DISCRETE SYMMETRIES IN NUCLEAR AND PARTICLE PHYSICS. Parity PHYS NUCLEAR AND PARTICLE PHYSICS

Similar documents
Discrete Transformations: Parity

Quantum Numbers. F. Di Lodovico 1 EPP, SPA6306. Queen Mary University of London. Quantum Numbers. F. Di Lodovico. Quantum Numbers.

Lecture 7. both processes have characteristic associated time Consequence strong interactions conserve more quantum numbers then weak interactions

Space-Time Symmetries

129 Lecture Notes More on Dirac Equation

Quantum Numbers. Elementary Particles Properties. F. Di Lodovico c 1 EPP, SPA6306. Queen Mary University of London. Quantum Numbers. F.

Gian Gopal Particle Attributes Quantum Numbers 1

Weak interactions and vector bosons

2.4 Parity transformation

2007 Section A of examination problems on Nuclei and Particles

Physics 4213/5213 Lecture 1

cgrahamphysics.com Particles that mediate force Book pg Exchange particles

The Standard Model (part I)

.! " # e " + $ e. have the same spin as electron neutrinos, and is ½ integer (fermions).

Properties of Elementary Particles

Option 212: UNIT 2 Elementary Particles

Electron-positron pairs can be produced from a photon of energy > twice the rest energy of the electron.

Invariance Principles and Conservation Laws

Weak interactions. Chapter 7

SECTION A: NUCLEAR AND PARTICLE PHENOMENOLOGY

Isospin. K.K. Gan L5: Isospin and Parity 1

Particle Physics Lectures Outline

( ) ( ) ( ) Invariance Principles & Conservation Laws. = P δx. Summary of key point of argument

Elementary Particle Physics Glossary. Course organiser: Dr Marcella Bona February 9, 2016

Outline. Charged Leptonic Weak Interaction. Charged Weak Interactions of Quarks. Neutral Weak Interaction. Electroweak Unification

FYS 3510 Subatomic physics with applications in astrophysics. Nuclear and Particle Physics: An Introduction

PhysicsAndMathsTutor.com 1

Weak interactions, parity, helicity

ψ(t) = U(t) ψ(0). (6.1.1)

Books: - Martin, B.R. & Shaw, G Particle Physics (Wiley) (recommended) - Perkins, D.H. Introduction to High Energy Physics (CUP) (advanced)

Matter: it s what you have learned that makes up the world Protons, Neutrons and Electrons

Nuclear and Particle Physics 3: Particle Physics. Lecture 1: Introduction to Particle Physics February 5th 2007

1. Introduction. Particle and Nuclear Physics. Dr. Tina Potter. Dr. Tina Potter 1. Introduction 1

PhysicsAndMathsTutor.com

Problem Set # 1 SOLUTIONS

PHYS 3446 Lecture #17

9.2.E - Particle Physics. Year 12 Physics 9.8 Quanta to Quarks

Contents. Preface to the First Edition Preface to the Second Edition

INTRODUCTION TO THE STANDARD MODEL OF PARTICLE PHYSICS

Introduction to Elementary Particles

Particles. Constituents of the atom

NATIONAL OPEN UNIVERSITY OF NIGERIA

Particle Physics. Dr Victoria Martin, Spring Semester 2012 Lecture 14: CP and CP Violation

Lecture 3: Quarks and Symmetry in Quarks

Physics 125 Course Notes Identical Particles Solutions to Problems F. Porter

The Quark Parton Model

III. Particle Physics and Isospin

Problem Set # 4 SOLUTIONS

Particle Physics. All science is either physics or stamp collecting and this from a 1908 Nobel laureate in Chemistry

FXA Candidates should be able to :

Lecture 8. CPT theorem and CP violation

Introduction to particle physics Lecture 12: Weak interactions

Weak Interactions Cabbibo Angle and Selection Rules

1 Introduction. 1.1 The Standard Model of particle physics The fundamental particles

Name : Physics 490. Practice Final (closed book; calculator, one notecard OK)

