A Review of Perturbation Theory

Similar documents
Approximation Methods in QM

11 Perturbation Theory

Phys460.nb Back to our example. on the same quantum state. i.e., if we have initial condition (5.241) ψ(t = 0) = χ n (t = 0)

So far, we considered quantum static, as all our potentials did not depend on time. Therefore, our time dependence was trivial and always the same:

Time-Independent Perturbation Theory

Physics 221A Fall 2018 Notes 22 Bound-State Perturbation Theory

Quantum Mechanics Solutions

NANOSCALE SCIENCE & TECHNOLOGY

Time dependent perturbation theory 1 D. E. Soper 2 University of Oregon 11 May 2012

Physics 137A Quantum Mechanics Fall 2012 Midterm II - Solutions

Degenerate Perturbation Theory

Perturbation Theory 1

Degenerate Perturbation Theory. 1 General framework and strategy

6. Molecular structure and spectroscopy I

Quantum Light-Matter Interactions

Quantum Physics II (8.05) Fall 2002 Outline

10 Time-Independent Perturbation Theory

The Particle in a Box

7 Three-level systems

Quantum Physics III (8.06) Spring 2008 Final Exam Solutions

B2.III Revision notes: quantum physics

Ψ t = ih Ψ t t. Time Dependent Wave Equation Quantum Mechanical Description. Hamiltonian Static/Time-dependent. Time-dependent Energy operator

Brief review of Quantum Mechanics (QM)

Section 5 Time Dependent Processes

-state problems and an application to the free particle

The quantum state as a vector

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

arxiv: v1 [quant-ph] 25 Apr 2008

Advanced Quantum Mechanics

Advanced Quantum Mechanics

Quantum Mechanics Solutions. λ i λ j v j v j v i v i.

Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension

P3317 HW from Lecture 7+8 and Recitation 4

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

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

Lecture 7. More dimensions

The 3 dimensional Schrödinger Equation

Light - Atom Interaction

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m.

Quantum Physics II (8.05) Fall 2002 Assignment 11

Perturbation Theory. Andreas Wacker Mathematical Physics Lund University

Berry s phase in Hall Effects and Topological Insulators

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

(Again, this quantity is the correlation function of the two spins.) With z chosen along ˆn 1, this quantity is easily computed (exercise):

14 Time-dependent perturbation theory

Chapter 6. Q. Suppose we put a delta-function bump in the center of the infinite square well: H = αδ(x a/2) (1)

E = φ 1 A The dynamics of a particle with mass m and charge q is determined by the Hamiltonian

Problem 1: A 3-D Spherical Well(10 Points)

1 Mathematical preliminaries

Phys 622 Problems Chapter 5

International Journal of Scientific and Engineering Research, Volume 7, Issue 9,September-2016 ISSN

A New Class of Adiabatic Cyclic States and Geometric Phases for Non-Hermitian Hamiltonians

Quantum Physics II (8.05) Fall 2002 Assignment 3

MP463 QUANTUM MECHANICS

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

26 Group Theory Basics

Physics 221B Spring 1997 Notes 31 Adiabatic Invariance, the Geometric Phase, and the Born-Oppenheimer Approximation

Semi-Classical Theory of Radiative Transitions

DEPARTMENT OF PHYSICS BROWN UNIVERSITY Written Qualifying Examination for the Ph.D. Degree January 23, 2009 READ THESE INSTRUCTIONS CAREFULLY

Section 6: Measurements, Uncertainty and Spherical Symmetry Solutions

I. Perturbation Theory and the Problem of Degeneracy[?,?,?]

The Einstein A and B Coefficients

6.1 Nondegenerate Perturbation Theory

QUANTUM MECHANICS I PHYS 516. Solutions to Problem Set # 5

2 The Density Operator

Group representation theory and quantum physics

Harmonic Oscillator I

1 Time-Dependent Two-State Systems: Rabi Oscillations

1 Planck-Einstein Relation E = hν

Radiating Dipoles in Quantum Mechanics

MATH325 - QUANTUM MECHANICS - SOLUTION SHEET 11

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

Lecture #1. Review. Postulates of quantum mechanics (1-3) Postulate 1

Helsinki Winterschool in Theoretical Chemistry 2013

Addition of Angular Momenta

Path integrals and the classical approximation 1 D. E. Soper 2 University of Oregon 14 November 2011

Physics 221A Fall 2017 Notes 27 The Variational Method

3. Quantum Mechanics in 3D

Time Independent Perturbation Theory Contd.

