30.01.2015 Views

QCD at low energies

QCD at low energies

QCD at low energies

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>QCD</strong> <strong>at</strong> <strong>low</strong> <strong>energies</strong><br />

K.Goeke, WS 1999/2000<br />

January 10, 2006<br />

Contents<br />

I Basic concepts of field theory 5<br />

1 Lagrangeans and action 5<br />

1.1 Hamilton principle . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.2 Free massive boson field with spin S=0 (Klein-Gordon): . . . . . 6<br />

1.3 Free massive fermion field with spin S=1/2 (Dirac): . . . . . . . 7<br />

1.4 Free massles bosonic field of Spin=1 (Maxwell) . . . . . . . . . . 8<br />

1.5 Fermion-Boson-coupling . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.6 Bilinear covariants: . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.7 Parity, time reversal, charge conjug<strong>at</strong>ion . . . . . . . . . . . . . . 10<br />

2 Canonical field quantiz<strong>at</strong>ion 11<br />

2.1 General remarks: . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.2 Quantized Klein-Gordon-Field: . . . . . . . . . . . . . . . . . . . 12<br />

2.3 Quantized Dirac-field: . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.4 Quantized Maxwell-field: . . . . . . . . . . . . . . . . . . . . . . . 16<br />

3 Symmetries and currents 18<br />

3.1 Elements of Lie-Group theory . . . . . . . . . . . . . . . . . . . . 18<br />

3.1.1 Lie-Groups SU(N) . . . . . . . . . . . . . . . . . . . . . . 18<br />

3.1.2 Represent<strong>at</strong>ions . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

3.1.3 Gell-Mann m<strong>at</strong>rices . . . . . . . . . . . . . . . . . . . . . 20<br />

3.2 Noether Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

3.2.1 General structure . . . . . . . . . . . . . . . . . . . . . . . 21<br />

3.2.2 Noether-Theorem and Lie-groups . . . . . . . . . . . . . . 23<br />

3.2.3 Altern<strong>at</strong>ive expression for several fields, Energy momentum<br />

tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

3.3 Current algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.4.1 Free massless fermions, chirality, helicity . . . . . . . . . . 28<br />

3.4.2 Free massive Dirac field, Fermion-number, U(1)-transform<strong>at</strong>ion 31<br />

1


3.4.3 Free massless Dirac field, Axial U A (1)-transform<strong>at</strong>ion . . 32<br />

3.5 Isospin symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.6 Linear (chiral) Sigma-model . . . . . . . . . . . . . . . . . . . . . 35<br />

4 Symmetry breaking 39<br />

4.1 Realiz<strong>at</strong>ion of symmetries . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.2 Principle of Spontaneous symmetry breaking . . . . . . . . . . . 40<br />

4.3 Spontaneous symmetry breaking in the linear sigma model . . . . 43<br />

4.3.1 The Goldstone Mode . . . . . . . . . . . . . . . . . . . . . 45<br />

4.3.2 Broken chiral symmetry and PCAC . . . . . . . . . . . . 47<br />

5 The gauge principle 50<br />

5.1 Abelian gauge theory: . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

5.2 Non-abelian gauge theory . . . . . . . . . . . . . . . . . . . . . . 55<br />

6 The <strong>QCD</strong>-Lagrangean 61<br />

6.1 The gauge transform<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . 61<br />

6.2 Strong coupling constant . . . . . . . . . . . . . . . . . . . . . . . 63<br />

6.3 Mandelstam variables . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

7 Symmetries and anomalies 70<br />

7.1 Mass terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

7.2 Vector symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

7.2.1 Global fermion symmetry . . . . . . . . . . . . . . . . . . 72<br />

7.2.2 Global (iso-)vectorial symmetry . . . . . . . . . . . . . . . 73<br />

7.3 Axial (non-)symmetries: . . . . . . . . . . . . . . . . . . . . . . . 75<br />

7.3.1 Flavour symmetry and U A (1) anomaly . . . . . . . . . . . 75<br />

7.3.2 Axial Vector symmetry and Axial Anomaly . . . . . . . . 76<br />

7.4 Other symmetries, Theta-vacuum . . . . . . . . . . . . . . . . . . 78<br />

7.4.1 Discrete symmetries . . . . . . . . . . . . . . . . . . . . . 78<br />

7.4.2 Scale invariance and trace-anomaly . . . . . . . . . . . . . 78<br />

7.4.3 Theta-Vacuum . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

8 Anomalies 85<br />

9 Electroweak Interactions in the Standard Model 85<br />

9.1 Electromagnetic Quark Currents . . . . . . . . . . . . . . . . . . 86<br />

9.2 Weak Quark Currents . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

9.3 Leptonic currents . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

2


10 Chiral symmetry breaking 90<br />

10.1 Chiral Symmetry and Current algebras . . . . . . . . . . . . . . . 90<br />

10.2 Spontaneous symmetry breaking . . . . . . . . . . . . . . . . . . 97<br />

10.3 Pion decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

10.4 PCAC: partial conserv<strong>at</strong>ion of axial current . . . . . . . . . . . . 103<br />

10.5 The chiral condens<strong>at</strong>e . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

10.6 LSZ-Reduction formulae . . . . . . . . . . . . . . . . . . . . . . . 107<br />

10.7 Pion-Nucleon coupling constant . . . . . . . . . . . . . . . . . . . 108<br />

10.8 Goldberger-Treiman rel<strong>at</strong>ion and pion pole . . . . . . . . . . . . 110<br />

10.9 Vacuum Sigma Term and Gell-Mann–Okubo . . . . . . . . . . . 112<br />

10.10Pion-Nucleon Sigma Term Σ πN . . . . . . . . . . . . . . . . . . . 115<br />

10.11Ward-Identities and Low energy pion-nucleon theorems . . . . . 123<br />

10.12Isospin-Viol<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

11 Group theory, Flavour structure and Quark model 128<br />

11.1 Elements of Lie-group theory . . . . . . . . . . . . . . . . . . . . 129<br />

11.2 SU(2)-algebra: . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

11.3 SU(2)-multiplets in n<strong>at</strong>ure . . . . . . . . . . . . . . . . . . . . . . 133<br />

11.4 SU(3)-algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

11.5 SU(3)-multiplets in n<strong>at</strong>ure . . . . . . . . . . . . . . . . . . . . . . 136<br />

11.6 Quarks and SU(3) . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

11.6.1 Quarks and fractional charges . . . . . . . . . . . . . . . . 139<br />

11.6.2 Construction of higher multiplets by triplets, Confinement<br />

(phenomenological) . . . . . . . . . . . . . . . . . . . . . . 141<br />

11.7 Fock st<strong>at</strong>es and non-rel<strong>at</strong>ivistic quark model . . . . . . . . . . . . 147<br />

11.7.1 Mass splittings . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

11.7.2 Quark model: Calcul<strong>at</strong>ions . . . . . . . . . . . . . . . . . 153<br />

12 Effective bosonic Lagrangean 155<br />

12.1 Review: Spontaneous breakdown of chiral symmetry . . . . . . . 156<br />

12.1.1 Current algebra and transform<strong>at</strong>ion properties . . . . . . 156<br />

12.1.2 Chiral Quark Condens<strong>at</strong>e . . . . . . . . . . . . . . . . . . 156<br />

12.2 Represent<strong>at</strong>ions of the chiral boson fields . . . . . . . . . . . . . 157<br />

12.2.1 Linear Sigma Model . . . . . . . . . . . . . . . . . . . . . 157<br />

12.2.2 Non-linear Sigma Model . . . . . . . . . . . . . . . . . . . 158<br />

12.2.3 Generaliz<strong>at</strong>ion to finite Goldstone Boson mass . . . . . . 159<br />

12.3 <strong>QCD</strong> and chiral fields . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

12.4 The minimal effective bosonic Lagrange-density . . . . . . . . . . 163<br />

12.4.1 Massless case . . . . . . . . . . . . . . . . . . . . . . . . . 163<br />

12.4.2 Massive case . . . . . . . . . . . . . . . . . . . . . . . . . 166<br />

12.4.3 Higher order terms . . . . . . . . . . . . . . . . . . . . . . 168<br />

12.5 Applic<strong>at</strong>ion: pion-pion sc<strong>at</strong>tering . . . . . . . . . . . . . . . . . . 174<br />

12.5.1 Calcul<strong>at</strong>ion: . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

12.5.2 Comparison with experiment . . . . . . . . . . . . . . . . 177<br />

12.6 Chiral perturb<strong>at</strong>ion theory and <strong>QCD</strong>: Functional integrals . . . . 180<br />

12.6.1 Internal symmetries and Ward-Identities . . . . . . . . . . 180<br />

3


12.6.2 External fields and Greens functions . . . . . . . . . . . . 183<br />

12.6.3 Local chiral symmetry . . . . . . . . . . . . . . . . . . . . 186<br />

12.7 Applic<strong>at</strong>ion: Pion decay . . . . . . . . . . . . . . . . . . . . . . . 189<br />

12.8 The chiral Lagrange-Density in ord(p 4 ) or ord(E 4 ) and renormaliz<strong>at</strong>ion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191<br />

12.8.1 The Lagrangean . . . . . . . . . . . . . . . . . . . . . . . 191<br />

12.8.2 Chiral perturb<strong>at</strong>ion theory, renormaliz<strong>at</strong>ion program . . . 192<br />

12.9 Applic<strong>at</strong>ion in order(p 4 ): Masses of Goldstone bosons . . . . . . 194<br />

12.9.1 Aim of this section . . . . . . . . . . . . . . . . . . . . . . 194<br />

12.9.2 Calcul<strong>at</strong>ion of the pole mass . . . . . . . . . . . . . . . . 195<br />

12.9.3 Calcul<strong>at</strong>ion of the self energy . . . . . . . . . . . . . . . . 198<br />

13 L<strong>at</strong>tice Gauge Theory 202<br />

13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203<br />

13.2 Quantum Mechanics: Transition amplitudes and p<strong>at</strong>h integrals . 203<br />

13.2.1 Free Motion (Propag<strong>at</strong>or) . . . . . . . . . . . . . . . . . . 205<br />

13.2.2 Propag<strong>at</strong>ors ( Harmonic oscill<strong>at</strong>or) . . . . . . . . . . . . . 209<br />

13.2.3 Extraction of inform<strong>at</strong>ion, Euklidean time . . . . . . . . . 212<br />

13.2.4 Correl<strong>at</strong>ors . . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

13.2.5 Partition function . . . . . . . . . . . . . . . . . . . . . . 217<br />

13.2.6 Metropolis Formalism . . . . . . . . . . . . . . . . . . . . 218<br />

13.3 Boson Quantum Field on the l<strong>at</strong>tice . . . . . . . . . . . . . . . . 222<br />

13.3.1 Quantum Field Theory with functional integrals . . . . . 222<br />

13.3.2 Euklidean Field Theory . . . . . . . . . . . . . . . . . . . 225<br />

13.3.3 Scalar boson field: Discretiz<strong>at</strong>ion of space-time . . . . . . 228<br />

13.4 Fermion Quantum fields on the l<strong>at</strong>tice . . . . . . . . . . . . . . . 233<br />

13.4.1 Grassmann-algebra . . . . . . . . . . . . . . . . . . . . . . 234<br />

13.4.2 Fermionic p<strong>at</strong>h integral . . . . . . . . . . . . . . . . . . . 235<br />

13.4.3 Wilson fermions, quenched l<strong>at</strong>tice. etc. . . . . . . . . . . . 237<br />

13.5 Gauge Fields on the L<strong>at</strong>tice . . . . . . . . . . . . . . . . . . . . . 238<br />

13.5.1 Abelian gauge fields: QED . . . . . . . . . . . . . . . . . 238<br />

13.5.2 Non-abelian gauge fields on the l<strong>at</strong>tice: <strong>QCD</strong> . . . . . . . 243<br />

13.5.3 Continuum limit . . . . . . . . . . . . . . . . . . . . . . . 244<br />

13.6 Practical applic<strong>at</strong>ions of l<strong>at</strong>tice <strong>QCD</strong> . . . . . . . . . . . . . . . . 247<br />

13.6.1 Metropolis algorithm in pure gauge theory . . . . . . . . . 247<br />

13.6.2 Pure gauge calcul<strong>at</strong>ions . . . . . . . . . . . . . . . . . . . 248<br />

13.6.3 Problems with Fermions . . . . . . . . . . . . . . . . . . . 254<br />

14 Linear chiral Sigma model (Gell-Mann–Levy) 255<br />

14.0.4 Fock st<strong>at</strong>es and vari<strong>at</strong>ional principle . . . . . . . . . . . . 255<br />

14.0.5 Projection techniques . . . . . . . . . . . . . . . . . . . . 257<br />

4


15 Restbestände 259<br />

Part I<br />

Basic concepts of field theory<br />

1 Lagrangeans and action<br />

1.1 Hamilton principle<br />

The dynamics of classical interacting fields<br />

φ i (x) :::::::: i = 1,.....,N<br />

with x = (t,x) is determined often by the Lagrangean density L:<br />

L = L(φ i (x),∂ µ φ i (x)) (1)<br />

This sort of Lagrangean is very common in connection with elementary particle<br />

theories, i.e. QED, <strong>QCD</strong>, Glashow-Salam-Weinberg theories and simple (linear)<br />

theories of strong interaction. Effective theories, of baryon-meson interaction<br />

etc. and field theories for many body systems are often more complic<strong>at</strong>ed and<br />

show higher deriv<strong>at</strong>ives. We will not consider them <strong>at</strong> the moment but l<strong>at</strong>er.<br />

The action of the fields is defined as:<br />

∫ ∫<br />

S = dtL = d 4 xL(φ i (x),∂ µ φ i (x))<br />

The Lagrange density (we refer to it simply as lagrangean) is an expression,<br />

where φ i (x) and ∂ µ φ i (x) are considered as independent variables. It is just a<br />

m<strong>at</strong>hem<strong>at</strong>ical expression, which has to fulfil certain general conditions. The<br />

actual motion of fields, i.e. evolution in time, is determined by the equ<strong>at</strong>ions<br />

of motion and by initial conditions. By Hamiltons principle the equ<strong>at</strong>ions of<br />

motion are fixed by the Euler-Lagrange equ<strong>at</strong>ions of the vari<strong>at</strong>ion principle<br />

without subsidiary conditions:<br />

δS = 0<br />

The explicit vari<strong>at</strong>ion of φ i (x) and ∂ µ φ i (x) is performed <strong>at</strong> a given but arbitrary<br />

space-time point x by varying the value of the field φ keeping the x fixed. The<br />

equ<strong>at</strong>ions of motion describe how a field , given <strong>at</strong> the time t 1 in the whole<br />

space x evolves to a l<strong>at</strong>er time t 2 in the whole space. The theory of vari<strong>at</strong>ion<br />

tells th<strong>at</strong> one has to perform the vari<strong>at</strong>ion (for simplicity in case of only one)<br />

field in the fol<strong>low</strong>ing way<br />

∫<br />

δS = d 4 x[L(φ + δφ,∂ µ φ + δ∂ µ φ) − L(φ,∂ µ φ)]<br />

5


∫<br />

=<br />

∫<br />

=<br />

∫<br />

=<br />

[ ∂L<br />

d 4 x<br />

∂φ δφ + ∂L ]<br />

∂(∂ µ φ) δ(∂ µφ)<br />

[ ∂L<br />

d 4 x<br />

∂φ δφ + ∂L ]<br />

∂(∂ µ φ) ∂ µ(δφ)<br />

[ ( ∂L ∂L<br />

d 4 x<br />

∂φ δφ − ∂ µ<br />

∂(∂ µ φ)<br />

)<br />

δφ + ∂ µ<br />

( ∂L<br />

∂(∂ µ φ) δφ )]<br />

The last term can be turned into a surface integral over the boundary of the<br />

four-dimensional space-time region of integr<strong>at</strong>eion. Since the initial and final<br />

field configur<strong>at</strong>ion φ(t 1 ,x) and φ(t 1 ,x) are assumed given the δφ vanishes <strong>at</strong><br />

the temporal beginning t 1 and temporal end t 2 in the whole integr<strong>at</strong>ion region.<br />

This lets the surface integral vanish: We obtain<br />

∫ [ ]<br />

∂L<br />

δS = d 4 x<br />

∂φ − ∂ ∂L<br />

µ δφ<br />

∂(∂ µ φ)<br />

Since the δφ is besides its boundary conditions <strong>at</strong> t 1 ,t 2 an arbitrary vari<strong>at</strong>ion,<br />

the contents of the bracket must be zero. This yields the equ<strong>at</strong>ion of motion:<br />

∂ µ<br />

∂L<br />

∂(∂ µ φ i ) − ∂L = 0 i = 1,...,N (2)<br />

∂φ i<br />

If one does a snapshot of the field <strong>at</strong> the time t in the whole space, then the<br />

equ<strong>at</strong>ions of motion (2) describe how the field propag<strong>at</strong>es in space and time.<br />

For the quantiz<strong>at</strong>ion of the fields one needs always the conjug<strong>at</strong>e momentum<br />

field<br />

π i (x) = ∂L<br />

∂ 0 φ i<br />

(x) (3)<br />

There are several basic sorts of classical Lagrangeans, which are used in<br />

elementary particle physics. These Lagrangeans describe fields, which we meet<br />

in n<strong>at</strong>ure.We will generally use = c = 1<br />

1.2 Free massive boson field with spin S=0 (Klein-Gordon):<br />

This is the simplest field. The famous higgs-field, or the scalar σ-field of the<br />

nucleon-nucleon interaction, or a neutral pion field is e.g. a free massive scalar<br />

boson field. After quantiz<strong>at</strong>ion (see l<strong>at</strong>er) the corresponding field particles have<br />

a finite mass and the spin zero and move freely without interaction. (In reality<br />

one is of course interested in the interaction of the particles with others. This<br />

will be discussed l<strong>at</strong>er): The Lagrangean of such a field, if it is not interacting<br />

with anything else, is given by<br />

L = 1 (<br />

∂λ φ∂ λ φ − µ 2 φ 2) (4)<br />

2<br />

The corresponding equ<strong>at</strong>ion of motion is the famous Klein-Gordon equ<strong>at</strong>ion:<br />

(<br />

∂ 2 + µ 2) φ = 0 (5)<br />

with ∂ 2 = ∂ λ ∂ λ = ∂ 2 0 − ∇ 2 .<br />

6


1.3 Free massive fermion field with spin S=1/2 (Dirac):<br />

The corresponding field quanta are Fermions with a mass m and Spin= 1 2 .The<br />

massive fermion field is described in the simplest form by the Lagrangean<br />

L = ψ [iγ µ ∂ µ − m]ψ (6)<br />

We have by definition<br />

¯ψ(x) = ψ † (x)γ 0<br />

The corresponding equ<strong>at</strong>ion of motion is the famous Dirac equ<strong>at</strong>ion:<br />

[iγ µ ∂ µ − m]ψ = 0 (7)<br />

Fermions have Spin=1/2, in a non-rel<strong>at</strong>ivistic theory they have to be described<br />

by a two-component spinor. Since we are dealing in particle physics with rel<strong>at</strong>ivistic<br />

systems the description with 2 components is not sufficient, but one<br />

needs <strong>at</strong> least a 4-component spinor (Dirac spinor)<br />

ψ(x) =<br />

⎛<br />

⎜<br />

⎝<br />

ψ 1 (x)<br />

ψ 2 (x)<br />

ψ 3 (x)<br />

ψ 4 (x)<br />

The γ µ are 4x4-m<strong>at</strong>rices. There are various equaivalent ways, one can express<br />

them. Usually the γ-m<strong>at</strong>rices are expressed as generaliz<strong>at</strong>ions of the Pauli-<br />

Spin-m<strong>at</strong>rices. Explicit forms are given in books on Dirac-theory (convention:<br />

µ,ν = 0,1,2,3 and i,j = 1,2,3:<br />

( ) ( )<br />

γ 0 I 0<br />

=<br />

γ i 0 σ<br />

i<br />

=<br />

0 −I<br />

−σ i (8)<br />

0<br />

One shoud note the commut<strong>at</strong>ion rules<br />

and the hermiticity properties:<br />

⎞<br />

⎟<br />

⎠<br />

{γ µ ,γ ν } = 2g µν (9)<br />

(<br />

γ<br />

0 ) †<br />

= γ<br />

0<br />

(<br />

γ<br />

i ) †<br />

= −γ<br />

i<br />

(10)<br />

The Pauli spin m<strong>at</strong>rices are given as<br />

( ) (<br />

σ 1 0 1<br />

=<br />

σ 2 0 −i<br />

=<br />

1 0<br />

i 0<br />

One often needs<br />

γ 5 = iγ 0 γ 1 γ 2 γ 3 =<br />

γ 0 γ µ γ 0 = (γ µ ) † (11)<br />

( 0 I<br />

I 0<br />

)<br />

)<br />

σ 3 =<br />

(<br />

1 0<br />

0 −1<br />

(<br />

γ<br />

5 ) †<br />

= γ<br />

5<br />

)<br />

(12)<br />

(13)<br />

7


with<br />

{<br />

γ 5 ,γ ν} = 0 (14)<br />

and furthermore<br />

σ µν = i 2 [γµ ,γ ν ] (15)<br />

Trγ µ = Trγ 5 = 0<br />

One often needs the fol<strong>low</strong>ing formulae: From the Dirac equ<strong>at</strong>ion we obtain<br />

γ µ ∂ µ ψ = −imψ (16)<br />

We get for the hermitian conjug<strong>at</strong>e with eq.(11) and multiplying from left with<br />

γ 0 :<br />

(iγ µ ∂ µ ψ) † = −i∂ µ ψ † (γ µ ) † = −i∂ µ ψ † γ 0 γ 0 (γ µ ) † = −i∂ µ ¯ψγ0 (γ µ ) †<br />

or<br />

−i∂ µ ¯ψγ0 (γ µ ) † γ 0 = mψ † γ 0<br />

∂ µ ¯ψγ µ = im ¯ψ (17)<br />

and<br />

and similarly we obtain from iγ µ ∂ µ ψ = mψ+ γ µ A µ<br />

Here we must put in final 12<br />

γ µ ∂ µ ψ = −imψ − iγ µ A µ ψ (18)<br />

∂ µ ¯ψγ µ = im ¯ψ + i ¯ψγ µ A µ (19)<br />

1.4 Free massles bosonic field of Spin=1 (Maxwell)<br />

The typical example of a free massless boson field of spin=1 is the electromagnetic<br />

field. The corresponding field quanta are the photons. The Lagrangean of<br />

the free field is<br />

L = − 1 4 F µνF µν (20)<br />

and it is expressed by the antisymmetric field tensor, which is directly rel<strong>at</strong>ed<br />

to the classical electric and magnetic fields:<br />

⎛<br />

⎞<br />

0 E x E y E z<br />

F µν (x) = ⎜ −E x 0 B z −B y<br />

⎟<br />

⎝ −E y −B z 0 B x<br />

⎠ (21)<br />

−E x B y −B x 0<br />

1 put here the free Dirac st<strong>at</strong>es consistent with the quantized Dirac field<br />

2 Here we must display plane wave st<strong>at</strong>es consistently with quantiz<strong>at</strong>ion of Dirac field<br />

8


The F µν is rel<strong>at</strong>ed to the 4-potential A µ by:<br />

F µν (x) = ∂ µ A ν (x) − ∂ ν A µ (x) (22)<br />

The electric and magnetic fields E and B have as sources charges and currents.<br />

We write them as<br />

j µ (x) = (ρ(x),j(x))<br />

and then we can note the Lagrangean of the classical electromagnetic field coupled<br />

to the charge current as<br />

L = − 1 4 F µνF µν − j µ A µ (23)<br />

The equ<strong>at</strong>ions of motion an be derived using (2) and they are the famous<br />

Maxwell equ<strong>at</strong>ions in covariant form:<br />

∂ ν F µν (x) = j µ (x) inhom.eq. (24)<br />

∂ λ F µν (x) + ∂ µ F νλ (x) + ∂ ν F λµ (x) = 0 hom.eq. (25)<br />

If one writes these equ<strong>at</strong>ions explicitely in terms of E and B one obtains the<br />

usual Maxwell equ<strong>at</strong>ions of classical electrodynamics in the non-covariant form.<br />

1.5 Fermion-Boson-coupling<br />

The free field lagrangeans are not very interesting because they are interacting<br />

with nothing. The most popular interaction, which occurs e.g. in all elementary<br />

gauge fields, is the interaction of a fermion field with a boson field. If we take<br />

the simplest case of scalar boson field coupled minimally to the fermion field we<br />

have<br />

L = ψ [iγ µ ∂ µ − m]ψ + 1 (<br />

∂λ σ∂ λ σ − µ 2 σ 2) − gψσψ<br />

2<br />

One can write down the equ<strong>at</strong>ions of motion immedi<strong>at</strong>ely:<br />

∂ µ<br />

∂L<br />

∂(∂ µ ψ) − ∂L<br />

∂ψ = 0 ⇒ [iγµ ∂ µ − m]ψ = gσψ<br />

∂ µ<br />

∂L<br />

∂(∂ µ σ) − ∂L<br />

∂σ = 0 ⇒ (<br />

∂ 2 + µ 2) σ = gψψ<br />

Apparently the bosonic field σ modifies the equ<strong>at</strong>ion of motion of the fermion,<br />

such th<strong>at</strong> this is no longer a free fermion. Vice versa the fermion field in the<br />

form of a scalar current gψψ serves as a source for the boson scalar field σ. We<br />

will meet this simple structure very often.<br />

9


1.6 Bilinear covariants:<br />

In the previous subsection we had a very simple coupling of the fermion field<br />

to a boson field. A realistic coupling of a fermion field with charge q to the<br />

electromagnetic field is given by the Lagrangean:<br />

with the electromagnetic current<br />

L = ψ [iγ µ ∂ µ − m]ψ − 1 4 F µνF µν − j µ A µ (26)<br />

j µ = −ieψγ µ ψ<br />

Apparently here the source for the A-field is of the form gψγ µ ψ, which is a<br />

vector-current. Actually there are more complic<strong>at</strong>ed currents known, which<br />

have other 4-tensorial characters. The structure of the field and the structure<br />

of the source must be consistent, since the source gener<strong>at</strong>es the bosonic field<br />

and the total Lagrangean incorpor<strong>at</strong>ing the coupling must be a scalar function.<br />

We have the fol<strong>low</strong>ing covariant bilinears, which form a complete set and which<br />

could serve as sources for the corresponding bosonic fields:<br />

Bilinear Covariant (27)<br />

field<br />

ψψ<br />

ψγ µ ψ<br />

ψγ 5 ψ<br />

ψγ µ γ 5 ψ<br />

ψσ µν ψ<br />

name<br />

S(x) scalar<br />

J µ (x) vector<br />

P(x) pseudoscalar<br />

(x) pseudo-vector<br />

T µν (x) tensor<br />

J µ 5<br />

Covariance means, th<strong>at</strong> these bilinears have in all inertial systems the same<br />

structure. I.e. if you perform a Lorentz-transform<strong>at</strong>ion from IS to IS’<br />

x ′ µ = Λ ν µx ν (28)<br />

then e.g. the 4-vector (ψγ µ ψ) ρσ in IS gets transformed into (ψγ µ ψ) ′ ρσ =<br />

(Λ ν µψγ ν ψ) ρσ in IS’. The ρσ are indices of the m<strong>at</strong>rixelements in the Gammam<strong>at</strong>rices.<br />

1.7 Parity, time reversal, charge conjug<strong>at</strong>ion<br />

There are some fundamental properties of space time, which play an important<br />

role in physics and field theory:<br />

Parity:<br />

x = (x 0 ,x) → x P = (x 0 , −x)<br />

Time reversal: x = (x 0 ,x) → x T = (−x 0 ,x)<br />

The effect of these transform<strong>at</strong>ion on a fermion field ψ(x) will be implemented<br />

by a unitary oper<strong>at</strong>or P for parity and an antiunitary oper<strong>at</strong>or T for<br />

time reversal. We have<br />

Pψ(x)P −1 = γ 0 ψ(x P ) ::::::::::::::::: Tψ(x)T −1 = iγ 1 γ 3 ψ(x T )<br />

10


There is a third transform<strong>at</strong>ion, which maps m<strong>at</strong>ter into antim<strong>at</strong>ter, this is<br />

Charge conjug<strong>at</strong>ion:<br />

Cψ(x)C −1 = iγ 2 γ 0 ψ(x) T<br />

The effects on a scalar boson field are<br />

Pφ(x)P −1 = φ(x P )<br />

Tφ(x)T −1 = φ(x T )<br />

The effects on a photon field are<br />

Cφ(x)C −1 = φ ∗ (x)<br />

PA µ (x)P −1 = A µ (x P )<br />

TA µ (x)T −1 = A µ (x T )<br />

CA µ (x)C −1 = −A µ (x)<br />

The response of the fermion bilinears to these transform<strong>at</strong>ions can be easily<br />

calcul<strong>at</strong>ed and is given in the table<br />

Discrete Transform<strong>at</strong>ions (29)<br />

C P T<br />

S(x) S(x P ) S(x T )<br />

P(x) −P(x P ) −P(x T )<br />

−j µ (x) J µ (x P ) j µ (x T )<br />

j µ 5 (x) −j 5µ(x P ) j 5µ (x T )<br />

−T µν (x) T µν (x P ) −T µν (x T )<br />

(Attention: moving of indices µ,ν because {γ µ ,γ ν } = 2g µν )<br />

The Standard Model involving electromagnetic, weak and colour-gauge fields,<br />

is invariant under CPT = 1. For the fol<strong>low</strong>ing it will be important to note the<br />

behaviour of those covariants.<br />

2 Canonical field quantiz<strong>at</strong>ion<br />

2.1 General remarks:<br />

The fundamental degrees of freedom are the interacting fields. In the above section<br />

the fields are assumed to be classical. In analogy to simple non-rel<strong>at</strong>ivistic<br />

quantum mechanics we proceed now to quantize the fields. This step, which is<br />

11


here only sketched shortly, is necessary if one wants to describe modern experiments<br />

in the field of particle and radi<strong>at</strong>ion physics. We review her the canonical<br />

quantiz<strong>at</strong>ion process. It consists in quantizing the free massive fermion and<br />

boson fields and in tre<strong>at</strong>ing interaction terms by means of perturb<strong>at</strong>ion theory<br />

lieading to Feynman-diagrams. There is also the quantiz<strong>at</strong>ion procedure of interacting<br />

fields by means of p<strong>at</strong>h integrals, which will primarily be used l<strong>at</strong>er in<br />

quantizing interacting non-abelian fields.<br />

The canonically quantized fields give rise to the definition of physical particles.<br />

These particles can be encountered in detectors by “click” or a flash of<br />

of light. The particles have angular momentum, parity etc. which are characteristics<br />

corresponding to the ones of the original fields and the corresponding<br />

Lagrangean and equ<strong>at</strong>ion of motion. As we will see, the quantiz<strong>at</strong>ion process<br />

takes in a n<strong>at</strong>ural way into account, th<strong>at</strong> the particles are identical and th<strong>at</strong> we<br />

have fermions (antisymmetrized) and bosons (symmetrized). The mass appearing<br />

in the Lagrangean is (after renormaliz<strong>at</strong>ion etc.) closely rel<strong>at</strong>ed to the mass<br />

of the freely propag<strong>at</strong>ing particle.<br />

The quantiz<strong>at</strong>ion of non-rel<strong>at</strong>ivistic quantum mechanics consisted in the<br />

fol<strong>low</strong>ing recipe: Consider a Lagrangean L(q i (t), ˙q(t)). Calcul<strong>at</strong>e the canonical<br />

momentum p i (t) canonical to q i (t) by<br />

p i (t) =<br />

∂L<br />

∂ ˙q i (t)<br />

and promote the canonical momentum to an oper<strong>at</strong>or ̂p i (t) with the demand<br />

[<br />

q i (t), ̂p<br />

]<br />

j (t) = iδ ij<br />

The quantiz<strong>at</strong>ion takes place for both quantities <strong>at</strong> the same time t.<br />

The quantiz<strong>at</strong>ion of fields is similar, but more complic<strong>at</strong>ed. In particular<br />

we have the situ<strong>at</strong>ion, th<strong>at</strong> the different sort of fields (Klein-Gordon, Dirac,<br />

Maxwell) require different quantiz<strong>at</strong>ion presriptions, because their field quanta<br />

are bosons and fermions and photons, repectively..<br />

2.2 Quantized Klein-Gordon-Field:<br />

Take the above Klein-Gordon Lagrangean density eq.(4)<br />

L = 1 2<br />

(<br />

∂λ φ∂ λ φ − µ 2 φ 2)<br />

and calcul<strong>at</strong>e the conjug<strong>at</strong>e momentum<br />

π(x) =<br />

∂L<br />

∂(∂ 0 φ)<br />

⇒<br />

π(x) = ∂ 0 φ(x)<br />

Then demand the equal time commut<strong>at</strong>ion rules<br />

[<br />

]<br />

φ i (t,x),π j (t,x ′ ) = iδ (3)<br />

ij (x − ) (30)<br />

x′<br />

12


[<br />

] [<br />

]<br />

φ i (t,x),φ j (t,x ′ ) = π i (t,x),π j (t,x ′ ) = 0<br />

Actually this demand is sufficient for a complete quantiz<strong>at</strong>ion. Mostly one<br />

formul<strong>at</strong>es this in another way using the plane wave solutions of the free Klein-<br />

Gordon equ<strong>at</strong>ion as a basis to expand the fields. (This is convenient, but by<br />

no means necessary, one also could use an expansion into partial waves.). Then<br />

one has for the classical field<br />

∫<br />

d 3 k<br />

]<br />

φ(t,x) = √<br />

[a(k)exp(i(kx − ω k t) + a(k) ∗ exp(−i(kx − ω k t)<br />

(2π)3 2ω k<br />

∫<br />

π(t,x) = −i<br />

√<br />

d 3 ωk<br />

[<br />

]<br />

k<br />

(2π) 3 a(k)exp(i(kx − ω k t) − a(k) ∗ exp(−i(kx − ω k t)<br />

2<br />

with ω k = + √ k 2 + µ 2 , such th<strong>at</strong> the φ(t,x) fulfills the Klein Gordon equ<strong>at</strong>ion<br />

(5) The demand to the above expressions, equivalent in both directions, is<br />

realized by promoting the coefficients a(k)and a ∗ (k) to oper<strong>at</strong>ors, i.e.to cre<strong>at</strong>ion<br />

and annihil<strong>at</strong>ion oper<strong>at</strong>ors of the field quanta. Thus we demand the well known<br />

commut<strong>at</strong>or rules of bosonic single particles<br />

[ ]<br />

a(k),a † (k ′ ) = δ (3) (k − k ′ ) (31)<br />

[ ] [ ]<br />

a(k),a(k ′ ) = a † (k),a † (k ′ ) = 0<br />

By this the fields φ and π become oper<strong>at</strong>ors as well (we do not indic<strong>at</strong>e this by<br />

a ”head”):<br />

∫<br />

φ(t,x) =<br />

d 3 k<br />

]<br />

√<br />

[a(k)exp(−ikx + a(k) † exp(+ikx)<br />

(2π)3 2ω k<br />

(32)<br />

∫<br />

π(t,x) = −i<br />

√<br />

d 3 ωk<br />

[<br />

]<br />

k<br />

(2π) 3 a(k)exp(−ikx) − a(k) † exp(+ikx)<br />

2<br />

(33)<br />

with k = (k 0 = ω k ,k). By construction the quantized fields (32) and (33)<br />

fulfill the commut<strong>at</strong>ion rules of the scalar boson fields (30). It is now clear<br />

how one proceeds in general. Any quantity, depending on the classical fields<br />

φ, π becomes now an oper<strong>at</strong>or by promoting the φ, π to oper<strong>at</strong>or fields. Hence<br />

in principle the full theory is now quantized. Calcul<strong>at</strong>ions become much more<br />

complic<strong>at</strong>ed.<br />

There are several important further developements in the quantized field<br />

theory, which all have their analogies in classical field theory where commut<strong>at</strong>ors<br />

are replaced by Poisson-brackets. We define the oper<strong>at</strong>or of the hamiltonian<br />

density<br />

H(x) = π(x,t)∂ 0 φ(x,t) − L<br />

13


and the hamiltonian oper<strong>at</strong>or<br />

∫<br />

H(t) =<br />

d 3 xH(x)<br />

One obtains by easy calcul<strong>at</strong>ions the fol<strong>low</strong>ing fe<strong>at</strong>ures<br />

i[H(t),π(x,t)] = ∂ 0 π(x,t)<br />

i[H(t),φ(x,t)] = ∂ 0 φ(x,t)<br />

One defines also the momentum oper<strong>at</strong>or<br />

∫<br />

P k (t) = d 3 x [ π(x,t)∂ k φ(x,t) + φ(x,t)∂ 0 π(x,t) ]<br />

with the properties<br />

i [ P k (t),π(x,t) ] = ∂ k π(x,t)<br />

i [ P k (t),φ(x,t) ] = ∂ k φ(x,t)<br />

The proofs of these equ<strong>at</strong>ions are simple: One inserts the quantized fields, uses<br />

the commut<strong>at</strong>or rules (30), gets zeros and delta-funtions, performs the integrals<br />

over space and gets the above formulae. If one combines H and P k to one<br />

4-Vector oper<strong>at</strong>or P µ and considers a general oper<strong>at</strong>or F which is a function of<br />

the fields and depends through them on (x,t) then one can write<br />

An often used fe<strong>at</strong>ure is<br />

i[P µ (t),F(x,t)] = ∂ µ F(x,t)<br />

F(x) = exp(iPx)F(0)exp(−iPx) (34)<br />

F(0) = exp(−iPx)F(x)exp(iPx)<br />

If one takes m<strong>at</strong>rix elements of such an oper<strong>at</strong>ion between many body st<strong>at</strong>es,<br />

gener<strong>at</strong>ed by the a(k) † , of given 4-momentum p and p ′ one obtains the often<br />

used formula<br />

〈p ′ |F(x) |p〉 = exp(i(p ′ − p)x)) 〈p ′ |F(0) |p〉 (35)<br />

One needs in the fol<strong>low</strong>ing the (bare) Feynman-Propag<strong>at</strong>or of the Klein-<br />

Gordon field. It is defined as<br />

with the time ordering oper<strong>at</strong>or given by<br />

i∆ F (x) = 〈0| T {φ(x)φ † (0) |0〉 (36)<br />

T {φ(x)φ † (y)} = θ(x 0 − y 0 )φ(x)φ † (y) + θ(y 0 − x 0 )φ † (y)φ(x) (37)<br />

The expression for the Feynman propag<strong>at</strong>or is<br />

∫<br />

d 4 k exp(−ikx)<br />

∆ F (x) =<br />

(2π) 4 k 2 − m 2 + iε<br />

(38)<br />

14


2.3 Quantized Dirac-field:<br />

The quantiz<strong>at</strong>ion of the Dirac-field is conceptually similar to the one of the<br />

Klein-Gordon field. We again calcul<strong>at</strong>e the spinor field π(t,x) conjug<strong>at</strong>e to the<br />

spinor ψ(t,x) using (3) applied to the Lagrangean of the free Dirac field (6).<br />

The conjug<strong>at</strong>e fermion momentum appears to be<br />

π(t,x) = iψ † (t,x)<br />

Technically the quantiz<strong>at</strong>ion prescription of the fermion fields (Dirac fields) is<br />

different from the one of the boson fields (Klein Gordon fields) because the<br />

corresponding field quanta are fermions r<strong>at</strong>her than bosons. We demand<br />

{<br />

}<br />

ψ i (t,x),π j (t,x ′ )<br />

= iδ (3)<br />

ij (x − x′ )<br />

{<br />

} {<br />

}<br />

ψ i (t,x),ψ j (t,x ′ ) = π i (t,x),π j (t,x ′ ) = 0<br />

where the first equ<strong>at</strong>ion can also be replaced by<br />

{<br />

ψi (t,x),ψ † (t,x) } = δ (3)<br />

ij (x − x′ )<br />

The indices i,j are Dirac indices of the spinor commponents. The expansion<br />

of the quantized fields in terms of the plane wave solutions u(k) and v(k) of<br />

the free Dirac equ<strong>at</strong>ion (7) with mass m reads with k = (k 0 = ω k ,k) and<br />

ω k = + √ k 2 + m 2 :<br />

ψ(t,x) = ∑ ∫<br />

d 3 √<br />

k m<br />

[<br />

]<br />

√ c r (k)u r (k) exp(−ikx) + d †<br />

(2π)<br />

3 ω<br />

r(k)v r (k)exp(+ikx)<br />

k<br />

r=1,2<br />

(39)<br />

with the commutaion rules for the fermionic cre<strong>at</strong>ion and annihil<strong>at</strong>ion oper<strong>at</strong>ors<br />

for particles c and anti-particles d:<br />

{ }<br />

c r (k),c † s(k ′ ) = δ rs δ (3) (k − k ′ ) (40)<br />

{ } { }<br />

c r (k),c s (k ′ ) = c † r(k),c † s(k ′ ) = 0<br />

The analogous expressions hold for the oper<strong>at</strong>ors d,d † of the antifermions. { In }<br />

addition any anticommut<strong>at</strong>or between a c,c † and a d,d † vanishes, e.g. c r (k),d † s(k ′ ) =<br />

0<br />

One needs in the fol<strong>low</strong>ing the (bare) Feynman-Propag<strong>at</strong>or of the Dirac field.<br />

It is defined as<br />

iS F (x) = 〈0|T {ψ(x) ¯ψ(0) |0〉<br />

with the time ordering oper<strong>at</strong>or given [with a different sign compared to (37)]<br />

by<br />

T {ψ(x) ¯ψ(y)} = θ(x 0 − y 0 )ψ(x) ¯ψ(y) − θ(y 0 − x 0 ) ¯ψ(y)ψ(x) (41)<br />

15


and it is given by<br />

or<br />

S F (x) = (iγ µ ∂ µ + m) ∆ F (x)<br />

∫<br />

S F (x) = d 4 k<br />

exp(−ikx)<br />

γ µ k µ − m + iε<br />

(42)<br />

2.4 Quantized Maxwell-field:<br />

In principle the quantiz<strong>at</strong>ion of the Maxwell-field proceeds similarly to the one<br />

of the Klein-Gordon- and the Dirac-field. There is, however, a fundamental<br />

difference.<br />

If we take the Lagrangean (20) of the free Maxwell field and calcul<strong>at</strong>e the<br />

conjug<strong>at</strong>e momentum field (3), we obtain<br />

π µ (x) =<br />

∂L<br />

∂(∂ 0 A µ ) = −F µ0 (x)<br />

This yields π 0 (x) = −F 00 (x) = 0, which does not al<strong>low</strong> to get a quantiz<strong>at</strong>ion<br />

presription for µ = 0. On the other hand one likes to preserve Lorentz-covariance<br />

so th<strong>at</strong> one cannot distinguish a particular Lorentz-indices µ. Thus one must<br />

proceed differently. A Lagrangean, which is suitable for the quantiz<strong>at</strong>ion procedure<br />

has been suggested by Fermi and is<br />

Here we get<br />

L = − 1 2 (∂ µA ν (x))(∂ ν A µ (x)) − j µ (x)A ν (x)<br />

π µ (x) =<br />

∂L<br />

∂(∂ 0 A µ ) = −∂ 0A µ (x)<br />

This can be quantized, demanding<br />

[<br />

A µ (t,x), A ˙<br />

]<br />

ν (t,x ′ ) = −ig µν δ (3) (x − x ′ )<br />

. The problem, however, is now, th<strong>at</strong> the equ<strong>at</strong>ions of motion of this Lagrangean<br />

do not give the Maxwell equ<strong>at</strong>ions. There appear additional terms which are<br />

only vanishing in the Laurentz-Gauge:<br />

∂ µ A µ (x) = 0 (43)<br />

This means, the equ<strong>at</strong>ions of motion reduce to Maxwell equ<strong>at</strong>ions only in a<br />

special gauge, which contradicts gauge invariance.<br />

All these problems have been solved in a quantiz<strong>at</strong>ion formalism invented by<br />

Gupta and Bleuler. In this theory one quantizes the photon field (i.e. Maxwell<br />

field) analogous to eq.(30) leading then to the quantized photon field oper<strong>at</strong>or<br />

A µ (x) =<br />

3∑<br />

∫<br />

r=0<br />

d 3 k [<br />

√ ε<br />

µ<br />

r (k)a r (k)exp(−ikx + ε µ r(k)a(k) † exp(+ikx) ]<br />

(2π)3 2ω k<br />

16


with the polariz<strong>at</strong>ion vectors given by<br />

ε µ 0 (k) = nµ = (1,0,0,0) ε µ r(k) = (0, ε r (k) for r = 1,2,3<br />

and ε r (k) being a orthogonal dreibein. In covarianter form one can write<br />

and<br />

One can write in a covariant way<br />

ε µ r(k)ε µr (k) = −ζ r δ rs<br />

∑<br />

ζ r ε µ r(k)ε ν r(k) = −g µν<br />

r<br />

ζ 0 = −1 ζ 1 = ζ 2 = ζ 3 = +1<br />

ε µ 3 (k) = kµ − (kn)n µ<br />

[(kn) 2 − k 2 ] 1/2<br />

and we call ε µ 1 (k),εµ 2 (k) transversal, εµ 3 (k) longitudial and εµ 0 (k) skalar or timelike.edvia<br />

plane waves to cre<strong>at</strong>ion and annihil<strong>at</strong>ion oper<strong>at</strong>ors for photons with<br />

the fol<strong>low</strong>ing commut<strong>at</strong>or rules:<br />

[ ]<br />

a r (k),a † s(k ′ ) = δ rs δ (3) (k − k ′ ) (44)<br />

[ ]<br />

a r (k),a s (k ′ ) = [ a † r(k),a † s(k ′ ) ] = 0<br />

for the four spinor components of A µ µ = 0,1,2,3. and the four polariz<strong>at</strong>ion<br />

components of the phon field r,s = 0,1,2,3. However the Lorentz-condition (43<br />

leads to a additional constraint for physical (measurable) photon st<strong>at</strong>es |Ψ〉:<br />

[a 3 (k) − a 0 (k)] |Ψ〉 = 0 for all k<br />

Das h<strong>at</strong> zur Folge, daß reale physikalische Photonen weder Longitudinalkomponente<br />

noch eine Zeitkomponente besitzen und transversal sind.<br />

The Gupta-Bleuler formalim has been formul<strong>at</strong>ed for abelian gauge theories.<br />

In principle one can generalize it to non-abelian ones, however, nowadays one<br />

prefers p<strong>at</strong>h-integral quantiz<strong>at</strong>ion procedures which can equally well be applied<br />

to abelian and non-abelian gauge fields like the gluon-field or the fields of the<br />

elctro-weak bosons.<br />

One needs in the fol<strong>low</strong>ing the (bare) Feynman-Propag<strong>at</strong>or of the Maxwell<br />

field. It is defined analogous to eqs.(37,38) as<br />

iD µν<br />

F (x) = 〈0|T {Aµ (x)A ν (0) |0〉<br />

and it is given by (in the limit of boson mass in the bosonic propag<strong>at</strong>or going<br />

to zero)<br />

D µν<br />

F<br />

(x) = − lim<br />

m→0 gµν ∆ F (x,m)<br />

or<br />

∫<br />

D µν<br />

F (x) = −gµν d 4 k exp(−ikx)<br />

k 2 (45)<br />

+ iε<br />

17


3 Symmetries and currents<br />

3.1 Elements of Lie-Group theory<br />

3.1.1 Lie-Groups SU(N)<br />

Symmetries play a tremendous role in elementary particle field theory and<br />

hadronic physics, particularly those associ<strong>at</strong>ed with Lie-Groups . Thus we discuss<br />

in this section some basic properties of Lie-groups. Details can be found in<br />

group theory books (e.g. F. Stancu). We consider the fol<strong>low</strong>ing transform<strong>at</strong>ion<br />

applied to an abstract Hilbert vector |Ψ〉:<br />

with<br />

|Ψ〉 → |Ψ ′ 〉 = U(θ a ) |Ψ〉<br />

U(θ a ) = exp(−iθ a F a ) (46)<br />

The θ a are real numbers, a = 1,2,...,N 2 − 1. The oper<strong>at</strong>ors U(θ a ) are<br />

members of a group, i.e. the Lie-group SU(N), if the oper<strong>at</strong>ors F a are hermitian<br />

and are the gener<strong>at</strong>ors of the Lie-group SU(N). By definition the F a fulfill a<br />

closed algebra obbeying the fol<strong>low</strong>ing commut<strong>at</strong>ion rules<br />

[<br />

F a ,F b] = if abc F c (47)<br />

with the normaliz<strong>at</strong>ion<br />

Tr(F a F b ) = 1 2 δab<br />

The numbers f abc are called structure functions of the Lie-group. They are<br />

totally antisymmetric ( i.e. f abc = −f bac = −f acb = −f cba etc.) and determine<br />

the group properties uniquely. For the SU(N) group one has<br />

f abc f abd = Nδ cd<br />

In SU(2) we jave f abc = ε abc . The anticommut<strong>at</strong>or of the gener<strong>at</strong>ors defines the<br />

symmetric coefficients d abc with d abc = d bac = d acb = d cba etc:<br />

{<br />

F a ,F b } = 1 3 Iδab + d abc F c<br />

The Lie-Group SU(N) can be shown to be of rank r = N − 1. This means<br />

th<strong>at</strong> among the gener<strong>at</strong>ors F a there exist r = N −1 gener<strong>at</strong>ors, which commute<br />

with each other. Each Lie-group SU(N) has N −1 Casimir-oper<strong>at</strong>ors C i i =<br />

1,2,...N − 1, which are defined by the property th<strong>at</strong> they commute with each<br />

of the gener<strong>at</strong>ors and hence with each of the group members:<br />

[C i ,F a ] = 0 i = 1,2,...N − 1 a = 1,....,N<br />

18


The Casimir oper<strong>at</strong>ors and the commuting gener<strong>at</strong>ors have common eigenvectors.<br />

They are grouped in multiplets |c 1 ,...c N−1 ,α 1 ,....,α r 〉 .The multiplets are<br />

characterized by the eigenvalues c i of the Casimir oper<strong>at</strong>ors C i , and all eigenvectors<br />

within a multiplet are characterized by r = N − 1 quantum numbers of<br />

the commuting gener<strong>at</strong>ors. The multiplets can be explicitely constructed from<br />

the commut<strong>at</strong>or rules of SU(N).<br />

3.1.2 Represent<strong>at</strong>ions<br />

In the fol<strong>low</strong>ing we need represent<strong>at</strong>ions of a Lie-group. One needs this represent<strong>at</strong>ion<br />

if one represents the Hilbert vector |Ψ〉 by a n-tupel with n components<br />

in some basis (n-spinor). A represent<strong>at</strong>ion of a group is a mapping of the group<br />

element U(θ a ) on a unitariy n · n m<strong>at</strong>rix U(θ a ):<br />

U(θ a ) ↦→ U(θ a )<br />

where the mapping is such th<strong>at</strong> the group properties are fulfilled. Th<strong>at</strong> means<br />

the represent<strong>at</strong>ion of a product of two oper<strong>at</strong>ors equals the product of the two<br />

represent<strong>at</strong>ion. Furthermore we have<br />

UU † = 1 and detU = 1<br />

Usually the represent<strong>at</strong>ion of a gener<strong>at</strong>or F a is called 1 2 λa . The m<strong>at</strong>rices 1 2 λa<br />

are traceless and fulfill the fol<strong>low</strong>ing commut<strong>at</strong>or rules<br />

[ λ<br />

a<br />

2 , λb<br />

2<br />

]<br />

abc λc<br />

= if<br />

2<br />

(48)<br />

with normaliz<strong>at</strong>ion<br />

The anticommut<strong>at</strong>or of the λ-m<strong>at</strong>rices fulfills:<br />

{<br />

λa<br />

2 , λ }<br />

b<br />

= 1 2 3 δab + d abc λ c<br />

2<br />

Tr(λ a λ b ) = 2δ ab (49)<br />

(50)<br />

The represent<strong>at</strong>ion of the oper<strong>at</strong>or (46) is given by the N · N m<strong>at</strong>rix U(θ a )<br />

with<br />

U(θ) = exp(−i λa θ a<br />

2 ) (51)<br />

For SU(N) with N > 1 we have a non-abelian group with<br />

U(θ 1 )U(θ 2 ) ≠ U(θ 2 )U(θ 1 )<br />

There exists a fundamental represent<strong>at</strong>ion of SU(N). It is the smallest nontrivial<br />

represent<strong>at</strong>ion and is formed by N-spinors, which transform by the N ·Nm<strong>at</strong>rix<br />

ψ α −→ Uβ α ψ β α,β = 1,..,N<br />

19


Complex conjug<strong>at</strong>e N-spinors (which should be presented by a row) transform<br />

by the hermitian-conjug<strong>at</strong>e m<strong>at</strong>rix<br />

ψ † α −→ ψ † β U †β<br />

α<br />

α,β = 1,..,N<br />

where the fol<strong>low</strong>ing conventions hold: ψ α † = (ψ α ) † and U α<br />

†β<br />

clear th<strong>at</strong> the combin<strong>at</strong>ion ψ αψ † α is an SU(N) invariant<br />

= ( U †) β<br />

. It is<br />

α<br />

ψ † αψ α −→ ψ † β U †β<br />

α U α γ ψ γ = ψ † αψ α = ψ † ψ<br />

There are also quantities transforming according to the (N 2 − 1)-dimensional<br />

adjoint represnt<strong>at</strong>ion:<br />

A a −→ O ab A b<br />

with<br />

O ab = 1 2 Tr ( U † λ a Uλ b) and O ab O ac = δ bc<br />

To verify the lawst eq. you need the Fiertz identity<br />

Fiertz identity<br />

( ) λ<br />

a α ( λ<br />

a<br />

β<br />

2 2<br />

) γ<br />

δ<br />

= − 1<br />

2N δα βδ γ δ + 1 2 δα δ δ γ β<br />

One can show th<strong>at</strong> the fol<strong>low</strong>ing two comintions are invariants of the SU(N)<br />

transform<strong>at</strong>ions<br />

A a A a = inv A a ψ † λ a ψ=inv<br />

3.1.3 Gell-Mann m<strong>at</strong>rices<br />

The represent<strong>at</strong>ions of SU(N) are not unique. The most used ones for SU(2)<br />

are the Pauli-m<strong>at</strong>rices (12). For SU(3) we have the Gell-Mann m<strong>at</strong>rices:<br />

The Gell-Mann-m<strong>at</strong>rices are given<br />

⎛<br />

λ 1 = ⎝<br />

⎛<br />

λ 4 = ⎝<br />

⎛<br />

λ 7 = ⎝<br />

0 1 0<br />

1 0 0<br />

0 0 0<br />

0 0 1<br />

0 0 0<br />

1 0 0<br />

0 0 0<br />

0 0 −i<br />

0 i 0<br />

⎞<br />

Gell-Mann-M<strong>at</strong>rices (52)<br />

⎛<br />

⎠ λ 2 = ⎝<br />

⎞<br />

⎛<br />

⎠ λ 5 = ⎝<br />

⎞<br />

⎠<br />

λ 8 = 1 √<br />

3<br />

⎛<br />

⎝<br />

0 −i 0<br />

i 0 0<br />

0 0 0<br />

0 0 −i<br />

0 0 0<br />

i 0 0<br />

⎞<br />

⎛<br />

⎠ λ 3 = ⎝<br />

⎞<br />

1 0 0<br />

0 1 0<br />

0 0 −2<br />

⎛<br />

⎠ λ 6 = ⎝<br />

1 0 0<br />

0 −1 0<br />

0 0 0<br />

0 0 0<br />

0 0 1<br />

0 1 0<br />

The structure coefficients f abc are totally antisymmetric, i.e if one interchanges<br />

two indices one obtains a minus-sign:<br />

⎞<br />

⎠<br />

Antisymmetric Coefficients (53)<br />

⎞<br />

⎠<br />

⎞<br />

⎠<br />

20


f 123 = 1<br />

f 147 = −f 156 = f 246 = f 257 = f 345 = −f 367 = 1 2<br />

f 458 = f 678 = 1 2√<br />

3<br />

All other f abc either obtained by a transposition of two arbitrary indices ( for<br />

each transpposition you have a minus-sign) pre they are Zero. The symmetric<br />

coefficients d abc are given by<br />

Symmetric coefficients (54)<br />

d 118 = 1 √<br />

3<br />

,d 146 = 1 2 ,d 157 = 1 2 ,d 228 = 1 √<br />

3<br />

d 247 = − 1 2 ,d 256 = 1 2 ,d 338 = √ 1 ,d 344 = 1 3 2<br />

d 355 = 1 2 ,d 366 = − 1 2 ,d 377 = − 1 2 ,d 448 = − 1<br />

2 √ 3<br />

d 558 = − 1<br />

2 √ 3 ,d 668 = − 1<br />

2 √ 3 ,d 778 = − 1<br />

2 √ 3 ,d 888 = −√ 1 3<br />

All other d abc are obtained by a tranpositon of two arbitrary indices (for each<br />

transposition you have a plus-sign) or are Zero. The SU(3)-Lie-group has the<br />

rank 2, because you find maximally 2 gener<strong>at</strong>ors, which commute with each<br />

other, as one can see <strong>at</strong> the Gell-Mann M<strong>at</strong>rices where λ 3 and λ 8 are diagonal.<br />

This corresponds to the oper<strong>at</strong>or rel<strong>at</strong>ions:<br />

[F 3 ,F 8 ] = 0 ⇒ F 3 , F 8 : can be diagonalized simultaneously<br />

3.2 Noether Theorem<br />

3.2.1 General structure<br />

For any field theory we have an important theorem: Noether Theorem:<br />

Noether Theorem: To every continuous symmetry of a Langrangean there<br />

exists a conserved current<br />

More precisely st<strong>at</strong>ed: Consider the<br />

Lagrangean (1), i.e. L = L(φ i (x),∂ µ φ i (x)) We assume th<strong>at</strong> L depends only<br />

on φ and ∂ µ φ, d.h. NOT on ∂ µ ∂ ν φ :::::: ∂ 2 φ ::::etc. We can have basically any<br />

power of ∂ µ φ, however, if we want a renormalizable theory we are restricted<br />

to the second power. There are also non-renormalizable theories and they also<br />

have conserved currents and they can also be derived in the way shown right<br />

now:<br />

21


In detail the Noether-theorem reads in the simplest case of one field as<br />

fol<strong>low</strong>s: The L be invariant under an infinitesimal change δφ i (x), i.e.<br />

L(φ + δφ) = L(φ)<br />

In fact one needs less, namely th<strong>at</strong> only the action S = ∫ d 3 xL is invariant.<br />

If the symmetry is continuous this implies the existence of a conserved current<br />

J µ (x)<br />

∂ µ j µ (x) = 0<br />

with the charge defined by<br />

which is a constant of motion:<br />

Q(t) = const<br />

∫<br />

Q(t) = d 3 xj 0 (x)<br />

or<br />

dQ(t)<br />

dt<br />

Actually the reasoning for the last equ<strong>at</strong>ion is trivial, because the surface term<br />

<strong>at</strong> infinity is negligeably small. On always assumes th<strong>at</strong> <strong>at</strong> infinity the fields are<br />

vanishing such th<strong>at</strong> any current vanishes. This is not correct for a simple plane<br />

wave, which is of course an idealiz<strong>at</strong>ion, but it is always correct for a physically<br />

gener<strong>at</strong>ed field with a source of finite size.:<br />

dQ(t)<br />

dt<br />

= 0<br />

∫<br />

∫ ∫ ∫<br />

= ∂ 0 d 3 xj 0 (x) = −∂ i d 3 xj i (x) = d 3 x∇j(x) = dfj(x) = 0<br />

since the integral goes along the surface <strong>at</strong> infinity, where the fields composing<br />

the current j are vanishing.<br />

The proof of the Noether theorem is simple: Consider the Lagrangean under<br />

the transform<strong>at</strong>ion<br />

φ → φ + δφ<br />

this means<br />

or<br />

δL = L(φ + δφ) − L(φ)<br />

δL = ∂L<br />

∂φ δφ + ∂L<br />

∂(∂ µ φ) δ(∂ µφ)<br />

we can rewrite the first term by the equ<strong>at</strong>ion of motion eq.(2) such th<strong>at</strong><br />

( ) ∂L<br />

δL = ∂ µ<br />

∂(∂ µ φ) δφ<br />

or<br />

with the current<br />

j µ (x) =<br />

∂ µ j µ (x) = δL (55)<br />

∂L<br />

δφ(x) (56)<br />

∂(∂ µ φ(x))<br />

22


There is often factor −i or +i, which is purely by convention, it is often omitted.<br />

This means, if the Lagrangean is invariant under the transform<strong>at</strong>ion, we have<br />

∂ µ j µ (x) = 0 (57)<br />

This is the general structure of the Noether-Theorem. The current gives rise<br />

to the definition of the charge, which is time independent if the lagragean is<br />

invariant under the transform<strong>at</strong>ion. If not, the charge is time dependent.<br />

Since we have<br />

∫<br />

Q(t) =<br />

d 3 xj 0 (x) :::::::::: dQ(t)<br />

dt<br />

π(x) =<br />

∂L<br />

∂(∂ 0 φ)<br />

the charge can be rewritten as<br />

∫<br />

Q(t) = d 3 xπ(x,t)δφ(x,t)<br />

= 0 (58)<br />

3.2.2 Noether-Theorem and Lie-groups<br />

One mostly uses the Noether Theorem if several fields are mixed by the transform<strong>at</strong>ion<br />

and the Lagrangean remains unchanged. This we will discuss now:<br />

Assume the Lagrangean invariant under the transform<strong>at</strong>ion<br />

φ i → φ i + δφ i<br />

i = 1,...,n<br />

with<br />

δφ i (x) = iɛ a t a ijφ j (x)<br />

and t a with a = 1,2,....,N are n-dimensional represent<strong>at</strong>ions of the gener<strong>at</strong>ors of<br />

the SU(N) Lie-group.This is the most often used case because the Lagrangeans<br />

of the elementary forces whe have in n<strong>at</strong>ure show several of those symmetries and<br />

the interaction of elementary particles with external fields can also be formul<strong>at</strong>ed<br />

in terms of those Noether currents.The fact th<strong>at</strong> we work with a Lie-group means<br />

th<strong>at</strong> the t a fulfill the commut<strong>at</strong>ion rules (48) We can derive easily the structure<br />

of the current and also its conserv<strong>at</strong>ion if the Lagrangean is invariant under the<br />

above transform<strong>at</strong>ion: We have the vari<strong>at</strong>ion<br />

δL = ∂L δφ i +<br />

∂L<br />

∂φ i ∂(∂ µ φ i ) δ(∂ µφ i )<br />

We can use now the equ<strong>at</strong>ion of motion eq(2) to modify the first term on RHS<br />

∂L<br />

∂φ i<br />

yielding<br />

∂L<br />

δL =∂ µ<br />

∂(∂ µ φ i ) δφ + ∂L<br />

∂(∂ µ φ i ) ∂ µ(δφ i )<br />

23


Since the ɛ a are arbitrary we have<br />

( ) ∂L<br />

δL = ɛ a ∂ µ<br />

∂(∂ µ φ i ) ita ijφ j<br />

Defining the Noether-current for each a as<br />

jµ(x) a δL<br />

= −i<br />

δ(∂ µ φ i (x)) ta ijφ j (x) (59)<br />

one can write analogously to eq(55)<br />

δL = ɛ a ∂ µ j a µ(x)<br />

or, if the Lagrangean is invariant, we have analogously to eq.(57):<br />

∂ µ j a µ(x) = 0<br />

a = 1,2,....,N<br />

On can define the charges, which are in general time dependent. If the Lagrangean<br />

is invariant under the considered Lie-group the charges are time independent:<br />

∫<br />

Q a (t) = d 3 xj0(x)<br />

a<br />

Since we have<br />

dQ a (t)<br />

dt<br />

π k (x) =<br />

= 0 a = 1,2,....,N (60)<br />

∂L<br />

∂(∂ 0 φ k )<br />

the charge oper<strong>at</strong>or can be rewritten as<br />

∫<br />

Q q (t) = −i d 3 xπ k (x,t)t a kjφ j (x,t) (61)<br />

3.2.3 Altern<strong>at</strong>ive expression for several fields, Energy momentum<br />

tensor<br />

Actually there is an equivalent and sometimes more convenient way to calcul<strong>at</strong>e<br />

the current of a Lagrangean under a symmetry transform<strong>at</strong>ion (without proof):<br />

This consists in performing an infinitesimal field transform<strong>at</strong>ion, and assuming<br />

the ɛ artificially and only for the moment to be a function of x, this beeing only<br />

a technical trick in order to get another and often more convenient deriv<strong>at</strong>ion<br />

of the current. Hence one considers the infinitesimal transform<strong>at</strong>ion<br />

̂φ i (x) = φ i (x) + ɛ a (x)F a i (φ j (x))<br />

24


The F is a function of the set φ ⃗ i . In the restiriction to constant ɛ the lagrangean<br />

becomes invariant. For an internal symmetry the Noether current is then obtained<br />

by<br />

j a(x) µ ∂<br />

=<br />

∂(∂ µ ɛ a (x)) L( ̂φ i ,∂ µ ̂φi )<br />

One also obtains immedi<strong>at</strong>ely<br />

∂ µ j µ a(x) =<br />

∂L<br />

∂ɛ a (x)<br />

In practice one simply expands the Lagrangean<br />

L( ̂φ i (x), ̂ ∂µ φ i (x)) = L(φ i (x),∂ µ φ i (x)) + ∂ µ ɛ a (x)F µ a (x)<br />

performs the deriv<strong>at</strong>ive and can read off the current.<br />

Proof: Under the above transform<strong>at</strong>ion we can use the fol<strong>low</strong>ing formula<br />

δL = ∂L δφ i +<br />

∂L<br />

∂φ i ∂(∂ µ φ i ) δ(∂ µφ i )<br />

= − ∂L ɛ a Fi a −<br />

∂L<br />

∂φ i ∂ (∂ µ φ i ) [(∂ µɛ a ) Fi a + ɛ a (∂ µ Fi a )]<br />

Using the Euler Eqs. of motion we have<br />

( ) ∂L<br />

δL = −∂ µ ɛ a Fi a −<br />

∂L<br />

∂(∂ µ φ i ) ∂(∂ µ φ i ) [(∂ µɛ a ) Fi a + ɛ a (∂ µ Fi a )]<br />

yielding<br />

( ∂L<br />

δL = −ɛ a ∂ µ<br />

∂(∂ µ φ i ) F i<br />

a<br />

Defining the current as<br />

j µa (x) = −<br />

∂L<br />

∂(∂ µ φ i ) F i<br />

a<br />

) ( ) ∂L<br />

− (∂ µ ɛ a )<br />

∂(∂ µ φ i ) F i<br />

a<br />

associ<strong>at</strong>ed with the above infinitesimal transform<strong>at</strong>ion with the x-dependent<br />

epsilon we have<br />

δL = = ɛ a ∂ µ j µa + (∂ µ ɛ a )j µa<br />

From this fol<strong>low</strong>s th<strong>at</strong> the current may also be defined as<br />

j µa =<br />

which is equivalent to eq.(59). and furthermore<br />

∂<br />

∂ (∂ µ ε a δL (62)<br />

)<br />

∂ µ j µa = ∂ δL (63)<br />

∂εa 25


Without proof, which one finds in any textbook, we quote the concept of<br />

energy momentum tensor. This is defined as<br />

T µν =<br />

∂L<br />

∂(∂ µ φ i ) ∂ν φ i − g µν L (64)<br />

For a system, which is invariant under a shift in space and time, i.e. timeindependent<br />

and transl<strong>at</strong>ionally invariant, we haave the fol<strong>low</strong>ing conserv<strong>at</strong>ion:<br />

∂ µ T µν = 0 (65)<br />

The various components have a clear meaning: The 00-component equals to the<br />

hamiltonian of the system: ∫<br />

H = d 3 xT 00 (66)<br />

And the other 0-components set up the oper<strong>at</strong>ors of the momentum, altogether<br />

a 4-momentum:<br />

∫<br />

P µ = d 3 xT µ0 (67)<br />

Actually we know an important property of P µ which was given alrady in<br />

eq.(34). There it was derived in the context of a free boson field. Here we<br />

see,however, th<strong>at</strong> it is also valid for a general Lagrangean:<br />

3.3 Current algebras<br />

F(x) = exp(iPx)F(0)exp(−iPx)<br />

We consider a = 1,....N gener<strong>at</strong>ors t a ij of the Lie-group G, which means th<strong>at</strong><br />

the t a fulfill the commut<strong>at</strong>ion rules of eq.(48). Fol<strong>low</strong>ing the previous section,<br />

we have also a = 1,...,N charges Q a (t), see eq.(60). By now the Noether<br />

theorem was completely on the classical level. If we quantize the fields φ i (x) in<br />

a canonical way then the charges get quantized as well and become oper<strong>at</strong>ors.<br />

From nowon we tre<strong>at</strong> them as oper<strong>at</strong>ors, omitting, however, any indic<strong>at</strong>ion of<br />

such a promotion. The assertion is th<strong>at</strong> the quantized charge oper<strong>at</strong>ors fol<strong>low</strong><br />

the algebra of the Lie-group considered and hence may be used themselves as<br />

gener<strong>at</strong>ors of the group fulfilling of course eq.(48):<br />

[<br />

Q a (t) ,Q b (t) ] = if abc Q c (t) (68)<br />

The proof is simple, and we demonst<strong>at</strong>e it for a boson field: The quantiz<strong>at</strong>ion<br />

is well defined and requires only the knowledge of the field π(x) conjug<strong>at</strong>e to φ(x)<br />

and the usual quantiz<strong>at</strong>ion procedure. We have (now all fields are oper<strong>at</strong>ors)<br />

eq.(61) from which we can write<br />

[<br />

Q a (t),Q b (t) ] ∫<br />

=<br />

d 3 xd 3 y [ π i (x)t a ijφ j (x),π k (y)t b kmφ m (y) ] x 0 =y 0<br />

26


with x 0 = y 0 .With the oper<strong>at</strong>or identity [AB,CD] = A[B,C]D − C [D,A] B<br />

(valid if[A,C] = [D,B] = 0) we have now<br />

[<br />

Q a (t),Q b (t) ] ∫<br />

= − d 3 xd 3 yπ i (x,t)t a [<br />

ij φj (x,t),π k (y,t) ] t b kmφ m (y,t)<br />

∫<br />

+<br />

d 3 xd 3 yπ k (x,t)t b [<br />

km πj (x,t),φ m (y,t) ] t a ijφ j (y,t)<br />

∫<br />

= −<br />

∫<br />

= −<br />

d 3 xπ k (x,t)i [ t a ,t b] kj φ j(x,t)<br />

d 3 xπ k (x,t)φ j (x,t)f abc t c kj<br />

Witlh the definition Q q (t) given in eq. (61) we get<br />

[<br />

Q a (t),Q b (t) ] = if abc Q c (t) qed<br />

Thus the charge oper<strong>at</strong>ors gener<strong>at</strong>e the changes of the fields like the gener<strong>at</strong>ors<br />

do: A simple proof using [AB,C] = A[B,C] + [A,C] B yields with the<br />

same techniques as above:<br />

[Q a (t),φ k (x,t)] = −t a kjφ j (x,t)<br />

[Q a (t),π k (x,t)] = −t a kjπ j (x,t)<br />

This means, if there is only one field, one immedi<strong>at</strong>ely determines the charge of<br />

a field by calcul<strong>at</strong>ing the commut<strong>at</strong>or of the field with the charge oper<strong>at</strong>or.<br />

He above algebra independent whether the Lagrangean is invariant under<br />

the considered Lie-group transform<strong>at</strong>ion or not. However, if the Lagrangean<br />

is invariant the charges commute with the Hamiltonian of the system.This is<br />

obvious, however it can be shown explicitely with the above techniques.<br />

The Hamiltonian is defined<br />

H = ∑ i<br />

π i (x)∂ 0 φ(x) − L(φ i (x),∂ µ φ i (x))<br />

giving (without proof) for all a<br />

[H,Q a (t)] = 0<br />

One can generalize the charge algebras with the same and simple oper<strong>at</strong>or<br />

techniques to current algebras. One obtains using the above techniques (without<br />

explicit but r<strong>at</strong>her simple proof)<br />

[<br />

Q a (t),j b 0(x,t) ] = if abc j c 0(x,t)<br />

or more generally [<br />

Q a (t),j b µ(x,t) ] = if abc j c µ(x,t) (69)<br />

27


and [<br />

j<br />

a<br />

0 (x,t),j b 0(y,t) ] = if abc j c 0(x,t)δ (3) (x − y) (70)<br />

However be careful: The analogy for space like componends is not valid due to<br />

the so called “Schwinger term”:<br />

[<br />

j<br />

a<br />

0 (x,t),j b i (y,t) ] = if abc j c i (x,t)δ (3) (x − y) + S ab<br />

ij (x) ∂<br />

∂y j<br />

δ (3) (x − y)<br />

The Schwinger term vanishes upon integr<strong>at</strong>ion over the whole space. By this it<br />

does not modify the charge algebra.<br />

3.4 Examples<br />

3.4.1 Free massless fermions, chirality, helicity<br />

We a consider a system of free massles fermions. Let ψ be the solution of this<br />

massles field<br />

[iγ µ ∂ µ ]ψ = 0<br />

We multiply this eu<strong>at</strong>ion from the left with γ 5 and use the {γ 5 ,γ µ } = 0 anticommut<strong>at</strong>or<br />

rules to obtain another solution:<br />

[iγ µ ∂ µ ]γ 5 ψ = 0<br />

We superimpose these solutions to form by construction the combin<strong>at</strong>ions of<br />

definite “chirality” (w<strong>at</strong>ch in books the definition of the γ 5 , they often differ in<br />

sign and hence the definitions of right- and left-handed look different. Here we<br />

use the definition of Bjorken-Drell, Peskin-Schroeder, Izykson-Zuber)<br />

ψ R = 1 2 (1 + γ 5)ψ ψ L = 1 2 (1 − γ 5)ψ (71)<br />

In the massles case the chirality of a free particle is a Lorentz-invariant<br />

concept. For example, particle which is left handed to one observer will appear<br />

left handed to all observers.<br />

On the Lagrangean level we can write<br />

or<br />

L = ψ R [iγ µ ∂ µ ]ψ R + ψ L [iγ µ ∂ µ ]ψ L (72)<br />

with<br />

L = L R + L L<br />

L R = ψ R [iγ µ ∂ µ ]ψ R<br />

L L = ψ L [iγ µ ∂ µ ]ψ L<br />

28


These Lagrange densities are both invariant under the global chiral phase transform<strong>at</strong>ions<br />

ψ R (x) → ψ ′ R(x) = exp(−iα R )ψ R (x)<br />

¯ψR (x) → ¯ψ ′ R(x) = exp(+iα R )ψ R (x)<br />

ψ L (x) → ψ ′ L(x) = exp(−iα L )ψ L (x)<br />

¯ψR (x) → ¯ψ ′ R(x) = exp(+iα R )ψ R (x)<br />

where the phase is real-valued and arbitrary. The associ<strong>at</strong>ed Noether currents<br />

for the present massless free lagrangean are<br />

j µ R (x) = ¯ψ R (x)γ µ ψ R (x) ∂ µ j µ R (x) = 0<br />

j µ L (x) = ¯ψ L (x)γ µ ψ L (x) ∂ µ j µ L (x) = 0<br />

For massless free fermions these chiral currents are conserved. We can construct<br />

chiral charges<br />

∫<br />

Q R,L (t) = d 3 xjR,L(x)<br />

0<br />

From the chiral currents we can construct the vector and axial vector currents<br />

by linear combin<strong>at</strong>ions:<br />

j µ (x) = j µ R + jµ L<br />

vector<br />

j µ 5 (x) = jµ R − jµ L<br />

axial vector<br />

. We also can define vector and axial vector charges<br />

∫<br />

Q(t) = d 3 j 0 (x)<br />

∫<br />

Q 5 (t) =<br />

d 3 j 0 5(x)<br />

The above properties hold if the considered fermions are massless or perhaps<br />

coupled to a field which does not mix the left-handed with the right handed<br />

fermion fields. Actually if we have a mass term in the Lagrangean represented<br />

by a real mass or a scalar field σ(x) then immedi<strong>at</strong>ely right- and left-handed<br />

components are mixed and the separ<strong>at</strong>e right- and left-handed currents are no<br />

longer conserved:<br />

L = ψ R [iγ µ ∂ µ ]ψ R + ψ L [iγ µ ∂ µ ]ψ L − m(ψ L ψ R + ψ R ψ L ) (73)<br />

In this lagrangean we have no more invariance separ<strong>at</strong>ely under SU(2)-L and<br />

SU(2)-R and the above charge oper<strong>at</strong>ors are time dependent. One might think<br />

th<strong>at</strong> the l<strong>at</strong>ter case is the realistic one and the one with massless fermions only<br />

29


an academic construction since in n<strong>at</strong>ure there are no massless fermions. In fact<br />

the massless case is important since in e.g. high energy lepton sc<strong>at</strong>tering one<br />

has the situ<strong>at</strong>ion th<strong>at</strong> the momentum transfer Q 2 of the exchanged photon to<br />

the quarks is very much larger than the mass of the quarks, i.e. Q 2 >> m 2 .In<br />

such a case the mass of the quark is negligible and the quarks can be tre<strong>at</strong>ed<br />

massless.<br />

There is another interesting difference between the massive and the massless<br />

case. This concerns the concept of both “helicity” and “chirality”. Generally<br />

a particle (with or without mass) with definite helicity is a particle, whose<br />

spin and momentum are aligned. This can always be achieved by chosing the<br />

momentum direction as axis of quantiz<strong>at</strong>ion. Positive helicity means then →⇒<br />

and neg<strong>at</strong>ive helicity means →⇐. In general a massive particle with positive<br />

helicity does not have definite chirality and vice versa. For a massive particle<br />

one can always find a Lorentz-transform<strong>at</strong>ion into a inertial frame where the<br />

helicity is positive and into another frame where the helicity is neg<strong>at</strong>ive. If we<br />

have a particle with mass, then the construction of ψ R does not yield a st<strong>at</strong>e<br />

with definite helicity, since helicity depends on the fram whereas chirality does<br />

not Only in the case of a massless particle, which has no rest frame and which<br />

always moves with speed of light the helicity is a lorentz invariant quantity.<br />

For a massless particle positive helicity implies right handedness and vice versa:<br />

This means, if we have a massles particle with definite momentum →then for<br />

ψ R the spin and momentum are aligned and for ψ L they are anti-aligned.<br />

massless: ψ R right-handed →⇒ positive helicity<br />

massless: ψ L left-handed →⇐ neg<strong>at</strong>ive helicity<br />

In the massless limit the concept of chirality is a Lorentz invariant concept and<br />

hence it is a n<strong>at</strong>ural label to be used for massless fermions and a collection of<br />

such particles may be characterized by the number of left-handed and righthanded<br />

particles.<br />

The proof, th<strong>at</strong> in the massless limit helicity and chirality are identical is<br />

simple: Consider the Dirac eq<strong>at</strong>ion for a massless quark and a plane wave<br />

solution in 3-direction<br />

Plane wave in 3-direction yields<br />

p 2 = m 2 = 0 = (p 0 ) 2 − (p 1 ) 2 − (p 2 ) 2 − (p 3 ) 2<br />

p 1 = p 2 = 0 :::::⇒:::: p 0 = p 3 = p<br />

ψ = exp (−ipx) ψ 0 ⇒ γ µ p µ ψ = 0 ⇒ p(γ 0 − γ 3 )ψ = 0 ::::⇒:::: γ 0 ψ = γ 3 ψ<br />

hence with<br />

J 3 = 1 2 σ12 = i 4<br />

[<br />

γ 1 ,γ 2] = i 2 γ1 γ 2<br />

30


J 3 ψ L = i 2 γ1 γ 2 ψ L = i 2 γ0 γ 0 γ 1 γ 2 ψ L = i 2 γ0 γ 1 γ 2 γ 0 ψ L = i 2 γ0 γ 1 γ 2 γ 3 ψ L<br />

where we have used<br />

γ 0 ψ = γ 3 ψ,γ 5 = i γ 1 γ 2 γ 3 , ( γ 5) 2<br />

= I, {γ µ ,γ ν } = 2g µν , {γ 1 ,γ 2 } = 0<br />

Reshuffling gives<br />

J 3 ψ L = + 1 2 γ5 ψ L = − 1 1<br />

2 2 (1 − γ 5)ψ = − 1 2 ψ L<br />

J 3 ψ R = + 1 2 ψ R :::::::::: qed<br />

3.4.2 Free massive Dirac field, Fermion-number, U(1)-transform<strong>at</strong>ion<br />

Consider the lagrangean for the free massive fermion field (6). Apparently this<br />

Lagrangean is invariant under the “global U(1)-transform<strong>at</strong>ion” with:<br />

ψ(x) → ψ ′ (x) = exp(−iα)ψ(x) (74)<br />

ψ(x) → ψ ′ (x) = ψ(x) exp(+iα)<br />

Here α is a real constant. The proof is trivial. We can immedi<strong>at</strong>ely calcul<strong>at</strong>e<br />

the corresponding Noether-current (since δψ = −iαψ) using eq.(56):<br />

yielding<br />

j µ = ∂L 0<br />

∂(∂ µ ψ) δψ<br />

j µ = ψγ µ ψ (75)<br />

The α is a free number, hence it is omitted in the expression for the current j.<br />

The Noether theorem says<br />

∂ µ j µ (x) = 0<br />

This current agrees with the vector current V µ = j µ R + jµ L<br />

. Apparently the<br />

divergence of the current vanishes for massless as well as massive fermions.<br />

Apparently we have<br />

j 0 (x) = ψγ 0 ψ = ψ † ψ<br />

This quantity is positive definite. If one interpretes ψ as a particle, then we<br />

have here the particle density and hence can use the probability interpret<strong>at</strong>ion.<br />

31


3.4.3 Free massless Dirac field, Axial U A (1)-transform<strong>at</strong>ion<br />

We consider the free massles fermion lagrangean eq.(6) with m = 0. We consider<br />

the axial U A (1)-transform<strong>at</strong>ion<br />

ψ(x) → ψ ′ (x) = exp(−iαγ 5 )ψ(x) (76)<br />

corresponding to (<strong>at</strong>tention: sign in the exponent does not change due to anticommut<strong>at</strong>ion<br />

rules (9) of the γ-m<strong>at</strong>rices)<br />

ψ(x) → ψ ′ (x) = ψ(x) exp(−iαγ 5 ) (77)<br />

Wit the anti-commut<strong>at</strong>ion rules of the γ-m<strong>at</strong>rices (9) one can show by direct<br />

calcul<strong>at</strong>ion, th<strong>at</strong> the free massles fermion lagrangean is invariant under the<br />

above transform<strong>at</strong>ion:<br />

L ′ 0 = ψ ′ [iγ µ ∂ µ ]ψ ′ = L 0 = ψ [iγ µ ∂ µ ]ψ<br />

and the corresponding Noether current is the axial current (with :<br />

with<br />

j µ 5 = ψγµ γ 5 ψ (78)<br />

∂ µ j µ 5 (x) = 0<br />

This current agrees with the axial vector current j µ 5 = jµ R − jµ L<br />

. It is interesting<br />

to consider axial transform<strong>at</strong>ions in the case of a massive fermion field eq.(6):<br />

L 0 = ψ [iγ µ ∂ µ − m]ψ<br />

If we now perform the transform<strong>at</strong>ion eq.(76,77) we see th<strong>at</strong> the Lagrangean is<br />

no longer invariant:<br />

L ′ = ψ [iγ µ ∂ µ ]ψ − mψ exp(−2iαγ 5 )ψ<br />

. When we calcul<strong>at</strong>e the current, using eq.(56) we see of course th<strong>at</strong> it is identical<br />

to the one calcul<strong>at</strong>ed with m = 0, since the definition of the current (56) uses<br />

only terms coming from the kinetic energy of the Lagrangean. Howvever if we<br />

calcul<strong>at</strong>e the divergence of the current we see from eq.(55) th<strong>at</strong> it is of course<br />

no longer zero:<br />

∂ µ j µ A (x) = ∂ µψγ µ γ 5 ψ = i2mψγ 5 ψ (79)<br />

Hence: A mass term destroys the invariance of the free fermion field under the<br />

U A (1)-transform<strong>at</strong>ion. Remember the mass term did not destroy the invariance<br />

of the free fermion field under the U(1)-transform<strong>at</strong>ion.<br />

32


3.5 Isospin symmetry<br />

There is a simple example of an effective model, which shows SU(2) isospin<br />

symmetry, which we are going to meet often and particulaly in <strong>QCD</strong>. We will<br />

practice some useful techniques here. The example considered is the nucleonpion<br />

system. Consider a doublet of nucleon fields each being a 4-spinor field<br />

( )<br />

p(x)<br />

ψ(x) =<br />

n(x)<br />

and a triplet of pion fields π a .......a = 1,2,3 with the lagrangean<br />

L = ψ [iγ µ ∂ µ − m]ψ + 1 (<br />

∂λ π a ∂ λ π a − m 2<br />

2<br />

ππ a π a) − gψiτ a π a γ 5 ψ − λ 4 (πa π a ) 2<br />

(80)<br />

with τ a being the Pauli-isospin m<strong>at</strong>rices and the mass m<strong>at</strong>rix m is given by the<br />

nucleon masses M N : ( )<br />

MN 0<br />

m =<br />

0 M N<br />

This Lagrangean describes a bare nucleon (i.e. proton and neutron described<br />

by the Dirac fields p(x) and n(x)) interacting with a pion field in the most<br />

simple way. The M prot = 938 MeV and the neutron is about two MeV heavier.<br />

Thus it is a good approxim<strong>at</strong>ion to neglect the mass difference between proton<br />

and neutron, which would be an isospin viol<strong>at</strong>ing effect. The Lagrangean is<br />

to be understood purely classical. This means, th<strong>at</strong> ψ describes nucleons with<br />

positive energy (i.e. valence nucleons), there does not exist a Dirac-sea, the<br />

vacuum is hence described by ψ = 0, the boson fields do not show any quantum<br />

fluctu<strong>at</strong>ion. For any value of M N the Lagrangean is invariant under the global<br />

