Antibracket, Antifields and Gauge-Theory Quantization
Antibracket, Antifields and Gauge-Theory Quantization
KUL-TF-94/12
UB-ECM-PF 94/15
UTTG-11-94
hep-th/9412228
May 1994
arXiv:hep-th/9412228v1 28 Dec 1994
Antibracket, Antifields
and Gauge-Theory Quantization
Joaquim Gomis∗1 , Jordi Parı́s♯2 and Stuart Samuel†3
∗Theory Group, Department of Physics
The University of Texas at Austin
RLM 5208, Austin, Texas
and
Departament d’Estructura i Constituents de la Matèria
Facultat de Fı́sica, Universitat de Barcelona
Diagonal 647, E-08028 Barcelona
Catalonia
†
Department of Physics
City College of New York
138th St and Convent Avenue
New York, New York 10031 U.S.A.
E-mail: jordi=paris%tf%[email protected]
3 E-mail: [email protected]
Abstract
The antibracket formalism for gauge theories, at both the classical and
quantum level, is reviewed. Gauge transformations and the associated
gauge structure are analyzed in detail. The basic concepts involved in
the antibracket formalism are elucidated. Gauge-fixing, quantum effects,
and anomalies within the field-antifield formalism are developed. The
concepts, issues and constructions are illustrated using eight gauge-theory
models.
Contents
1 Introduction 2
7 Gauge-Fixing Examples 89
7.1 The Spinless Relativistic Particle . . . . . . . . . . . . . . . . . . . . 89
7.2 Yang-Mills Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.3 Topological Yang-Mills Theory . . . . . . . . . . . . . . . . . . . . . . 91
7.4 The Antisymmetric Tensor Field Theory . . . . . . . . . . . . . . . . 93
7.5 Open String Field Theory . . . . . . . . . . . . . . . . . . . . . . . . 94
7.6 The Massless Relativistic Spinning Particle . . . . . . . . . . . . . . . 96
7.7 The First-Quantized Bosonic String . . . . . . . . . . . . . . . . . . . 97
References 171
This work is dedicated to Joseph and Marie,
to Pilar,
and to the memory of Pere and Francesca.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 2
1 Introduction
The known fundamental interactions of nature are all governed by gauge theo-
ries. The presence of a gauge symmetry indicates that a theory is formulated in a
redundant way, in which certain local degrees of freedom do not enter the dynamics.
Conversely, when there are degrees of freedom, which do not enter the lagrangian, a
theory possesses local invariances. Although one can in principle eliminate the gauge
degrees of freedom, there are reasons for not doing so. These reasons include manifest
covariance, locality of interactions, and calculational convenience.
The first example of a gauge theory was electrodynamics. Electric and magnetic
forces are generated via the exchange of photons. Being particles of spin 1, photons
involve a vector field, Aµ . However, not all four components of the electromagnetic
potential Aµ enter dynamically. Two degrees of freedom correspond to the two pos-
sible physical polarizations of the photon. The longitudinal degree of freedom plays
a role in interactions via virtual exchanges of photons. The remaining gauge degree
of freedom does not enter the theory. Consequently, electromagnetism is described
by a gauge theory. When it was realized that the weak interactions could be unified
with electromagnetism in an SU(2) × U(1) gauge theory [129, 266, 213] and that
this theory is renormalizable [243, 244], the importance of non-abelian gauge theo-
ries [276] grew enormously. The strong interactions are also governed by an SU(3)
non-abelian gauge theory. The fourth fundamental force is gravity. It is based on
Einstein’s general theory of relativity and uses general coordinate invariance. When
formulated in terms of a metric or any other convenient fields, gravity also possesses
gauge symmetries.
The quantization of gauge theories is not always straightforward. In the abelian
case, relevant for electromagnetism, the procedure is well understood. In contrast,
quantization of a non-abelian theory and its renormalization is more complicated.
Quantization generally involves the introduction of ghost fields. Typically, a gauge-
fixing procedure is used to render dynamical all degrees of freedom. Ghost fields
are used to compensate for the effects of the gauge degrees of freedom [101], so that
unitarity is preserved. In electrodynamics in the linear gauges, ghosts decouple and
can be ignored. In non-abelian gauge theories, convenient gauges generically involve
interacting ghosts. A major step in understanding these issues was the Faddeev-
Popov quantization procedure [98, 83], which relied heavily on the functional-integral
approach to quantization [102, 1, 165]. From this viewpoint, the presence of ghost
fields is understood as a “measure effect”. In dividing out the volume of gauge
transformations in function space, a Jacobian measure factor arises. This factor is
produced naturally by introducing quadratic terms in the lagrangian for ghosts and
then integrating them out. It was realized at a later stage that the gauge-fixed action
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 3
retains a nilpotent, odd, global symmetry involving transformations of both fields and
ghosts. This Becchi-Rouet-Stora-Tyutin (BRST) symmetry [36, 254] is what remains
of the original gauge invariance. In fact, for closed theories, the transformation law
for the original fields is like a gauge transformation with gauge parameters replaced
by ghost fields. In general, this produces nonlinear transformation laws. The relations
among correlation functions derived from BRST symmetry involve the insertions of
the BRST variation of fields. These facts require the use of composite operators and
it is convenient to introduce sources for these transformations. The Ward identities
[265] associated with the BRST invariance treated in this way are the Slavnov-Taylor
identities [233, 241]. The Slavnov-Taylor identities and BRST symmetry have played
an important role in quantization, renormalization, unitarity, and other aspects of
gauge theories.
Ghosts fields have been useful throughout the development of covariant gauge-
field-theory quantization [181, 182, 184, 205, 3]. It is desirable to have a formulation
of gauge theories that introduces them from the outset and that automatically in-
corporates BRST symmetry [32]. The field-antifield formulation has these features
[36, 277, 24, 25, 26, 27]. It relies on BRST symmetry as fundamental principle and
uses sources to deal with it [36, 254, 277]. It encompasses previous ideas and develop-
ments for quantizing gauge systems and extends them to more complicated situations
(open algebras, reducible systems, etc.) [113, 114, 172, 238, 81]. In 1975, J. Zinn-
Justin, in his study of the renormalization of Yang-Mills theories [277], introduced
the above-mentioned sources for BRST transformations and a symplectic structure
( , ) (actually denoted ∗ by him) in the space of fields and sources, He expressed the
Slavnov-Taylor identities in the compact form (Γ, Γ) = 0, where Γ, the generating
functional of the one-particle-irreducible diagrams, is known as the effective action
(see also [187]). These ideas were developed further by B. L. Voronov and I. V. Tyutin
in [263, 264] and by I. A. Batalin and G. A. Vilkovisky in refs.[24, 25, 26, 27, 28]. These
authors generalized the role of ( , ) and of the sources for BRST transformations and
called them the antibracket and antifields respectively. Due to their contributions,
this quantization procedure is often referred to as the Batalin-Vilkovisky formalism.
The antibracket formalism gained popularity among string theorists, when it was
applied to the open bosonic string field theory [56, 246]. It has also proven quite
useful for the closed string field theory and for topological field theories. Only within
the last few years has it been applied to more general aspects of quantum field theory.
In some sense, the BRST approach, which was driven, in part, by renormalization
considerations, and the field-antifield formalism, which was motivated by classical
considerations such as gauge structure, are not so different. When sources are in-
troduced for BRST transformations, the BRST approach resembles the field-antifield
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 4
one. Antifields, then, have a simple intepretation: They are the sources for BRST
transformations. In this sense, the field-antifield formalism is a general method for
dealing with gauge theories within the context of standard field theory.
The general structure of the antibracket formalism is as follows. One introduces
an antifield for each field and ghost, thereby doubling the total number of original
fields. The antibracket ( , ) is an odd non-degenerate symplectic form on the space
of fields and antifields. The original classical action S0 is extended to a new action
S, in an essentially unique way, to arrive at a theory with manifest BRST symmetry.
One equation, the master equation (S, S) = 0, reproduces in a compact way the
gauge structure of the original theory governed by S0 . Although the master equation
resembles the Zinn-Justin equation, the content of the two is different since S is a
functional of quantum fields and antifields and Γ is a functional of classical fields.
The antibracket formalism currently appears to be the most powerful method for
quantizing a gauge theory. Beyond tree level, order h̄ terms usually need to be added
to the action, thereby leading to a quantum action W . These counterterms are ex-
pected to render finite loop contributions, after a suitable regularization procedure
has been introduced. The master equation must be appropriately generalized to the
so-called quantum master equation. It involves a potentially singular operator ∆.
The regularization procedure and counterterms should also render ∆ and its action
on W well-defined. Violations of the quantum master equation are equivalent to
gauge anomalies [251]. To calculate correlation functions and scattering amplitudes
in perturbation theory, a gauge-fixing procedure is selected. This procedure elimi-
nates antifields in terms of functionals of fields. When appropriately implemented,
propagators exist, and the usual Feynman graph methods can be used. In addition,
for the study of symmetry properties, renormalization and anomalies, a modified ver-
sion of the gauge-fixing procedure is available which keeps antifields. In short, the
antibracket formalism has manifest gauge invariance or BRST symmetry, provides
the extra fields needed for covariant quantization, permits a perturbative expansion
of the quantum theory, and allows the study of quantum corrections to the symmetry
structure of the theory.
The field-antifield formalism can treat systems that cannot be handled by Faddeev-
Popov functional integration approach. This is particularly clear for theories in which
quartic ghost interactions arise [172, 81]. Faddeev-Popov quantization leads to an
action bilinear in ghost fields, and fails for the case of open algebras. An open
algebra occurs when the commutator of two gauge transformations produces a term
proportional to the equations of motion and not just another gauge transformation
[81, 27]. In other words, the gauge algebra closes only on-shell. Such algebras occur
in gravity [110] and supergravity [114, 172, 81, 258] theories. The ordinary Faddeev-
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 5
Popov procedure also does not work for reducible theories. In reducible theories, the
gauge generators are all not independent [67, 171, 6, 37, 228, 250, 260, 80, 242]. Some
modifications of the procedure have been developed by introducing ghosts for ghosts
[228, 178]. However, these modifications [228, 148, 178, 116] do not work for the
general reducible theory. Even for Yang-Mills theories, the Faddeev-Popov procedure
can fail, if one considers exotic gauge-fixing procedures for which “extraghosts” appear
[172, 199, 200]. The field-antifield formalism is sufficiently general to encompass
previously known lagrangian approaches to the quantization of gauge theories.
Perhaps the most attractive feature of the field-antifield formalism is its imita-
tion of a hamiltonian Poisson structure in a covariant way. In some instances, the
hamiltonian approach to quantization has the advantage of being manifestly unitary.
However, it is necessarily non-covariant since the time variable is treated in a manner
different from the space variables. In addition, the gauge invariances usually must be
fixed at the outset. In compensation for this, one needs to impose constraints on the
Hilbert space of states. In the field-antifield approach, the antibracket plays the role
of the Poisson bracket. As a consequence, hamiltonian concepts, such as canonical
transformations, can be formulated and used [262, 263, 264, 27, 105, 251]. At the
same time, manifest covariance and BRST invariance are maintained. Since the an-
tibracket formalism proceeds via the functional integral, the powerful techniques of
functional integration are available.
A non-trivial aspect of the field-antifield approach is the construction of the quan-
tum action W . When loop effects are ignored, W → S provides the solution to the
master equation. A straightforward but not necessarily simple procedure is available
for obtaining S given the classical action S0 and its gauge invariances for a finite-
reducible system. When quantum effects are incorporated, W must satisfy the more
singular quantum master equation. However, there is currently no known method that
guarantees the construction of W . The problem is that the field-antifield formalism
does not automatically provide the functional integration measure. These issues are
linked with those associated with unitarity, renormalization, quantum gauge invari-
ance, and anomalies. Because these aspects of gauge theories are inherently difficult,
it is not surprising that the field-antifield formalism does not provide a simple solu-
tion.
Another, less serious weakness, is that the antibracket formalism involves quite
a bit of mathematical machinery. Sometimes, a gauge theory is expressed in a form
which is more complicated than necessary. This can make computations somewhat
more difficult.
The organization of this article is as follows. Sect. 2 discusses gauge structure.
Some notation is presented during the process of introducing gauge transformations.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 6
The distinction between irreducible and reducible gauge theories is made. The latter
involve a redundant set of gauge invariances so that there are relations among the
gauge generators. As a result, there exists gauge invariances for gauge invariances,
and ghosts for ghosts. A theory is Lth-stage reducible if there are gauge invariances
for the gauge invariances for the gauge invariances, etc., L-fold times. The general
form of the gauge structure for a first-stage reducible case is determined. In Sect.
3, specific gauge theories are presented to illustrate the concepts of Sect. 2. The
spinless relativistic particle, non-abelian Yang-Mills theories, topological Yang-Mills
theory, the antisymmetric tensor field, free abelian p-form theories, open bosonic
string field theory, the massless relativistic spinning particle, and the first-quantized
bosonic string are treated. The spinless relativistic particle of Sect. 3.1 is also used to
exemplify notation. The massless relativistic spinning particle provides an example of
a simple supergravity theory, namely a theory with supersymmetric gauge invariances.
This system is used to illustrate the construction of supersymmetric and supergravity
theories. A review of the construction of general-coordinate-invariant theories is given
in the subsection on the first-quantized open bosonic string. These mini-reviews
should be useful to the reader who is new to these subjects.
The key concepts of the field-antifield formalism are elucidated in Sect. 4. An-
tifields are introduced and the antibracket is defined. The latter is used to define
canonical transformations. They can be quite helpful in simplifying computations.
Next, the classical master equation (S, S) = 0 is presented. When appropriate bound-
ary conditions are imposed, it reproduces, in a compact way, the gauge structure of
Sect. 2. A suitable action S satisfying the master equation is called a proper solution.
Given the gauge-structure tensors of a first-stage reducible theory, Sect. 4.4 presents
the generic proper solution. The last part of Sect. 4 defines and discusses the classical
BRST symmetry. Examples of proper solutions are provided in Sect. 5 for the gauge
field theories presented in Sect. 3.
Sect. 6 begins the passage from the classical to the quantum aspects of the
field-antifield formalism. The gauge-fixing procedure is discussed. The gauge-fixing
fermion Ψ is a key concept. It is used as a means of eliminating antifields in terms
of functions of fields. The result is an action that is suitable for use in the path
integral. Only in this context and in performing standard perturbative computations
are antifields eliminated. It is shown that results are independent of the choice of
Ψ, if the quantum action W satisfies the quantum master equation. To implement
gauge-fixing, more fields and their antifields must be introduced. How this works
for irreducible and first-stage reducible theories is treated first. Then, for reference
purposes, the general Lth-stage reducible case is considered. Delta-function type
gauge-fixing is treated in Sect. 6.3. Again, irreducible and first-stage reducible cases
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 7
are presented first. Again, for reference purposes, the general Lth-stage reducible
case is treated. Gauge-fixing by a gaussian averaging process is discussed in Sect.
6.4. After gauge-fixing, a classical gauge-fixed BRST symmetry can be defined. See
Sect. 6.5. The freedom to perform canonical transformations permits one to work in
any appropriate field basis. This freedom can be quite useful. Concepts tend to have
different interpretations in different bases. One basis, associated with Ψ and called
the gauge-fixed basis, is the last topic of Sect. 6. Examples of gauge-fixing procedures
are provided in Sect. 7. With the exception of the free p-form theory, the theories are
the ones considered in Sects. 3 and 5.
Quantum effects and possible gauge anomalies are analyzed in Sect. 8. The
key concepts are quantum-BRST transformations and the quantum master equa-
tion. Techniques for assisting in finding solutions to the quantum master equation
are provided in Sects. 8.2, 8.5 and 8.6. The generating functional Γ for one-particle-
irreducible diagrams is generalized to the field-antifield case in Sect. 8.4. This allows
one to treat the quantum system in a manner similar to the classical system. The
Zinn-Justin equation is shown to be equivalent to the quantum master equation.
When unavoidable violations of the latter occur, the gauge theory is anomalous. See
Sect. 8.5. Explicit formulas at the one-loop level are given. In Sect. 9, sample anomaly
calculations are presented. It is shown that the spinless relativistic particle does not
have an anomaly. In Sect. 9.2, the field-antifield treatment of the two-dimensional
chiral Schwinger model is presented. Violations of the quantum master equation are
obtained. This is expected since the theory is anomalous. A similar computation
is performed for the open bosonic string. For D 6= 26, the theory is anomalous, as
expected. Some of the details of the calculations are relegated to Appendix C.
Section 10 briefly presents several additional topics. The application of the field-
antifield formalism to global symmetries is presented. A review is given of the geo-
metric interpretation of E. Witten [273]. The next topic is the role of locality. This
somewhat technical issue is important for renormalizability and for cohomological
aspects. A summary of cohomological methods is given. Next, the relation between
the hamiltonian and antibracket approaches is discussed. The question of unitarity
is the subject of Sect. 10.6. One place where the field-antifield formalism has played
an essential role is in the D = 26 closed bosonic string field theory. This example is
rather complicated and not suitable for pedagogical purposes. Nevertheless, general
aspects of the antibracket formalism for the closed string field theory are discussed.
Finally, an overview is given of how to handle anomalous systems using an extended
set of fields and antifields.
Appendix A reviews the mathematical aspects of left and right derivatives, inte-
gration by parts, and chain rules for differentiation. Appendix B discusses in more
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 8
are treated in [253]. Pedagogical treatments are given in references [157, 253]. In
certain places, material from reference [206] has been used.
This review focuses on the key points and concepts of antibracket formalism.
There is some emphasis on applications to string theory. Our format is to first
present the material abstractly and then to supply examples. The reader who is
new to this subject and mainly interested in learning may wish to reverse this order.
Exercises can be generated by verifying the abstract results in each of the sample
gauge theories of Sects. 3, 5, 7, and 9. Other systems, which have been treated
by field-antifield quantization and may be of use to the reader, are the free spin 52
field [26], the spinning string [136], the 10-dimensional Brink-Schwarz superparticle
and superstring [123, 139, 173, 190, 211, 43, 232, 44, 227]1 , chiral gravity [77], W3
gravity [161, 45, 74, 162, 72, 257], general topological field theories [66, 185, 50, 49,
127, 164, 191, 158, 79], the supersymmetric Wess-Zumino model [33, 159] and chiral
gauge theories in four-dimensions [251]. The antibracket formalism has found various
interpretations in mathematics [125, 126, 127, 209, 189, 198, 218, 219, 237]. Some
other recent relevant work can be found in [261, 21, 22, 47, 147]. The referencing in
this review is thorough but not complete. A restriction has been made to only cite
works directly relevant to the issues addressed in each section. Multiple references
are done first chronologically and then alphabetically. The titles of references are
provided to give the reader a better indication of the content of each work.
We work in Minkowski space throughout this article. Functional integrals are
defined by analytic continuation using Wick rotation. This is illustrated in the com-
putations of Appendix C. We use ηµν to denote the flat-space metric with the signa-
ture convention (−1, 1, 1, . . . , 1). Flat-space indices are raised and lowered with this
metric. The epsilon tensor εµ0 µ1 µ2 ...µd−1 is determined by the requirement that it be
antisymmetric in all indices and that ε012...d−1 = 1, where d − 1 and d are respectively
the dimension of space and space-time. We often use square brackets to indicate a
functional of fields and antifields to avoid confusion with the antibracket, i.e., S[Φ, Φ∗ ]
in lieu of S(Φ, Φ∗ ).
1
See [227] for additional references.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 10
Here, εα (x) are infinitesimal gauge parameters, that is, arbitrary functions of the
space-time variable x, and Rαi are the generators of gauge transformations. These gen-
erators are operators that act on the gauge parameters. In kernel form, (Rαi (φ)εα ) (x)
R
can be represented as dyRαi (x, y) εα (y).
It is convenient to adopt the following compact notation [82, 83]. Unless otherwise
stated, the appearance of a discrete index also indicates the presence of a space-
time variable. We then use a generalized summation convention in which a repeated
discrete index implies not only a sum over that index but also an integration over the
corresponding space-time variable. As a simple example, consider the multiplication
of two matrices g and h, written with explicit matrix indices. In compact notation,
f A B = g A C hC B (2.2)
becomes not only a matrix product in index space but also in function space. Eq.(2.2)
represents
XZ
f A B (x, y) = dz g A C (x, z) hC B (z, y) (2.3)
C
in conventional notation. In other words, the index A in Eq.(2.2) stands for A and
x in Eq.(2.3). Likewise, B and C in Eq.(2.2) represent {B, y} and {C, z}. The
generalized summation convention for C in compact notation yields a sum over the
discrete index C and an integration over z in conventional notation in Eq.(2.3). The
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 12
∂r S0 [φ]
S0,i (φ, x) ≡ , (2.6)
∂φi (x)
where the subscript r indicates that the derivative is to be taken from the right (see
Appendix A). Henceforth, when a subscript index i, j, etc., appears after a comma
it denotes the right derivative with respect to the corresponding field φi , φj , etc.. In
compact notation, we write Eq.(2.6) as S0,i = ∂∂φr S0
i where the index i here stands for
a total derivative. We assume that surface terms can be dropped in such integrals
– this is indeed the case when Eq.(2.7) is applied to gauge parameters that fall off
sufficiently fast at spatial and temporal infinity. The Noether identities in Eq.(2.8)
are the key equations of this subsection and can even be thought of as the definition
of when a theory is invariant under a gauge transformation of the form in Eq.(2.1).
To commence perturbation theory, one searches for solutions to the classical equa-
tions of motion, S0,i (φ, x) = 0, and then expands about these solutions. We assume
there exists at least one such stationary point φ0 = {φj0 } so that
!
∂l ∂r S0
⇒ Rαj =0 , (2.10)
∂φi ∂φj
φ0
∂l ∂r S0
i.e., the hessian ∂φ i ∂φj of S0 is degenerate at any point on the stationary surface
i
Σ. The Rα are on-shell null vectors of this hessian. Since propagators involve the
inverse of this hessian, propagators do not exist for certain combinations of fields.
This means that the standard loop expansion cannot be straightforwardly applied. A
method is required to overcome this problem.
Technically speaking, to study the structure of the set of gauge transformations
it is necessary to assume certain regularity conditions on the space for which the
equations of motion S0,i = 0 hold. The interested reader can find these conditions
in Appendix B. A key consequence of the regularity conditions is that if a function
F (φ) of the fields φ vanishes on-shell, that is, when the equations of motion are
implemented, then F must be a linear combination of the equations of motion, i.e.,
where |Σ indicates the restriction to the surface where the equations of motion hold
[81, 28, 106, 103, 105]. Eq.(2.11) can be thought of as a completeness relation for the
equations of motion. We shall make use of Eq.(2.11) frequently.
Throughout Sect. 2, we assume that the gauge generators are fixed once and for
all. One could take linear combinations of the generators to form a new set. This
would change the gauge-structure tensors presented below. This non-uniqueness is
not essential and is discussed in Sect. 4.5.
To see explicit examples of the abstract formalism that follows, one may want to
glance from time to time at the examples of Sect. 3.
S0,i λi = 0 ⇔ λi = R0α
i
0
λ′α0 + S0,j T ji , (2.12)
T ij = −(−1)ǫi ǫj T ji . (2.13)
i
The R0α 0
are the gauge generators in Eq.(2.1). For notational convenience, we have
appended a subscript 0 on the gauge generator and the gauge index α. This subscript
indicates the level of the gauge transformation. The second term S0,j T ji in Eq.(2.12)
is known as a trivial gauge transformation. Such transformations are discussed in
the next subsection. It is easily checked that the action is invariant under such
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 15
is n − m0 . Define the net number of degrees of freedom ndof to be the number of fields
that enter dynamically in S0 , regardless of whether they propagate.2 Then for an
irreducible theory ndof is n − m0 since there are m0 gauge degrees of freedom. Note
that ndof matches the rank of the hessian in Eq.(2.15).
If, however, there are dependences among the gauge generators, and the rank
i
of the generators is less than their number, rank R0α0 < m0 , then the theory is
Σ
reducible. If m0 − m1 of the generators are independent on-shell, then there are m1
α0
relations among them and there exist m1 functionals R1α 1
such that
i α0 ji
R0α0
R1α 1
= S0,j V1α 1
, α1 = 1, . . . , m1 ,
α0
ǫ(R1α 1
) = ǫα0 + ǫα1 (mod 2) , (2.16)
ji ij ji
for some V1α 1
, satisfying V1α 1
= −(−1)ǫi ǫj V1α 1
. Here, ǫα1 is the statistical parity
α0 i
of the level-one gauge parameter. The R1α1 are the on-shell null vectors for R0α 0
i α0 ji
since R0α 0
R 1α1 Σ = 0. The presence of V 1α1 in Eq.(2.16) is a way of extending
this statement off-shell. Here and elsewhere, when a combination of field equations
appears on the right-hand side of an equation, it indicates the off-shell extension of an
on-shell statement; such an extension can be performed using the regularity postulate
of Appendix B. Note that, if εα = R1α α
εα for any εα1 , then δφi in Eq.(2.5) is zero on-
1 1
shell, so that no gauge transformation is produced. In Eq.(2.16) it is assumed that
i α0 α0
the reducibility of the R0α 0
is completely contained in R1α 1
, i.e., R1α 1
also constitute
a complete set
i
R0α0
λα0 = S0,j M0ji ⇒ λα0 = R1α
α0
1
λ′α1 + S0,j T0jα0 , (2.17)
2
In electromagnetism, ndof = 3, but there are only two propagating degrees of freedom corre-
sponding to the two physical polarizations.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 16
from which one concludes that the commutator of a trivial transformation with any
other transformation is a trivial transformation. Hence, the trivial transformations
are a normal subgroup H of the full group of gauge transformations, Ḡ.
The trivial gauge transformations are of no physical significance: They neither
lead to conserved currents nor do they prevent the development of a perturbative
expansion about a stationary point. They are simply a consequence of having more
than one degree of freedom. On these grounds, it would seem sensible to dispense with
them and restrict oneself to the quotient G = Ḡ/H. However, this is only possible in
certain cases. In general, the commutator of two non-trivial gauge transformations
produces trivial gauge transformations. Furthermore, for reasons of convenience,
particularly when it is desirable to have manifest covariance or preserve locality, one
sometimes wants to include trivial transformations. Hence, the full group Ḡ is used
for studying the gauge structure of the theory.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 18
On the other hand, since this commutator is also a gauge symmetry of the action it
satisfies the Noether identity so that, factoring out the gauge parameters εβ1 and εα2 ,
one may write
i
S0,i Rα,j Rβj − (−1)ǫα ǫβ Rβ,j
i
Rαj = 0 .
Taking into account Eq.(2.12) the above equation implies the following important
relation among the generators
i
Rα,j Rβj − (−1)ǫα ǫβ Rβ,j
i γ
Rαj = Rγi Tαβ ji
− S0,j Eαβ , (2.22)
γ ji γ ji
for some gauge-structure tensors Tαβ and Eαβ . This equation defines Tαβ and Eαβ .
β α
Restoring the dependence on the gauge parameters ε1 and ε2 , the last two equations
imply
γ β α ji β α
[δ1 , δ2 ]φi ≡ Rγi Tαβ ε1 ε2 − S0,j Eαβ ε1 ε2 , (2.23)
γ
where Tαβ are known as the “structure constants” of the gauge algebra. The words
γ
structure constants are in quotes because in general the Tαβ depend on the fields of
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 19
ji
the theory and are not “constant”. The possible presence of the Eαβ term is due to
the fact that the commutator of two gauge transformations may give rise to trivial
gauge transformations [81, 24, 27].
ij
The gauge algebra generated by the Rαi is said to be open if Eαβ 6= 0, whereas the
ij
algebra is said to be closed if Eαβ = 0. Moreover, Eq.(2.22) defines a Lie algebra if
ij γ
the algebra is closed, Eαβ = 0, and the Tαβ do not depend on the fields φi.