Notes on atomic and molecular energy levels and fundaments of spectroscopy

1 The postulates of quantum mechanics

A. Time dependence from separation of timescales

Adiabatic particle pumping and anomalous velocity

Adiabatic quantum computation a tutorial for computer scientists

5.61 Physical Chemistry Lecture #36 Page

Notes on x-ray scattering - M. Le Tacon, B. Keimer (06/2015)

UNIVERSITY OF MARYLAND Department of Physics College Park, Maryland. PHYSICS Ph.D. QUALIFYING EXAMINATION PART II

Further Quantum Mechanics Problem Set

ECE 487 Lecture 6 : Time-Dependent Quantum Mechanics I Class Outline:

Non-stationary States and Electric Dipole Transitions

Physics 342 Lecture 26. Angular Momentum. Lecture 26. Physics 342 Quantum Mechanics I

Physics 115C Homework 2

Two and Three-Dimensional Systems

Quantum Mechanics for Mathematicians: The Heisenberg group and the Schrödinger Representation

Quantum mechanics in one hour

2 Quantization of the Electromagnetic Field

Adiabatic Approximation

CONTENTS. vii. CHAPTER 2 Operators 15

i~ ti = H 0 ti. (24.1) i = 0i of energy E 0 at time t 0, then the state at afuturetimedi ers from the initial state by a phase factor

Transcription:

A Review of Perturbation Theory April 17, 2002 Most quantum mechanics problems are not solvable in closed form with analytical techniques. To extend our repetoire beyond just particle-in-a-box, a number of approximation techniques have been developed. A large class of these fall under the heading of perturbation theory, in which we consider our system to obey Hamiltonian H that may be written as H = H 0 + λh 1 + λ 2 H 2 +..., (1) where H 0 is an exactly solvable Hamiltonian, λ is a small parameter, and the other terms may therefore be taken as small corrections. These notes are a quick review of how to deal with systems that obey such Hamiltonians. We ll be using Dirac notation and the Schrödinger formalism, in which the states are timedependent. We begin with time-independent perturbation theory, and will then move on to consider time-dependent problems. For now, we ll only deal with single-particle problems, too. 1 Time-independent We re going to use the time-independent Schrödinger equation, and look for the energy eigenvalues and eigenstates of H. First, suppose that our unperturbed problem, H 0 ψ 0 = E 0 ψ 0 (2) may be solved exactly, giving an energy eigenvalue spectrum Ej 0 with corresponding eigenstates ψj 0. Remember, because H0 is a Hermitian operator, its eigenvalues are real, and its eigenvectors ψj 0 form a complete set. The strategy we re going to take is straightforward. We re going to assume that our perturbative corrections to the Hamiltonian, λh 1 + λ 2 H 2 +..., lead to corresponding perturbative corrections to the eigenvalues and eigenstates. That is, E j = Ej 0 + λej 1 + λ 2 Ej 2 +..., ψ j = ψj 0 + λ ψj 1 + λ 2 ψj 2 +... (3) For now, we ll plug into the time-independent Schrödinger equation, and use some linear algebra to solve for terms of interest. First, we need to check normalization. Taking the inner product of the proposed eigenstate ψ j with itself and setting equal to 1, and grouping terms by orders in λ, wesee ψj 0 ψj 0 = 1 ψj 1 ψj 0 + ψj 0 ψj 1 = 0, etc. (4) So, we can take ψj 1 ψ0 j = 0 without any problems. 1