SU(2) transform<strong>at</strong>ion (rot<strong>at</strong>ion in isospin space) of the fields<br />

ψ → ψ ′ = Uψ ::::::::::: U(α) = exp(−i τa α a<br />

2 ) (81)<br />

ψ → ψ ′ = ψU †<br />

::::::::::: U † (α) = exp(+i τa α a<br />

2 )<br />

where the are the SU(2) isospin m<strong>at</strong>rices or the Pauli m<strong>at</strong>rices (12). We have<br />

UU †<br />

= U † U = I. This unitarity can be shown explicitely by performing a<br />

Taylor expansion. We show it to second order:<br />

UU † =<br />

(<br />

1 − i τa α a<br />

2<br />

= 1−i τa α a<br />

2 +iτc α c<br />

2 −iτa α a<br />

2 iτc α c<br />

+ 1 ( 2 i<br />

τ<br />

2! 2) a π a τ b π<br />

)(1 b + i τc α c<br />

+ 1 ( )<br />

2 i<br />

τ<br />

2 2! 2) c π c τ d π d<br />

2 ++ 1 2!<br />

( i<br />

2) 2<br />

τ a π a τ b π b + 1 2!<br />

( i<br />

2<br />

) 2<br />

τ c π c τ d π d = 1<br />

The higher orders go similarly. The invariance of the mass tern and of<br />

the kinetic energy term of the fermions under the above transform<strong>at</strong>ion is obvious.<br />

The nucleon-pion interaction is invariant provided the pion fields are<br />

33


transformed as<br />

τ a π a → τ a π ′a = Uτ a π a U †<br />

In this case the action of U on the fermion- and on the boson-field compens<strong>at</strong>e<br />

each other. Thus we demand this transform<strong>at</strong>ion properties for the pionic field.<br />

In order to prove then th<strong>at</strong> also π a π a is invariant one should use the identity<br />

π a π a = 1 2 Tr(τa π a τ b π b ) (82)<br />

from which we see immedi<strong>at</strong>ely th<strong>at</strong> π ′a π ′a = π a π a ,because we have under the<br />

action of the transform<strong>at</strong>ion U.<br />

π ′a π ′a = 1 2 Tr(τa π ′a τ b π ′b ) = 1 2 Tr(Uτa π a U † Uτ b π b U † ) = 1 2 Tr(τa π a τ b π b U † U) = 1 2 Tr(τa π a τ b π b ) = π a π a<br />

Thus the total Lagrangean eq.(80)is invariant under (81). The transform<strong>at</strong>ion<br />

properties of the pion field are defined in a slightly indirect way, which is not<br />

th<strong>at</strong> simple as for the fermion field. However, the response of the individual<br />

pion components to the isospin transform<strong>at</strong>ion can be found from multiplying<br />

the transform<strong>at</strong>ion equ<strong>at</strong>ion for the pion field by τ b and taking the trace using<br />

a well known trace-property of all Lie-groups:<br />

We obtain explicitely:<br />

Tr(τ a τ b ) = 2δ ab (83)<br />

Tr(τ b τ a )π a → Tr(τ b τ a )π ′a = Tr(τ b Uτ a π a U † )<br />

and finally<br />

2δ ab π a → 2δ ab π ′a = Tr(τ b Uτ a π a U † )<br />

π ′a = R ab (α)π b with R ab (α) = 1 2 Tr(τa Uτ b U † )<br />

To determine the isospin current, one can directly apply the formulae for the<br />

conserved current. One also can consider the spacetime-dependent transform<strong>at</strong>ion<br />

with α a (x) now infinitesimal (the ɛ abc enters because of the commut<strong>at</strong>or<br />

rules of the isospin Pauli m<strong>at</strong>rices=:<br />

̂ψ = (1 − iτ a α a (x)/2)ψ ::::::::::: ̂π a = π a − ɛ abc π b α c (x)<br />

Performing this transform<strong>at</strong>ion on the lagrangean gives<br />

L( ̂ψ, ̂π a ) = L(ψ,π a ) + 1 2 ψγµ τ a ∂ µ α a (x)ψ − ɛ abc (∂ µ π a )π b ∂ µ α c (x)<br />

Aplying the general expression for a current eq.(62)<br />

j µ (x) =<br />

∂<br />

∂(∂ µ ɛ(x)) L(̂φ,∂ µ̂φ)<br />

34


in the present situ<strong>at</strong>ion yields<br />

V a µ =<br />

∂<br />

∂(∂ µ α a (x)) L(̂φ,∂ µ̂φ)<br />

Explicit and straight forward calcul<strong>at</strong>ion shows th<strong>at</strong> we obtain the triplet of<br />

conserved iso-vector currents:<br />

V a µ = ψγ µ<br />

τ a<br />

2 ψ + ɛabc π b ∂ µ π c (84)<br />

This current is conserved for each a, as one can explicitely show by direct<br />

calcul<strong>at</strong>ion using the equ<strong>at</strong>ions of motion for ψ and π a . We see th<strong>at</strong> the current<br />

is composed of a fermion part and a pion part, which equally contribute. The<br />

mesonic contribution is called ”meson exchange current”. One can calcul<strong>at</strong>e<br />

the charges60 and one can show (without proof) th<strong>at</strong> they s<strong>at</strong>isfy the charge<br />

commut<strong>at</strong>ion rules eq.(68):<br />

[<br />

Q a (t),Q b (t) ] = iε abc Q c (t)<br />

as the Isospin-Pauli-M<strong>at</strong>rices do, eq.(). This is expected from the general<br />

rules. Actually this current is not the charge current. The charge current<br />

involves only the upper component of ψ (because only the proton is charged)<br />

and only the charged pion fields π + and π − . The isospin current involves upper<br />

and <strong>low</strong>er component and all pion fields.<br />

Actually in case of the nucleon, i.e. the fermion field representing proton and<br />

neutron, the Lagrangean has been and still is very popular. The Lagrangean<br />

implies (without proof) fe<strong>at</strong>ures, which are well corresponding to experiment<br />

for the masses:<br />

m 0 = m p = m n m π + = m π − = m π 0<br />

and for the pion-nucleon-nucleon coupling constants<br />

g ppπ 0 = −g nnπ 0 = g pnπ + / √ 2 = g npπ − / √ 2 π ± = (π 1 ± iπ 2 )/ √ 2<br />

yielding a common value of g2<br />

4π = 14.3. In real life, howevere, we have m n −<br />

m p = 2MeV with m p = 938MeV , and with m π + = 139.5MeV and m π 0 =<br />

134.9MeV MeV, which means th<strong>at</strong> the isospin symmetry, described be<strong>low</strong>, is<br />

slightly broken.<br />

3.6 Linear (chiral) Sigma-model<br />

With a few modific<strong>at</strong>ions the above Lagrangean (80) becomes one of the most<br />

instructive of all field theoretical models for the nucleon and for the general<br />

concept of spontaneously broken symmetry. We note the Gell-Mann–Levy lagrangean,<br />

where we add a scalar field σ and remove the bare mass<br />

L = ψ [iγ µ ∂ µ ]ψ + 1 2<br />

(<br />

∂λ π a ∂ λ π a + ∂ λ σ∂ λ σ ) (85)<br />

35


−gψ(σ + iτ a π a γ 5 )ψ − µ2<br />

2 (σ2 + π a π a ) − λ 4 (σ2 + π a π a ) 2<br />

We shall see l<strong>at</strong>er th<strong>at</strong> this model exhibits the famous phenomenon of spontaneous<br />

symmetry breaking which is extremely important and interesting, since<br />

it is shown by the <strong>QCD</strong> as well. When it was invented it was first used to describe<br />

a nucleon, interpreting the ψ as a doublet of proton and neutron field as<br />

we did in the previous section. The nucleon was in the end a sysstem of interacting<br />

fields, where the bare nucleon field given by ψ was ”dressed” by the sigma<br />

and pion field. The result was moder<strong>at</strong>e since the mass of the nucleon was too<br />

small by a factor of two. Around 1985 the Gell-Man-Levy eq.(85) lagrangean<br />

was used by interpreting the ψ as a quark field doublet<br />

( ) u(¯r,t)<br />

ψ(¯r,t) =<br />

(86)<br />

d(¯r,t)<br />

In this form it was very successfull in the description of nucleon properties.<br />

Today the Lagrangean in appreci<strong>at</strong>ed since it has basic properties of <strong>QCD</strong>, i.e.<br />

it is completely field theoretical, it shows spontaneous chiral symmetry breaking<br />

(see l<strong>at</strong>er) and isovector symmetry (see right now).<br />

In tre<strong>at</strong>ing this Lagrangean and in investig<strong>at</strong>ing its symmetry properties it<br />

is useful to rewrite the mesons in terms of a mtrix field<br />

Σ(x) = σ(x) + iτ a π a (x)<br />

such th<strong>at</strong><br />

σ 2 + π a π a = 1 2 Tr(Σ† Σ)<br />

where we have used the identity (82).<br />

Then we obtain by simple and direct calcul<strong>at</strong>ion after separ<strong>at</strong>ing into rightand<br />

left-hand fermions for the Gelll-Man-Levy lagrangean (85) :<br />

L = ψ R [iγ µ ∂ µ ]ψ R + ψ L [iγ µ ∂ µ ]ψ L + 1 4 Tr(∂ µΣ∂ µ Σ † )<br />

− 1 4 µ2 Tr(ΣΣ † ) − λ [<br />

Tr(ΣΣ † ) ] 2 (<br />

− g ψL Σψ R + ψ R Σ † )<br />

ψ L<br />

16<br />

The lagrangean (85), written in this way, shows directly its invariance, as we<br />

see in the fol<strong>low</strong>ing: The left-handed and right-handed fermions are coupled to<br />

each other only in the interaction with the Σ-field, i.e the last tern on the rhs.<br />

The full lagrangean<br />

⊗<br />

has separ<strong>at</strong>e left- and right-invariances, i.e. it is invariant<br />

under SU(2) L SU(2)R :<br />

ψ R,L (x) → ψ ′ R,L(x) = U R,L ψ R,L (x)<br />

Σ ′ = U L ΣU † R<br />

with U R,L being two arbitrary SU(2) m<strong>at</strong>rices analogous to (81).<br />

U R,L (α) = exp(−i τa αR,L<br />

a )<br />

2<br />

36


For the fermions the SU(2) L<br />

⊗ SU(2)R transform<strong>at</strong>ion involves just a separ<strong>at</strong>e<br />

rot<strong>at</strong>ion of the left- and right-handed fields. For the mesons we have a<br />

combin<strong>at</strong>ion of a pure isospin rot<strong>at</strong>ion among the pionic fields together with a<br />

transform<strong>at</strong>ion between the σ and π fields. These invariances have the fol<strong>low</strong>ing<br />

conserved Noether-currents:<br />

J a Lµ = ψ L γ µ<br />

τ a<br />

2 ψ L − i 8 Tr ( τ a ( Σ∂ µ Σ † − ∂ µ ΣΣ †))<br />

JRµ a τ a<br />

= ψ R γ µ<br />

2 ψ R − i 8 Tr ( τ a ( Σ † ∂ µ Σ − ∂ µ Σ † Σ ))<br />

These can be rewritten into a conserved vector current and axial vector current:<br />

V a µ = J a Rµ + J a Lµ = ψγ µ<br />

τ a<br />

2 ψ + ɛabc π b ∂ µ π c (87)<br />

A a µ = JRµ a − JLµ a τ a<br />

= ψγ µ γ 5<br />

2 ψ + πa ∂ µ σ − σ∂ µ π a (88)<br />

In the infinitesimal form the corresponding symmetry transform<strong>at</strong>ions are:<br />

Isovector with β a = αR a + αa L :<br />

ψ → ψ ′ = (1 − iβ a τa<br />

)ψ (89)<br />

2<br />

σ → σ ′ = σ<br />

and axial with β a = α a R − αa L :<br />

π a → π ′a = π a + ɛ abc β b π c<br />

ψ → ψ ′ = (1 − iβ a τa<br />

2 γ 5)ψ (90)<br />

σ → σ ′ = σ + β a π a<br />

π a → π ′a = π a − β a σ<br />

As one sees, the vector current is identical to the isospin current (84) in the<br />

simple effective model (80), it contains only the fermion- and pion-fields. The<br />

axial current, however, is composed of the fermion-, and pion- and the sigmafields.<br />

One sees th<strong>at</strong> the axial transform<strong>at</strong>ion mixes σ- and π-field, whereas the<br />

(iso-)vector transform<strong>at</strong>ion does not mix thes two different types of fields.<br />

There are (without proof) closed charge algebras and current algebras for<br />

the left-hand and right-hand current analogous to eqs.(68) separ<strong>at</strong>ely (with<br />

f abc = ε abc ):<br />

37


[<br />

Q<br />

a<br />

R (t),Q b R(t) ] = if abc Q c R(t) (91)<br />

[<br />

Q<br />

a<br />

L (t),Q b L(t) ] = if abc Q c L(t)<br />

[<br />

Q<br />

a<br />

L (t),Q b R(t) ] = 0<br />

One can now write down the algebras of the vector and axial currents, and<br />

these expressions remind us of the expressions we had in the general chapter on<br />

Noether currents69,70:<br />

[<br />

Q<br />

V<br />

a (t),V µ<br />

b (xt)] = if abc V c µ (x) (92)<br />

[<br />

Q<br />

V<br />

a (t),A µ b (xt)] = if abc A µ c (x)<br />

[<br />

Q<br />

A<br />

a (x),A µ b (y)] = if abc V µ<br />

c (x)<br />

and<br />

[<br />

Q<br />

A<br />

a (x),V µ<br />

b (y)] = −if abc A µ c (x)<br />

δ(x 0 − y 0 ) [ V 0<br />

a (x),V 0<br />

b (y) ] = iδ(x − y)f abc V 0<br />

c (x) (93)<br />

δ(x 0 − y 0 ) [ V 0<br />

a (x),A 0 b(y) ] = iδ(x − y)f abc A 0 c(x)<br />

δ(x 0 − y 0 ) [ A 0 a(x),A 0 b(y) ] = iδ(x − y)f abc V 0<br />

c (x)<br />

These rel<strong>at</strong>ions are obtained directly by explicit calcul<strong>at</strong>ions (without proof).<br />

They hold, even if there are finite quark masses, because the definition of the<br />

currents involves only the kinetic energy of the quarks and the transform<strong>at</strong>ion<br />

done on the system. Of course, in such a case the charges are time dependent.<br />

The left hand and right hand currents form each a closed algebra, but the linear<br />

combin<strong>at</strong>ion does not: In fact, the vector currents provide a closed algebra,<br />

whereas the axial currents do not.One should note th<strong>at</strong> these rel<strong>at</strong>ions hold<br />

even if there is no pion field in the lagrangean, i.e. for a lagrangen where several<br />

fermion fields are ordered in a multiplet and being free or coupled to a gauge<br />

field. This simple fe<strong>at</strong>ure will be the case in <strong>QCD</strong>, where the quark fields are<br />

coupled to a flavour blind gluon field.<br />

38


4 Symmetry breaking<br />

4.1 Realiz<strong>at</strong>ion of symmetries<br />

If the Lagrangean shows a symmetry, then there are various ways this symmetry<br />

manifests itself. This is by no means trivial. We have actually the fol<strong>low</strong>ing<br />

possibilities:<br />

1. The symmetry is exactly and always fulfilled, on the classical as well as<br />

on the quantum level. Th<strong>at</strong> means, any solution of the Lagrangean will<br />

show it. This is e.g the case for electromagnetic U(1)-symmetry in QED<br />

and <strong>QCD</strong>, and also charge symmetry, flavour symmetry, baryon number<br />

symmetry in <strong>QCD</strong>.<br />

2. The symmetry may have an anomaly. This means, th<strong>at</strong> it is not a true<br />

symmetry, because it is only a symmetry of the classical lagrangean but<br />

not of the quantized lagrangean. It means th<strong>at</strong> the quantized symmetry<br />

current is not conserved, or better to say, the symmetry is not a symmetry<br />

of the quantized theory. This is actually not unsusual, so the name<br />

“anomaly” is misleading. However, as a m<strong>at</strong>ter of fact and based on<br />

fundamental field theoretical properties (and without proof), symmetry<br />

currents directly associ<strong>at</strong>ed to the gauge transform<strong>at</strong>ion must not have an<br />

anomaly.<br />

3. The symmetry may be explicitely broken. Example is the chiral symmetry<br />

of a massless Dirac-equ<strong>at</strong>ion, which is explicitely broken by small<br />

massterm in the lagrangean. Isospin symmetry is broken by mass differences<br />

of the upper and <strong>low</strong>er component of the field e.g. for proton-neutron<br />

doublets, or triplets of up-, down- and strange quarks, or by electromagnetic<br />

forces (charges). In real life this explicit breaking very often happens,<br />

but - depending on the process or structure considered - can often can be<br />

ignored if it is small.<br />

4. The symmetry may be ”hidden”, as one names this phenomenon. This<br />

means, the symmetry is an invariance of the quantized Lagrangean but<br />

not of the ground st<strong>at</strong>e of the system. Thus one does not recognize it<br />

by investig<strong>at</strong>ing the properties of the ground st<strong>at</strong>e or some of the excited<br />

st<strong>at</strong>es of the system. Such a fe<strong>at</strong>ure is also called “spontaneously broken<br />

symmetry”. Such a phenenon is only possible in infinite systems. Actually<br />

there are different physical mechanisms, which produce the spontaneous<br />

symmetry breaking<br />

(a) If in the lagrangean a scalar field occurs and if it acquires a nonvanishing<br />

vacuum expect<strong>at</strong>ion value (i.e. it is not dependent on<br />

space and time), then we have a spontaneously broken symmetry.<br />

An example is the Higgs-field in the electro-weak theory, whose nonvanishing<br />

vacuum value gives rise to finite values of the fermion and<br />

39


gauge boson masses due to the meson-fermion coupling term. Another<br />

example is the sigma-field in the linear chiral soliton model<br />

(85), which gives the fermions a dynamical mass..<br />

(b) Even in the absence of scalar fields, quantum effects can lead to the<br />

dynamical breaking of a symmetry. This is the case of the spontaneously<br />

broken chiral symmetry in <strong>QCD</strong>, one of the most important<br />

symmetries in the universe.<br />

4.2 Principle of Spontaneous symmetry breaking<br />

We consider a system with the Lagrangean (1) involving one field φ. Suppose<br />

the Lagrangean is invariant under a simple continuous symmetry which has a<br />

conserved Noether current as consequence. This means, if Q is the charge of<br />

the symmetry as obtained from the Noether-current, the Q can be used as the<br />

gener<strong>at</strong>or of the symmetry group.<br />

Nonbroken Symmetry: In the usual case of a non-broken field theory<br />

the ground st<strong>at</strong>e of the system |0〉 shows the symmetries of the Lagrangean and<br />

hence the ground st<strong>at</strong>e is invariant under<br />

|0〉 → exp(−iαQ) |0〉<br />

where α is a continuous parameter. Thus for such a (normal) sytem we have<br />

or<br />

exp(−iαQ) |0〉 = |0〉<br />

Q |0〉 = 0 (94)<br />

Obvioulsly we have only one ground st<strong>at</strong>e, and applic<strong>at</strong>ion of the symmetry<br />

oper<strong>at</strong>ion does not change it.<br />

Spontaneously broken symmetry: We have by definition spontaneous<br />

symmetry breaking, if under the above transform<strong>at</strong>ion the vacuum changes and<br />

or<br />

exp(−iαQ) |0〉 ≠ |0〉<br />

Q |0〉 ≠ 0 (95)<br />

Such a phenomenon does not occur in finite systems. One needs infinite systems<br />

in space-time and a field theory with its infinite number of particles and<br />

antiparticles provides this. In this case we have<br />

exp(−iαQ) |0〉 = |α〉 :::::::::::: |α〉 ≠ |0〉 (96)<br />

Examples for this are the ferromagnetism in solid st<strong>at</strong>es. The hamiltonian of<br />

the <strong>at</strong>om-<strong>at</strong>om interaction is rot<strong>at</strong>ionally smmetric and hence the hamiltonian<br />

for the many body system as well. Nevertheless we have in ferromagnets an<br />

orient<strong>at</strong>ion of the magnetic moments inside the Weiss-regions. If we rot<strong>at</strong>e all<br />

40


spins of a Weiss-region we get a set of degener<strong>at</strong>e st<strong>at</strong>es. Another example is<br />

the gravitaional field of the earth. It is rot<strong>at</strong>ionally symmetric, nevertheless for<br />

a pen in the gravit<strong>at</strong>ional field and on the spherical surface of the earth the<br />

<strong>low</strong>est energy st<strong>at</strong>e is not the pen pointing upright but lying on the ground.<br />

And one obtains degener<strong>at</strong>e st<strong>at</strong>es, if one rot<strong>at</strong>es the pen on the surface and<br />

changes by this its orient<strong>at</strong>ion. Although the st<strong>at</strong>es |α〉 of eq.(96) for different<br />

α are different, their energy is the same. This one can see easily: We have from<br />

Noethers theorem<br />

[H,Q] = −i dQ<br />

dt<br />

Since, however, Q is a charge oper<strong>at</strong>or of a symmetry current, and it is hence<br />

independent on time. Therfore the RHS is zere and hence it commutes with the<br />

Hamiltonian. Thus the energy of the st<strong>at</strong>e |α〉 is the one of the ground st<strong>at</strong>e<br />

|0〉:<br />

H |α〉 = H exp(−iαQ) |0〉 = exp(−iαQ)H |0〉 = exp(−iαQ)E 0 |0〉 = E 0 |α〉<br />

Because the symmetry transform<strong>at</strong>ion is continuous there must occur a continuous<br />

family of degener<strong>at</strong>e st<strong>at</strong>es |α〉.<br />

Actually there is an important consequence of the above consider<strong>at</strong>ions.<br />

Assertion: There are m<strong>at</strong>rixelements of the current, which connect the vacuum<br />

with one particular excited st<strong>at</strong>e of the system.This is a particular excit<strong>at</strong>ion of<br />

the ground st<strong>at</strong>e (vacuum), namely the so called Goldstone Boson. The logic is:<br />

dQ<br />

dt = 0 ⇒ d<br />

〈0|[Q(t),φ(0)] |0〉 = 0<br />

dt<br />

⇒ 〈0|[Q(t),φ(0)] |0〉 = η ≠ 0 (97)<br />

If we have no spontaneous symmetry breaking the η = 0, since according to<br />

eq.(94) we have Q |0〉 = 0. With spontaneous symmetry breaking we have η ≠ 0<br />

sincewith eq.(95) we have Q |0〉 ≠ 0 and η is time independent since Q(t) is time<br />

independent. This simple fe<strong>at</strong>ure eq.(97) has the immedi<strong>at</strong>e consequence th<strong>at</strong><br />

there exists a Goldstone boson. In order to see this, we represent the charge<br />

by an integral over the current and use the transl<strong>at</strong>ion oper<strong>at</strong>ion and insert a<br />

complete set of st<strong>at</strong>es |n〉, i.e. ground st<strong>at</strong>e of the system and excited ones. This<br />

looks in detail as fol<strong>low</strong>s<br />

η = 〈0|[Q(t),φ(0)] |0〉 =<br />

=<br />

∫<br />

d 3 x 〈0|[j 0 (x,t),φ(0)] |0〉 =<br />

= ∑ n<br />

∫<br />

d 3 x{〈0| j 0 (x,t) |n〉 〈n| φ(0) |0〉 − 〈0|φ(0) |n〉 〈n| j 0 (x,t) |0〉}<br />

41


The |n〉 are physical st<strong>at</strong>es, which can in principle be observed. They have the<br />

quantum numbers, which can be cre<strong>at</strong>ed by applying the charge oper<strong>at</strong>or to the<br />

vacuum. Now we use th<strong>at</strong> F(x) = exp(iPx)F(0)exp(−iPx) shown in eq.(34)<br />

and obtain<br />

η = ∑ n<br />

∫<br />

d 3 x {〈0|exp(iPx)j 0 (x = 0,t = 0)exp(−iPx) |n〉 〈n| φ(0) |0〉 −<br />

− 〈0| φ(0) |n〉 〈n|exp(iPx)j 0 (x = 0,t = 0)exp(−iPx) |0〉 }<br />

We have 〈0|exp(iPx) = 〈0|exp(iE 0 t − ip 0 x) with E 0 = 0 and p 0 = 0 because<br />

the vacuum is time independent and homogeneous. We have analogue<br />

expressions for the applic<strong>at</strong>ion of the shift oper<strong>at</strong>or on |n〉 .This yields then<br />

after integr<strong>at</strong>ion over d 3 x<br />

η = ∑ n<br />

(2π) 3 { δ (3) (p n ) 〈0|j 0 (0) |n〉 〈n| φ(0) |0〉 exp(−iE n t)−<br />

−δ (3) (−p n ) 〈0|φ(0) |n〉 〈n|j 0 (0) |0〉 exp(iE n t) }<br />

In order th<strong>at</strong> the right hand side is non-vanishing and time-independent we<br />

must have the fol<strong>low</strong>ing situ<strong>at</strong>ion:<br />

1) consider only st<strong>at</strong>es with p n = 0.<br />

2) Consider the vacuum st<strong>at</strong>e |n〉 = |0〉 with E 0 = 0. This yields a term,<br />

which does NOT contribute the the above sum for η, since this term is obviously<br />

zero. Thus in order to get something unequal zero <strong>at</strong> least one of the other st<strong>at</strong>es<br />

|n〉 must have a non-vanishing m<strong>at</strong>rix element.<br />

3) For an arbitrary st<strong>at</strong>e with E n ≠ 0 the time-exponentials are out of phase<br />

and hence yield a nonvanishing contribution, which is time dependent in contrast<br />

to η which is time independent. Thus all those “normal” st<strong>at</strong>es do not contribute<br />

in the sum, th<strong>at</strong> is their m<strong>at</strong>rix elements must be zero: 〈0| j 0 (0) |n〉 〈n| φ(0) |0〉 =<br />

0. There must be one particular st<strong>at</strong>e, call it |n = G〉, which must have the<br />

property:<br />

E G = 0<br />

〈0| j 0 (0) |G〉 ≠ 0 ::: and :::: 〈0|φ(0) |G〉 ≠ 0 (98)<br />

because then its contribution to η is not time dependent and unequal zero. This<br />

particular st<strong>at</strong>e is the st<strong>at</strong>e of the Goldstone Boson. It is an excit<strong>at</strong>ion of the<br />

vacuum and hence belongs to the set of st<strong>at</strong>es |n〉. Thus the current and the field<br />

connects the vacuum with the Goldstone boson |G〉. This is a very prominent<br />

fe<strong>at</strong>ure, which will l<strong>at</strong>er on in <strong>QCD</strong> give rise to concepts like conserved axial<br />

current, partially conserved axial current, pion decay constant, etc.<br />

Wh<strong>at</strong> does this mean physically In a quantum field theory the ground<br />

st<strong>at</strong>e of the system is the <strong>low</strong>est st<strong>at</strong>e. For e.g. baryon number zero it is the<br />

vacuum, for baryon number one it is the proton. In a quantum field theory<br />

42


any excit<strong>at</strong>ion of the ground st<strong>at</strong>e becomes quantized and describes in some<br />

way a particle. The minimum excit<strong>at</strong>ion energy corresponds to the particles<br />

mass. Thus the above excit<strong>at</strong>ions, having no excit<strong>at</strong>ion energy, correspond<br />

to the cre<strong>at</strong>ion of a particle with vanishing mass. These particles carry the<br />

quantum number of the symmetry which gener<strong>at</strong>es them (via the field or via<br />

the current), and they are called Goldstone Bosons. If the symmetry is not<br />

exact but slightly broken, then the Goldstone bosons acquire a small mass. As<br />

we will see, pions are the Goldstone bosons of the spontaneously broken chiral<br />

SU(2)-symmetry in <strong>QCD</strong>. Their mass should hence be zero, but in reality it<br />

is m π = 139 : MeV because m u = 6 : MeV and m d = 10 : MeV . For<br />

the spontaneously broken SU(3)-symmetry, involving also strange quarks, the<br />

kaons should also be Goldstone bosons, their mass is already m K = 493 : MeV<br />

due to the fact th<strong>at</strong> m s = 180 : MeV .<br />

Spontaneous symmetry breaking is a very general phenomenon. We will<br />

illustr<strong>at</strong>e it <strong>at</strong> one particular model, namely the chiral linear sigma model (85).<br />

4.3 Spontaneous symmetry breaking in the linear sigma<br />

model<br />

The lagrangean of the linear sigma model is given by eq.(85). It is an excellent<br />

example to illustr<strong>at</strong>e the spontaneous breakdown of thea chiral symmetry, as<br />

it in fact happens in <strong>QCD</strong>. Its symmetries and invariances (vector and axial<br />

vector) are known from previous sections an will play a prominent role. It<br />

seems th<strong>at</strong> there are mass terms for σ and π in the Lagrangean. However, the<br />

quartic term also gives quadr<strong>at</strong>ic terms in the boson fields and hence the mass is<br />

given as second deriv<strong>at</strong>ive of the bosonic potential, which collects all the bosonic<br />

terms which are neither coupled to a fermionic field nor have a deriv<strong>at</strong>ive. In<br />

the fol<strong>low</strong>ing we want to interpret the Lagrangean in a classical way. This is<br />

not really possible because the ψis a Grassmann variable and only functional<br />

integrals over it make sense. Nevertheless if we want to investig<strong>at</strong>e the properties<br />

of the vacuum one can first consider the properties of the bosonic part of the<br />

lagrangean using ψ = 0 for the vacuum and after having investig<strong>at</strong>ed th<strong>at</strong>, add<br />

the fermion. This procedure has its reason in the fact th<strong>at</strong> the bosonic field can<br />

be as large as you want whereas the fermionic one is basically restricted to one<br />

due to Pauli principle.<br />

We can bring the Lagrangean into a slightly modified but equivalent form<br />

L = ψ [iγ µ ∂ µ ]ψ + 1 2<br />

(<br />

∂λ π a ∂ λ π a + ∂ λ σ∂ λ σ ) ::::::::::::::::::::<br />

:::::::::::::::::::: −gψ(σ + iτ a π a γ 5 )ψ − C 2 (σ 2 + π a π a − A) 2<br />

To simplify the fol<strong>low</strong>ing consider<strong>at</strong>ions we consider classical limit of this lagrangean.<br />

We want to determine the ground st<strong>at</strong>e of the system, for this we<br />

need the hamilton function and its minimum. We obtain the hamilton function<br />

by considering the energy momentum tensor of the mesonic part, see eq.(64) for<br />

43


the definition. The fermionic part is ignored since we consider the vacuum:<br />

T 00 = −g 00 +<br />

= − 1 2<br />

= 1 2<br />

∂L<br />

∂ (∂ 0 π a ) ∂ 0π a +<br />

∂L<br />

∂ (∂ 0 σ) ∂ 0σ<br />

[<br />

(∂ µ σ) 2 + (∂ µ<br />

−→ π )<br />

2 − C 2 (σ 2 + π a π a − A) 2] + (∂ 0 σ) 2 + (∂ 0<br />

−→ π )<br />

2<br />

( ∂σ<br />

∂t<br />

) 2<br />

+ 1 2<br />

( ∂<br />

−→ ) 2 π<br />

+ 1 ( −→∇<br />

)<br />

−→ 2 1<br />

( −→∇σ<br />

) 2<br />

π + + C 2 (σ 2 + −→ π 2 − A) 2<br />

∂t 2 2<br />

Determining the minimum of the energy density means determining in a the energy<br />

per unit volume and minimize this. Thus we consider the bosonic potential<br />

energy of the system. This is<br />

V (σ, π a ) = C 2 (σ 2 + π a π a − A) 2<br />

and we look for the minimum of V (σ,π a ) with respect to σ and π a . In this<br />

simple picture masses of σ and π a are given by the second deriv<strong>at</strong>ive of V <strong>at</strong><br />

the vacuum values of σ and π. Obviously we have for the vacuum two different<br />

cases:<br />

.<br />

Case A < 0: Wigner Mode<br />

The minimum is <strong>at</strong> σ V = 0 and πV a = 0. And the masses are given by<br />

(<br />

1 ∂ 2 V<br />

2 m2 σ =<br />

∂σ<br />

)V<br />

2 = −2AC 2 > 0<br />

(<br />

1 ∂ 2 V<br />

2 m2 π =<br />

∂π<br />

)V<br />

2<br />

= −2AC 2 > 0<br />

Apparently σ and π are both massive and degener<strong>at</strong>ed. We have complete<br />

symmetry with respect to sigma and pi. This vacuum st<strong>at</strong>e is invariant under<br />

(iso-)vector transform<strong>at</strong>ion eq.(89):<br />

σ → σ ′ = σ → σ ′ V = σ V = 0<br />

π a → π ′a = π a + ɛ abc α b π c → (π a V ) ′ = π a V = 0<br />

and axial vector transform<strong>at</strong>ion eq.():<br />

σ → σ ′ = σ + α a π a → σ ′ V = σ V = 0<br />

π a → π ′a = π a − α a σ → (π a V ) ′ = π a V = 0<br />

Thus altogether the Wigner mode shows vanishing mesonic fields in the<br />

vacuum and this vacuum is invariant under the vector and axial vector transform<strong>at</strong>ion<br />

44


4.3.1 The Goldstone Mode<br />

This mode is given for A > 0. In such a case there is not one definite minimum<br />

but an infinite multitude of minima which all are only required to fulfill<br />

σ 2 + π a π a = A > 0<br />

All these minima are of equal right. If we demand positive parity from the<br />

vacuum we chose π a V = 0 and σ V = √ A. This vacuum is invariant under the<br />

isovector transform<strong>at</strong>ion eq.(89):<br />

σ → σ ′ = σ → σ ′ V = σ V<br />

π a → π ′a = π a + ɛ abc α b π c → (π a V ) ′ = π a V = 0<br />

However it is not invariant under the infinitesimal axial transform<strong>at</strong>ion eq.().<br />

In fact this consists in rot<strong>at</strong>ing along the chiral circle σ 2 + π a π a = A. For an<br />

infinitesimal transform<strong>at</strong>ion we have:<br />

σ → σ ′ = σ + α a π a → σ V → σ ′ V = σ V + α a π a V → σ ′ V = σ V<br />

π a → π ′a = π a −α a σ → π a V → π ′a<br />

V = π a V −α a σ V → (π a V ) ′ = −α a σ V ≠ π a V = 0<br />

The axial transform<strong>at</strong>ion changes all the possible vacua into each other, however<br />

for an infinitesimal transform<strong>at</strong>ion only the pion field gets an ifinitesimal<br />

contribution proportional to the vacuum value of the sigma field..<br />

We can now calcul<strong>at</strong>e the masses of the pion and sigma. They are given by<br />

the second deriv<strong>at</strong>ive of the potential, however not around the trivial vacuum<br />

but around the new vacuum. In order to do so we rewrite the Lagrangean by<br />

defining new fields, which describe the devi<strong>at</strong>ion from the vacuum field: Instead<br />

of σ we use now σ = ˜σ + σ V , such th<strong>at</strong> the vacuum value of ˜σ is zero. Doing<br />

this we obtain the same Lagrangean with the new fields as:<br />

L = ψ [iγ µ ∂ µ − gσ V ]ψ + 1 (<br />

∂λ˜σ∂ λ˜σ − 8C 2 σV 2 ˜σ 2) + 1 2<br />

2 ∂ λπ a ∂ λ π a<br />

− gψ(˜σ + iτ a π a γ 5 )ψ + 4C 2 σ V ˜σ(˜σ 2 + π a π a ) − C 2 (˜σ 2 + π a π a ) 2<br />

Again we can now calcul<strong>at</strong>e the masses of the pion and sigma:<br />

1<br />

2 m2 σ =<br />

( ∂ 2 V<br />

∂σ 2 )V<br />

= 4AC 2 = 4σ 2 V C 2 > 0<br />

(<br />

1 ∂ 2 V<br />

2 m2 π =<br />

∂π<br />

)V<br />

2 = 0<br />

Apparently there is no mass term for the pion. The pion appears as massless<br />

Goldstone boson and the mass-squared of the sigma has increased twice.<br />

45


There is also an interesting consequence for the fermions. In the original<br />

Lagrangean the fermions were assumed to be massles. Thus their energy<br />

spectrum is given by ɛ(k) = ± √ k 2 After transforming to the new vacuum the<br />

fermions appear as massive particles with a mass of m = gσ V . This process is<br />

called “dynamical mass gener<strong>at</strong>ion”. Thus their energy spectrum is given by<br />

ɛ(k) = ± √ k 2 + m 2 and hence there appears a gap of 2m between the occupied<br />

and unoccupied single particle st<strong>at</strong>es.<br />

ε (k)<br />

single quark <strong>energies</strong><br />

Wigner mode<br />

ε (k)<br />

M<br />

Single quark <strong>energies</strong>: Goldstone mode<br />

The Lagrangean looks actually quite different compared to the original one.<br />

However the smmetries are unchanged and the Lagrangean is still invariant<br />

under SU(2) L<br />

⊗ SU(2)R .On the other hand the σand πdo not form a degener<strong>at</strong>e<br />

multiplet. Thus the symmetry is ”hidden”, or spontaneously broken. Also, the<br />

full set of original symmetry currents remain conserved. The vector current 87is<br />

unchanged. The axial current, however, looks different. It was given by eq.(88)<br />

and is now is now<br />

A a µ = ψγ µ γ 5<br />

τ a<br />

2 ψ + πa ∂ µ σ − ˜σ∂ µ π a − σ V ∂ µ π a<br />

and, however, still has a vanishing divergence ∂ µ A a µ = 0. Thus the fact th<strong>at</strong> the<br />

fermion field got a mass due to canamical mass gener<strong>at</strong>ion has not change the<br />

divergence of the currents.<br />

[A comment for the advanced reder: The above procedure is in principle<br />

much more complex: If we want to solve this Lagrangean we have in<br />

priciple to perform functional integrals to obtain the partition function, from<br />

which one can derive basically everything. The partition function is given by<br />

Z = ∫ DψDψDσDπ exp [ i ∫ d 4 xL(x) ] .First we perform the integral over the<br />

grassmann variables. By this you integr<strong>at</strong>e out the fermionic fields yielding a<br />

preexponential factor, i.e. the fermion determinnt=Trlog. For the fol<strong>low</strong>ing<br />

consider<strong>at</strong>ions it is helpful to consider the socalled small coupling limit. To perform<br />

this limit we change the variables in the functional integral by introducing<br />

σ → √ σ λ<br />

and π a → √ πa<br />

λ<br />

. This does not affect the preexponential factor if we<br />

assume th<strong>at</strong> the fermion-boson coupling constant scales with √ λ.The st<strong>at</strong>ionary<br />

phase approximtion is now given by the minimum of the bosonic action. After<br />

46


having obtained the bosonic field in the minimum we insert this into the Trlog.<br />

The st<strong>at</strong>ionary phase approxim<strong>at</strong>ion is justified in the limit λ → 0 because then<br />

the bosonic terms domin<strong>at</strong>e.]<br />

4.3.2 Broken chiral symmetry and PCAC<br />

In the fol<strong>low</strong>ing section we move with the chiral linear sigma model a bit coloser<br />

to reality, since in experiment the pions have a small but finite mass of 139<br />

MeV. This will lead to the principle of partial conserv<strong>at</strong>ion of the axial current,<br />

which is very imprtant. In order to simul<strong>at</strong>e this <strong>QCD</strong>-fe<strong>at</strong>ure in the Gell-man<br />

Levy model (85)it is sufficinet to add to the Lagrangean a term<br />

L → L+L m = L + Dσ<br />

With this term the vector current is still conserved, however the divergence of<br />

the axial current does not vanish any more, as we see immedi<strong>at</strong>ely. In order<br />

to calcul<strong>at</strong>e the axial current we consider the infinitesimal axial transform<strong>at</strong>ion<br />

(90)<br />

σ → σ ′ = σ + ε a π a<br />

π a → π ′a = π a − ε a σ<br />

Therefore we have for the mesonic part δL m = Dπ a ε a (x) and hence with<br />

eq.(62) and (63) one obtains immedi<strong>at</strong>ely<br />

∂ µ j µa<br />

5 =<br />

∂<br />

∂(ε a (x)) L m =<br />

∂<br />

∂(ε a (x)) Dπb ε b (x) = Dπ a<br />

This equ<strong>at</strong>ion holds even if we consider the full Lagranean including the Ferionic<br />

fields. Since the divergence of the axial current is not zero, the chiral symmetry<br />

is ecplicitely broken. This has immedi<strong>at</strong>e effect on the classical ground st<strong>at</strong>e of<br />

the theory as we can see <strong>at</strong> the bosonic potential<br />

We obtain the minimum of U by<br />

U = C 2 ( σ 2 + −→ π 2 − A ) 2<br />

− Dσ<br />

∂U<br />

∂σ = 4C2 σ ( σ 2 + −→ π 2 − A ) − D = 0 (99)<br />

∂U<br />

∂π a = 4C2 π a ( σ 2 + −→ π 2 − A ) = 0<br />

Again we denote the ground st<strong>at</strong>e values with the index V: σ V and π V . The<br />

Wigner-mode is trivially changed. More interesting is the Goldstone mode with<br />

A > 0. There we obtain from eq.(99)<br />

σ 2 V − A =<br />

47<br />

D<br />

4C 2 σ V<br />

(100)


If we again expand the sigma field around the vacuum value σ V we obtain<br />

again th<strong>at</strong> the fermions get a mass of M = gσ V . The mesonic poten<strong>at</strong>ial is now<br />

changed into by elimin<strong>at</strong>ing A by eq.(100)<br />

U = C 2 ( σ 2 + −→ π 2 − A ) 2<br />

− Dσ<br />

= C 2 ( (˜σ + σ V ) 2 + −→ π 2 − A ) 2<br />

− D(˜σ + σV )<br />

(<br />

= C 2 ˜σ 2 + −→ π 2 +<br />

D ) 2<br />

4C 2 + 2˜σσ V − D(˜σ + σ V )<br />

σ V<br />

We find immedi<strong>at</strong>ely by taking the deriv<strong>at</strong>ive w.r. to π 2 th<strong>at</strong> the mass or the<br />

pion is now finite:<br />

m 2 π = D σ V<br />

Using this equ<strong>at</strong>ion and M = gσ V we can express the yet unknown value of D<br />

in the fol<strong>low</strong>ing way<br />

D = m2 πM<br />

g<br />

This rel<strong>at</strong>ion holds only in the Goldstone phase. Now we are able to rewrite the<br />

divergence of the current:<br />

∂ µ j µa<br />

5 = Dπ a = m2 πM<br />

π a (101)<br />

g<br />

Apparently the current is almost conserved because the pion mass is small and<br />

in practice the M = 350MeV and g = 5 roughly. Still we have to get to know<br />

the RHS better.<br />

The objective of the fol<strong>low</strong>ing lines is, to express the right hand side of<br />

eq.(101) in terms of the pion decay constant, since this is an important and<br />

well known quantity f π = 93MeV . For this consider the m<strong>at</strong>rixelement for the<br />

pion decay (detailled reasoning is given in sect on pion decay), i.e. in the view<br />

of strong interaction from a 1-pion st<strong>at</strong>e to the vacuum. The relevant process<br />

for this is for instance π − → µ − + ν µ .Since the pion has neg<strong>at</strong>ive parity only<br />

the axial current contributes and we have hence the definition of the pion decay<br />

constant<br />

< 0|j µa<br />

5 (0)|πb (k) >= ik µ δ ab f π<br />

This is equ<strong>at</strong>ion is the concretiz<strong>at</strong>ion of the more abstract expression (98).<br />

Using the shift oper<strong>at</strong>or eq.(34) one can write<br />

< 0|exp(−iPx)j µa<br />

5 (x)exp(+iPx)|πb (k) >= ik µ δ ab f π<br />

or<br />

< 0|j µa<br />

5 (x)|πb (k) >= ik µ δ ab f π exp(−ikx)<br />

and one gets for the divergence because of k 2 = m 2 π<br />

< 0|∂ µ j µa<br />

5 (x)|πb (k) >= m 2 πf π δ ab exp(−ikx)<br />

48


Using the expression for the quantized boson field (32) and the norm of the<br />

boson cre<strong>at</strong>ion oper<strong>at</strong>ors (31) one gets for the properly normalized one-pion<br />

st<strong>at</strong>e<br />

|π b (k) >= (2ω k ) 1 2<br />

a † (k)|0 > (102)<br />

so th<strong>at</strong><br />

Thus we obtain<br />

yielding<br />

< π a (k)|π b (k ′ ) >= 2ω k (2π) 3 δ(k − k ′ ) (103)<br />

< 0|π a (x) |π b (k) >= δ ab exp(−ikx)<br />

< 0|∂ µ j µa<br />

5 (x)|πb (k) >= m 2 πf π < 0|π a (x) |π b (k) ><br />

If we go back to the classical level we obtain<br />

∂ µ j µa<br />

5 (x) = m2 πf π π a (x) (104)<br />

This is an extremely important equ<strong>at</strong>ion, which expresses the principle of<br />

PCAC, i.e. partially conserved axial current.<br />

One can apply this in the fol<strong>low</strong>ing way. We had the divergence of the<br />

current as ∂ µ j µa<br />

5 = m2 π M<br />

g<br />

π a and ∂ µ j µa<br />

5 (x) = m2 πf π π a yielding M = gf π and<br />

hence<br />

σ V = f π<br />

and hence the vacuum value of the sigma-field equals the pion decay constant.<br />

If one takes M = gf π , as we have shown above, and interpretes the fermion<br />

field as the nucleon m<strong>at</strong>ter field, then the coupling constant g equals to the<br />

pion-nucleon coupling constant g πNN We have from experiment the numbers<br />

M N = 938MeV<br />

g πNN = 13.45<br />

f π = 93MeV<br />

which fits the equ<strong>at</strong>ion M = gf π within 30%. The fact th<strong>at</strong> such a simple effective<br />

model yields such an agreement with experiment is really an achievement.<br />

One should note: A small quantity as the pion decay constant, being rel<strong>at</strong>ed to<br />

the weak decay of the pion, is rel<strong>at</strong>ed via the pion-decay constant to a large<br />

quantity, i.e. the nucleon mass. This is an astonishing fe<strong>at</strong>ure.<br />

Actually one can improve this model by interpreting the fermion field as a<br />

field for quarks interacting with the pion field. In this way (without proof) one<br />

can incorpor<strong>at</strong>e the substructure of the nucleon. If one does it (see l<strong>at</strong>er) we<br />

obtain a modified equ<strong>at</strong>ion, the so called Goldberger-Treimann equ<strong>at</strong>ion:<br />

g A = 1.26<br />

M N g A = g πNN f π<br />

49


with the axial vector coupling constant of the nucleon with experimental value<br />

g A = 1.26. The Goldberger Treiman rel<strong>at</strong>ion is s<strong>at</strong>isfied <strong>at</strong> the 5% level. In<br />

this way one has a phenomenological description of the dynamic mass gener<strong>at</strong>ion<br />

caused by the spontaneous breaking or the chiral symmetry: The massless<br />

quarks (so called current quarks, or <strong>QCD</strong>-quarks) get via this breaking a mass<br />

of about 350 to 400 MeV, such th<strong>at</strong> three of them yield basically the mass of the<br />

nucleon. One sees th<strong>at</strong> without the breaking we would not have an explan<strong>at</strong>ion<br />

for the mass of the nucleon.<br />

5 The gauge principle<br />

The gauge principle has been established in the last 30 years. It is basically a<br />

recipe how to construct field theories, which describe the forces of n<strong>at</strong>ure. The<br />

principle has not been derived, but it has been proven extremely succesfull. We<br />

first consider for this the Maxwell theory, which is an abelian gauge theory, and<br />

then a Yang-Mills theory, which is a non-abelian gauge theory. The Quantum<br />

Chromodynamics, which is the theory for strong interaction is a non-abelian<br />

gauge theory, as well as the Glashow-Salam-Weinberg theory describing the<br />

electro-weak interaction. The gauging of a theory and the deriv<strong>at</strong>ion of a gauge<br />

field will be exemplified <strong>at</strong> the Dirac equ<strong>at</strong>ion.<br />

5.1 Abelian gauge theory:<br />

Global symmetry: Consider the lagrangean equ<strong>at</strong>ion (6) for the free electron.<br />

Apparently this Lagrangean is invariant under the “global U(1)-transform<strong>at</strong>ion”<br />

(74) Here α is a real constant. The symmetry is a trivial symmetry of the free<br />

Dirac Lagrangean, each free Dirac particle shows this. Actually the symmetry is<br />

an abelian symmetry since two successive symmetry transform<strong>at</strong>ions commute.<br />

Local symmetry: Now we want to generalize this symmetry into a “local<br />

U(1) symmetry” by demanding, th<strong>at</strong> the Lagrangean should be invariant under<br />

this symmetry even if α(x) is an arbitrary differentiable function. We will l<strong>at</strong>er<br />

see, th<strong>at</strong> this seemingly arbitrary demand is highly relevant and will gener<strong>at</strong>e<br />

in principle all present theories of elementary particles and their interaction:<br />

ψ(x) → ψ ′ (x) = U(α(x))ψ(x) = exp(−iα(x))ψ(x) (105)<br />

ψ(x) → ψ ′ (x) = ψ(x)U −1 (α(x)) = ψ(x) exp(+iα(x))<br />

This is a very strong demand, it says th<strong>at</strong> we can change <strong>at</strong> any space-time point<br />

the phase of the Dirac field in an arbitrary way and the Lagrangean, i.e. the<br />

dynamics of the Dirac field should be the same. This is a lot, in non-rel<strong>at</strong>ivistic<br />

quantum mechanics we can multiply the wave function only with a constant<br />

phase, never with a space- and time-dependent one. However, the demand is<br />

not stupid, in fact, this demand has appeared over the last 20 years as the basic<br />

50


tool to construct theories of elementary particles and their interactions. Electrodynamics,<br />

Quantum Chromodynamics and Elctroweak theory are constructed<br />

in this way.<br />

Immedi<strong>at</strong>ely we see th<strong>at</strong> the above Lagrangean does NOT fulfil the demand<br />

of a local U(1) symmetry. The mass term is indeed invariant, however the<br />

kinetic term is not since the deriv<strong>at</strong>ive acts on α(x):<br />

ψ(x)∂ µ ψ(x) → ψ ′ (x)∂ µ ψ ′ (x) = ψ(x) exp(+iα(x))∂ µ exp(−iα(x))ψ(x)<br />

= ψ(x)∂ µ ψ(x) − iψ(x)(∂ µ α(x))ψ(x)<br />

The second term on the RHS destroys the invariance of the Lagrangean. In<br />

order to fulfill our demand, we have to modify the Lagrangean. We do this by<br />

introducing a new field A µ (x) and the so-called gauge-covariant deriv<strong>at</strong>ive:<br />

D µ ψ(x) = (∂ µ + ieA µ (x))ψ(x) (106)<br />

The covariant deriv<strong>at</strong>ive contains the coupling constant e of the fermion field<br />

ψ(x) to the gauge field A µ (x). In the present U(1) theory this coupling constant<br />

e is the charge of the fermion. [The electron has the charge e = −e 0 , where e 0<br />

is the elementary charge.] We consider the new Lagrangean, where we have<br />

replaced ∂ µ by D µ :<br />

L = ψ [iγ µ D µ − m]ψ (107)<br />

and demand invariance of this Lagrangean (107) under the gauge transform<strong>at</strong>ion<br />

() of the fermion field. However, this demand cannot be fulfilled without<br />

demanding simultaneously the fol<strong>low</strong>ing transform<strong>at</strong>ion property of the A-field:<br />

A µ (x) → A ′ µ(x) = A µ (x) + i e (∂ µU(α))U −1 (α) = A µ (x) + 1 e ∂ µα(x) (108)<br />

Assertion: The Lagrangean (107) with the covariant deriv<strong>at</strong>ive (106) is invariant<br />

under simultaneous gauge transform<strong>at</strong>ion of the fermion field and the gauge field<br />

( 105,108).<br />

Proof: On can see immedi<strong>at</strong>ely in a direct calcul<strong>at</strong>ion th<strong>at</strong> we have under<br />

the new combined gauge transform<strong>at</strong>ion (108) the property<br />

D µ ψ(x) → [D µ ψ(x)] ′ = exp (−iα(x)) D µ ψ(x) (109)<br />

and hence the kinetic energy term is invariant:<br />

ψ(x)D µ ψ(x) → ψ ′ (x)[D µ ψ(x)] ′<br />

( [<br />

= ψ(x) exp(+iα(x)) ∂ µ + ie A µ (x) + 1 ])<br />

e ∂ µα(x) exp(−iα(x))ψ(x)<br />

51


= ψ(x) (∂ µ + ieA µ (x)) ψ(x) = ψ(x)D µ ψ(x) q.e.d.<br />

By now the A-field is an additional field with no dynamical properties. To<br />

make it a true dynamical variable we have to complement the Lagrandian density<br />

(107) by a term corresponding to the kinetic energy of the A-field. The<br />

additional term should not destroy the gauge invariance (108) and should contain<br />

deriv<strong>at</strong>ives in order to provide dynamics. A candid<strong>at</strong>e is the Lagrangean<br />

(20) of the free Maxwell field. We will show be<strong>low</strong> th<strong>at</strong> − 1 4 F µνF µν is indeed<br />

gauge invariant. Therefore we can note the full Lagrangean, being gauge invariant<br />

under U(1), for a fermion field of charge e coupled to a Maxwell field<br />

:<br />

L = ψ [iγ µ D µ − m]ψ − 1 4 F µνF µν (110)<br />

with the field tensor known from classical electrodynamics (22)<br />

F µν (x) = ∂ µ A ν (x) − ∂ ν A µ (x) (111)<br />

The proof of the gauge invariance of the kinetic energy of A is simple: By a<br />

direct calcul<strong>at</strong>ion one can easily show (first) th<strong>at</strong><br />

(D µ D ν − D ν D µ ) ψ = ieF µν ψ (112)<br />

The proof of eq.(112) is simple. [One should note for the deriv<strong>at</strong>ion th<strong>at</strong> ∂ µ acts<br />

on everything on the right, when it is situ<strong>at</strong>ed in D µ ]:<br />

(D µ D ν − D ν D µ ) ψ =<br />

= [(∂ µ + ieA µ (x)) (∂ ν + ieA ν (x)) − (∂ ν + ieA ν (x)) (∂ µ + ieA µ (x))]ψ<br />

= ie(∂ µ A ν (x) − ∂ ν A µ (x))ψ<br />

. Direct calcul<strong>at</strong>ion yields the above identity. Furthermore from eq.(109) we<br />

derive immedi<strong>at</strong>ely (second)<br />

[(D µ D ν − D ν D µ ) ψ] ′ = exp(−iα(x))(D µ D ν − D ν D µ ) ψ<br />

The proof of this equ<strong>at</strong>ion is simple because<br />

(D µ D ν ψ) ′ = D ′ µD ′ ν exp(−iα(x)ψ(x) = D ′ µ exp(−iα(x)D ν ψ(x)<br />

You can consider D ν ψ as some particular spinor and hence you get:<br />

D ′ µ exp(−iα(x)D ν ψ(x) = exp(−iα(x)D µ D ν ψ(x)<br />

which proofs the assertion. In summary we have<br />

F ′ µνψ ′ = exp(−iα(x))(F µν ψ)<br />

52


Because ψ is an arbitrary spinor and exp(−iα(x)) and F µν commutes since F µν is<br />

a function, we have finally the important gauge invariance of the field tensor<br />

F ′ µν = F µν qed (113)<br />

This fe<strong>at</strong>ure, remember our section on classical Maxwell field, is necessary. The<br />

F µν are directly rel<strong>at</strong>ed to the electric and magnetic fields E and B, which are<br />

observables, since they appear directly in the Lorentz force acting on a charged<br />

particle moving in these fields: K = q (vxB) .Du to the motion of the particle<br />

the electric and magnetic fields are observable quantities. The follwing remarks<br />

are in order:<br />

1. The above Lagrangean can be rewritten as<br />

with the current<br />

L = ψ [iγ µ ∂ µ − m]ψ − 1 4 F µνF µν − ej µ A µ (114)<br />

j µ = ψγ µ ψ (115)<br />

A comparison shows th<strong>at</strong> we have tacitly derived the Lagrangean of a<br />

Dirac field and a Maxwell-field, where the Dirac field provides the source<br />

(current) to gener<strong>at</strong>e the Maxwell field. If we put e = −e 0 we describe<br />

electrons.<br />

2. There is no mass term for the photon. It shoud have the form m phot A µ A µ ,<br />

which is obviously NOT gauge invariant. One sees this directly looking <strong>at</strong><br />

the gauge transform<strong>at</strong>ion property of the A-field eq .().<br />

3. In the covariant deriv<strong>at</strong>ive (106) we have a r<strong>at</strong>her simple coupling of the<br />

gauge field to the fermion field. This coupling is called “minimal coupling”.<br />

This coupling results from the wish to introduce a gauge field<br />

in the utmost simple way into the original lagrangean in order to make<br />

the lagrangean of the free fermion invariant under the U(1) gauge transform<strong>at</strong>ion.<br />

The coupling is therefore completely independent on further<br />

properties of the fermion field. Indeed any charged fermion field couples in<br />

the same way to the gauge field ( i.e. to the scalar Φ- and vector A- field<br />

of the Electrodynamics). This property is calles universality. Other<br />

higher dimensional gauge-invariant couplings such as ψs µν ψF µν are ruled<br />

out by the requirement of renormalizability of the resulting lagrangean.<br />

There is, however, one degree of freedom in the coupling, th<strong>at</strong> is the value<br />

of the charge. We could simply add to the lagrangean with the fermion<br />

field ψ another fermion field φ with a charge e ′ = be with an arbitrary<br />

b. As one sees in the next formulae has in fact one electromagnetic field<br />

coupled to the two different fermion fields<br />

4.<br />

L = ψ [iγ µ D µ − m]ψ + φ[iγ µ D µ − m]φ − 1 4 F µνF µν<br />

53


D µ ψ(x) = (∂ µ + ieA µ (x)) ψ(x)<br />

D µ φ(x) = (∂ µ + iebA µ (x))φ(x)<br />

ψ(x) → ψ ′ (x) = exp(−iα(x))ψ(x)<br />

φ(x) → φ ′ (x) = exp(−ibα(x))φ(x)<br />

A µ (x) → A ′ µ(x) = A µ (x) + i e (∂ µU(α))U −1 (α)<br />

= A µ (x) + i<br />

be (∂ µU(bα))U −1 (bα)<br />

= A µ (x) + 1 be ∂ µ[bα(x)] = A µ (x) + 1 e ∂ µα(x)<br />

and there would be exactly one electromagnetic field being gener<strong>at</strong>ed by<br />

both fermion fields. This will be different for non-abelian gauge fields.<br />

There the charges are not arbitrary.<br />

5. From the above Lagrangean (114) one derives immedi<strong>at</strong>ely the inhomogeneous<br />

Maxwell equ<strong>at</strong>ions (24). They are linear in the field tensor F µν or<br />

in the potential A µ . In the Maxwell-equ<strong>at</strong>ions the only coupling of A is<br />

to the current (115) in the form j µ A µ . They do not contain terms of the<br />

sort AA or AAA or AAAA. Those would correspond to a selfcoupling of<br />

the gauge field. This selfcoupling does not appear here and this is consistent<br />

with the fact th<strong>at</strong> the A-field is not charged. Thus a photon-photon<br />

interaction does not exist. It exists only in higher order quantum field<br />

theory.<br />

6. The fe<strong>at</strong>ures 1. to 3. occur in all field theories, abelian and non-abelian in<br />

one or the other form. The selfcoupling, however, is fundamentally different<br />

in non-abelian field theories: It really exists and makes the tre<strong>at</strong>ment<br />

and phenomenology of these theories often quite different.<br />

The coupling of the photon field A to the fermion field ψ can graphically be<br />

depicted as<br />

f<br />

γ<br />

54


5.2 Non-abelian gauge theory<br />

We assume the ψ-field to be an n-tupel<br />

⎛<br />

ψ(x) =<br />

⎜<br />

⎝<br />

ψ 1 (x)<br />

ψ 2 (x)<br />

...<br />

ψ n (x)<br />

The free field is described by the well known Lagrangean (6) and we assume the<br />

m to be a diagonal n · n-m<strong>at</strong>rix with identical masses as entries.<br />

Global symmetry: We consider now a SU(N)-transform<strong>at</strong>ion (51 ) and<br />

consider the represent<strong>at</strong>ion with dimension n i.e. the represent<strong>at</strong>ion of the<br />

gener<strong>at</strong>ors 1 2 τa are n · n-m<strong>at</strong>rices:<br />

⎞<br />

⎟<br />

⎠<br />

ψ(x) → ψ ′ (x) = exp(−i τa θ a<br />

)ψ(x) (116)<br />

2<br />

ψ(x) → ψ ′ (x) = ψ(x) exp(+i τa θ a<br />

2 )<br />

The 1 2 τa (a = 1,2,3...N 2 − 1) are n · n m<strong>at</strong>rices of the n-dimensional<br />

represent<strong>at</strong>ion of the SU(N)-group. In fact they are the represent<strong>at</strong>ions of<br />

the gener<strong>at</strong>ors of the SU(N). The θ a are a real constants, th<strong>at</strong> is why the<br />

transform<strong>at</strong>ion is called global, since the θ a are not dependent on space or time.<br />

The τ a s<strong>at</strong>isfy the commut<strong>at</strong>ion rules (48)<br />

[ ] τ<br />

a<br />

2 , τb τ c<br />

= if abc<br />

2 2<br />

As we know the structure coefficients f abc define uniquely the group SU(N).<br />

Due to the fact th<strong>at</strong> they do not commute with each other the SU(N)-Lie-group<br />

is non-abelian. This means th<strong>at</strong> two successive transform<strong>at</strong>ions give a different<br />

result when one interchanges the order. We denote<br />

and then we have<br />

U(θ) = exp(−i τa θ a<br />

2 )<br />

U(θ 1 )U(θ 2 )ψ ≠ U(θ 2 )U(θ 1 )ψ<br />

The invariance of the free Lagrangean under SU(N) is a remarkable one.<br />

It says th<strong>at</strong> one can start from the fields ψ k and is al<strong>low</strong>ed to perform in a<br />

well defined way linear combin<strong>at</strong>ions ot those fields to get new fields ψ ′ k and<br />

these new fields, combined again to a n-plet, obbey the same Lagrangean. This<br />

seemingly absurd thing is the SU(N)-symmetry of L 0 and it is really true.<br />

Local symmetry: We demand now th<strong>at</strong> the L 0 Lagrangean is locally<br />

invariant under SU(N), which means th<strong>at</strong> we demand invariance al<strong>low</strong>ing the<br />

55


θ a = θ a (x) to be differentiable functions depending on the space-time points<br />

x. This is a very strong demand, it says th<strong>at</strong> we can change <strong>at</strong> any space-time<br />

point the phase of the Dirac field independently of the phase <strong>at</strong> any other point<br />

and the Lagrangean, i.e. the dynamics of the Dirac field should be the same.<br />

We have<br />

U(θ) = exp(−i τa θ a (x)<br />

) (117)<br />

2<br />

The assertion, which we have to proof is the fol<strong>low</strong>ing: The Lagrangean L 0 =<br />

ψ [iγ µ ∂ µ − m]ψ as such is invariant under the global SU(N) transform<strong>at</strong>ion<br />

but NOT invariant under the local one. Similar to the abelian case we have<br />

to complement the Lagrangian by adding an additional field A µ (x). Since our<br />

invariance is now more complic<strong>at</strong>ed one gauge field is not enough and instead we<br />

have to consider A a µ(x) with a = 1,2,3...N 2 − 1. Apparently we have so many<br />

A a -fields, as there are gener<strong>at</strong>ors. For this we define the covariant deriv<strong>at</strong>ive<br />

analogously to eq.(106)<br />

(<br />

D µ ψ(x) = ∂ µ + ig τa A a )<br />

µ(x)<br />

ψ(x) (118)<br />

2<br />

and we demand for the field A a µ(x) something similar as in the non-abelian case<br />

eq.(109):<br />

D µ ψ(x) → [D µ ψ(x)] ′ = U(θ)[D µ ψ(x)] (119)<br />

or explicitely<br />

D µ ψ(x) → [D µ ψ(x)] ′ = exp(−i τa θ a (x)<br />

)[D µ ψ(x)] (120)<br />

2<br />

We damand this because (proof be<strong>low</strong>) then the more general Lagrangean<br />

is gauge invariant:<br />

L = ψ [iγ µ D µ − m]ψ<br />

This lagrangean is exactly the like the one from the abelian theory eq.(107)<br />

except th<strong>at</strong> now the ψ is a multiplet and the covarin<strong>at</strong> deriv<strong>at</strong>ive contains a<br />

more complic<strong>at</strong>ed field A a µ(x) ra<strong>at</strong>her than just A µ (x). The above demand is<br />

actually a demand for the behauviour of the gauge field A a µ(x) under gauge<br />

transform<strong>at</strong>ions, and it implies analogously to eq.(108):<br />

τ a A a µ(x)<br />

2<br />

→ τa A a µ(x) ′<br />

2<br />

( τ a A a µ(x)<br />

= U(θ(x))<br />

2<br />

Proof: The above demand implies<br />

(∂ µ + ig τa A a µ(x) ′<br />

2<br />

)<br />

(U(θ)ψ) = U(θ)<br />

or [<br />

(∂ µ U(θ)) + ig τa (A a µ(x)) ′<br />

2<br />

]<br />

U(θ)<br />

56<br />

)<br />

U −1 (θ(x))+ i g [∂ µU(θ(x))]U −1 (θ(x))<br />

(<br />

∂ µ + ig τa A a )<br />

µ(x)<br />

ψ<br />

2<br />

ψ = igU(θ) τa A a µ(x)<br />

ψ<br />

2<br />

(121)


or<br />

τ a (A a µ(x)) ′<br />

2<br />

= U(θ) τa A a µ(x)<br />

U −1 (θ) + i 2 g [∂ µU(θ(x))]U −1 (θ(x))<br />

qed<br />

Actually we have reached our goal somehow, since the Lagrangean L 0 = ψ [iγ µ D µ − m]ψ<br />

is now SU(N)-gauge invariant.<br />

We have now D ′ µψ ′ = UD µ ψ or D ′ µUψ = UD µ ψ from which fol<strong>low</strong>s, since<br />

ψ is arbitrary, th<strong>at</strong><br />

D ′ µ = UD µ U −1<br />

Analogously to the abelian theory (112) we define now<br />

(D µ D ν − D ν D µ ) ψ = ig τa<br />

2 F a µνψ (122)<br />

or explicitely and analogously to eq.(111)<br />

F a µν(x) = ∂ µ A a ν(x) − ∂ ν A a µ(x) − gɛ abc A b µ(x)A c ν(x) (123)<br />

The gauge transform<strong>at</strong>ion proterties of F µν can now easily be inferred: They<br />

are slightly more complic<strong>at</strong>ed than in the abelian case of eq.(113)<br />

Proof: We have from eq.(122)<br />

τ a<br />

2 F ′a<br />

µν = U τa<br />

2 F a µνU −1 (124)<br />

or<br />

and<br />

[D µ ,D ν ] = ig τa<br />

2 F a µν<br />

[D ′ µ,D ′ ν] = ig τa<br />

2 F ′a<br />

µν<br />

[D ′ µ,D ′ ν] = [UD µ U −1 ,UD ν U −1 ] = U[D µ ,D ν ]U −1 = igU τa<br />

2 F a µνU −1 qed<br />

Knowing this we can now write down the complete Lagrangean including kinetic<br />

terms for the boson fields and being invariant under SU(2) (local) gauge<br />

transform<strong>at</strong>ion. I looks to the QED lagrangean (110), however the meaning of<br />

the quanities is different:<br />

L = ψ [iγ µ D µ − m]ψ − 1 4 F a µνF µν<br />

a (125)<br />

[Note: Unlike the Lorentz indices, µ,ν there is no difference between color<br />

indices, a,b raised or <strong>low</strong>ered. The choice is dict<strong>at</strong>ed by conveninece in writing.]<br />

57


For this we have to proof, th<strong>at</strong> the kinetic energy term is indeed gauge invariant:<br />

Proof: We apply (49):<br />

F ′a<br />

µνF ′µν<br />

a = 1 2<br />

∑<br />

ab<br />

= 2 ∑ ab<br />

= 2 ∑ ab<br />

F ′a<br />

µνF ′µν<br />

b<br />

2δ ab = 1 2<br />

Tr(UFµν<br />

a τ a<br />

F a µνTr( τa<br />

2<br />

∑<br />

ab<br />

2 U −1 UF µν<br />

b<br />

τ b µν<br />

)F<br />

b<br />

= ∑ 2<br />

ab<br />

F µνF ′a ′µν<br />

b<br />

Tr[τ a τ b ] = 2 ∑ ab<br />

Tr(Fµν<br />

a τ a<br />

τ b<br />

2 U −1 ) = 2 ∑ ab<br />

Tr(F µν<br />

′a τ a<br />

2 F µν<br />

b<br />

F a µνF µν<br />

b<br />

δ ab qed<br />

τ b<br />

2 )<br />

2 F ′µν<br />

b<br />

The Lagrangean looks nearly identical to the corresponding one in the abelian<br />

case (110). There is, however, a fundamental difference. In the term FµνF a a<br />

µν<br />

there are factors contained, which are trilinear and quadrilinear in A:<br />

F a µνF µν<br />

a<br />

∼ −gf abc ∂ µ A a νA bµ A cν − g2<br />

4 fabc f ade A b µA c νA dµ A eν<br />

These terms correspond to a selfcoupling of the gauge field with itself. This<br />

means, th<strong>at</strong> the gauge field is ”charged” and hence interacts with itself. This<br />

property makes the theory fundamentally different from the abelian case, where<br />

there is no selfcoupling. Actually the Lagrangean can be rewritten in the fol<strong>low</strong>ing<br />

way<br />

L = ψ [iγ µ ∂ µ − m]ψ − 1 4 F a µνF µν<br />

a<br />

− igj a µA µ a<br />

with the current<br />

jµ a = ψγ µ τa<br />

2 ψ<br />

apparently this current gener<strong>at</strong>es the gauge field<br />

The interactions of this lagrangean can be depicted as fol<strong>low</strong>s:<br />

τ b<br />

2 )<br />

Several remarks are in order:<br />

1. The coupling of the non-abelian A-field to the current is universal as it is<br />

in the case of an abelian field. Th<strong>at</strong> means any fermion field cre<strong>at</strong>es the<br />

same structure of the A-field, if the lagrangean is to be SU(2) gauge invariant.<br />

However there is a restriction for the charge (with charge we mean<br />

the fermion-A-coupling constant) unlike to the abelian case: Suppose we<br />

simply add to the lagrangean with the fermion field ψ with coupling g<br />

another fermion field φ with a coupling ˜g = bg.<br />

L = ψ [iγ µ D µ − m]ψ + φ[iγ µ D µ − m]φ − 1 4 F a µνF µν<br />

a<br />

D µ ψ(x) =<br />

(<br />

∂ µ + ig τa A a )<br />

µ(x)<br />

ψ(x)<br />

2<br />

58


(<br />

D µ φ(x) = ∂ µ + igb τa A a )<br />

µ(x)<br />

φ ′ (x)<br />

2<br />

ψ(x) → ψ ′ (x) = exp(−i τa θ a<br />

2 )ψ(x)<br />

φ(x) → φ ′ (x) = exp(−ib τa θ a<br />

2 )φ(x)<br />

One likes to have one gauge field with the transform<strong>at</strong>ion properties (121)<br />

However, this provides immedi<strong>at</strong>ely problems: The gauge transform<strong>at</strong>ion<br />

of the field φ can be written as having the angle bθ a (x). This is identical<br />

to having gener<strong>at</strong>ors of the group not τ a but bτ a . This, however, has<br />

consequences since the bτ a have to fulfill commut<strong>at</strong>ion rules (in contrast<br />

to the abelian case). They are given by (48) and should fulfill also<br />

[ ] bτ<br />

a<br />

2 , bτb bτ c<br />

= iɛ abc<br />

2 2<br />

This, however, requires b 2 = b or b = 1. Hence all particles coupling to<br />

a given gauge field, must have the same charge. In an abelian theory the<br />

r.h.s of these equ<strong>at</strong>ions vanishes and hence there is no restriction on the<br />

charge.<br />

59


2. There is no mass term for the gauge field in the Lagrangean. If it was<br />

there it should have the structure A a µA µ a and would be not gauge invariant<br />

as one can see <strong>at</strong> eq.(121).<br />

6 The <strong>QCD</strong>-Lagrangean<br />

6.1 The gauge transform<strong>at</strong>ion<br />

The Lagrangean of the <strong>QCD</strong> results from starting from the free Lagrangean of<br />

fermions with flavour f:<br />

N f<br />

∑<br />

L 0 = q f [iγ µ (D µ ) − m f ]q f<br />

f=1<br />

Quark fields: Spin- 1 2 fermions, 6 flavours (N f = 6). We demand now th<strong>at</strong><br />

each quark of a particular flavour has three components with different colourquantum<br />

number. The <strong>QCD</strong>-Lagrangean can be constructed by gauging the<br />

colour degrees of freedom with a SU(3) − colour gauge transform<strong>at</strong>ion.<br />

The gauge transform<strong>at</strong>ion, under which this Lagrangean is invariant, acts<br />

on the colour degree of freedom and is ( for each flavour separ<strong>at</strong>ely) fol<strong>low</strong>ing<br />

eq.(51)<br />

U(θ(x)) = exp(−i λa θ a (x)<br />

) (126)<br />

2<br />

q f (x) → q ′ f(x) = U(θ(x))q f (x)<br />

The λ a are Gell-Mann M<strong>at</strong>rices of the SU(3)-algebra. They defined by<br />

eq.(48) and are given in eq.(52). The totally antisymmetric structure coefficients<br />

f abc and the totally symmetric ondes d abc can be found in eq.(53) and<br />

eq.(54), respectively. In the end the full <strong>QCD</strong>-Lagrangean is given by<br />

N f<br />

L <strong>QCD</strong> = − 1 2 Tr(G ∑<br />

µνG µν ) + q f [iγ µ D µ − m f ]q f (127)<br />

The colour-index is hidden, the flavour index is f . The covariant deriv<strong>at</strong>ive<br />

is defined as, with g being a dimensionless coupling constant:<br />

(<br />

D µ q f (x) = ∂ µ + ig λa A a )<br />

µ(x)<br />

q f (x)<br />

2<br />

Gluon field:<br />

Massless, Spin-1 bosons, colour octet because the group contains eight gener<strong>at</strong>ors,<br />

flavour singulet (i.e. gluons are flavour-blind, not distinguishing between<br />

f=1<br />

61


different flavours of wh<strong>at</strong>ever object it acts.<br />

A µ (x) =<br />

8∑<br />

a=1<br />

A a µ(x) λa<br />

2<br />

Thus explicitely with flavour index f and colour index c and Dirac index d one<br />

writes the <strong>QCD</strong>-lagrangean (127) as<br />

L <strong>QCD</strong> = − 1 N<br />

∑ f<br />

2 Tr(Ga µνG µν<br />

a )+ q fcd<br />

[i(γ µ ) dd ′<br />

f=1<br />

) ]<br />

(δ cc ′∂ µ + ig( λa<br />

2 ) cc ′Aa µ(x) − δ dd ′δ cc ′ q fd′ c ′<br />

We define a field strenght tensor in the usual nonabelian way<br />

or explicitely<br />

G µν (x) = ∂ µ A ν (x) − ∂ ν A µ (x) + ig [A µ (x),A ν (x)]<br />

G a µν(x) = ∂ µ A a ν(x) − ∂ ν A a µ(x) − gf abc A b µ(x)A c µ(x)<br />

with<br />

G µν (x) = λa<br />

2 Ga µν(x)<br />

The factor in front of the kinetic energy of the gluons comes from the normaliz<strong>at</strong>ion<br />

of the Gell-Mann m<strong>at</strong>rices 49:<br />

Tr(G µν G µν ) = 1 4 Tr(λ aλ b )G a µν(x)G µν<br />

b (x) = 1 2 Ga µν(x)G µν<br />

a (x)<br />

The classical equ<strong>at</strong>ions of motion are<br />

(iγ µ D µ − m f )q f (x) = 0<br />

D µ G a µν(x) = g ∑ f<br />

q f (x) λa<br />

2 γ νq f (x)<br />

In the limit of vanishing coupling constant g → 0 these equ<strong>at</strong>ions describe a very<br />

simple world, i.e. free massive quarks and independently from th<strong>at</strong> a gluonic<br />

field. The gluon field is, however, quite complic<strong>at</strong>ed and not comparable with<br />

the QED-field, because of the self interacing terms. The full theory is even<br />

more complic<strong>at</strong>ed since the quarks get dressed with gluons etc , dynamic mass<br />

gener<strong>at</strong>ion happens and the experimental spectrum has nothing to do with free<br />

quarks.<br />

It is interesting to note, th<strong>at</strong> the gluonic field tensor G a µν is not invariant<br />

under the gauge tranform<strong>at</strong>ion. This is in contrast to the gauge property of<br />

the QED. There the invariance of the field tensor was necessary since the corresponding<br />

electric and magnetic field was observable via the Lorentz force. Here<br />

in <strong>QCD</strong> it is different: The gluonic field is not observable directly. Observable<br />

62


are only glueballs and complex quark-gluon systems. Glueballs are stable gluon<br />

configur<strong>at</strong>ions which are observable since they are colour singuletts. They have<br />

been ”observed” in l<strong>at</strong>tice gauge calcul<strong>at</strong>ions of pure gluon field.<br />

One often finds a different form of the <strong>QCD</strong>-lagrangean. There one redefines<br />

the gluon field in such a way thqt it absorbs the strong coupling constant:<br />

gA µ → A µ . Then we have the lagrangean equivalent to (127) as<br />

L <strong>QCD</strong> = − 1<br />

N<br />

2g 2 Tr(G ∑ f<br />

µνG µν ) + q f [iγ µ (∂ µ + iA µ ) − m f ]q f<br />

6.2 Strong coupling constant<br />

f=1<br />

We know th<strong>at</strong> we have to renormalize the <strong>QCD</strong> as we have to renormalize QED<br />

(read Mandl-Shaw for an elementary introduction). This means th<strong>at</strong> e.g. the<br />

fol<strong>low</strong>ing graphs have to be summed up in order to describe properly the quark<br />

gluon vertex:<br />

This summ<strong>at</strong>ion (including higher order terms) results in an effective vertex<br />

with an effective coupling constant ḡ = ḡ(Q 2 ). Thus the interaction between<br />

63


the gluons themselves and the quarks and gluons is dependent on the process<br />

considered, in particular dependent on the momentum transfer Q 2 of the process.<br />

This is a result of renormaliz<strong>at</strong>ion techniques and the transition from g to<br />

gdepending on the scale chosen Q 2 .<br />

This is expressed in the “running coupling constant”. With the runing coupling<br />

constant is defined as<br />

α s (Q 2 ) = g2 (Q 2 )<br />

4π<br />

(128)<br />

.Here the effective coupling constant ḡ(Q 2 ) is given by means of the betafunction,<br />

which is calcul<strong>at</strong>ed perturb<strong>at</strong>ively:<br />

with the beta-function<br />

log( Q2 2<br />

Q 2 ) =<br />

1<br />

β(α s ) = α s(Q 2 )<br />

b 1 +<br />

π<br />

∫ ḡ(Q<br />

2<br />

2 )<br />

ḡ(Q 2 1 )<br />

dg<br />

β(g)<br />

( α<br />

2<br />

s (Q 2 )<br />

π<br />

)<br />

b 2 + ... (129)<br />

und<br />

b 1 = −[ 11 3 N c − 2 3 N f] (130)<br />

The fact th<strong>at</strong> the beta-function is neg<strong>at</strong>ive for small values α s is very important,<br />

because due to this fe<strong>at</strong>ure we have asymptotic freedom, i.e. for large<br />

Q 2 we have a loragithically vanishing effective coupling constant. A picture of<br />

α s (Q 2 ) is given as<br />

64


The running coupling constant is given in QED equally to (128) as<br />

α ( Q 2) = ē2 (Q 2 )<br />

4π<br />

(131)<br />

65


with α ( Q 2 → 0 ) = 1<br />

137<br />

. However the vari<strong>at</strong>ion of the electromagnetic running<br />

coupling constant with Q 2 is extremely small compared to the vari<strong>at</strong>ion of the<br />

strong running coupling constant (see figures and eq.(132)). For QED the change<br />

of α ( Q 2) is less than 0.1 % over the <strong>energies</strong> of the available acceler<strong>at</strong>ors.<br />

As an easy example to obtain α s (Q 2 ) consider the process e + e − → q¯q divided<br />

by e + e − → µ + µ − .<br />

e+<br />

q+<br />

e−<br />

q−<br />

e+<br />

µ+<br />

e−<br />

µ−<br />

This r<strong>at</strong>io is given by(Leader, Predazzi Vol.2, p.134)<br />

σ(e + e − → q¯q)<br />

σ(e + e − → µ + µ − ) = 3∑ Q 2 f(1 + α s(s)<br />

+ 1.411( α s(s)<br />

π π )2 )<br />

where s = (p e + + p e −) 2 , e.e. the energy of the colliding leptons. One performs<br />

the measurement <strong>at</strong> various values of s and determines the αs(s)<br />

π<br />

by comparison<br />

of the formula with experiment. Actually in practice there other, fully strong experiments,<br />

which determine the α s . From the particle d<strong>at</strong>a booklet one extracts<br />

for three-jet events in e + e − -collisions the value<br />

α s (s = m 2 Z) = 0.1134 ± 0.0035<br />

66


α s (s = 34GeV 2 ) = 0.1424 ± 0.018<br />

Usually one does not plot α s in dependence of the Mandelstam variable s but<br />

in dependence on the virtuality of the impinging particle (coming from deep<br />

inelastic collisions: There one has the asymptotic form i.e. for large values<br />

of Q 2 where large means much larger than 10GeV 2 .Perturbtion theory of first<br />

order in the <strong>QCD</strong>-coupling constant gives<br />

α s (Q 2 ) =<br />

4π<br />

( 11<br />

3 N c − 2 3 N f)<br />

ln<br />

(<br />

Q 2<br />

) (132)<br />

Figure on α s<br />

Λ 2 <strong>QCD</strong><br />

times correction terms. The Λ <strong>QCD</strong> is a parameter, which must be determined<br />

from the experimental values of α s . One obtains<br />

Λ <strong>QCD</strong> = (150 − 300)MeV (133)<br />

Apparently the value of Λ <strong>QCD</strong> is not well known, since it appears in the<br />

logarithmus, which is a very s<strong>low</strong>ly varying function. The values of α s (s) and<br />

of α s (Q 2 ) are rel<strong>at</strong>ed by analytical continu<strong>at</strong>ion. Actually the determin<strong>at</strong>ion of<br />

such an important constant as Λ <strong>QCD</strong> is a bit tricky:<br />

67


1) The processes to do experimentally are complic<strong>at</strong>ed and hence there are<br />

non-negligeable experimental errors. Furthermore the Λ <strong>QCD</strong> appears in the<br />

logarithm, where it is difficult to extract from the experimental d<strong>at</strong>a of α s .<br />

2) Each extraction of Λ <strong>QCD</strong> is based upon a pertub<strong>at</strong>ive p<strong>QCD</strong> calcul<strong>at</strong>ion<br />

to some order, e.g. leading order or next to leading order and a renormaliz<strong>at</strong>ion<br />

scheme. This means the α s used in the calcul<strong>at</strong>ion must be small, and so must<br />

be the α s resulting in the above formula. This is a consistency condition.<br />

3) The value of Λ <strong>QCD</strong> dependes on the number of active flavours in the<br />

process considered. Active flavours are those, whose mass is be<strong>low</strong> the scale<br />

of the reaction, i.e. be<strong>low</strong> the momentum transfers or <strong>energies</strong> of the reaction.<br />

This is analog to perturb<strong>at</strong>ion theory in non-rel<strong>at</strong>ivistic quantum mechanics,<br />

where there appears in the first order perturb<strong>at</strong>ion theory for the wave function<br />

an energy denomin<strong>at</strong>or. This denomin<strong>at</strong>or is, in our context, twice the mass of<br />

the quark. Hence the heavier the quarks, the less their Dirac sea is perturbed.<br />

This means there are often threshold effects, whose tre<strong>at</strong>ment deservs other<br />

techniques and where the above formula (132) cannot be used.<br />

For Q 2 → ∞ the interaction vanishes, which is called “asymptotic freedom”.<br />

At Q 2 ≤ 1GeV 2 the dependence of α s on Q 2 is very uncertain and in the limit<br />

Q 2 → Λ <strong>QCD</strong> the above expression even diverges. This is not a problem: The reason<br />

for divergence is th<strong>at</strong> in the construction of α s (Q 2 ) one needs perturb<strong>at</strong>ive<br />

<strong>QCD</strong> (i.e. Feynman perturb<strong>at</strong>ion theory). This perturb<strong>at</strong>ive theory assumes a<br />

coupling constant α s ≪ 1, which is for Q 2 ≤ 1GeV 2 not the case. Hence to<br />

use eq.(132) for small Q 2 is not consistent. There are modern theories, which<br />

have arguments for another dependence of α s (Q 2 ) on Q 2 <strong>at</strong> small Q 2 . There<br />

the α s (Q 2 ) grows monotonically for Q 2 → 0 but never exceeds a value of around<br />

1.5. A suggestion has been made by Solovtsov and Shirkov: The curves a and b<br />

show the 1-loop analytic results of their theory for Λ <strong>QCD</strong> = 200 and 400MeV .<br />

The cuves c and d the corresponding perturb<strong>at</strong>ive results.<br />

Important Remark: The very fe<strong>at</strong>ure of <strong>QCD</strong> is the fact th<strong>at</strong> <strong>at</strong> <strong>low</strong> <strong>energies</strong><br />

