Vibrations and Rotations of Diatomic Molecules

Similar documents
Molecular Physics. Attraction between the ions causes the chemical bond.

Vibrational and Rotational Analysis of Hydrogen Halides

5.1 Classical Harmonic Oscillator

Chem120a : Exam 3 (Chem Bio) Solutions

Lecture 5: Harmonic oscillator, Morse Oscillator, 1D Rigid Rotor

Molecular spectroscopy Multispectral imaging (FAFF 020, FYST29) fall 2017

eigenvalues eigenfunctions

Diatomic Molecules. 7th May Hydrogen Molecule: Born-Oppenheimer Approximation

r R A 1 r R B + 1 ψ(r) = αψ A (r)+βψ B (r) (5) where we assume that ψ A and ψ B are ground states: ψ A (r) = π 1/2 e r R A ψ B (r) = π 1/2 e r R B.

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

Introduction to Quantum Mechanics (Prelude to Nuclear Shell Model) Heisenberg Uncertainty Principle In the microscopic world,

Exercises 16.3a, 16.5a, 16.13a, 16.14a, 16.21a, 16.25a.

Chemistry 795T. NC State University. Lecture 4. Vibrational and Rotational Spectroscopy

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,

Structure of diatomic molecules

Model for vibrational motion of a diatomic molecule. To solve the Schrödinger Eq. for molecules, make the Born- Oppenheimer Approximation:

The Harmonic Oscillator: Zero Point Energy and Tunneling

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

CHEM3023: Spins, Atoms and Molecules

Intro/Review of Quantum

Problem Set 9: Solutions

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

Lecture 10 Diatomic Vibration Spectra Harmonic Model

CHAPTER 6 Quantum Mechanics II

2m 2 Ze2. , where δ. ) 2 l,n is the quantum defect (of order one but larger

Intro/Review of Quantum

Fun With Carbon Monoxide. p. 1/2

CHAPTER 13 LECTURE NOTES

A few principles of classical and quantum mechanics

Chemistry 483 Lecture Topics Fall 2009

The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r

1 Reduced Mass Coordinates

Rotations and vibrations of polyatomic molecules

Lecture 10. Central potential

CHAPTER 8 The Quantum Theory of Motion

A Quantum Mechanical Model for the Vibration and Rotation of Molecules. Rigid Rotor

Lecture 10. Born-Oppenheimer approximation LCAO-MO application to H + The potential energy surface MOs for diatomic molecules. NC State University

Molecular energy levels

Principles of Molecular Spectroscopy

Atoms, Molecules and Solids (selected topics)

ONE AND MANY ELECTRON ATOMS Chapter 15

More On Carbon Monoxide

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

V( x) = V( 0) + dv. V( x) = 1 2

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

van Quantum tot Molecuul

Spectra of Atoms and Molecules. Peter F. Bernath

Chemistry 881 Lecture Topics Fall 2001

1. For the case of the harmonic oscillator, the potential energy is quadratic and hence the total Hamiltonian looks like: d 2 H = h2

PHYS 3313 Section 001 Lecture # 22

Physics 342 Lecture 23. Radial Separation. Lecture 23. Physics 342 Quantum Mechanics I

The Hydrogen Atom Chapter 20

A Brief Introduction to the Quantum Harmonic Oscillator

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep

Molecular orbitals, potential energy surfaces and symmetry

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 10, February 10, / 4

NPTEL/IITM. Molecular Spectroscopy Lectures 1 & 2. Prof.K. Mangala Sunder Page 1 of 15. Topics. Part I : Introductory concepts Topics

Quantum mechanics (QM) deals with systems on atomic scale level, whose behaviours cannot be described by classical mechanics.

P3317 HW from Lecture 15 and Recitation 8

CHAPTER 6 Quantum Mechanics II

Angular Momentum Quantization: Physical Manifestations and Chemical Consequences

Total Angular Momentum for Hydrogen

We now turn to our first quantum mechanical problems that represent real, as

The one and three-dimensional particle in a box are prototypes of bound systems. As we

Harmonic Oscillator with raising and lowering operators. We write the Schrödinger equation for the harmonic oscillator in one dimension as follows:

Brief review of Quantum Mechanics (QM)

CHAPTER 6 Quantum Mechanics II

Vibrational states of molecules. Diatomic molecules Polyatomic molecules

Chemistry 2000 Lecture 1: Introduction to the molecular orbital theory

(Refer Slide Time: 5:21 min)

The Wave Function. Chapter The Harmonic Wave Function

Lecture #21: Hydrogen Atom II

The Hydrogen Atom. Chapter 18. P. J. Grandinetti. Nov 6, Chem P. J. Grandinetti (Chem. 4300) The Hydrogen Atom Nov 6, / 41

