Non-Perturbative Field Theory
Non-Perturbative Field Theory
Non-Perturbative Field Theory
Providing a new perspective on quantum field theory, this book gives a ped-
agogical and up-to-date exposition of non-perturbative methods in relativistic
quantum field theory and introduces the reader to modern research work in
theoretical physics.
It describes in detail non-perturbative methods in quantum field theory,
and explores two-dimensional and four-dimensional gauge dynamics using those
methods. The book concludes with a summary emphasizing the interplay
between two- and four-dimensional gauge theories.
Aimed at graduate students and researchers, this book covers topics from two-
dimensional conformal symmetry, affine Lie algebras, solitons, integrable models,
bosonization and ’t Hooft model, to four-dimensional conformal invariance, inte-
grability, large N expansion, Skyrme model, monopoles and instantons. Applica-
tions, first to simple field theories and gauge dynamics in two dimensions, and
then to gauge theories in four dimensions and quantum chromodynamics (QCD)
in particular, are thoroughly described. This title, first published in 2010 , has
been reissued as an Open Access publication on Cambridge Core.
†
Issued as a pap erback.
YITZHAK FRISHMAN
The Weizmann Institute of Science
JACOB SONNENSCHEIN
Tel Aviv University
www.cambridge.org
Information on this title: www.cambridge.org/9781009401647
DOI: 10.1017/9781009401654
© Yitzhak Frishman and Jacob Sonnenschein 2010
This work is in copyright. It is subject to statutory exceptions and to the provisions
of relevant licensing agreements; with the exception of the Creative Commons version the
link for which is provided below, no reproduction of any part of this work may take
place without the written permission of Cambridge University Press.
An online version of this work is published at doi.org/10.1017/9781009401654 under a
Creative Commons Open Access license CC-BY-NC-ND 4.0 which permits re-use,
distribution and reproduction in any medium for non-commercial purposes providing
appropriate credit to the original work is given. You may not distribute derivative works
without permission. To view a copy of this license, visit
https://creativecommons.org/licenses/by-nc-nd/4.0
All versions of this work may contain content reproduced under license from third parties.
Permission to reproduce this third-party content must be obtained from these third-parties directly.
When citing this work, please include a reference to the DOI 10.1017/9781009401654
First published 2010
Reissued as OA 2023
A catalogue record for this publication is available from the British Library.
ISBN 978-1-009-40164-7 Hardback
ISBN 978-1-009-40161-6 Paperback
Cambridge University Press & Assessment has no responsibility for the persistence or accuracy of
URLs for external or third-party internet websites referred to in this publication
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
Yitzhak Frishman
To my mother Hilda,
wife Nava
and children Nir, Ori and Tal
Jacob Sonnenschein
Preface page xv
Acknowledgements xviii
References 423
Index 433
Field theory is the framework with which one describes the theory of the standard
model of elementary particles and their interactions. The electromagnetic sector
(QED) of the standard model is understood extremely well using perturbation
theory, but the color interaction (QCD) which is responsible for hadron physics
can only be accounted for perturbatively for a limited set of observational data.
Due to the fact that at long distances the color interaction is strongly coupled,
one cannot reliably apply perturbative methods to extract, for instance, the
spectrum of the hadrons. The arsenal of tools to handle strongly coupled systems
is obviously much more limited than the one used for weakly coupled ones.
Nevertheless, several methods to handle non-perturbative field theories have been
developed. The main goal of this book is to expose the reader to those techniques
and to describe their applications in two-dimensional and four-dimensional field
theories and finally in QCD in four dimensions.
The topic of non-perturbative field theory is by itself very rich and it is clear
that one cannot cover it in a non superficial manner in one book. Thus we had
to make certain decisions about the flow of the book and about the topics that
should be addressed. As for the former issue we have decided to present the
book in three parts. In the first part we describe, in detail, the most impor-
tant non-perturbative techniques of two-dimensional field theory. The reason for
this is obvious since physical systems with one space dimension and one time
dimension are the simplest and hence it is easier to grasp the non-perturbative
tools when applied to these systems. In the second part of the book we study
two-dimensional gauge theories with the emphasis on employing the techniques
developed in the first part. The third part is devoted to the non-perturbative
aspects of gauge dynamics in four dimensions. In this part we elevate the tech-
niques of the first part to four dimensions and we examine to what extent gauge
theories in four dimensions behave like their two-dimensional simplified analogs.
There are several books on the shelves discussing non-perturbative methods in
general such as [66] and [182], there are books describing one particular method,
like conformal field theory in two dimensions for instance [77], there are books
that describe two-dimensional QCD, [2] and books that study various aspects of
four-dimensional QCD, for example [151] and of course there are books on the
basics of field theory, for example [37], [130], [173] and [215]. The aim of this
book is three-fold, to review a package of non-perturbative methods, to present
a picture which is close to the state-of-the-art in the topics described and to
into general conformal field theories. In dealing with more advanced topics, like
for instance instantons in four dimensions, the reader will need to consult with
specialized references to obtain a more complete and wider picture of the topic.
Some of the content of the book, mainly in Part 2, is based on the research
work of the authors, but most of the material is a review of the work of many
researchers in the field.
The book is aimed for advanced Ph.D. students, post-docs and other newcom-
ers to the arena of non-perturbative methods in field theory. The reader should
definitely be equipped with a basic knowledge of field theory, group theory and
algebra, differential equations, geometry and topology.
Throughout the book we refer to only a limited list of references. The number
of scientific contributions to the topics discussed in this book is enormous and
since we could not cover all of them we have referred to papers that initiated the
various topics, and to review papers and books where a much more exhaustive
list of references can be found.
We have made an attempt to keep the same notations throughout the book.
However in certain instances we have changed notations during the course of
the book, mainly to be in accordance with relevant literature. In these cases we
specified explicitly the change in notation made.
In this chapter we analyze the simplest field theory, which is the theory of a
free massless scalar field in two space-time dimensions, one space and one time.1
The rich symmetry and algebraic structure of this theory encapsulates the basic
concepts of two-dimensional conformal field theory, which will be the topic of
the next chapter.
1 The content of this chapter comprises the basics of massless scalar fields in two dimensions.
This is covered in many textbooks.
2 The use of complex coordinates in the context of the bosonic string theory is described by
Polyakov in [177].
x0
x1
energy-momentum tensor),
1
T ≡ Tξ ξ = (T00 − 2iT10 − T11 )
4
1
T̄ ≡ Tξ¯ξ¯ = (T00 + 2iT10 − T11 )
4
1
Tξ¯ξ = Tξ ξ¯ = (T00 + T11 ). (1.6)
4
Often, especially in the context of string theory, the space direction is no longer
R, but rather is compactified on S 1 so that x1 ≡ x1 + 2π. For such a geometry
it is convenient to introduce the following conformal map:
0
+ix 1
ξ → z = e ξ = ex ,
which maps the cylinder to the complex plane (see Fig. 1.1).
In particular the past x0 = −∞ is mapped into the origin and the future
x = ∞ into a circle with an infinite radius. It is clear that the relations between
0
¯ and (x0 , x1 ) derived above hold also between (z, z̄) and (Real(z), Im(z)).
(ξ, ξ)
The holomorphic and anti-holomorphic derivatives with respect to z will be
denoted by ∂ ≡ ∂z and ∂¯ ≡ ∂z̄ .
3 Affine Lie algebras describing a physical system were first discussed in [27]. More references
will be given in the next two chapters.
are automatically obeyed by X̂(z, z̄). For this case the mode expansion is
expressed in terms of a Laurent series,
∞
∞
αn ᾱn
∂X = −i ∂¯X̄ = −i . (1.18)
n =−∞
zn + 1 n =−∞
z̄ n +1
P = α0 = ᾱ0 . (1.20)
For open strings the boundary conditions are of Neumann type, namely
Unlike the situation in more than two dimensions, and due to the fact that the
symmetries (1.14) are in fact not only global ones but rather “half local”, the
currents
¯ ≡ ∂∂X
∂J ¯ = 0; ∂ J¯ ≡ ∂ ∂¯X̄ = 0. (1.25)
1 1
T = − ∂X∂X; T̄ = − ∂¯X̄ ∂¯X̄, (1.27)
2 2
where the coefficients were chosen in a way that will turn out to be convenient
when discussing the corresponding quantum generators.
These conditions yield the standard algebra of the creation and annihilation
operators for (1.16),
Substituting the mode expansion (1.16) into the expressions of the Noether
charges associated with the symmetries of the action (1.7) one finds that
the energy-momentum operators are proportional to a† (k)a(k) + a(k)a† (k) and
hence their vacuum expectation values are proportional to δ(0) ∼ L, where L is
the size of the space direction. It is thus clear that for the infinite Euclidean plane
(or a Minkowski space-time with space R) these expectation values diverge. One
then defines the normal ordered operators:
: O : ≡ O − <0|O|0> . (1.30)
For free fields this is equivalent to ordering annihilation operators to the right
of creation operators, and sufficient to make : O : finite.
Using the algebra of the creation and annihilation operators and the normal
ordered Hamiltonian, the construction of the Fock space is standard. One defines
the vacuum state |0> such that
The canonical quantization for the scalar field on a compact space direction,
with the boundary conditions of open or closed string, (1.21) and (1.17), respec-
tively, follows very similar steps. Imposing the quantization conditions (1.28)
above implies the following algebra for the αn operators of the open string and
4 The notion of radial quantization was introduced in [104]. This construction was used in the
context of complex geometry in [93].
z z
w
w w
1
δ X̂(w, w̄) = dz(z)R(∂X(z)X̂(w, w̄))
2πi
1 −1
= dz (z) = −(w), (1.40)
2πi z−w
1 1
δ ∂X(w) = dz(z)R − : ∂X(z)∂X(z) : ∂X(w)
2πi 2
1 1
= dz ∂X(z)(z) = ∂(w)∂X(w) + (w)∂ 2 X(w), (1.42)
2πi (z − w)2
where − y) are the coefficient functions which are singular in the limit
ckij (x
of x → y. Such expansions were proven to hold in renormalizable field theo-
ries. The OPEs are an essential tool in exploring quantum field theories. Recall
that all of the information on the QFT is encoded into the values of all possi-
ble correlation functions of the complete set of local operators Oi (x), namely,
< O1 (x1 )...On (xn ) >. In particular, one is interested in the behavior of these
correlation functions when two or more points approach each other, which is
encapsulated in the OPEs. For all applications discussed here the OPEs are
treated as asymptotic expansions and only their singular terms will be specified.
For the present case of two-dimensional free massless scalar field theory the OPE
converges and in fact, as will be discussed in Section 3.7.2, a similar situation
occurs in all 2d CFTs.
The OPEs of the free massless scalar can be deduced from its propagator,
which can be evaluated from the solution. It takes the form:
< X̂(z z̄)X̂(ww̄) >= −log|z − w|2 . (1.45)
In terms of the separation of the solution into holomorphic and anti-holomorphic
parts the two propagators read:
< X(z)X(w) >= − log(z − w); < X̄(z̄)X̄(w̄) >= − log(z̄ − w̄). (1.46)
By differentiating the last relation with respect to z and to w one finds the short
distance expansion of other operators like J(z), T (z) etc. In particular the OPE
of the currents is
1
J(z)J(w) = ∂X(z)∂X(w) = − + finite terms, (1.47)
(z − w)2
with a similar result for the anti-holomorphic currents.
A different, though equivalent, approach is to write the OPE as a Taylor
expansion in (z − w) and (z̄ − w̄) in the following form:
∞
1
X̂(z, z̄)X̂(ww̄) = −log|z − w|2 + [(z − w)k : (∂ k X̂(w, w̄))X̂(w, w̄) :
k!
k=1
+(z̄ − w̄)k : (∂¯k X̂(w, w̄))X̂(w, w̄) :]. (1.48)
5 Wilson introduced for the first time the concept of an operator product expansion [219]. It
was used for two-dimensional conformal field theories in [33].
This form of expansion is based on the property that the normal ordered product
of the scalar fields,
(i) The conformal properties of : eiα X (w ) are being determined by its OPE with
T (z) which takes the form
1
T (z) : eiα X (w ) : = − (: ∂X(z)∂X(z) :)(: eiα X (w ) :)
2 2
( α2 ) iα X (w ) 1
= e + ∂eiα X (w ) . (1.51)
(z − w) 2 (z − w)
In language that will be developed in Section 2.2 this result will mean that
2
: eiα X (w ) : has a conformal dimension of α2 .
(ii) The OPE of two operators of the form : eiα X (w ) is
: eiα X (w ) e−iβ X (w ) :
(: eiα X (z ) :)(: e−iβ X (w ) :) = . (1.52)
(z − w)α β
as can be deduced by using the fact that the path integral of a total derivative
vanishes:
δ
0 = DX̂(z, z̄) [e−S X̂(w, w̄)]
δ X̂(z,
z̄)
−S δS
= DX̂(z, z̄)e − X̂(w, w̄) + δ (z − w, z̄ − w̄)
2
δ X̂(z, z̄)
1
= < ∂ ∂¯X̂(z, z̄)X̂(w, w̄) > + < δ 2 (z − w, z̄ − w̄) > . (1.55)
2π
Alternatively one can use (1.45) and (1.46) directly. Note that in that case care
must be exercised, as naively we would get zero rather than the delta function,
since the expression is a sum of two terms, one depending on z only and the
other on z̄ only. The point is that the expressions (1.46) cannot be taken over at
the origin. A working rule is:
1 1
∂¯ =∂ = (2π)δ 2 (z, z̄)
z z̄
1 z̄
This can be derived by going over from z to z z̄ + 2 , to regulate the singularity
at the origin.
it is obvious from the mode expansion (1.18) that the algebra of the currents is
related to that of the αn operators, namely
This form of algebra will be shown in the next chapter to be associated with level
k = 1 abelian affine current algebra (or Kac–Moody algebra as it is sometimes
referred to).
This algebra translates into the following algebra of the currents:
Using the technique developed in Section 1.7 we can derive this result also from
the operator product expansion of two currents,
1
J(z)J(w) = + finite terms. (1.59)
(z − w)2
6 The Virasoro algebra was presented in [212]. More references will be given in the next two
chapters.
the result. Taking this into account, one gets a c-number shift in the commutation
rule,
If we now take fΛ (k) to 1, without being careful, we get the divergent sum,
∞
A(m) → m k. (1.67)
k=1
Only large k is relevant now, as for any finite k we can take Λ to infinity first,
obtaining zero on the right-hand side. We now take,
−p
fΛ (q) ≈ |q| |q| → ∞, (1.69)
where the sum is usually finite, and the operators Õ(n ) (w, w̄) have to be deter-
mined. Using radial quantization as in Section 1.7 and the OPE above, we get
for the transformation generated by T (z),
1
δ Õ(w, w̄) = (∂ n )Õ(n ) (w, w̄) . (2.7)
n
n!
∂z
h
∂ z̄
h̄
O(z, z̄) → O (z z̄ ) = O(z z̄ ). (2.8)
∂z ∂ z̄
An operator with such conformal transformations is a primary field or a ten-
sor operator with conformal weights (h, h̄), which are sometimes referred to as
the holomorphic and anti-holomorphic conformal dimensions.1 The sum of the
weights h + h̄ is the total dimension that determines the behavior under scal-
ing, whereas h − h̄ is the spin that controls the behavior under rotations. The
infinitesimal transformations that correspond to (2.8) are,
¯ + ¯∂)
δ,¯ O(z, z̄) = (h∂ + ∂) + (h̄∂¯ ¯ O(z z̄). (2.9)
This form of transformation implies that the singular part of the OPE of T and
O(w, w̄) reduces to,
h 1
T (z)O(w, w̄) = O(w, w̄) + ∂O(w, w̄). (2.10)
(z − ω)2 (z − ω)
Applying these notions to the free scalar field we find that ∂X(z) has (1, 0)
2 2
weights, ∂¯X̄(z̄) has (0, 1) and the weights of : eiα X (z , z̄ ) : are ( α2 , α2 ).
In Chapter 1 the notion of OPE was discussed in the context of scalar field the-
ory. The generalization to any CFT is straightforward. Normalize the operators
with fixed conformal weights as,
1 1
Oi (z, z̄)Oj (w, w̄) = δij , (2.11)
(z − w) (z̄ − w̄)2 h̄ i
2h i
then, for a complete set, the OPE of any pair of such operators is, to leading
singularity,
Oi (z, z̄)Oj (w, w̄) ∼ Cij k (z − w)h k −h i −h j (z̄ − w̄)h̄ k −h̄ i −h̄ j Ok (w, w̄), (2.12)
k
1 The notion of conformal primary field and its descendants was introduced in [33] and further
discussed in [236].
we can integrate the equation to get (2.16). The first term suggests integrating to
f /f , the variation of which gives 1/f (δf ) − f /(f )2 (δf ) , while the second
term suggests −3(f )2 /2(f )2 , the variation of which gives −3f /(f )2 (δf ) +
3(f )2 /(f )3 (δf ) .
For the massless scalar case T can be written as T (z) = − 12 : J(z)J(z) :, as
we saw in (1.5). In fact, as will be discussed in Chapter 3, there is a large class
of theories that share this so-called Sugawara form. For this type of theory the
proof that the finite transformation is of the form of (2.15) is as follows. Recall
that as a primary field of weights (1, 0), J(z) → ∂∂zz J(z ). If we write T (z) =
− 12 limz →w (J(z)J(w) + (z −w
1
)2 ) and substitute the transformation of the currents
we end up after some lengthy but straightforward calculation with (2.15).
so that,
1
Ln = dzz n + 1 T (z). (2.20)
2πi
The expansion is chosen such that Ln has scale dimension n under z → z/a,
namely, Ln → an Ln .
The Virasoro algebra2 can now be derived using the OPE of T (z)T (w) given
in (2.13),
2
1
[Ln , Lm ] = dz dw[z n + 1 wm + 1 − z m +1 wn +1 ]T (z)T (w). (2.21)
2πi
The double
integral is performed by fixing w and transforming the difference of
the two dz integrals into one integral around w,
1
2
[Ln , Lm ] = dz dw z n + 1 wm +1 − z m +1 wn +1
2πi
c/2 2T (w) ∂w T (w)
+ +
(z − w)4 (z − w)2 (z − w)
1
= dw c/12(n3 − n)wn +m −1
2πi
+ [2(n + 1) − (n + m + 2)] wn +m +1 T (w)
c 3
= (n − n)δ(n + m) + (n − m)Ln +m . (2.22)
12
Performing identical steps for L̄n we get that L̄n obeys the same infinite algebra,
with some central charge c̄, and that [Ln , L̄m ] = 0.
Any CFT is a representation of the Virasoro algebra characterized by c and c̄.
It is straightforward to identify the following properties of the algebra:
r The generators (L±1 , L0 ) span an SL(2, R) algebra,
2 The first use of the Virasoro algebra was by M. Virasoro in the context of the dual resonance
model [212]. Its application to two-dimensional CFT was presented in [33].
0 h φ
1 h+1 L−1 φ
2 h+2 L−2 φ, L2−1 φ
3 h+3 L−3 φ, L−2 L−1 φ, L3−1 φ
.. .. ..
. . .
N h+N P(N) fields
r For n > 0, L−n is a raising operator and Ln is a lowering one, since [L0 , Ln ] =
−nLn so that if |ψ> is an eigenstate of L0 , L0 |ψ> = h|ψ> then L0 |Ln ψ> =
(h − n)|Ln ψ>.
(2.26)
¯
where we denote by φlkl (w, w̄) the descendants L−l 1 . . . L−l n L̄−¯l 1 . . . L̄−¯l n φk (w, w̄)
{l ¯l}
with the normalization given in (2.11). The product coefficients Cij k are given
in terms of those of (2.12) Cij k as,
{l ¯l} k {l} k {¯l}
Cij k = Cij k βij β̄ij , (2.27)
k {l}
where βij are determined by conformal invariance and are functions of c and
k {¯l}
hi , hj , hk , and similarly for β̄ij . This follows from a detailed analysis that we
do not show here.
The OPEs of any pair of descendant fields can also be deduced from (2.12)
which implies in fact that all the information about the OPE is encoded in the
product coefficients Cij k . Moreover since the structure of (2.26) holds for all the
primaries and their descendants, one can write the so-called fusion algebra for
conformal, families, which takes the form,
[φi ][φj ] = Nijk [φk ]. (2.28)
k
It is thus clear that any state in a positive Hilbert space is a linear combination
of hws, and their descendants. The representation given in Table 2.1 is referred
to as the Verma module. Denoting it by V(c, h) and its analogous representation
for the anti-holomorphic Virasoro algebra by V̄(c̄, h̄), the Hilbert space of the
theory is a direct sum of the products V(c, h) ⊗ V̄(c̄, h̄), namely,
H= V(c, h) ⊗ V̄(c̄, h̄). (2.34)
h, h̄
The Verma module may be reducible in the sense that there is a submodule that
is by itself a Verma module. Such a submodule whose states transform amongst
themselves under any conformal transformation, is built from a |hnull>. The
latter is both an hws., namely Ln |hnull> = 0 for n > 0, as well as a descendant.
Such a state is called null state or null vector, motivated by what follows. It
generates its own Verma module which is included in the parent module. It is
orthogonal to the whole Verma module as well as to itself <hnull |hnull> = 0,
since <hnull |L−k 1 . . . L−k n |h> = <h|Lk n . . . Lk 1 |hnull>∗ = 0, and in particular it
has a zero norm <hnull |hnull> = 0 and similarly also its descendants. The null
state corresponds to a null operator which is simultaneously a primary and a
secondary field.
Let us now demonstrate the construction of a null vector. Consider a general
linear combination of the states of level 2,
we would like to check whether for certain values of the mixing coefficient a, this
state is a null state. If indeed it is |null>, then so is the state [Ln |null>]. In fact
it is easy to verify that at level 2, one has to check these consistency conditions
only for L1 and L2 . Now using the Virasoro algebra we find that,
[L1 , L−2 ] |h> + a L1 , L2−1 |h> = (3 + 2a(2h + 1))L−1 |h>,
c
[L2 , L−2 ]|h> + a L2 , L2−1 |h> = 4h + + 6ah |h> . (2.36)
2
It is thus clear that for the following values of a and c,
3 2h
a=− c= (5 − 8h), (2.37)
2(2h + 1) 2h + 1
the linear combination state (2.35) is a null state. In the unitary case we have
h and c positive (see next section). Hence in this example h < 58 .
An irreducible representation of the Virasoro algebra can be constructed from
a Verma module that contains a null vector by a quotient procedure, taking
out of the Verma module the null module. In the next section we discuss this
construction.
the matrix of inner products Iss = <s|s> is block diagonal with blocks I(N ) for
states at level N ( i ki = N ). For a given Verma module the elements of I are
functions of (h, c). It is easy to realize that unitarity dictates c > 0 and h > 0.
This follows from <h|Ln L−n |h> = [2nh + 1/12cn(n2 − 1)] <h|h>, which is pos-
itive for n = 1 only if h > 0 and for large enough n only for c > 0. To determine
the full set of constraints for unitarity let us analyze further the properties of I.
A general state |ŝ> = k ck |s> has a norm <ŝ|ŝ> = ĉ† Iĉ, with ĉ the vector of
the ck . Now since I is hermitian it can be diagonalized by a unitary matrix U
so that the norm can be written as <ŝ|ŝ> = k lk |tk |2 where t = U ĉ and lk are
the eigenvalues of I, which are real. It is thus clear that there are negative norm
states if and only if I has negative eigenvalues. A vanishing eigenvalue indicates
a null vector which means a reducible Verma module.
For the low lying levels these matrices take the following form:
I(0) = 1
I(1) = 2h
4h(2h + 1) 6h
I(2) = . (2.39)
6h 4h + c/2
The derivation of the various elements is straightforwad, for instance,
(2)
I11 = <h|L1 L1 L−1 L−1 |h> = <h|L1 L−1 L1 L−1 |h> + 2 <h|L1 L0 L−1 |h>
= 4 <h|L1 L−1 L0 |h> + 2 <h|L1 L−1 |h> = 8h2 + 4h. (2.40)
3 The proof of the Kac determinant is detailed in [89], [206] and [95].
where αN are constants independent of (c, h) and hp,q (c) can be expressed in
terms of m = − 21 ± 12 25−c
1−c as,
Note that we can choose either the plus or the minus sign in the expression for
m, as their interchange is like interchanging p with q, which does not change the
determinant. Note also that hp,q is invariant under p → m − p, q → m + 1 − q.
Let us also mention that for N = 2 the result is identical to (2.41).
In the (h, c) plane the determinant vanishes along the curves h = hp,q (c) which
are therefore called the vanishing curves. If the determinant (2.42) is negative it
means that there is an odd number of negative eigenvalues and hence the corre-
sponding Virasoro representation is not unitary. If the determinant is vanishing
or positive one needs to further analyze the determinant as follows:
r For c > 1 and h > 0 it is straightforward to show that the determinant does
not vanish.
In the domain 1 < c < 25 the value for m has an imaginary part. Thus hp,q are
complex for p = q, and as they come in complex conjugate pairs the product of
the appropriate two factors in the determinant is positive. For p = q the value
of hp,q is negative. Thus the determinant is positive in that domain.
For c > 25 the hp,q are negative.
For large h the matrix is dominated by its diagonal elements.
Since these elements are positive, the eigenvalues for large h are all pos-
itive. Now since the determinant never vanishes in the region considered
(h > 0, c > 1) all the eigenvalues have to be positive on the entire region.
Note that in I(2) the off-diagonal element is larger at large h than the 22
element, but still the determinant is dominated at large h by the diagonal
elements, and thus also the eigenvalues, as a 2 × 2 matrix.
r For c = 1 we have hp,q = (p − q)2 /4, and so the determinant is never negative.
However, it vanishes when h = n2 /4 for some integer n.
r For 0 < c < 1, h > 0 a closer look at the determinant is required. We draw
hp,q (c) in Fig. 2.1.
By expanding the curves around c = 1 one can show that any point in the
region can be connected to the right of c = 1 by crossing a single vanishing
curve. The vanishing of the determinant is due to one eigenvalue that reverses
its sign which implies that there are negative norm states at any point in the
region that are not on the vanishing curve. In fact it turns out that there are
additional negative norm states at points along the vanishing curve except at
h h4,2
h3,1
h1,3
h2,4
1
2
h3,2
h2,1
1
4
h1,2
h2,3
1
16 h3,3
h2,2
h1,1
1 7 1 c
0 1
2 10 5
2.8 Characters
The structure of the Verma module, and in particular the degeneracy of states
at each level, is captured in the generating function χ(c,h) (τ ), the character of
the Verma module, defined by,
∞
χ(c,h) (τ ) = Tr[q L 0 − 2 4 ] = dim(h + n)q h+n − 2 4 ,
c c
(2.45)
n =0
where q ≡ e 2πiτ
, τ is a complex number, and dim(n + h) is the number of linearly
independent states of the module at level n. The latter is equal to P (n) the
partitions of n in the generic case, but may be smaller when there are null
states. For |q| < 1, namely, τ in the upper half plane, the series is uniformly
convergent, since |q| < 1 is the domain of convergence of the inverse of the Euler
function ϕ(q) defined by,
∞ ∞
1 1
= = P (n)q n . (2.46)
ϕ(q) n = 1 1 − q n n =0
In terms of this function the character of a generic Verma module is given by,
c
q h− 2 4
χ(c,h) (τ ) = . (2.47)
ϕ(q)
The character can be expressed also in terms of the Dedekind η(τ ) function,
∞
1 1
η(τ ) ≡ q 2 4 ϕ(q) = q 2 4 (1 − q n ), (2.48)
n =1
in the form
1 −c
q h+ 2 4
χ(c,h) (τ ) = . (2.49)
η(τ )
To get the character of a minimal model one has to determine the irreducible
Verma module using the quotient procedure discussed in the previous section.
We do not give the derivation here, just the final result, which is,
c
q h− 2 4 ( 2 p p n + p r −p s ) 2 ( 2 p p n + p r + p s ) 2
χ(c(p, p ), hr s (p, p )) = = q 4 p p
−q 4 p p
, (2.50)
ϕ(q)
n ∈Z
where
(p − p )2
c(p, p ) = 1 − 6 , (2.51)
pp
and
(pr − p s)2 − (p − p )2
hr s (p, p ) = . (2.52)
4pp
Note that these are the non-unitary minimal models, except for the cases p − p =
±1, which coincide with the cases of the previous section with the identification
of p = m or p = m.
<0|U −1 φ1 (z1 , z̄1 )U . . . U −1 φn (zn , z̄n )U |0> = <0|φ1 (z1 , z̄1 ) . . . φ − n(zn , z̄n )|0> .
(2.53)
Recall that by definition a primary field of dimension h transforms under an
SL(2, C) transformation z → f (z) = az cz +d (with ad − bc = 1), as,
+b
Let us mention that SL(2, C) invariance holds for CFT in any dimension.
The invariance of the vacuum implies, in infinitesimal form,
<0|[Lk , φ1 (z1 , z̄1 )] . . . φn (zn , z̄n )|0> + . . . <0|φ1 (z1 , z̄1 ) . . . [Lk , φn (zn , z̄n )]|0> = 0,
(2.55)
for k = 0, ±1. Using [Lk , φ(z, z̄)] = h(k + 1)z k φ(z, z̄) + z k +1 ∂φ(z, z̄) we get Ward
identities in terms of differential equations:
k = −1 : ∂i <0|φ1 (z1 , z̄1 ) . . . φn (zn , z̄n )|0> = 0
i
k= 0: (zi ∂i + hi ) <0|φ1 (z1 , z̄1 ) . . . φn (zn , z̄n )|0> = 0
i
k = +1 : (zi2 ∂i + 2zi hi ) <0|φ1 (z1 , z̄1 ) . . . φn (zn , z̄n )|0> = 0. (2.56)
i
These Ward identities are associated with the invariance under translations,
dilations and special conformal transformations. Applying these equations to the
two point function one finds that,
c2
G2 (z1 , z̄1 , z2 , z̄2 ) ≡ <0|φ1 (z1 , z̄1 )φ1 (z2 , z̄2 )|0> = ,
(z1 − z2 )2h 1 (z̄1 − z̄2 )2 h̄ 1
(2.57)
where c2 is a constant, to be put to 1 in the normalization (2.11). Note also
that when taking two different fields φ1 and φ2 , SL(2, C) implies that h1 = h2
is necessary for a non-zero two-point function.
In a similar manner the three-point function is given by,
1
G3 (zi , z̄i ) = c123 h 1 +h 2 −h 3 h 1 +h 3 −h 2 h 2 +h 3 −h 1
z → z̄, h → h̄ , (2.58)
z12 z13 z23
w1 w1
w2 w2
w4 w4
w3 w3
⎡⎛ ⎞ ⎤
−(h i +h j )+h/3 ⎠
G4 (zi , z̄i ) = f (Z, Z̄) ⎣⎝ zij z → z̄, h → h̄ ⎦ , (2.59)
i< j
4
where h = i= 1 hi and the cross ratio Z is defined as Z = zz 11 23 zz 32 44 , which is an
SL(2, C) invariant.
For a general n-point function, denoting
the power of zij by −hij , we get,
2
hij = n −2 (hi + hj ) − (n −1)(n
2
−2) h for n ≥ 3.
So far we have implemented the global Ward identities. To get the local Ward
identity one performs a conformal transformation of an n-point function of pri-
mary fields Gn . This is achieved by integrating (z)T (z) along a contour C which
bounds a region that includes all the operators (see Fig. 2.2)
Now using analyticity one can deform the contour into a sum of countours
each of which encircles one operator. The result of the integral is therefore,
% &
dz
(z)T (z)φ(w1 , w̄1 ) . . . φ(wn , w̄n )
2πi
n % &
dz
= φ(w1 , w̄1 ) . . . (z)T (z)φ(wi , w̄i ) . . . φ(wn , w̄n )
i= 1
2πi
n
= φ(w1 , w̄1 ) . . . δ φ(wi , w̄i ) . . . φ(wn , w̄n ) . (2.60)
i= 1
Using (2.9) we substitute now for δ φ(wi , w̄i ) = (wi )∂ + h∂(wi )φ(wi , w̄i ).
Since this holds for arbitrary we can get a local form of the Ward identity,
n
hi 1 ∂
T (z)φ(w1 , w̄1 ) . . . φ(wn , w̄n ) = +
i= 1
(z − wi )2 (z − wi ) ∂w i
φ(w1 , w̄1 ) . . . φ(wn , w̄n ) , (2.61)
similar to the transition from (2.9) to (2.10). It is thus clear that the correlation
function above is a meromorphic function of z with singularities at the positions
of the operators.
A useful tool for computing correlators is the use of null vectors. Rather than
discussing this for a general null vector we demonstrate this procedure on a level
two null vector. Recall that in models with a primary of weight h such that c =
2h+ 1 (5 − 8h) there is a null vector at level two of the form (L−2 + aL−1 )Φ
2h 2 (h)
=0
where a = − 2(2h+1) . As L−1 φ (z) = ∂φ (z) one can trade L−2 φ (z) with
3 (h) (h) (h)
This exact differential equation will enable us to compute the four-point function
for the Ising model as we discuss in Section (2.13).
Next we would like to deduce the implications of the associativity on correla-
tion functions of primaries and descendant operators.
i l
i l
Σp Clmp Cijp p
= Σq Cilq Cjmq q
j m
j m
0 1
i l
−
Fijlm (p|x) F ijlm (p|x−) = p
j m
x ∞
Fig. 2.4. Single channel amplitude.
The requirement that the two ways of computing coincide, referred to as crossing
symmetry, is expressed in Fig. 2.3.5
Using conformal transformations, we can relate the diagram on the left-hand
side of Fig. 2.3 to the diagram drawn in Fig. 2.4, which corresponds to the
sum of the contributions of intermediate states belonging to the conformal fam-
ily [φp ] with the four-point function of operators located at (w1 , w2 , w3 , w4 ) =
(0, z, 1, ∞). Note that in such a situation, z is actually also the cross ratio Z. We
denote this amplitude by the conformal block Fijk l (m|z)F̄ijk l (m|z̄) which depends
on the Virasoro anomaly of the theory and the dimensions of all the operators
involved. In terms of conformal blocks the crossing symmetry condition takes
the form,
Cij m Ck lm Fijk l (m|z)F̄ijk l (m|z̄)
m
jl jl
= Cij n Ck ln Fik (n|1 − z)F̄ik (n|1 − z̄). (2.65)
n
For a given set of conformal blocks (2.65) is a set of equations that determine the
Cij k and the weights. The general set of solutions of these equations is not known,
but for a particular class of theories like the minimal models these equations can
be solved.
(n ) S̃ in
which means that λi = S̃ 0n
and therefore,
S̃jl S̃il (S̃ −1 )kl
Nijk = . (2.69)
l
S̃0l
Now, for the reader who knows about the τ parameter and the characters
(discussed in Section 2.8), we recall that under the S-transformation τ → − τ1
the characters of a given CFT transform as,
1
χj (− ) = Sjk χk (τ ). (2.70)
τ
k
Expressing the Dirac fermion in terms of chiral (or Weyl) fermions, a left ψ and
a right ψ̃, with Ψ ≡ (ψ, ψ̃), and using the fact that in two dimensions one can
take γ 0 = σ 2 and γ 1 = σ 1 , we rewrite the action as,
1
S= d2 z(ψ † ∂ψ
¯ + ψ̃ † ∂ ψ̃). (2.73)
4π
We remind the reader that ∂ ≡ ∂z and ∂¯ ≡ ∂z̄ . The equations of motion are,
¯ = 0 ∂ ψ̃ = 0 → ψ = ψ(z)
∂ψ ψ̃ = ψ̃(z̄). (2.74)
From this point on we discuss the theory of single Majorana fermions, namely χ
and χ̃ are a left and a right Weyl–Majorana fermion with the action,
1 ¯ + χ̃∂ χ̃).
= d2 z(χ∂χ (2.78)
8π
which implies that χ is a primary field of conformal dimensions of ( 12 , 0), and sim-
ilarly χ̃ with (0, 12 ). The Virasoro anomaly, which comes as usual from T (z)T (w),
is c = 12 , and c̄ = 12 from T̄ (z)T̄ (w).
Recall that the energy-momentum tensor of the scalar field (1.27) takes the
form of a bilinear of the “current algebra” currents (1.24). We want to examine
now if such a construction can be applied also for the fermionic fields. Since for
Weyl–Majorana fermions there are no such currents it is left only to check for
the T (z) of Weyl fermions. Let us note first the OPE of the currents and the
Weyl fermions that read,
ψ(z) ψ † (z)
J(z)ψ(w) = −i J(z)ψ † (w) = i , (2.84)
(z − w) (z − w)
{ψr , ψs } = δr +s . (2.88)
The form (2.82) holds for the periodic case. For the anti-periodic case there is
an extra factor of 12 ( wz + wz ), which tends to 1 as z → w.
The canonical quantization conditions in terms of real space-time coordinates
take the following form,
1
{ψ(x1 ), ψ(y 1 )}|x 0 =y 0 = δ(x1 − y 1 ), (2.89)
2
since ψ is the conjugate momentum of itself. Combining two Majorana fermions
into a Dirac one yields the following anti-commutation relations for the Dirac
fermions,
where Z = z1 2 z3 4
z1 3 z2 4 . Using also z1 4 z2 3
z1 3 z2 4 = 1 − Z, we can rewrite as,
−1/8 −1/24
G4 (z1 , z̄1 . . . z4 , z̄4 ) = f˜(Z, Z̄) (z13 z24 ) [Z (1 − Z)] (C.C.) . (2.96)
7 The two-dimensional Ising model has a long history. It was discussed in [137]. The relation
to Majorana fermions was discussed in [187].
If we now substitute this ansatz into (2.94) we find the following differential
equation for f ,
Z(1 − Z)∂ 2 + (1/2 − Z)∂ + 1/16 f (Z, Z̄) = 0. (2.98)
A similar equation holds for Z̄. The solutions of this differential equation are
√ 1/2
f1,2 (Z) = 1 ± 1 − Z and so finally the four-point function takes the form,
) )1/4
) z13 z24 )
G4 (z1 , z̄1 . . . z4 , z̄4 ) = )) )
) |f1 (Z)|2 + |f2 (Z)|2 , (2.99)
z12 z23 z34 z34
where the unique combination is dictated by the requirement for a single value.
This is identical to the result found in the Ising model for G4 .
Note also that f (1 − Z, 1 − Z̄) is a solution, a result of the symmetry under
the interchange of z1 with z3 .
Note also that although Φ(1/2) is a free fermion, Φ(1/16) cannot be constructed
in a local way from the fermion.
[T a , T b ] = ifcab T c , (3.1)
associated with a group G, namely, the set T a are the generators of the group
G.1 We will consider simple groups, namely those that do not contain invari-
ant subgroups. Denote the maximal set of commuting Hermitian generators by
H i , i = 1, . . . , r so that
[H i , H j ] = 0 i, j = 1, . . . , r. (3.2)
[H i , E α ] = αi E α , i = 1, . . . , r. (3.3)
The r-dimensional vector α is called a root associated with the step operator E α .
The roots are real and up to multiplication with a scalar there is a single E α
associated with α via (3.3). No multiple of a given root α is a root apart from
−α which is the root paired with E −α = E α † . The number of roots is obviously
(dimG − rankG).
The rest of the algebra are the commutation relations [E α , E β ] which follow
from the Jacobi identity,
[H i , H j ] = 0 i, j = 1, . . . , rank(G)
[H i , E α ] = αi E α α = 1, . . . , (dim G − rank(G))
α β α +β
[E , E ] = (αβ)E if α + β is a root
2α · H
= if α + β = 0
α2
= 0 otherwise. (3.5)
This basis of the algebra is a modified version of the Cartan–Weyl basis. The
constants (αβ) can be chosen to be ±1 if all the root vectors have the same
length.
1 Finite-dimensional Lie algebra is covered in many books, for instance Cahn [54].
where the ni are integers which are either all ni ≥ 0 or all ni ≤ 0. In the former
case α is positive, while in the latter it is negative. This base is called the basis
of simple roots. Associated with the simple roots one defines the simple Weyl
reflections σα ( i ) that generate the entire Weyl group.
The scalar products of simple roots define the Cartan matrix as follows:
2α(i) · α(j )
Aij = 2 . (3.10)
α(j )
The Cartan matrix is a rank(G) × rank(G) matrix with integer components and
with diagonal elements which take the value of 2. The off diagonal elements are
either negative or vanishing.
Ar su(r+1)
1 2 3 r−2 r−1 r
r−2 r−1 r
Br so(2r+1)
1 2 3 r−2 r−1 r
Cr sp(r)
1 2 3 r−2 r−1 r
r
Dr r−2 so(2r)
1 2 3 r−3 r−1
E6
E7
E8
F4
G2
Fig. 3.1. Dynkin diagrams for all the simple Lie algebras.
It can be shown that the roots of a simple Lie algebra can have at most two
different lengths, a long root and a short one. The ratio between their lengths
are either 2 or 3. When all the roots have the same length the algebra is a simply
laced algebra. From the Cartan matrix Aij one can reconstruct a basis of simple
roots up to a scale and overall orientation. In fact constructing all the roots from
the simple roots, one finds that full information on G is encoded in Aij .
The information contained in the Cartan matrix Aij may be encoded in a
planar diagram, the Dynkin diagram. The construction of such a diagram is as
follows:
r To each simple root α assign a node in the diagram.
(i)
r If a node represents a short root mark it by a black dot, and if a long one by
a white dot.
r Join the points α and α by Aij Aj i lines. For i = j, Aij Aj i can take the val-
(i) (j )
ues of 0, 1, 2, 3. In fact since Aij Aj i = 4cos2 θij , where θij is the angle between
the two roots, then orthogonal simple roots are not connected, and those with
an angle of 120, 135, 150 degrees are connected with one, two, or three lines.
r In some conventions one does not separate between black and white dots, but
2 2
rather one draws an arrow pointing from α(i) to α(j ) when α(i) > α(j ).
The Dynkin diagrams for all simple Lie algebras are given in Fig. [3.1].
H i |μ = μi |μ . (3.11)
2α · μ/α2 ∈ Z, (3.12)
E α |μ0 = 0, (3.14)
for every α > 0. All the states of a given irreducible representation can be built
by acting on the highest weight state with lowering operators, namely, each state
takes the form,
E −β 1 . . . .E −β n |μ0 . (3.15)
In fact every irreducible representation has a unique highest weight state |μ0 and
the other weights μ have the property that μ0 − μ is a sum of positive roots. The
highest weight state is a dominant weight. In the opposite direction for each dom-
inant weight there is a unique irreducible representation for which it is the highest
weight state. As was mentioned for the adjoint representation the weights are
s1 1
α2 α1+α 2
ω1
− α1 ω2 α1
s1 s 2 s2
− α1−α2 −α 2
s1 s2s1 s2 s1
the roots so that the corresponding highest weight state is the highest root. The
difference between the highest root and any other root is a sum of positive roots.
We end this subsection with an example. Consider the SU (3) algebra. The
Cartan matrix for this algebra takes the form,
2 −1
. (3.16)
−1 2
The simple roots are related to the fundamental weights as α(1) = 2λ(1) −
λ(2) and α(2) = −λ(1) + 2λ(2) . The scalar products between the fundamental
weights are (λ(1) , λ(1) ) = (λ(2) , λ(2) ) = 2/3 and (λ(1) , λ(2) ) = 1/3, using the stan-
dard normalization of (α(i) , α(i) ) = 2. The full Weyl group is given by W =
{1, σ1 , σ2 , σ1 σ2 , σ2 σ1 , σ1 σ2 σ1 }. The action of the different elements of the Weyl
group on the two simple roots gives all roots.
The root system and the Weyl chambers are given in Fig. 3.2. The Weyl cham-
bers are separated by the dashed lines, and are specified here by the elements of
the Weyl group, with the latter denoted in the figure by si . The Weyl chambers
are defined by,
Cω = {λ|(ωλ, αi ) ≥ 0, i = 1 . . . r}.
2 Affine Lie algebras were introduced into the physics literature by Bardacki and Halpern in
[27]. Independently V. Kac [136] and R.V. Moody [164] introduced them in the mathematical
literature.
Unlike the simple Lie algebra, here we have an r + 1 dimensional abelian sub-
algebra consisting of [H01 , . . . , H0r , k̂]. With respect to these generators, Em
α
are
step operators,
[H0i , Em
α
] = α i Em
α
[k̂, Enα ] = 0. (3.20)
Using the generators (H0i , k̂, −L0 ) as in the Cartan subalgebra we have as the
step operators Enα corresponding to a root (α, 0, n) and Hni corresponding to
(0, 0, n).
The root system of ALA is thus infinite but spans an (r + 1) dimensional space.
We can divide the roots into positive and negative according to the following rule:
where αi is the basis of simple roots of the Lie algebra, and θ is the highest root.
Thus an arbitrary root of the ALA can be expressed as,
r
α= ni α(i) . (3.24)
i= 0
It is positive if ni ≥ 0 and negative when ni ≤ 0, and these are the only two
posibilities.
is
ai · aj = αi · αj . (3.28)
The root that corresponds to Enα , a = (α, 0, n) has a norm a2 = α2 > 0 and hence
is referred to as a space-like root, whereas the root that is associated with Hni ,
nδ = (0, 0, n) has zero norm (light-like) and is orthogonal to all other roots. We
have used the “unit” of δ = (0, 0, 1).
The Cartan matrix of ĝ, which is an (r + 1) × (r + 1) matrix, is defined in a
similar way to one of the Lie algebra, namely,
2a(i) · a(j )
Aij = 0 ≤ i, j ≤ r. (3.29)
a2(j )
We add to the Cartan matrix of the Lie algebra, the extra row and column Ai0
and A0i which can be found using (3.10) with α0 = −θ. Now from the definition
r
of the fundamental weight it follows that θ = − i=0 A0i λi . Since θ is a long
root of g, namely, θ2 ≥ α(i)
2
one gets that,
−Ai0 = 1 if A0i = 0
−Ai0 = 0 if A0i = 0, (3.30)
provided that θ is not itself a simple root, as happens for SU (2). The Dynkin
diagram of ĝ is obtained using that of g appended with an extra point that
corresponds to α0 connected by −A0i lines to the points a(i) . If −A0i > 1 an
arrow is drawn which points toward a(0) . We demonstrate the construction in
the following example:
SUˆ (2) - There are only two simple roots a(0) = (−α, 0, 1) and a(1) = (α, 0, 0)
so that A0i = Ai0 = −2 and the Cartan matrix is
2 −2
. (3.31)
−2 2
Thus, there are two roots of equal length connected by four lines with arrows
pointing in both directions to indicate that a2(0) is equal to a2(1) .
The Dynkin diagrams of the affine simple algebra are shown in Fig. 3.3. The
point that corresponds to α(0) is marked by a zero. The black dots relate to the
notion of twisted affine Lie algebra which we do not discuss here (see for example
[111]).
A1 E6
1 3 4 5 6
0
Ar 2
r>2
E7
1 2 r−1 r 0 1 3 4 5 6 7
2
0
Br E8
r>3 1 2 3 r−2 r−1 r 1 3 4 5 6 7 8 0
Cr F4
r>2 0 1 2 r−2 r−1 r 0 1 2 3 4
0 r
Dr G2
r>3
1 2 3 r−3 r−2 r−1 0 1 2
Fig. 3.3. Dynkin diagrams for all the affine simple Lie algebras.
for a = (α, 0, nα ) and b = (β, 0, nβ ) a space-like root. Light-like roots are invari-
ant under such reflections σa (nδ) = nδ. In fact the Weyl group of ĝ is a semi-
direct product of the Weyl group of g and the coroot lattice of g which is the
lattice generated by the coroots αν = αα2 . The coroots form the root system of
the Lie algebra dual to g obtained by interchanging the root lengths. The simply
laced algebras An , Dn , En are obviously self-dual, as are F4 and G2 , whereas
Bnν = Cn and Cnν = Bn .
for n > 0, α > 0. The eigenvalue of this state is the highest weight vector μ̂0 =
(μi0 , k, h) given by,
r
level = ni mi . (3.37)
i= 0
Level 1 representations are thus associated with highest weights l(i) with all
mi = 1. Those are indicated by open points in Fig. 3.3.
From the definition of mi it follows that,
r
Aij mj = 0. (3.38)
i= 0
Since the Cartan matrix has the basic symmetry of the extended Dynkin dia-
gram, also the positions of the open dots have to preserve this symmetry. For
the classical groups Ar , Br , Cr , Dr the values of mi for the closed dots is 2. For
the exceptional groups the vector (m0 , . . . , mr ) is as follows
Ê 6 (1, 1, 2, 2, 3, 2, 1)
7
Ê (1, 2, 2, 3, 4, 3, 2, 1)
8
Ê (1, 2, 4, 6, 5, 4, 3, 2, 1)
4
F̂ (1, 2, 3, 2, 1)
2
Ĝ (1, 2, 1) (3.39)
We can now use the OPE to evaluate the infinitesimal transformation of the
current under ALA transformations,
1
δ J a (w) = dzb (z)J b (z)J a (w)
2πi w
1 kδ ab fcab J c (w)
= dzb (z) + i = ifbc J − k∂a .
a b c
(3.41)
2πi w (z − w)2 (z − w)
The same structure also holds for J¯a (z̄).
The OPE form of the ALA can be transformed into a commutator form of the
algebra. We introduce a Laurent expansion of the currents,
1
J a (z) = Jna z −(n + 1) Jna = dzz n J a (z). (3.42)
n
2πi
Substituting the OPE into the expression of the commuation relation we indeed
find the ALA of (3.17), namely
a
[Jm , Jnb ] = ifcab Jm
c
+n + k̂mδab δm +n ,0 . (3.43)
In free scalar theory there are two “currents” which are holomorphically con-
served, J and T , and moreover the energy-momentum tensor is bilinear in J,
as was shown in Section 1.5. We now elevate this special case into a general
construction of T for theories which admit ALA structure. The construction is
known as the Sugawara construction. One writes T as a normal ordered product
of the currents,
1
T (z) = : J a (z)Ja (z) : (3.44)
2κ
with a coefficient κ that has to be determined quantum mechanically. In fact,
one way to determine κ is by requiring that J a is a primary field of weight 1,
namely,
Ja ∂J a (w)
T (z)J a (w) = + . (3.45)
(z − w)2 (z − w)
Using the OPE (3.40) and the relation −fcab fbd
c
= 2Cδda , where C is the dual
(k +C )
Coxeter number, we find that κ = 1 so that the form of the Sugawara con-
structed T is,
1
T (z) = : J a (z)Ja (z) : (3.46)
2(k + C)
Note that the Casimir of the adjoint is 2C. Note also that in Section 1.5, for the
free scalar case, we had a relative minus sign, due to a difference of factor i in
defining the currents there.
In the WZW models discussed in the next section, classically one has T with a
1
coefficient of 2k . It is thus clear that for those models quantum mechanically, due
to the double contraction, we get a finite renormalization of the level k → k + C.
where here the normal ordering implies putting the currents with positive m to
the right. In fact normal ordering is required only for L0 .
Using this relation, we write down the full Virasoro algebra and ALA,
c
[Ln , Lm ] = (n − m)Ln +m + (n − 1)n(n + 1)δ(n + m)
12
a
[Ln , Jm ] = −mJm a
+n
a
[Jm , Jnb ] = ifcab Jm
c
+n + k̂mδab δm +n ,0 . (3.51)
manner ALA primaries Φl, ¯l (z, z̄) are defined via their OPE with J,
Tla Φl, ¯l (w, w̄)
J a (z)Φl, ¯l (w, w̄) =
(z − w)
a
T¯ Φl, ¯l (w, w̄)
J¯a (z)Φl, ¯l (w, w̄) = l , (3.52)
(z̄ − w̄)
where Tla , T¯la are the matrix T a in the l, ¯l representations, for the holomorphic
and antiholomorphic sectors, respectively. From here on we discuss only holo-
morphic properties. In terms of the Laurent components Jna the condition for a
primary field reads,
Jna Φl, ¯l (0, z̄) = 0 for n > 0; J0a Φl, ¯l (z, z̄) = Tla Φl, ¯l (z, z̄). (3.53)
J a (0) = J−1
a
I. (3.56)
Note however that J a (z) is a Virasoro primary. Apart from the distinguished
descendant J there are descendant operators of all the primaries. In fact all the
local operators can be written as,
ā
a1
J−n 1
aN
. . . J−n N
J¯−n̄
ā 1
1
. . . J¯−n̄
N̄
N̄
Φl, ¯l (z, z̄), (3.57)
and in the case of a Sugawara construction all the operators are of the form,
ā N̄
a1
L−m 1 . . . L−m M L̄−m̄ 1 . . . L̄−m̄ M̄ J−n 1
aN
. . . J−n N
J¯−n̄
ā 1
1
. . . J¯−n̄ N̄
Φl, ¯l (z, z̄). (3.58)
let us now define a character of the CFT and ALA module λ̂ as follows:
χλ̂ (z j , τ ) = e−im λ̂ δ T rλ̂ [e2π iτ L 0 e−2π i j zj hj
], (3.59)
where mλ̂ , δ and hj are the generators of the Cartan subalgebra associated with
the group, and zj are complex numbers. The character can also be expressed in
terms of the generalized theta function Θλ̂ in the following form:
w ∈W (w)Θw ( λ̂+ ρ̂)
χλ̂ (z , τ ) =
j
, (3.60)
w ∈W (w)Θw ρ̂
where the sums are over the elements of the finite Weyl group, (w) = (−1)l(w )
with l(w) the length of w.
Rather than defining the generalized theta function for any ALA at any level,
ˆ (2) level k. For this case we have,
we define here only the function for SU
λ2τ
Θλ 1 (z; τ ; t) = e−2π k t e−2π i[k n z + 2 λ 1 z −k n τ−
(k ) 1 2 1 ]
4k . (3.61)
n ∈Z
(k )
with Θλ 1 (z; τ ; o) ≡ Θ(k − λ1 , λ1 ) (see [77] for details).
In terms of this function the character of SU ˆ (2)k takes the form,
(k + 2) (k + 2)
Θλ 1 + 1 − Θ−λ 1 −1
χλ̂ = (2) (2)
, (3.62)
Θ1 − Θ−1
Now from the OPE (3.52) we know that δ φl i (wi , w̄i ) = a (wi )Tlai φl i (wi , w̄i ).
Since this holds for arbitrary we can get a local form of the Ward identity in
the form,
Tlai
n
J a (z)φl 1 (w1 , w̄1 ) . . . φl n (wn , w̄n ) = φl 1 (w1 , w̄1 ) . . . φl n (wn , w̄n ) .
i= 1
(z − wi )
(3.66)
As for the global Ward identity, we use the fact that the correlator has to be
invariant under global g transformations (constant a ), namely,
leading to
n
Tlai φl 1 (w1 , w̄1 ) . . . φl n (wn , w̄n ) = 0. (3.67)
i= 1
Null vectors of CFTs were found to be useful in Section 2.9, since they lead to
differential equations for certain correlators. In a similar manner one can write
down null vectors of ALA. In the context of the Sugawara construction, due
to the link between the Virasoro algebra generator T and the ALA generators
J a , there are null vectors that combine generators from both infinite algebras.