(few GeV) the physical degrees of freedom have nothing to do with the<br />

elementary degrees fo freedom. This is quite different from QED, where we<br />

always have electrons and photons. At <strong>low</strong> <strong>energies</strong> the physical degrees of freedom<br />

in <strong>QCD</strong> are baryons and mesons. The elementary degrees of freedom are<br />

quarks and gluons. The fact th<strong>at</strong> they are so different is a consequence of the<br />

non-abelian n<strong>at</strong>ure of the gauge symmetry and the fact th<strong>at</strong> the coupling constant<br />

is so large and th<strong>at</strong> the quantum fluctu<strong>at</strong>ions are really important. This<br />

fe<strong>at</strong>ure will it make necessary to consider current algebras, which are purely<br />

non-perturb<strong>at</strong>ive concelpts<br />

6.3 Mandelstam variables<br />

For general educ<strong>at</strong>ion it is interesting to consider the kinem<strong>at</strong>ics of the above<br />

process a bit more in detail. Here we have two real incoming particles and<br />

two real outgoing particles. Th<strong>at</strong> means the momenta <strong>at</strong> the particles are all<br />

timelike: p 2 i = m2 i . The coordin<strong>at</strong>es of the process are given by<br />

68


c<br />

d<br />

<br />

b<br />

<br />

a<br />

Q (GeV)<br />

<br />

<br />

69


Mandelstam Variables for the reaction a + b → A + B are generally defined<br />

as properties:<br />

s = (p a + p b ) 2 = (p A + p B ) 2 (134)<br />

t = (p a − p A ) 2 = (p b − p B ) 2<br />

with the subsidiary condition<br />

u = (p a − p B ) 2 = (p b − p B ) 2<br />

s + t + u = ∑ i<br />

m 2 i<br />

One has to distinguish this kinem<strong>at</strong>ics from the one, where one considers<br />

a virtual particle hitting a nucleon as e.g in the determin<strong>at</strong>ion of structure<br />

functions or form factors: There we have<br />

q raumartig<br />

virtual particle<br />

hitting a nucleon<br />

p zeitartig<br />

p’ zeitartig<br />

7 Symmetries and anomalies<br />

7.1 Mass terms<br />

It will be very important and convenient, do distinguish in <strong>QCD</strong> betweeen light<br />

and heavy quarks:<br />

⎫<br />

m u = mū = (4 ± 2)MeV ˜5MeV ⎬<br />

m d = m ¯d = (8 ± 4)MeV ˜10MeV<br />

⎭ m q ≪ Λ <strong>QCD</strong> :::: or ::::: M N ,M ρ<br />

m s = m¯s = (164 ± 33)MeV ˜180MeV<br />

(135)<br />

70


m c = 1.4 : GeV<br />

m b = 5.3GeV<br />

m t = 175GeV<br />

⎫<br />

⎬<br />

⎭ m q ≫ Λ <strong>QCD</strong> :::::: or :::::: M N ,M ρ (136)<br />

One should realize, th<strong>at</strong> the r<strong>at</strong>io of mu<br />

mt<br />

= 1/30000. The mass of the nucleon<br />

is M N = 938MeV and of the Rho-Meson is M ρ = 770MeV . (With numbers<br />

like this one must be a bit careful: The quark masses are scale dependent.)<br />

By energy reasons we see immedi<strong>at</strong>ely th<strong>at</strong> the nucleon is governed by up and<br />

down quarks and also a bit of strange quark antiquark admixtures. Thus for<br />

the nucleon <strong>at</strong> reactions, with momentum transfers or <strong>energies</strong> smaller than<br />

the masses of the heavy quarks, one can ignore the existence of the heavy<br />

quarks, since their quark-antiquark excit<strong>at</strong>ion requires too much energy and<br />

hence (by non-rel<strong>at</strong>ivistic simple perturb<strong>at</strong>ion theory) those excit<strong>at</strong>ions are suppressed.However,<br />

in changing the energy in an experiment one must be careful of<br />

threshold effects. Altogether: For <strong>low</strong> <strong>energies</strong> one can consider the Lagrangean<br />

with three flavours as the relevant <strong>QCD</strong>-Lagrangean:<br />

L <strong>QCD</strong> =<br />

3∑<br />

f=1<br />

q f [iγ µ D µ ]q f − 1 4 Ga µνG µν<br />

a<br />

−<br />

3∑<br />

¯q f m f q f (137)<br />

The fact, th<strong>at</strong> <strong>at</strong> <strong>low</strong> <strong>energies</strong> the heavy quarks do not play a role has the<br />

consequence th<strong>at</strong> certain symmeetries, which are valid for a massless theory,<br />

like e.g. the chiral symmetry, is approxim<strong>at</strong>ely valid for the nucleon. One has<br />

now to do a decision, if one considers the strange quark as a “light” quark or as<br />

a “heavy” quark. This means, one has to decide if one postul<strong>at</strong>es approxim<strong>at</strong>ely<br />

a SU(2) or SU(3) global flavour invariance. Thus one rewrites (137) as<br />

f=1<br />

L <strong>QCD</strong> = ψ [iγ µ D µ ]ψ − 1 4 Ga µνG µν<br />

a − ¯ψmψ (138)<br />

In SU(2)-flavour the ψ is a doublet consisting of the up-spinor and down-spinor<br />

( )<br />

ψu (x)<br />

ψ =<br />

ψ d (x)<br />

and the mass term is<br />

or<br />

with<br />

( )<br />

mu 0 ¯ψ<br />

0 m d<br />

ψ = m u + m d<br />

2<br />

(ūu + ¯dd) + m u − m d<br />

(ūu −<br />

2<br />

¯dd)<br />

( )<br />

mu 0 ¯ψ ψ =<br />

0 m ¯ψ [ m 0 I + m 3 τ 3] ψ (139)<br />

d<br />

m 0 = m u + m d<br />

2<br />

(140)<br />

m 3 = m u − m d<br />

2<br />

(141)<br />

71


In SU(3) we have<br />

⎛<br />

ψ = ⎝<br />

ψ u (x)<br />

ψ d (x)<br />

ψ s (x)<br />

and the mass term is<br />

⎛<br />

⎞<br />

m u 0 0<br />

¯ψ ⎝ 0 m d 0 ⎠ ψ = ¯ψ [m 1 I + m 3 λ 3 + m 8 λ 8 ]ψ (142)<br />

0 0 m s<br />

⎞<br />

⎠<br />

with (141) and<br />

m 1 = m u + m d + m s<br />

3<br />

m 8 = m u + m d − 2m s<br />

2 √ 3<br />

(143)<br />

(144)<br />

7.2 Vector symmetries<br />

7.2.1 Global fermion symmetry<br />

In the Lagrangean of <strong>QCD</strong> (137) there exists a symmetry of the form<br />

q f → q ′ f = exp(−iθ f )q f (145)<br />

for each flavour separ<strong>at</strong>ely and with constant θ f . This is obvious because the<br />

gluons are not affected by this transform<strong>at</strong>ion. The corresponding current is<br />

given by<br />

j µ f = q fγ µ q f (146)<br />

The corresponding charge of a given flavour corresponds to the number of quarks<br />

minus the number of antiquarks. Since the invariance property holds for each<br />

flavour separ<strong>at</strong>ely one can form linear combin<strong>at</strong>ions like e.g:<br />

with<br />

j µ em = ψQ em γ µ ψ = 2 3 uγµ u − 1 3 dγµ d − 1 3 sγµ s (147)<br />

⎛<br />

Q em = ⎝<br />

+ 2 3<br />

0 0<br />

0 − 1 3<br />

0<br />

0 0 − 1 3<br />

The current (147) is the electromagnetic current. Apparently it origin<strong>at</strong>es as<br />

conserved current from another symmetry of the Lagrangean, namely U(1) local,<br />

but is conserved in <strong>QCD</strong> as well, however as consequence of the global flavour<br />

symmetry.<br />

⎞<br />

⎠<br />

72


7.2.2 Global (iso-)vectorial symmetry<br />

SU(2): Consider up- and down-quarks in <strong>QCD</strong> (138) and assume them to have<br />

the same mass. Then we have the global isospin invariance of (138):<br />

(<br />

u<br />

ψ =<br />

d<br />

)<br />

→ ψ ′ = exp(−iθ a τa<br />

)ψ (148)<br />

2<br />

with the τ a being the tree Pauli-m<strong>at</strong>rices (12). We know this invariance from<br />

the studies around eqs. (81). The Noether current corresponding to the action<br />

of (148) to the <strong>QCD</strong> lagrangean is a triple:<br />

V µ = jµ a τ a<br />

= ψγ µ ψ a = 1,2,3 (149)<br />

2<br />

This Noether-current is conserved if the masses of up- and down-quarks are<br />

identical.<br />

SU(3): The same symmetry reads in SU(3):<br />

⎛<br />

ψ = ⎝<br />

u<br />

d<br />

s<br />

⎞<br />

⎠ → ψ ′ = exp(−iθ a λa<br />

)ψ (150)<br />

2<br />

Here the λ a are the Gell-mann-m<strong>at</strong>rices (), however not for color (as used to<br />

define the colour gauge transform<strong>at</strong>ion) but for flavour. The Noether current<br />

corresponding to the action of (150) to the <strong>QCD</strong> lagrangean is an octet<br />

λ a<br />

V µ = ψγ µ ψ a = 1,2,...,8 (151)<br />

2<br />

This Noether-current is conserved if the masses of up- and down- and s-<br />

quarks are identical.<br />

The SU(3) has three subgroups of SU(2)-n<strong>at</strong>ure.<br />

This is the T-spin:<br />

( ) ( )<br />

u<br />

ψ = → ψ ′ = exp(−iθ a τa u<br />

d<br />

2 ) d<br />

This is the U-spin:<br />

(<br />

d<br />

ψ =<br />

s<br />

This is the V-spin:<br />

(<br />

u<br />

ψ =<br />

s<br />

)<br />

→ ψ ′ = exp(−iθ a τa<br />

2 ) (<br />

d<br />

s<br />

)<br />

→ ψ ′ = exp(−iθ a τa<br />

2 ) (<br />

u<br />

s<br />

)<br />

)<br />

Let us now consider the mass terms eq.(135) and their effect on the divergence<br />

of the above currents (149,151). The masses are different from each<br />

73


other, nevertheless the currents are the same. However their divergence is not<br />

vanishing and one can derive from eq.(55) and (63) easily a general formula<br />

∂ µ V µ<br />

a = ¯ψi 1 2 [λ a,m]ψ a = 1,2,3,..,8 (152)<br />

from which we obtain in the SU(2)-case with m u = m d and m 0 = (m u +m d )/2<br />

given by eq.(140)<br />

∂ µ V a µ = m 0 ¯ψi τa ψ a = 1,2,3 (153)<br />

Compared to all masses of hadrons the up- and down-mass is always small,<br />

hence the SU(2) T-spin-symmetry (isospin-symmetry) is a generally good approxim<strong>at</strong>ion.<br />

The U-spin and V-spin symmetries are eplicitely broken due to<br />

the mass difference of about 170MeV between s-quarks and u,d-quarks. Hence<br />

predictions with isospin symmetry work <strong>at</strong> the level of 1%, whereas the corresponding<br />

ones with U-spin or V-spin symmetry in SU(3) work <strong>at</strong> 30% accuracy.<br />

Explicitely we can write with<br />

the fol<strong>low</strong>ing useful formulae<br />

and in case of al<strong>low</strong>ing m u ≠ m d<br />

V µ + = V µ<br />

1 + iV µ<br />

2<br />

V µ − = V µ<br />

1 − iV µ<br />

2<br />

V µ + = ¯ψ u γ µ ψ d<br />

V µ − = ¯ψ d γ µ ψ u<br />

∂ µ V µ + = i(m u − m d ) ¯ψ u ψ d<br />

∂ µ V µ + = i(m d − m u ) ¯ψ d ψ u<br />

One often defines also the isospin current V µ<br />

3 and the hypercharge current<br />

J µ Y in the fol<strong>low</strong>ing way (V µ as such is called isovector current):<br />

.<br />

V µ<br />

3 = 1 2 [ ¯ψ u γ µ ψ u − ¯ψ d γ µ ψ d ]<br />

J µ Y = 2 √<br />

3<br />

V µ<br />

8 = 1 8 [ ¯ψ u γ µ ψ u + ¯ψ d γ µ ψ d ] − 2 3 ¯ψ s γ µ ψ s<br />

In this business the fol<strong>low</strong>ing formula is useful<br />

[Aλ a ,Bλ b ] = 1 2 {A,B}[λ a,λ b ] + 1 2 [A,B]{λ a,λ b }<br />

74


7.3 Axial (non-)symmetries:<br />

7.3.1 Flavour symmetry and U A (1) anomaly<br />

The iso-vector symmetry holds, if the quark masses are identical. If the quark<br />

masses vanish, there are additional symmetries because in this limit left-handed<br />

and right-handed components of the quark-fields are decoupled. They are both<br />

coupled to the gluon field, however th<strong>at</strong> does not carry flavour quantum numbers.<br />

So we can write an expression for the quark fields in the <strong>QCD</strong>-lagrangean<br />

analogous to eq.(72) Thus, for the massless <strong>QCD</strong> Lagrangean one can derive a<br />

left and right hand flavour symmetry and also a global axial flavour symmetry<br />

(76,77) under the axial transform<strong>at</strong>ion<br />

with the current ()<br />

q f → q ′ f = exp(−iθ f γ 5 )q f (154)<br />

j µ 5f = q fγ µ γ 5 q f (155)<br />

However, this symmetry exists only for the classical lagrangean, it does not<br />

exist in a full quantum field theory (without proof). There the correspponding<br />

current is not conserved. Instead we have so called axial anomaly, where the<br />

divergence of the axial vector current is rel<strong>at</strong>ed to the gluon field:<br />

with the definition<br />

∂ µ j f 5µ = α s<br />

8π Ga µν<br />

µν ˜G a for eachflavour f separ<strong>at</strong>ely (156)<br />

˜G µν<br />

a = ɛ µναβ G a αβ (157)<br />

If one combines the fields with different flavour f into the usual dublets for<br />

SU(2)-flavour and triplets for SU(3)-flavour it makes sense to consider the singlet<br />

axial current and writes is in a different nomencl<strong>at</strong>ure for e.g. SU(3)-flavour<br />

as<br />

j (0)<br />

5µ = ¯ψγ µ γ 5 ψ = ūγ µ γ 5 u + ¯dγ µ γ 5 d + sγ µ γ 5 s (158)<br />

Here the anomaly equ<strong>at</strong>ion reads as fol<strong>low</strong>s and is is called the U A (1)−anomaly:<br />

∂ µ j (0)<br />

5µ = N fα s<br />

8π Ga µν<br />

µν ˜G a (159)<br />

with N f = 2 for SU(2) and N f = 3 for SU(3). The reason for this anomaly<br />

is the fol<strong>low</strong>ing: If the renormaliz<strong>at</strong>ion procedure of <strong>QCD</strong> one has to regularize<br />

the integrals in the Feynman diagrams and <strong>at</strong> the end of the procedure one lets<br />

the cut-off parameter go to infinity or, if one does a dimensional regulariz<strong>at</strong>ion,<br />

one lets the ε-parmaeter go to zero. Actually one can chose a regulariz<strong>at</strong>ion in<br />

such a way th<strong>at</strong> there is no U A (1)−anomaly, or exactly th<strong>at</strong> we have an axial<br />

singlet current (158) which vanishes in the massless limit. Unfortun<strong>at</strong>ely one<br />

cannot avoid in this case th<strong>at</strong> the singlet axial current j (0)<br />

5µ is no longer gauge<br />

invariant under the SU(3)-colour−transform<strong>at</strong>ion of <strong>QCD</strong>. Since, however, <strong>QCD</strong><br />

is constructed by demanding colour-gauge invariance such a situ<strong>at</strong>ion is not<br />

75


acceptable. Hence one must choose a regulariz<strong>at</strong>ion, in which the colour-gauge<br />

invariance is preserved but then one cannot avoid to have the U A (1)-anomaly.<br />

In fact this anomaly is responsible for the decay of the π 0 → γγ, which<br />

is experimentally well known. The axial anomaly is also responsible for the<br />

fact th<strong>at</strong> the mass of the η ′ -Meson is considerably larger than the mass of the<br />

η-meson (m η ′ = 958MeV,m η = 547MeV ). Without the U A (1)-Anomaly the<br />

massless <strong>QCD</strong> would have a chiral U(3) R ◦U(3) L -symmetry, and its spontaneous<br />

breakdown would lead to nine r<strong>at</strong>her than eight Goldstone bosons and this<br />

would include the η ′ -Meson. The axial anomaly removes the U A (1)-symmetry,<br />

keeping SU(3) R ◦SU(3) L ◦ U(1) intact, which is spontaneously broken down to<br />

SU(3) V ◦ U(1) V .<br />

In case we consider quark masses the formula modifies to<br />

∂ µ j (0)<br />

5µ = 3α s<br />

8π Ga µν<br />

µν ˜G a + 2i ( )<br />

m u ¯ψu γ 5 ψ u + m d ¯ψd γ 5 ψ d + m s ¯ψs γ 5 ψ s<br />

Proof: Using eqs.(18,19) we can write<br />

∂ µ j (0)<br />

5µ = ∂µ [ ¯ψγ µ γ 5 ψ] = (∂ µ ¯ψ)γµ γ 5 ψ + ¯ψγ µ γ 5 (∂ µ ψ)<br />

= (∂ µ ¯ψ)γµ γ 5 ψ − ¯ψγ 5 γ µ (∂ µ ψ)<br />

(160)<br />

= (im + iγ µ A µ ) ¯ψγ 5 ψ − ¯ψγ 5 (−im − iγ µ A µ )ψ = im ¯ψγ 5 ψ − ¯ψγ 5 (−imψ) qed<br />

We note here th<strong>at</strong> the finite quark masses do not modify the coefficient in font<br />

of the anomaly term.<br />

In reading al the arguments about regulariz<strong>at</strong>ion one realizes th<strong>at</strong> they apply<br />

equally well to QED if one replaces the strong coupling constant by the<br />

electromagnetic one and the gluonic field tensor by the electromagnetic field<br />

tensor.<br />

7.3.2 Axial Vector symmetry and Axial Anomaly<br />

Massless quarks: The <strong>QCD</strong>-Lagrangean is for massless quarks invariant under<br />

the axial vector transform<strong>at</strong>ion, often called also chiral transform<strong>at</strong>ion:<br />

ψ → ψ ′ = exp(−iθ a λa<br />

2 γ 5)ψ (161)<br />

It results in the axial current<br />

A a µ = ¯ψγ µ γ 5<br />

λ a<br />

2 ψ (162)<br />

For a = 0 we have the above case of a singlet axial current and it is U A (1)−anomalous<br />

(see prev. subsubsection). For the cases a = 1,2,4,5,6,7 the divergence is zero<br />

since the λ a are not diagonal and hence, written explicitely there are always<br />

non-diagonal terms (e.g. of the sort ūγ µ γ 5 d) whose divergence does not show<br />

an anomaly. For the cases a = 3,8 we also do not have an anomaly, but the<br />

76


eason is different. For these a = 3,8 the λ-m<strong>at</strong>rices are diagonal but traceless,<br />

so we have e.g. for λ 8 a term like ūγ µ γ 5 u + ¯dγ µ γ 5 dūγ − 2¯s µ γ 5 s where the<br />

U A (1)−anomaly cancels.<br />

However, explicit calcul<strong>at</strong>ions within a <strong>QCD</strong> and electromagnetic SU(3) color<br />

⊗ U(1)<br />

theory show, th<strong>at</strong> this axial vector for a = 3,8 has another anomaly, not with<br />

the gluon field G a µνbut with the photon field F µν :<br />

∂ µ A 3 µ = N cα<br />

12π F µν ˜F µν (163)<br />

∂ µ A 8 µ = N cα<br />

12π F µν ˜F µν (164)<br />

Actually a direct consequence of this anomaly is the fact th<strong>at</strong> the neutral pion<br />

can decay intotwo photons: π 0 → γγ. This decay is much faster, since it is<br />

connected with the electromagnetic coupling constant α, than the decay due<br />

to the electroweak interaction (small weak coupling constant of Fermi), under<br />

which the charged pions decay. So in practice the anomalous decay of the pion<br />

π 0 → γγ is most important. The details of the pion decay will come l<strong>at</strong>er.<br />

Explicitly we can write the divergence of the axial vector current in the<br />

presence of finite quark masses. In addition to the anomaly term for a = 3 and<br />

a = 8 one obtains for a = 1 → 8 :<br />

∂ µ A µ a = ¯ψiγ 5 1 2 {λ a,m} ψ (165)<br />

For isospin symmetry in the up-down sector this is very simple<br />

∂ µ A µ a = m 0 ¯ψiγ 5 τ a ψ (166)<br />

with m 0 = mu+m d<br />

2<br />

given in eq.(140). Explicitly we can write also the fol<strong>low</strong>ing<br />

useful formulae<br />

A µ + = A µ 1 + iAµ 2<br />

A µ − = A µ 1 − iAµ 2<br />

with<br />

A µ + = ¯ψ u γ µ γ 5 ψ d<br />

A µ − = ¯ψ d γ µ γ 5 ψ u<br />

and<br />

∂ µ A µ 3 = m ¯ψ u u iγ 5 ψ u − m d ¯ψd iγ 5 ψ d<br />

∂ µ A µ + = (m u + m d ) ¯ψ u iγ 5 ψ d<br />

∂ µ A µ − = (m u + m d ) ¯ψ d iγ 5 ψ u<br />

77


7.4 Other symmetries, Theta-vacuum<br />

7.4.1 Discrete symmetries<br />

The standard model is a Lorentz invariant local quantum field theory.Its Lagrangean<br />

is hermitean. It is invariant under the combined set of transform<strong>at</strong>ions<br />

CPT. The <strong>QCD</strong> separ<strong>at</strong>ely conserves C, P, and T separ<strong>at</strong>ely (we assume th<strong>at</strong><br />

there is no θ-term, see l<strong>at</strong>er).<br />

7.4.2 Scale invariance and trace-anomaly<br />

If the quarks were massless the <strong>QCD</strong> lagrangean would contain no dimensional<br />

quantity. The action (not the Lagrangean) would therefore be invariant under<br />

the scale transform<strong>at</strong>ions<br />

ψ(x) ↦→ ψ ′ (x) = λ 3 2 ψ(λx) A<br />

a<br />

µ (x) → λA a µ(λx) (167)<br />

Proof: Take a symbolic <strong>QCD</strong>-Lagrangean of the form L <strong>QCD</strong> = q [∂ + D µ ]q,<br />

where we write down the action, and in the action we perform a change of<br />

variable y = λx:<br />

∫<br />

S =<br />

∫<br />

=<br />

∫<br />

=<br />

∫<br />

=<br />

d 4 y[q(y) d q(y) + q(y)A(y)q(y)]<br />

dy<br />

λ 4 d 4 x[q(λx) 1 d<br />

q(λx) + q(λx)A(λx)q(λx)]<br />

λ dx<br />

λ 3 d 4 x[q(λx) d q(λx) + q(λx)λA(λx)q(λx)]<br />

dx<br />

d 4 x[λ 3 d<br />

2 q(λx)<br />

dx λ 3 3<br />

3<br />

2 q(λx) + λ 2 q(λx)λA(λx)λ 2 q(λx)]<br />

One sees immedi<strong>at</strong>ely th<strong>at</strong> the above replacements concerve the action. q.e.d.<br />

This is a continuous symmetry and hence it must have a conserved Noether<br />

current. This current is given by (without proof)<br />

J µ scale (x) = x νΘ µν (x) (168)<br />

There the Θ µ ν is the energy momentum tensor (64) of <strong>QCD</strong> and given by:<br />

Θ µν = −g µν (− 1 4 Gλσ a G a λσ + ψiγ λ D λ ψ) − G µλ<br />

a G ν aλ + A ν a∂ λ G µλ<br />

a + (169)<br />

+ i 2 ψγµ ( ←− ∂ + −→ ∂ ) ν ψ + g µν ψmψ<br />

The current is conserved on the classical level, because the above mentioned<br />

invariance under a continuous scale transform<strong>at</strong>ion holds:<br />

∂ µ J µ scale (x) = 0<br />

78


One can calcul<strong>at</strong>e the divergence immedi<strong>at</strong>ely:<br />

∂ µ J µ scale = ∂ µ(x ν Θ µν ) = ∂ µ (g νρ x ρ Θ µν ) = g νρ δ ρµ Θ µν +g νρ x ρ ∂ µ Θ µν = g νµ Θ µν = Θ ν ν<br />

where we have used the fact th<strong>at</strong> the energy momentum tensor is conserved,<br />

which is fulfilled for a time independent and homogeneous system, wh<strong>at</strong> <strong>QCD</strong><br />

is. .<br />

∂ µ Θ µν (x) = 0<br />

Thus a conserved scaling current has as consequence th<strong>at</strong> the trace of the<br />

energy momentum tensor vanishes:<br />

Θ ν ν(x) = 0<br />

If the scaling current would really be conserved such a situ<strong>at</strong>ion with a vanishing<br />

trace of the energy momentum tensor would have drastic consequences on the<br />

theory, since all particles of this world (now built of massless quarks) would<br />

then be massless, i.e. the proton would be massles, the pion as well, etc..To<br />

proof this st<strong>at</strong>emtne we forst show: For any hadron H the m<strong>at</strong>rix element of<br />

the energy-momentum tensor <strong>at</strong> zero momentum transfer is (assertion):<br />

〈<br />

H(p)<br />

∣ ∣ Θ µν (0) ∣ ∣ H(p)<br />

〉<br />

= 2p µ p ν (170)<br />

Proof:: We have on one hand for a given hadron<br />

〈H(p ′ )| ̂P µ |H(p)〉 = p µ 2p 0 δ (3) (p − p ′ )(2π) 3<br />

and on the other hand<br />

∫<br />

〈H(p ′ )| ̂P µ |H(p)〉 =<br />

〈p ′ |Θ µ0 (x) |p〉 d 3 x<br />

Using the shift oper<strong>at</strong>or (34) we obtain<br />

〈H(p ′ )| Θ µ0 (x) |H(p)〉 = exp(+ix(p ′ − p)) 〈H(p ′ )| Θ µ0 (0) |H(p)〉<br />

We can rewrite this by noting th<strong>at</strong> the integral over exponential gives the deltadistribution<br />

∫<br />

∫<br />

〈H(p ′ )| Θ µ0 (x) |H(p)〉 d 3 x = [ d 3 xexp(+ix(p ′ − p))] 〈H(p ′ )| Θ µ0 (0) |H(p)〉<br />

= exp(+ix 0 (p ′ 0 − p 0 ) 〈H(p ′ )|Θ µ0 (0) |H(p)〉 (2π) 3 δ (3) (p − p ′ )<br />

Since the 3-momenta are equal the 0-components are equal as well and the<br />

exponential function equals One. Then one gets by comparison<br />

〈H(p)| Θ µ0 (0) |H(p)〉 = p µ 2p 0<br />

and by Lorentzs-invariance this holds not only for ν = 0 but for all ν. Thus one<br />

gets the assertion eq.(170).q.e.d.<br />

79


A vanishing trace would mean<br />

〈<br />

H(p)<br />

∣ ∣ Θ<br />

µ<br />

µ (0) ∣ ∣ H(p)<br />

〉<br />

= 2p µ p µ = 0 = 2M 2 H<br />

This is direct a problem of <strong>QCD</strong> since the proton is built of up and down<br />

quarks, which are basically massless. On the contrary the experimental value<br />

of the proton mass is 938 MeV. The very fact is, th<strong>at</strong> this scale current J µ scale is<br />

anomalous. The divergence of the quantum current is given by (without proof)<br />

∂ µ J µ scale = Θν ν = β <strong>QCD</strong><br />

2g Ga µνG µν<br />

a + m u ūu + m d ¯dd + ms ss (171)<br />

Here β <strong>QCD</strong> is the beta-function of <strong>QCD</strong>, known from renormaliz<strong>at</strong>ion techniques<br />

and perturb<strong>at</strong>ion theory of <strong>QCD</strong>, see (129) with (130).This term is the<br />

anomalous term, which gives rise to the ”Scale anomaly”. The other terms are<br />

proportional to the quark masses and are the result of explicit breaking of the<br />

scale invariance. This means th<strong>at</strong> e.g. for the nucleon we have<br />

M N ū(p)u(p) = 〈 N(p) ∣ ∣ β <strong>QCD</strong><br />

2g Ga µνG µν<br />

a + m u ūu + m d ¯dd + ms ss ∣ ∣ N(p)<br />

〉<br />

(172)<br />

Since we have for the Dirac spinors the normaliz<strong>at</strong>ion<br />

ū(p,r)u(p,s) = 2M N δ rs<br />

we can write<br />

M N = 1<br />

2M N<br />

〈<br />

N(p)<br />

∣ ∣<br />

β <strong>QCD</strong><br />

2g Ga µνG µν<br />

a + m u ūu + m d ¯dd + ms ss ∣ ∣ N(p)<br />

〉<br />

(173)<br />

The term corresponding to the light masses m u ,m d is rel<strong>at</strong>ed to the famous<br />

Σ πN −Term. It is measured and shows, th<strong>at</strong> it contribute to about 45MeV or,<br />

according to more modern measurements, to about 65MeV to the total mass<br />

of the nucleon of m N = 938MeV . This leaves the bulk of the nucleons mass<br />

to the gluons and to the s-quarks. This is very interesting because it says, th<strong>at</strong><br />

the by far largest contribution to the nucleon mass comes from non-valence<br />

particles! This is a model independent st<strong>at</strong>ement and is based purely on <strong>QCD</strong><br />

and experiment. There is one problem in this argument. If we consider heavy<br />

quarks then their contribution is negligeable since < N(p)|m Q QQ ∣ ∣ N(p)<br />

〉<br />

/2MN<br />

is smaller than 1MeV and hende negligeable even for the heavy top-quark. In<br />

facht the condens<strong>at</strong>e in the nucleon is th<strong>at</strong> small Altogether, ist is fully justified<br />

to ignore the contributions of heavy quarks to the nucleon mass, nucleon angular<br />

momentum, nucleon momentum, etc.. The “intrinsic” charm, e.g., is small.<br />

One can support these arguments quantit<strong>at</strong>ively: The contribution of the s-<br />

quarks to the mass of the nucleon is still under deb<strong>at</strong>e, but more and more people<br />

agree, th<strong>at</strong> it is not small because of the large mass m s = 180MeV . Estim<strong>at</strong>es<br />

in the liter<strong>at</strong>ure are for nucleons with momentum p = 0, with m 0 = 1 2 (m u+m d )<br />

from (140)<br />

Σ πN = m 0<br />

2M N<br />

〈<br />

N(p)<br />

∣ ∣ mu ūu + m d ¯dd<br />

∣ ∣N(p)<br />

〉<br />

= (45 ± 5)MeV (174)<br />

80


1 〈 ∣<br />

N(p) β <strong>QCD</strong><br />

∣<br />

2M N 2g Ga µνG µν<br />

a<br />

∣N(p) 〉 = 634MeV :::: (764MeV ) (175)<br />

1<br />

2M N<br />

〈<br />

N(p)<br />

∣ ∣ ms ss ∣ ∣ N(p)<br />

〉<br />

= 260MeV ::::: (130MeV ) (176)<br />

In modern texts the value of the sigma-term is believed to be noticeaby larger,<br />

i.e. Σ πN˜(65 ± 8)MeV , but the question is not completely settlet yet. Note<br />

th<strong>at</strong> the bulk contribution to the energy comes from the gluons, but the strange<br />

quarks also contribute noticeably, in fact more clearly than the up- and downquarks.<br />

This fact provides a problem to a naive interpret<strong>at</strong>ion of the nucleon as<br />

composed of three constituent up- and down-quarks without internal structure.<br />

Actually it shows th<strong>at</strong> there is some gluon field and even ss-fields inside the<br />

constituent quarks, which are the dominant carrier of the energy density. Thus<br />

in the process of forming a constituent quark, the quark is “dressed” by gluonic<br />

and strange fields.<br />

The arguments leading to the above numbers can be understood after having<br />

studied the quark model. Basically they are the fol<strong>low</strong>ing: One can introduce<br />

a mass splitting oper<strong>at</strong>or führe dieses<br />

L m−s = 1 3 (m 0 − m s )(ūu + ¯dd − 2ss)<br />

später vor<br />

Actually the mass-splittings between the hyperons are governed by this oper<strong>at</strong>or.<br />

Thus one can show by very simple means involving SU(3)-symmetry and<br />

perturb<strong>at</strong>ion to first order in m s :<br />

δ s = 〈N|(m 0 − m s )(ūu + ¯dd − 2ss) |N〉<br />

2m N<br />

= 3 2 (m Ξ − m N ) = 574MeV<br />

One should note th<strong>at</strong> the structure of the nucleon is given by P = uud and<br />

N = udd and of the Xi by Ξ 0 = uss ::::: Ξ − = dss. Multiplying this with the<br />

small r<strong>at</strong>io m0<br />

m s<br />

and neglecting small terms yields the quantity<br />

δ = m 0<br />

〈N|ūu + ¯dd − 2ss |N〉<br />

2m N<br />

m 2 π<br />

= 3 2 m 2 K − (m Ξ − m N ) ∼ = 25MeV (35MeV )<br />

m2 π<br />

where the number in brackets includes higher order chiral corrections (Gasser<br />

87). Comparison of δ und σ yields, th<strong>at</strong> the strange quark m<strong>at</strong>rix element does<br />

not vanish. Indeed, one requires<br />

〈N|ss |N〉<br />

〈N|ūu + ¯dd + ss |N〉 ∼ = 0.18 ::: (0.09)<br />

if one uses these values one obtains the above numbers.<br />

81


7.4.3 Theta-Vacuum<br />

There are a gre<strong>at</strong> differences between an abelian gauge theory and a non-abelian<br />

one. One of those is the Theta-Vacuum. The gauge transform<strong>at</strong>ions one usually<br />

considers are those, which are connected to the identity oper<strong>at</strong>or in a continuous<br />

manner. They are called small gauge transform<strong>at</strong>ions.. There are however<br />

some, where this is not so, they are called large gauge transform<strong>at</strong>ions. For<br />

the fol<strong>low</strong>ing we consider the temporal gauge, which is given by A a 0 = 0 for<br />

a = 1,...,8. By definition the small gauge transform<strong>at</strong>ions do not change the<br />

winding number, the large ones do. To study this we remember the gauge degree<br />

of freedom (121), which is still there even if we have decided to work inside the<br />

temporal gauge. We restrict ourselve to the SU(2) subgroup a = 1,2,3 within<br />

the temporal gauge :<br />

τ a A a j (x)<br />

2<br />

→ τa A a j (x)′<br />

2<br />

( τ a A a j<br />

= U(θ(x))<br />

(x)<br />

2<br />

)<br />

U −1 (θ(x))+ i g [∂ jU(θ(x))] U −1 (θ(x))<br />

where the θ(x) has no time-dependence, since otherwise ba the gauge transform<strong>at</strong>ion<br />

an A 0 a ≠ 0 would be cre<strong>at</strong>ed.<br />

Consider now the fol<strong>low</strong>ing function:<br />

Λ 1 (x) = x2 − d 2<br />

x 2 + d 2 + 2idτx<br />

x 2 + d 2<br />

One can formul<strong>at</strong>e a gauge transform<strong>at</strong>ion in such a way th<strong>at</strong> we have<br />

A j a(x) λ a<br />

2 → U(α (x) Aj a(x) τ a<br />

2 U −1 (x) + ∆A j a(x)<br />

τ a<br />

2 ∆Aj a(x) = − i g [∇ jΛ 1 (x)]Λ −1<br />

1 (x)<br />

2d [<br />

= − τj<br />

g(x 2 + d 2 (d 2 − x 2 ) + 2x j (τ a x a ]<br />

) − 2d(x × τ) j<br />

)<br />

The interesting fact is, th<strong>at</strong> this gauge transform<strong>at</strong>ion transforms even a<br />

trivial vacuum field A j a(x) = 0 into a finite gauge field. The d is an arbitrary<br />

free parameter, thus even for d → 0 we do not meet the identity, since then<br />

the Λ 1 (x) becomes singular <strong>at</strong> x = 0. These gauge potentials carry a so called<br />

topological charge or the winding number (without proof):<br />

n = ig3<br />

24π 2 ∫<br />

d 3 xTr( τa<br />

2 Aa i (x) τb<br />

2 Ab j(x) τc<br />

2 Ac k(x))ɛ ijk (177)<br />

If one constructs the gauge potentials out of the trivial vacuum then one can<br />

show th<strong>at</strong> for all values of d the gauge potentials carries a conseraved topological<br />

charge for the above transform<strong>at</strong>ion<br />

n = 1<br />

82


If we apply the above gauge transform<strong>at</strong>ion several times one can successively<br />

construct A-fields with winding numbers n = 1,2,3,.....and also with n =<br />

−1, −2, −3,..... Since these are all obtained from the trivial vacuum by a gauge<br />

transform<strong>at</strong>ion, their energy is always zero. Hence one can plot the <strong>energies</strong> of<br />

the vacua against the winding number:<br />

E(A)<br />

0<br />

n<br />

Indeed, all possible vacua, having al vanishing energy, can be classified into<br />

disjunct sectors |n〉 labelled by their winding number n. Inside the sector we<br />

have small gauge transform<strong>at</strong>ions, if we transform from one sector to another<br />

one we need large gauge transform<strong>at</strong>ions. And there exist oper<strong>at</strong>ors,U 1 which<br />

do the transform<strong>at</strong>ion of the sort<br />

U 1 |n〉 = |n + 1〉<br />

This implies th<strong>at</strong> the real vacuum, being necessarily gauge invariant, requires<br />

contributions from all classes, such as the coherent superposition<br />

|θ〉 = ∑ n<br />

exp(−iθn) |n〉<br />

where θ is an arbitrary parameter. This vacuum is gauge invariant since we<br />

have<br />

U 1 |θ〉 = ∑ n<br />

exp(−iθn) |n + 1〉 = exp(−iθ) ∑ m<br />

exp(−iθm) |m〉 = exp(−iθ) |θ〉<br />

This Theta-vacuum is interesting because of several reasons.<br />

1) If θ ≠ 0 then the so formul<strong>at</strong>ed <strong>QCD</strong> is no longer invariant under parity<br />

transform<strong>at</strong>ion and also under time-reversal transform<strong>at</strong>ion. Thus it viol<strong>at</strong>es<br />

CP-invariance in addition to the usual mechanism caused by the Cabibbo-<br />

Kobayashi-Maskawa m<strong>at</strong>rix. CPT is always conserved. If these symmetries are<br />

not conserved, then the neutron should have a electric dipole moment. There are<br />

83


measurements of this but the upper limits are so small, th<strong>at</strong> certainly θ ≤ 10 −6 .<br />

Hence presently one does not know theta but only its upper limit.<br />

2) If one performs for any observable a p<strong>at</strong>h integral over the gauge-field<br />

A a µ(x) then the integral must extend over all gauge fields including those, which<br />

lead from one “minimum” with winding number n to another “minimum” with<br />

another winding number. Such fields do exist. They start off <strong>at</strong> t =-∞ as the<br />

zero potential A a µ(x) = 0 , have some interpol<strong>at</strong>ing A a µ(x) ≠ 0 for intermedi<strong>at</strong>e<br />

times, and end up <strong>at</strong> t =+∞ in the gauge equivalent configur<strong>at</strong>ion A j µ(x) =<br />

∆A j a(x) and a vanishing 0-component. For those fields one can show th<strong>at</strong> the<br />

fol<strong>low</strong>ing integral is non-vanishing:<br />

g 2 ∫<br />

24π 2<br />

d 3 xTr( τa<br />

2 ∆Aa i (x) τb<br />

2 ∆Ab j(x) τc<br />

2 ∆Ac k(x))ɛ ijk<br />

which means th<strong>at</strong> the interpol<strong>at</strong>ing field configur<strong>at</strong>ion gives a change in the<br />

winding number between asysmptotic gauge field configur<strong>at</strong>ions.<br />

Ε<br />

n<br />

The contributions from those fields are truly non-perturb<strong>at</strong>ive since they are<br />

not confined to one minimum, as Feynman diagrams are (this can be demonstr<strong>at</strong>ed,<br />

but will not be shown here)<br />

Actually the height of the barriers between the “minima” depends on the<br />

sp<strong>at</strong>ial extension we al<strong>low</strong> the intermedi<strong>at</strong>e fields A(x,t) in the functional integral.<br />

Such a rel<strong>at</strong>ion is necessary in order to fulfill the scale invariance of the<br />

(massless) <strong>QCD</strong>. Basically all sizes have to be considered, since the functional<br />

integral is not restricted. Now the interestign fe<strong>at</strong>ure is the fol<strong>low</strong>ing:<br />

If we consider small extensions of the interpol<strong>at</strong>ing field contributing to the<br />

functional integral (small compared to 0.1 fm) then the corresponding barrier is<br />

high and for those configur<strong>at</strong>ions it is sufficient to approxim<strong>at</strong>e the functional<br />

integral by a saddlepoint aproxim<strong>at</strong>ion and zero-mode fluctu<strong>at</strong>ions and small<br />

fluctu<strong>at</strong>ions around the saddle point. The configur<strong>at</strong>ion of the saddle point are<br />

the so called instantons. They are the solutions of the classical equ<strong>at</strong>ions of<br />

motion in the Euclidean space. The zero mode fluctu<strong>at</strong>ions around them correspond<br />

to symmetries of the instantons, are tre<strong>at</strong>ed via collective coordin<strong>at</strong>es<br />

84


and can be tre<strong>at</strong>ed exactly. The small fluctu<strong>at</strong>ions around the instantons correspond<br />

to a perturb<strong>at</strong>ive tre<strong>at</strong>ment along the non-perturb<strong>at</strong>ive intermedi<strong>at</strong>e<br />

field (instantons leading from one minimum to another). As long as they are<br />

small such a perturb<strong>at</strong>ive tre<strong>at</strong>ment is justified and it can be done in fact.<br />

If we consider large extensions of the interpol<strong>at</strong>ing field contributing to the<br />

functional integral, then the corresponding barrier is <strong>low</strong>. Still the saddlepoint<br />

approxim<strong>at</strong>ion consists in taking the instantons and zero-mode fluctu<strong>at</strong>ions and<br />

small fluctu<strong>at</strong>ions around them. However, for small barriers this is no longer a<br />

good approxim<strong>at</strong>ion: One has in principle to sove fully the functional integral<br />

and the techniques for this are unknown. Th<strong>at</strong> is the reason, why <strong>QCD</strong> is not<br />

fully solvable due to non-perturb<strong>at</strong>ive effects.<br />

Actually, given this nontrivial vacuum structure, one requires three ingredients<br />

to completely specify <strong>QCD</strong>: 1) the <strong>QCD</strong> lagrangean including all quark<br />

masses, 2) the coupling constant, i.e. Λ <strong>QCD</strong> , 3) the vacuum label θ. One can<br />

show th<strong>at</strong> one can modify the <strong>QCD</strong>-Lagrangean, known to us so far, by adding<br />

a simple term, which does not destroy the gauge invariance and leads for a given<br />

θ to a corresponding Theta-vacuum:<br />

˜G<br />

µν<br />

a<br />

L <strong>QCD</strong> = L (θ=0)<br />

<strong>QCD</strong> + θ g<br />

64π 2 Ga µν<br />

µν ˜G a<br />

with given by eq.(157). So without proof we st<strong>at</strong>e, taht a correct procedure<br />

for doing calcul<strong>at</strong>ions involving θ-vacua is to fol<strong>low</strong> the ordiönary p<strong>at</strong>h<br />

integra method but with this modified Lagrangean. One sees clearly: A different<br />

θ corresponds to a different Lagrangean and hence to a different theory<br />

altogether. One further sees: The oper<strong>at</strong>or F ˜F is P-odd and T-odd, because<br />

it corresponds to EB and this changes sign under Parity and Time-reversal.<br />

Hence, as mentioned above, a finite θ introduces CP-viol<strong>at</strong>ion.<br />

8 Anomalies<br />

9 Electroweak Interactions in the Standard Model<br />

By now we have investig<strong>at</strong>ed symmetries of the lagrangeans and the corresponding<br />

symmetry currents as e.g. V a<br />

µ or A µ a. Some of these currents were<br />

preserved,others had a non-vanishing divergence. The interactions of a hadronic<br />

system, consisting of quark and gluons, with external fields of electroweak characterm,<br />

is given by other currents, i.e. the electromagnetic current J µ EM , the<br />

weak charged current J µ CC and the weak neutral current Jµ NC<br />

. The interesting<br />

point is th<strong>at</strong> both sorts of currents are intim<strong>at</strong>ely rel<strong>at</strong>ed in such a way th<strong>at</strong> the<br />

interaction currents can be expressed in terms of the symmetry currents. To<br />

see this we first repe<strong>at</strong> in the fol<strong>low</strong>ing the basic interactions of the Standard<br />

Model. Actually, we do not consider colour currents interacting with an external<br />

glluon filed since the l<strong>at</strong>ter one does not exist as an external field due to the<br />

confinement mechnism.<br />

Here one<br />

should insert<br />

Mario´s notes<br />

on anomalies.<br />

Check them ,<br />

only very little<br />

problem with<br />

some technical<br />

specific<strong>at</strong>ions<br />

85


The Standard Model of electromagnetic and weak interactions is based on local<br />

gauge symmetry under the group SU(2) L ×U(1). After spontaneous symmetry<br />

breaking through a Higgs mechanism the interaction part of the Lagrangian<br />

is:<br />

L int = − eJ µ EM (x)A g<br />

µ(x) − J µ NC<br />

2cos θ (x)Z µ(x)<br />

W<br />

−<br />

g<br />

2 √ 2 Jµ CC (x)W† µ(x) + Hermitian conjug<strong>at</strong>e.<br />

(178)<br />

It involves J µ EM , Jµ CC and Jµ NC , and their couplings to the photon field A µ, the<br />

charged W-boson fields W ± µ and the neutral Z 0 -boson field Z µ . The electromagnetic<br />

and weak couplings are rel<strong>at</strong>ed by the Weinberg angle θ W :<br />

sin θ W = e g .<br />

Its cosine gives the r<strong>at</strong>io of W- and Z-boson masses:<br />

cos θ W = M W<br />

M Z<br />

.<br />

The weak coupling constant g is usually given interms of the Fermi coupling<br />

constant, G F , whic are rel<strong>at</strong>ed by<br />

G F<br />

√<br />

2<br />

=<br />

g2<br />

8M 2 W<br />

At present the best values of these parameters are:<br />

.<br />

G F =1.16639 ± 0.00001 × 10 −5 GeV −2 ,<br />

sin 2 θ W =0.2230 ± 0.0004,<br />

M W =80.42 ± 0.06GeV; M Z = 91.188 ± 0.002GeV.<br />

(179)<br />

The electroweak currents in Eq.(178) include sums over all quarks and leptons.<br />

We now specify the quark currents, starting from the electromagnetic ones.<br />

9.1 Electromagnetic Quark Currents<br />

The u-,d- and s-quarks are the building blocks of a fundamental represent<strong>at</strong>ion<br />

of the unitary flavor group SU(3) f . Together they form a triplet field ψ(x):<br />

⎛ ⎞<br />

ψ u (x)<br />

ψ(x) = ⎝ψ d (x) ⎠<br />

ψ s (x)<br />

where x ≡ x µ = (t,⃗x) represents time and space coordin<strong>at</strong>es. Each of the spin<br />

1<br />

2 quark fields ψ u, ψ d and ψ s is a four-component Dirac field which annihil<strong>at</strong>es a<br />

quark or cre<strong>at</strong>es an antiquark of a given flavor. (For the present discussion color<br />

86


plays no role and we shall drop explicit color labels. We simply note th<strong>at</strong> all<br />

electroweak currents (e.g. Eq.(178) , be<strong>low</strong>), being color blind, involve a trace<br />

over color.)<br />

The hypothesis th<strong>at</strong> the quarks are elementary, pointlike Dirac particles with<br />

charges Q u = + 2 3 and Q d = Q s = − 1 3<br />

(in units of e,electrons having the charge<br />

−e) immedi<strong>at</strong>ely implies th<strong>at</strong> their electromagnetic (EM) current is<br />

J µ EM (x) = ¯ψ(x)Qγ µ ψ(x) (180)<br />

with Dirac m<strong>at</strong>rices γ µ (µ = 0,1,2,3), ¯ψ = ψ † γ 0 , and our conventions for the<br />

metric, Dirac m<strong>at</strong>rices and so on are summarized in one fo the first sections.<br />

The charge m<strong>at</strong>rix ⎛<br />

2/3 0 0<br />

⎞<br />

Q = ⎝ 0 −1/3 0 ⎠ (181)<br />

0 0 −1/3<br />

takes into account the electric charges of the quarks. More explicitly,<br />

J µ EM (x) = (+2 3 ) ¯ψ u (x)γ µ ψ u (x) + (− 1 3 ) ¯ψ d (x)γ µ ψ d (x) + (− 1 3 ) ¯ψ s (x)γ µ ψ s (x).<br />

(182)<br />

The J µ EM<br />

(x) is conserved since each flavour component is conserved due to the<br />

flavour symmetry. Note th<strong>at</strong> the u- and d-quarks alone form an isospin I = 1/2<br />

doublet, with<br />

( )<br />

⃗t 2 ψu<br />

= 3 ( ) ( )<br />

ψu ψu<br />

, t 3 = 1 ( )<br />

ψu<br />

,<br />

ψ d 4 ψ d ψ d 2 −ψ d<br />

where ⃗t = 1 2 (τ 1,τ 2 ,τ 3 ) are the gener<strong>at</strong>ors of the isospin SU(2) group . Their<br />

EM current has an isoscalar and an isovector part. The charge oper<strong>at</strong>or for the<br />

u- and d-quarks is<br />

SU(2) : Q = 1 2 (B + τ 3) =<br />

(<br />

2/3 0<br />

0 −1/3<br />

)<br />

, (183)<br />

where the isoscalar part involves the quark baryon number, B = 1/3, and<br />

the isovector part is proportional to τ 3 . When the s-quark is included, the<br />

isoscalar part is generalized to incorpor<strong>at</strong>e its strangeness S = −1, and the<br />

baryon number is replaced by the hypercharge, Y = B + S and Q = Y 2 + I 3<br />

with<br />

⎛<br />

⎞<br />

⎛ ⎞<br />

Y = λ 8<br />

√<br />

3<br />

=<br />

1/3 0 0<br />

⎝ 0 1/3 0 ⎠ , I 3 = λ 1/2 0 0<br />

3<br />

√ = ⎝ 0 −1/2 0⎠ , (184)<br />

0 0 −2/3 2 0 0 0<br />

in terms of the flavor SU(3) m<strong>at</strong>rices λ 3 and λ 8 . The electromagnetic current<br />

(182) can now be written as<br />

J µ EM (x) = 1 2 Jµ Y + V µ<br />

3<br />

87


with the isospin (isovector) current<br />

and the hypercharge current<br />

V µ<br />

3 (x) = ¯ψ(x)I 3 γ µ ψ(x) = 1 ( ¯ψu γ µ ψ u −<br />

2<br />

¯ψ d γ µ )<br />

ψ d ,<br />

J µ Y (x) = ¯ψ(x)Y γ µ ψ(x) = 1 3<br />

( ¯ψu γ µ ψ u + ¯ψ d γ µ ψ d<br />

)<br />

−<br />

2<br />

3 ¯ψ s γ µ ψ s .<br />

The generaliz<strong>at</strong>ion including the heavy (c, b and t) quarks is straightforward but<br />

will not be of gre<strong>at</strong> relevance as long as we consider light mesons and baryons<br />

build of u-, d- and s-quarks.<br />

9.2 Weak Quark Currents<br />

Charged current (CC) weak interactions occur in processes like neutron beta<br />

decay, muon capture or neutrino sc<strong>at</strong>tering, i.e.:<br />

¯ν µ + p → µ + + n,<br />

Neutral current (NC) interactions can be studied, either through direct reactions<br />

such as<br />

ν µ + p → ν µ + p,<br />

or through interference effects such as the measurement of parity viol<strong>at</strong>ion in<br />

electron sc<strong>at</strong>tering (performed in order to learn something about the strange<br />

quark content of the nucleon). As a consequence we shall need all of the weak<br />

currents associ<strong>at</strong>ed with quarks.<br />

Where it is necessary to distinguish between the hadronic and the leptonic<br />

weak currents we shall add an appropri<strong>at</strong>e subscript – e.g. J µ had,W<br />

. This will be<br />

dropped where it is obvious from the context which current is intended. The<br />

full hadronic, weak current will be conveniently separ<strong>at</strong>ed into neutral (NC) and<br />

charged (CC) current pieces:<br />

J µ had,W = Jµ NC + Jµ CC .<br />

The NC is the easiest to write as it is diagonal in flavor:<br />

(<br />

J µ had,NC = ¯ψ 1<br />

u γ µ 2 − 4 3 sin2 θ W − 1 )<br />

2 γ 5 ψ u + (u → c) + (u → t)<br />

+ ¯ψ d γ<br />

(− µ 1 2 + 2 3 sin2 θ W + 1 )<br />

2 γ 5 ψ d + (d → s) + (d → b).<br />

(185)<br />

Whereas the neutral current couplings are identical (within experimental<br />

accuracy) for u,c and t as well as for d, s and b, the situ<strong>at</strong>ion for the charged<br />

couplings is considerably more complic<strong>at</strong>ed. The hadronic part of the charged<br />

current CC is given by<br />

( ) ( )<br />

J µ had,CC = ¯ψ u → c u → t<br />

u γ µ (1 − γ 5 )ψ d ′ +<br />

d ′ → s ′ +<br />

d ′ → b ′ + h.c. (186)<br />

88


where h.c. stands for the Hermitian conjug<strong>at</strong>e. The structure of this current<br />

is characterized by ”V − A”, i.e. vector current minus axial current. In this<br />

current there appear the the weak eigenst<strong>at</strong>es d ′ , s ′ and b ′ obtained through<br />

a unitary transform<strong>at</strong>ion involving the Cabibbo-Kobayashi-Maskawa (CKM)<br />

m<strong>at</strong>rix V applied to the quark mass eigenst<strong>at</strong>es with (weak) isospin − 1 2 (i.e.<br />

d, s and b), which are the st<strong>at</strong>es appearing in the propag<strong>at</strong>or. The Cabibbo-<br />

Kobayashi-Maskawa (CKM) m<strong>at</strong>rix V :<br />

⎛ ⎞ ⎛ ⎞<br />

d ′ d<br />

⎝s ′ ⎠ = V ⎝s⎠ , (187)<br />

b ′ b<br />

where V is expressed in terms of three ”Cabibbo angles”, θ 1,2,3 , and a phase<br />

angle δ. The m<strong>at</strong>rix V has the form<br />

⎛<br />

⎞ ⎛<br />

⎞<br />

c 1 −s 1 c 3 −s 1 s 3 V ud V us V ub<br />

V = ⎝s 1 c 2 c 1 c 2 c 3 − s 2 s 3 e iδ c 1 c 2 s 3 + s 2 c 3 e iδ ⎠ = ⎝V cd V cs V cb<br />

⎠ ,<br />

s 1 s 2 c 1 s 2 c 3 + c 2 s 3 e iδ c 1 s 2 s 3 − c 2 c 3 e iδ V td V ts V tb<br />

(188)<br />

where for ease of writing we have abbrevi<strong>at</strong>ed cosθ 1 as c 1 ,sin θ 2 as s 2 and so<br />

on. (This is one out of several equivalent parametriz<strong>at</strong>ions of the CKM m<strong>at</strong>rix.<br />

The CKM-m<strong>at</strong>rix is often also named as indic<strong>at</strong>ed in the RHS of the above<br />

formula.) The reason for the existence of the Cabibbo-Kobayashi-Maskawa<br />

m<strong>at</strong>rix is, th<strong>at</strong> in weak processes the W-bson couples only to the left handed<br />

particles and right handed anti-particles. This means, it can only decay into<br />

leptons or by W → u L ¯dR or W → d L ū R there is no decay of the sort W →<br />

u R ¯dL . Thus in weak processes the d-, s- and b-quarks are gener<strong>at</strong>ed only via<br />

their left-handed component (weak eigenst<strong>at</strong>es), whereas the eigenst<strong>at</strong>es of the<br />

hamiltonian, describing their free evolution in time (mass eigenst<strong>at</strong>es), involves<br />

the full wave function having both left- and right handed components. There<br />

is nothing particular in all these down quarks, one can formul<strong>at</strong>e the theory<br />

equivalently in such a way th<strong>at</strong> one mixes all quark st<strong>at</strong>es or one uses a CKMm<strong>at</strong>rix<br />

only for u-, c- and t-quarks instead for d-, s- and b-quarks. This one<br />

can see if one writes down a 6-plet Ψ = (ψ u ,ψ d ,ψ c ,..,ψ b ) and some sort of<br />

CKM-m<strong>at</strong>rix Ṽ yielding for the charged hadronic current<br />

J had,CC = ¯Ψγ µ (1 − γ 5 )Ṽ Ψ (189)<br />

One can apply the Ṽ to Ψ, or to ¯Ψ or write Ṽ = S† S and apply S to Ψ and S †<br />

to ¯Ψ.<br />

For practical purposes θ 2 and θ 3 are small. Thus for most applic<strong>at</strong>ions in<br />

this book we can ignore in the CKM-m<strong>at</strong>rix (188)all but the usual Cabibbo<br />

angle θ C = −θ 1 , writing<br />

( d<br />

′<br />

s ′ )<br />

( )( )<br />

cos θC sinθ C d<br />

∼= , (190)<br />

−sin θ C cos θ C s<br />

with sinθ C = 0.220 ± 0.002 and cos θ C = 0.975 ± 0.002. Sometimes we are<br />

ignoring even this mixing and assume θ C = 0.<br />

89


9.3 Leptonic currents<br />

The leptonic elcetromagnetic current is clear and simple<br />

J µ lep,EM = − ¯ψ e γ µ ψ e + (e → µ) + (e → τ) (191)<br />

Finally, we shall also need the leptonic weak current, J µ lep,W<br />

, which we summarize<br />

here:<br />

J µ lep,W = Jµ lep,NC + Jµ lep,CC .<br />

Here the charged current is<br />

J µ lep,CC = ¯ψ νe γ µ (1 − γ 5 )ψ e +<br />

and the neutral current is<br />

( )<br />

νe → ν µ<br />

+<br />

e → µ<br />

( )<br />

νe → ν τ<br />

+ h.c., (192)<br />

e → τ<br />

J µ lep,NC = ¯ψ e γ µ (g V − g A γ 5 )ψ e + (e → µ) + (e → τ)<br />

+ 1 2 ¯ψ νe γ µ (1 − γ 5 )ψ νe + (ν e → ν µ ) + (ν e → ν τ ),<br />

(193)<br />

with<br />

g A = − 1 2 ; g V = 2sin 2 θ W − 1 2 .<br />

Equipped with these basic tools we are now ready to explore the structure of the<br />

nucleon. We will in the fol<strong>low</strong>ing look in which way the interaction currents can<br />

be replaced with symmetry currents, whose algebraic rules (current algebras)<br />

we know.<br />

10 Chiral symmetry breaking<br />

10.1 Chiral Symmetry and Current algebras<br />

In the limit of massless quarks the <strong>QCD</strong> lagrangean has the symmetry rel<strong>at</strong>ed<br />

to the conserved right- or left-handedness (chirality) of zero mass spin-1/2 particles.<br />

From this one can derive the invariance under the global SU(3) (iso-)vector<br />

flavour transform<strong>at</strong>ion (89)<br />

ψ(x) → ψ ′ (x) = exp(−iα a λa<br />

2 )ψ(x)<br />

and the SU(3) axial flavour transform<strong>at</strong>ion (90)<br />

ψ(x) → ψ ′ (x) = exp(−iα a λa<br />

2 γ 5)ψ(x)<br />

One can now write down the algebras of the vector and axial currents, and<br />

these expressions remind us of the expressions we had in the general chapter<br />

90


on Noether currents: With the explicit expressions of the vector and axial currents()<br />

and ()<br />

V µ (a) = ¯ψγ τ a<br />

µ<br />

2 ψ<br />

A (a)<br />

µ = ¯ψγ τ a<br />

µ γ 5<br />

2 ψ<br />

We also can define vector and axial vector charges<br />

∫ ∫<br />

Q V a (t) =<br />

d 3 xV 0<br />

a (x) =<br />

∫ ∫<br />

Q A a (t) = d 3 xA 0 a(x) =<br />

d 3 xψ † (x) λa<br />

2 ψ<br />

d 3 xψ † (x)γ 5 λa<br />

2 ψ<br />

Actually one should note: The charge oper<strong>at</strong>or e.g. Q A a (t) is an oper<strong>at</strong>or in the<br />

Hilbert space spanned by the cre<strong>at</strong>ion and annihil<strong>at</strong>ion oper<strong>at</strong>ors of the field<br />

oper<strong>at</strong>ors ψ of all flavours. They are not oper<strong>at</strong>ors in the vlavour space where<br />

the λ a act nor in the spinorspace, where the γ µ act.The charge oper<strong>at</strong>ors are<br />

time independent since the chiral currents are conserved. This holds only if the<br />

fermions are considered massless. Actually we have a mass term in the <strong>QCD</strong>-<br />

Lagrangean, thus the time independence holds only approxim<strong>at</strong>ely. It can be<br />

used in particular processes, where the momentum transfer is th<strong>at</strong> large and<br />

sets such a large scale th<strong>at</strong> one can neglect the quark masses.<br />

We have a closed algebra for the vector charges and an open algebra for the<br />

axial charges:<br />

[<br />

Q<br />

V<br />

a ,Q V ]<br />

b = ifabc Q V c<br />

[<br />

Q<br />

V<br />

a ,Q A ]<br />

b = ifabc Q A c<br />

[<br />

Q<br />

A<br />

a ,Q A ]<br />

b = ifabc Q V c<br />

and similarly<br />

[Q a (t),V µ<br />

b (t,x)] = ifabc V µ<br />

c (x)<br />

[Q a (t),A µ b (t,x)] = ifabc A µ c (x)<br />

[<br />

Q<br />

5<br />

a (t),A µ b (t,x)] = if abc V µ<br />

c (x)<br />

and similarly<br />

[<br />

Q<br />

5<br />

a (t),V µ<br />

b (t,x)] = −if abc A µ c (x)<br />

δ(x 0 − y 0 ) [ V 0<br />

a (x),V 0<br />

b (y) ] = iδ(x − y)f abc V 0<br />

c (x)<br />

91


δ(x 0 − y 0 ) [ V 0<br />

a (x),A 0 b(y) ] = iδ(x − y)f abc A 0 c(x)<br />

δ(x 0 − y 0 ) [ A 0 a(x),A 0 b(y) ] = iδ(x − y)f abc V 0<br />

c (x)<br />

These rel<strong>at</strong>ions are obtained directly by explicit calcul<strong>at</strong>ions fol<strong>low</strong>ing the techniqus<br />

explained be<strong>low</strong>. They hold, even if there are finite quark masses, because<br />

the definition of the currents involves only the kinetic energy of the quarks and<br />

the transform<strong>at</strong>ions done on the system. We need the properties of SU(3)Lie<br />

groups with the totally antisymmetric structure coefficients f abc<br />

[ λ<br />

a<br />

2 , λb<br />

2<br />

]<br />

abc λc<br />

= if<br />

2 ::::::::: Tr(λa λ b ) = 2δ ab<br />

and the symmetric coefficients d abc with d abc = d bac = d acb etc:<br />

{<br />

λa<br />

2 , λ }<br />

b<br />

= 1 2 3 δab + d abc λ c<br />

2<br />

The proofs of the above formulae are based on the fol<strong>low</strong>ing equ<strong>at</strong>ions, which<br />

one can prove by writing them down explicitely:<br />

and<br />

[AB,C] = A[B,C] + [A,C] B<br />

[C,AB] = [C,A] B + A[C,B]<br />

[AB,C] = A{B,C} − {A,C}B<br />

[C,AB] = {C,A}B − A{C,B}<br />

We will prove now the fol<strong>low</strong>ing formulae<br />

[Q V a ,ψ(y)] = − λ a<br />

2 ψ(y)<br />

[Q V a , ¯ψ(y)] = λ a<br />

2 ¯ψ(y)<br />

[Q V a ,γ 5 ψ(y)] = − λ a<br />

2 γ 5ψ(y)<br />

[Q V a , ¯ψ(y)γ 5 ] = λ a<br />

2 ¯ψ(y)γ 5<br />

Before one starts to calcul<strong>at</strong>e one should note: The charge oper<strong>at</strong>or Q V a (t)<br />

is an oper<strong>at</strong>or in the Hilbert space spanned by the cre<strong>at</strong>ion and annihil<strong>at</strong>ion<br />

oper<strong>at</strong>ors of the field oper<strong>at</strong>ors ψ of all flavours. It is are not an oper<strong>at</strong>or in<br />

the flavour space where the λ a act nor in the spinorspace, where the γ µ act.<br />

92


Proof: At equal times and with A = ψ † (x),B = λa<br />

2<br />

ψ(x),C = ψ(y) we have with<br />

[AB,C] = A{B,C} − {A,C}B<br />

∫<br />

[Q V a ,ψ(y)] = d 3 x[ψ † (x) λa<br />

2 ψ(x),ψ(y)]<br />

∫<br />

= d 3 x(ψ † (x){ λa<br />

2 ψ(x),ψ(y)} − {ψ† (x),ψ(y)} λa<br />

2 ψ(x))<br />

∫<br />

= − d 3 xδ 3 (x − y) λa<br />

ψ(x) = −λa ψ(y) qed<br />

2 2<br />

Doing the hermitian conjug<strong>at</strong>e we obtain the second formula:<br />

[Q V a ,ψ(y)] † = −( λ a<br />

2 ψ(y))†<br />

−[Q V a ,ψ † (y)] = −(ψ † (y) λ a<br />

2 )<br />

−[Q V a ,ψ † (y)γ 0 ] = −ψ † (y)γ 0 λ a<br />

2<br />

−[Q V a , ¯ψ(y)] = − ¯ψ(y) λ a<br />

2<br />

[Q V a , ¯ψ(y)] = ¯ψ(y) λ a<br />

2<br />

We need the analogue formulae for axial quantities<br />

[Q A a ,ψ(y)] = − λ a<br />

2 γ5 ψ(y)<br />

[Q A a , ¯ψ(y)] = − ¯ψ(y)γ 5 λ a<br />

2<br />

qed<br />

and because γ 5 = γ † 5 also [Q A a ,γ 5 ψ(y)] = − λ a<br />

2 ψ(y)<br />

[Q A a , ¯ψ(y)γ 5 ] = − ¯ψ(y) λ a<br />

2<br />

The proof for this is similar to the above one and like th<strong>at</strong>. We take<br />

[AB,C] = A{B,C} − {A,C}B and use A = ψ † ,B = λa<br />

2 γ5 ψ(x),C = ψ(y)<br />

and equal times:<br />

∫<br />

]<br />

d 3 x<br />

[ψ † (x) λa<br />

2 γ5 ψ(x),ψ(y)<br />

∫<br />

= d 3 x(ψ † (x){ λa<br />

2 γ5 ψ(x),ψ(y)} − {ψ † (x),ψ(y)} λa<br />

2 γ5 ψ(x))<br />

∫<br />

= − d 3 xδ 3 (x − y) λa<br />

2 γ5 ψ(x) = − λa<br />

2 γ5 ψ(y) qed<br />

93


To get the commut<strong>at</strong>or of Q A a with ¯ψ we need the hermitian conjug<strong>at</strong>e (and<br />

the fact th<strong>at</strong> γ † 5 = γ 5and γ 5 anticommuting with γ 0 )<br />

[Q A a ,ψ(y)] † = −( λ a<br />

2 γ5 ψ(y)) †<br />

−[Q A a ,ψ † (y)] = −(ψ † (y)γ 5 λ a<br />

2 )<br />

−[Q A a ,ψ † (y)γ 0 ] = −ψ † (y)γ 5 λ a<br />

2 γ0<br />

−[Q A a ,ψ † (y)γ 0 ] = ψ † (y)γ 0 γ 5 λ a<br />

2<br />

−[Q A a , ¯ψ(y)] = ¯ψ(y)γ 5 λ a<br />

2<br />

[Q A a , ¯ψ(y)] = − ¯ψ(y)γ 5 λ a<br />

2<br />

qed<br />

And now we can derive two formulae, which we need l<strong>at</strong>er in the section on<br />

PCAC and the vacuum condens<strong>at</strong>es and sigma terms::<br />

[<br />

Q<br />

V<br />

a , ¯ψ(y)λ b ψ(y) ] = 2 ¯ψ(y)[ λ a<br />

2 , λ b<br />

2 ]ψ(y) = ifabc ¯ψλc ψ (194)<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)λ b γ 5 ψ(y) ] = −2 ¯ψ(y){ λ a<br />

2 , λ b<br />

2 }ψ(y) = −2 3 δ ab ¯ψ(y)ψ(y) − d abc ¯ψ(y)λc ψ(y)<br />

(195)<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)λ b ψ(y) ] = −2 ¯ψ(y){ λ a<br />

2 , λ b<br />

2 }γ5 ψ(y) = − 2 3 δ ab ¯ψ(y)γ 5 ψ(y) − d abc ¯ψ(y)λc γ 5 ψ(y)<br />

(196)<br />

The proof is the fol<strong>low</strong>ing using the above simple commut<strong>at</strong>ors: At equal times<br />

and with C = Q V a ,A = ¯ψ,B = λ b ψ one obtains from [C,AB] = [C,A] B +<br />

A[C,B]<br />

[<br />

Q<br />

V<br />

a , ¯ψ(y)λ b ψ(y) ] = [ Q V a , ¯ψ(y) ] λ b ψ(y) + ¯ψ(y) [ Q V a ,λ b ψ(y) ]<br />

= ¯ψ(y) λ a<br />

2 λ bψ(y) − ¯ψ(y)λ b<br />

λ a<br />

2 ψ(y)<br />

= 2 ¯ψ(y)[ λ a<br />

2 , λ b<br />

2 ]ψ(y) = λ c<br />

ifabc ¯ψ(y) ψ(y) qed<br />

2<br />

The proof for the second formula is the fol<strong>low</strong>ing: At equal times and with<br />

C = Q V a ,A = ¯ψ,B = λ b γ 5 ψ one obtains from [C,AB] = [C,A] B + A[C,B] the<br />

94


fol<strong>low</strong>ing equ<strong>at</strong>ions:<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)λ b γ 5 ψ(y) ] = [ Q A a , ¯ψ(y) ] λ b γ 5 ψ(y) + ¯ψ(y) [ Q A a ,λ b γ 5 ψ(y) ]<br />

= − ¯ψ(y) λ a<br />

2 γ5 λ b γ 5 ψ(y) − ¯ψ(y)λ b<br />

λ a<br />

2 γ5 γ 5 ψ(y)<br />

= −2 ¯ψ(y){ λ a<br />

2 , λ b<br />

2 }ψ(y)<br />

= − 2 3 δ ab ¯ψ(y)ψ(y) − d abc ¯ψ(y)λc ψ(y) qed<br />

if one now inserts the explicit expression for e.g. λ 1 and with the only<br />

non-zero relevant coefficient d 118 = 1 √<br />

3<br />

one obtains the above assertion:<br />

[<br />

Q<br />

A<br />

1 , ¯ψλ 1 γ 5 ψ ]<br />

⎛<br />

= − 2 3 ( ¯ψ u ψ u + ¯ψ d ψ d + ¯ψ s ψ s ) − d 118 ( ¯ψ u , ¯ψ s , ¯ψ s ) √ 1 ⎝<br />

3<br />

1 0 0<br />

0 1 0<br />

0 0 −2<br />

= − 2 3 ( ¯ψ u ψ u + ¯ψ d ψ d + ¯ψ s ψ s ) − ( ¯ψ u , ¯ψ s , ¯ψ s ) 1 3 ( ¯ψ u ψ u + ¯ψ d ψ d − 2 ¯ψ s ψ s )<br />

= −( ¯ψ u ψ u + ¯ψ d ψ d )<br />

⎞<br />

⎠<br />

⎛<br />

⎝<br />

⎞<br />

ψ u<br />

ψ d<br />

⎠<br />

ψ s<br />

.<br />

We need in the fol<strong>low</strong>ing the special cases of this<br />

[<br />

Q<br />

A<br />

1 , ¯ψγ 5 λ 1 ψ ] = − ( ¯ψu ψ u + ¯ψ d ψ d<br />

)<br />

[<br />

Q<br />

A<br />

8 , ¯ψγ 5 λ 8 ψ ] = − 1 3<br />

( ¯ψu ψ u + ¯ψ d ψ d − 4 ¯ψ s ψ s<br />

)<br />

(197)<br />

As example we give the deriv<strong>at</strong>ion of the first of them. When we inserts the<br />

explicit expression for e.g. λ 1 and with the only non-zero relevant coefficient<br />

d 118 = √ 1<br />

3<br />

one obtains:<br />

[<br />

Q<br />

A<br />

1 , ¯ψλ 1 γ 5 ψ ]<br />

⎛<br />

= − 2 3 ( ¯ψ u ψ u + ¯ψ d ψ d + ¯ψ s ψ s ) − 2d 118 ( ¯ψ u , ¯ψ s , ¯ψ 1<br />

s )<br />

2 √ ⎝<br />

3<br />

1 0 0<br />

0 1 0<br />

0 0 −2<br />

= − 2 3 ( ¯ψ u ψ u + ¯ψ d ψ d + ¯ψ s ψ s ) − ( ¯ψ u , ¯ψ s , ¯ψ s ) 1 3 ( ¯ψ u ψ u + ¯ψ d ψ d − 2 ¯ψ s ψ s )<br />

= −( ¯ψ u ψ u + ¯ψ d ψ d )<br />

One can also immedi<strong>at</strong>ely derive the formula<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)ψ(y) ] = − ¯ψ(y)λ a γ 5 ψ(y)<br />

⎞<br />

⎠<br />

⎛<br />

⎝<br />

⎞<br />

ψ u<br />

ψ d<br />

⎠<br />

ψ s<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)γ 5 ψ(y) ] = − ¯ψ(y)λ a ψ(y)<br />

95


Proof: We use [C,AB] = [C,A] B + A[C,B] with C = Q A a ,A = ¯ψ(y),B = ψ(y)<br />

and obtain immedi<strong>at</strong>ely with the above formulae<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)ψ(y) ] = [ Q A a , ¯ψ(y) ] ψ(y) + ¯ψ(y) [ Q A a ,ψ(y) ]<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)ψ(y) ] = − ¯ψ(y)γ 5 λ a<br />

2 ψ1944scri<br />

y) − ¯ψ(y) λ a<br />

2 γ5 ψ(y)<br />

= − ¯ψ(y)λ a γ 5 ψ(y)<br />

Actually, in his business the fol<strong>low</strong>ing formula is useful (without proof, which<br />

however can be obtained with the above techniques):<br />

[Aλ a ,Bλ b ] = 1 2 {A,B}[λ a,λ b ] + 1 2 [A,B]{λ a,λ b }<br />

We will need also the fol<strong>low</strong>ing formulae (preliminary)<br />

We have to proof the above expression of the double commut<strong>at</strong>or (do not<br />

use it,<br />

Σ ab<br />

πN = − [ Q a A, [ Q b A, L M<br />

]]<br />

Proof: We have [C,AB] = [C,A] B + A[C,B] and C = Q b A ,A = ¯ψ,B = ψand<br />

[Q A a ,ψ] = − λa<br />

2 γ5 ψand [Q A a , ¯ψ] 5 λa<br />

= − ¯ψγ<br />

2<br />

and hence<br />

[<br />

Q<br />

a<br />

A , [ Q b A, ¯ψψ ]] = [Q a A,[C,AB]] = [Q a A,[C,A]B + A[C,B]]<br />

= [Q a A,[Q b A, ¯ψ]ψ + ¯ψ[Q b A,ψ] ]<br />

= −[Q a A, ¯ψλ b γ 5 ψ]<br />

= 2 3 δ ab ¯ψ(y)ψ(y) + d abc ¯ψ(y)λc ψ(y) q.e.d.<br />

For the fol<strong>low</strong>ing we need[C,AB] = [C,A] B + A[C,B] and C = Q a A ,A =<br />

¯ψ,B = γ 5 ψ<br />

[Q a A, ¯ψ γ 5 ψ ] = [ Q a A, ¯ψ ] γ 5 ψ + ¯ψ [ Q a A,γ 5 ψ ]<br />

= − ¯ψγ 5 λ a<br />

2 γ5 ψ + ¯ψγ 5 [Q a A, ψ]<br />

= − ¯ψγ 5 λ a<br />

2 γ5 ψ + ¯ψγ 5 (− λ a<br />

2 γ5 ψ)<br />

= − ¯ψλ a ψ<br />

We have [C,AB] = [C,A] B + A[C,B] and C = Q b A ,A = ¯ψ,B = λ c ψ and<br />

[<br />

Q<br />

A<br />

a , ¯ψγ 5 ψ ] = − ¯ψλ a ψ and [ Q A a , ¯ψλ b γ 5 ψ ] = − 2 3 δ ab ¯ψψ − d abc ¯ψλ c ψ and hence<br />

96


[<br />

Q<br />

a<br />

A , [ Q b A, ¯ψλ c ψ ]] = [Q a A,[C,AB]] = [Q a A,([C,A]B + A[C,B])]<br />

= [Q a A,([Q b A, ¯ψ]λ c ψ + ¯ψ[Q b A,λ c ψ]) ]<br />

= [Q a A,(− ¯ψγ 5 λ b<br />

2 λ cψ − ¯ψ λ c<br />

λ b<br />

2 γ5 ψ) ]<br />

= −[Q a A,( ¯ψγ 5 λ b<br />

2 λ cψ + ¯ψ λ c<br />

λ b<br />

2 γ5 ψ) ]<br />

= − 1 2 [Qa A, ¯ψ{λ b , λ c }γ 5 ψ ]<br />

So altogether we have (preliminary)<br />

= − 1 2 [Qa A, ¯ψ(− 2 3 δbc )γ 5 ψ + ¯ψd bcd λ d γ 5 ψ ]<br />

= − 1 3 δbc [Q a A, ¯ψ γ 5 ψ ] − 1 2 dbcd [Q a A, ¯ψ λ d γ 5 ψ ]<br />

= − 1 3 δbc (− ¯ψλ a ψ) − 1 2 dbcd (− 2 3 δ ad ¯ψψ − d ade ¯ψλe ψ)<br />

= − 1 3 δbc (− ¯ψλ a ψ) + 1 3 dbcd δ ad ¯ψψ +<br />

1<br />

2 dbcd d ade ¯ψλe ψ<br />

= 1 6 ¯ψλ a ψδ bc + 1 3 dbca ¯ψψ +<br />

1<br />

2 dbcd d ade ¯ψλe ψ<br />

[<br />

Q<br />

a<br />

A , [ Q b A, ¯ψλ c ψ ]] = 1 6 ¯ψλ a ψδ bc + 1 3 dbca ¯ψψ +<br />

1<br />

2 dbcd d ade ¯ψλe ψ q.e.d.<br />

10.2 Spontaneous symmetry breaking<br />

We remember the situ<strong>at</strong>ion of a spontaneously broken symmetry without explicit<br />

breaking. There we have shown th<strong>at</strong> we must have an excited st<strong>at</strong>e of<br />

the vacuum with some particular properties, i.e. Goldstone bosonic st<strong>at</strong>e. We<br />

consider SU(2) and will call this st<strong>at</strong>e ∣ ∣ π(p)<br />

〉<br />

and have then<br />

m π = 0<br />

〈0|π j (0) ∣ ∣ π k (p) 〉 = aδ jk<br />

〈0|A j µ(0) ∣ ∣ π k (p) 〉 = bδ jk<br />

〈0|∂ µ A j µ(0) ∣ ∣ π k (p) 〉 = 0<br />

Attention: we denote field-oper<strong>at</strong>ors by π k and the one boson field st<strong>at</strong>es by<br />

∣<br />

∣π k〉 . The first equ<strong>at</strong>ion is merely an equ<strong>at</strong>ion which denotes the normaliz<strong>at</strong>ion<br />

of the boson field st<strong>at</strong>es, and we take for convenienct a = 1.We see wh<strong>at</strong> th<strong>at</strong><br />

97


means if we take the expression of the quantized boson field (32) and use it <strong>at</strong><br />

x = 0<br />

∫<br />

d 3 k<br />

[<br />

φ(0) = √ a(k) + a(k) ∗]<br />

(2π)3 2ω k<br />

The a = 1 corresponds to<br />

〈0| π j (0) ∣ ∣ π k (p) 〉 = δ jk<br />

To get this the one-pion st<strong>at</strong>e must be normalized with a(p) |0 pion 〉 = 0 as<br />

√<br />

|π(p)〉 = 2ω p (2π) 3 a(p) † |0〉<br />

or equivalently the overlap of two pion st<strong>at</strong>es is<br />

〈<br />

π k (p) ∣ ∣ π j (p ′ ) = 2ω p (2π) 3 δ (3) (p − p ′ )δ jk<br />

Because of the Lorentz structure of the LHS the b of the second equ<strong>at</strong>ion can<br />

be rewritten as<br />

〈0|A j µ(0) ∣ ∣ π k (p) 〉 = if π p µ δ jk (198)<br />

The f π is the pion decay constant. It is not obvious <strong>at</strong> all th<strong>at</strong> the f π is indeed<br />

the pion decay constant. However, why this is so, this will be discussed in the<br />

next subsection where we explicitely calcul<strong>at</strong>e the pion decay.<br />

The fact th<strong>at</strong> we have spontaneous chiral symmetry breaking has some very<br />

simple but also very basic consequences: We know th<strong>at</strong> the vacuum has the<br />

properties<br />

Q V a |0 >= 0 Q A a |0 >≠ 0<br />

from which one can immeditely conclude using eq.(194) th<strong>at</strong><br />

〈0| [ Q V a , ¯ψ(y)λ b ψ(y) ] |0〉 = if abc 〈0| ¯ψ(y) λ c<br />

ψ(y) |0〉 = 0<br />

2<br />

If one uses all combin<strong>at</strong>ions of a and b and takes the few f abc which are nonzero<br />

() one obtains immedi<strong>at</strong>ely the fol<strong>low</strong>ing st<strong>at</strong>ements about the vacuum<br />

condens<strong>at</strong>es:<br />

〈0| ¯ψ i ψ k |0〉 = 0 for i ≠ k (199)<br />

and<br />

〈0| ¯ψ u ψ u |0〉 = 〈0| ¯ψ d ψ d |0〉 = 〈0| ¯ψ s ψ s |0〉 (200)<br />

One obtains from Q A a |0 >≠ 0 and<br />

[<br />

Q<br />

A<br />

1 , ¯ψγ 5 λ 1 ψ ] = − ( ¯ψu ψ u + ¯ψ d ψ d<br />

)<br />

[<br />

Q<br />

A<br />

8 , ¯ψγ 5 λ 8 ψ ] = − 1 3<br />

( ¯ψu ψ u + ¯ψ d ψ d − 4 ¯ψ s ψ s<br />

)<br />

the fact th<strong>at</strong> the fol<strong>low</strong>ing condens<strong>at</strong>es exist<br />

〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉 ≠ 0<br />

〈0| ¯ψ u ψ u + ¯ψ d ψ d − 4 ¯ψ s ψ s |0〉 ≠ 0<br />

98


From these formulae we conclude th<strong>at</strong> all three condens<strong>at</strong>es are identical and<br />

non-zero. As we shall see l<strong>at</strong>er in the section about PCAC and Gell-Mann-<br />

Oakes-Renner we have<br />

〈0| ¯ψ u ψ u |0〉 = 〈0| ¯ψ d ψ d |0〉 = 〈0| ¯ψ s ψ s |0〉 = −(225 ± 25) 3 MeV 3 (201)<br />

The fact those condens<strong>at</strong>es tells a lot about the vacuum |0〉 . We know the<br />

expression for the fermion field (39). Due to transl<strong>at</strong>ional invariance of the<br />

vacuum we are al<strong>low</strong>ed to consider<br />

ψ(0) = ∑<br />

r=1,2<br />

∫<br />

d 3 k<br />

√<br />

(2π)<br />

3<br />

and hence we have with normal ordering<br />

〈0| ¯ψ u ψ u |0〉 = ∑ rs<br />

[<br />

]<br />

c r (k)u r (k) + d † r(k)v r (k) : |0〉<br />

√ m<br />

[<br />

]<br />

c r (k)u r (k) + d †<br />

E<br />

r(k)v r (k)<br />

k<br />

∫ d 3 kd 3 k ′ √ m m<br />

[<br />

]<br />

(2π) 3 〈0| : c †<br />

E k E<br />

r(k ′ )ū r (k ′ ) + d r (k ′ )¯v r (k ′ )<br />

k ′<br />

Suppose ∣ ∣˜0 〉 is be the vacuum of the oper<strong>at</strong>ors c r and d r with c r<br />

∣ ∣˜0 〉 = 0 and<br />

d r<br />

∣ ∣˜0 〉 = 0 then the above oper<strong>at</strong>or expression sandwiched between ∣ ∣˜0 〉 would<br />

be obviously zero. Since it does not vanish we have |0〉 ≠ ∣ ∣˜0 〉 , which means th<strong>at</strong><br />

|0〉 must be vacuum for another set of annihil<strong>at</strong>ion oper<strong>at</strong>ors. The particles<br />

which correspond to these new single particle oper<strong>at</strong>ors are constituent quarks,<br />

in contrast to the particles corresponding to the the oper<strong>at</strong>ors c r and d r in the<br />

quantized field expressions, which are the <strong>QCD</strong>-quarks or current quarks. The<br />

l<strong>at</strong>ter name is obvious since their field oper<strong>at</strong>ors ψ constitute the currents e.g.<br />