1.1 Nondegenerate, 1st and 2nd order Plugging into the Schrödinger equation up to second order, (H 0 + λh 1 + λ 2 H 2 )( ψ 0 + λ ψ 1 + λ 2 ψ 2 )=(E 0 j + λe1 j + λ2 E 2 j )( ψ0 + λ ψ 1 + λ 2 ψ 2 ). (5) For simpicity s sake, let s assume initially that our original, unperturbed eigenvalue spectrum is nondegenerate. That is, Ej 0 E0 k (6) for all j k. Now let s take the inner product of Eq. 5 with ψj 0. Again, let s look order by order in λ. To zeroth order, we get ψj 0 H0 ψj 0 = ψ0 j E0 ψj 0, (7) which we know is true because ψj 0 is the solution to the unperturbed problem with eigenvalue Ej 0. To first order, we find ψj 0 H 1 ψj 0 = Ej 1. (8) So, the first order correction to the energy eigenvalue of state j is the matrix element of the first order perturbation Hamiltonian between the unperturbed states. Now we want to find ψj 1, the first order correction to the state ψ0 j. Take the inner product of Eq. 5 with ψk 0 this time, with k j. The zeroth order term vanishes. The first order term gives: ψk 0 H1 ψj 0 =(E0 k E0 j ) ψ0 k ψ1 j. (9) Rearranging, ψk 0 ψ1 j = ψ0 k H1 ψj 0 (Ek 0 E0 j ) (10) Since the unperturbed eigenfunctions form a complete set, we can write the correction to the eigenfunction as ψ 1 j = k ψ 0 k ψ 0 k ψ 1 j = k j ψ 0 k ψ0 k H1 ψ 0 j (E 0 k E0 j ). (11) Nowthatwehaveanexpressionfor ψj 1, we can plug that into Eq. 5, take the inner product of both sides with ψj 0 as before, and look at the second order terms. E 2 j = ψ 0 j H1 ψ 1 j = ψj 0 H1 ψk 0 ψ0 k H1 ψj 0 (E 0 k j k E0 j ) = k j ψ 0 k H1 ψ 0 j 2 (E 0 k E0 j ). (12) We ve now found the first and second order corrections to the unperturbed energy spectrum, and the first order correction to the unperturbed states. 2

1.2 Degeneracies If the unpertrubed system has degenerate levels, things can get messy. The perturbation can lift the degeneracy, meaning that two unperturbed states with identical unperturbed energies can end up getting different perturbation corrections to their energies and the states themselves. A classical example: a compass in zero magnetic field has all orientations degenerate. However, an external magnetic field in the plane of the compass needle breaks the degeneracy, and picks out a unique ground state of the perturbed system. Suppose we start out with two unperturbed degenerate states, ξf 0 and ξ0 g. Rather than just using ξf 0 and ξ0 g as the basis for our perturbation calculation, we want to find linear combinations of the two to use instead. Ideally we want the perturbation to not mix the originally degenerate states. That is, we want to find a combination of coefficients c f,c g and use c f ξf 0 + c g ξg 0 (and the combination orthogonal to that) as the starting points for our perturbation calculation. We determine the values of c f,c g by requiring that the perturbation be diagonal in the new basis. 1.3 Summary We ve seen that perturbation theory may be used to solve problems that are close to exactly solvable unperturbed problems. The perturbation changes the energy eigenvalues from their unperturbed values. It also alters a particular unperturbed state by mixing it with the other unperturbed states by an amount related to the matrix element of the perturbation; see Eq. 8. That change to the states leads to a higher order correction to the energy, Eq. 12. There are systematic approaches to carry out perturbation theory to higher orders. The most well-known example is the diagrammatic technique developed in quantum field theory by Richard Feynman. 2 Time-dependent What happens if we have an explicitly time-dependent Hamiltonian? The classic example of this is behavior of a system under the influence of electromagnetic radiation. Here the timedependence of the electric and magnetic fields is harmonic, and the results can be profound (Stokes shifts in Raman spectroscopy; lasing; etc.). Here we briefly introduce a few ways of dealing with time-dependent problems. We discuss the sudden and adiabatic approximations, and time-dependent perturbation theory. We end with a derivation of Fermi s Golden Rule, an expression used commonly for determining rates of processes in solid state systems. 2.1 Sudden and adiabatic approximations Suppose we consider turning on some perturbing Hamiltonian, H, in addition to our unperturbed Hamiltonian, H 0. Suppose further that the system was initially in a state ψ(0). Consider the case where the perturbation is turned on instantaneously at t = 0andisleft on for all future times. This is a time when we can apply the sudden approximation. If the eigenfunctions of the new Hamitonian H 0 + H are labeled φ j with energies E j,thenwecan 3