Modern Physics. Luis A. Anchordoqui. Department of Physics and Astronomy Lehman College, City University of New York. Lesson XI November 19, 2015

Electron-Positron Annihilation

Parity violation. no left-handed ν$ are produced

PHY-105: Introduction to Particle and Nuclear Physics

Neutron Decay Disagree

Lecture 03. The Standard Model of Particle Physics. Part II The Higgs Boson Properties of the SM

Lecture 01. Introduction to Elementary Particle Physics

Elementary (?) Particles

Particle Physics Outline the concepts of particle production and annihilation and apply the conservation laws to these processes.

The Standard Model. 1 st 2 nd 3 rd Describes 3 of the 4 known fundamental forces. Separates particle into categories

The Scale-Symmetric Theory as the Origin of the Standard Model

Particles and Forces

An Introduction to the Standard Model of Particle Physics

COPYRIGHTED MATERIAL. Basic Concepts. 1.1 History

Intrinsic Parity of Neutral Pion. Taushif Ahmed 13th September, 2012

Kern- und Teilchenphysik I Lecture 13:Quarks and QCD

The Mirror Symmetry of Matter and Antimatter

Nuclear Physics and Astrophysics

Essential Physics II. Lecture 14:

DEVIL PHYSICS THE BADDEST CLASS ON CAMPUS IB PHYSICS

T 1! p k. = T r! k., s k. ', x' ] = i!, s y. ', s y

1. What does this poster contain?

3. Introductory Nuclear Physics 1; The Liquid Drop Model

Standard Model & Beyond

Modern Physics: Standard Model of Particle Physics (Invited Lecture)

Particle Physics Lecture 1 : Introduction Fall 2015 Seon-Hee Seo

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

Phys 328 Nuclear Physics and Particles Introduction Murat

Visit for more fantastic resources. AQA. A Level. A Level Physics. Particles (Answers) Name: Total Marks: /30

Lecture 02. The Standard Model of Particle Physics. Part I The Particles

3.3 Lagrangian and symmetries for a spin- 1 2 field

Particle Physics I Lecture Exam Question Sheet

Lecture PowerPoint. Chapter 32 Physics: Principles with Applications, 6 th edition Giancoli

Allowed beta decay May 18, 2017

For example, in one dimension if we had two particles in a one-dimensional infinite potential well described by the following two wave functions.

13. Basic Nuclear Properties

THE STANDARD MODEL OF MATTER

Outline. Charged Leptonic Weak Interaction. Charged Weak Interactions of Quarks. Neutral Weak Interaction. Electroweak Unification

Exotic Hadrons. O Villalobos Baillie MPAGS PP5 November 2015

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

Chapter 32 Lecture Notes

The Electro-Strong Interaction

The Physics of Particles and Forces David Wilson

The Building Blocks of Nature

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015

Transcription:

PHYS 30121 NUCLEAR AND PARTICLE PHYSICS DISCRETE SYMMETRIES IN NUCLEAR AND PARTICLE PHYSICS Discrete symmetries are ones that do not depend on any continuous parameter. The classic example is reflection in a mirror either we reflect an object in the mirror or we leave it unchanged. This is in contrast to continuous symmetries like rotations where we can turn the object through any angle. There are three discrete symmetries that play crucial roles in particle physics: parity (P ), charge conjugation (C), and time reversal (T ). This note collects together some of their main features. Parity The most familiar of these symmetries is parity, which is also important in nuclear physics and many other areas of quantum physics. It describes the effect on a system of reversing all coordinates: x x, y y and z z, or more compactly: r r. In terms of spherical polar coordinates, the transformation is: r r, θ π θ and φ π + θ. This can also be thought of as reflection in a mirror, z z, followed by a rotation through 180 in the xy plane. If the system is symmetric under rotations about the z axis, the rotation has no effect. Since this is often the case for systems of interest, you may hear the parity transformation referred to as reflection in a mirror. As well as reversing the coordinates of all particles, the parity transformation reverses their motions, and so momenta are transformed similarly: p p. In contrast, angular momenta are invariant under this transformation: J J. In the case of orbital angular momentum, this follows from the definition, L = r p, which shows that reversing both r and p leaves L unchanged. Alternatively, imagine a gyroscope spinning with its axis normal to a mirror; the sense of rotation of the mirror image is the same as that of the real thing. Quantities like angular momentum whose direction lies along an axis of rotation and so are unchanged under parity are known as axial vectors or pseudovectors. In quantum mechanics, we can define a parity operator P which, acting on a single-particle wave function, has the effect P ψ(r) = ψ( r). (1) An eigenstate of this operator satisfies P ψ(r) = P ψ(r), (2) where P is its eigenvalue. Since two reflections take the system back to where it started, 1