A a µ(x) we are dealing within our current algebra.<br />

10.3 Pion decay<br />

Looking <strong>at</strong> the formulae relevant for the spontanesous breakdown of the chiral<br />

symmetry we find th<strong>at</strong> the pion decay constant f π plays a dominant role. Thus<br />

we consider now the pion decay in n<strong>at</strong>ure. The purpose is to extract the f π<br />

from experimental d<strong>at</strong>a.<br />

The charged pions decay to 99% via the fol<strong>low</strong>ing channel<br />

π + → µ + ν µ π − → µ − ν µ (202)<br />

The neutral pion decays to 99% into two photons via the chiral anomaly<br />

of the axial current (163). The decay of the charged pion is induced by the<br />

electroweak interaction Lagrangean. At a scale of the W + -boson this can be<br />

written as<br />

99


u<br />

µ+<br />

W+<br />

d<br />

bar<br />

ν<br />

bar<br />

Electroweak interaction<br />

For Feynman-diagrams one should recall the Feynman amplitude of the process<br />

e + e − → µ + µ − which is given by<br />

M (2) (e + e − → µ + µ − ) = −iū (µ) (p ′ 2)γ α v (µ) (p ′ 1)D αβ<br />

F (p1 + p 2 )¯v (e) (p 1 )γ β u (e) (p 2 )<br />

For the present process the current ū (µ) (p ′ 2)γ α v (µ) is somewh<strong>at</strong> different since<br />

we have a different interaction, but the structure current-propag<strong>at</strong>or-current is<br />

the same. In the present case of pion decay one one needs the propag<strong>at</strong>or of the<br />

W-boson, which is given by<br />

(<br />

iD αβ<br />

−g αβ<br />

F (k,m + k α k β /m 2 )<br />

W<br />

W) = i<br />

k 2 − m 2 W + iε<br />

Since the W + 1<br />

is heavy (80 GeV) the propag<strong>at</strong>or of it is merely since its<br />

m 2 W<br />

momentum dependence k = p 1 + p 2 can be ignored <strong>at</strong> <strong>low</strong> <strong>energies</strong>. If we<br />

denote the fermion-W coupling constant by g ′ (analogous to e in the above<br />

) 2<br />

process) the diagram is proportional to (g′ , which is called in the definition<br />

of the interaction Lagrangean<br />

(<br />

g<br />

2 √ 2<br />

m 2 W<br />

) 2<br />

.With the definition of GF as GF √<br />

2<br />

=<br />

g 2<br />

one obtains with the charged currents (186192) the corresponding effective<br />

8MW<br />

2<br />

Hamiltonian:<br />

H weak = G F<br />

√<br />

2<br />

[J µ had,CC + Jµ lep,CC ][Jµ had,CC + Jµ lep,CC ]† (203)<br />

This is the famous current-current interaction. Altogether, if we concentr<strong>at</strong>e<br />

on the decay of the π + into a pair of leptons we have to consider the m<strong>at</strong>rix<br />

element<br />

M = 〈0 hadrons | 〈 l + ν l<br />

∣ ∣ Hweak<br />

∣ ∣π + 〉 |0 leptons 〉<br />

with |0〉 = |0 hadros 〉 |0 leptons 〉. The only possible combin<strong>at</strong>ion of fields is<br />

here:<br />

H weak = G F<br />

√<br />

2<br />

V ud ¯ψd γ λ (1 − γ 5 )ψ u<br />

[ ¯ψνe γ λ (1 − γ 5 )ψ e + (e ↔ µ) ]<br />

100


with ( in another nomencl<strong>at</strong>ure) V ud = cos θ C = 0.975.This can be diagramm<strong>at</strong>ically<br />

represented as:<br />

u<br />

µ+<br />

d<br />

bar<br />

ν bar<br />

Fermi−interaction (weak)<br />

The Lagrangean is now considered <strong>at</strong> a <strong>low</strong> scale and hence the value of G F is<br />

taken from the decay of the free neutron and given by G F = 1.17⋆10 −5 GeV −2 .<br />

Apparently the decay of the pion is governed by this very small coupling<br />

constant, whereas the decay of the π 0 is govened by the electric coupling constant<br />

α = 1<br />

137 , which is much larger. Thus we expect the liftetime of the π+ to be<br />

much larger than th<strong>at</strong> of π 0 . This is indeed the case as the experimental numbers<br />

indic<strong>at</strong>e:<br />

τ(π + ) = 2.6 ∗ 10 −8 s<br />

τ(π 0 ) = 8.4 ⋆ 10 −17 s<br />

The leptonic part of the above m<strong>at</strong>irxelement is<br />

〈<br />

l + ν l<br />

∣ ∣ Hweak |0 leptons 〉 = 〈 l + ν l<br />

∣ ∣ ¯ψνe γ λ (1 − γ 5 )ψ e |0 leptons 〉<br />

= ū ν γ λ (1 − γ 5 )v µ<br />

This cannot be evalu<strong>at</strong>ed further since the neutrino and the lepton are physical<br />

particles leaving the interaction region and being detected somewhere. The<br />

hadronic part of this m<strong>at</strong>rixelement is<br />

〈0 hadrons | H weak<br />

∣ ∣π + 〉 = G F<br />

√<br />

2<br />

V ud<br />

1 √2 〈0 hadrons | ¯ψ d γ λ (1 − γ 5 )ψ u [ ∣ ∣ π 1 (p) + iπ 2 (p) 〉 ]<br />

To evalu<strong>at</strong>e the non-trivial quark part one should use the fol<strong>low</strong>ing trivial identities:<br />

¯ψ d γ λ ψ u = V1 λ − iV2<br />

λ<br />

¯ψ d γ λ γ 5 ψ u = A λ 1 − iA λ 2<br />

The transition from the vacuum to the pion via the vector current vanishes,<br />

because the pion is pseudoscalar and the current is a vector. For, if we denote the<br />

101


parity transform<strong>at</strong>ion by P, we have 〈0 hadrons |P † = 〈0 hadrons |and P ∣ ∣ π k (p) 〉 =<br />

− ∣ ∣ π k (p) 〉 and hence:<br />

〈0 hadrons |V j µ(0) ∣ ∣ π k (p) 〉 = 〈0 hadrons |PV j µ(0)P † P ∣ ∣ π k (p) 〉 = − 〈0 hadrons |V j µ(0) ∣ ∣ π k (p) 〉<br />

since. Thus for the strong part of the transition m<strong>at</strong>rixelement the m<strong>at</strong>rixelement<br />

of the axial current is relevant, which we of course know (198). Thus we<br />

have<br />

Collecting terms we find<br />

〈0 hadrons |V j µ(0) ∣ ∣ π k (p) 〉 = 0<br />

〈0 hadrons |A j µ(0) ∣ ∣ π k (p) 〉 = if π p µ δ jk<br />

〈0 hadrons | H weak<br />

∣ ∣π + 〉 = G F<br />

√<br />

2<br />

V ud<br />

1 √2 〈0 hadrons |[A λ 1 − iA λ 2][ ∣ ∣ π 1 (p) + iπ 2 (p) 〉 ]<br />

= G F 1<br />

√ V ud √2 2if π p λ<br />

2<br />

This formula contains the f π simply as a number, which characterizes the<br />

transition m<strong>at</strong>rixelement from vacuum to pion via the axial current. For the<br />

present process this is the only unknown quantity coming from the strong part.<br />

Apparently this number can be fixed here by reproducing the decay of the real<br />

pion. In the present nomencl<strong>at</strong>ure the f π is defined such th<strong>at</strong> its experimental<br />

value will be 93MeV . The decay has then the invariant Feynman amplitude<br />

T π+ →µ + ν µ<br />

= G F<br />

√<br />

2<br />

V ud<br />

√<br />

2fπ p λ ū ν γ λ (1 − γ 5 )v µ<br />

Using now the Dirac equ<strong>at</strong>ion (γ µ p µ − m)u(p) = 0 one can simplify<br />

getting the final expression<br />

p λ ū ν γ λ (1 − γ 5 ) = mū ν (1 − γ 5 )<br />

T π+ →µ + ν µ<br />

= −G F V ud f π m µ ū ν (1 − γ 5 )v µ<br />

An analogous expression holds for the decay π − → µ − ν µ .<br />

One sees here the well known helicity suppression phenomenon. Besides the<br />

decay of the pion into muons (202) one also has decay into electrons like<br />

π + → e + ν e<br />

π − → e − ν e<br />

This decay is very much suppresed compared to the muon decay. The reason is,<br />

th<strong>at</strong> the weak leptonic current contains the left-handed chiral projection oper<strong>at</strong>or<br />

(1−γ 5 ) which in the limit of massles leptons produces only left-handed particles<br />

and right-handed antiparticles. In this limit “lefthandedness” means “neg<strong>at</strong>ive<br />

helicity” i.e. spin and momentum antiparallel, and “right-handedness”<br />

102


means “positive helicity” i.e. spin and momentum parallel. If a pion in rest decays,<br />

the resulting leptons must have opposite momenta. Suppose e.g. th<strong>at</strong> in<br />

the decay of π − the outgoing muon runs into the right direction. It is a particle,<br />

has neg<strong>at</strong>ive helicity, and so its spin points towards left. The antineutrino runs<br />

into the left direction in order to conserve momentum, as antiparticle it must<br />

have positive helicity and hence its spin points towards left. Thus the spins of<br />

the muon and antineutrino add up to ONE pinting left. Th<strong>at</strong> contradicts the<br />

fact th<strong>at</strong> the pion has ZERO spin. Hence the pion decay is only possible because<br />

the electron and myon are massive. Since the muon is 200 times more massive<br />

than the electron the decay goes primarily through the muon channel.<br />

Altogether one obtains for the decay r<strong>at</strong>e<br />

Γ π+ →µ + ν µ<br />

= G2 F<br />

4π |V ud| 2 fπm 2 2 µm π (1 − m2 µ<br />

)<br />

Before one can compare this with experiment one has to perform some radi<strong>at</strong>ive<br />

corrections. Taking in the end V ud from the Kobayashi-Maskawa-M<strong>at</strong>rix and<br />

using the experimental number<br />

one obtains<br />

Γ π+ →µ + ν µ<br />

= 3.841x10 7 s −1<br />

m 2 π<br />

f π = (92.4 ± 0.2)MeV (204)<br />

It is important to note, th<strong>at</strong> the f π appears as number chacterizing the decay<br />

of the pion. Simultaneously it characterzes the transition m<strong>at</strong>rix element of the<br />

axial current between vacuum and pion. The reason for this is: The pion is a<br />

Goldstone boson, therefore This is a strong interaction m<strong>at</strong>rix element.<br />

10.4 PCAC: partial conserv<strong>at</strong>ion of axial current<br />

The concept of PCAC is very important, since it provides a direct link between<br />

the <strong>QCD</strong> (quark fields) and effective theories (pion fields). The final PCAC<br />

formula is given by<br />

∂ µ A j µ(x) = m 0 ¯ψ(x)iγ 5 τ j ψ(x) = f π m 2 ππ j (x) (205)<br />

This equ<strong>at</strong>ion tells clearly th<strong>at</strong> the pion field is identical to a certain combin<strong>at</strong>ion<br />

of quark and antiquark fields, it is only another name. This has the<br />

consequence th<strong>at</strong> it makes sense to forget the quark fields and to consider only<br />

pionic fields. This is the philosophy of effecive theories. The main purpose for<br />

the effective theories is, to derive equ<strong>at</strong>ions for the π-field itself (meson field), by<br />

this avoiding to consider equ<strong>at</strong>ions for the ψ-field (fermion field), in such a way<br />

th<strong>at</strong> S-m<strong>at</strong>rix elements or more generally observables in the effective theory and<br />

in <strong>QCD</strong> are very similar.. Effective theories provide often advantages because<br />

they are often simpler and nevertheless hit the correct physics. We know such<br />

a theory, it is the chiral sigma model with Gell-Mann-Levy lagrangean. We will<br />

103


learn about others, for instance the model independent lagrangean used e.g. in<br />

chiral perturb<strong>at</strong>ion theory, or l<strong>at</strong>er on the Skyrme model, etc..<br />

For the proof of PCAC (205) we calcul<strong>at</strong>e the divergence of the axial vector<br />

current. In fact we had from spontaneous chiral symmetry breaking the basic<br />

rel<strong>at</strong>ion (198) In order to calcul<strong>at</strong>e now the m<strong>at</strong>rixelement of the divergence of<br />

the current explicitey we use the shift oper<strong>at</strong>or (34)<br />

〈0|A j µ(0) ∣ ∣ π k (p) 〉 = 〈0| exp(−iPx)A j µ(x)exp(+iPx) ∣ ∣ π k (p) 〉 = 〈0|A j µ(x) ∣ ∣ π k (p) 〉 exp(+ipx)<br />

or<br />

〈0|A j µ(x) ∣ ∣π k (p) 〉 = 〈0|A j µ(0) ∣ ∣π k (p) 〉 exp(−ipx) = if π p µ δ jk exp(−ipx)<br />

and hence<br />

〈0|∂ µ A j µ(x) ∣ ∣π k (p) 〉 = (∂ µ exp(−ipx))if π p µ δ jk = f π p µ p µ δ jk exp(−ipx) = f π m 2 πδ jk exp(−ipx)<br />

Apparently, for a conserved axial current (CAC) we have (this is an ideal case<br />

of course)<br />

conserved axial current ⇒ m 2 π = 0 and/or f π = 0<br />

The fact is, th<strong>at</strong> m 2 π = 0 agrees with having a massless Goldstone boson. Thus<br />

the pion decay constant can still be non-zero, as it is in n<strong>at</strong>ure and as it is<br />

required by sppontaneous symmetry breaking which requires a nonzere m<strong>at</strong>rixelement<br />

between the vacuum and the goldstone st<strong>at</strong>e.<br />

One can generalize this consider<strong>at</strong>ions to the case of non-vanishing divergence<br />

of the axial current because of the finite masses of the quarks. This<br />

comes to the concept of PCAC (partially conserved axial current). We just<br />

discussed the divergence of the axial current and can write it down again<br />

〈0|∂ µ A j µ(0) ∣ ∣ π k (p) 〉 = f π m 2 π 〈0| π j (0) ∣ ∣ π k (p) 〉<br />

Where we have replaced δ jk by the normaliz<strong>at</strong>ion expression of the pion field<br />

〈0|π j (0) ∣ ∣ π k (p) 〉 = δ jk .On the other hand we know from the section about <strong>QCD</strong><br />

with m 0 = mu+m d<br />

2<br />

:<br />

∂ µ A j µ(x) = m 0 ¯ψ(x)iγ 5 τ j ψ(x)<br />

from which we can take the m<strong>at</strong>rixelement<br />

and obtain by comparison<br />

〈0|∂ µ A j µ(0) ∣ ∣ π k (p) 〉 = m 0 〈0| ¯ψ(0)iγ 5 τ j ψ(0) ∣ ∣ π k (p) 〉<br />

m 0 〈0| ¯ψ(0)iγ 5 τ j ψ(0) ∣ ∣ π k (p) 〉 = f π m 2 π 〈0|π j (0) ∣ ∣ π k (p) 〉<br />

or generalized to an oper<strong>at</strong>or equ<strong>at</strong>ion and using transl<strong>at</strong>ion invariance of the<br />

vacuum<br />

∂ µ A j µ(x) = m 0 ¯ψ(x)iγ 5 τ j ψ(x) = f π m 2 ππ j (x)<br />

This completes the proof of the PCAC rel<strong>at</strong>ion.<br />

104


10.5 The chiral condens<strong>at</strong>e<br />

We can now easily derive the existence of vacuum condens<strong>at</strong>es and their values.<br />

We will prove the famous Gell-Mann–Oakes–Renner formula,<br />

f 2 πm 2 π = − 1 2 (m u + m d ) 〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉<br />

which rel<strong>at</strong>es observable quantities like pion decay constant and pion mass to<br />

<strong>QCD</strong>-quantities like the current masses of the quarks and the condens<strong>at</strong>es.We<br />

will derive from this the value of the condens<strong>at</strong>e (201). The Gell-Mann–Oakes–<br />

Renner formula makes a connection between the properties of the vacuum,<br />

i.e. condens<strong>at</strong>e, the properties of the quarks in the <strong>QCD</strong>-Lagrangean, i.e. the<br />

masses, and properties of the physical and observable hadronic properties, i.e.<br />

the mass and decay constant or the pion, being the goldstone boson of the<br />

spontaneously broken chiral symmetry of the vacuum. In fact we have a huge<br />

condens<strong>at</strong>e, connecting a number of the order of 10 8 MeV 4 on the LHS to 10MeV<br />

on the RHS. The condens<strong>at</strong>e is connected with an energy density, which is about<br />

1000 MeV per fm 3 .This energy has been released in the chiral phase transition<br />

about 10 −6 sec after the Big Bang, where the Universe changed from something<br />

like a quark-gluon plasma to our present hadronic phase. The energy was released<br />

in terms of pions, which radi<strong>at</strong>ed into photons and W-bosons. The chiral<br />

phase transition occured because the universe expanded and cooled such th<strong>at</strong> a<br />

hadronic vacuum st<strong>at</strong>e got energetically preferable.<br />

For the proof we start from the divergence of the axial current (166)<br />

∂ µ A 1 µ = m 0 ¯ψiγ 5 τ 1 ψ = 1 2 (m u + m d ) ¯ψiγ 5 λ 1 ψ<br />

We know from charge algebras the formula (197)<br />

[<br />

Q<br />

A<br />

1 , ¯ψγ 5 λ 1 ψ ] = − ( ¯ψu ψ u + ¯ψ d ψ d<br />

)<br />

and can replace the RHS of the commut<strong>at</strong>or yielding<br />

〈0| [ Q A 1 ,∂ µ A 1 ] i<br />

µ |0〉 = −<br />

2 (m u + m d ) 〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉<br />

We now insert a complete set of st<strong>at</strong>es into the commut<strong>at</strong>or, which however<br />

fol<strong>low</strong>ing the philosophy of spontaneous symmetry breaking, is exausted solely<br />

by the pion st<strong>at</strong>e (<strong>at</strong>tention: sum over a because the pionic goldstone st<strong>at</strong>e is<br />

an iso-triplet and there are in SU(3) also the kaon and the eta as Goldstone<br />

st<strong>at</strong>es).<br />

∑<br />

∫<br />

d 3 p<br />

2E p (2π) 3 |π a(p) >< π a (p)| = 1<br />

a<br />

105


In this way we obtain (only the term with a = 1 remains):<br />

− ∑ a<br />

∫<br />

d 3 p<br />

2E p (2π) 3 〈0| ∂µ A 1 µ(0)|π a (p) >< π a (p)|Q A 1 |0〉<br />

< 0| [ Q A 1 ,∂ µ A 1 µ(0) ] |0 ><br />

= ∑ ∫<br />

d 3 p<br />

2E<br />

a p (2π) 3 〈0| QA 1 |π a (p) >< π a (p)|∂ µ A 1 µ(0) |0〉<br />

One sees <strong>at</strong> this formula immedi<strong>at</strong>ely th<strong>at</strong> there is no contribution from the<br />

vacuum |0〉 ,which is of course also part of the complete set, since both terms<br />

are identical in this case. To calcul<strong>at</strong>e the m<strong>at</strong>rixelements in this expresseion<br />

we have used the above m<strong>at</strong>rixelement of the axial current between vacuum and<br />

pion st<strong>at</strong>e, being the definition of the pion decay constant,<br />

〈0| A j µ(x) ∣ ∣π k (p) 〉 = if π p µ δ jk exp(−ipx)<br />

and obtain from th<strong>at</strong><br />

〈0| Q j ∣<br />

A π k (p) 〉 ∫<br />

=<br />

d 3 x 〈0| A j 0 (x)∣ ∣ π k (p) 〉 = if π p 0 δ jk ∫<br />

d 3 xexp(−ipx)<br />

yielding<br />

Furthermore we use<br />

〈0|Q A a |π b (p) >= if π δ ab E p (2π) 3 δ 3 (p)<br />

〈0| ∂ µ A j µ(0) ∣ ∣ π k (p) 〉 = f π m 2 πδ jk<br />

One can continue now<br />

〈0| [ Q A 1 ,∂ µ A 1 µ(0) ] |0〉 =<br />

∑<br />

∫<br />

d 3 p<br />

2E<br />

a p (2π) 3 [if πE p (2π) 3 δ 3 (p) f π m 2 πδ a1<br />

− ∑ ∫<br />

d 3 p<br />

2E<br />

a p (2π) 3 f πm 2 πδ a1 (−i)f π E p (2π) 3 δ 3 (p)<br />

= if 2 πm 2 π<br />

The integral gives a ONE, lots of terms cancel, and we obtain the famous Gell-<br />

Mann–Oakes–Renner formula,<br />

f 2 πm 2 π = − 1 2 (m u + m d ) 〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉<br />

which rel<strong>at</strong>es observable quantities like pion decay constant and pion mass to<br />

<strong>QCD</strong>-quantities like the current masses of the quarks and the condens<strong>at</strong>es. One<br />

should note th<strong>at</strong> the current masses and the condens<strong>at</strong>e separ<strong>at</strong>ely depend on<br />

the renormaliz<strong>at</strong>ion scale, their product, however does not.:<br />

106


Using f π = 93MeV and m π = 139MeV and m u +m d ≈ 14MeV (we learn<br />

l<strong>at</strong>er how to get the masses of the current quarks) one obtains eq.(201)<br />

〈0| ¯ψ u ψ u |0〉 = 〈0| ¯ψ d ψ d |0〉 = 〈0| ¯ψ s ψ s |0〉 = −(225MeV ) 3 = −1.5fm −3<br />

This completes the proof of the Gell-Mann–Oakes–Renner formula and of<br />

the magnitude of the chiral condens<strong>at</strong>e.<br />

10.6 LSZ-Reduction formulae<br />

This section summarizes some useful formulae. In the fol<strong>low</strong>ing one normalizes<br />

the pion field as mentioned above. Then the Fourier transform of the two-point<br />

correl<strong>at</strong>ion function, considered as an analytic function of p 2 , has a simple pole<br />

<strong>at</strong> the mass of the one-particle st<strong>at</strong>e. Thus in the limit p 2 → m 2 a one can write<br />

(no sum over a):<br />

∫<br />

d 4 xexp(ipx) 〈0|T {π a (0)π a i<br />

(x)} |0〉 =<br />

p 2 − m 2 a + iɛ<br />

For the fol<strong>low</strong>ing we need some formulae, which are known under the name<br />

Lehman-Symanzik-Zimmerman reduction formulae, the deriv<strong>at</strong>ion of which is<br />

based on basic S-m<strong>at</strong>rix properties and one can find e.g. in Peskin-Schroeder.<br />

The formulae involve the invariant amplitude and physical particles, ie. those<br />

which are on the mass shell, i.e. q 2 1 = q 2 2 = m 2 π, and p 2 1 = p 2 2 = m 2 N . Instead<br />

of the nucleon st<strong>at</strong>es we can have also other st<strong>at</strong>es of physical particles, e.g.<br />

the vacuum as well. In the formulae there appear in- and out-st<strong>at</strong>es. They<br />

are rel<strong>at</strong>ivistic analogues of the sc<strong>at</strong>tering st<strong>at</strong>es of non-rel<strong>at</strong>ivistic quantum<br />

mechanics, which correspond e.g. to outgoing waves from the interaction region.<br />

Just to remind, each of these sets separ<strong>at</strong>ely are complete. We remember:<br />

Ψ out = exp(ikr) + f(φ) exp(+ikr)<br />

r<br />

Ψ in = exp(ikr) + f(φ) exp(−ikr)<br />

r<br />

We have the Lehmann-Symanzik-Zimmermann reduction formula (LSZ):<br />

< π a (q 2 )N(p 2 );out ∣ N(p1 );in 〉 = i(2π) 4 δ (4) (p 1 − p 2 − q 2 )TN→πN<br />

b<br />

< π b (q 2 )N(p 2 );out ∣ N(p1 );in 〉 ∫<br />

= i d 4 y exp(iq 2 y)(−q2+m 2 2 π) 〈 N(p 2 ) ∣ π b (y) ∣ N(p1 ) 〉<br />

and the correponding process with q 2 1 = m 2 π, and p 2 1 = p 2 2 = m 2 N<br />

< N(p 2 );out ∣ π a (q 1 )N(p 1 );in 〉 = i(2π) 4 δ (4) (p 1 + q 1 − p 2 )TπN→N<br />

a<br />

< N(p 2 );out ∣ π a (q 1 )N(p 1 );in 〉 ∫<br />

= i d 4 y exp(−iq 1 y)(−q1+m 2 2 π) 〈 N(p 2 ) ∣ π a (y) ∣ N(p1 ) 〉<br />

107


One can easily see the consistency in these formulae. Consider the fol<strong>low</strong>ing<br />

case<br />

∫<br />

i d 4 y exp(iq 2 y)(−q2 2 + m 2 π) 〈 N(p 2 ) ∣ π b (y) ∣ N(p1 ) 〉<br />

∫<br />

= i d 4 y exp(iq 2 y)(−q2 2 + m 2 π) 〈 N(p 2 ) ∣ exp(iPy)π b (0)exp(−iPy) ∣ N(p1 ) 〉<br />

∫<br />

= i d 4 y exp(iq 2 y)exp(ip 2 y)exp(−ip 1 y)(−q2 2 + m 2 π) 〈 N(p 2 ) ∣ π b (0) ∣ N(p1 ) 〉<br />

= i(2π) 4 δ (4) (p 1 − p 2 − q 2 )(−q2 2 + m 2 π) 〈 N(p 2 ) ∣ π b (0) ∣ N(p1 ) 〉<br />

and hence<br />

T b N→πN = (−q 2 2 + m 2 π) 〈 N(p 2 ) ∣ ∣ π b (0) ∣ ∣N(p 1 ) 〉<br />

T b πN→N = (−q 2 1 + m 2 π) 〈 N(p 2 ) ∣ ∣ π b (0) ∣ ∣N(p 1 ) 〉<br />

We get a reduction formula similarly for pion-nucleon sc<strong>at</strong>tering in the limit<br />

q 2 1 → m 2 a = m 2 π and q 2 2 → m 2 b = m2 π :<br />

< π b (q 2 )N(p 2 );out ∣ ∣ π a (q 1 )N(p 1 );in 〉 = i(2π) 4 δ (4) (p 1 + q 1 − p 2 − q 2 )T ab<br />

πN→πN<br />

< π b (q 2 )N(p 2 );out ∣ ∣ π a (q 1 )N(p 1 );in 〉<br />

= (i) 2 ∫<br />

d 4 xd 4 y exp(iq 2 y)exp(−iq 1 x)(−q 2 1 + m 2 a)(−q 2 2 + m 2 b) 〈 N(p 2 ) ∣ ∣ T {π b (y)π a (x)} ∣ ∣ N(p1 ) 〉<br />

and :<br />

We get a similar reduction formula :<br />

〉 ∫<br />

∣<br />

< 0 ∣Ô(0)|πa (q 1 ) = i d 4 xexp(−iq 1 x)(−q1 2 + m 2 b) 〈0|T {Ô(0)πa (x)} |0〉<br />

〉 ∫<br />

< π b ∣<br />

(q 2 ) ∣Ô(0)|0 = i<br />

d 4 y exp(iq 2 y)(−q2 2 + m 2 b<br />

b) 〈0|T {π (y)Ô(0)} |0〉<br />

The formulae are somehow understandable, since the field oper<strong>at</strong>ors consist of<br />

cre<strong>at</strong>ion and annihil<strong>at</strong>ion oper<strong>at</strong>ors of the field quanta, and the physical particles<br />

in the entrance and exit channels are cre<strong>at</strong>ed from the vacuum by applic<strong>at</strong>ion of<br />

the cre<strong>at</strong>ion oper<strong>at</strong>ors. The particles appearing in the formulae are particular<br />

since they are all on the mass shell . The detailled proof is given in field theory<br />

books (e.g. Peskin-Schroeder, or Izykson-Zuber §5-1-3) .<br />

10.7 Pion-Nucleon coupling constant<br />

Another thing is important, th<strong>at</strong> is the definition of the pion-nucleon-nucleon<br />

coupling constant.<br />

< π + (q 2 )N(p 2 ) ∣ ∣ N(p1 ) 〉 = i(2π) 4 δ (4) (p 1 − p 2 − q 2 )T + N→πN<br />

108


with the pion emission amplitude<br />

T + N→πN = i√ 2g πNN (q 2 )u(p 2 )γ 5 τ a u(p 1 )<br />

This definition implies q 2 = m 2 π. Due to symmetry consider<strong>at</strong>ions the coupling<br />

constants are independent on a. Actually one can rewrite this using the<br />

reduction theorems:<br />

(−q 2 + m 2 π) 〈p(k ′ )| π + (0) |n(k)〉 = √ 2g πNN (q 2 )u(k ′ )iγ 5 u(k)<br />

Again this implies q 2 = m 2 π, however it is generalized to the definition of the<br />

pion-nucleon-nucleon form factor. In this way the form factor g πNN (q 2 ) is defined<br />

even for space-like q 2 by analytic continu<strong>at</strong>ion. This is possible since<br />

the pion mass is small. Experimentally the pion-nucleon coupling constant is<br />

measured in pion-nucleon sc<strong>at</strong>tering and nucleon-nucleon sc<strong>at</strong>tering in forward<br />

direction. The varous kinem<strong>at</strong>ical regimes are connected by means of analytic<br />

continu<strong>at</strong>ion. Altogether one has experimentally<br />

with<br />

g πNN = g πNN (q 2 = m 2 π)<br />

g πNN = 14.6<br />

The value is not extremely accur<strong>at</strong>e since one must have pions as secondary<br />

beam or must look <strong>at</strong> NN-sc<strong>at</strong>tering. Pions are gener<strong>at</strong>ed by shooting protons<br />

on nuclei, then pions are gener<strong>at</strong>ed in forward direction. (They decay into<br />

muons, which are then used for muon-nucleon sc<strong>at</strong>tering).<br />

Pion−Nucleon Sc<strong>at</strong>tering<br />

π (q)<br />

π (q’)<br />

N(p)<br />

N(p’)<br />

We have the follolwing on-shell conditions:<br />

q 2 = (q ′ ) 2 = m 2 π<br />

p 2 = (p ′ ) 2 = m 2 N<br />

109


Hence the “nucleon” in the propag<strong>at</strong>or is off-shell:<br />

(p + q) 2 = 2m 2 N + 2pq ≠ m 2 N<br />

10.8 Goldberger-Treiman rel<strong>at</strong>ion and pion pole<br />

The Goldberger-Treiman rel<strong>at</strong>ion is<br />

f π g πNN = m N g A (0)<br />

It is an important rel<strong>at</strong>ion rel<strong>at</strong>ing strong and weak interactions and reflecting<br />

the chiral spontaneous symmetry breaking. If the pion were massles (chiral<br />

limit) then the Goldberger-Treiman rel<strong>at</strong>ion would hold exactly. Since the pion<br />

is massive it holds with 7% devi<strong>at</strong>ion. We had a similar rel<strong>at</strong>ion already in the<br />

effective linear chiral sigma model. However, now we do it fully.<br />

In order to derive the Goldberger-Treiman rel<strong>at</strong>ion consider the m<strong>at</strong>rix element<br />

of the axial current:<br />

〈p(k ′ )|(A 1 µ + iA 2 µ)(0) |n(k)〉 = u p (k ′ ) [ γ µ γ 5 g A (q 2 ) + q µ γ 5 h A (q 2 ) ] u n (k)<br />

One can show, th<strong>at</strong> from Lorenze covariance the LHS must have the structure<br />

as it is given on the RHS. Here is q = k − k ′ . Actually this m<strong>at</strong>rix element is<br />

measured in the beta decay of the nucleon. Experimentally one has g A (0) =<br />

1.26. On can derive from this the m<strong>at</strong>rix element of the divergence of the axial<br />

current between proton and neutron. This goes in the fol<strong>low</strong>ing way analogous<br />

to the m<strong>at</strong>rixelement of the axial current between vacuum and one-pion st<strong>at</strong>e:<br />

We know th<strong>at</strong> (see section about quantiz<strong>at</strong>ion of boson field) for any oper<strong>at</strong>or,<br />

depending on the elementary fields<br />

F(x) = exp(iPx)F(0)exp(−iPx)<br />

F(0) = exp(−iPx)F(x)exp(+iPx)<br />

Inserting F(0) into the expression with the current and using e.g.<br />

one obtains immedi<strong>at</strong>ely<br />

>From this we get<br />

exp(iPx) |n(k)〉 = exp(ikx) |n(k)〉<br />

〈p(k ′ )|(A 1 µ + iA 2 µ)(0) |n(k)〉<br />

= 〈p(k ′ )| exp(−iPx)(A 1 µ + iA 2 µ)(x)exp(+iPx) |n(k)〉<br />

= 〈p(k ′ )| (A 1 µ + iA 2 µ)(x) |n(k)〉 exp(+ikx − ik ′ x)<br />

〈p(k ′ )|(A 1 µ+iA 2 µ)(x) |n(k)〉 = exp(i(k ′ −k)x)u p (k ′ ) [ γ µ γ 5 g A (q 2 ) + q µ γ 5 h A (q 2 ) ] u n (k)<br />

110


Now we perform the divergence on both sides yielding:<br />

〈p(k ′ )|(∂ µ A 1 µ + i∂ µ A 2 µ)(x) |n(k)〉 =<br />

= i(k ′ − k) µ exp(i(k ′ − k)x)u p (k ′ ) [ γ µ γ 5 g A (q 2 ) + q µ γ 5 h A (q 2 ) ] u n (k)<br />

or <strong>at</strong> x = 0:<br />

〈p(k ′ )|(∂ µ A 1 µ + i∂ µ A 2 µ)(0) |n(k)〉 =<br />

= i(k ′ − k) µ u p (k ′ ) [ γ µ γ 5 g A (q 2 ) + q µ γ 5 h A (q 2 ) ] u n (k)<br />

We can rewrite this by using the Dirac equ<strong>at</strong>ion (γ µ k µ − m)u(k) = 0 obtaining<br />

〈p(k ′ )|(∂ µ A 1 µ + i∂ µ A 2 µ)(0) |n(k)〉 =<br />

= iu p (k ′ )γ 5 u n (k) [ 2m N g A (q 2 ) + q 2 h A (q 2 ) ]<br />

In order to rel<strong>at</strong>e this to the pion-nucleon coupling constant we have to consider<br />

physical pion fields:<br />

and using the PCAC rel<strong>at</strong>ion<br />

π + (x) = 1 √<br />

2<br />

(π 1 (x) + iπ 2 (x))<br />

∂ µ A j µ(x) = f π m 2 ππ j (x)<br />

one gets by sandwiching between proton and neutron st<strong>at</strong>es<br />

〈p(k ′ )| (∂ µ A 1 µ + i∂ µ A 2 µ)(0) |n(k)〉 = √ 2f π m 2 π 〈p(k ′ )| π + (0) |n(k)〉<br />

This can directly be rel<strong>at</strong>ed to the pion-nucleon coupling constant as it is known<br />

to us from the previous subsection. There we had for g πNN (q 2 ) the fol<strong>low</strong>ing<br />

√<br />

2<br />

〈p(k ′ )|π + (0) |n(k)〉 =<br />

−q 2 + m 2 g πNN (q 2 )u(k ′ )iγ 5 u(k)<br />

π<br />

Combing this with the above formulae yields<br />

〈p(k ′ )|(∂ µ A 1 µ + i∂ µ A 2 µ)(0) |n(k)〉 = √ √<br />

2<br />

2f π m 2 π<br />

−q 2 + m 2 g πNN (q 2 )u(k ′ )iγ 5 u(k)<br />

π<br />

Comparison with the above expression for the divergence yields<br />

f π m 2 2<br />

π<br />

−q 2 + m 2 g πNN (q 2 ) = [ 2m N g A (q 2 ) + q 2 h A (q 2 ) ]<br />

π<br />

This is a general rel<strong>at</strong>ionship valid for all momentum transfers q.<br />

111


If we set q 2 = 0 we have<br />

f π g πNN (0) = m N g A (0)<br />

Unfortun<strong>at</strong>ely the point g πNN (0) is not physical since neither pion nucleon<br />

sc<strong>at</strong>tering nor nucleon-nucleon sc<strong>at</strong>tering with pion exchange has a vanishing<br />

momentum transfer. One has therefore to assume a “smooth” and analytical<br />

behaviour from q 2 = 0 to q 2 = m 2 π (which is r<strong>at</strong>her small) such th<strong>at</strong> we have<br />

g πNN (0) ≃ g πNN (m 2 π) = g πNN<br />

which gives then the famous Goldberger-Treiman rel<strong>at</strong>ion:<br />

f π g πNN = m N g A (0)<br />

This rel<strong>at</strong>ion is s<strong>at</strong>isfied to about 7% in n<strong>at</strong>ure, wh<strong>at</strong> we can easily check using<br />

the numbers f π = 93MeV,g πNN = 14.6,m N = 938MeV,g A (0) = 1.26 for yielding<br />

1357˜1181. This example shows, th<strong>at</strong> one can get really rel<strong>at</strong>ions between<br />

physical observables by utilizing the PCAC rel<strong>at</strong>ion. On the other hand we<br />

needed an extrapol<strong>at</strong>ion g πNN (0) ≃ g πNN (m 2 π) = g πNN which is believed to be<br />

justified, because the pion mass is small on a hadronic scale. One sees clearly,<br />

th<strong>at</strong> one gets in trouble if one generalizes such an argument<strong>at</strong>ion to SU(3). The<br />

rel<strong>at</strong>ionship of PCAC to physical quantities in th<strong>at</strong> case requires extrapol<strong>at</strong>ion<br />

from q 2 = 0 to q 2 = m 2 K , where the kaon mass is about 500 MeV. This large<br />

kinem<strong>at</strong>ical region makes PCAC less usefull for SU(3).<br />

One can see here immedi<strong>at</strong>ely the famous pion pole term: Consider<br />

f π m 2 2<br />

π<br />

−q 2 + m 2 g πNN (q 2 ) = [ 2m N g A (q 2 ) + q 2 h A (q 2 ) ]<br />

π<br />

and consider then the limit q 2 → m 2 π. Since the first term on the RHS ˜g A<br />

remains finite and q 2 as well, we see th<strong>at</strong> the h A (q 2 ) is completely determined<br />

by the diverging term on the LHS. Thus in his limit we have the so called pion<br />

pole.<br />

Limes q 2 → m 2 π ::::⇒:::: h A (q 2 ) =<br />

2f π<br />

m 2 π − q 2 g πNN(q 2 ) :::<br />

In fact, if we have a reaction, which is governed by some form factors in the<br />

region of q 2˜m 2 π the reaction process is govened by h A (q 2 ) and this is given by<br />

the pion-nucleon coupling constant, known from many pion-nucleon sc<strong>at</strong>tering<br />

processes.<br />

10.9 Vacuum Sigma Term and Gell-Mann–Okubo<br />

The vacuum Sigma term al<strong>low</strong>s to characterize the vacuum and and its condens<strong>at</strong>e<br />

by rel<strong>at</strong>ing them to the Goldstone meson masses. Th<strong>at</strong> why it is an<br />

important quantity. WE first consider it in SU(2). There it is defined by<br />

SU(2):<br />

σ ab<br />

0 = − 〈0| [ Q a A, [ Q b A, L m<br />

]]<br />

|0〉<br />

112


In fact, as we will see immedi<strong>at</strong>ely, the diagonal numbers can be rel<strong>at</strong>ed directly<br />

to the pseudo scalar meson mass m a and decay constants f a , which are of<br />

course observable quantities. Thus one learns something about the vacuum by<br />

considering σ0 ab and the meson masses. One basically learns something about<br />

the product of current quark masses and the vacuum condens<strong>at</strong>e.<br />

The assertion is for a = 1,2,3<br />

σ ab<br />

0 = δ ab m 2 af a<br />

In order to see this we start with the PCAC rel<strong>at</strong>ion<br />

∂ µ A a µ(x) = f a m 2 aπ a (x)<br />

and the redction formula<br />

< 0 ∣ π a (0)|π b (k) 〉 ∫<br />

= i d 4 y exp(−iky)(−k 2 + m 2 b) 〈0|T {π a (0)π b (y)} |0〉<br />

We insert the PCAC rel<strong>at</strong>ion into this expression and obtain<br />

δ ab = i (−k2 + m 2 b ) ∫<br />

f a m 2 af b m 2 d 4 y exp(−iky) 〈0| T {∂xA ν a ν(0)∂ y µ A b µ(y)} |0〉<br />

b<br />

= i (−k2 + m 2 b ) ∫<br />

f a m 2 af b m 2 d 4 y exp(−iky)∗<br />

b<br />

∗ 〈0| θ(−y 0 )∂xA ν a ν(0)∂ y µ A b µ(y) + θ(y 0 )∂ y µ A b µ(y)∂xA ν a ν(0) |0〉<br />

Performing now partial integr<strong>at</strong>ion yields −∂ µ y exp(−iky) = ik µ exp(−iky) and<br />

apparently terms where the ∂ µ y is applied to the θ(y 0 ) and θ(−y 0 ).With the<br />

equ<strong>at</strong>ions<br />

one obtains immedi<strong>at</strong>ely<br />

∂ y µθ(y 0 − x 0 ) = g µ0 δ(y 0 − x 0 )<br />

∂ y µθ(x 0 − y 0 ) = −g µ0 δ(y 0 − x 0 )<br />

δ ab f a m 2 a = i (−k2 + m 2 b ) ∫<br />

f b m 2 {ik µ d 4 y exp(−iky) 〈0| T {∂xA ν a ν(0)∂ y µ A b µ(y)} |0〉<br />

b<br />

∫<br />

− d 4 y exp(−iky) 〈0| δ(y 0 )[A b 0(y),∂ µ A a µ(0)] |0〉}<br />

in the limit k → 0 we obtain a d 3 y-integral yielding an axial charge Q b A (x 0),<br />

and the integral over the time component yields Q b A (0). Thus we obtain<br />

δ ab f a m 2 a = − 〈0| [Q b A(0),∂ µ A a µ(0)] |0〉 for a = 1,2,3<br />

In SU(2), wh<strong>at</strong> we have presently assumed, one can develope it further, as it<br />

is often done. The divergence of the axial current is known for isospin symmetry<br />

in the up-down sector<br />

∂ µ A µ a = (m u + m d ) ¯ψiγ 5 1 2 τ aψ for a = 1,2,3<br />

113


together with the equ<strong>at</strong>ion from the section on chiral symmetry breaking<br />

[<br />

Q<br />

A<br />

a , ¯ψ(y)ψ(y) ] = − ¯ψ(y)τ a γ 5 ψ(y)<br />

we obtain th<strong>at</strong><br />

or<br />

∂ µ A a µ(0) = − i 2 (m u + m d ) [ Q A a , ¯ψ(y)ψ(y) ] for a = 1,2,3<br />

∂ µ A a µ(0) = −i [ Q A a , L u,d ]<br />

m<br />

such th<strong>at</strong> we altogether obtain in the limit k → 0<br />

δ ab f a m 2 a = − 〈0| [Q A b , [ Q A a (0), L u,d<br />

m (0) ] ] |0〉 = σ0<br />

ab<br />

This is a well known formula written in many books. Often one writes H(0) =<br />

−L u,d<br />

m (0).<br />

For SU(3) we have to go back to the formula with the divergence of the axial<br />

current. In fact we have<br />

δ ab f a m 2 a = − 〈0| [Q a A(0),∂ µ A a µ(0)] |0〉 for a = 1,...,8<br />

since we never assumed in the deriv<strong>at</strong>ion of this formula really SU(2). However<br />

the step from ∂ µ A a µ(0) to L m (0) cannot be done since the strange mass is much<br />

larger than the up- and down-mass. However one can directly insert the explicit<br />

expression for the divergence of the axial current<br />

∂ µ A µ a = ¯ψiγ 5 1 2 {λ a,m} ψ<br />

with the mass m<strong>at</strong>rix being m = [m 1 I + m 3 λ 3 + m 8 λ 8 ]. The commut<strong>at</strong>or of the<br />

axial charges with ¯ψiγ 5 1 2 {λ a,m} ψ can be explicitely calcul<strong>at</strong>ed using known<br />

formulae. This yields in the end important rel<strong>at</strong>ions between the vacuum condens<strong>at</strong>es<br />

and the masses and decay constants of the correspponding Goldstone<br />

bosons:<br />

f π m 2 π = m u + m d<br />

2<br />

f K m 2 K = m u + m s<br />

2<br />

f η m 2 η = m u + m d<br />

6<br />

〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉<br />

〈0| ¯ψ u ψ u + ¯ψ s ψ s |0〉<br />

〈0| ¯ψ u ψ u + ¯ψ d ψ d |0〉 + 4m s<br />

3 〈0| ¯ψ s ψ s |0〉<br />

We can estim<strong>at</strong>e now the current masses of the quarks in an approxim<strong>at</strong>e way:<br />

We have derived in the massless limit the property for the vacuum condens<strong>at</strong>es<br />

as<br />

〈0| ¯ψ u ψ u |0〉 = 〈0| ¯ψ d ψ d |0〉 = 〈0| ¯ψ s ψ s |0〉<br />

Assuming massles limit also for the decay constants we have<br />

f π = f K = f η<br />

114


and one obtains immedi<strong>at</strong>ely the famous Gell-Mann–Okubo mass rel<strong>at</strong>ion<br />

4m 2 K = 3m 2 η + m 2 π<br />

which is well fulfilled, sinced we have m K = 494MeV,m η = 548MeV,m π =<br />

139MeV.Using these numbers one obtains<br />

4m 2 K − (3m2 η + m 2 π)<br />

= 10%<br />

4m 2 K<br />

One can also obtain the quark mass r<strong>at</strong>io<br />

m u + m d<br />

2m s<br />

=<br />

m 2 π<br />

2m 2 K − m2 π<br />

≈ 1<br />

25<br />

Apparently the strange quark mass is very much larger than the up- and downquark<br />

mass. If one assumes for m s = 180MeV then m u + m d = 14, which is<br />

about right.<br />

10.10 Pion-Nucleon Sigma Term Σ πN<br />

The pion nucleon Sigma Term Σ πN is an important quantity. It rel<strong>at</strong>es the<br />

mass-terms of the <strong>QCD</strong> Lagrangean L M to the pion-nucleon sc<strong>at</strong>tering. This is<br />

interesting since the quarks are never free and hence we cannot measure their<br />

mass directly. We can measure, however, the pion-nucleon sc<strong>at</strong>tering. So a<br />

conclusion from such an observable quantity like the pion-nucleon sc<strong>at</strong>tering<br />

amplitude to an not-observable quantity like L M , which however appears in the<br />

<strong>QCD</strong>-Lagrangean, is extremely interesting. Therefore the Σ πN is discussed for<br />

years already.<br />

We consider the pion-nucleon sigma term first in SU(2). If we write down<br />

the <strong>QCD</strong> Lagrangean then the mass terms of the quarks are given by<br />

L m = − ¯ψMψ<br />

with M = diag(m u ,m d ). The mass term is rel<strong>at</strong>ed to the oper<strong>at</strong>or of the Sigma<br />

term Σ πN .This is defined as<br />

[ [ ]]<br />

Σ ab<br />

πN = − Q a A, Q b A, L u,d<br />

M<br />

with the mass term<br />

L M = −m 0 ¯ψ ψ<br />

and can be shown to have the structure (proof be<strong>low</strong>):<br />

Σ ab<br />

πN =<br />

8∑<br />

e=0<br />

C ab<br />

e (m u ,m d ,m s ) ¯ψλ e ψ<br />

The term is measured in the pion-nucleon sc<strong>at</strong>tering<br />

π b (q 1 ) + N(p 1 ) → π a (q 2 ) + N(p 2 )<br />

115


One usually takes the kinem<strong>at</strong>ic variables<br />

ν = 1 4 (p 1 + p 2 )(q 1 + q 2 )<br />

ν B = − 1 2 q 1q 2<br />

and we use from the section on Lehmann-Symanzik-Zimmerman reduction formulae<br />

the equ<strong>at</strong>ion for pion-nucleon sc<strong>at</strong>tering:<br />

< π a (q 2 )N(p 2 );out ∣ ∣π b (q 1 )N(p 1 );in 〉 = i(2π) 4 δ (4) (p 1 +q 1 −p 2 −q 2 )T ab<br />

πN→πN(ν,ν B )<br />

in the limit q 2 1 → m 2 b = m2 π and q 2 2 → m 2 a = m 2 π<br />

< π a (q 2 )N(p 2 );out ∣ ∣π b (q 1 )N(p 1 );in 〉<br />

= (i) 2 ∫<br />

d 4 xd 4 y exp(iq 2 x)exp(−iq 1 y)(−q 2 1 + m 2 b)(−q 2 2 + m 2 a) 〈 N(p 2 ) ∣ ∣ T {π a (x)π b (y)} ∣ ∣ N(p1 ) 〉<br />

We know th<strong>at</strong><br />

〈0|A a µ(x) ∣ ∣ π b (p) 〉 = ip µ f π δ ab exp(−ipx)<br />

〈0|∂ µ A a µ(x) ∣ ∣ π b (p) 〉 = im 2 πf π δ ab exp(−ipx)<br />

and PCAC<br />

∂ µ A a µ(x) = m 2 πf π π a (x)<br />

Take now the RHS of the above expression for < π a (q 2 )N(p 2 );out ∣ π b (q 1 )N(p 1 );in 〉<br />

and replace the pion field by the divergence of the axial current. The expression<br />

is defined <strong>at</strong> the points q1 2 = q2 2 = m 2 π . This is not sufficient for the fol<strong>low</strong>ing,<br />

therefor take the analytical continu<strong>at</strong>ion to any q1 2 ≠ q2 2 ≠ m 2 π. This yields<br />

i(2π) 4 δ (4) (p 1 + q 1 − p 2 − q 2 )TπN→πN(ν,ν ab<br />

B ,q1,q 2 2)<br />

2<br />

1 1 (<br />

= i m<br />

2<br />

m 2 af π m 2 b f b − q 2 ) (<br />

1 m<br />

2<br />

a − q 2 )<br />

2 ∗<br />

π<br />

∫<br />

∗ d 4 xd 4 y exp(iq 2 x)exp(−iq 1 y) 〈 N(p 2 ) ∣ T {∂<br />

x<br />

ν A νa (x)∂µA y µb (y)} ∣ N(p1 ) 〉<br />

Consider now the integral<br />

∫<br />

I = d 4 xd 4 y exp(iq 2 x)exp(−iq 1 y) 〈 N(p 2 ) ∣ T {∂<br />

x<br />

ν A νa (x)∂µA y µb (y)} ∣ N(p1 ) 〉<br />

We can reformul<strong>at</strong>e it (proof be<strong>low</strong>) yielding by a complic<strong>at</strong>ed but straight<br />

forward calcul<strong>at</strong>ion the Ward Identity:<br />

T {∂ x νA νa (x)∂ y µA µb (y)} = A + B + C<br />

= ∂µ∂ y νT x {A µb (y)A νa (x)}<br />

− δ(y 0 − x 0 ) [ A 0b (y),∂νA x νa (x) ]<br />

+ ∂µ<br />

y (<br />

δ(y0 − x 0 ) [ A µb (y),A 0a (x) ])<br />

116


the proof of this formula is given be<strong>low</strong>: Actually, although this looks like a<br />

complic<strong>at</strong>ion, we rewrite this because with this new expression we can have<br />

tremendous simplific<strong>at</strong>ions in the limit q µ 1 → 0 and qµ 2 → 0.The RHS is a sum of<br />

three terms I = A+B+C, which have different properties if we consider the soft<br />

pion limit, defined by q µ 1 → 0 and qµ 2 → 0. One sees this by shifting e.g. in A<br />

the deriv<strong>at</strong>ive ∂µ y in the integral over y by partial integr<strong>at</strong>ion on the exponential<br />

exp(−iq 1 y) yielding exp(−iq 1 y)∂µ y → −(−iq µ 1 )exp(−iq 1y) = iq µ 1 exp(−iq 1y) and<br />

the same for exp(iq 1 y)∂ν x → −iq2 ν exp(iq 1 y). Thus we have A˜q µ 1 qν 2 and hence<br />

A → 0 for q µ 1 → 0 or qµ 2 → 0. Similarly we have C˜ − iq 2ν, which also goes<br />

to zero in the soft pion limit. The term B does not al<strong>low</strong> for such a simple<br />

tre<strong>at</strong>ment, since the deriv<strong>at</strong>ives act also on the δ(x 0 − y 0 ) and hence do not<br />

yield a simple q µ 1 or so. Thus in the soft pion limit the only remaining term is<br />

∫<br />

I B = − d 4 xd 4 y exp(iq 2 x)exp(−iq 1 y)δ(y 0 − x 0) [A 0b (y),∂νA x νa (x)]<br />

This term will turn out to be basically the pion-nucleon sigma term. To see this<br />

we have first to show th<strong>at</strong> with equal up- and down-masses we have the useful<br />

expression (proven already in the section about vacuum sigma term)<br />

[<br />

Q<br />

a<br />

A (x 0 ), ¯ψ(x)ψ(x) ] = − ¯ψ(x)γ 5 τ a ψ(x)<br />

and<br />

∂ µ A µ a = (m u + m d ) ¯ψiγ 5 1 2 τ aψ<br />

which yields then after sandwiching with the nucleonic st<strong>at</strong>es<br />

∫<br />

B ′ = −i d 4 xd 4 y exp(iq 2 x)exp(−iq 1 y)δ(y 0 −x 0 ) 〈 N(p 2 ) ∣ [<br />

] ∣∣N(p1 [A 0b (y), Q a A(x 0 ), L u,d<br />

M (x) ) 〉<br />

This expression is already quite similar to the sigma term, which is basically<br />

the double commut<strong>at</strong>or of the mass term with the axial charges. In fact the<br />

integrals and the delta-function make out of the A 0b the axial charge as we will<br />

see now. We know the rel<strong>at</strong>ion F(x + y) = exp(iPx)F(y)exp(−iPx) which we<br />

apply to the above expression:<br />

〈<br />

N(p2 ) ∣ ]<br />

[A 0b (y),[Q a A(x 0 ), L u,d<br />

M (x) ] ∣ N(p1 ) 〉<br />

= 〈 N(p 2 ) ∣ [ ]<br />

exp(iPx)[A 0b (y − x), Q a A(0), L u,d<br />

M (0) ]exp(−iPx) ∣ ∣N(p 1 ) 〉<br />

= exp(i(p 2 − p 1 )x) 〈 N(p 2 ) ∣ [ ]<br />

[A 0b (y − x), Q a A(0), L u,d<br />

M (0) ] ∣ N(p1 ) 〉<br />

inserted in the above expression yields<br />

∫<br />

B ′ = −i d 4 xd 4 y exp(i(q 2 +p 2 −p 1 )x)exp(−iq 1 y)δ(y 0 −x 0 ) 〈 N(p 2 ) ∣ [ ]<br />

[A 0b (y−x), Q a A(0), L u,d<br />

M (0) ] ∣ N(p1 ) 〉<br />

117


which is reformul<strong>at</strong>ed by changing the variables y −x = z and dy = dz .yielding<br />

∫<br />

B ′ = −i d 4 xd 4 z exp(i(q 2 + p 2 − p 1 )x)exp(−iq 1 x − iq 1 z)δ(z 0 ) 〈 N(p 2 ) ∣ [ ]<br />

[A 0b (z), Q a A(0), L u,d<br />

M (0) ] ∣ ∣N(p 1 ) 〉<br />

∫<br />

= −i(2π) 4 δ (4) (q 2 + p 2 − p 1 − q 1 ) d 4 z exp(−iq 1 z)δ(z 0 ) 〈 N(p 2 ) ∣ [ ]<br />

[A 0b (z), Q a A(0), L u,d<br />

M (0) ] ∣ N(p1 ) 〉<br />

This can now be simplified to the axial charge, if we take again the limit q µ 1 → 0,<br />

because<br />

∫<br />

∫<br />

∫<br />

d 4 z exp(−iq 1 z)δ(z 0 )A 0b (z) = d 3 z exp(iq 1<br />

z ) dz 0 δ(z 0 )A 0b (z)<br />

∫<br />

= d 3 z A 0b (0,z) = Q b A(0)<br />

Apparently the last step is only possible in the soft pion limit, otherwise in the<br />

space integral we would have the phase and the result would be only approxim<strong>at</strong>ely<br />

the axial charge.<br />

We can now collect the terms and have in the soft pion limit<br />

with<br />

B ′ = i(2π) 4 δ (4) (p 2 − p 1 )Σ ab<br />

πN<br />

Σ ab<br />

πN = − 〈 N(p 2 ) ∣ [ ]<br />

[Q<br />

b<br />

A (0), Q a A(0), L u,d<br />

M (0) ] ∣ N(p1 ) 〉<br />

Comparison with the expression for the pion-nucleon sc<strong>at</strong>tering amplitude yields<br />

in the soft pion limit<br />

i(2π) 4 δ (4) (p 2 − p 1 ) T ab<br />

πN→πN(ν,ν B ,q 2 1 = 0,q 2 2 = 0) =<br />

i(2π) 4 δ (4) (p 2 − p 1 ) Σ ab<br />

or, since in the soft pion limit also the ν,ν B go to zero, we have finally<br />

πN<br />

TπN→πN(ν ab = 0,ν B = 0,q1 2 = 0,q2 2 = 0) = 1<br />

fπ<br />

2 Σ ab<br />

πN<br />

1 1<br />

m 2 πf π m 2 (m 2 π − q1)(m 2 2 π − q1)<br />

2<br />

πf π<br />

This is a direct rel<strong>at</strong>ion between the pion-nucleon sc<strong>at</strong>tering amplitude and the<br />

pion-nucleon sigma term. The unfortun<strong>at</strong>e thing is, th<strong>at</strong> one needs TπN→πN ab (0,0,0,0),which<br />

cannot be measured directly since it is in a kinet<strong>at</strong>ically forbidden region. The<br />

experiment has always q1 2 = m 2 π and q2 2 = m 2 π and hence the pions in the<br />

above TπN→πN ab (0,0,0,0) are off-shell. In addition we have in a real sc<strong>at</strong>tering<br />

process <strong>at</strong> the threshold ν = m π m N ,ν B = − 1 2 m2 π . Therefore the point with<br />

ν = 0,ν B = 0 has a special name, it is called the Cheng-Dashen-Point. Using<br />

Mandelstam coordin<strong>at</strong>es s = (p 1 + q 1 ) 2 and t = (q 2 − q 1 ) the Cheng Dashen<br />

point is <strong>at</strong> t = 2m 2 π and s = m 2 N .Thus it cannot be reached since in a real sc<strong>at</strong>tering<br />

process, one has <strong>at</strong> least s = m 2 N +m2 π. The extrapol<strong>at</strong>ion from the d<strong>at</strong>a<br />

<strong>at</strong> kinem<strong>at</strong>ically accessible points to the Cheng-Dashen point and the off-shell<br />

118


pions is approxim<strong>at</strong>ely possible as we shall see be<strong>low</strong>, since the pion mass is<br />

only 139 MeV and hence small. For kaon- or eta-sc<strong>at</strong>tering the extrapol<strong>at</strong>ion<br />

must be done over a larger mass scale and is therfore not th<strong>at</strong> reliable. Thus the<br />

sigma term involving components a > 3 and b > 3 is not th<strong>at</strong> well destermined<br />

by kano-nucleon or eta-nucleon sc<strong>at</strong>tering.<br />

The extrapol<strong>at</strong>ion from TπN→πN ab (ν = 0,ν B = 0,q1 2 = m 2 π,q2 2 = m 2 π) to<br />

T ab<br />

πN→πN (0,0,q2 1 = 0,q2 2 = 0) can be done by means of the Adlers consistency<br />

rel<strong>at</strong>ion. This rel<strong>at</strong>ion can be proved with the same techniques as above, but<br />

this will not be done here (see Cheng-Li for th<strong>at</strong>). In order to do th<strong>at</strong> we<br />

decompose the above amplitude in isospin space by (remember th<strong>at</strong> the TπN→πN<br />

ab<br />

as oper<strong>at</strong>or act on the spinor or the nucleon, which is a doublet of proton and<br />

neutron:<br />

TπN→πN ab = δ ab T (+)<br />

πN→πN + 1 [<br />

τ a ,τ b] T (−)<br />

πN→πN<br />

2<br />

The indices a,b act in the space of the pion and the 2x2-m<strong>at</strong>rices τ a act in the<br />

protono-neutron-space. Adler has shown the PCAC consistency conditions<br />

T (+)<br />

πN→πN (0,0,m2 π,0) = 0......in the limit q 2 2 → 0<br />

T (+)<br />

πN→πN (0,0,0,m2 π) = 0.....in the limit q 2 1 → 0<br />

We can use these conditions in the fol<strong>low</strong>ing way;<br />

T (+)<br />

πN→πN (0,0,m2 π,m 2 π) = T (+)<br />

πN→πN (0,0,0,0) + dT (+)<br />

m2 π<br />

dq1<br />

2 + m 2 dT (+)<br />

π<br />

dq2<br />

2 + o(m 4 π)<br />

>From Adlers PCAC consistency condition fol<strong>low</strong>s<br />

or<br />

T (+)<br />

πN→πN (0,0,m2 π,0) = T (+)<br />

πN→πN (0,0,0,0) + dT (+)<br />

m2 π<br />

dq1<br />

2 = 0<br />

m 2 dT (+)<br />

π<br />

dq1<br />

2 = −T (+)<br />

πN→πN (0,0,0,0)<br />

and the same for the deriv<strong>at</strong>ive w.r. to q 2 2.If we replace the deriv<strong>at</strong>ives in the<br />

expansion for T (+)<br />

πN→πN (0,0,m2 π,m 2 π) we altogether get<br />

T (+)<br />

πN→πN (0,0,m2 π,m 2 π) = −T (+)<br />

πN→πN (0,0,0,0) + o(m4 π)<br />

= − 1<br />

fπ<br />

2 Σ πN + o(m 4 π)<br />

with Σ ab<br />

πN = Σ πNδ ab , if one ignores isospin breaking effects, i.e. if one assumes<br />

m u = m d . If one calcul<strong>at</strong>es the double commut<strong>at</strong>or, we get out in SU(2) in the<br />

rest frame of the nucleon:<br />

Σ πN = m 0<br />

2M N<br />

〈N| ¯ψ u ψ u + ¯ψ d ψ d |N〉<br />

119


Thus the Σ πN measures the contribution of the non-vanishing quark masses<br />

m 0 = mu+m d<br />

2<br />

to the nucleon mass m N .(Remember the section on scaling<br />

anomaly, where this contribution was discussed.) Wh<strong>at</strong> is left in the context<br />

of pion-nucleon sc<strong>at</strong>tering is the extrapol<strong>at</strong>ion from the physical point to the<br />

Cheng-Dashen point. This has been done first in a famous paper by Gasser,<br />

Leutwyler and Sainio (chiral perturb<strong>at</strong>ion theory, 1-loop calcul<strong>at</strong>ion) yielding<br />

(this has to be done more in detail and show also th<strong>at</strong> the mass-term of the<br />

lagrangean comes out of the double commut<strong>at</strong>or, or do this last step <strong>at</strong> the<br />

beginning of the section)<br />

Σ πN (2m 2 π) − Σ πN (0) ≃ 15MeV<br />

this gives for the Sigma-term in the nomencl<strong>at</strong>ure of Gasser, Leutwyler and<br />

Sainio<br />

F 2 π ¯D (+) (2m 2 π) = Σ πN (2m 2 π) = (60 ± 8)MeV<br />

yielding then for the Σ πN (0) ≃ (45 ± 8) MeV. These are the well known values<br />

from 1991. There are new measurements in the last few years, which yield larger<br />

values Fπ 2 ¯D (+) (2m 2 π) = (75±5) MeV or Σ πN (0) ≃ (60±5) MeV. The final value<br />

is not settled yet.Nevertheless, since the Σ πN measures the contribution of the<br />

non-vanishing quark masses m 0 = mu+m d<br />

2<br />

to the nucleon mass m N ,.all these<br />

values show, th<strong>at</strong> this contribution is small compared to the nucleon mass of<br />

938 MeV.<br />

From the sigma-term we can learn something about the strange quark content<br />

of the nucleon. The term can exactly be written as<br />

Σ πN = m 0 〈N| ¯ψ u ψ u + ¯ψ d ψ d − 2 ¯ψ s ψ s |N〉<br />

2M N 1 − y<br />

where the quantity y is a measure of the scalar strang quark content of the<br />

nucleon:<br />

2 〈N|<br />

y =<br />

¯ψ s ψ s |N〉<br />

〈N| ¯ψ u ψ u + ¯ψ d ψ d |N〉<br />

Assuming th<strong>at</strong> the SU(3) symmetry breaking is small, one can derive from this<br />

expression and the octet mass rel<strong>at</strong>ions (see l<strong>at</strong>er) the expression<br />

(1 − y)Σ πN =<br />

m 0<br />

m s − m 0<br />

(M Ξ + M Σ − 2M N ) (206)<br />

An expression of this sort is not surprising <strong>at</strong> all. First it resembles the situ<strong>at</strong>ion<br />

in the vacuum, where we had pseudoscalar meson masses rel<strong>at</strong>ed to the<br />

vacuum sigma term. For the baryons the situ<strong>at</strong>ion is similar since the mass<br />

term 〈N| ¯ψ u ψ u + ¯ψ d ψ d − 2 ¯ψ s ψ s |N〉 is the only term in the <strong>QCD</strong>-lagrangean by<br />

which the masses of the baryons can differ. Without this all multiplets and all<br />

members in the multiplets (to be discussed in the section about quark model)<br />

would be identical. In one uses m s = 25m 0 one obtains with the known masses<br />

(1−y)Σ πN = 26MeV. Together with the empirical value of Σ πN = (45±8)MeV<br />

one gets y = 0.2 ± 0.2. With the new and larger values of Σ πN = 79MeV one<br />

120


obtains a value of y no longer comp<strong>at</strong>ible with Zero. However by chiral perturb<strong>at</strong>ion<br />

theory the connection between Σ πN and the masses of the baryons is<br />

somewh<strong>at</strong> different and one gets there (1 − y)Σ πN = (36 ± 7)MeV so th<strong>at</strong> one<br />

gets altogether somehow<br />

y = 0.2 → 0.6<br />

and from there a contribution to the mass of the strangeness contribution to<br />

nucleon mass: m s 〈N| ¯ψ s ψ s |N〉 = (150 → 480)MeV<br />

The reasoning for the last point goes as fol<strong>low</strong>s. We start from a SU(3) theory<br />

without symmetry breaking, i.e. where all quark masses are equal. To calcul<strong>at</strong>e<br />

the masses of the hyperons one does <strong>at</strong> least first order perturb<strong>at</strong>ion theory<br />

and hence one has to calcul<strong>at</strong>e something like < Σ nonbroken |m s¯ss|Σ nonbroken ><br />

where we can write<br />

⎛ ⎞<br />

¯ss = 1 2 ¯ψ<br />

1 0 0 √<br />

⎝ 0 1 0 ⎠ 3<br />

ψ −<br />

2 ¯ψλ 8 ψ<br />

0 0 1<br />

In order to calcul<strong>at</strong>e the mass splitting of the unbroken hyperons due to strange<br />

quark mass we need the Wigner Eckart theorem of the SU(3) group like<br />

〈a|O b |c〉 = B 1 f abc + B 2 d abc<br />

where each of the st<strong>at</strong>es is a members of a multiplet and B 1 and B 2 are dynamical<br />

numbers, which are valid for the corresponding whole multiplet. This is in<br />

the cartesian represent<strong>at</strong>ion and we need it in the spherical represent<strong>at</strong>ion since<br />

there the particles are defined with diagonal charge. If one does all this one is<br />

able to express the B 1 and B 2 through mass differences. This yields the above<br />

formula eq.(206).<br />

We have still to prove the Ward-Identity: For this we use<br />

And now we want to proof th<strong>at</strong>:<br />

∂ y µθ(y 0 − x 0 ) = g µ0 δ(y 0 − x 0 )<br />

∂ y µθ(x 0 − y 0 ) = −g µ0 δ(y 0 − x 0 )<br />

T {∂µA y µb (y)∂νA x νa (x)} = ∂µ∂ y νT x {A µb (y)A νa (x)}<br />

− δ(y 0 − x 0 ) [ A 0b (y),∂νA x νa (x) ]<br />

+ ∂µ<br />

y (<br />

δ(y0 − x 0 ) [ A µb (y),A 0a (x) ])<br />

121


we do it by evalu<strong>at</strong>ing the first term on the RHS:<br />

∂µ∂ y νT x {A µb (y)A νa (x)} =∂µ∂ y ν x {θ(y 0 − x 0 )A µb (y)A νa (x) + θ(x 0 − y 0 )A νa (x)A µb (y)}<br />

= ∂µ[∂ y νθ(y x 0 − x 0 )A µb (y)A νa (x) + θ(y 0 − x 0 )A µb (y)∂νA x νa (x)<br />

+ ∂νθ(x x 0 − y 0 )A νa (x)A µb (y) + θ(x 0 − y 0 )∂νA x νa (x)A µb (y)]<br />

= ∂µ∂ y νθ(y x 0 − x 0 )A µb (y)A νa (x) + ∂νθ(y x 0 − x 0 )∂µA y µb (y)A νa (x)<br />

+ ∂µθ(y y 0 − x 0 )A µb (y)∂νA x νa (x) + θ(y 0 − x 0 )∂µA y µb (y)∂νA x νa (x)<br />

+ ∂µ∂ y νθ(x x 0 − y 0 )A νa (x)A µb (y) + ∂νθ(x x 0 − y 0 )A νa (x)∂µA y µb (y)<br />

+ ∂µθ(x y 0 − y 0 )∂νA x νa (x)A µb (y) + θ(x 0 − y 0 )∂νA x νa (x)∂µA y µb (y)<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = ∂µ∂ y νθ(y x 0 − x 0 )A µb (y)A νa (x) + (−)g ν0 δ(y 0 − x 0 )∂µA y µb (y)A νa (x)<br />

+ g µ0 δ(y 0 − x 0 )A µb (y)∂νA x νa (x) + θ(y 0 − x 0 )∂µA y µb (y)∂νA x νa (x)<br />

+ ∂µg y ν0 δ(y 0 − x 0 )A νa (x)A µb (y) + g ν0 δ(y 0 − x 0 )A νa (x)∂µA y µb (y)<br />

+ (−)g µ0 δ(y 0 − x 0 )∂νA x νa (x)A µb (y) + θ(x 0 − y 0 )∂νA x νa (x)∂µA y µb (y)<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = ∂µ∂ y νθ(y x 0 − x 0 )A µb (y)A νa (x) + (−)δ(y 0 − x 0 )∂µA y µb (y)A 0a (x)<br />

+ δ(y 0 − x 0 )A 0b (y)∂νA x νa (x) + θ(y 0 − x 0 )∂µA y µb (y)∂νA x νa (x)<br />

+ ∂µδ(y y 0 − x 0 )A 0a (x)A µb (y) + δ(y 0 − x 0 )A 0a (x)∂µA y µb (y)<br />

+ (−)δ(y 0 − x 0 )∂νA x νa (x)A 0b (y) + θ(x 0 − y 0 )∂νA x νa (x)∂µA y µb (y)<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = −∂µδ(y y 0 − x 0 )A µb (y)A 0a (x) + (−)δ(y 0 − x 0 )∂µA y µb (y)A 0a (x)<br />

+ δ(y 0 − x 0 )A 0b (y)∂νA x νa (x) + θ(y 0 − x 0 )∂µA y µb (y)∂νA x νa (x)<br />

+ ∂µδ(y y 0 − x 0 )A 0a (x)A µb (y) + δ(y 0 − x 0 )A 0a (x)∂µA y µb (y)<br />

+ (−)δ(y 0 − x 0 )∂νA x νa (x)A 0b (y) + θ(x 0 − y 0 )∂νA x νa (x)∂µA y µb (y)<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = θ(y 0 − x 0 )∂µA y µb (y)∂νA x νa (x) + θ(x 0 − y 0 )∂νA x νa (x)∂µA y µb (y)<br />

+ (−)∂µδ(y y 0 − x 0 )A 0b (y) (y)A νa (x) + (−)δ(y 0 − x 0 )∂µA y µb (y)A 0a (x)<br />

+ δ(y 0 − x 0 )A 0b (y)∂νA x νa (x) + (−)δ(y 0 − x 0 )∂νA x νa (x)A 0b (y)<br />

+ ∂µδ(y y 0 − x 0 )A 0a (x)A µb (y) + δ(y 0 − x 0 )A 0a (x)∂µA y µb (y)<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = T {∂µA y µb (y)∂νA x νa (x)}<br />

− ∂µδ(y y 0 − x 0 )A µb (y)A 0a (x) + (−)δ(y 0 − x 0 )∂µA y µb (y)A 0a (x)<br />

+ δ(y 0 − x 0 )[A 0b (y),∂νA x νa (x)]<br />

+ ∂µδ(y y 0 − x 0 )A 0a (x)A µb (y) + δ(y 0 − x 0 )A 0a (x)∂µA y µb (y)<br />

122


We have and reformul<strong>at</strong>e<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = T {∂µA y µb (y)∂νA x νa (x)}<br />

− ∂µδ(y y 0 − x 0 )[A µb (y),A 0a (x)]<br />

+ δ(y 0 − x 0 )[A 0b (y),∂νA x νa (x)]<br />

+ δ(y 0 − x 0 )[A 0a (x),∂µA y µb (y)]<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = T {∂µA y µb (y)∂νA x νa (x)}<br />

− ∂µδ(y y 0 − x 0 )[A µb (y),A 0a (x)]<br />

+ δ(y 0 − x 0 )[A 0b (y),∂νA x νa (x)]<br />

− δ(y 0 − x 0 )[∂µA y µb (y),A 0a (x)]<br />

∂µ∂ y νT x {A µb (y)A νa (x)} = T {∂µA y µb (y)∂νA x νa (x)}<br />

− ∂µ(δ(y y 0 − x 0 )[A µb (y),A 0a (x)])<br />

+ δ(y 0 − x 0 )[A 0b (y),∂νA x νa (x)]<br />

or finally<br />

T {∂ y µA µb (y)∂ x νA νa (x)} = ∂ y µ∂ x νT {A µb (y)A νa (x)} =<br />

+ ∂ y µ(δ(y 0 − x 0 )[A µb (y),A 0a (x)])<br />

− δ(y 0 − x 0 )[A 0b (y),∂ x νA νa (x)]<br />

qe.d<br />

We get the above formula by realizing th<strong>at</strong> T {∂ x νA νa (x)∂ y µA µb (y)} = T {∂ y µA µb (y)∂ x νA νa (x)}<br />

The proof is by writing down<br />

T {∂ y µA µb (y)∂ x νA νa (x)} = θ(y 0 − x 0 )∂ y µA µb (y)∂ x νA νa (x) + θ(x 0 − y 0 )∂ x νA νa (x)∂ y µA µb (y)<br />

T {∂ x νA νa (x)∂ y µA µb (y)} = θ(y 0 − x 0 )∂ y µA µb (y)∂ x νA νa (x) + θ(x 0 − y 0 )∂ x νA νa (x)∂ y µA µb (y)<br />

This finalizes the proof for the above Ward identity.<br />

10.11 Ward-Identities and Low energy pion-nucleon theorems<br />

We remember the definition for isospin even and isospin-odd part of the pionnucleon<br />

sc<strong>at</strong>tering T-m<strong>at</strong>rix:<br />

TπN→πN ab = δ ab T (+)<br />

πN→πN + 1 [<br />

τ a ,τ b] T (−)<br />

πN→πN<br />

(207)<br />

2<br />

123


We repe<strong>at</strong> Adler’s PCAC consistency conditions<br />

T (+)<br />

πN→πN (0,0,m2 π,q 2 2) = 0......in the limit q 2 2 → 0<br />

T (+)<br />

πN→πN (0,0,q2 1,m 2 π) = 0.....in the limit q 2 1 → 0<br />

With similar techniques as used in the section about pion-nucleon sigma<br />

term one can show several other Ward identities and <strong>low</strong> energy theorems (for<br />

indic<strong>at</strong>ion of proofs see Cheng and Li):<br />

ig 2 Aν 1 2<br />

[<br />

τ a ,τ b] = −ifπT 2 πN→πN ab + iν 1 [<br />

τ a ,τ b] − iδ ab Σ πN<br />

2<br />

lim 1<br />

ν → 0 ν T (−)<br />

πN→πN (ν,0,0,0) = 1 − g2 A<br />

with the axial coupling constant g A = 1.26. We furthermore use the unsubtracted<br />

dispersion rel<strong>at</strong>ion<br />

from which we get<br />

or<br />

1<br />

ν T (−)<br />

πN→πN (ν,0,0,0) = 2 π<br />

1<br />

g 2 A<br />

1 − g 2 A<br />

f 2 π<br />

= 2 π<br />

∫ ∞<br />

ν 0<br />

∫ ∞<br />

= 1 + 2M2 N<br />

πg πNN<br />

∫ ∞<br />

ν 0<br />

f 2 π<br />

(−)<br />

′ Im T<br />

πN→πN<br />

dν (ν,0,0,0)<br />

ν ′2 − ν 2<br />

(−)<br />

′ Im T<br />

πN→πN<br />

dν (ν,0,0,0)<br />

ν ′2<br />

ν 0<br />

(−)<br />

′ ImT<br />

πN→πN<br />

dν (ν,0,0,0)<br />

ν ′2<br />

with ν 0 = m π m N .Using a smoothness assumption and from the optical theorem<br />

Im T (−)<br />

πN→πN<br />

(ν,0,0,0) = νσ(−)<br />

tot (ν) = ν[σ π− N<br />

tot<br />

(ν) − σ π+ N<br />

tot (ν)]<br />

we get the famous Adler-Weissberger rel<strong>at</strong>ion, which is well fulfilled.<br />

with<br />

1<br />

g 2 A<br />

= 1 + 2M2 N<br />

πg 2 πNN<br />

∫ ∞<br />

ν 0<br />

N<br />

dν [σπ− tot<br />

(ν) − σ π+ N<br />

tot (ν)]<br />

ν<br />

Finally we have the Kroll-Rudermann-Theorem:<br />

σ(γN → πN) = 4π |−→ q |<br />

ω<br />

(|E 0+ | 2 + 2 |M 1+ | 2 + |M 1− | 2 )<br />

E 0+ (γN → πN) = e√ 2g A<br />

8πf π<br />

We get also for neutral pion photoproduction<br />

E 0+ (γp → π 0 p) = eg [<br />

A<br />

− m π<br />

+ m2 π<br />

8πf π m N 2m 2 (3 + κ p ) + m2 π<br />

N<br />

16fπ<br />

2 + o( m3 π<br />

m 3 )<br />

N<br />

124<br />

]


These <strong>low</strong> energy theorems are all based on Ward-identities which are generaliz<strong>at</strong>ions<br />

of Noethers theorem. For the <strong>low</strong> energy theorems Ward identities are<br />

used which involve rel<strong>at</strong>ions between the m<strong>at</strong>rix elements of currents and the<br />

m<strong>at</strong>rix elements of divergences. One Ward Identity we know already and have<br />

used it in the section about the pion-nucleon sigma-term.<br />

There is also for s-wave pion -nucleon sc<strong>at</strong>tering another <strong>low</strong>-energy theorem:<br />

Weinberg-Tomozawa Theorem. If one starts from the decomposition of the<br />

t-m<strong>at</strong>rix eq.(207) then the corresponding s-wave sc<strong>at</strong>tering lengths are<br />

a ± = 1 (<br />

1 + m ) −1<br />

π<br />

T ± | threshold<br />

4π M N<br />

Since <strong>at</strong> threshold th isospin-odd amplitudebecomes<br />

T − = ω<br />

2f 2 π<br />

with q 2 = ω 2 + −→ q 2 . This implies <strong>at</strong> threshold th<strong>at</strong> the isovectors c<strong>at</strong>tering length<br />

is given by (<br />

1 + m )<br />

π<br />

a − = m π<br />

M N 8πfπ<br />

2 ∼ 0.13fm<br />

and hence it vanishes in the chiral soft-pion limit m π −→ 0. On the other hand<br />

the isoscalar sc<strong>at</strong>tering length<br />

(<br />

1 + m π<br />

M N<br />

)<br />

a + = m2 π<br />

4πf 2 π<br />

(<br />

ΣπN<br />

m 2 π<br />

)<br />

+ d − g2 A<br />

4M N<br />

is suppressed in comparison with a − by an additional power of m π and in volves<br />

a subtel cancell<strong>at</strong>ion bewtween the sigma term and the parameter d, which must<br />

be frixed by comparison with experiment. The empirical pion-nuclen sc<strong>at</strong>tering<br />

length can be taken from the analysis of Ericson, Loisier and Thomas and yields<br />

in agreement with the above estim<strong>at</strong>es<br />

10.12 Isospin-Viol<strong>at</strong>ion<br />

α + = −(0.003 ± 0.002) fm<br />

α − = +(0.128 ± 0.002) fm<br />

We remember the symmmetry breaking mass-term in the <strong>QCD</strong> hamiltonian<br />

H sbr−m<br />

<strong>QCD</strong><br />

= −[m uūu + m d ¯dd + ms¯ss]<br />

In the context of the sigma-term we considered only the dase of m u = m d and<br />

ignored completely the fact th<strong>at</strong> there are also electromagnetic self <strong>energies</strong>,<br />

which contribute to the mass of a particle. Such a electromagnetic contribution<br />

is by no means small. On the contrary, its contribution is of the same order<br />

of magnitude as the contribution due to the different quark masses. Thus one<br />

125


must carefully distinguish between the hadronic and electromagnetic isospin<br />

splitting of the masses. For example: The pion is built like π + = u ¯d and<br />

π 0 = 1 √<br />

2<br />

(uū + d ¯d). Thus the quark masses contribute basically equal to both<br />

sorts of pions, whereas the electromagnetic selfenergy of the π 0 is zero and of<br />

π + is nonzero. The mass difference of both pions can be read off from the table<br />

particle mass (MeV) structure<br />

π + 139.6 u ¯d<br />

π 0 135.0 uū + d ¯d<br />

K + 493.7 ¯su<br />

K 0 497.7 ¯sd<br />

η 547.5 ūu + ¯dd − 2¯ss<br />

p 938.3 uud<br />

n 939.6 udd<br />

Actually one has to coplement the above symmetry breaking term by an<br />

electromagnetic symmetry breaking term. It can be written as<br />

∫<br />

= e2 d 4 xT (J µ elm (x)Jν elm(0)) Dµν(x)<br />

γ<br />

H sbr−γ<br />

<strong>QCD</strong><br />

where Dµν(x) γ is the photon propag<strong>at</strong>or and J µ elm<br />

(x) is the electromagnetic current.<br />

We will not need in the fol<strong>low</strong>ing the particular expression. We can now<br />

directly generalize the expressions of the previous subsection. We obtain: insert the<br />

with<br />

f a m 2 aδ ab = σ0m ab + σ0γ<br />

ab<br />

σ ab<br />

0m = 〈0| [ Q 5a , [ Q 5b ,H m <strong>QCD</strong>]]<br />

|0〉<br />

[ [ ]]<br />

σ0γ ab = 〈0| Q 5a , Q 5b ,H γ <strong>QCD</strong><br />

|0〉<br />

with a = 1,2,3 corresponding to pion, a = 4,5,6,7 corresponding to kaon and<br />

a = 8 corresponding to eta. For the electrically neutral axial charge oper<strong>at</strong>ors<br />

we have [ ]<br />

Q 5a ,H γ <strong>QCD</strong><br />

= 0<br />

and hence we obtain (Dashen 1979)<br />

σ 0γ (π 0 ) = σ 0γ (K 0 ) = σ 0γ (K 0 ) = σ 0γ (η 0 ) = 0<br />

Since in the terms σ 0γ there are no masses involved we have for them exact<br />

U-spin-invariance one also, i.e. invariance under<br />

( )<br />

( )<br />

d<br />

ψ = → ψ ′ = exp(−iθ a τ a d<br />

)<br />

s<br />

s<br />

and hence one gets exactly (and this defines the constant B)<br />

σ 0γ (π + ) = σ 0γ (K + ) = B<br />

previous expressions<br />

without<br />

isosopin<br />

breakint<br />

126


For the estim<strong>at</strong>ion of the total isospin splitting, origin<strong>at</strong>ing from quark masses<br />

and electromagnetic effects, we proceed now similarly as we know. We assume<br />

〈0| ūu |0〉 = 〈0| ¯dd |0〉 = 〈0| ¯ss |0〉 = A<br />

Then a simple generaliz<strong>at</strong>ion of our previous calcul<strong>at</strong>ion yields:<br />

f π m 2 π + = (m u + m d )A + B<br />

f π m 2 π 0 = (m u + m d )A<br />

f K m 2 K + = (m u + m s )A + B<br />

f K m 2 K 0 = (m u + m s )A<br />

f η m 2 η 0 = 1 3 (m u + m d + 4m s )A<br />

Assuming all the decay constants to be identical we obtain by linear combin<strong>at</strong>ions<br />

the generalized Gell-Mann-Okubo rel<strong>at</strong>ion:<br />

4m 2 K + = 3m2 η + m 2 π +<br />

and<br />

and<br />

We obtain again<br />

and<br />

4m 2 K 0 = 3m2 η + m 2 π 0<br />

m 2 π 0<br />

m 2 K 0<br />

= m u + m d<br />

m s + m d<br />

m 2 K + − m 2 K 0 − m 2 π +<br />

m 2 K 0 − m 2 K + + m 2 π + − 2m 2 π 0<br />

m u + m d<br />

2m s<br />

≃ 1<br />

25<br />

m d<br />

m u<br />

∼ = 1.8<br />

= m d<br />

m u<br />

The consider<strong>at</strong>ions we just have done had a gre<strong>at</strong> impact about 30 years<br />

ago. At th<strong>at</strong> time it was not clear, th<strong>at</strong> the quarks up and down had different<br />

masses and one <strong>at</strong>tributed the mass difference of the physical particles solely<br />

to electromagnetic effects. The consequences would be: The charged pion has<br />

a larger mass than the neutral one, due to the repulsive electric forces in the<br />

charged pions. One expected the same for the kaon. This had lead to the so<br />

called Dashen-sum-rule m 2 K + − m 2 K 0 = m 2 π + − m 2 π 0 However, if one looks<br />

<strong>at</strong> the experimental masses one notices th<strong>at</strong> the neutral kaon is heavier than<br />