write down the state for all future times: φ(t >0) = j φ j φ j ψ(0) e ie j t/ h. (13) We project the original state onto the new eigenstates, and time-evolve those new eigenstates according to the new Hamiltonian. This approximation is valid provided the turn-on time of H is short compared to the relevant time scales of h/e j. Similarly, we consider the adiabatic limit, in which H is turned on infinitesmally slowly (or at least slowly compared to all relevant energy scales, given by h/ E jk,where E jk is the energy spacing between any two adjacent levels). If the adiabatic approximation is a good one, then the turn-on of the perturbation may be sufficiently gentle that each old eigenstate φ 0 j smoothly evolves into a new eigenstate φ j. If the perturbation would ve caused old levels to cross, instead one generally finds an avoided crossing. The detailed analysis of the probability of ending up in the other level after an avoided crossing as a function of perturbation rate was done be Landau and Zener back in the 1930s. See Landau and Lifschitz QM text, or Rubbermark et al., PRA23, 3107 (1981). 2.2 Transition probabilities What if the time dependence is more subtle than just turning on a perturbation and leaving it there? Then we need time-dependent perturbation theory. Begin with our unperturbed Hamiltonian, H 0, and again assume a nondegenerate eigenvalue spectrum Ej 0 (As above, the degenerate case is a reasonably straightforward development from the nondegenerate case). At time t = 0 we consider beginning to perturb our system with some H (t). At any time t we can expand the true, unknown Ψ(t) in the time-evolved versions of the unperturbed eigenstates, ψj 0.Thatis: Ψ(t) = j c j (t)e ie0 j t/ h ψj 0, (14) where the c j (t) are the time-dependent coefficients of the expansion. We can interpret this as saying that, if we turned the perturbation H off instantly at time t, the probability of finding the system in state ψj 0 is given by c j(t) 2. Note (without proof) that this expansion automatically preserves normalizalization as long as the perturbation H (t) isreal. Plug the state given in Eq. 14 into the time-dependent Schrödinger equation, using the total Hamiltonian, H 0 + H : i h j ċ j (t)e ie0 j t/ h ψ 0 j +i h j i h E0 j c j(t)e ie0 j t/ h ψ 0 j = j (H 0 +H (t))c j (t)e ie0 j t/ h ψj 0. (15) Drop identical terms from both sides after using the H 0 operator on the right hand side. Then take the inner product of both sides with ψ 0 k eie0 k t/ h,andget i hċ k (t) = j c j (t) ψ 0 k H (t) ψ 0 j ei(e0 k E0 j )t/ h. (16) This is exact, so far. Define ω kj E0 k E0 j, h H kj ψk 0 H (t) ψj 0. (17) 4

