where P a is a projector to the eigenspace of A corresponding to a. 4. Time evolution of states is governed by the Schrödinger equation

Similar documents
Attempts at relativistic QM

Quantum Field Theory

Quantization of Scalar Field

Quantum Mechanics II

The Klein-Gordon equation

Quantum Field Theory

Quantum Field Theory

1 Free real scalar field

7 Quantized Free Dirac Fields

ΑΒΓΔΕΖΗΘΙΚΛΜΝΞΟΠΡΣΤΥΦΧΨΩ αβγδεζηθικλμνξοπρςστυφχψω +<=>± ħ

QFT. Unit 1: Relativistic Quantum Mechanics

General Relativity in a Nutshell

Second quantization: where quantization and particles come from?

Page 684. Lecture 40: Coordinate Transformations: Time Transformations Date Revised: 2009/02/02 Date Given: 2009/02/02

2 Canonical quantization

Quantization of scalar fields

2 Quantization of the scalar field

CHAPTER 1. SPECIAL RELATIVITY AND QUANTUM MECHANICS

1 Quantum fields in Minkowski spacetime

3 Quantization of the Dirac equation

Quantum Field Theory Notes. Ryan D. Reece

Quantum Field Theory II

Quantum Theory and Group Representations

Preliminaries: what you need to know

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

Lecture notes for QFT I (662)

Quantization of a Scalar Field

1 Mathematical preliminaries

Second Quantization: Quantum Fields

The Dirac Equation. Topic 3 Spinors, Fermion Fields, Dirac Fields Lecture 13

The Dirac Field. Physics , Quantum Field Theory. October Michael Dine Department of Physics University of California, Santa Cruz

Prancing Through Quantum Fields

Relativistic Waves and Quantum Fields

Lorentz Invariance and Second Quantization

Tutorial 5 Clifford Algebra and so(n)

MSci EXAMINATION. Date: XX th May, Time: 14:30-17:00

4.3 Lecture 18: Quantum Mechanics

MP463 QUANTUM MECHANICS

Part I. Many-Body Systems and Classical Field Theory

The Dirac equation. L= i ψ(x)[ 1 2 2

As usual, these notes are intended for use by class participants only, and are not for circulation. Week 6: Lectures 11, 12

PRINCIPLES OF PHYSICS. \Hp. Ni Jun TSINGHUA. Physics. From Quantum Field Theory. to Classical Mechanics. World Scientific. Vol.2. Report and Review in

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer

1 The Quantum Anharmonic Oscillator

PHY 396 K. Problem set #5. Due October 9, 2008.

Vector Fields. It is standard to define F µν = µ ϕ ν ν ϕ µ, so that the action may be written compactly as

The Hamiltonian operator and states

Classical field theory 2012 (NS-364B) Feynman propagator

221B Lecture Notes Quantum Field Theory II (Fermi Systems)

Introduction to string theory 2 - Quantization

Quantum Physics 2006/07

Canonical Quantization

2.1 Green Functions in Quantum Mechanics

221B Lecture Notes Quantum Field Theory II (Fermi Systems)

Extending the 4 4 Darbyshire Operator Using n-dimensional Dirac Matrices

Spin Statistics Theorem

The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case

SECOND QUANTIZATION. notes by Luca G. Molinari. (oct revised oct 2016)

Lecture 4 - Dirac Spinors

3P1a Quantum Field Theory: Example Sheet 1 Michaelmas 2016

Second quantization (the occupation-number representation)

What s up with those Feynman diagrams? an Introduction to Quantum Field Theories

Quantum Field Theory I

Lecture 10. The Dirac equation. WS2010/11: Introduction to Nuclear and Particle Physics

What is a particle? Keith Fratus. July 17, 2012 UCSB

Rotations in Quantum Mechanics

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

On the Heisenberg and Schrödinger Pictures

(a p (t)e i p x +a (t)e ip x p

SECOND QUANTIZATION. Lecture notes with course Quantum Theory

Section 4: The Quantum Scalar Field

Classical Field Theory

msqm 2011/8/14 21:35 page 189 #197

We can instead solve the problem algebraically by introducing up and down ladder operators b + and b