127


the charged one. Today one explains this discrepancy with the mass-difference<br />

between up- and down-quark.<br />

Altogether we obtain for the current quark masses (<strong>QCD</strong>-quarks):<br />

m u ≃ 6MeV m d ≃ 10MeV m s ≃ 180MeV<br />

Comment: One should clearly distinguish these current masses from the<br />

constituten masses, which we find in simple non-rel<strong>at</strong>ivistic hadron models.<br />

Those masses are about 350-450 MeV<br />

11 Group theory, Flavour structure and Quark<br />

model<br />

We remember the Lagrangean of the <strong>QCD</strong> (127). In the fol<strong>low</strong>ing we consider<br />

the <strong>QCD</strong>-Lagrangean restricted to up-, down- and strange fields. This is of<br />

course an approxim<strong>at</strong>ion, but if one restricts oneself to the calcul<strong>at</strong>ion of light<br />

mesons and baryons this is in view of their masses (i.e. around and smaller<br />

than 1-2 GeV) probably not too bad. This mutil<strong>at</strong>ed <strong>QCD</strong>–Lagrangean is<br />

SU(3)-flavour invariant (vector symmetry), if all fields are assumed to have the<br />

same mass parameter in the Lagrangean. In fact this is true as far as up- and<br />

down-quark fields are concerned, where it is a good approxim<strong>at</strong>ion to assume<br />

the masses to be the same. If onc includes the strange quark mass in the Lagrangean<br />

one one has to tre<strong>at</strong> it in a different way. One way is, to incorpor<strong>at</strong>e<br />

it by perturb<strong>at</strong>ion theory. It will be shown l<strong>at</strong>er th<strong>at</strong> this is indeed a good<br />

way and th<strong>at</strong> first order is already enough. Thus in the fol<strong>low</strong>ing we will work<br />

out eigenst<strong>at</strong>es of the SU(3)-flavour symmetry for up- down- and strange quark<br />

fields (they form the so called SU(3)-multiplets) and after th<strong>at</strong> tre<strong>at</strong> the term<br />

m s¯ss by first order perturb<strong>at</strong>ion. In fact the particles in n<strong>at</strong>ure correspond to a<br />

good approxim<strong>at</strong>ion to the st<strong>at</strong>es in the multiplets. We particularly consider the<br />

isovector SU(2) and isovector SU(3)-symmetry and derive mesonic and baryonic<br />

st<strong>at</strong>es, which belong to multiplets of this symmetry. So far we are talking only of<br />

the <strong>QCD</strong>-Lagrangean and its approxim<strong>at</strong>e symmetries. However the quantum<br />

numbers of the multiplets can be equally well be constructed by a coupling of<br />

three quarks, if one considers them as the particles of the fundamental represent<strong>at</strong>ion<br />

of SU(3)-Flavour and if one associ<strong>at</strong>es with them three colour degrees of<br />

freedom in a totally antisymmetric st<strong>at</strong>e, and spin- and flavour-degrees of freedom<br />

in symmetric st<strong>at</strong>es. Then practically all currently known hadrons can be<br />

classified as either q¯q st<strong>at</strong>es (mesons) or qqq st<strong>at</strong>es (baryons). This is the reason<br />

for the success of the quark model. However, one should mention, th<strong>at</strong> all the<br />

SU(3)-algebra holds without assuming the existence of quarks. In fact, when<br />

Gell-Mann and Zweig invented the SU(3)-algebra they thought of an abstract<br />

classific<strong>at</strong>ion of the symmetries and not of particles. For the fol<strong>low</strong>ing, however,<br />

it often makes life easier, if one thinks of quarks as real particles forming a<br />

(small) many-body st<strong>at</strong>e with the structure as either q¯q st<strong>at</strong>es (mesons) or qqq<br />

st<strong>at</strong>es (baryons). In the first sections of this chapter this is not necessary since<br />

128


we deal solely with symmetries and fields. The last step builds a bridge to the<br />

quark model and there one makes assumptions going beyond simple symmetry<br />

algebra.<br />

The detailed structure of the baryons and mesons requires a study of the<br />

colour group, the rot<strong>at</strong>ional group and the flavour group. Colour turns out<br />

mostly as a trivial quantum number and requires mostly only a trace. The<br />

rot<strong>at</strong>ional group O(3) is known and it is known th<strong>at</strong> it is isomorphic to SU(2).<br />

It will be shortly reviewed. The flavour group, since it is an SU(3) group, will<br />

be discussed in detail.<br />

11.1 Elements of Lie-group theory<br />

We consider the transform<strong>at</strong>ion of the Lie group<br />

|ψ〉 → |ψ ′ 〉 = U(θ a ) |ψ〉<br />

U(θ a ) = exp(−iθ a T a )<br />

Here |ψ〉 is an abstract Hilbert vector. The θ a are real numbers, a =<br />

1,2,...,N 2 − 1. The oper<strong>at</strong>ors U(θ a ) are members of a the Lie-group SU(N).<br />

The oper<strong>at</strong>ors T a are hermitian and called “gener<strong>at</strong>ors of the Lie-group SU(N)”<br />

and by definition they fulfill the fol<strong>low</strong>ing commut<strong>at</strong>ion rules<br />

[<br />

T a ,T b] = if abc t c Tr(T a T b ) = 1 2 δab<br />

The numbers f abc are called structure functions of the Lie-group. They determine<br />

the group properties uniquely. The Lie-Group SU(N) can be shown to be<br />

of rank r = N − 1. This by definition means th<strong>at</strong> among the N 2 − 1 gener<strong>at</strong>ors<br />

T a there exist r = N − 1 gener<strong>at</strong>ors, which commute with each other.<br />

Each Lie-group has N −1 Casimir-oper<strong>at</strong>ors C i with 1 = 1,2,...N −1, which<br />

are functions of the gener<strong>at</strong>ors and which are defined by the property th<strong>at</strong> each<br />

of them commutes with each of the gener<strong>at</strong>ors and hence with each member of<br />

the group:<br />

[C i ,T a ] = 0 i = 1,2,...N − 1 a = 1,....,N 2 − 1<br />

The N − 1 Casimir oper<strong>at</strong>ors and the r commuting gener<strong>at</strong>ors have common<br />

eigenvectors. They are grouped in multiplets |C 1 ,...C N−1 ,α 1 ,....,α r 〉, where<br />

all eigenvectors within a multiplet are characterized by r = N − 1 quantum<br />

numbers. The multiplets will be explicitely constructed be<strong>low</strong>. If the Hamilton<br />

oper<strong>at</strong>or of a system shows the symmetry corresponding to the group considert,<br />

then it commutes with the group oper<strong>at</strong>ors U(θ a ), then it commutes also with<br />

the gener<strong>at</strong>ors<br />

[H,T a ] = 0<br />

and in particular with the r = N − 1 gener<strong>at</strong>ors, which commute with each<br />

other. Thus the st<strong>at</strong>es of the multiplet are also eigenst<strong>at</strong>es of the Hamiltonian.<br />

129


In this case they all have the same energy. Since they are eigenst<strong>at</strong>es of hermitian<br />

oper<strong>at</strong>ors they form a basis. In this basis the commuting gener<strong>at</strong>ors are diagonal,<br />

the non-commuting ones are non-diagonal.<br />

It is important to define a represent<strong>at</strong>ion of a Lie-group: A represent<strong>at</strong>ion<br />

of a group is a mapping of the group element U(θ a ) on a unitariy nxn m<strong>at</strong>rix<br />

D(U(θ a )):<br />

U(θ a ) ↦→ D(U(θ a ))<br />

where the mapping is such th<strong>at</strong> the group properties are fulfilled:<br />

and<br />

D(U 1 U 2 ) = D(U 1 )D(U 2 )<br />

D(U)D(U) † = 1<br />

The simplest represent<strong>at</strong>ion, with dimension larger than one, is called “fundamental<br />

represent<strong>at</strong>ion”.<br />

11.2 SU(2)-algebra:<br />

We know as example the rot<strong>at</strong>ional oper<strong>at</strong>ors:<br />

U(θ a ) = exp(−iθ a J a ) a = 1,2,3<br />

with<br />

[<br />

J a ,J b] = iɛ abc J c<br />

They form a SU(2) group. Its fundamental represent<strong>at</strong>ion has the dimension 2<br />

and it is given by the Pauli-m<strong>at</strong>rices. The rank of the SU(2) group is given by<br />

r = N − 1 = 1 and hence has one casimir oper<strong>at</strong>or, which we know, of course<br />

C = J 2 = J 2 1 + J 2 2 + J 2 3<br />

By definition J 2 commutes with all gener<strong>at</strong>ors J a , however for the construction<br />

of multiplets one choses always J 3 as the one, whose eigenst<strong>at</strong>es group into<br />

multiplets. The multiplets are charakterized by being eigenst<strong>at</strong>es of J 2 ,J 3 and<br />

indic<strong>at</strong>ed as |j,m〉, where the various multiplets are distinguished by different<br />

values of j and the st<strong>at</strong>es within a multiplet are distinguished by differen values<br />

of m with −j ≤ m ≤ +j. The various multiplets can be constructed easily. We<br />

repe<strong>at</strong> this simple exercise because in the case of SU(3) it is pricipally the same,<br />

but technically more complic<strong>at</strong>ed:<br />

Define for this raising and <strong>low</strong>ering oper<strong>at</strong>ors:<br />

with the fol<strong>low</strong>ing properties<br />

J ± = J 1 ± iJ 2<br />

(J + ) † = J −<br />

(J − ) † = J +<br />

130


Then we have immedi<strong>at</strong>ely the fol<strong>low</strong>ing formulae:<br />

J 2 = 1 2 (J +J − + J − J + ) + J 2 3<br />

[J + ,J − ] = 2J 3<br />

[J ± ,J 3 ] = ∓J ±<br />

One can immedi<strong>at</strong>ely construct the multiplets. Consider normalized eigenst<strong>at</strong>es<br />

|λ,m〉 of J 2 and J 3 :<br />

J 2 |λ,m〉 = λ |λ,m〉<br />

J 3 |λ,m〉 = m |λ,m〉<br />

One can easily show th<strong>at</strong> the st<strong>at</strong>es |λ,m〉 have the fol<strong>low</strong>ing properties under<br />

the action of J ± :<br />

J ± |λ,m〉 = a ± (λ,m) |λ,m ± 1〉<br />

because e.g:<br />

J 3 (J ± |λ,m〉) = ±J ± |λ,m〉 + J ± J 3 |λ,m〉<br />

= ±J ± |λ,m〉 + mJ ± |λ,m〉 = (m ± 1)J ± |λ,m〉<br />

Th<strong>at</strong> gives the name for the raising and <strong>low</strong>ering oper<strong>at</strong>ors since their applic<strong>at</strong>ion<br />

to |λ,m〉 raises and <strong>low</strong>ers the quantum number m by one. For a given λ the<br />

values of m are bounded:<br />

λ − m 2 > 0<br />

because 0 ≤ 〈λ,m| J 2 1 |λ,m〉 + 〈λ,m| J 2 2 |λ,m〉 = 〈λ,m| J 2 − J 2 3 |λ,m〉. For a<br />

given λ we consider the largest m and call this j. Then we have, since m is<br />

largest<br />

J + |λ,j〉 = 0<br />

From this we get by simple calcul<strong>at</strong>ion<br />

Or<br />

0 = J − J + |λ,j〉 = (J 2 − J 2 3 − J 3 ) |λ,j〉 = (λ − j 2 − j) |λ,j〉<br />

λ = j(j + 1)<br />

If onc starts from the smallest value of m, we call it j ′ , then we can show by<br />

similar means th<strong>at</strong><br />

λ = j ′ (j ′ − 1)<br />

From this fol<strong>low</strong>s th<strong>at</strong><br />

j(j + 1) = j ′ (j ′ − 1)<br />

131


or either j ′ = −j or j ′ = j + 1. Since we have assumed th<strong>at</strong> j is the largest<br />

value, we must have<br />

j ′ = −j<br />

By applic<strong>at</strong>ion of J − to |λ,m〉 one changes m → m − 1. By a finite number of<br />

applic<strong>at</strong>ions of J − to |λ,j〉 we must end up <strong>at</strong> |λ,j ′ 〉. Hence j − j ′ must be an<br />

integer. The j − j ′ is known and it is j − j ′ = 2j Hence<br />

j = integer or half-integer<br />

Now we can easily evalu<strong>at</strong>e the coefficients a ± (λ,m) for an arbitrary m in the<br />

fol<strong>low</strong>ing way: We know th<strong>at</strong><br />

We can calcul<strong>at</strong>e explicitely<br />

〈λ,m|J − J + |λ,m〉 = |a + (λ,m)| 2<br />

〈λ,m|J − J + |λ,m〉 = 〈λ,m|J 2 −J 2 3 −J 3 |λ,m〉 = 〈λ,m| j(j +1) −m 2 −m |λ,m〉<br />

with the result th<strong>at</strong><br />

a + (λ,m) = ((j − m) (j + m + 1)) 1 2<br />

a − (λ,m) = ((j + m) (j − m + 1)) 1 2<br />

We can summarize and have then the multiplet:<br />

J 2 |j,m〉 = j(j + 1) |j,m〉<br />

J 3 |j,m〉 = m |j,m〉<br />

J ± |j,m〉 = ((j ∓ m) (j ± m + 1)) 1 2<br />

|j,m ± 1〉<br />

The different multiplets are now those which we know well:<br />

j dim= 2j + 1<br />

0 1<br />

1<br />

2<br />

2<br />

1 3<br />

3<br />

2<br />

4<br />

2 5<br />

132


11.3 SU(2)-multiplets in n<strong>at</strong>ure<br />

We consider the <strong>QCD</strong>-Lagrangean and consider the SU(2)-flavour transform<strong>at</strong>ion<br />

between up- and down-quark fields or quarks. If we consider these quark<br />

fields to be of equal current mass, the Lagrangean is invariant, as it is well<br />

known. We then have the global isospin invariance of the system:<br />

( )<br />

u<br />

ψ = → ψ ′ = exp(−iθ a τ a )ψ<br />

d<br />

This SU(2) is a symmetry of the <strong>QCD</strong>-Hamiltonian and there should be eigenst<strong>at</strong>es<br />

of the <strong>QCD</strong>-Hamiltonian which form SU(2) multiplets. The eigenst<strong>at</strong>es<br />

of the <strong>QCD</strong>-Hamiltonian are particles, which we find in n<strong>at</strong>ure. Thus we expect<br />

the some of the particles to group into multiplets, whose members all have the<br />

same mass. Actually those multiplets are called isospin-multiplets and they are<br />

well known. The difference in mass of the various members of the multiplet is<br />

much less than 1%, only for the pion we have about 5%, because it is r<strong>at</strong>her<br />

light:<br />

particles dimension isospin-multiplet T J mass (MeV)<br />

1 1<br />

nucleon 2 p,n<br />

2 2<br />

938<br />

Sigma 3 Σ + ,Σ 0 ,Σ − 1<br />

1<br />

2<br />

1189<br />

Xi, Cascade 2 Ξ − ,Ξ 0 1 1<br />

2 2<br />

1315<br />

Delta 4 ∆ ++ ,∆ + ,∆ 0 ,∆ − 3 3<br />

2 2<br />

1230<br />

Lambda 1 Λ 0 1<br />

0<br />

2<br />

1115<br />

Pion 3 π + ,π 0 ,π − 1 0 139<br />

Rho 3 ρ + ,ρ 0 ,ρ − 1 1 768<br />

Kaon 2 K 0 ,K + 1 2<br />

0 494<br />

Kaon 2 K ∗0 ,K ∗+ 1 2<br />

1 892<br />

Eta 1 η 0 0 547<br />

Omega 1 ω 0 1 783<br />

Thus we see th<strong>at</strong> there are systems in n<strong>at</strong>ure, which fulfill the SU(2)-<br />

symmetry.<br />

11.4 SU(3)-algebra<br />

The SU(3)-Lie-Group is given by the commut<strong>at</strong>ion rel<strong>at</strong>ions<br />

[<br />

F a ,F b] = if abc F c Tr(F a F b ) = 1 2 δab<br />

with given and well documented structure coefficients f abc . The fundamental<br />

represent<strong>at</strong>ion is 3-dimensional. It is characterized by the Gell-Mann m<strong>at</strong>rices.<br />

The Gell-Mann m<strong>at</strong>rices are a basis for all traceless hermitean m<strong>at</strong>rices of dimensin<br />

3x3. Any hermitian 3x3 m<strong>at</strong>rix can be written as a linear combin<strong>at</strong>ion<br />

of the unity m<strong>at</strong>rix and the eight Gell-Mann m<strong>at</strong>rices. One can constrauct the<br />

133


multipletts in a way completely analogous to the SU(2) case. For this one needs<br />

the fol<strong>low</strong>ing step-up and step-down oper<strong>at</strong>ors:<br />

T ± = F 1 ± iF 2 T 3 (<br />

= F 3<br />

V ± = F 4 ± iF 5 V 3 = 1 3<br />

2 ( 2 Y + T 3)<br />

U ± = F 6 ± iF 7 U 3 = 1 3<br />

2 2 Y − T 3)<br />

Y = √ 2 3<br />

F 8<br />

with the properties<br />

[T 3 ,T ± ] = ±T ± [T + ,T − ] = 2T 3 T-Spin SU(2)<br />

[U 3 ,U ± ] = ±U ± [U + ,U − ] = 2U 3 U-Spin SU(2)<br />

[V 3 ,V ± ] = ±V ± [V + ,V − ] = 2V 3 V-Spin SU(2)<br />

and<br />

[Y,T ± ] = 0 [Y,U ± ] = ±U ± [Y,V ± ] = ±V ±<br />

[T + ,V + ] = [T + ,U − ] = 0 [U + ,V + ] = 0 [T 3 ,Y ] = 0<br />

[T + ,V − ] = −U − [T + ,U + ] = V + [U + ,V − ] = T −<br />

[T 3 ,V ± ] = ± 1 2 V ±<br />

Ther hermiticity properties are given by<br />

T + = (T − ) † U + = (U − ) † V + = (V − ) †<br />

The Casimir are defined as oper<strong>at</strong>ors which commute iwth each gener<strong>at</strong>or<br />

and hence with each group element. They are given in each group theory book<br />

(see F. Stancu). They are denoted as<br />

C 1 =<br />

8∑<br />

i=1<br />

F 2 i = 2i √<br />

3<br />

∑<br />

fijk F i F j F k<br />

C 2 = ∑ d ijk F i F j F k<br />

Here the d ijk are the totally symmetric structure coefficients.<br />

The SU(3) group has rank 2, thus there are 2 commuting gener<strong>at</strong>ors. These<br />

are Y and T 3 with<br />

[Y,T 3 ] = 0<br />

This corresponds to the fact th<strong>at</strong> λ 3 and λ 8 commute. Since there are also<br />

the two Casimir oper<strong>at</strong>ors, which commute with each gener<strong>at</strong>or, the multiplet<br />

is characterized by a complete system of commuting oper<strong>at</strong>ors in the fol<strong>low</strong>ing<br />

way.<br />

C 1 |αβT 3 Y 〉 = α |αβT 3 Y 〉<br />

C 2 |αβT 3 Y 〉 = β |αβT 3 Y 〉<br />

T 3 |αβT 3 Y 〉 = T 3 |αβT 3 Y 〉<br />

Y |αβT 3 Y 〉 = Y |αβT 3 Y 〉<br />

134


By direct calcul<strong>at</strong>ion, using the commut<strong>at</strong>or rules, one can show now th<strong>at</strong><br />

the fol<strong>low</strong>ing equ<strong>at</strong>ions hold<br />

V ± |αβT 3 Y 〉 = a(T 3 ,Y ) ∣ α,β,T3 ± 1 2<br />

,Y ± 1〉<br />

U ± |αβT 3 Y 〉 = b(T 3 ,Y ) ∣ ∣α,β,T 3 ∓ 1 2<br />

,Y ± 1〉<br />

T ± |αβT 3 Y 〉 = c(T 3 ,Y ) |α,β,T 3 ± 1,Y 〉<br />

The multiplets are now certain figures in the T 3 −Y −plane. Due to symmetries<br />

they show a 6-plet structure. Because the oper<strong>at</strong>ors T ± ,T 3 obbey a SU(2)<br />

algebra, the T 3 is integer or half-integer. The units and limits of Y are not<br />

fixed yet. Their difference must be Unity. One yields in the end a structure like<br />

this, which is characterized by several symmetries which all origin<strong>at</strong> from the<br />

sdymmetries of the SU(2) subgoups:<br />

Reflection symmetry w.r. to straight line given by T 3 = 0<br />

Reflection symmetry w.r. to straight line given by U 3 = 0 corresponding<br />

(see above formulae) to Y = 2 3 T 3<br />

Reflection symmetry w.r. to straight line given by V 3 = 0 corresponding (see<br />

above formulae) to Y = − 2 3 T 3<br />

In general a multiplet looks like this:<br />

U<br />

+<br />

Y<br />

V<br />

+<br />

T<br />

3<br />

The explicit construction of a multilplet works now in the fol<strong>low</strong>ing way:<br />

Chose a point, call this |ψ max 〉 and this should be the st<strong>at</strong>e of the largest T 3 -<br />

value and som Y -value. Apparently we have<br />

T + |ψ max 〉 = 0<br />

V + |ψ max 〉 = 0<br />

U − |ψ max 〉 = 0<br />

This point defines the right upper end of the multiplet. In order to draw the<br />

outer most borderlines one applies V − to |ψ max 〉. This is, say, possible p-times.<br />

135


Then we have<br />

(V − ) p+1 |ψ max 〉 = 0<br />

Then apply to this point T − as often as possible. This will go q-times, i.e:<br />

(T − ) q+1 (V − ) p |ψ max 〉 = 0<br />

The rest of the multiplet is given by symmetries. The inner points one can<br />

construct accordingly. The scheme is given in the fol<strong>low</strong>ing way:<br />

Y<br />

ψ<br />

max<br />

V<br />

−<br />

T<br />

3<br />

T<br />

−<br />

One can also calcul<strong>at</strong>e the multiplicity of each point, i.e. the number how<br />

often in this construction a certain point is gener<strong>at</strong>ed. In this way one can<br />

construct various multiplets, from which we note the smallest ones:<br />

notion dimension (p,q)<br />

[1] 1 (0,0)<br />

[<br />

[3]<br />

]<br />

3 (1,0)<br />

3 3 (0,1)<br />

[8] 8 (1,1)<br />

[6] 6 (2,0)<br />

[10] 10 (3,0)<br />

[<br />

10<br />

]<br />

10 (0,3)<br />

11.5 SU(3)-multiplets in n<strong>at</strong>ure<br />

One noteces immedi<strong>at</strong>ely th<strong>at</strong> the particles in n<strong>at</strong>ure group themselves according<br />

to the <strong>low</strong>est dimensional multiplets of SU(3) if one links the hypercharge with<br />

the charge of the particle by means of the famous Gell-Mann-Nishijima rel<strong>at</strong>ion,<br />

which is a phenomenological rel<strong>at</strong>ion <strong>at</strong> the time, when it was discovered:<br />

Q = 1 2 Y + T 3<br />

136


η<br />

0<br />

T<br />

3<br />

Actually one proceeds in the fol<strong>low</strong>ing way: One takes a particle, looks in<br />

the d<strong>at</strong>a for its isospin partners, which are nearly mass-degener<strong>at</strong>e, from the<br />

number of isospin partners one extracts the T 3 . Then one looks for the charge<br />

of the particle, extracts from it and from the now known T 3 the hypercharge<br />

Y according to Gell-Mann-Nishijima rel<strong>at</strong>ion and then one has its quantum<br />

numbers T 3 and Y and its place in the multiplet. In this way one constructs the<br />

fol<strong>low</strong>ing correspondences of particles with multiplets, where we give already<br />

the angular momentum J without telling yet, wlhere it comes from:<br />

Meson-singlet,J=0,T=0, bessere Bezeichnung η 1 , weil Singulett:<br />

Meson-octet: pseudoscalar mesons, J=0, fol<strong>low</strong>s as:<br />

K<br />

0<br />

Y<br />

+<br />

K<br />

−<br />

π<br />

0<br />

π<br />

8<br />

η<br />

+<br />

π<br />

T<br />

3<br />

−<br />

K<br />

0<br />

K<br />

Meson-octet, J=1, fol<strong>low</strong>s as:<br />

137


−<br />

ρ<br />

* 0 Y<br />

* +<br />

K<br />

K<br />

0<br />

ρ<br />

ρ<br />

+<br />

ω<br />

* 0<br />

*<br />

K<br />

− K<br />

T<br />

3<br />

Baryon-octet, J=1/2, fol<strong>low</strong>s as:<br />

Y<br />

0<br />

n<br />

p<br />

− 0<br />

+<br />

Σ Σ Σ<br />

−<br />

Ξ<br />

0<br />

Ξ<br />

T<br />

3<br />

Baryon-decuplet, J=3/2, fol<strong>low</strong>s as:<br />

138


Y<br />

−<br />

∆<br />

0<br />

∆<br />

∆<br />

+<br />

++<br />

∆<br />

∗ −<br />

Σ<br />

∗<br />

Σ<br />

Σ<br />

∗+<br />

T<br />

3<br />

*<br />

−<br />

Ξ<br />

0<br />

*<br />

Ξ<br />

Ω<br />

One should note th<strong>at</strong> in constructing the multiplets we never made anny<br />

assumption about the existence of quarks. In fact they are not needed <strong>at</strong> all in<br />

order to have multiplets. Th<strong>at</strong> is the reason why there are models (e.g. Skyrmemodel<br />

and all its variants), which work reasonably well and do not have quarks<br />

<strong>at</strong> all but are quantum theories of interacting meson fields with some particular<br />

topological structure.<br />

11.6 Quarks and SU(3)<br />

11.6.1 Quarks and fractional charges<br />

We have seen th<strong>at</strong> the mesons and baryons group themselves in a n<strong>at</strong>ural way<br />

in SU(3)-multiplets, if one interpretes Y as hypercharge, which is rel<strong>at</strong>ed to the<br />

charge by the Gell-Mann-Nishijima rel<strong>at</strong>ions:<br />

Q = 1 2 Y + T 3<br />

The smallest represent<strong>at</strong>ion of SU(3) is [1] = D(0,0). This trivial represent<strong>at</strong>ion<br />

is in fact realized in n<strong>at</strong>ure (e.g. Λ and η). If quarks are perhaps the building<br />

blocks of the light baryons and mesons from etsthetical points one would like<br />

them to be in the smallest represent<strong>at</strong>ion. This is not true, howevber one can<br />

m<strong>at</strong>hch the phenomenology if one assumes three quarks of different flavour to<br />

be represented by the smallest NON-trivial represent<strong>at</strong>ion, i.e. the fundamental<br />

represent<strong>at</strong>ion of SU(3)-flavour. Achtually SU(3) has rank 2 and hence 2<br />

139


fundamental repesent<strong>at</strong>ions:<br />

[3] = D(1,0)<br />

[¯3] = D(0,1)<br />

We will see, th<strong>at</strong> the quarks form [3] and the anti-quarks form [¯3]. This can<br />

graphically be represented as:<br />

ψ<br />

2<br />

ψ<br />

1<br />

ψ<br />

*<br />

3<br />

ψ<br />

3<br />

ψ * ψ *<br />

1 2<br />

Quarks: [3] = D(1,0)<br />

Antiquarks: [¯3] = D(0,1)<br />

We can proceed m<strong>at</strong>hem<strong>at</strong>ically as: Consider ψ 1 = ψ max and another ψ 2<br />

with (ψ 1 ,ψ 2 ) forming an isospin dublett, in which then by definition we have<br />

ψ 2 = T − ψ 1 . and<br />

T 3 ψ 1 = 1 2 ψ 1 T 3 ψ 2 = − 1 2 ψ 2<br />

The ψ 3 forms an isospin singlet, i.e. T 3 ψ 3 = 0ψ 3 .<br />

The ψ 2 ,ψ 3 form an U-Spin dublett, i.e. ψ 2 = U − ψ 3<br />

The ψ 1 forms an U-Spin singlet, i.e. U 3 ψ 1 = 0<br />

The ψ 1 ,ψ 3 form an V-Spin dublett, i.e. ψ 3 = V − ψ 1<br />

The ψ 2 forms an U-Spin singlet, i.e. V 3 ψ 2 = 0<br />

These simple fe<strong>at</strong>ures have direct consequences for the charge of the st<strong>at</strong>es<br />

ψ i and ( hence for the particles corresponding the these st<strong>at</strong>es: We have with<br />

U 3 = 1 3<br />

2 2 Y − T )<br />

3 from above the equ<strong>at</strong>ion<br />

and hence<br />

U 3 = 1 4 (3Y − 2T 3)<br />

or<br />

Y ψ 1 = 1 3 (4U 3 + 2T 3 )ψ 1 = 2 3 T 3ψ 1 = 1 3 ψ 1<br />

Qψ 1 = ( 1 2 Y + T 3)ψ 1 = ( 1 6 + 3 6 )ψ 1 = 2 3 ψ 1<br />

and in a similar way charges of ψ 2 and ψ 3 . If we recall now these results we have<br />

alltogether (where we have renamed the quark flavours in the usual way):<br />

Qψ 1 = 2 3 ψ 1<br />

ψ 1 = u<br />

140


Qψ 2 = − 1 3 ψ 2<br />

ψ 2 = d<br />

Qψ 3 = − 1 3 ψ 3<br />

ψ 3 = s<br />

Thus we get autom<strong>at</strong>ically fractional charges for the quarks. We remember:<br />

The assumption was th<strong>at</strong> they occupy the fundamental represent<strong>at</strong>ion fo the<br />

SU(3)-flavour. In the exactly similar way we construct the antiquark triplett:<br />

Qψ 1 = − 2 3 ψ 1<br />

ψ 1 = ū<br />

Qψ 2 = + 1 3 ψ 2<br />

ψ 2 = ¯d<br />

Qψ 3 = + 1 3 ψ 3<br />

ψ 3 = ¯s<br />

The fundamental represent<strong>at</strong>ions [3] and [¯3] are two totally different represent<strong>at</strong>ions.<br />

They cannot be transformed into each other by transform<strong>at</strong>ion of one<br />

or several gener<strong>at</strong>ors. Thus we obtain fractional charges of the quarks if we<br />

demand SU(3) symmetry and the Gellman-Nishijima rel<strong>at</strong>ion and assume th<strong>at</strong><br />

the quarks occupy the smallest non-trivial represent<strong>at</strong>ion.<br />

11.6.2 Construction of higher multiplets by triplets, Confinement<br />

(phenomenological)<br />

Coupling of Fundamental represent<strong>at</strong>ion We will see, th<strong>at</strong> one can construct<br />

the other multiplets in n<strong>at</strong>ure outgoing from these two fundamental represent<strong>at</strong>ions,<br />

if one couples these fundamental represent<strong>at</strong>ionsby means of SU(3)-<br />

Clebsh-Gordan coefficients and constructs successively the higher dimensional<br />

multiplets. We can give examples:<br />

In SU(2), which we take as an example, we couple<br />

−→ 1<br />

2 ⊕ −→ 1<br />

2 = −→ −→ −→<br />

1 1 ⊕ 1<br />

2 = −→ 3<br />

2<br />

−→ 1<br />

2<br />

−→ −→ −→<br />

0 0 ⊕<br />

1<br />

2 = −→ 1<br />

2<br />

In the usual nomencl<strong>at</strong>ure this is written as<br />

[2] ⊕ [2] = [3] + [1]<br />

[2] ⊕ [2] ⊕ [2] = [4] + [2] + [2]<br />

In SU(3) all multiplets can be constructed by coupling in a particular way<br />

the quark triplet and quark anti-triiplet. The general construction is<br />

141


[3] ⊕ [3] ⊕ [3] ⊕ [3] ⊕ [3] ⊕ [ 3 ] ⊕ [ 3 ] ⊕ [ 3 ] = D(p,q)<br />

p-times [3] and q-times 3<br />

The simplest coupling consists in coupling an quark triplet with an quarkantitriplet:<br />

[3] ⊕ [ 3 ] = [1] + [8]<br />

Although the fundamental triplets do not occur freely in n<strong>at</strong>ure, the coupled<br />

ones exist and their members are the well known pseudoscalar mesons, i.e. a<br />

meson octet and meson singlet.<br />

The interesting point is, th<strong>at</strong> not all possible multiplets are realized in n<strong>at</strong>ure.<br />

It is m<strong>at</strong>hem<strong>at</strong>ically possible to construct<br />

[3] ⊕ [3] = [6] + [ 3 ]<br />

However the corresponding particles do not exist. In terms of quarks the resulting<br />

particles should then consist of 2 quarks. This is not possible because of<br />

confinement: The 3 colour degrees of freedom for each of the 2 quarks cannot<br />

be coupled to a colour singlet. And the concept of confinement demands th<strong>at</strong><br />

physical st<strong>at</strong>es are exclusively colour singlets.<br />

The concept of confinement is not yet fully understood. One observes confinement<br />

everywhere and no experiment is known which contradicts the confinement.<br />

Confinement means first, th<strong>at</strong> physical st<strong>at</strong>es, i.e. st<strong>at</strong>es which can be<br />

detected in a detector, must be in a colour singlet. This is realized in the quark<br />

model (see be<strong>low</strong>) by explicitely constructing a colour singlet from three quarks<br />

or from quark-antiquarks. It means second, th<strong>at</strong> all three quarks must be close<br />

together in a small volume of about 1 fm diameter. Actually, the demand of<br />

colourless physical st<strong>at</strong>es is not a consequence of the colour gauge invariance<br />

of the <strong>QCD</strong>-Lagrangean. One can see this immedi<strong>at</strong>ely by an analogy to the<br />

QED, which is invariant under abelian gauge transform<strong>at</strong>ion. There we have<br />

charged st<strong>at</strong>es of different charges and apparently this is comp<strong>at</strong>ible with gauge<br />

invariance.<br />

Thus, if we couple the above non-existing multiplets of [3] ⊕[3] with another<br />

quark triplet, we get physical baryonic (B=1) st<strong>at</strong>es, because three quarks with<br />

different colours can couple to a colour singlet as it is given e.g. in eq.(211) for<br />

the ∆ ++ :<br />

[3] ⊕ [3] ⊕ [3] = [6] ⊕ [ 3 ] ⊕ [3] = [8] + [8] + [10] + [10]<br />

Thus the phenomenological baryons and mesons are properly grouped in<br />

SU(3) flavour -multiplets if those are constructed by an appropri<strong>at</strong>e coupling of<br />

of quark triplets and antiquark-triplets, where the quarks and anti-quarks have<br />

the fol<strong>low</strong>ing quantum numbers:<br />

142


quark T T 3 Q Y S B<br />

1 1 2 1 1<br />

q 1 = u<br />

2 2 3 3<br />

0<br />

3<br />

1<br />

q 2 = d<br />

2<br />

− 1 2<br />

− 1 1 1<br />

3 3<br />

0<br />

3<br />

q 3 = s 0 0 − 1 3<br />

− 2 1<br />

3<br />

−1<br />

3<br />

1<br />

q 1 = u<br />

2<br />

− 1 2<br />

− 2 3<br />

− 1 3<br />

0 − 1 3<br />

1 1 1<br />

q 2 = d<br />

2 2 3<br />

− 1 3<br />

0 − 1 3<br />

1 2<br />

q 3 = s 0 0<br />

3 3<br />

1 − 1 3<br />

Thus we obtain finally the flavour structure of the mesons and baryons. In<br />

fact one obtains Clebsh-Gordan coefficients which yield the combin<strong>at</strong>ions of the<br />

group elements in the coupling process to gener<strong>at</strong>e higher multiplets.<br />

Mesons:<br />

J=0 singlet η 1 = ūu + ¯dd + ¯ss<br />

J=0 octet K 0 ∼ ¯sd K + ∼ ¯su 494MeV<br />

π − ∼ ūd π 0 ∼ ūu − ¯dd π + ∼ ¯du 139MeV<br />

η 8 ∼ ūu + ¯dd − 2¯ss<br />

K − ∼ ūs K 0 ∼ ¯ds J=0 494MeV<br />

ii<br />

J=1 singlet ω 1 = ūu + ¯dd + ¯ss(= φ)<br />

J=1 octet K ∗0 ∼ ¯sd K ∗+ ∼ ¯su 892MeV<br />

ρ − ∼ ūd ρ 0 ∼ ūu − ¯dd ρ + ∼ ¯du 768MeV<br />

ω 8 ∼ ūu + ¯dd − 2¯ss<br />

K ∗− ∼ ūs K ∗0 ∼ ¯ds J=1 892MeV<br />

Baryons:<br />

J=1/2 octet n ∼ udd p ∼ uud 938MeV<br />

Σ − ∼ dds Σ 0 ∼ (ud + du)s Σ + ∼ uus 1189MeV<br />

Λ 0 ∼ (ud − du)s<br />

1115MeV<br />

Ξ − ∼ dss Ξ 0 ∼ uss J=1/2 1315MeV<br />

∆ − ∼ ddd ∆ 0 ∼ udd ∆ + ∼ uud ∆ ++ ∼ uuu 1230MeV<br />

Σ ∗− ∼ dds Σ ∗0 ∼ uds Σ ∗+ ∼ uus 1385MeV<br />

Ξ ∗− ∼ dss Ξ ∗0 ∼ uss 1530MeV<br />

Ω − ∼ sss octet J=3/2 1672MeV<br />

In the above tables there are the masses of the particles given as well. This<br />

is an antiacip<strong>at</strong>ion: At the present level of the discussion the members of a<br />

multiplet are degener<strong>at</strong>e. We will calcul<strong>at</strong>e the splitting due to the finite current<br />

mass of the strange quark in one of the next sections.<br />

Mixing of η and ω There is one comment in order about the pseudoscalar<br />

η-mesons. Actually we have one η as a member of the meson octet: η 8 =<br />

143


(uū+d ¯d−2s¯s) and another η as the only member of a singlet: η 1 = uū+d ¯d+s¯s.<br />

Since both particles have the same quantum numbers and since the pure SU(3)-<br />

symmetry is only an approxim<strong>at</strong>ion (since we neglect the fact th<strong>at</strong> the strange<br />

quark has a finite mass), in n<strong>at</strong>ure these two eta-st<strong>at</strong>es mix, such th<strong>at</strong> there are<br />

two new eta-st<strong>at</strong>es of the fol<strong>low</strong>ing form:<br />

η = η 8 cos θ − η 1 sin θ<br />

η ′ = η 8 sinθ + η 1 cos θ<br />

However one should note: There is the axial anomaly, and due to this we get<br />

an additional mixing mechanism between η and η ′ cre<strong>at</strong>ing two new particles,<br />

which as a surprise are again closer to the original ones, i.e. η 8 and η 0 yielding<br />

η = 0.58(uū + d ¯d) − 0.57s¯s<br />

η ′ = 0.40(uū + d ¯d) + 0.82s¯s<br />

547MeV<br />

958MeV<br />

Thus after these two mixings the s¯s-content of η is similar to its (uū + d ¯d)-<br />

component, which bears a strong similarity to the original structure of η 1 . Similarly<br />

s¯s-content of η ′ is about twice as large as its (uū + d ¯d)-componen, which<br />

reminds of the structure of η 8 .<br />

A similar mixing happens for the pseudoscalar vectormesons, however without<br />

the effect of the anomaly. There is the singlet ω 1 and the octett ω 8 . These<br />

mesons show the so called ideal mixing, which is defined in such a way, th<strong>at</strong> the<br />

s¯s component decouples completely from the (uū+d ¯d)-component. Thus in the<br />

end we have for the physical mesons in a good approxim<strong>at</strong>ion the structures<br />

ω = uū + d ¯d<br />

φ = s¯s<br />

782MeV<br />

1020MeV<br />

These are the famous multiplets of mesons and baryons, which have been<br />

popular for more than 40 years.<br />

OZI-rule In the 1960’s an empirical property, called the Okubo-Zweig-Iizuka<br />

rule (OZI-rule) was developed for mesonic decays and coupling constants. Its<br />

usual st<strong>at</strong>ement is th<strong>at</strong> flavour disconnected pocesses are suppressed compared<br />

to those in which quark lines are nonnected. Unfortuntely the phenomenological<br />

and theoretical st<strong>at</strong>us of this so-called rule is abibuous and several viol<strong>at</strong>ions are<br />

known. Nevertheless in the present language we can express it by the fol<strong>low</strong>ing<br />

picture<br />

Each straight line represents a quark. Thus OZI-al<strong>low</strong>ed transitions are<br />

those, where the quarks of the decaying particle end up as quarks of the daughter<br />

particles. If the flavour structure of the daughter particles is so different from<br />

the mother particle, then intermedi<strong>at</strong>e gluons are required. However, each of<br />

the gluons couples with the strong coupling constant α s (Q 2 ), which is mostly<br />

144


145


¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡<br />

s<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥ ¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

s<br />

u−bar<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥ ¥<br />

¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

¤ ¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥<br />

¥<br />

¥ ¥ ¥ ¥<br />

¥ ¤ ¤ ¤ ¤ ¤<br />

K−minus<br />

¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥ ¥<br />

¤<br />

¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥ ¥<br />

¤<br />

¤ ¤ ¤ ¤<br />

¥ ¥ ¥ ¥ ¥<br />

¤<br />

Phi<br />

¡ ¡ ¡ ¡ ¡<br />

u<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

s−bar<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢ ¢<br />

£ £ £ £<br />

£<br />

£ £ £ £<br />

£ ¢ ¢ ¢ ¢ ¢<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

K−plus<br />

s−bar<br />

¢ ¢ ¢ ¢<br />

£ £ £ £ £<br />

¢<br />

<br />

<br />

<br />

<br />

<br />

<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § § §<br />

§ § § § §<br />

¦ ¦ ¦ ¦ ¦<br />

¦ § ¦ § ¦ § ¦ § ¦ §<br />

§ § § § § ¦ ¦ ¦ ¦ ¦<br />

§ § § § § ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

d<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

s<br />

u−bar<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Pi−minus<br />

<br />

<br />

<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

Phi<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § § §<br />

u<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¨ ¨ ¨ ¨ © © ©<br />

©<br />

© © ©<br />

© ¨ ¨ ¨ ¨<br />

¨ ¨ ¨ ¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

© © © ©<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¨ ¨ ¨ ¨ © © ©<br />

©<br />

© © ©<br />

© ¨ ¨ ¨ ¨<br />

¨ ¨ ¨ ¨ © © ©<br />

©<br />

© © ©<br />

© ¨ ¨ ¨ ¨<br />

Pi−plus<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

¦ ¦ ¦ ¦ ¦<br />

§ § § §<br />

§<br />

§ § § §<br />

§ ¦ ¦ ¦ ¦ ¦<br />

s−bar<br />

d−bar<br />

¨ ¨ ¨ ¨<br />

© © © ©<br />

¨ ¨ ¨ ¨ © © ©<br />

©<br />

© © ©<br />

© ¨ ¨ ¨ ¨<br />

¨ ¨ ¨ ¨ © © ©<br />

©<br />

© © ©<br />

© ¨ ¨ ¨ ¨<br />

¨ ¨ ¨ ¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

© © © ©<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¦ ¦ ¦ ¦<br />

§ § § § §<br />

¦<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

¨ ¨ ¨<br />

© © © ©<br />

¨<br />

smaller than one. Thus the right graph is suppressed compared to the left graph.<br />

This argument is not very convincing since the α s (Q 2 ) is not small for small<br />

Q 2 . Thus kinem<strong>at</strong>ic properties enter as well. In fact there are several exceptions<br />

known, where this OZI rule does not work. There is some justific<strong>at</strong>ion also in<br />

the limit of large N c .<br />

An example is the decay of the φ-meson, which is basically s¯s:<br />

and<br />

It decays to 49% into K + K − and only to 10 −3 % into π + π − , which fol<strong>low</strong>s<br />

the OZI rule as the two pictures show.<br />

M<strong>at</strong>rix represent<strong>at</strong>ion of octets This group represent<strong>at</strong>ion is not always<br />

favourable. There exist effective theories which are not <strong>QCD</strong> in terms of gluons<br />

and quarks but tre<strong>at</strong> solely the mesonic and baryonic degrees of freedom. The<br />

Lagrangean of these theories expressed in terms of U = exp[iλ a φ a ] and this<br />

gives rise to the fol<strong>low</strong>ing set of m<strong>at</strong>rices depending separ<strong>at</strong>ely on the parity<br />

and angular momentum. This yields then for pseudoscalar mesons the field<br />

m<strong>at</strong>rix<br />

φ PS (x) = 1 √<br />

2<br />

λ a φ a PS(x) =<br />

⎛<br />

⎜<br />

⎝<br />

π 0<br />

√<br />

2<br />

+ η8 √<br />

6<br />

π + K +<br />

π −<br />

− π0 √<br />

2<br />

+ η8 √<br />

6<br />

K 0<br />

K − ¯K0 − 2η8<br />

√<br />

6<br />

⎞<br />

⎟<br />

⎠ (208)<br />

146


Here the η meson field is the octet component of the η field, i.e.η 8 , and not the<br />

observed η meson. This is a bit confusing as far as the not<strong>at</strong>ion is concerned.<br />

All upper indices indic<strong>at</strong>e the charge of the particle but in case of the eta, where<br />

it indic<strong>at</strong>es the group component. For the vector mesons we have similarly<br />

φ V (x) = 1 √<br />

2<br />

λ a φ a V (x) =<br />

⎛<br />

⎜<br />

⎝<br />

ρ 0<br />

√<br />

2<br />

+ ω8 √<br />

6<br />

ρ + K ∗+<br />

− √ ρ0<br />

2<br />

+ √ ω8<br />

6<br />

K ∗− ¯K∗0 − 2ω0<br />

ρ −<br />

and for the octet baryons the field m<strong>at</strong>rix<br />

⎛<br />

Ψ B (x) = 1 √<br />

2<br />

λ a ψ a B(x) =<br />

⎜<br />

⎝<br />

Σ 0<br />

√<br />

2<br />

+ Λ0 √<br />

6<br />

Σ + p<br />

Σ −<br />

⎞<br />

K ∗0<br />

√<br />

6<br />

− √ Σ0<br />

2<br />

+ √ Λ0<br />

6<br />

n<br />

Ξ − Ξ 0 − 2Λ0<br />

⎟<br />

⎠ (209)<br />

√<br />

6<br />

⎞<br />

⎟<br />

⎠ (210)<br />

If one is interested in properties of a particular particle field, e.g. the field Σ + ,<br />

one easily finds out the combin<strong>at</strong>ion of λ a which yields this field. Actually one<br />

should note th<strong>at</strong> the normaliz<strong>at</strong>ions of these m<strong>at</strong>rices differ in the liter<strong>at</strong>ure<br />

and often the √ 2 are hidden somewhere else. Here we take the normaliz<strong>at</strong>ion<br />

of Donoghue, Go<strong>low</strong>ich, Holstein. There is a clear transform<strong>at</strong>ion between the<br />

above multiplets and these new fields.<br />

11.7 Fock st<strong>at</strong>es and non-rel<strong>at</strong>ivistic quark model<br />

We have considered fields by now. We also can consider single particle quantized<br />

st<strong>at</strong>es of the fields. This is not exactly done, since the field can e.g. be<br />

interacting, but it is done in the sense of the Lehman-Symanzik-Zmmermann<br />

reduction formalismus. In this sense we can have |π + 〉 = b † π + |0〉 , which can be<br />

considered as the result of canonical quantiz<strong>at</strong>ion of the field π + and the applic<strong>at</strong>ion<br />

of the cre<strong>at</strong>ion oper<strong>at</strong>or b † π + to the vacuum. Since, however, the field<br />

π + is obtained by the product 3x¯3 and one can associ<strong>at</strong>e with the fundamental<br />

triplet a quark, the final hadron Fock st<strong>at</strong>e is <strong>at</strong> least composed of a quark and<br />

a anti-quark cre<strong>at</strong>ion oper<strong>at</strong>or. If one assignes phenonemological masses to the<br />

quarks and anti-quarks one can couple quark-antiquark paires to J = 0,1 and<br />

three quarks can be coupled to J = 1/2,3/2. In all practical quark models an<br />

assumption is made which gre<strong>at</strong>ly simplifies subsequent steps in the analysis,<br />

namely th<strong>at</strong> the sp<strong>at</strong>ial, spin, flavour and colour degrees of freedom factorize, or<br />

in other words, the total wave function is a product of a sp<strong>at</strong>ial wave function,<br />

times a spin-flavour Hilbert vector, and times a colour Hilbert vector. This<br />

assumption al<strong>low</strong>s us to write the cre<strong>at</strong>ion and annihil<strong>at</strong>ion oper<strong>at</strong>ors or the<br />

quarks in terms of the sp<strong>at</strong>ial (n) spin (s = 1/2,m s ),flavour(q = u,d,s) and<br />

color (k = 1,2,3) degrees of freedom. Furthermore it is assumed th<strong>at</strong> the colour<br />

degrees of freedom are fully antisymmetrized, such th<strong>at</strong> the product of the other<br />

degrees must be symmetric, in order to reflect the fermionic character of the<br />

quarks. In the fol<strong>low</strong>ing we consider the quarks as sitting in the same orbital 1sst<strong>at</strong>e<br />

e.g. of some potential. Therefor the sp<strong>at</strong>ial degree of freedom is given by<br />

147


an orbital wavefunction with n = 0,which is by construction symmetric. Thus<br />

the spin-flavour wave function must be symmetric.<br />

Altogother we have cre<strong>at</strong>ion oper<strong>at</strong>ors for quarks and antiquarks of the form<br />

∆ ++<br />

3/2<br />

q † k,m s<br />

= b † (n = 0,q,m s ,k)<br />

¯q † k,m s<br />

= d † (n = 0,q,m s ,k)<br />

All this means th<strong>at</strong> e.g. the fol<strong>low</strong>ing non-rel<strong>at</strong>ivistic st<strong>at</strong>e of the delta-isobar<br />

is written as (sum over double colour indices i,j,k)<br />

∣<br />

∣∆ ++<br />

3/2<br />

〉<br />

= 1 6 ε ijku † i↑ u† j↑ u† k↑<br />

|0〉 (211)<br />

Due to the factoriz<strong>at</strong>ion only in this way a reasonable non-rel<strong>at</strong>ivistic wave<br />

function for the ∆ ++<br />

3/2<br />

can be constructed, which is known from experiment to<br />

be in all respects consistent with a st<strong>at</strong>e of of three up-quarks only. From this<br />

one concludes even th<strong>at</strong> we have exactly 3 colours. In a similar way we obtain<br />

the st<strong>at</strong>e of the rho-+ meson (J=1,T 3 = 1) as<br />

∣ ρ<br />

+<br />

1<br />

〉<br />

=<br />

1 √3 u † i↑ ¯d † i↑ |0〉<br />

In a coordin<strong>at</strong>e represent<strong>at</strong>ion the cre<strong>at</strong>ion oper<strong>at</strong>ors cre<strong>at</strong>e a st<strong>at</strong>e with a certain<br />

wave function in coordin<strong>at</strong>e space, given by e.g. a harmonic oscill<strong>at</strong>or or<br />

a bag or some other apropri<strong>at</strong>e potential. Since this coupling does not change<br />

the flavour structure one can explicitely write down the non-rel<strong>at</strong>ivistic wave<br />

function of the meson or baryon constructing it in a way th<strong>at</strong> the colour is totally<br />

antisymmetrized and hence the spin-flavour part fully symmetrized. The<br />

flavour part is already constructed properly, the spin part has still to be tre<strong>at</strong>ed.<br />

This can be done th<strong>at</strong> one performs Clebsh-Gordan sums over the 3-component<br />

of the spin of the quark fields, coupling first two spins to J=0 and J=1 and<br />

coupling these st<strong>at</strong>es then with the third quark to J=1/2. In fact it does not<br />

m<strong>at</strong>ter in which way the coupling is performed as long as one considers symmetric<br />

spin-flavour wave functions. Inthis way one obtains e.g. for some of the<br />

mesons with J = 0 and some of the baryons with J = 1/2 the fol<strong>low</strong>ing 3-quark<br />

wave functions (sum over colour indices i,j,k included):<br />

st<strong>at</strong>e vectors of the pseudoscalar octet and singlet mesons (J=0)<br />

∣ π<br />

+ 〉 = √ 1 [u † ¯d †<br />

6<br />

i↑ i↓ − u† ¯d † i↓ i↑ ] |0〉<br />

∣ π<br />

− 〉 = 1 √<br />

6<br />

[d † i↑ū† i↓ − d† i↓ū† i↑ ] |0〉<br />

∣ K<br />

+ 〉 = 1 √<br />

6<br />

[u † i↑¯s† i↓ − u† i↓¯s† i↑ ] |0〉<br />

etc.<br />

148


und<br />

st<strong>at</strong>e vectors of baryon spin-1/2 octet<br />

|p ↑ 〉 = √ 1 ε ijk [u †<br />

18<br />

i↓ d† j↑ − u† i↑ d† j↓ ]u† k↑ |0〉<br />

|n ↑ 〉 = 1 √<br />

18<br />

ε ijk [u † i↓ d† j↑ − u† i↑ d† j↓ ]u† k↑ |0〉<br />

|Λ ↑ 〉 = 1 √<br />

12<br />

ε ijk [u † i↓ d† j↑ − u† i↑ d† j↓ ]s† k↑ |0〉<br />

〉<br />

∣<br />

∣Ξ − ↑<br />

= √ 1 ε ijk [s †<br />

18<br />

i↓ d† j↑ − s† i↑ d† j↓ ]s† k↑ |0〉<br />

etc.<br />

As one sees the sum over colours (indices=i,j,k)corresponds for the mesons<br />

to a trace, where colour in the quark is paired with anti-colour in the antiquark,<br />

and corresponds for the baryons to an expression involving the totally<br />

antisymmetrized tensor ε ijk . The orbital wave functions are obtained in the<br />

simplest way by the s-st<strong>at</strong>e of a harmonic oscill<strong>at</strong>or, whose width is adjusted<br />

to some experiments and where the masses of the quarks are about 300-400<br />

MeV. The outcome is then the famous non-rel<strong>at</strong>ivistic quark-model. Such a<br />

149


model had many succsses and we just describe briefly its magnetic properties in<br />

comparison with experiment. Let us assume th<strong>at</strong> the quark magnetic moments<br />

are given by their standard Dirac values<br />

µ q = e q<br />

2M q<br />

whith the quark electric charges e u = 2 3e, etc. (electron has charge −e). The<br />

baryonic magnetic moments are easily evalu<strong>at</strong>ed as m<strong>at</strong>rix elements of the nonrel<strong>at</strong>ivistic<br />

magnetic moment oper<strong>at</strong>or<br />

µ B = 〈B ↑|m z |B ↑〉<br />

−→ m = µu<br />

∑<br />

kss ′ 〈s| −→ σ |s ′ 〉 u † sk u s ′ k + (d) + (s)<br />

where k indic<strong>at</strong>es the colour over which is summarized. Using the above Fockst<strong>at</strong>es<br />

the results for the proton and neutron are simply<br />

µ p = 4 3 µ u − 1 3 µ d<br />

µ n = 4 3 µ d − 1 3 µ u<br />

and in the limit of exact isospin symmetry with m u = m d one finds the famous<br />

rel<strong>at</strong>ion<br />

µ p<br />

µ n<br />

= − 3 2<br />

150


which is close to the observed value µp<br />

µ n<br />

= −1.46, and was and is considered<br />

as a gre<strong>at</strong> success of the quark model. Fixing the constituent mass M u of the<br />

up-quark and the (identical) down quark by the proton magnetic moment<br />

µ p = 4 e u<br />

− 1 e d<br />

=<br />

e [ 4 2<br />

3 2M u 3 2M d 2M u 3 3 + 1 ]<br />

1<br />

3 3<br />

gives a value for the quark masses<br />

= e<br />

2M u<br />

=<br />

[<br />

2.79 e<br />

2M p<br />

]exp<br />

M u = M d = 938MeV = 340MeV<br />

2.79<br />

Obviously these values are quite different from the numbers of 7MeV or 10MeV,<br />

which we have quotet previously. In fact these big masses are called ”constituent<br />

quark masses” and they canbe understood as the result of the spontaneous chiral<br />

symmetry breaking, which is connected with dynamical mass gener<strong>at</strong>ion.<br />

Similar consider<strong>at</strong>ions for the Λ give us an estim<strong>at</strong>e for the strange constituent<br />

quark masse. The magnetic moment µ Λ of the Λ equals the magnetic<br />

moment of the strange quark, µ s , since the ud-pair in the Λ is coupled to spin<br />

zero (see its wavefunction). The µ s depends on the constituent mass of the<br />

strange quark, and we obtain:<br />

µ Λ = µ s = e q<br />

= − 1 e<br />

=<br />

2M q 3 2M s<br />

which implies with M s = 938MeV<br />

3∗0.61<br />

the values<br />

[<br />

−0.61 e<br />

2M p<br />

]exp<br />

M s = 510MeV and M s − M u = 150MeV ∼ m s (212)<br />

One sees here already th<strong>at</strong> one expects a mass splitting between quark-st<strong>at</strong>es<br />

consisting of up- and down-quarks only and those which contain one or two s-<br />

quarks. This mass splitting will be presently ignored, however it will be tre<strong>at</strong>ed<br />

in the next subsection.<br />

The quark model also establishes a simple rel<strong>at</strong>ionship between the proton<br />

magnetic moment and the m<strong>at</strong>rix element for the magnetic diple (M 1 ) transition<br />

between the nucleon and the Delta-isobar. The m<strong>at</strong>rix element<br />

µ p∆ = 〈p ↑|m z<br />

∣ ∣∣∆ +<br />

J z=+ 1 2<br />

enters the photoproduction amplitude for the γp → ∆ + transition. THis is<br />

easily evalu<strong>at</strong>ed with the above non-rel<strong>at</strong>ivistic wave functions yielding<br />

µ p∆ = 2 3√<br />

2µp<br />

This results holds only if one assumes th<strong>at</strong> the momentum transfer is small<br />

compared to the nucleon or delta mass. However in practice the photo excit<strong>at</strong>ion<br />

of the delta requires an energy transfer of about 300 MeV since the mass of the<br />

delta is 1232 MeV and the mass of the nucleon is 938 MeV. So in fact recoil<br />

corrections are not negligeable. A full comparison of the magnetic moments of<br />

the octet baryons with experiment is given in the table:<br />

〉<br />

151


11.7.1 Mass splittings<br />

By now we have considered pure SU(3)-flavour times SU(2)-rot<strong>at</strong>ion and in<br />

principle all st<strong>at</strong>es within a given multiplet are degener<strong>at</strong>e. This is because<br />

we have ignored any mass difference between up-, down- and strange-quarks.<br />

For up- and down-quarks this is well fulfilled in n<strong>at</strong>ure, since e.g. the mass<br />

splitting between Neutron and Proton is only 2 MeV compared to the proton<br />

mass of 938 MeV. For the strange quarks this is no longer a good approxim<strong>at</strong>ion<br />

since its mass is about 150-180 MeV, as we have seen <strong>at</strong> eq.(212). Therefore<br />

we have to take it into account. Today we know th<strong>at</strong> the term m s¯ss in the<br />

<strong>QCD</strong>-Lagrangean is the only term, which can be responsible for the splitting<br />

inside the octet and inside the decuplet. The simplest way to tree<strong>at</strong> it, and it<br />

will be shown th<strong>at</strong> this is indeed sufficient, consists in first order perturb<strong>at</strong>ion<br />

theory, wh<strong>at</strong> we will do now. This will lead to the famous Gell-Mann-Okubo<br />

mass formula.<br />

One can easily show th<strong>at</strong> we have<br />

⎛<br />

m s¯ss = m s¯q ⎝<br />

0 0 0<br />

0 0 0<br />

0 0 1<br />

⎞<br />

⎠ q = m s¯q<br />

[ 1<br />

3 I − √ 1 ] [ ] 1<br />

λ 8 q = m s¯q 3 3 I − Y q<br />

thus the perturb<strong>at</strong>ion is proportional to m s and consists of a constant term and<br />

a term linear in the hypercharge. Thus we have equal spacings between the<br />

various isospin multiplets, since they differ linearly by Y . Thus we have for the<br />

<strong>low</strong>est baryon masses<br />

M Σ = M N + M s<br />

M Ξ = M N + 2M s<br />

M Λ = M N + M s<br />

152


or the Gell-Mann-Okubo rel<strong>at</strong>ion:<br />

M Σ + 3M Λ = 2(M N + M Ξ )<br />

[2.23GeV ] exp<br />

= [2.25GeV ] exp<br />

This rel<strong>at</strong>ion is experimentally well s<strong>at</strong>isfied as one sees. Of course there is also<br />

equal spacing between the members of the decuplet. This actually lead to the<br />

prediction of the Ω − particle, which was then identified experimentally and for<br />

which Gell-Mann obtained the Nobel prize. One should note, th<strong>at</strong> this this<br />

rel<strong>at</strong>ion was discovered <strong>at</strong> a time, where no <strong>QCD</strong> was known yet. Th<strong>at</strong> means<br />

the perturb<strong>at</strong>ion oper<strong>at</strong>or was not known. Thus its algebraic structure had to<br />

be inferred from the experiment.<br />

Fol<strong>low</strong>ing SU(3) and its breaking by the strange mass there is no rel<strong>at</strong>ionship<br />

between the splitting in the octet and the splitting in the decuplet. This just<br />

does not fol<strong>low</strong> from the quark model. However, Guadagnini has extracted<br />

from the Skyme model and phenomenological consider<strong>at</strong>ions an equ<strong>at</strong>ion, which<br />

rel<strong>at</strong>es the masses of the octed and decuplett to each other. The Guadagnini<br />

formula is<br />

8(M Ξ ∗ + M N ) + 3M Σ = 11m Λ + 8M Σ ∗<br />

and both sides differ only by less than 1%. The chiral quark soliton model gets<br />

this result again by first order perturb<strong>at</strong>ion theory in m s in a n<strong>at</strong>ural way. This<br />

is by no means trivial, since e.g. the Skyrme model does not provide it.<br />

All these consider<strong>at</strong>ions show, th<strong>at</strong> the strange mass m s can be considered<br />

as ”small” such th<strong>at</strong> first order perturb<strong>at</strong>ion theory is justified.<br />

11.7.2 Quark model: Calcul<strong>at</strong>ions<br />

The quark model has a long history and it has been improved several times<br />

by using wave functions of more and more sophistic<strong>at</strong>ed potentials. This was<br />

altogether r<strong>at</strong>her successfull in describing the <strong>energies</strong> of the various baryons.<br />

Not only those with the quarks in the orbital s-st<strong>at</strong>e of some 3-dimensional<br />

harmonic oscill<strong>at</strong>or but also st<strong>at</strong>es which correspond to quarks in one or two<br />

single particle excited st<strong>at</strong>es. We quote be<strong>low</strong> the first pioneering calcul<strong>at</strong>ions<br />

by Isgur and Karl:<br />

Their hamiltonian is given by<br />

with<br />

H =<br />

3∑<br />

i=1<br />

(<br />

M i + p2 i<br />

2M i<br />

)<br />

+ K 2<br />

∑<br />

i


The first figure shows the spectrum of 1ħω-neg<strong>at</strong>ive parity nucleon and delta<br />

resonances. the shaded regions indic<strong>at</strong>e the empirical positions and widths, the<br />

solid bars are the predictions of the Isgur-Karl model.<br />

The second figure shows the spectrum of the 2ħω-positive parity nuccleon<br />

and delta resonances.<br />

One should note, however, th<strong>at</strong> there is no fundamantal reason for the potentials<br />

used in those models and also for the use of the 3-quark wave functions.<br />

There are several inner inconsistencies in the model:<br />

First, the average momentum of a constituent quark inside the harmonic<br />

oscill<strong>at</strong>or potential can easily be calcul<strong>at</strong>ed and is of the same order of magnitude<br />

as its constituent mass, which means th<strong>at</strong> a rel<strong>at</strong>ivistic description is<br />

needed. There are those extensions, however their conceptual background is<br />

not there, since they are not rel<strong>at</strong>ivistic field theories. Second: The assumption<br />

of a three-quark st<strong>at</strong>e with constituent masses is a pure assumption which<br />

cannot be derived from <strong>QCD</strong>. In <strong>QCD</strong> one has dynamical mass gener<strong>at</strong>ion due<br />

to spontaneous breakdown of chiral symmetrie, however this mass is momentum<br />

dependent and serves also as a quark-pion coupling constant. Furthermore,<br />

effects like the strange content of the nucleon - a r<strong>at</strong>her topical problem - or<br />

strange contributions to parton distribution functions are not described in the<br />

quark models.<br />

However, the virtue of those crude models lies in their phenomenological<br />

simplicity. They work to some extent, in particular for excit<strong>at</strong>ion <strong>energies</strong> of<br />

the mesons and baryons, although one does not really know, why. Altogether no<br />

rasoning in terms of a quantum field theory is known which justifies the whole<br />

154


non-rel<strong>at</strong>ivistic quark model or its ”rel<strong>at</strong>ivised” variants.<br />

12 Effective bosonic Lagrangean<br />

The objective of this section is to construct an effective theory, which describes<br />

the interaction between Goldstone-bosons. The term “effective” means here,<br />

th<strong>at</strong> the fields entering the Lagrangean are not the quark or gluon fields but<br />

the fields of the Goldstone bosons, which are in fact composite particles. The<br />

construction of the effective Lagrangean will be based purely on the concept<br />

of spontaneously broken chiral symmetry. Whenever useful we will make connections<br />

to <strong>QCD</strong>. However, one should note, th<strong>at</strong> never the de<strong>at</strong>illed kowledge<br />

of <strong>QCD</strong> in terms of quarks and gluons is required, only the notion of chiral<br />

symmetry and its spontaneous breakdown. At the end we want a Lagrangean<br />

of the form o<br />

L = L eff (GoldstoneFields,ExternalFields)<br />

and a clear prescription how to use this Lagrangean in order to calcul<strong>at</strong>e a certain<br />

process between Goldstone Bosons. The Goldstone Bosons, although being<br />

composite particles, are tre<strong>at</strong>ed as fields. On the level of <strong>QCD</strong> the oper<strong>at</strong>ions<br />

we will do correspond to a SU(3)-flavour <strong>QCD</strong>, i.e. a <strong>QCD</strong> where the heavy<br />

quark degrees of freedom have already been integr<strong>at</strong>ed out.<br />

155


12.1 Review: Spontaneous breakdown of chiral symmetry<br />

12.1.1 Current algebra and transform<strong>at</strong>ion properties<br />

We recollect the spontaneous chiral symmetry breaking of the <strong>QCD</strong> and as we<br />

see it in n<strong>at</strong>ure: The <strong>QCD</strong> Lagrangean is in the massless limit invariant under<br />

SU(3) L ⊗SU(3) R ⊗U(1) V . This yields conserved left-handed and right-handed<br />

currents and the fermion current (which will be ignored in the fol<strong>low</strong>ing). We<br />

have then left-handed and right-handed charges Q a L and Qa R and vector- and<br />

axial-vector combin<strong>at</strong>ions:<br />

Q a V = Q a R + Q a L<br />

with the vector algebra<br />

Q a L = Q a R − Q a L<br />

[<br />

Q a (t),Q b (t) ] = if abc Q c (t)<br />

The ground st<strong>at</strong>e of the system is invariant under the vector transform<strong>at</strong>ion<br />

exp(−iα a Q a V ) |0〉 = |0〉 Q a V |0〉 = 0 a = 1....8<br />

>From Coleman´s theorem it fol<strong>low</strong>s then th<strong>at</strong> the Hamilton oper<strong>at</strong>or H of the<br />

system commutes with the vector charge:<br />

[Q a V ,H] = 0<br />

and hence the energy eigenst<strong>at</strong>es can be grouped into SU(3)-multiplets where<br />

all members of a certain multiplet have the same mass.<br />

However, the ground st<strong>at</strong>e is not invariant under the corresponding axial<br />

transform<strong>at</strong>ions<br />

exp(−iα a Q a A) |0〉 ≠ |0〉 Q a V |0〉 ≠ 0 a = 1....8<br />

In fact for each Q a A there exists a Goldstone-Boson, which is massless, has the<br />

quantum numbers of the axial current (pseudoscalar, spinn zero). Actually<br />

there are 8 gener<strong>at</strong>ors of SU(3) L and also 8 gener<strong>at</strong>ors of SU(3) L i.e. altogether<br />

16 gener<strong>at</strong>ors. Since the subgroup SU(3) V has 8 gener<strong>at</strong>ors there must be<br />

(8 + 8) − 8 = 8 Goldstone bosons. The goldstone bosonic fields π a (x) a =<br />

1....8 transform under the subgroup SU(3) V of SU(3) L ⊗SU(3) R like an octett<br />

represent<strong>at</strong>ion: [<br />

Q<br />

a<br />

V ,π b (x) ] = if abc π c (x)<br />

12.1.2 Chiral Quark Condens<strong>at</strong>e<br />

Actuall there is a direct connection between the appearance of the Goldstoe<br />

bosons and the scalar quark condens<strong>at</strong>e 〈0| ¯qq |0〉 = 〈0|ūu+ ¯dd+ ¯ss |0〉. In order<br />

to recall this we define the scalar quark density<br />

s a (x) = ¯q(x)λ a q(x)<br />

156


√<br />

with λ a being the flavour Gell Mann m<strong>at</strong>rices and λ 0 =<br />

now show by direct calcul<strong>at</strong>ion th<strong>at</strong> we have<br />

[<br />

Q<br />

a<br />

V ,s 0 (x) ] = 0 a = 1....8<br />

2<br />

3<br />

diag(1,1,1) One can<br />

[<br />

Q<br />

a<br />

V ,s b (x) ] = if abc s c (x) a,b,c = 1.....8<br />

The proof for this needs the simple facts th<strong>at</strong> Q a V = ∫ d 3 x¯q(x)λ a q(x) and<br />

the simple fact th<strong>at</strong> for an oper<strong>at</strong>or A(x) = q †(x)Âq(x) and B(x) = q† (x) ̂Bq(x)<br />

with  and ̂Bsome Dirac-Flavour-Colour-m<strong>at</strong>rices, ]<br />

we have [A(x,t),B(y,t)] =<br />

δ (3) (x − y)q † (x)[Â, ̂B q(x). With help of the rel<strong>at</strong>ion f abc f abd = 3δ cd we immedi<strong>at</strong>ely<br />

obtain<br />

s a (x) = − i 3 f [<br />

abc Q<br />

b<br />

V ,s c (x) ]<br />

For an SU(3) V invariant vacuum we have therefore<br />

〈0| s a (x) |0〉 = 〈0| − i 3 f [<br />

abc Q<br />

b<br />

V ,s c (x) ] |0〉 = 0 a = 1.....8<br />

For 〈0|s 0 (x) |0〉 there does not exist an equ<strong>at</strong>ion like this and hence in general<br />

we have 〈0|s 0 (x) |0〉 ≠ 0. Because of〈0|s 3 (x) |0〉 = 〈0|s 8 (x) |0〉 = 0 fol<strong>low</strong>s then<br />

a finite condens<strong>at</strong>e.<br />

0 ≠ 〈0| ¯q(x)q(x) |0〉 = 3 〈0|ū(x)u(x) |0〉 = 3 〈0| ¯d(x)d(x)s |0〉 = 3 〈0| ¯s(x)s(x) |0〉<br />

This is the direct connection between spontaneous symmetry breaking and the<br />

existence of a chiral condens<strong>at</strong>e.<br />

12.2 Represent<strong>at</strong>ions of the chiral boson fields<br />

12.2.1 Linear Sigma Model<br />

We know the linear Sigma-Model without explicit fermion mass term:<br />

L = ψ [iγ µ ∂ µ ]ψ + 1 2<br />

(<br />

∂λ π a ∂ λ π a + ∂ λ σ∂ λ σ ) ::::::::::::::::::::<br />

:::::::::::::::::::: −gψ(σ + iτ a π a γ 5 )ψ − µ2<br />

2 (σ2 + π a π a ) − λ 4 (σ2 + π a π a ) 2<br />

We know th<strong>at</strong> this model exhibits the spontaneous symmetry breaking. In fact<br />

it has been constructed to do so In tre<strong>at</strong>ing this Lagrangean it is useful to<br />

rewrite the mesons in terms of a m<strong>at</strong>rix field<br />

Σ = σ + iτ a π a<br />

157


such th<strong>at</strong><br />

σ 2 + π a π a = 1 2 Tr(Σ† Σ)<br />

Then we obtain the well known form<br />

L = ψ R [iγ µ ∂ µ ]ψ R + ψ L [iγ µ ∂ µ ]ψ L + 1 4 Tr(∂ µΣ∂ µ Σ † )<br />

+ 1 4 µ2 Tr(ΣΣ † ) − λ [<br />

Tr(ΣΣ † ) ] 2 (<br />

− g ψL Σψ R + ψ R Σ † )<br />

ψ L<br />

16<br />

The above lagrangean<br />

⊗<br />

has separ<strong>at</strong>e Left- and right-invariances, i.e. it is invariant<br />

under SU(2) L SU(2)R :<br />

ψ R (x) → ψ ′ R(x) = Rψ R (x)<br />

ψ L (x) → ψ ′ L(x) = Lψ L (x)<br />

Σ ′ = LΣR †<br />

with R and L being two arbitrary SU(2) m<strong>at</strong>rices.<br />

R(α R ) = exp(−i τa αR<br />

a )<br />

2<br />

L(α L ) = exp(−i τa αL<br />

a )<br />

2<br />

In the case of spontaneous symmetry breaking we separ<strong>at</strong>e the vacuum value v<br />

of the sigma field and write<br />

L = ψ [iγ µ ∂ µ − gv]ψ + 1 2<br />

(<br />

∂λ˜σ∂ λ˜σ − 4C 2 v˜σ 2) + 1 2 ∂ λπ a ∂ λ π a + :::::::::<br />

::::::::::::::::::::::: −gψ(˜σ + iτ a π a γ 5 )ψ − vC 2˜σ(˜σ 2 + π a π a ) + C 2 (˜σ 2 + π a π a ) 2 . The<br />

interactions in the Lagrangean are polynomic couplings without deriv<strong>at</strong>ives.<br />

We will se immedi<strong>at</strong>ely th<strong>at</strong> this structure of the Lagrangean corresponds to a<br />

particular choice of parametrizing the fields. One can write down a Lagrangean<br />

with a completely different form with completely different interactions, still<br />

describing the same physics.<br />

12.2.2 Non-linear Sigma Model<br />

We can rewrite the linear Sigma model to the non-linear one by using an exponential<br />

represent<strong>at</strong>ion, which in fact corresponds to a polar represent<strong>at</strong>ion:<br />

Σ = σ + iτ a π a = v + ˜σ + iτ a π a = (v + S)U<br />

with<br />

U = exp( i v ⃗τ⃗π′ )<br />

158


If you do a Taylor expansion of the U-field then one obtains an expansion in<br />

povers of ( 1 v )n . In detail:<br />

Σ = v + ˜σ + iτ a π a = (v + S)(1 + i v ⃗τ⃗π′ + 1 2 ( i v )2 (⃗τ⃗π ′ ) 2 + ....<br />

= v + S + i −→ τ −→ π + higher − Terms − in − 1 v<br />

Or<br />

π a = (π ′ ) a + ∑ ( 1 v )n F n ((π ′ ) a ,S)<br />

with F n ((π ′ ) a ,S = 0) = 0. Hence, asymptotically in a large distance from the<br />

interaction region, where S → 0, we have π a = (π ′ ) a . This means th<strong>at</strong> the free<br />

particles in their asysmptotic sc<strong>at</strong>tering st<strong>at</strong>es are identical. The lagrangean<br />

of the linear and non-linear Sigma model are only rewritten in a way, th<strong>at</strong> the<br />

free particles are identical. A particular S-m<strong>at</strong>rix element must not depend on<br />

the way we define the fields, as long as the physical particles are not redefined.<br />

Thus both represent<strong>at</strong>ions describe the identical physics.<br />

Actually the non-linear represent<strong>at</strong>ion will play in the fol<strong>low</strong>ing a dominant<br />

role, since it simplifies certain parts of the formalism tremendously. In fact, it<br />

will turn out to be the corner stone of the effective <strong>low</strong> energy theory of the<br />

<strong>QCD</strong>. It is interesting to note the fol<strong>low</strong>ing point. If we perform a left- and<br />

right-handed transform<strong>at</strong>ion the U-field changes according to<br />

Is this correct<br />

U → U ′ = RUL †<br />

and we obtain explicitely<br />

L = 1 2 (∂ µS∂ µ S − 2µ 2 S 2 ) +<br />

(v + S)2<br />

Tr(∂ µ U∂ µ U † ) − λvS 3 − λ 4<br />

4 S4 +<br />

¯ψiγ µ ∂ µ ψ − g(v + S)( ¯ ψ L U † ψ R + ¯ ψ R Uψ L )<br />

Apparently the bosonic part of the Lagrangean is written down in a r<strong>at</strong>her<br />

compact form, i.e. Tr(∂ µ U∂ µ U † ) and th<strong>at</strong> is in the end one of the reasons<br />

why one prefers this represent<strong>at</strong>ion. On the other hand, if one expands the<br />

exponentials, one obtains immedi<strong>at</strong>ely deriv<strong>at</strong>ives acting on the physical pion<br />

fields. Nevertheless, as mentioned above already, both Lagrangeans describe the<br />

same physics, S-m<strong>at</strong>rixelements for a certain process are bound to be identical.<br />

This has been explicitely shown in the book by Donoghue et al p.100.<br />

12.2.3 Generaliz<strong>at</strong>ion to finite Goldstone Boson mass<br />

We will l<strong>at</strong>er spply the chiral perturb<strong>at</strong>ion theory in a realistic way and hence<br />

we have to consider finite pion masses. We want to study this in the nonlinearrepresent<strong>at</strong>ion,<br />

which we write down without the scalar field S:<br />

L = 1 4 Tr(∂ µΣ∂ µ Σ † )<br />

159


+ 1 4 µ2 Tr(ΣΣ † ) − λ [<br />

Tr(ΣΣ † ) ] 2<br />

16<br />

This Lagrngean is invariant under the transform<strong>at</strong>ion<br />

Σ → LΣR †<br />

We add now a symmetry breaking term, wich will do wh<strong>at</strong> we want, i.e.<br />

L sb = ɛTr(Σ + Σ † )<br />

Apparently the term is invariant only under a vector transform<strong>at</strong>ion, i.e. if<br />

we have then e.g.<br />

L = R = V<br />

Tr(Σ + Σ † ) → Tr(V ΣV † ) = Tr(ΣV † V ) = Tr(Σ)<br />

However the term is variant under the axial transform<strong>at</strong>ion, where L † = R = A:<br />

Tr(Σ) → Tr(AΣA) = Tr(ΣA 2 ) ≠ Tr(Σ)<br />

We again change to the non-linear or polar represent<strong>at</strong>ion)<br />

Σ = σ + iτ a π a = v + ˜σ + iτ a π a = (v + S)U<br />

for which we can write down the mass term as<br />

L sb = ɛ 4 (v + S)Tr(U + U † )<br />

If one expands now the symmetry breaking Lagrangean in powers of π a one<br />

obtains<br />

L sb = ɛ 4 (v + S)Tr(2 − τa π a<br />

) 2 + ....<br />

F 0<br />

or<br />

L sb = ɛ (v + S) −<br />

ɛ<br />

4<br />

which correspondes to a pion mass of<br />

m 2 π = ɛ<br />

F 0<br />

2F 0<br />

π a π a ....<br />

Hence the total Lagrangean for the Goldstone fields (non-linear Sigma model)<br />

can be written in the simple form<br />

L = f2 π<br />

4 Tr(∂ µU∂ µ U † ) + m2 π<br />

4 f2 πTr(U + U † )<br />

We will see l<strong>at</strong>er th<strong>at</strong> this is indeed the <strong>low</strong>est order term of the effective Lagrangean<br />

160


12.3 <strong>QCD</strong> and chiral fields<br />

Let us consider <strong>QCD</strong> and let us perfom the usual transform<strong>at</strong>ions<br />

R(α R ) = exp(−i τa αR<br />

a )<br />

2<br />

L(α L ) = exp(−i τa αL<br />

a )<br />

2<br />

acting on the left- and right-handed quark fields separ<strong>at</strong>ely. In doing the transform<strong>at</strong>ions<br />

the Goldstone boson fields transform themeselves like an octett, as<br />

we know. One can now easily define the <strong>QCD</strong>-analogue of the m<strong>at</strong>rix U of the<br />

non-linear sigma model:<br />

U(x) = exp i φ(x)<br />

F 0<br />

with<br />

φ(x) = λ a π a (x)<br />

and a = 1,2,3 corresonding to pion, a = 4,5 corresonding to K + ,K 0 , and<br />

a = 6,7 corresponding to K − , ¯K 0 , and a = 8 corresponding to η. Then we have<br />

⎛<br />

√ √<br />

π 0 + 1<br />

⎞<br />

√<br />

3<br />

η 8 2π<br />

+ 2K<br />

+<br />

⎜ √ √<br />

φ(x) = ⎝ 2π<br />

−<br />

−π 0 + √ 1 3<br />

η 8 2K<br />

0 ⎟<br />

√ √<br />

⎠<br />

2K<br />

− 2K<br />

0<br />

− √ 2<br />

3<br />

η 8<br />

Actually: If the quarks are transformed as described above, the transform<strong>at</strong>ion<br />

law of U(x) turns out to be extremely simple. In fact we have<br />

U(x) → U ′ (x) = RU(x)L †<br />

The proof goes as fol<strong>low</strong>s: We will first show, how the bosonic fields π a in<br />

the above expression for U transform under the vector transform<strong>at</strong>ion if this is<br />

applied to U. Then we compare this with the old expressions for the transform<strong>at</strong>ion<br />

of boson field (we know: like an octet) and show th<strong>at</strong> they are identical.<br />

Thus:<br />

For a simple vector transform<strong>at</strong>ion we have R = L. Put for simplicity<br />

R = L = V and then we have<br />

U → U ′ (x) = V U(x)V †<br />

The expansion of the exponential function yields<br />

U(x) = 1 + i φ(x) − 1 F 0 2<br />

→ U ′ = V<br />

(<br />

1 + i φ(x) − 1 F 0 2<br />

φ 2<br />

F0<br />

2<br />

φ 2<br />

F0<br />

2<br />

+ ...<br />

)<br />

+ ... V †<br />

161


→ U ′ = V V † + iV φ F 0<br />

V † − 1 2 V φ F 0<br />

V † V φ F 0<br />

V † + ...<br />

→ U ′ = 1 + iV φ V † − 1 F 0 2 V φ2<br />

F0<br />

2 V † + ...<br />

>From the lase expression we can read off th<strong>at</strong> the φ(x) transformes as<br />

φ(x) → V φ(x)V †<br />

>From this one can show easily, th<strong>at</strong> the π a -fields transform like an octet. Take<br />

for this the usual parametriz<strong>at</strong>ion<br />

V (α V ) = exp(−i λa α a V<br />

2<br />

Then we obtain<br />

[ ] λ<br />

φ(x) = λ b π b (x) → V φ(x)V † = φ(x) − iαV<br />

a a<br />

2 ,λb π b (x) + ...<br />

= φ(x) + F abc α a V π b (x)λ c + ....<br />

This expression can be compared with the one, we had already: Because we<br />

know<br />

[<br />

Q<br />

a<br />

V ,π b (x) ] = if abc π c (x)<br />

Th<strong>at</strong> actually ment, th<strong>at</strong> the Goldstone fields transform like members of the<br />

octett. This simple and dimportant expression can be immedi<strong>at</strong>ely generalized<br />

to<br />

]<br />

exp(i αa V Qa V<br />

2<br />

)(λ b π b (x))exp(−i αa V Qa V<br />

2<br />

)<br />

) = λ b π b (x) + iα a V<br />

= φ(x) + f abc α a V π b (x)λ c + ....<br />

[ λ<br />

a<br />

2 ,λb π b (x)<br />

+ ...<br />

Apparently both expressions , one derived from U and the other one derived<br />

from the known octett transform<strong>at</strong>ion, are identical henc the transform<strong>at</strong>ion<br />

law of U is proven by this.<br />

We can also consider an axial transform<strong>at</strong>ion. There we have R = L † and<br />

we set R = L † = A Now we can repe<strong>at</strong> the above steps and transform<br />

U(x) → Aφ(x)A = AA + iA φ(x) A − 1 F 0 2 Aφ(x) φ(x)<br />

A + ...<br />

F 0 F 0<br />

Here we have AA ≠ 1 in contrast to above V V † = 1 and hence the transfom<strong>at</strong>ion<br />

properties of φ(x) and of the pionic filds π a (x) are not th<strong>at</strong> simple. This<br />

corresponds to the fact th<strong>at</strong> the vacuum is not invariant under the axial transform<strong>at</strong>ion.<br />

We also cannot insert a term AA between the two φ-fields, because<br />

AA ≠ 1.<br />

162


Altogether we can summarize this subsection: We have an appropri<strong>at</strong>e field<br />

U(x) whose transform<strong>at</strong>ion properties are simple and known. The U(x) is<br />

parametrized in the physical fields (Goldstone fields). The ground st<strong>at</strong>e of<br />

the system is described by U(x) = 1. This st<strong>at</strong>e is invariant under vector<br />

transform<strong>at</strong>ion, since V V † = 1, however it is not invariant under the axial<br />

transform<strong>at</strong>ion since AA ≠ 1.<br />

12.4 The minimal effective bosonic Lagrange-density<br />

12.4.1 Massless case<br />

We consider a system of Goldstone bosons and their interactions. The simplest<br />

Lagrange density describing this is given by<br />

L eff = F 2 0<br />

4 Tr(∂ µU∂ µ U † )<br />

with<br />

U = exp( i τ a π a )<br />

F 0<br />

In fact it is the simplest Lagrangean, if we restrict ourselves to 2 deriv<strong>at</strong>ives.<br />

