arxiv:hep-th/ v1 14 Jan 1997

Similar documents
arxiv:hep-th/ v2 8 Jun 2000

BFT quantization of chiral-boson theories

arxiv:hep-th/ v2 15 Jul 1999

The Dirac Propagator From Pseudoclassical Mechanics

arxiv:hep-th/ v2 13 Sep 2001

arxiv:hep-th/ v1 1 Dec 1998

Improved BFT embedding having chain-structure arxiv:hep-th/ v1 3 Aug 2005

arxiv:hep-th/ v2 11 May 1998

arxiv:hep-th/ v1 21 Jan 1997

arxiv:hep-th/ v1 13 Feb 1992

arxiv:hep-th/ v1 7 Nov 1998

arxiv:quant-ph/ v2 28 Nov 2000

arxiv:gr-qc/ v2 6 Apr 1999

Snyder noncommutative space-time from two-time physics

Symmetry transform in Faddeev-Jackiw quantization of dual models

BFT embedding of noncommutative D-brane system. Abstract

Duality between constraints and gauge conditions

arxiv:hep-th/ v2 11 Jan 1999

arxiv:hep-th/ v2 6 Mar 2000

arxiv: v2 [hep-th] 4 Sep 2009

Quantum Field Theory I Examination questions will be composed from those below and from questions in the textbook and previous exams

Path Integral Quantization of the Electromagnetic Field Coupled to A Spinor

arxiv:hep-th/ v2 11 Sep 1996

arxiv:hep-th/ v1 2 Oct 1998

arxiv:hep-th/ v1 10 Apr 2006

Topological DBI actions and nonlinear instantons

A Lax Representation for the Born-Infeld Equation

arxiv:hep-th/ v1 23 Mar 1998

arxiv: v1 [hep-th] 23 Mar 2015

Topologically Massive Yang-Mills field on the Null-Plane: A Hamilton-Jacobi approach

Improved BFT quantization of O(3) nonlinear sigma model

GRANGIAN QUANTIZATION OF THE HETEROTIC STRING IN THE BOSONIC FORMULAT

arxiv:hep-th/ v1 1 Mar 2000

arxiv:hep-th/ v1 28 Jan 1999

arxiv:hep-th/ v1 17 Jan 2007

HIGHER SPIN PROBLEM IN FIELD THEORY

arxiv:hep-th/ v1 16 Aug 1996

Generalization of the Hamilton-Jacobi approach for higher order singular systems

Path Integral Quantization of Constrained Systems

138. Last Latexed: April 25, 2017 at 9:45 Joel A. Shapiro. so 2. iψ α j x j

arxiv:gr-qc/ v1 11 Nov 1996

Hamilton-Jacobi Formulation of A Non-Abelian Yang-Mills Theories

LAGRANGIANS FOR CHIRAL BOSONS AND THE HETEROTIC STRING. J. M. F. LABASTIDA and M. PERNICI The Institute for Advanced Study Princeton, NJ 08540, USA

Quantization of scalar fields

Hamiltonian Embedding of SU(2) Higgs Model in the Unitary Gauge

Faddeev-Yakubovsky Technique for Weakly Bound Systems

Department of Physics and Basic Science Research Institute, Sogang University, C.P.O. Box 1142, Seoul , Korea. (November 26, 2001) Abstract

SUSY QM VIA 2x2 MATRIX SUPERPOTENTIAL

Remarks on Gauge Fixing and BRST Quantization of Noncommutative Gauge Theories

arxiv:hep-ph/ v1 19 Mar 1998

d 2 Area i K i0 ν 0 (S.2) when the integral is taken over the whole space, hence the second eq. (1.12).

Light-Cone Quantization of Electrodynamics

Maxwell s equations. based on S-54. electric field charge density. current density

Supergravity in Quantum Mechanics

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

1 Quantum fields in Minkowski spacetime

arxiv:gr-qc/ v1 11 Oct 1999

Coordinate/Field Duality in Gauge Theories: Emergence of Matrix Coordinates

arxiv:hep-th/ v1 16 Jun 1999

arxiv:math-ph/ v1 10 Nov 2003

3 Quantization of the Dirac equation

arxiv:hep-ph/ v1 16 Mar 2004 GLUEBALL-GLUEBALL INTERACTION IN THE CONTEXT OF AN EFFECTIVE THEORY

arxiv:gr-qc/ v1 22 Jul 1994

Solving Systems of Linear Equations Symbolically

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

1 Canonical quantization conformal gauge

ON ABELIANIZATION OF FIRST CLASS CONSTRAINTS

A Lattice Approximation of Dirac Equation

b quark Electric Dipole moment in the general two Higgs Doublet and three Higgs Doublet models

Attempts at relativistic QM

arxiv:hep-ph/ v1 30 Oct 2005

Hamilton-Jacobi Formulation of Supermembrane

Representation theory of SU(2), density operators, purification Michael Walter, University of Amsterdam

arxiv:quant-ph/ v1 5 Mar 2001

Lorentz-covariant spectrum of single-particle states and their field theory Physics 230A, Spring 2007, Hitoshi Murayama

Emergence of Yang Mills theory from the Non-Abelian Nambu Model

Star operation in Quantum Mechanics. Abstract

Some Quantum Aspects of D=3 Space-Time Massive Gravity.