The gauge-structure tensors have the following symmetry properties under the
interchange of indices
ij ji ij
Eαβ = −(−1)ǫi ǫj Eαβ = −(−1)ǫα ǫβ Eβα ,
γ γ
Tαβ = −(−1)ǫα ǫβ Tβα . (2.24)
ij
In other words, Eαβ is graded-antisymmetric both in lower indices and in upper indices
γ
and Tαβ is graded-antisymmetric in lower indices. The statistical parity of structure
tensors is determined by the sum of the parities of the tensor indices,
so that ǫ (Rαi ) =
γ ij
(ǫα + ǫi ) (mod 2), ǫ Tαβ = (ǫα + ǫβ + ǫγ ) (mod 2), and ǫ Eαβ = (ǫi + ǫj + ǫα + ǫβ )
(mod 2).
The next step determines the restrictions imposed by the Jacobi identity. In
general, it leads to new gauge-structure tensors and equations [172, 258, 84, 28]. The
identity X
[δ1 , [δ2 , δ3 ]] = 0 ,
cyclic over 1, 2, 3
Again, the completeness of the generators implies that the general solution of the
preceding equation is of the form
ji
Bαβγ + (−1)ǫi ǫδ Rδj Dαβγ
iδ jδ
− (−1)ǫj (ǫi +ǫδ ) Rδi Dαβγ kji
= −S0,k Mαβγ , (2.30)
kji
where Mαβγ is graded antisymmetric in i, j, and k. In this way, the Jacobi identity
jδ kji
leads to the existence of two new gauge-structure tensors Dαβγ and Mαβγ which, for
a generic theory, are different from zero and must satisfy Eqs.(2.28) and (2.30).
Continuing in the same way, that is to say, commuting more and more gauge
transformations, new structure tensors with increasing numbers of indices are ob-
tained. These tensors are the so-called structure functions of the gauge algebra and
they determine the nature of the set of gauge transformations of the theory. The
reader may not be aware of the higher-order tensors because in the simplest gauge
theories, such as Yang-Mills, they vanish.
ji jδ kji
The tensors Aδαβγ , Bαβγ , Dαβγ and Mαβγ are all graded-antisymmetric in α, β
ji kji
and γ. In addition Bαβγ is graded-antisymmetric in i and j while Mαβγ is graded
antisymmetric in i, j and k. Here we summarize these properties
ǫ(C α ) = ǫα + 1 , (2.32)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 21
and to replace gauge parameters by ghosts, as is done in the BRST formalism [32].
The ghost fields obey the same boundary conditions as gauge parameters. The ghosts
can be used as a compact way of writing the gauge-structure equations. However, in
γ ij jδ kji
order to do this, the symmetry properties of Tαβ , Eαβ , Dαβγ , Mαβγ , etc., need to be
correctly incorporated. Note that these tensors are graded anti-symmetric in lower-
index gauge indices α, β, etc., whereas the ghosts satisfy C α C β = (−1)(ǫα +1)(ǫβ +1) C β C α .
If one is given a graded anti-symmetric tensor Tα1 α2 α3 α4 ... , then a way to make it into
a graded symmetric tensor with symmetry factors ǫα1 + 1, ǫα2 + 1, etc., associated
with indices α1 , α2 , etc., is to multiply by a factor of (−1)ǫαi for every other index αi
in Tα1 α2 α3 α4 ... . In other words, one replaces Tα1 α2 α3 α4 ... by (−1)ǫα2 +ǫα4 +...Tα1 α2 α3 α4 ... .
Using this device, one arrives at a compact way of writing the Noether identity (2.8),
the gauge commutator relation (2.22), as well as Eqs.(2.28) and (2.30) which arise
from the Jacobi identity:
S0,i Rαi C α = 0 , (2.33)
i
2Rα,j Rβj − Rγi Tαβ
γ ji
+ S0,j Eαβ (−1)ǫα C β C α = 0 , (2.34)
jδ
Aδαβγ − S0,j Dαβγ (−1)ǫβ C γ C β C α = 0 , (2.35)
ji
Bαβγ + (−1)ǫi ǫδ Rδj Dαβγ
iδ jδ
− (−1)ǫj (ǫi +ǫδ ) Rδi Dαβγ kji
+ S0,k Mαβγ (−1)ǫβ C γ C β C α = 0 ,
(2.36)
δ ji
where Aαβγ and Bαβγ are defined in Eqs.(2.26) and (2.27). The graded-anticommuting
nature of the ghosts automatically produces the appropriate graded-cyclic sums.
Equations (2.33) through (2.36) are key equations for an irreducible algebra.
Now let us consider a first-stage reducible gauge theory. In this case, the existence
of non-trivial relations among the generators in Eq.(2.16) leads to the appearance of
new tensor quantities.
For first-stage reducible theories there are on-shell null vectors for the generators
Rα . Let Zaα denote these null vectors. In Eq.(2.16), the Zaα are called R1a
i α
when α = α0
and a = α1 . The null vectors are independent on-shell. Their presence modifies the
solutions of the Jacobi identities in Eqs.(2.35) and (2.36) as well as higher-commutator
structure equations. In addition there are new structure equations. One of these is
Eq.(2.16) itself:
Rαi Zbα = S0,j Vbji . (2.37)
Another is derived as follows. Take relation (2.22) and multiply it by Zaβ to obtain
i
Rα,j Rβj − (−1)ǫα ǫβ Rβ,j
i γ
Rαj − Rγi Tαβ ji
+ S0,j Eαβ Zaβ = 0 .
Use Eq.(2.37) to express Rβj Zaβ as a term proportional to equations of motion. Also
i
do the same with Rβ,j Rαj Zaβ and make use of the Noether identity in Eq.(2.8). After
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 22
a little algebra, one finds that the previous equation can be written in the form
Rγi (−1)ǫa ǫβ Za,j
γ
Rβj − Tβδ
γ ji
Zaδ = S0,j Mβa ,
ji
for some quantity Mβa . Terms proportional to the equations of motion have been
ji
collected into Mβa . Using the completeness of the null vectors Zaα , the general solution
to this equation is
(−1)ǫa ǫβ Za,j
γ
Rβj − Tβδ
γ
Zaδ = −Zdγ Adaβ − S0,j Gjγ
aβ . (2.38)
Eq.(2.38) is a new gauge-structure equation for the first-stage reducible case. Two
new structure tensors Adaβ and Gjγ aβ arise.
The null vectors also lead to modifications of the solution of the Jacobi identity.
Eq.(2.25) still holds but its solution is different. Instead of Eq.(2.35), one obtains
Aδαβγ + Zcδ Fαβγ
c jδ
− S0,j Dαβγ (−1)ǫβ C γ C β C α = 0 , (2.39)
where we have made use of the completeness of the null vectors Zaα . In this equation
Aδαβγ stands for the combination of terms in Eq.(2.26).
Multiplying Eq.(2.39) by Rδi and using the Jacobi identity result of Eq.(2.25) lead
ij
to a modification of Eq.(2.29) involving Bαβγ . The new result reads
ji
S0,j Bαβγ − (−1)ǫj (ǫi +ǫδ ) Rδi Dαβγ
jδ
+ Vcji Fαβγ
c
(−1)ǫβ C γ C β C α = 0 ,
and
(−1)ǫa ǫβ Za,j
γ
Rβj − Tβδ
γ
Zaδ + Zdγ Adaβ + S0,j Gjγ a β
aβ η C = 0 . (2.42)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 23
To summarize, key equations for first-stage reducible theories are Eqs.(2.33), (2.34)
and (2.39) – (2.42). Besides the null vectors Zaβ , the new structure tensors are Vaji ,
Adaβ , Gjγ a
aβ , Fαβγ as well as higher-level tensors.
Needless to say, for a higher-order reducible theory the number of quantities and
equations increases considerably. The complexity of the formalism makes the study
of the gauge structure at higher levels quite complicated. A more sensible approach is
to have a generating functional whose expansion in terms of auxiliary fields produces
the generic gauge-structure tensors. In addition, it is desirable to have a simple single
equation which, when expanded in terms of auxiliary fields, generates the entire set
of gauge-structure equations. The field-antifield method [24, 25, 26] provides such
a formalism. The generating functional for structure tensors is a generalized action
subject to certain boundary conditions and the classical master equation contains
all the gauge-structure equations. Before presenting the abstract machinery, it is of
pedagogical value to consider some examples of the formalism of this section.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 24
where a dot over a variable indicates a derivative with respect to proper time. The
variations of the action with respect to the fields, Eq.(2.6), are
!
d ẋµ 1 ẋ2
S0,µ =− , S0,e = − 2 − m2 , (3.2)
dτ e 2 e
where S0,e is the variation of the action with respect to field e; in other words, we also
use e as a field index for the einbein e. If the equation of motion S0,e = 0 is used to
solve for e, and this solution is substituted into the action in Eq.(3.1), one finds that
R √
the action becomes the familiar one: S0 = − dτ m −ẋ2 . Classically, this action
and the one in Eq.(3.1) are equivalent.
The infinitesimal gauge transformations for this system can be written as
ẋµ
δxµ = ε, δe = ε̇ . (3.3)
e
It is straightforward to verify that Eq.(3.3) is a symmetry of Eq.(3.1). The Noether
identity in Eq.(2.8) reads
Z ( ! )
ẋµ d ẋµ 1 ẋ2 d
dτ − − 2
+ m2 δ (τ − τ ′ ) = 0 , (3.4)
e dτ e 2 e dτ
where δ(ε1 ) indicates a gauge transformation with parameter ε1 . These abelian gauge
transformations (3.3) are related to the standard reparametrization transformations
d
δR xµ = ẋµ ε , δR e = (eε) (3.6)
dτ
through the following redefinition of the gauge parameter
ε −→ εe .
Eqs.(3.7) and (3.8) correspond to the usual diffeomorphism algebra. Since the com-
mutator of two gauge transformations is a gauge transformation, the algebra is closed
ji
and Eαβ in Eq.(2.23) is zero. This example illustrates the effect of field-dependent
redefinitions of the gauge parameters or, equivalently, of the gauge generators: An
abelian algebra can be transformed into a non-abelian one. The converse of this also
holds. One can transform any given non-abelian algebra into an abelian algebra using
field-dependent redefinitions, a result known as the abelianization theorem [27]. The
fact that this process can spoil the locality of the transformations is one of the reasons
for using the non-abelian version.
It may appear unusual that a single family of gauge transformations produces
non-abelian commutation relations. This is due to the local non-commutativity of
reparametrization transformations that arises from the time derivatives in Eq.(3.6).
Indeed, when ε1 and ε2 have non-overlapping support, i.e., ε1 (τ ) = 0 where ε2 (τ ) 6= 0
and vice-versa, ε12 = 0.
γ
It is instructive to see how a non-zero structure constant Tαβ for the diffeomor-
phism algebra arises using compact notation. In what follows, gauge indices, α, β,
etc. are replaced by proper time variable ρ, σ, τ , υ. From Eq.(3.6) one sees that the
transformation operators R for reparametrizations are
Rσµτ = ẋµ (τ ) δ (τ − σ) ,
d d
Rσeτ = e (τ ) δ (τ − σ) + ė (τ ) δ (τ − σ) = [e (τ ) δ (τ − σ)] .
dτ dτ
For the xµ degrees of freedom, a straightforward computation yields
µτ d
Rσ,νυ = δνµ δ (τ − σ) δ (τ − υ) ,
dτ
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 27
XZ d
µτ
Rσ,νυ Rρνυ = dυδνµ δ (τ − σ) δ (τ − υ) ẋν (υ) δ (υ − ρ)
ν dτ
d
= ẍµ (τ ) δ (τ − σ) δ (τ − ρ) + ẋµ (τ ) δ (τ − σ) δ (τ − ρ) .
dτ
Antisymmetrizing in ρ and σ and comparing with Eq.(2.22) one finds
τ d d
Tσρ = δ (τ − σ) δ (τ − ρ) − δ (τ − ρ) δ (τ − σ) , (3.9)
dτ dτ
which is in agreement with Eqs.(3.7) and (3.8). For e, straightforward computation
produces
eτ d d d
Rσ,eυ = δ (τ − υ) δ (τ − σ) + δ (τ − σ) δ (τ − υ) = [δ (τ − υ) δ (τ − σ)] ,
dτ" #" dτ # dτ
eτ d d d
Rσ,eυ Rρeυ = e (τ ) δ (τ − σ) δ (τ − ρ) + ė (τ ) δ (τ − ρ) δ (τ − σ) +
dτ dτ dτ
2
d d
2ė (τ ) δ (τ − σ) δ (τ − ρ) + e (τ ) δ (τ − σ) 2 δ (τ − ρ) + ë (τ ) δ (τ − σ) δ (τ − ρ) .
dτ dτ
Antisymmetrizing in ρ and σ and using Eq.(2.23), one finds T is again given by
Eq.(3.9).
Although compact notation is useful to represent the formalism of gauge theories
in full generality, it is cumbersome for specific theories, especially for those in which
more natural notation has already been established. In the examples that follow,
we do not explicitly display equations in compact form but use more conventional
notation.
where fab c are the structure constants of G. They are real and antisymmetric in lower
indices fab c = −fba c and they must satisfy the Jacobi identity
It is useful to verify the key equations for an irreducible closed algebra given in
Eqs.(2.33)-(2.36). The generator of gauge transformations is the covariant derivative
in Eq.(3.17). Using Eqs.(3.14) and (3.15), the Noether identity in Eq.(2.33) reads
Z
dd x (D µ Fµν )b (D ν C)b = 0 .
To verify this equation, integrate by parts, use the antisymmetry of Fµν in µ and ν,
use Eq.(3.19), and then make use of the antisymmetry of fcd a in c and d:
Z Z
d µ ν b
d x (D Fµν )b (D C) = − dd x (D ν D µ Fµν )b C b =
Z Z
1 1
− dd x ([D ν , D µ ] Fµν )b C b = − dd x fbd a Faνµ Fµν
d b
C =0 .
2 2
i
A straightforward computation of the 2Rα,j Rj C β C α term in the commutator alge-
β
bra equation of Eq.(2.34) produces −2fba c ∂ µ C b − Adµ fde b C e C a which, after a lit-
tle algebra that makes use of Eq.(3.11), leads to D µc d fab d C b C a . Using this for
i
2Rα,j Rβj C β C α in Eq.(2.34), one concludes, as expected, that Tabc ji
= fab c and Eαβ = 0.
c c
When Tab = fab is used, the gauge-structure Jacobi equation in Eq.(2.35) leads to
3Adabc = −( LHS of Eq.(3.11)) so that Adabc = 0. Finally, the other consequence of
the Jacobi identity, namely Eq.(2.36), produces the tautology 0 = 0. All terms are
ji iδ kji
zero because the tensors Bαβγ , Dαβγ and Mαβγ are all zero. Higher-level equations
(which were not displayed in Sect. 2.4) are automatically satisfied because higher-level
tensors are identically zero.
leaves the action invariant, as a short calculation verifies. The two gauge transforma-
tions form a closed algebra since
Rbaµ = D µa b , aµ
Rbν = δba δνµ . (3.23)
The number of null vectors is equal to the number of gauge generators. It is easily
verified that Eq.(2.41) holds since Rcbµ Zac η a +Rcν Za η = D µb c δac η a +δcb δνµ (−D νc a η a ) =
bµ cν a
jδ ji
Eq.(2.39) holds because Aδαβγ = Fαβγc
= Dαβγ = 0. Eq.(2.40) holds because Bαβγ
kji
and Mαβγ are also zero. Finally, Eq.(2.42) is valid as long as Gjγ
aβ = 0 and
∂r S0 ∂r S0
= −Faµν = 0 , = Aaµ + D νa b Bνµ
b
=0 . (3.29)
∂Bµνa ∂Aµa
In spite of the presence of Lie-algebra structure constants fab c , the model has an
γ
abelian gauge algebra, i.e., Tαβ = 0, due to the fact that the vector field, which
appears in the covariant derivative, does not transform.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 32
The gauge transformations (3.28) have an on-shell null vector. Indeed, taking
Λbν = D νb c ξ c , (3.30)
Zabλ = D λb a . (3.32)
cµν bλ
Using Eq.(2.41), one finds Rbλ Za = 21 ǫµνρσ fab c Fρσ
b
so that
Vabρσ cµν
= −ǫµνρσ δ bd fad c ,
Vaji = 0 , if j = bν or i = bν , (3.33)
that is, Vaji = 0 if i or j corresponds to the field index of Abν . The derived quantities
ij
Aδαβγ and Bαβγ are zero. Other gauge-structure tensors also vanish:
γ ji jδ c kji
Tαβ = Eαβ = Dαβγ = Fαβγ = Mαβγ = Adaβ = 0 . (3.34)
The gauge-structure equations are all satisfied. Eq.(2.33) holds because the action
is invariant under gauge transformations, as is easily checked. Eq.(2.41) is satisfied. In
fact, it was used above in Eq.(3.33) to obtain Vaji . Eqs.(2.39) and (2.40) are satisfied
because all the tensors entering these equations are zero. Each term in Eq.(2.34) is
i
zero: The first term Rα,j Rβj is zero because Rα,ji
= 0 when j = aµν, i.e., when j is
j
a field index of B , and Rβ = 0 when j = aµ, i.e., when j is a field index of Aaµ .
aµν
The other terms vanish because the structure tensors vanish. Likewise each term in
γ
Eq.(2.42) is zero: The first term Zα,j Rβj is zero because Zα,j
γ i
like Rα,j is zero when
j = aµν.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 33
Let N be the number of generators of G, i.e., the dimension of the Lie algebra
G. The number of degrees of freedom ndof is 4N for Aaµ plus 6N for Bµν a
minus
the number of gauge transformations 4N plus the number of null vectors N, so that
Eq.(2.20) reads ndof = 7N.
δA = dλp−1 , (3.36)
where λp−1 is a p − 1 form. This gauge transformation has its own gauge invariance.
In fact, there is a tower of gauge invariances for gauge invariances given by
δλp−1 = dλp−2 ,
..
.
δλ1 = dλ0 ,
αs −1
The gauge generators at the s-th stage, Rsα s
, correspond to the exterior deriva-
(p−1−s)
tive d acting on the space of (p − 1 − s) forms, d :
R0 ↔ d(p−1) ,
(p−2)
R1 ↔ d ,
..
.
Rp−1 ↔ d(0) .
iαs−2
Because dd = 0, Eq.(2.19) holds off-shell and all Vsα s
are zero. With the excep-
tion of the gauge generators, all gauge-structure tensors are zero.
If n(p + 1) = d = dimension of space time, a topological term can be added to the
action Z
∆S0 = F | ∧F ∧ {z. . . ∧ F} . (3.38)
n terms
For n = 2 one can consider the quantization of ∆S0 alone if d = 2 (p + 1). This would
be another example of a topological theory.
Abelian p-form theories provide a good background for covariant open string field
theory, a non-abelian generalization of p-form theory which is infinite-stage reducible.
The ghost number of a star product of fields is the sum of their ghost numbers:
g(A ∗ B) = g(A) + g(B). The derivation Q increases the ghost number by 1: g(QA) =
g(A) + 1. In some circles, including refs.[270, 271], the ghost number is shifted by
−3/2 so that A has ghost number −1/2 instead of 1.
The axioms are satisfied for non-abelian Chern-Simons theory in three dimensions.
The field is a non-abelian vector potential which is converted into a Lie-algebra-
valued 1-form by multiplying by dxµ : Aa b ≡ Aaµb dxµ ≡ Asµ (Ts )a b dxµ , where the Ts
are a set of matrix generators for the Lie group. The derivation Q is the exterior
derivative d. The star product is the wedge product and a matrix multiplication:
(A ∗ B)a b = Aa c ∧ B c b . Integration is an integral over a three-dimensional manifold
R R
M without boundary and a trace over the Lie-algebra indices: A ≡ M Aa a . Axiom
(1) is satisfied because dd = 0. Axiom (2) holds because M has no boundary. Axiom
(3) is satisfied because the exterior derivative d is graded distributive across the wedge
product. Axiom (4) holds because both the wedge product and matrix multiplication
are associative. Finally, axiom (5) is satisfied because of the graded antisymmetry of
the wedge product and the cyclic property of a trace: T r(MN) = T r(NM) for any
R
two matrices M and N. In this example and in open string field theory, B = 0
unless B is a string 3-form.
The action for the string 1-form A is given by
Z Z
1 1
S0 [A] = A ∗ QA + A∗A∗A . (3.39)
2 3
Using the above five axioms it is straightforward to show that the action is invariant
under the gauge transformation
δA = QΛ0 + A ∗ Λ0 − Λ0 ∗ A , (3.40)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 36
F ≡ QA + A ∗ A = 0 . (3.41)
For the open bosonic string, Q is the BRST charge QBRST of the first-quantized
theory promoted to an operator by second quantization (see Eq.(7.51) with ρ =
1). The ∗ product is intuitively described as follows. Let σ be the parameter that
determines a point on the string, so that σ = 0 corresponds to one endpoint and
σ = π corresponds to the other endpoint. Divide the string in two at σ = π/2 and
call the two halves the left and right halves. Let C = A ∗ B. Then ∗ “glues” the
left half of A to the right half of B by delta functions and what remains is C so that
the left half of C is the right half of A and the right half of C is the left half of B.
See Fig. 1. The star operator can be thought of as matrix multiplication if the range
0 ≤ σ ≤ π/2 of points of the string is associated with one matrix index and the range
π/2 ≤ σ ≤ π is associated with the other matrix index. The integral operator is a
delta function equating the left and right halves. See Fig. 2. In the matrix analogy,
R
it is the trace. Precise definitions of Q, ∗ and in terms of the vibrational modes of
the string, i.e., particle excitations, can be found in refs.[68, 214, 142, 143].
σA = π σB = 0
A B
@
σA = 0 C @@
σB = π
σC = π σC = 0
Figure 1. The String Star Product
π
σ=
2
∫ ←→
σ=0 σ=π
Figure 2. The String Integral
Unlike the three-dimensional Chern-Simons theory, the open bosonic string theory
possesses q-forms for q < 0. As a consequence there are on-shell gauge invariances
for gauge invariances at all levels:
Rs ↔ D(−s) , (3.43)
when acting on any q-form Bq and A is the string gauge field. Eq.(2.19), when applied
to a −s form ghost η (s) , reads
The gauge field is contained in the einbein e of Sect. 3.1. It transforms as in Eq.(3.6).
The covariant time derivative Dτ is
1
Dτ xµ (τ ) ≡ ∂τ xµ (τ ) . (3.48)
e
Using Eq.(3.6) it is a simple exercise to verify that Dτ xµ (τ ) transforms as in Eq.(3.47).
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 39
δF = ε (τ ) ∂τ F , (3.49)
is invariant since
Z Z Z
dτ δ (eF ) = dτ (∂τ (εe) F + εe∂τ F ) = dτ ∂τ (εeF ) −→ 0 ,
xµ (τ ) → X µ (τ, θ) ≡ xµ (τ ) + θψ µ (τ ) , (3.52)
Dθ ≡ ∂θ − iθ∂τ ,
Q ≡ ∂θ + iθ∂τ . (3.53)
The 1-dimensional super-algebra is
[H, Q] = 0 . (3.54)
These equations are easily verified using Eq.(3.53). In theories for which a super-
symmetry charge Q exists, the hamiltonian necessarily has a non-negative spectrum
as a consequence of H = QQ ≥ 0. In addition, [H, Q] = 0 implies that there is
a fermionic state for every bosonic state and vice-versa, except possibly for a zero
energy state. The (anti)commutators involving Dθ are {Q, Dθ } = 0, [H, Dθ ] = 0, and
{Dθ , Dθ } = −2H.
The supersymmetry transformations are defined by δA = iξQA where ξ is an
anticommuting parameter. For X µ , a simple calculation gives δX µ = iξQX µ =
iξψ µ + θξ∂τ xµ . Given that δX µ is defined by δX µ ≡ δxµ + θδψ µ , one obtains for the
transformation rules for the components of X µ
δxµ = iξψ µ ,
δψ µ = ξ∂τ xµ . (3.55)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 41
The supersymmetric generalization of the action for the massless relativistic particle
consists in taking F in (3.56) as F = 2i Dθ X µ ∂τ Xµ since, when written in component
fields, it contains the correct kinetic energy term for xµ :
Z Z Z
i µ 1
S0 = dτ dθDθ X ∂τ Xµ = dτ (∂τ xµ ∂τ xµ − iψ µ ∂τ ψµ ) . (3.57)
2 2
It is easily checked that Eq.(3.57) is invariant under the transformations in Eq.(3.55).
One could add interactions by using a more general F (∂τ X µ , Dθ X µ , X µ ).
The final goal of this subsection is to implement both local translational and local
supersymmetry and construct a supergravity theory. Following the procedures above,
we promote the einbein e(τ ) to a superfield E: e (τ ) → E (τ, θ) where
E (τ, θ) ≡ e (τ ) + θχ (τ ) . (3.58)
τ , the right-hand side of Eq.(3.61) becomes (iξ∂θ + θξ∂τ ) A = iξQA, which is the flat
space supersymmetry transformation. Hence Eq.(3.61) reduces to expected results in
these two limits. The transformation law for E must be generalized from δe = ∂τ (εe).
As will become clear below when we construct invariant actions, one needs
δE = ∂τ η t E − ∂θ η θ E . (3.62)
δF (Dτ X µ , Dθ X µ , X µ ) = η t ∂τ F + η θ ∂θ F .
is a total derivative. The transformation rule for E in Eq.(3.62), including the mi-
nus sign on the right-hand side, was deliberately chosen so that EF would lead to
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 43
δA = εm ∂m A . (3.70)
One would like to find covariant derivatives Da A that transform in the same way:
δ(Da A) = εm ∂m (Da A). The vielbein ea m allows one to do this. Letting
Da A = ea m ∂m A , (3.71)
and requiring the correct transformation law for Da A, one finds that ea m must trans-
form as
δea m = εn ∂n ea m − ea n ∂n εm . (3.72)
Any function F of A and Da A will transform like A: δF (A, Da A) = εm ∂m F . Suppose
one can find a density e whose transformation law is
δe = ∂m (εm e) . (3.73)
The density e can be constructed from the determinant of the vielbein det (e. . ).
Note that det (e. . ) transforms as
δ det (e. . ) = δea m × cofactor matrix of ea m = δea m ea m det (e. . ) .
The derivative of det (e. . ) is
∂n det (e. . ) = ∂n ea m × cofactor matrix of ea m = ∂n ea m ea m det (e. . ) .
Using these results, a short calculation reveals that
1
e≡ .
= det (e. . ) (3.74)
det (e. )
transforms as in Eq.(3.73). The relativistic particle, presented as the example in Sect.
3.1, is a (0 + 1)-dimensional theory where the field e in that section is the same as
the density e defined here: e = det (e. . ) = e0 0 = 1/e0 0 .
In contrast to the metric gmn , which has d(d+1)/2 degrees of freedom, the vielbein
ea has d2 degrees of freedom. To compensate for this difference, one requires that the
m
theory be invariant under local Lorentz transformations that act on Lorentz indices
a, b, etc.. The transformation laws are
δea m = (εM)a b eb m , δA = 0 , (3.75)
b
where (εM)a b stands for 21 (εcd Mcd )a . Here, the d(d−1)/2 parameters εcd satisfy εcd =
−εdc and characterize the size of the infinitesimal-local-Lorentz transformations. The
d(d − 1)/2 matrix generators Mab satisfy the Lorentz algebra [Mab , Mcd ] = −ηac Mbd +
ηad Mbc + ηbc Mad − ηbd Mac and are antisymmetric in the lower indices c and d: Mcd =
−Mdc . When we write (Mcd )a b , a and b are matrix indices whereas c and d label
the different Lorentz transformations: Mcd produces an infinitesimal transformation
in the c–d plane. One explicit realization of Mcd is (Mcd )a b = ηac δdb − ηad δcb . The
extra d(d − 1)/2 gauge invariances guarantees that the physical numbers of degrees
of freedom in ea m and gmn are the same.
Given a number of fields Vai , with Lorentz indices, local invariants are constructed
by using the flat space metric η ab via η ab Vai Vbj or the flat space antisymmetric εa0 a1 ...ad−1
symbol via εa0 a1 ...ad−1 Vai00 Vai11 . . . Vaid−1
d−1 . A locally general coordinate and Lorentz in-
We thus have three types of local transformations: translations, Lorentz rotations and
scaling. Let us respectively associate ε. , ε.. and ε with each of these. In other words,
we use the number of indices on the infinitesimal parameter to distinguish the trans-
formations. The algebra of gauge transformations is as follows. The commutator of
two local translations is a local translation so that they form a subalgebra. Likewise
the commutator of two local Lorentz transformations is a local Lorentz transforma-
tion. The commutator of a translation and a Lorentz rotation is a Lorentz rota-
tion. Local translations and Lorentz rotations also form a subalgebra. Local scaling
transformations commute. They also commute with local Lorentz transformations.