For more deriv<strong>at</strong>ives see l<strong>at</strong>er. The F 0 = f π is the pion decay constant in<br />

the massles limit. It is taken equal for pions and kaons, which correspondes<br />

approxim<strong>at</strong>ely to the empirical values of f π = 93MeV adn f K = 113MeV . The<br />

term F0 2 takes care th<strong>at</strong> the kinetic energy of the Goldstone bosons have the<br />

standardform (remember: Tr(λ a λ b ) = 2δ ab )<br />

T = 1 2 ∂ µπ a ∂ µ π a<br />

One sees this explicitely if one expands the exponentials in U and writes down<br />

the various terms in the Goldstone fields π a . One obtains in SU(2):<br />

L eff = 1 2 ∂ µπ a ∂ µ π a 1<br />

1 + 1 π a π a<br />

2 F0<br />

2<br />

= 1 2 ∂ µπ a ∂ µ π a (1 − 1 π a π a<br />

2 F0<br />

2 + −....)<br />

Apparently there are many more terms behind the kinetic energy. They will be<br />

used indeed in order to dewscribe properly e.g. pion-pion-sc<strong>at</strong>tering or pionkaon-sc<strong>at</strong>tering<br />

etc.<br />

We will prove all these st<strong>at</strong>ements, but before we do so we will generalize<br />

the above Lagrangean by including a term, which gives the Goldstone bosons a<br />

finite mass. We know from the non-linear Sigma model, how to do this. Thus<br />

the final simplest effective Lagrangean with finite Goldstone boson mass is (this<br />

will be proven immedi<strong>at</strong>ely):<br />

L eff = F 2 0<br />

4 Tr(∂ µU∂ µ U † ) + F 2 0<br />

2 B 0Tr(MU † + UM † )<br />

Actually the B 0 has to do with the chiral vacuum condens<strong>at</strong>es and the current<br />

quark masses. If we consider SU(2) and the case m u = m d because then the<br />

163


mass term can be written interms of the pion mass and is then identical to the<br />

one of the non-linear Sigma model<br />

F 2 0<br />

2 B 0Tr(MU † + UM † ) → m2 π<br />

Tr(U + U † )<br />

We will proceed with the prof by considering first the massless limit.<br />

Proof: There are other expressions of second order in the deriv<strong>at</strong>ives, they<br />

are however equivalent to the above one and can berewritten as e.g:<br />

Tr((∂ µ ∂ µ U)U † ) = ∂ µ<br />

[<br />

Tr(∂ µ UU † ) ] − Tr(∂ µ U∂ µ U † )<br />

The first term is a total deriv<strong>at</strong>ive and hence irrelevant for the Lagangean.<br />

A term of the sort Tr(U + U †) contributes to the masses of the Goldstone<br />

bosons and will be considered separ<strong>at</strong>ely and l<strong>at</strong>er. A term of the sort Tr(U−U †)<br />

does not contribute to the mass, however it is not al<strong>low</strong>ed since it has the wrong<br />

behaviour under parity transform<strong>at</strong>ion. Thus:<br />

We have to show finally, th<strong>at</strong> terms with ONE deriv<strong>at</strong>ive do not contribute.<br />

The reason is simple because<br />

Tr ( ∂ µ UU † = 0 )<br />

However we will show this explicitely. Consider for this the case F 0 = 1. Then<br />

we have<br />

U = exp(iφ) = 1 + iφ + 1 2 (iφ)2 + ...<br />

f 2 π<br />

Tr [∂ µ U] = exp(iφ)i∂ µ φ<br />

Tr [ ∂ µ UU †] = Tr [ i∂ µ φUU †] = tr [i∂ µ φ]<br />

= Tr [i∂ µ π a (x)λ a ] = 0<br />

since we have Tr(λ a ) = 0.<br />

Currents:<br />

We construct explicitely the Noether-currents for the above effective Lagrangean<br />

We will show th<strong>at</strong><br />

L eff = F 2 0<br />

4 Tr(∂ µU∂ µ U † )<br />

J µ,a<br />

L<br />

(x) = if2 π<br />

4 Tr [ λ a ∂ µ U † U ]<br />

and hence<br />

J µ,a<br />

V<br />

J µ,a<br />

R<br />

(x) = if2 π<br />

4 Tr [ λ a U∂ µ U †]<br />

= J µ,a<br />

R<br />

+ Jµ,a L<br />

= −if2 π<br />

4 Tr ( λ a [ U,∂ µ U †])<br />

164


J µ,a<br />

A<br />

= Jµ,a R<br />

− Jµ,a L<br />

= −if2 π<br />

4 Tr ( λ a { U,∂ µ U †})<br />

The proof for these expressions for the currents goes as fol<strong>low</strong>s: We parametrize<br />

in the konwn way<br />

L = exp(−i θa L (x) λ a )<br />

2<br />

and we take θR a = 0 and we take these parameters x-dependent in order to derive<br />

currents. For the infinitesimal transform<strong>at</strong>ion we have<br />

)<br />

U → U ′ = RUL † = U<br />

(1 + i λa<br />

2 θa L<br />

and therefrom<br />

U † → (U ′ ) † = LU † R † =<br />

) (1 − i λa<br />

2 θa L U †<br />

∂ µ U → ∂ µ U ′ = (∂ µ U)(1 + i λa<br />

2 θa L) + 1 2 U(i∂ µθ L )λ a<br />

∂ µ U † → ∂ µ (U ′ ) † = (1 − i λa<br />

2 θa L)(∂ µ U † ) + 1 2 (−i∂ µθ L )λ a U †<br />

Herewith we obtain for δL (only terms linear in θL a are kept):<br />

(<br />

δL = f2 π<br />

4 Tr Ui∂ µ θL<br />

a λ a<br />

)<br />

2 ∂µ U † + ∂ µ U(−i∂ µ θL) a λa<br />

2 U †<br />

If one uses trace-properties one obtains<br />

(<br />

δL = f2 π λ<br />

4 i(∂ µθL)Tr<br />

a a [<br />

∂ µ UU † − U † ∂ µ U ])<br />

2<br />

= f2 π<br />

4 i(∂ µθ a L)Tr ( λ a ∂ µ U † U )<br />

For an internal symmetry the Noether current is then obtained by<br />

J µ,a<br />

L (x) = ∂<br />

∂(∂ µ θ a L (x))L( ̂φ i ,∂ µ ̂φi )<br />

which yields immedi<strong>at</strong>ely the above formula.<br />

165


12.4.2 Massive case<br />

We will now consider the case with massive Goldstone bosons due to some<br />

explicit symmetry breaking by quark masses. In the reality we have to add<br />

terms to the above effective Lagrangean, which correspond to the mass terms<br />

of the <strong>QCD</strong>. We know those;<br />

L <strong>QCD</strong><br />

sb<br />

= −q¯<br />

R Mq L − q¯<br />

L M † q R<br />

where the mass m<strong>at</strong>rix is given by M = diag(m u ,m d ,m s ). Apparently in <strong>QCD</strong><br />

this term is invariant under SU(2)xSU(2). We know from the non-linear Sigma<br />

model which term to add to the effective Lagrangean in order to obtain a proper<br />

mass term:<br />

L sb<br />

eff = f2 π<br />

4 B 0Tr ( MU † + UM †)<br />

Actually the constant B 0 is rel<strong>at</strong>ed to the chiral quark condens<strong>at</strong>e (without<br />

proof):<br />

3f 2 πB 0 = − 〈0 |¯qq|0〉 = − 〈 0 ∣ ∣ūu − ¯dd − ¯ss<br />

∣ ∣ 0<br />

〉<br />

One can easily see th<strong>at</strong> this mass term is correct and does wh<strong>at</strong> it should. We<br />

expand the exponential function in U to second order in φ with φ = λ a π a and<br />

obtain:<br />

L sb<br />

eff = fπB 2 0 (m u + m d ) − B 0<br />

2 Tr ( φ 2 M )<br />

We use the fact th<strong>at</strong> M is diagonal and we take the explicit form of φ:<br />

⎛<br />

√<br />

2K<br />

+<br />

φ(x) =<br />

⎜<br />

⎝<br />

√<br />

π 0 + √ 1<br />

⎞<br />

3<br />

η 8 2π<br />

+<br />

√ √<br />

2π<br />

−<br />

−π 0 + √ 1 3<br />

η 8 2K<br />

0 ⎟<br />

√ √<br />

⎠<br />

2K<br />

− 2K<br />

0<br />

− √ 2<br />

3<br />

η 8<br />

giving<br />

Tr(φ 2 M) = m u<br />

[<br />

(π 0 + 1 3 η)2 + 2π + π − + 2K + K − ]<br />

[<br />

+m d 2π − π + + (−π 0 + √ 1 ]<br />

η) 2 + 2K 0 ¯K0 3<br />

+m s<br />

[<br />

2K − K + + 2 ¯K 0 K 0 + 4 3 η2 ]<br />

= 2(m u + m d )π + π − + 2(m u + m s )K + K − + 2(m d + m s )K 0 ¯K0<br />

+2(m u + m d )π 0 π 0 + 2 √<br />

3<br />

(m u − m d )π 0 η + 1 3 (m u + m d + m s )η 2<br />

166


Apparently we can identify the mass terms and compare them with the experimental<br />

d<strong>at</strong>a.<br />

Assume first isospin symmetry, i.e. m u = m d . In the mesonic language<br />

wesay “We ignore the pion-eta mixing”. In this approxim<strong>at</strong>ion we note down<br />

the quadr<strong>at</strong>ic terms of the above expression and we obtain up to corrections<br />

quadr<strong>at</strong>ic in ¯m.:<br />

m 2 π = 2B 0 ¯m<br />

m 2 K = B 0 ( ¯m + m s )<br />

m 2 η = 2 3 B 0 ( ¯m + 2m s )<br />

Apparently these are exactly the Gell-Mann-Oakes-Renner-Rel<strong>at</strong>ions. The masses<br />

also fulfill the Gell-Mann-Okubo-Rel<strong>at</strong>ion<br />

4m 2 K == 3m 2 η + m 2 π<br />

Apparently the above expression for the symmetry breaking term in the bosonic<br />

language works.<br />

Without further inform<strong>at</strong>ion on B 0 we cannot extract absolute values for<br />

the quark masses. The reason is simple, because we meet always the product<br />

B 0 ¯m and B 0 m s and never the B 0 or the mass as such. However we know the<br />

empirical values for the r<strong>at</strong>ios<br />

m 2 K<br />

m 2 π<br />

= ¯m + m s<br />

2 ¯m → m s<br />

m π<br />

≃ 25.9<br />

m 2 η<br />

m 2 π<br />

= ¯m + 2m s<br />

3 ¯m → m s<br />

m π<br />

≃ 24.3<br />

Several comments are in order: 1) In the above expansion<br />

L sb<br />

eff = f 2 πB 0 (m u + m d ) − B 0<br />

2 Tr ( φ 2 M )<br />

we have the first term without boson field φ It describes the vacuum energy<br />

which is caused by the spontaneous symmetry breaking<br />

E vac = f 2 πB 0 (m u + m d )<br />

2) In the expansion of the effective symmetry breaking Lagrangean<br />

L sb<br />

eff = f2 π<br />

4 B 0Tr ( MU † + UM †)<br />

in terms of the goldstone boson fields there appear terms of higher than second<br />

order in the boson field. They are of the sort (⃗π⃗π) 2 and hence contribute to the<br />

boson-boson interaction. Thus the mass terms change the interaction of the<br />

bosons amongst each other.<br />

167


12.4.3 Higher order terms<br />

We have considered by now the most general effective Lagrangean exhibiting<br />

spontaneous broken chiral symmetry but restricted to the <strong>low</strong>est number of<br />

deriv<strong>at</strong>ives. For a complete theory of the interaction between goldstone bosons<br />

one should envisage the most general effective Lagrangean without such a restriction.<br />

So one should consider higher numbers of deriv<strong>at</strong>ives and hence one<br />

should add to the simple 2.deriv<strong>at</strong>ive effective Lagrangean terms of the form<br />

Tr ( ∂ µ U∂ ν U †) Tr ( ∂ µ U∂ ν U †)<br />

Tr ( ∂ µ U∂ µ U †) Tr ( ∂ ν U∂ κ ∂ κ ∂ ν U †)<br />

and other terms. There are several similar terms with 4 deriv<strong>at</strong>ives and several<br />

with 6 deriv<strong>at</strong>ives etc etc. In the end there is a growing number of unknnown<br />

coefficients, which must be determined by experiment. For a theoretical consider<strong>at</strong>ion,<br />

however, it is more appropri<strong>at</strong>e and even sufficient to write down the<br />

Lagrangean purely organized by the dimensionality of the oper<strong>at</strong>ors:<br />

There appear unknown coefficients:<br />

L eff = L 2 + L 4 + L 6 + ....<br />

L 2 → g 2 = f π<br />

L 4 → g (1)<br />

4 ,g(2) 4 ,g(3) 4 ,.....<br />

L 6 → g (1)<br />

6 ,g(2) 6 ,g(3) 6 ,.....<br />

which have, like e.g. the masses of the mesons, to be determined by experiment.<br />

It would be useless, if the above series would never end, because then one would<br />

need innumerable many coefficients g (k)<br />

i which must result from innumerable<br />

numbers of experiments. Th<strong>at</strong> means the above formalism makes only applicable<br />

if the above series concverges more or less rapidly such th<strong>at</strong> we have only few<br />

coefficients to determine and few terms to consider in our calcul<strong>at</strong>ion (which<br />

we still do not yet know how to do!). Tlhis is in fact true as the fol<strong>low</strong>ing<br />

consider<strong>at</strong>ion will lshow us:<br />

Consider e.g. pion-pion sc<strong>at</strong>tering and consider the mandelstam variables:.<br />

168


u<br />

B<br />

s<br />

a<br />

t<br />

A<br />

Mandelstam−Variables<br />

In the follwing all Mandelstam variables of this process are called q and<br />

are all assumed to be small compared to a typical hadronic scale Λ hadr , i.e.<br />

for reactions involving mesons the mass or the Rho-Meson (700 MeV) or for<br />

reactions involving baryons the mass of the proton (900 MeV). The mass of the<br />

pion or kaon is NOT taken, because these bosons are Goldstone bosons and have<br />

mass zero. Since U is dimensionless and f π has the dimension Λ 2 the effective<br />

Lagrangean L 2 has the dimension Λ 4 :<br />

[L 2 ] = Λ 4<br />

This is consistent with<br />

∫<br />

S =<br />

d 4 xL 2 = 1<br />

If we consider the dimension of the higher terms then each term of the Lagrangean<br />

has obviously the dimension Λ 4 but if we write<br />

L 2n = C 2n Tr ( ∂ µ U......∂ ν U †)<br />

with 2n deriv<strong>at</strong>ive terms, then we can write<br />

[C 2n ] =∼ Λ 4−2n<br />

Suppose we consider now an arbitrary vertex with 2n deriv<strong>at</strong>ives, then the<br />

contribution of this vertex to an S-m<strong>at</strong>rix element is proportional to<br />

q 2n<br />

Λ 2n−4 ∼ ( q Λ )2n 1 Λ 4<br />

Thus, if the reaction is of <strong>low</strong> energy, i.e. if each Mandelstam variable is smaller<br />

than Λ then the vertex contributes less, if it has more deriv<strong>at</strong>ives. This is<br />

the basic fe<strong>at</strong>ure: We can chose the momentum of the reaction partners in<br />

the experiment small enough such th<strong>at</strong> higher order vertices do not contribute<br />

much. In this way we can force the system to converge.<br />

169


The argument above is not quite correct: If the Mandelstam variable of the<br />

reaction are small, it does not mean th<strong>at</strong> the momenta of loops are small too<br />

and hence it can be th<strong>at</strong> we have few external lines with small q and several<br />

internal lines with large q and hence our argument is no longer true.<br />

Fortun<strong>at</strong>ely we need not worry, because there is the famous power counting<br />

theorem by Weinberg:<br />

Consider an arbitrary diagram based on the effective Lagrangean. if one<br />

rescales the momenta of the externl mesons, i.e. the mesons, which are physical<br />

and take part in the reaction, then we have for the invariant Feynman amplitude<br />

the fol<strong>low</strong>ing fe<strong>at</strong>ure:<br />

M(tp,t 2 M 2 ) = t D M(p,M 2 )<br />

D = 2 + ∑ 2(n − 1)N 2n + 2N L<br />

Here M is the mass of the Goldstone boson, the N 2n is the number of vertices<br />

in the diagram, which come from L 2n herrühren. and N L is the number of loops<br />

in the diagram. Hence, if in a certain reaction we reduce the momenta, then the<br />

diagrams with vertices of a higher number of deriv<strong>at</strong>ives contribute less and less<br />

to the Feynman amplitude and the diagrams with more loops contribute also<br />

less and less to the amplitude. In the ende the <strong>low</strong>er the incoming momenta<br />

the more domin<strong>at</strong>ing are the TREE DIAGRAMS coming from L 2 ! Th<strong>at</strong> is the<br />

secret and th<strong>at</strong> is why perturb<strong>at</strong>ion theory works even in the non-perturb<strong>at</strong>ive<br />

region. One does perturb<strong>at</strong>ion with effective and physical fields, pions, and not<br />

with quarks and gluons. uvertices withof higher of the effeem, becausere is only<br />

one pointcto be able to apply the morezs nzHence, If we have hence a higher<br />

ordnn, whiched. The incoming pions have the eWe are saved<br />

To illustr<strong>at</strong>e the above arguments we consider a simple exemple: Take<br />

L 2 = gφ 2 ∂ µ φ∂ µ φ<br />

Taking the usual rules for Feynman diagrams and considering a tree- and a loopdiagram<br />

we can write down the fol<strong>low</strong>ing formulae for the Feynman amplitude:<br />

The corresponding Feynman amplitude is:<br />

M(p 1 ,p 2 ,p 3 ,p 4 ) = 4ig [(p 1 + p 2 )(p 3 + p 4 ) − p 1 p 2 − p 3 p 4 ]<br />

170


which obviously scales as<br />

M(tp 1 ,tp 2 ,tp 3 ,tp 4 ) = t 2 M(p 1 ,p 2 ,p 3 ,p 4 )<br />

Comparison with Weinbergs formula yields:<br />

M(tp,t 2 M 2 ) = t D M(p,M 2 )<br />

D = 2 + ∑ 2(n − 1)N 2n + 2N L<br />

Apparently we have here: D = 2 corresponding to N L = 0 and N 2 = 1 since we<br />

have considered ONE vertex, which comes from L 2 . The scaling fe<strong>at</strong>ure can be<br />

contrasted with the one of a digaram, which contains one loop:<br />

L<br />

2<br />

D=4<br />

The Feynman rules give here:<br />

M(p 1 ,p 2 ,p 3 ,p 4 ) =<br />

= 16g 2 ∫<br />

d 4 k<br />

(2π) 4 [(p 1<br />

1 + p 2 )(p 3 + p 4 ) − (p + p 2 − k)k − p 3 p 4 ]<br />

k 2 − M 2 + iɛ<br />

1 [<br />

(p1<br />

(p 1 + p 2 − k) 2 − M 2 + p 2 ) 2 − p 1 p 2 − (p 1 + p 2 − k)k ]<br />

+ iɛ<br />

If we do now the scaling like p i → tp i and M 2 → t 2 M 2 and the variablesubstitution<br />

k = tl we can write down<br />

M(tp 1 ,tp 2 ,tp 3 ,tp 4 ) =<br />

= 16g 2 ∫ t 4 d 4 l<br />

(2π) 4 [<br />

t 2 (p 1 + p 2 )(p 3 + p 4 ) − t 2 (p + p 2 − l)l − t 2 p 3 p 4<br />

] 1<br />

t 2 (l 2 − M 2 + iɛ)<br />

1 [<br />

t 2<br />

t 2 ((p 1 + p 2 − l) 2 − M 2 (p 1 + p 2 ) 2 − t 2 p 1 p 2 − t 2 (p 1 + p 2 − l)l ]<br />

+ iɛ)<br />

We obtain a scaling with D = 4 which corresponds to Weinbergs formula with<br />

N L = 1 and N 2 = 2. Weinbergs formula yields the same value:<br />

D = 2 + ∑ 2(n − 1)N 2n + 2N L<br />

171


If we replace in this diagram in the right vertex the L 2 by and L 4 then we<br />

have the fol<strong>low</strong>ing picture:<br />

L<br />

2<br />

L<br />

4<br />

D=6<br />

And if both vertices haven an L 4 we get<br />

L<br />

4<br />

D=8<br />

Thus with Weinbergs rule we can immedi<strong>at</strong>ely estim<strong>at</strong>e the relevance of the<br />

graph for the calcul<strong>at</strong>ion of the process.<br />

Here we have done tacitley an important step. We have considered the mass<br />

of the Goldstone Boson in the scaling procedure lieke an external momentum.<br />

This makes indeed sense, because the momenta of the external particles obbey<br />

p 2 = M 2 . This step is essenttial, because without this we would not have a<br />

clear scaling and hence no clear decision, which diagrams to consider and which<br />

to ignore. In this step it is also clear, th<strong>at</strong> we cannot have mesons in the game,<br />

which are not goldstone bosons, like e.g. σ-mesons or ρ-mesons. They would<br />

occur only in inner loops and would destroy the power counting.<br />

172


Instead of saying “the diagram scales with D = 4 or t 4 one say also “the<br />

diagram is of order E 4 . And one sees immedi<strong>at</strong>ely: The diagrams can be ordered<br />

according to energy, we have, in fact an energy expansion. And vertices, which<br />

origin<strong>at</strong> from L 2 , which contribute e.g. as a tree diagram to O(E 2 ) contribute<br />

to O(E 2 ) via a loop and to O(E 6 ) via two loops etc.<br />

Hence: The <strong>low</strong>er the energy of the process considered, the more domin<strong>at</strong>e<br />

tree diagrams with <strong>low</strong> dimensional L 2n i.e. <strong>low</strong> n.<br />

We can summarize the points:<br />

1) L-Loop diagrams are suppressed by powers E 2L .<br />

2) From the series<br />

L eff = L 2 + L 4 + L 6 + ....<br />

we have the fol<strong>low</strong>ing contributions<br />

O(E 2 ): tree diagrams with L 2 -insertions<br />

O(E 4 ): tree diagrams with one L 4 -insertion plus 1-loop graphs with only L 2<br />

insertions.<br />

O(E 6 ): tree diagrams one with L 6 -insertion plus 1-loop graphs with one<br />

L 4 -insertion plus 2-loop graphs with only L 2 -insertions.<br />

Altogether we have: ENERGY EXPANSION = EXPANSION IN LOOPS<br />

Take an example: ππ-sc<strong>at</strong>tering: Again we verify Weinbergs counting rule.<br />

For exampe the <strong>low</strong>er left diagram has 3L 2 -vertices (none of them contributes<br />

to the D of the Weinberg rule) and 2 Loops, such th<strong>at</strong> D = 2 + 2 ∗ 2:<br />

L<br />

2<br />

L<br />

4<br />

L<br />

6<br />

D=2<br />

D=4<br />

D=6<br />

The more accur<strong>at</strong>e one wants to calcul<strong>at</strong>e the higher one goes in the energy<br />

expansion, the more graphs one has to consider. However, one is clear: it is a<br />

system<strong>at</strong>ic expansion which can be improved in a system<strong>at</strong>ic way. The higher<br />

one goes in energy expansion the more unknown coefficients are there, see above:<br />

:<br />

L 2 → g 2 = f π<br />

L 4 → g (1)<br />

4 ,g(2) 4 ,g(3) 4 ,.....<br />

L 6 → g (1)<br />

6 ,g(2) 6 ,g(3) 6 ,.....<br />

173


which have to be determined by experiment. I.e. one calcul<strong>at</strong>es the cross section<br />

of a certin experiment in a certain degree of energy expansion and adiusts the<br />

coefficients g (i)<br />

k<br />

to the experimental d<strong>at</strong>a. One should obtain the result, th<strong>at</strong> this<br />

expansion is well converging. However, it is not always the case, for example in<br />

the vicinity of resonances.<br />

12.5 Applic<strong>at</strong>ion: pion-pion sc<strong>at</strong>tering<br />

12.5.1 Calcul<strong>at</strong>ion:<br />

The sc<strong>at</strong>tering process<br />

π + π → π + π<br />

is the purest process to test the ideas of ppure chiral dynamics. It is also the<br />

easiest process on the theoretial side, however, it is not easy to measure. We<br />

consider L 2 with mass term<br />

L 2 = F 2 0<br />

4 Tr(∂ µU∂ µ U † ) + F 2 0<br />

2 B 0Tr(MU † + UM † )<br />

For the calcul<strong>at</strong>ion we have to select th<strong>at</strong> part of L 2 , which will appear in the<br />

graphs for pion -pion sc<strong>at</strong>tering. These are the ones which al<strong>low</strong> 2 incoming and<br />

2 outgoing pions. Apparently L 2 has only even powers of φ so we can write<br />

L 2 = L 2φ<br />

2 + L4φ 2 + L6φ 2 + ...<br />

Vertices with 3 bosons do not exist so for the pi-pi-sc<strong>at</strong>tering with D = 2 we<br />

must only consider a contact interaction with 4 goldstone bosons: To indentify<br />

this we expand<br />

U = 1 + i φ F 0<br />

− 1 2<br />

φ 2<br />

F0<br />

2<br />

− i 6<br />

φ 3<br />

F0<br />

3<br />

+ 1<br />

24<br />

φ 4<br />

F0<br />

4<br />

+ ....<br />

U † = 1 − i φ F 0<br />

− 1 2<br />

φ 2<br />

F0<br />

2<br />

+ i 6<br />

φ 3<br />

F0<br />

3<br />

∂ µ U = i<br />

F 0<br />

∂ µ φ + ....<br />

+ 1 24<br />

∂ µ U † = − i<br />

F 0<br />

∂ µ φ + .....<br />

This yields in the expansion to fourth order<br />

φ 4<br />

F0<br />

4<br />

+ ....<br />

L 2φ<br />

2 = 1 4 Tr(∂ µφ∂ µ φ − 2BMφ 2 )<br />

and then after some calcul<strong>at</strong>ion we can identify<br />

L 4φ<br />

2 = 1<br />

24F0<br />

2 Tr ([φ,∂ µ φ]φ∂ µ φ) + 1<br />

24F0<br />

2 B 0 Tr(Mφ 4 )<br />

174


For the pi-pi-sc<strong>at</strong>tering we have to reduce this on expressions involving solely<br />

pion fields π a and for this we have to use φ(x) = τ a π a (x). To tre<strong>at</strong> this we need<br />

some simple formulae like<br />

φ = ⃗τ⃗π<br />

τ a τ b = δ ab + iɛ abc τ c<br />

This yields then<br />

φ 2 = π a π b τ a π b = ⃗π 2 + iɛ abc τ c π a π b<br />

∂ µ φ 2 = 2π a ∂ µ π a + iɛ abc τ c (∂ µ π a π b + π a ∂ µ π b )<br />

φ 3 = τ a π a π b π b + iɛ abc π a π b π c<br />

∂ µ φ 3 = iɛ abc (∂ µ π a π b π c + π a ∂ µ π b π c + π a π b ∂ µ π c ) + τ a (∂ µ π a π b π b + 2π a π b ∂ µ π b )<br />

We insert these expressions in the above formula and use the identities<br />

Tr ([(⃗τ⃗π),(⃗τ∂ µ ⃗π)] (⃗τ⃗π)(⃗τ∂ µ ⃗π)) = 4(⃗π∂ µ ⃗π)(⃗π∂ µ ⃗π) − (⃗π⃗π)(∂ µ ⃗π∂ µ ⃗π)<br />

Tr ( Mφ 4) = mTr ( (⃗τ⃗π) 4) = 2m(⃗π⃗π) 2<br />

Hence with m 2 π = 2B 0 m with m u = m d = m we get th<strong>at</strong> part of L 2 which<br />

enters the pion-pion sc<strong>at</strong>tering:<br />

L 2π<br />

2 = 1 [<br />

∂µ π a ∂ µ π a − m 2<br />

2<br />

ππ a π a]<br />

L 4π<br />

2 = 1<br />

6F 2 0<br />

((⃗π∂ µ ⃗π)(⃗π∂ µ ⃗π) − (⃗π⃗π)(∂ µ ⃗π∂ µ ⃗π)) + m2 π<br />

24F0<br />

2 (⃗π⃗π) 2<br />

One can rewrite this to cartesian isospin-indices and then we obtain for the<br />

sc<strong>at</strong>tering<br />

π a (p a ) + π b (p b ) → π c (p c ) + π d (p d )<br />

the Feynman-amplitude M<br />

with<br />

−i6F 2 0 M(p a ,p b ,p c ,p d ) = 2A + B<br />

A = δ ab δ cd (−ip a − ip b ) (ip c + ip d ) + δ ac δ bd (−ip a + ip c )(−ip b + ip d )<br />

+δ ad δ bc (−ip a + ip d ) (−ip b + ip c ) − 4(δ ab δ cd ((−ip a )(−ip b ) + (ip c )(ip d ))<br />

175


+δ ac δ bd ((−ip a )(ip c ) + (−ip b )(ip d )) + δ ad δ bc ((−ip a )(ip d ) + (ip b )(ip c )))<br />

B = m2 π<br />

4 8( δ ab δ cd + δ ac δ bd + δ ad δ bc)<br />

If one takes into account th<strong>at</strong> the incoming and outgoing particles are all on the<br />

mass shell one can change to<br />

−i3F 2 0 M(p a ,p b ,p c ,p d ) =<br />

δ ab δ cd ((p a + p b ) 2 + 2p a p c − 2p c p d + m 2 π)<br />

+δ ac δ bd ((p a − p c ) 2 − 2p a p c − 2p b p d + m 2 π)<br />

+δ ad δ bc ((p a − p d ) 2 − 2p a p d − 2p b p c + m 2 π)<br />

and then one obtains finally after some direct and explicit calcul<strong>at</strong>ion:<br />

M(p a ,p b ,p c ,p d ) = i<br />

F 2 0<br />

[<br />

δ ab δ cd (s − m 2 π) + δ ac δ bd (t − m 2 π) + δ ad δ bc (u − m 2 π) ]<br />

where we have the Mandelstam variables (where the arrow indic<strong>at</strong>es, how they<br />

look in the center of mass system:<br />

s = (p a + p b ) 2 = (p c + p d ) 2 → 4(q 2 + m 2 π) = W<br />

t = (p a − p c ) 2 = (p d − p b ) 2 → −2q 2 (1 − cos(Θ sc<strong>at</strong>t ))<br />

u = (p a − p d ) 2 → −2q 2 (1 + cos(Θ sc<strong>at</strong>t ))<br />

s + t + u = ∑ p 2 i = 4m 2 π<br />

Here Θ sc<strong>at</strong>t is the sc<strong>at</strong>tering angle in the cm-system and q is the momentum of<br />

the incoing pion in the cm-system. We have used rel<strong>at</strong>ions like 2p a p b = 2p c p d =<br />

s − 2m 2 π. We have of course assumed th<strong>at</strong> the incoming and outgoing pions are<br />

on the mass shell. This result has also been obtained purely by means of current<br />

algebra by Weinberg (1966), where the formalism was much more complic<strong>at</strong>ed.<br />

176


12.5.2 Comparison with experiment<br />

One writes the sc<strong>at</strong>tering amplitude (or Feynman amplitude) in the fol<strong>low</strong>ing<br />

way:<br />

T abcd = δ ab δ cd A(s,t,u) + δ ac δ bd A(t,s,u) + δ ad δ bc A(u,t,s)<br />

It is remarkable th<strong>at</strong> the pion-pion sc<strong>at</strong>tering amplitude depends on one<br />

function A(s,t,u) only:<br />

A(s,t,u) = s2 − m 2 π<br />

f 2 π<br />

Actually pions have isospin 1 and hence it is customary to characterize the<br />

input- and output channels by the total isospin. Since the isospin is a conserved<br />

quantum number (assuming equal masses of up and down quarks) the isosopin<br />

of the entrance channel and the one of the exit echannel are identical. We can<br />

have isospins of T = 0,1,2. Starting from the cartesian represent<strong>at</strong>ion we can<br />

change it into a represent<strong>at</strong>ion with good isospin yielding then for the sc<strong>at</strong>tering<br />

amplitudes<br />

T (0) = 3A(s,t,u) + A(t,s,u) + A(u,t,s)<br />

T (1) = A(t,s,u) − A(u,t,s)<br />

T (2) = A(t,s,u) + A(u,t,s)<br />

In practice one decomposes these sc<strong>at</strong>tering amplitudes into partial waves with<br />

definite isospin and dfinite angular momentum l:<br />

T (I)<br />

l<br />

(s) = 1<br />

64π<br />

∫ +1<br />

−1<br />

d(cosΘ)P l (cosΘ)T (I) (s,t,u)<br />

(Attention: t,u depend on the sc<strong>at</strong>tering angle Θ). We have also<br />

T (I) (s,t,u) = 32π<br />

∞∑<br />

P l (cosΘ)T (I) (s)<br />

l=0<br />

The sc<strong>at</strong>tering amplitudes T (I) and the S-m<strong>at</strong>rix are both charakterized by the<br />

phase shifts:<br />

S = exp{2iδ (I)<br />

T (I)<br />

l<br />

(s) =<br />

√<br />

s<br />

s − 4m 2 π<br />

√ s 1<br />

=<br />

s − 4m 2 π 2i<br />

l<br />

)<br />

exp[iδ (I)<br />

l<br />

]sin δ (I)<br />

l<br />

[<br />

exp(2iδ (I)<br />

l<br />

) − 1<br />

For a comparison with experiment one investig<strong>at</strong>es the behaviour of the sc<strong>at</strong>tering<br />

amplitude T (I)<br />

l<br />

in the vicinity of the threshold by doing an expansion in<br />

terms of the pion momentum q 2 . This yields<br />

[ ]<br />

ReT (I)<br />

l<br />

= q 2l a (I)<br />

l<br />

+ q 2 b (I)<br />

l<br />

+ ...<br />

]<br />

177


where a (I)<br />

l<br />

is called “sc<strong>at</strong>tering length” and b (I)<br />

l<br />

is called range parameter, and<br />

both are called threshold parameters. The predictions, as calcul<strong>at</strong>ed above, are<br />

for the s- and p-waves (i.e. l = 0,1):<br />

T (0)<br />

0 (s) = 1<br />

32πfπ<br />

2 (2s − m 2 π).................s − wave<br />

T (1)<br />

1 (s) = 1<br />

96πfπ<br />

2 (s − 4m 2 π)...............p − wave<br />

T (2)<br />

0 (s) = 1<br />

32πfπ<br />

2 (2m 2 π − s)................s − wave<br />

There are no d, f,.... waves due to the simple form of A(s,t,u) ∼ αP 0 (cosΘ) +<br />

βP 1 (cosΘ) Since the s = 4(q 2 + m 2 π) we can evalu<strong>at</strong>e the threshold parameters<br />

yielding:<br />

a (0)<br />

0 = 7m2 π<br />

32πfπ<br />

2<br />

b (0)<br />

0 = 1<br />

4πfπ<br />

2<br />

a (1)<br />

1 = 1<br />

24πfπ<br />

2<br />

a (2)<br />

0 = − m2 π<br />

16πfπ<br />

2<br />

b (2)<br />

0 = − 1<br />

8πfπ<br />

2<br />

Now we can perform the comparison with the d<strong>at</strong>a. In the table we include for<br />

infrom<strong>at</strong>ion the results of the calcul<strong>at</strong>ion to <strong>low</strong>est order and to first two orders<br />

(book: Donoghue, Go<strong>low</strong>ich, Holstein): For the sc<strong>at</strong>tering length and the range<br />

parameter we get<br />

ππ-sc<strong>at</strong>t. a (0)<br />

0 b (0)<br />

0 a (1)<br />

1 b (1)<br />

1 a (2)<br />

0 b (2)<br />

0 a (0)<br />

2 a (2)<br />

2<br />

L 2 0.16 0.18 0.030 0 −0.045 −0.089 0<br />

L 4 0.20 0.26 0.036 0.043 −0.041 −0.070 20x10 −4 3.5x10 −4<br />

Experiment 0.26 0.25 0.038 − −0.028 −0.082 (17 ± 3)10 −4 1.3 ± 3)10 −4<br />

exp. error ±0.05 ±0.03 ±0.002 − ±0.012 ±0.008 ±3x10 −4 ±3x10 −4<br />

Here the values origin<strong>at</strong>e e.g. from a (0)<br />

0 = 7<br />

32π (139 93 )2 i.e. they are given in<br />

appropri<strong>at</strong>e units of m π . The agreement between theory and experiment is<br />

amazing! The calcul<strong>at</strong>ion assumes just chiral symmetry and is fully analytic. It<br />

does not include any higher order terms in the number of deriv<strong>at</strong>ives.<br />

Actually the deriv<strong>at</strong>ion of the a (0)<br />

0 and b =(0) is easy: We have<br />

Tl (I) = 1<br />

64π<br />

∫ +1<br />

−1<br />

dxP l (x)T (I) (s,t,u)<br />

178


and<br />

and we have<br />

>From this we get<br />

which gives<br />

T (0)<br />

0 = 1<br />

64π<br />

ReT (I)<br />

l<br />

= q 2l [ a (I)<br />

l<br />

]<br />

+ q 2 b (I)<br />

l<br />

T (0) = 3A(s,t,u) + A(t,s,u) + A(u,t,s)<br />

∫ +1<br />

−1<br />

dx 1<br />

fπ<br />

2 P 0 (x) [ 3s − 3m 2 π + t − m 2 π + u − m 2 ]<br />

π<br />

T (0)<br />

0 = 1 [ ]<br />

3s + t + u − 5m<br />

2<br />

64πfπ<br />

2 π 2<br />

,where the 2 comes from the integral over P 0 (x) = 1 . Using s + t + u = 4m 2 π<br />

yields<br />

T (0)<br />

0 = 1 [ ]<br />

2s − m<br />

2<br />

32πfπ<br />

2 π<br />

. Further we have s = 4(q 2 + m 2 π) and hence<br />

T (0)<br />

0 = 1 [<br />

7m<br />

2<br />

32πfπ<br />

2 π + 8q 2]<br />

which yields a (0)<br />

0 = 7m2 π<br />

32πf<br />

and b (0)<br />

π<br />

2 0 = 1<br />

4πf<br />

. One can easily show, th<strong>at</strong> in <strong>low</strong>est<br />

π<br />

2<br />

order we have b (1)<br />

1 = 0 .<br />

Above the threshold, which is of course <strong>at</strong> 2m 2 π we only have the phase shift<br />

of the s-wave and this starts <strong>at</strong> zero exacxtly <strong>at</strong> the threshold. If we there<br />

approxim<strong>at</strong>e exp(iδ (I)<br />

l<br />

sin(δ (I)<br />

l<br />

) ≃ δ (I)<br />

l<br />

( √ s) then we obtain<br />

√<br />

δ (I)<br />

l<br />

≃<br />

1 − 4m2 π<br />

s T (I)<br />

l<br />

(s)<br />

For the s-wave we see the comparison with the experimental d<strong>at</strong>a (Rosselet et<br />

al.)in the fol<strong>low</strong>ing graph:<br />

179


12.6 Chiral perturb<strong>at</strong>ion theory and <strong>QCD</strong>: Functional integrals<br />

12.6.1 Internal symmetries and Ward-Identities<br />

We are interested generally in theories described by lagrangeans L = L 0 + L 1 ,<br />

where L 0 is invari<strong>at</strong> under global internal symmetry G and L 1 explicitely breaks<br />

the symmetry. Furtherore we assume the bare fields φ i in L form a basis for<br />

some definite represent<strong>at</strong>ion R of the symmetry group of L 0 . Th<strong>at</strong> means, if we<br />

perform an infinitesimal transform<strong>at</strong>ion of the fields we have<br />

δφ i = −iθ a T a ijφ j = θ a δ a φ i<br />

with δ a being an oper<strong>at</strong>or. The bare currents<br />

jµ(x) a ∂L<br />

= i<br />

∂(∂ µ φ i ) T ijφ a j<br />

which are conserved currents of the G symmetric classical theory described by<br />

L 0 remain unchanged when L 0 → L if L 1 does not contain deriv<strong>at</strong>ives of the<br />

fields present in the j a µ(x), and we assume this to be the case.<br />

180


Under an infinitesimal global symmetry transform<strong>at</strong>ion<br />

φ ′ i(x) = φ i (x) + θ a δ a φ i (x)<br />

the lagrangean L transformes as fol<strong>low</strong>s<br />

L ′ (x) = L(x) + θ a δ a L(x) = L(x) + θ a δ a L 1 (x)<br />

since the lagrangean L is considered invariant. Using the classical equ<strong>at</strong>ions of<br />

motion<br />

∂ µ j a µ(x) = δ a L 1 (x)<br />

In the fol<strong>low</strong>ing we shall also need the vari<strong>at</strong>ion of the same lagrangean L under<br />

the local transform<strong>at</strong>ion<br />

φ ′ i(x) = φ i (x) + θ a (x)δ a φ i (x)<br />

Remembering th<strong>at</strong> L is invariant under global transform<strong>at</strong>ions and th<strong>at</strong> L 1 does<br />

not contain deriv<strong>at</strong>ives of the fields we get<br />

an hence<br />

δL(x) = θ a (x)δ a L(x) + j a µ(x)∂ µ θ a (x)<br />

j a µ(x) = ∂(δL)<br />

∂(∂ µ θ a ) (x)<br />

We want to derive Ward identities for bare regularized or for renormalized<br />

Greens functions fol<strong>low</strong>int from the symmetries of the lagrangean. We assume in<br />

this section th<strong>at</strong>h the UV regfulariz<strong>at</strong>ion of the theory does preserve its classical<br />

symmetries. Thus we ignore the problem of anomalies. They can be tre<strong>at</strong>ed in<br />

fact.<br />

We can now proceed to derive Ward identities using the functional integral<br />

formul<strong>at</strong>ion of quantum field theory. Consider the gener<strong>at</strong>ing functional W [J]<br />

with external sources J i (x) for the fields φ i (x):<br />

∫ [ ∫<br />

]<br />

Z [J] = N Dφ i exp i d 4 x(L(φ i ) + J i φ i )<br />

This integral is invariant under the change of integr<strong>at</strong>ion variables defined above<br />

(local transform<strong>at</strong>ion). If the integr<strong>at</strong>ion measure is invariant under the change<br />

of integr<strong>at</strong>ion variables (no anomaly)<br />

φ i (x) → φ ′ i(x) = φ i (x) + θ a (x)δ a φ i (x)<br />

we get with the above expression for δL(x) = θ a (x)δ a L(x) + j a µ(x)∂ µ θ a (x)<br />

∫<br />

0 =<br />

∫<br />

d 4 x<br />

( ∫<br />

Dφ i exp iS + i<br />

d 4 xJ i φ i<br />

)(θ a (x)δ a L<br />

+j a µ(x)∂ µ θ a (x) + J i (x)θ a (x)δ a φ i (x)<br />

181


One defines the Greens functions by<br />

{<br />

〈0|T A(x) ∏ } ∫<br />

φ i (y i ) |0〉 ∼<br />

Dφ i A(x) ∏ φ i (y i )exp(iS)<br />

where A(x) denotes one of the jµ(x),δL,δ a a φ i (x). This are here more general<br />

Greens functions, namely not only of the fields but also of composite oper<strong>at</strong>ors.<br />

One should remember th<strong>at</strong> the lhs of this equ<strong>at</strong>in is just a short hand not<strong>at</strong>ion<br />

of the rhs. One obtains now the general expression for the Ward identity:<br />

∂ µ 〈0|T<br />

−i ∑ i<br />

{<br />

j a µ(x) ∏ φ i (y i )<br />

}<br />

|0〉 = 〈0|T<br />

{<br />

δ a L ∏ φ i (y i )<br />

⎧<br />

⎫<br />

⎨<br />

δ(x − y) 〈0| T<br />

⎩ δa φ i (y i ) ∏ ⎬<br />

φ j (x j )<br />

⎭ |0〉<br />

j≠i<br />

}<br />

|0〉<br />

One obtains the Ward identity by the fol<strong>low</strong>ing steps: by 1) integr<strong>at</strong>ing the<br />

above eq. 0 = ∫ d 4 x... by parts to get rid of the ∂ µ θ a (x) and 2) differenti<strong>at</strong>ing<br />

it functionally with respect to the sources J i and 3) setting the J i (x) to zero.<br />

One sees th<strong>at</strong> this Ward identity is general concept, which can be formul<strong>at</strong>ed<br />

for any field theory, in particular for <strong>QCD</strong>. This ward identity is valid<br />

irrespective of whether the symmetry is spontaneously broken or not. If the<br />

symmetry G is an exact global symmetry of the lagrangean thei first term of<br />

the rhs vanishes. Up tonow the Ward identity is a concept which is based on<br />

LOCAL symmetry. They represent the symmetry properties of the theory on<br />

the level of the Green functions.<br />

The Ward Identities are intim<strong>at</strong>ely connected with the global symmetries of<br />

the Lagrangean. One can intim<strong>at</strong>ely connect them also to the local symmetries<br />

of the system. In fact one can express the existence of the Ward Identities in<br />

this way in a very compact forem. For this we conswider even Ward identities<br />

for Greens functions invelving several symmetry currents jµ(x) a and/or the symmetry<br />

breaking oper<strong>at</strong>or L 1 . We take L 1 to have well defined transform<strong>at</strong>ion<br />

properties under the symmetry group. The most compact way to derive such<br />

Ward identities is based on the general method to introduce background fields to<br />

get, by construction, an action which is exactly invariant under the considered<br />

symmetry transform<strong>at</strong>ions. In our case the background fields are the sources<br />

J i ,A a µ and K for the oper<strong>at</strong>ors φ i ,jµ a and L 1 , respectively. The new step consists<br />

now in introducing them in such a way th<strong>at</strong> the action S [ φ i ,J,A a µ,K ] is<br />

invariant under the LOCAL transform<strong>at</strong>ion θ a (x). This is achieved if<br />

∫<br />

S [φ,J,A,K] = S 0 [φ,A] + d 4 x ( J i )<br />

φ i + KL 1<br />

where S 0 (φ,A) = ∫ d 4 xL 0 (φ i (x),D µ φ i (x)) with the covariant deriv<strong>at</strong>ive D µ =<br />

∂ µ − iA a µT a . This action can be shown to be invariant under the local transform<strong>at</strong>ion<br />

if simultaneous transform<strong>at</strong>ions on A µ (x), J(x),K(x) are defined as<br />

δA a µ(x) = −∂ µ θ a (x) + f abc θ b (x)A c µ(x)<br />

182


δJ i (x)φ i (x) = −J i (x)δφ i (x)<br />

δKL 1 = −KδL 1<br />

The gener<strong>at</strong>ing functional is then invariant under the above transform<strong>at</strong>ions of<br />

the background fields, i.e.<br />

Z [A,J,K] = Z [A ′ ,J ′ ,K ′ ]<br />

Actually: One can show (Leutwyler) th<strong>at</strong> this equ<strong>at</strong>ion contains all the inform<strong>at</strong>ion<br />

which is contained in the Ward identities (which originally come from the<br />

global symmetries). The interesting point is, th<strong>at</strong> these Ward identities (which<br />

fol<strong>low</strong> from the global symmetry of the original theory) can be equivalently expressed<br />

by demanding LOCL symmetry Z [A,J,K] = Z [A ′ ,J ′ ,K ′ ] . One can<br />

see this by performing an infinitesimal transform<strong>at</strong>ion of Z [A,J,K] yielding<br />

∫<br />

0 = d 4 x<br />

[ δZ<br />

δJ i (x) δJi (x) +<br />

δZ<br />

δA a µ(x) δAa µ(x) +<br />

δZ ]<br />

δK(x) δK(x)<br />

The proof will not be given here.<br />

Important is: If one demands for the background fields invariance under local<br />

transform<strong>at</strong>ions, then the Ward identities are autom<strong>at</strong>ially s<strong>at</strong>isfied. Or in<br />

other words: The invariance of the gener<strong>at</strong>ing functional under “gauge transform<strong>at</strong>ions”<br />

of the external fields expresses the symmetry properties of the theory<br />

on the level of the Green functions, i.e. the Ward identities. This fe<strong>at</strong>ure represents<br />

the basic ingredient of the follwoing analysis, while the specific properties,<br />

which the theory may otherwise have, do not play any role.<br />

12.6.2 External fields and Greens functions<br />

In order to transl<strong>at</strong>e the above ideas to our concrete problem of embedding the<br />

effective Lagrangean into <strong>QCD</strong> we first introduce certain external hermitean<br />

fields into the <strong>QCD</strong>-Lagrangean. This is a method which goes back to Gasser<br />

and Leutwyler Ann.Phys. 1984.<br />

L <strong>QCD</strong> = L 0 <strong>QCD</strong> − ¯qγ µ 1 2 (1 + γ 5)l µ q − ¯qγ µ 1 2 (1 − γ 5)r µ q<br />

or<br />

−[ q¯<br />

L (s + ip)q R − q¯<br />

R (s − ip)q L ]<br />

L <strong>QCD</strong> = L 0 <strong>QCD</strong> − q¯<br />

L γ µ l µ q L − q¯<br />

R γ µ r µ q R<br />

or<br />

−[ q¯<br />

L (s + ip)q R − q¯<br />

R (s − ip)q L ]<br />

L <strong>QCD</strong> = L 0 <strong>QCD</strong> − ¯qγ µ v µ q − ¯qγ µ γ 5 a µ q − [ q¯<br />

L (s + ip)q R − q¯<br />

R (s − ip)q L ]<br />

183


with<br />

v µ = r µ + l µ<br />

a µ = r µ − l µ<br />

In the present SU(3) formalism all these fields are decomposed according to<br />

v µ (x) = λa<br />

2 va µ(x)<br />

a µ (x) = λa<br />

2 aa µ(x)<br />

s(x) = λa<br />

2 sa (x)<br />

p(x) = λa<br />

2 pa (x)<br />

where the sum goes over a = 0,1,...,8<br />

If one considers the gener<strong>at</strong>ing functional with the external fields one calcul<strong>at</strong>es<br />

basically how the system developes in the presence of the external field.<br />

Th<strong>at</strong> means one observs the response of the strong system to the external field<br />

and how the responding system developes in time. This is the way, how one can<br />

learn something about the system and th<strong>at</strong> is why the deriv<strong>at</strong>ives of the gener<strong>at</strong>ing<br />

function with respect to the external fields contain all the inform<strong>at</strong>ion on<br />

the system, which is then formul<strong>at</strong>ed in terms of Greens functions.<br />

The embedding of the previous formalism into <strong>QCD</strong> consists now in writing<br />

down the master integral for the gener<strong>at</strong>ing function of <strong>QCD</strong>:<br />

∫<br />

∫<br />

Z <strong>QCD</strong> [l µ ,r µ ,s,p] = DqD¯qDA a µ exp d 4 xL <strong>QCD</strong> (q, ¯q A a µ,l µ ,r µ ,s,p)<br />

and identifying this with<br />

∫<br />

Z eff [l µ ,r µ ,s,p] =<br />

∫<br />

DU exp<br />

d 4 xL eff (U,l µ ,r µ ,s,p)<br />

The effective chiral action L eff can be expanded in energy terms as we know<br />

from above:<br />

L eff = L 2 + L 4 + L 6 + ....<br />

One has to specify wh<strong>at</strong> “identifying” in the above formula. Suppose one<br />

calcul<strong>at</strong>es a correl<strong>at</strong>ion function of axial currents A a µ(x) and vector currents<br />

V a µ (x) like e.g:<br />

δ δ<br />

〈0| T {V µ (x 1 )A ν (x 2 )} |0〉 <strong>QCD</strong><br />

=<br />

δv µ (x 1 ) δa ν (x 2 ) Z <strong>QCD</strong> [l µ ,r µ ,s,p]<br />

184


δ δ<br />

〈0| T {V µ (x 1 )A ν (x 2 )} |0〉 eff<br />

=<br />

δv µ (x 1 ) δa ν (x 2 ) Z eff [l µ ,r µ ,s,p]<br />

Then we perform fourier transform<strong>at</strong>ions of the correl<strong>at</strong>ion functions:G(<br />

∫<br />

G <strong>QCD</strong> (p 1 ,p 2 ) = d 4 x 1 d 4 x 2 exp(−ip 1 x 1 − ip 2 x 2 ) 〈0|T {V µ (x 1 )A ν (x 2 )} |0〉 <strong>QCD</strong><br />

∫<br />

G eff (p 1 ,p 2 ) = d 4 x 1 d 4 x 2 exp(−ip 1 x 1 − ip 2 x 2 ) 〈0|T {V µ (x 1 )A ν (x 2 )} |0〉 eff<br />

Then we expand in Z eff [l µ ,...] the L eff = L 2 +L 4 +.... and denote the functional<br />

integral by means of Feynman diagrams with L 2 -insertions and L 4 -insertions<br />

etc. and with 1-Loops and 2-Loops etc etc. Thus altogether we have expanded<br />

G eff (p 1 ,p 2 ) in powers of the momenta and the mass of the goldstone bosons,<br />

all this in the way discussed above, a technically well defined expansion, which<br />

can and is really perfored (remember the pion-pion sc<strong>at</strong>tering, think th<strong>at</strong> the<br />

functional integral is only a short hand writing of the Feynman perturb<strong>at</strong>ion series).<br />

We then perform (only in our head!) an expansion of the G <strong>QCD</strong> (p 1 ,p 2 ) in<br />

poweres of the momenta, which al<strong>low</strong>s us to compare momentum by momenta.<br />

All this we do only for axial currents A a µ(x) and vector currents V a µ (x), not<br />

for any arbitrary current. Then the fol<strong>low</strong>ing is ment with “identifying”: The<br />

smaller the external momenta, the higher the expansion in L 2n , the higher the<br />

scaling properties t D the more the effective calcul<strong>at</strong>ion agrees with the <strong>QCD</strong>calcul<strong>at</strong>ion.<br />

And in the limit of decreasing external momenta we obtain <strong>at</strong> the<br />

threshold really an identity. Actually this sort or argument cannot only be applied<br />

to the above two-point green functiosn but to any Ward identity involving<br />

axial and vector currents and symmetry breaking mass terms. This very fe<strong>at</strong>ure<br />

is a consequence of demanding for the external background fields LOCAL symmetries,<br />

as we know from above. Thus the identity of Ward identities, achieved<br />

by demanding local symmetry is th<strong>at</strong>, wh<strong>at</strong> is called embedding of the effective<br />

theory into <strong>QCD</strong>.<br />

Besides embedding the effective theory fully into <strong>QCD</strong> the procedure of<br />

coupling external fields to the <strong>QCD</strong>-Lagrangean or the effective Lagrangean has<br />

distinct advantageous properties:<br />

1) Electromagnetic interactions are autom<strong>at</strong>ically included with<br />

and Q = diag( 2 3 , −1<br />

3 , −1<br />

3<br />

r µ = l µ = eqA µ (x)<br />

) because then we have<br />

−eA µ ( q¯<br />

L Qγ µ q L + q¯<br />

R Qγ µ q R ) = −eA µ¯qQγ µ q =<br />

−eA µ ( 2 u − 1 3ūγµ 3 ¯dγ µ q − 1 3¯sγµ s)<br />

2) Semileptonic weak interactions are autom<strong>at</strong>ically included with<br />

e<br />

l µ = −√ (W µ + T + + h.c.)<br />

2 sin ϑW<br />

185


with<br />

⎛<br />

T + = ⎝<br />

0 V ud V us<br />

0 0 0<br />

0 0 0<br />

here the V ud ,V us are Cabibo-Kobayashi-Maskawa-Elements, the Weinberg-Angle<br />

is rel<strong>at</strong>ed to the Fermi-weak coupling constant by<br />

√<br />

2e<br />

2<br />

G F =<br />

8MW 2 sinϑ W 2<br />

⎞<br />

⎠<br />

and we have<br />

W + µ = 1 √<br />

2<br />

(W 1 µ + iW 2 µ)<br />

If you insert this in the Lagrange density of the <strong>QCD</strong> we have the interaction<br />

term, well known:<br />

2 [<br />

−√ W<br />

+<br />

µ (V ud ūγ µ (1 − γ 5 )d + V us ūγ µ (1 − γ 5 )s + h.c.) ]<br />

2 sin ϑW<br />

The numbers are known from various experiments:<br />

V ud = 0.9744 ± 0.0010<br />

V us = 0.220 ± 0.004<br />

3) S-m<strong>at</strong>rix elements and general Greens functions of quark currents can be<br />

obtained directly from the partition function by functional differenti<strong>at</strong>ion.<br />

Thus the well known gauge symmetries U(1) and SU(2) L of the Standard<br />

Model are autom<strong>at</strong>ically transferred to the effective theory and we know there,<br />

how to couple external fields to the effective theory. Actually we have demanded<br />

local chiral symmetry of the full lagrangean including the external fields and this<br />

has given us the transform<strong>at</strong>ion properties of the external fields. One may ask<br />

if we have demanded too much, since altogether only the global symmetry is<br />

known to exist. In some sense we have done more than we needed since it<br />

would be sufficient to couple only the electromagnetic field (U(1)) and the weak<br />

field SU(2)-left as real gauge fields. This would provide technical difficulties<br />

and hence it is more conveninet to demand local chiral symmetry. Since in the<br />

end we consider only physical processes which are by definition associ<strong>at</strong>ed with<br />

physical fields, i.e. photon or W ± , the final formulae only invovle those fields<br />

and notheing of those, which exist only due to the larger chiral local symmetry.<br />

Thus: One does more than needed, this is technically simpler, and one does<br />

nothing wrong, if on calcul<strong>at</strong>es physical processes.<br />

12.6.3 Local chiral symmetry<br />

We know th<strong>at</strong> the <strong>QCD</strong>-Lagrangean is invariant under global chiral transform<strong>at</strong>ions.<br />

Consider now the coupling of an external field, lets say r µ (x) to the<br />

186


<strong>QCD</strong>-Lagrangean. Since the Lagrangean has the symmetries, it can happen th<strong>at</strong><br />

we meet the situ<strong>at</strong>ion th<strong>at</strong> two different [ external ] fields [ r µ (1) (x) and ] r µ (2) (x) have<br />

the same gener<strong>at</strong>ing functional Z l µ ,r µ (1) ,s,p = Z l µ ,r µ (2) ,s,p . Apparently<br />

the gener<strong>at</strong>ing functional Z [l µ ,r µ ,s,p] must have a very particular structure, in<br />

order to reflect this property. The transition r µ (1) (x) → r µ (2) (x) can in this case<br />

be compens<strong>at</strong>ed by a transition U (1) (x) → U (2) (x), which is a chiral transform<strong>at</strong>ion<br />

such th<strong>at</strong> there exists a R(x) and L(x) with U (2) (x) = R(x)U (1) (x)L † (x).<br />

One can revert this argument: For a transform<strong>at</strong>ion U (1) (x) → U (2) (x) =<br />

R(x)U (1) (x)L † (x) there exists a transform<strong>at</strong>ion r (1)<br />

µ (x) → r (2)<br />

µ (x) which compens<strong>at</strong>es<br />

the effect of the chiral transform<strong>at</strong>ion on U and leaves the gener<strong>at</strong>ing<br />

functional Z [l µ ,r µ ,s,p] invariant. In a famous paper by Gasser and Leutwyler<br />

Ann.Phys. Vol.158 (1984)p.142 und Ann.Phys. Vol.235 (1994) p.165 it has been<br />

shown th<strong>at</strong> the compens<strong>at</strong>ing transform<strong>at</strong>ion looks formally like some sort of a<br />

chiral gauge transform<strong>at</strong>ion for the external fields: Under the transform<strong>at</strong>ions<br />

applied to U<br />

R(α R (x)) = exp(−i τa αR a (x) )<br />

2<br />

L(α L (x)) = exp(−i τa αL a(x)<br />

)<br />

2<br />

the external fields behave like<br />

r µ → Rr µ R † + iR∂ µ R †<br />

l µ → Ll µ l † + iL∂ µ L †<br />

s + ip → R(s + ip)L †<br />

s − ip → L(s − ip)R †<br />

and then the Z [l µ ,r µ ,s,p] remains unchanged. Now: We know from the previous<br />

section th<strong>at</strong> this particular structure of Z [l µ ,r µ ,s,p] guarantees, th<strong>at</strong> the<br />

Ward-identities, which provide links between the divergences of currents and<br />

the currents, are fulfilled, if one concentr<strong>at</strong>es on vector and axial currents. In<br />

detail:<br />

In <strong>QCD</strong> you have Ward-Identities. These are identities, which connect the<br />

divergences of currents with the currents either <strong>at</strong> certain momenta or <strong>at</strong> certain<br />

coordin<strong>at</strong>es. One can derive those ward identities directly from the <strong>QCD</strong> (see<br />

above). There are certain Ward identities connected with vector currents and<br />

with axial currents, i.e. directly with the symmetries which we consider here and<br />

which are connected with spontaneous broken chiral symmetry. In Gasser and<br />

Leutwylers famous paper (1984) it has been shown, th<strong>at</strong> gener<strong>at</strong>ing functionals,<br />

where the external fields fol<strong>low</strong> the above transform<strong>at</strong>ion laws, reproduce these<br />

187


Ward identities. Thus in order to handle this formally we introduce like in a<br />

gauge theory covariant deriv<strong>at</strong>ives:<br />

which obbeys the law<br />

(D µ U)(x) = ∂ µ U(x) + iU(x)l µ (x) − ir µ (x)U(x)<br />

D µ U → RD µ UL †<br />

We also define field strength tensors (not to be used for the definition of kinetic<br />

terms fo the background fields) as:<br />

F R µν(x) = ∂ µ r ν (x) + ∂ ν r µ (x) − i[r µ (x),r ν (x)] → R(x)F R µν(x)R † (x)<br />

F L µν(x) = ∂ µ l ν (x) + ∂ ν l µ (x) − i[l µ (x),l ν (x)] → L(x)F L µν(x)L † (x)<br />

With Tr(l µ ) = Tr(r µ ) = 0 (trace over the lambda m<strong>at</strong>rices) fol<strong>low</strong>s immedi<strong>at</strong>ely<br />

th<strong>at</strong><br />

Tr(F L µν) = Tr(F R µν) = 0<br />

Fol<strong>low</strong>ing Gasser and Leutwyler we introduce the combin<strong>at</strong>ion<br />

χ(x) = 2B 0 (s(x) + ip(x))<br />

We write down the chiral counting rules: We define the U as of order<br />

U = Ord(E 0 )<br />

and the other terms according their number of deriv<strong>at</strong>ives and each deriv<strong>at</strong>ive<br />

we count as Ord(E 1 ). This means th<strong>at</strong> in the process considered we must have<br />

all Mandelstam variables small compared to a hadronic scale, i.e. Rho-Mass.<br />

U = Ord(E 0 )<br />

D µ U r µ l µ = Ord(E 1 )<br />

L µν R µν s r = Ord(E 2 )<br />

Altogether we can summarize: The simplest effective Lagrangean with two<br />

deriv<strong>at</strong>ives reads<br />

L 2 = F 2 0<br />

4 Tr(D µUD µ U † ) + F 2 0<br />

2 Tr(χU † + Uχ † )<br />

188


12.7 Applic<strong>at</strong>ion: Pion decay<br />

We consider in this section the decay of the ion in the chiral perturb<strong>at</strong>ion theory<br />

using the simplest approach, i.e. L 2 with covariant deriv<strong>at</strong>ives. The decay of<br />

the pion (see previous section) happens via an intermediar W − -boson, which<br />

is coupled one one side to the leptons, and on the other side to the pion. The<br />

coupling to the leptons reads<br />

L coupl =<br />

e<br />

2 √ 2sin(θ W ) (W + µ ν (µ) γ µ (1 − γ 5 )µ − + W − µ ¯µ − γ µ (1 − γ 5 )ν (µ)<br />

The coupling of the W-bosons to the pion is obtained by the fact th<strong>at</strong> the W-<br />

bosonic field appears in the field l µ (x) , which is in the covariant deriv<strong>at</strong>ive. We<br />

have r µ (x) = 0 and<br />

and<br />

e<br />

l µ = −√ (W µ + T + + h.c.)<br />

2 sin ϑW<br />

L 2 = F 2 0<br />

4 Tr(D µUD µ U † ) + F 2 0<br />

2 B 0Tr(MU † + UM † )<br />

D µ U = ∂ µ + iUl µ<br />

Similar as in the pion-pion-sc<strong>at</strong>tering we have to extract from this lagrangean<br />

the term which corresponds to the pion decay. The simplest term is just linar<br />

in l µ (x) and we disentengle the L 2 in order to find this:<br />

F 2 0<br />

4 Tr(D µU(D µ U) † ) = F 2 0<br />

4 Tr ( (∂ µ + iUl µ )(∂ µ − il µ U † ) )<br />

= .... + i F 0<br />

2 Tr ( l µ ∂ µ U † U )<br />

We set<br />

l µ (x) = λa<br />

2 la µ(x)<br />

If we compare with the well known expression for the left handed current of the<br />

effective Lagrangean, i.e.<br />

then we can write<br />

J µ,a<br />

L<br />

(x) = if2 π<br />

4 Tr [ λ a ∂ µ U † U ]<br />

L coupl = lµ(x)J a µ,a<br />

L<br />

(x)<br />

We have now the term linear in l µ (x) and want to expand it in order to find the<br />

term linear in the pion field. For this we expand the current J µ,a<br />

L<br />

(x) up to first<br />

order in φ:<br />

J µ,a<br />

L = iF 2 0<br />

4 T ( λ a ∂ µ U † U ) = F 0<br />

4 Tr (λa ∂ µ φ) + ord(φ 2 )<br />

189


Since we have Tr(λ a λ b ) = 2δ ab we can simplify<br />

With this we have in particular<br />

J µ,a<br />

L = F 0<br />

2 ∂µ φ + ord(φ 2 )<br />

< 0|J µ,a<br />

L (0)|φb (p) >= F 0<br />

2 < 0|∂µ φ a (0)|φ b (p) >= ip µ F 0<br />

2 δab<br />

Now we insert l µ<br />

e<br />

l µ = −√ (W µ + T + + h.c.)<br />

2 sin ϑW<br />

and we obtain<br />

L coupl = F 0 e F<br />

2 ∂µ φ = −√ 0<br />

2 sin ϑW 2 Tr((W µ + T + + Wµ − T − )∂ µ φ)<br />

= − e<br />

sinϑ W<br />

F 0<br />

2 (W + µ (V ud ∂ µ π − + V us ∂ µ K − ) + W − µ (V ud ∂ µ π + + V us ∂ µ K + ))<br />

We know the propag<strong>at</strong>or for the W-bosons, which can be approxim<strong>at</strong>ed due to<br />

the large mass of the W-boson:<br />

−g µν + kµkν<br />

M 2 W<br />

k 2 − M 2 W<br />

= g µν<br />

M 2 W<br />

+ ord( k2<br />

MW<br />

4 )<br />

we get then for the invariant amplitude of the pion decay:<br />

( )<br />

e<br />

M = −i√ ū (µ) γ µ ig µν e F0<br />

(1 − γ 5 )v¯ν(µ) −i<br />

2 sinϑW sin(θ W ) 2 V ud(−ip ν )<br />

Here the p is the four-momentum of the pion and G F = 1.166x10 −5 GeV −2 .<br />

Fol<strong>low</strong>ing Bjorken-Drell sec. 10.14 one obtains for the decay r<strong>at</strong>e (replace √ a 2<br />

=<br />

V ud F 0 ) the expression<br />

( )<br />

1<br />

τ = G2 F V ud<br />

2<br />

4π F 0 2 m π m 2 µ 1 − m2 µ<br />

m 2 π<br />

Actually, we had this expression already in the previous section on pion decay,<br />

however, there it was simply taken from the liter<strong>at</strong>ure, here we have derived it<br />

in the chiral perturb<strong>at</strong>ion theory. Aplparently the constant F 0 is the pion decay<br />

constant, shich in the present papproxim<strong>at</strong>ian is identical to the Kaon decay<br />

constant. In n<strong>at</strong>ure we have 93MeV for the pion one and 113MeV for the Kaon<br />

decay constant.<br />

M 2 W<br />

190


12.8 The chiral Lagrange-Density in ord(p 4 ) or ord(E 4 ) and<br />

renormaliz<strong>at</strong>ion.<br />

12.8.1 The Lagrangean<br />

We quote the most general Lagrange-Density of ord(p 4 ) as it is presented by<br />

Gasser and Leutwyler in Nucl.Phys. B250 (1985) 465:<br />

with<br />

L = L 2 + L 4<br />

L 4 = L 1 Tr(D µ U(D µ U) † ) 2 + L 2 Tr(D µ U(D ν U) † )Tr(D µ U(D ν U) † )<br />

+L 3 Tr(D µ U(D µ U) † D ν U(D ν U) † ) + L 4 Tr(D µ U(D µ U) † Tr(χU † + Uχ † )<br />

+L 5 Tr(D µ U(D µ U) † (χU † + Uχ † )) + L 6 (Tr(χU † + Uχ † )) 2<br />

+L 7 (Tr(χU † − Uχ † )) 2 + L 8 Tr(Uχ † Uχ † + χU † χU † )<br />

−iL 9 Tr(F R µνD µ U(D ν U) † + F L µν(D µ U) † D ν U) + L 10 Tr(UF L µνU † F µν<br />

R )<br />

+H 1 Tr(F R µνF µν<br />

R + F L µνF µν<br />

L ) + H 2Tr(χχ † )<br />

The terms H 1 and H 2 are chirally invariant without involving the m<strong>at</strong>rix<br />

U. They do not gener<strong>at</strong>e any couplings to goldstone bosons and hence are not<br />

of gre<strong>at</strong> phenomenological interest. However, if one were to use the effective<br />

Lagrangean to describe correl<strong>at</strong>ion funcitions of the external sources, these two<br />

oper<strong>at</strong>ors can gener<strong>at</strong>e contact terms (see be<strong>low</strong> the renormaliz<strong>at</strong>ion program).<br />

The values of the so called <strong>low</strong> energy constants (Leutwyler coefficients) L i<br />

cannot be determined purely from chiral symmetry breaking, as it was possible<br />

in L 2 with F 0 = f π and B 0 rel<strong>at</strong>ed to m π . They are dynamical coefficients,<br />

whcih should be determined from <strong>QCD</strong> directly, it this was possible. They<br />

should perhaps be calcul<strong>at</strong>ed in some <strong>QCD</strong>-inspired model as it is possible with<br />

the Chiral Quark Soliton MOdel. In practice they are determined by adjusting<br />

them to experimental d<strong>at</strong>a as described be<strong>low</strong>.<br />

Remark: The effective lagrangean may be used in the context of either<br />

chiral SU(2) of SU(3). Because SU(2) is a subgroup of SU(3) the general SU(3)<br />

Lagrangean is also valid for chiral SU(2). However, the SU(2) version has fewer<br />

<strong>low</strong> energy constants, so th<strong>at</strong> only certain combin<strong>at</strong>ions of the L r i will apear in<br />

purely pionic processes. If one is dealing with pions <strong>at</strong> <strong>low</strong> energy the kanos<br />

and the eta are considered as heavy particles and may be integr<strong>at</strong>ed out. THis<br />

procedure produces a shift in the values of the <strong>low</strong> energy renormalized constants<br />

such tha the coefficients of a purely SU(2) lagrangean and an SU(3) one will<br />

differ by a finite amount, which can be exactly calcul<strong>at</strong>ed. Actually the SU(2)<br />

coefficients can be found by first performing calcul<strong>at</strong>ions in the SU(3) limit and<br />

then tre<strong>at</strong>ing the masses of kaons and eta as going to infinity.<br />

191


12.8.2 Chiral perturb<strong>at</strong>ion theory, renormaliz<strong>at</strong>ion program<br />

The way physics is extracted from the effective Lagrangean is called “Chiral<br />

perturb<strong>at</strong>ion theory”. It works in order(E 4 ) in the fol<strong>low</strong>ing way, which contains<br />

three ingredients:<br />

• The general Lagrangean L 2 which is to be used both <strong>at</strong> tree level and in<br />

loop diagrams.<br />

• The general lagrangean L 4 which is to be used only <strong>at</strong> tree-level.<br />

• The renormaliz<strong>at</strong>ion program which describes how to make physical predictions<br />

<strong>at</strong> one-loop level.<br />

• Generaliz<strong>at</strong>ion of all th<strong>at</strong> to order(E 6 )<br />

D=4<br />

Above: Renormaliz<strong>at</strong>ion of loops with trees. Attention: There are various<br />

different 1-loop-graphs with L 2 insertions since L 2 has to be expanded in the<br />

physical fields. There are also various different tree graphs of L 4 due to the<br />

same reason.<br />

Actually for the last point needs some explan<strong>at</strong>ion: Consider the calcul<strong>at</strong>ion<br />

of an actual process. In <strong>low</strong>est order one should use L 2 for th<strong>at</strong>. When expanded<br />

in terms of the meson fields π a it specifies a set of interaction vertices. These<br />

can be used to calcul<strong>at</strong>e treee-level and one-loop diagrams for any process of<br />

interest. This result is added to the contribution which comes from the vertices<br />

contained in the lagrangean L 4 , tre<strong>at</strong>ed <strong>at</strong> tree-level only (because then it has<br />

the same dimension in Weinbergs counting). If we consider the one-loop graphs,<br />

we realize th<strong>at</strong> they are infinite.<br />

One sees this immedi<strong>at</strong>ely if one considers a typical 1-loop-diagram, which<br />

always contains the fol<strong>low</strong>ing term<br />

∫<br />

d 4 k<br />

(2π) 4<br />

∫<br />

i<br />

k 2 − M 2 + iɛ ∼<br />

k 3 1<br />

dk<br />

k 2 − M 2 + iɛ → ∞<br />

Divergent parts in loops are not a problem, because there are well known<br />

techniques to handle those infinities. One is the dimensional regulariz<strong>at</strong>ion,<br />

where one considers the Feynman integrals not in a space with dimension d = 4<br />

192


ut in a general space with dimension d > 4, does the integrals, and then performs<br />

the limit d → 4. This procedure al<strong>low</strong>s to identify from each Feynman<br />

diagram the part which is finite and which contains the physics, and the part<br />

which is infinite and is just there, because the integrals extend to infinite momenta,<br />

where th<strong>at</strong> part of the physics, which is described by the <strong>low</strong> energy<br />

Lagrangean, is no longer valid. Actually we will use this dimensional regulariiz<strong>at</strong>ion.<br />

There are various conventions for doing th<strong>at</strong>, we take the one of<br />

Donoghue, Go<strong>low</strong>ifh and Holstein and write<br />

∫<br />

d 4 ∫<br />

k<br />

(2π) 4 → µ4−d<br />

d d k<br />

(2π) d<br />

where the scale µ has to be introduced in order to keep the dimension of the<br />

expression for any d. The µ is an auxiliary parameter, the so called renormaliz<strong>at</strong>ion<br />

point or subtraction point. The final observalbe does not depend on it.<br />

Thus we will meet always in the loops the fol<strong>low</strong>ing integral,<br />

∫<br />

I(M 2 ,µ 2 ) = µ 4−d d d [ ]<br />

k i<br />

(2π) d k 2 − M 2 + iɛ = M2<br />

16π 2 R + log( M2<br />

µ 2 )<br />

with<br />

R = 2<br />

d − 4 − [log(4π) + Γ′ (1) + 1]<br />

where the Γ ′ (x) is the deriv<strong>at</strong>ive of the Γ-function. Actually the result of the<br />

integr<strong>at</strong>ion is proportional to a Gamma function Γ(− d 2<br />

), which diverges for<br />

d<br />

2<br />

= 1,2,3,... but converges for values in between. Using an expansion of Γ(x)<br />

<strong>at</strong> the point x = − d 2<br />

and <strong>at</strong> d = 4 one obtains the deriv<strong>at</strong>ive of the Gamma<br />

function due to the expansion and then the final result from above. One shoud<br />

note th<strong>at</strong> Γ ′ (1) = −γ = −0.5772 = d dz Γ(z + 1)| z=0.<br />

Taking this into account <strong>at</strong> this stage (after calcul<strong>at</strong>ing tree and 1-loop for<br />

L 2 and tree for L 4 ) the result contains both bare parameters L i and divergent<br />

loop integrals. One needs to determine the parameters from experiment. If the<br />

Lagrangea is indeed the most general one possible, rel<strong>at</strong>ions between observables<br />

will be FINITE when expressed in terms of physical quantities. Thus all the<br />

divergences can be absorbed in redefining the <strong>low</strong> energy coefficients L i . (We<br />

will show this explicitely in the example be<strong>low</strong>, where we calcul<strong>at</strong>e the meson<br />

masses in one-loop form). This is called “renormaliz<strong>at</strong>ion” of the <strong>low</strong> energy<br />

coefficients, i.e. one says: The coefficient in L 4 consists of two terms, one is<br />

the coefficient L r i , containing the physics and determining the observables and<br />

being finite, plus a term which compens<strong>at</strong>es the divergent part from the one-loop<br />

diagrams of L 2 . Thus one has<br />

L i = L r i + Γ i<br />

32π 2 R i = 1,.....8<br />

H i = H r i + ∆ i<br />

32π 2 R i = 1,2<br />

193


The constants Γ i and ∆ i are in the table be<strong>low</strong>. The renormalized <strong>low</strong> energy<br />

coefficients are dependent on the renormaliz<strong>at</strong>ion scale µ. The coefficients for<br />

two different scales are rel<strong>at</strong>ed to each other by<br />

L r i(µ 2 ) = L r i(µ 1 ) + Γ i<br />

16π 2 log(µ 1<br />

µ 2<br />

)<br />

The <strong>low</strong> energy coefficients are given vor SU(3) in the fol<strong>low</strong>ing table. They<br />

are real numbers and given be<strong>low</strong> in units of 10 −3 <strong>at</strong> the renormaliz<strong>at</strong>ion scale<br />

of the Rho-mass (see Bijnens, Ecker and Gasser, The socond DAΦNE physics<br />

Handbook Nucl. Phys. The conventions are taken from the book by Donoghue,<br />

Go<strong>low</strong>ich and Holstein “Dynamics of the Standard model”.<br />

coeff emp. value experiment Γ i<br />

L r 3<br />

1 0.4 ± 0.3 ππ → ππ<br />

32<br />

L r 3<br />

2 1.35 ± 0.3 ππ → ππ<br />

16<br />

L r 3 −3.5 ± 1.1 ππ → ππ 0<br />

L r 1<br />

4 −0.3 ± 0.5 ∼ 0 Zweig-Regel<br />

f K<br />

8<br />

3<br />

fπ 8<br />

L r 5 1.4 ± 0.5<br />

L r 11<br />

6 −0.2 ± 0.3 Zweig-Regel<br />

144<br />

L r 7 −0.4 ± 0.2 Gell-Man-Okube,L 5 ,L 8 0<br />

L r 8 0.9 ± 0.3 M K 0 − M K +,L 5 ,(2m s − m u − m d) : (m d − m u )<br />

5<br />

48<br />

L r 9 6.9 ± 0.7 isovector radius of the pion<br />

1<br />

4<br />

L r 10 −5.5 ± 0.7 π → eνγ − 1 4<br />

Remark: There exists always an ambiguity of wh<strong>at</strong> finite constants should<br />

be absorbed into the renormalized <strong>low</strong> energy coefficients. This lambiguity<br />

does not affect the rel<strong>at</strong>ionship between observables, but only influences the<br />

numerical values quaoted for the <strong>low</strong> energy constants. One should keep this<br />

in mind, if one compares sets of constants of different authors. Also there are<br />

various regulariz<strong>at</strong>ion procedures for handling the divergent integrals. They also<br />

influence not the observalbes but the numerical values of L i .<br />

Erklaere was<br />

die Zweig<br />

Regel ist.<br />

12.9 Applic<strong>at</strong>ion in order(p 4 ): Masses of Goldstone bosons<br />

12.9.1 Aim of this section<br />

We have used once the L 2 in tree level to calcul<strong>at</strong>e the masses of the Goldstone<br />

bosons. There it meant th<strong>at</strong> we looked <strong>at</strong> the Lagrangean of the Goldstone boson<br />

and we identified the mass of the particle with the coefficient of the quadr<strong>at</strong>ic<br />

term. We learned th<strong>at</strong> Gell-Mann-Okubo and Gell-Mann Oaks Renner etc. were<br />

all fulfilled, which shows th<strong>at</strong> such a procedure is not stupid. In this section<br />

we calcul<strong>at</strong>e the masses of the Goldstone bosons again, however, we will do it<br />

system<strong>at</strong>ically better. We will use tree- and 1-loop-diagrams of L 2 and treediagrams<br />

of L 4 . This corresponds system<strong>at</strong>ically to all terms with d = 4. We<br />

will demsonstr<strong>at</strong>e in this calcul<strong>at</strong>ion the fol<strong>low</strong>ing fe<strong>at</strong>ures, which are the basic<br />

points in the chiral perturb<strong>at</strong>ion theory:<br />

194


• The connection between the bare coefficients L i and the renormalized ones<br />

L r i is done in such a way (see above formula) th<strong>at</strong> the infinities of the 1-<br />

loop diagrams of L 2 are compens<strong>at</strong>ed by the tree-diagrams of L 4 .<br />

• The scale dependence (dpendence on µ 2 ) is such, th<strong>at</strong> the expressions for<br />

observables are independent on the choice of µ.<br />

We consider L =L 2 +L 2 but do not take into account external fields. We<br />

consider also isospin symmetry, i.e. assume up and down quarks to be degener<strong>at</strong>e.<br />

First we have to learn, how to calcul<strong>at</strong>e masses beyond tree-level. Actually<br />

there are several definitions of a mass of a particle. One is the so called “Pole<br />

mass”. This means: Consider the particular example of the Propag<strong>at</strong>or of a<br />

meson field . The propag<strong>at</strong>or is the Fourier-transform of the Greens 2-point<br />

function:<br />

∫<br />

i∆(p) = d 4 xexp(−ipx) < 0|T [Φ(x)Φ(0)] |0 ><br />

In <strong>low</strong>est order of perturb<strong>at</strong>ion theory this propag<strong>at</strong>or reads<br />

i∆ 0 (p) =<br />

i<br />

p 2 − M 2 0 + iɛ<br />

If we calcul<strong>at</strong>e from the Lagrangean it we find out th<strong>at</strong> the M 0 of the propag<strong>at</strong>or<br />

is exactly the quadr<strong>at</strong>ic term in the Lagrangean, because this part defines the<br />

“free” lagrangean and higher powers of the field φ are considered as perturb<strong>at</strong>ion,<br />

which we tre<strong>at</strong> in this example in <strong>low</strong>est order. Thus in this approxim<strong>at</strong>ion<br />

(<strong>low</strong>est order) the mass in the Lagrangean and the mass M 0 in the propag<strong>at</strong>or<br />

are the same. The M 0 is called pole mass, since the propag<strong>at</strong>or has a pole <strong>at</strong><br />

p 2 = M 2 0.<br />

12.9.2 Calcul<strong>at</strong>ion of the pole mass<br />

We want now to calcul<strong>at</strong>e the propag<strong>at</strong>or in a better approxim<strong>at</strong>ion. If one<br />

fol<strong>low</strong>s standard books on field theory one can write down the full propag<strong>at</strong>or<br />

i∆(p) of the interacting field as a sum of all connected diagrams. In practice<br />

one does not calcul<strong>at</strong>e this explicitely bit uses a simple rel<strong>at</strong>ionship between<br />

this and the self-energy Σ(p), which is ths sum of all one-particle-irreducible<br />

diagrams.<br />

∆(p) −1 = ∆ −1<br />

0 (p) − Σ(p)<br />

or equivalently<br />

i∆(p) =<br />

i<br />

p 2 − M 2 0 + iɛ +<br />

i<br />

i<br />

p 2 − M0 2 + ))<br />

iɛ(−iΣ(p2 p 2 − M0 2 + iɛ<br />

i<br />

i<br />

i<br />

+<br />

p 2 − M0 2 + ))<br />

iɛ(−iΣ(p2 p 2 − M0 2 + ))<br />

iɛ(−iΣ(p2 p 2 − M0 2 + ......<br />

+ iɛ<br />

i<br />

=<br />

p 2 − M0 2 − Σ(p2 ) + iɛ<br />

195


Since Σ(p 2 ) consists of one-particle-irreducible diagrams we must apply to it the<br />

chiral counting. Th<strong>at</strong> means it consits of a sum of tree-diagram with L 4 plus a<br />

1-loop-diagram with L 2 since the tree diagram with L 2 is already contained in<br />

M 0 .<br />

So in the chiral counting the Σ(p 2 ) is of order ord(E 4 ) or ord(p 4 ) because<br />

D = 2 + ∑ 2(n − 1)N 2n + 2N L = 4. If we look to the above expansion, we have<br />

the series<br />

ord(E −2 )+ord(E −2 )ord(E 4 )ord(E −2 )+ord(E −2 )ord(E 4 )ord(E −2 )ord(E 4 )ord(E −2 )+......<br />

or ord(E −2 ) + ord(E 0 ) + ord(E 2 ) + .... Thus we have to stop the series after<br />

the first non-trivial term. However, this is correct, if we are interested in the<br />

propag<strong>at</strong>or <strong>at</strong> momenta noticeably different from the physical mass, i.e. p 2 ≠<br />

M 2 . When we are interested in the pole, th<strong>at</strong> is in the behaviour <strong>at</strong> p 2 ≈ M 2<br />

then we must be more careful. Then we have to consider th<strong>at</strong> between the poleposition<br />

to zeroth order (i.e. p 2 = M 2 0) and the pole position to first order (i.e.<br />

p 2 = M 2 0 + Ωord(E 4 ) with Ω being a constant, which should be determined.<br />

This corresponds exactly to the chiral counting, because the D = 4 for the self<br />

energy insertion Σ(p 2 ). Thus we have<br />

i∆ 0 (p) =<br />

i<br />

p 2 − M0 2 + iɛ = i<br />

M0 2 − M2 0 − ord(E4 ) = ord(E−4 )<br />

This has a drastic consequence: Now we have the order counting of the series<br />

for the full propag<strong>at</strong>or like:<br />

ord(E −4 )+ord(E −4 )ord(E 4 )ord(E −4 )+ord(E −4 )ord(E 4 )ord(E −4 )ord(E 4 )ord(E −4 )+......<br />

