INTERMOLECULAR INTERACTIONS

Similar documents
Lecture 33: Intermolecular Interactions

5.61 Physical Chemistry Exam III 11/29/12. MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Chemistry Chemistry Physical Chemistry.

( r) = 1 Z. e Zr/a 0. + n +1δ n', n+1 ). dt ' e i ( ε n ε i )t'/! a n ( t) = n ψ t = 1 i! e iε n t/! n' x n = Physics 624, Quantum II -- Exam 1

PAPER No. : 8 (PHYSICAL SPECTROSCOPY) MODULE No. : 5 (TRANSITION PROBABILITIES AND TRANSITION DIPOLE MOMENT. OVERVIEW OF SELECTION RULES)

5.61 Physical Chemistry Final Exam 12/16/09. MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Chemistry Chemistry Physical Chemistry

Electric properties of molecules

Consider Hamiltonian. Since n form complete set, the states ψ (1) and ψ (2) may be expanded. Now substitute (1-5) into Hψ = Eψ,

LS coupling. 2 2 n + H s o + H h f + H B. (1) 2m

Lecture: P1_Wk1_L1 IntraMolecular Interactions. Ron Reifenberger Birck Nanotechnology Center Purdue University 2012

$ +! j. % i PERTURBATION THEORY AND SUBGROUPS (REVISED 11/15/08)

5.61 Physical Chemistry Lecture #35+ Page 1

5.80 Small-Molecule Spectroscopy and Dynamics

CE 530 Molecular Simulation

Lecture 3: Helium Readings: Foot Chapter 3

1s atomic orbital 2s atomic orbital 2s atomic orbital (with node) 2px orbital 2py orbital 2pz orbital

and Discrete Charge Distributions

WKS Name Intermolecular Forces Period Date

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

9 STRUCTURE & BONDING

Pauli Deformation APPENDIX Y

The Physics Behind the Cosmic Microwave Background

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

An Introduction to Quantum Chemistry and Potential Energy Surfaces. Benjamin G. Levine

Intermolecular forces: Background

only two orbitals, and therefore only two combinations to worry about, but things get

0 belonging to the unperturbed Hamiltonian H 0 are known

α β β α β β 0 β α 0 0 0

Molecular Compounds Compounds that are bonded covalently (like in water, or carbon dioxide) are called molecular compounds

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

Covalent Bonding. Chapter 8. Diatomic elements. Covalent bonding. Molecular compounds. 1 and 7

Chapter 6. Chemical Bonding

Chem120a : Exam 3 (Chem Bio) Solutions

Chapter 3. Acids and Bases

221B Lecture Notes Many-Body Problems II Molecular Physics

Quadratic Equations Part I

5.61 Physical Chemistry Lecture #36 Page

Brief review of Quantum Mechanics (QM)

18.01 Single Variable Calculus Fall 2006

Alkali metals show splitting of spectral lines in absence of magnetic field. s lines not split p, d lines split

MODERN ELECTRONIC STRUCTURE THEORY: Basis Sets

conventions and notation

Multi-Electron Atoms II

5.80 Small-Molecule Spectroscopy and Dynamics

3/30/2015. Third energy level. Second energy level. Energy absorbed. First energy level. Atomic nucleus. Energy released (as light)

Lecture #21: Hydrogen Atom II

Name Date Class MOLECULAR COMPOUNDS. Distinguish molecular compounds from ionic compounds Identify the information a molecular formula provides

Stochastic Histories. Chapter Introduction

X. Assembling the Pieces

Chapter 2 Approximation Methods Can be Used When Exact Solutions to the Schrödinger Equation Can Not be Found.

(Refer Slide Time: 5:21 min)

Chapter 8 : Covalent Bonding. Section 8.1: Molecular Compounds

Chapter 6. Preview. Objectives. Molecular Compounds

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

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,

32 Electrostatics. Electrostatics involves electric charges, the forces between them, and their behavior in materials.

CHEM-UA 127: Advanced General Chemistry I

Liquids & Solids. Mr. Hollister Holliday Legacy High School Regular & Honors Chemistry

5.74 Introductory Quantum Mechanics II

Identical Particles in Quantum Mechanics

Chem 344 Final Exam Tuesday, Dec. 11, 2007, 3-?? PM

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

CHEMISTRY XL-14A CHEMICAL BONDS

CHAPTER 8 The Quantum Theory of Motion

7.2 Dipolar Interactions and Single Ion Anisotropy in Metal Ions

Cosmology and particle physics

Chemistry 2. Lecture 1 Quantum Mechanics in Chemistry

HOW IS THE AP CHEMISTRY EXAM STRUCTURED?

