4
Physics Letters B 680 (2009) 195–198 Contents lists available at ScienceDirect Physics Letters B www.elsevier.com/locate/physletb Finite temperature effective actions Ashok Das a,b,, J. Frenkel c a Department of Physics and Astronomy, University of Rochester, Rochester, NY 14627-0171, USA b Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Calcutta 700064, India c Instituto de Física, Universidade de São Paulo, 05508-090, São Paulo, SP, Brazil article info abstract Article history: Received 30 July 2009 Accepted 24 August 2009 Available online 29 August 2009 PACS: 11.10.Wx 11.15.-q We present, from first principles, a direct method for evaluating the exact fermion propagator in the presence of a general background field at finite temperature, which can be used to determine the finite temperature effective action for the system. As applications, we determine the complete one loop finite temperature effective actions for (0 + 1)-dimensional QED as well as the Schwinger model. These effective actions, which are derived in the real time (closed time path) formalism, generate systematically all the Feynman amplitudes calculated in thermal perturbation theory and also show that the retarded (advanced) amplitudes vanish in these theories. © 2009 Elsevier B.V. All rights reserved. The effective action for a system of fermions interacting with a background field, which incorporates all the one loop corrections in the theory, is an important fundamental concept in quantum field theory. At zero temperature we know that the n-point ampli- tudes (involving the background fields) at one loop are, in general, divergent and, consequently, the evaluation of the effective action at T = 0 needs a regularization. Effective actions can, of course, be evaluated perturbatively. However, a beautiful method due to Schwinger [1], also known as the proper time formalism, is quite useful in evaluating one loop effective actions at zero temperature with a gauge invariant regularization (in the case of gauge back- grounds). It involves solving dynamical equations in the proper time, which are not always trivial, to determine the effective ac- tion. When the dynamical equations can be solved in a closed form, the gauge invariant regularized effective action can be writ- ten in a closed form or at least in an integral representation. This has been profitably used to calculate the imaginary part of the ef- fective action for fermions interacting with a constant background electromagnetic field which describes the decay rate of the vacuum [1]. When the dynamical equations cannot be solved in a closed form, the method due to Schwinger leads to a perturbative deter- mination of the effective action. In the past couple of decades, there have been several attempts [2,3] to generalize the method due to Schwinger to finite temper- ature [4,5] and to determine the imaginary part of the effective * Corresponding author at: Department of Physics and Astronomy, University of Rochester, Rochester, NY 14627-0171, USA. E-mail addresses: [email protected] (A. Das), [email protected] (J. Frenkel). action leading to conflicting results [2]. Here we present an alter- native method for determining finite temperature effective actions for fermions interacting with an arbitrary background field. We believe that since the amplitudes at finite temperature are ultra- violet finite unlike those at zero temperature, it is not necessary to generalize the method due to Schwinger to finite temperature. Af- ter all, the proper time method was designed to provide a (gauge invariant) regularization which is not necessary at finite tempera- ture. Therefore, we propose a direct method for evaluating finite temperature effective actions based mainly on the general proper- ties of systems at finite temperature. In this connection, we believe that the real time formalism [5] (we use the closed time path for- malism due to Schwinger [6]) is more suited for this purpose. We note that, in general, the imaginary time formalism (the Matsubara formalism [7]) leads naturally to retarded and advanced ampli- tudes, but the Feynman (time ordered) amplitudes (beyond the two point function) cannot be consistently generated in this for- malism [8]. On the other hand, the effective action that we are interested in is precisely the one that generates Feynman ampli- tudes. In contrast to the imaginary time formalism, the effective action, when evaluated properly in the real time formalism, leads naturally not only to the Feynman amplitudes, but also to the re- tarded and the advanced amplitudes as we will show in examples. Furthermore, as we have emphasized earlier in [5,9], the real time calculations can be carried out quite easily in the mixed space where the spatial coordinates have been Fourier transformed as we will describe in the following examples. Let us consider a system of fermions interacting with an exter- nal field which we generically denote by A. This can be a scalar or a vector background field and we suppress the Lorentz index (structure) of the background field for simplicity. If the fermion has 0370-2693/$ – see front matter © 2009 Elsevier B.V. All rights reserved. doi:10.1016/j.physletb.2009.08.054

Finite temperature effective actions

Embed Size (px)