Finally, the commutator of a scaling transformation and a translation produces a
scaling transformation. These last few statements correspond to
There are four gauge invariances, two translations, one Lorentz boost and one
scaling transformation. It is necessary for the target space to be 26 dimensions in
order to avoid an anomaly in one of these invariances. This is discussed in Sect. 9.3.
Hence in D =!26, the components of ea m can be fixed to constant values. Often
1 0
ea m = is used. Then Eq.(3.79) becomes a free field theory. However, in
0 1
gauge fixing, one should introduce ghosts. These ghosts play a somewhat minor role
in the first-quantized theory. However, in the second-quantized formulation, namely
the string field theory of Sect. 3.6, the ghosts as well as X µ (σ) become coordinates
of the string field A in Eq.(3.39).
Eq.(3.79) is classically equivalent to the usual Nambu-Goto action [197, 138].
Using the equations of motion of ea m to eliminate the vielbein, one finds, after some
algebra, that
q
1 ab
eη ηµν Da X µ Db X ν → (∂τ X µ ∂σ Xµ )2 − (∂τ X µ ∂τ Xµ )(∂σ X ν ∂σ Xν ) ,
2
which is the familiar Nambu-Goto action density.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 48
(ii) On the space of fields and antifields, one defines an odd symplectic
structure ( , ) called the antibracket.
(v) Finally, one finds solutions to the classical master equation subject to
certain boundary conditions.
The key result of this section is that the solution in steps (iv) and (v) leads to a
set of equations containing all relations defining the gauge algebra and its solution.
The action S is the generating functional for the structure functions. Hence, the field-
antifield formalism is a compact and efficient way of obtaining the gauge structure
derived in Sect. 2.
An additive conserved charge, called ghost number, is assigned to each of these fields.
The classical fields φi are assigned ghost number zero, whereas ordinary ghosts have
ghost number one. Ghosts for ghosts, i.e., level-one ghosts, have ghost number two,
etc.. Similarly, ghosts have opposite statistics of the corresponding gauge parameter,
but ghosts for ghosts have the same statistics as the gauge parameter, and so on,
with the statistics alternating for higher-level ghosts. More precisely,
Next, one introduces an antifield Φ∗A , A = 1, . . . , N, for each field ΦA . The ghost
number and statistics of Φ∗A are
h i
gh [Φ∗A ] = −gh ΦA − 1 , ǫ (Φ∗A ) = ǫ ΦA + 1 (mod 2) , (4.3)
Many properties of (X, Y ) are similar to a graded version of the Poisson bracket, with
the grading of X and Y being ǫX + 1 and ǫY + 1 instead of ǫX and ǫY . The antibracket
satisfies
(Y, X) = −(−1)(ǫX +1)(ǫY +1) (X, Y ) ,
((X, Y ) , Z) + (−1)(ǫX +1)(ǫY +ǫZ ) ((Y, Z) , X) + (−1)(ǫZ +1)(ǫX +ǫY ) ((Z, X) , Y ) = 0 ,
gh[(X, Y )] = gh[X] + gh[Y ] + 1 ,
3
This definition differs somewhat from the one of refs.[252, 257].
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 50
when one groups the fields and antifields collectively into z a : z a = {ΦA , Φ∗A }, a =
1, . . . , 2N. The expression for the antibracket in Eq.(4.8) is sometimes useful in ab-
stract proofs. The antibracket formalism can be developed in an arbitrary coordinate
system, in which case ζ ab is replaced by an odd closed field-dependent two-form. For
more details, see Sect. 10.7. However, locally there always exists a basis for which ζ ab
is of the form of Eq.(4.8) [218].
The antibracket in the space of fields and antifields plays a role analogous to
the Poisson bracket. Whereas the Poisson bracket is used at the classical level in a
hamiltonian formulation, the antibracket is used at the classical or quantum level in a
lagrangian formalism. One can use the antibracket in a manner similar to the Poisson
bracket. The antifield Φ∗A can be thought of as the conjugate variable to ΦA since
ΦA , Φ∗B = δBA . (4.9)
Infinitesimal canonical transformations [24] in which ΦA → Φ̄A and Φ∗A → Φ̄∗A , where
Φ̄A and Φ̄∗A are functions of the Φ and Φ∗ , are defined by
Φ̄A = ΦA + ε ΦA , F + O(ε2) , Φ̄∗A = Φ∗A + ε (Φ∗A , F ) + O(ε2) , (4.10)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 51
where F is an arbitrary function of the fields and antifields with gh[F ] = −1 and
ǫ(F ) = 1. A short calculation reveals that the canonical structure in Eq.(4.9) is
maintained:
Φ̄A , Φ̄∗B = δBA + O(ε2) , Φ̄A , Φ̄B = 0 + O(ε2 ) , Φ̄∗A , Φ̄∗B = 0 + O(ε2 ) .
(4.11)
Canonical transformations preserve the antibracket, i.e.,
∂r X ∂l Y ∂r X ∂l Y ∂r X ∂l Y ∂r X ∂l Y
∗
− ∗
= − + O(ε2 ) .
A
∂ Φ̄ ∂ Φ̄A ∂ Φ̄A ∂ Φ̄ A ∂Φ ∂ΦA ∂Φ∗A ∂ΦA
A ∗
is called the classical master equation. This simple looking equation is the main topic
of this subsection.
Not every solution to Eq.(4.16) is of interest. It is necessary to satisfy certain
boundary conditions. A relevant solution plays a double role. On one hand, a solu-
tion S is the generating functional for the structure functions of the gauge algebra. All
relations among structure functions are contained in Eq.(4.16), thereby reproducing
the equations in Sect. 2.4 and generalizing them to the generic L-th stage reducible
theory. On the other hand, S is the starting action to quantize covariantly the the-
ory. After a gauge-fixing procedure is implemented, one can commence perturbation
theory. The latter aspects are treated in Sect. 6.
Regard S as the action for fields and antifields. The variations of S with respect
Φ and Φ∗A are the equations of motion:
A
∂r S
=0 , (4.17)
∂z a
where the collective variables z a in Eq.(4.8) are used. We assume there exists a least
one stationary point for which Eq.(4.17) holds. We let Σ denote this subspace of
stationary points in the full space of fields and antifields:
( )
∂r S
Σ ←→ =0 . (4.18)
∂z a
An action S, satisfying the master equation (4.16), possesses its own set of gauge
invariances. Indeed, by differentiating Eq.(4.16) with respect to z b , one finds after a
little algebra that
∂r S a
R =0, a = 1, . . . , 2N , (4.20)
∂z a b
where
∂l ∂r S
Rab ≡ ζ ac c b . (4.21)
∂z ∂z
Although there appears to be 2N gauge invariances, not all of them are independent
on-shell as we now demonstrate. Differentiating Eq.(4.20) with respect to z d , mul-
tiplying by ζ cd, using the definition of Rab in Eq.(4.21) and imposing the stationary
condition in Eq.(4.17), one finds that
Rca Rab |Σ = 0 ,
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 53
where Σ denotes the space satisfying the on-shell condition in Eq.(4.18). One con-
cludes that Rab is on-shell nilpotent. It is an elementary result of matrix theory that a
nilpotent 2N × 2N matrix has rank less than or equal to N. Hence, at the stationary
point there exist at least N relations among the gauge generators Rab and therefore
the number 2N − r of independent gauge transformations on-shell is greater than or
equal to N, where r is the rank of the hessian of S at the stationary point:
∂l ∂r S
r ≡ rank . (4.22)
∂z a ∂z b Σ
Necessarily,
r≤N .
A solution to the master equation is called proper if
r=N , (4.23)
where r is given in Eq.(4.22). When r < N, there are solutions to ∂∂zr Sa δz a = 0 other
than the ones given by Eqs.(4.20) and (4.21).
Usually, only proper solutions are of interest. The reason is simple. The action S
contains the physical fields φi and the ghost fields necessary for quantization. How-
ever, the antifields Φ∗A , A = 1, . . . , N, are unphysical. If r = N, then the number of
independent gauge invariances of the type in Eq.(4.20) is the number of antifields. As
a consequence, one can remove the N non-physical variables Φ∗A , while maintaining
the N fields ΦA . At a later stage, the antifields can be eliminated through a gauge-
fixing procedure. Throughout the rest of this article, we restrict the discussion to the
proper case.
We have not yet specified the relation between S0 and S. To make contact with
the original theory, one requires the proper solution to contain the original action
S0 [φ]. This requirement ensures the correct classical limit. It corresponds to the
following boundary condition on S
∗ α
s−1
where Cs−1,αs−1
is the antifield of Cs−1 :
∗
Cs,αs
≡ (Csαs )∗ . (4.26)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 54
In the next subsection we obtain the leading terms in the solution for a first-stage
reducible theory. We also demonstrate how the gauge algebra is encoded in the
classical master equation, when the classical basis is used.
1 ∗ ∗
φj φk (−1)ǫj Eαβ kj
(−1)ǫα C β C α + . . . ,
4 ,i
∂l S i α ∗ ǫi ji a 1 ji ǫα β α
= R α C + φ j (−1) V a η + E (−1) C C
∂φ∗i 2 αβ
ǫδ ǫi ∗ iδ a β 1 iδ ǫβ γ β α
+ (−1) Cδ Gaβ η C − Dαβγ (−1) C C C +
2
1
+ (−1)ǫj φ∗j φ∗k Mαβγ kji
(−1)ǫβ C γ C β C α + . . . ,
4
∂r S 1
α
= φ∗i Rαi + Cγ∗ Tαβγ
(−1)ǫα C β + (−1)ǫi φ∗i φ∗j Eαβ ji
(−1)ǫα C β + . . . ,
∂C 2
∂l S 1 α 1
∗
= Zaα η a + Tβγ (−1)ǫβ C γ C β − (−1)ǫα φ∗i Dβγδ iα
(−1)ǫγ C δ C γ C β + . . . ,
∂Cα 2 2
∂r S 1
a
= Cα∗ Zaα + φ∗i φ∗j (−1)ǫi Vaji + . . . ,
∂η 2
∂l S 1 a
∗
= Aabα η b C α + Fαβγ (−1)ǫβ C γ C β C α + . . . .
∂ηa 2
Although ghosts do not depend on φi , we write ( ),i in the first equation to avoid
some minus signs.
Given the above equations, it is straightforward to form the sum over products of
derivatives in Eq.(4.16). We leave this step to the reader and simple quote the result
γ ji
[24, 26]: The classical master equation (4.16) is satisfied if the tensors, Tαβ , Eαβ , etc.
in Eq.(4.29) are the ones in Sect. 2.4. In other words, Eq.(4.29) with the tensors
identified as the ones in Sect. 2.4 is a proper solution to the master equation. One
finds that (S, S) = 0 implies the gauge structure in equations Eqs.(2.33), (2.34) and
(2.39) – (2.42). The reason why one equation (S, S) = 0 is able to reproduce many
equations is that the coefficients of each ghost and antifield term must separately be
zero. Up to overall factors, Eq.(2.33) is the term independent of antifields, Eq.(2.34)
is the coefficient of φ∗i which is bilinear in the C α , Eq.(2.39) is the coefficient linear in
Cδ∗ and trilinear in the C α , Eq.(2.40) is the coefficient of φ∗i φ∗j which is trilinear in the
C α , Eq.(2.41) is the coefficient linear in φ∗i and in η a , and Eq.(2.42) is the coefficient of
Cγ∗ which is linear in both η a and C β . The coefficients of higher-order terms produce
the higher-order gauge structure equations. For an irreducible system, the proper
solution is the above with η a and ηa∗ set to zero.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 56
Summarizing, the antibracket formalism using fields and antifields allows a simple
determination of the relevant gauge structure tensors. The proper solution to the
classical master equation is a compact way of expressing the relations among the
structure tensors.
can be added to S. Adding such a bilinear to S does not ruin the classical master
equation (S, S) = 0. When a non-minimal set of fields is employed, the proper
solution to the master equation is unique up to canonical transformations and the
addition of trivial pairs.
δB X ≡ (X, S) . (4.32)
δB S = 0 , (4.34)
as a consequence of (S, S) = 0.
The BRST operator δB is a nilpotent graded derivation4 : Given two functionals
X and Y ,
δB (XY ) = XδB Y + (−1)ǫY (δB X) Y , (4.35)
and
δB2 X = 0 . (4.36)
4
An alternative definition for δB given by
δB X ≡ (−1)ǫX (X, S) ,
is used by some authors. In this case, the δB continues to be a nilpotent graded derivation, but acts
from the left to right, i.e., (4.35) is replaced by
The nilpotency follows from two properties of the antibracket: the graded Jacobi iden-
tity and graded antisymmetry (see Eq.(4.5)). These properties imply ((X, S) , S) =
− ((S, S) , X) + (−1)ǫX +1 ((S, X) , S) = − ((S, S) , X) − ((X, S) , S) which leads to
((X, S) , S) = − 12 ((S, S) , X) = 0. Therefore, δB2 X = ((X, S) , S) = 0.
A functional O is a classical observable if δB O = 0 and O 6= δB Y for some Y . Two
observables are considered equivalent if they differ by a BRST variation [103, 105]. A
linear combination of observables is an observable. Because of the graded derivation
property of δB in Eq.(4.35), the product of two classical observables is BRST invariant.
Thus, observables form an algebra.
For a closed irreducible theory, the BRST transformation rules for ΦA and Φ∗A ,
which depend on the original action S0 and the structure tensors R and T , can be
obtained using Eqs.(4.29) and (4.33):
δB φi = Rαi C α + . . . ,
1 α
δB C α = Tβγ (−1)ǫβ C γ C β + . . . ,
2
1
δB φ∗i = − (−1)ǫi S0,i − (−1)ǫαǫi φ∗j Rα,i
j
C α − (−1)(ǫα +ǫβ +1)ǫi +ǫα Cγ∗ Tαβ,i
γ
Cβ Cα + . . . ,
2
δB Cα∗ = (−1)ǫα φ∗i Rαi + Cγ∗ Tαβ
γ
Cβ + . . . , (4.37)
where we have displayed the terms involving S0 , R and T . When the gauge-structure
iδ kji
equations hold off-shell at all levels, then the tensors Dαβγ , Mαβγ , etc. in Eq.(4.29) are
zero, and the proper action is linear in antifields. In this case, there are no additional
terms in Eq.(4.37) for irreducible systems. We illustrate BRST symmetry in a few of
the examples in the next section.
Because Sect. 4 has introduced many ideas, it is worth enumerating the most
important ones. After reading this section, one should know the following tools and
concepts: antifields and the antibracket, canonical transformations, the classical mas-
ter equation, properness and proper solution, classical BRST symmetry, and classical
observables.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 59
Two versions of the gauge transformations were presented. In the first, given in
Eq.(3.3), the only non-zero structure tensors are the gauge generators Rαi given in
Eq.(3.5). Using this and the general solution in Eq.(4.29), one finds
Z
C
∗ µ
S [Φ, Φ ] = S0 [x , e] + dτ x∗µ ẋµ + e C˙
∗
, (5.4)
e
where S0 [xµ , e] is given in Eq.(3.1). In the second version, given in Eq.(3.6), the
gauge generators Rαi and the commutator structure constant Tβγ α
are non-zero. Using
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 60
By the uniqueness theorem of the previous section, the solutions in Eqs.(5.4) and
(5.5) must be related by a canonical transformation. In fact, Eq.(5.4) is mapped to
Eq.(5.5) if xµ , x∗µ and e are unchanged but
1 1
C → eC , C∗ → C∗ , e∗ → e∗ − C ∗ C .
e e
One can check that this is a canonical transformation by verifying that the antibracket
structure in Eq.(4.11) is preserved. The infinitesimal version of the transformation is
generated by the fermion
F = ln (e) C ∗ C ,
when F is used in Eq.(4.10). Infinitesimally,
1
δC = ε ln (e) C , δC ∗ = −ε ln (e) C ∗ , δe∗ = −ε C ∗ C .
e
The finite transformation is obtained by iterating the infinitesimal transformation N
times, requiring Nε = 1 and then letting N → ∞. Alternatively, the full transforma-
tion can be generated using
Z !
C˜∗ C
F2 = dτ x̃∗µ xµ ∗
+ ẽ e + (5.6)
e
and Eq.(4.13).
It is straightforward to obtain the BRST transformation rules using Eq.(4.33).
From Eq.(5.4), one finds
ẋµ C
δB e = C˙ , δB xµ = , δB C = 0 ,
e
!
∗ 1 ẋ2 2 x∗µ ẋµ C
δB e = + m + ,
2 e2 e2
!
d ẋµ + x∗µ C
δB x∗µ = , (5.7)
dτ e
x∗µ ẋµ
∗
δB C = − ė∗ .
e
It is a useful exercise to verify the nilpotency of δB when acting on any field or
antifield. One must be careful of signs. In this regard Eq.(4.35) is useful.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 61
δB Aaµ = D µa b C b ,
1
δB C a = fbc a C c C b ,
2
δB A∗aµ = − (D ν Fνµ )a + fab c A∗cµ C b , (5.10)
δB Ca∗ = − D µ A∗µ + Cc∗ fab c C b .
a
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 62
The non-zero gauge tensors are the gauge generators Rαi given in Eq.(3.23), the com-
γ
mutator structure constants Tαβ given in Eq.(3.25), the null vectors Zaα given in
Eq.(3.24), and Aabα given in Eq.(3.26). Other tensors are zero. Inserting the non-zero
tensors in Eq.(4.29), one finds [137]
Z
1 a ∗ µν 1
S= 4
dx Fµν Fa + A∗aµ D µa b C b + C aµ + Cc∗ fab c C b C a
4 2
o
∗
+Ccµ fab c C bµ C a + Cb∗ η b − Caµ
∗
D µa b η b + ηc∗ fba c η a C b . (5.12)
The non-zero structure tensors Rαi , Zaα and Vaij are given in Eqs.(3.31) – (3.33). The
proper solution is
Z
1 1 ∗
S = S0 + d x Ba∗κλ εκλµν D µa b C bν + Cbµ
4 ∗
D µb a η a − Bcρσ ∗b µνρσ
Bµν ε fab c η a , (5.14)
2 8
where S0 is given in Eq.(3.27). An effect of the on-shell reducibility is the appearance
of terms quadratic in the antifields.
where ∧ is the wedge product and a “ ∗ ” in front of a field or antifield indicates the
dual star operation. In Eq.(5.16), some antifields have been redefined by a minus
sign factor compared to the definitions in Sect. 2. As a consequence of integration
by parts and dd = 0, as well as the definitions of the Hodge star operation and the
wedge product, one can check that (S, S) = 0. Indeed,
Z Z
∂r S ∂l S
∝ (∗ d∗ F ) ∧ ∗ dC0 ∝ d (∗ F ∧ dC0 ) = 0 ,
∂Aµ1 µ2 ...µp ∂A∗µ1 µ2 ...µp
∂r S ∂l S
∗µ1 µ2 ...µp−i−1 ∝
∂ (Ci )µ1 µ2 ...µp−i−1 ∂ (Ci )
Z Z
∗ ∗
( d (Ci∗ )) ∗
∧ dCi+1 ∝ d (∗ (Ci∗ ) ∧ dCi+1 ) = 0 , i = 0, 1, . . . , p − 1 ,
∗
where C−1 ≡ A∗ .
Here, g (ϕi ) is the ghost number of ϕi which is the same as the order of the string
form, and ǫ (ϕi ) is given by Eq.(4.2).
The quantization of the bosonic string was carried out by C. Thorn [246] and by
M. Bochicchio [56]. Since this subject has been reviewed in ref.[247], we keep the
discussion brief. To gain some insight in finding the proper solution, let us compute
the non-zero terms in Eq.(4.29). The classical action S0 is given in Eq.(3.39). The
γ
Rαi , Zaα , Vaij and Tαβ terms in Eq.(4.29) are respectively
Z
∗
A∗ ∗ (QC0 + A ∗ C0 + C0 ∗ A) ,
Z
∗ ∗
C0 ∗ (QC1 + A ∗ C1 + C1 ∗ A) ,
Z
∗
A∗ ∗ ∗ A∗ ∗ C1 , (5.20)
Z
∗ ∗
C0 ∗ C0 ∗ C0 ,
where ∗ before a field is the string analog of the dual star operation determined by
R
the bilinear form A ∗ B. It takes p-forms into 3 − p forms. Note that the structure
of the above terms is similar to the classical action S0 evaluated using various fields
and antifields. The sign of the term C0 ∗ A in the first equation is opposite to that of
Eq.(3.42) because C0 has odd statistics. The field-antifield statistics, ghost number
and total statistics for the fields are
Note that the total statistics of Ci and the ∗ Ci∗ is odd, so that it makes sense to define
a field which is the formal sum
s∗+∗3 3 2 1 0 −s
Ψ ≡ ...+ Cs + . . . + ∗ C0∗ + ∗ A∗ + A + C0 + . . . + Cs + . . . ,
∞
X
Ψ= ψp , where ψ−p ≡ Cp , for p ≥ −1 , ψp ≡ ∗ Cp−3
∗
, for p ≥ 2 . (5.23)
p=−∞
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 65
In the first equation, the order of the string form is denoted above the field. The
total statistics of the field-antifield Ψ is odd: s (Ψ) = 1. The terms in Eq.(5.20) and
the classical action are both contained in the following ansatz for the proper solution
Z Z
1 1
S= Ψ ∗ QΨ + Ψ∗Ψ∗Ψ=
2 3
∞ Z ∞ Z
1 X 1 X
ψ2−p ∗ Qψp + ψp ∗ ψq ∗ ψ3−p−q . (5.24)
2 p=−∞ 3 p=−∞
q=−∞
By obtaining the L-th stage reducible proper solution, computing the non-zero
structure tensors, one could derive Eq.(5.24). Instead let us check the ansatz. One
simply needs to verify the classical master equation
Z
?
0 = (S, S) ∝ (QΨ + Ψ ∗ Ψ) ∗ (QΨ + Ψ ∗ Ψ) . (5.25)
where axioms (1), (2) and (3) of Sect. 3.6 have been used. Note that
Z
0= Q (Ψ ∗ Ψ ∗ Ψ) =
Z Z Z
(QΨ ∗ Ψ ∗ Ψ) − (Ψ ∗ QΨ ∗ Ψ) + (Ψ ∗ Ψ ∗ QΨ) =
Z Z
3 (QΨ ∗ Ψ ∗ Ψ) = 3 (Ψ ∗ Ψ ∗ QΨ) ,
R R
where axioms (2)–(5) are used. This implies that (QΨ ∗ Ψ ∗ Ψ) = (Ψ ∗ Ψ ∗ QΨ) =
0, leading to the vanishing of two of the terms in Eq.(5.25). The last term in Eq.(5.25)
is zero when axioms (4) and (5) are combined to give
Z Z Z
(Ψ ∗ Ψ) ∗ (Ψ ∗ Ψ) = (Ψ ∗ Ψ ∗ Ψ ∗ Ψ) = − (Ψ ∗ Ψ ∗ Ψ ∗ Ψ) ,
where Eq.(5.18) is used in the last step. Since the master equation is satisfied and
the suitable boundary conditions are correctly implemented, Eq.(5.24) is a proper
solution. The classical action in Eq.(3.39) and the proper solution in Eq.(5.24),
although structurally identical, differ in that the field entering the action is different.
In S0 it is the string one-form A, while in S it is the tower Ψ, given in Eq.(5.23),
which includes ghosts and antighosts as well as A.
The BRST transformation rules for the fields and antifields are
δB Ψ = QΨ + Ψ ∗ Ψ . (5.26)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 66
δB2 Ψ = QδB Ψ + Ψ ∗ δB Ψ − δB Ψ ∗ Ψ =
Q2 Ψ + Q (Ψ ∗ Ψ) + Ψ ∗ QΨ + Ψ ∗ (Ψ ∗ Ψ) − QΨ ∗ Ψ − (Ψ ∗ Ψ) ∗ Ψ = 0 .
The last equality holds due to the axioms of open string field theory: The first term
is zero because of axiom (1), the nilpotency of Q; the second, third and fifth terms
cancel because of axiom (3), the graded distributive property of Q; and the fourth
and sixth terms cancel by axiom (4), the associativity of the star product.
values σ and τ and C ba = −C ab . Combining these ghosts with the original fields X µ
and ea m , one arrives at the following set of fields and antifields:
n o n o
ΦA = X µ , ea m , C m , C ab , C , Φ∗A = Xµ∗ , e∗a m , Cm
∗ ∗
, Cab , C∗ , (5.29)
where µ ranges from 0 to D − 1 = 25. The only non-zero terms in Eq.(4.29) are S0 ,
γ
φ∗i Rαi C α and 12 Cγ∗ Tαβ (−1)ǫα C β C α . These tensors are computed in Sect. 3.8. One thus
obtains Z Z π n
S = S0 + dτ dσ Xµ∗ C m ∂m X µ +
0
∗a
e m C ∂n ea − ea ∂n C m + ea m C + eb m ηac C cb
n m n
o
∗ n ∗
−Cm C ∂n C m + Cab C ac C bd ηcd − C n ∂n C ab − C ∗ C n ∂n C , (5.30)
where S0 is given in Eq.(3.79).
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 68
6.1 Generalities
Although ghost fields have been incorporated into the theory, the field-antifield
action S still possesses gauge invariances (See Eq.(4.20)). Hence it is not yet suitable
for quantization via the path integral approach. A gauge-fixing procedure is needed.
The theory also contains many antifields that usually one wants to eliminate before
computing amplitudes and S-matrix elements. One cannot simply set the antifields to
zero because the action would reduce to the original classical action S0 , which is not
appropriate for commencing perturbation theory due to gauge invariances. Following
ref.[24], antifields are eliminated by using a gauge-fixing fermion Ψ (Φ) via
∂Ψ
Φ∗A = . (6.1)
∂ΦA
Note that Ψ is a functional of fields only. It does not matter whether right or left
∂l Ψ ∂r Ψ
derivatives are used in Eq.(6.1) because ∂Φ A = ∂ΦA since ǫ (Ψ) = 1 (see Eq.(A.4) of
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 69
To quantize the theory, let us use the path integral approach with the constraint
in Eq.(6.1) implemented by a delta function:
Z !
∗ ∂Ψ i
IΨ (X) = [dΦ] [dΦ ] δ Φ∗A − A
exp W [Φ, Φ∗ ] X [Φ, Φ∗ ] , (6.5)
∂Φ h̄
Z
∂r I ∂l δΨ 2
= [dΦ] + O (δΨ)
∂Φ∗A ∂ΦA
Z
= [dΦ]∆IδΨ + O (δΨ)2 ,
where integration by parts (see Eq.(A.5) in Appendix A) has been used, and where
the operator ∆ is defined to be [24, 26]
∂r ∂r
∆ ≡ (−1)ǫA+1 . (6.7)
∂ΦA ∂Φ∗A
It is a kind of “nilpotent divergence operator” in the space of fields and antifields. It
formally satisfies
∆∆ = 0 ,
∆ (XY ) = X∆Y + (−1)ǫY (∆X) Y + (−1)ǫY (X, Y ) ,
∆ ( (X, Y ) ) = (X, ∆Y ) − (−1)ǫY (∆X, Y ) . (6.8)
Note that
gh [∆] = ǫ (∆) = 1 . (6.9)
According to the above calculation, the integral IΨ (X) is infinitesimally independent
of Ψ if
∆I = 0 . (6.10)
Eq.(6.10) is a requirement of a “good” integrand.