One can now try writing H (t) asapowerseriesinsomesmallparameterλ, and parallel our earlier analysis. However, a more common and useful approach is an interative one, described below. Let s assume that at t = 0 the system was in a particular unperturbed eigenstate, ψa 0. Therefore, c a (t = 0) = 1, and c j a (t = 0) = 0. That is, c a (t =0)=δ ja. PlugthisintoEq.16, and we get c a (t) = 1 t i h c j (t) = 1 i h 0 t 0 H aa(τ)dτ +1, j = a, H ja(τ)e iω jaτ dτ, j a. (18) We can define a transition probability, the probability of finding the system in (approximately) state j at time t after starting out in state a, as P ja (t) = c j (t) 2 = 1 t h 2 H ja (τ)eiωjaτ dτ 2. (19) 0 From this expression it is already possible to see where selection rules come from: if the matrix element of the perturbation between the initial and (approximate) final states, H ja, is zero, then to first order the transition is forbidden. Consider something like our sudden approximation case from before, where H is turned on at t = 0. In particular, let s assume [ e H iω rt + e iωrt ] (t) =V 0 cos(ω r t)=v 0. (20) 2 for t>0. Now we re representing the true new eigenstates approximately by the old unperturbed states. If we actually plug this into our probability expression, we find, first for ω r =0, P ja = V 0,ja 2 sin 2 (ω ja t/2) h 2 (ω ja /2) 2. (21) Remember, sin 2 (x)/x 2 is strongly peaked around x = 0. That implies that transitions between initial and final states are favored if energy is conserved, thatis,ifω ja 0. Let s define: F (t, ω ja ) 2sin2 ω ja t/2 ωja 2. (22) This quantity will be useful in a bit. If we do the ω r 0 case, we find: c j (t) = i V [ ] 0,ja 2 e i(ω ja +ω r)t 1 h 2 + ei(ω ja ω r)t 1 (23) i(ω ja + ω r ) i(ω ja ω r ) If we take the magnitude-squared of this to find P ja,weget P ja (t) =F (t, ω ja + ω r ) V 0,ja 2 + F (t, ω ja ω r ) V 0,ja 2 + crossterms. (24) The function F (t, ω) has some interesting properties. For fixed ω 0,it st-dependence is just sin 2. However, at fixed t, it speakatω =0ist 2 /2, and its half-width is 2π/t. So, as t increases, F (t, ω) becomes more and more strongly peaked around ω =0. Infact, + F (t, ω)dω = πt, (25) 5

and one can show that F (t, ω) πtδ(ω) ast. For us, this means that at long times (but not times so long that the initial state gets depleted!) transitions only occur to states that conserve energy. For our ω r 0 case above, we find transitions are probable between states j and a only if they differ in energy by an amount hω r. These two processes correspond to absorption and stimulated emission. 2.3 Fermi s Golden Rule Consider the ω r 0 case from above, and suppose that there are a number of states ν(e) that satisfy the conservation of energy condition between initial and final states. Also, suppose those final states also share the same matrix element with the initial state, V 0,ja. In that case, the rate at which transitions occur is given by: dp ja (t) = 2π h dt V 0,ja 2 ν(e). (26) This relation is called Fermi s Golden Rule, and is commonly used to determine the rates of many quantum mechanical processes. Remember that we re doing perturbation theory here, and asking the question, if a system starts out in (unperturbed) state a, what is the probability of finding it later in (approximately) unperturbed state p. The key to our approximations is that the time-dependent perturbation is small, that we wait long enough for our F functions to act like energy-conserving delta functions, but not so long that the initial state s population is changed much. These are conditions for the validity of Fermi s Golden Rule, which is fundamentally a first-order time dependent perturbation theory result. Note that: If the initial and final states in question are not connected by the perturbing potential (that is, the matrix element in question is zero), the rate of the process is also zero. This is an example of the origin of selection rules. If the number of available final states is zero, the rate of the process is also zero. The importance of this point can t be overestimated. As we ll see, it s the reason that semiconductors are semiconductors, that band insulators are band insulators, etc. 3 Conclusion This concludes our review of perturbation theory. The main results to take away from this are the first and second order corrections to the energies in time-independent problems, and Fermi s Golden Rule. For a more complete discussion of perturbation methods, most quantum mechanics textbooks will be useful references. Particularly valuable are Cohen-Tannoudji s and Sakurai s texts. 6