Citation preview

Page 1: Finite temperature effective actions

Physics Letters B 680 (2009) 195–198

Contents lists available at ScienceDirect

Physics Letters B

www.elsevier.com/locate/physletb

Finite temperature effective actions

Ashok Das a,b,∗, J. Frenkel c

a Department of Physics and Astronomy, University of Rochester, Rochester, NY 14627-0171, USAb Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Calcutta 700064, Indiac Instituto de Física, Universidade de São Paulo, 05508-090, São Paulo, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history:Received 30 July 2009Accepted 24 August 2009Available online 29 August 2009

PACS:11.10.Wx11.15.-q

We present, from first principles, a direct method for evaluating the exact fermion propagator in thepresence of a general background field at finite temperature, which can be used to determine the finitetemperature effective action for the system. As applications, we determine the complete one loop finitetemperature effective actions for (0+1)-dimensional QED as well as the Schwinger model. These effectiveactions, which are derived in the real time (closed time path) formalism, generate systematically allthe Feynman amplitudes calculated in thermal perturbation theory and also show that the retarded(advanced) amplitudes vanish in these theories.

© 2009 Elsevier B.V. All rights reserved.

The effective action for a system of fermions interacting with abackground field, which incorporates all the one loop correctionsin the theory, is an important fundamental concept in quantumfield theory. At zero temperature we know that the n-point ampli-tudes (involving the background fields) at one loop are, in general,divergent and, consequently, the evaluation of the effective actionat T = 0 needs a regularization. Effective actions can, of course,be evaluated perturbatively. However, a beautiful method due toSchwinger [1], also known as the proper time formalism, is quiteuseful in evaluating one loop effective actions at zero temperaturewith a gauge invariant regularization (in the case of gauge back-grounds). It involves solving dynamical equations in the propertime, which are not always trivial, to determine the effective ac-tion. When the dynamical equations can be solved in a closedform, the gauge invariant regularized effective action can be writ-ten in a closed form or at least in an integral representation. Thishas been profitably used to calculate the imaginary part of the ef-fective action for fermions interacting with a constant backgroundelectromagnetic field which describes the decay rate of the vacuum[1]. When the dynamical equations cannot be solved in a closedform, the method due to Schwinger leads to a perturbative deter-mination of the effective action.

In the past couple of decades, there have been several attempts[2,3] to generalize the method due to Schwinger to finite temper-ature [4,5] and to determine the imaginary part of the effective

* Corresponding author at: Department of Physics and Astronomy, University ofRochester, Rochester, NY 14627-0171, USA.

E-mail addresses: [email protected] (A. Das), [email protected](J. Frenkel).

0370-2693/$ – see front matter © 2009 Elsevier B.V. All rights reserved.doi:10.1016/j.physletb.2009.08.054

action leading to conflicting results [2]. Here we present an alter-native method for determining finite temperature effective actionsfor fermions interacting with an arbitrary background field. Webelieve that since the amplitudes at finite temperature are ultra-violet finite unlike those at zero temperature, it is not necessary togeneralize the method due to Schwinger to finite temperature. Af-ter all, the proper time method was designed to provide a (gaugeinvariant) regularization which is not necessary at finite tempera-ture. Therefore, we propose a direct method for evaluating finitetemperature effective actions based mainly on the general proper-ties of systems at finite temperature. In this connection, we believethat the real time formalism [5] (we use the closed time path for-malism due to Schwinger [6]) is more suited for this purpose. Wenote that, in general, the imaginary time formalism (the Matsubaraformalism [7]) leads naturally to retarded and advanced ampli-tudes, but the Feynman (time ordered) amplitudes (beyond thetwo point function) cannot be consistently generated in this for-malism [8]. On the other hand, the effective action that we areinterested in is precisely the one that generates Feynman ampli-tudes. In contrast to the imaginary time formalism, the effectiveaction, when evaluated properly in the real time formalism, leadsnaturally not only to the Feynman amplitudes, but also to the re-tarded and the advanced amplitudes as we will show in examples.Furthermore, as we have emphasized earlier in [5,9], the real timecalculations can be carried out quite easily in the mixed spacewhere the spatial coordinates have been Fourier transformed aswe will describe in the following examples.