Physics 221A Fall 2010 Notes 1 The Mathematical Formalism of Quantum Mechanics

PHY2403F Lecture Notes

6.6 Charge Conjugation Charge Conjugation in Electromagnetic Processes Violation of C in the Weak Interaction

Problem Set No. 1: Quantization of Non-Relativistic Fermi Systems Due Date: September 14, Second Quantization of an Elastic Solid

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department 8.323: Relativistic Quantum Field Theory I PROBLEM SET 2

A Note On The Chern-Simons And Kodama Wavefunctions

Request for Comments Conceptual Explanation of Quantum Free-Field Theory

Lecture 4: Equations of motion and canonical quantization Read Sakurai Chapter 1.6 and 1.7

Quantum Physics II (8.05) Fall 2002 Outline

φ µ = g µν φ ν. p µ = 1g 1 4 φ µg 1 4,

Advanced Quantum Mechanics, Notes based on online course given by Leonard Susskind - Lecture 8

Canonical Quantization C6, HT 2016

Quantum Field Theories for Quantum Many-Particle Systems; or "Second Quantization"

Introduction to Electronic Structure Theory

Week 5-6: Lectures The Charged Scalar Field

SUPERSYMMETRIC HADRONIC MECHANICAL HARMONIC OSCILLATOR

Introduction to particle physics Lecture 3: Quantum Mechanics

Particle Physics. Michaelmas Term 2011 Prof. Mark Thomson. Handout 2 : The Dirac Equation. Non-Relativistic QM (Revision)

Quantum Mechanics for Mathematicians: Energy, Momentum, and the Quantum Free Particle

Lecture Notes 2: Review of Quantum Mechanics

Lecture 4 - Relativistic wave equations. Relativistic wave equations must satisfy several general postulates. These are;

Summary of free theory: one particle state: vacuum state is annihilated by all a s: then, one particle state has normalization:

Quantum Mechanics - I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 7 The Uncertainty Principle

Quantum Electrodynamics Test

Hamiltonian Field Theory

Chapter 2 The Group U(1) and its Representations

Transcription:

1 Content of the course Quantum Field Theory by M. Srednicki, Part 1. Combining QM and relativity We are going to keep all axioms of QM: 1. states are vectors (or rather rays) in Hilbert space.. observables are Hermitian operators and their values are the spectrum. 3. probability of measuring a particular value a of an observable A in a state Ψ is P a Ψ Ψ, where P a is a projector to the eigenspace of A corresponding to a. 4. Time evolution of states is governed by the Schrödinger equation i dψ(t) dt = HΨ(t), where H is the Hamiltonian (energy operator). 5. Symmetries are unitary or anti-unitary operators preserving the Hamiltonian. 6. etc. For a nonrelativistic particle, we let H = L (R 3 ) and let the momentum operator (generator of translations) be ˆP = i. Since E = P /m classically, it is natural to define H = ˆP /m = /m. From now on, I will let = 1, so H = /m. For a relativistic particle, so can try E = P c + m c 4, H = c + m c 4 This expression is problematic: treats time and space asymmetrically and appears nonlocal. Alternatively, we can quantize the squared dispersion relation E = P c 4 + m c 4 to get the Klein-Gordon equation t Ψ = ( c + m c 4) Ψ. 1

This equations is more reasonable, as it is more symmetric w.r. to exchange of time and space. To see relativistic invariance better, let x 0 = ct. From now I will let c = 1, so in such units x 0 = t. Also, x 0 = t, and x i = x i, i = 1,, 3. Greek indices will run over the set 0, 1,, 3. Minkowski metric: g µν = diag( 1, 1, 1, 1) = g µν. x µ = g µν x ν, where we use Einstein s convention (summation over repeated indices). Similarly, x µ = g µν x ν. Minkowski interval: x = x µ x µ = x µ x µ g µν = (x 0 ) + i (xi ). Lorenz transformations are x µ = Λ µ νx ν, where Λ is any real matrix such that x µ x µ = x µ x µ. Now we can check relativistic invariance of the KG equation, i.e. that φ(x) and φ( x) satisfy the same equation. Let µ = x, µ µ = g µν ν. Then µ = Λ µ ν ν. and therefore =. Hence the KG operator is Lorenz-invariant. Problems: 1. d 3 x Ψ is not conserved. Moreover, it has wrong transformation properties under the Lorenz transformation: Ψ is not a time component of a 4-vector, so we do not expect a continuity equation to hold (and it does not). One can write down something which is a component of a conserved 4-vector: j µ = i(ψ µ Ψ µ Ψ Ψ) satisfies µ j µ = 0, and so d 3 xj 0 is conserved. But j 0 is not positive-definite, so cannot be interpreted as probability density.. Negative-energy solutions. Dirac tried to solve these problems by looking for a first-order equation, but for a multicomponent wavefunction. This solved problem 1, but not problem.