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

SECOND-ORDER LAGRANGIAN FORMULATION OF LINEAR FIRST-ORDER FIELD EQUATIONS

arxiv:hep-th/ v2 29 Aug 2003

Feynman rules for fermion-number-violating interactions

Constrained Dynamical Systems and Their Quantization

Helicity conservation in Born-Infeld theory

Testing a Fourier Accelerated Hybrid Monte Carlo Algorithm

The Mandelstam Leibbrandt prescription and the Discretized Light Front Quantization.

arxiv:hep-th/ v1 2 Jul 2003

arxiv: v1 [hep-th] 30 Jan 2009

Bloch Wilson Hamiltonian and a Generalization of the Gell-Mann Low Theorem 1

15.2. Supersymmetric Gauge Theories

Lagrangian. µ = 0 0 E x E y E z 1 E x 0 B z B y 2 E y B z 0 B x 3 E z B y B x 0. field tensor. ν =

Introduction to string theory 2 - Quantization

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

arxiv: v2 [hep-th] 14 Aug 2008

arxiv: v1 [math.ds] 18 Nov 2008

1 Systems of Differential Equations

arxiv:hep-th/ v1 16 Jun 1993

The Hamiltonian operator and states

Quantum Field Theory Example Sheet 4 Michelmas Term 2011

The descent of off-shell supersymmetry to the mass shell

Transcription:

Transformation of second-class into first-class constraints in supersymmetric theories arxiv:hep-th/9701058v1 14 Jan 1997 J. Barcelos-Neto and W. Oliveira Instituto de Física Universidade Federal do Rio de Janeiro RJ 21945-970 - Caixa Postal 68528 - Brasil Astract We use the method due to Batalin, Fradkin, Fradkina, and Tyutin (BFFT) in order to convert second-class into first-class constraints for some quantum mechanics supersymmetric theories. The main point to e considered is that the extended theory, where new auxiliary variales are introduced, has to e supersymmetric too. This leads to some additional restrictions with respect the conventional use of the BFFT formalism. PACS: 03.65.-w, 11.10.Ef, 11.30.P e-mail: arcelos@if.ufrj.r e-mail: wilson@fisica.ufjf.r Permanent address: UFJF - Departamento de Física - Juiz de Fora, MG 36036-330 1

1 Introduction The method developed in series of papers y Batalin, Fradkin, Fradkina, and Tyutin (BFFT) [1, 2] has as main purpose the transformation of second-class into firstclass constraints [3]. This is achieved with the aid of auxiliary fields that extend the phase space in a convenient way. After that, one has a gauge-invariant system which matches the original theory when the so-called unitary gauge is chosen. The BFFT method is quite elegant and operates systematically. One application we could immediately envisage for it is the covariant quantization of systems with the Green-Schwarz (GS) supersymmetry [4]. In these systems, the existence of the κ-symmetry [5] inhiits a covariant separation of first and second-class constraints. The use of the method there would transform the second-class constraints into firstclass and one could operate with them in the conventional way. However, this might not e an easy and simple task as one may confirm y looking at the works of covariant quantization of superstrings [6]. We intend that there are some prolems that should e solved first. One of the crucial question is how the BFFT method can operate consistently in supersymmetric theories. This is a pertinent question ecause when the phase space is extended y introducing auxiliary variales a new theory is otained, and this must also e supersymmetric. The purpose of the present paper is to address the BFFT method to this particular prolem related to supersymmetric theories. We show that the freedom we have in choosing some parameters in the application of the method can e conveniently used in order that a final supersymmetric theory should e verified. We consider two examples in supersymmetric quantum mechanics and we show that the method is only consistently applied when an infinite numer of first-class constraints is introduced. Our paper is organized as follows. In Sec. 2 we make a rief review of the BFFT method in order to emphasize and clarify some of its particularities that we shall use in the forthcoming sections. In Sec. 3 we consider one of the simplest supersymmetric quantum mechanical model, where the supersymmetric space is descried y just one time and one Grassmannian variale. This system does not permit us to introduce interaction and has to e analyzed with osons and fermions disconnected. In Sec. 4 we consider a more realistic model where it is possile to have interactions etween osons and fermions. This system is descried in a space with one time and two Grassmannian variales. Finally, Sec. 5 contains some concluding remarks and we also introduce an appendix in order to explain some details concerning the choice of the extended space. 2

2 Brief review of the BFFT formalism Let us consider a system descried y a Hamiltonian H 0 in a phase-space (q i,p i ) with i = 1,...,N. Here we suppose that the coordinates are osonic (extension to include fermionic degrees of freedom and to the continuous case can e done in a straightforward way). It is also supposed that there just exist second-class constraints (at the end of this section we refer to the case where first-class constraints are also present). Denoting them y T a, with a = 1,...,M < 2N, we have {T a, T } = a, (2.1) where det( a ) 0. The general purpose of the BFFT formalism is to convert second-class constraints into first-class ones. This is achieved y introducing canonical variales, one for each second-class constraint (the connection etween the numer of secondclass constraints and the new variales in a one-to-one correlation is to keep the same numer of the physical degrees of freedom in the resulting extended theory). We denote these auxiliary variales y y a and assume that they have the following general structure {y a, y } = ω a, (2.2) where ω a is a constant quantity with det (ω a ) 0. The otainment of ω a is discussed in what follows. It is emodied in the calculation of the resulting firstclass constraints that we denote y T a. Of course, these depend on the new variales y a, namely T a = T a (q,p;y), (2.3) and satisfy the oundary condition T a (q,p;0) = T a (q,p). (2.4) Another characteristic of these new constraints is that they are assumed to e strongly involutive, i.e. { T a, T } = 0. (2.5) The solution of (2.5) can e achieved y considering T a expanded as T a = n=0 T (n) a, (2.6) 3