Let us consider a system of fermions interacting with an exter-nal field which we generically denote by A. This can be a scalaror a vector background field and we suppress the Lorentz index(structure) of the background field for simplicity. If the fermion has

Page 2: Finite temperature effective actions

196 A. Das, J. Frenkel / Physics Letters B 680 (2009) 195–198

a mass m, from the definition of the effective action it is straight-forward to obtain

∂Γeff

∂m=

∫dt dx tr S(t,x; t,x), (1)

where S(t,x; t′,x′) denotes the complete Feynman propagator forthe fermion in the presence of the background field and “tr” standsfor the trace over the spinor indices. However, keeping in mindthat the fermion may not always have a mass (say, for example, inthe Schwinger model [10]), we use alternatively the fact that thevariation of the effective action with respect to the backgroundfield leads to the generalized fermion “propagator” at coincidentpoints (even though we use the same symbol as in (1), the exactmeaning of S below depends on the nature of the background fieldas we explain),

δΓeff

δA(t,x)= g tr S(t,x; t,x), (2)

where g denotes the coupling to the background field and we aresuppressing the Lorentz structure of the background field as wellas that of the generalized “propagator”. We note that for a scalarbackground, S in (2) indeed denotes the complete fermion prop-agator of the interacting theory at coincident coordinates. On theother hand, for a gauge field background, the right-hand side in(2) determines the current density of the theory which is relatedto the complete fermion propagator of the theory through a Diractrace involving the Dirac matrix. In either case, we note that it isthe fermion propagator that is relevant in (2) for the evaluation ofthe effective action. In the mixed space (where the coordinates xhave been Fourier transformed), we can write (2) as

δΓeff

δA(t,−p)= g tr S(t, t;p). (3)

Since the effective action is so intimately connected with thefermion propagator and since we are not interested in the zerotemperature part of the effective action, our proposal is to deter-mine the complete fermion propagator at finite temperature di-rectly such that

(i) it satisfies the appropriate equations for the complete propa-gator of the theory,

(ii) it satisfies the necessary symmetry properties of the theorysuch as the Ward identity,

(iii) and most importantly, it satisfies the anti-periodicity propertyassociated with a finite temperature fermion propagator [5].

In fact, it is the third requirement that is quite important in adirect determination of the propagator. We note that this last con-dition is missing at zero temperature which makes it difficult todetermine the complete propagator (independent of the problemof divergence). When the theory is divergence free (so that it doesnot need a regularization at zero temperature), this propagator willbe the exact fermion propagator of the theory and would lead tothe complete effective action including the correct zero tempera-ture part. On the other hand, if the theory needs to be regularizedat zero temperature, this propagator will not yield the correct zerotemperature effective action, but the finite temperature part of theeffective action, which does not need to be regularized, will be de-termined correctly. We illustrate the method with two examples.

Let us start with the (0 + 1)-dimensional QED described by theLagrangian

L = ψ(t)(i∂t − m − e A(t)

)ψ(t), (4)

where the fermion mass can be thought of as a chemical poten-tial and in 0 + 1 dimension, the fermion field as well as the gauge

Fig. 1. The closed time path contour in the complex t plane. Here T → −∞, whileT ′ → ∞ and β denotes the inverse temperature (in units of the Boltzmann constantk) [5]. The two branches labelled by C± lead to the doubling of the degrees offreedom while the branch C⊥ along the imaginary axis decouples from any physicalamplitude.

potential have only a single component. This is a simple modelwhich has been studied exhaustively [11,12] in connection withlarge gauge invariance [13] at finite temperature, but it is alsoquite useful in clarifying what is involved in our proposal beforewe generalize it to higher dimensions. As we noted earlier, we usethe closed time path formalism where the path in the complextime plane has the form shown in Fig. 1. In the closed time pathformalism (in any real time formalism) [5], the degrees of free-dom need to be doubled and we denote the background fields onthe C± branches of the contour as A±(t), respectively. Since t isthe only coordinate on which field variables depend in this the-ory, there is no need for a mixed space propagator. We note thatthe complete fermion propagator of the theory (ordered along thecontour in Fig. 1) satisfies the equations(i∂t − m − e Ac(t)

)Sc(t, t′) = iδc(t − t′),

Sc(t, t′)(i←−∂ t′ + m + e Ac(t

′)) = −iδc(t − t′), (5)

where the subscript “c” characterizes a function on the contour. Onthe contour, the step function is defined naturally as [5]