We discuss now an important example of this class that leads to the Knizhnik–
Zamolodchikov equations. Consider, at Virasoro level one, the following null
vector,
* +
1
|null> = L−1 − a
J−1 Tlai |Φl i> . (3.68)
k+C
It is easy to see that this is indeed a null state, following (3.52). If we insert
the corresponding null operator into a correlation function of primary fields, like
<Φ1 (z1 ) . . . null(zi ) . . . Φn (zn )>, the latter must vanish and hence we get,
⎧ ⎫
⎨ 1 Tia Tja ⎬
∂i − <Φ1 (z1 ) . . . Φn (zn )> = 0. (3.69)
⎩ k+C (zi − zj ) ⎭
j = i
For the case of four-point functions, as the correlator depends only on the cross-
ratio coordinate Z = zz 11 23 zz 32 44 , the partial differential equations reduce to an ordi-
nary differential equation. In Section 4.4 we will demonstrate a solution of the
Knizhnik–Zamolochikov equation for a four-point function.
2
3.7.1 Free Majorana fermions and SO(N)
Consider a generalization of the action given in Section 2.11 for N Majorana
fermions,
N
1 ¯ i + ψ̃i ∂ ψ̃i },
S= d2 z {ψi ∂ψ (3.71)
8π i= 1
where ψ and ψ̃ are left and right Weyl–Majorana fermions, respectively. Note
that this is possible in 2d, and in any other dimension that is 2 modulu 8. In 4d,
for example, we do not have a Weyl–Majorana fermion, as in the case of a single
Majorana fermion, due to the equations of motion,
3 The free fermion realization of ALA was presented for the first time in [27].
However, unlike the case of a single fermion, here the action is invariant under
2 ) affine algebra generated by the following
transformations associated with SO(N
holomorphically (anti-holomorphically) conserved currents,
1 1
J a (z) = ψ i Tija ψ j J¯a (z̄) = ψ̃ i Tija ψ̃ j , (3.73)
2 2
where T a are SO(N ) matrices which can be expressed as,
(k l)
Tija ≡ tij = i(δik δjl − δjk δil ). (3.74)
The coefficients (halfs) are not determined by the Noether procedure, but are
chosen in a manner that will be explained below.
The Tija matrices obey the relations,
T r[T a T b ] = 2δ ab
Tija Tkal = −δik δj l + δil δj k
a
fabc fabd = 2(N − 2)δcd . (3.75)
ab
The anticommutation relation and the OPE generalize in an obvious way those
of the single Majorana fermion, namely,
1
{ψ i (x0 , x1 )ψ j (y0 , y1 )}|x 0 =y 0 = δ ij δ(x1 − y1 ) (3.76)
2
and
δ ij δ ij
ψ i (z)ψ j (w) = ψ̃ i (z)ψ̃ j (w) = . (3.77)
z−w z̄ − w̄
Now using this OPE one can derive the OPE of two currents and verify that
they take the form of (3.40),
1
J a (z)J b (w) = : ψ i (z)Tija ψ j (z) :: ψ k (w)Tkbl ψ l (w) :
4
1 δ ik
= Tija Tkbl − : ψ i (z)ψ k (w) : + ψ j (z)ψ l (w)
4 z−w
δ il 1 1
+ : ψ i (z)ψ l (w) : ψ j (z)ψ k (w) = Tija Tkbl
z−w 4 z−w
[ −δ ik : ψ j (z)ψ l (w) : +δ il : ψ j (z)ψ k (w) : +δ j k : ψ i (z)ψ l (w) :
1 1 ik j l
− δ j l : ψ i (z)ψ k (w) :] + Tija Tkbl −δ δ + δ il δ j k . (3.78)
4 (z − w)2
By expanding the fields that are functions of z around w and using the relations
above one finds that indeed the OPE of the two currents take the form of (3.40),
namely,
1δ ab fcab J c (w)
J a (z)J b (w) = + + finite terms (3.79)
(z − w)2 (z − w)
It is thus clear that this is a realization of an SO(N ) ALA of level k = 1.
which implies that ψ i are N primary fields of conformal dimensions ( 12 , 0), and
similarly ψ̃ i has dimension (0, 12 ).
Is the primary field Φ(1/2,1/2) (z z̄) = ψ(z)ψ̃(z̄) the only primary operator (in
addition to the identity operator that corresponds to the vacuum state)? For the
2 )1 we find (see (2.13)) that there is also one primary
primaries of the ALA SO(N
N
operator with dimension 16 for odd N , and two primary operators for even N .
These additional primaries transform in the spinor representation of SO(N 2 )1 .
Can one construct these primaries in terms of the fermionic fields ψ and ψ̃ ?
The situation here is similar to the one in the Ising model. In fact, using the spin
operator σ(z, z̄) or its dual, one indeed gets from the N independent Majorana
N
fermion theories, the dimension 16 and the number of degrees of freedom 2N ,
which are identical to the dimension of the spinor representation.
So far we have shown the free fermion construction of SO(N 2 )1 , namely, of
the ALA at level 1. We would now like to investigate the possibility of having
free fermion realization also to the affine Lie algebra at higher levels. Going
through our previous derivation it is clear that the ALA structure of the OPE
of two currents (3.40), applies to fermions at any representation. For a given
representaion ρ the corresponding level k is determined from the first term on
the right-hand side, namely T r(T a T b ) = 2kδ ab . Now since for a representation ρ,
T r(Tρa Tρb ) = 2D2 (ρ)δ ab where D2 (ρ) is the Dynkin index of the representation,
2 ) at level D2 (ρ).
it is clear that free fermions constitute a realization of SO(N
3.8 Free Dirac fermions and the U(N)
Consider the theory of N Dirac fermions described by the following action,
1 ¯ i + ψ̃ i† ∂ ψ̃i }.
S= d2 z{ψ i† ∂ψ (3.83)
4π
In terms of symmetries, the difference between this theory and the one of a single
Dirac fermion, is that now there is an invariance under U (N ) left holomorphic
and right anti-holomorphic transformations, namely,
where g(z), ḡ(z̄) ∈ U (N ). The associated holomorphic currents are given by,
† †
J a = ψ i T a ji ψj J = ψ i ψi , (3.85)
where J is the U (1) current, J a (z) are the SU (N ) currents and T a ji are matrices
in the adjoint of SU (N ), that obey
T r[T a T b ] = δ ab
1
Tija Tkal = δil δj k − δij δk l
a
N
fabc fabd = N δcd . (3.86)
ab
Using the OPEs of the fermions, it is straightforward to realize that the cur-
rents indeed constitute the OPEs that correspond to a U (N ) of level k = 1,
1δ ab f ab J c (w)
J a (z)J b (w) = + c + finite terms
(z − w) 2 (z − w)
1
J(z)J(w) = + finite terms
(z − w)2
J a (z)J(w) = finite terms. (3.87)
Similar to the case of Majorana fermions, the Noether current T is given by,
1 † †
T (z) = T (z)U (1) + T (z)S U (N ) = − [ψ i ∂ψi − ∂ψ i ψi ], (3.88)
2
and can be reexpressed in terms of a Sugawara form,
1 i j
T (z)U (1) = : ψ † ψi ψ † ψj :
4N
1 i j
T (z)S U (N ) = : ψ † T a ψi ψ † T a ψj : . (3.89)
2(N + 1) a
Since a Dirac fermion has a c = 1 Virasoro anomaly, it is clear that the theory of
N Dirac fermions has c = N . This is also the outcome of the Virasoro anomaly
associated with the Sugawara form as follows,
N2 − 1
cU (1) + cS U (N ) = 1 + = N. (3.90)
N +1
with
¯ = ∂ J¯ = 0
∂J
and
u ∈ G, G = SO(N ) or SU (N ), (4.2)
and the currents corresponding to the left and right multiplications take the
form,
1 −1 μ 1
Jμ = u ∂ u J˜μ = − ∂ μ uu−1 . (4.6)
4π 4π
Both currents are conserved. Note that the conservation of one implies the con-
servation of the other. However, unlike the massless free scalar theory, now we
do not have an ALA structure associated with a separate holomorphic and anti-
holomorphic conservation. The latter would have taken the form of JL (z) corre-
sponding to left transformation of the form u → g(z)u and JR (z̄), corresponding
to right transformation of the form u → ug(z̄). In a similar manner one finds
that the energy-momentum tensor,
Moreover, it can be shown that for an action whose variation takes the form of
(4.7) the energy-momentum takes the form,
T ∼ T r[JJ], T̄ ∼ T r[J¯J],
¯ (4.11)
Fig. 4.1. The map between the space-time S 2 and the group manifold.
maps from S 2 to the group space G vanishes, namely, π2 (G) = 01 for any non-
abelian group G.
On the other hand since π3 (G) = Z, there are topologically inequivalent ways
to extend the map u to a map from the ball to the group manifold. This implies
that there is an ambiguity in SWZ and it is well defined only modulo SWZ →
SWZ + 2π. Thus the coefficient of this term must be an integer k, and to have a
variation of the form (4.12) it is clear that the sigma term has to have the same
coefficient.
Let us now summarize. The classical action of the WZW model is,
k ¯ −1 ]
SWZW = d2 zT r[∂u∂u
4π
k
+ d3 σij k T r[(u−1 ∂i u)(u−1 ∂j u)(u−1 ∂k u)]. (4.14)
12π
The variation of this action is given by,
k k
δS = d2 zT r u−1 δu∂(u−1 ∂u)
¯ = d2 zT r δuu−1 ∂(u∂u
¯ −1
) , (4.15)
4π 4π
so that the equation of motion takes the form,
∂(u−1 ∂u)
¯ = ∂(u∂u
¯ −1
) = 0. (4.16)
where clearly u ∈ G, ū ∈ G.
We should state, that the form (4.14), with a term extended to one dimension
higher, follows from general properties. Equations of motion that we want, in
even space-time dimensions, imply a term in the action with one dimension
higher, otherwise the action will involve singular terms, like the introduction of
Dirac strings in the case of elementary monopoles.
The symmetries of the action are the ALA transformations,
it is easy to realize that the OPE that is compatible with such a transformation
is,
kδ ab ifcab J c
J a (z)J b (w) = + , (4.25)
(z − w)2 (z − w)
which is the OPE associated with the ALA discussed in Section 3.2.
Next we want to determine the conformal properties of u, in particular its
confomal dimension. The classical form of the currents (4.20) is elevated to the
quantum expression via the equations,
where κ, which is a renormalized level, will be determined shortly, and the nor-
mal ordering refers as usual to subtracting the singular parts of the product.
Assuming that u is an ALA primary field, the OPE takes the form,
∞
c2
J a (z)T a u(w, w̄) = u(w, w̄) + κ∂u(w, w̄) + (z − w)n −1 T a J−n
a
u(w, w̄),
z−w n =2
(4.27)
where (4.26) was inserted as the (z − w)0 term, and c2 is the quadratic Casimir
a a
operator in the representation of u, a T T = c2 . If we assume that u is a
Virasoro primary as well, then L−1 u = ∂u, so that combined with (4.27) one can
write down the following null vector,
vnull ≡ (J−1
a
T a − κL−1 )u = 0. (4.28)
L0 Vnull = (h + 1)Vnull
J0a Vnull = T a Vnull
Ln Vnull = Jna Vnull = 0 for n > 0, (4.29)
where L0 u = hu. For n = 1 the conditions L1 Vnull = J1a Vnull = 0 imply that the
renormalized level and the conformal dimension of u take the form,
1 c2
κ= (C2 + k) h = , (4.30)
2 C2 + k
where C2 is the quadratic Casimir in the adjoint representation, defined as
f acd f bcd = C2 δ ab .
The use of null vectors and the differential equations that determine cor-
relators of primary fields were introduced in the landmark paper of Knizhnik
and Zamolodchikov [143]. An elaboration of the application of these equations
appears in [77]. This direction was further developed by Gepner and Witten
[108].
The WZW has an in-built Sugawara construction. In fact it is very often taken
as the prototype model for this structure. According to the discussion in Section
3.3 the quantum version of the classical energy-momentum tensor (4.21) takes
the form of (3.46),
1
T (z) = : J a (z)J a (z) :, (4.31)
2(k + C2 )
and the Virasoro anomaly of the model is,
k dim G
c= . (4.32)
k + C2
The Sugawara construction is described in [203]. This paper, however, does
not have the correct expression that includes the finite renormalization. This
was done later in the paper of Dashen and Frishman [73].
Recall from (2.59) that in general due to the conformal Ward identity the four-
point function can be written as,
where
z12 z34 z̄12 z̄34
x= x̄ = , (4.36)
z14 z32 z̄14 z̄32
+k ) .
Now G(x, x̄) can be decomposed into a sum of terms, each one representing a
conformal block, the latter having the form of a product of a holomorphic and
anti-holomorphic function,
G(x) = I1 G1 + I2 G2 , (4.38)
I1 ≡ δm 1 ,m 2 δm 3 ,m 4 I2 ≡ δm 1 ,m 3 δm 2 ,m 4 . (4.39)
If we now substitute this decomposed form of the four-point function into the
Knizhnik–Zamolodchikov equation (3.69) we find,
⎛ ⎞
t a
⊗ t a
⎝ ∂z i + 1 i j ⎠
[z14 z23 ]−4h (I1 G1 + I2 G2 ) = 0. (4.40)
k+N zi − zj
j = i
z1 = x, z2 = 0, z3 = 1, z4 = ∞, (4.41)
Extracting G2 from the first equation and plugging it back into the second
equation, the latter translates into a hypergeometric differential equation,
x(1 − x) 2 2 2
[N κ ∂x + A(x)∂x + B(x)]g1 (x) = 0, (4.44)
N2
where κ = k + N and with the following two possible values for A(x), B(x) and
the relation between g1 and G1 as
N (N + κ) N2 N4 − N2 + 2
A(x) = − Nκ B(x) = − ,
x 1−x x(1 − x)
1
G1 = [x(1 − x)] κ N g1+
or
−N (N − κ) N2 2(N 2 − 1)
A(x) = − Nκ B(x) = − ,
x 1−x x(1 − x)
N 2 −1
G1 = x− (1 − x) κ N g1−
1
κN (4.45)
The solutions of the differential equations are the following hypergeometric func-
tions,
1 1 N N −1 N +1 N
g1− = F ,− ,1 − ;x g1+ = F , ,1 + ;x . (4.46)
κ κ κ κ κ κ
With parametrization,
(m ) (n )
G(x, x̄) = Ii I¯j Gi,j (x, x̄) Gi,j = ξm n Gi Gj . (4.48)
i,j = 1,2 n ,m =+,−
which follows from the fact that under crossing symmetry I1 ↔ I2 . Single valued-
ness implies that ξ+− = ξ−+ = 0. To obey the crossing symmetry requirement
F (a, b, c; x) = A1 F (a, b, a + b − c; 1 − x)
+ A2 (1 − x)c−a−b F (c − a, c − b, c − a − b + 1; 1 − x), (4.50)
where
Γ(c)Γ(c − a − b) Γ(c)Γ(a + b − c)
A1 = A2 = . (4.51)
Γ(c − a)Γ(c − b) Γ(a)Γ(b)
Finally we find that
(−) (−) c2−− − 1 (+) (+)
Gij = Gi (x)Gj (x̄) + Gi (x)Gj (x̄), (4.52)
c2+−
where
Γ( Nκ )Γ(− Nκ ) [Γ( Nκ )]2
c−− = N c+− = −N . (4.53)
Γ( κ1 )Γ(− κ1 ) Γ( N κ+1 )Γ( N κ−1 )
For k = 1 we have c−− = −1, and hence1 the second term in (4.52) vanishes.
Using F N 1+ 1 , − N 1+ 1 , N 1+ 1 ; x = (1 − x) N + 1 the four-point function takes the
form,
1 1 1 ¯ 1 ¯ 1
G(x, x̄) = [xx̄(1 − x)(1 − x̄)] N I1 + I2 I 1 + I2 . (4.54)
x 1−x x 1−x
In Section 6.3 it will be shown that the WZW theory of U (N ) at level k = 1 is
a bosonized equivalent to that of N Dirac fermions controlled by the action,
¯ α + ψ̃ † ∂ ψ̃β ].
Sf = d2 z [ψα† ∂ψ (4.55)
β
where M is a mass scale and where ũ denotes the U (N ) group element. In the
theory of free Dirac fermions the four-point function of the fermion bilinears
takes the form,
1 1 1 1
G(x, x̄) = I1 + I2 I¯1 + I¯2 . (4.57)
x 1−x x 1−x
It is easy to see that by converting (4.54) to a similar correlator of U (N ) we find
exactly the same answer. This is done as follows: define the U (N ) group element
to be,
√
ũ(z, z̄) = ei 4π/N γ ϕ(z , z̄ ) u(z, z̄). (4.58)
The tangent space to the group G at the point u can be split into Tu G =
|| ||
Tu⊥ G ⊕ Tu G, where Tu G consists of vectors tangential to the orbit of Ad through
u. On Tu G, (1 − Ad(u)) = 0 and (1 + Ad(u)) = 2, so that (u−1 ∂x u)⊥ = 0 and
⊥
the corresponding D branes, namely the submanifolds where the condition (4.62)
is obeyed, coincide with the conjugacy classes. In the case that (1 − Ad(u)) is
it follows that,
It is also obvious that a similar relation holds with Lhm which is the Virasoro
generator built by a Sugawara construction from the currents of ĥ, namely,
L(g
m
/h)
≡ Lm − Lhm . (4.80)
[L(g
m
/h)
, L(g
n
/h)
] = [Lm , Ln ] − [Lhm , Lhn ]
(g /h) (m3 − m)
= (m − n)Lm +n + [c(ĝk ) − c(ĥk h )] δm +n ,0 . (4.81)
12
(g /h)
Thus we have just found that the Virasoro generators of the coset Lm obey a
Virasoro algebra with a central charge of
kdimg kh dimh
c= − . (4.82)
k + C2 (g) kh + C2 (h)
A special class of coset models are the diagonal coset models ĝ ⊕ĝ ĝ . The gen-
erators of the coset in this case are given by Jha = J(1)a a
+ J(2) , namely the sum
of the generators of each copy. It thus follows that the level of the coset must
a a
be the sum of the two levels since [J(1) , J(2) ] = 0. The coset therefore takes the
form,
ĝk 1 ⊕ ĝk 2
,
ĝk 1 +k 2
where CG is the second Casimir of the adjoint representation.3 The second term
can be viewed as S−(k + 2C G ) (h). Since the Hilbert space of the model decomposes
into holomorphic and anti-holomorphic sectors we restrict our discussion only to
the former.
3 This was C 2 (g) in the previous section; notation has changed according to the literature.
4 dG is what we called dim g in the previous section; changed according to the literature.
of this book. We thus do not discuss here the topological quantum field theory
aspects of the G
G models.
By construction of the BRST procedure the space of physical states of a
G model is given by the cohomology of Q. That is to say, a physical state |phys>
G
has to be closed under Q, namely Q|phys> = 0 and not exact, namely |phys> =
Q|state> where |state> is any other state. It can be shown that taking the trace
over those states one finds the torus partition function of the G G model which
is based on the decomposition into WZW characters, discussed in Section 3.5.
The torus partition function can be expressed as
∗
−r
Z G = cτ2 duZ g (τ, u)Z hh (τ, u)Z g h (τ, u), (4.96)
G
where du is the measure over the flat gauge connections on the torus and r is
the rank of G; Z g (τ, u) is the torus partition function of the Gk WZW model,
−c
Z g (τ, u) = (q q̄) 2 4 χk ,λ L (τ, u)χ̄k ,λ R (τ, u)Nλ R ,λ L , (4.97)
λ L ,λ R
where q = e2iπτ , λR , λL denote the Gk highest weights, and for the diagonal
invariant Nλ R ,λ L = δλ R ,λ L . The character can be written as,
Mk ,λ (τ, u)
χk ,λ (τ, u) = , (4.98)
M0,0 (τ, u)
∗
with Mk ,λ defined explicitly for the SU (2) case below. Z hh (τ, u) in (4.96) is
c
the contribution of h ∈ GG at level k + 2CG or equivalently h ∈ G at level −(k +
∗ c
2cG ). This takes the form Z hh (τ, u) ∼ |M0,0 (τ, u)|−2 indicating that GG contains
just one conformal block. It is straightforward to calculate Z g h (τ, u), the ghost
contribution to the partiton function Z g h (τ, u) ∼ |M0,0 (τ, u)|4 . Thus there is a
cancellation of the |M0,0 (τ, u)| factors and the resulting character is given by
the numerator of the character of the “matter” sector. In the G G model it is
Mk ,λ . This cancellation property leads to an index interpretation for Mk ,λ . For
G = SU (2) it was found that one can express,
∞
j+ 1
* +
2 2 1
Mk ,j (τ, θ) = q (k + 2)(l+ ( k + 2 ) ) sin πθ (k + 2)l + j + , (4.99)
2
l=−∞
as
2
1 ( j( k++122)) iπθ (j + 1 ) iπθ Jˆ(0tot )
Mk ,j (τ, θ) = q e 2 T r[(−)Ĝ q L̂ 0 e ], (4.100)
2i
where θ = Reu, Ĝ is the ghost number, L̂0 is the excitation level and Jˆ(tot) 0
is
0
the J(tot) eigenvalue of the excitation. Note that Mk ,j (τ, θ) is obtained from the
torus Mk ,j (τ, u) by restricting to just one angle. This amounts to considering the
propagation along a cylinder rather than around the torus. As long as we are
interested in the spectrum it is sufficient to consider Mk ,j (τ, θ). This index inter-
pretation enables us to read important information about the physical spectrum
5.1 Introduction
In the previous chapters we have addressed 2D field theories with no scale. As
we discussed in Chapter 2, one cannot define an S-matrix for such theories.
Generically physical systems are characterized by certain energy scales and the
notion of S-matrix plays an important role. It is thus time to move forward
and examine non-conformal field theories. Again we start our journey with the
theory of a free scalar field, but now a massive one. We then move on and discuss
interacting theories equipped with infinite numbers of conserved charges, the so-
called integrable models, that resemble the free massive theory in a way that will
be explained below.
∂ μ ∂μ φ + m2 φ = 0, (5.2)
is solved for the case of uncompactified space-time by the following Fourier trans-
form,
dk 1 1 −ik ·x
0 1
φ(x , x ) = √ √ a(k )e + a† (k 1 )eik ·x , (5.3)
2π k 0
where (k 0 )2 − (k 1 )2 = m2 .
A dramatic difference between the massless field discussed in Chapter 2, and
the massive one we discuss here, shows up when analyzing the symmetries of the
system.
The only transformations that leave the action invariant are the ISO(1, 1)
Poincare transformations, namely, the space and time translations and a single
Lorentz transformation. These are,
x0 → x0 + a0 x1 → x1 + a1
x0 → x0 + a01 x1 x1 → x1 + a10 x0 , (5.4)
where the transformation parameters are constants and a01 = a10 . The fact that
the parameters are constants, and not holomorphic and anti-holomorphic func-
tions of the complex coordinates, has a tremendous impact, since it implies the
absence of the powerful infinite-dimensional Virasoro algebra.
The corresponding Noether currents associated with the Poincare transforma-
tions are,
Tμν = ∂μ φ∂ν φ − gμν L,
μ μ01
JLor ≡ JLor = ρν T μρ xν . (5.5)
However, since the space-time is two dimensional, there is an additional conserved
current, the so-called topological current,
μ
Jtop = μν ∂ν φ, (5.6)
μ
which is conserved regardless of the equations of motion, since obviously = ∂μ Jtop
μν ∂μ ∂ν φ = 0. In fact this current is conserved for any interacting scalar field in
2d, and as we will see later on it plays an important role in the analysis of soliton
solutions of integrable models.
The theory of a free massive scalar field, as well as other scalar theories that
will be addressed in this chapter, are obviously invariant under the discrete
symmetry of,
φ → −φ. (5.7)
The canonical quantization was described in Section 1.6. The normal ordered
Hamiltonian and momentum expressed in terms of the creation and annihilation
operators take the form,
H = dk 1 (k 1 )2 + m2 a† (k 1 )a(k 1 ) P = dk 1 k 1 a† (k 1 )a(k 1 ). (5.8)
1 So far in the context of conformal field theory we have denoted the conformal dimension by
h. Here in the chapter on integrable models it will be denoted by Δ.
where ˙ and refer to time and space derivatives and V (φ) is a positive semi-
definite function of φ. The corresponding equation of motion is given by,
Let us assume that the potential has a set of N absolute minima at which it
vanishes, namely V (φi ) = 0 for i = 1, . . . , N . If φi are constants independent of
space-time, then the corresponding energy vanishes, and in fact E(φ) = 0 if and
only if φ(x, t) = φi .
Static solutions of the equation of motion are determined by,
where x0 the integration constant is any arbitrary point where the field has
the value of φ(x0 ). The integral is non-singular apart from the end-points since
everywhere else V (φ) is positive.
where φsol (x) is a time-independent soliton solution and δ(x, t) is a small pertur-
bation. Substituting this configuration into the equation of motion and retaining
only linear terms in the perturbation we get,
Since the equation is invariant under time translation, we express the perturba-
tion as a superposition of normal modes in the following form,
δφ(x, t) = Re[an eiw n t δn (x)]. (5.30)
n
Often one refers to φ± as the topological indices. In fact for theories with a
potential that has a discrete number (finite or infinite) of vacua, non-singular
field configurations of finite energy have both φ+ and φ− separately conserved.
This results from the following argument. Finite energy implies that both φ+
and φ− are at absolute minima of the potential. Now since the non-singular
configurations are continuous in time, and the potential has a set of discrete
(finite or infinite) vacua, φ(t, ∞) must be stationary at φ+ , or ∂0 φ(t, ±∞) =
∂0 φ± = 0, namely the indices are conserved.
In fact this conservation can be used to show the existence of non-dissipative
solutions. For instance in the φ4 theory we can show that a configuration with
φ+ = −φ− is non dissipative. By continuity in x there must be, for any t, some x
for which φ = 0. At this point T00 ≥ V (0) and since the definition of a dissipative
solution is that the limt→∞ maxx T00 = 0 it is clear that it is non-dissipative.
Similar arguments hold for other cases of solitons.
Thus, one can divide the space of finite-energy non-singular solutions into
topologically disconnected sub-spaces that are characterized by the two indices
φ± . Such a sub-space cannot be continuously deformed into another one unless
the finite energy condition is violated. For instance, in the φ4 theory, the potential
has two minima so that φ+ = mλ and φ− = − mλ . Hence, there are four subspaces
(−, +), (+, −), (−, −), (+, +) associated with the soliton, the anti-soliton and the
two trivial constant vacuum solutions. For the sine-Gordon the solitons belong
to the subspaces characterized by φ− = 2πn √mλ and φ+ = 2π(n + 1) √mλ .
Obviously, non-trivial topological charges require multiple vacua. The latter
situation occurs if and only if there is a spontaneous breaking of a symmetry.
For instance in φ4 and sine-Gordon it is the discrete φ → −φ symmetry which
is broken.
Derrick’s theorem
Consider a scalar field theory in D + 1 space-time dimensions described by the
Lagrangian density,
1
L = ∂ μ φ∂μ φ − V (φ), (5.33)
2
where the potential V (φ) is non-negative and vanishes at its minima. The the-
orem states that for D ≥ 2 the only non-singular time-independent solutions of
finite energy are the vacua.
Let us denote by φ(x) a time-independent solution of the equation of motion.
We now introduce a one-parameter family of field configurations defined as,
φ(λ, x) = φ(λx), (5.34)
where λ is a positive real number. The energy associated with the configuration
φ(λ, x) is,
1
E(λ) = λ−D dD x λ2 (∇φ) + V (φ) .
2
(5.35)
2
By Hamilton’s principle the energy as a function of λ is stationary at λ = 1 so
that,
1 2
dD x (D − 2)(∇φ) + DV (φ) = 0. (5.36)
2
For D > 2 the two terms in the integral have to vanish separately, which occurs
only for the vacua. For D = 2, the potential term has to vanish, which again
occurs only for the vacua. This proves the theorem.
The following remarks are in order:
(i) Derrick’s theorem applies only to time-independent configurations.
(ii) It applies to field theories with only scalar fields. Once one introduces
additional fields like gauge fields or fermions the theorem is not valid (see
Section 20.3).
which looks like a soliton to the left. Similarly, it looks like an anti-soliton to
the right. The soliton and anti-soliton move further apart as sin(wt) increases to
one, and then when sin(wt) decreases they approach each other. As sin(wt) → 0
the approximation that lead to (5.38) is no longer valid, in accordance with the
fact that in this region the soliton and anti-soliton are on top of each other.
A similar discussion applies also for negative sin(wt). It is thus clear that the
solution (5.37) describes an oscillation of a soliton anti-soliton pair around their
common center of mass.
Revealing a bound state solution implies that the system must be attractive
at least in a certain region of the “coupling constant”. Indeed if one uses the
coupling constant,
πβ 2
ξ= , (5.39)
8π − β 2
then
where Msol is the mass of the soliton. This verifies the existence of a binding
energy since the mass of the breather is less than twice the mass of the soliton.
In Section 5.5.1 the quantum description of the bound states will be addressed,
and their scattering processes in Section 5.6.
where Vct (φs ) is a counter term that one has to add to the Lagrangian.
In two dimensions the only source of UV divergences in any order of pertur-
bation theory are diagrams that contain a loop consisting of a single internal
line. Stated differently, UV divergences are due to the fact that the action is not
normal ordered. The corresponding diagrams are depicted in Figure 5.1. In fact
the corrections (a) and (b) cancel and the only corrections follow from (c).
Let us first recall the normal ordering of φ2 (x), namely,
1 dk
φ2 (x) =: φ2 (x) :m + √ (5.45)
4π k 2 + m2
where : :m indicates that the normal ordering is performed for a scalar of mass
m. The last integral in obviously divergent so one introduces a cutoff Λ m
such that
Λ
1 dk 1 4Λ2 m2
√ = ln 2 + O . (5.46)
4π −Λ k 2 + m2 4π m Λ2
For a general potential V (φ), Wick’s theorem tells us that,
2 2
1
ln 4 Λ d
V (φ) =: e 8 π m 2 d φ 2 V (φ) :m . (5.47)
If one expands the exponent one finds a term with no contraction, next, with
one contraction etc. We can pass from normal ordering at mass m to normal
ordering at m̃. This transformation is independent of the cutoff, since
1 m2
: φ2 :m =: φ2 :m̃ + ln 2 , (5.48)
4π m̃
and hence
1 m 2 d2
ln
: V (φ) :m =: e 8 π m̃ 2 d φ 2 V (φ) :m̃ . (5.49)
When applied to the sine-Gordon case, the normal-ordered potential takes the
form,
m2 (m2 − δm2 )
: [cos(βφ) − 1] :m [cos(βφ) − 1] , (5.50)
β2 β2
and where to the lowest order in β,
Λ
m2 β 2 dk
δm2 = − √ . (5.51)
4π k2 + m2
Thus the counterterm potential reads,
δm2 ∞
Vcounter (φ) = − 2 dx[(1 − cos(βφs )] − Evacuum , (5.52)
β −∞
where we further subtracted the energy of the vacuum. Finally the quantum
mass takes the form
1 δm2 ∞ 1 2
Mquantum = Ms + wn − 2 dx[(1 − cos(βφs )] − kn + m2 ,
2 n β −∞ 2 n
(5.53)
where kn = L , with L the size of the quantization length, to be sent to ∞ at
2π
where N is an integer.
Using the relation between the Hamiltonian and Lagrangian densities, H =
π∂0 φ − L, we find after integrating over one period that,
τ
Eτ = 2πN − dt dxL. (5.56)
0
where the expression for w in terms of E and M follows from the calculation of
the energy which can be performed most conveniently at t = 0, as was done in
(5.41).
Integrating this equation and using a natural boundary condition that N = 0
for E = 0 the Bohr–Sommerfeld procedure predicts the following spectrum,
N β2 8π
MN = 2Msol sin N = 1, 2, . . . , < . (5.58)
16 β2
Next we would like to describe the quantization procedure of Dashes, Hass-
lacher and Neveu (DHN). The classical action of the breather solution per period
τ = 2π
w is determined by substituting the breather solution (5.37) into the action
and integrating,
32π w
Scl (φb ) = 2 cos−1 −η . (5.59)
β m
The stability of the breather solution is determined by the requirement that
there are no negative eigenmodes to the stability equation,
[∂ μ ∂μ + m2 cos(βφb )]δn (x, t) = 0, (5.60)
where δn (x, t), which obeys δn (x, t + τ ) = eiν n δn (x, t), is the fluctuation of the
breather solution. The set of all the solutions of this equation was written down
by DHN [74].
The corresponding spectrum of νn reads,
2π 2
ν 0 = 0 ν1 = 0 νn = m + qn2 , (5.61)
w
where Ect and Sct are the energy and action associated with the counterterm.
In the limit of L → ∞ the sum over qn turns into an integral. The integral has
a quadratic as well as logarithmic divergences. These divergences will be can-
celled out by the contribution of the counterterm such that Scl (φb ) + Sct (φb ) −
∞
1/2 0 νn is the same as the Scl (φb ) given above with the renormalization of
πβ 2
the coupling constant β 2 → ξ = 8π−β 2 . Using this result it is easy to determine
the energy,
∞
d 2wη
E=− Scl (φb ) + Sct (φb ) − 1/2 νn (5.64)
dτ 0
ξ
Substituting this into the second equation of (5.63) one finds that the energy
levels take the form,
2m Nξ π
MN = sin N = 1, 2, . . . , < . (5.65)
ξ 2 ξ
Note that in spite of the fact that the quantization condition permits any N ,
only if it is smaller than πξ the classical breather solution exists. Thus the inter-
pretation of this result is that there is a finite number of quantum bound states
corresponding to the classical breather solution. Even though the derivation of
the mass spectrum of the bound states was based on a Wentzel, Krames and
Brillonin (WKB) approximation, the final result turns out to be exact. This
statement follows from the analysis of the physical poles of soliton anti-soliton
scattering, and already indicated in perturbation theory from two loops. The
latter, though, works for mass ratios only, in view of scale dependence for the
normal ordering of each individual mass.
The spectrum (5.65) can be re-written in terms of the mass of the quantum
soliton. Using (5.54) this takes the form,
Nξ π
MN = 2Msol sin N = 1, 2, . . . , < , (5.66)
2 ξ
which indicates that the quantum breather states are indeed bound states of a
quantum soliton anti-soliton pair.
At weak coupling N β 2 1, the mass spectrum reads,
2
1 N β2
MN = N m 1 − 2 6
+ O(N β ) . (5.67)
6 16
Thus at weak coupling the lowest bound state has a mass of,
2
1 β2
M1 = m 1 − + O(β 6 ) , (5.68)
6 16
showing that first bound state is in fact the “elementary” boson of the theory.
2
Moreover the higher bound states have a mass which is N M1 + O[(β 2 ) N (1 −
N 2 )], namely bound states of N elementary bosons. These bound states are
2
loosely bound with a binding energy of m6 ( β16 )2 N (N 2 − 1). Using perturbation
theory one can show that each of these states is stable against decay to states
with lower N . In fact the stability turns out to be an exact statement. The source
of the stability of these states is the set of infinitely conserved charges as will be
discussed in the following section.
where p is the momentum of the particle and (a) denotes its type. For the sine-
(a)
Gordon case the eigenvalues wn (p) are given by,
(a) (a)
w2n + 1 (p) = p2n + 1 , w2n (p) = p2n p2 + m2a . (5.70)
(a)
In general one assumes that the wn (p) form a set of independent functions. A
multiple particle in or out state obeys
(a 1 ) (a k )
k
(a 1 ) (a k )
Qn |p1 . . . pk , in> = wn(a i ) (pi )|p1 . . . pk , in>, (5.71)
i= 1
p2
p1
p3
p4
p4
p3 p1
p2
These two rules, that should apply also to intermediate states where particles
are far enough from each other, together with the special kinematics of two
dimensions, are behind the assertion that the multi-particle S-matrix of theories
equipped with infinitely many conserved charges, can be expressed in terms of
two-particle ones.
The factorized S-matrix corresponds to the following scattering process:
r In the infinite past a set of N particles with momenta p1 > p2 > . . . > pN are
spatially arranged in the opposite order, namely, x1 < x2 < . . . < xN .
r In the interaction region the particles collide in pairs. In each collision the
momenta are conserved and in between collisions the particles move as free
particles.
r The final state of the outgoing particles, achieved after N (N −1) pair collisions,
2
is built from the N particles arranged along the x coordinate in the order of
increasing momenta.
p2
p2
p1
p1
p3 p3
p3 p3
p1
p1
p2
p2
a b
Fig. 5.3. Two possible ways of three-particle scattering.
r Each line corresponds to a given value of the momentum associated with the
slope of the line.
r Each vertex corresponds to a two-particle collision. The two-particle amplitude
Sij (pi , pj ) has to be attached to each vertex.
r The total S-matrix element of the process is the product of all the N (N −1)
2
two-particle amplitudes ij Sij , and then a sum over the different kinds of
particles in the internal lines.
Take for example the case of N = 3. The same scattering can be represented in
two ways, as is shown in Fig. 5.3. These two differ only by a parallel translation
of a line, and thus represent the same process.
p± = m exp(±β). (5.73)
s = (p1 + p2 )2 , (5.74)
is given by
where i, j, k, l = 1, 2 so that the particles are in doublets of O(2), and the ± refers
to bosons (+) and fermions (−). This can be generalized to O(N ) in a straightfor-
ward way. Here we analyze only the case of the doublet. The amplitudes S2 , S3
are the transition and reflection amplitudes, respectively, and S1 corresponds
to the process Ai + Ai → Aj + Aj for (i = j). The Ai (β), non-commutative
2 Please note, that the β we had before, in the term cos(βφ) in the action, is not to be confused
with the present one, which denotes the rapidity.
where,
∞
8π 8β 8π 8β Rn (β)Rn (iπ − β)
U (β) = Γ Γ 1+i Γ 1− −i ,
γ γ γ γ n =1
Rn (0)Rn (iπ)
Γ + 2n 8π
γ i 8β
Γ 1 + 2n 8π
γ +γi 8β
γ )
Rn (β) = . (5.83)
Γ (2n + 1) γ + i γ Γ 1 + (2n − 1) γ + i 8β
8π 8β 8π
γ
where S(β), ST (β) and SR (β) are the scattering amplitude of identical solitons,
transition and reflection amplitude for soliton anti-soliton, which are related to
S1 (β), S2 (β) and S3 (β) as,
X X X X X X
X X X X X X X X X X
0
X X X X X X
X X X
0
Fig. 5.4. The zeros (crosses) and poles (dots) of ST (β) (upper) and SR (β)
(lower). All the singularities have pure imaginary values even though some of
them are displaced from the imaginary axis for clarity [235].
where,
1 8π2
SR (β) = sin U (β). (5.89)
π γ
The zeros and the poles of ST (β) and SR (β) are shown in Fig. 5.4.
The solution (5.88) is in fact the exact solution of the S-matrix of the SG
model. This assertion is supported by the following properties:
r The poles of ST (β) are located at equidistance, and their values are in accor-
dance with the semi-classical mass spectrum if one equates γ = 8ξ.
r Note that for the value of ξ = π where the coupling of the associated Thirring
model vanishes, and the SG model is a bosonized version of a free Dirac fermion,
ST (β) ≡ S(β) = 1 SR (β) = 0. (5.90)
r At ξ ≥ π all bound states, including the “elementary particle” associated with
the field of the SG model, become unstable and the spectrum includes only
soliton and anti-soliton. This situation follows from the fact that at this region
the Thirring coupling is negative and there is a repulsion between the solition
and anti-soliton.
r At ξ = π the reflection amplitude vanishes identically.
n
r Expanding (5.88) in powers of [( 8π ) − 1], which means small coupling of the
γ
massive Thirring model, matches the perturbative expansion of the latter
model.
r The limit β 2 → 0 of the exact result (5.88) is equal to the semi-classical expres-
sion of the two-particle S-matrix.
The explicit expression for the two-particle S-matrix (5.88) enables one to also
write down the S-matrix for any number of solitons and anti-solitons and the
scattering of any number of bound states. This general approach to solving the
S-matrix can be applied to the various integrable models. Here we demonstrate
it on the sine-Gordon model. For soliton and anti-solitons we find the following
S-matrix elements:
1
,1 − 1 ,− 1
S 12 , 12 (β) = S− 12 ,− 12 (β) = S(β)
2 2 2 2
π
1
,− 12 − 12 , 12
sh ξβ
S 2
1
,− 12
(β) = S − 12 , 12
(β) = − S(β)
π
2
sh ξ (β − πi)
π2 i
1
,− 12 − 12 , 12
sh ξ
S 2
− 12 , 12
(β) = S 1 1 (β) = S(β)
2 ,− 2 π
sh ξ (β − πi)
,
S 11 , 22 (β) = 0 1 + 2 = 1 + 2 . (5.91)
Note that just as for the conformal field theories we use here complex
two-dimensional coordinates. The Φi (z, z̄) are primary fields of conformal
dimension Δi + Δ̄i < 2, namely relevant operators. Since these operators are
super-renormalizable, they do not affect the short distance behavior but do affect
the structure of the IR domain. In the analogy with statistical mechanical sys-
tems, where the CFT describes the behavior of the system at its fixed point, the
perturbation with the relevant primary fields describes the scaling region around
the fixed point.
A system described by an action of the form (5.93), is integrable provided that
one can identify a set of infinitely many conserved charges just as for the sys-
tems described previously. An important class of such theories are the integrable
minimal models, for example the tricritical Ising model M4,5 perturbed by Φ 35 , 35
and the tricritical Potts model M6,7 perturbed by Φ 17 , 17 .
The renormalization group (RG) flow of the integrable systems follows a tra-
jectory that starts at a fixed point and may end on another one in the IR, or
on a point that corresponds to a massive QFT. An important property of these
flows is the c-theorem3 which states the following:
Quantum field theories which possess rotational invariance, reflection pos-
itivity, and conservation of the energy momentum tensor admit a function
c(λi ) of the coupling constants λi which is non-increasing along the RG
trajectories and is stationary only at fixed points.
The proof of the theorem is as follows. Consider the correlators of T ≡ Tz z
and Tz z̄ ,
F (z z̄) G(z z̄)
<T (z, z̄)T (0, 0)> = , <T (z, z̄)Tz z̄ (0, 0)> = ,
z4 z 3 z̄
H(z z̄)
<Tz z̄ (z, z̄)Tz z̄ (0, 0)> = 2 2 . (5.94)
z z̄
We now use the conservation law,
¯ + ∂Tz z̄ = 0,
∂T ¯ z z̄ = 0,
∂ T̄ + ∂T (5.95)
to deduce the following differential equations for F, G and H,
Ḟ + (Ġ − 3G) = 0 Ġ − G + Ḣ − 2H = 0, (5.96)
where Ȧ = ddAlog(x)x . Since the positivity condition implies that H ≥ 0, the following
c function is non-increasing,
c = 2F − 4G − 6H, ċ = −12H. (5.97)
At the fixed points Tz z̄ = 0, hence G = H = 0 and c = 2F , as indeed it should
(c )
be. Recall that the OPE in CFT is of the form T (z z̄)T (0, 0) = z24 + . . .
One can further write down an expression for the integral of ċ, namely for
the difference of c in the UV and IR regions. For the case of a perturba-
tion with a single operator Φ, the trace of the energy-momentum tensor is
Tz z̄ = πλ(1 − Δ)Φ, where the total conformal dimension is 2Δ. Using (5.97) it
was shown that,
cUV − cIR = 12πλ (1 − Δ)
2 2
d2 x|x|2 <Φ(x)Φ(0)> . (5.98)
This result has been applied to integrable minimal models yielding the cor-
rect difference in the Virasoro anomalies. Another example where this relation
between the conformal data and the properties of the non-conformal theory can
be tested is the sine-Gordon model. The model can be thought of as a pertur-
bation on a free massless scalar field which has c = 1, and the massive model
has c = 0. Let us check for this case the outcome of the relation (5.97). The
perturbation now is,
m2 β2
λΦ = : (cos βφ − 1) : 2Δ = . (5.99)
β2 4π
which verifies our data about the Virasoro anomalies at the two ends of the
trajectory. One can further show that higher-order terms in β are in accordance
with the fact that cUV − cIR is β independent.
The reason for rewriting the equation of motion in this form is the fact that the
spectrum of L is conserved. To realize this property of the Lax pair, notice first
that the solution of (5.104) can be parameterized as,
Since x+ is the light-cone time direction, this implies the conservation of the
eigenvalues. This conservation of the spectrum is the origin of the infinite set of
conserved charges.
For the case that L can be represented by a finite matrix, it is obvious from
(5.105), using the cyclicity of the trace, that Qn = T r[Ln ] are conserved charges.
In general, and in particular in our case, L does not act only in a space of
finite-dimensional matrices but also in the continuous space whose base vectors
∞
are |x−>. Thus the trace takes the form of the integral −∞ dx− . Due to the
+
unitarity of S(x ) the cyclicity property of the trace is maintained and hence
Qn = T r[Ln ] are indeed conserved charges.
It turns out that one can map the Lax pair of the sine-Gordon system to that
of the Korteweg-deVries (KdV) equation ∂0 u(x, t) = 6u∂1 u − ∂13 u. This map is
useful since it is more convenient to express the conserved charges of the sine-
Gordon system in terms of the L operator of the KdV equation. In this format
the first four charges of the set of infinite charges which are classically conserved
are,
∞
1
Q1 = − (∂− φ̃)2 dx−
4 −∞
2 ∞
1
Q2 = + [(∂− φ̃)4 − 4(∂−2
φ̃)2 ]dx−
4 −∞
3 ∞
1
Q3 = − [(∂− φ̃)6 − 20(∂− φ̃)2 (∂−2
φ̃)2 + 8(∂−
3
φ̃)2 ]dx−
4 −∞
4 ∞
1 112 2 4
Q4 = + [(∂− φ̃)8 − (∂− φ̃) − 56(∂− φ̃)4 (∂− 2
φ̃)2
4 −∞ 5
224 64 4 2
+ (∂− φ̃)2 (∂−
3
φ̃)2 − (∂− φ̃) ]dx− . (5.108)
5 5
∂− ∂+ n + un = 0, n2 = 1. (5.115)
So u is actually the action density u = (∂+ n) · (∂− n). Instead we can write the
equation of motion as,
∂μ ∂ μ n + n(∂μ n · ∂ μ n) = 0. (5.116)
It is easy to check, using the equations of motion, that this current is indeed
conserved. In addition the energy-momentum is conserved,
1 1
∂+ T−− = ∂+ (∂− n)2 = ∂− n∂− ∂+ n = −u∂− (n)2 = 0. (5.118)
2 2
Thus classically the trace of the energy-momentum tensor vanishes T+− = 0.
In addition there is an infinite set of currents, the simplest of which takes the
form,
2
1 ∂− n u
J− = ∂− , J+ = − . (5.119)
2|∂− n| |∂− n| |∂− n|
We can now rewrite the differential equation (5.120) in terms of the integral
equation,
x
w
U (t, x) = 1 + dy j0i − wj1i (t, y)σ i U (t, y). (5.123)
1 − w −∞
2
into the integral equation for U (t, x), we find the following recurrence relation,
x
Un (t, x) = i dy j0i (t, y)σ i Un −2k −1 (t, y)
−∞ ( n −1 )
0≤k ≤
2
− j1i (t, y)σ i Un −2l (t, y) , (5.125)
(n )
1≤l≤ 2
The algebraic structure associated with these charges is the Yangian symmetry,
the description of which is beyond the scope of this book. In the reference list we
mention several that deal with these algebras. In Section 5.12 it will be further
shown that both the charges of the form (5.119) as well as (5.126) are quantum
mechanically conserved.
Jμ = g −1 ∂μ g, ∂ μ Jμ = 0, (5.127)
DJ ≡ dJ + J ∧ J = 0 d ∗ J = 0, (5.128)
or directly,
(1) (1)
= − dxj1 (t, x) + dxj0 (t, x)χ(1) (t, x). (5.134)
We can now re-express χ(1) as χ(1) (t, x) = dx j0 (t, x ). When substituting
(1)
In fact one can show that the constant β̂ is the one-loop beta function.
Consider now the next conservation law. Classically it reads,
1 2 1 2 2 3
∂− u(∂− n) + ∂+ (∂− n) = (∂− n)2 (∂− u). (5.136)
2 2 2
One can show that this classical conservation is directly related to the conserva-
tion of the current in (5.119). For this make use of the classical scale invariance
x− → f (x− ) and the classical conservation ∂+ (∂− n)2 = 0, to choose a gauge
where (∂− n)2 = 1. Now the classical conservation law takes the form,
1 2 2
∂+ (∂− n) = (∂− u). (5.137)
2
This is exactly the conservation of the current (5.119), when inserting the gauge
(∂− n)2 = 1. To get to the form in general coordinates, namely the form as in
The reason that the conserved charge on an asymptotic state has to be pro-
portional to i P− 3i is that its tensorial structure is of the form −−− , the only
conserved quantity with − Lorentz index is P− and there are no higher tensorial
charges that are not products of P− .
In a similar manner one has a similar conservation law for P+ so that,
P−3 = P−3 P+3 = P+3 . (5.143)
in out in out
charges. Here we will establish the algebraic structure of these charges. It will
be shown that they involve non-trivial braiding and that they obey the algebra
associated with “quantum groups”.
Rather than discussing the generalities of this algebraic structure we analyze it
in the context of our laboratory model, the sine-Gordon model. We now rewrite
the Lagrangian density of the model as,
1 2 ¯ λ̂
S= d z∂φ∂φ + d2 z : cos(β̂φ) : . (5.144)
4π π
It is straightforward to relate λ̂ of this formulation to m, β of (5.23), and β̂ here
equals β there.
Recall (Section 5.9) that this action can be considered as a conformal field
theory plus a relevant perturbation of the form (5.93). In such a case one can
identify a conserved current that obeys the relation,
¯
∂J(z, ¯ z̄) = ∂¯H̄(z, z̄),
z̄) = ∂H(z, z̄) ∂ J(z, (5.145)
¯ = ∂ J¯ = 0, and H, H̄ are defined via,
where for the conformal limit one has ∂J
Resz =w (φp ert (w)J a (z)) = ∂ha (z) H a (z, z̄) = 2λ̂ha (z)φ̄p ert (z̄), (5.146)
where the perturbation term is written, in the conformal limit, as
φp ert (z)φ̄p ert (z̄), and all this under the condition that the Res above are indeed
total derivatives. A similar construction applies also for the anti-holomorphic
current.
We can now identify a pair of non-local fields
x
1
φ̃± (t, x) = φ(t, x) ± dy∂0 φ(t, y) , (5.147)
2 −∞
with which we can write a pair of conserved currents J± of the form (5.145) as,
± 2 i φ̃ + (t,x)
J± = e β̂
β̂ 2 ±i 2
−β̂ φ̃ + (t,x)∓ i φ̃ − (t,x)
H± = λ e β̂ β̂ . (5.148)
β̂ 2 − 2
The conserved charges associated with the pair of currents are
1 1
Q± = dzJ± + dz̄H± Q̄± = dz̄ J¯± + dz H̄± . (5.149)
2πi 2πi
The charges Q± are non-local, as a consequence of being built from the non-local
field φ̃.
Using the basic canonical commutation relation [φ(t, x), ∂0 φ(t, y)] = 4πiδ(x −
y) and eA eB = e[A ,B ] eB eA (when [A,B] commutes with A and B) one finds the
following braiding relations,
1
J± (t, x)J¯∓ (t, y) = 2 J¯∓ (t, y)J± (t, x)
q
J± (t, x)J± (t, y) = q 2 J¯± (t, y)J± (t, x),
¯ (5.150)
where,
− 2 πi
q=e 2
β̂ . (5.151)
These non-trivial braiding relations of the currents imply similar relations for
the conserved charges,
Q+ Q̄+ − q 2 Q̄+ Q+ = 0
Q− Q̄− − q 2 Q̄− Q− =0
1
Q+ Q̄− − 2 Q̄− Q+ = a(1 − q 2Q t o p )
q
1
Q− Q̄+ − 2 Q̄+ Q− = a(1 − q −2Q t o p )
q
[Qtop , Q± ] = ±2Q± [Qtop , Q̄± ] = ±2Q̄± , (5.152)
where a = λ 2
2πi γ , γ −1 = Δ = −Δ̄(Q± ) and the topological charge Qtop =
β̂
= ∞) − φ(x = −∞)) (compare with (5.6)).
2π (φ(x
This algebra of the charges is referred to as “q-deformation” ŜLq (2) of the
SL(2) affine Lie algebra with zero center. Recall the basic SL(2) algebra in the
Chevaley basis (3.5),
[H, E± ] = ±2E± [E+ , E− ] = H. (5.153)
Introducing the spectral parameter w the infinitely many generators of the SL(2)
affine Lie algebra are defined via J a = n Jna wn with J a = H, E± . We then
ˆ as,
define the Chevaley basis of the affine algebra SL(2)
E+ 1 = wE+ E−1 = w−1 E−
E+ 0 = wE− E−0 = w−1 E+
H1 = H H0 = −H. (5.154)
In terms of these generators the ŜLq (2) algebra reads,
[Hi , E+j ] = aij E+j
[Hi , E−j ] = −aij E−j
q H i − q −H i
[E+i , E−j ] = δij E−j , (5.155)
q − q −1
and aij is the Cartan matrix of SL(2).
The relations between the non-local charges Q± and Q̄± and the generators
of the ŜLq (2) algebra are,
H1 H0
Q+ = cE+ 1 q 2 Q− = cE+0 q 2
H1 H0
Q̄− = cE−1 q 2 Q̄+ = cE−0 q 2
5.14 Integrable spin chain models and the algebraic Bethe ansatz
The discussion of the algebraic Bethe ansatz follows closely the pedagogical paper
of Faddeev [88] and also Beisert [31]. The use of the ansatz in a continuous system
that we present follows that of Zamolodchikov [234].
A very useful class of two-dimensional integrable models are the spin chain
models. In these models the space is divided into a discrete number of sites
where spin variables are placed. So far, and in fact also in the rest of this book,
we do not discuss discretized field theories. In this chapter we do since the spin
chain models will be shown, in Section 18, to be intimately related to integrable
sectors of gauge theories in four dimensions. We will demonstrate the techniques
used to solve the spin chain models by applying them to a prototype model,
the XXX1/2 model. We will describe the model, write down the Bethe ansatz
equations associated with it, solve them and extract the spectrum of the model.
We then apply the technique to the discretized sine-Gordon model.
A discrete circle with N ordered points is taken to be the space direction.
The “space” is periodic so that each site is identified with i ≡ i + N . The formal
continuum limit can be taken by introducing a lattice spacing Δ such that Δ →
0, N → ∞ while x = N Δ is kept finite.
At each site there is a dynamical variable Xiα where i denotes the site and α
is a set of finite number of values. One defines a quantum algebra of observables
A by fixing a set of commutation relations between the Xiα . When [Xiα , Xjβ ] = 0
for any i = j the algebra is called ultra local. Examples are canonical variables,
and spin variables which will be used in the XXX1/2 model we are about to
describe.