Organic Chemistry. Second Edition. Chapter 3 Acids and Bases. David Klein. Klein, Organic Chemistry 2e

Welcome to PHYS2002!

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

Electric Fields, Dipoles and Torque Challenge Problem Solutions

General Physical Chemistry II

26 Group Theory Basics

8.04: Quantum Mechanics Professor Allan Adams Massachusetts Institute of Technology Tuesday March 19. Problem Set 6. Due Wednesday April 3 at 10.

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 20, March 8, 2006

Decoupling Theory and Practice

Structure of diatomic molecules

Lab #20: Observing the Behavior of Electrons

CHAPTER 11 MOLECULAR ORBITAL THEORY

8.04 Spring 2013 March 12, 2013 Problem 1. (10 points) The Probability Current

States of Matter. Intermolecular Forces. The States of Matter. Intermolecular Forces. Intermolecular Forces

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

Lecture Notes 1: Physical Equilibria Vapor Pressure

5.80 Small-Molecule Spectroscopy and Dynamics

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

Chemical Formulas and Equations

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

16.1 Introduction to NMR Spectroscopy. Spectroscopy. Spectroscopy. Spectroscopy. Spectroscopy. Spectroscopy 4/11/2013

The role of atomic radius in ion channel selectivity :

Example questions for Molecular modelling (Level 4) Dr. Adrian Mulholland

8.04 Spring 2013 April 09, 2013 Problem 1. (15 points) Mathematical Preliminaries: Facts about Unitary Operators. Uφ u = uφ u

7. Arrange the molecular orbitals in order of increasing energy and add the electrons.

11 Perturbation Theory

Outline for Today. Monday, Nov. 26. Chapter 11: Intermolecular Forces and Liquids. Intermolecular Foces. Comparing States of Matter

1.1 The Fundamental Chemistry of life

Grade 8 Chapter 7: Rational and Irrational Numbers

Structure and Bonding of Organic Molecules

Chemical Bonding. Section 1 Introduction to Chemical Bonding. Section 2 Covalent Bonding and Molecular Compounds

18 VALENCE BOND THEORY

Transcription:

5.61 Physical Chemistry Lecture #3 1 INTERMOLECULAR INTERACTIONS Consider the interaction between two stable molecules (e.g. water and ethanol) or equivalently between two noble atoms (e.g. helium and neon). Call the two species A and, and suppose they are oriented as Now, according to our simple MO pictures, there will not be any chemical bonds between A and ; the MOs will be fairly localized either on A or on and we should not have significant hybridization of the orbitals. Thus, according to the MO picture, these molecules will not interact. However, we know that they do interact. If they did not, we would never be able to form liquids or solids, as it is the attraction between molecules that holds such things together. Of course, our intuition also tells us that the interactions between molecules are much weaker than the forces that hold molecules together, and so we immediately guess that the intermolecular interactions can be treated as a perturbation. Toward this end, we write the Hamiltonian for the A- system as: H ˆ = H ˆ A + H ˆ + V ˆ A where H ˆ A ( H ˆ ) describes the isolated interactions within molecule A () and V ˆ A contains all the interaction terms between A and. Now, V ˆ A is a fairly complicated object: it contains all the interactions between electrons and/or nuclei on A and electrons and/or nuclei on. Rather than deal with the full V ˆ A (which would be very hard) we will note that as long as A and are far apart (i.e. as long as R is large) we can approximate V ˆ A using a classical dipole-dipole interaction ˆ R μˆ μ ˆ A 3( μ ˆ ) A R ( R μ ˆ ) V A 4πε 0 R 5 Eq. 1 Here, μˆ A ( μˆ )is an operator that measures the dipole moment on molecule A (). We won t particularly care about the form of this operator in this lecture, but we will use it quite a bit later on. For reference, μˆ A e rˆ R A μ Nuclear A The first part measures the dipole moment of the electron charge distribution, while the second subtracts the dipole of the nuclear charges.

5.61 Physical Chemistry Lecture #3 Now, Eq. 1 is still too complicated for us. The dipole is a vector quantity it has a magnitude and a direction. As a result, the dipole-dipole interaction depends on the directions of the dipoles involved: While the orientation dependence of the μ-μ interaction can be important in many situations, we will not be interested in this level of detail for now. Hence, we will assume that the molecular dipoles are oriented in their most favorable head-to-tail orientation (the first situation above) in which case ˆ ˆ μ ˆ V A Aμ 3 πε0r where the non-boldface operator ˆ μ A ( ˆ μ ) returns the magnitude of the dipole moment on A(). In practice, this will overestimate the true interactions, because sometimes the dipoles will be in less favorable orientations relative to one another, but it will suffice for qualitative purposes. Next, we split the Hamiltonian into a zeroth order part and a perturbation in the logical way: H ˆ = ˆ + ˆ ˆ 0 H A H V = V ˆ A ecause the zeroth order Hamiltonian is separable, we immediately recognize that the zeroth order eigenstates will factorize as products, and that the energies will add: ˆ ˆ + ˆ E A + E Ψ A Ψ H Ψ A Ψ = ( H H ) Ψ A Ψ = 0 α β A α β α β α β