Introduction to Computational Chemistry

Numerical Solution of a Potential Final Project

Brief introduction to molecular symmetry

Problem Set 5: Solutions

Introduction to Vibrational Spectroscopy

Quantum Theory of Angular Momentum and Atomic Structure

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom

we have to deal simultaneously with the motion of the two heavy particles, the nuclei

Computational Spectroscopy III. Spectroscopic Hamiltonians

Physics 216 Spring The Variational Computation of the Ground State Energy of Helium

VIBRATION-ROTATION SPECTRUM OF CO

Joel Broida UCSD Fall 2009 Phys 130B QM II. Homework Set 4

Physics 113!!!!! Spring 2009!! Quantum Theory Seminar #9

PHYS3113, 3d year Statistical Mechanics Tutorial problems. Tutorial 1, Microcanonical, Canonical and Grand Canonical Distributions

1.3 Harmonic Oscillator

( ) electron gives S = 1/2 and L = l 1

Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine

Physical Chemistry I Fall 2016 Second Hour Exam (100 points) Name:

Atoms, Molecules and Solids (selected topics)

Physical Chemistry Laboratory II (CHEM 337) EXPT 9 3: Vibronic Spectrum of Iodine (I2)

23 The Born-Oppenheimer approximation, the Many Electron Hamiltonian and the molecular Schrödinger Equation M I

Chem 442 Review of Spectroscopy

Lecture 6 Quantum Mechanical Systems and Measurements

Physics 342 Lecture 22. The Hydrogen Atom. Lecture 22. Physics 342 Quantum Mechanics I

Physics 228 Today: April 22, 2012 Ch. 43 Nuclear Physics. Website: Sakai 01:750:228 or

1 Commutators (10 pts)

Lecture 18 Long Wavelength Spectroscopy

Transcription:

Chapter 6 Vibrations and Rotations of Diatomic Molecules With the electronic part of the problem treated in the previous chapter, the nuclear motion shall occupy our attention in this one. In many ways the motion of two nuclei in a potential well formed by the electron cloud is among the simpler of all quantum mechanical problems. The reason is that it readily reduces to the motion of a single particle in a potential well. The derivation of how that comes about is given in detail in case some have not seen it before. A single particle in a potential well is also what the hydrogen atom is and it is useful to reflect on the differences. For hydrogen, one is interested in the motion of the electron in the coulombic field of the nucleus. It too is a two body problem, but the reduced mass is almost the same as the electron mass. For the diatomic molecule the reduced mass is half the mass of either nucleus for homonuclear molecules but always on the order of nuclear masses, not the electronic mass. More importantly, the attractive force is not coulombic but rather harmonic; it becomes stronger as the distance between the nuclei gets larger. As a field theory professor of mine remarked to his class, all of quantum mechanics is the simple harmonic oscillator! He was not thinking of the diatomic molecule when he said that but the diatomic molecule affords an excellent example of a system for which the simple harmonic oscillator is a good approximation. When studying an electron in the hydrogen atom one can readily imagine the spherical coordinates used to locate the electron from the origin which is positioned at the proton (more properly at the center of mass but that is very close to where the proton is). For a diatomic molecule one should 144

picture a dumbbell with the center of mass located midway between the nuclei for a homonuclear molecule, otherwise closer to the more massive nucleus. The dumbbell can now rotate about the center of mass and only two angular parameters are needed to describe that. They are the same two as used for the angular description of the electron in hydrogen and hence the rotational motion is a solved problem. Imagining a diatomic molecule as a rigid dumbbell when rotating, and as two masses connected to a spring when vibrating is simplistic on the one hand and remarkably good on the other. Your task is to keep in mind when and how these approximations break down and what can be done about that. One of the easier ways to picture the situation is to realize that the simple harmonic oscillator demands a potential curve that looks like a parabola while the real potential curve looks more like the Morse potential, to be treated subsequently. The real potential curve supports only a finite number of vibrational levels (often fewer than 0) while a parabola supports an infinite number. As one gets closer to the top of the finite potential, vibrational spacings are nothing like that predicted by the simple expressions derived for the harmonic oscillator even when corrected for anharmonic effects. Perhaps one surprising feature that results from applying quantum mechanics to diatomic molecules is that the spin of the nuclei plays a role. This is a consequence of spin statistics, is a purely quantum effect, and one that I have always found nonintuitive. This chapter tries to present the core material that one requires when doing spectroscopy of diatomic molecules. Far more has been left out than included but the student should be in a position to read the classic monograph, Spectra of Diatomic Molecules by Gerhard Herzberg 1. 6.1 Basic Considerations In the previous chapter the solution of the electronic motion in a diatomic molecule was discussed, which represents the first part in the two-part Born- Oppenheimer approximation. The second part considers the motion of the nuclei, which is represented by (5.3) on page 10: [ h i + E(X i ) v(x i ) = Ev(X i ) (5.3) M i i 1 Spectra of Diatomic Molecules, nd ed. Gerhard Herzberg, Van Nostrand Reinhold, N.Y. 1950. 145