the eigenvalue must satisfy P 2 = 1. Its possible values are thus P = ±1. We say that functions with P = +1 have even parity while those with P = 1 have odd parity (since ψ( r) = P ψ(r)). A wave function with definite orbital angular momentum also has definite parity. For example, Y 0,0 (θ, φ) = 1 4π (3) is obviously even under a parity transformation whereas 3 Y 1,+1 (θ, φ) = 4π sin θ eiφ (4) is odd. More generally, any wave function with even l has P = +1 whereas one with odd l has P = 1. (This is most easily checked for the cases with m = 0 which are even or odd polynomials of cos θ.) This means that for a single particle with orbital angular momentum l we can write P = ( 1) l. (5) Most quantum numbers are additive: for a system of several particles we simply add the quantum numbers for the individual particles to get the total for the systems (remembering that, in the case of angular momenta, we need to use the rules that follow from addition of vectors). In contrast, parity is multiplicative. For a state with two particles described by wave functions ψ 1 (r 1 ) and ψ 2 (r 2 ) with definite parities P 1 and P 2, the parity transformation gives P ψ 1 (r 1 )ψ 2 (r 2 ) = ψ 1 ( r 1 )ψ 2 ( r 2 ) = P 1 P 2 ψ 1 (r 1 )ψ 2 (r 2 ), (6) and so the overall parity is the product P = P 1 P 2. This multiplication of parities for individual particles is useful in situations where the particles move independently in a spherical potential, such electrons in an atom or nucleons in a (nondeformed) nucleus. For example, the ground state of 41 Ca has even numbers of protons and neutrons filling s, p and d orbitals, plus one neutron in an f orbital. Since there is an even number of particles in each filled orbital, their parities multiply to give +1. The overall parity of the state is then given by that of the last neutron, which has l = 3 and hence P = 1. We cannot use this approach for a system of two particles interacting with each other because they do not move independently. Instead we need to consider their relative and centre-of-mass motions, described by the coordinate vectors r = r 2 r 1 and R = (m 1 r 1 + m 2 r 2 )/(m 1 + m 2 ). These motions are independent and so, if both have definite orbital angular momenta, L r and L cm, we could use ( 1) L to find the parity for each and then multiply these to get the overall parity. 2