The Hilbert space of the representations of the ultra local algebra has a natural
tensor product,
N 6 6 6 6
H= hi = h1 h2 . . . hi . . . hN , (5.157)
i= 1
and is conserved,
[H, S α ] = 0. (5.160)
The notion XXX associates with the fact that the coefficient of (Siα Si+1 α
− 14 )
is a constant independent on i and α. In the case where the coefficient is α
1 − 4 ), the model is referred as the XYZ model.
1
dependent, namely J α (Siα Si+α
To extract the spectrum of the model we make use of the Lax operator defined
by,
6 6
Lk ,a (λ) = λIk Ia + i Skα σα , (5.161)
α
Pa 1 ,a 2 = Pa 2 ,a 1
Pn ,a 1 Pn ,a 2 = Pa 1 ,a 2 Pn ,a 1 = Pn ,a 2 Pa 2 ,a 1 . (5.165)
The idea now is to relate the Hamiltonian and other conserved charges to the
monodromy of a string of Lax operators along the whole chain. For that we
have to analyze the commuting structure of the Lax operators. This structure is
controlled by the fundamental commutation relations (FCR) which will be shown
later to be part of the Yang–Baxter relations and read,
where,
To prove this relation one makes use of the relations of the permutation operator
(5.165). The Lax operator can be interpreted as a connection along the chain in
the sense,
ψi+ 1 = Li ψi . (5.169)
From this relation it follows that the trace of the monodromy operator,
is commuting, namely [F (λ), F (μ)] = 0. We can now expand both TN and F (λ)
as a polynomial of order N in λ as,
6
Ta,N (λ) = λN + iλN −1 Sα σα + . . .
α
N −2
F (λ) = 2λN + Ql λl . (5.175)
l= 0
We will see shortly that the set of N − 1 operators Ql are commuting and con-
stitute the set of conserved charges, including the Hamiltonian. Next we expand
the monodromy at λ = 2i ,
i
Ta,N = iN PN ,a PN −1,a . . . P1,a = iN P1,2 P2,3 . . . PN −1,N PN ,a . (5.176)
2
whereˆmeans that the corresponding factor is absent. Using the same procedure
as above we find that,
d
Fa (λ)|λ=i/2 = iN −1 P1,2 P2,3 . . . PN −1,N . (5.180)
dλ i
Recalling the expression we found earlier for the Hamiltonian (5.166) we can now
see that,
i d N
H= ln(Fa (λ))|λ=i/2 − . (5.182)
2 dλ 2
We have just shown that the Hamiltonian is part of a set of N − 1 commuting
operators generated by F (λ), the trace of the monodromy. In fact there are N
such conserved charges if we add also one component, say S 3 , of the spin. The
model is characterized by its N degrees of freedom and is equipped with N
conserved charges and hence is (at least classically) integrable.
of the harmonic oscillator. To derive the above one uses an explicit 4 × 4 matrix
7
formulation for the operators in V V . A natural basis for these matrices is
6 6 6 6
e1 = e+ e+ , e2 = e+ e− , e3 = e− e+ , e4 = e− e− , (5.185)
where,
1 0
e+ = , e− = . (5.186)
0 1
In this basis the permutation operator and the R matrix read,
⎛ ⎞
1 0 0 0
⎜0 1 0 0⎟
P =⎜ ⎝0
⎟ (5.187)
0 1 0⎠
0 0 0 1
⎛ ⎞
λ+i 0 0 0
⎜ 0 λ i 0 ⎟
R(λ) = ⎜⎝ 0
⎟. (5.188)
i λ 0 ⎠
0 0 0 λ+i
The matrices Ta 1 (λ) and Ta 2 (μ) read,
⎛ ⎞
A(λ) a(λ)
⎜ A(λ) a(λ) ⎟
Ta 1 (λ) = ⎜
⎝ a† (λ)
⎟
⎠ (5.189)
D(λ)
a† (λ) D(λ)
⎛ ⎞
A(μ) a(μ)
⎜ a† (μ) D(μ) ⎟
Ta 2 (λ) = ⎜
⎝
⎟. (5.190)
A(μ) a(μ) ⎠
a† (μ) D(μ)
Explicit multiplication of these matrices yields (5.183).
Similarly to the case of the harmonic oscillator we now define the ground state,
6
a(λ)|0> = a(λ) |0>i = 0. (5.191)
i
and hence,
i N
λ+ ∗
T (λ)|0> = 2
N |0>, (5.194)
0 λ − 2i
which means that |0> is an eigenstate of both A(λ) and D(λ) and thus also of
F (λ). Higher excited states are created from the ground state |0> by a successive
action with creation operators,
Requiring that the state Φ({λ}) is an eigenstate of F (λ) imposes a set of relations
on the λ1 , . . . , λl . In particular using the FCR relation (5.183) we find,
The first term on the right-hand side of the equation has the form of an eigenstate
equations but the rest of the terms do not. The idea is to choose the set {λ} such
that these terms will cancel out against similar terms in D(λ)a† (λ1 ) . . . a† (λl )|0>.
To get the value of the coefficient M1 (λ, {λ}) we use the second term on the right-
hand side of the second equation in (5.183) when interchanging A(λ) and a† (λ1 )
and in all other exchanges we use the first term. In this way we find that,
l N
i λ1 − λk − i i
M1 (λ, {λ}) = λ1 + . (5.197)
λ − λ1 λ1 − λk 2
k=2
Interchanging now λ1 → λj we get similarly the expressions for all Mj (λ, {λ}).
The same type of manipulations yield,
with,
l N
i λ1 − λk − i i
Nj (λ, {λ}) = λ1 − . (5.199)
λ − λ1 λ1 − λk 2
k =2
then the undesirable terms in (5.196) cancel out and we end with an eigenstate
of F (λ) and hence of the Hamiltonian. For the former it takes the form,
N l N l
i λ − λj − i i λ − λj − i
F (λ)Φ({λ}) = λ+ + λ− Φ({λ}).
2 λ − λj 2 λ − λj
j=1 k = j
(5.201)
These conditions can be rewritten as,
N l
λj + i/2 λj − λk + i
= . (5.202)
λj − i/2 λj − λk − i
k = j
[S 3 , a† ] = −a† [S + , a† ] = A − D. (5.204)
S + Φ({λ}) = 0. (5.206)
The last expressions calls for a “quasi particle” interpretation. The operator
a† (λ) creates a quasi particle which reduces the spin S 3 by one unit and adds to
the momentum and energy p(λ) and e(λ), respectively. We further observe that,
1 d
e(λ) = p(λ). (5.210)
2 dλ
It is also evident that we can eliminate the dependence on the rapidity of the
energy and momentum and read directly the dispersion relation,
Since this energy is always non-positive, the highest weight state |0> can be
considered a ground state only if we take −H as the Hamiltonian rather than
H. In fact it will be shown shortly that the latter corresponds to a system of an
antiferromagnet whereas the former corresponds to that of a ferromagnet.
where Qi are integers 0 ≤ Qj ≤ N − 1 that define the branch of the log and ϕ(λ)
λ+i
is a fixed branch of log( λ−i ). For large N and Qj and fixed l one finds the usual
expression for the momentum of a free particle on the chain,
Qj
pj = 2π , (5.213)
N
since the ϕ(λ) is negligible.
The second term in (5.212) associates with the scattering of these particles.
In fact by comparison with the quantum mechanics of a particle in a box we
see that ϕ(λi − λj ) stands for the corresponding phase shift of the particles with
rapidity λi and λj . Using this analogy we can now identify the S-matrix element
with,
λ−μ+i
S(λ − μ) = . (5.214)
λ−μ−i
The S-matrix also enters the large N commutation relations of the normalized
creation operators ㆠ(λ) = a† (λ)A−1 (λ),
In addition to the quasi-particle states in the Hilbert space, there are also
bound states of the quasi-particles. These states correspond to complex solutions
of the BAE. The simplest case is with two quasi-particles l = 2. In this case the
two BAE read,
N N
λ1 + i/2 λ1 − λ2 + i λ2 + i/2 λ2 − λ1 + i
= . (5.216)
λ1 − i/2 λ1 − λ2 − i λ2 − i/2 λ2 − λ1 − i
λM ,m = λM + im − M ≤ m ≤ M, (5.220)
where λM is real and m are integers and half integers. The corresponding momen-
tum and energy are given by,
1 λ + i(M + 1/2) 1 2M + 1
pM (λ) = ln eM (λ) = . (5.221)
i λ − i(M + 1/2) 2 λ2 + (M + 1/2)2
where,
λ + iM λ + i(M + 1)
S0,M (λ) = . (5.223)
λ − iM λ − i(M + 1)
and the higher conserved charges that render the model integrable are,
i
l
1 1
Qr = − , (5.226)
r−1 (λk + 2i s)r −1 (λk − 2i s)r −1
k=1
In the thermodynamic limit, N → ∞ states with low energy and zero momen-
tum can be dealt with by introducing the scaling,
1 s 1 s
l l l
2
s 1
Ẽ = N E = 2πn = 2πnk − = ,
N λ̃2
k =1 k
N λ̃
k=1 k j =1,j = k
λ̃ −
λ̃kλ̃kN
(5.227)
where λk = N λ̃k and where the second and third expressions were derived by
taking the log of the zero momentum condition U = 1 and of the BAE, respec-
tively. Using the same scaling one finds for the higher charges and the transfer
matrix the results,
1 s
l l
Qr s
1
Q̃r = = − i log T̃ (λ̃) = −i log T (N λ̃) = .
N r −1 N λ̃r λ̃ − λ̃k
N
k=1 k j =1,j = k j
(5.228)
For N → ∞ it is plausible to assume that the Bethe roots accumulate on
smooth contours (C1 , . . . CA ) ≡ C which are referred to as “Bethe strings”. Thus
we replace the discrete λ̃k locations of the roots by a continuum variable λ̃
described by a density ρ(λ̃) so that the sum of the root translates into the
integral,
1
l
→ dλ̃ρ(λ̃), (5.229)
N C k=1
where
It can be shown that Tf (λ, w|a, i, μ) is subjected to the FCR relations in a similar
manner to T (λ) (13.77). Due to the commutativity of F we get a zero curvature
condition on the transport around an elementary plaquette of the space-time
lattice,
−1 −1
L2i,a (λ + w)L2i−1,a (λ − w) = U− L2i−1,a (λ − w)U− U+ L2i,a (λ + w)U+ .
(5.237)
The light-like shifts U± are related to the shift in space and time in the usual
form, namely,
As for the case of only space dimension discretized, here too one finds that the
condition of having an eigenvector of the energy and momentum is the BAE
which takes the form,
N
α(λj + w)α(λj − w)
= S(λj − λk ), (5.239)
δ(λj + w)δ(λj − w)
j = k
where α(λ), δ(λ) are local eigenvalues and S(λ) is the quasi particle phase factor.
To reduce the number of degrees of freedom per site from two to one we impose
the constraint,
−1
u2i u2i−1 v2i v2i−1 = 1. (5.242)
zN
zW zE
zS
Fig. 5.6. Elementary plaquette.
With this operator at hand we can now determine the equation of motion of the
model and show that it corresponds in the continuuum limit to the sine-Gordon
equation. To accomplish this we define now zi such that,
zi+1
wi = . (5.248)
zi−1
It is easy to see that zi does not commute only with one w, namely wi for which
we have,
zi wi = q 2 wi zi . (5.249)
(q −1 zN zW − zS zE ) = κ2 (q −1 zS zW − zN zE ). (5.252)
In the quantum version one modifies the scaling to take into account the mass
renormalization.
C
L
R B
y
Alternatively, one can consider the space of states B along the contour b with
the time direction along −x, with the Hamiltonian,
1
Hb = Txx dy, (5.260)
2π b
and with the momentum,
1
Pb = − Ty x dy, (5.261)
2π b
where now the quantization condition is that the eigenvalues of Pb are quantized
in units of 2πn
L .
Let us consider the cylindrical geometry via the limit L → ∞ (L R). For
this case the partition function Z(L, R) is dominated by the ground state of Ha
with the ground state energy E0 (R),
Z(R, L) ∼ e−L E 0 (R ) . (5.262)
On the other hand,
Z(R, L) = T rB [e−R H b ]. (5.263)
In B, the thermodynamic limit, namely infinite space L → ∞ which is the
analog of the large N limit in Section 5.14.3, gives the free energy f (R) at
temperature 1/R, via log Z(R, L) ∼ −Lβf (β), where β = R is the inverse of the
temperature. Hence,
E0 (R) = Rf (R). (5.264)
The ground state energy E0 (R), which can be referred to also as the Casimir
energy, can be related in the limit of conformal field theory, to the Virasoro
anomaly. Define the scaling factor c̃(r) via,
πc̃(r)
E0 (R) = −
, (5.265)
6R
where the dimensionless quantity r = m1 R, with m1 the lowest mass in the
theory. The scaling factor will be determined by the TBA. On the other hand,
in the case of conformal invariance, when R → 0, using the relation between the
Hamiltonian and L0 and L̄0 we have,
2π 1
E0 (R) = Δm in + Δ̄m in − c , (5.266)
R 12
where Δm in denotes the conformal dimension of the lowest state. This means
that the scaling factor should reduce to the effective Virasoro anomaly,
and that the amplitudes obey the unitarity and crossing symmetry (5.78),
For regions of configuration space where the particles are highly separated,
namely where |xi − xj | Rc , i, j = 1, . . . , N with Rc denoting the scale of the
interaction, the particles can be treated as approximately free. In these regions
it makes sense to introduce the wavefunction of the system Φ(x1 , . . . , xN ) (in
regions where the particles are not well separated, particle creation and annihi-
lation prevent the use of a single wavefunction).
For integrable systems at any free region the number of particles will be the
same, namely N , as well as the set of momenta pi . The set of particles in a
free region will be denoted by (i1 , i2 , . . . , iN ) where xi 1 < xi 2 < . . . < xi N . A
scattering process that yields a transition between (i1 , i2 , . . . , ik , ik +1 , . . . , iN )
and (i1 , i2 , . . . , ik +1 , ik , . . . , iN ) affects the wavefunctions, such that the latter
wavefunction is given by the former multiplied by the scattering amplitude
S(βk − βk + 1 ). These matching conditions on the wavefunctions combined with
the quantization of the momenta due to the fact that the “space” coordinate is
compact, lead to the relation,
eip i L S(βi − βj ) = 1, Or mL sinhβi + δ(βi − βj ) = 2πni , (5.270)
j = i j = i
N
N
Hb = m coshβi , Pb = m sinhβi . (5.271)
i= 1 i=1
If one further defines the level density ρ(β) such that n = ρ(β)dβ we get,
mL coshβ + ϕ(β − β )ρ1 (β )dβ = 2πρ(β), (5.273)
At this point one has to distinguish between bosons and fermions. From the
unitarity condition one can have either S(0) = 1 or S(0) = −1. In the former case
bosons occupy each rapidity value in any number whereas for fermions the occu-
pation number is at most one. In the latter case the situation is in some sense
the opposite. S(0) = −1 implies that for two particles with the same rapidity
the wavefunction is antisymmetric in their coordinates, and since this is incom-
patible with bose statistics, this state should be excluded. Hence the bosonic
particles for S(0) = −1 behave like fermions, and indeed we will refer to this
case as “fermionic”. On the other hand identical particles that are fermion states
of identical rapidity are allowed, and will be referred to as part of a “bosonic”
system.
Now we would like to address the issue of the entropy for both the bosonic
and fermionic cases. For small intervals of the rapidity Δβα 1 but such that
m L Δβα there is a large number of levels Nα ∼ ρ(βα )Δβα and large number of
1
particles nα ∼ ρ1 (βα )Δβα that are distributed among these levels. The number
of different distributions for the two cases are,
(Nα )! (Nα + nα − 1)!
“fermionic” = , “bosonic” = . (5.275)
(nα )!(Nα − nα )! (nα )!(Nα − 1)!
The entropy S(ρ, ρ1 ) = log N (ρ, ρ1 ), where N (ρ, ρ1 ) is the total number of states,
is given by,
SFerm i = dβ[ρ log ρ − ρ1 log ρ1 − (ρ − ρ1 ) log(ρ − ρ1 )],
SBose = dβ[−ρ log ρ − ρ1 log ρ1 + (ρ + ρ1 ) log(ρ + ρ1 )]. (5.276)
Computing the partition function by performing the trace over all the states
of the system translates to the minimization of the free energy,
with respect to the densities ρ and ρ1 , subjected to the constraints (5.273). Using
(5.273) the extremum equations take the form,
dβ
−Rm coshβ + (β) ± ϕ(β − β ) log(1 ± e−(β ) ) = 0, (5.278)
2π
where the upper sign is for the fermionic case, the lower for the bosonic, and the
“pseudoenergies” (β) are defined via,
ρ1 e−(β )
= . (5.279)
ρ 1 ± e−(β )
The extremal free energy is,
dβ
Rf (R) = ∓m (coshβ) log(1 ± e−(β ) ) . (5.280)
2π
Comparing (5.273) with (5.280) determines a useful relation,
L ∂(β)
ρ(β) = . (5.281)
2π ∂R
Finally we can also write down the TBA expressions for the expectation values
of Tμ,ν . Using the last relation and (5.279) we get,
dE(R) 2π
< Txx > = 2π = m ρ1 (β) coshβdβ,
dR L
2π d[RE(R)]
< Tμμ > =
R dR
2π 1 ρ1 ∂(β)
=m [ρ1 (β) coshβ − sinhβ]dβ. (5.282)
L R ρ ∂β
These relations generalize in a straightforward manner to the more general case
of N types of particles with masses ma , a = 1, . . . , N̂ and scattering amplitudes
Sab which are now N̂ × N̂ matrices.
For integrable models where one can take the limit of rRm1 → 0, with m1
being the lowest mass of the system, one can determine the scaling function c̃(r).
In this limit (β) become constant in the regime where − ln(2/r) β ln(2/r).
Their constant value is determined by the limit of equation (5.273) which now
where c̃a is the scaling factor associated with the particle of type a and L(x) is
the dilogarithm function,
1 x ln t ln(1 − t)
L(x) = − + dt. (5.285)
2 0 1−t t
Thus the determination of the Virasoro anomaly of the underlying CFT of our
massive integrable theory with a purely elastic S-matrix follows from the solution
for the pseudo-energies (5.283) and plugging it into the expression of the scaling
factor. One can proceed with the continuous TBA and determine not only the
ground state energy but also the full spectrum of energies as well as the spectrum
of the eigenvalues of all the conserved charges, as was described for the discretized
case discussed in Section 5.14. This is beyond the scope of this book. We refer
the reader to the relevant papers in the reference list.
In one space dimension there are obviously no rotations and hence no angular
momentum. This raises the possibility of equivalence relations between scalar
fields and fields of higher tensorial structure like spinors, vectors etc. However,
spinors and scalars seem to be distinct even in two dimensions due to their
different statistics. An equivalence between these two types of fields should
therefore incorporate the identification of operators, made out of scalars, that
are anti-commuting and vice versa. It is well known that a bilinear of fermi fields
is a commuting field, but it is less obvious how to construct a field that obeys
the Fermi–Dirac statistics from scalars. This is precisely what the bosonization
procedure does.
Coleman [63] and Mandelstam [159] introduced the concept of bosonization.
Their construction is now referred to as the “abelian bosonization”. An anti-
commuting Fermi field, constructed from the exponential of a boson, was given
explicitly by Mandelstam [159].
The fact that the theories of a free massless scalar and a free Dirac fermion are
equivalent can be proven by showing that they fall into the same representation
of the affine current algebra and the Virasoto algebra. The bosonic–fermionic
duality can also be further elevated to the free massive theories and also to
interacting ones.
It turns out that the original abelian bosonization is not convenient to accom-
modate color (or flavor) degrees of freedom and hence is inconvenient to address
systems like QCD2 . A breakthrough in that direction was achieved by Witten,
in his non-abelian bosonization [224].1
The equivalence enables one to use, as convenient, either the fermionization
of scalar fields or the bosonization of fermions. The latter is useful in several
cases. For instance in the case of duality between the Thirring model [205] and
the sine-Gordon model,2 which will be discussed in Section 6.2, the bosonization
takes the form of a strong–weak duality. For strong fermionic interactions one
finds a weak bosonic coupling. In applications to gauge theories (Section 9) it will
be shown that the one loop anomaly behavior is encoded in classical bosonized
theory. In QCD2 , as will be discussed in Section 9.3.2, the bosonic version of
1 This paper discusses Majorana fermions. The construction for Dirac fermions was done in [7].
2 This was proved by Coleman [63].
the theory admits a separation between the color and flavor degrees of freedom,
which is very useful in describing the low energy color singlet states.
We start this chapter by introducing the set of rules that span abelian
bosonization, including the rules for mass terms and the equivalence of the inter-
acting Thirring model and the sine-Gordon model. We then describe Witten’s
non-abelian bosonization of Majorana fermions. This is further generalized to the
case of massless Dirac fermions. We discuss the subtleties of the massive case and
present two methods of handling the non-abelian bosonization of massive Dirac
fields.3 We then discuss in detail the action formulation of the chiral bosonization
both abelian and non-abelian. We then depart from the applications that will be
found to be relevant to QCD2 and present topics in bosonization which are more
relevant to conformal field and string theories like the bosonization of ghost fields
and the Wakimoto bosonization [213]. We do not discuss bosonization on higher
Riemann surfaces. The interested reader can consult for instance [211] and [84].
The topic of bosonization in two-dimensional field theories has been reviewed
in several papers and books, like that of Stone [202]. Here we mainly follow the
review of Frishman and Sonnenschein [101] for the basic ingredients, and update
it to include more recent topics.
3 A bosonization prescription for the mass term in the flavored case was suggested in [75] and
[99].
which admit a Virasoro algebra with c = 1 and affine Lie algebra with level
k = 1.
Recall also that the theory of a free massless Dirac field with the action,
1
S= d2 z(ψ † ∂ψ
¯ + ψ̃ † ∂ ψ̃) (6.5)
4π
admits conserved currents,
J(z) = ψ † ψ ¯ = ψ̃ † ψ̃,
J(z̄) (6.6)
and its energy-momentum tensor can be expressed as a bilinear of the currents
using the Sugawara construction,
1 1 1
T (z) = − [ψ † ∂ψ − ∂ψ † ψ] = − : ψ † ψψ † ψ := − : J(z)J(z) : . (6.7)
2 2 2
The correponding level of the affine algebra and of the Virasoro anomaly are
again k = 1, and c = 1, respectively.
Due to the uniqueness of the irreducible unitary k = 1 representation of the
affine Lie algebras, and the fact that the infinite-dimensional algebraic structure
fully determines the theories, we conclude that in two space-time dimensions the
theories of massless free scalar field and Dirac field are equivalent.
The equivalence implies that every operator of one theory should have a part-
ner in the other theory, in such a way that the OPEs of these dual operators
should be identical. We have just realized such correspondence for the currents
and energy-momentum tensor, namely,
Jb (z) = ∂φ(z) ↔ Jf (z) =: ψ † ψ(z) : ,
1 1
Tb (z) = − : ∂φ∂φ : ↔ Tf (z) = − [ψ † ∂ψ − ∂ψ † ψ], (6.8)
2 2
and similarly for the anti-holomorphic counterparts.
For completeness we now redescribe the currents using the “old” terminology
of vector and axial currents. The vector current reads,
1
JVμ =: ψ̄γ μ ψ := − √ μν ∂ν φ. (6.9)
π
This identification of J μ leads automatically to a conserved current,
∂μ JVμ = 0, (6.10)
independent of the equations for φ. This is a “topological” conservation, con-
nected with choosing the “vector conservation” scheme. In the applications to
follow, we will demand more freedom in the scheme choice of interacting theories,
in particular the possibility to have a vector current anomaly. The bosonization
procedure will therefore be somewhat modified. The modification will correspond
to a change of regularization scheme.
The overall coefficient of the current is such that the fermion number charge,
∞
Q= dxj0 (x) = 1, (6.11)
−∞
for the ψ-field. In addition to the “topologically” conserved vector current, the
bosonic theory has an axial current, which is equivalent to the fermionic axial
current,
1
JAμ =: ψ̄γ μ γ 5 ψ := √ ∂ μ φ. (6.12)
π
The bosonic current is the Neother current associated with the invariance of
the bosonic action under the global shift δφ = . The holomorphic and anti-
holomorphic conserved currents discussed above are (in real coordinates) nothing
μ
but the left and right chiral currents J± = Jvμ ± JAμ , which correspond to shifts
with (x+ ) and (x− ). Using the commutation relation (8.4) the ALA reads,
2i
J± (x± ), J± (x± ) = δ (x± − x± ). (6.13)
π
This is the same algebra as that of the fermionic chiral currents. The Sugawara
construction in this terminology reads,
T±± = π : J± J± : . (6.14)
α2
where . . . stands for non-singular terms, Hence the conformal dimension is 2 .
Thus we conclude that the following equivalence should hold,
ψ(z) ↔ eiφ(z ) , ψ † (z) ↔ e−iφ(z ) . (6.17)
To confirm this bosonization rule we compute the OPEs in both descriptions and
verify that they are indeed identical,
1
ψ(z)ψ † (−z) = + :ψ(0)ψ † (0): +2z : [ψ(0)∂ψ † (0): −∂ψ(0)ψ † (0))]+O(z 2 )
2z
1
ψ(z)ψ † (−z) = + J(0) + 2zT (0) + O(z 2 )
2z
1
: eiφ(z ) :: e−iφ(−z ) : = + J(0) + 2zT (0) + O(z 2 )
2z
1
: eiφ(z ) :: e−iφ(−z ) : = + i∂φ(0) + 2z(∂φ∂φ(0)) + O(z 2 ). (6.18)
2z
The bosonic version of the fermion ψ was originally proposed by Mandelstam.
His formulation was done in terms of real coordinates and cannonical quanti-
zation. For completeness we now also present the “old” construction and proof
of equivalence. The bosonized chiral fermion in the latter formulation takes the
form,4
5 x
cμ √
ψL = : exp − i π dξ[π(ξ) + φ(x)] :
2π −∞
5 x
cμ √
ψR = : exp − i π dξ[π(ξ) − φ(x)] :, (6.19)
2π −∞
the left and right chiral componenets of the Dirac fermion takes the following
well-knows form,
mf [ψ̃ † (z̄)ψ(z) + ψ † (z)ψ̃(z̄)] = mf (ψL† ψR + ψR† ψL ). (6.21)
Again for completeness we write down the expression both in the complex coor-
dinates as well as in real coordinates.
Using the bosonization rules for chiral fermions (6.17) (to be justified below),
we deduce the map of the fermion bilinear to the equivalent bosonic operator,
mf [: ei φ̄( z̄ ) :: eiφ(z ) : + : e−iφ(z ) :: e−i φ̄( z̄ ) :] = mf μ : cos(φ̂(z, z̄) :, (6.22)
where we have made use of φ̂(z, z̄) = φ(z) + φ̄(z̄) and of the fact that there is no
non-trivial OPE between φ(z) and φ̄(z̄). Note that we write down the bosonic
equivalent of the mass term operator in the context of the massless theory and
hence the factorization to holomorphic and anti-holomorphic parts of the scalar
field holds. Once we identify this operation relation we will then use it to add a
fermion mass term to the bosonized action. The additional parameter which has
a dimension of mass μ is the normal ordering scale.
The derivation of the bosonized mass term in the “old language” is somewhat
more involved. We will return to this after we address the bosonization duality
between the fermionic Thirring model and the bosonic sine-Gordon model.
We now summarize the equivalence relations between the bosonic and
fermionic operators of the free theories, in both the “modern” complex coor-
dinate formulation, as well as the “old” formulation in terms of real coordinates:
J+ (x+ ) : ψL† ψL : ∂+ φ
−
J− (x ) : ψR† ψR : ∂− φ
T+ + (x+ ) − 12 : [ψL† ∂ψL − ∂ψL† ψL ] : − 12 : ∂+ φ∂+ φ(x+ ) :
T−− (x− ) − 12 : [ψR† ∂ψR − ∂ψR† ψR ] : − 12 : ∂− φ∂− φ(x+ ) :
cμ √ x
fermionL ψL (x )+
2π
: exp −i π dξπ(ξ) + φ(x) :
−∞
−
cμ √ x
fermionR ψR (x ) 2π
: exp −i π dξπ(ξ) − φ(x) :
−∞
where Jμ = : ψ̄γμ ψ : . The model is exactly solvable and meaningful for g > −π.
The corresponding equation of motion reads,
i ∂ψ(x) = gγμ J μ (x)ψ(x). (6.23)
The theory is invariant under vector and axial U (1) global transformations.
The corresponding conserved currents are,
ν
JμV = J μ = : ψ̄γμ ψ : JμA = μν J V . (6.24)
The model can be studied by means of the operator product expansion on the
light-cone [76]. The fermionic bilinears of the model are expressed as a function
of the current, and the expressions obtained turn out to be very natural in the
light of the bosonization procedure, which we now describe.
We start with the following generalization of the bosonization formula (6.19):
⎛ ⎛ ⎞⎞
5 √ x
cμ √ 2 π β
ψL = : exp ⎝−i π ⎝ dξπ(ξ) + √ φ(x)⎠⎠ :
2π β 2 π
−∞
⎛ ⎛ ⎞⎞
5 √ x
cμ √ 2 π β
ψR = : exp ⎝−i π ⎝ dξπ(ξ) − √ φ(x)⎠⎠ :. (6.25)
2π β 2 π
−∞
The meaning of the new parameter β will be clarified shortly. In a similar man-
ner to the derivation of (6.20) we can verify that the equal-time anti-commutation
relations are still obeyed. Furthermore one can show that the Dirac operator
built from (6.25) obeys the equation of motion (6.23) provided that the bosonic
field φ(x) obeys the equation of motion of the sine-Gordon model discussed in
Section 5.3, namely,
μ2
∂μ ∂ μ φ(x) + : sin(βφ(x)) := 0, (6.26)
β
and that the parameter β is related to the coupling constant g in the following
way,
β2 1
= g . (6.27)
4π 1+ π
From this last relation it follows that the special value β 2 = 4π corresponds to
g = 0 and hence a free Dirac fermion. Indeed as we shall see below for that value
of β the sine-Gordon potential translates into the bosonized mass term. It is
interesting to note a remarkable property of (6.27) which relates the coupling
constants of the Thirring model and its bosonic equivalent, the sine-Gordon
model. The weak coupling of one theory is the strong coupling of the other.
This property often occurs in bosonized theories and hints at the usefulness
of the method in dealing with theories for strong coupling, where perturbative
methods fail.
The bosonization dictionary of the vector fermion number current is the fol-
lowing,
β μν
JμV = : ψ̄γμ ψ : ↔ J μ = − ∂ν φ. (6.28)
2π
This expression differs from the bosonized current of the free Dirac theory (6.9)
in its normalization factor. We immediately realize that for the special value
β 2 = 4π we precisely reproduce (6.9). The normalization factor can be deter-
mined from the assignment of the fermion number charge of a soliton that should
be equal to the charge of the field ψ. Recall from Section 5.3 that the classical
sine-Gordon model admits a finite energy soliton solution. It is time-independent
and interpolates between adjacent wells of the scalar potential. In quantum the-
ory this classical solution becomes a particle. The static soliton solution is given
by (5.26),
4
φ= tan−1 [expμ(x − x0 )],
β
where x0 is the “center” of the soliton. Substituting this into the integral of the
current we find the fermion number of this solution to be,
β
Q= [φ(∞) − φ(−∞)] = 1. (6.29)
2π
Thus we see that indeed the normalization factor in the vector current (6.28) is
the right one.
β2
The relation (6.28) implies that the level in the affine Lie algebra will be 4π ,
as compared to 1 for the free case.
Let us address again the issue of the bosonization of the fermion mass bilinear
(6.22). The definition of the mass term in the “old formulation”, as that of the
current, requires some care due to the appearance of the products of operators
at the same point. In fact when x approaches y one gets the following OPEs,
cμ
ψR† (x)ψL (y) = |cμ(x − y)|δ : e−iβ φ :
2π
cμ
ψL† (x)ψR (y) = |cμ(x − y)|δ : eiβ φ : (6.30)
2π
β2
with δ = − 2π
g
(1 + 4π ). The proper fermion mass term will therefore be defined
by,
∞ ∞
−δ cμ
limy →x dx |cμ(x − y)| mψ̄(x)ψ(y) = m dx : cosβφ(x) : .
π
−∞ −∞
μπ
With μ chosen such that m = cβ 2 , the mass term transforms in the bosonic
language to,
μ2
ΔL = : cosβφ : .
β2
The normal ordering is with respect to μ.5
where ψL , ψR are left and right Weyl–Majorana spinor fields, ∂± = √12 (∂0 ± ∂1 )
and k = 1, . . . , N . The corresponding bosonic action is the Wess–Zumino–Witten
The associated Virasoro central charges of the two descriptions are identical,
as follows,
1 k[dim O(N )] 1/2N (N − 1) N
cf = N × cb = = = . (6.34)
2 k+N −2 1+N −2 2
For the fermions it is just N times the central charge of a single Majorana
fermion, whereas for the bosonized version we make use of the fact that the dual
Coexter number of O(N ) is N − 2. The conformal invariance of the action (15.2)
can be also shown by realizing that the corresponding β function vanishes. If one
generalizes (15.2) by taking a coupling 4λ1 2 as a coefficient of the first term and
k
24π of the WZ term (k integer), the β function associated with λ is given at the
one loop level (in the sense of expanding around u = 1), by,
2
dλ2 (N − 2)λ2 λ2 k
β≡ =− 1− ,
dlnΛ 4π 4π
Ji j (z) : ψi ψj (z) : iN
4π
[u−1 ∂u]i j (z)
J¯i j (z̄) : ψ̃i ψ̃j (z̄) : iN ¯ −1 ]i j (z̄)
[u∂u
4π
T (z) − 12 N i = 1 [ψi ∂ψi − ∂ψi ψi ]: − 2 (N1−1 ) : J a J a (z) :
N
T̄ (z̄) − 2 i = 1 : [ψ̃∂ ψ̃i − ∂ ψ̃i ψ̃i ]:
1
− 2 (N1−1 ) : J¯a J¯a : (z̄)
One way to prove the equivalence of the fermionic and bosonic theories now,
for N free massless Dirac fermions and the k = 1 WZW theory on U (N ) group
manifold, is by showing that the generating functionals of the current Green
functions of the two theories are the same. For the fermions we have,
−iW ψ (A μ ) 2
ψ
e = (dψ+ dψ− dψ̄+ dψ̄− )ei d x ψ̄ i D , (6.38)
(1)
μ ( 2 T ) + Aμ × 1 and ( 2 T ) generators of
1 A 1 A
where Dμ = ∂μ + iAμ , Aμ = AA
SU (N ). The term Wψ (Aμ ) was calculated by Polyakov and Wiegmann in a
regularization scheme which preserves the global chiral SU (NL ) × SU (NR ) sym-
metry and the local U (1) diagonal symmetry, leading to,
1
Wψ (Aμ ) = S[Ã] + S[B̃] + d2 xA(1)
μ A
μ(1)
, (6.39)
4πN
where Ã, B̃ ⊂ SU (N ) are related to the gauge fields AA
μ by
−1 −1
iAA
+ = (Ã ∂+ Ã)A , iAA − = (B̃ ∂− B̃)A .
In the bosonic theory one calculates,
e−iW B (A μ ) = [du]eiS [u ]+i d x(J − A + +J + A − )
A 2 B B B B
2 2 (1) (1)
e−iW B (A μ ) = [dφ]e 2 d x[(∂ φ) +(J − A + +J + A − )]
(1) i
(6.40)
(1) i −1
where J+B AB− and J+ A− are the appropriate parts of 4π T r[(g ∂+ g)A− ], and
(1)
similarly for the (− +) case and with A± = T r(A± ). These functional integrals
can be performed exactly, leading to,
1
WB (AA μ ) = S[Ã] + S[B̃] W B (A (1)
) = d2 xA(1)
μ A
μ(1)
.
4πN
Thus the bosonic current Green functions are identical to those of the fermionic
theory, the latter regulated in the way mentioned above.
where Nμ denotes normal ordering at mass scale μ. The fermion mass term
mq ψ̄ i ψi is therefore,
m Nμ d2 xT r(g + g † ),
2
where m 2 = mq c̃μ, mq is the quark mass, and c is the same constant as in
(6.19). It is straightforward to show that the above bosonic operator transforms
N −1 1
Δ = Δg + Δφ = + = 1, (6.42)
N N
G(zi , z̄i ) = <g(z1 , z̄1 )g −1 (z2 , z̄2 )g −1 (z3 , z̄3 )g(z4 , z̄4 )>, (6.43)
is given by,
G(zi , z̄i ) = [(z1 − z4 )(z2 − z3 )(z̄1 − z̄4 )(z̄2 − z̄3 )]−Δ g G(x, x̄), (6.44)
where I1 , I2 , I¯1 , I¯2 are group invariant factors. This result for the correlation
function, combined with the U (1) part gives an expression identical to that for
the fermionic bilinears. Moreover the result can be generalized to an n-point
function.
The bosonic fields g and h take their values in O(NF ) and O(NC ), respectively
and S[u] is the WZW action given in (15.2).
iNF −1 iNF
J+ab =: ψ+ai ψ+bi := (h ∂+ h)ab J−ab =: ψ−ai ψ−bi := (h∂− h−1 )ab ,
2π 2π
(6.48)
where : : stands for normal ordering with respect to fermion creation and anni-
hilation operators. As for the bosonic expressions for the currents, regularization
is obtained by subtracting the appropriate singular parts.
In terms of the complex coordinates z = ξ1 + iξ2 , z̄ = ξ1 − iξ2 (where ξ1
and ξ2 are complex coordinates spanning C 2 , and the Euclidian plane (ξ1 → x,
ξ2 → −t) and Minkowski space-time (ξ1 → x, ξ2 → −it) can be obtained as
appropriate real sections), one can express the currents as
iNC iNC −1
J(z)ij ≡ πJ−ij = (g∂z g −1 )ij J(z̄)ij
¯ ≡ πJ+ij = (g ∂z̄ g)ij ,
2 2
and similarly for the flavored currents.
In a complete analogy the theory of NF × NC Dirac fermions can be expressed
4π
NFNC −i φ
in terms of the bosonic fields g, h, e now in SU (NF ), SU (NC ) and
U (1) group manifolds respectively. The corresponding action is now,
1
S[g, h, φ] = NC S[g] + NF S[h] + d2 x∂μ φ∂ μ φ. (6.49)
2
4π
NCNF −i φ
This action is derived simply by substituting ghe instead of u in (6.31).
As for the equivalence between the bosonic and fermionic theories, we note
that in both theories the commutators of the various currents have the same
current algebra, and the energy-momentum tensor is the same when expressed in
terms of the currents. But the situation changes when mass terms are introduced
(see next section). The bosonization rules for the color and flavor currents are
obtained from (6.47) and (6.48) by replacing the Weyl–Majorana spinors with
Weyl ones, and in addition we have the U (1) current,
5
√ (1) † NF NC
J (z) ≡ πJ− = : ψ−ai ψ−ai :=
(1)
∂−φ
π
5
√ (1) † NF NC
¯
J (z̄) ≡ πJ+ = : ψ+ai ψ+ai :=
(1)
∂+φ . (6.50)
π
The affine Lie algebras are given by,
A B i
Jn , Jm = if A B C JnC+m + knδ A B δn +m ,0 ,
2
where J A = T r(T A J), T A the matrices of SU (NC ), k = NF for the colored cur-
rents and J(z) is expanded in a Laurent series as J(z) = z −n −1 Jn . A similar
expression will apply for the flavor currents with T I the matrices of SU (NF ),
¯
and the central charge k = NC instead of NF . The commutation relation for J(z̄)
will have the same form.
Generalizing the case of SU (N ) × U (1) to our case, the Sugawara form for the
energy-momentum tensor of the WZW action is given by,
1 A 1 I
T (z) = : J (z)J A (z) : + : J (z)J I (z) :
2κC 2κF
A I
1
+ : J (1) (z)J (1) (z) :, (6.51)
2κ
where the dots denote normal ordering with respect to n (n > 0 meaning anni-
hilation). The κs are constants yet to be determined. In terms of the affine Lie
generators this can be written as,
∞
∞
1 A A 1 I I
Ln = : Jm Jn −m : + : Jm Jn −m :
2κC m =−∞
2κF m =−∞
∞
1 (1)
+ : J (1) J : (6.52)
2κ m =−∞ m n −m
Now, by applying the last expression on any primary field φl we can get a set of
infinitely many “null vectors” of the form,
1
0
χnl = Ln − : JA JA :
2κC m =n m n −m
1 1
0 0
(1) (1)
− :J JI I
:− :J J : φl ,
2κF m =n m n −m 2κ m =n m n −m
for any n ≤ 0 (for n > 0 holds immediately). Since each of these vectors must
certainly be a primary field, Lm χn = Jm A n
χ = Jm I n
χ = Jm χn = O, which holds
trivially for m > 0. When checking for m ≤ 0, it leads to expressions for the vari-
ous κ, for the central charge c of the Virasoro Algebra, and for the dimensions of
the primary fields Δl = Δl+ + Δl− , in terms of NC , NF and the group properties
of the primary fields,
1
κC = κF = (NC + NF ), κ = NF N C
2
NC (NF2 − 1) NF (NC2 − 1)
c= + + 1 = NF N C
(NC + NF ) (NC + NF )
(c2l± )F (c2l± )C (c2 )(1)
Δl± = + + l± (6.53)
(NF + NC ) (NF + NC ) NC NF
where (c2l± )C is the eigenvalue of the SUR,L (NC ) second Casimir operator in
the representation of the primary field φl , namely ( 12 T A )( 12 T A ) = (c2l )C I, and
similarly for the flavor group. In the cases of SU (NC ) and SU (NF ) the discussion
2 2
applies to Δl+ or Δl− separately, with Cl+ and Cl− , respectively. Note that the
expressions for κF and κC of equation (6.53) are an immediate generalization
of the case of the group SU (N ) with the central term equal to one. There the
factor was N + 1, the N being the second Casimir of the adjoint representation,
and the 1 being the central term.
The equivalence of the bosonic and fermionic Hilbert spaces was demonstrated
by showing that the two theories have the same current algebra (affine Lie alge-
bra), and that the energy-momentum tensor can be constructed from the cur-
rents in a Sugawara form. Goddard et al. [110] showed that a necessary and
sufficient condition for such a construction of the fermionic T (z), in a theory
with a symmetry group G, is the existence of a larger group G ⊂ G such that
G /G is a symmetric space with the fermions transforming under G just as the
tangent space to G /G does. Based on this theorem they found all the fermionic
theories for which an equivalent WZW bosonic action can be constructed. The
cases stated above fit in this category. Note in passing that this does not hold for
cases where the symmetry group includes more non-abelian group factors, like
for example SU (NA ) × SU (NF ) × SU (NC ) × U (1).
The prescription equation (6.49) described above , for the bosonic action that is
equivalent to that of colored and flavored Dirac fermions, is by no means unique.
In fact it will be shown that this prescription will turn out to be inconvenient
once mass terms are introduced. Another scheme, based on the WZW theory of
U (NF NC ) will be recommended.
Consequently, the bosonic form of the fermion mass term mq ψ̄ ia ψia is,
4π
2 2 † † −i N F N C φ
m Nμ d x(T rgT rh + T rh T rg )e , (6.55)
with m 2 = mq c̃μ. Once again the bosonic operator (6.54) has the correct chiral
transformations and the proper dimension,
NF2 − 1 NC2 − 1 1
Δ = Δg + Δh + Δφ = + + = 1.
NF (NF + NC ) NC (NC + NF ) NC NF
Unfortunately, the explicit calculation of the four-point function reveals a dis-
crepancy between the fermionic and bosonic terms in (6.54). This can actually
be understood directly. Since g and h are fields defined on entirely independent
group manifolds, then (ignoring for a moment the U (1) factor) the four-point
function of the mass term can be written as,
function and in fact any Green’s function will now reproduce the results of the
fermionic calculation.
The currents constructed from u obey the Affine Lie algebra with k = 1.
The color currents, for instance, are J A = T r(T A J), where T A are expressed as
(NC NF ) × (NC NF ) matrices defined by λA ⊗ 1, with λA the Gell–Mann matri-
ces in color space and 1 stands for a unit NF × NF matrix. The central charge
is k = NF . The same arguments will apply for the flavor currents, now with
k = NC . The central charge for the U (1) current is NC NF .
To see the difference between the present theory and the previous one let us
express u in terms of (NF NC ) × (NF NC ) matrices g̃, h̃ and ˜l in SU (NF ), SU (NC )
and the coset-space,
respectively, through,
4π
−i φ
u = g̃ h̃˜le NC NF
.
Using the formula for expressing an action of the form S[AgB −1 ] we get
˜ 1
S[u] = S[g̃ h̃l] + d2 x∂μ φ∂ μ φ
2
1
S[g̃ h̃˜l] = S[g̃] + S[˜l] + S[h̃] + d2 xT r(g̃ † ∂+ g̃˜l∂− ˜l† + h̃† ∂+ h̃˜l∂− ˜l† ).
2π
We can now choose ˜l = l so that l∂− l† will be spanned by the generators that
are only in SU (NF × NC )/{SU (NF ) × SU (NC ) × U (1)}. This can be achieved
by taking ũ = g̃ h̃˜l, which is u but without the U (1) part, and then taking for
h̃ = h ⊗ 1 a solution of the equation ∂− hh† = N1F T rF [(∂− ũ)ũ† ], and similarly for
g with N1C T rC . These are also the conditions that the flavor currents should be
expressed in terms of g̃ and the color currents in terms of h̃. For this choice, the
mixed term in the above action, the term involving products of ˜ls with g̃s or h̃s,
is zero, and so the new action is,
1
S[u] = NC S[g] + NF S[h] + d2 x(∂μ φ∂ μ φ + S[l]).
2
Note that l is still an SU (NC NF ) matrix, while g and h are expressed as SU (NF )
and SU (NC ) matrices respectively, but the matrix l involves only products of
color and flavor matrices (not any of them separately).
action of a fermion with one given chirality. It is also easy to factorize a scalar
field into its left φL (x− )(φ(z)) and right moving φR (x+ )(φ̄(z̄)) parts since the
solution of equation of motion of a scalar field in real and complex coordinates
takes the form
However, as will be discussed shortly it turns out that it is quite subtle to write
down an action of a chiral boson which is equivalent to that of a left or a right
chiral fermion. Once we establish an action for a chiral boson, the question is
how can one couple it to abelian and non-abelian gauge fields?
In this section we will construct two seemingly independent constructions of
the action of a chiral boson. In fact, it will turn out that one formulation is a
special case of the other. We start with Siegel’s formulation which is based on
the coupling of a scalar field to fictitious gravity in a light-cone gauge [194] and
then we describe a manifestly non-Lorentz invariant action [92] which is a special
case of the former.7 In [36], [199], and in [100] the two formulations were related
and further generalizations were discussed. We follow the latter paper here.
Chiral bosons play an important role in string theories and in particular chiral
bosons on Riemann surfaces of any genus. Here we will not enter into discussions
on these constructions and describe chiral bosons only on a two-dimensional flat
Minkowsky space-time.
g + + = 0 g +− = 1 g −− = 2λ. (6.58)
Since we have fixed only part of the local symmetries of (6.57) it is straightfor-
ward to realize that the last action is still invariant under the following local
transformation
These equations imply classically the chiral nature of the boson, namely a left
moving boson φ(x+ ) and the conservation of the axial current. Note that unlike
an ordinary scalar field the chiral scalar action (6.59), admits only the “holomor-
phic” conservation of the left current but not the “anti-holomorphic” conserva-
tion of the right one namely,
which implies that there is only left affine symmetry but not a right one.
So far we have discussed the classical system. Quantum mechanically it turns
out that the symmetry (6.60) is anomalous. There are several ways to verify this
anomaly. Probably the easiest way is to realize the resemblance of the action
(6.57) to that of the bosonic string. It is well known that the latter is consistent
only in 26 dimensions and not, as our case seems to be, in one dimension. Tech-
nically this follows from the fact that the ghost system associated with the fixing
of the fictitious diffeomorphism and Weyl invariance have a Virasoro anomaly
(see Section 6.5.1) equal to −26. In fact following the discussion in Section 6.5.1
we know that there is another way to cancel this anomaly and that is to add to
the action a background charge term of the following form,
where q is the background charge and R(2) is the fictitious scalar curvature.
This modification of the action yields a modified energy-momentum tensor as
form of gauge fixing χ3 = λ(x) − λ0 (x) with λ0 (x) a given function, than all the
constraints are second order with the constraint algebra,
cij (x, y) = {χi (x), χj (y)} c22 = −2δ (x − y) c13 = −δ(x − y). (6.71)
(6.72)
where cij (x, z)c−1
jk (z, y) = δ ik (x − y). The Dirac bracket rather than the Poisson
bracket is then elevated to the commutator in the quantum theory [ ] = i{ }.
Using this prescription one finds the desired result,
1
[φ(x), φ(y)] = (x − y). (6.73)
4i
Implementing the constraint quantization in the path integral formulation, one
has,
Z(J) = [dφ][dπ][dλ][dπλ ]δ(χ1 )δ(χ2 )δ(χ3 )
2
× Det[Cij (x, y)]ei d x(πφ̇+πλ λ̇−H−J φ) . (6.74)
Thus we see that this procedure yields the action (6.59) in the gauge λ = −1
which will be the topic of the following section.
where ρ(x) is a local bosonic field. The system can also be described in terms of
the non-local bosonic field,
1
φ(x) = dy(x − y)ρ(y), φ (x) = ρ(x), (6.77)
2
with a local Lagrangian density,
L = φ φ̇ − φ = ∂− φ(∂+ φ − ∂− φ).
2
(6.78)
As we said the Lagrangian density is in fact (6.59) in the gauge 2λ = −1. The
classical equation of motion which corresponds to (6.76) is,
∂− φ = 0. (6.79)
As usual from the vector and the axial currents we can write down the left
and right currents J(lr) = 12 [j(v ) ± j(ax) ], respectively. They have the following
expressions,
Note however that only the left current is holomorphically conserved namely
∂− J(l) + = 0, while the right current is not antiholomorphically conserved. This
property is related to the invariance of the Lagrangian under δφ = α(x+ ) and
not under δφ = α(x− ). As will be explained below only the left U (1) affine Lie
algebra current exists in the quantum theory. Similarly, the Lagrangian (6.76) is
invariant under the conformal transformations δφ = (x+ )φ and δφ = (t)∂− φ.
The associated Sugawara type Noether currents are,
Note that only k ≤ 0 appears in the decomposition of φ(x), which expresses the
chiral nature of the field. The single-particle Hilbert space is then a continuum
of states with energy E = |k|, k ≤ 0. Hence the Hamiltonian formalism has cor-
rectly implemented the chirality constraint ∂− φ = 0. Furthermore, this property
can also be deduced from the Hamiltonian equation of motion φ̇ = i[H, φ] = φ .
Note that to get the chiral solution ∂− φ = 0 as a solution of the equation of
the motion, we had to assume chiral boundary conditions. Here it looks at first
that the chirality property was derived with no assumptions, but in passing
from (6.83) to (6.84 we assumed that (π − φ )(x = ∞) = −(π − φ )(x = −∞),
so together with choosing zero surface terms we in fact assumed chiral boundary
conditions.
For the path integral quantization of the system we use the method developed
for Hamiltonian systems with constraints. The generating functional is given by,
Z[J] = dφdπδ(χ)e d x(πφ̇−Hc −J φ)
2
d 2 x(L−J φ)
= dφe , (6.87)
J+ =: ψ † ψ T+ + = i : ψ † ∂+ ψ := π : J+ J+ :, (6.90)
obey the left U (1) affine Lie and Virasoro algebras, respectively with the well-
known central charges k = 1 and c = 1.
It is now straightforward to realize that the J(l) = φ and T(l) = (φ )2 have the
same k = 1 and c = 1 central charges. Using the operator algebra we can now
evaluate the commutators of the chiral current and of the energy-momentum
tensor. The results are as follows
i
[J(l) (x), J(l) (y)] = [φ (x), φ (y)] = δ (x − y),
2
[T(l) (x), T(l) (y)] = [: (φ (x))2 , (φ (y))2 ] = i(T(l) (x) + T(l) (y))δ (x − y)
i
− δ (x − y). (6.91)
24π
From the general discussion in Sections 2.4 and 3.3 it follows now that these
commutation relations correspond to central charges of c = k = 1. Hence our
bosonic theory furnishes the same irreducible representation of the affine Lie
and Virasoro algebras as one free left-handed chiral fermion. Since for k = 1
the affine Lie algebra has a unique irreducible unitary representation the two
theories on flat two-dimensional space-time are therefore equivalent. Below it
will be shown that the anomalies of the bosonized theory in coupling to gauge
and gravitational background are the same as for the fermionic theory.
L = LL + LR (6.92)
φ = φL + φ R π = π L + πR (6.94)
Partition function
We would now like to compare the one loop partition function of a chiral fermion
and that of our chiral boson. We therefore pass to a two-dimensional space-time
domain with 1 ≥ x ≥ 0. The mode expansion for φ previously given by equation
(6.85) now takes the form,
1 +
a†n e2πin x + an e−2πin x .
+
φ(x, t) = φ0 + p(x + t) + √ (6.97)
n>0
2n
The one loop partition function which corresponds to this mode expansion is
where q = eiτ π and we use the fact that H = −P . Let us first calculate the
contribution to the trace of the oscillation modes,
n a †n a n − 112 )
q − 1 2 [1 − q 2n ]−1 = η −1 (τ ),
H 1
Tr[q π ] = Tr[q ( n ]= (6.99)
n=1
where L0 is the uncoupled Lagrangian density and V stands for the vector con-
serving scheme. The vector current is still obviously conserved, but the divergence
of the axial current,
is now equal to the anomaly deduced in the fermionic theory from the one loop
diagram.
Next we discuss the bosonization in the left-right scheme. For that purpose a
term bilinear in the gauge fields has to be added to the J(l) + A(l) − term. The
Lagrangian then takes the form:
√
2
L(LR) = L0 + J(l) + A(l) − + A(l) − A(l) 1 , (6.103)
4
where (LR) indicates the left-right scheme. The divergence of the left current
∂ L( L , R )
which is derived from ∂ A ±
now has the form,
1 1
∂− (J(l) − )(L,R) + ∂+ (J(l) + )(L,R) = (∂− A+ − ∂+ A− ) = μν Fμν . (6.104)
4 4
The Lagrangian (6.103) is therefore really the bosonized fermionic action regu-
larized in the left-right scheme. Obviously, a similar prescription for the (V) and
(LR) schemes can be applied to the right chiral boson. It is straightforward to
show that the vector as well as the left-right actions are invariant under “curved
space-time” Lorentz transformations as discussed above.
In a similar manner to the abelian case this is a WZW action coupled to fictitious
chiral gravity in the gauge h+ + = −1. In fact we can consider a generalization
of this action to the so-called k-level chiral WZW namely S+k [U ] = kS+ [u]. The
equation of motion that follows from the variation of (6.105) with respect to the
variation of u can be expressed as,
The action S+k [u] is invariant under the left affine transformation δu = (x+ )u.
The corresponding energy momentum tensor has again a Sugawara form and
Virasoro central charge which are,
2π a a kdimG
T(l) = : J(l) J(l) : c = , (6.108)
(c2 + k) a c2 + k
a
where as usual J(l) = J(l) T a and T a are hermitian matrices representing the alge-
bra of the group, C2 is the second Casimir operator in the adjoint representation
and dimG, is the dimension of the group.