Thus one sees th<strong>at</strong> the combin<strong>at</strong>ion (−iΣ(p 2 )) is of order unity. The<br />

p 2 −M0 2 +iɛ<br />

consequence is, th<strong>at</strong> it is consistent to sum up all contributions like this and<br />

cannot chop of the series after the first non-trivial term. To make it clear, this<br />

is consistent order counting in chiral perturb<strong>at</strong>ion theory. Actually, as we will<br />

see be<strong>low</strong>, one can handle with this fact, because one indeed can sum up the<br />

series exactly and even easily.<br />

We will explicitely do the calcul<strong>at</strong>ion be<strong>low</strong>, but lets first consider, wh<strong>at</strong> we<br />

do with it once we know the Σ(p 2 )., i.e. how do we extract the pole-mass of the<br />

field. This is actually very simple, because we have to look for the zero of the<br />

i<br />

196


inverse pole. We look for th<strong>at</strong> p 2 for which ∆ −1 (p) = p 2 − M 2 0 − Σ(p 2 ) = 0. If<br />

we call this solution p 2 = M 2 , the equ<strong>at</strong>ion to solve reads<br />

M 2 − M 2 0 − Σ(M 2 ) = 0<br />

This is an implicit equ<strong>at</strong>ion. In order to solve it and to get en explicit expression<br />

for the physical mass we bring the expression of ∆(p) as close to a form of ∆ 0 (p 2 )<br />

as possible, because from this one we know how to read off the mass.<br />

In order to achieve this we first expand as a m<strong>at</strong>hem<strong>at</strong>ical exercise the Σ(p 2 )<br />

around a point µ 2 , which is considered arbitrary for the moment.<br />

Σ(p 2 ) = Σ(µ 2 ) + (p 2 − µ 2 )Σ ′ (µ 2 ) + ˜Σ(p 2 )<br />

where we collect all the rest in the ˜Σ(p 2 ), which can be easily shown to have<br />

the property<br />

˜Σ(µ 2 ) = ˜Σ ′ (µ 2 ) = 0<br />

The propag<strong>at</strong>or can now be written as<br />

i<br />

i∆(p) =<br />

p 2 − M0 2 − Σ(µ2 ) − (p 2 − µ 2 )Σ ′ (µ 2 ) − ˜Σ(p 2 ) + iɛ<br />

We replace now the M 2 0 by the pole-equ<strong>at</strong>ion and choose as renormaliz<strong>at</strong>ion<br />

point µ 2 = M 2 . Then the propag<strong>at</strong>or reads<br />

i<br />

i∆(p) =<br />

p 2 − M 2 − (p 2 − M 2 )Σ ′ (M 2 ) − ˜Σ(p 2 ) + iɛ<br />

The popag<strong>at</strong>or has still not yet the convenient form. However we get this if we<br />

introduce the wave function renormaliz<strong>at</strong>ion constant<br />

1<br />

Z φ =<br />

1 − Σ ′ (p 2 )<br />

and then we can rewrite<br />

i<br />

i∆(p) =<br />

(p 2 − M 2 ) (1 − Σ ′ (p 2 )) − ˜Σ(p 2 ) + iɛ = iZ φ<br />

p 2 − M 2 − Z φ˜Σ(p2 ) + iɛ<br />

If we define now renormalized fields like<br />

φ R = √ φ<br />

Zφ<br />

then we have for the renormalized propag<strong>at</strong>or<br />

∫<br />

i∆(p) = d 4 xexp(−ipx) < 0|T [Φ R (x)Φ R (0)] |0 ><br />

with<br />

i<br />

=<br />

p 2 − M 2 − Z φ˜Σ(p2 ) + iɛ<br />

˜Σ(M 2 ) = 0<br />

Hence in the vicinity of the physical mass the full propag<strong>at</strong>or has the same<br />

structure as the the free propagartor with the bare mass.<br />

197


12.9.3 Calcul<strong>at</strong>ion of the self energy<br />

We have to calcul<strong>at</strong>e the self energy applying strictly the chiral counting rules.<br />

Th<strong>at</strong> is we have to use<br />

L 4φ<br />

2 for the vertices in the loop diagram<br />

L 2φ<br />

4 for tree diagram<br />

It is clear wh<strong>at</strong> L 4φ<br />

2 is, because we have derived it in the section about elastic<br />

pion sc<strong>at</strong>tering<br />

L 4φ<br />

2 = 1<br />

24F0<br />

2 Tr ([φ,∂ µ φ]φ∂ µ φ) + 1<br />

24F0<br />

2 B 0 Tr(Mφ 4 )<br />

For L 2φ<br />

4 we have to consider in detail the structure of L 4 . Since we want to<br />

have expressions with two fields, we see immedi<strong>at</strong>ely th<strong>at</strong> only the terms with<br />

L 4 ,L 5 ,L 6 and L 7 have to be considered. Take e.g. the L 4 -term. We can rewrite<br />

it as<br />

L 4 Tr(∂ µ U∂ µ U † Tr(χU † + Uχ † )<br />

= L 4<br />

2<br />

F 2 0<br />

(<br />

∂µ η∂ µ η − ∂ µ π 0 ∂ µ π 0 + 2∂ µ π + ∂ µ π + + 2∂ µ K + ∂ µ K + + 2∂ µ K 0 ∂ µ ¯K0 ) (4B 0 (2m+m s )<br />

We proceed analogously for the other terms and obtain in the end<br />

L 2φ<br />

4 = 1 2 (a η∂ µ η∂ µ η − b η η 2 ) + 1 2 (a π∂ µ π 0 ∂ µ π 0 − b π π 0 π 0 ) + a π ∂ µ π + ∂ µ π −<br />

−b π π + π − + a k ∂ µ K + ∂ µ K − − b k K + K − + a K ∂ µ K 0 ∂ µ ¯K0 − b K K 0 ¯K0<br />

with m = 1 2 (m u + m d ) and the the constants<br />

b η = 64B 0<br />

3F 2 0<br />

a η = 16B 0<br />

F 2 0<br />

((2m + m s )L 4 + 1 3 (m + 2m s)L 5<br />

)<br />

(<br />

(2m + ms )(m + 2m s )L 6 + 2(m − m s ) 2 L 7 + (m 2 + m 2 s)L 8<br />

)<br />

a π = 16B 0<br />

F0<br />

2 ((2m + m s )L 4 + mL 5 )<br />

b π = 64B2 (<br />

0 (2m + ms )mL 6 + m 2 )<br />

L 8<br />

F 2 0<br />

a K = 16B 0<br />

((2m + m s )L 4 + 1 )<br />

2 (m + m s)L 5<br />

b K = 32B2 0<br />

F 2 0<br />

F 2 0<br />

((2m + m s )(m + m s )L 6 + 1 2 (m + m s) 2 L 8<br />

)<br />

With these expressions the self <strong>energies</strong> are of the form<br />

Σ φ (p 2 ) = A φ + B φ p 2<br />

198


Here indic<strong>at</strong>es the index φ the various contributions separ<strong>at</strong>ely from pions, kaons<br />

and eta. Apparently we have only diagonal terms in the self energy, where the<br />

incoming and outgoing boson are the same. One can show this structure easily:<br />

Each of the coefficients A and B consists of two contributions: One coming<br />

from the Tree-graph with the vertex based upon L 2φ<br />

4 and another coming from<br />

the 1-loop graph with a vertex based on L 4φ<br />

2 . If we look <strong>at</strong> the tree graph of<br />

L 2φ<br />

4 and <strong>at</strong> the explicit expression for it (see above) we see th<strong>at</strong> the terms in<br />

L 2φ<br />

4 there are either two deriv<strong>at</strong>ives or none, corresponding to p 2 or p 0 . So for<br />

this the structure of Σ(p 2 ) is obvious. Indeed, the fol<strong>low</strong>ing example shows this<br />

explicitely for the η-terms of L 2φ<br />

4 , it yields<br />

1<br />

−iΣ tree<br />

η (p 2 ) = i2(<br />

2 a η(ip µ )(−ip µ ) − 1 )<br />

2 b η = i(a η p 2 − b η )<br />

For the terms of the 1-loop contribution of L 4φ<br />

2 the argument for the structure<br />

Σ(p 2 ) = A + Bp 2 goes as fol<strong>low</strong>s: L 4φ<br />

2 has either two deriv<strong>at</strong>ives (symbolically<br />

φφ∂φ∂φ) or no deriv<strong>at</strong>ive (symbolically φ 4 ). The first term is proportional to<br />

M 2 if the φ are contracted to external lines, and it is proportional to p 2 , if the<br />

∂φ are connected with external lines. The second term does not yield external<br />

momenta. Thus we obtain the structure Σ(p 2 ) = A + Bp 2 .<br />

We show now, how the loop diagrams are actually calcul<strong>at</strong>ed. For this<br />

we consider first the vertex of L 2 which we actually know from the pion-pion<br />

sc<strong>at</strong>tering: there we had for the Feynman-amplitude M the fol<strong>low</strong>ing expression<br />

(We do not use the second expression of the pion-pion-vertex, as we have done<br />

in the elastic pion sc<strong>at</strong>tering, because th<strong>at</strong> expression assumes physical particles<br />

in all four legs). Thus we have<br />

with<br />

−i6F 2 0 M(p a ,p b ,p c ,p d ) = 2A + B<br />

A = δ ab δ cd (−ip a − ip b ) (ip c + ip d ) + δ ac δ bd (−ip a + ip c )(−ip b + ip d )<br />

+δ ad δ bc (−ip a + ip d ) (−ip b + ip c ) − 4(δ ab δ cd ((−ip a )(−ip b ) + (ip c )(ip d ))<br />

+δ ac δ bd ((−ip a )(ip c ) + (−ip b )(ip d )) + δ ad δ bc ((−ip a )(ip d ) + (ip b )(ip c )))<br />

B = m2 π<br />

4 8( δ ab δ cd + δ ac δ bd + δ ad δ bc)<br />

Let us consider this for the pion-loop to the self energy of π 0 , i.e. with<br />

a = 3 p a = p b = j p b = k c = 3 p c = p p d = k d = j<br />

199


We obtain for the 1.loop contribution<br />

Loop = 1 2<br />

∫<br />

d 4 k<br />

(2π) 4<br />

i<br />

3F 2 0<br />

3∑<br />

i<br />

(X j + Y j + Z j )<br />

k 2 − m 2 π + iɛ<br />

j=1<br />

X j = δ 3j δ 3j [ (p + k) 2 + 2pk + m 2 ]<br />

π<br />

Y j = δ 33 δ jj [ (p − p) 2 − 2p 2 − 2k 2 + m 2 ]<br />

π<br />

Z j = δ 3j δ 3j [ (p − k) 2 − 2pk − 2kp + 5m 2 ]<br />

π<br />

For the evalu<strong>at</strong>ion we also need the divergent loop integral and the expression<br />

∫<br />

d n ∫<br />

k k µkν<br />

i<br />

(2π) n k 2 − M 2 + iɛ = M2<br />

n g d n k i<br />

µν<br />

(2π) n k 2 − M 2 + iɛ = M2<br />

n µn−4 I(M 2 ,µ 2 )<br />

performing the integrals yields:<br />

loop = 1 ∫<br />

d 4 k i<br />

2 (2π) 4<br />

3F 2 0<br />

[<br />

−4p 2 − 4k 2 + 5m 2 π] i<br />

k 2 − m 2 π + iɛ<br />

= i<br />

6F0<br />

2 (−4p 2 + m 2 π)I(m 2 π,µ 2 )<br />

Apparently we have the structure, which we predicted i.e. Σ φ (p 2 ) = A φ +B φ p 2 .<br />

If one performs the calcul<strong>at</strong>ions with these 1-loop integrals, one obtains:<br />

(<br />

A π = m2 π<br />

F0<br />

2 − 1 6 I(m2 π) − 1 6 I(m2 η) − 1 )<br />

3 I(m2 K) + 32[(2m + m s )B 0 L 6 + mB 0 L 8 ]<br />

(<br />

A K = m2 K 1<br />

F0<br />

2 12 I(m2 η) − 1 4 I(m2 π) − 1 2 I(m2 K) + 32<br />

[(2m + m s )B 0 L 6 + 1 ])<br />

2 (m + m s)B 0 L 8<br />

A η = m2 η<br />

(− 2 )<br />

3 I(m2 η) + 16m 2 ηL 8 + 32(2m + m s )B 0 L 6<br />

+ m π<br />

F 2 0<br />

B K = 1 I(m 2 η)<br />

4<br />

F 2 0<br />

( 1<br />

6 I(m2 η) − 1 2 I(m2 π) + 1 3 I(m2 K))<br />

+ 128<br />

9<br />

B0(m 2 − m s ) 2<br />

(3L 7 + L 8 )<br />

B π = 2 I(m 2 π)<br />

3 F0<br />

2 + 1 I(m 2 K )<br />

3 F0<br />

2 − 16B 0<br />

F0<br />

2 [(2m + m s )L 4 + mL 5 ]<br />

F 2 0<br />

+ 1 I(m 2 π)<br />

4 F0<br />

2 + 1 I(m 2 K )<br />

2 F0<br />

2 − 16B 0<br />

F0<br />

2<br />

B η = I(m2 K )<br />

F 2 0<br />

F 2 0<br />

[(2m + m s )L 4 + 1 2 (m + m s)L 5<br />

]<br />

− 8m2 η<br />

F0<br />

2 L 5 − 16B 0<br />

F0<br />

2 [2m + m s ]B 0 L 4<br />

Here the appearing masses (e.g. m 2 π are “abbrevi<strong>at</strong>ions” for the terms<br />

m 2 π = m 2 π,2 = 2B 0 m<br />

200


m 2 K = m 2 K,2 = B 0 (m + m s )<br />

m 2 η = m 2 η,2 = 2 3 B 0(m + 2m s )<br />

We determine the masses by solving the equ<strong>at</strong>ion<br />

M 2 − M 2 0 − Σ(M 2 ) = 0<br />

We do this in the fol<strong>low</strong>ing careful way for each boson separ<strong>at</strong>ely (omit the<br />

index φ):<br />

M 2 = M 2 0 + A + Bp 2<br />

or<br />

Then<br />

or approxim<strong>at</strong>ely<br />

This yields<br />

M 2 = M 2 0 + A + BM 2<br />

M 2 (1 − B) = M 2 0 + A<br />

M 2<br />

1 + B ≈ M2 0 + A<br />

M 2 ≈ M 2 0(1 + B) + A(1 + B)<br />

We can (and must) simplify this expression further by considering carefully<br />

the ordering in the chiral perturb<strong>at</strong>ion approach. We have M 2 = ord(E 2 ),<br />

M 2 0 = ord(E 2 ), and since Σ(p 2 ) = A + Bp 2 = ord(E 4 ) we have A = ord(E 4 )<br />

and B = ord(E 2 ). Thus, if we collect in the above equ<strong>at</strong>ion all terms up to<br />

ord(E 4 ) we obtain the result:<br />

M 2 = M 2 0(1 + B) + A<br />

Now we are close to the end: With this prescription we write down the masses<br />

and replace simultaneously the bare <strong>low</strong> energy coefficients by the renormalized<br />

ones (plus the divergent integral of course), i.e. inserting<br />

L i = L r i + Γ i<br />

32π 2 R i = 1,.....8<br />

H i = H r i + ∆ i<br />

32π 2 R i = 1,2<br />

This yields then the final expression up to terms of ord(E 4 ):<br />

[<br />

m 2 π,4 = m 2 π,2<br />

1 + m2 π,2<br />

32π 2 F 2 0<br />

log( m2 π,2<br />

µ 2 ) − m2 η,2<br />

96π 2 F 2 0<br />

]<br />

log( m2 η,2<br />

µ 2 )<br />

+ 16m2 π,2<br />

F 2 0<br />

[(2m + m s )B 0 (2L r 6 − L r 4) + mB 0 (2L r 8 − L r 5)]<br />

201


+ 16m2 K,2<br />

F 2 0<br />

+m 2 π,2<br />

m 2 η,4 = m 2 η,2<br />

+ 16m2 η,2<br />

F 2 0<br />

[<br />

m<br />

2<br />

η,2<br />

96π 2 F 2 0<br />

m 2 K,4 = m 2 K,2<br />

[<br />

1 + m2 η,2<br />

48π 2 F 2 0<br />

]<br />

log( m2 η,2<br />

µ 2 )<br />

[<br />

(2m + m s )B 0 (2L r 6 − L r 4) + 1 ]<br />

2 (m + m s)B 0 (2L r 8 − L r 5)<br />

[<br />

[<br />

1 + m2 K,2<br />

16π 2 F 2 0<br />

log( m2 K,2<br />

µ 2 ) − m2 η,2<br />

24π 2 F 2 0<br />

]<br />

log( m2 η,2<br />

µ 2 )<br />

(2m + m s )B 0 (2L r 6 − L r 4) + 8 m2 η<br />

F0<br />

2 (2L r 8 − L r 5)<br />

log( m2 η,2<br />

µ 2 ) − m2 π,2<br />

32π 2 F 2 0<br />

+ 128<br />

9F 2 0<br />

log( m2 π,2<br />

µ 2 ) + m2 K,2<br />

48π 2 F 2 0<br />

[<br />

(m − ms )B 2 0(3L r 7 + L r 8) ]<br />

]<br />

]<br />

log( m2 K,2<br />

µ 2 )<br />

These are the final formulae for the masses of the Goldstone bosons up to order<br />

ord(E 4 ) in terms of the masses of the Goldstone bosons up to order ord(E 2 ),<br />

where the l<strong>at</strong>ter ones are given in terms of unknown <strong>QCD</strong>-quark-masses (see<br />

above). One notices the fol<strong>low</strong>ing important fe<strong>at</strong>ures:<br />

The masses of the Goldstone bosons vanish if the quark masses go to zero,<br />

since then also the masses to ord(E 2 ) go to zero. This is gr<strong>at</strong>ifying, since in case<br />

of no symmetry breaking this is actually wh<strong>at</strong> should happen. We woud have a<br />

bad theory if in the limit of exact chiral symmetry the goldstone bosons would<br />

have mass. If one looks into the liter<strong>at</strong>ure it often happens th<strong>at</strong> this point is<br />

not done right.<br />

The expressions for the masses contain analytic terms ∝ m q and non-analytic<br />

terms ∝ m q log(m q ) . The l<strong>at</strong>ter ones are multiplied with the <strong>low</strong> energy coefficients<br />

and involve no new paramters. This illustr<strong>at</strong>es the general theorem of Li<br />

and Pagels th<strong>at</strong> a symmetry, which is realized in the Nambu-Goldstone mode,<br />

leads in perturb<strong>at</strong>ion theory to analytic as well as non-analytic trms.<br />

Looking to the above formulae it seems, th<strong>at</strong> the physical masses are dependent<br />

on the scale µ 2 (scale dependence). However the <strong>low</strong> energy coefficients L r i<br />

are also scale dependent, and their scale dependence is such th<strong>at</strong> it is exactly<br />

compens<strong>at</strong>ed by the scale dependence of the chiral logarithms. This is most<br />

easily shown by performing the deriv<strong>at</strong>ive of the above formulae with respect<br />

to µ , which turns out to vanish. Thus the physical observables are not scale<br />

dependent, as it should be.<br />

13 L<strong>at</strong>tice Gauge Theory<br />

13.1 Introduction<br />

202


We assume th<strong>at</strong> the <strong>QCD</strong> is the theory for strong interactions and th<strong>at</strong> mesons<br />

and baryons are mainly determined by this forces. In the previous chapters<br />

we have never considered this fact, except in studying symmetries, but looked<br />

always for phenomenological descriptions in terms of hadronic currents, quark<br />

degrees of freedom, etc. In this chapter we will try to solve <strong>QCD</strong> directly as<br />

a non-abelian gauge theory. The idea for this goes back to K. Wilson und<br />

Wegener. The idea is to formul<strong>at</strong>e <strong>QCD</strong> on a discrete space-time l<strong>at</strong>tice (1+3<br />

dimensions) written down in terms of altogether 4 Euklidean dimensions with<br />

l<strong>at</strong>tice constant a.<br />

Hereby the inform<strong>at</strong>ion on the quark fields will reside on littice sites (points)<br />

and the inform<strong>at</strong>ion on the gluon fields will reside on links between sites. The<br />

final formul<strong>at</strong>ion maintains exact local gauge invariance. There is a n<strong>at</strong>ural<br />

cut-off <strong>at</strong> p ∼ 1 a<br />

. Th<strong>at</strong> means things smaller than a cannot be described and<br />

momenta larger than the cut-off cannot be described either. The formul<strong>at</strong>ion is<br />

suitable for (r<strong>at</strong>her extensive!!) computer simul<strong>at</strong>ions in large supercomputers<br />

and require usually large collabor<strong>at</strong>ions. The idea is to calcul<strong>at</strong>e observables,<br />

in order to compare with experiment, or to provide benchmark results in accur<strong>at</strong>e<br />

calcul<strong>at</strong>ions which simpler models with properly chosen effective degrees<br />

of freedom have to reproduce. On the first view the l<strong>at</strong>tice <strong>QCD</strong> techniques<br />

sound very fundamental and exact. However in practice several shortcomings<br />

and technical problems are encountered (finite a, finite l<strong>at</strong>tice, too large current<br />

quark mass) such th<strong>at</strong> often the approach is by no means th<strong>at</strong> fundamental as<br />

it seems and often a comparison with experiment is difficult.<br />

13.2 Quantum Mechanics: Transition amplitudes and p<strong>at</strong>h<br />

integrals<br />

We consider 1-dimensional quantum mechanics as an explicit and didactic example.<br />

There the system is described in the Schrödinger picture by the wave<br />

function ψ(x,t) which corresponds to a time dependent st<strong>at</strong>e vector in a Hilbert<br />

203


space |ψ(t)〉. We know the st<strong>at</strong>e vector |x〉 which is eigenst<strong>at</strong>e of the coordin<strong>at</strong>e<br />

oper<strong>at</strong>or ˆx |x〉 = x |x〉 and similarly the st<strong>at</strong>e vector |p〉 which is eigenst<strong>at</strong>e of<br />

the momentum oper<strong>at</strong>or ˆp |p〉 = p |p〉. The oper<strong>at</strong>ors fulfill the commut<strong>at</strong>ion rule<br />

[ˆx, ˆp] = iħ and since ˆx and ˆp are hermition we have the completeness rel<strong>at</strong>ions<br />

∫<br />

1 = dx|x >< x|<br />

∫<br />

1 = dp|p >< p|<br />

. The wave functions are defined as ψ(x,t) =< x |ψ(t)〉 and ψ(p,t) =< p |ψ(t)〉<br />

.The Fourier transform rel<strong>at</strong>es both:<br />

∫<br />

< x |ψ(t)〉 = dp < x |p〉 < p |ψ(t)〉<br />

with<br />

< x |p〉 = 1 √<br />

2πħ<br />

e ipx/ħ (213)<br />

The wavefunctions are described by the Schrödinger-eq. with Hamiltonian H<br />

and the time evolution is given by (if the H is time-independent)<br />

|ψ(t)〉 = e −iHt/ħ |ψ(0)〉<br />

The fact th<strong>at</strong> we assume H to be time independent is no limit<strong>at</strong>ion. We<br />

will chose l<strong>at</strong>eron small time steps such th<strong>at</strong> <strong>at</strong> each time intervall the H can<br />

be considered time independen. For the formul<strong>at</strong>ion of quantum mechanics<br />

and field theory on the l<strong>at</strong>tice the concept of a propag<strong>at</strong>or or the correl<strong>at</strong>or is<br />

important. On gets in a n<strong>at</strong>ural way like this: We have<br />

< x f |ψ(t f )〉 =< x f |exp [−iHt f /ħ] |ψ(0)〉<br />

=< x f |exp [−iHt f /ħ] exp[+iHt i /ħ] |ψ(t i )〉<br />

and inserting the completeness rel<strong>at</strong>ion we get<br />

∫<br />

< x f |ψ(t f )〉 = dx i < x f |exp [−iH(t f − t i )/ħ] |x i 〉 〈x i |ψ(t i )<br />

or<br />

∫<br />

ψ(x f ,t f ) =<br />

dx i K( x f , t f ,x i ,t i )ψ(x i ,t i )<br />

with the propag<strong>at</strong>or K given by<br />

K( x f , t f ,x i ,t i ) =< x f |exp [−iH(t f − t i )/ħ] |x i 〉 =< x f (t f ) |x i (t i )〉<br />

204


13.2.1 Free Motion (Propag<strong>at</strong>or)<br />

In this subsection we calcul<strong>at</strong>e te propag<strong>at</strong>or of a free particle with H = p2<br />

2m<br />

and we will introduce a formul<strong>at</strong>ion with discrete space and time points. We<br />

have:<br />

∫<br />

]<br />

< x f |exp [−iH(t f − t i )/ħ] |x i 〉 = dp < x f |p〉 exp<br />

[−i p2<br />

2mħ (t f − t i ) < p |x i 〉<br />

We can insert eq.(213) and perform the integr<strong>at</strong>ion. For this we use the well<br />

known formulae<br />

∫ +∞<br />

√ ∫ π +∞<br />

√ π<br />

dx exp(−λx 2 ) = or dx exp(−λx 2 +2λx¯x) =<br />

λ<br />

λ exp( λ¯x 2)<br />

−∞<br />

−∞<br />

however, we use analytical continu<strong>at</strong>ion to get rid of the i , integr<strong>at</strong>e and continue<br />

analytically back. Then we obtain (H for free motion)<br />

( ) [ ]<br />

1/2<br />

m<br />

im (x f − x i ) 2<br />

< x f |exp [−iH(t f − t i )/ħ] |x i 〉 =<br />

exp<br />

2πiħ(t f − t i ) 2ħ (t f − t i )<br />

(214)<br />

The exponenet on RHS can be written as<br />

exp<br />

[<br />

im<br />

2ħ<br />

(x f − x i ) 2<br />

(t f − t i )<br />

]<br />

[ ]<br />

i 1<br />

= exp<br />

ħ 2 mv2 (t f − t i ) = i ∫ tf<br />

Ldt = i ħ t i<br />

ħ S cl(x f t f ;x i t i )<br />

Here is S cl the action (classical, i.e. no oper<strong>at</strong>ors) between the space-time points.<br />

Hence in this example of free motion of a massive point the propag<strong>at</strong>or is known.<br />

To get closer to the formul<strong>at</strong>ion on a l<strong>at</strong>tice we rewrite the above expression<br />

in a trivial way. For this we introduce an arbitrary intermedi<strong>at</strong>e space-time<br />

point x 1 ,t 1 with t f > t 1 > t i :<br />

∫<br />

ψ(x f ,t f ) = dx 1 K( x f , t f ;x 1 ,t 1 )ψ(x 1 ,t 1 )<br />

∫<br />

ψ(x 1 ,t 1 ) = dx i K( x 1 , t 1 ;x i ,t i )ψ(x i ,t i )<br />

and with th<strong>at</strong> we have<br />

∫<br />

K( x f , t f ;x i ,t i ) =<br />

dx 1 K( x f , t f ;x 1 ,t 1 )K( x 1 , t 1 ;x i ,t i )<br />

where the integrand is proportional to<br />

]<br />

i<br />

∝ exp[<br />

ħ (S cl(x f t f ;x 1 t 1 ) + S cl (x 1 t 1 ;x i t i ))<br />

We introduced one intermedi<strong>at</strong>e point in the time, which resulted in an integral<br />

over the x-coordin<strong>at</strong>e <strong>at</strong> th<strong>at</strong> time. We will introduce now N intermedi<strong>at</strong>e<br />

205


points, with N large, and will have <strong>at</strong> each time-point the integral over the<br />

x-coordin<strong>at</strong>e:<br />

ε = t f − t i<br />

N<br />

t j = t i + jε<br />

j = 0,1,...,N<br />

Then we have for large N a small ε and can write down explicitely<br />

[ ]<br />

< x f |exp −iĤ(t f − t i )/ħ |x i 〉 =<br />

[<br />

< x f |exp f − t N−1 )/ħ<br />

] [<br />

exp N−1 − t N−2 )/ħ<br />

∫<br />

]<br />

= dx 1 dx 2 ....dx N−1 < x f |exp<br />

... < x 2 |exp<br />

[−iĤε/ħ ] [−iĤε/ħ ]<br />

|x 1 〉 < x 1 |exp |x i 〉<br />

]<br />

|x N−1 〉 < x N−1 |exp<br />

[ ]<br />

...exp −iĤ(t 1 − t i )/ħ |x i 〉<br />

[−iĤε/ħ<br />

]<br />

|x N−2 〉 ...<br />

These are N factors and N − 1 integr<strong>at</strong>ions. We cannot put them together<br />

without a little thinking. We have (not for our free motion but) in general the<br />

Hamiltonian written in terms of the coordin<strong>at</strong>e and momentum oper<strong>at</strong>ors<br />

and we know th<strong>at</strong><br />

Ĥ = ˆp2<br />

2m + V (ˆx)<br />

eÂe<br />

ˆB =<br />

eÂ+<br />

ˆB+ 1<br />

2[Â, ˆB]<br />

The  and ˆB are ∝ ε, whereas the commut<strong>at</strong>or is in our example ∝ ε 2 . Thus<br />

for sufficiently small ε we can neglect the commut<strong>at</strong>or. In this case (wh<strong>at</strong> must<br />

206


always be guaranteed by proper l<strong>at</strong>tice techniques) we have<br />

[−iĤε/ħ<br />

]<br />

exp ≈ exp [ −iεˆp 2 /2mħ ] exp[−iεV (ˆx)/ħ]<br />

and hence<br />

< x k+1 |exp<br />

[−iĤε/ħ ]<br />

|x k 〉 ≈< x k+1 |exp [ −iεˆp 2 /2mħ ] [<br />

|x k 〉 exp −iεV ( 1 ]<br />

2 (x k+1 + x k )/ħ<br />

The first term on RHS is known from our example of free motion eq.(214) and the<br />

second term is a number. So we get altogether with ε = t f −t i<br />

N<br />

= t k+1 − t k = ∆t<br />

the expressiion<br />

[−iĤε/ħ [<br />

] ( m<br />

) ( ) 1/2 2<br />

imε xk+1 − x k<br />

< x k+1 |exp |x k 〉 = exp − iεV (x k ) /ħ]<br />

2πiħε 2ħ ε<br />

(215)<br />

and with x 0 = x i and x N = x f we can write:<br />

[ ]<br />

< x f |exp −iĤ(t f − t i )/ħ |x i 〉 = lim<br />

(216)<br />

N→∞<br />

⎡<br />

∫<br />

{<br />

D (N) x exp ⎣ i N−1<br />

ħ ε ∑<br />

( ) } ⎤ 2<br />

m xj+1 − x j<br />

− V (x j ) ⎦<br />

2 ε<br />

j=0<br />

where we have abbrevi<strong>at</strong>ed as (we recognize N factors and N − 1 integr<strong>at</strong>ions)<br />

(<br />

) N/2<br />

D (N) m<br />

x =<br />

dx 1 dx 2 ....dx N−1 (217)<br />

2πiħ(t f − t i )/N<br />

For vanishing potential we discover in the exponent the Lagrangean of the free<br />

motion. If the potential is time dependent one has to take the time corresponding<br />

to the coordin<strong>at</strong>e, i.e. V (x j ) → V (x j ,t j ). The expression looks formally<br />

difficult, however, it is easier as a working prescription: One cuts the time intervall<br />

(t f − t i ) into many equidistant time subintervalls, <strong>at</strong> each intermedi<strong>at</strong>e<br />

time t k one integr<strong>at</strong>es along the x k −axis and multiplies with a weight according<br />

to the exponent in eq.(216).<br />

One can view this much simpler, and this is the view in practice: Suppose<br />

you perform all these integr<strong>at</strong>ions simultaneously and let the x j run from −∞<br />

to +∞, then do <strong>at</strong> a certain moment a snapshot. At this moment the snapshot<br />

shows you a set of coordintes ¯x 1 ,...¯x N−1 between x i (t i ) and x f (t f ) which appears<br />

to be in the looking glas a certain zick-zack-p<strong>at</strong>h ¯x(t). Since ε = ∆t the<br />

sum in the exponent appears to be equal to i ħ ∆t ∑ { ( ) 2<br />

N−1 m d¯xj<br />

j=0 2 dt − V (¯xj )}<br />

=<br />

∫<br />

i<br />

ħ dtL(¯x(t)) =<br />

i<br />

ħS [¯x(t)] . Thus the action along this particular zick-zack-p<strong>at</strong>h<br />

has to be calcul<strong>at</strong>ed. See the figure for th<strong>at</strong>:<br />

The multitude of snapshots yields in the end all possible zick-zack-p<strong>at</strong>hs<br />

between x i (t i ) and x f (t f ). Thus, because of the integrals between -∞ and +∞<br />

207


one has to sum over all these p<strong>at</strong>hs giving each p<strong>at</strong>h of them in this huge sum a<br />

weight exp ( i<br />

ħ S[x]) . This sum over all possible zick-zack-p<strong>at</strong>hs is called ∫ t f<br />

t i<br />

Dx<br />

. The result is the propag<strong>at</strong>or, which in fact contains the full inform<strong>at</strong>ion about<br />

the system.<br />

Hence we can write down finally<br />

∫ tf<br />

[ ] i<br />

< x f |exp [−iH(t f − t i )/ħ] |x i 〉 = lim D (N) xexp<br />

N→∞ t i<br />

ħ S [x] (218)<br />

where x(t) is the p<strong>at</strong>h for t moving from t i −→ t f with<br />

x(t i ) = x i<br />

x(t f ) = x f<br />

and the classical action (no oper<strong>at</strong>ors)<br />

S [x] =<br />

∫ tf<br />

t i<br />

dtL(x,ẋ) =<br />

∫ tf<br />

t i<br />

[ 1<br />

dt<br />

2 mẋ(t)2 − V (x(t))]<br />

Presently it does not look very efficient for studying a quantum mechanical<br />

system. However, when we consider quantized field theories, we will see, th<strong>at</strong><br />

there are tremendous advantages of this formalism over a standard canonical<br />

one. In fact, for non-abelian gauge theories as e.g. the <strong>QCD</strong> the p<strong>at</strong>h integral<br />

formalism is the only one, which yields a quantized field theory and is<br />

managable.<br />

As we have seen the p<strong>at</strong>h integrals give us simple transition amplitudes<br />

208


[which are often written as < x f (t f ) |x i (t i )〉]<br />

< x f |exp [−iH(t f − t i )/ħ] |x i 〉 =<br />

∫ xf<br />

[ ∫ i<br />

tf<br />

]<br />

= lim D (N) xexp dtL[x,ẋ]<br />

N→∞ x i<br />

ħ<br />

This important result generalizes to more complic<strong>at</strong>ed amplitudes (without<br />

proof), from which one can altogether extract inform<strong>at</strong>ion about our system.<br />

We have e.g. for t i < t 1 < t 2 < t f the fol<strong>low</strong>ing identity<br />

∫ xf<br />

[ ] i<br />

< x f (t f )|x(t 2 )x(t 1 ) |x i (t i )〉 = lim D (N) xx(t 2 )x(t 1 )exp<br />

N→∞<br />

ħ S [x] (219)<br />

13.2.2 Propag<strong>at</strong>ors ( Harmonic oscill<strong>at</strong>or)<br />

x i<br />

We will calcul<strong>at</strong>e another example, namely the harmonic oscill<strong>at</strong>or, which is<br />

less trivial, however one can learn more. We will learn th<strong>at</strong> one can discretize<br />

the time and will nevertheless get the proper continum limit, and th<strong>at</strong> one can<br />

extract directly observables and not only abstract quantities like propag<strong>at</strong>ors.<br />

We will not do it any more in such a technical detail. There the action for the<br />

one-dimensional harmonic oscill<strong>at</strong>or is<br />

∫ tf<br />

[ 1<br />

S[x] = dt<br />

2 mẋ(t)2 − 1 ]<br />

2 mω2 x(t) 2<br />

t i<br />

In order to solve the p<strong>at</strong>h integral (218) it is advisable to select from all possible<br />

p<strong>at</strong>hs in the p<strong>at</strong>hintegral only the classical p<strong>at</strong>h, i.e. the p<strong>at</strong>h which is the result<br />

of solving the classical equ<strong>at</strong>ions of motion, which in turn arise from demanding<br />

δS = 0. For this particular p<strong>at</strong>h we have ẍ(t) + ω 2 x(t) = 0 . If we rewrite the<br />

above S[x] by partial integr<strong>at</strong>ion, i.e. using ∫ dtẋ 2 = xẋ − ∫ dtẍx, and if we<br />

replace ẍ by ẍ(t) = −ω 2 x(t) we have simply<br />

t i<br />

S[x] = 1 2 mx(t)ẋ(t)|t f<br />

ti<br />

Since the classical p<strong>at</strong>h is known, namely<br />

x cl (t) = x f sin[ω(t − t i )] − x i sin[ω(t − t f )]<br />

sin(ωT)<br />

we obtain easily<br />

S[x cl ] =<br />

mω [(<br />

x<br />

2<br />

2sin(ωt) i + x 2 ]<br />

f)<br />

cos(ωT) − 2xi x f<br />

with T = t f − t i<br />

To calcul<strong>at</strong>e the full p<strong>at</strong>h integral it is convenient to write the p<strong>at</strong>hs by devi<strong>at</strong>ion<br />

from the classical p<strong>at</strong>h: x(t) = x cl (t) + y(t) with y(t i ) = y(t f ) = 0. Thus we<br />

obtain<br />

∫ tf<br />

[ 1<br />

S[x] = S[x cl ] + dt<br />

t i<br />

2 mẏ(t)2 − 1 ]<br />

2 mω2 y(t) 2<br />

209


and therefore (please have a look <strong>at</strong> the limits of the integrals)<br />

∫ xf<br />

x i<br />

[ ] [ ] ∫ i i 0<br />

[ ] i<br />

Dxexp<br />

ħ S[x] = exp<br />

ħ S[x cl] Dy exp<br />

0 ħ S[y]<br />

We can discretise now in the well known way fol<strong>low</strong>ing eq.(216) way modifying<br />

slightly the potential part in a justified way for small ε = (t f − t i )/N = T/N:<br />

∫ 0<br />

0<br />

[ ] i<br />

( m<br />

Dy exp<br />

ħ S[y] = lim<br />

N→∞ 2πiħε<br />

⎡<br />

∫ N−1 ∏<br />

dy k exp<br />

k=1<br />

) N/2<br />

⎣ 1 N−1<br />

ħ ε ∑<br />

j=0<br />

{<br />

m<br />

2<br />

( ) } ⎤ 2 yj+1 − y j<br />

− 1 1 2 mω2 2 (y2 j + yj+1) 2 j<br />

⎦<br />

where y 0 = y N = 0. (Remember: N factors and N − 1 integr<strong>at</strong>ions). This<br />

can be written in a vector and m<strong>at</strong>rix form with y = (y 1 ,y 2 ,...,y N−1 ) and<br />

M = (N − 1) · (N − 1)-m<strong>at</strong>rix as<br />

lim<br />

N→∞<br />

∫ 0<br />

0<br />

⎡<br />

⎤<br />

[ ] i<br />

( m<br />

) N/2<br />

∫ N−1 ∏<br />

N−1<br />

D (N) y exp<br />

ħ S[y] ∑<br />

= lim<br />

dy k exp ⎣<br />

−m<br />

N→∞ 2πiħε<br />

2iħε yT My⎦<br />

k=1 j=0<br />

Explicitely we have<br />

⎛<br />

⎞<br />

2 − ε 2 ω 2 −1 0 0 ... 0<br />

−1 2 − ε 2 ω 2 −1 0 ... 0<br />

0 −1 2 − ε 2 ω 2 −1 ... 0<br />

M =<br />

0 −1 2 − ε 2 ω 2 −1 ... 0<br />

⎜ ... ... ... ... ... ...<br />

⎟<br />

⎝ 0 0 0 −1 2 − ε 2 ω 2 −1 ⎠<br />

0 0 0 0 −1 2 − ε 2 ω 2<br />

with M being real and symmetric, i.e. we can find a m<strong>at</strong>rix A which diagonalizes<br />

M:<br />

from which fol<strong>low</strong>s<br />

M = ADA −1 where A T A = 1 and det(A) = 1<br />

exp [ −αy T My ] = exp [ −αy T ADA −1 y ] = exp [ −αu T Du ]<br />

This is simple since D is diagonal with eigenvalues d i . So we√ have Gaussian<br />

integrals over u yielding (if α were real...) a product of terms π<br />

αd i<br />

. Thus we<br />

get<br />

lim<br />

N→∞<br />

∫ 0<br />

0<br />

[ ] √<br />

i<br />

D (N) y exp<br />

ħ S[y] = lim<br />

N→∞<br />

ε<br />

m<br />

2πiħεdet(M)<br />

210


Actually the calcul<strong>at</strong>ion of det(M) is not simple, one gets (without proof) with<br />

the abbrevi<strong>at</strong>ion δ = ωε the recursion<br />

and hcnce<br />

det(M) nxn = (2 − δ 2 )det(M) (n−1)x(n−1) − det(M) (n−2)x(n−2)<br />

det(M) 1x1 = 2 − δ 2<br />

det(M) 2x2 = 3 − 4δ 2 + δ 4<br />

...<br />

det(M) nxn = (n + 1) − δ2<br />

6<br />

n(n + 1)(n + 2) +<br />

δ4<br />

120 (n − 1)n(n + 1)(n + 2)(n + 3) + O(δ6 )<br />

In the end we have to go with N to very large values. We use this to simplify the<br />

above expression collecting only those terms which are leading in N,compared<br />

to which all subleading terms are negligeable:<br />

det(M) NxN = N − δ2<br />

6 N3 + δ4<br />

120 N5<br />

We recall now th<strong>at</strong> δ = ωε and rewrite this leading term using δN = ωεN = ωT<br />

)<br />

det(M) (N−1)x(N−1) = N<br />

(1 − (ωT)2 + (ωT)4<br />

6 120 + O((ωT)6 )<br />

= N sin(ωT)<br />

ωT<br />

Thus we obtain finally (with N = T/ε ) for the p<strong>at</strong>h integral over the fluctu<strong>at</strong>ions<br />

around the classical p<strong>at</strong>h of the harmonic oscill<strong>at</strong>or<br />

∫ 0<br />

[ ] √ i<br />

Dy exp<br />

ħ S[y] mω<br />

=<br />

2πiħsin(ωT)<br />

0<br />

and thus finally<br />

∫ f<br />

[ ] i<br />

K(x f ,T;x i 0) = lim D (N) xexp<br />

N→∞ i<br />

ħ S[x] = (220)<br />

√<br />

[<br />

mω<br />

=<br />

2πiħsin(ωT) exp imω [(<br />

x<br />

2<br />

2ħsin(ωt) i + x 2 ) ] ]<br />

f cos(ωT) − 2xi x f<br />

This result is exact and extremely didactical. We have obtained it by discretizing<br />

space-time and considered its continuum limit (by using the terms leading in<br />

N). In the present approach it was simple to do so, however when we consider<br />

<strong>QCD</strong> the continuum limit will be a problem.<br />

We remember th<strong>at</strong> our p<strong>at</strong>h integral calcul<strong>at</strong>es the propag<strong>at</strong>or K(x f t f ;x i t i ).<br />

We can check our result by applying it e.g. to a wave function of Gaussian type<br />

<strong>at</strong> t=0:<br />

1<br />

ψ(x i ,0) = √√ exp<br />

[− (x i − ¯x) 2 ]<br />

2πσ 4σ 2<br />

211


yielding<br />

ψ(x f ,T) =<br />

∫ +∞<br />

−∞<br />

dx i K(x f T;x i 0)ψ(x i ,0)<br />

Inserting our p<strong>at</strong>h integral eq.( 220) for the propag<strong>at</strong>or we obtain again a gaussian,<br />

which writes (apart of a phase).<br />

[<br />

1<br />

ψ(x f ,T) = √√ exp − (x i − ¯x cos(ωT) 2 ]<br />

2πσ<br />

′ 4σ ′2<br />

with<br />

σ ′ = σ<br />

√<br />

cos 2 (ωT) + ħ2<br />

2σ 2 sin2 (ωT)<br />

One can easily check, th<strong>at</strong> this ψ(x f ,T) s<strong>at</strong>isfies indeed the Schroedinger equ<strong>at</strong>ion<br />

and is hence a true time evolved st<strong>at</strong>e from our initial gaussian packet.<br />

13.2.3 Extraction of inform<strong>at</strong>ion, Euklidean time<br />

By now we had integr<strong>at</strong>ions with imaginary exponents, which are highly oscill<strong>at</strong>ing.<br />

They are not appropi<strong>at</strong>e for a numerical tre<strong>at</strong>ment. To achieve this we<br />

go to Euclidean time. This not only makes life easier we also will show th<strong>at</strong><br />

one can extract directly physical inform<strong>at</strong>ion from the Euklidean correl<strong>at</strong>ion<br />

functions. In fact there is some inform<strong>at</strong>ion, we can extract from the Euklidean<br />

formul<strong>at</strong>ion (wave functions, transition elements, certain m<strong>at</strong>rixelements), there<br />

is another type of inform<strong>at</strong>ion (processes, cross sections) which require formul<strong>at</strong>ion<br />

in Minkowski space and hence are not accessible by present time techniques.<br />

Euklidean time is defined as<br />

Then we have<br />

t = −iτ (τ > 0)<br />

〈x f |exp[−iH(t f − t i )/ħ] |x i 〉 −→ 〈x f |exp [−H(τ f − τ i )/ħ] |x i 〉 (221)<br />

The transition from Minkowski to Euklidean space is easy, take as example the<br />

propagtor of free motion eq.(214), which reads in Euklidean coordin<strong>at</strong>ew<br />

( ) [ ]<br />

1/2<br />

m<br />

im (x f − x i ) 2<br />

exp<br />

2πiħ(t f − t i ) 2ħ (t f − t i )<br />

( ) [<br />

]<br />

1/2<br />

m<br />

→<br />

exp − m (x f − x i ) 2<br />

2πħ(τ f − τ i ) 2ħ (τ f − τ i )<br />

and the exponent is equal to − 1 ħ S E(x f ,τ f ;x i ,τ i ). If we have in general a Hamiltonian<br />

with kinetic and potential terms ge get<br />

∫ f<br />

[<br />

〈x f |exp [−H(τ f − τ i )]/ħ |x i 〉 = Dxexp − 1 ]<br />

ħ S E[x]<br />

212<br />

i


with<br />

S E [x] =<br />

∫ τf<br />

τ i<br />

[ 1<br />

dτ<br />

2 mẋ(τ)2 + V (x)]<br />

In order to see the equivalence of a continuum formul<strong>at</strong>ion and a l<strong>at</strong>tice formul<strong>at</strong>ion<br />

we will compare now the RHS with the LHS of the fol<strong>low</strong>ing formula,<br />

where the LHS is calcul<strong>at</strong>ed analytically (continuum) and the RHS numerically<br />

(l<strong>at</strong>tice):<br />

∫ f<br />

[<br />

〈x f |exp [−H(τ f − τ i )/ħ] |x i 〉 | analyt = lim D (N) xexp − 1 ]<br />

N→∞ i<br />

ħ S E[x] | numerical<br />

(222)<br />

We take now a special case, <strong>at</strong> which we can see, how to extract physics out<br />

of the above equ<strong>at</strong>ions. We have with the choice x = x f = x i and T = τ f − τ i<br />

the fol<strong>low</strong>ing formula after inserting a the complete set of eigenst<strong>at</strong>es of the<br />

Hamiltonian H:<br />

LHS = 〈x|exp [−HT/ħ] |x〉 = ∑ [<br />

< x|n > exp − E ]<br />

nT<br />

< n|x ><br />

ħ<br />

n,m<br />

If we integr<strong>at</strong>e over all x we obtain the famous partition function Z, which we<br />

know from st<strong>at</strong>istical mechanics:<br />

∫<br />

Z = dx 〈x|exp[−HT]/ħ |x〉 = ∑ [<br />

exp − E ]<br />

nT<br />

ħ<br />

n<br />

In order to obtain explicit inform<strong>at</strong>ion about the system take now the limit<br />

T → ∞.This makes all excited st<strong>at</strong>es contribute less and less with increasing T<br />

to the partition function than the ground st<strong>at</strong>e. So in the end <strong>at</strong> T → ∞ only<br />

the ground st<strong>at</strong>e remains in the sum:<br />

〈x|exp[−HT]/ħ |x〉 −→ exp[−E 0 T/ħ] | < x|0 > | 2 (223)<br />

Thus we can extract immedi<strong>at</strong>ely the energy of the ground st<strong>at</strong>e E 0 and the<br />

wave function of the ground st<strong>at</strong>e < x|0 > from the Euklidean form of the<br />

correl<strong>at</strong>ion function.<br />

So far the LHS, i.e. the analytic tre<strong>at</strong>ment of eq.(222). The numerical<br />

tre<strong>at</strong>ment (RHS)goes in a well known way, τ j = τ i + ja with j = 0,1,...,N and<br />

a = T N<br />

. Then we have with eq.(217)<br />

∫<br />

D (N) x −→<br />

( m<br />

) N/2<br />

∫ +∞<br />

2πħa<br />

−∞<br />

...<br />

∫ +∞<br />

−∞<br />

dx 1 dx 2 ...dx N−1<br />

These are N factors and an N − 1 dimensional integr<strong>at</strong>ion, where x j = x(τ j )<br />

and with the choice x = x i = x f . In each intervall we have to perform the<br />

213


integr<strong>at</strong>ion over the Euklidean Lagrangean L E = −L in the exponential, in<br />

order to yield the action, i.e.<br />

∫ τj+1<br />

∫ τj+1<br />

[ 1<br />

dτL E = dτ<br />

τ j τ j<br />

2 mẋ(τ)2 + V (x)]<br />

≈<br />

[ ( ) 2<br />

1<br />

a<br />

2 m xj+1 − x j<br />

+ 1 a 2 (V (x j+1) + V (x j ))]<br />

With this we obtain the Euklidean action of the l<strong>at</strong>tice<br />

S l<strong>at</strong>t [x] = − 1 ħ<br />

N−1<br />

∑<br />

j=0<br />

and hence we obtain for the RHS of eq.(222)<br />

RHS =<br />

∫ f<br />

i<br />

⎡<br />

exp ⎣− 1 ħ<br />

[ m<br />

2a (x j+1 − x j ) + aV (x j )]<br />

[<br />

Dxexp − 1 ] ( m<br />

) N/2<br />

∫ +∞<br />

ħ S E[x] =<br />

2πħa<br />

N−1<br />

∑<br />

j=0<br />

[ m<br />

] ⎤<br />

2a (x j+1 − x j ) + aV (x j ) ⎦<br />

−∞<br />

...<br />

∫ +∞<br />

−∞<br />

dx 1 dx 2 ...dx N−1<br />

Actually we have integrals to solve with r<strong>at</strong>her simple integrands, but it is a multiple<br />

integral. There are sophistic<strong>at</strong>ed techniques to do multiple integrals, which<br />

we will not discuss now. In fact these techniques correspond as well to select<br />

all possible p<strong>at</strong>hs beween the fixed end points x i and x f and to integr<strong>at</strong>e over<br />

these p<strong>at</strong>hs weighted with the exponent of the corresponding classical action.<br />

One might worry about approxim<strong>at</strong>ing ẋ with (x j+1 − x j )/a in our formula<br />

for the l<strong>at</strong>tice action S l<strong>at</strong>t [x]. It is not obvious th<strong>at</strong> this is a good approxim<strong>at</strong>ion<br />

given th<strong>at</strong> x j+1 − x j can be arbitrarily large in our p<strong>at</strong>h integral; th<strong>at</strong> is,<br />

p<strong>at</strong>hs can be arbitrarily rough. While not so important for our one-dimensional<br />

problem, this becomes a crucial issue for four-dimensional field theories. It is<br />

dealt with using renormaliz<strong>at</strong>ion theory, which we discuss in l<strong>at</strong>er sections.<br />

If we consider x i = x f = x, and integr<strong>at</strong>e over x then we have a formula<br />

for the partition function, when we slightly change the definition of (217) from<br />

a (N − 1)- dimensional integral to a N-dimensional one, i.e. N factors and<br />

an N-dimensional integr<strong>at</strong>ions. Thus we obtain with this new definition of Dx<br />

without the index N: ∫ [<br />

Z = Dxexp − 1 ]<br />

ħ S E[x]<br />

13.2.4 Correl<strong>at</strong>ors<br />

With the example of the harmonic oscill<strong>at</strong>or we show how one can calcul<strong>at</strong>e<br />

correl<strong>at</strong>ors, which are a bit more complic<strong>at</strong>ed than propag<strong>at</strong>ors, however, each<br />

of them al<strong>low</strong>s to extract other inform<strong>at</strong>ion from the system. By now we have<br />

extracted from the propag<strong>at</strong>or the ground st<strong>at</strong>e energy and the ground st<strong>at</strong>e<br />

214


wave function. Now we consider (we put ħ = 1) the fol<strong>low</strong>ing correl<strong>at</strong>or as<br />

example (1-particle Euklidean Greens function)<br />

N >= 〈x f |exp [−Hτ f ]ˆx(τ ′ )exp[+Hτ i ] |x i 〉<br />

with a normaliz<strong>at</strong>ion factor N to be defined l<strong>at</strong>er. In the Heisenberg-picture we<br />

have<br />

ˆx(τ ′ ) = exp[Hτ ′ ]ˆx exp[−Hτ ′ ]<br />

and hence<br />

N >= 〈x f |exp[−H(τ f − τ ′ )]ˆx exp[−H(τ ′ − τ i )] |x i 〉<br />

∫<br />

= dy 〈x f |exp [−H(τ f − τ ′ )]ˆx|y >< y|exp[−H(τ ′ − τ i )] |x i 〉<br />

∫<br />

= dy 〈x f |exp[−H(τ f − τ ′ )]|y > y < y|exp[−H(τ ′ − τ i )] |x i 〉<br />

Under the integral we have the product of two correl<strong>at</strong>ors, each of the being<br />

represented by a p<strong>at</strong>h integral. Thus we obtain<br />

∫ {∫<br />

} {∫<br />

}<br />

N >= dy Dxexp[−S y−→x f<br />

E<br />

] y Dxexp[−S xi−→y<br />

E<br />

]<br />

We can rewrite this in a familiar way, since in fact we have just added to all the<br />

discretized p<strong>at</strong>hs another discrete point <strong>at</strong> the time τ ′ . Thus we obtain<br />

∫<br />

N >= Dxx(τ ′ )exp[−S xi−→x f<br />

E<br />

]<br />

In full analogy we obtain an expression for 2-particle Greens function. We have<br />

for τ 2 > τ 1<br />

∫<br />

〈x f |exp [−Hτ f ]ˆx(τ 2 )ˆx(τ 1 )exp[+Hτ i ] |x i 〉 = Dxx(τ 2 )x(τ 1 )exp[−S xi−→x f<br />

E<br />

] for τ 2 > τ 1<br />

In order to extract inform<strong>at</strong>ion about the system we proceed as in the case of<br />

the propag<strong>at</strong>or. There we considered the limit of large time separ<strong>at</strong>ions. To do<br />

this we put x i = x f = x and define accordingly<br />

∫<br />

Dxx(τ2 )x(τ 1 )exp[−S E [x]]<br />

>= ∫<br />

Dxexp[−SE [x]]<br />

where we have integr<strong>at</strong>ed also over x. This defines also the above normaliz<strong>at</strong>ion<br />

factor<br />

∫<br />

N = Dxexp[−S E [x]]<br />

More in general we define<br />

>=<br />

∫<br />

DxΓ[x]exp[−SE [x]]<br />

∫<br />

Dxexp[−SE [x]]<br />

215


These correl<strong>at</strong>ors help to extract inform<strong>at</strong>ion on the system from the p<strong>at</strong>h<br />

integrals. To see this we consider the complete set of eigenst<strong>at</strong>es of the Hamiltonian<br />

(see e.g. harmonic oscill<strong>at</strong>or) H|n >= E n |n > . Then we can insert<br />

∫ completeness rel<strong>at</strong>ions into the expression for the correl<strong>at</strong>or yielding with<br />

dx < n|x >< x|m >= δnm e.g.:<br />

∫<br />

dx 〈x|exp [−H(τ f − τ ′ )]ˆx exp[−H(τ ′ − τ i )] |x〉<br />

∫<br />

= dx 〈x|n > exp [−E n (τ f − τ ′ )] < n|ˆx|m > exp[−E m (τ ′ − τ i )] < m|x ><br />

= ∑ n<br />

exp[−E n (τ f − τ i )] < n|ˆx|n ><br />

For symmetric potentials, like e.g. the harmonic oscill<strong>at</strong>or, this expression is<br />

not very interesting since due to parity reasons it vanishes.<br />

However we obtain a very relevent expression if we take the 2-particle Greens<br />

function:<br />

∫<br />

G E (τ 2 ,τ 1 ) = dx 〈x|exp [−H(τ f − τ 2 )]ˆx exp[−H(τ 2 − τ 1 )] ˆxexp[−H(τ 1 − τ i )] |x〉<br />

= ∑ n<br />

exp[−E n T < n|ˆx exp[−(H − E n )t]ˆx|n > with T = τ f − τ i and t = x 2 − x 1<br />

oder<br />

∑<br />

n<br />

>=<br />

exp[−E nT] < n|ˆx exp[−(H − E n )t]ˆx|n ><br />

∑<br />

n exp[−E nT]<br />

Now we can consider the limit for large T. The leading term is the one which<br />

survives. It is<br />

lim >= exp[−E 0T] < 0|ˆx exp[−(H − E 0 )t]ˆx|0 ><br />

T −→∞ exp[−E 0 T]<br />

There is some cancell<strong>at</strong>ion and we can write after inserting the completeness<br />

rel<strong>at</strong>ion<br />

lim<br />

T −→∞ >= ∑ m<br />

exp[−(E m − E 0 )t < 0|ˆx|m >< m|ˆx|0 ><br />

When we take also t −→ ∞, but always with T ≫ t. In this particular limit we<br />

obtain<br />

>= exp[−(E 1 − E 0 )t] |< 0|ˆx|1 >| 2 (224)<br />

Actually from this expression one can extract immedi<strong>at</strong>ely the energy of the<br />

first excited st<strong>at</strong>e. Denote for a moment the LHS as G(t). Then one obtains<br />

immedi<strong>at</strong>ely<br />

[ ] G(t)<br />

log<br />

≈ a(E 1 − E 0 ) (225)<br />

G(t + a)<br />

216


One can also obtain inform<strong>at</strong>ion on the transition m<strong>at</strong>rix element < 0|ˆx|1 >, if<br />

one inserts the energy difference into the correl<strong>at</strong>or <strong>at</strong> large t.<br />

One can redo the analysis for many different correl<strong>at</strong>ors als there are >, or >, or >.<br />

All these correl<strong>at</strong>ors contain some specific inform<strong>at</strong>ion on the system. We should<br />

remind, th<strong>at</strong> this physical inform<strong>at</strong>ion origin<strong>at</strong>es from the Euklidean form of the<br />

correl<strong>at</strong>ors.<br />

13.2.5 Partition function<br />

We know from the 1-dimensional example the partition function<br />

Z = ∑ ∫ [<br />

exp (−E n T) = Dx(τ) exp − 1 ]<br />

ħ ST E[x]<br />

n<br />

(226)<br />

with<br />

S T E [x] =<br />

∫ T<br />

0<br />

dτ<br />

[<br />

m<br />

2<br />

( ) 2 dx<br />

+ V (x(τ))]<br />

dτ<br />

This expression can be used to illustr<strong>at</strong>e several important things.<br />

First, there is some similarity to st<strong>at</strong>istical mechanics. There one implies<br />

1<br />

th<strong>at</strong> T =<br />

kT T emp<br />

= β Temp , where T Temp is the temper<strong>at</strong>ure. In particle physics<br />

T is the observ<strong>at</strong>ion (Euklidean) time. In one is interested only in the ground<br />

(or vacuum) st<strong>at</strong>e E 0 , one gets it by taking the temper<strong>at</strong>ure T Temp −→ 0 or<br />

observ<strong>at</strong>ion time T −→ ∞.<br />

Second. The various p<strong>at</strong>hs x(τ) enter the Z with a weight given by their<br />

action S [x(τ)] . In the classical limit, i.e. for ħ −→ 0, only the p<strong>at</strong>h with the<br />

smallest action survives, i.e. the one with δS = 0. This is Hamiltons principle<br />

and the corresponding vari<strong>at</strong>ional Euler-equ<strong>at</strong>ion are the Lagrange equ<strong>at</strong>ions.<br />

Third: For small but finite ħ one expects naively in eq.(226) th<strong>at</strong> the p<strong>at</strong>h<br />

with the smallest (minus sign in the exponent !) action in fact domin<strong>at</strong>es the<br />

p<strong>at</strong>h integral . However this is only one single p<strong>at</strong>h, which corresponds in the<br />

figure just to the point (remember x 1 = x 2 ) <strong>at</strong> the bottom of the potential. If one<br />

al<strong>low</strong>s more p<strong>at</strong>hs which in some random way ”oscill<strong>at</strong>e” around the minimum<br />

then one has immedi<strong>at</strong>ely lots of p<strong>at</strong>hs contributing to the p<strong>at</strong>h integral. Those<br />

p<strong>at</strong>hs have actions larger than the minimal one, however since there are many<br />

of them one devi<strong>at</strong>es immedi<strong>at</strong>ely from the classical limit. Thus one loses in<br />

action but one gains in entropy, th<strong>at</strong>s why quantum fluctu<strong>at</strong>ions are important.<br />

Fourth: The p<strong>at</strong>hs, which contribute, e.g. for an harmonic oscill<strong>at</strong>or and a<br />

two-well harmonic oscill<strong>at</strong>or can be depicted as shown in the figure (the figure<br />

does not really show th<strong>at</strong> x 1 = x 2 ). One realizes th<strong>at</strong> p<strong>at</strong>hs extend from one<br />

valley to the other. These p<strong>at</strong>hs are called ”Instantons”. They describe tunneling<br />

processes between the valleys. They play a tremendous role in <strong>QCD</strong>, where<br />

they are known to be responsible for the spontaneous chiral symmetry breaking.<br />

217


13.2.6 Metropolis Formalism<br />

We could evalu<strong>at</strong>e the p<strong>at</strong>h integrals in 〈〈Γ[x]〉〉 using a standard multidimensional<br />

integr<strong>at</strong>ion code, <strong>at</strong> least for one-dimensional systems. However this is<br />

unfeasable for realistic l<strong>at</strong>tices. For eample a l<strong>at</strong>tice gauge calcul<strong>at</strong>ion in <strong>QCD</strong><br />

with 40 l<strong>at</strong>tice points in every direction we have 4 ⋆ 10 4 so called link-variables<br />

and due to the group SU(3) we have 81920000 real variables. Th<strong>at</strong> should be<br />

intractable for conventional quadr<strong>at</strong>ures even in the future. Here, instead, we<br />

employ a more generally useful Monte Carlo procedure. Noting th<strong>at</strong><br />

∫<br />

DxΓ[x]exp[−S[x]]<br />

〈〈Γ[x]〉〉 = ∫<br />

Dxexp[−S[x]]<br />

is a weighted average over p<strong>at</strong>hs with weight exp(−S[x]), we gener<strong>at</strong>e a large<br />

number, N cf , of random p<strong>at</strong>hs or configur<strong>at</strong>ions,<br />

x (α) ≡ {x (α)<br />

0 x (α)<br />

1 ...x (α)<br />

N−1 }<br />

α = 1,2...N cf,<br />

on our grid in such a particular way th<strong>at</strong> the probability P[x (α) ] for obtaining<br />

any particular p<strong>at</strong>h x (α) is<br />

P[x (α) ] ∝ exp[−S[x (α) ]] (227)<br />

218


e -S(x)<br />

x<br />

Then an unweighted average of Γ[x] over this particular set of p<strong>at</strong>hs approxim<strong>at</strong>es<br />

the weighted average over uniformly distributed p<strong>at</strong>hs:<br />

〈〈Γ[x]〉〉 ≈ Γ ≡ 1<br />

N<br />

∑ cf<br />

Γ[x (α) ].<br />

N cf<br />

Γ is our “Monte Carlo estim<strong>at</strong>or” for 〈〈Γ[x]〉〉 on our l<strong>at</strong>tice. In fact this provides<br />

us with a large number of points in the important regions of the integral,<br />

improving the accuray drastically<br />

Of course the estim<strong>at</strong>e will never be exact since the number of p<strong>at</strong>hs N f will<br />

never be infinite. The Monte Carlo uncertainty σ¯Γ in our estim<strong>at</strong>e is a potential<br />

source of error; it is estim<strong>at</strong>ed in the usual fashion:<br />

{<br />

σ 2¯Γ ≈ 1<br />

N<br />

}<br />

1 ∑<br />

Γ 2 [x (α) ] − Γ 2 . (228)<br />

N cf<br />

This becomes<br />

N cf<br />

α=1<br />

α=1<br />

σ 2¯Γ = 〈〈Γ2 〉〉 − 〈〈Γ〉〉 2<br />

N f<br />

for large N f . Since the numer<strong>at</strong>or in this expression is independent of N f (in<br />

principle, it can be determined directly from quantum mechanics), the st<strong>at</strong>istical<br />

uncertainties vanish as 1/ √ N f when N f increases.<br />

We need some sort of specialized random-vector gener<strong>at</strong>or to cre<strong>at</strong>e our set<br />

of random p<strong>at</strong>hs x (α) with probability (227). Possibly the simplest procedure,<br />

though not always the best, is the Metropolis Algorithm . In this procedure, we<br />

start with an arbitrary p<strong>at</strong>h x (0) and modify it by visiting each of the sites on<br />

the l<strong>at</strong>tice, and randomizing the x j ’s <strong>at</strong> those sites, one <strong>at</strong> a time, in a particular<br />

fashion th<strong>at</strong> is described be<strong>low</strong>. In this way we gener<strong>at</strong>e a new random p<strong>at</strong>h<br />

from the old one: x (0) → x (1) . This is called “upd<strong>at</strong>ing” the p<strong>at</strong>h. Applying the<br />

algorithm to x (1) we gener<strong>at</strong>e p<strong>at</strong>h x (2) , and so on until we have N cf random<br />

p<strong>at</strong>hs. This set of random p<strong>at</strong>hs has the correct distribution if N cf is sufficiently<br />

large.<br />

The algorithm for randomizing x j <strong>at</strong> the j h site is:<br />

219


• gener<strong>at</strong>e a random number ζ, with probability uniformly distributed between<br />

−δ and δ for some constant δ;<br />

• replace x j → x j + ζ and compute the change ∆S in the action caused by<br />

this replacement (generally only a few terms in the l<strong>at</strong>tice action involve<br />

x j , since lagrangians are local; only these need be examined);<br />

• if ∆S < 0 (the action is reduced) retain this new value for x j , and proceed<br />

to the next site;<br />

• if ∆S > 0 accept the new value x j + ζ with probability exp(−∆S). This<br />

means: gener<strong>at</strong>e a random number η unformly distributed between 0<br />

and 1; retain the new value for x j if exp(−∆S) > η, otherwise restore<br />

the old value; proceed to the next site. Acutally, to gener<strong>at</strong>e a random<br />

number on the computer is a non-trivial task, since the computer cannot<br />

gener<strong>at</strong>e a real random number. There are sophistic<strong>at</strong>ead techniques to<br />

do th<strong>at</strong>, which we will not discuss here.<br />

There are two important details concerning the tuning and use of this algorithm.<br />

First, in general some or many of the x j ’s will be the same in two<br />

successive random p<strong>at</strong>hs. The amount of such overlap is determined by the parameter<br />

δ: when δ is very large, changes in the x j ’s are usually large and most<br />

will be rejected; when δ is very small, changes are small and most are accepted,<br />

but the new x j ’s will be almost equal to the old ones. Neither extreme is desirable<br />

since each leads to very small changes in x, thereby s<strong>low</strong>ing down the<br />

220


numerical explor<strong>at</strong>ion of the space of all important p<strong>at</strong>hs. Typically δ should<br />

be tuned so th<strong>at</strong> 40%–60% of the x j ’s are changed on each pass (or “sweep”)<br />

through the l<strong>at</strong>tice. Then δ is of order the typical quantum fluctu<strong>at</strong>ions expected<br />

in the theory. Wh<strong>at</strong>ever the δ, however, successive p<strong>at</strong>hs are going to be quite<br />

similar (th<strong>at</strong> is “highly correl<strong>at</strong>ed”) and so contain r<strong>at</strong>her similar inform<strong>at</strong>ion<br />

about the theory. Thus when we accumul<strong>at</strong>e random p<strong>at</strong>hs x (α) for our Monte<br />

Carlo estim<strong>at</strong>es we should keep only every N cor -th p<strong>at</strong>h; the intervening sweeps<br />

erase correl<strong>at</strong>ions, giving us configur<strong>at</strong>ions th<strong>at</strong> are st<strong>at</strong>istically independent.<br />

The optimal value for N cor depends upon the theory, and can be found by trial.<br />

It also depends on the l<strong>at</strong>tice spacing a, going roughly as<br />

N cor ∝ 1 a 2 .<br />

Other algorithms exist for which N or grows only as 1/a when a is reduced, but<br />

since our interest is in large a’s we will not discuss these further.<br />

The second detail concerns the procedure for starting the algorithm. The<br />

very first configur<strong>at</strong>ion used to seed the whole process is usually fairly <strong>at</strong>ypical.<br />

Consequently we should discard some number of configur<strong>at</strong>ions <strong>at</strong> the beginning,<br />

before starting to collect x (α) ’s. Discarding 5N cor to 10N cor configur<strong>at</strong>ions is<br />

usually adequ<strong>at</strong>e. This is called “thermalizing the l<strong>at</strong>tice.”<br />

To summarize, a computer code for a complete Monte Carlo calcul<strong>at</strong>ion of<br />

〈〈Γ[x]〉〉 for some function Γ[x] of a p<strong>at</strong>h x consists of the fol<strong>low</strong>ing steps:<br />

• initialize the p<strong>at</strong>h, for example, by setting all x j ’s to zero;<br />

• upd<strong>at</strong>e the p<strong>at</strong>h 5N cor –10N cor times to thermalize it;<br />

• upd<strong>at</strong>e the p<strong>at</strong>h N cor times, then compute Γ[x] and save it; repe<strong>at</strong> N cf times.<br />

• average the N cf values of Γ[x] saved in the previous step to obtain a Monte<br />

Carlo estim<strong>at</strong>e Γ for 〈〈Γ[x]〉〉.<br />

A Monte Carlo estim<strong>at</strong>e Γ of some expect<strong>at</strong>ion value 〈〈Γ〉〉 is never exact;<br />

there are always st<strong>at</strong>istical errors th<strong>at</strong> vanish only in the limit where infinitely<br />

many configur<strong>at</strong>ions are employed (N cf → ∞). An important part of any<br />

Monte Carlo analysis is the estim<strong>at</strong>ion of these st<strong>at</strong>istical errors. There is a<br />

simple but very powerful method, called the “st<strong>at</strong>istical bootstrap,” for making<br />

such estim<strong>at</strong>es.<br />

In the previous exercises, for example, we assemble an “ensemble” of measurements<br />

of the propag<strong>at</strong>or G (α) , one for each configur<strong>at</strong>ion x (α) . These are<br />

averaged to obtain G, and, from it, an estim<strong>at</strong>e for ∆E n (generaliz<strong>at</strong>ion of<br />

Eq. (224)). An obvious way to check the st<strong>at</strong>istical errors on this estim<strong>at</strong>e for<br />

∆E n is to redo the whole calcul<strong>at</strong>ion, say, 100 times, each time with different<br />

random numbers to gener<strong>at</strong>e different random p<strong>at</strong>hs. With 100 copies of the<br />

entire calcul<strong>at</strong>ion, we could analyze the distribution of the 100 random ∆E n ’s<br />

obtained, and deduce the st<strong>at</strong>istical uncertainty in our original estim<strong>at</strong>e. This,<br />

however, is exceedingly expensive in computer time. The bootstrap procedure<br />

221


provides new, almost zero-cost random ensembles of measurements by synthesizing<br />

them from the original ensemble of N f measurements.<br />

Given an ensemble {G (α) ,α = 1...N cf } of Monte Carlo measurements, we<br />

assemble a “bootstrap copy” of th<strong>at</strong> ensemble by selecting G (α) ’s <strong>at</strong> random<br />

from the original ensemble, taking N cf in all while al<strong>low</strong>ing duplic<strong>at</strong>ions and<br />

omissions. The resulting ensemble of G’s might have two or three copies of<br />

some G (α) ’s, and no copies of others. This new ensemble can be averaged and a<br />

new estim<strong>at</strong>e obtained for ∆E n . This procedure can be repe<strong>at</strong>ed to gener<strong>at</strong>ed<br />

as many bootstrap copies of the original ensemble as we wish, and from each<br />

we can gener<strong>at</strong>e a new estim<strong>at</strong>e for ∆E n . The distribution of these ∆E n ’s<br />

approxim<strong>at</strong>es the distribution of ∆E n ’s th<strong>at</strong> would have been obtained from<br />

the original Monte Carlo, and so can be used to estim<strong>at</strong>e the st<strong>at</strong>istical error in<br />

our original estim<strong>at</strong>e.<br />

Another useful procedure rel<strong>at</strong>ed to st<strong>at</strong>istical errors is “binning.” At the<br />

end of a large simul<strong>at</strong>ion we might have 100’s or even 100,000’s of configur<strong>at</strong>ions<br />

x (α) , and for each a set of measurements like G (α) , our propag<strong>at</strong>or. The<br />

measurements will inevitably be averaged, but we want to save the separ<strong>at</strong>e<br />

G (α) ’s for making bootstrap error estim<strong>at</strong>es and the like. We can save a lot of<br />

disk space, RAM, and CPU time by partially averaging or binning the measurements:<br />

For example, instead of storing each of<br />

we might instead store<br />

G (1) G (2) G (3) G (4) G (5) ...<br />

G (1) ≡ G(1) + G (2) + G (3) + G (4)<br />

4<br />

G (2) ≡ G(5) + G (6) + G (7) + G (8)<br />

4<br />

...<br />

The G (β) ’s are far less numerous but have the same average, standard devi<strong>at</strong>ion,<br />

and other st<strong>at</strong>istical properties as the original set. Typically the bin size is<br />

adjusted so th<strong>at</strong> there are only 50–100 G (β) ’s.<br />

13.3 Boson Quantum Field on the l<strong>at</strong>tice<br />

13.3.1 Quantum Field Theory with functional integrals<br />

Now we are going to transl<strong>at</strong>e the above represent<strong>at</strong>ion of quantum mechanics<br />

in terms of p<strong>at</strong>h integrals to field theory. We consider a scalar field φ(x), where<br />

x = (⃗x,t) labels space-time coordin<strong>at</strong>es, and the time evolution of φ(⃗x,t) is<br />

given by<br />

φ(⃗x,t) = e iHt φ(⃗x,t = 0)e −iHt .<br />

The objects of interest in field theory are vacuum expect<strong>at</strong>ion values of (time<br />

ordered) products of field oper<strong>at</strong>ors, i.e. the Greens functions:<br />

〈0|φ(x 1 )φ(x 2 )...φ(x n )|0〉, t 1 > t 2 > · · · > t n .<br />

222


Prominent examples are propag<strong>at</strong>ors<br />

〈0|φ(x)φ(y)|0〉.<br />

The Greens functions essentially contain all physical inform<strong>at</strong>ion. In particular,<br />

S-m<strong>at</strong>rix elements are rel<strong>at</strong>ed to Greens functions, e.g. the 2-particle sc<strong>at</strong>tering<br />

elements can be obtained from<br />

〈0|φ(x 1 )...φ(x 4 )|0〉.<br />

Instead of discussing the functional integral represent<strong>at</strong>ion for quantum field<br />

theory from the beginning, we shall restrict ourselves to transl<strong>at</strong>ing the quantum<br />

mechanical concepts to field theory by means of analogy. To this end we would<br />

like to transl<strong>at</strong>e the basic variables x i (t) into fields φ(⃗x,t). The rules for the<br />

transl<strong>at</strong>ion are then<br />

x i (t) ←→ φ(⃗x,t)<br />

i ←→ ⃗x<br />

∏<br />

dx i (t) ←→ ∏ dφ(⃗x,t) ≡ Dφ<br />

t,i<br />

t,⃗x<br />

∫<br />

∫<br />

S = dt L ←→ S = dtd 3 x L,<br />

where S is the classical action.<br />

For scalar field theory we might consider the fol<strong>low</strong>ing Lagrangian density<br />

(φ 4 -theory):<br />

L = 1 2<br />

(<br />

( ˙φ(x)) 2 − (∇φ(x)) 2) − m2 0<br />

2 φ(x)2 − g 0<br />

4! φ(x)4<br />

= 1 2 (∂ µφ)(∂ µ φ) − m2 0<br />

2 φ(x)2 − g 0<br />

4! φ(x)4 .<br />

The mass m 0 and coupling constant g 0 bear a subscript 0, since they are bare,<br />

unrenormalized parameters. This theory plays a role in the context of Higgs-<br />

Yukawa models, where φ(x) is the Higgs field.<br />

In analogy to the quantum mechanical p<strong>at</strong>h integral we now write down<br />

a represent<strong>at</strong>ion of the Greens functions in terms of wh<strong>at</strong> one calls functional<br />

integrals:<br />

〈0|φ(x 1 )φ(x 2 )...φ(x n )|0〉 = 1 ∫<br />

Dφ φ(x 1 )φ(x 2 )...φ(x n )e iS<br />

Z<br />

with<br />

∫<br />

Z =<br />

Dφ e iS .<br />

These expressions involve integrals over all classical field configur<strong>at</strong>ions.<br />

As mentioned before, we do not <strong>at</strong>tempt any deriv<strong>at</strong>ion of functional integrals<br />

but just want to motiv<strong>at</strong>e their form by analogy. Furthermore, in the case<br />

223


of quantum mechanics we considered the transition amplitude, whereas now we<br />

have written the formula for Greens functions, which is a bit different.<br />

The formulae for functional integrals give rise to some questions. First of all,<br />

how does the projection onto the groundst<strong>at</strong>e |0〉 arise Secondly, these integrals<br />

contain oscill<strong>at</strong>ing integrands, due to the imaginary exponents; wh<strong>at</strong> about their<br />

convergence Moreover, is there a way to evalu<strong>at</strong>e them numerically<br />

In the fol<strong>low</strong>ing we shall discuss, how the introduction of imaginary times<br />

helps in answering these questions.<br />

Procedure: Summary In fact the task to compute observables on a l<strong>at</strong>tice<br />

inolves the fol<strong>low</strong>ing steps<br />

i) Any renormalizable quantum field theory, such as <strong>QCD</strong> requires an ultraviolet<br />

(short distance) regulariz<strong>at</strong>ion in order to elimin<strong>at</strong>e infinities from calcul<strong>at</strong>ed<br />

physical quantities. Here this cutoff procedure is introduced by defining<br />

the fields on a mesh of discrete l<strong>at</strong>tice points which replaces the space-time<br />

continuum. Ultim<strong>at</strong>ly one must ensure th<strong>at</strong> the dependence of any calcul<strong>at</strong>ed<br />

observable on the l<strong>at</strong>tice spacing s<strong>at</strong>isfies the appropri<strong>at</strong>e renormaliz<strong>at</strong>ion group<br />

equ<strong>at</strong>ion. Th<strong>at</strong> is, the calcul<strong>at</strong>ed values of physical observables should scale correctly<br />

with a −→ 0. This meas, if one identifies the inverse l<strong>at</strong>tice spacing a −1<br />

with the UV-cut-off Λ cutoff , then the calcul<strong>at</strong>ed values should have the proper<br />

Λ cutoff -dependence given by renormaliz<strong>at</strong>ion group equ<strong>at</strong>ions.<br />

ii) An infrared (long-wavelength) cutoff is imposed by working on a finite<br />

l<strong>at</strong>tice volume. This is a techical necessity in order to be able to perform the<br />

numerical calcul<strong>at</strong>ion of the integrals, which unavoidably requires a finite number<br />

of variables. Again, one must extrapol<strong>at</strong>e the value of any observable to the<br />

infinite volume limit.<br />

iii) In general the functional integrals in Minkowski space are complex and oscill<strong>at</strong>e<br />

rapidly, so th<strong>at</strong> they cannot be tre<strong>at</strong>ed numerically in this form. In order<br />

to provide a sound basis, both for formal m<strong>at</strong>hem<strong>at</strong>ical reasons and for practical<br />

numerical purposes, it is normal to calcul<strong>at</strong>e the p<strong>at</strong>h integrals for imaginary<br />

times: t = x 0 = −ix 4 = iτ. This change of variables, together with the appropri<strong>at</strong>e<br />

analytic continu<strong>at</strong>ion of the Green functions, defines a Euclidean quantum<br />

field theory: th<strong>at</strong> is, the fields now live on a four dimensional Euclidean space.<br />

L<strong>at</strong>tice QFT is then represented in terms of well defined functional integrals<br />

taken over the Euclidean l<strong>at</strong>tice hypercube Λ = { x; xµ a<br />

∈ Z;µ = 1,2,3,4}<br />

iv) In practice all these limits have to be studied and justified.<br />

v) In fact there are even more problems if one wants to calcl<strong>at</strong>e e.g. nucleon<br />

properties in terms of <strong>QCD</strong>: There are so called fermion doublings which are<br />

caused by the fact th<strong>at</strong> on a discrete l<strong>at</strong>tice autom<strong>at</strong>ically several fermions<br />

are calcul<strong>at</strong>ed simultaneously and special techniques have to be developed to<br />

minimize the errors caused by th<strong>at</strong>. Furthermore present day l<strong>at</strong>tices have<br />

tremendous problems using quarks on the l<strong>at</strong>tice which have physical values<br />

of their masses. In practice all the used quark masses are too large and one<br />

has to think about extrapol<strong>at</strong>ion methods towards physical values of current<br />

quark masses. Very often not full l<strong>at</strong>tice calcul<strong>at</strong>ions involving fermions are<br />

224


t<br />

t t 1<br />

2<br />

performed but so-called quenched calcul<strong>at</strong>ions. For large fermion masses they<br />

seem to be alright, however for values closer to their pysical value the quenched<br />

approxim<strong>at</strong>ion does not seem to work well.<br />

13.3.2 Euklidean Field Theory<br />

Let us return to quantum mechanics for a moment. Here we have introduced<br />

Greens functions, e.g.<br />

G(t 1 ,t 2 ) = 〈0|ˆx(t 1 )ˆx(t 2 )|0〉, t 1 > t 2 .<br />

and we have demonstf<strong>at</strong>ed th<strong>at</strong> these Greens functions are rel<strong>at</strong>ed to quantum<br />

mechanical amplitudes <strong>at</strong> imaginary times by analytic continu<strong>at</strong>ion. We had<br />

expressed the Greens function <strong>at</strong> imaginary times,<br />

G E (τ 1 ,τ 2 ) = 〈0|ˆxe −H(τ1−τ2)ˆx|0〉,<br />

can be expressed as a p<strong>at</strong>h integral<br />

G E (τ 1 ,τ 2 ) = 1 ∫<br />

Z<br />

Dx x(τ 1 )x(τ 2 )e −SE ,<br />

where<br />

∫<br />

Z =<br />

Dx e −SE<br />

The Greens function <strong>at</strong> real times, which we were interested in originally,<br />

can be obtained from G E by means of analytical continu<strong>at</strong>ion, G(t 1 ,t 2 ) =<br />

G E (it 1 ,it 2 ). The analytical continu<strong>at</strong>ion has to be done in such a way th<strong>at</strong><br />

all time arguments are rot<strong>at</strong>ed simultaneously counter-clockwise in the complex<br />

t-plane. This is the so-called Wick rot<strong>at</strong>ion, illustr<strong>at</strong>ed in thei Fig..<br />

Now we turn to field theory again. The Greens functions<br />

G(x 1 ,...,x n ) = 〈0|Tφ(x 1 )...φ(x n )|0〉,<br />

225


continued to imaginary times, t = −iτ, are the so-called Schwinger functions<br />

G E ((⃗x 1 ,τ 1 ),...,(⃗x n ,τ n )) = G((⃗x 1 , −iτ 1 ),...,(⃗x n , −iτ n )).<br />

In analogy to the quantum mechanical case their functional integral represent<strong>at</strong>ion<br />

reads<br />

G E (x 1 ,...,x n ) = 1 ∫<br />

Dφ φ(x 1 )...φ(x n )e −SE<br />

Z<br />

with<br />

∫<br />

Z =<br />

Dφ e −SE<br />

and e.g. in the above example of φ 4 -theory:<br />

∫ {<br />

S E = d 3 1<br />

xdτ<br />

2<br />

∫<br />

=<br />

( ) }<br />

2 dφ<br />

+ 1 dτ 2 (∇φ)2 + m2 0<br />

2 φ2 + g 0<br />

4! φ4<br />

{ 1<br />

d 4 x<br />

2 (∂ µφ) 2 + m2 0<br />

2 φ2 + g }<br />

0<br />

4! φ4 . (229)<br />

As can also be seen from the kinetic part contained in S E , the metric of<br />

Minkowski space<br />

−ds 2 = −dt 2 + dx 2 1 + dx 2 2 + dx 2 3<br />

has changed into<br />

dτ 2 + dx 2 1 + dx 2 2 + dx 2 3,<br />

which is the metric of a Euclidean space. Therefore one speaks of Euclidean<br />

Greens functions G E and of Euclidean functional integrals. They are taken<br />

as starting point for non-perturb<strong>at</strong>ive investig<strong>at</strong>ions of field theories and for<br />

constructive studies.<br />

As S E is real, the integrals of interest are now real and no unpleasant oscill<strong>at</strong>ions<br />

occur. Moreover, since S E is bounded from be<strong>low</strong>, the factor exp(−S E )<br />

in the integrand is bounded. Strongly fluctu<strong>at</strong>ing fields have a large Euclidean<br />

action S E and are thus suppressed by the factor exp(−S E ). (Strictly speaking,<br />

this st<strong>at</strong>ement does not make sense in field theory unless renormaliz<strong>at</strong>ion<br />

is taken into account.) This makes Euclidean functional integrals so <strong>at</strong>tractive<br />

compared to their Minkowskian counterparts.<br />

To illustr<strong>at</strong>e the coordin<strong>at</strong>e transform<strong>at</strong>ion to imaginary time, there is a<br />

little exercise. Consider the Feynman propag<strong>at</strong>or and it is easy to show th<strong>at</strong><br />