Ultimately, the problem is that relativistic QM can be consistently developed only if we do not work in a theory with a fixed number of particles. Hence we need to understand how to describe systems where particle creation and annihilation is allowed. 3 Fock space methods (second quantization) 3.1 Bosons A single particle has H 1 = L (R 3 ) as its Hilbert space. Two particles have H = L (R 3 R 3 ) sym Sym H 1. And so on. The Hilbert space without any particles is one-dimensional (the vacuum state). Thus H = C H 1 H... = k=0sym k (H 1 ). This is called the bosonic Fock space F(H 1 ) associated to H 1. The Fock space is always infinite-dimensional, even if H 1 is not. Let us look at the extreme case, H 1 C. Then F(C) = C C C.... Thus a vector in F(C) is an infinite sequence of numbers or vector (a 0, a 1, a,...) such that k a k <. It is often convenient to think of such a sequence as Taylor coefficients of an analytic function f(z) = a 0 + a 1 z + a z +... Degree is then identified with the particle number. Polynomials form a dense set in this space of functions and correspond to states with a finite number of particles. Two natural operations on polynomials are z and. They satisfy [, z] = 1. One calls the annihilation operator a, and calls z the creation operator a. They are indeed conjugate to each other if we define the scalar product to be f(z) = 1 d z f(z) e z. π 3

Using this scalar product, one can compute z n = n!. Thus a normalized n-particle state is n = 1 z n = 1 (a ) n 0. n! n! Thus a n = n + 1 n + 1, a n = n n 1 This can serve as a definition of creation and annihilation operators. The particle number operator can be expressed as N = z z = a a. Eigenstates of N are homogeneous polynomials. Polynomials are Fock space states which involve only a finite number of particles. More generally, suppose H 1 C N. Let us choose a basis ψ i, i = 1,..., N in H 1 and introduce N variables z 1,..., z N. Then one can identify Sym p (H 1 ) with the space of polynomials in N variables of total degree p: a state with k 1 particles in the state ψ 1, k particles in the state ψ, etc. can be identified with the polynomial z k 1 1... z k N N. We define a i = i, a i = z i so that [a i, a j ] = δi j. The Fock space is then the space of all polynomials in variables z 1,..., z N. If we change the basis in H 1, creation-annihilation operators also change: if ψ i = B j i ψ j, where B is a unitary matrix, then a i = B i j a j The particle number operator is N = i z i i = i a i a i. Eigenstates of N are homogenous polynomials. If H 1 is infinite-dimensional, but has a countable basis, we can still think of its Fock space as a completion the space of polynomials in variables z 1, z,.... But usual bases on L (R 3 ) (momentum eigenstates p and coordinate eigenstates x are not like that. Still, one can define analogues of creation and annihilation operators: Ψ(x) = i a i ψ i (x), Ψ (x) = i a i ψ i (x). 4