In practice, we are interested in the angular momentum and parity of just the internal structure of a system. (The fact that the centre-of-mass motion is independent means that its momentum, orbital angular momentum and parity are separately conserved.) The wave function describing the relative motion of the particles is a function of a single vector, r. Its parity is given in terms of the relative orbital angular momentum L r in the usual way: P = ( 1) Lr. Systems of three particles are more complicated in general, since there are two relative coordinate vectors. However their ground-state wave functions typically have zero orbital angular momenta and hence even parity. Finally, in calculating the parity of a state, we need to multiply by the intrinsic parities of the particles involved. Intrinsic parity is a property of a particle, like intrinsic spin, which does not rely on it having any moving constituents. It is an extra contribution to the phase of the particle s wave function under the parity transformation. For example, a particle with intrinsic parity P a in a state of orbital angular momentum l transforms to P ψ l (r) = P a ψ l ( r) = P a ( 1) l ψ l (r), (7) and so has overall parity P = P a ( 1) l. The photon provides an example of this. As ultrarelativistic particles, photons require quantum field theory the relativistic version of quantum mechanics to describe them. In this, the polarization state of a photon is described by its electric field. This is a regular vector, transforming under parity as E(r, t) E( r, t). The minus sign here indicates the photon has odd intrinsic parity, P γ = 1. A more subtle result of relativistic quantum mechanics is that fermions and their antiparticles must have opposite intrinsic parities. As an example, the intrinsic parities of the electron and positron satisfy P e P e + = 1. (8) By convention, we choose the constituents of normal matter (electrons, protons, quarks etc.) to have positive intrinsic parity, and so their antiparticles have negative parity: P e = +1, P e + = 1. (9) This choice means that when dealing with states of atoms or nuclei, we need to consider only the parities of the orbital wave functions of their electrons or nucleons. Note that this applies only to fermions; bosons and their antiparticles have the same parity. In contrast, when dealing with hadrons, we need to be more careful about intrinsic parities. For example, a pion consists of a quark and an antiquark with relative L = 0. The opposite intrinsic parities of its constituents mean that its overall parity is odd (P = (+1)( 1)( 1) L = 1). 3

Elementary particles, states of hadrons and states of nuclei are often labelled by J P where J is the total internal angular momentum of the state and P = ± is the corresponding parity. (Since P is always unity, we just need to specify the sign of P.) For example, the ground states of all even-even nuclei have J P = 0 +. Note that J is often referred to as the spin of a state, even when the object is composite and J denotes the (vector) sum of the orbital angular momenta and intrinsic spins of all its constituents. Returning to the example of the pion, the spins of the quark and antiquark add up to zero and so the total internal angular momentum of the state is J = 0. This composite particle therefore has spin and parity J P = 0. A spin-0 object like this, which is invariant under rotations but has odd parity is known as a pseudoscalar to distinguish it from an ordinary scalar, which is even. Charge conjugation Charge conjugation describes the effect on a system of replacing every particle by its antiparticle. This can be thought of as reflecting it in a mirror that reverses the signs of all charges, not just electric charge but also the generalised charges of baryon number, lepton number and flavour quantum numbers. Completely neutral particles like photons and π 0 mesons, which carry none of these charges, are their own antiparticles. They remain the same under charge conjugation, except for possible changes of phase of their wave functions. We can introduce a charge conjugation operator Ĉ. For a completely neutral state ψ, we define its charge-conjugation eigenvalue C by Ĉψ = Cψ. (10) This eigenvalue is often referred to as the C parity of the state. As with ordinary parity, acting with twice Ĉ returns the system to its original state, and so the possible values are C = ±1. Like parity, C parity is a multiplicative quantum number. Similarly to the case of parity, completely neutral particles have intrinsic C parities. For example, the photon is its own antiparticle and it therefore has a definite C parity, corresponding to the phase change of its wave function, or rather electric field, under charge conjugation. Since charge conjugation reverses all charges of all particles in a system, including their electric charges, it reverses the sign of the electric field they produce, E(r, t) E(r, t). This implies that the photon has odd intrinsic C parity, C γ = 1. A state consisting of a particle and its antiparticle also carries no net charge of any kind and so has a definite C parity. Since charge conjugation has the effect of swapping the particle and antiparticle, the C parity can be found from the symmetry of their overall wave function. Assuming that the particles have definite orbital angular momentum L and total spin S, their wave function factorises into orbital and spin states, and we can consider the symmetry of each state separately. Swapping the positions of the particle 4