The path integral itself should be gauge independent. This corresponds to setting
X = 1 in IΨ (X), from which one finds
i i i 1
∆ exp W = exp W ∆W − 2 (W, W ) = 0 , (6.11)
h̄ h̄ h̄ 2h̄
so that one needs
1
(W, W ) = ih̄∆W . (6.12)
2
Eq.(6.12) is known as the quantum master equation [24, 26]. When W satisfies
Eq.(6.12), X in Eq.(6.6) must satisfy
the quantum master equation (Eq.(6.12)) implies that the first term S and quantum
correction terms Mp must satisfy
1 n−1
X
(Mn , S) = i∆Mn−1 − (Mp , Mn−p ) , for n ≥ 2 . (6.15)
2 p=1
By the correspondence principle, S in Eq.(6.14) is identified as the classical field-
antifield action and is consistent with Eqs.(6.12) and (6.15) because it satisfies the
classical master equation (S, S) = 0.
Likewise, when X in Eq.(6.13) is expanded in an h̄ series
∞
X
X= h̄p Xp , (6.16)
p=0
(X0 , S) = 0 ,
n
X
(Xn , S) = i∆Xn−1 − (Xn−p , Mp ) , for n ≥ 1 . (6.17)
p=1
The first equation says that the “classical” part of the quantum operator X is clas-
sically BRST invariant.
C¯sα
0
s
≡ C¯sαs , 0
π̄sαs
≡ π̄sαs . (6.23)
The following equations also apply to the minimal sector fields when k = −1 and 0.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 73
..
.
Diagram 2. The Triangular Field Tableau
The index s of Cskαs or C¯sα
k
s
labels different horizontal rows, whereas the index k
labels different rows slanting to the right and downward.
The new proper solution Snm of the classical master equation, involving the non-
minimal set of fields, is given by
where S is the proper solution for the minimal set of fields presented in Sect. 4 and
Saux is the trivial-pair proper solution for the auxiliary fields given by
2 ¯2∗α2
π̄2α C
2 2
1∗
+ C2α π 1α2 + π̄2α2 C¯2∗α2 +
2 2
2 ¯2∗α3
3∗
C3α π 3α3 + π̄3α
3 3
C
3 3
1∗
+ C3α π 1α3 + π̄3α3 C¯3∗α3 +
3 3
...
or
L
X L
X L
X L
X
k ¯k∗αs k∗ kαs
Saux = π̄sα C
s s
+ Csα π
s s
. (6.28)
k=0 s=k k =1 s=k
k even k odd
The auxiliary antighost fields at level s are eliminated from the theory using
Eq.(6.1) which leads to
∂Ψ s
C¯sk∗αs = ¯k , k = 0, 2, 4, . . . , 2 ,
∂ Csαs 2
k∗ ∂Ψ s−1
Csα = , k = −1, 1, 3, 5, . . . , 2 +1 , (6.29)
s
∂Cskαs 2
at each level s.
∂Ψ ∂Ψ
C¯sk∗αs = ¯k = 0 , k∗
Csα = =0 , (6.30)
∂ Csαs s
∂Cskαs
would not exist. In the rest of this section, we determine the conditions on Ψ so that
an admissible gauge-fixing procedure arises. The discussion is somewhat technical so
that a reader may wish to skip to Sect. 6.5. For the irreducible case, one should read
to Eq.(6.40).
It is useful to define ns as the net number of non-gauge degrees of freedom for the
ghost Csαs at level s:
L
X
ns ≡ (−1)t−s mt . (6.31)
t=s
indicates that the gauge degrees of freedom in ΦA are eliminated by the equation
∂Ψ
∂Φ∗B
= 0. To ensure their elimination, two points on the gauge slice determined
by Ψ should not be related by an infinitesimal gauge transformation. This leads to
conditions on Ψ that are presented below. A downward-slanting arrow going from
ΦA to ΦB such as
ΦA
/
ΦB
Diagram 4. Elimination of Non-Gauge Degrees of Freedom
to indicate that non-gauge degrees of freedom are eliminated. In contrast, there are
upward-sloping lines meaning that the gauge degrees of freedom in φi , Csαs and C¯sαs
are fixed. The fields Cskαs , for k ≥ 1 and odd k, and C¯sαk
s
, for k ≥ 2 and even k, are
non-propagating, i.e., there is no quadratic form in the action for these fields. When
gauge-fixing is fully implemented, they are completely fixed. Like Csαs and C¯sαs , it will
be useful to think of Cskαs and C¯sαk
s
as having ns independent (or non-gauge) degrees
of freedom and ns+1 gauge degrees of freedom. Since there are both upward-sloping
and downward-sloping lines terminating on these fields, both gauge and independent
degrees of freedom are fixed. Hence, all ms = ns+1 + ns fields are removed from the
theory. The difference between Csαs and C¯sαs and the other ghosts and antighosts is
that the independent degrees of freedom in these fields are propagating. The fields
αs−1
C¯sαs play a dual role: They fix the gauge degrees of freedom in Cs−1 and serve as
αs
the antighosts of Cs , meaning that they enter quadratically in the action with these
fields.
Let Φ0 be a solution to the equations of motion for the action SΨ fixed by the
gauge conditions in Eq.(6.1). Since one is interested in performing a perturbative
expansion about this solution, write
ΦA = ΦA
0 + δΦ
A
, (6.32)
∂l ∂r Ψ
The matrix ∂ C̄0α0 ∂φi
that multiplies δφi must eliminate n0 gauge degrees of free-
0
dom. This requires that Ψ be such that
!
∂l ∂r Ψ
rank = n0 . (6.35)
∂ C¯0α0 ∂φi 0
The gauge modes of φi are proportional to Rαi 0 . Hence one also needs
!
∂l ∂r Ψ i
rank R = n0 , (6.36)
∂ C¯0α0 ∂φi β0 0
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 77
to ensure that the correct degrees of freedom in φi are fixed. Using Eq.(4.28) and
∂Ψ
φ∗i = ∂φ ¯
i , one sees that the operator in Eq.(6.36) is the quadratic form for C0α0 and
Hence Eq.(6.36) is consistent with the fact that C¯0α0 and C0β0 should have n0 propa-
gating degrees of freedom.
If the system is irreducible, n0 = m0 and no further constraints on the gauge-
fixing fermion Ψ are necessary. Eqs.(6.35) and (6.36) are the only requirements on
Ψ. There are n − m0 propagating modes for φi and m0 propagating modes for C¯0α0
and C0α0 . Of the original n degrees of freedom of φi , m0 have been gauge fixed. The
simplest gauge-fixing fermion Ψδ is [26]
Ψδ = C¯0α0 χα0 (φ) , (6.38)
where χα0 is an arbitrary functional of φ. The subscript δ on Ψ indicates that this is
a delta-function gauge-fixing scheme. The condition 0 = C¯0∗α0 in Eq.(6.33) leads to
χα0 (φ) = 0 , (6.39)
so that χα0 is the gauge-fixing condition on φ. Eqs.(6.35) and (6.36) respectively
become
rank χα,i0 (φ0 ) = n0 ,
rank χα,i0 Rβi 0 (φ0 ) = n0 , (6.40)
where n0 = m0 for the irreducible case.
Suppose the theory is at least first-stage reducible. Then C¯0α0 and C0α0 have m0
degrees of freedom but n1 of these are gauge modes and thus m0 − n1 = n0 are
independent. According to the Triangular Field Tableau of Diagram 2, the equation
C¯1∗α1 = 0 is used to fix the n1 gauge degrees of freedom in C0α0 , and C1α
1∗
1
= 0 is used to
¯
fix the n1 gauge degrees of freedom in C0α0 . Expanding the former condition about
the perturbative solution, one obtains
! !
∂Ψ ∂Ψ ∂l ∂r Ψ
0 = C¯1∗α1 = ¯ = + δC0α0 + . . . . (6.41)
∂ C1α1 ∂ C¯1α1 0
¯
∂ C1α1 ∂C0α0 0
Actually, the first term on the right-hand side of Eq.(6.41) vanishes, but we display it
to emphasize the idea that we are expanding about the perturbative solution. Since
one wants to fix n1 gauge modes of C0α0 , a necessary condition on Ψ is that
!
∂l ∂r Ψ
rank ¯ = n1 . (6.42)
∂ C1α1 ∂C0α0 0
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 78
The modes to gauge-fix are those that are non-propagating in Eq.(6.37). These modes
are proportional to
−1α0 α0
Z1β 1
≡ R 1β 1
, (6.43)
0
of Eq.(2.16). Since there are n1 such modes,
−1α0
rank Z1β 1
= n1 . (6.44)
Finally one needs to fix the n1 independent or “non-gauge” modes of C11α1 . From
Diagram 2, one sees that this is done using
! !
∂Ψ ∂Ψ ∂l ∂r Ψ
0= C¯0∗α0 = ¯ = + δC11α1 + . . . , (6.51)
∂ C0α0 ∂ C¯0α0 0 ∂ C¯0α0 ∂C11α1 0
A necessary condition that the ns+1 gauge modes of Cskαs can be fixed is
!
∂l ∂r Ψ
rank = ns+1 , (6.55)
∂ C¯s+1αs+1 ∂Cskαs
k+1
0
where k is odd and ranges between s ≥ k ≥ −1, and s is restricted to L−1 ≥ s ≥ −1.
When there is a diagram such as
C¯s−1α
k−1
s−1
/
Cskαs
Diagram 6. Elimination of Independent Degrees of Freedom in Cskαs
kαs
A sufficient condition that these gauge modes Zs+1β s+1
can be removed from Cskαs via
Eq.(6.54) is that
!
∂l ∂r Ψ
rank Z kαs = ns+1 . (6.60)
∂ C¯s+1αs+1 ∂Cskαs s+1βs+1
k+1
0
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 81
When k = −1 the operator in Eq.(6.60) is the quadratic form of SΨ for C¯s+1αs+1 and
βs+1
Cs+1 , as can be seen from Eqs.(4.28) and (6.29). The condition that these fields have
ns+1 propagating modes is consistent with Eq.(6.60) for the case k = −1.
The discussion for C¯sα k
s
where k is even works similarly. The ns independent
degrees of freedom of C¯sαs , via the diagram
k−1αs−1
Cs−1
/
C¯sα
k
s
Diagram 7. Elimination of Independent Degrees of Freedom in C¯sα
k
s
are fixed by
∂Ψ ∂l ∂r Ψ
k−1∗
0 = Cs−1α = k−1αs−1 = . . . + δ C¯sα
k
k−1αs−1
+ ... , (6.62)
∂ C¯sα
s−1 s k ∂C
∂Cs−1 s s−1 0
which requires
∂l ∂r Ψ
rank = ns .
k ∂C k−1αs−1
∂ C¯sαs s−1 0
kαs+1
More precisely, the Z̄s+1αs
are defined by
kα
s+1
∂l ∂r Ψ
Z̄s+1α =0 , (6.64)
s
k ∂C k−1αs−1
∂ C¯sαs s−1 0
C¯sα
k
s
7
k+1α
Cs+1 s+1
Diagram 8. Elimination of Gauge Degrees of Freedom in C¯sα
k
s
are fixed by
∂Ψ ∂l ∂r Ψ
k+1∗
0 = Cs+1α = k+1α = . . . + δ C¯sα
k
k+1α
+ ... , (6.65)
s+1
∂Cs+1 s+1
s
¯k
∂ Csαs ∂Cs+1 s+1 0
which requires
∂l ∂r Ψ
rank k+1αs+1
= ns+1 .
∂ C¯sα
k ∂C
s s+1 0
where, for each upward-sloping line (like in Diagram 5 with k → k − 1 and s → s − 1),
kαs
one associates the matrix ωsα s−1
and where, for each downward-sloping line (like in
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 83
kαs−1
Diagram 6), one associates the matrix ω̄sα s
. In addition to Eq.(6.40), one requires
Eqs.(6.55), (6.60), Eq.(6.57), and (6.66), that is,
kαs
rank ωsαs−1
(φ0 ) = ns ,
kαs k−1αs−1
rank ωsαs−1
(φ0 ) Zsβs = ns , (6.68)
for k even, s ≥ k ≥ 0 and L ≥ s ≥ 1, as well as
kαs−1
rank ω̄sαs
(φ0 ) = ns ,
k−1βs kαs−1
rank Z̄sαs−1
ω̄sαs (φ0 ) = ns , (6.69)
for k odd, s ≥ k ≥ 1 and L ≥ s ≥ 1.
1 L−1
X L
X
Ψπ = C¯sα
k
ρkαs (φ) πsk+1βs + π̄sα
s sβs
k
ρkαs (φ) Csk+1βs
s sβs
2 k=0 s=k+1
k even
L
1 X
+ C¯kα
k
σ kαk βk (φ) π̄kβ
k k
k
, (6.77)
2 k=0 k
k even
kαk βk
where ρkα
sβs and σk
s
are arbitrary matrices. If one eliminates antifields by Eq.(6.1)
and inserts the result into Eq.(6.28), one obtains the following quadratic terms for π
fields
L−1
X L
X
Saux |ΣΨ = . . . + k
π̄sα ρkαs πsk+1βs
s sβs
k=0 s=k+1
k even
L
1 X
+ k
π̄kα σ kαk βk π̄kβ
k k
k
+ ... , (6.78)
2 k=0 k
k even
thereby allowing a gaussian average procedure for gauge invariances at all levels.
In some instances, in lieu of Eq.(6.1), one might want to eliminate certain fields
rather than antifields. The simplest way to accomplish this is to first perform a
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 85
canonical transformation that interchanges some fields for antifields [104, 230, 259,
44, 252, 257]. Then, when Eq.(6.1) is used, elimination of certain antifields will
correspond to the elimination of some original fields. Given the freedom to first
perform canonical transformations, Eq.(6.1) is quite general.
This is the same result as first performing the non-gauge-fixed BRST transformation
in Eq.(4.32) and then imposing the gauge-fixing fermion condition in Eq.(6.1). The
BRST transformation is a global symmetry of the classical gauge-fixed action since
∂r SΨ ∂l S
δBΨ SΨ = (SΨ , S)|ΣΨ = =
∂ΦA ∂Φ∗A ΣΨ
!
∂r S ∂l S ∂r S ∂r ∂r Ψ ∂l S
+ =0 , (6.81)
∂ΦA ∂Φ∗A
ΣΨ
∂Φ∗B ∂ΦA ∂ΦB ∂Φ∗A ΣΨ
where the first term vanishes as a consequence of the classical master equation and
the second term vanishes due to statistical symmetry properties of the factors.
Due to the elimination of antifields, δBΨ is not necessarily off-shell nilpotent. For
example, consider
∂r ∂l S ∂l S ∂r ∂l S ∂r ∂l Ψ ∂l S
δB2 Ψ ΦA = + ,
∂ΦC ∂Φ∗A ∂Φ∗C ΣΨ ∂Φ∗C ∂Φ∗A ∂ΦD ∂ΦC ∂Φ∗D ΣΨ
where the second term arises from the chain rule (see Eq.(A.7) of Appendix A) via
the implicit dependence on ΦD through Φ∗C , which is now a functional of the Φ via
Eq.(6.1). By using the identity 0 = ∂l∂Φ
(S,S)
∗ , the first term can be rearranged and then
A
the two terms combine if one notes that
∂r (SΨ ) ∂r S ∂r S ∂r ∂l Ψ
= + . (6.82)
∂ΦC ∂ΦC ΣΨ ∂Φ∗D ∂ΦC ∂ΦD ΣΨ
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 86
The equations of motion for the gauge-fixed action are ∂r∂Φ (SΨ )
C = 0. Hence, the gauge-
fixed BRST symmetry is on-shell nilpotent. From Eq.(4.29), one sees that a necessary
ji
condition for off-shell nilpotency is that the algebra is closed, Eαβ = 0, and that there
ji
is off-shell reducibility, Va = 0.
Another way of obtaining the on-shell nilpotency is to note that the gauge-fixed
BRST transformation agrees with the ordinary BRST transformation evaluated at
ΣΨ , up to equations of motion. For functionals which do not depend on antifields
this follows from Eq.(6.79). For functionals, which depended on antifields before
gauge-fixing but are now evaluated at ΣΨ , the result follows because
∂l (SΨ )
δBΨ Φ∗A (Φ) = (δB Φ∗A )|ΣΨ + . (6.84)
∂ΦA
The last term is a gauge-fixed equation of motion. In other words, the processes
of performing the off-shell BRST transformation in Eq.(4.32) and gauge-fixing via
Eq.(6.1) commute up to equations of motion.
!
∂Ψ BA
+Φ̃∗A A
R + R + ...
∂ΦB
1
+ Φ̃∗A Φ̃∗B RBA + . . . + . . .
2
1 h i
≡ S̃0 [Φ] + Φ̃∗A R̃A + Φ̃∗A Φ̃∗B R̃BA + . . . ≡ S̃ Φ, Φ̃∗ . (6.87)
2
The terms in the Φ̃∗ expansion have an interpretation. The Φ̃∗ -independent term
∂r X A
is the generator of gauge-fixed BRST transformations since δBΨ X = ∂Φ A R̃ . Hence,
which is the invariance of the gauge-fixed action under δBΨ . The vanishing of the
second coefficient in Eq.(6.90) leads to the equation
∂r R̃B A BA ∂r SΨ
R̃ = −R̃ . (6.92)
∂ΦA ∂ΦA
Eq.(6.92) is related to the on-shell nilpotency of δBΨ :
!
∂r ∂r SΨ A
0 = δBΨ (δBΨ SΨ ) = R̃ R̃B =
∂ΦB ∂ΦA
of Eq.(6.88). The proper condition in Eqs.(4.22) and (4.23) becomes [259, 44, 252, 133]
∂l ∂r SΨ [Φ]
rank =N . (6.94)
∂ΦA ∂ΦB { ∂r SΨ =0}
∂ΦD
Eq.(6.94) ensures that propagators for the ΦA are defined so that the usual pertur-
bation theory can be developed.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 89
7 Gauge-Fixing Examples
In this subsection, we present some gauge-fixing procedures for the examples
considered in Sects. 3 and 5. The first step is to introduce the auxiliary fields and
antifields necessary for gauge-fixing, as specified in Sect. 6.2. The second step is to
choose an appropriate Ψ. Here, there is quite a bit of freedom and we make specific
choices. The third step is to eliminate antifields using Ψ and substitute the results
into S to obtain the gauge-fixed action SΨ .
The reader interested in doing exercises can try the following. (i) Derive the
equations of motion of the gauge-fixed action. (ii) Determine the effect of the gauge-
fixed BRST transformation δBΨ on the fields. (iii) Check the gauge-fixed BRST
invariance of the gauge-fixed action. (iv) Compute δB2 Ψ and verify that the non-
zero terms, when present, vanish if the equations of motion are used. (v) Perform a
gauge-fixing procedure with a different gauge-fixing fermion Ψ.
Because the original algebra is closed, the gauge-fixed BRST transformation δBΨ ,
given by
ẋµ C
δBΨ e = C˙ , δBΨ xµ = , δBΨ C = 0 ,
e
δBΨ π̄ = 0 , δBΨ C¯ = π̄ , (7.5)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 90
satisfies δBΨ δBΨ = 0 off-shell. After integrating over π̄, the gauge e = 1 is implemented
and Eq.(7.4) becomes
Z
1 2
SΨ → dτ ẋ − m2 + C¯C˙ . (7.6)
2
Since the equations of motion for π̄ and e have been used, Eq.(7.6) is no longer
invariant under Eq.(7.5). To derive the modified BRST transformations, start with
Eq.(7.1), and perform the shift, C¯∗ → C¯∗ + e − 1, e∗ → e∗ + C, ¯ to the gauge-fixed
∂r SΨ
basis. Then, determine the equation of motion for π̄ using ∂e = 0 and perform
another canonical transformation to shift π̄ by this solution: π̄ → π̄ + 12 (ẋ2 /e2 + m2 ).
The gauge-fixed BRST transformations of the transformed action at the saddle point,
e = 1, π̄ = 0, given by
1 2
δB̃Ψ xµ = ẋµ C , δB̃Ψ C = 0 , δB̃Ψ C¯ = ẋ + m2 , (7.7)
2
constitute a symmetry of the action in Eq.(7.6), as can easily be checked. Because
equations of motion have been used, δB̃Ψ is not longer nilpotent off-shell. Indeed, a
computation of δB̃2 Ψ reveals that it is, in general, nonzero; it is zero if the equations
of motion, ẍµ = 0 and C˙ = 0 are used.
δBΨ Aaµ = D µa b C b ,
1
δBΨ C a = fbc a C c C b ,
2
δBΨ C¯a = Ba , (7.12)
δBΨ Ba = 0 .
The nilpotency of δBΨ holds off-shell because the original gauge algebra is closed. The
gaussian integration over Ba can be performed for Eq.(7.11) to give
Z ( )
1 a µν ξ
SΨ → d
d x − Fµν Fa − ∂µ C¯a D µa b C b + ∂ µ Aaµ ∂ ν Aaν , (7.13)
4 2
which is known as the Yang-Mills action fixed in the Rξ gauge [1]. The case ξ = 1
is the Feynman gauge. When ξ → ∞, the π̄a = Ba dependence in Ψ of Eq.(7.9)
disappears and the Landau gauge ∂ µ Aaµ = 0 is imposed as a delta-function condition.
Because the quadratic forms in Eq.(7.13) are non-degenerate, Ψ is an admissible
gauge-fixing fermion. Propagators exist and Eq.(7.13) can be used as an action for
the Yang-Mills perturbation series [1, 165].
The BRST symmetry of Eq.(7.13) is determined by the procedure described at
the end of Sect. 7.1. One performs canonical transformations to the gauge-fixed basis
and then to the classical solution for B a . The latter is determined by varying the
action with respect to B a itself, and leads to the shift B a → B a + ξ∂ µ Aaµ . One finds
that the BRST symmetry of Eq.(7.13) is given by
1
δB̃Ψ Aaµ = D µa b C b , δB̃Ψ C a = fbc a C c C b , δB̃Ψ C¯a = ξ∂ µ Aaµ . (7.14)
2
The gauge-fixed BRST generator δB̃Ψ is only nilpotent on-shell.
The integration over C¯aµ and η̄a leads to δ (C aµ ) δ (η a ) which can be used to do the C aµ
and η a integrals. All the functional integrals over fields have been performed, leaving
no local degrees of freedom in the action. Not too surprisingly, the integration over
the topological action leads to a trivial lagrangian. The functional integral produces
a finite number. It is important that all the terms in Ψ in Eq.(7.16) be present. If the
last term η 1a C¯a were dropped from Ψ, then the integrations over Λ1a and C¯a would
produce singular contributions.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 93
to Ψ in Eq.(7.21). Here, ξ, ξ¯ and ξ 1 are constants. One could also replace ordinary
derivatives ∂µ by covariant ones Dµ in Eq.(7.21). Then Dµ replaces ∂µ in Eqs.(7.22)
and (7.23). In this case, A∗aµ 6= 0, but since A∗aµ does not enter the proper solution in
Eq.(5.14), the gauge-fixed action is as in Eq.(7.23) with ∂µ → Dµ .
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 94
..
.
Diagram 9. The Triangular String Field Tableau
In addition, there are the trivial-pair partners πsk for k = 1, 3, 5, . . . and s ≥ k, and
π̄sk for k = 0, 2, 4, . . . and s ≥ k. Although there are no αi type indices on Csk , C¯sk ,
πsk and π̄sk , they are string fields and represent an infinite tower of ordinary particle
fields. The ghost numbers of the fields are
h i h i
gh Csk = gh C¯sk∗ = s − k , g Csk = g C¯sk∗ = 1 + k − s ,
h i h i
gh C¯sk = gh Csk∗ = k − s − 1 , g C¯sk = g Csk∗ = 1 + k − s ,
h i
gh π̄sk = k − s , g π̄sk = 1 + k − s ,
h i
gh πsk = s − k + 1 , g πsk = 1 + k − s , (7.24)
where gh[ ] is the field-antifield ghost number and g( ) is the string ghost number.
In Eq.(7.24), we use the abbreviations in Eq.(6.23) for the cases when k = −1 and
k = 0. All Csk , C¯sk∗ and π̄sk have odd total statistics, while C¯sk , Csk∗ and πsk have even
total statistics.
The auxiliary action Saux in Eq.(6.28) is
∞
X ∞ Z
X ∞
X ∞ Z
X
Saux = ∗ k
π̄s ∗ C¯sk∗ + ∗ k∗
Cs ∗ πsk , (7.25)
k=0 s=k k=1 s=k
k even k odd
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 95
where ∗ before a field is the string Hodge star operation. For an arbitrary field or
antifield ϕ, gh [ ∗ ϕ] = gh [ϕ] but g ( ∗ ϕ) = 3 − g (ϕ). The action in Eq.(7.25) is to be
added to the minimal-field proper solution S in Eq.(5.24).
The most convenient gauge for the open bosonic string field theory is the Siegel-
Feynman gauge [229] which imposes the condition c̄0 A = 0 on the string field A. When
R
implemented, the quadratic action of the massless vector Aµ becomes 12 d26 x Aµ ∂ν ∂ ν Aµ .
Above, c̄0 is the zero mode of the antighost of first-quantized open string theory in
Sect. 3.8 and Sect. 5.8. It is the conjugate momentum of the zero mode c0 of the
first-quantized ghost. These modes satisfy
and have string ghost numbers of g (c0 ) = 1 and g (c̄0 ) = −1. See Sect. 7.7 for more
discussion. The gauge fermion Ψsf , which implements the Siegel-Feynman gauge, has
an expansion that, for the first few levels, reads
Z
Ψsf = ∗
C¯0 ∗ c0 c̄0 A
Z Z
∗
C¯1 ∗ c0 c̄0 C0 + ∗ 1
C1 ∗ c0 c̄0 C¯0 +
Z Z Z
∗
C¯2 ∗ c0 c̄0 C1 + ∗ 1
C2 ∗ c0 c̄0 C¯1 + ∗ ¯2
C2 ∗ c0 c̄0 C11 + . . . .
For all levels,
∞
X ∞ Z
X ∞ Z
∞ X
X
∗ ¯k
Ψsf = Cs ∗ k−1
c0 c̄0 Cs−1 + ∗ k
Cs ∗ c0 c̄0 C¯s−1
k−1
. (7.27)
k=0 s=k k=1 s=k
k even k odd
Since Ψsf is not a functional of πsk or π̄sk , a delta-function type gauge-fixing pro-
cedure is implemented. Elimination of the antifields via Eq.(6.1) leads to
Csk∗ = 0 ⇒ c0 C¯s−1
k−1
= 0 and c̄0 C¯s+1
k+1
=0, k = 1, 3, 5, . . . , (7.29)
which follow from Eqs.(7.26) and (7.28). For k ≥ 1, Csk and C¯sk are set equal to zero
because any field Φ, for which c̄0 Φ = 0 and c0 Φ = 0, must be identically zero, as
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 96
Eq.(7.26) implies. On the other hand, for Cs−1 and C¯s , only half these conditions are
imposed:
c̄0 Cs−1 = 0 , s ≥ 0 ,
c0 C¯s = 0 , s≥0 . (7.30)
When these results are used in Cs∗ ≡ Cs−1∗ = c0 c̄0 C¯s+1
0
≡ c0 c̄0 C¯s+1 , for s ≥ −1, one
finds Cs−1∗ = C¯s+1 , so that the gauge-fixed string field ΨΣsf in Eq.(5.23) is
The condition c0 C¯s = 0 implies c̄0 ∗ C¯s = 0. Hence, the constraints in Eq.(7.30) can be
written succinctly as
c̄0 ΨΣsf = 0 . (7.32)
Eq.(7.32) implies that ΨΣsf has no term proportional to c0 . Since the string integral
is zero unless there is a c0 factor present, Q must supply it. Since Q = c0 L0 + . . .,
Q → c0 L0 , (7.33)
The fermion Z
Ψ= dτ C¯ (e − 1) + Γ̄χ (7.37)
imposes the delta-function conditions
G ≡ C + θΓ , Ḡ = Γ̄ + θC¯ ,
eτ τ = eσ σ = ρ−1/2 , eτ σ = eσ τ = 0 . (7.43)
∂τ ∂τ X µ − ∂σ ∂σ X µ = 0 ,
∂τ C τ − ∂σ C σ = 0 , ∂τ C σ − ∂σ C τ = 0 ,
∂τ C¯τ − ∂σ C¯σ = 0 , ∂τ C¯σ − ∂σ C¯τ = 0 , (7.47)
6
When D 6= 26, one cannot do this due to Weyl anomaly. See Sect. 9.3.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 99
and
∂σ X µ = C σ = C¯σ = 0 , at σ = 0, π , (7.48)
which follow from a careful treatment of boundary conditions. The solutions are
∞
X αnµ −inτ
X µ (τ, σ) = xµ + pµ τ + i e cos (nσ) ,
n=−∞ n
n6=0
∞
X
C τ (τ, σ) = cn e−inτ cos (nσ) ,
n=−∞
∞
X
C σ (τ, σ) = −i cn e−inτ sin (nσ) ,
n=−∞
∞
X
C¯σ (τ, σ) = −i c̄n e−inτ cos (nσ) ,
n=−∞
∞
X
C¯τ (τ, σ) = c̄n e−inτ sin (nσ) , (7.49)
n=−∞
theory of Sects. 3.6, 5.6 and 7.5 are such functionals. Alternatively, one may expand
the string fields as a linear combination of first-quantized states whose coefficients are
ordinary particle fields. For more details, see the reviews in refs.[215, 247].