where T (n) a is a term of order n in y. Compatiility with the oundary condition (2.4) requires that T (0) a = T a. (2.7) The replacement of (2.6) into (2.5) leads to a set of equations, one for each coefficient of y n. We list some of them elow {T a (0) (0),T } (q,p) + {T a (1),T (1) } (y) = 0, (2.8) {T (0) a,t (1) } (q,p) + {T (1) a,t (0) } (q,p) + {T (1) a,t (2) } (y) + {T (2) a,t (1) } (y) = 0,(2.9) {T (0) a,t (2) } (q,p) + {T (1) a,t (1) } (q,p) + {T (2) a,t (0) } (q,p) + {T (1) + {T a (2) (2),T. } (y) + {T a (3) a,t (3) } (y),t (1) } (y) = 0, (2.10) These correspond to coefficients of the powers y 0, y 1, y 2,..., respectively. The notation used aove, {, } (q,p) and {, } (y), represents the parts of the Poisson racket {, } relative to the variales (q,p) and (y). Equations aove are used iteratively in the otainment of the corrections T (n) (n 1). Equation (2.8) shall give T (1). With this result and (2.9), one calculates T (2), and so on. Since T (1) is linear in y we may write T (1) a = X a (q,p)y. (2.11) Introducing this expression into (2.8) and using the oundary condition (2.4), as well as (2.1) and (2.2), we get a + X ac ω cd X d = 0. (2.12) We notice that this equation does not give X a univocally, ecause it also contains the still unknown ω a. What we usually do is to choose ω a in such a way that the new variales are unconstrained. It is opportune to mention that it is not always possile to make a choice like this, that is to say, just having unconstrained variales [7]. In consequence, the consistency of the method requires an introduction of other new variales in order to transform these constraints also into first-class. This may lead to an endless process. However, it is important to emphasize that ω a can e fixed anyway. After fixing ω a, we pass to consider the coefficients X a. They are also not otained in a univocally way. Even after fixing ω a, expression (2.12) leads to less equations than variales. The choice of X s has to e done in a convenient way [8]. We shall see in the following sections how to use this freedom in order to keep the supersymmetry in the final theory. 4

The knowledge of X a permits us to otain T a (1). If X a do not depend on (q,p), it is easily seen that T a + T a (1) is already strongly involutive. When this occurs, we succeed in otaining T a. If this is not so, we have to introduce T a (1) into equation (2.9) to calculate T a (2), and so on. Another point in the BFFT formalism is that any dynamic function A(q, p) (for instance, the Hamiltonian) has also to e properly modified in order to e strongly involutive with the first-class constraints T a. Denoting the modified quantity y Ã(q,p;y), we then have { T a, Ã} = 0. (2.13) In addition, à has also to satisfy the oundary condition Ã(q,p;0) = A(q,p). (2.14) The otainment of à is similar to what was done to get T a, that is to say, we consider an expansion like à = A (n), (2.15) n=0 where A (n) is also a term of order n in y s. Consequently, compatiility with (2.14) requires that A (0) = A. (2.16) The comination of (2.6), (2.13) and (2.15) gives {T (0) a,a (0) } (q,p) + {T (1) a,a (1) } (y) = 0, (2.17) {T a (0),A (1) } (q,p) + {T a (1),A (0) } (q,p) + {T a (1),A (2) } (y) + {T a (2),A (1) } (y) = 0, (2.18) {T a (0),A (2) } (q,p) + {T a (1),A (1) } (q,p) + {T a (2),A (0) } (q,p) + {T a (1),A (3) } (y) + {T (2) a,a (2) } (y) + {T (3) a,a (1) } (y) = 0, (2.19). which correspond to the coefficients of the powers y 0, y 1, y 2, etc., respectively. The expression (2.17) aove gives us A (1) A (1) = y a ω a X c {T c, A}, (2.20) where ω a and X a are the inverses of ω a and X a. 5

In the otainment of T a (1) we had seen that T a + T a (1) was strongly involutive if the coefficients X a do not depend on (q,p). However, the same argument does not necessarily apply here. Usually we have to calculate other corrections to otain the final Ã. Let us discuss how this can e systematically done. We consider the general case first. The correction A (2) comes from equation (2.18), that we conveniently rewrite as where Thus {T (1) a, A (2) } (y) = G (1) a, (2.21) G (1) a = {T a, A (1) } (q,p) + {T (1) a, A} (q,p) + {T (2) a, A (1) } (y). (2.22) A (2) = 1 2 ya ω a X c G (1) c. (2.23) In the same way, other terms can e otained. The final general expression reads where A (n+1) = 1 n + 1 ya ω a X c G (n) c, (2.24) G (n) a = n n 2 {T a (n m), A (m) } (q,p) + {T a (n m) m=0 m=0, A (m+2) } (y) + {T (n+1) a, A (1) } (y). (2.25) For the particular case when X a do not depend on (q,p) we have that the corrections A (n+1) can e otained y the same expression (2.24), ut G (n) a simplifies to G (n) a = {T a, A (n) } (q,p). (2.26) At this stage, it is opportune to mention that when X a do not depend on (q,p) and the second-class constraints are linear in the phase-space variales, there is an alternative and more direct way of calculating the modified function à [9]. Let us riefly discuss how this can e done. The comination of (2.24) and (2.26) gives A (n+1) = 1 n + 1 ya ω a X c {T c, A (n) } (q,p), = 1 n + 1 ya ω a X c {T c, Ω α } A(n) Ω α, 1 n + 1 ya K α a αa (n), (2.27) 6