∆ E F (x) =<br />

∫ d 4 p<br />

(2π) 4<br />

e ipx<br />

p 2 + m 2 ,<br />

0<br />

(where px is to be understood as a Euclidean scalar product), is obtained by<br />

correct Wick rot<strong>at</strong>ion. To be more precise,<br />

∆ F (⃗x,t) = lim<br />

φ→π/2 ∆E F (⃗x,te iφ ),<br />

226


p 2 + m 2<br />

p 0<br />

p<br />

4<br />

with ∆ F the Feynman propag<strong>at</strong>or in Minkowski-space<br />

∫ d 4 p e −ip∗x<br />

∆ F (⃗x,t) = i<br />

(2π) 4 p 2 − m 2 0 + iɛ,<br />

where all scalar products in the last expression are defined with Minkowski<br />

metric. An important fe<strong>at</strong>ure of the Wick-rot<strong>at</strong>ed propag<strong>at</strong>or is the absence of<br />

singularities on the p 4 -axis in Euclidean space, see Fig.<br />

One might think th<strong>at</strong> in the Euclidean domain everything is unphysical and<br />

there is no possibility to get physical results directly from the Euclidean Greens<br />

functions. But this is not the case, as we might guess already from the simple<br />

quantum mechanical exerzises. For example, the spectrum of the theory can be<br />

obtained in the fol<strong>low</strong>ing way. Let us consider a vacuum expect<strong>at</strong>ion value of<br />

the form<br />

〈0|A 1 e −Hτ A 2 |0〉,<br />

where the A i ’s are formed out of the field φ, e.g. A = φ(⃗x,0) or A = ∫ d 3 x φ(⃗x,0).<br />

Now, with the familiar insertion of a complete set of energy eigenst<strong>at</strong>es, we have<br />

〈0|A 1 e −Hτ A 2 |0〉 = ∑ n<br />

〈0|A 1 |n〉e −Enτ 〈n|A 2 |0〉.<br />

In case of a continuous spectrum the sum is to be read as an integral. On the<br />

other hand, representing the expect<strong>at</strong>ion value as a functional integral leads to<br />

∫<br />

1<br />

Dφ e −SE A 1 (τ)A 2 (0) = ∑ 〈0|A 1 |n〉〈n|A 2 |0〉e −Enτ .<br />

Z<br />

n<br />

This is similar to the ground st<strong>at</strong>e projection <strong>at</strong> the beginning of this chapter.<br />

For large τ the <strong>low</strong>est energy eigenst<strong>at</strong>es will domin<strong>at</strong>e the sum and we can thus<br />

227


obtain the <strong>low</strong>-lying spectrum from the asymptotic behaviour of this expect<strong>at</strong>ion<br />

value. One should note, th<strong>at</strong> the excited st<strong>at</strong>es of a field theory are the<br />

1-particle-2-particle st<strong>at</strong>es, which are theFock-st<strong>at</strong>es compose by the cre<strong>at</strong>ion<br />

oper<strong>at</strong>ors of the quantized field oper<strong>at</strong>or. Inorder to obtain the mass of the<br />

<strong>low</strong>est lying excited st<strong>at</strong>e, i.e. the lightest particle in rest, one has to choose<br />

A 1 ,A 2 suitably, e.g. for<br />

∫<br />

A ≡ A 1 = A 2 = d 3 x φ(⃗x,0),<br />

such th<strong>at</strong> 〈0|A|1〉 ≠ 0 for a one-particle st<strong>at</strong>e |1〉 with zero momentum ⃗p = 0<br />

and mass m 1 , we will get<br />

∫<br />

1<br />

Dφ e −SE A(τ)A(0) = |〈0|A|1〉| 2 e −m1τ + ...,<br />

Z<br />

which means th<strong>at</strong> we can extract the mass of the lightest particle.<br />

From now on we shall remain in Euclidean space and suppress the subscript<br />

E, so th<strong>at</strong> S ≡ S E means the Euclidean action.<br />

L<strong>at</strong>tice discretiz<strong>at</strong>ion<br />

One central question still remains: does the infinite dimensional integr<strong>at</strong>ion<br />

over all classical field configur<strong>at</strong>ions, i.e.<br />

Dφ = ∏ x<br />

dφ(x),<br />

make sense <strong>at</strong> all How is it defined<br />

13.3.3 Scalar boson field: Discretiz<strong>at</strong>ion of space-time<br />

Remember the way we derived the p<strong>at</strong>h integral represent<strong>at</strong>ion of quantum<br />

mechanics. It was obtained as a limit of a discretiz<strong>at</strong>ion in time τ. As in field<br />

theory the fields depend on the four Euclidean coordin<strong>at</strong>es instead of a single<br />

time coordin<strong>at</strong>e, we may now introduce a discretized space-time in form of a<br />

l<strong>at</strong>tice, for example a hypercubic l<strong>at</strong>tice, specified by<br />

see Fig.<br />

x µ = an µ , n µ ∈ Z,<br />

The quantity a is called the l<strong>at</strong>tice spacing for obvious reasons. The scalar<br />

field<br />

φ(x), x ∈ l<strong>at</strong>tice, −→ φ j<br />

is now defined on the l<strong>at</strong>tice points only. Partial deriv<strong>at</strong>ives are replaced by<br />

finite differences,<br />

∂ µ φ −→ ∆ µ φ(x) ≡ 1 a (φ(x + aˆµ) − φ(x)) = 1 a<br />

(<br />

φj+1µ − φ j<br />

)<br />

,<br />

228


and<br />

∂ µ ∂ µ φ −→ ∑ µ<br />

1 [ ]<br />

a 2 φj+1µ − 2φ j φ j − φ j−1µ<br />

and space-time integrals by sums:<br />

∫<br />

d 4 x −→ ∑ x<br />

a 4 = a 4 ∑ j<br />

.<br />

The action of our discretized φ 4 -theory, Eq. (229), can be written as<br />

S = ∑ {<br />

}<br />

a 4 1<br />

4∑<br />

(∆ µ φ(x)) 2 + m2 0<br />

2<br />

2 φ(x)2 + g 0<br />

4! φ(x)4 .<br />

x<br />

µ=1<br />

In the functional integrals the measure<br />

Dφ = ∏ x<br />

dφ(x) −→ ∏ j<br />

dφ j<br />

involves the l<strong>at</strong>tice points x only. So we have a discrete set of variables to<br />

integr<strong>at</strong>e. If the l<strong>at</strong>tice is taken to be finite, we just have finite dimensional<br />

integrals.<br />

Discretiz<strong>at</strong>ion of space-time using l<strong>at</strong>tices has one very important consequence.<br />

Due to a non-zero l<strong>at</strong>tice spacing a cutoff in momentum space arises.<br />

The cutoff can be observed by having a look <strong>at</strong> the Fourier transformed field<br />

˜φ(p) = ∑ x<br />

a 4 e −ipx φ(x).<br />

The Fourier transformed functions are periodic in momentum-space, so th<strong>at</strong> we<br />

can identify<br />

p µ<br />

∼ = pµ + 2π a<br />

229


and restrict the momenta to the so-called Brillouin zone<br />

− π a < p µ ≤ π a .<br />

The inverse Fourier transform<strong>at</strong>ion, for example, is given by<br />

φ(x) =<br />

We recognize an ultraviolet cutoff<br />

∫ π/a<br />

−π/a<br />

d 4 p<br />

(2π) 4 eipx ˜φ(p).<br />

|p µ | ≤ π a .<br />

Therefore field theories on a l<strong>at</strong>tice are regularized in a n<strong>at</strong>ural way.<br />

In order to begin in a well-defined way one would start with a finite l<strong>at</strong>tice.<br />

Let us assume a hypercubic l<strong>at</strong>tice with length L 1 = L 2 = L 3 = L in every<br />

sp<strong>at</strong>ial direction and length L 4 = T in Euclidean time,<br />

x µ = an µ , n µ = 0,1,2,...,L µ − 1,<br />

with finite volume V = L 3 T. In a finite volume one has to specify boundary<br />

conditions. A popular choice are periodic boundary conditions<br />

φ(x) = φ(x + aL µ ˆµ),<br />

where ˆµ is the unit vector in the µ-direction. They imply th<strong>at</strong> the momenta are<br />

also discretized,<br />

p µ = 2π a<br />

l µ<br />

L µ<br />

with l µ = 0,1,2,...,L µ − 1,<br />

and therefore momentum-space integr<strong>at</strong>ion is replaced by finite sums<br />

∫ d 4 p<br />

(2π) 4 −→ 1 ∑<br />

a 4 L 3 .<br />

T<br />

l µ<br />

Now, all functional integrals have turned into regularized and finite expressions.<br />

Of course, one would like to recover physics in a continuous and infinite<br />

space-time eventually. The task is therefore to take the infinite volume limit,<br />

L,T −→ ∞,<br />

which is the easier part in general, and to take the the continuum limit,<br />

a −→ 0.<br />

Constructing the continuum limit of a l<strong>at</strong>tice field theory is usually highly nontrivial<br />

and most effort is often spent here.<br />

230


The Eucclidean continuum action of the Klein-Gordon-Field is given by<br />

S E [φ] = 1 ∫<br />

d 4 xφ(x) ( −∂ µ ∂ µ + M 2) φ(x)<br />

2<br />

If we define<br />

˜φ j = aφ j<br />

˜Mj = aM<br />

then we get<br />

with<br />

S E = 1 ∑<br />

˜φ j K jk ˜φk<br />

2<br />

jk<br />

[ ]<br />

δj+1µ,k − 2δ jk + δ j−1µ,k + ˜M2 δ jk (230)<br />

K jk = − ∑ µ<br />

An importante technique in the quantum field theory is to use the gener<strong>at</strong>ing<br />

functional, since by functional deriv<strong>at</strong>ives one can obtain the greens functions :<br />

∫ [ ∫ ]<br />

Z [J] = Dφexp −S E [φ] + d 4 xJ(x)φ(x) (231)<br />

which reads in the discretized case<br />

⎡<br />

∫<br />

Z [J] = Π l d˜φ l exp ⎣− 1 ∑<br />

˜φ j K jk ˜φk + ∑ 2<br />

j<br />

jk<br />

J j ˜φj<br />

⎤<br />

⎦<br />

and now the Greens function becomes just<br />

∫ [<br />

Πl d˜φ l ˜φr ˜φs exp<br />

∑jk ˜φ<br />

]<br />

j K jk ˜φk<br />

< 0|˜φ r ˜φs | >=<br />

∫<br />

Πl d˜φ l exp[<br />

− 1 2<br />

− 1 2<br />

∑ ˜φ<br />

] = ∂2 Z [J]<br />

jk j K ∂J jk ˜φk r ∂J s Z [0] | J=0<br />

In order to solve functional integral we remember the harmonic oscill<strong>at</strong>or formulae,<br />

which looked identical. There is indeed no difference, since the integral<br />

over the fields is here reduced to an integral over the magnitude of the field <strong>at</strong><br />

a certain space-time point. Thus we obtain analogously to eq.() the integral<br />

⎡<br />

⎤<br />

∫<br />

Π l d˜φ l exp⎣− 1 ∑<br />

˜φ j K jk ˜φk + ∑ (√ )<br />

⎡<br />

⎤<br />

N<br />

2π<br />

J j ˜φj ⎦ = √ exp ⎣− 1 ∑<br />

J j (K −1 ) jk J k<br />

⎦<br />

2<br />

jk<br />

j det K 2<br />

jk<br />

(232)<br />

The proof of eq.(232) is simple: We can write the LHS as<br />

∫ [<br />

LHS = Π l d˜φ l exp −<br />

2−→˜φ 1 T K −→˜φ ] ∫ [<br />

−→ + J<br />

−→˜φ = Π l d˜φ l exp −<br />

2−→˜φ 1 T ADA −→˜φ ]<br />

−→ + J<br />

−→˜φ<br />

231


since K is real and symmetric and we have then det(A) = 1. We define −→ u =<br />

A −1−→˜φ and we can rewrite this expression as<br />

∫ [<br />

LHS = Π l du l exp −<br />

2−→ 1 u T D −→ u + −→ J T A −→ ]<br />

u<br />

We use now the well known expression<br />

∫ ∞<br />

−∞<br />

du exp ( −αu 2 + βu ) =<br />

√ ( ) π<br />

a exp − β2<br />

4α<br />

and obtain immedi<strong>at</strong>ely eq.(232).<br />

From second order deriv<strong>at</strong>ives with respect to J k of eq.(232) one obtains<br />

immedi<strong>at</strong>ely<br />

< 0|˜φ r ˜φs | >= ( K −1) rs<br />

The m<strong>at</strong>rix K −1 has to be calcul<strong>at</strong>ed on the l<strong>at</strong>tice. The best is to do this in<br />

momentum space. There we have the represent<strong>at</strong>ion of the delta-function<br />

δ nm =<br />

∫ +π<br />

−π<br />

d 4ˆk [ ]<br />

(2π) 4 exp iˆk(n − m)<br />

where ˆk = (ˆk 1 , ˆk 2 , ˆk 3 , ˆk 4 ) = k/a is dimensionless. Thus we have to calcul<strong>at</strong>e the<br />

K(ˆk) of<br />

∫ +π<br />

d 4ˆk [ ]<br />

K nm =<br />

(2π) 4 K(ˆk)exp iˆk(n − m)<br />

−π<br />

Starting from eq.(230) a straightforward and explicit but slightly complic<strong>at</strong>ed<br />

calcul<strong>at</strong>ion yields<br />

[ ]<br />

δj+1µ,k − 2δ jk + δ j−1µ,k + ˜M2 δ jk<br />

K jk = − ∑ µ<br />

=<br />

and we find from<br />

then also<br />

∫ +π<br />

−π<br />

(K −1 ) rs =<br />

d 4ˆk<br />

[<br />

(2π) 4 4<br />

) ]<br />

4∑<br />

(ˆkµ<br />

sin 2 +<br />

2<br />

˜M<br />

[ ]<br />

2 exp iˆk(j − k)<br />

µ=1<br />

∑<br />

K rq (K −1 ) qs = δ rs<br />

q<br />

∫ +π<br />

−π<br />

d 4ˆk<br />

(2π) 4<br />

[ ]<br />

exp iˆk(r − s)<br />

4 ∑ 4<br />

µ=1 sin2 ( ˆkµ<br />

2<br />

)<br />

+ ˜M 2 (233)<br />

It is useful and easy to check th<strong>at</strong> this expression on the l<strong>at</strong>tice has the proper<br />

continuum limit. To show this we go back to dimensional quantities, defining<br />

r = x/a and s = y/a and ˜M = aM and k = ˆk/a then we have<br />

(K −1 ) rs = a 4 ∫ +π/a<br />

−π/a<br />

d 4 k<br />

(2π) 4<br />

exp [ik(x − y)]<br />

4 ∑ 4<br />

µ=1 sin2 ( ak µ<br />

2<br />

)<br />

+ a 2 M 2<br />

232


In the continuum limit a → 0 we can expand the sinus and retain only the first<br />

term. This yields then<br />

< 0|˜φ r ˜φs | >= ( K −1) rs → a2 ∫ +∞<br />

−∞<br />

d 4 k exp[ik(x − y)]<br />

(2π) 4 k 2 + M 2<br />

Except of the normaliz<strong>at</strong>ion factor a 2 this is the well known expression for the<br />

Greens function in Euklidean space.<br />

13.4 Fermion Quantum fields on the l<strong>at</strong>tice<br />

First we consider free massive fermions without any gauge field. The free Dirac<br />

equ<strong>at</strong>ion reads<br />

[iγ µ ∂ µ − M]ψ(x) = 0<br />

or [<br />

iγ 0 ∂ t + i −→ γ −→ ]<br />

∇ − M ψ(x) = 0<br />

Their Minkowsky-action is given by<br />

∫<br />

S F = d 4 x ¯ψ(x)(iγ µ ∂ µ − M) ψ(x)<br />

Again it is convenient to go to Euklidean space, t → −iτ, i.e. changing from<br />

iγ 0 ∂ t = −γ 0 ∂ τ . Thus the Dirac equ<strong>at</strong>ion reads<br />

[<br />

−γ 0 ∂ τ + i −→ γ −→ ]<br />

∇ ψ(x) = 0<br />

For a consistent writing it is useful to introduce Euklidean gamma m<strong>at</strong>rices:<br />

γ E 0 = γ 0 γ E 1 = −iγ 1 γ E 2 = −iγ 2 γ E 3 = −iγ 3<br />

which s<strong>at</strong>isfy<br />

{<br />

γ<br />

E<br />

µ ,γν<br />

E }<br />

= 2δµν<br />

Then the Euklidean Dirac equ<strong>at</strong>ion reads<br />

[<br />

−γ E 0 ∂ τ − −→ ∇ −→ γ E] ψ(x) = 0<br />

and the Euklidean action is<br />

∫ ( 4∑<br />

)<br />

SE F = d 4 x ¯ψ(x) γµ E ∂ µ + M ψ(x)<br />

µ=1<br />

From now on we will drop the index E and will always work in the Euklidean<br />

space. The Greens functions are given by p<strong>at</strong>h integrals as (α,β,... are Dirac<br />

indices)<br />

∫<br />

DψD ¯ψψα (x).....ψ β (x)......exp [ [<br />

−SE<br />

F ψ, ¯ψ]]<br />

< 0|ψ α (x).....ψ β (x)......|0 >= + ∫ [ [<br />

DψD ¯ψ exp −S<br />

F<br />

E<br />

ψ, ¯ψ]]<br />

233


However, one should notice, due to the anticommut<strong>at</strong>ion rules of the fermion<br />

fields the p<strong>at</strong>h integral must be formul<strong>at</strong>ed in a special way. We descretize again<br />

∫<br />

x µ → j µ a d 4 x → a ∑ 4 j<br />

and<br />

and<br />

ψ α (x) → ψ αj<br />

¯ψα (x) → ¯ψ αj j=Euklidean 4-vector<br />

∂ µ ψ α (x) → ψ α,j+1 µ−ψ α,j−1µ<br />

2a<br />

D ¯ψ(x)Dψ(x) → ∏ ∏<br />

¯ψ αj ψ βk<br />

α,j<br />

Actually the ψ αj , ¯ψ αj are Grassmann variables, which takes into account th<strong>at</strong><br />

the fermionic quantum fields are anticommuting. Its properties will be discussed<br />

now.<br />

13.4.1 Grassmann-algebra<br />

The reasoning is as fol<strong>low</strong>s: First remember scalar fields in the continuum.<br />

Classical fields are just ordinary functions and s<strong>at</strong>isfy<br />

[φ(x),φ(y)] = 0,<br />

which can be considered as the limit → 0 of the quantum commut<strong>at</strong>ion rel<strong>at</strong>ions.<br />

Fermi st<strong>at</strong>istics implies th<strong>at</strong> fermionic quantum fields have the well-known<br />

equal-time anticommut<strong>at</strong>ion rel<strong>at</strong>ions<br />

β,k<br />

{ψ(⃗x,t),ψ(⃗y,t)} = 0.<br />

Motiv<strong>at</strong>ed by this, we might introduce a classical limit in which classical fermionic<br />

fields s<strong>at</strong>isfy<br />

{ψ(x),ψ(y)} = 0<br />

for all x,y. Classical fermionic fields are therefore anticommuting variables,<br />

which are also called Grassmann variables.<br />

We would like to point out th<strong>at</strong> the argument above is just a heuristic<br />

motiv<strong>at</strong>ion. More rigorous approaches can be found in the liter<strong>at</strong>ure.<br />

In general, a complex Grassmann algebra is gener<strong>at</strong>ed by elements η i and<br />

¯η i , which obey<br />

{η i ,η j } = 0<br />

{η i , ¯η j } = 0<br />

{¯η i , ¯η j } = 0.<br />

234


An integr<strong>at</strong>ion of Grassmann variables can be defined by<br />

∫<br />

dη i (a + bη i ) = b<br />

for arbitrary complex numbers a,b.<br />

In fermionic field theories we have Grassmann fields, which associ<strong>at</strong>e Grassmann<br />

variables whith every space-time point. With the above rules of integr<strong>at</strong>ion<br />

and differenti<strong>at</strong>ion one can easily derive the important result<br />

⎡ ⎤<br />

∫<br />

∏ N N∑<br />

fermion fields: d¯η k dη k exp ⎣− ¯η i F ij η j<br />

⎦ = det(F) (234)<br />

k=1<br />

which is to be compared to<br />

⎡<br />

∫<br />

∏ N<br />

boson fields: du k exp ⎣−<br />

k=1<br />

N∑<br />

i,j=1<br />

i,j=1<br />

⎤<br />

u i F ij u j<br />

⎦<br />

1<br />

∼ √ (235)<br />

det(F)<br />

Thus the difference between Fermion- and Boson fields is just the power of<br />

det(F).<br />

13.4.2 Fermionic p<strong>at</strong>h integral<br />

We write down now the Fermionic action in descretized form and proceed similarly<br />

to the case of bosonic fields.<br />

∫ ( 4∑<br />

)<br />

SE F = d 4 x ¯ψ(x) γµ E ∂ µ + M ψ(x)<br />

µ=1<br />

We rescale<br />

∂ µ ψ α (x) → ψ α,j+1 µ−ψ α,j−1µ<br />

2a<br />

˜ψ α,j = a 3/2 ψ αj<br />

˜¯ψα,j = a 3/2 ¯ψαj<br />

˜M = aM<br />

and we obtain the discretized action<br />

S F E = ∑<br />

and<br />

α,β,j,k<br />

˜¯ψ αj K αβ<br />

jk ˜ψ βk<br />

4∑<br />

K αβ<br />

jk = 1<br />

2 (γ [ ]<br />

µ) αβ δj+1µ,k − δ j−1µ,k + ˜Mδjk δ αβ<br />

µ=1<br />

With the experience we have, we can do now the p<strong>at</strong>h integral. We have to go<br />

to the same steps as we did in the scalar field theory case and determine the<br />

235


Greens functions. If we do this we find expressions very similar to those we had<br />

before:<br />

< 0| ˜ψ α,r˜¯ψβ,s |0 >= ( K −1) αβ<br />

rs<br />

the inverse m<strong>at</strong>rix is defined by<br />

∑<br />

qλ<br />

K αλ<br />

rq<br />

(<br />

K<br />

−1 ) λβ<br />

qs = δ rsδ αβ<br />

r,s=Euklidean 4-index<br />

We can calcul<strong>at</strong>e it by going to the Fourier space and we find for our fermion<br />

field<br />

[<br />

∫ +π<br />

(K −1 ) αβ d 4ˆk −i ∑ 4<br />

µ=1 γ µ sin(ˆk µ ) + ˜M<br />

]<br />

[ ]<br />

αβ<br />

rs =<br />

−π (2π) 4 ∑ (ˆkµ )<br />

4<br />

µ=1 sin2 + ˜M<br />

exp iˆk(r − s) (236)<br />

2<br />

compared to the expression of the scalar field eq.(233)<br />

[ ]<br />

∫ +π<br />

(K −1 d 4ˆk exp iˆk(r − s)<br />

) rs =<br />

(2π) 4<br />

−π<br />

4 ∑ 4<br />

µ=1 sin2 ( ˆkµ<br />

2<br />

)<br />

+ ˜M 2<br />

Although both expressions for fermions and bosons look r<strong>at</strong>her similar there is a<br />

fundamental difference. Actually in the continuum limit the fermion expression<br />

(236) goes not into the usual fermion Greens function. which is to be seen if<br />

one compares the denomin<strong>at</strong>ors. If we go back for (236) to non-scaled variables<br />

we have the fol<strong>low</strong>ing relevant expressions<br />

sin (ak µ )<br />

a<br />

(fermion)<br />

vs.<br />

( )<br />

2sin a kµ 2<br />

a<br />

(boson)<br />

The argument of the sine-function of the boson case is only half of th<strong>at</strong> of the<br />

fermion case. As we shall see immedi<strong>at</strong>ely, this makes a big difference and is the<br />

origin of the so called ”fermion doubling” problem, which has hampered l<strong>at</strong>tice<br />

calcul<strong>at</strong>ions with fermions for decades and still provides problems. The relevant<br />

intervall is [−π,+π] and we can plot it<br />

Consider a case with M = 0, which is plotted in the figure.Then the propag<strong>at</strong>or<br />

of the field should have a zero <strong>at</strong> k = 0, sinc the pole of the propag<strong>at</strong>or<br />

indic<strong>at</strong>es the mass of the particle (pol mass). The bosonic case has indeed only a<br />

zero <strong>at</strong> k = 0 representing the physical particle. This pole exists in the fermionic<br />

case as well. However there are poles also <strong>at</strong> the edges of the Briolloin zone <strong>at</strong><br />

k µ = ± π a<br />

due to the periodicity of the denomin<strong>at</strong>or. These two poles correspond<br />

to a particle with mass ( π<br />

a) 2<br />

and do not correspond to a physical particle. So<br />

the discretized version of the fermionic action corresponds (in four dimensions)<br />

to 2 4 = 16 particles r<strong>at</strong>her than 1 particle. This problem is called euphemistically<br />

”fermion doubling”. These 16 fermions are pure discretiz<strong>at</strong>ion artefacts,<br />

because of which the fermion theory does not have a proper continuum limit, if<br />

one does not modify it properly.<br />

236


13.4.3 Wilson fermions, quenched l<strong>at</strong>tice. etc.<br />

There have several remedies been considered to get rid of the problem of fermion<br />

doubling. The first relevant one was the introduction of Wilson femions. Wilson<br />

added an ”irrelevant” term to the fermion action, which cured the problem i<br />

the continuum limit.<br />

S F,Wilson = S F − ar<br />

2<br />

∫<br />

d 4 x ¯ψ(x)∂ µ ∂ µ ψ(x)<br />

with r a dimensionless quantity. The discretized version of the action becomes<br />

now<br />

S F,Wilson = ∑ ˜¯ψ αj W αβ ˜ψ jk βk (237)<br />

with<br />

W αβ<br />

jk = 1 2<br />

4∑<br />

µ=1<br />

α,β,j,k<br />

[ ]<br />

(γµ − r) αβ δ j+1µ,k − (γ µ + r) αβ δ j−1µ,k +( ˜M +4r)δjk δ αβ (238)<br />

The corresponding inverse m<strong>at</strong>rix is now<br />

[<br />

∫ +π<br />

(W −1 ) αβ d 4ˆk −i ∑ 4<br />

µ=1 γ µ sin(ˆk µ ) + ˜M(ˆk)<br />

]<br />

[ ]<br />

αβ<br />

rs =<br />

−π (2π) 4 ∑ (ˆkµ )<br />

4<br />

µ=1 sin2 + ˜M(ˆk)<br />

exp iˆk(r − s)<br />

2<br />

with<br />

˜M(ˆk) = ˜M + 2r ∑ µ<br />

sin 2 (ˆk µ<br />

2 )<br />

237


going to unscaled quantities with ˜M = aM and k = ˆk/a we get<br />

M(k) = M + 2r<br />

a<br />

∑<br />

µ<br />

sin 2 ( ak µ<br />

2 )<br />

In the continuum limit a → 0 the second term vanishes for any k µ ≠ π a as<br />

one sees <strong>at</strong> the expansion in powers of akµ<br />

2<br />

,where all powers are high enough<br />

to compens<strong>at</strong>e the 1/a in front. In these cases M(k) approaches the given M<br />

for vanishing a. However the mass M(k) diverges for k µ = π a<br />

where for all a<br />

the sinus gives a finite value, i.e. one. Thus the doublers <strong>at</strong> the border of the<br />

Brillouin zone become infinite heavy in the continuum limit, so th<strong>at</strong> they do not<br />

disturb the rest of the l<strong>at</strong>tice. At a first glance this looks nice, there are however<br />

problems with these Wilson-Femions. For a genuin vanishing fermion mass the<br />

original fermion action was chirally symmetric. This is no longer the case due<br />

to the additional Wilson term, since this one is proportional to ¯ψ(x)∂ µ ∂ µ ψ(x)<br />

which behaves like a mass term under chiral transform<strong>at</strong>ions, and we know, th<strong>at</strong><br />

mass terms destroy chiral symmetry. Thus the use of Wilson fermions is not<br />

very suitable for problems where chiral symmetry is important. Unfortun<strong>at</strong>ely<br />

all <strong>low</strong> energy hadronic physics is of this sort. Furthermore Wilson fermions<br />

show large discretiz<strong>at</strong>ion errors, which however in practice can be compens<strong>at</strong>ed<br />

by adding further terms.<br />

There are also other and more advanced suggestions to cure the doubling<br />

problem: Kogut-Susskind fermions (or staggered fermions), and Ginsparg-Wilson<br />

fermions (or domain wall fermions). We will not discuss these points.<br />

13.5 Gauge Fields on the L<strong>at</strong>tice<br />

We derive the formalism first for abelian gauge fields and then generalize in a<br />

straight forward way to non-abelian ones.<br />

13.5.1 Abelian gauge fields: QED<br />

We have been considering simple scalar boson fields, then fermion fields, and<br />

now we consider a field theory which has a gauge degree of freedom. First we<br />

take an abelian gauge theory and repe<strong>at</strong> quickly the basics. The action for a<br />

free Dirac field is ∫<br />

S F = d 4 xψ(x)[iγ µ ∂ µ − M]ψ(x)<br />

it becomes invariant under local U(1) transform<strong>at</strong>ions by introducing a fourvector<br />

potential A µ (x) changing to the covariant deriv<strong>at</strong>ive<br />

D µ ψ(x) = (∂ µ + ieA µ (x))ψ(x)<br />

THen we obtain for the action<br />

∫<br />

S F = d 4 xψ(x)[iγ µ D µ − M]ψ(x)<br />

238


which is now invariant under simultaneous transform<strong>at</strong>ions<br />

with<br />

Under this transform<strong>at</strong>ion we have<br />

ψ(x) → ψ ′ (x) = G(x)ψ(x)<br />

¯ψ(x) → ¯ψ ′ (x) = ψ(x)G −1 (x)<br />

A µ (x) → A ′ µ(x) = A µ (x) + i e ∂ µα(x)<br />

G(x) = exp[−iα(x)]<br />

D µ → G(x)D µ G −1 (x)<br />

We need also the action of the kinetic term of the gauge field<br />

S A = − 1 ∫<br />

d 4 xF µν F µν<br />

4<br />

with<br />

F µν = ∂ µ A ν − ∂ ν A µ<br />

which behaves under the gauge transform<strong>at</strong>ion as<br />

F µν → F ′ µν = G(x)F µν G −1 (x)<br />

Going to Euklidean time we get<br />

∫ [ 4∑<br />

]<br />

SE F = d 4 xψ(x) γ µ D µ + M ψ(x)<br />

µ=1<br />

S A E = + 1 4<br />

∫<br />

d 4 x<br />

4∑<br />

µ,ν=1<br />

F µν F µν<br />

So far we have repe<strong>at</strong>ed simple abelian gauge theory (in the continuum limit)<br />

rewritten in the Euklidean space.<br />

To formul<strong>at</strong>e a gauge theory on the l<strong>at</strong>tice one does not proceed by taking<br />

the above Lagrangean and place all the fields on the l<strong>at</strong>tice points. Instead one<br />

proceeds in the way th<strong>at</strong> one takes the discretized Fermion theory and gauges<br />

it directly in the discretized form, i.e. on the l<strong>at</strong>tice. Thus, in order to obtain<br />

gauge ivariance we are led to modify the expression<br />

S F,Wilson = ∑ j<br />

( ˜M + 4r)˜¯ψj ˜ψj + 1 2<br />

∑<br />

j<br />

4∑<br />

˜¯ψ j (γ µ − r) ˜ψ j+1µ − ˜¯ψj+1µ (γ µ + r) ˜ψ j<br />

µ=1<br />

(239)<br />

There we realize immedi<strong>at</strong>ely something special: Consider the term ˜¯ψj (γ µ −<br />

r) ˜ψ j+1µ , which is corresponding to something like ¯ψ(x)ψ(y) . This is in fact a<br />

non-local product which, however, is not gauge invariant, but transforms as<br />

¯ψ(x)ψ(y) → ¯ψ ′ (x)ψ ′ (y) = ¯ψ(x)G −1 (x)G(y)ψ(y)<br />

239


C<br />

x<br />

y<br />

Actually such terms will occur, if we write down a kinetic term in l<strong>at</strong>tice field<br />

theory. Therefore, we need m<strong>at</strong>rices U(x,y) ∈ SU(N) which transform as<br />

U(x,y) −→ G(x)U(x,y)G −1 (y), such th<strong>at</strong> ¯ψ(x) · U(x,y)ψ(y) would be invariant.<br />

There is a solution to this problem. Take a arbitrary p<strong>at</strong>h C from x to y<br />

and define<br />

U(x,y; C) ≡ exp<br />

[<br />

ie<br />

∫ y<br />

where the integral is taken along the p<strong>at</strong>h C.<br />

x<br />

] [∫ sy<br />

A µ (z)dz µ = exp dsA µ (c(s)) dc ]<br />

µ(s)<br />

, (240)<br />

s x<br />

ds<br />

Then U(x,y; C) transforms as desired, i.e. it is inself an element of the gauge<br />

group U(1), i.e. it transforms under the gauge transform<strong>at</strong>ion of the A-field as<br />

since<br />

U ′ (x,y; C) = exp{ie<br />

or (q.e.d.):<br />

U(x,y) → U ′ (x,y) = G(x)U(x,y)G −1 (y)<br />

∫ y<br />

x<br />

[A µ (z)+ 1 e ∂ µα(z)]dz µ } = exp{ie<br />

U ′ (x,y; C) = exp [−iα(x)] exp{ie<br />

∫ y<br />

Due to this it fulfills the goal th<strong>at</strong> we have the invariance<br />

x<br />

∫ y<br />

x<br />

A µ (z)dz µ } exp [+iα(y)]<br />

A µ (z)dz µ + iα(y) − iα(x)}<br />

¯ψ(x)U(x,y)ψ(y) → ¯ψ(x)G −1 (x)G(x)U(x,y)G −1 (y)G(y)ψ(y) = ¯ψ(x)U(x,y)ψ(y)<br />

The expression in eq.(240) is called gauge link or parallel transporter. The<br />

l<strong>at</strong>ter expression is in analogy to similar objects in differential geometry, which<br />

map vectors from one point to another along curves. The parallel transporters<br />

depend not only on the points x and y but also on the chosen curve C. They<br />

obey the composition rule<br />

U(x,y; C) = U(x,u; C 1 ) · U(u,y; C 2 ),<br />

240


where the p<strong>at</strong>h C is split into two parts C 1 and C 2 .<br />

We will apply the above gauge procedure to the Wilson action for Fermions<br />

eq.(239) based on (237,238). Thus we change<br />

˜¯ψ j (γ µ − r) ˜ψ j+1µ<br />

˜¯ψ j+1µ (γ µ + r) ˜ψ j<br />

→<br />

→<br />

˜¯ψ j (γ µ − r)U(j,j + 1 µ ) ˜ψ j+1µ<br />

˜¯ψ j+1µ (γ µ + r)U(j + 1 µ ,j) ˜ψ j<br />

with U(j +1 µ ,j) = U(j,j +1 µ ) † . Hence the discretized Fermionic action is now<br />

invariant under simultaneous transform<strong>at</strong>ions<br />

˜ψ j → G j ˜ψj (241)<br />

˜¯ψ j → ˜¯ψj G −1<br />

j<br />

U(j,j + 1 µ ) → G j U(j,j + 1 µ )G −1<br />

j+1 µ<br />

U(j + 1 µ ,j) → G j+1µ U(j + 1 µ ,j)G −1<br />

j<br />

where G j is an arbitrary local gauge transform<strong>at</strong>ion specified <strong>at</strong> each l<strong>at</strong>tice<br />

site j (being a 4-vector). Thus the Fermionic action is written in terms of<br />

˜¯ψ j , ˜ψ j ,U(j,j +1 µ ),U(j +1 µ ,j). Obviously the fermion fields are defined <strong>at</strong> the<br />

l<strong>at</strong>tice sites, wheras the gauge fields are so called ”link variables”, which are<br />

definde on links connecting the l<strong>at</strong>tice sites. We can picture them<br />

It is useful to have a slightly condensed way of writing these link variables:<br />

link from site j in µ-direction by 1 step: U µ (j) = U(j,j + 1 µ )<br />

link to site j in − µ-direction by 1 step: U † µ(j) = U(j + 1 µ ,j) = U −µ (j + 1 µ )<br />

Using the trapezoidal rule for integr<strong>at</strong>ion [ ∫ b<br />

a f(x)dx ≈ 1 2<br />

(f(b) + f(a)) (b − a)]<br />

and with the definition of the gauge link U(x,y) = exp [ ie ∫ y<br />

x A µdz µ] in eq.(240)<br />

we can write for small a<br />

U µ (j) = exp<br />

[<br />

ie a ]<br />

2 (A µ,j + A µ,j+1µ )<br />

(242)<br />

When we will analyse the discretized and gauge invariant Wilson action<br />

we will find out th<strong>at</strong> the so called ”plaquettes” are the relevant quantities to<br />

describe the gauge field. The simplest one is given by<br />

241


with in the end (in fact we have 4 directions ond not only two as in the plot)<br />

P µν = U µ (j)U ν (j + 1 µ )U † µ(j + 1 ν )U † ν(j) (243)<br />

The reason for the importance of the plquettes is simple and very basic for<br />

the l<strong>at</strong>tice gauge theory: The plaquettes play the role of the guage field tensors<br />

F µν , as we will show immedi<strong>at</strong>ely.<br />

The plaquette is gauge invariant because of eq.(241) and hence<br />

U µ (j) → G j U µ (j)G −1<br />

j+1 µ<br />

and U µ(j) † → G j+1µ U µ(j)G † −1<br />

j<br />

using eq.(242) one obtains for P µν<br />

P µν = exp{ie a 2<br />

[<br />

Aµ,,j + A µ,,j+1µ + A ν,j+1µ+A ν,,j+1µ+1 ν<br />

− A µ,,j+1µ+1 ν−A µ,,j+1ν−A ν,j+1ν−A ν,j<br />

]<br />

}<br />

To interpret the plaquette use now the fact th<strong>at</strong> we always consider small l<strong>at</strong>tice<br />

constant a, such th<strong>at</strong><br />

A µ,j − A µ,j+1ν → −a∂ ν A µ,j<br />

such th<strong>at</strong><br />

F µν (j) = 1 a<br />

[<br />

(Aµ,j − A µ,j+1ν ) − (A ν,j − A µ,j+1µ ) ]<br />

and by Taylor expansion to first order in a<br />

P µν (j) → exp [ iea 2 (∂ µ A ν,j − ∂ ν A µ,j ) + O(a 4 ) ]<br />

= 1 + iea 2 F µν,j − a4 e 2<br />

2 F µν,jF µν,j + O(a 6 )<br />

where the last line ist obtained by expanding the exponential function. Apparently<br />

the plaquette starting from the same point j but with all arrows inverted<br />

242


yields instead of P µν (j) now a P νµ (j), which is identical to P µν (j) except in the<br />

sign of the imaginary part. Hence adding both these plaquettes yields only a<br />

real part if one neglects of order O(a 6 ). And this real part is exactly the kinetic<br />

energy of the gauge field 1 4 F µν,jF µν,j . Thus one obtains an important result:<br />

One has to sum up the contributions over all plaquettes to get the action of the<br />

gauge field.<br />

So finally the l<strong>at</strong>tice action for gauge fields becomes<br />

S A = 1 ∑ ∑<br />

[<br />

e 2 1 − 1 (<br />

Pµν (j) + P †<br />

2<br />

µν(j) )]<br />

j<br />

P<br />

j<br />

µ


This is in fact not a complic<strong>at</strong>ion since we had written the plaquettes (243)<br />

already in coorrect ordering: P µν = U µ (j)U ν (j + 1 µ )U µ(j † + 1 ν )U ν(j), † see also<br />

figure. Furthermore one has to take into account th<strong>at</strong> for the quark fields an<br />

additional sum appears since they have an additional colour index. If one writes<br />

things carefully down the l<strong>at</strong>tice action of QED (244) modifies into<br />

S <strong>QCD</strong><br />

[<br />

U,ψ, ¯ψ] =<br />

6<br />

g 2 ∑<br />

with<br />

+ ∑ j,ρ<br />

+ 1 2<br />

j<br />

∑<br />

P<br />

[<br />

1 − 1 6 Tr ( P(j) + P † (j) )] (245)<br />

( ˜M (ρ)<br />

+ 4r) ˜ψ j<br />

˜ψ (ρ)<br />

j<br />

∑<br />

4∑<br />

j,ρ,σ µ=1<br />

TrP µν = ∑ U (κ,ρ)<br />

[ ˜ψ(ρ) j (γ µ − r)U µ (ρ,σ) (σ) (ρ)<br />

(j) ˜ψ j+1 µ<br />

− ˜ψ j+1 µ<br />

(γ µ + r)U µ (ρ,σ)† (j)<br />

µ (j)U ν<br />

(ρ,σ)<br />

(j + 1 µ )U µ (σ,τ)† (j + 1 ν )U ν<br />

(τ,κ)† (j)<br />

Altogether the non-abelian theory on the l<strong>at</strong>tice is just technically more complic<strong>at</strong>ed<br />

than the abelian one without bringing new conceptual problems. The<br />

trace in the pure Yang-Mills part S l<strong>at</strong>t denotes taking the trace of U µ i.e. the<br />

sum of the 3 diagonal elements. S l<strong>at</strong>t sums over all plaquettes of all orient<strong>at</strong>ions<br />

on the l<strong>at</strong>tice.<br />

Actually the coupling constnt g is the so called ”bare coupling constant”. It<br />

is dependent on the l<strong>at</strong>tice spacing since this is connected with a momentum<br />

cut-off. Usually one does not write g(a) but β. This is<br />

β = 6 g 2 g = g(a) = bare coupling constant (246)<br />

Actually the β the single input parameter for a <strong>QCD</strong> calcul<strong>at</strong>ion involving only<br />

gluon fields (whether on the l<strong>at</strong>tice or not). Notice th<strong>at</strong> the l<strong>at</strong>tice spacing is<br />

not explicit anywhere, and we do not know its value until after the calcul<strong>at</strong>ion.<br />

(This is a difference from the quantum mechanical example of the previous<br />

section where we had to choose and input a value for a.) This is explained<br />

in the next subsection. The value of the l<strong>at</strong>tice spacing depends on the bare<br />

coupling constant, or vice versa, the value of the bare coupling constant depends<br />

on the l<strong>at</strong>tice spacing. Typical values of β for current l<strong>at</strong>tice calcul<strong>at</strong>ions using<br />

the Wilson plaquette action are β ≈ 6. One should note th<strong>at</strong> the continuum<br />

limit is reached for β → ∞.<br />

13.5.3 Continuum limit<br />

Although in the la<strong>at</strong>tice formul<strong>at</strong>ion of the quantum field theory all quantities<br />

have precise m<strong>at</strong>hem<strong>at</strong>ical meaning, one must not foget the goal is to define<br />

a quantum system incontinuous space-time, since there the experiments are<br />

done. For this one must proceed to the continuum limit, letting the l<strong>at</strong>tice<br />

spacing a go to zero. Here one should note th<strong>at</strong> the l<strong>at</strong>tice spacing a plays<br />

]<br />

(σ) ˜ψ j<br />

244


a double role. First it is there in order to have a discrete theory such th<strong>at</strong><br />

one can handle it on the computer. Furthermore the finite a means also th<strong>at</strong><br />

one can handle only finite momenta smaller than Λ cutoff ∼ 1/a in the l<strong>at</strong>tice,<br />

which means, th<strong>at</strong> all quantities get autom<strong>at</strong>ically regularized with a finite<br />

momentum cut-off Λ cutoff . In fact the l<strong>at</strong>tice is a non-perturb<strong>at</strong>ive way of<br />

regularizing and renormaliz<strong>at</strong>ion. In this context it is important to understand<br />

the connection of β = 6<br />

g<br />

of eq.(246) to physics in n<strong>at</strong>ure and to the l<strong>at</strong>tice<br />

2<br />

spacing a in the fol<strong>low</strong>ing way (we do this in the jargon of <strong>QCD</strong> including quarks<br />

although we formally consider a pure gluon theory in the first place): The g is<br />

the dimensionless bare coupling constant of <strong>QCD</strong>, it is a number and it is the<br />

only quantity in the Lagrangean which tells anything about the magnitude of<br />

any physical quantity to be calcul<strong>at</strong>ed. Suppose we want to calcul<strong>at</strong>e the mass<br />

M N of the nucleon. We do it <strong>at</strong> finite ´a and hence we have M N = M N (g,a),<br />

however in the limit a −→ 0 we should have M N (g,a −→ 0) = 938MeV , <strong>at</strong><br />

least this value should be finite [i.e. not zero and not infinite, such th<strong>at</strong> by a<br />

proper normaliz<strong>at</strong>ion one could adjust it to 938MeV]. To achieve this one must<br />

have the g a function of a, i.e. g = g(a), and in the limit a −→ 0 the function<br />

g(a) should be such th<strong>at</strong> ”M N (g(a),a −→ 0) = finite” is achieved. This<br />

fe<strong>at</strong>ure should hold for any physical observable, namely th<strong>at</strong> it has a finite limit<br />

for a → 0.Actually several functions g(a) can be thought of, which do th<strong>at</strong>.<br />

However for a → 0 we also have Λ cutoff → ∞, and in this limit perturb<strong>at</strong>ive<br />

<strong>QCD</strong> holds. Thus one can read off the g(a) from a few Feynman diagrams of<br />

first order in g. This is possible since <strong>QCD</strong> is an asymptotically free theory<br />

whose effective coupling constant ḡ(µ) or α s (µ) (132) goes logarithmically to<br />

zero for µ ∼ Λ cutoff ∼ 1/a → ∞. For small g first order perturb<strong>at</strong>ion theory is<br />

sufficient and the rel<strong>at</strong>ion between g and a can be shown to be governed by the<br />

equ<strong>at</strong>ion<br />

a dg(a)<br />

da = β 0g 3 (247)<br />

where β 0 is the famous beta-function (129)<br />

β 0 = 11 3 N 1<br />

c<br />

16π 2<br />

with N c equal the number of colours in the theory. The differential equ<strong>at</strong>ion<br />

(247) is solved by<br />

a = 1 (<br />

exp − 1 )<br />

Λ L 2β 0 g 2 (248)<br />

We can rewrite<br />

g(a) 2 1<br />

= −<br />

2β 0 log(Λ L a)<br />

The Λ L is unknown in principle. It has to be adjusted for one observable to<br />

its physical value, e.g. for the nucleon mass 938MeV: Then it is fixed and<br />

one can calcul<strong>at</strong>e with this Λ L and the above g(a) other physical quantities.<br />

Typical values of β = 6/g 2 for current l<strong>at</strong>tice calcul<strong>at</strong>ions (e.g. using the Wilson<br />

plaquette action to get the area law and the quark-antiquark potential) are<br />

245


β ≈ 6. This corresponds to a ≈ 0.1fm. Smaller values of β give coarser l<strong>at</strong>tices,<br />

larger ones, finer l<strong>at</strong>tices. 3<br />

In fact one should go even further. One knows from renormaliz<strong>at</strong>ion theory<br />

the asymptotic (renormaliz<strong>at</strong>ion scale µ −→ ∞) expression of any observable.<br />

E.g. for the nucleon mass one has<br />

8π 2 ]<br />

or expressed in terms of β :<br />

M N = C a exp [−<br />

M N a = C exp<br />

11<br />

3 N cg 2 (a)<br />

] [− 4π2 β<br />

11N c<br />

This functional dependence on β one can in principle explicitely check by performing<br />

calcul<strong>at</strong>ions for various β, and in fact one should do so to see if one is<br />

really close enough to the continuum limit to al<strong>low</strong> comparison of l<strong>at</strong>tice results<br />

with experiment. However the values of β and its range are mostly too small<br />

by technical reasons and hence this check is not always done in a convincing<br />

way. Usually people prefer to plot r<strong>at</strong>ios of observables in dependence on β, and<br />

they proceed in the fol<strong>low</strong>ing way: One chooses a value of β and the number of<br />

l<strong>at</strong>tice points, e.g. N 4 , and one performs a calcul<strong>at</strong>ion for a correl<strong>at</strong>ion function<br />

corresponding e.g. to the nucleon mass. The only input is this β and the<br />

number of l<strong>at</strong>tice sites N. In fact one has not to specify a l<strong>at</strong>tice spacing a, it is<br />

always finite and in fact equal ONE. If the l<strong>at</strong>tice is large enough the correl<strong>at</strong>ion<br />

function <strong>at</strong> large T is given by exp (−M N T) = exp (−[M N a]T/a) , where T/a<br />

is the number of l<strong>at</strong>tice points N and M N a is a dimensionless number. [This is<br />

basically the only thing you can calcul<strong>at</strong>e, namely dimensionless numbers.] Now<br />

one does this also for another observable mass, e.g. for the ρ-meson mass, and<br />

one can extract the r<strong>at</strong>io MNa<br />

M ρa = MN<br />

M ρ<br />

. One does this for several increasing values<br />

of β and if beyond a certain β one obtains for the r<strong>at</strong>io MN<br />

M ρ<br />

no β-dependence<br />

any more then one assumes th<strong>at</strong> one is in the asymptotic region and close to<br />

the continuum limit. In principle for sufficiently large β one is always in the<br />

asymptotic region, however in practice it may happen th<strong>at</strong> one cannot perform<br />

such a calcul<strong>at</strong>ion since the number of l<strong>at</strong>tice points N is limited and one may<br />

not be in the region where the exponential form of the correl<strong>at</strong>ion function holds<br />

already.<br />

If one goes beyond pure glue theory (Yang Mills) and includes quarks, i.e.<br />

performs a full <strong>QCD</strong> calcul<strong>at</strong>ion on the l<strong>at</strong>tice, one can rewrite the above formulae<br />

assuming now in eq.(248) a β 0 = ( 11<br />

3 N c − 2 3 N f)<br />

/16π 2 . If one identifies the<br />

inverse of the l<strong>at</strong>tice spacing with the renormaliz<strong>at</strong>ion scale µ [i.e. if one identifies<br />

g(1/a) with ḡ(µ)] and if one identifies the Λ L with Λ <strong>QCD</strong> one obtains after<br />

3 Actually one can go beyond first order perturb<strong>at</strong>ion theory for the rel<strong>at</strong>ion between a and<br />

g. If one does th<strong>at</strong> one obtains more complic<strong>at</strong>ed expressions (see Rohde sect. 9.2).<br />

246


a direct rewriting from eq.(248) the expression for the strong <strong>QCD</strong>-coupling<br />

constant (132). In fact these identific<strong>at</strong>ions are correct and correspond to the<br />

role of a as a regulariz<strong>at</strong>ion parameter.<br />

Altogether one has to remember th<strong>at</strong> the fol<strong>low</strong>ing limiting cases have to be<br />

under control for a l<strong>at</strong>tice simul<strong>at</strong>ion:<br />

Continuum limit a → 0<br />

L<strong>at</strong>tice size V → ∞<br />

Euklidean time T → ∞<br />

Light quark masses or pion mass agains the physical value m π → 139MeV<br />

In addition one has to cope with<br />

Fermion doubling<br />

13.6 Practical applic<strong>at</strong>ions of l<strong>at</strong>tice <strong>QCD</strong><br />

13.6.1 Metropolis algorithm in pure gauge theory<br />

The applic<strong>at</strong>ion of the Metropolis algorithm to a pure Yang-Mills action is<br />

completely analogous to simple quantum mechanics. The only difference is th<strong>at</strong><br />

we now have link variables, which are SU(3) elements. The upd<strong>at</strong>e process<br />

here means to propose a change in a link variable by multiplying it by a random<br />

SU(3)-m<strong>at</strong>rix and then to calcul<strong>at</strong>e the the difference in the action and to accept<br />

it with some probability.<br />

The algorithm for randomizing the link variable U µ (j) <strong>at</strong> the j h site in<br />

direction µ is:<br />

• gener<strong>at</strong>e a random SU(N c )−m<strong>at</strong>rix M. Analog to the 1-dimensional case<br />

in quantum mechanics one can chose in some way the size ζ of this m<strong>at</strong>rix.<br />

The ζ must be a random number, with probability uniformly distributed<br />

between some limits −δ and δ for some given constant δ;<br />

• replace U µ (j) → MU µ (j) and compute the change ∆S in the action caused<br />

by this replacement. Generally only a few terms in the l<strong>at</strong>tice action<br />

involve U µ (j). Since the Yang-Mills lagrangians are local, only these need<br />

to be examined. One should note, th<strong>at</strong> each of the 4 ∗ N 4 links is shared<br />

by 6 plaquettes. S, if one link changes, there are 6 terms in the action<br />

th<strong>at</strong> change.<br />

• if ∆S < 0 (the action is reduced) retain this new value for U µ (j), and<br />

proceed to the next link;<br />

• if ∆S > 0 accept the new value MU µ (j) with probability exp(−∆S).<br />

This means: gener<strong>at</strong>e a random number η unformly distributed between 0<br />

and 1; retain the new value for U µ (j) if exp(−∆S) > η, otherwise restore<br />

the old value; proceed to the next link. Acutally, to gener<strong>at</strong>e a random<br />

SU(3)-m<strong>at</strong>rix on the computer is a non-trivial task, since the computer<br />

cannot gener<strong>at</strong>e a real random number. There are sophistic<strong>at</strong>ead techniques<br />

to do th<strong>at</strong>, which we will not discuss here.<br />

247


13.6.2 Pure gauge calcul<strong>at</strong>ions<br />

Wilson loop, area law. string constant Pure gauge calcul<strong>at</strong>ions are obtained<br />

by the previous formalism by putting all fermion fields to zero i.e. using<br />

as action<br />

S Y M [U] = 6 ∑ ∑<br />

[<br />

g 2 1 − 1 6 Tr ( P + P †)] (249)<br />

j<br />

P<br />

which corresponds to a pure Yang-Mills theory. The most interesting quantity<br />

to calcul<strong>at</strong>e is the Wilson loop given in the figure<br />

To ‘measure’ this in pure gluon <strong>QCD</strong>, we gener<strong>at</strong>e configur<strong>at</strong>ions of gluon<br />

fields with probability e −SY M where S Y M is the discretis<strong>at</strong>ion of the pure gluon<br />

<strong>QCD</strong> action given by Equ<strong>at</strong>ion (249). On each of these configur<strong>at</strong>ions we calcul<strong>at</strong>e<br />

the r ×t Wilson loop, averaging over all positions of it on the l<strong>at</strong>tice. We<br />

then average over the results on each configur<strong>at</strong>ion in the ensemble to obtain a<br />

final answer with st<strong>at</strong>istical error, for lots of values of r and t. We have<br />

W(r,t) = 1 ∫<br />

DU<br />

1<br />

Z < 0|1 3 Tr(UU..U) 3<br />

rt|0 >=<br />

Tr(UU..U) ∫<br />

rte −SY M<br />

DUe<br />

−S Y M<br />

Typically we need many hundreds of configur<strong>at</strong>ions in an ensemble for a small<br />

st<strong>at</strong>istical error <strong>at</strong> large r and t.<br />

This quantity is interesting since one can show (only plausible arguments<br />

given be<strong>low</strong>) th<strong>at</strong> it is rel<strong>at</strong>ed to the st<strong>at</strong>ic potential between a heavy quark and<br />

a heavy anti-quark <strong>at</strong> a distance r in the limit of infinite mass of the quarks. In<br />

fact, it is interesting to note th<strong>at</strong> one can calcul<strong>at</strong>e this quark-quark-potential<br />

without having explicit quarks in the l<strong>at</strong>tice. We have<br />

W(r,t) ∝ exp [−V (r)t]<br />

The practical calcul<strong>at</strong>ions yield curves like<br />

In the figure a parameter R 0 appears. This is the so called Sommer-scale,<br />

which is explained be<strong>low</strong>, it is about 0.5 fṁ. It is The β = 6<br />

g 2 , where g is the bare<br />

248


coupling constant as it appears in the <strong>QCD</strong> lagrangean. In the limited region<br />

of coupling constant considered all curves lie on top of each other. In the region<br />

where V (r) ∝ r we see the so called Wilsons area law, which is interpreted as a<br />

sign of confinement.<br />

W(r,t) ∝ exp [−σA]<br />

Here A = rt is the area of the Wilson loop and σ is the so called string constant.<br />

In general we have [ ] W(r,t)<br />

log = aV (r)<br />

W(r,t + a)<br />

At small r we have a Coulomb like behaviour, i.e. altogether we have<br />

V (r) = − b r + σr<br />

The ensemble average of O is rel<strong>at</strong>ed to the heavy quark potential by similar<br />

arguments to those used for the oper<strong>at</strong>or x(t 1 )x(t 1 ) in section about harmonic<br />

oscill<strong>at</strong>or. One end of the Wilson loop cre<strong>at</strong>es a set of eigenst<strong>at</strong>es of the Hamiltonian<br />

th<strong>at</strong> are based on a massive quark-antiquark pair. These eigenst<strong>at</strong>es<br />

are produced with different amplitudes by the Wilson loop oper<strong>at</strong>or and have<br />

different <strong>energies</strong>. In this case there is no kinetic energy, so the <strong>energies</strong> are<br />

those of the heavy quark potential. The different eigenst<strong>at</strong>es propag<strong>at</strong>e for time<br />

249


t and, if t is large, the ground st<strong>at</strong>e eventually domin<strong>at</strong>es.<br />

W(r,t) = Ce −aV (r)t + C ′ e −aV ′ (r)t + ....<br />

By fitting the results as a function of the euklidean time length, t, in l<strong>at</strong>tice<br />

units, the heavy quark potential in l<strong>at</strong>tice units, aV (r), is obtained. The V ′ is<br />

some kind of excit<strong>at</strong>ion of the potential which we will not be interested in here.<br />

The heavy quark potential <strong>at</strong> short distances should behave perturb<strong>at</strong>ively and<br />

take a Coulomb form. At large distances we expect a ‘string’ to develop which<br />

confines the quark and antiquark and gives a potential which rises linearly with<br />

separ<strong>at</strong>ion. We can therefore fit the l<strong>at</strong>tice potential to the form<br />

aV (ρ = ra) = − 4 α s (ρ)<br />

+ σa 2 r +<br />

3 r<br />

˜C<br />

where ˜C is a ‘self-energy’ constant th<strong>at</strong> appears in the l<strong>at</strong>tice calcul<strong>at</strong>ion. If<br />

the results for aV (R) are plotted against R, the slope <strong>at</strong> large r is the ‘string<br />

tension’, σ, in l<strong>at</strong>tice units, i.e. σa 2 . Phenomenological models of the heavy<br />

quark potential give values for √ σ of around 440 MeV. Using this value for σ<br />

and the result from the l<strong>at</strong>tice of σa 2 , gives a value for a. This is often quoted<br />

as a value for a −1 in GeV. Note th<strong>at</strong> a −1 in GeV = 0.197/(ainfm).<br />

Figure 1: The heavy quark potential in units of the parameter, r 0 , as a function<br />

of distance, r, also in units of r 0 . The calcul<strong>at</strong>ions were done in quenched (pure<br />

glue) <strong>QCD</strong> <strong>at</strong> a variety of different values of the l<strong>at</strong>tice spacing, corresponding<br />

to the different values of β quoted. (Bali 2000)<br />

250


The results for aV can then be multiplied by a −1 to convert them to physical<br />

units of GeV and, having removed the constant ˜C, the results can be plotted as<br />

a function of the physical distance, r in fm. If discretis<strong>at</strong>ion errors are small,<br />

then results <strong>at</strong> different values of the l<strong>at</strong>tice spacing should be the same. 1<br />

shows results for the heavy quark potential on rel<strong>at</strong>ively fine l<strong>at</strong>tices <strong>at</strong> different<br />

values of β using the Wilson plaquette action for S g (Bali 2000). The V (r) is<br />

not in fact given in GeV here, nor is r in fm, but both are given in terms of a<br />

parameter called r 0 (Sommer 1994). This is the value of r <strong>at</strong> which r 2 ∂V/∂r<br />

= 1.65 and is a commonly used quantity to determine the l<strong>at</strong>tice spacing (or <strong>at</strong><br />

least rel<strong>at</strong>ive l<strong>at</strong>tice spacings), r<strong>at</strong>her than the string tension. The r 0 is not a<br />

physical parameter, and as such is not available in the Particle D<strong>at</strong>a Tables. We<br />

shall see l<strong>at</strong>er th<strong>at</strong> there are good hadron masses to use for the determin<strong>at</strong>ion<br />

of a and these can also be used to determine the value of r 0 . Meanwhile,<br />

r 0 ≈ 0.5fm. The results <strong>at</strong> different values of β lie on top of each other and this<br />

gives us confidence th<strong>at</strong> the discretis<strong>at</strong>ion errors are small.<br />

The fact th<strong>at</strong> The above curve is indeed some sort of potential energy between<br />

quark and anti-quark with the separ<strong>at</strong>ion r. This fe<strong>at</strong>ure can be understood<br />

in a Born-Oppenheimer concept: Consider as example a molecule with<br />

two nuclei <strong>at</strong> distance R with charges Z 1 and Z 2 . An electron moving in the<br />

field gener<strong>at</strong>ed by the two nuclei obbeys the Schroedinger equ<strong>at</strong>ion<br />

{<br />

− 1<br />

}<br />

2m ∇2 r + −e2 Z 1<br />

+ −e2 Z 1<br />

ψ R (r) = E R ψ R (r) (250)<br />

|r| |r − R|<br />

Due to the existence of the electrons and due to the rel<strong>at</strong>ive repulsion of the<br />

two nuclei we have a potential energy between the nuclei and for various R<br />

altogether a potential energy surface,<br />

V (R) = E R + Z 1Z 2 e 2<br />

in which the two nuclei move and can vibr<strong>at</strong>e against each other. Apparently<br />

the two nuclei are hold together due to the existence of E R .The st<strong>at</strong>ionary st<strong>at</strong>es<br />

of the motion of the nuclei is described by<br />

{<br />

− 1 }<br />

2M ∇2 R + V (R) Ψ(R) = EΨ(R)<br />

yielding the ground st<strong>at</strong>e and the excited st<strong>at</strong>es of the molecule.<br />

In the present case we have the analogon ”Two nuclei ⇐⇒Heavy quark<br />

and antiquark” and ”Electrons ⇐⇒Gluons, Glueballes, any gluonic structure<br />

binding the heavy quarks to each other”. The analogue equ<strong>at</strong>ion to eq.(250)<br />

is an equ<strong>at</strong>ion which describes the gluonic field between two heavy quarks of<br />

rel<strong>at</strong>ive distance R. If those are infinitely heavy they can be replaced by two<br />

coloured charges cre<strong>at</strong>ing the gluon field. In this case one really does not need<br />

quarks but only a gluon field facing the two colour charges, which makes life<br />

very much simpler. In the present field theoretical formalism this is not a<br />

Schroedinger equ<strong>at</strong>ion but a p<strong>at</strong>h integral which calcul<strong>at</strong>es the ground st<strong>at</strong>e of<br />

R<br />

251


t<br />

p<br />

1<br />

p<br />

2<br />

eq.(250). To get the ground st<strong>at</strong>e of the gluonic field one has to evalu<strong>at</strong>e p<strong>at</strong>h<br />

integrals with large Euklidean time t, and see th<strong>at</strong> there is some exponential<br />

fall of with increasing t. Th<strong>at</strong> means one has to consider a Wilson loop with<br />

space width r and time width t, as it is shown in the above picture. This yields<br />

then the analogue of E r . In the present case V (r) = E r because there is no<br />

gluon-less interaction between the heavy quarks. Thus the analogue of Z1Z2e2<br />

R<br />

is missing in the <strong>QCD</strong> case.<br />

Actually the above picture is only correct if one considers the quarks as<br />

heavy. For light quarks we have a different situ<strong>at</strong>ion. There instanton effects<br />

appear which exist only in the limit of vanishing quark mass and we have spontaneous<br />

chiral symmetr breaking which does not exist with heavy qurks. As<br />

a consequence with increasing distance r the quarks cre<strong>at</strong>e a pion field around<br />

them which shields them and the potential does not rise any more linearly but<br />

fl<strong>at</strong>tens out.<br />

Glueball-masses In a pure Yang-Mills theory we can calcul<strong>at</strong>e so called plaquettes<br />

correl<strong>at</strong>ions. These are expect<strong>at</strong>ion values of the product of two sp<strong>at</strong>ial<br />

plaquettes p 1 and p 2 , separ<strong>at</strong>ed by a time t as illustr<strong>at</strong>ed in the Fig.<br />

2.<br />

file=plaqucor.eps,height=2cm<br />

Figure 2: Plaquette–plaquette correl<strong>at</strong>ions<br />

In non-abelian l<strong>at</strong>tice gauge theories one finds th<strong>at</strong> these correl<strong>at</strong>ions fall off<br />

exponentially according to<br />

〈Tr(U(p 1 ))Tr(U(p 2 ))〉 c ∼ exp(−mt)<br />

From our general discussion about correl<strong>at</strong>ors e.g. in eq.(224) it fol<strong>low</strong>s<br />

th<strong>at</strong> m is the <strong>low</strong>est particle mass in the theory since the ground st<strong>at</strong>e has<br />

vanishing energy. Since there are only gluonic degrees of freedom present, the<br />

corresponding massive particle is called glueball. In fact those calcul<strong>at</strong>ions have<br />

been performed for decades an now the numbers stabilize. Some results are<br />

given in the figure<br />

252


253


Although this is an extremely interesting object, it has not unambiguously<br />

be identified experimentally. The reason is, th<strong>at</strong> there are quark-antiquarkexcit<strong>at</strong>ions<br />

of the vacuum, which have the same quantum numbers, and hence<br />

mix with the pre glueball st<strong>at</strong>es. Thus the identific<strong>at</strong>ion of the l<strong>at</strong>tice results<br />

with the experimental d<strong>at</strong>a (Crystal Barrel) is a bit dubious.<br />

13.6.3 Problems with Fermions<br />

Quenched-unquenched: If one looks <strong>at</strong> the actions of QED (244) or <strong>QCD</strong><br />

(245) one realizes th<strong>at</strong> the Fermion fields in the exponent are quadr<strong>at</strong>ic or<br />

linear. Hence one can perform an exact integral over the Fermion fields fol<strong>low</strong>ing<br />

eq.(234): This integr<strong>at</strong>ion can be done analytically and one obtains the so called<br />

fermion determinant det(F), th<strong>at</strong> depends only on the gauge fields U. This<br />

means we have<br />

∫<br />

DUD ¯ψDψ exp [ −S [ U, ¯ψ,ψ ]] ∫<br />

= DU det (F [U]) exp[−S eff [U]]<br />

So far these expressions are equivalent. However p<strong>at</strong>h integrals over fermion<br />

fields are costly and so are costly also the evalu<strong>at</strong>ion of the fermion determinant,<br />

since F(U) involves the whole l<strong>at</strong>tice simultaneously. There fore people were<br />

very happy to formul<strong>at</strong>e a so called quenched approxim<strong>at</strong>ion:<br />

quenched approxim<strong>at</strong>ion: det (F [U]) = 1<br />

This apprexim<strong>at</strong>ion corresponds to ignore all diagrams with virtual quarks.<br />

Missing out sea quarks entirely is known as the quenched approxim<strong>at</strong>ion. It<br />

is clearly wrong, but for a long time the presence of other system<strong>at</strong>ic errors<br />

and poor st<strong>at</strong>istics obscured this fact. More recently it has become clear th<strong>at</strong><br />

the system<strong>at</strong>ic error in the quenched approxim<strong>at</strong>ion is around 10-20%. When<br />

light quark vacuum polaris<strong>at</strong>ion (det(F)) is included the calcul<strong>at</strong>ion is said to<br />

be ‘unquenched’ or ‘dynamical’. The sea quarks are then also called dynamical<br />

quarks. We will see in the results section th<strong>at</strong> it is now possible to include<br />

realistic quark vacuum polaris<strong>at</strong>ion effects and the quenched approxim<strong>at</strong>ion can<br />

be laid aside <strong>at</strong> last.<br />

Light quarks Manipul<strong>at</strong>ions of the m<strong>at</strong>rix F are comput<strong>at</strong>ionally costly.<br />

Even though it is a sparse m<strong>at</strong>rix (with only a few non-zero entries) it is very<br />

large. There are various comput<strong>at</strong>ional techniques for calcul<strong>at</strong>ing the F −1 factors<br />

and the detF factor is included by repe<strong>at</strong>ed determin<strong>at</strong>ion of the calcul<strong>at</strong>ion<br />

of F −1 . This makes the inclusion of detF very costly indeed. It becomes<br />

increasingly hard as the quark mass becomes smaller because F becomes illconditioned.<br />

(The eigenvalues of F range between some fixed upper limit and<br />

the quark mass, so this range increases as m q → 0). In the real world the u<br />

and d quarks have very small mass of a few MeV and so in the past there has<br />

not been sufficient computer power available to include them as sea quarks or,<br />

if they have been included, their masses have been much heavier than their<br />

254


eal values. Chiral symmetry is spontaneously broken in the real world giving<br />

rise to a Goldstone boson, the π, whose mass consequently vanishes <strong>at</strong> zero<br />

quark mass (m 2 π ∝ m q ). For small quark masses, where m π is small, we have a<br />

well-developed chiral perturb<strong>at</strong>ion theory which tells us how hadron masses and<br />

properties should depend on the u/d quark masses (or equivalently m 2 π) and we<br />

can make use of this to extrapol<strong>at</strong>e down to physical u/d quark masses from the<br />

results of our l<strong>at</strong>tice <strong>QCD</strong> simul<strong>at</strong>ion provided th<strong>at</strong> we are able to work <strong>at</strong> small<br />

enough u/d quark masses to be in the regime where chiral perturb<strong>at</strong>ion theory<br />

works. In general m u/d < m s /2 is necessary for an accur<strong>at</strong>e extrapol<strong>at</strong>ion.<br />

Actually in earlier days the heavy mass of the up- and down quarks was<br />

evenwelcome, in order to get a reasonable calcul<strong>at</strong>ion for the heavier mesons.<br />

E.g the Rho-meson would not be the outcome of a calcul<strong>at</strong>ion but two pions,<br />

whose quantum numbers can couple to the Rho. Similarly with the Delta-Isobar,<br />

it exists in l<strong>at</strong>tice codes only if the pion mass is heavier than the delta-nucleon<br />

mass splitting.<br />

Heavy quarks Heavy quarks (b and c) represent a r<strong>at</strong>her different set of<br />

issues to those for light quarks in l<strong>at</strong>tice <strong>QCD</strong>. They have large quark masses,<br />

m Q , and therefore large values of m Q a <strong>at</strong> any value of a <strong>at</strong> which we are able<br />

to do <strong>QCD</strong> simul<strong>at</strong>ions. This means th<strong>at</strong> we risk large discretis<strong>at</strong>ion errors<br />

when handling these quarks if we use a formalism in which the errors are set by<br />

the size of ma. If m Q a > 1 then no amount of improvement will give a good<br />

discretis<strong>at</strong>ion for these quarks. This is particularly true for the b quark which<br />

has a mass of around 5 GeV. To reduce m Q a be<strong>low</strong> 1 requires a −1 > 5 GeV or<br />

a l<strong>at</strong>tice spacing < 0.04 fm which is incredibly expensive to simul<strong>at</strong>e (as well as<br />

being wasteful).<br />

14 Linear chiral Sigma model (Gell-Mann–Levy)<br />

The model must be understood in a special way. One cannot tre<strong>at</strong> it as a fully<br />

field theoretical model which includes the full Dirac sea because then instabilities<br />

occur (See Ripka and Kahana etc. and Sieber). One must understand it as<br />

a classical one with some mild quantum approxim<strong>at</strong>ions. The best way to<br />

understand it is described in the procedure, how to solve it for the properties<br />

of a baryon. there one considers only the valence quarks and believes th<strong>at</strong> the<br />

sea quarks are absorbed into the mesonic fields, which then also have a kinetic<br />

energy term.<br />

14.0.4 Fock st<strong>at</strong>es and vari<strong>at</strong>ional principle<br />

The Lagrangean is given, the only unknown parameters are the value of the<br />

coupling constant g, the vacuum value of the sigma field is taken to be f π .<br />

Evalu<strong>at</strong>e first the Hamiltonian density H(§)of the system of the system,<br />

which is well defined in terms of the fields ψ,σ,π t . Write down the quantized<br />

255


quark-, sigma- and pion field oper<strong>at</strong>ors:<br />

ψ(r) = ∑ {<br />

}<br />

ψ nljm (r)c nljm exp(−iɛ nlj t) + χ nljm (r)d † nljm exp(+iɛ nljt)<br />

nljm<br />

̂σ(r) = (2π) − 3 2<br />

∫<br />

d 3 k(2ω σ (k)) − 1 2 (a(k)exp(ikr) + a † (k)exp(−ikr))<br />

̂π t (r) = (2π) − 3 2<br />

∫<br />

d 3 k(2ω π (k)) − 1 2 (bt (k)exp(ikr) + b † t(k)exp(−ikr))<br />

Then you define coherent st<strong>at</strong>es as quantum st<strong>at</strong>es for the boson fields<br />

|Σ〉 = N − 1 2 exp(<br />

∫<br />

d 3 kη(k)a † (k)) |0〉<br />

with the properties<br />

∑<br />

∫<br />

|Π〉 = N − 1 2 exp( d 3 kζ t (k)b † t(k)) |0〉<br />

t<br />

a(k) |Σ〉 = η(k) |Σ〉<br />

and e.g.<br />

b t (k) |Π〉 = ζ t (k) |Π〉<br />

∫<br />

〈Σ 1 |Σ 2 〉 = N − 1 2<br />

1 N − 1 2<br />

2 exp(<br />

d 3 kη ∗ 1(k)η 2 (k))<br />

and you define the quark st<strong>at</strong>e as<br />

∣ q<br />

3 〉 〉<br />

∣<br />

= ∣c † 1 c† 2 c† 3<br />

with (for the <strong>low</strong>est positive single plarticle st<strong>at</strong>e)<br />

ψ nljm (r) = (r|c † nljm |0〉 = 1 ( )<br />

u(r)<br />

4π iv(r)σ a ̂r a |χ〉<br />

with the hedgehog structure<br />

|χ〉 = 1 √<br />

2<br />

(|u ↑〉 − |d ↓〉)<br />

The hedgehog baryon has the Fock-st<strong>at</strong>e:<br />

〉<br />

∣<br />

|Ψ〉 = ∣c † 1 c† 2 c† 3 |Π〉 |Σ〉<br />

Here one realizes the approxim<strong>at</strong>ions. The boson fields are quantized using<br />

the cre<strong>at</strong>ion oper<strong>at</strong>ors of the free fields as basis and assuming coherent st<strong>at</strong>ess.<br />

256


The fermion fields do not show any Dirac sea. This is achieved technically by<br />

considering only positive energy st<strong>at</strong>es and<br />

Since the quarks are the sources for the boson fields those have the structure<br />

〈Σ| ̂σ(r) |Σ〉 = σ(r) ::::::::: σ(r) ⇔ η(k)<br />

corresponding to<br />

⎛<br />

〈Π| ̂π t (r) |Π〉 = π t (r) = ⎝<br />

⎛<br />

ζ t (k) = ⎝<br />

x<br />

r<br />

y<br />

r<br />

z<br />

r<br />

⎞<br />

⎠ Φ(r)<br />

⎞<br />

ik x<br />

ik y<br />

⎠ A(k) :::::::::: A(k) ⇔ Φ(r)<br />

ik z<br />

Thus the inform<strong>at</strong>ions in η(k),ζ t (k) and σ(r),Φ(r) are equivalent.<br />

The energy of the system is given by<br />

E = 〈Ψ |HΨ〉<br />

in terms of the functions σ(r),Φ(r),u(r),v(r). The fields are assumed timeindependent.<br />

One determines the functions σ(r),Φ(r),u(r),v(r) by the vari<strong>at</strong>ional<br />

procedure:<br />

δE<br />

δu(r) = δE<br />

δv(r) = δE<br />

δσ(r) = δE<br />

δΦ(r) = 0<br />

This results in a set of four differential equ<strong>at</strong>ions with 4 unknown functions.<br />

The equ<strong>at</strong>ions are non-linear. Their result corresponds to a soliton. In the end<br />

the full Fock-st<strong>at</strong>e is known. There are, however, no quantum numbers yet.<br />

14.0.5 Projection techniques<br />

The quantum numbers are determined in the fol<strong>low</strong>ing way using Peierls-Yoccoz<br />

perojection techniques:<br />

One defines the projection oper<strong>at</strong>or:<br />

P J MK = 2J + 1<br />

8π 2<br />

P T M T K T<br />

= 2T + 1<br />

8π 2<br />

∫<br />

∫<br />

dΩD J∗<br />

MK(Ω)R(Ω)<br />

dΩ T D T ∗<br />

M T K T<br />

(Ω T )R T (Ω T )<br />

where R and R T are rot<strong>at</strong>ion oper<strong>at</strong>ors in spin and isospin SU(2)<br />

R(α,β,γ) = exp(−iαJ 3 )exp(−iβJ 2 )exp(−iγJ 3 )<br />

257


and analogously for R T . The projection oper<strong>at</strong>ors have the properties<br />

(<br />

P<br />

J<br />

M,K<br />

) †<br />

= P<br />

J<br />

K,M<br />

P J ′<br />

K ′ ,M ′P J M,K = δ J,J ′δ M,M ′P J K ′ ,K<br />

One can show th<strong>at</strong> the most general st<strong>at</strong>e with given angular momentum and<br />

isospin , obtained by rot<strong>at</strong>ing Ψ in space and isospin space, can be written as<br />

|J,T,M,α,M T 〉 = ∑<br />

PM,KP J M T T K T<br />

|Ψ〉<br />

with<br />

∑<br />

g (J,T,α)<br />

K,K T<br />

K,K T<br />

J,T,M,α,M T<br />

|J,T,M,α,M T 〉 〈J,T,M,α,M T | = I<br />

which means, th<strong>at</strong> Ψ is a linear combin<strong>at</strong>ion of the projected st<strong>at</strong>es, th<strong>at</strong> is, it<br />

contains the st<strong>at</strong>es with good angular momentum and isospin as components.<br />

∑<br />

|Ψ〉 = |J,T,M,α,M T 〉 〈J,T,M,α,M T | Ψ〉<br />

J,T,M,α,M T<br />

The coefficients g (J,T,α)<br />

K,K T<br />

are the results of a non-orthogonal diagonaliz<strong>at</strong>ion procedure:<br />

∑ (<br />

)<br />

h (J,T)<br />

K,K T ,K ′ ,K<br />

− E (J,T,α) n (J,T)<br />

T<br />

′ K,K T ,K ′ ,K<br />

g (J,T,α)<br />

T<br />

′ K ′ ,K<br />

= 0<br />

T<br />

′<br />

with the kernels<br />

h (J,T)<br />

K,K T ,K ′ ,K ′ T<br />

= 〈Ψ|HPKK J ′P K T T ,K |Ψ〉<br />

T<br />

′<br />

= 〈Ψ|PKK J ′P K T T ,K |Ψ〉<br />

T<br />

′<br />

The normaliz<strong>at</strong>ion of the coefficients is done by<br />

∑<br />

(J,T,α g ′ ) ∗<br />

g (J,T,α)<br />

K ′ ,K<br />

n (J,T)<br />

T<br />

′ K,K T ,K ′ ,K<br />

= δ<br />

T<br />

′ α′ α<br />

n (J,T)<br />

K,K T ,K ′ ,K ′ T<br />

K ′ ,K ′ T<br />

Apparently the eigenvalues E (J,T,α) of these equ<strong>at</strong>ions are the <strong>energies</strong> (masses)<br />

of the baryon with the corresponding quantum numbers. One has some simple<br />

sum rules:<br />

∣ ∣<br />

∑ ∣∣∣∣∣<br />

∑<br />

∣∣∣∣∣<br />

2<br />

g (J,T,α)<br />

K,K T<br />

n (J,T)<br />

M,K,M T K T<br />

= 1<br />

JTαMM T K,K T<br />

This sum rule tells, how much of the st<strong>at</strong>e JTαMM T is contained in the st<strong>at</strong>e<br />

Ψ. And the next sum rule tells, how much of the mean field energy 〈Ψ|H |Ψ〉<br />

comes from the st<strong>at</strong>eJTαMM T .<br />

∣ ∣∣∣∣∣<br />

∑ ∑<br />

E (J,T,α)<br />

JTαMM T<br />

g (J,T,α)<br />

K,K T<br />

K,K T<br />

n (J,T)<br />

M,K,M T K T<br />

∣ ∣∣∣∣∣<br />

2<br />

= 〈Ψ|H |Ψ〉<br />

258


To calcul<strong>at</strong>e a nucleon property one has to write an observable Ô in terms of<br />

̂ψ, ̂σ, ̂π t and has then to calcul<strong>at</strong>e<br />

O (α)<br />

J,T,M,M T<br />

= 〈J,T,α,M,M T |Ô |J,T,α,M,M T 〉<br />

which is now well defined. The calcul<strong>at</strong>ions of the form factors are done in this<br />

way. The calcul<strong>at</strong>ions are succesfull in several respects. If one adjusts the coupling<br />

constant g of the fermion-pion coupling such th<strong>at</strong> the mass of the nucleon<br />

is reproduced (M = 938MeV ), then the electric form factor of the proton is<br />

reasonably well reproduced. The absolute values of the magnetic moment are<br />

correct to 30%. One has also disadvantages: One obtains g A = 1.8 and the<br />

nucleon-delta-splitting comes out about about 150 MeV. The experimental values<br />

are g A = 1.25 and 300 MeV. The nucleon mass is reproduced because one<br />

autom<strong>at</strong>ically obtains some sort of correl<strong>at</strong>ion energy such th<strong>at</strong> in the limit of<br />

strong deform<strong>at</strong>ion in flavour space we have approxim<strong>at</strong>ely<br />

〈Ĵ2〉<br />

E J,T = E 0 +<br />

J(J + 1)<br />

2Θ<br />

One can also ask for the component a certain st<strong>at</strong>e|J,T,α,M,M T 〉 is contained<br />

1 1<br />

2 2<br />

20%<br />

3 3<br />

in the soliton. One obtains for the <strong>low</strong>est st<strong>at</strong>e for a given J,T,%:<br />

2 2<br />

50%<br />

5 5<br />

2 2<br />

25%<br />

One sees <strong>at</strong> the numbers th<strong>at</strong> higher components are only contained in Ψin<br />

a negligible way. This is interesting, compared to the outcome of semiclassical<br />

methods, where we have rigid rot<strong>at</strong>ions in the coordin<strong>at</strong>e and isospin space and<br />

hence all quantum nunbers are present.<br />

In the above procedure one performs first the vari<strong>at</strong>ion and after th<strong>at</strong> the<br />

projection (vari<strong>at</strong>ion before projection). However, one obtains only good results<br />

if one applies the vari<strong>at</strong>ion after projection:<br />

−<br />

2Θ<br />

δ 〈J,T,α,M,M T |Ĥ |J,T,α,M,M T 〉 = 0<br />

This means, one varies the already projeted st<strong>at</strong>es. In this case one has different<br />

functions σ(r),Φ(r),u(r),v(r) for each st<strong>at</strong>e |J,T,M,α,M T 〉 . This is an<br />

interesting result, which has not fully been exploited yet. Show form<br />

factors of<br />

gen hh with<br />

15 Restbestände<br />

projections<br />

Current algebra, Schwinger terms A simple argument shows, th<strong>at</strong> the<br />

Schwinger terms do not vanish. Asume for a moment th<strong>at</strong> they are absent. Then<br />

we have from the above formulae (because of antisymmetry of the structure<br />

coefficients) [<br />

j<br />

a<br />

0 (x,t),j a i (y,t) ] = 0<br />

which implies [<br />

j<br />

a<br />

0 (x,t),∂ i j b i (y,t) ] = 0<br />

259


Assuming current conserv<strong>at</strong>ion we obtain<br />

[<br />

j<br />

a<br />

0 (x,t),∂ 0 j b 0(y,t) ] = 0<br />

Taking the vacuum expect<strong>at</strong>ion value and inserting a complete set of energy<br />

eigenst<strong>at</strong>es we have<br />

〈0| [ j a 0(x,t),∂ 0 j b 0(y,t) ] |0〉 =<br />

= ∑ n<br />

(〈0| J 0 (x,t) |n〉 〈n| ∂ 0 J 0 (x,t) |0〉 − 〈0| ∂ 0 J 0 (x,t) |n〉 〈n| J 0 (x,t) |0〉<br />

= i ∑ n<br />

(<br />

exp(ipn (x − y)) + exp(−ip n (x − y)) ) E n |〈0| J 0 (0) |n〉| 2<br />

because the time deriv<strong>at</strong>ive, applied to the st<strong>at</strong>es, cancels the terms from the<br />

transl<strong>at</strong>ion oper<strong>at</strong>ors (used to change vom J(x) to J(0). In the limit x → y this<br />

means th<strong>at</strong> ∑<br />

E n |〈0| J 0 (x,t) |n〉| 2 = 0<br />

n<br />

>From the fact, th<strong>at</strong> all E n ≥ 0 fol<strong>low</strong>s th<strong>at</strong> 〈0|J 0 (x,t) |n〉 = 0 for all n. Thus<br />

we would have J 0 = 0 identically. Thus the Schwinger terms dissappear only if<br />

the current j 0 = 0, which is the trivial case.<br />

Since M N has the dimension of an energy and since a is the only dimensionful<br />

quantity we must have the rel<strong>at</strong>ion<br />

M N =<br />

1<br />

aλ(g(a))<br />

and hence there must be a critical value g cr such th<strong>at</strong><br />

lim λ(g) = ∞<br />

g−→g cr<br />

We can see the effect immedi<strong>at</strong>ely <strong>at</strong> the harmonic oscill<strong>at</strong>or. The eigenfunctions<br />

are known<br />

( mω<br />

) (√ ) ]<br />

1/4 1 mω<br />

< x|n >= √<br />

πħ 2n n! H n<br />

ħ x exp<br />

[− mωx2<br />

2ħ<br />

260

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!