The gauge-fixed BRST symmetry of the action in Eq.(7.46) is
δBΨ X µ = C n ∂n X µ ,
δBΨ C m = −C n ∂n C m ,
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 100
δBΨ C¯τ = −ρ1/2 ∂τ Xµ ∂σ X µ + ρ1/2 ∂n ρ−1/2 C¯τ C n + C¯σ (∂σ C τ + ∂τ C σ ) , (7.51)
1
δBΨ C¯σ = −ρ1/2 (∂τ Xµ ∂τ X µ + ∂σ X µ ∂σ Xµ ) + ρ1/2 ∂n ρ−1/2 C¯σ C n + C¯τ (∂σ C τ + ∂τ C σ ) ,
2
where m takes on the values τ and σ and the sum over n is over τ and σ. A
straightforward computation confirms that
∂r SΨ
δBΨ SΨ ≡ δB ΦA = 0 , (7.52)
∂ΦA Ψ
where SΨ is given in Eq.(7.46). A useful exercise is to perform this computation.
Because some fields have been eliminated using equations of motion, δBΨ δBΨ = 0 only
on-shell.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 101
this can be written as δB̂ X = 0. The operator δB̂ is the quantum BRST transforma-
tion. It is the quantum generalization of the classical BRST transformation δB : As h̄
goes to zero, δB̂ becomes δB in Eq.(4.32). A functional is said to be quantum-BRST
invariant if
δB̂ X = 0 . (8.2)
Although δB̂ is nilpotent as a consequence of the quantum master equation
1
δB̂ δB̂ X = −( (W, W ) − ih̄∆W, X) = 0 , (8.3)
2
it no longer is a graded derivation since
The first term produces zero since it is a surface term and the second term is zero
for symmetry reasons. Hence, an interesting functional O is one that is non-trivial in
the quantum-BRST cohomology, i.e., O is quantum-BRST invariant but cannot be
written as the quantum-BRST transform of some functional:
WΨ ≡ W |ΣΨ , (8.9)
and a similar identity holds for X → W . The fact that the functional in Eq.(8.7)
produces a zero expectation suggests that it is quantum-BRST trivial. Indeed, this
is the case since
!
∂l X|ΣΨ i ∂l (WΨ ) (−1)ǫ(X) ∂Ψ
+ X|ΣΨ = δ
B̂ Φ∗A − X , (8.10)
∂ΦA h̄ ∂Φ A ih̄ ∂ΦA
ΣΨ
as a short calculation reveals. One obtains the interesting result that the Schwinger-
Dyson equations are a consequence of the quantum-BRST symmetry of the field-
antifield formalism [152]. Reference [2] argued that antifields originate as the antighosts
of collective fields that ensure the Schwinger-Dyson equations. This differs from one
viewpoint that antifields are the sources of BRST transformations. Equations (8.6)
and (8.10) show that the Schwinger-Dyson equations are certain quantum-BRST
Ward identities.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 104
1
(S, S) = 0 , (8.12)
2
which yields the structure equations. The nilpotency of the quantum BRST operator
depends on Eq.(8.11) being satisfied, as can be seen from Eq.(8.3). The quantum
cohomology requires the existence of a nilpotent quantum BRST operator. It is used
to define the quantum observables of a gauge theory and to determine when two
functionals are considered equivalent. In Sect. 8.5, it is argued that the existence of
gauge anomalies is related to the violation of the quantum master equation. In short,
the field-antifield formalism at the quantum level depends crucially on Eq.(8.11) being
satisfied.
The potential difficulty is due to the operator ∆ in Eq.(6.7) which is singular
when acting on S or W because usually they are local functionals. Often terms
proportional to delta functions and derivatives of delta functions are produced. One
therefore needs to regularize ∆. If one can find a regularization such that (∆S)reg = 0,
while maintaining the classical master equation, then one can simply let W = S.7
Then the quantum master equation is satisfied. An example of this situation is a
theory without gauge invariances: Antifields are absent in S since the proper solution
is given by S = S0 . Consequently, ∆S = 0.
The quantum master equation has a symmetry given by quantum BRST trans-
formations, in which W → W − εδB̂ F , where ǫ(F ) = 1 and gh[F ] = −1 [27, 103, 152,
75, 253]. If Eq.(8.11) is true, and
′
′
W = W + ε W , F + ih̄ε∆F , (8.13)
′
then W also satisfies
Eq.(8.11) to order ε2 . Eq.(8.13) has a simple interpretation. The
′
term ε W , F is the change in W due to a canonical transformation (see Eq.(4.12)).
The extra term ih̄ε∆F is due to the non-invariance of the functional integral measure
7
Strictly speaking, this statement holds for a finite or a regularized theory. Renormalization
may require one to add additional terms to S.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 105
under a canonical transformation. The measure effect can be exponentiated and leads
to the extra term.
A solution to the quantum master equation ensures that the gauge symmetries
survive the process of quantization. Although finding a solution is an important
objective, a solution does not necessarily guarantee good behavior of the theory. The
theory may still have difficulties in regard to other issues such as locality, unitarity and
the presence of Feynman-diagram infinities. These difficulties may be insurmountable
or it may be necessary to add new counterterms to achieve the desired properties.
There are systems for which the quantum master equation is satisfied but the theory
is non-renormalizable (see the next subsection). In short, the quantum theory is not
determined solely by the quantum master equation.
The question of whether the quantum master equation is satisfied and the ques-
tion of renormalizability are separate issues (although somewhat related). As argued
above, a renormalizable theory without gauge invariances is guaranteed to satisfy the
quantum master equation. Regularization and renormalization must be performed,
even if the quantum master equation is satisfied at all stages of the renormalization
process. When gauge invariances are present, the interesting issue is whether the
quantum master equation can be maintained after renormalization. A theory can be
non-renormalizable but satisfy the quantum master equation: A non-renormalizable
theory without gauge invariances is an example. An infinite number of countert-
erms must be added to the action, but the quantum master equation is satisfied
trivially because no antifields enter. In the other extreme, the two-dimensional chiral
Schwinger model is anomalous, yet the theory still makes sense and is renormaliz-
able. It is an example of a renormalizable system that does not satisfy the quantum
master equation. For more discussion on regularization and renormalization in the
field-antifield formalism, which are the natural generalizations of ideas contained in
earlier approaches [277, 187, 167], see refs.[262, 264, 249, 4, 75, 156, 191].
The Zinn-Justin equation [277] has played an important role in analyzing the
renormalizability of Yang-Mills theories. The generalization of this equation within
the antibracket formalism is the subject of the next section.
original system.
Consider the functional integral in the presence of sources JA , i.e.,
Z
i
Z [J, Φ∗c ] ≡ [dΦ] exp W [Φ, Φ∗c ] + JA ΦA , (8.14)
h̄
∗
where JA are independent
variables that do not depend on Φ and Φc , and have
statistics ǫ (JA ) = ǫ ΦA = ǫA . For reasons that become clear below, we have
replaced the Φ∗ by classical antifields Φ∗c . Actually, to compute perturbatively, one
∂Ψ
needs to shift the antifields Φ∗cA by ∂Φ A , where Ψ is an appropriate gauge-fixing
∗ h̄ ∂l ln Z [J, Φ∗c ] D A EJ
ΦA
c [J, Φc ] ≡ = Φ , (8.15)
i ∂JA
∂r Γ ǫA ǫB ∂r JcB B ∂r ln Z [J, Φ∗c ] ∂r JcB
= −J cA − (−1) Φc − ih̄ .
∂ΦA ∂ΦA ∂JB ∂ΦA
c c c J=Jc
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 108
Using Eq.(8.15), one finds that the last two terms cancel so that
∂r Γ
= −JcA . (8.19)
∂ΦAc
At this stage, we have made a transition from fields to classical fields. Essentially
∗
all quantities are now functionals of ΦAc and ΦcA . Given two functionals X and Y of
∗
ΦAc and ΦcA , define the classical antibracket ( , )c by
∂r X ∂l Y ∂r X ∂l Y
(X, Y )c ≡ − . (8.20)
∂Φc ∂ΦcA ∂Φ∗cA ∂ΦA
A ∗
c
Since the classical antibracket is defined in the same manner as the antibracket, it
satisfies the same identities (4.5) – (4.7). Using Eq.(8.19), one obtains
1 ∂l Γ 1
(Γ, Γ)c = −JcA ∗ = −ih̄∆W + (W, W ) , (8.21)
2 ∂ΦcA 2 c
where the final equality is obtained after some algebra, which makes use of integration
by parts. In Eq.(8.21), h ic denotes the expectation value in the presence of J but
expressed in terms of Φc and Φ∗c . More precisely, if X is a functional of Φ and Φ∗
then
hXic ≡ hXiJ , (8.22)
J=Jc
J
where h i is defined in Eq.(8.16). Because of the quantum master equation (8.11),
Eq.(8.21) becomes
(Γ, Γ)c = 0 , (8.23)
a result known as the Zinn-Justin equation [277].
Equation (8.22) allows one to pass from a functional X of the original fields Φ and
Φ to a classical functional hXic of classical fields Φc and Φ∗c by taking the “classical
∗
The process X → hXic conforms to the idea that a classical variable is the expectation
value of the corresponding quantum functional.
Since Γ satisfies the Zinn-Justin equation and plays the role of S in the classical
antibracket formalism, one can construct by analogy a “classical-quantum” BRST
transformation δBcq having the same properties as δB . Define
where X is any functional of Φc and Φ∗c . The nilpotency δB2 cq = 0 of δBcq follows from
the Zinn-Justin equation. Define X to be cq-BRST invariant if δBcq X = 0. According
to Eq.(8.23), the effective action Γ is cq-BRST invariant since
1 ∂r (Γ, Γ)c
δBcq JcA ≡ (JcA , Γ)c = (−1)ǫA +1 =0 . (8.27)
2 ∂ΦAc
Consider
∂l Γ
−ih̄δBcq ln (Zc ) = Γ + JcA ΦA
c ,Γ = JcA ΦA
c ,Γ = JcA =0 ,
c c ∂Φ∗cA
where the first equality holds because of Eqs.(8.26) and (8.27), and the last equality
follows from Eqs.(8.11) and (8.21). Hence,
The cq-BRST operator δBcq is the effective classical version of the quantum-BRST
operator δB̂ X. To understand this statement, consider
The last step follows because h ic has dependence on Φc and Φ∗c through Jc . After
some algebra which makes use of integration by parts, one finds that
i
δBcq (Zc hXic ) = Zc δB̂ X + X ∆W + (W, W ) + (−1)ǫB Zc hXΦB
c i (JcB , Γ)c .
2h̄ c
(8.29)
If δBcq (hXic ) is computed instead, one obtains the connected part of Eq.(8.29) without
a Zc factor. Since Eq.(8.27) and the quantum master equation (8.11) are satisfied,
In other words, the cq-BRST variation of the classical version of X is the classical
version of the quantum-BRST variation of X [4].
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 110
The effective action Γ in the classical antibracket formalism plays a role analogous
to the proper solution S in the ordinary antibracket formalism. Properties obeyed by
Γ are the same as those obeyed by S. Therefore, one can define a BRST structure
associated with Γ [133]. The BRST structure tensors are encoded in Γ and the
relations among them are given by (Γ, Γ) = 0. Expanding in a Taylor series in Φ∗c ,
one has
1
Γ [Φc , Φ∗c ] = Γ0 (Φc ) + Φ∗cA ΓA (Φc ) + Φ∗cA Φ∗cB ΓBA (Φc ) + . . . . (8.31)
2
Recalling that there is an undisplayed dependence on the gauge-fixing fermion Ψ, the
above terms have the following interpretation: Γ0 (Φc ) is the one-particle-irreducible
generating functional for the basic fields including all loop corrections for the action
gauge-fixed using Ψ, i.e., SΨ in Eq. (6.88), and ΓA (Φc ) is the generator of gauge-fixed
cq-BRST transformations. The gauge-fixed cq-BRST operator δBcqΨ , the analogy of
δBΨ , is defined by
δBcqΨ X ≡ (X, Γ)|ΣΨ , (8.32)
where X is a functional of the ΦA c only. In the gauge-fixed basis, ΣΨ in Eq.(8.32)
means that classical antifields are set to zero. Thus, one has δBcqΨ ΦA A
c = Γ (Φc ).
The tensor ΓBA (Φc ) is related to the on-shell nilpotency of δBcqΨ . In summary, the
quantum aspects of the classical theory described by S are reproduced by an effective
classical theory governed by Γ.
where Ω is a local functional of the fields and antifields. The last term may seem
surprising but it is necessary if A is to satisfy Eq.(8.34): When the quantum master
equation is violated as in Eq.(8.33), the nilpotency of the quantum BRST operator
no longer holds, as can be seen from Eq.(8.3). The last term in Eq.(8.35) is required
to compensate for this effect and ensures Eq.(8.34). Let
′
W = W + h̄Ω . (8.36)
′
Then, using Eqs.(4.5) and (6.8), one finds that W satisfies the quantum master
equation:
1 ′ ′
′
W , W − ih̄∆W = 0 .
2
Since W has the same classical limit as W , namely, its order h̄0 term is S, one can use
′
′
W in lieu of W for the quantum action. Then, since the quantum master equation
is satisfied, a quantum gauge theory can be defined.
When A cannot be expressed as in Eq.(8.35) for a local functional Ω, there is an
anomaly in the quantum master equation and an obstruction to maintaining gauge
symmetries at the quantum level. Since anomalies involve subtleties and singular
expressions, they are usually not too easy to compute. The usual approach to this
subject uses a loop expansion:
∞
X
A= Al h̄l−1 = A1 + h̄A2 + . . . . (8.37)
l=1
Using Eqs.(6.14) and (8.37), one finds the following expression for the anomaly at the
one-loop level
A1 ≡ ∆S + i (M1 , S) . (8.38)
To this order, the condition for the absence of an anomaly in Eq.(8.35) is that A1 is
a classical BRST variation, i. e.,
the quantum master equation in Eq.(8.11) is satisfied to order h̄. In words, Eq.(8.39)
says that if A1 is expressible as a local BRST variation then effectively there is
no one-loop anomaly. Since the second term in Eq.(8.38) is already of this form,
the requirement becomes that ∆S should be a classical-BRST variation of a local
functional.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 112
δB A1 = (A1, S) = 0 . (8.41)
iA = aα (φ) C α + . . . , (8.42)
where the omitted terms involve antifields. Sometimes these terms are absent so that
Eq.(8.42) gives the structure of the quantum-master-equation anomaly. Although not
obvious, the coefficients aα (φ) are the usual gauge anomalies [251]. In other words,
a quantum-master-equation anomaly and a gauge anomaly are equivalent.
where the jacobian factor J is the berezinian governing the change from {Φ, Φ∗ }
variables to tilde variables. Then, Wf satisfies the quantum master equation exactly
in the tilde variables if W satisfies it in the {Φ, Φ∗ } variables. A detailed proof of
this result and a formula for J can be found in ref.[251]. Here, we provide a few key
steps. Define ∆e as the analog of ∆ in the transformed variables, i.e.,
e ≡ (−1)ǫA +1 ∂r ∂r
∆ .
∂ Φ̃A ∂ Φ̃∗A
If G is an arbitrary functions of Φ̃ and Φ̃∗ , then
h i h 1 i
e
∆G Φ̃, Φ̃∗ = ∆G Φ̃ [Φ, Φ∗ ] , Φ̃∗ [Φ, Φ∗ ] −
(G, ln J) . (8.45)
2
When ∆ acts on G on the right-hand side of Eq.(8.45), the tilde fields should be
regarded as functionals of Φ and Φ∗ , as indicated. The chain rule for derivatives is
then used. This produces ∆Ge plus an extra term, which is equal to 21 (G, ln J) and
needs to be subtracted to obtain the identity in Eq.(8.45). Using Eqs.(8.44) and
(8.45), one finds,
eWf− 1 f f
f −1 W f , ln J
1 f f
ih̄∆ W , W = ih̄ ∆W − W, W
2 2 2
( !) !
ih̄ 1 ih̄ 1 ih̄ ih̄
= ih̄ ∆W − ∆ ln J − W − ln J, ln J − W − ln J, W − ln J
2 2 2 2 2 2
1 h̄2 1
= ih̄∆W − (W, W ) + ∆ ln J − (ln J, ln J) .
2 2 4
It can be shown [27, 251] that ∆ ln J = 14 (ln J, ln J), so that the last term is zero.
Hence, if W satisfies the quantum master equation, then W f satisfies the tilde version
of the quantum master equation. For infinitesimal transformations,
Φ̃A = ΦA − ε ΦA , F + O ε2 ,
Φ̃∗A = Φ∗A − ε (Φ∗A , F ) + O ε2 ,
J = 1 − 2ε∆F + O ε2 . (8.46)
Then,
f [Φ, Φ∗ ] = W [Φ, Φ∗ ] + ε (W, F ) + ih̄∆F + O ε2
W ,
so that
f [Φ, Φ∗ ] = W − εδ F + O ε2
W [Φ, Φ∗ ] → W ,
B̂
and one recovers Eq.(8.43). The identity ∆ ln J − 14 (ln J, ln J) ∼ O (ε2 ) follows from
Eq.(8.46) and ∆2 = 0.
One can take advantage of canonical transformations in analyzing potential anoma-
lies by going to a basis for which the computation is simpler.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 114
assume that an admissible Ψ has been selected and that the shift to the gauge-fixed
basis has been performed. According to the result in Sect. 8.6, if the quantum master
equation is satisfied and a canonical transformation is performed to a new basis then,
by appropriately adjusting the action, the quantum master equation is satisfied in
the new basis. Hence, the existence or non-existence of an anomaly is independent of
the choice of basis, although the form of the anomaly may depend on this choice.
There are different ways of obtaining the anomaly. We mostly follow the approach
of ref.[251] and briefly mention other methods at the end of this subsection. Reference
[251] obtained general formulas for the antifield-independent part of the one-loop
anomaly using a Pauli-Villars regularization scheme. Since the derivation is somewhat
technical, we present only the final results. For more details, see refs.[87, 251, 259,
78, 75, 252, 257, 253]. In particular, refs.[75, 253] have an extensive discussion of
Pauli-Villars regularization in the antibracket formalism to which we refer the reader.
The goal of the next few paragraphs is to obtain a regularized expression for ∆S,
denoted by (∆S)reg . The anomaly A1 is essentially (∆S)reg since the M1 term in
Eq.(8.38) is eliminated as a possible violation of the quantum master equation via
Eq.(8.40) with Ω1 = −M1 , i.e., the counterterm −h̄M1 is added to the action. Define
∂l ∂r
K AB ≡ ∗ B
S [Φ, Φ∗ ] , (8.47)
∂ΦA ∂Φ
∂l ∂r
QAB ≡ S [Φ, Φ∗ ] . (8.48)
∂ΦA ∂ΦB
Note that Q involves derivatives with respect to fields and not antifields. If expanded
about a stationary point, Q becomes the quadratic form for the fields. In such an
expansion, the inverse of Q is the propagator. The properness condition in the gauge-
fixed basis guarantees that propagators exist. An operator O used to regulate ∆S is
related to Q by
AC
OA B ≡ T −1 QCB , (8.49)
where TAC is an arbitrary invertible matrix satisfying TBA = (−1)ǫA +ǫB +ǫA ǫB TAB .
The inverse of T obeys (T −1)BA = (−1)ǫA ǫB (T −1 )AB . The Grassmann statistics of
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 115
TAB and (T −1 )AB are ǫA + ǫB (mod 2). Eq.(8.49) implies QAB = TAC OC B . In the
regularization scheme of ref.[251], the matrix T appears in the mass term for the
regulating Pauli-Villars fields. In that approach, the violation of the quantum master
equation is shifted from the ∆ term to the (S, S) term.
The regulated expression for (∆S)reg is [251]
!B
1
(∆S)reg = F A B O
, (8.50)
1− M A 0
where
1
F A C ≡ K A C + (T −1 )AD (δB T )DC (−1)ǫC , (8.51)
2
and where δB T denotes the classical BRST transform of T : δB T = (T, S). In
Eq.(8.50), the sum over A and B, leading to the trace, involves the quadratic form
of fields only and not that of antifields. The subscript 0 on the square brackets in
Eq.(8.50) indicates that the term independent of M is to be extracted. Here, M de-
notes a regulator mass. The one-loop nature of Eq.(8.50) is evident by the presence
of the propagator factor −i/ (M − O) → (iM)−1 / (1 − O/M ) and the sum over the
index A indicating a trace.
When O is quadratic in space-time derivatives, one lets
R=O , M2 = M , (8.52)
where R denotes the quadratic regulator operator and M denotes the regulating
mass or cutoff. When O is linear in space-time derivatives, it is convenient to mul-
tiply on the right in the trace in Eq.(8.50) by 1/ (1 + O/M) (1 + O/M) and carry
out the multiplication of 1/ (1 − O/M) with 1/ (1 + O/M). Eq.(8.50) can then be
manipulated into the form [251]
!B
A 1
(∆S)reg = F ′ B R
, (8.53)
1− M 2 A 0
where
R = O2 , M=M , (8.54)
and
1
F ′A C ≡ F A C − (δB O)A C (−1)ǫC (8.55)
2M
with δB O = (O, S).
Summarizing, in the quadratic momenta case, (∆S)reg is given by Eqs.(8.50) and
(8.51). This is the same as using Eq.(8.53) with R and M given in Eq.(8.52) and
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 116
Z ∞
" !n #
∞ X λ
dλ exp (−λ) fn = f0 =
0 n=−p M2 0
∞ n " B #
X 1 R
A
fn = F B exp ,
n=−p M2 0 M2 A 0
1
where a Laurent expansion in M2
has been performed. The resulting expression,
" B #
A R
(∆S)reg = F B exp , (8.56)
M2 A 0
corresponds to the Fujikawa form [120] of the regularization [251] . In the limit
M → ∞, terms of order 1/Mn for n > 0 vanish, whereas terms with n < 0 blow up.
The regularization scheme consists of dropping the terms that blow up. In the Pauli-
Villars regularization, this is achieved by adding fields with appropriate statistics,
couplings and masses to cancel all n < 0 terms. As a consequence, only the n = 0
term remains.
In the quadratic momentum case, Eq.(8.50) can be manipulated into the following
supertrace form [133]
!B
1 1
(∆S)reg = − (R−1 δR)A B R (−1)ǫA , (8.57)
2 1− M 2 A 0
which only involves the quadratic regulator R. Eq.(8.57) shows that if δR = [R, G]
for some G then the anomaly vanishes, as a consequence of the cyclicity of the trace
[133]. Formally, Eq.(8.57) is
" A #
1 −R ǫA
(∆S)reg = δB − ln (−1) , (8.58)
2 M2 − R A 0
regulator R that is negative definite after a Wick rotation to Euclidean space. Mod-
ifying T changes the form of the anomaly. In particular, when more than one gauge
symmetry is present, varying T changes the coefficients aα in Eq.(8.42). If some non-
zero aα are made zero and vice-versa, the anomaly is shifted from being associated
with one type of gauge symmetry to another. This is analogous to the well known
situation for anomalous chiral gauge theories in four dimensions: The anomaly can be
moved from the axial vector sector to the vector sector, if so desired. See, for example,
Sect. 4.1 of ref.[251]. Although we do not present any examples of this effect, it is
well illustrated in ref.[251]. In performing anomaly calculations, it is useful to choose
T to render a computation as simple as possible. For a similar reason, it is also useful
to perform certain canonical transformations before commencing a calculation.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 119
1 ẋ2 1 2
π̄ − − m =0 . (9.3)
2 e2 2
The relevant canonical transformation is
1 ẋ2 1 2
π̄ → π̄ + + m ,
2 e2 2
∗ ∗ ∗ ẋ
2
d π̄ ∗ ẋµ
e →e + ,π̄ 3 → + x∗µ x∗µ
, (9.4)
e dτ e2
with the other fields and antifields left unchanged. The action becomes
Z
1 ρ 1
S→ dτ − 2 ẋ2 − ρm2 + C¯C˙ + π̄ (e − ρ) +
e 2e 2
! )
C ¯∗ 1 ẋ2 π̄ ∗ C ẋµ d ẋµ
x∗µ ẋµ + e C˙ + C
∗
+ m2
− . (9.5)
e 2 e2 e2 dτ e
Let us first determine the overall structure of the computation. From Eq.(9.5),
one finds that the non-zero entries of the matrix K A B are
xµ e π̄ C¯ C
x∗µ ∗ ∗ 0 0 ∗
e∗ 0 0 0 0 ∗
∗
K = π̄
∗ ∗ 0 0 ∗
, (9.6)
C¯ ∗
∗ ∗ 0 0 0
C∗ 0 0 0 0 0
where the columns and rows are labelled by the corresponding fields and antifields.
We select TAB to be proportional to the identity matrix, except in the ghost sector
for which
C¯ C
!
C¯ 0 −1
¯ =
TCC , (9.7)
C 1 0
and in the x sector for which
(Tx )µν = ηµν . (9.8)
In perturbation theory, the regulator R is evaluated at the stationary point of the
gauge-fixed action. For e, this corresponds to
e|Σ = ρ , (9.9)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 121
xµ e π̄ C¯ C
xµ ∗ 0 0 0 0
e 0 ∗ ∗ 0 0
R|Σ = π̄ 0 ∗ 0 0 0 . (9.10)
C¯
0 0 0 ∗ 0
C 0 0 0 0 ∗
Eq.(9.10) shows that propagation is diagonal within three sectors: the xµ sector, the
e-π̄ sector and the ghost sector. As expected, the canonical transformation in Eq.(9.4)
decouples e from xµ : For the shifted action in Eq.(9.5), one has
!
∂l ∂r S d 1 µ d ρ µ
= −2 ẋ + 2 ẋ =0 . (9.11)
∂e∂xµ Σ dτ e2 dτ e3
Σ
Because a constant TAB matrix has been selected, the non-zero entries of F A B in
Eq.(8.51) are the same as in Eq.(9.6). From the structure of Eqs.(9.6) and (9.10), one
sees that the anomaly computation separates into contributions from the xµ sector,
the e-π̄ sector and the ghost sector. The propagating fields are xµ , C¯ and C. The
field π̄ serves as a Lagrange multiplier for setting e equal to ρ. Hence, e and π̄ are
non-propagating and should not contribute to the anomaly according to the analysis
in Sect. 8.7 [253]. For this particular system, the contribution is zero because F A B
in the e-π̄ sector is off-diagonal. It is also clear that C and C¯ do not contribute to the
anomaly since F A B is zero for all ghost entries. One only needs to consider the xµ
sector.
In what follows we use a subscript x for quantities associated with xµ . Applying
Eqs.(8.47) and (8.48) to Eq.(9.5), one arrives at
d
(Kx )µ ν = δ µ ν e−1 C , (9.12)
dτ
and
(Qx )µν = ηµν Qx , (9.13)
where Qx without µν subscripts is defined by
d −1 d d d
Qx ≡ −2 e + ρe−2 . (9.14)
dτ dτ dτ dτ
Here and below, the derivative dτd acts on everything to the right including e, ρ and
the function to which Qx is applied. Some contributions to Eq.(9.14) come from the
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 122
shifts in Eq.(9.4). We have also dropped terms proportional to π̄ ∗ and C¯∗ because
they will not contribute, when expanding about the stationary point.