where Ω α is generically representing the phase-space coordinates (q,p) and K α a ω ax c {T c, Ω α } (2.28) is a constant matrix. By using expression (2.27) iteratively, we get A (n) = ( 1)n n! ( y a K α a α ) n A. (2.29) Introducing (2.29) into (2.15), we otain à in terms of A, namely Ã(Ω, y) = ( 1) n n=0 = exp ( y a K α a α) n A, n! ) ( y a Ka α α A. (2.30) Since exp ( y a K α a α) is a translation operator in the coordinate Ω, we have à (Ω α,y a ) = A(Ω α y a K α a ) (2.31) So, when X a do not depend on the phase-space coordinates y α and the constraints T a are linear in Ω α, the dynamic quantity à can e directly otained from A y replacing Ω α y a shifted coordinate Ω α = Ω α y a Ka α. (2.32) To conclude this rief report on the BFFT formalism, we refer to the case where there are also first-class constrains in the theory. Let us call them F α. We consider that the constraints of the theory satisfy the following involutive algera (with the use of the Dirac racket definition to strongly eliminate the second-class constraints) {F α, F β } D = U γ αβ F γ + I a αβ T a, {H 0, F α } D = V β α F β + K a α T a, {F α, T } D = 0 (2.33) In the expressions aove, U γ αβ, Ia αβ, V β α and K a α are structure functions of the involutive algera. The BFFT procedure in this case also introduces one auxiliary variale for each one of the second-class constraints (this is also in agreement with the counting of the physical degrees of freedom of the initial theory). All the constraints and the Hamiltonian have to e properly modified in order to satisfy the same involutive algera aove, namely, 7

{ F α, F β } = U γ αβ F γ + I a αβ T a, { H 0, Fα } = V β α F β + K a α T a, { F α, T } = 0. (2.34) Since the algera is now weakly involutive, the iterative calculation of the previous case cannot e applied here. We have to figure out the corrections that have to e done in the initial quantities. For details, see ref. [2]. 3 An example in N=1 supersymmetry Let us discuss here a simple model developed on a parameter space with just one Grassmannian variale. The supersymmetry transformations are defined y δt = iǫθ, δθ = ǫ, (3.1) where ǫ is a constant anticommuting parameter that characterizes the supersymmetry transformations (in this section we consider oth θ and ǫ as real quantities). The transformations aove can e expressed in terms of a supersymmetric Q- generator as follows δt = (ǫq)t, δθ = (ǫq)θ. (3.2) The comination of (3.1) and (3.2) permit us to directly infer that Q = θ + iθ t (3.3) and Q satisfies the anticommuting algera [Q,Q] + = 2i t. (3.4) In this space, a (real) supercoordinate has the general form φ i (t,θ) = q i (t) + iθ ψ i (t), (3.5) 8

where q i are osonic coordinates and ψ i are fermionic ones. The index i (i = 1,...,N) corresponds to the dimension of the configuration space. The quantity φ i given y (3.5) is actually a supercoordinates if it satisfies the supersymmetry transformation δφ i = (ǫq)φ i. (3.6) The comination of (3.3), (3.5), and (3.6) yields the transformation laws for the component fields, namely δq i = iǫψ i, δψ i = ǫ q i. (3.7) Theories that are expressed in terms of supercoordinates and operators that commute or anticommute with Q are manifestly supersymmetric invariant. An operator which anticommutes with Q is D = θ iθ t. (3.8) In order to construct supersymmetric Lagrangians, it is necessary to introduce another operator. Here, this is nothing other than the time derivative. For questions of hermicity, we conveniently write D t = i t. (3.9) We then consider the Lagrangian L = 1 2 D tφ i Dφ i, (3.10) that in terms of components gives L = 1 2 qi ψ i + 1 2 θ ( qi q i i ψ i ψ i ). (3.11) The θ-component of the Lagrangian aove transforms as a total derivative. So, it is considered to e the Lagrangian in terms of the component coordinate, namely L = 1 2 qi q i i 2 ψ i ψ i. (3.12) In order to identify the constraints, we calculate the canonical momenta 9

p i = L q i = qi, π i = L ψ i = i 2 ψi, (3.13) where we have used left derivative for fermionic coordinates. We notice that the second relation aove is a constraint, that we write as T i = π i + i 2 ψi 0, (3.14) where means weakly equal [3]. This is the only constraint of the theory and it is second-class. Its Poisson racket relation reads = {T i, T j }, = iδ ij. (3.15) The implementation of the BFFT formalism requires the introduction of auxiliary variales ξ i (one for each second-class constraint T i ). This coordinate is fermionic ecause the corresponding constraint is also fermionic. Since there is just one kind of auxiliary coordinates, we have to assume that they are constrained. This is so in order to otain strong first-class constraints. We then consider ω ij = {ξ i,ξ j } = iδ ij. (3.16) For this particular example, we do not need to go into details in the BFFT formalism to conclude that the first class constraints T i are given y T i = π i + i 2 ψi + iξ i 0. (3.17) Further, it is also immediately seen that the Lagrangian that leads to this theory reads L = 1 2 qi q i i 2 ψ i ψ i i ξ i ξ i + iξ i ψ i. (3.18) We oserve that it will e supersymmetric invariant in the sense of the following transformations δq i = iǫ(ψ i + 2ξ i ), δψ i = ǫ q i, δξ i = ǫ q i. (3.19) 10

