type GTO. When exactly one of the indices is one, the function has axial symmetry about a single cartesian axis and is called a p-type GTO. There are three possible choices for which index is one, corresponding to the p x, p y, and p z orbitals. When the sum of the indices is equal to two, the orbital is called a d-type GTO, etc. Although they are convenient from a computational standpoint, GTOs have specific features that do diminish their utility as basis functions. One issue of key concern is the shape of the radial portion of the orbital. For s type functions, GTOs are smooth and differentiable at the nucleus (r = 0), but real hydrogenic AOs have a cusp (see Figure at end of Lecture 19). In addition, all hydrogenic AOs have a radial decay that is exponential in r while the decay of GTOs is exponential in r 2 ; this results in too rapid a reduction in amplitude with distance for the GTOs.

27-7 As we saw previously in discussing variational atomic calculations, in order to combine the best feature of GTOs (computational efficiency) with that of STOs (proper radial shape), most of the first basis sets developed with GTOs used them as building blocks to approximate STOs. That is, the basis functions ϕ used for SCF calculations were not individual GTOs, but instead a linear combination of GTOs fit to reproduce as accurately as possible a STO, i.e., ( ) = c a φ a x, y, z;α a,i, j, k ϕ x, y, z; { α},i, j, k M ( ) (27-7) where M is the number of gaussians used in the linear combination, and the coefficients c are chosen to optimize the shape of the basis function sum and ensure normalization (perhaps by variational calculations on the atom, as we did earlier in the course). When a basis function is defined as a linear combination of gaussians, it is referred to as a contracted basis function, and the individual gaussians from which it is formed are called primitive gaussians. Thus, in a basis set of contracted GTOs, each basis function is defined by the contraction coefficients c and exponents α of each of its primitives. The degree of contraction refers to the total number of primitives used to make all of the contracted functions, as described in more detail below. Contracted GTOs when used as basis functions continue to permit analytical evaluation of all of the four-index integrals. Optimal contraction coefficients and exponents for mimicking STOs with contracted GTOs have been developed for a large number of atoms in the periodic table using different choices of M in eq. 27-7. These different basis sets are called STO-MG, for Slater-Type Orbital approximated by M Gaussians. Obviously, the more primitives that are employed, the more accurately a contracted function can be made to match a given STO. However, note that a four-index two-electron integral becomes increasingly complicated to evaluate as each individual basis function is made up of increasingly many primitive functions, according to ( µν λσ ) = ϕ µ ( 1)ϕ ν 1 ( ) 1 r 12 ϕ λ 2 ( )ϕ σ ( 2)dr 1 dr 2 a=1 M µ M ν = c aµ φ aµ ( 1) c aν φ aν 1 ( ) 1 c aλ φ aλ ( 2) c aσ φ aσ r 12 a µ =1 a ν =1 a λ =1 a σ =1 M λ M σ 2 ( )dr 1 dr 2 (27-8) M µ M ν = c aµ c aν c aσ c aλ φ aµ ( 1) φ aν 1 a µ =1 a ν =1 M λ M σ a λ =1 a σ =1 ( ) 1 r 12 φ aλ 2 ( )φ aσ ( 2)dr 1 dr 2 The optimum combination of speed and accuracy (when comparing to calculations using STOs) was achieved for M = 3. The figure at the end of Lecture 19 compares a 1s function using the STO-3G formalism to the corresponding STO and shows also the 3 primitives from which the contracted basis function is constructed. STO-3G basis functions have been defined for most of the atoms in the periodic table.

27-8 Gaussian functions have another feature that would be undesirable if they were to be used individually to represent atomic orbitals: they fail to exhibit radial nodal behavior! That is, no choice of variables permits eq. 27-6 to mimic a 2s orbital, which is negative near the origin and positive beyond a certain radial distance. Use of a contraction scheme, however, alleviates this problem; contraction coefficients c in eq. 27-7 can be chosen to have either negative or positive sign, and thus fitting to functions having radial nodal behavior poses no special challenges. To make a 2s function you would simply subtract a tighter gaussian from a looser one, in which case it would indeed be negative near the nucleus and positive further away. Homework To be solved in class: Consider the water molecule. If we decide to do a calculation on water and use the STO- 3G basis set, how many contracted basis functions will we need in order to minimally represent the total number of atomic orbitals spanned by the core and valence electrons of the oxygen atom and the two hydrogen atoms? How many one-electron integrals will there be that require evaluation? How many two-electron integrals will require evaluation? In each of the last two cases, how many primitive integrals will need to be evaluated? Do you see anything that makes the workload slightly less onerous than your formal analysis? How many occupied orbitals will there be in the final Slater determinant? Repeat the above problem but now for phosphine (PH 3 ) instead of water.