Since the Tx matrix is (Tx )µν = ηµν , the regulator matrix in Eq.(8.52) is
(Rx )µ ν = δ µ ν Qx . (9.15)
Because δB Tx = 0,
(Fx )µ ν = (Kx )µ ν , (9.16)
where (Kx )µ ν is given in Eq.(9.12).
All the relevant matrices of Sect. 8.7 for the computation of the anomaly have
been obtained. At this stage, one expands about the stationary point of the gauge-
fixed action. The field e is set equal to the function ρ according to Eq.(9.9). The
regulator matrix becomes
d −1 d
(Rx )µ ν |Σ = −δ µ ν ρ , (9.17)
dτ dτ
so that on-shell
(∆S)reg =
!
Z Z∞
dk d − dτd ρ−1 dτd
D dτ exp (−ikτ ) ρ−1 C exp exp (ikτ ) , (9.18)
2π dτ M2
−∞ 0
where we have used the form of (∆S)reg in Eq.(8.56). The trace over field indices
R
A produces a factor of dτ and a factor of D (because the number of xµ fields is
D and each contributes equally). The operator trace is evaluated using a complete
set of momentum-space functions, thereby generating the factors exp (±ikτ ). The
calculation in Eq.(9.18) is performed in Appendix C, where it is shown that the
integrand is an odd function of k. Consequently,
(∆S)reg = 0 . (9.19)
so that
(δB Rx )µ ν |Σ = 0 . (9.22)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 123
(∆S)reg = 0 ,
Although this model has not been discussed in previous sections, it is straightfor-
ward to apply the field-antifield formalism [61, 63]. The proper solution for the gauge
sector corresponds to the solution for the Yang-Mills example given in Sect. 5.2 for
d = 2 and for a U(1) group. In addition to the antifield A∗µ of the photon, one has
commuting antifields ψ ∗ and ψ̄ ∗ for the fermions. The proper solution of the master
equation is
Z h i
S = S0 + d2 x A∗µ ∂ µ C + i(ψ ∗ )t P− ψ C − iψ̄ ∗ P+t ψ̄ t C , (9.27)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 124
where C is the ghost field associated! with the gauge parameter ε. The superscript !
ψ1 ψ̄ 1
1 2 t
t stands for transpose: ψ = and ψ̄ = ψ̄ , ψ̄ , so that ψ̄ = and
ψ2 ψ̄ 2
(ψ ∗ )t = ((ψ 1 )∗ , (ψ 2 )∗ ).
A formal computation using the expression for ∆ in Eq.(6.7) reveals that only the
fermion sector contributes to ∆S. A more detailed analysis using a regularization
procedure confirms this.12 Therefore, we focus on the contribution to ∆S from ψ and
ψ̄. Gauge-fixing is not needed because propagators for the fermions already exist.
For these reasons, it is not necessary to consider gauge-fixing auxiliary fields, nor a
non-minimal proper solution.
Using Eqs.(8.47) and (9.27), one finds that the K matrix is given by
!
P− 02
K = −iC , (9.28)
02 −P+t
!
0 0
where all entries are 2×2 matrices, e. g., 02 = , so that K is a 4×4 matrix. In
0 0
Eq.(9.28) and throughout this subsection, we label the rows and columns of matrices
in the order ψ 1 , ψ 2 , ψ̄ 1 , ψ̄ 2 . From Eqs.(8.48) and (9.23), the matrix Q for the fermion
sector is !
02 −iD /̃
Q= , (9.29)
i/
D 02
R R
where D
/̃ is defined by d2 xψ̄i/
Dψ = − d2 xψ t iD
/̃ ψ̄ t . More precisely,
iD ∂t −A
/̃ = i/ / t P+t , (9.30)
where, here, the superscript t indicates the transpose of a matrix in Dirac-index space.
For the matrix T , we choose
!
02 12
T = , (9.31)
−12 02
!
1 0
where 12 = . Because δB T = 0, the matrix F in Eq.(8.51) is equal to the
0 1
matrix K in Eq.(9.28). Using Eq.(8.49), one finds
!
i/
D 02
O= . (9.32)
02 iD
/̃
12
The computation made in ref.[78] for a pure Yang-Mills theories in four dimensions supports
the idea that gauge fields and ghosts produce a BRST trivial contribution to (∆S)reg .
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 125
Additional terms are produced in the total action Stotal = S + Saux of Eqs.(5.30) and
(7.41) given by
Z Zπ n
S→S+ dτ dσ C¯ + C¯σ (C n ∂n eτ τ − eτ n ∂n C τ ) + C¯ − C¯σ (C n ∂n eσ σ − eσ n ∂n C σ )
0
+ C¯τ σ + C¯τ (C n ∂n eτ σ − eτ n ∂n C σ ) + C¯τ σ − C¯τ (C n ∂n eσ τ − eσ n ∂n C τ )
+ C¯ + C¯σ eτ τ + C¯ − C¯σ eσ σ + C¯τ σ + C¯τ eτ σ + C¯τ σ − C¯τ eσ τ C
o
− C¯ + C¯σ eσ τ + C¯ − C¯σ eτ σ + C¯τ σ + C¯τ eσ σ + C¯τ σ − C¯τ eτ τ C τ σ , (9.46)
and
Z Zπ n
Saux → dτ dσ π̄ C¯∗ + eτ τ + eσ σ − 2ρ−1/2 +
0
o
π̄τ C¯∗τ + eτ σ − eσ τ + π̄σ C¯∗σ + eτ τ − eσ σ + π̄τ σ C¯∗τ σ + eτ σ + eσ τ . (9.47)
At this stage, it is desirable to shift fields by the solutions to the equations of
motion of the ea m , by using a canonical transformation. Such a shift guarantees that
the quadratic form QAB is on-shell diagonal in the ea m sector. This avoids mixing of
the X µ and ghost sectors with the π̄-ea m sector. Variations of Stotal in the gauge-fixed
basis with respect to the ea m produce the equations for the four π̄. Using a subscript
0 to denote the solutions, and ignoring terms proportional to antifields, one finds that
!
∂S ∂S
2 (π̄)0 ≡ − τ
+ − 2π̄ =
∂eτ ∂eσ σ
2∂n C¯τ C n − C¯ (∂τ C σ − ∂σ C τ ) − C¯τ σ (∂τ C τ − ∂σ C σ )
+C¯τ (∂n C n − 2C) + C¯σ (∂τ C σ + ∂σ C τ − 2C τ σ ) ,
!
∂S ∂S
2 (π̄σ )0 ≡ − τ
− − 2π̄σ =
∂eτ ∂eσ σ
(eσ σ − eτ τ )eLX − e (∂τ X µ Dτ Xµ + ∂σ X µ Dσ Xµ ) +
2∂n C¯σ C n + C¯ (∂τ C τ − ∂σ C σ ) + C¯τ σ (∂τ C σ − ∂σ C τ )
+ C¯τ (∂τ C σ + ∂σ C τ − 2C τ σ ) + C¯σ (∂n C n − 2C) , (9.48)
where the X µ -lagrangian density LX is defined to be
e
LX ≡ (Dτ X µ Dτ Xµ − Dσ X µ Dσ Xµ ) . (9.49)
2
The canonical transformation of interest is given by
3
(KC¯)τ τ = (KC¯)σ σ = −C n ∂n − (∂n C n ) + C ,
2
1 1
(KC )τ τ = −C n ∂n + (∂n C n ) + (∂τ C τ − ∂σ C σ ) ,
2 2
1 1
(KC )σ σ = −C n ∂n + (∂n C n ) − (∂τ C τ − ∂σ C σ ) , (9.51)
2 2
where the presence of a parenthesis around a derivative indicates that it acts only on
fields within the parenthesis. Throughout this subsection, the absence of a parenthesis
means that the derivative acts on everything to the right. The terms for KX and KC
arise from differentiating Eq.(5.30), whereas those for KC¯ come from Saux in Eq.(9.47)
after the π̄n shifts in Eq.(9.50) have been performed.
The quadratic forms QAB are
(QX )µν = ηµν QX , (9.52)
where
QX = Dτ† e 1 − eρ−1/2 (eτ τ + eσ σ ) Dτ − Dσ† e 1 − eρ−1/2 (eτ τ + eσ σ ) Dσ
on the vielbeins in Dτ† and Dτ , and on any function to which the operator QX is
applied. Likewise, for the other derivatives. The terms in Eq.(9.53) come from the
original action S0 in Eq.(3.79), as well as from the π̄ shifts of Eq.(9.50) in Saux of
Eq.(9.47). For the ghost sector, let
C¯τ
¯
Cσ
V =
Cτ . (9.54)
Cσ
Then the ghost part of the action is
Z Zπ
1
Sghost = dτ dσ V t Qghost V , (9.55)
2
0
where the superscript t on V t stands for transpose. The ghost quadratic form is
0 0 ρ−1/2 ∂σ −ρ−1/2 ∂τ
0 0 −ρ−1/2 ∂τ ρ−1/2 ∂σ
Qghost =
∂ ρ−1/2
. (9.56)
σ −∂τ ρ−1/2 0 0
−∂τ ρ−1/2 ∂σ ρ−1/2 0 0
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 130
The terms in Qghost originate from Eq.(9.46) and from the π̄ shifts of Eq.(9.50) in Saux
of Eq.(9.47). The dependence on ea m cancels between the two contributions leaving
only a dependence on ρ.
For the matrix TAB of Sect. 8.7, we choose
and
0 e−2 0 0
−e−2 0 0 0
Tghost = −i
0
, (9.58)
0 0 e2
0 0 −e2 0
where Tghost acts in the space of antighosts and ghosts given in Eq.(9.54). In the
X µ sector the operator O, defined in Eq.(8.49), is e−1 δ µ ν QX and is equal to (RX )µ ν ,
where QX is given in Eq.(9.53). In the ghost sector, O is
!
02 e2 ρ−1/2∂/
Oghost = i , (9.59)
e−2∂/ρ−1/2 02
where each entry is a two by two matrix. It is somewhat accidental that the Dirac
operator in two-dimensional Minkowski space
∂/ ≡ γ n ∂n = γ τ ∂τ + γ σ ∂σ , (9.60)
where
RC¯ = −e2 ρ−1/2∂/e−2∂/ρ−1/2 , RC = −e−2∂/e2 ρ−1∂/ . (9.64)
To compute F A C in Eq.(8.51), one needs the BRST variation of TAB . Using
AD
and Eqs.(9.57) and (9.58), it is straightforward to calculate (T −1 ) (δB T )DC . The
result is combined with K A C in Eq.(9.51) to arrive at
(FX )µ ν = δ µ ν FX ,
where
1 1
FX = C n ∂n +(∂n C n ) + e−1 C n (∂n e) − C , (9.66)
2 2
µ
for the X sector. For the ghost sector
1
(FC¯)τ τ = (FC¯)σ σ = −C n ∂n − (∂n C n ) + e−1 C n (∂n e) − C ,
2
1 1
(FC )τ τ = −C n ∂n − (∂n C n ) − e−1 C n (∂n e) + 2C + (∂τ C τ − ∂σ C σ ) ,
2 2
1 1
(FC )σ σ = −C n ∂n − (∂n C n ) − e−1 C n (∂n e) + 2C − (∂τ C τ − ∂σ C σ ) . (9.67)
2 2
Although δB Oghost is non-zero, it does not contribute because the regulator matrix
R is block diagonal in the C¯n and C n sectors. It turns out that, even in each sector,
only diagonal terms contribute because of the nature of R. Hence, in Eq.(9.67) we
display only the diagonal part of F A C . Finally, since the contribution from Cτ turns
out to be equal to the contribution from Cσ , the two ∂τ C τ − ∂σ C σ terms in Eq.(9.67)
for FC cancel. For the rest of this section, we drop these terms.
The final step is the computation of (∆S)reg . Since δB O does not contribute, we
may use Eq.(8.56). At this stage, we expand about the classical saddle point, denoted
by Σ, corresponding to the solution to the equations of motion. We set the π̄ equal
to zero and evaluate the ea m as in Eq.(7.43).
To determine the regulators and F A B at Σ, note that
e|Σ = ρ . (9.68)
where
R̃X = ρ−1/2 ∂ n ∂n ρ−1/2 . (9.70)
For ghosts,
RC¯|Σ = ρR̃C¯ρ−1 , RC |Σ = ρ−1 R̃C ρ , (9.71)
where
R̃C¯ = −ρ1/2∂/ρ−2∂/ρ1/2 , R̃C = −ρ−1∂/ρ/
∂ ρ−1 . (9.72)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 132
The conjugating factors in Eqs.(9.69) and (9.71) can be commuted past the oper-
ators F A B to arrive at the following equivalent form for (∆S)reg
!!ν
µ R̃X
(∆S)reg = F̃X exp +
ν M2 µ
!!n !!n #
n R̃C¯ n R̃C
F̃C¯ exp + F̃C exp , (9.73)
n M2 n
n M2 n 0
where µ
1
F̃X µ
= δ ν C ∂n + (∂n C n ) − C n
,
ν 2
τ σ 1
F̃C¯ = F̃C¯ = −C n ∂n − (∂n C n ) − C ,
τ σ 2
τ σ 1
F̃C = F̃C = −C n ∂n − (∂n C n ) + 2C . (9.74)
τ σ 2
The TAB matrix has been judiciously chosen so that the violation in the quantum
master equation is proportional to C. The coefficient of C n in Eq.(9.73) vanishes.
To see this, let ψr and −Er2 be the eigenfunctions and eigenvalues of any of the R̃
operators:
R̃ψr = −Er2 ψr . (9.75)
The ψr can be chosen to be real. The terms in (9.74) involving C n enter in the
combination C n ∂n + 12 (∂n C n ). Such a combination gives a zero contribution to (∆S)reg
since !!
n 1 n R̃
C ∂n + (∂n C ) exp =
2 M2
Z Zπ !
X 1 E2
dτ dσ ψr (τ, σ) C ∂n + (∂n C n ) ψr (τ, σ) exp − r2
n
=
r 2 M
0
Z Zπ !!
1 X E2
dτ dσ ∂n ψr2 (τ, σ) C exp − r2
n
→0 ,
2 r M
0
where, in the last step, we assume that quantities fall off sufficiently fast at large τ
and obey appropriate boundary conditions at σ = 0 and σ = π. The reader can also
verify the absence of a C n anomaly directly by using the methods in Appendix C.
To evaluate the terms in (∆S)reg proportional to C, let
Hr ≡ −ρ−(r+1)/2∂/ρr∂/ρ−(r+1)/2 . (9.76)
except that, for X µ , R̃X acts in a D-dimensional space, whereas H0 acts in a two-
dimensional space, since −/ ∂∂/ = I2 ∂ n ∂n = I2 (−∂τ ∂τ + ∂σ ∂σ ), where I2 denotes the
two-dimensional unit matrix. The non-zero terms in (∆S)reg in Eq.(9.73) all involve
Z Zπ
Tr Hr
exp ≡ dτ dσ κr . (9.78)
2 M2 0
0
The trace T r is over both 2 by 2 gamma space and function space. The coefficients
κr , which are computed in Appendix C, are
1
κr = (3r + 1) ∂ n ∂n ln (ρ) . (9.79)
24πi
The contributions to the violation of the quantum master equation from the X µ ,
C¯n , and Cn sectors are respectively (−1)Dκ0 , (−1)2κ−2 , and (+2)2κ1 , where the
factors in parentheses are the coefficients of C in the F̃ in Eq.(9.74). The total
anomaly A on-shell is [210, 121]
π
(D − 26) Z Z
(∆S)reg =i dτ dσ C∂ n ∂n ln (ρ)
24π
0
Z Zπ
(D − 26)
=i dτ dσ C ∂ n ∂n ln (e)|Σ . (9.80)
24π
0
∂
z A ↔ dΦA , wA ↔ . (10.1)
∂ΦA
Consider the Clifford algebra for z A and wA determined by the quadratic form h , i,
namely
{z A , wB } = δBA , {z A , z B } = 0 , {wA , wB } = 0 , (10.2)
where { , } denotes the anticommutator: {x, y} = xy + yx. A possible representation
of the Clifford algebra regards the z A as creation operators and the wA as destruction
operators. Then, the most general state at a point Φ on the manifold is created by
Ω (Φ, z) = Ω0 (Φ) + ΩA (Φ) z A + 21 ΩAB (Φ) z B z A + . . . acting on a Fock-space vacuum
|0i, defined by wA |0i = 0 for all A, i. e., it is annihilated by all the wA . Representing
the wA as ∂z∂A , |0i can be taken to be 1 when considered as a function of the z A .
With the association z A ↔ dΦA , one sees that Ω is equivalent to an element of the
exterior algebra of differential forms on M, in which differential forms are multiplied
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 136
by using the wedge product ∧. When supplemented with the exterior derivative d,
this structure becomes the de Rham complex [107, 96]. In summary, one has an
irreducible representation of the Clifford algebra in Eq.(10.2) at each point of the
manifold.
The Clifford algebra in (10.2) is symmetrical in its treatment of the elements z
and w. Hence, one can reverse the above viewpoint and regard the wA as creation
operators and the z A as annihilation operators. In this picture, let us identify the
antifields Φ∗A of the antibracket formalism with the vector-field-like objects wA . The
most general state at a point Φ is created by
1
F [Φ, Φ∗ ] = F0 (Φ) + F A (Φ)Φ∗A + F AB (Φ)Φ∗B Φ∗A + . . . (10.3)
2
acting on a state |0i′ that is annihilated by all z A . In this picture, denoted by R′ by
∂r ∗ ′
E. Witten, z A = ∂Φ ∗ , wA = ΦA , and |0i can be taken to be 1 when regarded as a
A
function of the Φ∗A . Two elements F and G in the form of Eq.(10.3) are multiplied
∂r
using {Φ∗A , Φ∗B } = 0. Exploiting the association dΦA → z A = ∂Φ ′
∗ in the R picture,
A
∂r A ∂r ∂r
the exterior derivative d ≡ ∂Φ A dΦ becomes ∂Φ A ∂Φ∗ F = −∆F when acting on a
A
general functional F of the type in Eq.(10.3). Here, we have used the definition of ∆
given in Eq.(6.7). In computing ∆F , one treats ΦA and Φ∗B as independent variables.
Because the exterior derivative is nilpotent, ∆ satisfies ∆2 = 0. In short, one arrives
at a dual picture of the de Rham complex. It is isomorphic to the standard de Rham
complex but not in a natural way because there is no preferred manner of associating
the above two Fock-space vacuums |0i and |0i′. The state |0i′ of the R′ picture is
Q
represented as (f12... ) A z A in the first picture, where f12... is arbitrary. A natural
choice does exist if M is endowed with a measure dµ = (µ12... ) dΦ1 ∧ dΦ2 ∧ . . .. Then,
one can take f = µ.
If fermionic fields are present, M is a supermanifold. Then, { , } appears as a
ǫx ǫy
graded commutator: {x, y} = xy − (−1) yx. Note that ǫ z A = ǫA + 1 = ǫ (wA ),
so that z A and wA have the opposite statistics of ΦA . For the bosonic case, ǫA = 0
for all A, and { , } becomes the usual anticommutator.
In the R′ picture, the Clifford algebra in Eq.(10.2) is satisfied when
∂r
wA = Φ∗A , z A = (−1)ǫA . (10.4)
∂Φ∗A
Then, −d becomes, when acting on F of Eq.(10.3),
∂r ∂r
(−1)ǫA +1 F = ∆F . (10.5)
∂ΦA ∂Φ∗A
The nilpotent operator ∆ of the field-antifield formalism is identified with minus the
exterior derivative.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 137
10.3 Locality
An important but technical aspect of quantum field theories is locality. Here,
we study this issue in the antibracket formalism [154, 157, 206, 132, 11, 257]. In
going from the classical action S0 to the proper solution S and to the quantum action
W , lagrangian terms are added. In a theory defined by a local classical action, the
question is whether these terms are also local. Local interactions involve fields and
derivatives, up to a finite order, of fields multiplied at the same space-time point.
Nonlocal terms are likely to lead to difficulties such as non-renormalizability, non-
unitarity or violations of causality.
The discussion of the gauge structure algebra in Sect. 2 used extensively the
consequences of the regularity condition in Eq.(2.11). An examination of the proof of
Eq.(2.12) reveals that certain operators need to be inverted so that nonlocal effects are
possible. Indeed, it is easy to find a λi so that the solution to S0,i λi = 0 in Eq.(2.12)
involves a nonlocal operator T ji or a function λ′α0 that does not fall off fast at large
space-time distances. Nonlocality often occurs when the quantity of interest vanishes
because it is an integral of a total derivative. As an example, consider n free quantum
R
mechanical particles governed by the action S0 = 12 dt q̇ i q̇i . Note that S0,i = −q̈i .
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 138
R
Let λi = q̇ i . Then, S0,i λi = − 12 dt dtd (q̇ i q̇i ) → 0. The solution in Eq.(2.12) is λ′α0 = 0
and ( δji
ji ′ 2
, for t > t′ ,
T (t , t) = ji
− δ2 , for t < t′ ,
R
since λi (t) = dt′ S0,j (t′ ) T ji (t′ , t). The trivial gauge transformation governed by
T ji (t′ , t) obeys the correct symmetry property T ij (t, t′ ) = −T ji (t′ , t).
Hence, an important concept is local completeness [154]. Local completeness
holds, when solutions to equations, such as in Eqs.(2.12), (2.17) and (2.19), can
be satisfied for local functionals or more precisely, in non-integrated versions. The
difficulty is that sometimes these equations are valid due to total derivatives.
In principle, it is possible that the gauge structure tensors involve nonlocal opera-
tors. This issue has been analyzed in refs.[92, 154, 132]. Given the locality of S0 and
that the gauge generators Rαi are local operators, then the proper solution S of the
classical master equation is local. Reference [154] used cohomological arguments to
obtain this result. The gauge-fixed classical action SΨ is then guaranteed to be local
if the gauge-fixing fermion Ψ is. Under these conditions, the classical BRST opera-
tors δB and δBΨ produce local variations. The question of quantum locality is more
involved. Since this must be analyzed on a case by case basis, no general statements
about the locality of W can be made.
The cohomology Hk (δ) is equivalent to the elements in Fk that are closed but not
exact.
A standard example is the de Rham cohomology on an n-dimensional manifold M.
The spaces Fk consist of the differential forms of order k on M and δ is the exterior
derivative d. The dimension of Hk (d) is the kth Betti number for M. In this example,
more structure can be defined. Differential forms can be multiplied using the wedge
product ∧. One can add differential forms so that the formal sum of the Fk spaces
constitutes an algebra. The exterior derivative respects addition: d(α + β) = dα + dβ,
and it is a graded derivation from the left of the wedge product: d (αj ∧ βk ) = dαj ∧
βk + (−1)j αj ∧ dβk . In applications within the antibracket formalism δ has these
properties, except, with our conventions, δ is a graded derivation from the right:
δ (αβ) = αδβ + (−1)ǫβ (δα) β, where multiplication is denoted by juxtaposition of
elements.
Cohomological methods can be powerful. However, they often involve subtle is-
sues so that one must proceed with strict rigor. The question of whether a closed
element is expressible as δ( of something ) often involves global issues; usually, it can
be done “locally”. Hence, if one is not careful, one can miscalculate the cohomology.
In regard to the antibracket formalism, the pitfalls are more severe: The spaces Fk are
almost always infinite dimensional, and, in quantizing the system, the multiplication
operation becomes singular. Furthermore, an ambiguity concerning the issue of lo-
cality in Sect. 10.3 enters: one needs to decide whether local or non-local functionals
are permitted.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 140
One cohomology in the antibracket formalism uses the classical BRST operator δB
for δ. The spaces Fk consist of smooth functionals of fields with ghost number k. Note
that k ranges over all integers, both positive, negative and zero. Functionals form an
algebra since they can be added and multiplied. Furthermore, δB satisfies the correct
properties: it is nilpotent and is a graded derivation from the right. The cohomology
Hk (δB ) is the classical space of observables in the sector with ghost number k.
For proving certain results, two other cohomologies are useful. The first uses the
Koszul-Tate differential δkt [179, 58, 240]. Let G+ be the condition of setting all ghost
fields to zero: n h i o
G+ ≡ ΦA = 0 | gh ΦA ≥ 1 . (10.14)
Then, the Koszul-Tate differential is defined by
where δB X = (X, S). When acting on fields (and ghosts), δkt produces zero
A ∂l S
δkt Φ = =0 , (10.16)
∂Φ∗A G+
since δkt ΦA , having ghost number greater than one, must be proportional to ghost
fields. Hence, the interesting action is on antifields:
ǫA +1 ∂r S
δkt Φ∗A = (−1) . (10.17)
∂ΦA G+
2 ∗
To check the nilpotency of δkt , it suffices to compute δkt ΦA . One finds
! !!
∂r ∂r S ∂r S
ǫA +ǫB
2
δkt Φ∗A = (−1) ,
∂Φ∗B ∂ΦA ∂ΦB
G+
The Koszul-Tate differential is zero because S satisfies the classical master equation.
Because δkt is constructed using the BRST operator it is a graded derivation, i.e.,
The action of δkt for the antifields with ghosts numbers −1, −2 and −3 is respec-
tively
δkt φ∗i = (−1)ǫi +1 S0,i ,
δkt Cα∗ = (−1)ǫα φ∗i Rαi ,
∗ ǫα1 +1 ∗ α 1 ǫi ∗ ∗ ji
δkt C1α1 = (−1) Cα R1α1 + (−1) φi φj V1α1 , (10.19)
2
where the tensors R and V are given in Sect. 2.
An important result is that Hk (δkt ) = ∅ for k ≤ −1, where ∅ denotes the empty
set [105, 152]. Let us formally verify this for k = −1 and k = −2. One can set terms
involving ghosts to zero because they either transform to zero or are set to zero. The
most general element α−1 , with ghost number −1 and constructed from antifields, is
α−1 = φ∗i λi ,
where λi are functionals of the φ, with ǫ (λi ) = ǫi so that ǫ (α−1 ) = 1. Since δkt α−1 =
−S0,i λi , α−1 is closed if S0,i λi = 0. Accordingly, λi must be expressible as in Eq.(2.12),
so that
′
α−1 = φ∗i Rαi λ α + φ∗i S0,j T ji .
This is the most general form for a closed element with ghost number −1. The key
?
question is whether α−1 is exact, i.e., α−1 = δkt β−2 . Let
1
β−2 = Cα∗ λ α + (−1)ǫi +1 φ∗i φ∗j T ji .
′
2
Then, a short computation reveals that δkt β−2 = α−1 . In the k = −1 sector, there are
no closed elements that are not exact, so that the cohomology is trivial H−1 (δkt ) = ∅.
In the k = −2 sector, the most general element is
1
α−2 = Cα∗ λα + (−1)ǫi φ∗i φ∗j M0ji ,
2
where ǫ (λα ) = ǫα and ǫ M0ji = ǫi + ǫj (mode 2), so that ǫ (α−2 ) = 0. The functional
M0ji obeys M0ji = (−1)ǫi ǫj +1 M0ij . A computation reveals that
δkt α−2 = φ∗i Rαi λα − S0,j M0ji ,
so that
Rαi λα − S0,j M0ji = 0 , (10.20)
?
if α−2 is to be closed. Is α−2 = δkt β−3 . It takes a little work to show that α−2 is exact.
α
For α−2 to be closed, λ must obey Eq.(2.12). Substituting the solution of Eq.(2.12)
for λα into Eq.(10.20) gives
λ α1 + S0,j T0jα − S0,j M0ji = 0 ,
′
Rαi R1α
α
1
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 142
Suppose that a theory has no gauge invariances. Then the classical observables cor-
respond to functionals taking on distinct values on Σ. This space is H0 (δkt ).
If a theory has gauge invariances, then the observables should be the gauge-
invariant elements of H0 (δkt ). To facilitate the issue of gauge invariance, one in-
troduces the vertical differential δg [106, 152]. An alternative name for δg is the
“exterior derivative along the gauge orbit”. It is defined as
where G− corresponds to the condition of setting antifields to zero and going on-shell
with respect to the original fields, i.e.,
( )
∂r S
G− = Φ∗A = 0, i = 0 . (10.23)
∂φ
Antifields can be ignored in evaluating the vertical differential because they are either
set to zero or transformed to zero since
!