and antiparticle reverses their relative position vector and so has the same effect as parity reversal. This contributes a factor ( 1) L to the C parity of the state. The spin state of two bosons is even if their total spin is even, and odd if the total is odd. Hence swapping the spins of the particle and antiparticle contributes a factor ( 1) S. In contrast, the spin state of two fermions is odd if their total spin is even, and even if the total is odd. This contributes a factor of ( 1) S+1. Finally, there is also a factor of 1 which arises in quantum field theory when a fermion and its antifermion are swapped. The bottom line is that, for both bosons and fermions, the C parity of a particleantiparticle pair is given by C = ( 1) L+S, (11) where L is their relative angular momentum and S their total spin. For example, as described above, the neutral pion consists of a superposition of uū and d d with zero orbital angular momentum and total spin. Its C parity is therefore C π 0 = +1. For neutral particles and hadrons, we can extend the labelling mentioned above to J P C where C = ± indicates the particle s C parity. For example, you will see the photon listed in data tables as having J P C = 1 and the π 0 as J P C = 0 +. Time reversal The final discrete symmetry is time reversal, T, which reverses the flow of time: t t. This leaves the positions of all particles unchanged but reverses their motions and spins: p p, J J. Unlike P and C, this transformation is not represented by a standard quantum operator since it swaps initial and final states. This means that we cannot define a corresponding eigenvalue and then make use of its conservation. One place where time reversal does play an important role is in the CP T theorem. This is a rigorously provable result of quantum field theory (one of rather few such results; another is the connection between spin and statistics). It states that the interactions between particles should be invariant under the combination of all three discrete symmetries, C, P and T. If we take some physical process, replace all particles involved by their antiparticles, reflect it in a mirror, and then play the movie of it backwards, we get another physically allowed process. One of the consequences of this theorem is that particles and antiparticles must have identical masses. Any violation of CP T would indicate a serious flaw in our understanding of either quantum mechanics or relativity. Conservation laws and violation of these symmetries The symmetries of parity and charge conjugation are respected by the strong and EM interactions. As a result they are very good symmetries in much of atomic, nuclear and hadronic physics. They allow us to label states by their eigenvalues P and C. Conservation of these in strong and electromagnetic reactions and decays means that the parity and C parity of the final state of any process must match those of the initial state. We can use 5

them to determine whether a process is allowed or not, provided we know or can deduce the relative orbital angular momenta of all the particles involved. These conservation laws, along with conservation of angular momentum, can often be used to determine the quantum numbers of new particles. In nuclear physics, similar detective work using conservation of angular momentum and P can be used to deduce the quantum numbers of excited states of nuclei. However there is no fundamental reason why laws of physics should be invariant under these transformations individually. (After all, your image in a mirror is not a real object and indeed it differs in many little ways from the original.) As just noted, quantum mechanics and relativity require only that the combined CP T symmetry is respected. The weak interaction provides our only current example of breaking of these discrete symmetries. Its left-handedness means that it violates both parity and charge conjugation as much as possible. For example, β decays emit right-handed antineutrinos, but never their mirror images, left-handed antineutrinos, or their charge conjugates, right-handed neutrinos. On the other hand, β + decays emit only left-handed neutrinos. This implies that the combination CP is a very good symmetry of the weak interaction. Nonetheless, even this combination is broken in weak interactions, by a CP -violating phase in the couplings to the Higgs field. However this is a very small effect, which shows up only in rather special circumstances, in particular decays of neutral K and B mesons. The CP T theorem implies that time-reversal symmetry must also be violated in a way that exactly compensates for the breaking of CP. Direct signals of this would include electric dipole moments for particles like neutrons or electrons. These are expected to be extremely small and, despite intensive efforts, they have not so far been observed. Further reading B. R. Martin, Nuclear and particle physics: an introduction (2nd edn.), Sec. 1.3 B. R. Martin and G. Shaw, Particle physics (3rd edn.), Secs. 5.3 5.6 D. H. Perkins, Introduction to high-energy physics (4th edn.), Secs. 3.2 3.7, 3.9 3.11 E. M. Henley and A. Garcia, Subatomic physics (3rd edn.), Ch. 9 Mike Birse (September 2017) 6