Here X i are the coordinates of the nuclei of mass M i. E(X i ) is the potential curve obtained in the previous chapter which will look like one of the two in figure 5.4. The first step towards a solution of (5.3) is to separate out the center of mass solution for a system with two nuclei. On the off chance you have not seen this done before, it will be done here. Let RCM = M 1 r 1 + M r M 1 + M (6.1) R rel = r 1 r (6.) r 1 and r are the vectors from the origin to nucleus 1 and respectively. In the present case r 1 and r are oppositely directed, but that is not necessary for this derivation. Remember, all coordinates are now nuclear ones. The electron coordinates have all been integrated over to obtain the potential energy curve for nuclear motion. Equation (5.3) may be written [ h 1 h + E( R M 1 M rel ) v( r 1, r ) = Ev( r 1, r ) (6.3) It is easiest to work in Cartesian coordinates and the idea is to reexpress the above second order differential equation in center of mass and relative coordinates defined above. The following is a bit tedious but not difficult. 1 = x + 1 y1 = x + y + z 1 + z r 1 = ˆxx 1 + ŷy 1 + ẑz 1 r = ˆxx + ŷy + ẑz X CM = M 1x 1 + M x M 1 + M X rel = x 1 x and similarly for Y CM, Z CM, Y rel and Z rel. R CM = ˆxX CM + ŷy CM + ẑz CM R rel = ˆxX rel + ŷy rel + ẑz rel 146

Then = X CM x 1 x 1 x 1 = x 1 = M 1 + X rel X CM x 1 X CM + M 1 + M M1 (M 1 + M ) XCM X rel X rel (chain rule) + Xrel + M 1 M 1 + M X rel X CM and similarly for y 1 and. Also z1 x = M (M 1 + M ) XCM + Xrel M M 1 + M X rel X CM and similarly for y and. z Looking now at the x-components and remembering that precisely the same relations hold for y- and z-components, it is possible to write 1 1 M 1 M 1 x ; 1 1 M M x Adding these gives 1 M 1 + M 1 M 1 + M XCM + M 1 + M M 1 M X rel Let M M 1 + M and µ M 1M M 1 +M, and, realizing that the as-yet-unsolved wavefunction can just as easily be considered a function of R rel and R CM, yields [ h M CM h µ rel + E( R rel ) v( R rel, R CM ) = Ev( R rel, R CM ) Assume separability, that is that v may take the form v rel v( R rel, R CM ) = v CM ( R CM )v rel ( R rel ) and E = E rel + E CM The equation itself separates into the form: 1 [ h µ rel + E(R rel) v rel E rel = 1 v CM 147 [ h M CMv CM + E CM

Each side is clearly a function of different variables, so each may be set equal to a constant. Including that constant with the energy leads to the two equations [ h µ rel + E(R rel) v rel = E rel v rel (6.4) h M CMv CM = E CM v CM (6.5) The second equation is just the Schrödinger equation of a free particle of mass M, and is of no particular interest. The first equation is the Schrödinger equation for a particle of mass µ in the potential given by E(R rel ) and represents our looked for result. The subscript rel can be dropped and (6.4) can be considered in some detail. Spherical coordinates may be used, r, θ and φ, where r R rel is the distance between the nuclei, and θ, φ are the usual spherical coordinates which determine the direction of the dumbbell molecule. The notation can be simplified somewhat by redefining E(R rel ) V (r) v rel (r, θ, φ) = Z(r)Y m l (θ, φ) This second equation is just a separation of coordinates which will look very similar to the hydrogen atom in that there is a single particle in a central potential. (The form of the potential is very different from hydrogen, so a very different radial solution can be expected. The angular solution, however, must be the same.) Then writing out the Laplacian in spherical coordinates (6.4) becomes [ h µ ( 1 r r [ r r L h r = EZ(r)Y m l ) + V (r) Z(r)Yl m (θ, φ) (θ, φ) Once again separating variables, calling the separation constant l (l + 1) as was done for hydrogen, yields ( [ h 1 d µ r r d ) + V (r) + h l (l + 1) dr dr µr Z(r) = EZ(r) (6.6) and L Y m l ((θ, φ)) = l (l + 1) h Y m l ((θ, φ)) (6.7) Equation (6.6) is the DE for the radial variable. The term h l(l+1) µr comes from the angular part of the equation, so it is natural to associate this term 148