∂l S
δg Φ∗A =− =0 , (10.25)
∂ΦA
G−
which follows from ghost number considerations or Eq.(10.23). On fields, one has
!
∂l S
A
δg Φ = . (10.26)
∂Φ∗A
G−
With the above insights, one realizes that observables should roughly correspond
to H0 (δkt ) ∩ H0 (δg ). However, a difficulty arises. The intersection H0 (δkt ) ∩ H0 (δg )
does not make sense unless δkt δg + δg δkt = 0. The situation is analogous to the one
in quantum mechanics where one seeks a state that is simultaneously the eigenvector
of two different operators. Such a state is possible if the two operators commute.
Because of the graded nature of δkt and δg , the analog condition is that δkt and δg
anticommute. The difficulty can be posed as a question: Should one take the gauge-
invariant elements of H0 (δkt ) or should one take the elements of H0 (δg ) modulo
the equivalence relation of Eq.(10.21)? If δkt δg + δg δkt = 0, then the above two
procedures yield the same result. In such a case, one can define a nilpotent BRST
operator δB ≡ δkt + δg , and the observables correspond to the BRST cohomology.
Unfortunately, δkt δg +δg δkt 6= 0 in general. An inspection of δB , δkt and δg reveals that
δB = δkt + δg + extra terms. The extra terms render δB nilpotent, by compensating
for the failure of the anticommutivity of δkt and δg . The BRST operator is the
natural extension of δkt + δg . The elements of the cohomology of δB are the classical
observables [105, 152]. They are the definition of what one means by the “gauge-
invariant functionals on Σ”.
When quantum effects are incorporated, the quantum BRST transformation δB̂
is relevant. As discussed in Sect. 8.1, the quantum observables correspond to the
elements of the cohomology of δB̂ .
Because canonical transformations preserve the antibracket, the cohomology of
δB is independent of the basis, as can be seen as follows. Given a proper solution
S[Φ, Φ∗ ] in one (untilde) basis, a solution S̃[Φ̃, Φ̃∗ ] in another (tilde) basis is given
by S[e Φ̃, Φ̃∗ ] ≡ S[Φ, Φ∗ ]. Likewise, given any functional X[Φ, Φ∗ ], one can define a
functional X f of tilde fields using X[
f Φ̃, Φ̃∗ ] ≡ X[Φ, Φ∗ ]. The tilde antibracket of Xf
and S, e as a function of tilde fields and antifields, equals (X, S) as a function of
untilde fields. Hence, X f is closed if and only if X is, and Xf is exact if and only if X
is. Consequently, there is an exact isomorphism of the cohomologies.
Since the gauge-fixed BRST transformation δBΨ is not nilpotent, one cannot di-
rectly define a cohomology associated with δBΨ . However, according to Eq.(6.83),
δB2 Ψ is proportional to the equations of motion of the gauge-fixed action SΨ . Define
an equivalence relation, denoted by ≈, that equates two quantities if they differ by
terms proportional to the equations of motion for SΨ . Then, one has δB2 Ψ ≈ 0 and a
gauge-fixed BRST cohomology can be defined. What is the relation between Hn (δBΨ )
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 145
and Hn (δB )? The connection is best seen by going to the gauge-fixed basis for δB .
Let
Y (Φ̃, Φ̃∗ ) = y(Φ̃) + y A (Φ̃)Φ̃∗A + . . . (10.29)
be the antifield expansion of a functional Y in the gauge-fixed basis Φ̃A and Φ̃∗A of
Sect. 6.6. Eqs.(6.79) and (6.84) imply
∂l SΨ
δB Y = δBΨ y − y A + O(Φ̃∗ ) , (10.30)
∂ Φ̃A
so that
(δB Y )|{Φ̃∗ =0} ≈ δBΨ y , (10.31)
since the last term in Eq.(10.30) is proportional to gauge-fixed equations of motion.
Eq.(10.31) implies that if Y is δB -closed then y is δBΨ -closed, and that if Y is δB -exact
then y is δBΨ -exact. This is not enough to establish any relation between Hn (δBΨ ) and
Hn (δB ). Given a element of y of Hn (δBΨ ), one must uniquely construct an element
Y of Hn (δBΨ ), under the condition that Y |{Φ̃∗ =0} = y. In other words, one must find
the higher-order terms in Eq.(10.29). References [151, 105, 103] succeeded in doing
this. For additional discussion, see refs.[157, 253]. The cohomologies governed by δBΨ
and δB are equivalent.
where a dot over a field indicates a time derivative, where the conjugate momentum
of φi is πi ≡ ∂∂lφ̇Si0 , i = 1, . . . , n, and where HS0 is obtained as a function of the φ and π
by solving for φ̇ in terms of the π and φ. For some systems, this velocity-momentum
inversion process is not possible due to the presence of primary constraints. Even
in this case, a hamiltonian HS0 can be uniquely constructed on the surface of these
primary constraints. We symbolically represent the procedure of obtaining a hamil-
tonian from an action diagrammatically as
S0
y
H S0
For a wide class of gauge theories, HS0 is of the form
where the original n variables φi are split into dynamical degrees of freedom ϕa , a =
1, . . . , m ≤ n and Lagrange multipliers λα for the (secondary) constraints Tα . In
Eq.(10.33), H0 and Tα are functions of the ϕ and their momenta only. The velocities
λ̇α are usually assumed not to appear in S0 . This means that the momenta of the
λα , namely πα , are primary constraints and do not enter in HS0 . For example, in
a Yang-Mills theory, the hamiltonian density H0 is H0 = 21 Eai Eia + 41 Fija Faij , where
Eai = −Fa0i = Fa0i are the canonical momenta for the potentials Aai , the constraints
Tα correspond to Gauss’s law: Ta = −Dia b Ebi , and the Lagrange multipliers λα are
Aa0 .
For simplicity, assume that the constraints Tα and the hamiltonian H0 are first
class, i.e,
γ
{Tα , Tβ }P B = Tαβ Tγ , {H0 , Tα }P B = Vαβ Tβ , (10.34)
where { , }P B denotes the graded Poisson bracket defined by
∂r F ∂l G ∂r F ∂l G
{F, G}P B = (−1)ǫi − . (10.35)
∂φi ∂πi ∂πi ∂φi
Here, the sum over i is such that all fields and momenta are included. If the constraints
are second class, Dirac brackets [88] must be used. Note that {πi , φj }P B = −δij .
The BFV program is based on BRST invariance. One introduces ghosts and their
conjugate momenta. The ghosts needed correspond to the minimal set, introduced
in Eq.(4.1). For the irreducible case, they are the C α . We use P̄a to denote the
momentum associated with a ghost C a , where a is a label that enumerate all ghosts.
The Poisson bracket in Eq.(10.35) is then extended to include a sum over ghosts.
With these conventions, {P̄b , C a }P B = −δba . The ghost numbers and statistics of the
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 147
h i
BFV ghosts are the same as in Sect. 4.1. For momenta, gh P̄a = −gh [C a ] and
ǫ P̄a = ǫ (C a ). A canonical generator of the BRST transformations QB and an
extended hamiltonian H are constructed, using the requirement that QB be nilpotent
and that H be BRST invariant:
They can be expanded as a power series in ghost fields, for which the first few terms
are
1
H = H0 + C α Vαβ P̄β + . . . , QB = C α Tα − (−1)ǫβ C β C γ Tγβ
α
P̄α + . . . . (10.37)
2
It turns out that the requirement of {QB , QB }P B = 0 reproduces the relations defining
the structure of the gauge algebra at hamiltonian level. In other words, QB plays a
role analogous to the proper solution S of the antibracket formalism.
In this approach, O is an observable if it is BRST-invariant, i.e., {O, QB }P B = 0.
Thus, the hamiltonian is an observable. Two observables O1 and O2 are considered
′ ′
equivalent if O2 = O1 + {O , QB }P B , for some O . A state | ψ i is called physical if
QB | ψ i = 0. Two states | ψ1 i and | ψ2 i are considered equivalent if | ψ2 i = | ψ1 i +
QB | ψ ′ i, for some | ψ ′ i.
Given a suitable hamiltonian H, a lagrangian can be constructed via
Z Z
i [dΠ] i i
exp SH [Φ, Φ̇] = exp dt (Φ̇ Πi − H[Φ, Π]) , (10.38)
h̄ 2πih̄ h̄
where Φ denotes all degrees of freedom and Π denotes the corresponding momenta.
We indicate the process of constructing an action SH from a hamiltonian H by the
following diagram
SxH
H
In the BFV formalism, to obtain a hamiltonian HΨ , which is appropriate for
insertion in the functional integral, a fermion Ψ with ghost number minus one is
used. As in the antibracket formalism, BRST trivial pairs exist. Given two fields Λ
R
and Σ, and their conjugate momenta, P̄Λ and P̄Σ , a term dd−1 x Σ P̄Λ can be added
to QB without ruining nilpotency. The next step in the BFV program is to introduce
additional fields and their momenta and add them as trivial pairs to QB . These
fields are the analogs of the auxiliary gauge-fixing fields of Sect. 6.2. They include
antighosts, extraghosts, and the Lagrange-multiplier fields of Eq.(6.22). The fermion
Ψ in the hamiltonian formulation must satisfy conditions similar to those in Sect. 6.3
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 148
for the Ψ in the antibracket formalism. We denote the BRST charge extended by the
inclusion of the additional trivial terms by Qnm
B . The hamiltonian HΨ is given by
HΨ = H − {Ψ, Qnm
B }P B . (10.39)
R
Let ZΨ = [dΦ] exp h̄i SHΨ , where SHΨ is constructed from HΨ via Eq.(10.38). The
Fradkin-Vilkovisky theorem [109] states that ZΨ is independent of Ψ.
The equivalence of the BFV hamiltonian and antibracket methods is established
if the remaining leg of the following diagram
Sx0 −→ SΨ
y
H S0 −→ HΨ
can be completed. In other words, is SHΨ , as constructed from HΨ via Eq.(10.38),
equivalent to SΨ as obtained from the antibracket formalism? Likewise, is HSΨ , as
constructed from SΨ via Eq.(10.32), equivalent to the BFV hamiltonian HΨ ? Another
question is whether the gauge-fixed BRST charge QNoether , as constructed from SΨ
using Noether’s theorem, coincides with the BRST charge QnmB for the BFV formalism.
The affirmative answer to the above questions, obtained in refs.[104, 91], implies that
construction processes in
Sx0 −→ SxΨ
y y
HS0 −→ HΨ
commute to give equivalent results.
One can also ask whether an equivalence occurs before the introduction of gauge-
fixing and Ψ:
S
y?
H
Clearly, a straightforward correspondence cannot exist because S contains antifields.
However, at least for closed irreducible theories, if certain antifields are set to zero and
others are identified with ghost momenta, then an equivalence of HS , as constructed
from S via Eq.(10.32), and the BFV H is achieved [31]. Similar results have been
obtained in refs.[230, 91].
If sources for the BRST transformations are included at the hamiltonian level, the
above correspondence can be made clearer. Then, the sources in the hamiltonian for-
mulation can be identified with antifields in the antibracket formalism. This method
was used in refs.[140, 75, 76] to establish the equivalence in the gauge-fixed basis.
An open problem is to extend all of the above analysis to the quantum case in a
rigorous manner. That situation is more difficult due to operator ordering problems
and the singular character of field theories.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 149
10.6 Unitarity
The difficulty in proving unitarity in covariant approaches to quantizing gauge
theories is due to the presence of ghosts and of unphysical degrees of freedom with
negative norms. One often deals with indefinite-metric Hilbert spaces. Unitarity
can be spoiled in theories with kinetic energy terms of the wrong sign and/or non-
hermitian interaction terms. Wrong sign kinetic energy terms almost always arise in
gauge theories with particles of spin one or higher. Due to the sign of the metric
component η00 , there are potential difficulties with the temporal components, such as
A0 in electromagnetism, Aa0 in Yang-Mills theories, and g0i in gravity. Faddeev-Popov
and other gauge-fixing ghosts enter in loops with the wrong sign, and would lead to
a violation of unitarity, if their contributions were considered in isolation.
Let us summarize how unitarity is established in certain covariant quantization
procedures. First of all, one needs to assume that there are not any non-hermitian
interactions in the original theory and that the spatial components of tensors have the
correct sign in kinetic energy terms. In other words, the theory should be “naively”
unitary.
The first approach is as follows. In some theories, there exists a unitary gauge,
in which it is evident that the unphysical excitations are not present. If one can
establish the gauge invariance of the S-matrix, then unitarity can be proven by going
from a covariant gauge to an unitary one [100]. Unfortunately, this method is only
well developed for irreducible theories with closed algebras. For reducible systems,
this approach often encounter difficulties, although for some specific examples it has
been successfully implemented [116].
Another method for checking unitarity is in perturbation theory via Feynman
diagrams [243, 245]. Using the Ward-Takahashi [265, 239] or Slavnov-Taylor iden-
tities [241, 233], as well as the Landau-Cutkosky rules [95], one tries to show that
contributions from the unphysical polarizations of the classical fields are cancelled by
contributions from ghost fields or from other sources.
A third approach proceeds via canonical quantization. The “physical sector”
is selected out by imposing some subsidiary conditions that remove negative norm
states. The physical sector should be stable under time evolution and should involve
a non-negative metric. A well-known example of this approach is the Gupta-Bleuler
procedure [53, 144] for quantizing QED. All components of the electromagnetic field
Aµ are used; however only states | ϕ iphys satisfying
QB | ϕ iphys = 0 , (10.41)
where QB is the hermitian nilpotent BRST operator. Eq.(10.41) is the basis for BRST
quantization. We use Vphys to denote the space of states annihilated by QB . In the
BRST approach, the hamiltonian is automatically hermitian so that the S-matrix is a
unitarity operator in Vphys . However, there is a possible difficulty with Vphys . Despite
the fact that QB commutes with the hamiltonian, the positive semidefiniteness of
the norm of Vphys is not ensured. The question of unitarity in BRST quantization
becomes that of proving the positive semidefiniteness of Vphys , and must be analyzed
model by model.
However, T. Kugo and I. Ojima [181, 182] (see also [196]), obtained criteria under
which unitarity does hold. They established a connection with the metric structure
of Vphys and the multiplets of the algebra generated by the conserved BRST charge
QB and the conserved ghost number charge QC .16 With our conventions, QC is
antihermitian: Q†C = −QC . These generators satisfy
1
[QC , QB ] = QB , [QC , QC ] = 0 , {QB , QB } = Q2B = 0 . (10.42)
2
Three types of multiplets are possible:
(a) “True physical states”: BRST singlets with zero ghost number.
(b) Doublets: pairs of BRST singlets related by ghost conjugation.
(c) Quartets: pairs of BRST doublets related by ghost conjugation.
Roughly speaking, ghost conjugation is the operation that interchanges ghosts and
antighosts. Under this operation, the sign of the ghost number of a state is flipped.
In the next three paragraphs, we explain the classification of the multiplets.
One can choose states to be eigenfunctions of Qc . Let | g i be a state with a non-
zero ghost number g. Then | g i has zero norm since h g |Qc | g i = gh g | g i = −gh g | g i,
the first equality arising when Qc acts to the right, and the second equality arising
when Qc acts to the left. Non-zero matrix elements occur only when bra and ket
states have opposite ghost numbers. Under application of QB , the ghost number of
16
We assume there are no anomalies associated with QB and QC . For the first-quantized string,
this is actually not the case for QC , but BRST quantization is still possible [163, 174, 115]. For
similar analyses in other models see refs.[59, 42, 48, 60].
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 151
a state is increased by one. Such states | s i = QB | s′ i also have null norms since
h s | s i = h s′ |QB QB | s′ i = 0.
Due to the nilpotency of QB , the representations are either BRST singlets or
BRST doublets. A BRST singlet | s i satisfies QB | s i = 0 and | s i = 6 QB | s′ i for any
′ ′ ′
| s i. If QB | s i = | s i = 6 0, then | s i is a member of the BRST doublet consisting
′
of | s i and | s i. The upper member of a doublet | s i is annihilated by QB , since
QB | s i = QB QB | s′ i = 0, and it carries one unit of ghost number more than | s′ i:
Qc | s i = Qc | s′ i + 1.
If | s i is a BRST singlet and carries ghost number zero, then it is of type (a). If
| s i is a BRST singlet and carries non-zero ghost number g, then it is of type (b).
Under ghost conjugation, another BRST single with ghost number −g is created, thus
forming the pair. If | s i and | s′ i constitute a BRST doublet, then ghost conjugation
produces another BRST doublet and a type (c) multiplet is obtained.
For an irreducible gauge theory, T. Kugo and I. Ojima in [181, 182] proved that (i)
if type (a) states have positive definite norm and (ii) if type (b) states are absent, then
quartets only appear in Vphys through zero norm combinations. Consequently, when
(i) and (ii) are satisfied, Vphys has a positive semidefinite norm. To obtain a unitary
theory, one mods out the null-norm states: Two states are identified if they differ by
a null-norm vector. Clearly, null-norm states are identified with the null state. The
modding-out procedure automatically restricts states to the zero-ghost number sector,
since states with non-zero-ghost number have zero norms. Furthermore, because
BRST-trivial states QB | s′ i are null-norm vectors, all that remains after modding out
are the non-trivial elements of the g = 0 BRST cohomology, i.e., states with ghost
number zero that are annihilated by QB and that cannot be expressed as QB | s′ i for
any state | s′ i. This sector is preserved under time evolution because QB and QC
commute with the hamiltonian.
In the g = 0 sector, it makes sense to identify null-norm states with the null
vector because they decouple from matrix elements involving observables, such as the
hamiltonian. Observables O are BRST-invariant operators: [O, QB ] = 0. If | t i is a
BRST-trivial state, so that | t i = QB | s′ i, and if | s i is any element of Vphys , so that
QB | s i = 0, then h s |O| t i = h s |OQB | s′ i = h s |QB O| s′ i = 0.
For reducible systems, ghosts for ghosts and extraghosts arise, some of which
have zero ghost number. Hence a third condition arises for reducible theories: (iii)
a state of Vphys involving ghosts in the g = 0 sector must be a member of a quartet
multiplet. This guarantees that they are null vectors and do not ruin the positive
semidefiniteness of Vphys .
The above conditions provide criteria for establishing the positivity of the norm
and hence unitarity in a covariant formulation. Reference [70, 181, 182] established
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 152
unitarity for Yang-Mills theories by proving (i) and (ii) for this case.
In perturbation theory and in a Fock space representation, A. Slavnov in [234]
used (i)–(iii) to obtain simpler criteria. The most important requirements, apart
from the positivity of the norm of type (a) states, were that QB be nilpotent and that
it have nontrivial action on all ghost fields or their conjugate momenta. Under these
conditions, Vphys has a positive semidefinite norm. Then, S. A. Frolov and A. Slavnov
[117] using the hamiltonian BFV-BRST formalism for lagrangians L of the form in
Eqs.(10.32) and (10.33), verified the above-mentioned conditions perturbatively. The
(0)
analysis was simplified because one could use the free BRST charge QB . The criteria
(0) (0) (0)
became that QB QB = 0 and that QB have non-trivial action on all ghosts. Given
the validity of perturbation theory, their result on the unitarity of a gauge theory
holds for the finite reducible case.
S. A. Frolov and A. Slavnov in ref.[118, 235], were able to translate the above
program into a lagrangian approach, by using an effective action Aef f . The action
(0)
and BRST charge were perturbatively expanded in a series: Aef f = Aef f + . . . and
(0) (0)
QB = QB + . . .. The term Aef f was the leading order part of the general gaussian
gauge-fixed action SΨ of the field-antifield formalism presented in Sect. 6.4. Requiring
nilpotency and BRST invariance of the action lead to a series of recursion relations
for the higher order terms in Aef f and QB . The action Aef f , thus obtained, is
constructed using unitarity requirements. Finally, when certain conditions on the
rank of the gauge generators are imposed, the free BRST charge is seen to act non-
trivially on ghosts fields and unphysical polarizations of the classical fields, thereby
yielding a unitary theory if the classical gauge-invariant degrees of freedom have a
positive norm.
The problem of unitarity in the field-antifield formalism was addressed in [130,
206, 203, 204]. A perturbative solution of the proper solution S was obtained in
[130, 206, 132] (see also ref.[13]). Then, a general gaussian gauge-fixing procedure
was performed, using a fermion Ψ of the type given in Sects. 6.3 and 6.4. It was
shown that BRST invariance of the gauge-fixed action and nilpotency of the gauge-
fixed BRST transformation lead to the same recursion relations obtained in [118, 235],
and that the leading two terms of SΨ agree with Aef f . The conclusion is that the field-
antifield formalism produces an action SΨ that coincides with Aef f of ref.[118, 235]
obtained by unitarity considerations.
The above approaches to unitarity are formal in that the difficulties with field-
theoretic infinities are not addressed. The renormalizability or non-renormalizability
is not used. To proceed rigorously, one needs to regulate the theory with a cutoff,
verify unitarity, and then make sure that unitarity remains as the cutoff is removed.
The issue of locality also enters here. For example, it may happen that Aef f or S
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 153
contains non-local terms. This does not necessarily ruin unitarity, but might signal
that the theory is non-renormalizable or ill-defined. Studies of unitarity without
using perturbation theory for general systems with finite degrees of freedom, such as
in quantum mechanics, have been carried out in ref.[194].
Q = QL + QR , (10.53)
generator for rigid rotation of the first-quantized parameter σ, which labels the points
along the string. Since, in the closed-string
case,
σ is periodic, there is no preferred
L R
choice of an origin. The condition L0 − L0 C = 0 does not lead to an equation of
motion for C since the 21 ∂µ ∂ µ terms in LL0 and LR 0 cancel. From Eq.(7.34), one sees
that it implies that the mass ML of the left-sector must equal the mass MR of the
right-sector, a well-known constraint of first-quantized closed-string states. However,
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 157
the operation ◦ does not preserve the constraint. One can modify ◦ to ˆ◦ by averaging
over a rigid rotation that rotates the product string over angles ranging from 0 to 2π.
The new ◦ˆ no longer is associative, as can be checked by drawing some pictures.
The ghost number problem can be fixed by inserting a factor of c− 0 . Define the
quadratic form Z
hA, Bi = Aˆ◦ c−
0B . (10.57)
closed
For on-shell external states, this interaction correctly produces three-point interac-
tions.
(2) (3)
Tree-level gauge invariance is violated for the theory described by S0 + S0
due to the violation of the associativity axiom. However, by adding higher-order
terms gauge invariance can be restored [168, 169, 170, 180, 212, 183, 280]. The new
interactions can be defined by relatively simple geometrical constraints [280, 281].
This leads to a tree-level non-polynomial closed-string field theory. Unfortunately,
the classical theory needs further modification at the quantum level. One-loop and
higher-loop amplitudes are not produced by using only tree-level vertices. It is at
this stage where the antibracket formalism has been of great utility. Interaction
terms proportional to powers of h̄ need to be added in a manner similar to Eq.(6.14).
To ensure that the theory is quantum-mechanically gauge invariant, the work in
refs.[280, 224, 281, 282, 149, 225] has relied on the antibracket formalism. The guiding
principle is that the quantum closed-string field theory must satisfy the quantum
master equation.
The antibracket is defined using the quadratic form in Eq.(10.57) [224, 281, 282].
As in the open string field theory, the system is infinitely reducible so that there are
ghosts for ghosts ad infinitum. The fields can be collected into one object Ψc in a
manner similar to the open string case in Eq.(5.23)
s∗+∗4 4 3 2 1 −s + 1
Ψc ≡ . . . + Cs + . . . + ∗ C0∗ + ∗ C ∗ + C + C0 + . . . + Cs + . . . , (10.60)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 158
where the ghost number is indicated above the field. The close-string Hodge opera-
tion, denoted by a superscript ∗ in front of a field, is defined using the quadratic form
in Eq.(10.57). It takes a p-form into a 5 − p, where the order of string form is the
same as the ghost number.
Let ϕs be a complete set of normalized first-quantized states for g (ϕs ) ≤ 2. Let
∗ s
ϕ denote the corresponding state transformed by the Hodge star operation. The
∗ s
ϕ are a normalized complete set of states for ghost numbers greater than 2. With
these definitions, Ψc in Eq.(10.60) can be written in terms of ordinary particle fields
P
ψ s via Ψc = s (ϕs ψ s + ∗ ϕs ψs∗ ) where ψs∗ are the antifields of ψ s . The quantum
master equation is then the same as the particle case, namely 12 (W (Ψc ), W (Ψc )) =
r X ∂l Y ∂r X ∂l Y
ih̄∆W (Ψc ), where the antibracket is (X (Ψc ) , Y (Ψc )) = ∂∂ψ s ∂ψ ∗ − ∂ψ ∗ ∂ψ s and ∆ =
s s
(−1)ǫs +1 ∂ψ
∂r ∂r
s ∂ψ ∗ .
s
The solution of the quantum master equation for the closed-string field theory
is presented in refs.[281, 282]. This tour-de-force work goes beyond the goals of our
review. The reader interested in this topic can consult the above references for more
discussion.
The current formulation of string fields theories is developed around a particular
space-time background. Any background is permitted, as long as it leads to a nilpo-
tent first-quantized BRST charge. Such BRST charges correspond to two-dimensional
conformal field theories with the total central charge of the Virasoro algebra equal
to zero. Usually, the flat space-time background in 26 dimensions is used. Since
string theories contain gravity, it should be possible to pass from one background to
another. It is an interesting question of whether there is background independence
of string field theory [223, 222, 274, 224, 275, 225]. A proof for bosonic string field
theories has been obtained in refs.[224, 225], for backgrounds infinitesimally close.
The antibracket formalism has played an important role in the analysis. The basic
idea is that string field theory, formulated about a particular background B, corre-
sponds to a particular solution SB of the classical master equation. Reference [224]
demonstrated that, for two nearby backgrounds B1 and B2 , SB1 and SB2 are related
by a canonical transformation of the antibracket. The conclusion is that string field
theory is background independent, although not manifestly. Barring difficulties with
singular expressions, ref.[225] has extended the result to the quantum case. For the
quantum system, a particular background B corresponds to a solution WB of the
quantum master equation.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 159
Let φb∗î be the antifield of φbî . Given Eq.(4.29), the classical gauge structure of the
extended theory should be governed by an action Se = S + φb∗î R b î C α . Since field indices
α
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 160
h i Z 1 h i
M1 φ, φb = −i b
dsAî F φ, φs φbî , (10.62)
0
where the Aî are the antifield-independent part of the anomalies and F i is the finite
gauge transformation of the classical fields φi under the anomalous part of the group.
The BRST variation of M1 gives the original anomaly: (M1 , S) e = iA C α .
α
e
However, in the extended antibracket formalism the action S is not proper [77].
To have a well defined perturbation expansion, it is necessary to modify Se to a new
extended action Sb that satisfies the classical master equation
b S
S, b =0 , (10.63)
The final stage is to find a solution W c to the quantum master equation in the ex-
tended space. Because of the above-mentioned scaling of φ, c is a series in h̄ 21 [77, 257]
b W
c =S
rather than h̄: W b + h̄ 21 M
c1 + . . .. After regulating the extended theory, ∆S b
2 reg
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 161
the complete one-loop obstruction Ab1 to the quantum master equation in the extended
space can be cancelled. In terms of the unscaled extra variables, this adjustment cor-
responds to a finite renormalization of the original expression of the Wess-Zumino
term in Eq.(10.62), M1 → M c . Note that the locality of M c1 cannot be guaranteed.
1
2
Likewise, locality of the renormalized Wess-Zumino term is not assured due to the
integral over the variable s in expressions like (10.62). In some cases, such as the
first-quantized bosonic string [131, 133] or the abelian Schwinger model [207], only
local action terms are generated. Then, the anomalous theory makes sense at the
quantum level. However, for chiral QCD in two dimensions [131], the integral over s
remains. Even in these cases, the violation of locality is in some sense not severe: The
equations of motion are local, a situation referred to as quasilocal. When the quan-
tum extended theory is well-defined, the final stage, namely gauge-fixing, proceeds in
a manner similar to the non-anomalous case [131].