This is a very simple model and of course cannot tell us everything aout the BFFT formalism for supersymmetric theories. But it is important to emphasize that the otained effective theory is also constrained in the auxiliary variales. If we would like to go on with the BFFT method, we would have to introduce new auxiliary variales to take care of these new constraints. It is easily seen that these new variales would also have to e constrained and that this procedure would lead to an endless process, with an infinite numer of auxiliary variales. We shall see that this characteristic will also e manifested in the example we shall discuss in the next section. 4 An example with N=2 supersymmetry Let us here consider the same supersymmetric quantum mechanical system discussed in ref. [10]. It is developed in a space containing two Grassmannian variales, θ and θ. The supersymmetry transformations are defined y δt = i ǫθ + iǫ θ, δθ = ǫ, δ θ = ǫ, (4.1) where in this section we use the following convention for complex conjugation: θ = θ and ǫ = ǫ. There are two supersymmetric generators, Q and Q, defined y δt = (ǫq + ǫ Q) t, δθ = (ǫq)θ, δ θ = ( ǫ Q) θ. (4.2) From (4.1) and (4.2) one can directly infer Q = θ + i θ t, Q = θ + iθ t. (4.3) These operators satisfy the algera [Q, Q] + = 2i t, [Q,Q] + = 0 = [ Q, Q] +. (4.4) 11

Here, the (real) supercoordinate acquires the form φ i (t,θ, θ) = q i (t) + iθ ψ i (t) + i θ ψ i (t) + θθ d i (t), (4.5) that must satisfy the supersymmetry transformation δφ i = (ǫq + ǫ Q)φ i. (4.6) The coordinates ψ i and ψ i have the same complex conjugation as θ and θ. The comination of (4.3), (4.5) and (4.6) gives the following transformation for the component coordinates: δq i = i(ǫψ i + ǫ ψ i ), δψ i = ǫ( q i id i ), δ ψ i = ǫ( q i + id i ), δd i = ǫ ψi ǫ ψ i. (4.7) There are two operators that anticommute with Q and Q, which are D = θ i θ t, D = θ iθ t. (4.8) With these operators and the supercoordinates we may construct the following general Lagrangian density L = 1 2 Dφ i Dφ i V (φ). (4.9) The component Lagrangian is otained y taking the θθ part of the Lagrangian aove. The result is L = 1 2 qi q i + i ψ i ψ i + 1 2 di d i d i v q i + ψi ψj 2 v q i q j, (4.10) where v = V (q). The canonical momenta conjugate to q i, d i, ψ i and ψ i are given respectively y 12

p i = L q i = qi, s i = L d i = 0, π i = L ψ i = i ψ i, π i = L = 0. (4.11) ψi From these equations, one identifies the constraints T i 1 = π i + i ψ i 0, T i 2 = π i 0, T i 3 = s i 0. (4.12) Constructing the total Hamiltonian and imposing the consistency condition that constraints do not evolve in time [3] we get a new constraint T4 i = d i v 0, (4.13) qi We oserve that it is the equation of motion with respect the coordinates d i. We also mention that there are no more constraints. These all constraints are second-class. The matrix elements of their Poisson rackets read ( ij ) 12 = iδ ij, = ( ij ) 21, ( ij ) 34 = δ ij, = ( ij ) 43. (4.14) Other elements are zero. To implement the BFFT formalism, we introduce auxiliary coordinates, one for each second-class constraint. Let us generically denote them y y ai, where a = 1, 2, 3, 4. These are anticommuting coordinates for a = 1, 2 and commuting for a = 3,4. Here, contrarily to the previous example, the numer of constraints is even. So, it is natural that we try to start y supposing that the new variales are canonical 13