The coupling of non-abelian gauge fields to a chiral WZW action, is a straight-
forward generalization of the coupling of the abelian gauge fields. Again there are
several ways to couple gauge fields corresponding to the various regularization
schemes in the fermionic theory. The bosonized action (for k = 1) related to the
vector conserving regularization scheme is given by,
− +
SV [u, A− , A+ ] = S+ [u] + d2 xTr[J(v ) A− + J(v ) A+ ]
√
2
− d2 xTr[u−1 A1 uA− − A− A1 ], (6.109)
2π
where J(v ) and J(ax) are constructed from the left and right currents of (6.107)
in the usual way. Using the equation of motion one finds the conservation of the
vector current and the anomalous divergence of the axial current,
μ μ 1 μν
Dμ (J(v ) )V = 0 Dμ (J(ax) )V = π Fμν (6.110)
The coupling of a left non-abelian gauge field that corresponds to the fermionic
description in the left-right regularization scheme is given by,
√
2 1
S(LR) [u, A− , A1 ] = S+ [u] + d2 xTr[(u−1 ∂1 u + A1 )A− ]. (6.111)
2π 2π
The associated current divergence is,
1 μν
μ
Dμ (J(l) )LR = D− Jl− (LR) + D+ J−
+
l(LR) = Fμν . (6.112)
π
This expression for the anomalous divergence of the left current is identical
to the result of the loop calculation in the fermionic version regularized in the
left-right scheme.
scalar fields. In this section it will be shown that indeed such a map exists for
two families of theories, one with anti-commuting fields and the other with com-
muting ones with arbitrary integer and half-integer dimensions.
The bosonization of the b, c and β, γ systems was introduced in [94]
hb = λ hc = (1 − λ). (6.114)
imply that both fields are holomorphic, namely b(z), c(z). Similar to the deriva-
tion of the operator equations of motion using the path integral (1.55) for the
scalar field we find for the (b, c) system that,
¯
∂b(z)c(0) = 2πδ 2 (z, z̄). (6.116)
δb = ¯∂b + λ(∂¯
)b,
δc = ¯∂c + (1 − λ)(∂¯
)b. (6.118)
For holomorphic ¯(z) this is a symmetry transformation. Taking now ¯(z, z̄) we
read the energy-momentum tensor from the variation of the action as follows,
1 ¯T = 1 ¯[(∂b)c − λ∂(bc)].
δSb,c = − d2 z ∂¯ d2 ∂¯ (6.119)
2π 2π
operator cannot describe the (b, c) systems which are a family of CFTs. We
also need to identify a family characterized by the parameter λ of scalar field
theories. A simple way to achieve this is to realize that the energy-momentum
of the general (b, c) system can be written in terms of the spin 1/2 fermion as
follows,
1
T b,c = T ψ − λ − ∂(bc). (6.127)
2
Thus following (6.126) the scalar energy momentum has to have the form,
1
Tλφ = T φ − λ − ∂ 2 φ. (6.128)
2
This is the energy-momentum of a scalar field with a background charge, or the
linear dilaton theory with q = −i(λ − 12 ). The central charge of these theories
was shown to be,
Moreover using the fact that the dimension of an operator : eik φ(z ) : was shown
2
to be k2 + ikq it is easy to check that,
he i φ = λ = hb he −i φ = 1 − λ = hc , (6.130)
which verifies that the operators mapped by the bosonization indeed have the
same conformal dimensions.
The anti-commutative nature of the b, c system is obeyed also by their bosonic
duals, just as for the case of spin 1/2 fields, namely,
at equal times, namely |z| = |w| since for that case [φ(z), φ(w)] = ±iπ.
where . . . stands for non-singular terms. The corresponding scalar theories have a
background charge of (λ − 1/2) for φ and i/2 for χ so that the energy-momentum
tensor of the full bosonic reads,
1 1 1 1
Tφ,χ = − ∂φ∂φ + ∂χ∂χ + (1 − 2λ )∂ 2 φ + ∂ 2 χ, (6.135)
2 2 2 2
which yields the desired Virasoro anomaly c = 1 + 3(2λ − 1)2 = −cb,c . The
bosonic operators that correspond to β and γ are,
Alternatively, as was discussed in Section 6.4 one can add to this Lagrangian
density also a term of the form − √ i R2 φ. The corresponding energy-
2(k + 2)
momentum tensor T (z) is given by,
1 i
T (z) = −β∂γ − ∂ϕ∂ϕ − ∂ 2 ϕ, (6.138)
2 2(k + 2)
and the non-trivial OPEs of these field are given by,
1
γ(z)β(w) = + O(z − w) ϕ(z)ϕ(w) = −ln(z − w) + O(z − w). (6.139)
z−w
7.1 Introduction
The number of approximation techniques in quantum field theory is very limited.
1
Perturbation expansion in small interaction coupling, like αem = 137 of QED, is
obviously the most important one. Other methods include dimensional expan-
sion, high temperature expansion and large radius expansion. Quite surprisingly,
one of the most useful approximation techniques is expansion in the number of
degrees of freedom. A priori one would tend to think that the larger the number
of degrees of freedom, the more complex the system. However, it turns out that
theories with infinitely many degrees of freedom are much easier to solve than
those with a finite number of degrees of freedom. Once the system with N → ∞
is known, a systematic expansion in N1 provides an approximation procedure for
computing quantities that describe systems of finite N .
Large N methods have been applied in a very wide range of physical sys-
tems. Starting from non-critical phenomena in spin systems like the Heisenberg
ferromagnet (discussed in Section 5.14), then SU (N ) QCD theories in various
dimensions, and later matrix models associated with either string models or
two-dimensional models.
The large N approximation in field theory or correspondingly the planar
expansion of Feynman diagrams was introduced by ’t Hooft in his seminal paper
[122].
In this book we will focus on four arenas where large N approximations are
being used:
(i) Two-dimensional quantum field theory models, which include the Gross–
Neveu model and the CP N models, will be addressed in this chapter.
(ii) Quantum chromodynamics with large N SU (N ) gauge theory in two dimen-
sions. In Chapter 10 of Part 2 the solution of two-dimensional QCD, follow-
ing ’t Hooft [124], will be described,1 together with a certain generalization
of it.
(iii) The approach to four-dimensional QCD based on the N1 expansion.
(iv) Baryons in large N QCD.
The last two topics will be described in the third part of the book in Chapters
19 and 20.
In Nature the number of colors is three, and thus one may wonder whether
it makes sense to in expand a not-so-small parameter 1/3. Even though there
is no general proof that this expansion is indeed reliable, a vast literature on
the subject brings out a large amount of evidence that indeed this is the case.
It turns out that for certain quantities the N1 term vanishes and the correction
starts as N12 , and hence puts the approximation on a more solid base.
As an example of the accuracy of√the large N limit consider the Stirling formula
for N , where the leading term is 2πN (N/e)N for large N. But the correction
1
is actually 12N as compared to 1, making it only an 8 percent correction even
for N = 1.
In the next sections we describe the O(N ) model, the Gross–Neveu model and
the CP N models.
There are several review articles on large N expansions, for instance [46], [160],
[165]. In this chapter we make use of [160].
The discrete chiral symmetry forbids a mass term. In fact this is the most general
action invariant under these symmetry transformations, with terms of dimension
two or less, and hence it is a renormalizable action.
Let us now check whether in this formulation of the Lagrangian, where λ0 is
fixed, one can make sense of a large N limit. Consider the scattering process of
two fermions with flavor index a that turn into a pair of fermions with a different
flavor index b. The leading Feynman diagrams that contribute to this process
are given in Fig. 7.1. The first diagram which is the basic interaction vertex is
of order λ0 , the second one is of order λ20 , the third is of order λ20 N due to the
N different flavors of the fermions that can run in the loop, and the last two
diagrams are of order λ30 N 2 . It is thus clear that the perturbation expansion
a b a b
a b
a b a b
(a) (b)
a b a b a d b
c
c d c
c d c
c
a b a b a b
(c) (d) (e)
expressed in terms of the fixed coupling λ0 does not have a sensible large N
expansion. However, one can easily cure this problem by defining the coupling
λ ≡ λ0 N , which is taken to be fixed when N → ∞. The Lagrangian now reads,
λ
LG N = iψ̄ a ∂ψ a + (ψ̄ a ψ a )2 . (7.3)
2N
N 2
LGN = iψ̄ a ∂ψ a + σ ψ̄ a ψ a − σ . (7.4)
2λ
i
p/ 1
N
We would now like to integrate over the fermions and derive an effective
Lagrangian, Leffective (σ). Since the Lagrangian is quadratic in the fermion fields,
Leffective (σ) is given by the sum of terms (Fig. 7.4).
Note that all diagrams are with an even number of σ only, since σ is odd under
the transformation (7.2). The first term is the tree level contribution − 2λN 2
σ and
the rest are the one loop contributions. Both are of order N , the latter due to the
N fermions that can run in the loop. The N dependence of Leffective therefore
has the form,
This makes the counting of powers of N1 very easy. Consider a graph with E
external σ lines, I internal σ lines, V vertices and L independent loops. The
parameters (E, I, V, L) are not independent. For each internal line there is a
momentum integration and hence a loop. However each vertex introduces a delta
function in momenta that cancels one momentum apart from an overall delta
function associated with the momentum conservation.
Thus one has,
L = I − V + 1. (7.6)
1
Recall that each σ external or internal line carries a N factor, while each vertex
contributes a factor of N .
Thus the net power N of each graph is,
N −I +V −E = N −E −L +1 . (7.7)
It is obvious from this expression that adding loops and external σ lines sup-
presses the corresponding contribution due to additional powers of N1 . Since the
minimal number of σ external lines is two the leading behavior is of order N1 .
For the purpose of investigating the possibility of spontaneous breaking of
the discrete symmetry of (7.2), it is enough to compute the effective potential
V (σ) rather than the effective action, namely, the limit where all the external
lines carry zero momentum. The effective potential is given by the sum of the
diagrams in Fig. 7.4:
∞ 2n
N σ2 1 d2 p − pσ
−iV = −i −N Tr , (7.8)
2λ n=1
2n (2π)2 p2 + i
1
where 2n is the symmetry factor of the graph, N comes from summing over all
possible flavors, –1 from the fermion loop and the expression in the bracket is the
p/
product of the propagator and (i p 2 +i ) and the vertex (iσ). Using the identity,
∞
x2n 1
= − log(1 − x2 ), (7.9)
n=1
2n 2
The fact that the cutoff disappears obviously implies that the theory is indeed
renormalizable. The β function and the anomalous dimension can be determined
by substituting the effective potential into the renormalization group equation
(see Section 17.6),
We thus find the exact (to all orders of λr ) expression of β(λr ) and γσ (λr ), which
take the form,
λr 3
β(λr ) = − γσ (λr ) = 0. (7.15)
2π
are at
π
σ=0 and σ = ±σ0 = ±μe1− λ r , (7.17)
where
σ02
V (0) = 0 and V (σ0 ) = −N < 0. (7.18)
4π
in the Gross–Neveu model. However, since the interaction has the form of a
vector times a vector, the auxiliary field should also be a vector. This shifts the
Lagrangian according to,
2
λ N
LC P N → LC P N + Jμ + Aμ
N λ
N
= ∂μ Z † ∂ μ Z + 2Jμ Aμ + Aμ Aμ
λ
= (∂μ − iAμ )Z † [(∂ μ + iAμ )Z] , (7.27)
where in the third line we have used the constraint. It is clear from its last form
that the Lagrangian is invariant under the U (1) local transformation,
Z → eiα Z Aμ → Aμ − ∂μ α. (7.28)
We now incorporate the fact that the Z are constraint variables by introducing
another Lagrange multiplier into the Lagrangian,
N
LC P N = (∂μ − iAμ )Z † [(∂ μ + iAμ )Z] + σ Z † Z − . (7.29)
λ
Obviously the path integral over σ, or equivalently using its equation of motion,
implies that Z † Z = Nλ .
The action is now quadratic in Z, so we integrate out the Z fields similarly to
what was done in the Gross–Neveu model. The Feynman diagrams that consti-
tute the leading contributions to the effective action, which is now a functional
of σ and Aμ , are drawn in Figure 7.5.
These diagrams, which include pure σ, pure Aμ and mixed diagrams, are all
proportional to N . The computation of V (σ) is similar to that in the GN model,
leading to,
σ σ σ
V (σ) = −N + log 2 − 1 , (7.30)
λ 4π Λ
where Λ is the cutoff. Again similar to the GN model the cutoff can be eliminated
by performing a renormalization at a scale μ,
1 1 dV 1 1 μ2
=− |μ 2 + log , (7.31)
λr N dσ λ 4π Λ2
so that the potential takes the form,
σ σ σ
V (σ) = −N + log −1 . (7.32)
λr 4π μ2
It is also evident that the model has a negative β function, or differently stated,
for fixed λr and μ, when Λ → ∞, λ vanishes, namely, the model is asymptotically
free.
(ii) The Z fields can be interpreted as bosonic “quarks” in the fundamental rep-
resentation of the group, though transforming in a non-linear way. These Z
quarks are confined due to the dynamically generated abelian gauge inter-
action. As will be shown in Chapter 8, in two dimensions the abelian force
between a quark anti-quark pair is linear in separation distance.
Two-dimensional non-perturbative
gauge dynamics
In the first part of the book we have developed several non-perturbative tools
for analyzing two-dimensional field theories. These include methods associated
with conformal invariance, with affine Lie algebras and in particular the WZW
model, techniques of integrable massive theories including solitons, S-matrix and
the Yang–Baxter equation, sets of infinitely many conserved charges, the thermal
Bethe ansatz, methods of bosonization and the large N limit approximation.
In this second part of the book the main idea is to implement those meth-
ods in the context of two-dimensional gauge dynamics aiming at extraction of
the mesonic and baryonic spectra of two-dimensional QCD, decoding the con-
fining behavior, versus a screening one, and analyzing models with other quark
representations and models of generalized QCD.
The most important tool that will be used for this purpose will be bosoniza-
tion. It will enable us to easily solve the Schwinger model, derive the baryonic
spectrum of QCD2 using a strong coupling limit, determine the screening nature
of massless QCD models and compute the string tension for the massive ones.
Using the large N approximation we will extract the mesonic spectrum of QCD.
In two respects this part is an intermediate stage on the way to four-
dimensional gauge dynamics. Firstly, as was just mentioned, we will gain expe-
rience from applying non-perturbative methods on the simpler two-dimensional
models, which will serve us when using them in the context of “real” physi-
cal systems. Secondly, two-dimensional gauge systems will serve as a toy model
laboratory of four-dimensional ones. As will be shown in the third part of the
book, certain aspects of two-dimensional physics will survive the transition to
four dimensions. Obviously the challenge will be to identify those phenomena
and devise some additional tools to handle the other cases.
Fμν = ∂μ Aν − ∂ν Aμ (8.2)
where in the second line the action is expressed in terms of the light-cone deriva-
tives and components of the gauge fields, and the Dirac fermion is decomposed
into its left and right chiral fermions, as discussed in Section 3.8. It is evident
that with the gauge field having a vanishing dimension, the gauge coupling e has
a dimension of mass. Thus the action is not invariant any more under the two-
dimensional global conformal symmetry, but rather only under the ISO(1, 1)
Poincare group. For the massless case, the action is classically invariant under
the global transformations,
Ψ → eiα Ψ Ψ → ei α̃ γ 5 Ψ
ψ → ei(α + α̃ ) ψ ψ̃ → ei(α −α̃ ) ψ̃. (8.6)
In fact the left and right chiral transformations, for the massless case, can be
lifted also into holomorphic and anti-holomorphic transformations, as was dis-
cussed in Section 3.7.1. The corresponding vector and axial currents
Again by construction the action is also invariant under the gauge transfor-
mation,
1
Ψ → e−iΛ(x,t) Ψ Aμ (x, t) → Aμ (x, t) + ∂μ Λ(x, t). (8.8)
e
Quantum mechanically the axial current is not conserved even for the massless
case due to an anomaly,
e
∂μ J5μ = μν F μν . (8.9)
2π
We will derive this result using the bosonized version, see Section 9.1. Unlike
Maxwell’s theory, this theory has non-trivial degrees of freedom. However, once
again the gauge field is not dynamical. This phenomenon can be easily demon-
strated in the axial gauge A1 = 0, where the other component A0 can be solved
1 The Schwinger model was introduced in [190] and further analyzed in [68] and [64].
2 The θ angle was introduced by Lowenstein and Swieca [152] and also by Coleman [64].
3 The Yang–Mills non-abelian gauge theory was introduced in the seminal paper [229].
but only with respect to the ISO(1, 1) Poincare transformations. The action is
invariant under a non-abelian gauge transformation, which in infinitesimal form
is,
∂0 F 01 = 0 ∂1 F 10 + i[A1 , F 01 ] = 0. (8.15)
From the first we get that ∂02 A1 = 0, and thus A1 = f1 (x1 ) + x0 f2 (x1 ). Using
the residual gauge invariance, of gauge transformations that depend only on x1 ,
we can go to f1 = 0, and then the second equation implies that f2 is a constant
C, which yields F01 = C, and then again the requirement of finite energy results
in C = 0. This will also be shown in a complicated way using a BRST approach
in Chapter 15.
When the underlying manifold has a non-trivial topology like that of a torus
then the theory is not totaly empty but instead has topological degrees of free-
dom. This will be described in Section 16.
Finally, the non-abelian case is different from the abelian in higher dimensions,
as the former is not free there. While the abelian case represents free photons,
the non-abelian case represents interacting gluons, which turn to interacting glue
balls in the physical space.
1 The treatment of the bosonized Schwinger model was done in [68] and [64].
where c is the constant of bosonization and the normal ordering is with respect
to the mass m, as was explained in Section 6.1.1.
After a shift in the definition of φ,
1 θ
√ , φ→φ+
2 π
√
and normal ordering with respect to μ = e/ π, one finds,
1 2 1 1 cmμ √
H =: π + (∂1 φ)2 + μ2 φ2 − cos(θ + 2 πφ) :. (9.2)
2 2 2 π
From this expression the periodicity in θ is manifest. The angle θ is the conjugate
to the winding number, appearing in two dimensions for the abelian case, since
Π1 [U (1)] = Z (looking at a circle of large radius in the two-dimensional plane).
Physics is invariant under θ → θ + 2π. From (8.10) it is clear that 2π
eθ
corresponds
to a background electric field. The periodicity is due to the ability to produce
electron-positron pairs in the vacuum when | 2π eθ
| > 12 e, and these pairs create
their own electric field which reduces the original one.
When we set m = 0 we discover that the massless Schwinger model is in fact
a theory of one free bosonic field with a mass equal to μ.
In the strong coupling limit, the bosonized form of the Hamiltonian is very
useful. The theory contains a meson of a mass that is approximately μ, and the
number of bound states depends on the value of θ. It can be shown that there
are no bound states for |θ| > π/2. For 0 < |θ| ≤ π/2 there is a stable two-body
bound state, while for θ = 0 there is also a three-body bound state.
Note that even though the Hamiltonian density (9.2) resembles that of a sine-
Gordon model, it does not admit soliton solutions due to the mass term 12 μ2 φ2 .
We will come back later to analyze this bosonized Hamiltonian, in the con-
text of the question whether the system admits screening or confinement in
Chapter 14.
Finally, let us show how the anomaly arises in the bosonized version. The
equation of motion for the electromagnetic field is,
Bosonizing now the various parts of the Hamiltonian one then gets,2
H = HΨ0 + HE − H I
1 2 cmμ √
0
HΨ = Σai [π + (∂1 φai ) ] +2
: (1 − cos(2 πφai )) :
2 ai π
2
e2c
HE = (φai − φbi )
8πNc i
ab
x
2c2 μ2 √
I
H = 3 Σa= b Σij Kij,ab Nμ cos π (πai − πaj + πbj − πbi )(ξ)dξ
π2 −∞
−1
√
sin( π(φai + φaj − φbj − φbi )(ξ)) (φak − φbk ) , (9.12)
ab
2 Abelian bosonization of two-dimensional QCD was discussed in [24] and [201] and was further
elaborated in [62].
HΨ0 is the free “fermionic” part, after bosonization, thus in terms of bosonic
variables; HE is the first term of the Hamiltonian (9.9) rewritten in terms of the
boson variables corresponding to the fermions, by eliminating the electric fields
through the Gauss law. Thus although originally coming from the kinetic part of
the gauge potentials, it actually involves the interactions. This is a result of the
ab
fact that there are no transverse vectors in 1 + 1 dimensions. Kij is a properly
3
generalized ordering operator.
In the case of one flavor, i = j = 1, H I does not involve the π variables.
The interaction involves non-local terms which relate to color non-singlets.
For static and ec → ∞ approximations one finds that for NF = 2 the interaction
is field independent. For NF ≥ 3, on the other hand, the limit is singular. This
singularity should not be there in the predictions of physical quantities, but it
renders any further treatment very complicated.
It is thus clear that a different method of bosonization is required for the
treatment of flavored QCD2 . In the following it will be shown that the “non-
abelian bosonization”, based on the WZW model discussed in Section 6.3, is an
adequate tool for this purpose.
Before proceeding to non-abelian bosonization and in particular to gauge the
color symmetry group of the colored-flavored WZW model, we describe briefly
another approach, in which the flavor sector appears in the form of a WZW
model, but for the color degrees of freedom the gauged abelian bosonization is
invoked. As we have seen above one can use the Gauss law to express the gauge
fields in terms of the appropriate fermionic bilinear, which translate into bosonic
group elements as,
i
2∂1 ea = Ψ†ia Ψia = ∂1 TrF (log ga ), (9.13)
i
π
√
where ga ∈ U (NF ) is one out of NC such matrices, and ea = 2 πEaa , the diagonal
element. One can also express Aba for a = b in terms of fermion densities. Inserting
these into the QCD2 Hamiltonian one gets,
H = H0 + HI
(ec )2 −1 2 2 Tr(ga gb−1 )
HI = − Tr log ga gb − πμ
a,b
32π2 NC
a,b
Tr log(ga gb−1 )
+ mcμ NF Tr(ga ), (9.14)
a
H0 includes the fermion kinetic term. For NF = 2 the potential is free from
singularities, for NF ≥ 3 it is not. In the case of NF = 2 the low lying baryonic
spectrum can be extracted. Here we will not follow this approach further and
instead will move on to the fully non-abelian bosonization.
The variation of S (1) is derived using the infinitesimal variation of the gauge field
δAμ = −Dμ = −(∂μ + i[Aμ , ]). J μ is found to be,
−1 †
Jμ = {[h Aμ h + hAμ h† − 2Aμ ] − εμν [h† Aν h − hAν h† ]}. (9.22)
4π
The second iteration will be given by adding S (2) , where now J μ is replacing
Jμ,
S = d xTr(Aμ J ), δ S [h] = −2 d2 xTr(∂μ J μ ).
(2) 2 μ (2)
(9.23)
will contain (D-1) derivatives, and is gauged by replacing the ordinary derivatives
with covariant derivatives and by adding terms which contain products of Fμν
with powers of h and h† and also covariant derivatives Dμ h and Dμ h† . In two
dimensions, however, there is no room for such terms in the gauge covariant
current, as these involve μ 1 ...μ D in D dimensions, with one free index and the
others contracted with Fμν s and Dμ s, and in two dimensions they cannot be
constructed. Therefore the covariantized current is given by,
i
Jμ (h, Aμ ) = {[h† Dμ h + hDμ h† ] − εμν [h† Dν h − hDν h† ]}. (9.27)
4π
Knowing the current we deduce the action via, Jμ = δδASμ , getting (9.26) directly.
Finally, we combine the gauged WZW action of the color group manifold, the
WZW of the flavor group manifold and the action term for the gauge fields, to
get the bosonic form of the action of massless QCD2 . The well known fermionic
form of the action is (a mass term will be added later),
< 1 =
SF [Ψ, Aμ ] = d2 x − 2 Tr(Fμν F μν ) − Ψ̄ai [(i∂+ A )Ψi ]a , (9.28)
2ec
where ec is the coupling constant to the color potentials (note it has mass dimen-
sions in 1+1 space-time), and,
where we have also added a mass term with m 2 = mq cm̃. Now since we are
interested in gauging only the SU (NC ) subgroup of U (NF NC ), we take Aμ to
be spanned by the generator T D ⊂ SU (NC ) via Aμ = ec AD D
μ T . We then add
to this action the kinetic term for the gauge fields − 2e 2 d xTr(Fμν F μν ). The
1 2
c
√
coupling ec is related to the color gauge coupling ec by ec = N F ec , so that
after taking the trace over flavor we get the expected kinetic term with coupling
ec . The resulting action is invariant under local color and global flavor,
u → V (x)uV −1 (x), Aμ → V (x)(Aμ − i∂μ )V −1 (x); V (x) ⊂ SUV (NC )
u → W uW −1 ; W ⊂ U (NF ).
The symmetry group is now SUV (NC ) × U (NF ), just as for the gauged fermionic
theory. We choose the gauge A− = 0, so now the action takes the form,
S[u, A+ ] = S[u] + 1
e c 2
d2 xTr(∂− A+ )2 + i
2π d2 xTr(A+ u∂− u† )
+ m 2 Nm̃ d2 xTr(u + u† ). (9.32)
4π
NCNF −i φ
Upon the decomposition u = g̃ h̃le , we see that the current that cou-
ples to A+ is h̃∂− h̃ . In terms of u it is the color projection (u∂− u† )C =
†
† †
N F TrF [u∂− u − N C TrC u∂− u ]. Thus the coupling of the current to the gauge
1 1
i 2 †
field 2π d xTr(A+ h̃∂− h̃ ).
We can further manipulate the action to a form which will be convenient
for taking the strong coupling limit (see Chapter 13). We define H̃(x) by
∂− H̃ = ih̃∂− h̃† . We take the boundary conditions to be H̃(−∞, x− ) = 0 and
then integrate out A+ obtaining,
S̃[u] = S[u] − ( 4π
ec 2
) NF d2 xTr(H̃ 2 )
+ m 2 Nm̃ d2 xTr(u + u† ). (9.33)
In Chapter 13 this form of action will constitute the starting point of determining
the baryonic spectrum of QCD2 in the strong coupling limit. In Chapter (14)
we will use this action to analyze the string tension and the confining behavior
of massive QCD2 .
1 1
p2 p2
i mi k+ +k ik
m2i +2 k+ k i m2i +2 k+ k
j
2g
i
Fig. 10.1. The Feynman rules of QCD2 in the light cone.
Since the vertex is proportional to γ− , only that part of the propagator that is
proportional to γ+ can contribute. As a consequence we can eliminate all the γ
dependence from the Feynman diagrams. Thus the double line, representing the
−ik −
gluon propagator, is p12 , the fermion line is m 2 +2k + k − −i
, and the coupling is 2g.
−
Note that for the gauge field propagator one makes use of the principal value
such that,
1 1 1 1
D+ + (p) = P ≡ + . (10.5)
p2− 2 (p− + i)2 (p− − i)2
The dressed quark propagator and the quark self-energy, given in terms of the
diagrams in Fig. 10.2, obey the coupled equations,
ip−
S(p) =
2p+ p− − m2 − p− Σ(p) + i
dk+ dk− 1
Σ(p) = 4g 2 2
S(p − k)P , (10.6)
(2π) (k− )2
where Σ is the γ+ part, the only part that appears in the self-energy in our gauge.
If we shift the integration variables p+ − k+ → −k+ we eliminate the dependence
on p+ . Hence Σ is only a function of p− . Due to its Lorentz structure it implies
that Σ must be a constant times p1− , namely m2 + p− Σ ≡ M 2 . Thus in the
leading large N the sole effect of the interaction, for the propagator, is to replace
the quark mass m by a renormalized quark mass M .
Integrating over k+ we get,
g2 1 g2
Σ= dk− sgn(p− − k− )P = − , (10.7)
2π (k− )2 πp−
and hence,
g2
M 2 = m2 −
. (10.8)
π
In the original treatment of ’t Hooft, the regularization employed was not of
principal value, but rather of a sharp cutoff, namely integrating over |p− | > λ.
This avoids the infrared divergence as well, but introduces a new scale, which is
not gauge invariant. Obviously, one has to check that Green’s functions of gauge
invariant operators are independent of λ when λ → 0. Thus we find that,
g2 sgn(p) 1
Σ(p) = Σ(p− ) = − − , (10.9)
π λ p−
and correspondingly the dressed quark propagator is,
ip−
S(p) = 2 2 |p | . (10.10)
2p+ p− − m + gπ − g πλ
2 −
+ i
Now the pole of the quark propagator is shifted towards k+ → ∞ and hence
there is no physical single quark state.
Let us consider now the spectrum of the mesonic bound states. The propagator
of the meson is given by the sum of diagrams as is shown in Fig. 10.3.
This ladder sum is exact in the planar limit that follows from the large N
approximation. If the propagator has a meson pole, then the ladder diagrams
have to obey the Bethe–Salpeter equation as in Fig. 10.4.
The “blob” is the Fourier transform of the matrix element,
φ̃(p, q) = F.t. <meson|T ψ̄(x)ψ(0)|0>,
with external legs of a quark of mass m, momentum p, and an anti-quark of mass
m and momentum p − q (for simplicity, we take one flavor, and so the same mass
for the quark and anti-quark). The Bethe–Salpeter equation reads,
d2 k 1
φ̃(p, q) = −4ig S(p − q)S(p)
2
P φ̃(k− , q). (10.11)
(2π)2 (k− − p− )2
Defining
φ(p− , q) = dp+ φ̃(p, q),
q = q
p k q
we get,
g2 1
φ(p− , q) = −i dp+ S(p − q)S(p) dk− P φ(k− , q). (10.12)
π2 (k− − p− )2
The integral over p+ can be done explicitly
I(p− , q) ≡ dp+ S(p − q)S(p)
1 1
=− dp+ 2 −i
. (10.13)
M 2 −i
2(p+ − q+ ) − M
p − −q − 2p + − p−
If p− is outside the interval [0, q− ] then the two poles are on same side of the
real axis, and the integral vanishes. When p− is inside the interval, the integral
is (taking q− > 0),
M2 M2
−iπ 2q+ − − ,
p− (q− − p− )
so that,
M2 M2 g2 q − 1
2q+ − − φ(p− , q) = − dk− P φ(k− , q).
p− q− − p− ) π 0 (k− − p− )2
(10.14)
p− = xq− , k− = yq− ,
and,
2q+ q− = μ2 ,
Since the integral in (10.15) gets its main contribution from y close to x and
since for a periodic function we have,
1 ∞
eiw y eiw y
P dy P dy = −π|w|eiw x , (10.22)
0 (x − y)2
−∞ (x − y)2
then the configurations given in (10.17) are a good approximation of the eigen-
states of the system. The numerical solutions of eqn. (10.15) are drawn in
Fig. 10.5.
In this figure the mass spectrum of mesons is shown for various values of quark
mass. In cases when the mass of the quark and anti-quark are not equal, the term
2
M2 M2 M 22
[ Mx + 1−x ] in (10.15) is replaced by [ x1 + 1−x ].
The masses and wavefunctions cannot be determined in general in an analytic
form. However in certain limits one can write down approximate expressions. In
[52] it was shown that the highly excited states n 1, where n is the excitation
number have masses given by,
3 1
(Mm es )2n ∼ πg 2 N n+ + m2q 1 + m2q 2 ln(n) + C m2q 1 + C m2q 1 + O ,
4 n
(10.23)
μ2
61
11
1.
2.
=
=
m
m
4π2
27
1.
=
m
0
1.
=
m
3π2
56
0.
=
m
18
0.
=
m
2π2
0
=
m
π2
0
0 1 2 n 3
g2
Fig. 10.5. The spectrum of mesons. The squared masses are in units of π
[124].
where mq i are the masses of the quark and anti-quark and where the functions
C(m2q ) are given in [52].
The opposite limit of low-lying states and in particular the ground state can
be deduced in the limit of large quark masses, namely mq g and small quark
masses g mq . For the ground state in the former limit one finds,
Mm0 es ∼
= mq 1 + mq 2 . (10.24)
μ2−nπ2
a
33
π2
23 32
22 31
13
21
12
11
0
b
Fig. 10.6. Meson nonets for Nc = 3. In case (a) the masses of the triplet
are m1 = 0.00, m2 = 0.20, m3 = 0.4 and in (b) m1 = 0.80, m2 = 1.00, m3 = 1.2
[124].
is shown, in units of √gπ . Then the ground state is at 2.7, the first excited state
at 4.16, and level n = 10 is at 20.55. It is obvious from these cases that for larger
n, the wavefunction gets more and more sharply picked around x = 0.5. For the
case of unequal masses, the wavefunction ceases to be symmetric, as can be seen
from Fig. 10.7 for m1 = 1, m2 = 5.
1.5
φ(X)
0.5
0
0 0.2 0.4 0.6 0.8 1
X
q q’ q q’
i i i i
T
where,
φ(q− , q− , p) = dq+ SE (q)SE (q − p)T (q, q , p). (10.29)
Similar to the equation for the “wave function” φ(x) we now get the generaliza-
tion to φ(x, x , p) which reads,
2
M M2
μ2 φ(x, x , p) = + φ(x, x , p)
x 1−x
1
π2 [φ(x, x , p) − φ(y, x , p)]
+
+ dy . (10.30)
N p− (x − x )2
0 (x − y)2
(a)
(b)
(c)
Fig. 10.9. (a) Three-particle vertex function. (b) Hadronic exchange contribu-
tion to two-particle scattering amplitude. (c) Quark exchange contribution to
two-particle scattering amplitude.
order λ1 cancel, leaving a finite remainder. In this way we have verified unitarity
of the model to the first non-trivial order.
where φi , φf 1 , φf 2 are the wave functions of the initial meson and first and second
final mesons, respectively. The quark ends up being in the second final meson and
the anti-quark in the first final meson. The vertex function Φ(x), with x not ∈
[0, 1], is related to the wave function as,
1
1
Φ(x) = dy φ(y). (10.34)
0 (x − y)2
The kinematic parameter w takes the values,
μ2i + μ2f 1 − μ2f 2 ∓ (μ2i + μ2f 1 − μ2f 2 )2 − 4μ2i μ2f 2
w± = , (10.35)
2μ2i
where w+ and w− correspond to the right and left moving final state f1 . The
decay can take place only provided μi ≥ μf 1 + μf 2 . It is clear that for fixed g 2 N
the amplitude is of order A ∼ O( √1N ). The amplitude (10.33) is for a partial
decay and for full-on shell amplitude one has to add the partial decays
A = (1 − (−1)σ i +σ f 1 +σ f 2 )(A(i, f1 , f2 ; w+ ) + A(i, f1 , f2 ; w− )), (10.36)
with σ+ for even parity state and σ− for odd parity state. It was found that
numerically these amplitudes for various excited states do not vanish. This also
shows that the model is not integrable.
It was further found that the amplitudes for mesons made out of massless quark
anti-quark pairs differ significantly from those of mesons made out of massive
ones. An interesting result that follows from the computations of these ampli-
tudes is that the amplitude for decay of an exited meson into a pion and another
meson vanishes, in the case of massless quarks. This is actually to be expected,
as for massless quarks the two-dimensional pion is massless and decoupled, since
there is no chiral symmetry breaking in two dimensions.
11.1 Introduction
In this chapter we study the mesonic spectrum of various QCD2 theories. The
main idea is to use the current algebra of the underlying ungauged theories. In
addition we combine the bosonization techniques developed in Chapter 6 with
that of a large N expansion of Chapter 7 and a light-front quantization as in
Chapter 10. We will focus our attention on the massive mesonic spectrum of con-
formal field theories coupled to non-abelian gauge fields. In particular massless
multi-flavor fundamental quarks and adjoint quarks that will be shown to
correspond to the particular case of Nf = Nc .
First a universality theorem, that states that the massive mesonic spectrum
does not depend on the representation of the matter field but rather only on its
ALA level, will be derived, following Kutasov and Schwimmer [148].
We then present a detailed determination of the massive mesonic spectrum
using a ’t Hooft-like equation for the wave functions of “currentballs” states. We
will discuss in particular the special cases of Nf = 1, Nf = Nc and Nf Nc . The
last section is devoted to the spectrum of states built by the action of a single
current creation operator on the adjoint vacuum. In both cases it will be shown
that the bosonization approach leads to the introduction of current quanta as
the basic degrees of freedom. Once the mass operator P + P − = M 2 is expressed
in terms of the current quanta, the bosonization has already left the scene.
The main content of this chapter, the mesonic spectrum from current algebra,
is based on [17].1 The spectrum based on the adjoint vacuum was introduced
in [3].
affine Lie algebra, are candidates for coupling to non-abelian gauge fields of the
group G. A particular family of such theories are the WZW models, invariant
under G × G of level k. We have discussed in Chapter 6 the gauging of such
models. In this chapter we would like to address the issue of the spectrum of
such gauged conformal field theories, and in particular the massive sector of the
spectrum. In general the Lagrangian density of such a theory reads,
1 2
L = LC F T − 2 Tr Fμν + LI
2e
1
= LC F T − Tr (∂− A+ )2 + Tr [A+ J− ]
2e2
e2 1 e2 1
= LC F T − Tr J + 2 J + = LC F T − Tr J 2 J , (11.1)
2 ∂− 2 ∂
where we have used the light-cone gauge A− = 0. We will be using the notation
of J for J− and J¯ for J+ , and similar for other holomorphic and anti-holomorphic
quantities.
A conformal field theory invariant under the symmetry generated by a G
ALA has holomorphic currents J a in the adjoint representation of G, as well
as anti-holomorphic currents also in the adjoint representation of G. In general
the holomorphic currents obey an ALA with level k and the anti-holomorphic
currents an ALA of level k̄. However, gauging the conformal theory requires
vanishing of the chiral anomaly, namely it requires that,
k = k̄. (11.2)
Next we quantize the system on the light-front. This framework is very conve-
nient since both momenta P − and P + , or equivalently P and P̄ , can be expressed
in terms of J only (with no reference to J).¯ This decoupling of one sector (the
anti-holomorphic one) can be attributed to the fact that in a frame moving to the
right with the speed of light there is no way to interact with massless left-moving
particles. The light-cone Hamiltonian is given by,
1
+
P = dx− : J a (x− )J a (x− )
[C(G) + k]
∞
1 a a
= J J , (11.3)
n=1
n2 −n n
where in the last line we have assumed that the light-cone space direction x− = z
has been put on a circle. Thus the Hamiltonian acts inside current blocks, and
the problem of finding the massive spectrum splits into diagonalizing the decou-
pled blocks of P + on global G singlets. We want to emphasize again that the
light-front dynamics is fully independent of the anti-holomorphic sector, apart
from the constraint that k = k̄. This clearly means that we can replace the
anti-holomorphic sector with another anti-holomorphic sector, provided that
the latter has a level that equals k. Obviously we could have fixed the oppo-
site gauge A+ = 0, leaving only the anti-holomorphic sector with currents J. ¯ In
that gauge we could have replaced the holomorphic sector with another one,
again provided that it has level k. Thus we conclude that the massive spectrum
does not depend on the representations r and r̄, but only on the gauge group G
and the level k.
We would like to demonstrate this universality in the context of a gen-
eralization of Schwinger’s model, which contains nR right moving fermions
ψiR i = 1 . . . nR and nL left-moving fermions ψiL i = 1 . . . nL [120]. Both the right-
and left-moving fermions are charged with respect to an abelian U (1) gauge
symmetry with charges qiR and qiL respectively. The system is described by the
Lagrangian density,
†
¯ R + ψ L † ∂ψ L + ĀJ − 1 (∂ Ā)2 ,
L = ψiR ∂ψ (11.4)
i i i
4e2
n R †
where J = i qiR ψiR ψiR and we are using the gauge A = 0. Upon integrating
Ā we get,
†
¯ R + ψ L † ∂ψ L − e2 J 1 J.
L = ψiR ∂ψ (11.5)
i i i
∂2
We can now bosonize the system. Note that the fermions at hand are not Dirac
fermions but rather nR right and nL left chiral fermions. The system is consistent
in the sense that there is no chiral anomaly when,
R L
N
N
k ≡
R
qiR k ≡
L
qiL k R = k L = k. (11.6)
i= 1 i= 1
One can use the prescription for chiral bosonization described in Section 6.4.
In fact it is enough to note that the interaction term takes the form,
1
Lint = −e2 J J = e2 (φ)2 , (11.7)
∂2
N R N L L L
where φ = i qiR φR i = i qi φi and φL and φL are the right and left chiral
bosons that corresponds to the right and left chiral fermions. Thus we conclude
that the spectrum includes one massive mode corresponding to φ plus nR − 1 and
nL − 1 massless right- and left-moving particles, respectively. It is now evident
that indeed in accordance with the universality theorem, the massive sector
does not depend on the explicit sequence of charges qiR and qiL but only on the
combination expressed in φ.
Another example of the universality theorem is the case of adjoint fermions.
The ALA associated with the currents built from the adjoint fermions J ab =
ψ ac ψ cb is of level Nc . The CFT based on a WZW model of SU (Nc ) of level
k = Nc is another theory with the same ALA, and hence the massive sector of
the spectrum of these theories should, according to the theorem, be the same.
In the next section we describe the massive spectrum of such models based on a
’t Hooft-like equation for the currents.
Since our basic idea is to solve the system in terms of the “quanta” of the col-
ored currents, it is natural to introduce bosonization descriptions of the various
fields.
(i) As was discussed in Section (9.3.2), the bosonized action of colored-flavored
Dirac fermions in the fundamental representation is expressed in terms of a WZW
action of a group element u ∈ U (Nc × Nf ), with an additional mass term that
couples the color, flavor and baryon number sectors. In the massless case when
the latter term is missing, the action takes the form,
1
S0fund = S(N
W ZW
f )
(g) + S W ZW
(N c ) (h) + d2 x∂μ φ∂ μ φ, (11.11)
2
4π
i φ
where g ∈ SU (Nc ), h ∈ SU (Nf ) and e N c N f ∈ UB (1), with UB (1) denoting
the baryon number symmetry, and the WZW action was given in Section 4.1.
(ii) The current structure of free Majorana fermions in the adjoint repre-
sentation can be recast in terms of a WZW action of level k = Nc , namely
S0adj = S(N
W ZW
c)
(g), where now g is in the adjoint representation of SU (Nc ), so
that it carries a conformal dimension of 12 . Multi-flavor adjoint fermions can be
described as SN W ZW W ZW
(g) + SN (h) where g ∈ SO(Nc2 − 1) and h ∈ SO(Nf ). In
c −1
2
f
the present work we discuss only gauging of SU (Nc ) WZW so the latter model
would not be considered.
Substituting now S0fund or S0adj for S0 the action that corresponds to 11.10
becomes,
e2 1
S = S0 − d2 xJ + 2 J + , (11.12)
2 ∂−
k
where the current J + now reads J + = i 2π g∂− g † , and the level k = Nf and k =
Nc for the multi-flavor fundamental and adjoint cases, respectively.
The light-front quantization scheme is very convenient because the correspond-
ing momenta generators P + and P − can be expressed only in terms of J + . We
would like to emphasize that this holds only for the massless case.
Using the Sugawara construction, the contribution of the colored currents to
the momentum operator P + takes the simple form,
1
+
P = dx− : Jji (x− )Jij (x− ) :, (11.13)
Nc + k
√
where J ≡ πJ + , Nc in the denominator is the second Casimir operator of
the adjoint representation and the level k takes the values mentioned above.
Note that for future purposes we have added the color indices i, j = 1 . . . Nc to
the currents. In the absence of the interaction with the gauge fields the second
momentum operator P − vanishes. For the various QCD2 models it is given by,
e2 1
P− = − dx− : Jji (x− ) 2 Jij (x− ) : . (11.14)
2π ∂−
In order to find the massive spectrum of the model we should diagonalize
the mass operator M 2 = 2P + P − . Our task is therefore to solve the eigenvalue
equation,
2P + P − |ψ = M 2 |ψ . (11.15)
e2 ∞
P− = π 0
dp p12 Jji (−p)Jij (p). (11.16)
Recall that the light-cone currents Jji (p) obey a level k, SU (Nc ) affine Lie algebra,
1 1 1 n
Jik (p), Jln (p ) = kp δin δlk − δik δln δ(p+p )+
Ji (p+p )δlk −Jlk (p+p )δin .
2 N 2
(11.17)
We can now construct the Hilbert space. The vacuum |0, R is defined by the
annihilation property,
The commutator of the second term on the right-hand side of (11.21) with
J b (k − 1) yields,
∞
1 1
dp − if J (−p)J (p − k), J (k − 1)
abc a c b
(11.22)
k p2 (p − k)2
∞
1 1
= Nc dp − (J a (−p)J a (p − 1) − J a (p − k)J a (k − p − 1)).
k p 2 (p − k)2
with,
1
2e2 (Nc Nf ) 2 Φ(l) − Φ(k)
Ψ(k, p, l) = , (11.24)
π p2
and,
e2 1 1 2Nc
Φ̃(k) = (Nf − Nc ) + Φ(k) + Φ(k) (11.25)
π k 1−k
k − 1 k
Φ(p) Φ(p) 1 1
−Nc dp −Nc dp +Nc 2 − dpΦ(p)
0 (p−k)2 k + (p−k)2 k (1−k)2 0
Ignoring the three-current term (see below), we get that Φ(k) obeys the eigen-
value equation,
M2 1 1
Φ(k) = (Nf − Nc ) + Φ(k) (11.26)
2
e /π k 1−k
1 k
Φ(p) 1 1
−Nc P dp + Nc − dp Φ(p).
0 (p − k)2 k 2 (1 − k)2 0
1
We assumed that 0 dp Φ(p) = 0, which we will justify shortly.
For general Nc and Nf , discarding the three-current term is unjustified. How-
1
ever, since the length of Ψ is |Ψ(k, p, l)|∼ e2 (Nc Nf ) 2 , in the limit of large Nc
with fixed e Nc and fixed Nf , or large Nf with fixed e2 Nf and fixed Nc , the
2
with the ordinary ’t Hooft equation, it is useful to integrate eqn. (11.26) with
k
respect to k and rewrite it in terms of ϕ(k) ≡ 0 dp Φ(p), to get,
1
M2 1 1 ϕ(p)
ϕ(k) = (Nf − N c ) + ϕ(k) − N c P dp
e2 /π k 1−k (p − k)2
k 1 0
ϕ(p) ϕ(p)
+ Nf dp 2 + Nf dp . (11.27)
0 p k (1 − p)2
The derivation goes as follows. First, integrating eqn. (11.20) we get ϕ(k) =
−ϕ(1 − k) + const. Then taking ϕ(1) = 0 we get,
1
Now ϕ(1) = 0 implies 0 dkΦ(k) = 0, which was our assumption above. Then,
differentiating (11.27) we do get (11.26), and by the last equation we also get
that there is no extra integration constant.
We would like to comment on the issue of the Hermiticity of the “Hamiltonian”
M 2 . Naively, it seems that M 2 is not Hermitian with respect to the scalar product
1
<ψ|ϕ> = 0 dkψ (k)ϕ(k), since the Hermitian conjugate of (11.27) is,
†
M2 1 1
ϕ(k) = (Nf − Nc )+ ϕ(k)
e2 /π k 1−k
1 k 1
ϕ(p) 1 1
−Nc P dp − Nf 2 dpϕ(p) − Nf dpϕ(p).
0 (p − k)2 k 0 (1 − k)2 k
(11.29)
However, as we shall see in the next subsection, the numerical solution yields
real eigenvalues and eigenfunctions. Therefore, at least on the subspace which is
spanned by the eigenfunctions, namely real functions that are zero at k = 0, 1 and
anti-symmetric with respect to k = 12 , the operator M 2 is Hermitian. Note that
(11.29) is “more regular” than (11.27), as in (11.27) it is ϕ(p)/p2 that appears
in the integration from zero.
Equation (11.27) is similar to the ’t Hooft equation for a massive single flavor
e2 N
large Nc QCD2 , with m2 = π f . It differs from ’t Hooft’s equation by having two
additional terms (the two last terms in (11.27)). It suggests that the dynamics
that governs the lowest state of the multi-flavor model is given, approximately, by
a model of a massive “glueball” with an SU (Nc ) gauge interaction and additional
terms which are proportional to Nf .
Before we present our solution of (11.27) it is important to note that it is only
an approximate solution. We neglected the three-current state with, a priori, no
justification. We shall see, however, that the restriction to the truncated two-
current sector is an excellent approximation for the lowest massive meson.
The value of β is chosen so that the Hamiltonian will not be singular near k → 0
or k → 1. This consideration leads to the equation,
Nf Nf /Nc
−1 − + βπ cot βπ = 0, (11.31)
Nc β+1
as derived from eqn. (11.29). Had we started with (11.27), it would have been
−β replacing β in (11.31), and constrained to β larger than 1.
Upon truncating the infinite sum in (11.30) to a finite sum, the eigenvalue
problem reduces to a diagonalization of a matrix. So, the problem can be
reformulated as,
λNij Aj = Hij Aj , (11.32)
with,
1 2
1 2β +i+j
Nij = dk k − (k(1 − k)) , (11.33)
0 2
and,
1 2
Nf 1 2β +i+j −1
Hij = −1 dk k − (k(1 − k))
Nc 0 2
1 k
Nf 1 β +i 1 1 β +j
− dk k − (k(1 − k)) p− (p(1 − p))
Nc 0 2 k2 0 2
1
Nf 1 1 β +i 1 1 β +j
− dk k − (k(1 − k)) p− (p(1 − p))
Nc 0 2 (1 − k) k
2 2
1 β +i β +j
k − 2 (k(1 − k))
1
p − 2 (p(1 − p))
1
− dkdp (11.34)
0 (k − p) 2
Hence,
B(2β + i + j + 2, 2β + i + j + 2)
Nij = , (11.35)
2(2β + i + j + 1)
and,
Nf B(2β + i + j + 1, 2β + i + j + 1)
Hij = −1
Nc 2(2β + i + j)
Nf B(2β + i + j + 1, 2β + i + j + 1)
−
Nc 2(2β + i + j)(β + j + 1)
β Nf /Nc M2
0.0000 0 5.88
0.0573 0.2 6.91
0.1088 0.4 7.91
0.1552 0.6 8.91
0.1978 0.8 9.89
0.2366 1.0 10.86
0.2725 1.2 11.83
0.3050 1.4 12.77
0.3360 1.6 13.73
0.3645 1.8 14.67
Γ(x)Γ(y)
B(x, y) = . (11.37)
Γ(x + y)
In practice, the process converges rapidly and a 5 × 5 matrix yields the ‘contin-
uum’ results.
N
The lowest eigenvalues of (11.27) as a function of the ratio N fc are listed in
Table 11.1 (see also Fig. 11.1). Note that by β = 0, Nf /Nc = 0 we mean the limit
β → 0, Nf /Nc → 0.
These values are in excellent agreement with recent DLCQ calculations, as
will be given in the next chapter.
The typical error is less than 0.1 %.
An interesting observation is that the eigenvalues depend linearly on Nf ,
Fig. 11.1. The dependence is,
e2 Nc Nf
M2 = 5.88 + 5 . (11.38)
π Nc
We do not have a good understanding of this observation. It is not clear why the
lowest eigenvalue sits on a straight line.
In the following sections we will consider some special cases.
Mass Eigenvalues
16
14
12
Mass
Square
10
6
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Nf /Nc
which is just the ’t Hooft equation for the massless case. Note that (11.39) is
−1
exact, since in the small Nf limit the three-current state is suppressed by Nc 2
with respect to the two-current state and therefore we can neglect it. Note also
that in this eqn. (11.29) looks the same too.
Since the wave function ϕ(k) is anti-symmetric, we will recover only the odd
states in the spectrum of QCD2 (the even states can be recovered by considering
other sectors of the Hilbert space which decouple from the two-current state).
Though eqn. (11.39) is formally the same as the ’t Hooft equation, the interpre-
tation of ϕ(k) should be different. It is the integral of the function Φ(k) which
corresponds to the two-current state, namely to a mixture of 4-fermions and
2-fermions. What is the relation between the states that we find here and the
mesons in ’t Hooft’s model?
In order to answer this question let us expand the currents in terms of fermions.
It is useful to denote the current in double index notation
∞
1 i k
J (k) → Jj (k) =
a i
dq Ψ̄i (q)Ψj (k − q) − δj Ψ̄ (q)Ψk (k − q) . (11.40)
−∞ N c
We do not bother about normal ordering, as no problem for k non zero, and we
have to treat the k = 0 part in a limiting way. The state |Φ can be written as,
1
1
|Φ = dk Φ(k)Jji (−k)Jij (k − 1)|0 (11.41)
2Nc 0
1 ∞
1 1
= dk Φ(k) dq dp Ψ̄i (−q)Ψj (−k + q)− δji Ψ̄k (−q)Ψk (−k + q)
2Nc 0 −∞ N c
1 j
× Ψ̄j (−p)Ψi (k + p − 1) − δ Ψ̄k (−p)Ψk (k + p − 1) |0 .
Nc i
Note that the above expression (11.41) contains creation and annihilation
fermionic operators. Written in terms of creation operators only; (11.41) reads,
1 k 1−k
1
|Φ = dk dq dp Φ(k)Ψ̄i (−q)Ψj (−k + q)Ψ̄j (−p)Ψi (k+p−1)|0
2Nc 0 0 0
1 k 1−k
1
− dk dq dp Φ(k)Ψ̄i (−q)Ψi (−k + q)Ψ̄j (−p)Ψj (k + p − 1)|0
2Nc2 0 0 0
1 k
1
− 1− 2 dk dq Φ(k)Ψ̄i (−q)Ψi (q − 1)|0 . (11.42)
Nc 0 0
The last term in (11.42) corresponds to a meson. It can be written also as,
1 1 1
dq dk Φ(k)Ψ̄ (−q)Ψi (q − 1)|0 = −
i
dq ϕ(q)Ψ̄i (−q)Ψi (q − 1)|0 ,
0 q 0
(11.43)
which is exactly the ’t Hooft meson. We conclude that the two-current state has
an overlap with the ’t Hooft meson and this is why (11.27) reproduces exactly
the (odd part of the) spectrum of the ’t Hooft model.
Large Nf Nc limit
In the limit Nf Nc , with e2 Nf fixed, the truncation to two-current state
should again predict exact results. The reason is that the three-current state
−1
is suppressed by Nf 2 with respect to the two-current state.
In this limit eqn. (11.26) takes the form,
e2 Nf 1 1
M2 = + . (11.44)
π k 1−k
e2 N
It describes a continuum of states with masses above 2m, where m2 = π f .
The interpretation is clear: in this limit the spectrum of the theory reduces to a
single non-interacting meson (or “currentball”) with mass m.
The “adjoint vacuum” is created from the singlet vacuum by applying the
adjoint zero mode, which is taken as the limit → 0 of the product of quark and
anti-quark creation operators, each one at momentum . Hence in our case,
†
|0, R = lim ψ−1
i
()ψ−1,j () |0 , (11.45)
→0
i †
where ψ−1 and ψ−1,j are the creation operators of a quark and anti-quark respec-
tively. We can represent the action of the above adjoint zero mode on the vacuum
by the derivative of a creation current taken at zero momentum. Differentiating
the current with respect to k, and acting on the vacuum we get,
5 ∞ ∞
π d †
Jj i (k) |0 k = 0 − = dp i
dqδ(k + p + q)ψ−1 (p)ψ−1,j (q) |0 k =0 −
2 dk 0 0
5
π i
=− ψ ()ψj† () |0 →0 . (11.46)
2 −1
As the currents are traceless, we have to subtract the trace part for i = j. The
latter can be neglected for large Nc . For any given Nc , results that follow are
also the same after the trace is subtracted.