Acknowledgements
We thank G. Barnich, C. Batlle, F. De Jonghe, M. Henneaux, A. K. Kost-
elecký, J. M. Pons, J. Roca, R. Siebelink, A. Slavnov, P. Townsend, W. Troost, S.
Vandoren, A. Van Proeyen and F. Zamora for discussions. This work is supported
in part by the Comisión Interministerial para la Ciencia y la Technología (project
number AEN-0695), by a Human Capital and Mobility Grant (ERB4050PL930544),
by the National Science Foundation (grant number PHY-9009850), by a NATO Col-
laborative Research Grant (0763/87), by the Robert A. Welch Foundation, and by
the United States Department of Energy (grant number DE-FG02-92ER40698).
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 162
∂l X ∂r X
dX(φ) = dφ = dφ . (A.2)
∂φ ∂φ
The formula for the variation δX of X with respect to φ resembles Eq.(A.2):
∂l X ∂r X
δX(φ) = δφ = δφ . (A.3)
∂φ ∂φ
What is the relation between left and right derivatives? If φ is commuting (ǫ (φ) =
0), then ∂∂φ
lX
= ∂∂φ
rX
, so that one only needs to be careful when φ is anticommuting
(ǫ (φ) = 1). Assume φ is anticommuting. Then φφ = 0. Without loss of generality
we may assume that X = φY + Z where Y and Z have no φ dependence. The left
and right derivatives of X are then ∂∂φlX
= Y and ∂∂φ rX
= (−1)ǫY Y = (−1)ǫ(φ)(ǫX +1) Y .
For all cases,
∂l X ∂r X
= (−1)ǫ(φ)(ǫX +1) . (A.4)
∂φ ∂φ
As a pedagogical exercise, let us derive the graded antisymmetry property of the
bracket in Eq.(4.5). Start with the definition of (Y, X) in Eq.(4.4) and interchange
the order of derivatives to obtain
∂r Y ∂l X ∂r Y ∂l X
(Y, X) = − =
∂ΦA ∂ΦA ∂Φ∗A ∂ΦA
∗
As another exercise, let us verify the formulas for (F, F ) and (B, B) in Eq.(4.6)
When Y = X, the second term in Eq.(4.4) can be written as
∂r X ∂l X ∂r X ∂l X
∗
= (−1)(ǫX +1)(ǫX +1) ,
∂ΦA ∂Φ A ∂ΦA ∂Φ∗A
′
= (−1)ǫ(φ)ǫ(φ ) ′ ,
∂φ∂φ ∂φ ∂φ
∂r ∂r X ǫ(φ)ǫ(φ′ ) ∂r ∂r X
= (−1) . (A.6)
∂φ∂φ′ ∂φ′ ∂φ
In the first equation, derivatives act from different directions and hence commute. In
the second and third equations, one must be careful of the order.
If X is a functional of Y which is a function of φ, one has the chain rules
∂r X (Y (φ)) ∂r X ∂r Y
= ,
∂φ ∂Y ∂φ
∂l X (Y (φ)) ∂l Y ∂l X
= . (A.7)
∂φ ∂φ ∂Y
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 164
S0,i |Σ = 0 . (B.2)
In other words, the on-shell degeneracy of the hessian in Eq.(B.1) is completely due
to the n − ndof independent null vectors Rαi associated with gauge transformations
and does not come from some other source [27, 28]:
∂l ∂r S0 i
R =0 . (B.3)
∂φi ∂φj α Σ
S0,a = 0 , identically ,
∂l S0,s
rank = ndof . (B.4)
∂φi φ=φ0
The regularity condition assumes that the equations of motion S0,i constitute a
regular representation of the stationary surface Σ. This means that the functions S0,i
can be locally split into independent, Gs , and dependent ones, Ga , in such a way that
In other words, the regularity condition ensures that locally the changes of variables
φi → [ϕs , χa ] or φi → [Gs , Ga ] makes sense [27, 28, 105].
When the regularity condition is fulfilled, it can be shown that any smooth func-
tion that vanishes on the stationary surface Σ can be written as a combination of the
equations of motion [28, 103, 105, 106], i.e.,
where the λj may be functions of the φi . No restrictions are made on the λj (φ).
Putting restrictions can lead to violations of (B.5). An example is presented in [257].
By considering only local functionals, ref.[257] found cases for which Eq.(B.5) could
not be satisfied as a local combination of the equations of motion.
For more details on regularity conditions as well as derivations of the above results
consult references [81, 27, 28, 106, 103, 105, 152].
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 166
Z1 Zu
du ds exp [(1 − u) R0 ] V exp [(u − s) R0 ] V exp [sR0 ] + . . . . (C.1)
0 0
Typically, R0 is independent of the cutoff M, and V goes like inverse powers of M
so that only a few terms in Eq.(C.1) need to be kept.
The anomaly equation Eq.(8.53) involves a functional trace [120]. If one uses
momentum-space eigenfunctions to saturate the sum, then expressions such as
When derivatives in O (∂µ + ikµ ) act on the function 1, they produce zero.
For the spinless relativistic particle system, we begin by taking Eq.(9.18) and
commuting exp (ikτ ) through the expression, using Eq.(C.2), to obtain
(∆S)reg =
!
Z Z∞ − d
+ ik ρ−1 d
+ ik
dk −1 d dτ dτ
D dτ ρ C + ik exp 1 .
2π dτ M2
−∞ 0
Rescaling k by M produces
Z Z∞ " !
dk d
(∆S)reg = D dτ Mρ−1 C + iMk ×
2π dτ
−∞
! ! #
k2 k d d 1 d −1 d
exp −i ρ−1 + ρ−1 − ρ 1 . (C.3)
ρ M dτ dτ M2 dτ dτ 0
Eq.(C.3) is in the form of Eq.(C.1) where R0 = k 2 /ρ. We use the Dyson-like expansion
in Eq.(C.1) and pick out the M-independent term to arrive at
(∆S)reg =
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 167
! ! !
Z Z∞ d Z1
dk −1 k2 −1 d d −1 k2
−iD dτ kρ C ds exp (1 − s) ρ + ρ exp s
2π dτ ρ dτ dτ ρ
−∞ 0
Z1 ! ! !
k2 d −1 d k2
+ ds exp (1 − s) ρ exp s
ρ dτ dτ ρ
0
Z1 Zu ! !
k2 −1 d d
+k 2 du ds exp (1 − u) ρ + ρ−1 ×
ρ dτ dτ
0 0
! ! !)
k2 d d k2
exp (u − s) ρ−1 + ρ−1 exp s . (C.4)
ρ dτ dτ ρ
The first term in Eq.(C.4) is zero because it is a total derivative. To calculate the
other two terms rotate to Euclidean space using k → −ikE . Then the expression
k 2 /ρ → −kE2 /ρ in the exponents yields gaussian damping factors, so that the integrals
are convergent. Even before evaluating the derivatives dτd , it is clear that the integrand
is an odd function of k and hence produces a zero integral.
For the chiral Schwinger model, one starts with Eq.(9.41). Using momentum-space
wave functions, the functional trace is
Z
∆S = −i d2 x C×
"Z #
d2 k R+ R−
2 exp (−ik · x) exp 2
− exp exp (ik · x) , (C.5)
(2π) M M2 0
Equation (C.2) is used to eliminate the exp (±ik · x) factors. Then, one scales the
momentum k by M. The expression for ∆S becomes
Z Z
2 d2 k h 2 i
∆S = −i d xC M exp R̃+ − exp R̃− 1 , (C.6)
(2π)2 0
where
−2ik µ ∂µ ± A− k+ ∂ µ ∂µ ± i (∂+ A− ) ± iA− ∂+
R̃± = −k µ kµ − + . (C.7)
M M2
The parenthesis around (∂+ A− ) indicates that ∂+ acts only on A− .
The notation []0 selects the term in Eq.(C.6) independent of M. Hence, one
expands in the factors in the exponentials proportional to M−1 and M−2 . This
results in two terms, ∆S1 from the leading Taylor series term in M−2 , and ∆S2 from
the second order term in M−1 : ∆S = ∆S1 + ∆S2 , where
Z Z
2 d2 k µ
∆S1 = 2 d x C (∂+ A− ) 2 exp (−k kµ ) ,
(2π)
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 168
Z Z
d2 k
∆S2 = i d2 x C exp (−k µ kµ ) 2ik µ (∂µ A− ) k+ . (C.8)
(2π)2
Integrals are defined by analytic continuation using Wick rotation, that is, the k0
integral is performed by going to Euclidean space. The line integral contained in
R∞
dk0 can be rotated counterclockwise by 90o . Then, one sets k0 = −ik0E :
−∞
Z∞ Zi∞ −∞
Z Z∞
dk0 = dk0 = −i dk0E =i dk0E . (C.9)
−∞ −i∞ ∞ −∞
Hence,
2i Z 2
d x C (∂+ A− ) ,
∆S1 =
4π
Z
i
∆S2 = − d2 x C (∂+ A− ) . (C.12)
4π
Adding the two contributions,
i Z 2
∆S = d x C (∂+ A− ) . (C.13)
4π
The factor ∂+ A− can be written in Lorentz covariant form using
A ≡ ρ−(r+1)/2 (/
∂ ρrk/ + k/ρr∂/) ρ−(r+1)/2 , (C.16)
B ≡ −ρ−(r+1)/2∂/ρr∂/ρ−(r+1)/2 . (C.17)
For the next step, we use the Dyson-like expansion in Eq.(C.1). Recalling that
[ ]0 indicates that the M-independent term is to be selected, κr becomes a sum of
two terms
κr = B-term + AA-term ,
where
Z Z1 Zu
tr d2 k
AA-term = − du ds ×
2 (2π)2 0 0
" # " # " #
k2 k2 k2
exp − (1 − u) A exp − (u − s) A exp −s 1 ,
ρ ρ ρ
and " # " #
Z Z1
tr d2 k k2 k2
B-term = ds exp − (1 − s) B exp −s 1 .
2 (2π)2 0
ρ ρ
In AA-term, carry out the differentiations ∂/ in both A operators using the defini-
tion of A in Eq.(C.16) to obtain
Z Z1 Zu (
d2 k h i ∂ n ∂n ρ 2
AA-term = − du ds k 2 exp −k 2 2sk − 1 +
(2π)2 0 0
ρ
)
∂ n ρ∂n ρ 1 2
− r − 5 − 7sk 2 − uk 2 + 2suk 4 ,
ρ2 2
where another rescaling k → kρ1/2 has been performed. Next, rotate the k integration
∞
R
from Minkowski space to Euclidean space. The line integral contained in dkτ can
−∞
be rotated counterclockwise by 90o . One sets kτ = −ikτE and uses Eq.(C.9) for
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 170
k0 =h kτ . The iexponential factor exp [−k 2 ] = exp [+kτ2 − kσ2 ] in AA-term becomes
exp −kτE2 − kσ2 . One can then do the s, u and k integrations since the latter are
now convergent. The result is
!!
1 ∂ n ∂n ρ 1 ∂ n ρ∂n ρ r2 + 1
AA-term = − . (C.18)
4πi ρ 6 ρ2 4
The B-term is treated similarly. One carries out the differentiations ∂/ in the B
operator and rescales k by ρ1/2 to arrive at
1 (
tr Z d2 k Z h i ∂n∂ ρ r+1
2 n 2
B-term = ds exp −k sk − +
2 (2π)2 0 ρ 2
!)
∂ n ρ∂n ρ (3 − r) (r + 1)
− 3sk 2 + s2 k 4 .
ρ2 4
After rotating to Euclidean space, all integrations can be formed. The B-term is
!!
1 ∂ n ∂n ρ r ∂ n ρ∂n ρ r2 r 1
B-term = + − + . (C.19)
4πi ρ 2 ρ2 4 2 12
The sum of the AA- and B-terms in Eqs.(C.18) and (C.19) is the quoted result for
κr in Eq.(9.79).
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 171
References
[1] E. S. Abers and B. W. Lee, Gauge Theories, Phys. Rep. C9 (1973) 1.
[11] G. Barnich, F. Brandt and M. Henneaux, Local BRST Cohomology in the An-
tifield Formalism: I. General Theorems, preprint ULB-th-94/06, NIKHEF-H
94-13, hep-th/9405109.
[15] F. Bastianelli, Ward Identities for Quantum Field Theories with External Fields,
Phys. Lett. B263 (1991) 411.
[24] I. A. Batalin and G. A. Vilkovisky, Gauge Algebra and Quantization, Phys. Lett.
102B (1981) 27.
[25] I. A. Batalin and G. A. Vilkovisky, Feynman Rules for Reducible Gauge Theories,
Phys. Lett. 120B (1983) 166.
[30] C. Batlle and J. Gomis, Lagrangian and Hamiltonian BRST Structures of the
Antisymmetric Tensor Gauge Theory, Phys. Rev. D38 (1988) 1169.
[31] C. Batlle, J. Gomis, J. Parı́s and J. Roca, Lagrangian and Hamiltonian BRST
Formalisms, Phys. Lett. B224 (1989) 288;
Field-Antifield Formalism and Hamiltonian BRST Approach, Nucl. Phys.
B329 (1990) 139.
[34] L. Baulieu, E. Bergshoeff and E. Sezgin, Open BRST Algebras, Ghost Unification
and String Field Theory, Nucl. Phys. B307 (1988) 348.
[36] C. Becchi, A. Rouet and R. Stora, The Abelian Higgs Kibble Model, Unitarity of
the S-Operator, Phys. Lett. 52B (1974) 344;
Renormalization of the Abelian Higgs-Kibble Model, Comm. Math. Phys. 42
(1975) 127;
Renormalization of Gauge Theories, Ann. Phys. 98 (1976) 287.
[38] F. A. Berezin, The Method of Second Quantization, (Academic Press, New York,
1966).
[40] F. A. Berezin and D. A. Leites, Supermanifolds, Sov. Math. Dokl. 16 (1975) 1218.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 174
[44] E. Bergshoeff, R. Kallosh and A. Van Proeyen, Superparticle Actions and Gauge
Fixings, Clas. Quantum Grav. 9 (1992) 321.
[45] E. Bergshoeff, A. Sevrin and X. Shen, A Derivation of the BRST Operator for
Non-Critical W -Strings, Phys. Lett. B296 (1992) 95.
[46] N. Berkovits, M. T. Hatsuda and W. Siegel, The Big Picture, Nucl. Phys. B371
(1992) 434.
[53] K. Bleuler, A New Method for the Treatment of Longitudinal and Scalar Pho-
tons, Helv. Phys. Acta 23 (1950) 567.
[54] R. Bluhm and S. Samuel, The Off-Shell Koba-Nielsen Formula, Nucl. Phys.
B323 (1989) 337.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 175
[55] R. Bluhm and S. Samuel, Off-Shell Conformal Field Theory at the One-Loop
Level, Nucl. Phys. B325 (1989) 275.
[56] M. Bochicchio, String Field Theory in the Siegel Gauge, Phys. Lett. 188B
(1987) 330;
Gauge Fixing for the Field Theory of the Bosonic String, Phys. Lett. 193B
(1987) 31.
[58] A. Borel, Sur la Cohomologie des Espaces Fibres Principaux et Espaces Homo-
genes de Groupes de Lie Compacts, Ann. Math. 57 (1953) 115.
[59] P. Bouwknegt, J. McCarthy and K. Pilch, BRST Analysis of Physical States for
2-d Gravity Coupled to c ≤ 1 Matter, Comm. Math. Phys. 145 (1992) 541.
[64] F. Brandt, Antifield Dependence of Anomalies, Phys. Lett. B320 (1994) 57.
[70] G. Curci and R. Ferrari, An Alternative Approach to the Proof of Unitarity for
Gauge Theories, Nuovo Cim. 35A(1976) 273.
[72] O. F. Dayi, BV and BFV Formulation of a Gauge Theory of Quadratic Lie Al-
gebras in 2-D and a Construction of W3 Topological Gravity, Tubitak Research
Institute preprint, (Jan., 1994), hep-th/9401148.
[74] J. de Boer and J. Goeree, The Effective Action of W3 to All Orders, Nucl. Phys.
B401 (1993) 348.
[77] F. De Jonghe, R. Siebelink and W. Troost, Hiding Anomalies, Phys. Lett. B306
(1993) 295.
[81] B. de Wit and J. W. van Holten, Covariant Quantization of Gauge Theories with
Open Algebra, Phys. Lett. B79 (1979) 389.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 177
[82] B. DeWitt, Dynamical Theory of Groups and Fields, in Relativity, Groups and
Topology, Les Houches Summer School, Session XIII, edited by C. DeWitt and
B. DeWitt, (Gordon Breach, 1964).
[83] B. DeWitt, Quantum Theory of Gravity. I. The Canonical Theory, Phys. Rev.
160 (1967) 1113;
Quantum Theory of Gravity. II. The Manifestly Covariant Theory, Phys. Rev.
162 (1967) 1195;
Quantum Theory of Gravity. III. Applications of the Covariant Theory,
Phys. Rev. 162 (1967) 1239.
[86] A. H. Diaz, The Nonabelian Antisymmetric Tensor Field Revisited, Phys. Lett.
B203 (1988) 408.
[96] T. Eguchi, P. B. Gilkey and A. J. Hanson, Gravitation, Gauge Theories and Dif-
ferential Geometry, Phys. Rep. 66 (1980) 213.
[97] L. D. Faddeev, Operator Anomaly for the Gauss Law, Phys. Lett. 145B (1984)
81.
[98] L. D. Faddeev and V. N. Popov, Feynman Diagrams for the Yang-Mills Field,
Phys. Lett. 25B (1967) 29.
[101] R. P. Feynman, Quantum Theory of Gravitation, Acta Phys. Pol. 24 (1963) 697.
[105] J. M. L. Fisch and M. Henneaux, Homological Perturbation Theory and the Al-
gebraic Structure of the Antifield-Antibracket Formalism for Gauge Theories,
Comm. Math. Phys. 128 (1990) 627.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 179
[118] S. A. Frolov and A. A. Slavnov, Construction of the Effective Action for General
Gauge Theories Via Unitarity, Nucl. Phys. B347 (1990) 333.
[121] K. Fujikawa, Path Integral of Relativistic Strings, Phys. Rev. D25 (1982) 2584.
[122] K. Fujikawa, T. Inagaki and H. Suzuki, BRS Current and Related Anomalies in
Two Dimensional Gravity and String Theories, Nucl. Phys. B332 (1990) 499.
[124] S. J. Gates Jr., M. T. Grisaru, M. Roček and W. Siegel, Superspace or One Thou-
sand and One Lessons in Supersymmetry, (Benjamin/Cummings, Reading, MA,
1983).
[128] S. Giddings, The Veneziano Amplitude from Interacting String Field Theory,
Nucl. Phys. B278 (1986) 242
[130] J. Gomis and J. Parı́s, Unitarity and the Field-Antifield Formalism, Nucl. Phys.
B368 (1992) 311.
[131] J. Gomis and J. Parı́s, Field-Antifield Formalism for Anomalous Gauge Theo-
ries, Nucl. Phys. B395 (1993) 288.
[132] J. Gomis and J. Parı́s, Perturbation Theory and Locality in the Field-Antifield
Formalism, J. Math. Phys. 34 (1993) 2132.
[134] J. Gomis and J. Parı́s, Anomalous Gauge Theories Within the Field-Antifield
Formalism, in Proc. of the International EPS Conference on High Energy
Physics, HEP93, Marseille, (July, 1993), p. 101-102.
[135] J. Gomis, J. Parı́s, K. Rafanelli and J. Roca, BRST and Anti-BRST Symmetries
for the Spinning Particle, Phys. Lett. B246 (1990) 435.
[136] J. Gomis, J. Parı́s and J. Roca, BRST Structures of the Spinning String,
Nucl. Phys. B339 (1990) 711.
[137] J. Gomis and J. Roca, The Anti-BRST Symmetry in the Field-Antifield Formal-
ism, Nucl. Phys. B343 (1990) 152.
[142] D. Gross and A. Jevicki, Operator Formulation of Interacting String Field The-
ory (I), Nucl. Phys. B283 (1987) 1.
[143] D. Gross and A. Jevicki, Operator Formulation of Interacting String Field The-
ory (II), Nucl. Phys. B287 (1987) 225.
[146] H. Hata, Construction of the Quantum Action for Path Integral Quantization
of String Field Theory, Nucl. Phys. B339 (1990) 663.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 182
[147] H. Hata, “Theory of Theories” Approach to String Theory, Phys. Rev. D50
(1994) 4079.
[148] H. Hata, T. Kugo and N. Ohta, Skew-Symmetric Tensor Gauge Field Theory
Dynamically Realized in the QCD U(1) Channel, Nucl. Phys. B178 (1981)
527.
[150] M. Henneaux, Hamiltonian Form of the Path Integral for Theories with a Gauge
Freedom, Phys. Rep. 126 (1985) 1.
[154] M. Henneaux, Spacetime Locality of the BRST Formalism, Comm. Math. Phys.
140 (1991) 1.
[162] C. M. Hull, W Gravity Anomalies with Ghost Loops and Background Charges,
Int. J. Mod. Phys. A8 (1993) 2419.
[164] H. Ikemori, Extended Form Method of Antifield BRST Formalism for Topological
Quantum Field Theories, Clas. Quantum Grav. 10, (1993) 233.
[165] C. Itzykson and J. B. Zuber, Quantum Field Theory, (McGraw-Hill, New York,
1980).
[169] M. Kaku, Geometrical Derivation of String Field Theory from First Principles:
Closed Strings and Modular Invariance, Phys. Rev. D38 (1988) 3052;
Nonpolynomial Closed-String Field Theory, Phys. Rev. D41 (1990) 3734.
[170] M. Kaku and J. Lykken, Modular Invariant Closed String Field Theory,
Phys. Rev. D38 (1988) 3067.
[171] M. Kalb and P. Ramond, Classical Direct Interstring Action, Phys. Rev. D9
(1974) 2273.
[172] R. E. Kallosh, Modified Rules in Supergravity, Nucl. Phys. B141 (1978) 141.
[174] M. Kato and K. Ogawa, Covariant Quantization of String Based on BRS In-
variance, Nucl. Phys. B212 (1983) 443.
[177] T. W. Kibble, Lorentz Invariance and the Gravitational Field, J. Math. Phys. 2
(1961) 212.
[183] T. Kugo, and K. Suehiro, Nonpolynomial Closed String Field Theory: Action
and Gauge Invariance, Nucl. Phys. B337 (1990) 343.
[184] T. Kugo and S. Uehara, General Procedure of Gauge Fixing Based on BRS In-
variance Principle, Nucl. Phys. B197 (1982) 378.
[187] B. W. Lee, Gauge Theories, in Methods in Field Theory, Les Houches Summer
School 1975, edited by R. Balian and J. Zinn-Justin (North-Holland, Amster-
dam, 1976).
[192] J. Lykken and S. Raby, Non-Commutative Geometry and the Closed Bosonic
String, Nucl. Phys. B278 (1986) 256.
[194] R. Marnelius, General State Spaces for BRST Quantizations, Nucl. Phys. B391
(1993) 621.
[199] N. K. Nielsen, Ghost Counting in Supergravity, Nucl. Phys. B140 (1978) 494.
[201] Kh. S. Nirov and A. V. Razumov, Field-Antifield and BFV Formalisms for
Quadratic Systems with Open Gauge Algebras, Int. J. Mod. Phys. A7 (1992)
5719.
J. Gomis, J. Parı́s and S. Samuel — Antibracket, Antifields and . . . 186
[202] I. Ojima, Another BRS Transformation, Prog. Theo. Phys. 64 (1981) 625.
[203] C. Ordóñez, J. Parı́s, J. M. Pons and R. Toldrà, BV Analysis for Covariant and
Noncovariant Actions, Phys. Rev. D48 (1993) 3818.
[205] F. R. Ore, Jr. and P. van Nieuwenhuizen, Local BRST Symmetry and the Ge-
ometry of Gauge-Fixing, Nucl. Phys. B204 (1982) 317.
[207] J. Parı́s, Faddeev-Popov Method for Anomalous Quasigroups, Phys. Lett. B300
(1993) 104.
[208] J. E. Paton and H.-M. Chan, Generalized Veneziano Model with Isospin,
Nucl. Phys. B10 (1969) 516.
[214] S. Samuel, The Physical and Ghost Vertices in Witten’s String Field Theory,
Phys. Lett. B181 (1986) 255.
[216] S. Samuel, Covariant Off-Shell String Amplitudes, Nucl. Phys. B308 (1988)
285;
Off-Shell String Physics from Conformal Field Theory, Nucl. Phys. B308
(1988) 317.
[217] S. Samuel, Solving the Open Bosonic String in Perturbation Theory, Nucl. Phys.
B341 (1990) 513.
[223] A. Sen, On the Background Independence of String Field Theory, Nucl. Phys.
B345 (1990) 551;
On the Background Independence of String Field Theory II. Analysis of On-
shell S-Matrix Elements, Nucl. Phys. B347 (1990) 270;
On the Background Independence of String Field Theory III. Explicit Field
Redefinitions, Nucl. Phys. B391 (1993) 550.
[229] W. Siegel, Covariantly Second-Quantized String, Phys. Lett. 142B (1984) 276;
Covariantly Second-Quantized String II, Phys. Lett. 151B (1985) 391;
Covariantly Second-Quantized String III, Phys. Lett. 151B (1985) 396.
[234] A. A. Slavnov, Physical Unitarity in the BRST Approach, Phys. Lett. B217
(1989) 91.
[237] J. Stasheff, Closed String Field Theory, Strong Homotopy Lie Algebras and the
Operad Actions of Moduli Space, Univ. of North Carolina preprint UNC-MATH-
93-1, (April, 1993), hep-th/9304061.
[239] Y. Takahashi, On the Generalized Ward Identity, Nuovo Cim. 6 (1957) 371.
[240] J. Tate, Homology of Noetherian Rings and Local Rings, Illinois J. Math. 1
(1957) 14.
[246] C. B. Thorn, Perturbation Theory for Quantized String Fields, Nucl. Phys.
B287 (1987) 61.
[250] P. K. Townsend, Gauge Invariance for Spin 1/2, Phys. Lett. 90B (1980) 275.
[251] W. Troost, P. van Nieuwenhuizen and A. Van Proeyen, Anomalies and the
Batalin-Vilkovisky Lagrangian Formalism, Nucl. Phys. B333 (1990) 727.
[254] I. V. Tyutin, Gauge Invariance in Field Theory and Statistical Mechanics, Lebe-
dev preprint no 39 (1975), unpublished.
[255] I. V. Tyutin and Sh. Shvartsman, BRST Operator and Open Algebra, Phys. Lett.
B169 (1986) 225.
[261] E. Verlinde, The Master Equation of String Theory, Nucl. Phys. B381 (1992)
141.
[269] P. L. White, Anomaly Consistency Conditions to All Orders, Phys. Lett. B284
(1992) 55.
[270] E. Witten, Non-Commutative Geometry and String Field Theory, Nucl. Phys.
B268 (1986) 253.
[272] E. Witten, Topological Quantum Field Theory, Comm. Math. Phys. 117 (1988)
353.
[273] E. Witten, A Note on the Antibracket Formalism, Mod. Phys. Lett. A5 (1990)
487.
[274] E. Witten, On Background Independent Open String Field Theory, Phys. Rev.
D46 (1992) 5467.
[276] C. N. Yang and R. L. Mills, Considerations of Isotopic Spin and Isotopic Gauge
Invariance, Phys. Rev. 96 (1954) 191.
[278] J. Zinn-Justin, Quantum Field Theory and Critical Phenomena, (Oxford Uni-
versity Press, Oxford, England, 1989).
[280] B. Zwiebach, Quantum Closed Strings from Minimal Area, Int. J. Mod. Phys.
A5 (1990) 275;
Consistency of Closed String Polyhedra from Minimal Area, Phys. Lett. B241
(1990) 343;
How Covariant Closed String Theory Solves a Minimal Area Problem,
Comm. Math. Phys. 136 (1991) 83.
[281] B. Zwiebach, Closed String Field Theory: An Introduction, MIT preprint MIT-
CTP-2206, (May, 1993), hep-th/9305026.
[282] B. Zwiebach, Closed String Field Theory: Quantum Action and the B-V Master
Equation, Nucl. Phys. B390 (1993) 33.