and that there are no constrains involving them. If we proceeded in this way, we would actually get a final theory with just first-class constraints, ut it would not e supersymmetric. In fact, we mention that it would e completely inconsistent (see appendix). Also here we have to consider that the fermionic coordinates are not canonical (ut there is no prolem concerning the osonic sector) in order to get a final theory also supersymmetric. Let us particularly choose y ai = (ξ i, ξ i ; η i, ρ i ). (4.15) We shall consider the following racket structure among these variales {ξ i, ξ j } = iδ ij, {η i,ρ j } = δ ij. (4.16) Other rackets are zero. We notice that we have considered that ρ i are the canonical momenta conjugate to η i. With these elements, the matrix ω defined at (2.2) reads [(ω ij ) a ] = 0 i 0 0 i 0 0 0 0 0 0 1 0 0 1 0 δij, (4.17) where rows and columns follow the order ξ, ξ, η, and ρ. Introducing this result into eq. (2.12), as well as the ones given y (4.14), and considering that coefficients (X ij ) a are osonic for the cominations (a = 1,2; = 1,2) and (a = 3,4; = 3,4), and fermionic for (a = 1,2; = 3,4) and (a = 3,4; = 1,2), we otain the following independent equations iδ ij + i(x ik ) 11 (X jk ) 22 + i(x ik ) 12 (X jk ) 21 (X ik ) 13 (X jk ) 24 + (X ik ) 14 (X jk ) 23 = 0, δ ij + i(x ik ) 31 (X jk ) 42 + i(x ik ) 32 (X jk ) 41 (X ik ) 33 (X jk ) 44 + (X ik ) 34 (X jk ) 43 = 0, i(x ik ) 11 (X jk ) 32 + i(x ik ) 12 (X jk ) 31 (X ik ) 13 (X jk ) 34 + (X ik ) 14 (X jk ) 33 = 0, i(x ik ) 11 (X jk ) 42 + i(x ik ) 12 (X jk ) 41 (X ik ) 13 (X jk ) 44 + (X ik ) 14 (X jk ) 43 = 0, i(x ik ) 21 (X jk ) 32 + i(x ik ) 22 (X jk ) 31 (X ik ) 23 (X jk ) 34 + (X ik ) 24 (X jk ) 33 = 0, i(x ik ) 21 (X jk ) 42 + i(x ik ) 22 (X jk ) 41 (X ik ) 23 (X jk ) 44 + (X ik ) 24 (X jk ) 43 = 0. (4.18) 14

The set of equations aove does not univocally fix (X ij ) a. There are less equations than variales. We are then free to choose some of these coefficients in a convenient way. First, we notice that it is a good simplification to consider them osonic. We thus take equal to zero the fermionic ones, i.e. (X ij ) a = 0 for (a = 1,2; = 3,4) and (a = 3,4; = 1,2). (4.19) With this choice, the set of equations aove simplifies to δ ij + (X ik ) 11 (X jk ) 22 + (X ik ) 12 (X jk ) 21 = 0, δ ij (X ik ) 33 (X jk ) 44 + (X ik ) 34 (X jk ) 43 = 0. (4.20) Looking at the expressions of the constraints T i a, we notice que a choice that might preserve the supersymmetry is (X ij ) 12 = iδ ij, (X ij ) 21 = iδ ij, (X ij ) 11 = 0 = (X ij ) 22, (X ij ) 34 = δ ij, (X ij ) 43 = δ ij, (X ij ) 33 = 0 = (X ij ) 44, (4.21) ecause this would lead to convenient translations in the old coordinates. Since it was possile to otain the quantities X s independently of the phase space coordinates, we have that T (1), given y (2.11), is the entire contriution for T. Then the set of new constraints reads T i 1 = π i + i ψ i + i ξ i, T i 2 = π i + iξ i, T i 3 = s i ρ i, T i 4 = d i v q i + ηi. (4.22) Notice that the choice of the coefficients X s were made such that the auxiliary coordinates ξ i and η i figure as translations of ψ i and d i respectively. This will make easier our goal of otaining a final supersymmetric theory. The next step is to look for the Lagrangian that leads to this new theory. A consistently way of doing this is y means of the path integral formalism, where the 15

Faddeev-Senjanovic procedure [11] has to e used. Let us then otain the canonical Hamiltonian of the initial theory. We first write down the total Hamiltonian H = q i p i + d i s i + ψ i π + ψi π i L + λ a T a. (4.23) The order of velocities and momenta (mainly for the Grassmannian coordinates) is due to the left derivative convention. Using the first equation (4.11), one eliminates de velocities q i. The remaining ones are eliminated y just redefining some Lagrange multipliers. The final result reads H = 1 2 pi p i 1 2 di d i + d i v q i + ψ i ψ j 2 v q i q j + λ i 1 (πi + i ψ i ) + λ i 2 π i + λ i 3 s i + λ i 4 (d i v qi), (4.24) where λ s are the redefined Lagrange multipliers. From the expression aove one can identify the initial canonical Hamiltonian H c = 1 2 pi p i 1 2 di d i + d i v q i + ψ i ψ j 2 v q i q j. (4.25) To calculate the modified Hamiltonian we can use relations (2.15) and (2.24), where (ω ij ) a is the inverse of (4.17) and (X ij ) a is the inverse of the matrix formed y relations (4.21). On the other hand, it is just a matter of algeraic calculation to otain the quantities G (n)i a, defined y (2.25). The result is G (0)i 1 = ψ j 2 v q i q j, G (0)i 2 = ψ j 2 v q i q j, G (0)i 3 = d i v q i, G (0)i 4 = p j 2 v q i q j, G (1)i 1 = ξ j 2 v q i q j, G (1)i 2 = ξ j 2 v q i q j, G (1)i 3 = η i, G (1)i 4 = ρ j 2 v q i q k 16 2 v q j q k. (4.26)