The adjoint vacuum we have is a bosonic one, constructed from fermion-
antifermion zero modes, and as we show it can be written as the derivative
of the current acting on the singlet vacuum. In the case of adjoint fermions there
is another adjoint vacuum, a fermionic one, obtained by applying the adjoint
fermion zero mode on the singlet vacuum.
As we showed already, (J a ) (0) |0 represents the adjoint zero mode
b† (0)d† (0) |0 (indices suppressed), for any Nf and Nc , so in particular also
for Nf = Nc . But in the latter case the theory is equivalent to that of adjoint
fermions, as follows from the equivalence theorem discussed in Section 11.2. As
also stated there, states built on the adjoint vacuum above, cannot be distin-
guished from those built on the fermionic adjoint vacuum, the latter obtained
by applying the adjoint fermions on the singlet vacuum.
The adjoint bosonic vacuum can also have flavor quantum numbers, when the
fermion has flavor. This does not change our results about the mass of the new
state we have. Our “currentball” will have flavor too in such a case. In our scheme
of bosonization, which is the “product scheme”, especially convenient when the
quarks are massless, the flavor sector is decoupled, and so the flavor multiplets
are given by the action of flavor zero modes, not changing the mass values.
Let us introduce the notation,
5
2 a
Z ≡−
a
(J ) (0).
π
The state we have in mind is,
|k = J b (−k)Z b |0 .
This state is obviously a global color singlet, but in our light-cone gauge A− = 0
it is also a local color singlet, as the appropriate line integral vanishes.
Now,
5
π a 1
J (p), Z b = Nf δ ab δ(p) − if abc (J c ) (p), (11.47)
2 2
and thus, for p > 0,
5
2 abc c
J (p)Z |0 = Z J (p)|0 − i
a b b a
f (J ) (p)|0 = 0.
π
Hence the state Z b |0 is annihilated by all the annihilation currents, and so it is
indeed a colored vacuum.
Using,
+ b
P , J (−k) = kJ b (−k), (11.48)
the continuum and needs to be regularized. This subtlety does not affect our
calculation as we work in the singlet sector only.
Actually, the argument connected with P − acting on singlets should be some-
what sharpened. Let us put the lower limit at , and let it go to zero at the end.
This IR cutoff is similar to the one introduced in the derivation of ’t Hooft’s
model discussed in Section 10. Then J(), when acting on a singlet, would go
like . We have two currents in the integral, so we get 2 . But then we have 1/2
from the denominator, so a finite integrant. But the region for integration is of
order , so indeed the total contribution goes to zero.
Now apply P − on our state,
P − J b (−k)Z b |0 = P − , J b (−k) Z b |0 , (11.49)
Note that we use the fact that annihilation currents do annihilate the colored
vacuum also.
Let us apply the operator M 2 to our one-current state,
relative to 0|Z 1 Z 1 |0 .
The normalization of the second term is more complicated. A lengthy but
straightforward calculation gives,
> k >2
> >
> >
> if kdef
dqΦ(q)J (−q)J (q − k) Z e |0
d f
>
> 0 >
1
2
3 ) ) 4
= Nc Nc2 − 1 Nf kδ(0) 0 )Z 1 Z 1 ) 0 (11.53)
2
k k k −p
Nc
×k dpp(k − p)Φ(p) (Φ(p) − Φ(k − p)) − dpΦ(p) dqqΦ(q) .
0 Nf 0 0
where,
√
2
1 Nf N c3
C1 , C2 = . (11.56)
1 Nc
Nc 2 Nf
R1 + R2 N f
ψL
where Ψ = , with ψL and ψ̄R are Weyl fermions, the trace is over the
ψ̄R
color indices which are not written explicitly. To simplify the analysis we restrict
ourselves to the case of a single flavor.
The corresponding equations of motion take the form,
2 The application of the discrete light-front quantization to two-dimensional QCD was done
in [128] and in [127].
When one substitutes into these expressions the expansion of the fields in anti-
commuting modes, subjected to anti-periodic boundary conditions, one gets,
2π
P+ = n(b†n bn + d†n dn ), (12.7)
L 1 3
n = 2 , 2 ,...
where,
1 n πx − n πx −
ψR (x− ) = √ bn e−i L + d†n ei L . (12.8)
2L n = 1 , 3 ,...
2 2
The creation and annihilation operators b†n , d†n , bn , dn are all taken to be in the
fundamental representation of SU (N ) and obey the usual algebra,
where we introduce the dimensionless coupling λ̂. The rationale behind this
parameterization is that the spectrum and wave function depend, apart from
an overall mass scale, only on the ratio of mg . The Hamiltonian H is decomposed
into a free kinetic term H0 and the potential V ,
where,
1 †
H0 = (b bn + d†n dn ), (12.13)
n n
n = 12 , 32 ,...
and,
∞
1 a 1
V = J (k) 2 J a (−k), k=0 (12.14)
π k
k =−∞
λ̂2 CF In †
V =: V : + (b bn + d†n dn ), (12.16)
π n n
n = 12 , 32 ,...
n +1/2
tion, and In is the self-induced inertia In = − 2n
1
+ m =1 m12 . The self-induced
inertia terms cancel the infrared singularity in the interaction term in the con-
tinuum limit. : V : involves a sum of eight quartic terms in the fermionic creation
and annihilation operators. For instance one such term is,
1 c1 c3 1 c4 c3
− N δc 2 δc 4 − δcc41 δcc23 δn −n +n −n ,0 b† n 4 b† n 3 bn 2 ,c 2 bn 1 ,c 1 , (12.17)
4N (n4 − n2 )2 4 2 3 1
where ci are the color indices that have been suppressed before and there is an
implicit summation over the half integers ni such that the momentum is con-
served. Now since P − and P + (or H and K) commute they can be diagonalized
simultaneously. One fixes the value of K = 1, 2, 3 . . . and the corresponding Fock
space is finite dimensional. One then diagonalizes H in the restricted subspace
of gauge singlets such that the masses are given by,
2m2
M 2 = 2P + P − = KH(K). (12.18)
1 − λ̂2
Notice that the dependence of the invariant masses on L the size of the space
drops out.
In Fig. 12.1 the DLCQ spectrum of low-lying mesons is drawn as a function
of m/g for N = 2,3,4 and compared with the t’ Hooft large N calculation. A
comparison with lattice calculation is presented in Fig. 12.2.
In performing these calculations it was found that, except for very small quark
masses, there is a quick convergence of the numerics. This is a manifestation of
the fact that the lowest Fock states dominate the hadronic state. It was found
out that typically the momentum carried by sea quarks is less than one percent.
SU(2)
SU(3)
SU(4)
Hamer: SU (2) Lattice
M/g
m/g
Fig. 12.1. Comparison of the DLCQ meson spectra for N = 2,3,4 and the
spectrum derived from lattice calculations [127].
LARGE N
SU(4)
SU(3)
SU(2)
(2π/N)1/2 M/g
Fig. 12.2. Comparison of the DLCQ meson spectra for N = 2,3,4 and the
’t Hooft large N spectrum [127].
Several further properties were extracted from the DLCQ spectrum and wave
functions:
r The scaling of the lightest mesonic and baryonic masses with N . It was found
that there is fair agreement with the result deduced from bosonization for small
m
g , namely, that
Mm eson π
= 2 sin . (12.19)
Mbaryon 2(2N − 1)
The results “measured” were found to be 1, .62(5), .46(4) for N = 2, 3, 4 com-
paring with bosonization result 1, .618, .445. In the large N limit this result
implies that the baryon mass is proportional to N times the mass of the meson.
r The mesonic form factors were shown to be in accordance with analytical work.
r The “deuteron”, a loosely bound state of two nucleons, was shown to be stable,
in QCD2 with two colors and two flavors.
r The “anti Pauli-blocking” effect, for which the sea quarks with the same flavor
as that of the majority of the valence ones, are not suppressed in spite of their
fermionic nature.
3 Two-dimensional QCD with adjoint fermions was analyzed in several papers. Here we follow
[38], [72], [147].
Denoting by Φ the physical states of the system, that obey the zero charge
condition,
dx− J + |Φ> = 0, (12.24)
Next we introduce the mode expansion and transform the expressions from
the configuration space to the space of momenta. In Section 12.2 we have done
this directly in the discretized formalism. Here for completeness we first consider
a continuous momentum and then perform the discretization.
The mode expansion reads,
∞
1 + − + −
−
ψij (x ) = √ dk + [bij (k + )e−ik x + b†ij (k + )eik x ], (12.26)
2 π 0
and the non-trivial part of the algebra of the creation and annihilation operators
is given by,
1 1
{bij (k + ), bk l (q + )} = δ(k + − q + ) δil δj k − δij δk l . (12.27)
2 N
From here on we will omit the + of k + and denote it as k. Plugging the mode
expansion into (12.23) we get,
∞
P+ = dk kb†ij (k)bij (k), (12.28)
0
and
m2 ∞ dk † g2 N ∞
P =−
b (k)bij (k) + dkC(k)b†ij (k)bij (k)
2 0 k ij π 0
g2 N ∞
+ dk1 dk2 dk3 dk4 [A(ki )δ(k1 + k2 − k3 − k4 )b†k j (k3 )b†j i (k4 )bk l (k1 )bli (k2 )
2π 0
+ B(ki )δ(k1 + k2 + k3 − k4 )(b†k j (k4 )bk l (k1 )bli (k2 )bij (k3 )
− b†k j (k1 )b†j l (k2 )b†li (k3 )bk i (k4 ))], (12.29)
where
1 1
A(ki ) = − ,
(k4 − k2 )2 (k1 + k2 )2
1 1
B(ki ) = − ,
(k2 − k3 )2 (k1 + k2 )2
k
k
C(k) = dp . (12.30)
0 (p − k)2
The bosonic and fermionic states of the system take the following form,
∞ P+
2j
|Φb (p )> =
+
dk1 . . . dk2j δ ki − P +
j=1 o i=1
∞ P+
2j
|Φf (p )> =
+
dk1 . . . dk2j +1 δ ki − P +
j=1 o i=1
−j
f2j (k1 , k2 , . . . k2j )N Tr[b (k1 ) . . . b† (k2j +1 )]|0>, (12.32)
†
where the wave functions obey the cyclicity relation due to the fermionic nature
of the creation and annihilation operators,
Unlike the case of fundamental fermions, pairs of adjoint fermions are not sup-
pressed by additional factor of N1 and hence the eigenstates are generated by
applying operators on the vacuum with a mixture of different numbers of cre-
ation operators. This renders the extraction of the spectrum for adjoint fermions
much harder to determine than that of the fundamental ones. These states are
obviously eigenstates of P + . We will have to ensure that they are also eigenstates
of P − . Following the same procedure as for ’t Hooft’s model of Chapter 10 and
of Chapter 11 one derives a set of equations for the wavefunctions fi by applying
(12.25) on the bosonic and fermionic eigenstates which take the form,
m2
M 2 fi (x1 , x2 , . . . xi ) = fi (x1 , x2 , . . . xi )
x1
x 1 +x 2
g2 N
+ dyfi (y, x1 + x2 − y, x3 , . . . xi )
π(x1 + x2 )2 0
g 2 N x 1 +x 2 dy
[fi (x1 , x2 , . . . xi ) − fi (y, x1 + x2 − y, x3 , . . . xi )]
π 0 (xi − y)2
x1 x 1 −y
g2 N 1 1
dy dzfi+ 2 (y, z, x1 − y − z, x2 , . . . xi ) −
π 0 0 (y + z)2 (x1 − y)2
1 1
− ± cyclic, (12.34)
(x1 + x2 )2 (x2 + x3 )2
k+
where xi = P i+ and the last term of the equation stands for cyclic permutations
of (x1 , x2 , . . . , xi ) which for odd i comes with a + sign and for even i with
alternating signs. Similar to what happens in the ’t Hooft model, the equation
does not have an ambiguity once we incorporate a principal value prescription
to the Coulomb double pole since at x1 = y the numerator also vanishes.
At this point we implement the idea of discretizing the light-cone momenta in
the following way,
1
2
K
n
x→ dx → , (12.35)
K 0 K
o dd n > 0
Similar to (12.8), the creation and annihilation operators of (12.27) also take
discretized values, and obviously the Dirac delta function in (12.27) is replaced
by a Kronecker delta function. The eigenvalue problem now reads,
2
g N
2P + P − = K T + m2 V , (12.37)
π
where the mass term is given by,
1 †
V = b (n)bij (n), (12.38)
n
n ij
and,
n −2
T =4 n b†ij (n)bij (n) m
1
(n −m ) 2 +
2
N m {δn 1 +n 2 ,n 3 +n 4
1
(n 4 −n 2 ) 2 − 1
(n 1 +n 2 ) 2 b†k j (n)b†j i (n)bk l (n)bli (n)
+ δn 1 +n 2 +n 3 ,n 4 1
(n 3 +n 2 ) 2 − 1
(n 1 +n 2 ) 2
b†k j (n4 )bk l (n1 )bli (n2 )bij (n3 ) − b†k j (n1 )b†j l (n2 )b†li (n3 )bk i (n4 )}, (12.39)
20
15
M 10
5 10
<N>
1
i
√ Tr[b† (n1 ) . . . b† (ni )]|0> nj = K. (12.40)
N i/2 s j =1
The states are defined by ordered partitions of K into i positive odd integers,
modulo cyclic permutations. If (n1 , n2 , . . . , ni ) is taken into itself by s out of i
possible cyclic permutations, then the corresponding state receives a normaliza-
tion factor √1s . Otherwise s = 1. For even i, however, all partitions of K where
i/s is odd do not give rise to states.
Using the discretized Hamiltonian and the basis of states (12.40) one can
diagonalize the Hamiltonian and compute the spectrum for a range of values
of K and then extrapolate the results to infinite K, the continuum limit. One
can extract certain properties of the spectrum also from the results at a fixed
large K. In particular the dependence of the spectrum on the mass of the adjoint
quark m is also of interest and the special cases of m = 0 and m2 = g 2 N/π where
the model is supersymmetric.
The fermionic spectrum found by diagonalizing the system with K = 25 for
2
the massless case and for m2 = g πN is described in Figs. 12.3 and 12.4 in the
form of the mass of the bound state as a function of the expectation value of
the parton number. The bosonic spectrum using K = 24 for the two masses is
drawn in Figs. 12.5 and 12.6.
15
10
M
5 10
<N>
20
15
M
10
5 10
<N>
ρ(m) ∼ mα eβ m , (12.41)
where ρ(m) is the density and from the data it follows that β ∼ 0.7 g 2πN .
15
10
M
5
5 10 15
<N>
r The mass increases roughly linearly with the average number of partons. Such a
behavior characterizes a system of large N non-relativistic particles connected
into a closed string by harmonic springs.
r For the low-lying states the wave function strongly peaks on states with a
definite number of partons. For instance, for K = 25 the ground state has a
probability of 0.9993 of consisting of 3 partons, and the first excited state has
a probability of 0.99443 of consisting of 5 partons.
r Thus the low-lying states can be well approximated by truncating the diago-
nalization to a single parton number sector. For instance the bosonic ground
state can be derived from a truncation of (12.34) to a two-parton sector which
yields the following equation,
2 2 1 1 2g 2 N 1
φ(x) − φ(y)
M φ(x) = m φ(x) + + dy , (12.42)
x 1−x π 0 (y − x)2
with φ(x) = f2 (x, 1 − x). Note that this equation is the ’t Hooft equation dis-
cussed in Chapter 10 with the replacement of g 2 → 2g 2 . This difference stems
from the fact that unlike for mesons built from fundamental quarks, here there
are two color flux tubes connecting two partons.
r Due to the fermionic statistics φ(x) = −φ(1 − x) half of the states of the ’t
Hooft model including the ground state are now excluded. In particular for
m = 0 the state φ(x) = 1 which associates with a massless bound state is
missing. The absence of a massless ground state even in the limit of m → 0
can be explained huristically as follows. For m = 0 the mass of the states is
measured in units of the coupling constant g and hence the massless limit can
be achieved in the strong coupling limit g → ∞ for which the action takes the
form,
S = d2 xTr[iΨT γ 0 γ μ ∂μ Ψ + Aμ J μ ]. (12.43)
Now the left and right currents Jij± constitute two independent level N affine
Lie algebras for which we have seen in Chapter 3 the corresponding Virasoro
anomaly is,
k N2 − 1 k
c = c0 − (N 2 − 1) = − (N 2 − 1) , (12.44)
k+N 2 k+N
where c0 is the central charge before gauging and k is the ALA level. Since
k = N it is obvious that c = 0 and hence there is no massless bound state. For
fundamental quarks in the same limit we get, by taking k = 1 and c0 = N ,
that c = 1, which means that for this case there is a massless bound state.
We are now going to compute the baryonic spectrum of QCD2 , for which as
it turns out the bosonic formulation is very convenient. The mesonic spectrum
was found earlier, using large N with quark fields as variables in Chapter 10,
as well as using currents as building blocks in Section 11.3. For the baryon
spectrum, however, the large N limit, in terms of fermionic fields, is not the
natural framework to use since in such a picture the baryon is a bound state of
a large number N of constituents. Instead, it will be shown in this chapter that
the bosonized version of QCD2 in the strong coupling limit provides an effective
description of the baryons.1 We will start by deriving the effective action at
the strong coupling limit. It will be argued that for the purpose of extracting
the low-lying baryons, one can in fact use the product scheme instead of the
U (Nc × Nc ) scheme, with the former being more suitable for our purposes. Once
the effective action is written down we will search for soliton solutions that carry
a baryon number. It will be shown that for a static configuration the effective
action reduces to a sum of sine-Gordon actions. Using the knowledge acquired
on solitons, in Chapter 5, it will be easy to write down the classical baryonic
configuration. We will then semi-classically quantize these solitons. This problem
will be mapped into a quantum mechanical model on a CP (N f −1) manifold.
The energy and charges of the quantized soliton can be derived and thus the
spectrum of the baryons is determined. We then analyze the quark flavor content
of the baryons and discuss multi-baryon states. Finally, we include meson-baryon
scattering, this time also for the case of any coupling.
1 The spectrum of baryons of two-dimensional QCD extracted in the strong coupling limit was
derived in [75].
Our starting point is the last equation of Chapter 9. In the strong coupling limit
ec
mq → ∞, the fields in h̃ which contribute to H̃ will become infinitely heavy. The
sector g̃l ⊂ S US U(N(NF NC )C ) , however, will not acquire mass from the gauge interaction
term. Since we are interested only in the light particles we can, in the strong
coupling limit, ignore the √ heavy fields, if we first normal order the heavy fields
e c√ N F
at the mass scale μ̃ = 2π
. Using the relation, for a given operator O,
Δ
μ̃
Nμ̃ O = Nm̃ O, (13.1)
m̃
to perform the change in the scale of normal ordering, and then substituting
hab = δba , we get for the low energy effective action,
1
Seff [u] = S[g̃] + S[l] + d2 x∂μ φ∂ μ φ
2
−i N 4 π φ +i N 4 π φ † †
+ cmq μ̃Nμ̃ d2 xTr(e CNF g̃l + e CNF l g̃ ). (13.2)
We can now replace the two mass scales mq and μ̃ by a single scale, by normal
ordering at a certain m so the final form of the effective action becomes,
1
Seff [u] = S[g̃] + S[l] + d2 x∂μ φ∂ μ φ
2
m2 2 −i N 4 π
NF φ +i N 4 π
CNF
φ ˜† †
+ Nm d xTr(e C g̃l + e l g ), (13.3)
NC
with m given by,
√ ΔC
1 + 1Δ
C
ec NF
m = NC cmq √ , (13.4)
2π
N C2 −1
here ΔC , the dimension of h̃, is N C (N C +N F ) . For the l = 1 sector, defining g =
4π
−i φ
g̃e NCNF
⊂ U (NF ) one gets the effective action,
Seff [g ] = NC S[g ] + m2 Nm d2 xTrF (g + g † ). (13.5)
Thus, the low energy effective action in the l = 1 sector coincides with the result
of the “naive” approach of the product scheme.
In the strong coupling limit ec /mq → ∞ the low energy effective action reads,2
S[g] = NC S[g] + m2 Nm d2 x(Trg + Trg † ), (13.6)
with g in U (NF ). Note that the analog of our strong coupling to the case of 3+1
space-time, would be that of light current quarks compared to the QCD scale
ΛQ C D .
and for g that has only spatial dependence δW Z = 0. Without loss of generality
we may take, for the lowest energy, a diagonal g(x),
4π 4π
−i NC ϕ1 −i NC ϕN F
g(x) = e ,...,e . (13.8)
For this ansatz and with a redefinition of the constant term, the action density
reduces to,
5
NF
1 dϕi
2
4π
S̃d [g] = − dx − 2m cos
2
ϕi − 1 . (13.9)
i= 1
2 dx NC
these charges are determined solely by the boundary values of ϕ(x), which are,
5 5
4π 4π
ϕ(∞) = 2π, ϕ(−∞) = 0. (13.15)
NC NC
Under a general UV (NF ) global transformation g◦ (x) → g̃◦ (x) = Ag◦ (x)A−1
the energy of the soliton is obviously unchanged, but charges other than QB and
QY will be turned on. Let us introduce a parametrization of A that will be useful
later,
⎛ ⎞
z1
⎜ .. ⎟
⎜ Aij . ⎟
⎜ ⎟
A=⎜ ⎜ . ⎟. (13.16)
.. ⎟
⎜ ⎟
⎝ z(N F −1) ⎠
Y1 . . . . . . Y(N F −1) zN F
Now,
4π
−i ϕ
g̃◦ = 1 + (e NC
− 1)z, (13.17)
N F
where (z)α β = zα zβ∗ , and from unitarity α =i zα zα∗ = 1. The charges with g̃◦ (x)
are,
1
(Q̃◦ )A = NC Tr(T A z). (13.18)
2
Only the baryon number is unchanged. The discussion of the possible U (NF )
representations cannot be done yet, since we are dealing so far with a classical
system. We will return to the question of possible representations after quantizing
the system.
and to derive the effective action for A(t).3 Quantization of this action cor-
responds to doing the functional integral over g(x, t) of the above form. The
effective action for A(t) is derived by substituting g(x, t) = A(t)g◦ (x)A−1 (t) in
the original action. Here we use the following property of the WZ action,
S AgB −1 = S AB −1 + S g, Ãμ , (13.20)
where S[g] is the WZW action and S[g,Ã] is given by (9.25), respectively, with
the gauge field Ãμ given as,
iÃ+ = A−1 ∂+ A, iÃ− = B −1 ∂− B; A, B ∈ U (NF ). (13.21)
Using the above formula for A = B, noting that S(1) = 0, and taking A = A(t),
Ȧ
∂+ A = ∂ − A = √ , (13.22)
2
we get,
< =
NC
S̃ A(t)g◦ (x)A−1 (t) − S̃[g◦ ] = d2 xTr [A−1 Ȧ, g◦ ][A−1 Ȧ, g◦† ]
8π
NC < =
2 −1 †
+ d xTr (A Ȧ)(g◦ ∂1 g◦ ) . (13.23)
2π
This action is invariant under global U (NF ) transformations A → U A, where
U ∈ G = U (NF ). This corresponds to the invariance of the original action under
g → U gU −1 . On top of this it is also invariant under the local changes A(t) →
A(t)V (t), where V (t) ∈ H = SU (NF − 1) × UB (1) × UY (1), with the last two
U (1) factors corresponding to baryon number and hypercharge, respectively. This
subgroup H of G is nothing but the invariance group of g◦ (x). In terms of g◦ (x)
and A(t) the charges associated with the global U (NF ) symmetry, eqn. (13.13),
have the form,
<
NC =
B
Q =i dxTr T B A g◦† ∂1 g◦ − g◦ ∂1 g◦† + g◦ , A−1 Ȧ, g◦† A−1 .
8π
(13.24)
The effective action, eqn. (13.23), is an action for the coordinates describing the
coset space,
G/H = SU (NF ) × UB (1)/SU (NF − 1) × UY (1) × UB (1)
= SU (NF )/SU (NF − 1) × UY (1) = CP N · (13.25)
To see this explicitly we define the Lie algebra valued variables q A through
A−1 Ȧ = i T A q̇ A . In terms of these variables (13.23) takes the form (the part
that depends on q A ),
⎡ ? ⎤
2(N F −1)
1 2(N − 1)
Sq = dt ⎣ q̇ Y ⎦
F
(q̇ A )2 − NC
2M NF
A=1
∞
5 √
1 NC 4π 2 NC 3/2
= (1 − cos ϕ)dx = ( ) . (13.26)
2M 2π −∞ NC m π
The sum is over those q A which correspond to the G/H generators and q Y
is associated with the hypercharge generator. Although the q A seem to be a
“natural” choice of variables for the action (13.23), which depends only on the
combination A−1 Ȧ, they are not a convenient choice of variables. The reason for
that is the explicit dependence of the charges (13.24) on A−1 (t) and A(t) as well
as on A−1 Ȧ(t).
physical degrees of freedom. This is exactly the dimension of the coset space
S U (N F )
S U (N F −1)×U (1) . The corresponding phase space should have a real dimension
of 4(NF − 1). Naively, however, we have a phase space of 4NF dimensions and,
therefore, we expect four constraints.
There are several methods of quantizing systems with constraints. Here we
choose to eliminate the redundancy in the z variables and then invoke the canon-
ical quantization procedure.4
But before following these lines let us briefly describe another method, through
the use of Dirac’s brackets. We outline the classical case. The quantum case is
obtained by replacing { , } with i[ , ].
The first step in this prescription is to add to the Lagrangian a term of
the form λ( α zα zα∗ − 1), in which case the conjugate momentum πλ of the
Lagrange multiplier vanishes. By requiring that this condition be preserved in
time one gets the secondary constraint Φ1 = ( α zα zα∗ − 1) = 0. Further impos-
ing Φ̇1 = {Φ1 , H}P = 0, where { }P denotes a Poisson bracket, one finds another
second-class constraint Φ2 = Π · z + z † · Π† . In addition there is a first-class con-
straint Φ3 = Π · z − z † · Π† , which corresponds to the local U (1) invariance of the
model. Fixing this symmetry one gets an additional constraint Φ4 . For instance
∗
one can choose the unitary gauge Φ4 = zN F − zN F
. The next step is to compute
the constraint matrix {Φi , Φj }P = cij . In the constrained theory, the brack-
ets between F and G are replaced by the Dirac brackets of those operators,
given by
{F, G}D = {F, G}P − {F, Φi }P (c−1
ij ){Φj , G}P , (13.35)
where c−1 ij is the inverse of the constraint matrix. Imposing the constraints as
operator relations it is easy to see that zN F , ΠN F and their complex conjugates
can be eliminated. The brackets for the rest of the fields coincide with the results
we derive below, when eliminating the constraints explicitly.
We now describe in some detail the quantization of the system using uncon-
strained variables. We want to choose a set of new variables so that the constraint
N F ∗
α = 1 zα zα = 1 is automatically fulfilled. There is a standard choice of such vari-
ables, namely (f or i = 1, . . . , NF − 1),
ki ki∗ eiχ
zi = √ , zi∗ = √ , zN F = √
1+X 1+X 1+X
F −1
N
X= ki∗ ki . (13.36)
i= 1
The ki , ki∗ and χ are 2NF − 1 real variables with no constraints on them. The
phase space will now have dimension 2(2NF − 1) and we still have two extra
4 The quantization of the system including its constraint was done in [75]. For an alternative
procedure of quantization in the presence of constraints see [181].
5 Quark solitons as constituents of hadrons were discussed in [86]. Following that, the flavor
content of the baryons was discussed in [97].
is only the isospin 32 representation. This is what we would expect from naı̈ve
quark model considerations. The total wave function must be antisymmetric.
Baryon is a color singlet, so the wave function is antisymmetric in color and
it must be symmetric in all other degrees of freedom. There is no spin, so the
baryon must be in a totally symmetric representation of the flavor group, a 10
for three flavors. Therefore, strictly speaking there is no state analogous to the
proton. On the other hand, there is a state which is the analog of the Δ+ , namely
the charge 1 state in the 10 representation, z12 z2 .
The 10 is the lowest baryon multiplet in QCD2 . In the following we shall be
dealing with the relative weight of a given flavor in some baryon state. Thus,
q̄q B will henceforth stand for the ratio,
q̄q B
3 4
¯ + s̄s .
ūu + dd
(13.64)
B
as well as,
1 3 4 1
ūu = , ¯
dd = . (13.66)
Δ+ Δ+
2 3
In evaluating the integral in the numerator in eqn. (13.65) we have used |z3 |2 =
1 − |z1 |2 − |z2 |2 , which follows from the unitarity of the matrix A in (13.19).
Similarly, for Δ+ + ∼ z13 we have,
2 3 4
ūu Δ + + = , ¯ + + = 1 , s̄s + + = 1 .
dd (13.67)
Δ Δ
3 6 6
In the constituent quark picture Δ+ + contains just three u quarks. Both the
d-quark and the s-quark content of the Δ++ come only 3 4from virtual quark
¯ + + , and s̄s + =
pairs. Therefore in the SU (3)-symmetric case s̄s Δ + + = dd Δ Δ
s̄s Δ + + , as expected.
From eqn. (13.67) one can also read the results for Ω− ∼ z33 , by replacing
u ↔ s. In the general case of NF flavors and NC colors, one obtains,
1
(q̄q)sea B = , (13.68)
NC + N F
where (q̄q)sea refers to the non-valence quarks in the baryon B. Moreover, one can
also compute flavor content of valence quarks. Consider a baryon B containing
k quarks of flavor v. The v-flavor content of such a baryon is,
k+1
v̄v B = . (13.69)
NC + N F
This implies an “equipartition” for valence and sea, each with a content of
1/(NC + NF ). It also follows that the total sea content of NF flavors is,
NF
NF
(q̄q)sea B = , (13.70)
q=1
NC + N F
13.6 Multibaryons
Let us now explore the possibility of having multi-baryons states.6 The pro-
cedure follows similar lines to that of the baryonic spectrum, namely, we look
for classical solution of the equation of motions with baryon number kNC , and
then we semiclassically quantize this. The ansatz for the classical solution of the
This means that the Nz number is equal to (kNC ). Thus for any wave function,
written as a polynomial in z and z , the number of zs minus the number of z s
must equal (kNC ). Note that for k = 1 the transformation (13.79) represents also
the NFth flavor number. Thus (13.80) entails that the representation contains a
state with NC boxes of the NF flavor, and therefore must be the totally symmetric
representation.
Now, the effective action (13.76) is invariant under a larger group of local
transformations. In fact, we have extra (k 2 − 1) generators, which correspond to
SU (k) under which (13.76) is locally invariant. This can be exhibited by defining
“local gauge potentials”,
so that,
Dz = ż + z Ã. (13.82)
which implies,
ψ̃(z, z ) = ψk (z)[
(ziα zj α )n i j ], (13.86)
{i,j }
computation in the strong coupling limit [98] and then we discuss the general
case of any coupling [87].
Our starting point is the soliton solution that describes the static classical
baryon gc (x) = exp[−iΦc (x)] where,
⎛ ⎞
φc (x)
⎜ 0 ⎟
⎜ ⎟
⎜ . ⎟
⎜
Φc (x) = ⎜ ⎟, (13.94)
. ⎟
⎜ ⎟
⎝ . ⎠
0
and,
5
μx 8π
φc (x) = 4arctg (e ) , μ=m . (13.95)
NC
Note that we have shifted the non-trivial phase factor to the upper left-hand
corner, whereas in (13.77) it was put in the lower right-hand one.
We introduce a fluctuation around it of the form,
1
−iΦ c (x)
g≈e −i dτ e−iτ Φ c (x) δφ(x, t)e−i(1−τ )Φ c (x) . (13.97)
0
where we have denoted by δ̃φ the new variation, different from the δφ of eqn.
(13.97), but still a fluctuation about the classical solution. Now,
Nc
∂+ e−iΦ c (x) ∂− δ̃φ(x, t) eiΦ c (x)
4π
+ m2 e−iΦ c (x) δ̃φ(x, t) + δ̃φ(x, t)eiΦ c (x) = 0. (13.99)
Obviously the two expressions coincide in the abelian case. In fact, the relation
between δφ and δ̃φ is
1
δ̃φ(x, t) = dτ eiτ Φ c (x) δφ(x, t) e−iτ Φ c (x) . (13.100)
0
Physical quantities should obviously come out to be the same for both types of
fluctuation.
with,
ω 2 = k 2 + μ2 , (13.110)
the two asymptotic solutions are,
χab (x) ∼ A(ω) sin kx + B(ω) cos kx, (13.111)
and the S-matrix is,
1
Sforward = (B − iA) (13.112)
2
1
Sbackward = (B + iA), (13.113)
2
for an incoming wave eik x from x = −∞.
Fig. 13.1. The phase shift δ = 2ctg−1 (k/μ), eqn. (13.117), as a function of the
normalized momentum k/μ, for the potential (13.105), governing the small
fluctuations around the soliton in the abelian case. The phase shift is smooth
and monotonically decreasing with momentum, indicating that no resonance
is present. Note logarithmic momentum scale.
We can now proceed to derive the scattering matrix, using the standard
procedure. The solution for x → ∞ contains only the transmitted wave,
ψ(x → ∞) ∼ eik x .7
It turns out that for the particular potential (13.105) there is no reflection at
all, i.e. the wave function for x → −∞ contains only the incoming wave,
1 + ik/μ
ψ(x → −∞) ∼ eik x−δ = − eik x . (13.114)
1 − ik/μ
Thus,
1 1 + ik/μ
=− (13.115)
T 1 − ik/μ
1 − ik/μ
T =− = eiδ (13.116)
1 + ik/μ
1 k
ctg δ = . (13.117)
2 μ
As shown in Fig. 13.1, δ varies smoothly and decreases monotonically from δ = π
at k = 0 to δ = 0 at k = ∞, indicating that there is no resonance.
The no-reflection potential we found is a special case of a well-known class of
reflectionless in quantum mechanics.
7 We take the convention where the scattering phase is taken to be zero at x → ∞ and is
therefore extracted from the wave function at x → −∞.
resulting in,
uj (x) − iφc (x)uj (x) + ω 2 + ωφc (x) − 12 μ2 1 + eiφ c (x) uj (x) = 0. (13.122)
Defining,
i
u j ≡ e 2 φ c vj , (13.123)
we find,
vj + ω 2 + ωφc − 12 μ2 (1 + cos φc ) + (φc )
1 2
4 vj = 0. (13.124)
Using,
(φc ) = μ2 (1 − cos φc ) ,
1 2
2 (13.125)
we get,
vj + ω 2 + ωφc − μ2 cos φc vj = 0. (13.126)
where,
Fig. 13.2. The normalized potential V (ω/μ; x)/μ2 of eqn. (13.128), for ω/μ =
1.01 (upper), 2 (middle) and 3 (lower).
with ω = k 2 + μ2 , as before. Note that the potential depends on the momen-
tum of the incoming particle, as shown in Fig. 13.2.
Next we proceed to solve numerically for the reflection and transmission coef-
ficient. It turns out that for numerical solution of the scattering problem it is
more convenient to take the coefficient of the outgoing wave at x ∼ +∞ to be
1, instead of the T prefactor, and integrate eqn. (13.127) backward, reading off
the T and R amplitudes from the solution at x ∼ −∞.
We thus use,
vj (x) = eik x , x → +∞
−ik x
(13.129)
vj (x) = 1
T e ik x
+ R
T e , x → −∞.
Defining,
1
δ± = (δS ±δA ) , (13.131)
2
we find that,
T = eiδ + cos δ−
(13.132)
R = ieiδ + sin δ− .
Fig. 13.3. Scattering by the potential eqn. (13.128) as a function of the nor-
malized energy ω/μ. Upper plot: transmission probability |T |2 ; lower plot:
phase of T , δ+ (continuous line). Also shown is the approximate result for δ+
from WKB (dot-dashed line).
Note that R/T is purely imaginary. The transmission and reflections probabilities
are,
|T |2 = cos2 δ−
(13.133)
|R|2 = sin2 δ− .
The numerical results for the transmission probability |T |2 and for the phase of
T , δ+ are presented in Fig. 13.3. For comparison and as an extra check we also
plot the WKB result for δ+ . Note that no resonance appears.
Note that the asymptotic value of the phase shift is π. This can also be obtained
from a WKB calculation, which becomes exact at infinite energies.
M ≡ u∂− u† =
= exp(−iΦc ) ∂− (exp iΦc ) + exp(−iΦc ) exp(−iδΦ) [∂− exp(iδΦ)] exp(iΦc ),
(13.138)
and obtain the equations of motion for the meson field by varying with respect
to δΦ. The variation of (13.135) with respect to δΦ is proportional to,
δMc −2
∂ Mc . (13.139)
δ(δΦ)
To compute its variation with respect to δΦ, we need only the second term M2
of M , as the first term M1 is independent of δΦ.
We take for the soliton a diagonal ansatz (13.94) now in the form of a u matrix
rather then a g one,
√
[exp(−iΦc )]aa j j = δaa δj j exp (−i 4πχ α j ) :
a = 1, . . . , Nc , (13.140)
j = 1, . . . , Nf ,
so that,
The part of M that contributes to the effective action is its color projection
(13.136). We note that Trf &c M2 = 0, and thus,
√ √
[(M2 )c ]a,a = exp(−i 4πχaj )[exp(−iδΦ) ∂− exp(iδΦ)]aj,a j exp(i 4πχa j ).
j
(13.142)
The mesons δΦ have to be diagonal in color, so,
[(M2 )c ]a,a = [ exp(−iδΦ) ∂− exp(iδΦ) ]aj,aj δa,a . (13.143)
j
We recall that the flavor structure of the mesons is independent of their color
indices, and restrict our attention to mesons that have no U (1) flavor part. In
this way, we may be sure that classical solutions lead to stable particles, since
their non-vanishing flavor quantum numbers put them in a different sector from
the vacuum. We then have,
[ exp(−iδΦ) ∂− exp(iδΦ) ]α j,α j = 0, (13.144)
j
scale ambiguity in m, since, when the masses are different, there is a question of
which normal-ordering scale to use. The resulting equation of motion for δΦ is,
Thus δΦaj j is a free field with squared mass given by the average of m2j and
m2j in this case, which we do not discuss further.
r The second case is that of j = j , with a such that (Φc ) is a quark soliton
aj
inside the baryon. In this case,
r The third case is when j is different from j , now with one of the Φc being a
soliton and the other vanishing. Taking (Φc )aj to be the soliton, we obtain,
1
δΦaj j − i(∂+ Φc )aj (∂− δΦ)aj j + {μ2 j + μ2 j [exp(iΦc )]aj }δΦaj j = 0,
2
(13.152)
where j = j and (Φc )α j = 0.
Next we want to proceed and evaluate the meson-baryon scattering. For that
purpose we need to analyze the equations that determine the static solution
(Φc )aj . First one defines,
√
(Φc )aj = 4π(χc )aj , (13.153)
where the (χc )aj are canonical fields, whose equations of motion are,
⎛ ⎞
1 √ √
χα j − 4αc ⎝ χα − χ β l ⎠ − 2 4πm2j sin 4πχ α j = 0.
Nc
l βl
Note the extra factor 2 in front of the mass term, as compared with eqn. (22)
of [86], due to an error in this reference.
Choosing the boundary conditions χaj (−∞) = 0, we get as constraints for
χaj (+∞), denoted hereafter simply by χaj ,
1
√ χ α j = nα j integers, (13.154)
π
and
nα = n independent of a. (13.155)
l
The baryon number8 associated with any given flavor l is given by,
Bl = nα .
a
we find,
ωj2 = k 2 + μ2j , (13.158)
and the equation for uj (x) is,
uj (x) + ωj 2 uj − μ2 j [cos (Φc )j ]uj = 0. (13.159)
We define the potential Vj for this scattering process via,
uj (x) + ωj 2 uj − Vj uj = 0, (13.160)
and find,
Vj = μ2 j [cos (Φc )j ]. (13.161)
In our normalization the outgoing wave has coefficient 1, which is more conve-
nient for numerical calculations, and the wave for x → −∞ is now,
1 ik x Rj −ik x
uj (x) = e + e , x → −∞, (13.162)
Tj Tj
in this case.
In the second non-trivial case (13.152), we put,
δΦj j = e−iω j j t uj j (x), (13.163)
so that,
uj j (x) − i(Φc )j (x)uj j (x) + {ωj2j + ωj j (Φc )j (x)
1
− {μ2 j + μ2 j [exp(iΦc )]j }}uj j = 0. (13.164)
2
To eliminate the first derivative term in u, we substitute,
i
uj j = exp Φc vj j . (13.165)
2 j
Taking again,
vj j (x) −→ eik x , (13.169)
x→∞
we get,
1 2
ωj j = (μ j + μ2 j ), (13.170)
2
and the wave for x → −∞ is,
1 ik x Rj j −ik x
vj j (x) = e + e , x → −∞, (13.171)
Tj j Tj j
in this case.
To summarize we have shown that meson-baryon scattering in QCD2 in the
large-Nc limit is non-trivial for non-zero quark masses, and is described by two
distinct effective potentials when the quark masses are unequal. These effective
potentials are not of the sine-Gordon type found in previous cases, and we expect
the scattering amplitudes also to be non-trivial. Their calculation will require
numerical analysis.
V = σL. (14.1)
The coefficient in this linear dependence is the string tension. Thus, a non-
confining behavior, which will be referred to as a screening behavior, implies
a vanishing string tension. Whereas in four dimensions the computation of the
string tension is a formidable task, in two dimensions, as will be shown in this
chapter, it is a fairly easy one. In this chapter we describe the extraction of the
string tension in various two-dimensional gauge systems.
We start by calculating the string tension for the massive Schwinger model in
both the fermionic and the bosonic languages. This is done in the small mass
limit and then we discuss the corrections due to going beyond this limit. We then
discuss the short range corrections to the confining potential. We focus on the
abelian case, believing that the non-abelian case is very similar. Next we com-
ment on the behavior of the string tension when finite temperature is introduced.
Then we move to non-abelian generalization. We compute the string tension for
the cases of matter in the fundamental and adjoint representations, followed by
the symmetric and anti-symmetric representations.
Much of this chapter is based on [15] and [16].
The string tension of the massive Schwinger model was calculated using
bosonizaton in [68]. The massless cases in gauge theories were analyzed in [116].
The next-to-leading order in small mass was computed by [4].
Z= (14.2)
1 2
DAμ DΨ̄DΨ exp i d2 x − 2 Fμν + Ψ̄i∂Ψ − mΨ̄Ψ − qdyn Aμ Ψ̄γ μ Ψ ,
4e
q L q
where qdyn is the charge of the dynamical fermions. Gauge fixing terms were not
written explicitly. Let us add an external pair with charges ±qext at ±L, namely,
so that the change of L is −jμext Aμ (x). Note that by choosing jμext which is
conserved, ∂ μ jμext = 0, the action including the coupling to the external current
is also gauge invariant.
Now, one can eliminate this charge by performing a local, space-dependent
left-handed rotation,
1
Ψ → eiα (x) 2 (1−γ 5 ) Ψ (14.4)
−iα (x) 12 (1+γ 5 )
Ψ̄ → Ψ̄e , (14.5)
where |0>0 is the vacuum state with no external sources. The change in the
vacuum energy is due to the mass term. The change in the kinetic term which
appears in (14.6) does not contribute to the vacuum energy.
Thus,
qext qext
σ = m cos 2π <Ψ̄Ψ> −m sin 2π <Ψ̄iγ5 Ψ> −m <Ψ̄Ψ>0 .
qdyn qdyn
(14.13)
The values of the condensates <Ψ̄Ψ> and <Ψ̄γ5 Ψ> are needed. The easiest
way to compute these condensates is bosonization, but it can also be computed
directly in the fermionic language. We state here the final result for the m = 0
case (the derivation can be found in the references of this chapter),
exp(γ)
<Ψ̄Ψ>m = 0 = −e (14.14)
2π3/2
<Ψ̄γ5 Ψ>m = 0 = 0. (14.15)
Equation (14.15) is due to parity invariance (with our choice θ = 0). The
resulting string tension, to first order in m,
exp(γ) qext
σ = me 1 − cos 2π . (14.16)
2π3/2 qdyn
Though this expression is only the leading term in a m/e expansion and might
be corrected, when qext is an integer multiple of qdyn the string tension is exactly
zero, since in this case the rotated action (14.10) is not changed from the original
one (14.2).
The value of the first coefficient is C1 = 1 and the next was found to be
C2 = −8.9 exp(γ
8π1 / 2
)
. Higher coefficients are not calculated yet.
Note that for finite me we have to minimize the potential,
√ qext 1
V = M2 1 − cos 2 πφ + 2π + μ2φ φ2 . (14.23)
qdyn 2
where k is the Fourier transform of the space coordinate. We will discuss the
validity of our approximation for φ later in this section. The last equation can
be rewritten as,
m21 1 m21 1
A0 (k) = + 1 − e2 jext (k), (14.28)
m22 k 2 m22 k2 + m22
where,
The first term is the confining potential which exists whenever the quark mass
is non-zero. On top of this, there is always a screening potential.
The string tension which results from the above potential is,
2
qext
σ = mμ × 2π2 2 , (14.34)
qdyn
form is,
e2
φ1 (x) = √ qdyn qext (θ(x + L) − θ(x − L)), (14.36)
πm22
m
which for small e reduces to,
√ qext
φ1 (x) ∼ π (θ(x + L) − θ(x − L)). (14.37)
qdyn
√
Thus 2 πφ small means,
qext
(2π) 1, (14.38)
qdyn
This result indicates that the string is not torn even at very high temperatures.
The explicit expression shows that <Ψ̄Ψ>T is non-zero for all T . Thus, the system
does not undergo a phase transition. It is just energetically favorable to have the
electron-positron pair confined.
The action (14.43) differs from (14.42) by a factor of one half in front of the
WZW and interaction terms, because g is real and represents Majorana fermions.
Another difference is that g now is an (N 2 − 1) × (N 2 − 1) orthogonal matrix.
The two actions (14.42) and (14.43) can be schematically represented by one
action,
1
S = S0 + mμR d2 x tr g + g † (14.44)
2
ikdyn a
− d2 x g∂− g † Aa+ ,
4π
where A− = 0 gauge was used, S0 stands for the WZW action and the kinetic
action of the gauge field, kdyn is the level (the chiral anomaly) of the dynamical
charges (k = 1 for the fundamental representation of SU (N ) and k = N for the
adjoint representation).
Let us add an external charge to the action. We choose a static charge (with
respect to the light-cone coordinate x+ ) and therefore we can omit its kinetic
term from the action. Thus an external charge coupled to the gauge field would
be represented by,
ikext a
− d2 x u∂− u† Aa+ .
4π
Suppose that we want to put a quark and an anti-quark at a very large separation.
A convenient choice of the charges would be a direction in the algebra in which
the generator has a diagonal form. The simplest choice is a generator of an
SU (2) subalgebra. Since a rotation in the algebra is always possible, the results
are insensitive to this specific choice. As an example we write down the generator
in the case of fundamental and adjoint representations,
⎛ ⎞
1 1
3
Tfund = diag ⎝ , − , 0, 0, . . . , 0⎠
2 2 D EB C
N −2
⎛ ⎞
⎜ 1 1 1 1 1 1 ⎟
3
Tadj = diag ⎜
⎝1, 0, −1, 2 , − 2 , 2 , − 2 , . . . , 2 , − 2 , 0, 0, . . . , 0⎟⎠.
D EB C D EB C
2 (N −2)
2(N −2) dou blets
T 3 = diag(λ1 , λ2 , . . . , λi , . . . , 0, 0, . . .),
where {λi } are the ‘isospin’ components of the representation under the SU (2)
subgroup.
We take the SU (N ) part of u as,
3
u = exp −i4π θ(x− + L) − θ x− − L Text , (14.45)
3
for N > 2, and a similar expression but with a 2π factor for N = 2. Text represents
the ‘3’ generator of the external charge and u is static with respect to the light-
cone time coordinate x+ . The theta function is used as a limit of a smooth
function which interpolates between 0 and 1 over a very short distance. In that
limit u = 1 everywhere except at isolated points, where it is not well defined.
The form of the action (14.44) in the presence of the external source is,
*
1
S = S0 + mμR d2 x tr g + g †
2
+
ikdyn a a
+ − g∂− g † + kext δ a3 δ x− + L − δ x− − L A+ .
4π
The external charge can be eliminated from the action by a transformation of
the matter field. A new field g̃ can be defined as follows,
ikdyn a ikdyn a
− g̃∂− g̃ † = − g∂− g † + kext δ a3 δ(x− + L) − δ(x− − L) .
4π 4π
Since the Haar measure is invariant (and finite, unlike the fermionic case) with
respect to unitary transformations, the form of the action in terms of the new
variable g̃ reads,
ikdyn a
S = SWZW (g̃) + Skinetic (Aμ ) − d2 x g̃∂− g̃ † Aa+
4π
1 i4π
k ext
T3
k
−i4π k e x t T d3y n †
+ mμR d2 x tr g̃e k d y n d y n + e dyn g̃ , (14.48)
2
which is QCD2 with a chiraly rotated mass term.
The string tension can be calculated easily from (14.48). It is simply the vac-
uum expectation value (v.e.v.) of the Hamiltonian density, relative to the v.e.v.
of the Hamiltonian density of the theory without an external source,
We expect that similar corrections as those in eqn. (14.22) will occur also in
non-abelian systems. For the fundamental/adjoint case, the following expression
may correct the leading term,
∞
l−1
m kext
σQ C D = mμR C̃l 1 − cos 4πλj l . (14.50)
ekdyn j
kdyn
l= 1
(i) The string tension (14.49) reduces to the abelian string tension (14.21) when
abelian charges are considered. It follows that the non-abelian generalization
is realized by replacing the charge q with the level k.
(ii) The string tension was calculated in the tree level of the bosonized
action. Perturbation theory (with m as the coupling) may cause changes,
eqn. (14.49), since the loop effects may add O(m2 ) contributions. However,
we believe that it would not change its general character. In fact, one fea-
ture is that the string tension vanishes for any m when kkde yx nt is an integer,
as follows from eqn. (14.48), since the action does not depend then on kext
at all.
(iii) When no dynamical mass is present, the theory exhibits screening. This is
simply because non-abelian charges at the end of the world interval can be
eliminated from the action by a chiral transformation of the matter field.
(iv) When the test charges are in the adjoint representation kext = N , eqn.
(14.49) predicts screening by the fundamental charges (with kdyn = 1).
(v) String tension appears when the test charges are in the fundamental rep-
resentation and the dynamical charges are in the adjoint. The value of the
string tension is
4π 2π
σ = mμadj 2 1 − cos + 4 (N − 2) 1 − cos (14.51)
N N
The case of SU (2) is special. The 4π which appears in eqn. (14.49) is replaced
by 2π, since the bosonized form of the external SU (2) fundamental matter differs
by a factor of a half with respect to the other SU (N ) cases. Hence, the string
tension in this case is 4mμadj .
(vi) We would like to add, that when computing the string tension in the pure
YM case with external sources in representation R, the Wilson loop gives
1 2 1 2 2
2 e C2 (R), while our way of defining external source gives 2 e kext . Thus we
C 2 (R )
need a factor k 2 to bring our result to the Wilson loop case. Analogous
ext
factors should be computed for the other cases, when dynamical matter is
also present.
which includes the arbitrary integer l. What is the value of l that we should pick?
The dynamical charges are attracted to the external charges in such a way
that the total energy of the configuration is minimal. Therefore the value of l
which is needed, is the one that guarantees minimal string tension.
Thus the extended expression for string tension is the following,
@ A
kext
σ = min mμR 1 − cos 4πλi (1 + lN ) . (14.56)
l
i
kdyn
15.1 Introduction
In Chapter 9 we realized that the structure of the bosonized non-abelian massless
QCD2 is that of a gauged WZW model with an additional F 2 term of the gauge
fields. Apart from the pure gauge term, this is therefore a special form of a
two-dimensional coset model, discussed in Section 4.6. This naturally calls for
a treatment of the system similar to that for a coset model. Using the form of
the gauge fields in terms of scalars f and f¯ as A = if −1 ∂f , Ā = if¯∂¯f¯−1 with
f (z, z̄), f¯(z, z̄) ∈ H c , the complexification of H ≡ SU (NC ), leads to a convenient
formulation of the model.1 The main advantage of this approach is that one can
then easily decouple the “matter” and the gauge degrees of freedom.
In this chapter we point out that the F 2 term requires a special treatment.
The formulation of pure YM theory in terms of the f variables seems naively to
contain unexpected “physical” massive color singlet states. This result is obvi-
ously neither in accordance with our ideas of the degrees of freedom of the model
nor with the lattice and continuum solution of the theory. We show that similar
“naive” manipulations in the case of QED2 do reproduce the Schwinger model
results. Using a coupling constant renormalization we show that in the limit of no
matter degrees of freedom the coupling constant is renormalized to zero. In this
case the unexpected states turn into unphysical massless “BRST” exact states.
In the flavored QCD2 case a similar analysis shows the existence of physical
flavorless states of mass m2 = N2πF e2c .
This chapter is based on [96].
where l1 and l2 are in the adjoint representation and Stwist is a twist term given,
Nc ¯ −1 f −1 ∂f ].
Stwist = − d2 zTr[l∂l (15.4)
2π
SG F = SQ C D 2 + S (g f ) + S (g h) = SQ C D 2 + δB R S T (bĀ) =
= SQ C D 2 + TrH [B Ā] + TrH [bD̄c], (15.7)
where SG F , S (g f ) and S (g h) are, respectively, the gauge fixed action, the gauge
fixing term and the ghost action. The (b, c) fields are yet another (1, 0) ghost
system and B is a dimension-one auxiliary field, all in the adjoint representation
of SU (NC ). The integration over B introduces a delta function of the gauge
choice to the measure of the functional integral. In addition we integrate over
the ghosts b and c.
It is interesting to note that the QCD2 action can be related to a “perturbed”
topological HH coset model. To realize this face of QCD2 we parameterize u as
4π
i NCNF φ
ghle and rewrite (15.7) accordingly. The Polyakov–Wiegmann relation
implies,
1
S[u] = S[ghl] + d2 x∂μ φ∂ μ φ,
2
1
S[ghl] = S[g] + S[l] + S[h] + d2 xTr(g † ∂+ gl∂− l† + h† ∂+ hl∂− l† ). (15.8)
2π
Since l is a dimension-zero field with an associated zero central charge we have
S[l] = 0 and thus,
(H ) i ¯
SG F = SN F (h) + S−(N F + 2N C ) (f ) + d2 zTrH [ρ̄∂ χ̄ + ρ∂χ]
2π
1 ¯
+ SN C (g) + d2 z[∂φ∂φ]
2π
m2 2 −1 i N 4πNF φ −i N 4 π N F φ −1 −1 −1
+ d zTrG : f ghle C +e C l h g f :
2π
1 ¯ −1 ∂f ))2 ].