They satisfy [Ψ(x), Ψ (y)] = δ 3 (x y). All operators in Fock space can be expressed in terms of Ψ(x) and Ψ (x). Examples: 0. The particle number operator N = d 3 xψ (x)ψ(x). 1. One-particle operators. A one-particle operator is an operator of the form k k 1... 1 O 1... 1 = O i, k=1 i=1 where O is an operator on H 1. It can be written as d 3 xd 3 yψ (x) x O y Ψ(y). For example, the kinetic energy operator is a one-particle operator with O = /m, so the corresponding operator in Fock space is d 3 xψ (x) ( /m ) Ψ(x). The particle number operator is a one-particle operator with O = 1.. Two-particle operators. These are operators of the form k=1 1 i<j k where O ij is an operator on H which acts only on the i-th and j-th particle. The corresponding operator in Fock space is 1 d 3 xd 3 yd 3 zd 3 tψ (x)ψ (y) x, y O z, t Ψ(t)Ψ(z). For example, the potential energy operator is of this form, with x, y O z, t = V (x y)δ 3 (x z)δ 3 (y t). The corresponding operator in Fock space is 1 d 3 xd 3 yψ (x)ψ (y)v (x y)ψ(y)ψ(x). How do we formulate dynamics in the Fock space? Since the emphasis is on creation-annihilation operators, it is often convenient to work in the 5 O ij k=0 i=1

Heisenberg picture and write EOMs for Ψ and Ψ, instead of the Schrödinger equation. For free particles, we get Ψ t = i[h, Ψ] = i m Ψ. This looks like Schrödinger equation, but for a field operator. Hence the name second quantization. Let us find a solution. Go to momentum space: Ψ(x) = d 3 p(π) 3 b(p)e ipx Then and Thus [b(p), b (q)] = (π) 3 δ 3 (p q). H = Ψ(t, x) = d 3 3 p p(π) m b (p)b(p). b(p, t) = e iept b(p, 0). d 3 p (π) 3 b(p, 0)e Ept+ip x. This completely determines the evolution of all observables. For an interacting system, get the following equation: i Ψ t = 1 m Ψ + d 3 yψ (y)ψ(y)v (x y)ψ(x). This is nonlinear and cannot be regarded as second-quantized Schrodinger equation. Its classical analogue is a PDE for an ordinary function Ψ(t, x), which is NOT interpreted as a quantum-mechanical wavefunction. Remark: the quantum harmonic oscillator corresponds to the Fock space for H = 1. A collection of N harmonic oscillators is equivalent to the bosonic Fock space for H 1 = C N.Thus the quantization of a system of harmonic oscillators can be interpreted in terms of free bosonic particles. The energies of 1-particle states are ω i. 6

3. Fermions Now consider fermionic particles which obey the Pauli principle. Fermionic wavefunctions are antisymmetric with respect to the exchange of any two particles. Let us again begin with the case H 1 = C. Then the Fock space is F(H 1 ) = C C. This is two-dimensional, and there are many ways to think about it. E.g., we can identify it with the states of a spin-1/ particle. But we will choose a more esoteric viewpoint. Consider a variable θ which has a multiplication rule θ = 0. Then analytic functions of θ are linear functions f(θ) = a + bθ. The space of such functions can be identified with F(H 1 ): th vacuum state is 1, while the 1-particle state is θ. Creation-annihilation operators are defined as before: c = θ, c = θ. Note that c = (c ) = 0. It is also easy to check that cc + c c = 1. Note the crucial plus sign. The particle number operator is N = c c. We can define the scalar product so that c is indeed the adjoint of c. Details of this are left as an exercise. Now consider N-dimensional H 1. Introduce N variables θ i which satisfy θ i θ j + θ j θ i = 0. Consider an analytic function of θ i. Again, the series terminates in degree N. The total dimension of the space of functions is N. The k-ht term in the expansion is i 1...i k f i1...ik θ i1... θ ik. Here the coefficient functions are totally anti-symmetric, as required by the Fermi statistics. The creation operators are c i = θ i, the annihilation operators are c i = i. They satisfy c i c j + c j c i = δ ij. Note that the fermionic Fock space has a symmetry which replaces the vacuum with the filled state θ 1... θ N and exchanges c i and c i. There is nothing analogous in the bosonic case. 7

The rest proceeds as before. We can choose a countable basis in L (R 3 ) and define Ψ(x) = ψ i (x)c i, Ψ (x) = ψi (x)c i. i i They satisfy Ψ(x)Ψ (y) + Ψ (y)ψ(x) = δ 3 (x y). These are called canonical anti-commutation relations. In the noninteracting case, the EOM is linear and solved exactly as in the bosonic case. 4 Classical field theory There is something special about differential equations which come from dequantizing the Heisenberg equations of motion: they come from a variational principle. 4.1 Classical mechanics Recall classical mechanics. Action: S = T 0 dtl(q i (t), q i (t)). Euler-Lagrange variational principle: δs = 0 with q(0) and q(t ) fixed. Equations of motion: L q = d ( ) L. i dt q i Alternatively, we can introduce p i = L/ q i, the Hamiltonian H = p q L, and write the action as S = dt(p q H(p(t), q(t))). The equation δs = 0 then gives q i = H p i, ṗ i = H q i. 8