Using these quantities, we otain the extended canonical Hamiltonian H c = H c + H c (1) + H c (2) = 1 2 pi p i 1 2 di d i 1 2 ηi η i d i η i + (d i + η i ) v q i ( + ψi ψ j + ξ i ξ j + ξ i ψ j + ψ i ξ j + ρ i p j) 2 v q i q j + 1 2 ρi ρ j 2 v q i q k 2 v q j q k. (4.27) Since the quantities (X ij ) a do not depend on the initial phase-space coordinates, we can check the result aove y considering the extended Hamiltonian (4.27) could also have een otained from shifted coordinates defined in (2.32), i.e., H c (Ω,y) = H c ( Ω). (4.28) The calculation of the shifted coordinates gives q i = q i, p i = p i + ρ j 2 v q i q j, d i = d i + η i, s i = s i ρ i, ψ i = ψ i + ξ i, ψ i = ψ i + ξ i, π i = π i, π i = π i + iξ i. (4.29) Notice once more that the choice we have made for the X s quantities at (4.21) led to convenient shifts in the old coordinates. We shall see that this will e very important to demonstrate de supersymmetry invariance of the final theory. It is just a matter of algeraic calculation to show that the comination of (4.28) and (4.29) gives the extended canonical Hamiltonian (4.27). To use the path integral formalism to otain the Lagrangian that leads to this theory we have to pay attention for the fact that there are constraints with respect the auxiliary variales ξ i and ξ i. It is not difficult to see that a possile set of constraints is πξ i = i ξ i, π ī ξ = 0. (4.30) 17

The general expression for the propagator reads Z = N { t ( [dµ] exp i dt q i p i + d i s i + ψ i π i + ψi π i + η i ρ i + ξ i πξ i + ξi π ī ξ H )} c, (4.31) t 0 with the measure [dµ] given y [dµ] = [dq][dp][dd][ds][dψ][d ψ][dπ][d π][dξ][d ξ][dη][dρ][dπ ξ ][dπ ξ] det {, } δ [π i ξ + i ξ i ]δ [π ī ξ ] a δ [ T a ]δ [ Λ a ], (4.32) where Λ a are gauge-fixing functions corresponding to the first-class constraints T a and the term det {, } represents the determinant of all constraints of the theory, including the gauge-fixing ones. The quantity N that appears in (4.32) is the usual normalization factor. Using the delta functions, we can directly eliminate de momenta π i, π i, s i, πξ i, and πī. The result is ξ Z = N [dq][dp][dd][dψ][d ψ][dξ][d ξ][dη][dρ] det {, } δ [ T 4 ] a δ [ Λ a ] { t [ exp i dt q i p i + ( d i + η i ) ρ i 1 t 0 2 pi p i + 1 2 di d i + 1 2 ηi η i + η i d i i( ψi ψ i + ξi ξ i + ψi ξ i + ξi ψ i ) ( ψi ψ j + ξ i ψ j + ψ i ξ j + ξ i ξ j + ρ i p j) 2 v q i q j 1 2 ρi ρ j Finally, integrating over p i and ρ i we otain 2 v 2 v ]} q i q k q k q j. (4.33) Z = N [dq][dd][dψ][d ψ][dξ][d ξ][dη] det {, } δ [ d i + η i q j δ [ T 4 ] a { t δ [ Λ a ] exp i dt t 0 2 v q i q j ] [ 1 2 qi q i + 1 2 di d i + 1 2 ηi η i + η i d i (d i + η i ) v q i i( ψi ψ i + ξi ξ i + ψi ξ i + ξi ψ i ) ( ψi ψ j + ξ i ψ j + ψ i ξ j + ξ i ξ j) 2 v ]} q i q j, (4.34) where the new δ-function aove came from the integrations over ρ i. We notice that it is nothing other than the time derivative of constraint T 4. It is then just a consistency condition and does not represent any new restriction over the coordinates of the theory. 18

From the propagator (4.34), we identify the extended Lagrangian L = 1 2 qi q i + 1 2 di d i + 1 2 ηi η i + η i d i (d i + η i) v q i i( ψi ψ i + ξi ξ i + ψi ξ i + ξi ψ i ) ( ψi ψ j + ψ i ξ j + ξ i ψ j + ξ i ξ j) 2 v q i q j (4.35) The Lagrangian aove is supersymmetric invariant if we consider the transformations δq i = i[ǫ(ψ i + ξ i ) + ǫ( ψ i + ξ i )], δψ i = ǫ( q i id i ), δ ψ i = ǫ( q i + id i ), δd i = ǫ ψi ǫ ψ i, δξ i = i ǫη i, δ ξ i = iǫη i, δη i = ǫ ξi ǫ ξi. (4.36) 5 Conclusion We have used the method developed y Batalin, Fradkin, Fradkina, and Tyutin in order to quantize two supersymmetric quantum mechanical models, y transforming second-class into first-class constraints. We have shown that final supersymmetric theories are only achieved y taking constrained auxiliary variales. This procedure leads to an endless process of introducing auxiliary constrained variales and transforming them into first-class ones. This appears to e a general characteristic of the use of the BFFT method in supersymmetric theories. Although we have just considered supersymmetric quantum mechanical models, the same could e directly done for field theories. Of course, the algeraic work would e much igger, ut these would give no new information comparing what we have already found here. 19