+ 2 d2 zTrH [(∂(f (15.9)
ec
It is now easy to recognize the first line in the action as the action of SS UU (N C)
(N C )
topological theory.
It is interesting to note that a WZW term S−2N C (f ) appears in the action
even without the introduction of quarks. We therefore digress to an analysis of
the pure YM theory in the formulation introduced above.
Here again we remind the reader that the coupling constant undergoes a mul-
tiplicative renormalization. This will be discussed in Section 15.7. Let us first
discuss the corresponding equations of motion for f and f¯,
varying the action with respect to the gauge fields, one gets two derivatives of
F. In deriving the above, it is convenient to remember that,
1 : ;
δSW Z W (f ) = Tr (f −1 δf )∂(f
¯ −1 ∂f ) .
2π
The YM action equation (15.10) is obviously invariant under the original gauge
transformations,
f → f v(z, z̄) f¯ → v −1 (z, z̄)f¯,
with v ∈ SU (NC ). In addition the action is invariant separately under the holo-
morphic and anti-holomorphic “color” transformations,
f → u(z̄)f f¯ → f¯w(z),
where u, w ∈ SU (NC ). These are “spurious” transformations since they leave
A and Ā invariant. The corresponding holomorphic and anti-holomorphic color
currents are,
¯ f¯)−1 −
J¯s = − Nπc [i(f f¯)∂(f 2
f D̄F f −1 ],
m 2A
where is the standard antisymmetric step function. Notice that the massless
degree of freedom has commutation relations which correspond to a negative
metric on the phase space. These relations can also be translated to the following
OPEs (choosing the part ϕ1 (z) of ϕ1 ),
ϕ1 (z)ϕ1 (w) = log(z − w),
ϕ2 (z, z̄)ϕ2 (w, w̄) = − log |(z − w)|2 + O(μ2 |z − w|2 ),
ϕ1 (z)ϕ2 (w, w̄) = O(z − w). (15.17)
It is thus clear that the model is invariant under a U (1) affine Lie algebra of level
k = −1 since JG (z)JG (w) = (z −w 1
) 2 , as JG = ∂ϕ1 with no contribution from ϕ2 .
The physical states of the model have to be in the cohomology of the BRST
charge. Due to the fact that the current is holomorphically (and the other anti-
holomorphically) conserved, it follows that the same property holds for the BRST
charge, and thus the space of physical states is an outer product of the cohomol-
ogy of Q and Q̄. The latter are given by,
Q = χJ = χ(i∂ X̃ + ∂ϕ1 ),
Q̄ = χ̄J¯ = χ̄(−i∂¯X̃ + ∂ϕ
¯ 1 ). (15.18)
Expanding the fields i∂ X̃ and ∂ϕ1 in terms of the Laurent modes X̃n and
(ϕ̃1 )n with [Xn , Xm ] = nδn +m and [(ϕ1 )n , (ϕ1 )m ] = nδn +m we have
Q= χ̃n X̃−n − i(ϕ1 )−n .
n
then, in a complete analogy with the abelian case, JG = − Nπc J1 . The lat-
ter is an affine current of level k = −2NC . One can in fact show that
(15.5) can be assumed without a loss of generality. ∂J ¯ G = 0 implies that
2 ¯ ¯
J2 + m 2 (∂ ∂J2 + i[∂J2 , J1 + J2 ]) = u(z), where u(z) is some holomorphic func-
A
tion. We then introduce the shifted currents J˜2 = J2 − u(z), J˜1 = J1 + u(z). Now
∂¯J˜1 = 0 as does J1 , and J˜2 obeys eqn (15.5) with J˜1 replacing J1 . It is easy to
check that the shifts in the currents correspond to f1 → v(z)f1 , f2 → f1 v(z)−1
with u(z) = if1−1 (v(z)−1 ∂v(z))f1 .
Note that the equation for J2 involves a coupling to J1 . This is related to
the fact that, unlike the abelian case, one cannot write the action as a sum of
decoupled terms which are functions of J1 and J2 separately.
Once the color current JG is expressed in terms of the holomorphic current
J1 , the analysis of the space of physical states is directly related to that of the
topological GG model at k = 0. The physical states have to be in the cohomologyof
the BRST charge, which corresponds to the following holomorphically conserved
BRST current,
a
1 NC a 2 ¯ i a b c
Q(z) = χa JGa + Jgah = − χ A + 2 D(∂A) + fbc ρ χ .
2 π mA 2
An anti-holomorphic BRST current Q̄(z) determines the condition for physical
states in the analogous manner to Q. From here on we restrict our description
to the latter. We define now the zero level affine Lie algebra current,
a
J(tot) = JGa + J(g
a a a
h) = JG + i, fbc χb ρc ,
where J(tot) in , L̃n and Wn are the Laurent modes of J(tot) i the Cartan sub-algebra
currents, T and W , respectively. In fact the BRST cohomology of the present
model is a special case of the set of G/G models.
We therefore refer the reader to those works [9], [200], [229] and present here
only the result. On the plane where no ghost zero modes are allowed, the only
state in the cohomology is the zero ghost number vacuum state of J1 .
This state can be a tensor-product with oscillators of the massive modes of J2 .
Unlike the abelian case, JG does not commute with J2 so that in general the J2
modes are not obviously in the BRST cohomology. However, there is no reason
to believe that all the J2 modes will be excluded by the BRST condition. Those
J2 modes that remain are by definition color singlets.
This result contradicts previous results on Y M2 . Usually one believes that
pure gluodynamics on the plane is an empty theory since all local degrees of
freedom can be gauged away.
Z= DADĀDBeiS (A , Ā ,B ) ,
S = − d2 zT rH [ e1c F B + 14 B 2 ], (15.21)
It is thus clear that in the present formulation there is no trace of the massive
“physical modes” discussed in the previous section.
where k is an arbitrary level and Γk (L) = Sk (w) for L = iw∂w ¯ −1 . The renor-
e(−k −2N C ) k
malization factor e(k) has to satisfy e(k ) = k +2N C . In addition it is clear
from eqn. (15.24) that it has
to be singular at the origin. It can be shown that
e(k) takes the form e(k) = k + 2Nk
C
. Implementing this renormalization in our
case, eqn. (15.23) takes the form,
It is now clear that the action vanishes at NF = 0 and hence again, on triv-
ial topology, the theory is empty. Notice, however, that the implementation of
renormalization modifies also the result of the previous section.
The final conclusion is that in both methods one finds that indeed the pure
YM theory has an empty space of physical states as of course is implied by the
original formulation in terms of A. We have demonstrated that in this formulation
it follows only after taking subtleties of renormalization into account.
(15.7) or by eqn. (15.9). In the past the low-lying baryonic spectrum in the strong
m
coupling limit e cq → 0 was extracted using a semi-classical quantization. In this
chapter our analysis was based on switching off the mass term, mq = 0. This
limit cannot be treated by the semi-classical approach, as the soliton solution
is not there for mq = 0. In our case here one finds a decoupled W ZW action
for the flavor degrees of freedom SN C (g) and a decoupled free field action for
the baryon degree of freedom, in addition to the action of the colored degrees of
freedom which is given in eqn. (15.26) or eqn. (15.7). The general structure of a
physical state in this case is that of a tensor product of g and φ with the colored
degrees of freedom f , h and the ghosts. The structure of QCD2 which emerged
from the semi-classical quantization for mq = 0 involves g and φ only. In our case
here the f colored degrees of freedom acquire mass mA = ec N2πF while the h
degrees of freedom remain massless. In the limit ec → ∞ the f degrees of freedom
decouple. It is thus clear that one has to introduce the mass term which couples
the three sectors. The massless limit of QCD2 can then be derived by taking the
limit mq → 0 after solving for the physical states. Indeed, it was shown in the
limit of ec → ∞ that turning on mq = 0 results in a hadronic spectrum where
the flavor representation and the baryon number were correlated. The analysis
of the spectrum of the massive multi-flavor QCD2 in the approach of this work
remains to be worked out.
16.1 Introduction
Pure gauge theory in two dimensions is locally trivial and has no propagating
degrees of freedom. This was discussed in the first chapter of this part of the
book and now in the last chapter we will describe the global properties of gauge
theories in two dimensions. For the latter to be non-trivial we will either take
the underlying manifold to be a compact Riemann surface or introduce Wilson
loops external sources. We will show that in those cases the gauge theory has a
rich structure and in fact is almost a topological field theory, which is a theory
with no propagating degrees of freedom (see also Section 4.7). Moreover, it will
be shown that the theory has an interpretation in terms of a string theory.
It is easy to realize that in two dimensions the pure YM theory is in fact the
simplest member of a wide class of renormalizable theories that incorporate only
gauge fields. These will be referred to as the generalized gauge theories gYM.
In Chapter 15 we introduced an alternative formulation of the YM theory using
the action1
S = − d2 zTr[iFB + g 2 B 2 ]. (16.1)
1 In fact in (16.1) we have used a slightly different formulation which is however equivalent to
the one used in Chapter 15.
2 The notion of the generalized QCD theory in two dimensions was introduced and analyzed
in [82].
3 A lattice version of two-dimensional Yang–Mills theory was shown to be exactly solvable by
Migdal in [161]. Correlators of Wilson lines on this formulation were computed in [139].
Wilson loops in the gYM 2 case. Since this chapter is based on non-trivial two-
dimensional topology which has not been dealt with intensively in this book,
to fully understand its content the reader will need to consult the references to
this chapter. The reader who is not interested in the topological aspects of two-
dimensional gauge dynamics may skip this chapter and proceed directly to the
third part of the book.
This chapter is based mainly on [115], [118], [119] and [105].
4 The partition function on any Riemann surface of the discretized theory was written down
in [184]. An identical result was found also in the continuum formulation [228].
U2
U3
U1
− g12 tr(U +U † )
action rather than the Wilson action (which is Z [U ] = e ), i.e.
Z [U ] = dR χR (U )e−tc 2 (R ) , (16.5)
R
1. As t goes to zero (and, therefore, for finite g also a goes to zero) we want the
holonomy to be close to 1,
g ab = tr(ta tb ), (16.9)
we see that (16.5) is the correct answer up to terms of the order of O(t3/2 ) which
drop in the continuum limit. Using (16.5) as the starting point, we finally find
the following form of the partition function,
λ A c 2 (R )
2−2H − 2 N
Z(N, H, λA) = dR e . (16.12)
R
To get from (16.5) to (16.12) we take the following steps. First we make use of
the additivity property of the heat-kernel action. Consider two triangles glued
along U1 as depicted in Fig. 16.2.
U3 U4
U1
U2 U5
Fig. 16.2. Integrating over U1 on a link which is common to the two triangles.
U1
V 2+ V1
U 2+ U +1
V2 V 1+
U2
we find,
dU1 Z1 (U2 U3 U1 )Z2 (U1† U4 U5 ) = Z1 +2 (U2 U3 U4 U5 ). (16.14)
This relation can be used to argue that the lattice representation is exact and
independent of the triangulation since using this we can add as many trian-
gles as desired, thus reaching the continuum limit. We can also use this rela-
tion to reduce the number of triangles to the minimum needed to capture
the topology of M. Describing a genus H manifold in term of a 4H-polygon
with identified sides as described in Fig. 16.2 for a genus-two Reimann surface
a1 b1 a−1 −1 −1 −1
1 b1 . . . aH bH aH bH . The partition function on such a manifold can be
written as,
λ A c 2 (R )
ZM = dR e − 2N
DUl DVl χR [U1 V1 U1† V1† . . . UH VH UH† V1† ]. (16.15)
R
We can simplify this expression using again the orthogonality of the characters
and the relation,
1
DU χR [AU BU † ] = χR [A]χR [B], (16.16)
dimR
to arrive at (16.12).
where F = F μν with ij being the antisymmetric tensor 12 = −21 = 1. B is
μν
(16.21)
5 1
In principle, we could perturb the ordinary Y M 2 with operators of the form g 2 k −2
tr(F k ),
without the need for an auxiliary field.
Note that the index of C{·} will always pertain to a partition. Thus C{p} =
C{r 1 +r 2 +···+r j } even if p = r1 + r2 + · · · + rj . The brackets in the t-s mean a
total symmetrization ( (r1 + r2 + · · · + rj )! terms).
Cρ can easily be seen to commute with all the group elements and so, by
Schur’s lemma, is a constant matrix in every irreducible representation.
We claim that the correct lattice generalization of (16.5) is,
dR χR (U )e−tΛ(R ) , (16.22)
R
where,
Λ(R) = a{k i } C{k 1 ·1+k 2 ·2+k 3 ·3+···} (R). (16.23)
{k i }
where the path-ordered product around the closed curve γi is taken in the rep-
resentation Ri . Using loop equations, one can derive an algorithm to compute
Wilson loops on the plane. This can be further generalized into a prescription for
c1 = 6 c2 = 6 c3 = 4 c4 = 2 c5 = 2 c6 = 1 c7 = 1 c8 = 1
n1 = 8
n2 = 8
n3 = 3
n4 = 3
n5 = 2
n6 = 2
1 1 Nw
−λ A i C 2 ( R i )
W (R1 , γ1 , . . . Rn γn ) = .... DR 1 ...R n d2−2G i
e 2N ,
ZN n
i=1
Ri
R1 Rn
(16.28)
where Nw is the number of windows, 2 − 2Gi is the Euler number associated with
the window i and DR 1 ...R n is the product of the Wigner coefficients for neigh-
boring windows. For the case of intersecting loops a set of differential equations
provides a recursion relation by relating the average of a loop with n intersections
to those of loops with m < n intersections.
Generalizing these results to the gYM 2 is straightforward. The only alteration
−λ A C 2 ( R )
that has to be invoked is to replace the e 2 N factors that show up in those
algorithms with similar factors where the second Casimir operator is replaced
by the generalized Casimir operator (16.23). For instance the expectation value
of a simple Wilson loop on the plane is given by,
−λ A γ Λ ( R )
W (R, γ) = e 2N , (16.29)
It is interesting to note that for odd Casimir operators the expectation values
of real representations (like the adjoint representation) equal unity due to the
fact that the corresponding Casimirs vanish.
6 The stringy description of Yang–Mills theory in two-dimensional Riemann surfaces was intro-
duced by D. Gross and W. Taylor in [115], [118] and [119]. The formulation of the two-
dimensional Yang–Mills theory in terms of topological string theories was done in [126] and
[69].
MH
Mh
r
with n = i= 1 ni . For every representation R there is a conjugate representa-
tion R̄ whose Young tableau Y (R̄) has its rows and columns interchanged. To
determine the Casimir operator of the conjugate representation we use (16.30)
with 2P̂{2} (R) = −2P̂{2} (R)U (N ) (R).
Using the Frobenius relations between representations of the symmetric group
Sn and representations of SU (N ) (U (N )), the coefficients of this asymptotic
expansion were written in terms of characters of Sn . The latter can be shown
to correspond to permutations of the sheets covering the target space. The final
result takes the form of,
∞ (n + + n − )(2 H −2 )+ (i + + i − )
1
Z(A, H, N ) ∼
N
n ± , i ± = 0 p ± , . . . , p ± ∈T 2 ⊂S ± s ± , t ± , . . . , s ± , t ± ∈S ±
1 i± n 1 1 H H n
+ −
(−1)(i +i ) − − − 2 + −
(λA)(i +i ) e− 2 (n +n )λA e 2 ((n ) +(n ) −2n n )λA /N
+ 1 + 1 + 2 2
+ − + −
i !i !n !n !
H H
+ − − 2−2H
1 · · · pi + p1 · · · pi − Ωn + ,n −
δS n + ×S n − p+ [s+ +
j , tj ] [s− −
k , tk ] , (16.31)
j =1 k =1
where [s, t] = sts−1 t−1 . Here δ is the delta function on the group algebra of
the product of symmetric groups Sn + × Sn − , T2 is the class of elements of Sn ±
consisting of transpositions, and Ω−1 n + ,n − are certain elements of the group algebra
of the symmetric group Sn + × Sn − .
The formula (16.31) nearly factorizes, splitting into a sum over n+ , i+ , · · · and
n− , i− , · · ·. The contributions of the (+) and (−) sums were interpreted as arising
from two “sectors” of a hypothetical worldsheet theory. These sectors correspond
to orientation reversing and preserving maps, respectively. One views the n+ = 0
and n− = 0 terms as leading order terms in a 1/N expansion. At higher orders
the two sectors are coupled via the n+ n− term in the exponential and via terms
in Ωn + n − .
Thus, the conventional Y M2 theory has an interpretation in terms of sums of
covering maps of the target space, see Fig. 16.5. Those maps are weighted by the
factor of N 2−2h e− 2 n λA where h is the genus of the world-sheet and A is the area
1
of the target space. The power of N 2−2h is obtained from the Riemann–Hurwitz
formula,
where P̂2 (R) is the value of the scalar matrix representing the sum of transposi-
tions i< j ≤n (ij) in the representation R of Sn (the matrix commutes with all
permutations and thus is scalar). In the partition function, C2 (R) was multiplied
λA
by 2N . The resulting term 12 nλA arises from the action and is proportional to
the string tension. The term λAP̂{2} arises from the measure and is interpreted
as the contribution of branch points to the weight of a map.
Our task is, therefore, to express the generalized Casimirs Cρ of (16.21) in
terms of P̂ρ , the generalizations of P̂{2} (R). This is expressed as,
ρ
Cρ (R) = αρρ N h ρ P̂ρ (R), (16.34)
ρ
where αρρ are coefficients that are independent of R and the power factors hρρ ,
are adjusted so that a string picture is achieved.
The P̂{ρ } factors are associated with ρ which is an arbitrary partition of
certain numbers, namely,
k1 k2
B CD E B CD E
ρ : ki · i = 1 + 1 + · · · + 1 + 2 + 2 + · · · + 2 + · · · , (16.35)
i
P̂{ρ } (R) is the product of two factors. The first is the sum of all the permutations
in Sn (n is the number of boxes of R) which are in the equivalence class that
is characterized by having ki cycles of length i for i ≥ 2. Just like the case of
P̂{2} (R), the matrix P̂{ρ } (R) commutes with all permutations and thus is a
scalar. The sum is taken in the representation R of Sn . The second factor is,
n − i=2 iki
, (16.36)
k1
which can be interpreted later as the number of ways to put k1 marked points
on the remaining sheets that do not participate in the branch points.
16.7 Examples
A complete diagrammatic expansion of the operators was determined in [105].
Using this expansion one can write down the stringy description of the partition
function for any generalized YM theory. We end this chapter with a few examples
of the Casimir factors for various choices of Φ(B) for both U (N ) and SU (N )
groups.
λ
1. For N tr(B 2 ) which is the conventional Y M2 theory we get (16.33),
2λ
P̂{2} + λP̂{1} . (16.37)
N
The first term means that we give a factor of 2λA
N for each branch point, and
the second term means that we have a factor of λ for each marked point (i.e.
this is the string tension).
2. For αN −2 tr(B 3 ) in U (N ) we get,
1 1
3αN −2 P̂{3} + 3αN −1 P̂{2} + 3αN −2 P̂{1+1} + αP̂{1} + αN −2 P̂{1} . (16.38)
2 2
The first term is the contribution from branch points of degree 2 (the simple
branch points are of degree 1). The next term is a modification to the weight
of the usual branch points. The third is the weight of two marked points at
the same point (but different sheets), which will translate into n+ (n+ − 1)
+ n− (n− − 1) in the weight of a map for which (n+ , n− ) are the numbers
of sheets of each orientability. The last two terms are modifications to the
6 12 12
− P̂{2+ 1} − 3 P̂{2} + 4 P̂{1+1+1}
N3 N N
6 12 3 2
− 2
− 4 P̂{1+ 1} − 2
− 4 P̂{1} . (16.39)
N N N N
These terms and the terms in the previous example (16.38) do not mix chi-
ralities (i.e. sheets of opposite orientations). In the full theory (chiral and
anti-chiral sectors) there is the corresponding anti-chiral term:
6 12 12
+ 3
P̂{2̄+ 1̄} + 3 P̂{2̄} − 4 P̂{1̄+ 1̄+ 1̄}
N N N
6 12 3 2
+ − 4 P̂{1̄+ 1̄} + − 4 P̂{1̄} . (16.40)
N2 N N2 N
For SU (N ) there are additional terms that do mix chiralities. They are,
6 12
− 3
(P̂{2̄+ 1} − P̂{2+ 1̄} ) + 4 (P̂{1̄+ 1̄+1} − P̂{1̄+ 1̄+1} ). (16.41)
N N
The first term is the contribution of maps that have a branch point in one
orientability and a marked point in the other (at the same target space point).
The second term is the contribution of maps with three marked points – two
for one orientability and one for the other.
To illustrate the content of these formulae in terms of representations, we
will calculate the value of the third Casimir 16 dabc t(a tb tc) for a totally anti-
symmetric representation of SU (N ) with k boxes. The term P̂{3} is translated
into the sum of all the permutations of the k indices of a totally antisymmetric
tensor that are 3-cycles, this gives 13 k(k − 1)(k − 2). The term P̂{2} gives the
sum of all the permutations that are 2-cycles, that is − 21 k(k − 1) (a minus
16.8 Summary
In this chapter we studied the generalized two-dimensional Yang–Mills theory
on Riemann surfaces. We reviewed the exact formulae for the partition function
and Wilson loop averages of the conventional YM theory. We then presented
the generalization of these results in the context of the generalized YM theories.
These expressions are based on a replacement of the second Casimir operator
with more general Casimir operators depending on the particular model. There
is another method to obtain these results [228], i.e. by regarding the general
Yang–Mills actions as perturbations of the topological theory at zero area.
Using the relations between SU (N ) representations and representations of the
symmetric groups Sn , we wrote down the generalizations that have to be made
in the Gross–Taylor string rules for 2D Yang–Mills theory, so as to make the
generalized Yang–Mills theory for SU (N ) or U (N ) a local string theory as well.
The extra terms are special weights for certain maps with branch points of a
degree higher than one.
An obvious extension of the results presented in this chapter is to consider
other gauge groups. The conventional YM 2 theory with gauge groups O(N ) or
Sp(N ) which were shown to be related to maps from non-orientable world-sheets.
One can further couple the gYM 2 theories to fermionic matter in analogy to
’t Hooft’s analysis presented in Chapter 10. This domain of research is far from
being fully explored. A particularly interesting question is to find out certain
Φ(B)s that lead to a special behaviour of the coupled system. For example,
in the U (1) case, the representations R are labeled by an integer n and for
Φ(B) = −α log(1 + λB 2 ) we get,
Z(U (1), A) = (1 + λn2 )−α A
n
in the study of four dimensional QCD which is conformal only at the classical
level. We analyze the non-local operators built from a quark and an anti-quark
and expand them in terms of Gegenbauer polynomials. We use the COPE to write
down the operator product of two electromagnetic currents. Finally we deter-
mine, in the limit of large momentum exchange, the pion distribution amplitude.
Conformal invariance in four dimensions was described in several review papers
and books. The original studies on conformal symmetry are summarized in [207]
and [66]. A modern review that we follow in this chapter is [43].
where μ0 , μν ˜μ0 are vector, antisymmetric tensor, scalar and vector infinites-
0 , 0 ,
imal constants, respectively. The corresponding finite transformations are,
xμ → xμ = xμ + aμ
xμ → xμ = aμν xν
xμ → xμ = axμ
xμ + ãμ x2
xμ → xμ = , (17.6)
1 + 2ãμ xμ + ã2 x2
where the various forms of a are the finite parameters of transformation that
correspond to the infinitesimal ones above. The last transformation is referred
to as the special conformal transformation. It can in fact be decomposed into
μ
an inversion transformation xμ → − xx 2 , a space-time shift transformation and
another inversion. The sum of all these transformations is d + d(d−1)
2 +1+d=
(d+ 1)(d+ 2)
2 , which is the dimensions of SO(2, d), the algebra of the conformal
group in Minkowski space-time.
Let us analyze now the generators of the conformal transformations and show
that indeed they obey the SO(2, d) algebra. The generators are,
[Pμ , Pν ] = 0
[Pμ , Lν ρ ] = i(ημν Pρ − ημρ Pν )
[Lμν , Lρσ ] = −i(ημρ Lν σ − ημσ Lν ρ + ην σ Lμρ − ην ρ Lμσ )
[D, Pμ ] = −iPμ [D, Kμ ] = iKμ
[Pμ , Kν ] = 2i(Lμν − ημν D)
[D, Lμν ] = 0
[Kμ , Kν ] = 0, (17.8)
which is indeed the SO(2, 4) algebra. The first three lines constitute the Poincare
algebra in four dimensions. It is well known that (17.8) is not the most general
form of the SO(2, 4) algebra. One can further generalize the construction of the
generators by modifying Lμν in the following way,
where Σμν does not act on the space-time points and obeys,
Shortly the role of these generators in the conformal transformation of fields will
be discussed.
is composed of two parts, the one due to that of the space-time point with
δxμ given in (17.5) and an “internal transformation” δI Φ(x), which vanishes
for space-time translations, while for Lorentz transformations, dilatations and
special conformal transformations, takes the form,
where l is the conformal dimension of the field and the internal Lorentz generators
Σμν are given by,
i μ ν
Σμν = [γ , γ ] − Dirac spinors [Σμν ]βα = η μβ δαν − η ν β δαμ − gauge f ields
4
(17.13)
Recall that all the parameters of transformations are global, namely space-time
independent. To determine the Noether currents associated with the various
symmetry transformations, one elevates the transformations into local ones and
reads the currents from the variation of the action,
δS = d4 xJμa ∂ μ ea (17.14)
∂L
where Πμ = ∂ (∂ μ Φ) . In fact the variation of the action with respect to dilatations
μ
may lead in addition to the divergence of J (D ) , to another total derivative
term ΔD . However for Lagrangians that are polynomials in the fields and their
derivatives, this term vanishes. For the special conformal transformations the
additional term is defined by
δK S = d4 xeμ (x) [−∂ν K μν + 2xμ ΔD + Δμk ] , (17.16)
with Δμk = 2Πν Φ(lg μν + Σμν )Φ. For invariance we need ΔD = 0 and Δμk =
μ
2∂ν σ μν . For l = 1 and l = 3/2, σμν vanishes, so in these cases J (K ) ν are really
the generators of conformal transformations.
An interesting observation is that all the Noether currents associated with the
full conformal group can be expressed in terms of a modified energy-momentum
tensor. First note that the energy-momentum tensor defined above in not
μ
necessarily symmetric. In fact from the conservation of J (M ) ν ρ the antisymmetric
part of Tμν can be determined since,
μ
∂μ J (M ) ν ρ = Tρν − Tν ρ − ∂μ (Πμ Σμν Φ) = 0. (17.17)
Using this result it is now easy to define a modified conserved symmetric energy-
momentum tensor,
1
(S )
Tμν = Tμν + ∂ ρ (Πρ Σμν Φ − Πμ Σρν Φ − Πν Σρμ Φ). (17.18)
2
The current associated with the Lorentz transformations can be expressed in
(S )
terms of Tμν as,
μ μ μ
J (M ) ν ρ = xν T (S ) ρ − xρ T (S ) ν . (17.19)
2 The use of the SL(2, R) group in applications of conformal symmetry to QCD was introduced
in [150] and [83]
The SL(2, R) collinear subgroup is particularly useful for collinear fields, where
for instance Φ(x) takes the form Φ(α) ≡ Φ(αnμ+ ), with α a real number and nμ+
the light-cone direction defined above. In particular, as will be shown below,
this will apply to parton description of quarks. The field Φ(α) is taken to be an
eigenstate of the spin operator Σ+− ,
so that s is the spin projection to the n+ direction. The collinear subgroup of the
conformal group now acts on the coordinate α as an SL(2, R) transformation,
aα + b
α → α = , (17.30)
cα + d
where a, b, c, d are real numbers with ad − bc = 1, and correspondingly the field
Φ(α) transforms as,
aα + b
Φ(α) → (cα + d)−2j Φ , (17.31)
cα + d
with,
1
j= (l + s). (17.32)
2
Thus Φ(α) is a representation of SL(2, R) or an SL(2, R) form of degree j.
The generators of the SL(2, R) group and E act on the collinear field as,
In a similar manner the primary operator and the highest weight state of the
four-dimensional collinear group are defined [171], [42] as,
An interesting and useful map relates the descendant operators and polyno-
mials. Consider for instance the descendent operator defined in (17.39), which
can be re-expressed as,
L+ → L̃− = −u
L− → L̃+ = (u∂u2 + 2j∂u )
L0 → L̃0 = (u∂u + j). (17.43)
∂α → u α → ∂u , (17.44)
then interchanging the + and − components, and finally some “normal ordering”
of taking the derivatives with respect to u to the right of the factors of u.
The representation in terms of polynomials is referred to as the ‘adjoint repre-
sentation’. Note that since in the original algebra L− includes a term proportional
to α2 , the new algebra includes a second-derivative term ∂u2 in L̃+ . This can be
avoided by introducing a different argument of the polynomials defined as,
un
→ κn , (17.45)
Γ(n + 2j)
so that the action L0 , L± on P̃(κ) is the same as (17.33) with α → κ and the
interchange of L− and L+ .
We now discuss composite operators built from two “elementary” operators of
the form,
4 COPE in four dimensions was introduced in [90] and used in QCD in [49], [50], [51].
5 Conformal Ward identities which were studied in [168] are identical to the Callan–Symanzik
equation [55] and [204].
This relation can be derived straightforwardly using the path integral formulation
of correlation functions. The Ward identity takes a simpler form when integrating
over y μ ,
<T δφ(x1 )...φ(xN )> +...+ <T φ(x1 )...δφ(xi )...φ(xN )> +...
<T φ(x1 )...δφ(xN )> + <T iδSφ(x1 )...φ(xN )> = 0. (17.57)
In particular in analogy with (2.56) the Ward identities associated with dila-
tion and special conformal transformation take the form,
N
(lφ + xi ∂i ) <T φ(x1 )...φ(xN )> = −i d4 x <T ΔD (x)φ(x1 )...φ(xN )>
i
N
(2xμi (lφ + xi ∂i ) − 2Σμν xνi − x2i ∂iμ ) <T φ(x1 )...φ(xN )>
i
= −i d4 x2xμ <T ΔD (x)φ(x1 )...φ(xN )>, (17.58)
where lφ is the canonical dimension, namely that of the free field. Similarly
to the way we extracted information about the structure of correlators in 2d
CFT in Section 2.9, we can now constrain the form of correlators in 4d. The
Ward identities associated with the Poincare transformations imply that any
correlation function is in fact not a general function of the N coordinates xμi ,
but only of the invariants x2ij ≡ (xi − xj )2 .
To understand the implication of the dilatation transformation on the correla-
tion function let us first study the theory at its fixed point, namely at a coupling
g ∗ = g(μ∗ ) such that β(g ∗ ) = 0. Recall that the β function is defined as,
∂
β(g(μ)) = μ g(μ), (17.59)
∂μ
and hence the vanishing β function implies a fixed point of the coupling constant
g. This will be further discussed below for the case of 4d QCD. In this case the
dilatation Ward identity takes the form of that of a free theory, like the one given
in (17.58), apart from the change of scaling dimension,
N
(lΦ + γ(g ∗ ) + xi ∂i ) <T φ(x1 )...φ(xN )> = 0, (17.60)
i
It is easy to realize that the right-hand side of the conformal Ward identity can
be rewritten in the form,
∂
i d4 x <T ΔD (x)Φ(x1 )...Φ(xN )> = −M <T Φ(x1 )...Φ(xN )>, (17.66)
∂M
as follows from,
∂
ΔD (x) = −M Leff , (17.67)
∂M
for the cases with no explicit dimension-full parameters. We will show this explic-
itly for the effective theory of 4d QCD below. On the other hand the dependence
of the correlator on M follows from the dependence of the field renormalization
factor and the dependence of the coupling constant so that,
∂ ∂ N
−M <T Φ(x1 )...Φ(xN )> = β(g) + γΦ <T Φ(x1 )...Φ(xN )> .
∂M ∂g i=1 i
(17.68)
Combining this together with (17.58) and (17.60) we get the Callan–Symanzik
equation,
∂ ∂ N
μ + β(g) + γΦ <T Φ(x1 )...Φ(xN )> = 0. (17.69)
∂μ ∂g i= 1 i
The invariance under these transformations manifests itself in the form of con-
servation of the corresponding Noether current,
μν i ↔
Dμ = xν T (T L ) = xν F μρ a Fρν a + ψ̄(D)(μ γ ν ) ψ , ∂ μ Dμ = 0, (17.72)
2
↔ → ←
− −
where (D) ≡ D − D. The classical invariance is not maintained quantum
mechanically. This situation of having classical conformal symmetry but not
a corresponding quantum mechanical one is referred to as the conformal
loop beta function. It is easy to check that this one loop renormalized action is
not invariant under the scale transformations of (17.71). The variation of the
action under those transformation reads,
1 1 a μν a
δS = − β0 lnλ d4
x F F + ... . (17.74)
32π2 g02 μν low
Thus the quantum mechanically (unlike the classical case) dilatation Noether
current is not conserved,
1
∂μ Dμ ≡ ΔD = − 2
a
β(g)Fμν F μν a low , (17.75)
32π
and in deriving the right-hand side of the equation we have used the equations
of motion.
The effective action admits also an anomaly with respect to the special con-
formal transformations.
In (17.46) we discussed the general structure of non-local operators of four-
dimensional conformal field theory. In QCD in many cases we encounter a non-
local operator built from a quark and an anti-quark at light-like separation, with
where
↔ →
− ←
− →
− ←
−
D+ = D + − D + d+ = D + + D + , (17.82)
and where the Jacobi polynomials with two identical indices were replaced by
3/2 1/2
the Gegenbauer polynomials P (1,1) ∼ Cn and P (0,0) ∼ Cn .
A similar analysis can be carried out for the gluons. The various components
of the gluon field have the following properties,
F+T (s = +1, j = 3/2, t = 1) FT T , F+− (s = 0, j = 1, t = 2)
F−T (s = −1, j = 1/2, t = 3). (17.83)
Local operators built from two-gluon fields with leading twist are,
↔
Gn3/2,3/2 (α) = (i∂+ )n F+T (α)Cn5/2 D+ /d+ F+T (α) . (17.84)
charges of the u, d and s quarks. At the tree level only the transverse components
are of interest. The latter have spin sj = 0 and twist tj = 3. The quark operators
Qn1,1 are the relevant basis for the expansion, with conformal spin jn = (ln + 1 +
n)/2 = n + 2 and tn = (ln − 1 − n)/2 = 2. As Δ = 1 we find
J T (x)J T (0) ∼
∞ (6−t n )/2 1
n=0 Cn 1
x2 (−ix− )n + 1 Γ(jΓ(2j n)
n )Γ(j n ) 0
du[u(1 − u)]j n −1 Qn1,1 (ux− ).
(17.85)
The coefficients Cn can be extracted from deep inelastic scattering via the fol-
lowing matrix element of forward scattering,
∞
(6−t n )/2
1
<P |J T (x)J T (0)|P> ∼ Cn (−ix− )n +1 <P |Qn1,1 (0)|P> .
n=0
x2
(17.86)
Another application of the COPE is the determination of the short-distance
expansion of the operator Q+ ( 17.76). This case is characterized by sA = sB =
s1 = s2 = 12 so that Δ = 0 and lA = lB = l1 = l2 = 32 and we find,
∞
1
n
Q+ (α1 , α2 ) ∼ C̃n (−i) (α1 −α2 )n duun + 1 (1−u)n +1 Qn1,1 (uα1 +(1−u)α2 ),
n=0 0
(17.87)
C n Γ(n +2) 2
where C̃n = Γ(2n +4) which can be determined again from forward matrix ele-
ments and are found to be C̃n = 2(2n + 3)
(n + 1)! .
Conformal invariance can be used at short distances to give predictions for the
quark distribution amplitudes for flavor non-singlet mesons, namely the wave
functions which control the behavior of the exclusive mesons processes at large
momentum transfer. Here we discuss as an example the pion distribution ampli-
tude in the leading twist order.
The basic ingredient in computing exclusive reactions including a large
momentum transfer to a pion is the matrix element of a quark anti-quark between
the vacuum and a one pion state. By using the light-cone gauge A+ = 0 the
Wilson line (17.76) is set to unity. We choose a frame where pμ = p+ n−μ and
xμ = x− nμ+ + xμT , x+ = 0 so that x2 = x2T . The matrix element can then be
written as,
¯
<0|d(0)[0, ∞n]γ+ γ5 [∞n + x, x]u(x)|π+ (p)> =
1
ifπ p+ 0
dye−iy (p·x) f (y, lnx2 ) + O(x2 ). (17.88)
This matrix element is the probability amplitude to find the pion in the valence
state consisting of a u-quark carrying a momentum y and an anti-d quark of
momentum ȳ = 1 − y and have a transverse separation xT . This amplitude is
intimately related to the pion electromagnetic form factor for large momentum
transfer Q2 and small separation distance of the order xT ∼ 1/Q2 . To approach
this limit, one defines the pion distribution amplitude taken at exactly light-like
separation where xT = 0. This amplitude reads,
1
¯
<0|d(0)[0, α]γ+ γ5 u(α)|π+ (p)> = ifπ p+ dye−iy (α p +) φπ (y, μ). (17.89)
0
The distribution amplitude φπ (y, μ) is scale and scheme dependent. In fact the
small transverse distance behavior of the valence component of the pion wave
function is traded for the scale dependence of the distribution amplitude.
It can be shown that the evolution equation of φπ (y, μ) is given by,
1
d
μ2 2 φπ (y, μ) = dỹV (y, ỹ, αs (μ)φπ (y, μ), (17.90)
dμ 0
The mixing matrix is in fact triangular since operators with fewer total deriva-
tives can only mix with operators with more total derivatives but not the other
way around. The components of the matrix on the diagonal are true anomalous
dimensions, which are identical to those of inelastic scattering,
2
n+1
1
γn = CF 1 −
(0)
+4 , (17.95)
(n + 1)(n + 2) m =2
m
where,
(0)
γn
αs (μ) 11 2
β0
<P |On ,0 (μ)|P> = <P |On ,0 (μ0 )|P> Nc − Nf . β0 =
αs (μ0 )3 3
(17.96)
Conformal invariance is useful in finding the eigenvectors of the mixing matrix
since conformal operators with different conformal spins cannot mix under renor-
malization to leading order. This happens since to leading order the renormaliza-
tion is determined by counter terms of the tree level which is conformal invariant.
Thus the mixing eigenvector operators are Q1,1 (x) defined in (17.80) with the
right flavor and γ matrices structure,
↔
¯
Qn1,1 (x) = (i∂+ )n d(x)γ 3/2
+ γ5 Cn D+ /d+ u(x) . (17.97)
Note that because of their flavor content these operators cannot mix with oper-
ators made out of gluons and they also cannot mix with operators with more
fields since they have higher twist. Thus the operators (17.97) are the only rel-
evant ones and they must be multiplicatively renormalized. Comparing (17.93)
with (17.97) one concludes that the Gegenbauer moments of the pion distri-
bution amplitudes are given in terms of reduced matrix elements of conformal
operators,
1
ifπ pn++ 1 dyCn3/2 (2y − 1)φπ (y, μ) = <0|Qn1,1 (0)|π+ (p)> . (17.98)
0
of the integrable aspects of the theory will be incomplete and will be missing
certain essential parts. However, since N = 4 SYM is the simplest interacting
four-dimensional non-abelian gauge theory we start with this and then proceed
to a certain limit in non-supersymmetric QCD.
There are several review papers on integrability in four-dimensional gauge
dynamics. In this chapter we follow [31] about the integrability of N = 4 SYM
theory and [34] for the scale dependence of composite operators of QCD.
(18.1)
where Fμν is the field strength associated with an SU (N ) gauge group, Φm is a
set of six m = 1, . . . , 6 scalar fields, and ψ and ψ̇ are doublets of SU (2) × SU (2).
Both the scalars and the spinors are in the adjoint representation of SU (N ). The
matrices σ μ and σ m are the chiral projections of the gamma matrices in four and
six dimensions, respectively and is the totally antisymmetric tensor of SU (2).
It is convenient to write the corresponding action as,
4
d x
S=N LN =4 , (18.2)
4π 2
g2 N
where the coupling constant is taken to be g 2 ≡ Y8π M
2 .
It is well known that the theory, on top of being invariant under SU (N )
gauge symmetry and SO(6) global symmetry, is also conformal invariant and in
fact superconformal invariant. The β function of the theory which vanishes to
all orders in perturbation theory is believed to vanish also non-perturbatively
and hence the theory is assumed to be conformal also in the quantum level. In
Section 17.1 we have described the conformal symmetry algebra in four dimen-
sions. Recall the SO(2, 4)2 conformal transformations (see 17.7) which are being
generated by P μ , S μν , D, K μ , the generators of space-time translations, Lorentz
transformations, dilation and special conformal transformation, respectively.
A major player in the structure of the N = 4 is the dilatation operator D.
Whereas the generators of the Poincare group do not get quantum corrections,
the dilatation operator does so that in fact,
D = D0 + δD(g), (18.3)
2 In fact the N = 4 SYM admits a superconformal algebra of psu(2, 2|4) which we do not
discuss here.
where D stands for the covariant derivative and F ≡ Fμν is the field strength.
The Hilbert space of states is built, as for any conformal field theory, from
Verma modules each characterized by a highest weight state or a primary state,
which were defined in (2.8). An example of a highest weight state is |K> =
η m n Tr[Φm Φn ]. The rest of the Verma module includes the descendant states
which are derived by acting with lowering operators on primary states. Needless
to say the general structure of correlation functions of four-dimensional conformal
field theories discussed in Section 17.6 applies also for the case of the N = 4 SYM.
In particular recall (2.8) that the two-point function of two operators is given by,
M (g)
<O(x1 )O(x2 )> = . (18.5)
|x1 − x2 |2D(g )
The anomalous dimension can be computed perturbatively as a power series
in g. As was discussed in Chapter 7 the perturbation expansion becomes much
more tractable in the large N limit, namely, in the planar limit. In this limit the
dominant diagram has a vanishing Euler number χ = 2C − 2G − T = 0 where
C, G, T stand for the number of components, genus, namely the number of
handles, and the number of traces, respectively. Since each component requires
two traces, one incoming and one outgoing, it implies that the planar limit
projects onto diagrams with G = 0 and T = 2C. This means that only single
trace operators are relevant.
We have seen above that in the planar limit we deal with single trace operators.
Pictorially, (see Fig. 18.1) a single trace operator looks like a cyclic spin chain.
This map can be made precise. Spin chain as integrable models were discussed
in Section 5.14. Recall that a spin chain includes a set of L spins with cyclic
adjacency property.3 The spin at each site is a module of the symmetry algebra
of the system. The Hilbert space of the whole system is the tensor product
of L modules. In Section 5.14 we discussed only chains with a fixed number
3 In Section 5.14 we denoted the number of spins by N . Here to avoid confusion with the rank
of the gauge group we will refer to the number of spins as L.
of spins. One can generalize this situation to incorporate also a dynamic spin
chain with an unfixed number of spins. In this case the Hilbert space is a tensor
product of all Hilbert spaces of a fixed length. In the Heisenberg model each
(L )
spin has two possible states and the Hilbert space is therefore C (2 ) . In general
the spin in the chain can point in more than two directions and in particular
also in infinitely many directions, as is the case for the spin chain of the N = 4
SYM theory. In the latter case the spin is mapped into a field operator and the
possible spin states to the components of the gauge symmetry multiplet. The
cyclicity of the single trace operators maps into a constraint on the spin chain
so that states that differ by a trivial shift are identified and hence states with
non-trivial momentum are unphysical. In the language of Section 5.14 we have to
impose U = 1 as a constraint. In the Heisenberg model this renders the Hilbert
(2(L ) )
space into C Z L . The Hamiltonian of the spin chain model translates into the
dilatation operator and the energy eigenvalues to the anomalous dimensions. The
full correspondence between the spin chain and the planar limit of the N = 4
SYM theory is summarized in Table 18.1.
Once the correspondence with a spin chain model has been established, one
can proceed in a similar way as for the Heisenberg spin chain model. The next
step is to write down the algebraic Bethe ansatz which now corresponds to an
SO(6) symmetry if one considers operators constructed only from the fields Φm
or in general the psu(2, 2|4) for the full N = 4 SYM theory. The algebraic Bethe
ansatz, the analog of (18.6) now reads as follows,
L K
λk + i/2Vj k λk − λl + iMj k ,j l
= , (18.6)
λk − i/2Vj k λk − λl − iMj k ,j l
l= k
where L is the size of the chain (N in (5.224)), the total number of excitation
is K (l in (5.224)) and where for each of the corresponding Bethe roots λk one
specifies which of the simple roots is excited by jk which takes the values of
1, . . . ., #sr with #sr being the number of simple roots which for the SO(6) case
is three and for the psu(2, 2|4) is seven. M is the Cartan matrix of the algebra
(1 in (5.224)) and V are the Dynkin labels of the representation (s in (5.224)).
This is of course the analog of (5.225) and the higher conserved charges are,
i
K
1 1
Qr = 2 r −1 − 2 r −1 . (18.9)
r−1 λk + 2i Vj k λk − 2i Vj k
k=1
where the integration contour goes to the right of the poles of A(w, t) in the
w complex plane. The high energy asymptotic of A(s, t) is determined by
the poles of the partial wave amplitudes, namely if Ã(w, t) ∼ (w −w1 0 (t)) , then
A(s, t) ∼ is1+w 0 (t) . Poles in the w plane are referred to as reggeons and the posi-
tion of the pole is called the reggeon trajectory.
The partial wave amplitude Ã(w, t) can be written using the impact parameter
representation as follows,
Ã(w, t) = d b0 e
2 i(q b 0 )
d2 bA d2 bB ΦA (bA − b0 )Tw (bA , bB )ΦB (bB )
≡ d2 b0 ei(q b 0 ) <Φ(b0 )|Tw |Φ(0)>, (18.12)
where the impact factors ΦA (bA ) and ΦB (bB ) are the parton distributions which
are functions of the transverse coordinates bA = b1A , b2A , . . . , bnA for the A col-
liding hadron and bB = b1B , b2B , . . . , bnB for the B hadron, and Tw (bA , bB ) is the
scattering (partial wave) amplitude for a given parton configuration. The idea
behind this representation of the amplitude is that the transverse coordinates
of the partons can be considered as “frozen” during the interaction. It implies
that the structure of the poles in the w-plane does not depend on the parton
distribution in the colliding hadrons but rather on the general properties of the
gluon interaction of the t-channel. It was shown [145], that the propagators of
the t-channel gluons develop their own Regge trajectory due to interactions. A
t-channel gluon “dressed” by the virtual corrections is referred to as reggeized
gluon. The reggeized gluons are the relevant degrees of freedom of the high energy
scattering. The partial waves Tw (bA , bB ) can be classified according to the num-
ber of the reggeized gluons propagating in the t-channel. The minimal number
required to get a colorless exchange is two gluons. We will discuss here only this
case. It can be shown that the amplitude Tw (b1A , b2A b1B b2B ) satisfies the so-called
BFKL equation that reads [23], [145],
αs Nc
wTw = Tw(0) + HB F K L Tw , (18.13)
π
(0)
where Tw corresponds to the free exchange of two gluons. Formally one can
write the solution as,
−1
αs Nc
Tw = w − HB F K L Tw(0) , (18.14)
π
where Ψα is the eigenstate. The high energy behavior of the scattering ampli-
tude is dominated by the right-most singularity of Tw , namely on the max-
imal eigenvalue (Eα )m ax . The equation (18.15) has the interpretation of the
two-dimensional Schrödinger equation of two interacting particles. The inter-
acting particles can be identified with reggeized gluons and Ψα (b1 , b2 ) is the
wavefunction of a colorless bound state of them. Defining the holomorphic and
where Ji,i+ 1 is related to the sum of two spins of the neighboring sites, Si2 =
s(s + 1) and H(x) is the following harmonic function,
2s−1
1
H(x) = = ψ(2s + 1) − ψ(x + 1). (18.29)
1+l
l=x
To connect it to the analysis of Section 5.14 we check this for s = 1/2. For this
case Ji,i+ 1 can take one of the two values 0, 1 for which we have H(0) = 1 and
H(1) = 0, so that the Hamiltonian is a projection into Ji,i+1 = 0 subspace with
H(Ji,i+ 1 ) = 14 − Si · Si+ 1 which is identical to (5.158).
One can generalize the exchange of colorless boundstates of two reggeized
gluons to exchange of multireggeon boundstates built from Nr reggeized gluons.
This is beyond the scope of this book and can be found for instance in [34].
(a) (b)
Fig. 19.1. Four-dimensional Feynman rules in the usual form and in the double
line notation.
which obviously is not suitable for a large N expansion whereas if one replaces
g → √gN the β function takes the form,
3
dg 11 2Nf g
μ =− − + O(g 5 ). (19.3)
dμ 3 3N 16π2
The rules of the Feynman diagrams in two-dimensional QCD (see Fig. 10.1)
include the fermion propagator, the gluon propagator and the quark gluon vertex,
all expressed in the double line notation. Using the light-cone in two dimensions
one eliminates the three- and four-gluon vertices. In four dimensions due to the
transverse directions the gluon vertices cannot be eliminated by choosing a gauge.
Thus all together the four-dimensional Feynman rules are expressed in Fig. 19.1.
The figures a,b,c are identical to the two-dimensional ones (Fig. 10.1) whereas
figures d and e are the three- and four-gluon vertices. The quark propagator
(19.1a) is given by,
Fig. 19.2. Color flow in the double line notation associated with two traces.
λ N2
λ2N 2
λ3N 2
λ2
where,
χ ≡ V − E + F = 2 − 2h − b, (19.9)
is the Euler character of the surface, h is the genus, namely, the number of
handles and b is the number of boundaries. E(G ) and V(G ) are the appropriate
qualities for gluons. For instance the sphere has χ = 2 since it has no handles and
no boundaries, the disk has χ = 1 since it has no handles and one boundary and
the torus has one handle and no boundaries and hence it has χ = 0. Thus the
Feynman diagrams look like triangulated two-dimensional surfaces. In fact all
possible gluon exchange may fill the holes of the triangulated structure forming
a smooth surface with no boundaries for gluon only diagrams, and with bound-
aries for diagrams that include quark loops. It was conjectured that the two-
dimensional surface is the world sheet of a string theory which is dual to QCD.
There has been tremendous progress in this string/gauge duality in recent years
following the seminal AdS/CFT duality of Maldacena [158]. This is beyond the
scope of this book and we refer the reader to the relevant literature, for instance
the review [10].
To further demonstrate the determination of the order of a diagram consider
first the diagrams that involve only gluons which appear in Fig. 19.3. The dia-
gram in (a) has V = 2, E = 3, F = 3 and thus it behaves as N 2 λ. Similarly in
(b) and (c) V = 4, E = 6, F = 4 and V = 5, E = 8, F = 5 so that they behave as
N 2 λ2 and N 2 λ3 , respectively. The three diagrams (a), (b) and (c) are all planar
diagrams and have a topology of a sphere. Note however that diagram (d) which
is non-planar behaves as N 0 λ2 , namely of genus one. In the large N limit this
last diagram is obviously suppressed.
So far we have only discussed diagrams with gluons. Quarks propagators are
represented (see Fig. 19.1a) by a single line. A closed quark loop is a boundary
and hence using (19.9) it contributes to the diagram a factor of N1 . Consider for
example the diagram drawn in Fig. 19.4.
Fig. 19.4. Four-dimensional Feynman rules in the usual form and in the double
line notation.
It is a diagram of order N . This follows trivially from the fact that it has zero
genus, h = 0 and one boundary b = 1. Alternatively we have one gluon vertex,
three gluon propagators and three loops and hence N 1−3+3 = N . Obviously this
is also the result when one uses the unrescaled operators where each vertex con-
tributes √1N and each index loop N , so that we get ( √1N )4 × N 3 = N . Similar to
the non-planar gluon diagram (19.3d), Fig. 19.5 describes a non-planar diagram
that includes both gluons and quarks. This diagram scales like N1 since there is
no gluon vertex, two gluon propagators and one index loop N 0−2+1 = N1 .
As was mentioned above the difference between the SU (N ) case versus the
U (N ) can be accounted by adding a ghost U (1) gauge field whose role is to cancel
the extra U (1) part of the U (N ). The U (1) commutes with the U (N ) gauge fields
and therefore does not interact with them and hence one has to incorporate only
the coupling of the quark fields to the U (1) gauge field. When we consider a con-
nected diagram with gluons and U (1) ghost gauge fields the contribution to the
Fig. 19.6. U (1) ghost propagator connecting two otherwise disconnected diagrams.
counting of orders of N due to the gluons is not affected, whereas each U (1) ghost
field contributes a factor of N12 . The latter follows from the fact that the ghost
U (1) propagator contributes a factor of 1/N and another 1/N factor due to the
two coupling constants at the end of the propagator. This can easily be seen from
the unrescaled action (19.1). For diagrams that are connected with the ghost field
and otherwise disconnected as is depicted in Fig. 19.6 the situation is different.
For instance that diagram is of order N 0 since it has N × N × N12 . Note however
that even in this case there is a difference of order N12 between the SU (N ) and
U (N ) cases, (actually, in Fig. 19.6 the contribution of SU (N ) vanishes).
Correlator N counting
with a unit amplitude. In particular we read from the table that the glueball
meson interaction is of order √1N .
since single meson exchange dominates in the large N limit. The logarithmic
dependence on q 2 of the left-hand side can be recast only provided that the
sum on the right-hand side includes infinitely many terms. The resonances are
narrow since their decay width goes to zero in the large N limit. This follows
from the fact that the phase space factor is N independent and the coupling
constant behaves like √1N .
r In Chapter 17 we encountered the pion decay constant fπ . Let us check how
it scales with N . Recall its definition <0|ψ̄γ5 T A√
ψ|π b (p)>= ifπ pμ δ ab . The cor-
responding gauge invariant correlator is <N M̂1 N M̂2 >, where√the first oper-
ator N M̂1 corresponds to the axial current and the second N M̂2 to the
q q
q q
q q
q q
Fig. 19.7. Zweig’s rule for the decay of a meson into two mesons.
pion√produced from the vacuum with a unit amplitude. This correlator scales
like N and hence,
√
fπ ∼ N . (19.12)
r The suppression of exotic states of the form q q̄q q̄ and the fact that the meson
is almost a pure q q̄ state with little impact of the q q̄ sea are straightforward
consequences of large N . Since a quark loop, as we have seen above, is sup-
pressed by a factor of N1 the q q̄ sea is irrelevant. Since in the leading order the
mesons are non interacting in large N , there is no interaction that will bind
two mesons into a q q̄q q̄ exotic state.
r Consider the two diagrams of Fig. 19.7.
Using the counting rules it is obvious that the right-hand diagram is N1 sup-
pressed in comparison to the one on the left. Correspondingly the meson will
preferably decay into two mesons of the left-hand side of the figure, what is
referred to as Zweig rule conserving decay, and not to the two mesons on the
right which is a Zweig rule suppressed decay. In this sense large N predicts the
Zweig rule. The same mechanism is in charge of the fact that there is almost
flavor singlet and octet degeneracy. In the leading order in large N the whole
nonet is degenerate since the diagrams that split singlets from octets involve
a q q̄ annihilation which is order of N1 . In the large N for instance the vector
mesons (ρ, w, φ, K ∗ ) are degenerate.
r It is known that meson decay proceeds mainly via decay into two body states
and not into states of more mesons. Large N tells us that the decay into two
mesons behaves as √1N , whereas a decay into four mesons is of order N 13 / 2 .