5.61 Physical Chemistry Lecture #3 3 α and β are quantum numbers for A and respectively. The reason we have two quantum numbers here instead of just one is exactly the same as why we had two quantum numbers (n x and n y ) for the D harmonic oscillator: when we add degrees of freedom, we always introduce new quantum numbers. In the present case, when α=7 and β=3, we are looking at an excited state where molecule A is in its seventh excited state and is in its third excited state. It isn t enough to consider just excited states of A or individually one also has to allow for the possibility that both molecules might get excited at the same time. That being said, in chemistry we are usually interested in the ground electronic state of the system. For the A- system, the ground state implies that both A and are in their ground state, in which case Ψ ( 0 ) =Ψ A Ψ ( 0 ) 0 0 0 E 0 = E A 0 + E 0. Now, as discussed above, this zeroth order energy does not contain any interactions between A and. It is easy to see this in the equation because there are no terms that depend on A and simultaneously. Thus A doesn t know that exists, and vice versa. As a result, at zeroth order the molecules will never stick to one another. To introduce interactions, we apply perturbation theory. At first order, we have: (1) A* * ˆ A 1 E 0 = Ψ 0 Ψ 0 V A Ψ 0 Ψ 0 dτ A dτ = Ψ A* Ψ * ˆ μ ˆ A μ 1 = Ψ A* ˆ μ Ψ A dτ πε R 3 0 0 A Ψ 0 Ψ 0 dτ A dτ 0 Ψ * ˆ μ Ψ dτ πε 0 0 R3 A 0 A 0 0 where on the second line we have grouped terms so that the integral clearly factorizes into a product of an integral over A and an integral over. These two integrals have physical meaning: Ψ A* ˆ 0 μ A Ψ A 0 dτ ˆ A μ A (The ground state dipole of A) Ψ * ˆ μ ˆ 0 Ψ 0 dτ μ (The ground state dipole of ) Thus, the first order energy takes on an intuitive form: ˆ μ ˆ μ E (1) A 0 = πε 0 R 3. This is just what we would expect. As in classical physics, the average dipole on A interacts with the average dipole on. This interaction has a characteristic R -3 dependence, which will dominate at long range as all higher terms decay with R to some higher power (e.g R -6 below). At this point, we can make a physical connection with the abstract parameter λ used in

5.61 Physical Chemistry Lecture #3 4 perturbation theory: here R -3 plays the role of λ. Thus, if we increase the separation between the fragments, the effective value of λ gets smaller and we expect perturbation theory to work perfectly for large enough R. Likewise, if we decrease R the strength of the perturbation gets stronger and we eventually expect perturbation theory to break down once the molecules get too close together. Nonetheless, we have arrived at what we sought: an expression for the interaction between two molecules. This dipole-dipole interaction will typically dominate the intermolecular attraction at long range, except in one circumstance: if either A or has a dipole moment of zero, the first order term vanishes identically and we are left once again with no intermolecular attraction. Thus, we would still have no interaction between methane and water, because methane has no dipole. To rectify this, we must go to second order in the expansion: Ψ (0)* ˆ Ψ A* m V A Ψ(0) β V A Ψ A dτ A d 0 Ψ 0 τ E 0 = E (0) (0) = m E E A ( 0,0) 0 m ( α,β ) ( 0,0) 0 + E 0 E A α E β Here we clarify that in this expression the single index m always specifies all of the quantum numbers for the system. Thus, in this case m=(α,β), contains two quantum numbers. The exemption m ( 0,0 ) that restricts the sum eliminates only one term: the case were both α=0 and β=0. It does not remove terms where only α=0 or only β=0. We can rationalize this by recalling that only the state whose eigenvalue we are computing (in this case the A ground state) is eliminated from the sum. All other terms are to be included. Since the state with α=0 and β=1 is not the ground state (because molecule is excited) it is allowed in the sum. 0 dτ α Ψ* ˆ As a result, we can break down the sum into three pieces. If α=0 we get Ψ A* Ψ * V ˆ Ψ A Ψ dτ dτ Ψ A* Ψ * ˆ μ ˆ μ Ψ A Ψ dτ dτ 0 β A 0 0 A 1 0 β A 0 0 A E,ind = + E E E = E A A 4π ε R 6 E A + E E A β 0 0 0 0 β 0 0 0 0 E β 0 β Ψ A* ˆ μ Ψ A 0 d * ˆ Ψ 0 dτ = 1 0 A τ A Ψ β μ 4π ε R 6 0 E 0 E β 0 β where on the second line we have again rearranged things so that it is clear we have a product of an A integral times a integral. The A integral does not depend on β, so we can move it outside the sum:

5.61 Physical Chemistry Lecture #3 5 E ˆ ( ) μ Ψ * ˆ μ β Ψ 0 dτ ˆ ˆ 0 A β μ A μ,ind = ε 6 = 0R Eq. 0 E 0 E 4π ε R 6 E β β 0 β 0 0 E β 4π where we have defined the transition dipole by ˆ μ = Ψ * ˆ β μ Ψ 0 dτ. The interpretaion of this integral is a bit challenging at first. There are a number of ways to understand it: 1) If we put in an electric field, ˆ μ reflects the importance of the state β in the new ground state of the system ) In matrix language, ˆ μ is the off-diagonal coupling between state 0 and state β produced by the dipole operator 3) In spectroscopy, we will find that ˆ μ relates to the intensity of the optical transition between 0 and β. However you slice it, the important point is that ˆ μ need not be zero even if ˆ μ vanishes. Thus the second order term we have labeled E,ind will typically give a non-zero contribution as long as A has a dipole moment. The contribution will be attractive, because the numerator in Eq. is positive and the denominator is negative. Further, E,ind has a characteristic R -6 dependence on the A- separation. If both molecules have a dipole, this contribution will be totally swamped by the dipole-dipole contribution (which only decays as R -3 ), and so E,ind only really becomes important in cases where has no dipole. Physically, we can interpret E,ind as an induction effect (hence the subscript ind ). If molecule A has a permanent dipole, this dipole exerts a field on and induces a dipole there because the electrons on can polarize in the presence of the field. If we instead have the scenario where has a permanent dipole but A does not, we still get a contribution from the β=0 terms in the sum where the roles of A and are interchanged ˆ μ ΨA* ˆ μ Ψ A α A 0 dτ A ˆ μ ˆ 0 α E μ A A,ind = π ε A E A = 4 R 6 E 4π ε 6 A A 0 α 0 0 α 0R α 0 E 0 Eα This term describes induction where the dipole moment on induces a dipole on A. However, in the case where neither A nor has a dipole, we appear to still be out of luck. How do non-polar molecules stick to one another? The final piece to the puzzle is provided by the terms in the second order energy where neither α nor β is zero. In this case

5.61 Physical Chemistry Lecture #3 6 * A 1 Ψ A * ˆ μ ˆ α Ψβ Aμ Ψ 0 Ψ 0 dτ A dτ E disp = 4π ε 0 R6 A A α, β = 1 E 0 + E 0 E α E β 1 Ψ A* ˆ α μ A Ψ A 0 dτ A Ψ* ˆ β μ Ψ = 4π ε R 6 A A 0 α, β = 1 E 0 + E 0 E α E β Using our definitions of transition dipoles, this reduces to 0 α A ( ) 1 μ μ E disp = 4π ε 0 R6 E A, 1 0 + E A E α β = 0 E α β 0 dτ. Eq. 3 In general the summand will be a small number, because the denominator involves an energy difference between electronic states and we have seen that electronic energy differences are big numbers. However, it need not be zero even if neither molecule has a permanent dipole. The only thing involved are the transition dipoles, and we have seen that in general these need not be zero. We have thus discovered a universal force between molecules. It is attractive, because the numerator in Eq. 3 is positive, while the denominator is negative. Further, it has no classical counterpart it only exists because of quantum interactions. This attractive force was discovered by Frtiz London and is called the London dispersion force in his honor. In physics, the analogous interaction between two uncharged plates is called the Casimir force. In either situation, one typically rationalizes the interaction using the following quantum logic: While neither molecule has a dipole on average, the two molecules can still cooperate so that half the time molecule A will have a + dipole while will have a dipole (creating an attractive interaction) while the other half the time A is and is + (also attractive. On average, the dipoles are still zero, but there is still a net interaction. The dispersion interaction has a characteristic R -6 dependence on distance, and the coefficient is typically much smaller than that for dipole-induced dipole forces. Thus, it is most important in systems with no dipole moment. An important example of this is graphite, where the graphene sheets are held together solely by the dispersion interaction between the layers. Putting everything together, perturbation theory gives us a physical hierarchy of intermolecular forces and also justifies why all kinds of molecules tend to stick together.

MIT OpenCourseWare http://ocw.mit.edu 5.61 Physical Chemistry Fall 013 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.