with the rotational energy of the molecule. noting that This is more clearly seen by r = r 1 r = [r 1 + r r 1 r cos(180 ) 1 / = r 1 + r Since the origin is at the center of mass, and r 1 and r are oppositely directed, it follows that M 1 r 1 = M r = µr So M 1 r1 + M r = µ r + µ r = µr M 1 M So µr = M 1 r1 + M r = I, which is the moment of inertia of the dumbbell molecule. Then E rot = h l (l + 1) (6.8) I It is tempting to treat I as a constant which would certainly simplify things. In fact, this is done as a first approximation to a solution. Remember that r changes but little about some equilibrium value. Such small changes in r produce changes in E rot much smaller than changes in V (r). To a first approximation if E rot is dropped from the Hamiltonian, the energy due to vibrations could be found, and then the rotational energy could be added back in. This shall be done as a first attempt at a solution. Later it will be seen that inclusion of this term does make second-order but easily measurable changes to molecular energy levels. With those comments in mind return to (6.6) and write this for E rot = 0. (This is no approximation when l = 0.) What this produces is the differential equation for vibrational motion which will next be considered. The rotational energy will be added back in afterward which is equivalent to treating it as a constant. ( [ h 1 d µ r r d ) + V (r) Z(r) = EZ(r) (6.9) dr dr Consider solutions close to the equilibrium separation r 0, where V (r) achieves a minimum. Taking the zero of potential to be at infinite nuclear separation, then V (r 0 ) D e, which is the equilibrium dissociation energy. Next expand V (r) in a Taylor series about r 0 : V (r) V (r 0 ) + dv dr (r r 0 ) + d V (r r 0 ) r=r0 dr... r=r0! 149

But V (r) has a minimum at r = r 0 so dv dr r=r0 Then = 0. Call d V dr r=r0 k. V (r) D e + k (r r 0) = D e + k x (6.10) and x r r 0. What is about to be done is the transformation of the original differential equation, written for a function Z(r), into one written for a function H(x), with x defined above. The reason is simply to try to put the differential equation into a recognizable form with a known solution. In this case, the solution shall involve the Hermite polynomials. The differential operation in equation (6.9) may now be written 1 d r dr ( r dz dr ) = d Z dr + dz r dr = d Z dx + dz x + r 0 dx As was done for hydrogen, let H(x) (x + r 0 )Z(x + r 0 ). Then and dh dx = (x + r 0) dz dx + Z d H dx = (x + r 0) d Z dx + dz dx This then leads to 1 d H x + r 0 dx = d Z dx + dz x + r 0 dx d H dx + µ h ( E 1 / kx ) H = 0 (6.11) where E E + D e. Equation (6.11) is the Schrödinger equation of a one dimensional simple harmonic oscillator in the displacement x from equilibrium. The functions H n (x) are Hermite polynomials times an exponential and the solution for the eigenvalues is E n = h k µ (n + 1 / ) = hω(n + 1 / ) with ω k/µ. So the solution for the vibrational energy levels is given by E vib = D e + hω(n + 1 / ) (6.1) 150

Including the rotational energy levels the full solution may be written E tot = D e + hω(n + 1 / ) + h l (l + 1) I (6.13) A review of the levels of approximation which were used to obtain this result would be useful. The Born-Oppenheimer separation is fundamental, is really quite good, and gives V (r) as a solution to the electronic problem. Next a centrifugal force contribution to the potential was identified as the rotational energy. This is not all that bad an approximation. Finally the potential curve V (r) was approximated by a parabola which, while not bad for very small n, is expected to be rather poor as n becomes large. Doing better than making a parabolic approximation to the potential will be the next order of business. 6. The Anharmonic Oscillator and Non-rigid Rotator There exists an analytic potential function with a single adjustable parameter (besides the well depth) which can reasonably approximate most ground state potential curves. In addition it has the property that the Schrödinger equation can be exactly solved for this potential. It is called the Morse potential and is given by V (r) = D e [e a(r r 0) e a(r r 0) + A (6.14) Here a is an adjustable shape parameter and D e is the depth of the well at the equilibrium separation below the asymptotic value of A. Figure 6.1 shows a Morse potential for H with r 0 = 1.4011a 0, a = 0.9790a 1 0, D e = 0.1745a.u. and A = 1.0a.u.. It really is quite good for rough work but one should take notice that any Morse potential does poorly for small values of r. At r = 0 the potential is finite rather than infinite as Coulomb repulsion would demand. The Morse potential yields the following values for the vibrational energy in place of (6.1) of the previous section: E vib = D e + hω(n + 1 / ) h ω 4D e (n + 1 / ) (6.15) It is well known in quantum mechanics that such a quadratic arises from the treatment of the anharmonic oscillator. In such a treatment, V (r) is 151