This can also be compared to the decay of a meson via creation of a quark
anti-quark pair in the mesonic flux tube [60].
r The N counting rules tell us also that meson scattering amplitudes are given by
an infinite sum of tree diagram of exchange of physical mesons. This fits nicely
the so-called Regge phenomenology, where strong interactions are interpreted
as an infinite sum of tree diagrams with hadron exchange.
r Another very important phenomenon is related to the axial U (1), the theta
term and the mass of the η . This will be described in detail in Section 22.5,
but here we present the picture in the large N .
(19.16)
It is easy to check that in perturbation theory U (k) is of order N 2 due to
the contribution of the N 2 degrees of freedom of the gluons. However, perturba-
tively, limk →0 U (k) = 0 since F F̃ is a total derivative. One concludes that per-
turbatively the vacuum energy is θ independent. To better understand (19.16)
we rewrite U (k) in terms of a sum over intermediate single particle states,
N 2 (agb )2 N (am es )2
n n
U (k) = + , (19.17)
k 2 − (mgb )2n m es
k 2 − (m
m es )2n
gb
√
where gb stands for glueball and mes for meson. N agb and N am es are the
amplitudes for Tr[F F̃ ] to create a glueball and meson state, respectively. This
result again follows from the N counting rules,
√
<0|Tr[F F̃ ]|mes>∼ N <0|Tr[F F̃ ]|gb>∼ N. (19.18)
The fact that only single states and not multi-states are taken in the interme-
diate states is since the latter are suppressed in the large N . In the pure YM
2
without quarks the first term vanishes and hence U (0) ∼ N 2 and ddθE2 |θ =0 ∼ 1.
In the presence of massless quarks we know that there could not be any θ depen-
dence and thus we should be able to show that the first term is cancelled out.
However, it seems that there is no way that the second term can cancel the first.
In fact it is possible if there is one meson state with mass mm es ∼ √1N and if the
two terms have opposite sign. This obviously can cancel the k = 0 term in U (k)
and does not cancel for non-trivial k, but this is exactly what enters (19.16). The
opposite sign follows from the fact that an additional equal time commutator
term has to be added to (19.15) (see the appendix of [221]). Assuming such a
state with mass mm es ∼ √1N the form of U (0) is,
a2η
U (0) = N . (19.19)
Mη2
Using the axial anomaly equation which will be further discussed in (22.5),
g2
∂μ J5μ = Nf Tr[F F̃ ], (19.20)
4π2 N
we get,
g2 N N
<0|Tr[F F̃ ]|η > = <0|∂μ J5μ |η >= fη Mη2 , (19.21)
8π2 2Nf 2Nf
From this relation we get the Veneziano–Witten formula or the mass of the η ,
2
2N(f ) d2 E
Mη2 = |θ =0 . (19.22)
f(π ) dθ2
The picture that emerges from this discussion is that the η is a Goldstone boson
in the large N limit. It has a mass of the order of Mη ∼ √1N . The dependence
on η of nonzero amplitudes can be obtained from the dependence on θ in the
theory without quarks by the following replacement,
2N(f )
θ→θ+ η . (19.23)
f(η )
1
Note that f(η ) = f(π ) to leading order in N .
diagram scales as N1 . However since there are 12 N (N − 1) possible pairs the net
effect is of order N . In a similar manner the exchange of two gluons is of order
( √1N )4 N 4 ∼ N 2 , where the first factor comes from the four vertices and the sec-
ond from the number of ways to choose the four quarks. Higher-order exchange
diagrams will have higher order divergence in large N . We will now show, that
in spite of this fatal obstacle, there is a large N approximation to the problem of
the baryons. The idea is to divide the problem into two parts, in the first one uses
diagrammatic methods to study the problem of n quark interaction, and then
the effect of these forces on an N quark state. Let us first apply this approach
for determining the dependence of the mass of the baryon on N in the quark
model. Assuming that the mass gets contributions from the quark masses, quark
kinetic energy and quark–quark potential energy, the mass of the baryon reads,
1
MB = N m q + T q + V q , (19.24)
2
where mq is the quark mass, Tq is the kinetic energy of the quark and Vq is
the quark–quark potential energy. Thus we observe that the mass of the baryon
scales as N . This result will be shown to hold even beyond the quark model.
Again we have made use of the fact that the potential energy is combined from
the N 2 combinatorial factor and the N1 factor that comes from the vertices, or
gluon propagator.
Leaving aside the quark model, we want to address first the baryonic system
made out of very heavy quarks.
where we have suppressed the flavor and spin degrees of freedom, V n stands
for the n body interaction and is independent of N and its strength is of order
N 1−n , as explained above. In fact since the number of clusters of n quarks is of
order N n , each term in the Hamiltonian is proportional to N .
Next one takes for the ground state wave function a product of the wave
functions of each of the particles, namely,
where the particle wave functions are determined by a variational method. The
expectation value of the Hamiltonian,
1 3 2 1
<ψ|H|ψ> = N d r|∇φ| + d3 r1 d3 r2 V 2 (r1 , r1 )|φ(r1 )φ(r2 )|2 +
2m 2
1 3 3 3 3 2
+ d r1 d r2 d r3 V (r1 , r2 , r3 )|φ|(r1 )φ(r2 )φ(r3 )| + . . . , (19.27)
6
has to be minimized with respect to φ(r) subjected to the constraint that,
d3 r|φ|2 = 1. (19.28)
In spite of the fact that the interaction potential behaves like N1 , we can-
not treat this term as a perturbation since each quark interacts with N other
quarks and hence the total quark–quark interaction of each quark is of order one.
This situation calls for a Hartree approximation where, as explained above, the
quark is exposed to an average effective potential. The fluctuations of the effec-
tive potential are negligible and hence we can consider a background c-number
potential.
For heavy quarks where the potential is taken to be a Coulomb potential we
find,
)| |φ(r
d3 r |φ(r |r
2
g2 )|2
<ψ|H − N |ψ>= N [M + 1
2m d3 r|∇φ|2 − 2 d3 r −r |
− d3 r|φ(r)|2 ]. (19.32)
The main point here is that each of the terms is proportional to N and hence
the result of the minimization is N independent. The variation with respect to
φ∗ results in the following Schrodinger equation,
∇2 φ |φ(r )|2
− − g φ d3 r
2
= φ. (19.33)
2m |r − r |
One can convert this integro-differential equation into a fourth-order differential
equation
1 2 ∇2 φ
− ∇ + 4πg 2 |φ|2 = 0. (19.34)
2m φ
For radial solutions, for instance, the ground state of this equation takes the
form,
2
1 d 2 1 d2 2
− + ddr + ddr φ + 4πg 2 |φ|2 = 0, (19.35)
2m dr2 r φ dr2 r
q q
heavy quark baryons, also for the baryons made out of light quarks, the mass is
linear with N , whereas the size and the shape of the baryon are N independent.
A major difference between the case of heavy quarks versus that of light ones
is that in the latter case one has to introduce on top of the two-body interaction
also a three-body interaction and, in general, n-body interactions. In addition one
has to use a relativistic analog of the Hartree approximation. In two dimensions
one can solve the relativistic Hartree approximation. Unfortunately, the four-
dimensional analog is not known. Let us first discuss a non relativistic Hartree
approximation with n-body interactions and then argue about the relativistic
analog. The Hamiltonian for the case with any n-body interaction takes the form,
1 1 2 1 3
H= |pa |2 + V (ra , rb ) + V (ra , rb , rc ) + . . . ,
2m a 2N 6N 2
a= b a= b= c
(19.36)
n
where we have suppressed the flavor and spin degrees of freedom, V stands
for the n-body interaction and is independent of N . The strength of V n is of
order N 1−n since breaking the n quark line costs a factor of N −n and since the
baryon is in a totally antisymmetric representation, each quark line carries a
different color index.
We now substitute this Hamiltonian into <ψ|H|ψ> and use a variational
method as above. Since for each V n term there are N n ways to choose a set
of n quarks, here again the expectation value of the Hamiltonian is linear in N .
Next we have to introduce a four-dimensional relativistic Hartree approxima-
tion. In two dimensions in the large N limit the Hartree approximation is exact.
The generalization to four dimensions, however, is not known and hence one can
make only the qualitative statement that even in this case the mass is linear in
N and the size and shape are independent of N .
The Hartree approximation of light quarks moving in an effective potential
can be also related to a string model of the baryon. In this model, the N quarks
are attached to a common junction as can be seen in Fig. 19.9 for the case of
N = 3.1 In the large N approximation the junction can be regarded as a heavy
object and its motion can be ignored. The interaction of the quarks with the fixed
junction can be thought of as an interaction with an effective Hartree potential.
Inserting this ansatz into the expectation value of the Hamiltonian one gets a set
of two coupled nonlinear equations for φ0 and φ1 . This structure can obviously
be generalized to states with higher single particle excited states.
Another approach to studying excited states is to apply a time-dependent
Hartree approximation. It is easy to check that starting with the Hartree ansatz
for the wave function but now with single particle wave functions that are
also time dependent, one finds instead of (19.33) the following time-dependent
Schrodinger equation,
∇2 φ |φ(r )|2
− − g 2 φ d3 r = i∂t φ(r, t). (19.38)
2m |r − r |
This equation is solved by φ(t, r) = e−it φ(r) where φ(r) is a solution of the time-
independent equation. By Galilean boosting along, for instance the x direction,
a static baryon solution, we find the solution,
1 2
φ(r, t) = φ(x − vt)ei(M v x−t− 2 M v t)
, (19.39)
where n is some integer number. Recall that this condition follows from the fact
that the solutions are invariant under time translations, so from ψ(t) we can
also generate a solution ψ(t − t0 ) for any t0 and also any linear combinations of
T
them, and in particular a harmonic varying solution 0 dt0 e−it 0 E ψ(t − t0 ).
i j i i
i j i i
In analogy to the discussion in Section 5.5.1 here as well one can introduce a
non-abelian flavor group, namely construct baryons made out of several flavors.
In this case one introduces into the Hartree wave function a separate single
particle wave function for each flavor.
For very heavy quarks one can neglect the spin-dependent forces, and hence
anticipate that the baryons are spherically symmetric. However, for less heavy
quarks this is no longer the case. For a baryon made out of a single flavor, in the
ground state all the spins are aligned and hence the total spin is 12 N . Due to the
fact that the total spin is very large, the effect of the coupling of this large spin
to the orbit is significant and hence the ground state will no longer be spherically
symmetric. If one takes the large N analog of the baryon to be composed of N 2+1
quarks of one flavor and N 2+ 1 − 1 of the other flavor, then the net spin will be
1
2 since the spin–spin interaction will align the spin of the different flavors in
an antiparallel way. Unlike the one flavor case where the spin is N2 , the spin 12
will be too small to affect the spherically symmetric ground state via spin orbit
interaction, and hence for that case it will remain symmetric.
i j
i j
Fig. 19.11. Annihilation of a quark coming from the baryon and anti-quark
coming from the anti-baryon.
another factor of N from the other baryon and a factor of N1 from the quark
gluon vertices.
This fact that the amplitude is order N is behind why there is a smooth large
N limit to the baryon–baryon scattering. Recall that the mass of the baryon and
hence also the non-relativistic kinetic energy of the baryon are linear in N . Thus
the total Hamiltonian can be written as H = N Ĥ where Ĥ is N independent.
The eigenvectors of Ĥ and hence of the scattering process are N independent.
Quantitatively one addresses the question of baryon–baryon scattering using a
non relativistic time-dependent Schrodinger equation for a system of 2N quarks.
Due to the exclusion principle the total wave function should be a product of
two orthonormal space and spin wave functions φi (x, t) where i = 1, 2 in the
following way,
N N
ψ(x1 , . . . , x2N , t) = (−1)P φ1 (xi , t) φ2 (xj , t). (19.41)
P i= 1 j =1
Using again the time-dependent variational principle, we find that in the case
where all the spins of the quarks are parallel so we can ignore them,
∇2 dyφ∗1 φ1 (y, t)
i∂t φ1 (x, t) = φ1 (x, t) − g 2 φ1 (x, t)
2M |x − y|
∗
dyφ 2 φ1 (y, t)
−g 2 φ1 (x, t) , (19.42)
|x − y|
and another equation where φ1 ↔ φ2 . Apart from the last term this equation
is identical to the one describing a single baryon (19.38), hence the last term
obviously describes the interaction between the two baryons. To describe baryon–
baryon scattering we start with inital conditions where the wave functions φ are
localized at two far away regions of space, but heading for a collision. When the
two wave functions overlap the interaction term is important and determines the
scattering via (19.42).
The baryon anti-baryon scattering is dominated by an annihilation of a quark
coming from the baryon and an anti-quark from the anti-baryon. The amplitude
of this process is of order N since choosing one quark is order N , choosing an
anti-quark is order N and the coupling is order N1 (see Fig. 19.11). Again this
is like the scaling of the kinetic term and hence there is a smooth limit. The
where φ̄(r̄b ) is the wave function of a single anti-quark. The minimization now
yields a pair of coupled equations for φ and φ̄.
The meson–baryon scattering is described in a diagram like Fig. 19.12. The
diagram is of order N 0 since there is a factor of N1 from the coupling and N
from the number of ways to choose the quark from the baryon. Recall that the
baryon kinetic energy is order N and that of the meson is order one. Hence the
interaction term is negligible from the point of view of the baryon and it does
not feel the meson but the meson motion is affected by the interaction and thus
there is a meson baryon non-trivial scattering. Denoting again the wave function
of a quark of the baryon as φ(x, t) and that of the meson as φM (xM , yM , t), the
trial many body wave function reads,
ψ(ra , . . . , rN , xM , yM , t) = φ(ra , t)φM (xM , yM , t). (19.44)
a
20.1 Introduction
Low energy effective actions associated with four-dimensional QCD, and in par-
ticular the Skyrme model have been very thoroughly studied with an emphasis
on both formal aspects such as anomalies as well as phenomenological ones like
the spectrum of baryons. This chapter is devoted to the Skyrme model. We first
derive the various terms of the Skyrme action. These include the sigma term,
the WZ term, the mass term and the Skyrme term. The first three terms we
have encountered already in the two-dimensional analog, the bosonized QCD
(Chapter 13) whereas the fourth one, the Skyrme term, shows up as a stabi-
lization term only in the four-dimensional case. We then construct the classical
soliton solution, the Skyrmion, of the corresponding equations of motion. Next
we determine the classical mass and radius of the baryon. In a similar manner to
the procedure taken in the two-dimensional model, we quantize the system semi-
classically. This yields mass splitting between the nucleons and the delta particles
and the axial coupling of the nucleons. Most of the discussion will be done for
SU (Nf = 2) but we will also discuss certain properties of the three-flavor case.
The Skyrme model was introduced in [195], [196], [197] and [91].
The topic of baryons as Skyrmions was discussed and reviewed in several
books [53], [157] and reviews [22], [186], [231]. In several sections of this chapter
we follow the latter review.
the corresponding action, we will now consider various terms that eventually
construct the full Skyrme model.
in terms of the Goldstone bosons π(x) and then consider Nf = 3. It turns out
that P0 and (−1)N B are not conserved separately but only the combination
P = P0 (−1)N B . This is demonstrated by the process K + K − → π + π 0 π − . Obvi-
ously the number of bosons modulo two is not conserved as well as the parity
P0 since the π i are pseudo scalars. It is thus clear that the action (20.3) cannot
describe the effective action of QCD and another term that does conserve P and
not P0 and (−1)N B separately. It is well known that the parity transformation
P0 is violated by a term which is proportional to the Levi Civita tensor which in
four dimensions reads μν ρσ . However, it is very easy to verify that the only term
proportional to μν ρσ , namely μν ρσ Tr [g −1 ∂μ ggg −1 ∂ν gg −1 ∂ρ gg −1 ∂σ g] vanishes
due to the antisymmetric nature of μν ρσ and the cyclicity of the trace.
M Q Q
where we have used the fact that the sum of the two disks (Dn2 + Ds2 ) is topo-
logically equivalent to S 2 , that the five cycles in the group manifold SU (N ) of
the topology S 2 × S 3 can be represented by the cycles of topology S 5 and that
π5 (SU (N )) = Z and hence any S 5 in SU (N ) is topologically a multiple of the
basic S 5 on which w can be normalized such that S 5 w50 = 2π.
0
We thus conclude that the coefficient λS has to be an integer. As we mentioned
above in two dimensions we have shown that this integer has to be Nc . We will
show below when discussing the phenomenology of the Skyrme model that this
is the case also in four dimensions. Thus we will take from here on that λS = Nc .
For Nf > 2 one can use a mass term that breaks flavor symmetry by assigning
different masses to different flavors. One can also generalize the mass term by
using general functions of g which in the limit of g → 1 approach g.
As usual the local transformation can be a symmetry of the action only provided
we add to the action gauge fields that transform under the EM gauge transfor-
mation as Aμ → Aμ − 1e ∂μ (x) where e is a unit EM charge. For the sigma term
and the Skyrme term it is obvious that gauge invariance is achieved by replac-
ing the ordinary derivative with covariant ones, namely ∂μ → Dμ = ∂μ + ie∂u .
The gauging of the WZ term Γ ≡ SW Z is more subtle and as was done for the
two-dimensional case; we use a trial and error method. First we compute the
variation of the term under the global U (1) symmetry. We find that,
Γ → Γ − d4 x∂μ (x)J μ
1 μν ρσ
Jμ = { Tr [−Q(Rν Rρ Rσ )]
48π2
+ Tr [Q(Lν Lρ Lσ )]} (20.22)
where Q is defined in (20.20). The next step in gauging the WZ term is to replace
the original term with,
Γ → Γ − e d4 xAμ J μ . (20.23)
It turns out that the action after this replacement is still not gauge invariant but
it is invariant with the following addition,
fπ2
S= d4 x Tr [Dμ gDμ g −1 ] + Nc Γ̃,
16
ie
Γ̃(g, Aμ ) = Γ(g) − e d4 xAμ J μ + d4 xμν ρσ ∂μ Aν Aρ
24π2
× Tr [Q2 (Lσ − Rσ ) − QgQg −1 Rσ ]. (20.24)
In a similar manner one can gauge the full global symmetry or its subgroups.
Since we will not need it in this chapter we refer the reader to references, for
instance [231].
1 The classical properties of the SU (2) baryonic Skyrmion were analyzed in [133], [5] and
afterwards in many other papers.
where x = ef√π 2 r is dimensionless. The value of the integral is ∼ 11.7. One can
either use the mass of the proton combined with the mass of the delta, to deter-
mine fπ and the coefficient of the Skyrme term, or use the experimental values
of fπ and the axial coupling to be discussed shortly.
Let us now analyze the radial profile of the soliton F (r). Asymptotically for
r → ∞ only the terms inside the square brackets can be neglected leading to
2
A2
a solution of the form F (r) → 16e f π2 r 2 , where again A can be determined by
comparing to experimental data and is found to be A ∼ 1.08. On the other limit
around the origin it is easy to see that the equation is solved by F (r) ∼ nπ − ar.
The numerical solution of F (r) that interpolates between these two boundary
conditions is drawn in Fig. 20.2.
In addition to the mass, we have also extracted in two dimensions from the
classical soliton the flavor properties and baryon number. For the Skyrmion we
should also be able to determine these properties as well as its spin. When we
insert the classical soliton solution in the baryon density we get,
i ij k 1 2 F
B0 = L i L j Lk = sin F , (20.31)
24π2 2π2 r2
so that the baryonic charge is,
∞
1
B = 4π drr2 B 0 (r) = (F (0) − F (∞)) + 12π[sin(2F (∞)) − sin(2F (0))] = 1.
0 π
(20.32)
where we have used the boundary conditions of F (r) specified above. Using the
distribution of the baryonic charge, one can define the rms radius of the baryon
as follows,
∞ 1/2
e
rrm s = − dxx2 sin2 F F , (20.33)
πfπ 0
F (r)
0
0 1 2 3 4
r (fm)
The hedgehog configuration, see Fig. 20.3, used as an ansatz for the Skyrmion,
is by construction invariant under the operation of,
K ≡ J + I = (L + S) + I, (20.34)
where L, S and J are the orbital angular momentum, the spin and the total
angular momentum, and I is the isospin.
It follows from,
τ
[K, g(x)] = i sin F [(r × −∇) , τ · r̂] + , τ · r̂ = 0. (20.35)
2
Hence the Skyrmion carries a charge of K = 0. It is straightforward to notice
that it is also an invariant under the parity operator defined in (20.7), so that
altogether it is K = 0+ state.
Nothing in this prescription is two dimensional and hence we now use the same
ansatz also for the four-dimensional soliton. In two dimensions we discussed the
general Nf case, here we start with the simplest case of Nf = 2 and then we
discuss the Nf = 3 and comment about the general case. As discussed above for
SU (Nf = 2) there is no WZ term, thus we have to substitute (20.36) into the
action that includes the sigma term and the Skyrme term (for simplicity we do
not add the mass term). The collective coordinates A(t) can be parameterized
in the following ways either,
or,
i
A−1 Ȧ ≡ σ · w. (20.38)
2
It is easy to verify that in terms of A(t) the Lμ defined in (20.5),
fπ2 1 1
Lσ = Tr [∂μ g∂ μ g −1 ] ∼ ∂μ π · ∂ μ π+ 2 (π · ∂ μ π)2 − π 2 ∂μ π · ∂ μ π +O(π 6 ).
4 2 6fπ
(20.58)
The quadrilinear coupling behaves like f12 . If we expand the Skyrme term in a
π
similar manner we find that in that case the coupling behaves like e 21f 4 , and
π
hence we conclude in agreement with (19.12) that,
1
fπ ∼ Nc e∼ √ . (20.59)
Nc
This enables us to check the Nc dependence of the classical Skyrmion mass and
its semi-classical extension,
fπ 1 1
Mcl ∼ ∼ Nc Msc ∼ 2
∼ e3 fπ ∼ . (20.60)
e α Nc
Recall for comparison that the two-dimensional solitonic baryons were shown to
have classical mass which is also order Nc , but the semi-classical correction term
behaves like Nc0 and not N1c .
where λa are the SU (3) Gell–Mann matrices. In terms of the pions, kaons and
η we have,
⎛ √1 0 √1 8 ⎞
2
π + 6η π+ K+
Φ≡⎝ π− − √12 π 0 + √16 η 8 K0 ⎠ . (20.62)
− 0 2
√ η8
K K̄ 6
Next we have to choose an ansatz for the static classical configuration g0 (x).
Recall that in the two-dimensional case we took
an embedding of the U (1) in
−i 4π
φ(x)
U (Nf ) of the form g0 (x) = Diag(1, 1, . . . , e Nc
). In analogy in the four-
dimensional case we embed the SU (2) hedgehog configuration in the SU (3)
group element as follows,
⎛ ⎞
0
g0 (x) = ⎝ eiF (r ) τ ·r̂ 0⎠. (20.63)
0 0 1
Since the WZ term vanishes for an SU (2) group the solution for F (r) is iden-
tical to that discussed in Section 20.3 and hence the elevation to Nf = 3 shows
up basically only in the semi-classical quantization of the collective coordinates.
Recall that the latter are introduced via g0 (x) → A(t)g0 (x)A−1 (t). In two dimen-
sions we parameterized the quantum fields A(t) in terms of the Zi , i = 1, . . . , Nf
variables which was adequate for the CP N f −1 that the collective coordinates
span in that case. Clearly in the present case since the g0 (x) ∈ SU (2) a differ-
ent ansatz is required. The most straightforward one is in terms of the angular
velocities wa that generalize those of (20.38) as follows,
i a
8
A−1 (t)Ȧ(t) = λ wa . (20.64)
2 a=1
21.1 Introduction
Solitons play an important role in non-perturbative two-dimensional fields as
we have seen in the first part of this book. They are intimately related to non-
trivial topology, they are an essential ingredient in integrable models, and they
enable the phenomenon of fermion boson duality-bosonization. When passing
to four-dimensional field theories the topology may be even richer and thus
we would anticipate having topological solitons as static solutions also in four-
dimensional space-time. As we have seen in Section 5.3, Derrik’s theorem does
not permit the existence of solitons of scalar field theory in space dimensions
higher than one, however, they are not prohibited in theories that include higher
spin fields, in particular in theories of scalar fields coupled to non-abelian gauge
fields. Indeed as we will see in this section certain theories of this type that admit
spontaneous symmetry breaking, admit soliton solutions. These configurations
carry a conserved topological charge which guarantees their stability against
decay to the vacuum. As it will turn out this charge is in fact a magnetic charge
and hence these solitons are magnetic monopoles, or in the more general case
dyons with both magnetic and electric charge. The construction of dyons from
static solutions will be the analog process of building up breathers from two-
dimensional solitons.
In the next section we present the basics of the Yang–Mills Higgs theory. We
then show the relation between magnetic monopoles and topological solitons both
for the simplest case of SU (2) (and SO(3)) as well as for a general non-abelian
gauge group. The next topic is the seminal solution of ’t Hooft and Polyakov.
Then we discuss zero modes, time-dependent solutions and dyons. In the follow-
ing section we discuss the very important limit of BPS. We then describe the
construction of multi-monopole solutions that was proposed by Nahm. We show
its application to the construction of BPS monopoles of charge one and two. The
next topic is the moduli space of monopoles. We determine the metric on this
space for the case of widely separated monopoles.
The topic of magnetic monopoles and dyons has been covered by several review
papers, proceedings of meetings and books, for instance [21], [214], [67], [7], [182]
and [193], respectively. Here in this chapter we made use of mainly the former
two references.
Aμ = T a Aaμ Φ = T a Φa , (21.2)
where T a are the generators of G (see Section 3.1), the covariant derivative reads,
where λ is taken to be positive so that the energy is bounded from below, and
we also take μ2 > 0. In general one can discuss a similar system where Φ is
in any representation of G but here we consider only the case of the adjoint
representation.
Let us start with the simplest case where G = SU (2) and Φ is in the triplet
(adjoint) representation. For such a case the vacuum solution can be put in the
form,
5
σ3 μ2
Φ(x) = v ≡ Φ0 v ≡
2 λ
Aμ (x) = 0. (21.5)
In this case the vacuum expectation value of the Higgs field breaks the SU (2)
symmetry down spontaneously to a U (1) symmetry along the a = 3 direction.
The physical fields will be denoted as follows,
A1μ + iA2μ
Aμ = A3μ Wμ = √ ϕ = Φ3 , (21.6)
2
which associate with the massless “photon”, pair of mesons W, W ∗ with a mass
√ of
eV and charges ±e and an electrically neutral scalar boson with mass mH = 2μ,
respectively.
For a general group G which we take to be a simple Lie group of rank r (for
the basic definitions see Section 3.1). In this case the expectation value of φ = Φ0
can be taken to lie in the Cartan subalgebra of the group G. Using the notation
H for the r-dimensional vector of the elements of the Cartan subalgebra, Φ0 is
characterized by a vector h such that,
Φ0 = h · H. (21.7)
The generators of the unbroken subgroup are those generators of G that commute
with Φ0 . These are all the generators of the Cartan subalgebra together with
ladder operators associated with roots orthogonal to h. If none of the γ are
orthogonal to h the unbroken symmetry is U (1)r , whereas if there are some
roots γ such that γ · h = 0 then the unbroken symmetry is U (1)r −r × K where
K is of rank r and it has γ as its root diagram.
of integration that do not cross any of the zeros of Φ. In fact a configuration with
Q = n must have at least |n| zeros of the Higgs field. If we distinguish between
a + zero and a – anti-zero then the net number is precisely n. Obviously if
we consider Higgs configurations of Q = n = 1 there must be one with minimum
energy. This cannot be smoothly deformed to a vacuum since the winding number
is quantized. Hence such a configuration must be a local minimum of the energy
and therefore a static classical solution.
The next question we want to address is what is the connection between these
non trivial soliton solutions and magnetic monopoles? The field strength of the
abelian gauge field Aμ defined in (21.6), Fμν = ∂μ Aν − ∂ν Aμ is the outcome of,
1
F̃μν = Φ̂a Fμν
a
− abc Φ̂a Dμ Φ̂b Dν Φ̂c , (21.11)
g
when we take Φ̂ = (0, 0, 1), namely Φ = ϕ. What distinguishes F̃μν from an ordi-
nary abelian field strength is that it does not obey the Biachi identity,
∗ 1 4π
dF̃ = eμν ρσ ∂ ν F̃ ρσ = ekμ = kμ , (21.12)
2 g
where g is the magnetic charge which will be shown to be equal to 4π/e. Defining
now the magnetic field Bi associated with F̃μν as usual as,
1
Bi ≡ ij k F̃ j k , (21.13)
2
we find that,
4π 1 4π 4π
∇·B = k0 QM = d3 xk0 = Q= n. (21.14)
g g e e
We have thus realized that the non-trivial soliton configurations carry a magnetic
charge and hence are magnetic monopoles. We can further determine the classical
mass of the monopole since the total energy of such a solution is,
E = d3 x Tr[Ei2 ] + Tr[(D0 Φ)2 ] + Tr[Bi2 ] + Tr[(Di Φ)2 ] + V (Φ) . (21.15)
For a static configuration that does not carry electric charge, the first two terms
are expected to vanish. Then one can show that the form of the mass has to be,
4πv λ
M= f , (21.16)
e e2
where f ( eλ2 ) should be of order one.
So far we have discussed the topological charges of the group G = SU (2) case.
Let us now address the general case. Instead of the map (21.10), the asymptotic
Higgs field constitutes in the general case a map,
G
Φ(∞) : ∂M → , (21.17)
H
where ∂M is the boundary of the space which for ordinary flat Minkowski space-
time is S2 and G/H is the coset of the unbroken symmetry group H and the
original group G, which in the discussion above was SU (2)/U (1) = S2 . These
maps from the boundary of space to the coset, fall into equivalent classes which
form the homotopy group π2 (G/H). For simply connected group G, namely with
π1 (G) = 0 the classification of the maps is in fact done by π1 (H) since for this
type of G,
The image of a given homomorphism equals the kernel of the next one in the
sequence. It is well known that for any semi-simple group G, π2 (G) = 0 and
hence,
π2 (G/H) ∼
= Ker[π1 (H) → π1 (G)]. (21.20)
Now since for a simply connected group π1 (G) = 0 we find (21.18). Let us
describe now several cases of physical interest:
r The ’t Hooft–Polyakov solution that will be discussed in the next section, is
slightly different since in that case G = SO(3) which is not simply connected
π1 (SO(3)) = Z2 and hence only the even elements of π1 (H = U (1)) are in the
kernel of the homomorphism of (21.20). This is the source of the fact that
the quantization condition is twice the one given by Dirac (see (21.27)) even
though in both cases H = U (1).
r For a simply connected G of rank r and with H which is the full Cartan sub-
algebra, namely H = U (1)r , the homotopy group that classifies the magnetic
monopoles is π1 (H = U (1)r ) = Z r .
r In the spontaneous symmetry breaking of the electro-weak theory we have
G = SU (2) × U (1) and H = U (1) such that G/H ∼ = S 3 . Since π2 (S 3 ) = 0 mag-
netic monopoles are excluded in this theory.
r On the other hand a wide class of grand unified theories do admit mag-
netic monopoles. The most prominent example is the G = SU (5) grand unified
model with H = SU (3) × SU (2) × U (1). This is an example of the case that
H = U (1) × K where K is a semi-simple and simply connected group. In this
case there is only a single component of the magnetic charge that is topologi-
cally conserved.
r Another interesting scenario is the case where the group G twice undergoes
a spontaneous symmetry breaking namely, G → H1 ⊂ G → H2 ⊂ H1 . This is
relevant to an evolution of the universe where at as early stage magnetic
1 The reader who is not familiar with the notion of an exact sequence can refer to any text
book on topology or alternatively to the book of Coleman [66] where an elegant proof of this
theorem is presented.
monopoles associated with π2 (G/H1 ) are being created and then the question
is what is their fate when the universe undergoes a second phase transition to
H2 ? This can be determined by an exact sequence similar to (21.19).
J = L + I, (21.21)
where L = −ir × ∇ is the ordinary spatial part of the angular momentum and
I are the generators of the SU (2) gauge group. With this definition of J the
ansatz should obey,
Using this gauge ’t Hooft and Polyakov suggested the following ansatz for the
fields,
1 − u(r)
Aai = iam r̂m Φa = r̂a h(r). (21.23)
er
To write down the equations that determine u(r) and h(r) one can substitute
these expressions into the Lagrangian density (21.1) and vary with respect to
u(r) and h(r). This is easier than the usual procedure of substituting (21.23)
into the equations of motion derived from (21.1). The resulting equations are,
(u2 − 1)u
u − + e2 uh2 = 0
r2
2 2u2 h
h + h − 2 + λ(v 2 − h2 )h = 0. (21.24)
r r
Fig. 21.1. The hedgehog configuration of the Higgs field. The orientation in
isospace is aligned with the position vector.
The primes denote derivatives with respect to r. Finiteness of the energy asso-
ciated with the solution requires that,
Φ0 (∞) = 0 → u(∞) = 0
Di Φ0 (∞) = 0 → h(∞) = v. (21.25)
Similarly requiring that the solutions are non-singular at the origin implies that,
Qualitatively, the profile of the Higgs field is that of a hedgehog, as can be seen
in Fig. 21.1. The orientation in isospace is aligned with the position vector.
Analytic solutions of these equations will be derived in the next section using
a special (BPS) limit. In general one has to solve these equations numerically.
The physical picture that comes out from these calculations is that there is a
central core of radius Rcore ∼ ev
1
, outside of which u(r) and |h − u|(r) decrease
exponentially. The mass of the monopole takes the form M = 4πv λ
e f ( e 2 ) where
f (0) = 1 and f (∞) = 1.787.
eg = 2πn, (21.27)
γ2
γ1
one unfixed global gauge U (1) phase. This adds up to the parameters associated
with the translation of the monopole. Infinitesimal transformations of this set of
four parameters generate field variations δAi and δΦ that preserve the equations
of motion and leave the energy unchanged. These variations in general will be
referred to as zero modes.
Time-dependent excitations of the translational zero modes can be derived by
substituting r → r − vt into the static solution. This has to be done together
with ensuring the Gauss law constraint,
which also implies that for most choices of gauge A0 = 0. Substituting the solu-
tions of the Gauss law into the expression of the energy (21.15) one can show
that the change of energy for non-relativistic velocities is as one expects, given
by,
1
ΔE = M |v|2 . (21.33)
2
Next we want to describe the excitation of the zero mode associated with the
fourth parameter, that of the global gauge phase. The Noether charge associated
with this transformation is the electric charge that corresponds to the unbroken
U (1) gauge symmetry. In terms of the physical variables defined in (21.6) this
takes the form,
∗
QE = −ie d3 x[W j D0 (Wj − Dj W0 ) − W j D0 (Wj∗ − Dj W0∗ )], (21.34)
where the U (1) covariant derivative is defined by Dμ Wν = (∂μ − ieAμ )Wν and
where a string gauge is used where the Higgs field direction is uniform.
For this case the static field equations, which are the analog of (21.24) become,
2 2u2 h
h + h − 2 + λ(v 2 − h2 )h = 0,
r r
(u 2
− 1)u
u − − e2 u(h2 − j 2 ) = 0,
r2
2 2u2 j
j + j − 2 = 0, (21.38)
r r
where the first equation is identical to the one in (21.24), in the second there is
a 2e2 uj 2 addition and the third equation is the Gauss law.
The dependence of QE on w can be determined by substituting the ansatz
into (21.34) and recalling that W0 = 0 we find,
8πw j(r)
QE = dru(r)2 ≡ Iw. (21.39)
e j(∞)
The integral can be estimated since u(r) falls off exponentially outside a region
of radius ∼ 1/v so that,
4π ¯
I= I, (21.40)
e2 v
where I¯ is of order unity. In analogy to (21.33) one can show that the correction
to the energy is,
Q2E e 2 Q2 M Q2E
ΔE = = E
¯ ∼ M. (21.41)
2eI 4π 2If 2Q2M
The passage from the first line to the rest is of course an identity for arbitrary
α. Next we perform an integration by parts in the last line and use the Gauss
law D · E − ie[Φ, D0 Φ] = 0 and D · B = 0 we find,
E = d3 xTr[(E ± sinαDΦ)2 + (B ± cosαDΦ)2 + (D0 Φ)2
where the magnetic and electric charges QM = vQM and QE = vQE are given
by (21.14) and (21.35)
QM = 2 d2 S · Tr[ΦB] QE = 2 d2 S · Tr[ΦE]. (21.45)
Recall that so far α is arbitrary. The most stringent bound is achieved when
QE
one takes tan α = Q M
, for which the bound reads,
E ≥ Q2M + Q2E . (21.46)
for given magnetic and electric charges QM , QE respectively and hence are also
solutions of the (second-order) equations of motion of the system. In particular
the magnetic monopole which carries no electric charge and is static and hence
QE = 0 and D0 Φ = 0 (in the A0 = 0 gauge), obeys the Bogomolny equation,
B = DΦ. (21.49)
The BPS limit seems to be unnatural and artificial since we have introduced a
potential to give a non-trivial expectation value to the field Φ and then we tuned
the potential to zero keeping the expectation value. It turns out that in certain
suspersymmetric models the BPS equations follow from the requirement of the
invariance under supersymmetry. Since supersymmetry is not discussed in this
book we refer the reader to the literature, for instance [214].
If we go back to the equations of motion and substitute λ = 0 the equations
take the form,
2 u2 h
h + h − 2 2 = 0,
r r
u(u 2
− 1)
u − − e2 uh2 = 0. (21.50)
r2
The solution of this set of equations is given by,
evr 1
u(r) = h(r) = v coth(evr) − (21.51)
sinh(evr) er
The fact that Φ falls off as 1/r and not exponentially is due to the fact that in
the BPS limit it has a vanishing mass and hence associates with a long range
force.
The BPS equations can also be solved for the case of a dyon, namely a config-
uration that carries both a magnetic as well as an electric charge. In that case
the solution reads,
eṽr
u(r) =
sinh(eṽr)
Q2M + Q2E 1
h(r) = ṽ coth(eṽr) −
QM er
QM 1
j(r) = − ṽ coth(eṽr) − . (21.52)
QE er
Mass QE QM
photon 0 0 0
φ 0 0 0
W± ev ±e 0
4 πv
Monopole e
0 ± 4eπ
3. It can be proved that for the SU (2) case there are only two normalizable
solutions wa (s, r). The space-time fields are given in terms of these as follows,
v /2
Φab (r) = dswa† (s, r)wb (s, r),
−v /2
v /2
Aab
j (r) = dswa† (s, r)∂j wb (s, r). (21.60)
−v /2
We do not bring here the the details of the proof of this construction (see for
instance [71]) we just mention that it includes the following elements: (i) Using
of the solution. Thus we end up with one non-trivial parameter. Intuitively this
parameter should relate to the separation distance between the two centers of
the solution. Let us see if the Nahm construction verifies this intuition and to
what extent one can write an explicit solution for this case. The Ti (s) which now
are not constants can be decomposed into the following form,
1 v
Ti (s) = T (s) · σ + Tis (s)I2 . (21.66)
2 i
Substituting this into the Nahm equation implies that the Tiv (s) have to obey,
dTiv (s) 1
= ij k Tjv (s) × Tkv (s). (21.67)
ds 2
It is easy to realize that there is a relation between the Tkv (s), namely the fol-
lowing matrix,
1
Tij = Tiv (s) · Tjv (s) − δij Tkv (s) · Tkv (s), (21.68)
3
is s independent. After some tedious algebra one can show that the most general
form of Tkv (s) takes the form,
1
Tiv (s) = Aij fj (s + v/2, κ, d)τj + Tis I2 , (21.69)
2 i
2 The Euler top function can be expressed in terms of the elliptic functions SN κ (x), C N κ and
DN κ (x) as follows:
C N (D x )
f1 (x, κ, D) = −D S N κ (D x )
κ
D N κ (D x )
f2 (x, κ, D) = −D S N κ (D x )
f3 (x, κ, D) = −D S N 1(D x ) (21.70)
κ
For monopoles that are separated by large distances the task of determining
the metric of the moduli is much easier. Consider a system of N fundamental
monopoles all well separated from each other. In such a layout only abelian inter-
actions are relevant and there is an enhanced gauge symmetry. Instead of having
a conserved electric charge for each unbroken U (1), there is an effective con-
served charge for each monopole core. The moduli space is spanned in this case
by 3N coordinates of the positions of the cores and N angles ξj , j = 1, . . . , N .
The enhance symmetry is the translation along each of the ξj . The approximated
Lagrangian then takes the form,
1 1
L = Mij (x)ẋi · ẋj + g̃ij (x)[ξ˙i + wki (x)ẋk ][ξ˙j + wli (x)ẋl ]. (21.76)
2 2
By computing the pairwise interactions between the separated dyons, one can
determine the functions Mij (x), g̃ij (x) and wji (x). We refer the reader to [214]
for the derivation and we cite here the results,
4π α ∗ ·α ∗
Mij = mi − k = 1 e 2 ri i k k , i = j (21.77)
∗
4π α ·α ∗
Mij = k = 1 e 2 ri i j j , i = j (21.78)
Wij = − k = 1 αi∗ · αk∗ wik , i = j (21.79)
Wij = αi∗ · αi∗ wij , i=j (21.80)
2
−1
and K = (4π)
e4 M . It can be further shown that the metric of the moduli space
of a two monopole BPS solution can be determined exactly and it takes the
form of a Taub–Nut metric or Atiya–Hitchin metric [19] depending whether the
monopoles are distinct or the same. This is beyond the scope of this book and
we refer the reader to for instance the review [214].
In Chapter 5 we saw that solutions of the classical equations of motion, which are
characterized by a topological number, play an important role in two-dimensional
QFT. Derick’s theorem (5.36) forbids scalar field soliton solutions in higher than
two-dimensional space-time. However, for gauge fields one can bypass the theo-
rem, and indeed, as we have seen in Chapter 21, there are solitons in the form of
magnetic monopoles in four-dimensional gauge theories. The topic of this chap-
ter will be solutions of the Yang–Mills theory defined on a Euclidean space-time
which have finite action and are topological in their nature, the instantons. We
will start with a description of the basic properties of one instanton solution
including the topological charge that characterizes it. We then describe the con-
struction of multi-instanton solutions and the moduli space of instantons includ-
ing its dimension, complex nature, singularities and symmetries. When Wick
rotated to Minkowski space-time the instanton describes a tunneling process
between different vacua. We will elaborate on this phenomenon in the context of
the four-dimensional YM theory. Various properties of QCD and hadron physics
were thought to be related to instantons. In certain cases like confinement, the
relation to instantons is still a mystery. One case where the role of instantons
is clear is the U (1A ) problem. This will be described in the last section of this
chapter.
The one instanton solution was derived in [32]. The basic properties of instan-
tons were worked out by many authors including [125], [57]. The instantons of
SU (N ) gauge symmetry were derived in [218]. There are several review papers
such as [65], [208] and [189] and books [182] and [188] that describe the basic
instanton solutions.
we found,
1 −12xμ
Q = dσμ Kμ = − dσμ
24π2 |x4 |
1 xμ 1
= dΩxμ |x|2 4 = dΩ = 1. (22.6)
2π2 |x| 2π 2
Thus we have shown that Q is indeed the winding number that measures how
many times we wind Sg3 when we integrate over Ss3 .
Let us now return to (22.1). Since it is a sum of a topological charge and a
positive semi-definite quantity, it is clear that it is minimized when the latter
vanishes namely,
Dμ Fμν = 0. (22.8)
A solution of the equation of motion is not necessary self-dual but it can be shown
that the non-self-dual configurations are saddle points and not local minima of
the action.
Comparing the expression of the action (22.1) and the topological charge (22.2)
it is clear that a (anti) self-dual configuration that carries an instanton number
(Q = −k), Q = k has an action of,
8π2
S= |k|. (22.9)
g2
One can add the topological charge as an additional term to the action. To be
more precise one adds a θ term,
θ
Sθ = i d4 x Tr [F μν ∗ Fμν ] = iθk. (22.10)
16π2
Whereas the ordinary YM action is the same for the instanton and anti-instanton,
the θ term obviously distinguishes between them by assigning opposite charges
to them. We will further discuss the theta term in Section 22.5. For the self-dual
solution up to a constant the action is equal to the topological charge which by
definition does not depend on the metric. This exhibits the topological nature of
the instanton. Another indication of this nature is the fact that it has vanishing
energy-momentum tensor as follows from,
2 1
Tμν = − 2 Tr Fμρ Fν ρ − δμν Fα β F α β = 0. (22.11)
g 4
This clearly implies that instantons do not curve the space-time they
reside in.
The one instanton solution for the SU (2) can be constructed from U (x) given
in (22.5) via,
ρ2
Aaμ (x) = U −1 ∂μ U , (22.12)
(x − X)2 + ρ2
which yields the explicit form,
a
ημν (x − X)ν
Aaμ (x) = 2 , (22.13)
(x − X)2 + ρ2
a
where ημν is the ’t Hooft antisymmetric symbol defined by,
a
ημν = aμν μ, ν = 1, 2, 3 a
ημ4 = −η4μ
a
= δμa
a
η̄μν = aμν μ, ν = 1, 2, 3 a
η̄μ4 = −η̄4μ
a
= −δμa . (22.14)
a a
It is easy to check that ημν and η̄μν are self-dual and anti self-dual respectively.
The corresponding field strength takes the form,
ρ2
a
Fμν = −4ημν
a
. (22.15)
[(x − X)2 + ρ2 ]2
a a
Obviously since ημν is self-dual so is Fμν .
The one instanton solution is characterized by eight parameters, four corre-
spond to the center of the instanton Xμ , one to the size of the instanton ρ and
three to three global SU (2) gauge transformations. Recall that fixing a guage
we fix only the local gauge transformations. The space of parameters of the k
instanton solutions will be further addressed in Section 22.3 where it will be
shown that in general for SU (N ) the dimension of the moduli space is 4kN .
The instanton solution (22.13) falls off asymptotically as x1 and hence it con-
tributes a finite amount to the integral of the topological charge (winding num-
ber) which is of the form A3 x3 dΩ. The field strength falls off as 1/x4 such
that indeed the corresponding action is finite. However due to the x1 asymptotic
behavior it is difficult to form square integrable expressions that contain it. For
that purpose one can use the following singular gauge transformation,
σμ† (x − X)μ
U= , (22.16)
|x − X|
1
which renders the instanton to have a x3 fall off as can be seen from,
1 2ρ (x − X)ν η̄νa μ
2
Aaμ = . (22.17)
g (x − X)2 [(x − X)2 + ρ2 ]
The singular instanton is obviously singular at the location of the instanton
xμ = Xμ . This singularity is not physical and can be removed by a gauge trans-
formation or by puncturing the Euclidean space with the singular point being
removed.
Instantons of SU (N ) gauge theory can be constructed by embedding SU (2)
instantons in SU (N ) for instance,
S U (N ) 0 0
Aμ = S U (2) , (22.18)
0 Aμ
where the instanton is the 2 × 2 matrix on the lower right. The most general
SU (N ) one instanton configuration can be derived from (22.18) by the following
transformation,
† 0 0 SU (N )
S U (N )
Aμ =U S U (2) U U∈ . (22.19)
0 Aμ SU (N − 2) × U (1)
3 Multi-instanton solutions were presented in [220], [132] and other papers. Our discussion of
the construction of multi-instanton is based on the paper of ADHM [20]. This approach was
further discussed in [70]. We follow the description of the construction given in [81].
where we made use of the fact that a must be Hermitian. In this canonical
parametrization the matrix f reads,
α̇
f = 2((↠) âα̇ + (aμ + xμ 1[k ]×[k ] )2 )−1 . (22.35)
The field strength Fμν that corresponds to the ADHM k instanton configura-
tion (22.26) can be written in the following form,
1 1
Fμν = ∂μ Aν − ∂ν Aμ + g[Aμ , Aν ] = ∂[μ (U † ∂ν ] U ) + (U † ∂[μ U )(U † ∂ν ] U )
g g
1 1
= ∂[μ U † (1 − U † U )∂ν ] U ) = ∂[μ U † Δf Δ† ∂ν ] U
g g
1 † 1 1
= U ∂[μ Δf ∂ν ] Δ† U = U † bσ[μ σ̄ν ] f b+ U = 4 U † bσμν f b+ U, (22.36)
g g g
where we have made use of,
J α̇ J
PIJ ≡ UIi U † = δIJ − ΔII α̇ fI J Δ† J . (22.37)
We next show that for the particular case of k = 1 the ADHM solution (22.26)
is identical to (22.25). For k = 1 we have to drop the indices I, J. One can verify
that in that case the parameters aμ can be identified with the center of the
instanton Xμ . From the ADHM constraint (22.34) we get that,
β̇
(↠) âα̇ = ρ2 δα̇β̇ , (22.40)
and,
1
f= , (22.41)
(xμ − Xμ )2 + ρ2
where ρ will naturally be the size of the instanton.
From the relation (22.39) we deduce that,
?
1 (x − X)2
Û = 1[N ]×[N ] + 2 − 1 (↠)β̇ âα̇ ,
ρ (x − X)2 + ρ2
(x − X)α α̇ (↠)α̇
U = − . (22.42)
|x − X| (x − X)2 + ρ2
Plugging these expressions into U we finally get the singular form of the one
instanton solution (22.25).
In general for k = 1 one can show that the gauge configuration given in (22.36)
indeed carries an instanton of charge k. To derive this result one makes use of
the following relation,
g 2 Tr [Fμν
2
] = ∂μ ∂ μ trk [log f ]. (22.43)
This relation can be proven by expanding the two sides of the equations using
the explicit expression for Fμν (22.36). Upon integrating (22.43) over the whole
1
Euclidean space-time divided by 16π 2 and making use of the fact that asymp-
where we have made use of an integration by parts to derive the last expression.
α̇ α
In the quaternionic notation this condition takes the form of D † δAα α̇ = 0
which combined with (22.45) is given by,
α̇ α
D† δAα β̇ = 0. (22.47)
4 The properties of the moduli space of instantons were discussed by many authors. In partic-
ular [155], [141] and [80]. The review about the moduli space that we are using is [81].
The fluctuation that obeys this equation is referred to as a zero mode since it is
a physical fluctuation that does not change the action. Note that this is exactly
the equation of motion of a Weyl spinor in the background of the instanton
Aμ (x). The zero modes defined by (22.47) are the collective coordinates of the
instantons.
Since the YM instantons are characterized by the topological charge defined
in (22.2) so is the corresponding moduli space. We thus discuss the moduli space
of instantons of charge k which we denote by Mk .
It can be proven that the moduli space of instantons is a manifold. In fact we
will see below that Mk has some conical singular points associated with zero size
instantons. The coordinates on the moduli space are the collective coordinates
that were just shown to be equivalent to the zero model (22.47). We denote by Xn
the collective coordinates where n = 1, . . . , dim Mk . A trivial set of coordinates
are the space-time coordinates of the center of the instanton Xμ accordingly the
moduli space is a product of the form,
Mk = R4 × M̂k . (22.48)
The collective coordinates Xμ follow from the fact that the instanton solu-
tion breaks the symmetry of the action under space-time translations. There are
other collective coordinates that associate with symmetries of the theory that the
instanton configuration breaks. However, not all symmetries yield non equivalent
collective coordinates and not all the coordinates associate with broken symme-
tries.
From the ADHM construction it follows that the moduli space is identified
with the variable aα̇ subject to the ADHM constraints (22.31) quotiented by the
residual U (K) symmetry transformation (22.32) with,
1[N ]×[N ] 0
A= , B=C C ∈ U (k), (22.49)
0 C1[2]×[2]
This result can be derived also by using an index theorem that counts the zero
modes at a point in the moduli space. In fact as we will see shortly the space Mk
is a hyper-Kahler quotient of the flat space R4k (k +N ) by the U (k) group. The
one instanton solution of SU (2) is indeed characterized by the four coordinates
of its center, its size and three global SU (2) gauge transformations.
The moduli space is a complex manifold. A complex manifold is an even-
dimensional manifold that admits a complex structure I a linear map of the
tangent space to itself such that I2 = 0. There are always local holomorphic
coordinates (Z i , Z̄ i ), i = 1, . . . , n (see Section 1.1) for which,
iδj̄ī 0
I= g = gi j̄ dZ i dZ̄ j̄ w = igi j̄ dZ i ∧ dZ̄ j̄ . (22.52)
0 −iδji
where g is an Hermitian metric and w is referred to as the fundamental 2 form.
In the case that the fundamental 2 form is closed namely dw = 0 it is called the
Kähler form and the associated manifold is a Kähler manifold. The latter is also
characterized by the fact that the complex structure is covariantly constant and
the Kähler metric can be derived from a Kähler potential,
∇μ I = 0 gi j̄ = ∂i ∂j̄ K. (22.53)
The moduli space of instanton is not only a Kähler manifold but in fact a hyper
Kähler manifold which means that it admits three linearly independent complex
structures, I(c) , c = 1, 2, 3 that satisfies the algebra
I(c) I(d) = −δ cd + cde I(e) . (22.54)
The four-dimensional Euclidean space R4 is hyper Kähler and the three complex
structures are,
μν = −ημν
I(c) c
(I · x)α α̇ = ixα β̇ σα̇β̇ , (22.55)
c
where ημν is the ’t Hooft η symbol defined in (22.14) and the expression in the
left-hand side is the quaterionic formulation. Now recall that by the definition
of the zero modes (22.47), if δn Aα α̇ is a zero mode so is also δn Aα α̇ Cα̇β̇ for any
constant matrix C and in particular also to σ and hence if δn Aα α̇ so is also
(I · δn A)α α̇ = iδn Aα α̇ σα̇β̇ . Since the zero modes form a complete set there must
exist Inm such that,
(I · δm A)α α̇ = δn Aα α̇ Inm , (22.56)
from which it implies that Inm satisfies the algebra (22.54).
The Kähler potential which is common to the three complex structures of the
moduli space of instantons takes the form,
g2
K=− d4 xx2 TrFμν F μν . (22.57)
4
Using the form of I(c) on R4 given in (22.52) for instance I (3) associated with
the complex coordinates on R4 ix3 + x4 and ix1 − x2 , we find that,
(I(c) · ∂Z i A)α α̇ = i∂Z i Aα α̇ , (22.58)
for instance for I(3) we get ∂Z i Aα 2 = ∂Z̄ ī Aα 1 = 0. Furthermore the derivative
with the respect to the holomorphic and anti-holomorphic coordinates of the
gauge fields obey the equations of zero modes namely,
δi Aμ ≡ ∂Z i Aμ δ̄ī Aμ ≡ ∂Z̄ ī Aμ . (22.59)
Next let us now discuss the symmetries of the moduli space, in particular the
realization of symmetries of the gauge theory which are broken by the instanton
configuration. We start with the four-dimensional conformal group (see Section
17). In the quaternionic formulation the basic variable of the ADHM construction
Δ is transformed as follows,
AB
x → x = (Ax + B)(cX + D)−1 det =1
CD
Δ(x; a, b) → Δ(x ; a, b) = Δ(x; aD + bB, aC + bA)(Cx + D)−1 (22.62)
In fact the term (Cx + D)−1 in the right-hand side of the last equation is irrele-
vant since the gauge field depends on U and U † defined in (22.26) is redundant.