Acknowledgment: This work is supported in part y Conselho Nacional de Desenvolvimento Científico e Tecnológico - CNPq, Financiadora de Estudos e Projetos - FINEP and Fundação Universitária José Bonifácio - FUJB (Brazilian Research Agencies). Appendix Let us discuss in this appendix the result we would otain if we had considered the two auxiliary fermionic coordinates as non constrained and taken one as the momentum conjugate of the other, say ξ i and χ i, satisfying the Poisson racket relation {ξ i,χ j } = δ ij. (A.1) The matrix [(ω ij ) a ] would e [(ω ij ) a ] = 0 1 0 0 1 0 0 0 0 0 0 1 0 0 1 0 δij. (A.2) Also considering the condition (4.19) we would otain the following equations involving the coefficients X s iδ ij + (X ik ) 11 (X jk ) 22 + (X ik ) 12 (X jk ) 21 = 0, δ ij (X ik ) 33 (X jk ) 44 + (X ik ) 34 (X jk ) 43 = 0, (A.3) and a good choice for X s could e (ecause it leads to translations in the old coordinates) (X ij ) 11 = iδ ij, (X ij ) 22 = δ ij, (X ij ) 12 = 0 = (X ij ) 21, (X ij ) 34 = δ ij, (X ij ) 43 = δ ij, (X ij ) 33 = 0 = (X ij ) 44. (A.4) The new constraints T read 20

T i 1 = π i + i ψ i + iξ i, T i 2 = π i χ i, T i 3 = s i ρ i, T i 4 = d i v q i + ηi. (A.5) We actually notice that we have chosen X s in such a way that ξ and η appear as translations of the ψ and d, respectively. Following the same procedure as efore, we otain the extended Hamiltonian H c = 1 2 pi p i 1 2 di d i 1 2 ηi η i η i d i + (d i + η i ) v q i ( + ψi ψ j + i ψ i χ j + ξ i ψ j + iξ i χ j + p i ρ j) 2 v q i q j + 1 2 ρi ρ j 2 v q i q k 2 v q k q j. (A.6) Apparently, there is no prolem with this Hamiltonian, ut considering the general expression of the propagator and using the δ functional to eliminate the momenta π, π, and s, as well as performing the integration over χ, we otain Z = N [dq][dp][dd][dψ][d ψ][dη][dρ][dξ] det {, } δ [ T 4 ] a δ [ Λ a ] [ δ ψi + ξ i i( ψ j + ξ j 2 v ] { t [ ) a i q j exp i dt q i p i + d i ρ i + η i ρ i t 0 + 1 2 di d i + 1 2 ηi η i + d i η i (d i + η i ) v q i + iψ i ( ψi + ξ i i( ψ j + ξ j ) 1 2 (p i + ρ j 2 v a i q j ) ( p i + ρ k 2 v ) q i q j 2 v )]} a i q k. (A.7) The δ function that appear into the expression aove was otained after integrating over χ. We notice that the constraint imposed y this δ function is not a consistency condition for other constraints, as occurred with the δ-function that appears in (4.35). More than that, we also notice that it causes the elimination of all terms with ψ and ψ in the propagator. Consequently, this would lead to a final effective Lagrangian without these variales, what is oviously a non consistent result ecause it does not matches the initial theory when the auxiliary fields are 21

turned off. It is also important to emphasize that this result is not related to the particular choice we have made for the coefficients X s given at (A.5). We mention that this kind of inconsistency would persist for any choice we make for X s in equations (A.3). References [1] I.A. Batalin and E.S. Fradkin, Phys. Lett. B180 (1986) 157; Nucl. Phys. B279 (1987) 514; I.A. Batalin, E.S. Fradkin, and T.E. Fradkina, iid. B314 (1989) 158; B323 (1989) 734. [2] I.A. Batalin and I.V. Tyutin, Int. J. Mod. Phys. A6 (1991) 3255. [3] P.A.M. Dirac, Can. J. Math. 2 (1950) 129; Lectures on quantum mechanics (Yeshiva University, New York, 1964). [4] M.B. Green and J.H. Schwarz, Phys. Lett. B136 (1984) 367; Nucl. Phys. B243 (1984) 285. For an in-depth review, see M.B. Green, J.H. Schwarz and E. Witten, Superstring theory (Camridge U.P., Camridge, 1987). [5] W. Siegel, Phys. Lett. B128 (1983) 397; Class. Quantum Grav. 2 (1985) L95. [6] See, for example, R. Kallosh, Phys. Lett. B195 (1987) 369 and E. Nissimov, S. Pacheva and S. Solomon, Nucl. Phys. B297 (1988) 349. [7] R. Amorim and J. Barcelos-Neto, Phys. Lett. B333 (1994) 413; Phys. Rev. D53 (1996) in press. [8] N. Banerjee, S. Ghosh, and R. Banerjee, Nucl. Phys. B417 (1994) 257; Phys. Rev. D49 (1994) 1996. [9] R. Amorim and A. Das, Mod. Phys. Lett. A9 (1994) 3543. R. Amorim, Z. Phys. C67 (1995) 695. [10] E. Witten, Nucl. Phys. B188 (1981) 513. See also, P. Salomonson and J.W. Van Holten, Nucl. Phys. B196 (1982) 509; S.P. Misra and T. Pattnaik, Phys. Lett. B129 (1983) 401; J. Barcelos-Neto and A. Das, Phys. Rev. D33 (1986) 2863; J. Barcelos-Neto and E.S. Che-Terra, Z. Phys. C54 (1992) 133. [11] L.D. Faddeev, Theor. Math. Phys. 1 (1970) 1; P. Senjanovick, Ann. Phys. (NY) 100 (1076) 277. [12] E.S. Fradkin and G.A. Vilkovisky, Phys. Lett. B55 (1975) 224; I.A. Batalin and G.A. Vilkovisky, Phys. Lett. B69 (1977) 309. 22