These are Hamilton equations. Finally, if we introduce the Poisson bracket {F, G} = F p i G q i F q i G p i for any two functions F, G, the Hamilton equations of motion can be written as q i = {H, q i }, ṗ i = {H, p i }. We also have {p i, q j } = δ j i, {qi, q j } = {p i, p j } = 0. Under quantization, Poisson bracket becomes i times the commutator. 4. Nonrelativistic field theory Now we want to have a similar formalism where i is replaced with a continuous index x. Instead of q i (t) will have Ψ(t, x). Action: S = dtl(ψ, Ψ). EOM: δl δψ(t, x) = d ( ) δl dt δ Ψ(t,. x) Here the variational derivative is defined by δl = d 3 δl x δψ(t, x). δψ(t, x) In the free case, it is sufficient to take ( ) L = L 0 = d 3 x iψ 1 Ψ m iψ i Ψ. Note that L is an integral of a local expression, L = d 3 xl, so S = dtd 3 xl(ψ, Ψ). 9

This is very nice, but is not obligatory in a nonrelativistic situation: in an interacting case one finds L = L 0 1 d 3 xd 3 y Ψ(x) Ψ(y) V (x y). This is local in some very special cases. For example, when V (x) = δ 3 (x) ( contact interaction ). In the relativistic case only such interaction are allowed. Note that this fits better with the second version of the variational principle: iψ is the momentum conjugate to Ψ. So one has Poisson brackets {Ψ (x), Ψ(y)} = iδ 3 (x y). The Hamiltonian is then given by the same expression as before, but Ψ s are now ordinary functions, not Fock-space operators. Quantization now is easy: we get the standard commutation relations for Ψ and Ψ and realize them as operators in Fock space. How do we get fermionic Fock space in this way? There is no good way of doing so. Reason: classical limit makes sense only when a large number of particles are in the same state. For clarity, consider discrete case. In order for the commutator term to be negligible, need to consider a state where a has a large expectation value (and small variance). Hence N = a a will have a large expectation value. This is not possible in the fermionic case. Formally, we can still consider the same equations, but with Ψ and Ψ satisfying anticommutation relations. This means that they are not ordinary functions, but generators of a Grassmann algebra. We will use this trick later. 4.3 Relativistic field theory Main idea: interpret the KG equation not as an equation for a wavefunction, but an equation for a field operator. That is, let us make relativistic not the one-particle Schrodinger equation, but the Heisenberg equation of motion for the Fock space operator. To understand it, we need to specify commutation relations for Ψ in such a way, that the KG equation is the Heisenberg equation for some Hamiltonian. We can do this like this: first solve an analogous classical problem, and then quantize everything. 10

The classical KG equation comes from the action S = 1 dtd 3 x ( µ φ µ φ m φ ). This looks more like the first version of the variational principle. The momentum is p(x) = φ(x), and the Hamiltonian is H = d 3 xh = 1 The Poisson brackets are Hence quantization will give d 3 x ( p(x) + ( φ) + m φ ) {p(x), φ(y)} = δ 3 (x y}. [φ(x), p(y)] = iδ 3 (x y). This is just like [q.p] = i, but with continuous indices. Reason: the classical system describes the continuum limit of a system of particles connected with springs, and φ(x) is the continuum limit of the coordinate of a particle. Classical excitations are waves. What about quantization? Expect that we get a system of free bosonic particles with a relativistic dispersion law. Two reasons: (1) that is what we set out to describe; () classical system can be Fourier-analyzed into a collection of harmonic oscillators; each oscillator is equivalent to a Fock space (for a one-dimensional vector space), so the whole thing is equivalent to a Fock space (for an infinite-dimensional 1-particle space), so describes free bosonic particles. 11