We can now use transformations of (22.32) that keeps the canonical structure
of b (22.33). Upon applying this transformation a goes into,
a → A(aD + bB)B −1 . (22.63)
A particular example of transformations which belong to the conformal group
are the translations. For this case,
Δ(x; a, b) → Δ(x; a + b, b), (22.64)
from which it follows that,
aμ → aμ + μ 1[k ]×[k ] âα̇ → âα̇ . (22.65)
It is thus clear that indeed the components aμ are proportional to the coordinates
of the center of the instanton,
trk aμ = kXμ . (22.66)
Global gauge transformations act non trivially on the ADHM variables if N ≤
2k, while if N ≥ 2k there are transformations that leave the instantons fixed.
This is the stability group of the instanton. One can embed the k instanton
solution in an SU (2k) subgroup of SU (N ) and show that the stability group is
S(U (N − 2k) × U (1)).
The moduli space M̂k is in fact not a smooth manifold due to certain singu-
larities. However, these singularities do not signal any pathology of the moduli
space and integrals over the moduli space are well defined. It can be shown that
M̂k is a cone. For the moduli space of single instanton k = 1 the apex of the
cone is the point ρ = 0 where the instanton has a zero size. This structure can
be generalized also to the k = 1 instantons.
Thus the vacuum gauge configuration is that of a pure gauge. Prior to a dis-
cussion of how to tunnel between two vacuum states, we have to classify and
enumerate the vacua namely following (22.68) the group elements U (x). This
translates to the equivalence classes of maps from S 3 to the SU (N ) group man-
ifold. This is done by the topological charge or winding number or Pontryagin
number defined in (22.2). Since this step is very essential in the discussion of the
tunneling let us clarify this point. Let us analyze the tunneling between a vacuum
state Ai (x, t1 ) at t = t1 into another vacuum state Ai (x, t2 ) at t = t2 . On top of
fixing A0 (x, t) = 0 we can use the residual gauge symmetry to set Ai (x, t1 ) = 0.
Next we consider a path in the space of gauge configurations that connects the
two vacua points and has a finite energy H (in Minkowski space-time). Finite
energy implies that for large |x| → ∞ it has to be a pure gauge,
Aμ →|x|→∞ U ∂μ U † . (22.69)
Aj = e–α je
α
t = t2
Aj = 0
t = t1
Aj = 0
where in the last expression the integration is over the three sphere at t = t2
since at all other parts of the surface of the hyper-cylinder Ai = 0. Thus
the configurations in Minkowski space-time that connect a vacuum state at
t = t1 to one at t = t2 are classified by the winding number of the maps of
S 3 (space) → S 3 (group) just as the maps of instanton in Euclidean space-time.5
In the latter case S 3 (space) is the boundary of R4 whereas in the former it is a
compactification of R3 at t = t2 .
It is easy to realize that there is no way to interpolate a vacuum at t = t1 of
zero winding number with a one at t = t2 with non vanishing winding number
with a configuration of zero energy. The latter corresponds to a pure gauge
configuration which has Fμν = 0 everywhere and hence also vanishing topological
number. Thus the energy of the tunneling configuration as a function of time
should look as in Fig. 22.2.
To identify the configuration that has the largest tunneling rate we consider a
family of gauge configurations characterized by the collective coordinates asso-
ciated with a coordinate transformation from t to λ(t) such that,
(λ)
Ai (x, t) = Ai (x, λ(t)), (22.71)
with the requirement that λ(t1 ) = t1 and λ(t2 ) = t2 . Next we compute the elec-
tric and magnetic fields,
∂Ai
Ei = Fi0 = −∂0 A(λ) (x, t) = (x, λ(t))λ̇
∂λ
1 1
Bi = ij k Fj k = ij k (∂j Ak (x, λ(t)) + Aj (x, λ(t)(Ak (x, λ(t)) − (j ↔ k) (22.72)
2 2
5 The role of instantons in tunneling between different vacua was proposed in [131]. It was also
discussed in [26], [40] and [44]. This topic is reviewed in [185] and in [209]. We follow the
latter.
H (t)
t
t1 t2
The Lagrangian (22.73) is the Lagrangian of a particle that moves from one
vacuum at t = t1 where V (λ) = m(λ) = 0 to a vacuum at t = t2 , where again
V (λ) = m(λ) = 0. The quantum mechanical tunneling rate is proportional to
e−2R where R is given by,
λ2
R= 2m(λ)(V (λ) − E)
dλ
λ1
F
λ2 G G 2
1 ∂A 1
dλH
i
= d3 x Tr d3 x Tr (Bi )2
λ1 g2 ∂λ g2
t2 ?
2 1
= 2 dt d3 x Tr (Ei )2 d3 x Tr (Bi )2 . (22.74)
g t1 g2
Using the triangle inequality we can relate the tunneling rate to the winding or
instanton number as follows,
t2
2 8π2
R ≥ 2| d4 x Tr [E i B i ]| = 2 |k|, (22.75)
g t1 g
where the instanton number is Q = k. Thus we see that the tunneling rate is
bounded by,
2
− 8gπ2 |k |
e−R ≤ e , (22.76)
The generator of large gauge transformation that changes the winding number by
one unit, namely, T |k> = |k + 1> has to be a symmetry generator that commutes
with the Hamiltonian so that T |vac> = eiϕ |vac> for some phase ϕ. Indeed for
the θ vacuum we get T |θ> = k eik θ |k + 1> = e−iθ |θ>.
The energy associated with the θ vacua given by,
This follows from the following steps. Consider the amplitude to tunnel from a
vacuum |i> to a vacuum |j> is given by,
δN −N −(j −i)
<j|e−H t |i> = N± + − (Kte−S )N + +N − , (22.79)
N+ !N− !
when the instantons are sufficiently dilute and where K is the pre-exponential
factor in the tunneling amplitude, and N± are the number of instantons and anti-
instantons. We introduce the parameter θ via a representation of the Kroneker
delta function,
2π
1
δN + −N + + (i−j ) = eiθ (N + −N + +(i−j )) . (22.80)
2π 0
Upon performing the summations over N+ and N− we get,
2π
−S
<j|e−H t |i> = eiθ (i−j ) e2K t cos(θ )e , (22.81)
0
which implies that the energy of the θ vacuum is as given in (22.78). Note that
this does not imply that the YM theory has a continuous spectrum without mass
gap since the θ parameter is fixed for a given theory and it cannot be changed.
Fixing the value of θ can be achieved by adding a θ term to the action (22.10),
1 1 θ
SYM = 2 d x Tr Fμν F → 2 d x Tr Fμν F +
4 μν 4 μν
d4 x Tr [Fμν ∗ F μν ].
2g 2g 16π2
(22.82)
The additional θ term is a surface term and hence does not affect the equations
of motion, however it is not invariant under CP or T transformations.6 As will be
discussed below, with no massless quarks indeed the θ term implies a strong CP
violation. The most severe restriction on CP violation comes from the electric
dipole moment of the neutron. This sets the upper bound to theta to be,
θ < 10−9 . (22.83)
The puzzle of why θ is so tiny is referred to as the strong CP problem. One
proposal to handle this problem is the introduction of the axion, χ, a pseudo
scalar field with a coupling of the form χ Tr [Fμν ∗ F μν ] so that the effective θ is
√
the sum of < χ > and the θ term. As will be discussed below there is in fact
an even simpler mechanism to resolve the strong CP problem and that is having
a massless u quark. This brings us to the next topic which is the incorporation
of light quarks to the game.
In the presence of light quarks there is a simple physical observable
that distinguishes between the different topological vacua, the axial current.
Recall that for Nf massless quarks the theory is classically invariant under
global UL (Nf ) × UR (Nf ) ≡ SUL (Nf ) × SUR (Nf ) × UB (1) × UA (1) symmetry.
The SU (Nf ) × UV (1) symmetry group factors are realized in nature also quan-
tum mechanically. The invariance under the axial SU (Nf ) transformations is
broken spontaneously and there are Nf2 − 1 Goldstone bosons. For Nf = 2 these
are the pions. The UA (1) axial symmetry is not conserved quantum mechanically.
In analogy to the anomaly of the axial symmetry in two dimensions discussed
in Section 9.1, in four dimensions as well one can show using various different
methods that,
Nf
∂ μ Jμ5 = Tr [Fμν ∗ F μν ], (22.84)
8π2
where Jμ5 = i ψ̄i γμ γ5 ψi is the axial current and ψi i = 1, . . . , Nf are the fields of
the various flavored quarks. This resolves the so-called UA (1) puzzle, namely the
absence of the fourth Goldstone boson for Nf = 2 or the ninth one for Nf = 3.
Indeed for the former case one could associate the η pseudo-scalar meson with
the fourth Goldstone boson, however it has a mass of 478M eV whereas a current
6 A proposal for resolving the strong CP problem was proposed in [172]. The U A (1) problem
has been resolved by ’t Hooft. The mass of the η was proposed by Witten [221] and by
Veneziano [216]. Our discussion of this topic follows the review [185].
where S(x, y) is the fermion propagator S(x, y) = <x|(i D )−1 |y> that can be
determined from the eigenfunction equation i D ψλ = λψλ in the form S(x, y) =
ψ λ (x)ψ λ† (y )
λ . Substituting this expression we get,
λ
ψλ (x)ψ † (y)
ΔQ5 = Nf 4 μ
d x∂ Tr λ
2λγ5 = 2Nf (nL − nR ), (22.86)
λ
λ
where we have used the fact that ψλ and γ5 ψλ are orthogonal so only the nL (nR )
left (right) zero modes contribute.
Integrating the left-hand side of (22.84) we get the topological charge Q which
is thus related to ΔQ5 . The latter counts the number of left-handed zero modes
minus the number of right-handed zero modes. This is obviously associated with
instantons. Each instanton contributes one unit to the topological charge and
has a left-handed zero mode, whereas an anti-instanton has a right-handed zero
mode and Q = −1. This is the way the instantons contribute to the axial anomaly
and hence to the resolution of the UA (1) problem. For the case of Nf = 3 this
implies that this would be the ninth Goldstone boson, the η is massive even if
the quark masses vanish. It was shown that the mass of the η is related to the
topological susceptibility in the following form,
2Nf 2Nf
χ top = d4 x <Q(x)Q(0)>= m2η + m2η − 2m2K . (22.87)
fπ2 fπ2
The combination of masses on the right-hand side corresponds to the part of the
η mass which is not due to the strange quark mass.
23.1 General
Relativistic quantum field theory has been very successful in describing strong,
electromagnetic and weak interactions, in the region of small couplings by per-
turbation theory, within the framework of the standard model.
However, the region of strong coupling, like the hadronic spectrum and various
scattering phenomena of hadrons within QCD, is still largely unsolved.
A large variety of methods have been used to address this question, includ-
ing lattice gauge simulations, light-cone quantization, low energy effective
Lagrangians like the Skyrme model and chiral Lagrangians, large N approxima-
tion, techniques of conformal invariance, the integrable model approach, super-
symmetric models, string theory approach, QCD sum rules, etc. In spite of this
major effort the gap between the phenomenology and the basic theory has only
been partially bridged, and the problem is still open.
The goals of this book are to provide a detailed description of the tool box
of non-perturbative techniques, to apply them on simplified systems, mainly of
gauge dynamics in two dimensions, and to examine the lessons one can learn
from those systems about four-dimensional QCD and hadron physics.
The study of two-dimensional problems to improve the understanding of four-
dimensional physical systems was found to be fruitful. This follows two directions,
one is the utilization of non-perturbative methods on simpler setups and the
second is extracting the physical behavior of hadrons in one space dimension.
Obviously, physics in two dimensions is simpler than that of the real world
since the underlying manifold is simpler and since the number of degrees of
freedom of each field is smaller. There are some additional simplifying features
in two-dimensional physics. In one space dimension there is no rotation symmetry
and no angular momentum. The light-cone is disconnected and is composed of
left moving and right moving branches. Therefore, massless particles are either
on one branch or the other. These two properties are the basic building blocks
of the idea of transmutation between systems of different statistics. Also, the
ultra-violet behavior is more convergent in two dimensions, making for instance
QCD2 a superconvergent theory.
In this summary chapter we go over several notions, concepts and methods
with emphasis on the comparison between the two- and four-dimensional worlds
and what one can deduce about the latter from the former. In particular we deal
where the definitions of the various quantities are in Chapter 17. Again there
is a striking difference between the simple formula in two dimensions and the
complicated one in four dimensions.
r As an example let us compare the OPE of two currents. Recall from the dis-
cussion of Chapter 3 the expression in two dimensions reads,
kδ ab fcab J c (w)
J a (z)J b (w) = + i + finite terms, (23.5)
(z − w)2 (z − w)
for any non-abelian group, and in particular for the abelian case the second
term on the right-hand side is missing. For comparison the OPE of the trans-
verse components of the electromagnetic currents given in Chapter 17 takes
the form,
J T (x)J T (0) ∼
∞ 1 (6−t n )/2 1
n = 0 Cn x 2 (−ix− )n + 1 Γ(jΓ(2j n)
n )Γ(j n ) 0
du[u(1 − u)]j n −1 Qn1,1 (ux− ).
(23.6)
r The conformal Ward identity associated with the dilatation operator in four
dimensions (17.60),
N
(lφ + γ(g ∗ ) + xi ∂i ) <T φ(x1 )...φ(xN )>= 0, (23.7)
i
where lφ is the canonical dimension and γ(g ∗ ) is the anomalous dimension,
seems very similar to the one in two dimensions,
(zi ∂i + hi ) <0|φ1 (z1 , z̄1 )...φn (zn , z̄n )|0>= 0. (23.8)
i
In both cases one has to determine the full quantum conformal dimensions of
the various operators. However, as was shown in Section 2.7, in certain CFT
models, like the unitary minimal models, there are powerful tools based on
unitarity which enable us to determine exactly the dimensions hi of all the
primary operators and hence all the operators of the model. On the other
hand, it is a non-trivial task to determine the anomalous dimensions in other
models in two dimensions, and of course four-dimensional operators. In certain
supersymmetric theories there are operators whose dimension is protected, but
generically one has to use perturbative calculations to determine the anomalous
dimensions of gauge theories to a given order in the coupling constant.
Using the Ward identity one can extract the form of the two-point function
of operators of spin s in four dimensions. It is given by,
l φ +γ (g ∗ ) s
∗ ∗ −2γ (g ∗ ) 1 (x1 − x2 )+
<φ(x1 )φ(x2 )>= N2 (g )(μ ) .
(x1 − x2 )2 (x1 − x2 )−
(23.9)
The corresponding two-point function in two dimensions, which depends only
on the conformal dimension of the operator h, reads,
c2
G2 (z1 , z̄1 , z2 , z̄2 ) ≡<0|φ1 (z1 , z̄1 )φ1 (z2 , z̄2 )|0> = .
(z1 − z2 ) (z̄1 − z̄2 )2 h̄ 1
2h 1
(23.10)
r As for higher point functions, we have seen in Section 2.9 that one can use the
local Ward identities together with Virasoro null vectors to write down partial
differential equations. The result for a four-point function (2.63) was later used
to determine the four-point function of the Ising model (2.94). Two dimensional
conformal field theories are further invariant under affine Lie algebra transfor-
mations, and as we have shown in Section 3.6 those can be combined with
null vectors to derive the so-called Knizhnik–Zamolodchikov equations (3.69),
which were later used to solve for the four-point function of the SU (N ) WZW
model in Section 4.4. These types of differential equations that fully determine
correlation functions are obviously absent in four-dimensional interacting con-
formal field theories.
23.3 Integrability
Integrability was discussed in Chapter 5 in the context of two-dimensional mod-
els and in Chapter 18 in four-dimensional gauge theories. For systems with a
finite number of degrees of freedom, like spin chain models, there is a finite num-
ber of conserved charges, equal of course to the number of degrees of freedom.
For integrable field theories there is an infinite countable number of conserved
charges. Furthermore, the scattering processes of those models always involve a
conservation of the number of particles.
In two dimensions we have encountered continuous integrable models like the
sine-Gordon model as well as discretized ones like the XXX spin chain model.
The integrable sectors of gauge dynamical systems discussed in Chapter 18 are
based on identifying an exact map between certain properties of the systems and
a spin chain structure. In two dimensions the spin chain models follow from a
discretization of the space coordinate, by placing a spin variable on each site
that can take several values and imposing periodicity. In the four-dimensional
N = 4 super YM theory discussed in Section 18.1 the spin chain corresponds to
a trace of field operators and in the process of high-energy scattering of Section
18.2 it is a “chain” of reggeized gluons exchanged in the t-channel of a scattering
process. A summary of the comparison among the basic two-dimensional spin
chain, the “spin chains” associated with the planar N = 4 SYM, and the high-
energy scattering in QCD, is given in Table 23.1. A powerful method to solve
all these spin chain models is the use of the algebraic Bethe ansatz. This was
discussed in detail for the the XXX1/2 model in Section 5.14. The solutions of
the energy eigenvalues needed for the high-energy scattering process was based
on generalizing this method to the case of spin s Heisenberg model (see Section
18.2) and for the N = 4 to the case of an SO(6) invariance.
There is one conceptual difference between the spin chains of the two-
dimensional models and those associated with the N = 4 SYM in four dimen-
sions. In the former the models are non conformal, involving a scale, and hence
Table 23.1. Spin chain structure of the two-dimensional model and the
four-dimensional gauge systems of N = 4 SYM and of high-energy behavior
of scattering amplitudes in QCD.
23.4 Bosonization
Bosonization is the formulation of fermionic systems in terms of bosonic variables
and fermionization is just the opposite process. The study of bosonized physical
systems offers several advantages:
(1) It is usually easier to deal with commuting fields rather than anti-commuting
ones.
(2) In certain examples, like the Thirring model, the fermionic strong coupling
regime turns into the weak coupling one in its bosonic version, the sine-
Gordon model (see Section 6.2).
(3) The non-abelian bosonization, especially in the product scheme (see Section
6.3.4), offers a separation between colored and flavored degrees of freedom,
which is very convenient for analyzing low lying spectrum.
(4) Baryons composed of NC quarks are a many-body problem in the fermion
language, while simple solitons are in the boson language.
(5) One loop fermionic computations involving the currents turn into tree level
consideration in the bosonized version. The best-known example of the latter
are the chiral (or axial) anomalies (see Section 9.1).
Recall that for a system that admits, for instance an abelian case, also a current
Jμ = ∂μ φ that is conserved upon the use of the equations of motion, one can
then replace the two currents with left and right conserved currents J± = ∂± φ
or J = ∂φ and J¯ = ∂φ,¯ as was discussed in Chapter 1. The charge associated
with the topological conserved current is given by,
Qtop = dxφ = [φ(t, +∞) − φ(t, −∞)] ≡ φ+ − φ− , (23.12)
Table 23.2. Topological classical field configurations in two and four dimensions
r The topological charges, for compact spaces, are the winding numbers of the
corresponding topological configurations. For a compact one space dimension,
we have the map of S 1 → S 1 related to the homotopy group π1 (S 1 ). In two
G /H
space dimensions, the windings are associated with the map S2 → S2 , as
for the magnetic monopoles. For three space dimensions, it is S → S 3 for
3
the Skyrmions at Nf = 2, and the non-abelian instantons for the gauge group
SU (2). The topological data of the various models is summarized in Table 23.2.
r According to Derrick’s theorem (see Section 5.3), for a theory of a scalar field
with an ordinary kinetic term with two derivatives, and any local potential at
D ≥ 2, the only non singular time-independent solutions of finite energy are the
vacua. However, as we have seen in Chapters 20, 21 and 22, there are solitons
in the form of Skyrmions and monopoles and instantons. Those configurations
bypass Derrick’s theorem by introducing higher derivative terms or including
non-abelian gauge fields.
r As was emphasised in Chapter 20, the extraction of the baryonic properties in
the Skyrme model is very similar to the one for the baryons in the bosonized
theory in two dimensions. Unlike the latter which is exact in the strong cou-
pling limit, one cannot derive the former starting from the underlying theory.
Another major difference between the two models is of course the existence of
angular momentum only in the four-dimensional case.
r A non-trivial task associated with topological configurations is the construc-
tion of configurations that carry multipole topological charge, for instance a
multi-baryon state both of the bosonized QCD2 as well as of the Skyrme
model, a multi-monopole solution and a multi-instanton solution. For the
two-dimensional baryons (as discussed in Section 13.6) the construction is a
straightforward generalization of the configuration of baryon number one. For
the multi-monopole solutions we presented Nahm’s construction, and for the
multi-instantons the ADHM construction. These constructions, which are in
fact related, are much more complicated than that for the two-dimensional
muti-baryons.
r A very important phenomenon that occurs in both two and four dimensions
is the strong-weak duality, and the duality between a soliton and an elemen-
tary field. In two dimensions we have encountered this duality in the relation
between the Thirring model and the sine-Gordon model, where the coupling
of the latter β is related to that of the former g as (6.27),
β2 1
= g . (23.14)
4π 1+ π
This also relates the elementary fermion field of the Thirring model with the
soliton of the sine-Gordon model. In particular for g = 0 corresponding to
β 2 = 4π, the Thirring model describes a free Dirac fermion, while the soliton
of the corresponding sine-Gordon theory is the same fermion in its bosonization
disguise. An analog in four dimensions is the Olive–Montonen duality discussed
in Section 21.8, which relates the electric charge e with the magnetic one
±
eM = 4π e , where the former is carried by the elementary states W and the
latter by the magnetic monopoles.
with L in both two and four dimensions. However, that does not explain the
difference between two and four dimensions, it merely means that in two dimen-
sions the coulomb and confining potentials behave in the same manner.2 The
determination of the string tension in two dimensions cannot be repeated in four
dimensions. The reason is that in the latter case the anomaly is not linear in the
gauge field and thus one cannot use the chiral rotation to eliminate the external
quark anti-quark pair. That does not imply that the situation in four dimensions
differs from the two-dimensional one, it just means that one has to use different
methods to compute the string tension in four dimensions.
What are the four-dimensional systems that might resemble the two-
dimensional case of dynamical adjoint matter and external fundamental quarks?
A system with external quarks in the fundamental representation in the context
of pure YM theory seems a possible analog since the dynamical fields, the gluons,
are in the adjoint representation, though they are vector fields and not fermions.
An alternative is the N = 1 SYM where in addition to the gluons there are also
gluinos which are Majorana fermions in the adjoint representation. Both these
cases should correspond to the massless adjoint case in two dimensions. The lat-
ter admits a screening behavior where as the four-dimensional models seem to be
in the confining phase. This statement is supported by several different types of
calculations in particular for the non supersymmetric case this behavior is found
in lattice simulations.
At this point we cannot provide a satisfactory intuitive explanation why the
behavior in two and four dimensions is so different. There is also no simple
picture of how the massless adjoint dynamical quarks in two dimensions are able
to screen external charges in the fundamental representation.
It is worth mentioning that there is ample evidence that four-dimensional
hadronic physics is well described by a string theory. This is based for instance
on realizing that mesons and baryons in nature admit Regge trajectory behavior
which is an indication of a stringy nature. Any string theory is by definition a
two-dimensional theory and hence a very basic relation between four-dimensional
hadron physics and two-dimensional physics.
In addition to the ordinary string tension which relates to the potential
between a quark and anti-quark in the fundamental representation, one defines
the k string that connects a set of k quarks with a set of k anti-quarks. This
object has been examined in four-dimensional YM as well as four-dimensional
N = 1 SYM. These two cases seem to be the analog of the two-dimensional QCD
theory with adjoint quarks and with external quarks in a representation that is
characterized by k boxes in the Young tableau description. In Chapter 14 we have
derived an expression for the string tension as a function of the representation
of the external and dynamical quarks and in particular for dynamical adjoint
2 Note that the linear potential in two dimensions is already there at lowest order, while
obviously in four dimensions it is a highly non-perturbative effect.
23.7.1 Mesons
As was just mentioned the two-dimensional mesonic spectrum was extracted
using the large NC approximation in the fermionic formulation for Nf = 1
(’t Hooft model), the currentization for massless quarks and the DLCQ approach
for both cases of quarks in the fundamental and the adjoint representation. For
the particular region of NC >> Nf and m = 0 the fermionic large Nc and the
currentization treatments yielded identical results. In fact this result is achieved
also using the DLCQ method for adjoint fermions upon a truncation to a single
parton and replacing g 2 with 2g 2 (see (12.42)). For massive fundamental quarks
the DLCQ results match very nicely those of lattice simulations and the large
Nc calculations as can be seen from Figs (12.1) and (12.2).
In all these methods the corresponding equations do not admit exact ana-
lytic solutions for the whole range of parameters and thus one has to resort to
numerical solutions. However, in certain domains one can determine the analytic
behavior of the wavefunctions and masses.
The spectrum of mesons in two dimensions is characterized by the dependence
of the meson masses Mm es on the gauge coupling g, the number of colors Nc ,
the number of flavors Nf , the quark mass mq and the excitation number n. In
four-dimensional QCD the meson spectra depend on the same parameters apart
from the fact that ΛQ C D , the QCD scale, is replacing the two-dimensional gauge
coupling and of course some additional quantum numbers. The following lines
summarize the properties of the spectrum
r The highly excited states n 1, are characterized by,
Mm2 es ∼ πg 2 Nc n. (23.19)
Mm0 es ∼
= mq 1 + mq 2 , (23.20)
where mq i are the masses of the quark and anti-quark. In the opposite limit of
mq g
5
0 2 ∼ π g 2 Nc
(Mm es ) = (m1 + m2 ). (23.21)
3 π
For the special case of massless quarks we find a massless meson. This is very
reminiscent of the four-dimensional picture for the massless pions. For small
Mm2 es ∼ Nf . (23.23)
23.7.2 Baryons
In Chapter 13 we have described the spectrum of baryons in multiflavor two-
m
dimensional QCD in the strong coupling limit e cq → 0. The four-dimensional
baryonic spectrum was discussed in the large Nc limit in Chapter 19 and using
the Skyrme model approach in Chapter 20. We would like now to compare these
spectra and to investigate the possibility of predicting four-dimensional baryonic
properties from the simpler two-dimensional model. In the former case the mass
is a function of the QCD scale ΛQ C D , the number of colors Nc and the number
ūu Δ+ 1
p 2
3 4 2 5
¯
dd Δ + 1
p 11
3 30
s̄s Δ + 1
6
p 7
30
s̄s Δ+ + 1
6
Δ 7
24
s̄s Ω− 5
24
is no exact analog of the proton. Instead we take the charge = +1Δ+ as the
two-dimensional analog of the proton. In the Skyrme model one can compute
in a similar manner the flavor content of the four-dimensional baryons. The
two- and four-dimensional states compare as is summarized in Table 23.4.
23.8 Outlook
We can imagine future developments associated with the topics covered in the
book in three different directions: Further progress in the application of the
methods discussed in the book to unravel the mysteries of gauge dynamics in
nature; applications of the methods in other domains of physics not related to
four-dimensional gauge theories; and improving our understanding of the strong
interaction and hadron physics due to other non-perturbative techniques that
are not discussed in the book. Let us now briefly fantasize on hypothetical devel-
opments in those three avenues.
one can start with the Seiberg Witten solution of N = 2 [192] where the struc-
ture of vacua is known and extract confinement behavior in N = 1 and non
supersymmetric theories.
r A breakthrough in the understanding of gauge theories in the strong coupling
regime took place with the discovery by Maldacena of the AdS/CFT holo-
graphic duality [158]. The strongly coupled N = 4 in the large N and large
’t Hooft parameter λ is mapped into a weakly curved supergravity background.
Thousands of research papers that followed develop this map in many different
directions and in particular also in relation to the pure YM theory and QCD
in four dimensions. There is very little doubt that further exploration of the
duality will shed new light on QCD and on hadron physics.
r String theory has been born as a possible theory of hadron physics. It then
underwent a phase transition into a candidate for the theory of quantum grav-
ity and even a unifying theory for everything. In recent years, mainly due to
the AdS/CFT duality there is a renaissance of the idea that hadrons at low
energies should be described as strings. This presumably combined with the
duality seems to be a useful tool that will improve our understanding of gauge
dynamics.
r The computations of scattering amplitudes in gauge theories has been boosted
in recent years due to various developments including the use of techniques
based on twistors, on a novel T-duality in the context of the Ads/CFT duality
and on a conjectured duality between Wilson lines and scattering amplitudes.
One does not need a wild imagination to foresee further progress in the industry
of computing scattering amplitudes.
To summarize, non-perturbative methods have always been very important tools
in exploring the physical world. We have no doubt that they will continue to be a
very essential ingredient in future developments of science in general and physics
in particular.
[1] E. Abdalla and M. C. B. Abdalla, “Updating QCD in two-dimensions,” Phys. Rept. 265,
253 (1996) [arXiv:hep-th/9503002].
[2] E. Abdalla, M. C. B. Abdalla and K. D. Rothe, Nonperturbative Methods in Two-
Dimensional Quantum Field Theory, Singapore: World Scientific (2001).
[3] A. Abrashkin, Y. Frishman and J. Sonnenschein, “The spectrum of states with one
current acting on the adjoint vacuum of massless QCD2,” Nucl. Phys. B 703, 320 (2004)
[arXiv:hep-th/0405165].
[4] C. Adam, “Charge screening and confinement in the massive Schwinger model,” Phys.
Lett. B 394, 161 (1997) [arXiv:hep-th/9609155].
[5] G. S. Adkins, C. R. Nappi and E. Witten, “Static properties of nucleons in the Skyrme
model,” Nucl. Phys. B 228, 552 (1983).
[6] S. L. Adler, J. C. Collins and A. Duncan, “Energy momentum tensor trace anomaly in
spin 1/2 quantum electrodynamics’” Phys. Rev. D 15, 1712 (1977).
[7] I. Affleck, “On the realization of chiral symmetry in (1+1)-dimensions,” Nucl. Phys. B
265, 448 (1986).
[8] O. Aharony, O. Ganor, J. Sonnenschein and S. Yankielowicz, “On the twisted G/H
topological models,” Nucl. Phys. B 399, 560 (1993) [arXiv:hep-th/9208040].
[9] O. Aharony, O. Ganor, J. Sonnenschein, S. Yankielowicz and N. Sochen, “Physical states
in G/G models and 2-d gravity,” Nucl. Phys. B 399, 527 (1993) [arXiv:hep-th/9204095].
[10] O. Aharony, S. S. Gubser, J. M. Maldacena, H. Ooguri and Y. Oz, “Large N field theories,
string theory and gravity,” Phys. Rept. 323, 183 (2000) [arXiv:hep-th/9905111].
[11] A. Y. Alekseev and V. Schomerus, “D-branes in the WZW model,” Phys. Rev. D 60,
061901 (1999) [arXiv:hep-th/9812193].
[12] D. Altschuler, K. Bardakci and E. Rabinovici, “A construction of the c < 1 modular
invariant partition function,” Commun. Math. Phys. 118, 241 (1988).
[13] L. Alvarez-Gaume, G. Sierra and C. Gomez, “Topics in conformal field theory,” contri-
bution to the Knizhnik Memorial Volume, L. Brink, et al., World Scientific. In Brink, L.
(ed.) et al.: Physics and Mathematics of Strings 16-184 (1989). Singapore.
[14] F. Antonuccio and S. Dalley, “Glueballs from (1+1)-dimensional gauge theories with
transverse degrees of freedom,” Nucl. Phys. B 461, 275 (1996) [arXiv:hep-ph/9506456].
[15] A. Armoni, Y. Frishman and J. Sonnenschein, “The string tension in massive QCD(2),”
Phys. Rev. Lett. 80, 430 (1998) [arXiv:hep-th/9709097].
[16] A. Armoni, Y. Frishman and J. Sonnenschein, “The string tension in two dimensional
gauge theories,” Int. J. Mod. Phys. A 14, 2475 (1999) [arXiv:hep-th/9903153].
[17] A. Armoni, Y. Frishman and J. Sonnenschein, “Massless QCD(2) from current con-
stituents,” Nucl. Phys. B 596, 459 (2001) [arXiv:hep-th/0011043].
[18] A. Armoni and J. Sonnenschein, “Mesonic spectra of bosonized QCD in two-dimensions
models,” Nucl. Phys. B 457, 81 (1995) [arXiv:hep-th/9508006].
[19] M. F. Atiyah and N. J. Hitchin, “The geometry and dynamics of magnetic monopole.
M.B. Porter lectures” Princeton, USA: Princeton University Press (1988) 133p.
[20] M. F. Atiyah, N. J. Hitchin, V. G. Drinfeld and Yu. I. Manin, “Construction of instan-
tons,” Phys. Lett. A 65, 185 (1978).
[21] F. A. Bais, “To be or not to be? Magnetic monopoles in non-abelian gauge theories,”
in ’t Hooft, Hackensack, G. ed. Fifty years of Yang–Mills Theory, New Jersey World
Scientific, C 2005.
[48] S. J. Brodsky, H. C. Pauli and S. S. Pinsky, “Quantum chromodynamics and other field
theories on the light cone,” Phys. Rept. 301, 299 (1998) [arXiv:hep-ph/9705477].
[49] S. J. Brodsky, Y. Frishman, G. P. Lepage and C. T. Sachrajda, “Hadronic wave functions
at short distances and the operator product expansion,” Phys. Lett. B 91, 239 (1980).
[50] S. J. Brodsky, Y. Frishman and G. P. Lepage, “On the application of conformal symmetry
to quantum field theory,” Phys. Lett. B 167, 347 (1986).
[51] S. J. Brodsky, P. Damgaard, Y. Frishman and G. P. Lepage, “Conformal symmetry:
exclusive processes beyond leading order,” Phys. Rev. D 33, 1881 (1986).
[52] R. C. Brower, W. L. Spence and J. H. Weis, “Effects of confinement on analyticity in
two-dimensional QCD,” Phys. Rev. D 18, 499 (1978).
[53] G. E. Brown, ‘Selected papers, with commentary, of Tony Hilton Royle Skyrme,” Sin-
gapore: World Scientific (1994) 438 p. (World Scientific series in 20th century physics,
3).
[54] R. N. Cahn, Semisimple Lie Algebras and their Representations, Menlo Park, USA:
Benjamin/Cummings (1984) 158 p. (Frontiers In Physics, 59)
[55] C. G. Callan, “Broken scale invariance in scalar field theory,” Phys. Rev. D 2, 1541
(1970).
[56] C. G. Callan, N. Coote and D. J. Gross, “Two-dimensional Yang-Mills theory: a model
of quark confinement,” Phys. Rev. D 13, 1649 (1976).
[57] C. G. Callan, R. F. Dashen and D. J. Gross, “Toward a theory of the strong interactions,”
Phys. Rev. D 17, 2717 (1978).
[58] J. L. Cardy, “Boundary conditions, fusion rules and the Verlinde formula,” Nucl. Phys.
B 324, 581 (1989).
[59] J. L. Cardy, “Conformal invariance and statistical mechanics,” Les Houches Summer
School 1988:0169-246.
[60] A. Casher, H. Neuberger and S. Nussinov, “Chromoelectric flux tube model of particle
production,” Phys. Rev. D 20, 179 (1979).
[61] A. Chodos, “Simple connection between conservation laws in the Korteweg-De Vries and
sine-Gordon systems,” Phys. Rev. D 21, 2818 (1980).
[62] E. Cohen, Y. Frishman and D. Gepner, “Bosonization of two-dimensional QCD with
flavor,” Phys. Lett. B 121, 180 (1983).
[63] S. R. Coleman, “Quantum sine-Gordon equation as the massive Thirring model,” Phys.
Rev. D 11, 2088 (1975).
[64] S. R. Coleman, “More about the massive Schwinger model,” Annals Phys. 101, 239
(1976).
[65] S. R. Coleman, “The uses of instantons,” Subnucl. Ser. 15, 805 (1979).
[66] S. Coleman, Aspects of Symmetry, selected Erice lectures of Sidney Coleman, Cambridge,
UK, Cambridge Univ. Press (1985).
[67] S. R. Coleman, The magnetic monopole fifty years later, In the Unity of Fundamental
Interactions, ed. A. Zichichi: New York, Plenum (1983).
[68] S. R. Coleman, R. Jackiw and L. Susskind, “Charge shielding and quark confinement in
the massive Schwinger model,” Annals Phys. 93, 267 (1975).
[69] S. Cordes, G. W. Moore and S. Ramgoolam, “Large N 2-D Yang-Mills theory and topo-
logical string theory,” Commun. Math. Phys. 185, 543 (1997) [arXiv:hep-th/9402107].
[70] E. Corrigan, D. B. Fairlie, S. Templeton and P. Goddard, “A Green’s function for the
general selfdual gauge field,” Nucl. Phys. B 140, 31 (1978).
[71] E. Corrigan and P. Goddard, “Construction of instanton and monopole solutions and
reciprocity,” Annals Phys. 154, 253 (1984).
[72] S. Dalley and I. R. Klebanov, “String spectrum of (1+1)-dimensional large N QCD with
adjoint matter,” Phys. Rev. D 47, 2517 (1993) [arXiv:hep-th/9209049].
[73] R. F. Dashen and Y. Frishman, “Four fermion interactions and scale invariance,” Phys.
Rev. D 11, 2781 (1975).
[74] R. F. Dashen, B. Hasslacher and A. Neveu, “Nonperturbative methods and extended
hadron models in field theory. 2. Two-dimensional models and extended hadrons,” Phys.
Rev. D 10, 4130 (1974).
[75] G. Date, Y. Frishman and J. Sonnenschein, “The spectrum of multiflavor QCD in two-
dimensions’” Nucl. Phys. B 283, 365 (1987).
[76] G. F. Dell-Antonio, Y. Frishman and D. Zwanziger, “Thirring model in terms of currents:
solution and light cone expansions,” Phys. Rev. D 6, 988 (1972).
[77] P. Di Francesco, P. Mathieu, D. Senechal Conformal Field Theory, Series:Graduate Texts
in Contemporary Physics. New York: Springer-Verlag (1997).
[78] P. A. M. Dirac, “Theory Of Magnetic Monopoles,” In *Coral Gables 1976, Proceedings,
New Pathways In High-energy Physics, Vol. I*, New York: Plenum Press, 1976, 1-14.
[79] F. M. Dittes and A. V. Radyushkin, “Two loop contribution to the evolution of the pion
wave function,” Phys. Lett. B 134, 359 (1984); M. H. Sarmadi, Phys. Lett. B 143, 471
(1984); S. V. Mikhailov and A. V. Radyushkin, Nucl. Phys. B 254, 89 (1985); G. R.
Katz, Phys. Rev. D 31, 652 (1985).
[80] N. Dorey, T. J. Hollowood, V. V. Khoze, M. P. Mattis and S. Vandoren, “Multi-instanton
calculus and the AdS/CFT correspondence in N = 4 Nucl. Phys. B 552, 88 (1999)
[arXiv:hep-th/9901128].
[81] N. Dorey, T. J. Hollowood, V. V. Khoze and M. P. Mattis, “The calculus of many
instantons,” Phys. Rept. 371, 231 (2002) [arXiv:hep-th/0206063].
[82] M. R. Douglas, K. Li and M. Staudacher, “Generalized two-dimensional QCD,” Nucl.
Phys. B 420, 118 (1994) [arXiv:hep-th/9401062].
[83] A. V. Efremov and A. V. Radyushkin, “Factorization and asymptotical behavior of pion
form-factor in QCD,” Phys. Lett. B 94, 245 (1980).
[84] T. Eguchi and H. Ooguri, “Chiral bosonization on a Riemann surface,” Phys. Lett. B
187, 127 (1987).
[85] S. Elitzur and G. Sarkissian, “D-branes on a gauged WZW model,” Nucl. Phys. B 625,
166 (2002) [arXiv:hep-th/0108142].
[86] J. R. Ellis, Y. Frishman, A. Hanany and M. Karliner, “Quark solitons as constituents of
hadrons,” Nucl. Phys. B 382, 189 (1992) [arXiv:hep-ph/9204212].
[87] J. R. Ellis, Y. Frishman and M. Karliner, “Meson baryon scattering in QCD(2) for any
coupling,” Phys. Lett. B 566, (2003) 201 [arXiv:hep-ph/0305292].
[88] L. D. Faddeev, “How algebraic Bethe ansatz works for integrable model,” arXiv:hep-
th/9605187.
[89] B. L. Feigin and D. B. Fuchs “Skew symmetric differential operators on the line and
Verma modules over the Virasoro algebra” Funct. Anal. Prilozhen 16, (1982) 47.
[90] S. Ferrara, A. F. Grillo and R. Gatto, “Improved light cone expansion,” Phys. Lett. B
36, 124 (1971) [Phys. Lett. B 38, 188 (1972)].
[91] D. Finkelstein and J. Rubinstein, “Connection between spin, statistics, and kinks,” J.
Math. Phys. 9, 1762 (1968).
[92] R. Floreanini and R. Jackiw, “Selfdual fields as charge density solitons,” Phys. Rev. Lett.
59, 1873 (1987).
[93] D. Friedan, “Introduction To Polyakov’s string theory,” Published in Les Houches Sum-
mer School 1982:0839.
[94] D. Friedan, E. J. Martinec and S. H. Shenker, “Conformal invariance, supersymmetry
and string theory,” Nucl. Phys. B 271, 93 (1986).
[95] D. Friedan, Z. a. Qiu and S. H. Shenker, “Conformal invariance, unitarity and two-
dimensional critical exponents,” Phys. Rev. Lett. 52, 1575 (1984).
[96] Y. Frishman, A. Hanany and J. Sonnenschein, “Subtleties in QCD theory in two-
dimensions,” Nucl. Phys. B 429, 75 (1994) [arXiv:hep-th/9401046].
[97] Y. Frishman and M. Karliner, “Baryon wave functions and strangeness content in QCD
in two-dimensions,” Nucl. Phys. B 344, 393 (1990).
[98] Y. Frishman and M. Karliner, “Scattering and resonances in QCD(2),” Phys. Lett. B
541, 273 (2002). Erratum-ibid. B 562, 367, (2003). [arXiv:hep-ph/0206001].
[99] Y. Frishman and J. Sonnenschein, ”Bosonization of colored-flavored fermions and QCD
in two-dimenstions,” Nucl. Phys. B 294, 801 (1987).
[100] Y. Frishman and J. Sonnenschein, “Gauging of chiral bosonized actions’” Nucl. Phys. B
301, 346 (1988).
[128] K. Hornbostel, S. J. Brodsky and H. C. Pauli, “Light cone quantized QCD in (1+1)-
dimensions,” Phys. Rev. D 41, 3814 (1990).
[129] C. Imbimbo and A. Schwimmer, “The Lagrangian formulation of chiral scalars,” Phys.
Lett. B 193, 455 (1987).
[130] C. Itzykson and J. B. Zuber, Quantum Field Theory, New York, USA: McGraw-Hill
(1980) 705 P.(International Series In Pure and Applied Physics)
[131] R. Jackiw and C. Rebbi, “Vacuum periodicity in a Yang-Mills quantum theory,” Phys.
Rev. Lett. 37, 172 (1976).
[132] R. Jackiw, C. Nohl and C. Rebbi, “Conformal properties of pseudoparticle configura-
tions”, Phys. Rev. D 15, 1642 (1977).
[133] A. D. Jackson and M. Rho, “Baryons as chiral solitons,” Phys. Rev. Lett. 51, 751 (1983).
[134] B. Julia and A. Zee, “Poles with both magnetic and electric charges in nonabelian gauge
theory,” Phys. Rev. D 11, 2227 (1975).
[135] N. Ishibashi, “The boundary and crosscap states in conformal field theories,” Mod. Phys.
Lett. A 4, 251 (1989).
[136] V. G. Kac, “Simple graded algebras of finite growth,” Funct. Anal. Appl. 1, 328 (1967).
[137] L. P. Kadanoff, “Correlators along the line of two dimensional Ising model,” Phys. Rev.
188, 859 (1969).
[138] M. Kaku, Strings, Conformal Fields, and M-Theory, New York, USA: Springer (2000)
531 p.
[139] V. A. Kazakov, “Wilson loop average for an arbitrary contour in two-dimensional U(N)
gauge theory,” Nucl. Phys. B 179, 283 (1981).
[140] S. V. Ketov, Conformal Field Theory, Singapore, Singapore: World Scientific (1995) 486
p.
[141] V. V. Khoze, M. P. Mattis and M. J. Slater, “The instanton hunter’s guide to supersym-
metric SU(N) gauge theory,” Nucl. Phys. B 536, 69 (1998) [arXiv:hep-th/9804009].
[142] E. Kiritsis, String Theory in a Nutshell, Princeton, USA: Univ. Pr. (2007) 588 p.
[143] V. G. Knizhnik and A. B. Zamolodchikov, “Current algebra and Wess-Zumino model in
two dimensions,” Nucl. Phys. B 247, 83 (1984).
[144] W. Krauth and M. Staudacher, “Non-integrability of two-dimensional QCD,” Phys. Lett.
B 388, 808 (1996) [arXiv:hep-th/9608122].
[145] E. A. Kuraev, L. N. Lipatov and V. S. Fadin, “The Pomeranchuk singularity in non-
abelian gauge theories,” Sov. Phys. JETP 45, 199 (1977) [Zh. Eksp. Teor. Fiz. 72, 377
(1977)]. “Multi-reggeon processes in the Yang-Mills theory,” Sov. Phys. JETP 44, 443
(1976) [Zh. Eksp. Teor. Fiz. 71, 840 (1976)].
[146] D. Kutasov, “Duality off the critical point in two-dimensional systems with nonabelian
symmetries,” Phys. Lett. B 233, 369 (1989).
[147] D. Kutasov, “Two-dimensional QCD coupled to adjoint matter and string theory,” Nucl.
Phys. B 414, 33 (1994) [arXiv:hep-th/9306013].
[148] D. Kutasov and A. Schwimmer, “Universality in two-dimensional gauge theory,” Nucl.
Phys. B 442, 447 (1995) [arXiv:hep-th/9501024].
[149] T. D. Lee and Y. Pang, “Nontopological solitons,” Phys. Rept. 221, 251 (1992).
[150] G. P. Lepage and S. J. Brodsky, “Exclusive processes in quantum chromodynamics:
evolution equations for hadronic wave functions and the form-factors of mesons,” Phys.
Lett. B 87, 359 (1979).
[151] L. N. Lipatov, The Creation of Quantum Chromodynamics and the Effective Energy,
Bologna, Italy: Univ. Bologna (1998) 367 p.
[152] J. H. Lowenstein and J. A. Swieca, “Quantum electrodynamics in two-dimensions,”
Annals Phys. 68, 172 (1971).
[153] M. Luscher, “Quantum nonlocal charges and absence of particle production in the two-
dimensional nonlinear Sigma model,” Nucl. Phys. B 135, 1 (1978).
[154] D. Lust and S. Theisen, “Lectures on string theory,” Lect. Notes Phys. 346, 1 (1989).
[155] A. Maciocia, “Metrics on the moduli spaces of instantons over Euclidean four space,”
Commun. Math. Phys. 135, 467 (1991).
[156] G. Mack and A. Salam, “Finite component field representations of the conformal group,”
Annals Phys. 53, 174 (1969).
[185] T. Schafer and E. V. Shuryak, “Instantons in QCD,” Rev. Mod. Phys. 70, 323 (1998)
[arXiv:hep-ph/9610451].
[186] J. Schechter and H. Weigel, “The Skyrme model for baryons,” arXiv:hep-ph/9907554.
[187] T. D. Schultz, D. C. Mattis and E. H. Lieb, “Two-dimensional Ising model as a soluble
problem of many fermions,” Rev. Mod. Phys. 36, 856 (1964).
[188] M. A. Shifman, Instantons in Gauge Theories, Singapore, Singapore: World Scientific
(1994) 488 p.
[189] E. V. Shuryak, “The role of instantons in quantum chromodynamics. 1. Physical vac-
uum,” Nucl. Phys. B 203, 93 (1982).
[190] J. S. Schwinger, “Gauge invariance and mass. 2,” Phys. Rev. 128, 2425 (1962).
[191] N. Seiberg and E. Witten, “Monopoles, duality and chiral symmetry breaking in N=2
supersymmetric QCD,” Nucl. Phys. B 431, 484 (1994) [arXiv:hep-th/9408099].
[192] N. Seiberg and E. Witten, “Monopole condensation, and confinement in N=2 super-
symmetric Yang-Mills theory,” Nucl. Phys. B 426, 19 (1994) [Erratum-ibid. B 430, 485
(1994)] [arXiv:hep-th/9407087].
[193] Y. M. Shnir, Magnetic Monopoles, Berlin, Germany: Springer (2005) 532 p.
[194] W. Siegel, “Manifest Lorentz invariance sometimes requires nonlinearity,” Nucl. Phys.
B 238, 307 (1984).
[195] T. H. R. Skyrme, “A Nonlinear theory of strong interactions,” Proc. Roy. Soc. Lond. A
247, 260 (1958).
[196] T. H. R. Skyrme, “Particle states of a quantized meson field,” Proc. Roy. Soc. Lond. A
262, 237 (1961).
[197] T. H. R. Skyrme, “A unified field theory of mesons and baryons,” Nucl. Phys. 31, 556
(1962).
[198] Smirnov, F. A. “Form factors in completely integrable models of quantum field theory,”
Adv. Ser. Math. Phys. 14:1–208 (1992).
[199] J. Sonnenschein, “Chiral bosons” Nucl. Phys. B 309, 752 (1988).
[200] M. Spiegelglas and S. Yankielowicz, “G/G topological field theories by cosetting G(K),”
Nucl. Phys. B 393, 301 (1993) [arXiv:hep-th/9201036].
[201] P. J. Steinhardt, “Baryons and baryonium in QCD in two-dimensions,” Nucl. Phys. B
176, 100 (1980).
[202] M. Stone, Bosonization, Singapore: World Scientific (1994) 539 p.
[203] H. Sugawara, “A Field theory of currents,” Phys. Rev. 170, 1659 (1968).
[204] K. Symanzik, “Small distance behavior in field theory and power counting,” Commun.
Math. Phys. 18, 227 (1970).
[205] W. E. Thirring, “A soluble relativistic field theory,” Annals Phys. 3, 91 (1958).
[206] C. B. Thorn, “Computing the Kac determinant using dual model techniques and more
about the no-ghost theorem,” Nucl. Phys. B 248, 551 (1984).
[207] S. B Treiman, R. Jackiw and D. Gross Current Algebra and its Applications (Princeton,
University Press, New Jersey, 1972).
[208] A. I. Vainshtein, V. I. Zakharov, V. A. Novikov and M. A. Shifman, “ABC of instantons,”
Sov. Phys. Usp. 25, 195 (1982) [Usp. Fiz. Nauk 136, 553 (1982)].
[209] S. Vandoren and P. van Nieuwenhuizen, “Lectures on instantons,” arXiv:0802.1862
[hep-th].
[210] E. P. Verlinde, “Fusion rules and modular transformations in 2d conformal field theory,”
Nucl. Phys. B 300, 360 (1988).
[211] E. P. Verlinde and H. L. Verlinde, “Chiral bosonization, determinants and the string
partition function,” Nucl. Phys. B 288, 357 (1987).
[212] M. S. Virasoro, “Subsidiary conditions and ghosts in dual resonance models,” Phys. Rev.
D 1, 2933 (1970).
[213] M. Wakimoto, “Fock representations of the affine Lie algebra A1(1),” Commun. Math.
Phys. 104, 605 (1986).
[214] E. J. Weinberg and P. Yi, “Magnetic monopole dynamics, supersymmetry, and duality,”
Phys. Rept. 438, 65 (2007) [arXiv:hep-th/0609055].
[215] S. Weinberg, The Quantum Theory of Fields. Vol. 1: Foundations, Cambridge, UK:
Univ. Pr. (1995) 609 p.
[216] G. Veneziano, “U(1) Without instantons,” Nucl. Phys. B 159, 213 (1979).
[217] J. Wess and B. Zumino, “Consequences of anomalous Ward identities,” Phys. Lett. B
37, 95 (1971).
[218] F. Wilczek, “Inequivalent embeddings of SU(2) and instanton interactions,” Phys. Lett.
B 65, 160.
[219] K. G. Wilson, “Nonlagrangian models of current algebra,” Phys. Rev. 179, 1499 (1969).
[220] E. Witten, “Some exact multipseudoparticle solutions of classical Yang-Mills theory,”
Phys. Rev. Lett. 38, 121 (1977).
[221] E. Witten, “Instantons, the Quark model, and the 1/N expansion,” Nucl. Phys. B 149,
285 (1979).
[222] E. Witten, “Baryons in the 1/N expansion,” Nucl. Phys. B 160, 57 (1979).
[223] E. Witten, “Large N chiral dynamics,” Annals Phys. 128, 363 (1980).
[224] E. Witten, “Nonabelian bosonization in two dimensions,” Commun. Math. Phys. 92,
455 (1984).
[225] E. Witten, “Global aspects of current algebra,” Nucl. Phys. B 223, 422 (1983).
[226] E. Witten, “Current algebra, baryons, and quark confinement,” Nucl. Phys. B 223, 433
(1983).
[227] E. Witten, “On holomorphic factorization of WZW and coset models,” Commun. Math.
Phys. 144, 189 (1992).
[228] E. Witten, “Two-dimensional gauge theories revisited,” J. Geom. Phys. 9, 303 (1992)
[arXiv:hep-th/9204083].
[229] C. N. Yang and R. L. Mills, “Conservation of isotopic spin and isotopic gauge invariance,”
Phys. Rev. 96, 191 (1954).
[230] C. N. Yang, “Some exact results for the many body problems in one dimension with
repulsive delta function interaction,” Phys. Rev. Lett. 19, 1312 (1967).
[231] I. Zahed and G. E. Brown, “The Skyrme model,” Phys. Rept. 142, 1 (1986).
[232] A. B. Zamolodchikov, “Renormalization group and perturbation theory near fixed points
in two-dimensional field theory,” Sov. J. Nucl. Phys. 46, 1090 (1987) [Yad. Fiz. 46, 1819
(1987)].
[233] A. B. Zamolodchikov, “Exact solutions of conformal field theory in two-dimensions and
critical phenomena,” Rev. Math. Phys. 1, 197 (1990).
[234] A. B. Zamolodchikov, “Thermodynamic Bethe anzatz in relativistic models, scaling three
state Potts and Lee-Yang models,” Nucl. Phys. B 342, 695 (1990).
[235] A. B. Zamolodchikov and A. B. Zamolodchikov, “Factorized S-matrices in two dimensions
as the exact solutions of certain relativistic quantum field models,” Annals Phys. 120,
253 (1979).
[236] A. B. Zamolodchikov, A. B. Zamolodchikov and I. M. Khalatnikov, “Physics Reviews,
Vol. 10, pt. 4: Condormal field theory and critical phenomena in two-dimensional sys-
tems,” London, UK: Harwood (1989) 269–433. (Soviet scientific reviews, Section A, 10.4)
[237] B. Zwiebach, A First Course in String Theory, Cambridge, UK: Univ. Pr. (2004) 558 p