θc(t − t′) =

⎧⎪⎪⎨⎪⎪⎩

θ(t − t′) if both t, t′ ∈ C+,

θ(t′ − t) if both t, t′ ∈ C− or ∈ C⊥,

1 if t ∈ C− and t′ ∈ C+,

0 if t ∈ C+ and t′ ∈ C−,

(6)

and we have δc(t − t′) = ∂tθc(t − t′).Eqs. (5) can be solved exactly subject to our three requirements

leading to the contour ordered propagator of the form

Sc(t, t′) = e−imt− ie2

∫dt sgnc(t−t)Ac(t)

×(

θc(t − t′) − nF

(m + ie

β(a+ − a−)

))

× eimt′+ ie2

∫dt sgnc(t

′−t)Ac(t), (7)

where nF denotes the Fermi distribution function with β corre-sponding to the inverse temperature (in units of the Boltzmannconstant k) and we have identified

a± =∞∫

−∞dt A±(t). (8)

Furthermore, we have identified

sgnc(t − t′) = θc(t − t′) − θc(t′ − t). (9)

Although the phases in (7) can be combined to write them in asimpler form in this case, we have chosen to write them in thissuggestive form which generalizes naturally to higher dimensionswhere the propagator will carry spinor indices. When t, t′ are re-stricted to the appropriate branches of the contour, (7) determinesall the components of the full 2 × 2 matrix propagator of the the-ory. Furthermore, it can be checked that this propagator satisfies

Page 3: Finite temperature effective actions

A. Das, J. Frenkel / Physics Letters B 680 (2009) 195–198 197

the Lippmann–Schwinger equation (the perturbation expansion)[14] for the propagator. However, we do not go into the detailsof this analysis here, which will be discussed elsewhere.

The (0 + 1)-dimensional theory is free from divergences and,therefore, (7) represents the complete fermion propagator of thetheory in the presence of a background gauge field. We can nowtake the coincident limit (t = t′) and integrate (2) to obtain thenormalized effective action of the theory which has the form

Γeff[a+,a−] = −i ln

[cos

e(a+ − a−)

2

+ i tanhβm

2sin

e(a+ − a−)

2

]. (10)

This is the complete effective action of the theory which reduces tothe well studied action [11,12] on C+ when we set a− = 0. How-ever, being the complete effective action, (10) contains all the in-formation about retarded, advanced and other amplitudes as well.For example, let us note from (10) that since Γeff = Γeff[a+ − a−],the retarded n-point amplitude of the theory can be shown to van-ish (for a definition of retarded amplitudes, see [15]), namely,

Γ(n)R =

n−1∑m=0

(n−1Cm

dn−1−m

dan−1−m+

dm

dam−

)dΓeff[a+ − a−]

da+

∣∣∣∣= (1 − 1)n−1 dnΓeff[a+ − a−]

dan+

∣∣∣∣ = 0, n � 2, (11)

where the restriction stands for setting all the background fields tozero. Therefore, all the retarded (advanced) amplitudes vanish inthis theory. (The one point amplitude is by definition a Feynmanamplitude.)

With this brief derivation of the complete effective action in the(0 + 1)-dimensional theory, let us next consider the fermion sectorof the Schwinger model [10] or massless QED in 1 + 1 dimensionsdescribed by the Lagrangian density

L = ψ(t, x)γ μ(i∂μ − e Aμ(t, x)

)ψ(t, x). (12)

At zero temperature, this model is soluble and describes free mas-sive photons. The effective action for this model (for an arbitrarygauge background) has also been studied perturbatively at finitetemperature [16] even in the presence of a chemical potential [17].Here we will derive the closed form expression of the finite tem-perature effective action following our method. We note here thatthe two point function in the Schwinger model needs to be regu-larized at zero temperature and, consequently, the zero tempera-ture part of the effective action following from our propagator willnot coincide with the regularized zero temperature effective ac-tion. However, our interest is in the finite temperature part of theeffective action which is free from ultraviolet divergences. For com-pleteness we note that the simple point-splitting regularization ofthe fermion propagator is sufficient to regularize the theory andcan be carried out even in our method. However, we will not dothis here since our main interest is in the finite temperature partof the effective action.

The theory (12) is best studied in the natural basis of right-handed and left-handed fermion fields (although everything thatwe say can be carried out covariantly as well as in the presence ofa chemical potential). Defining [18,19]

ψR = 1

2(1 + γ5)ψ, ψL = 1

2(1 − γ5)ψ,

x± = x0 ± x1

2, p± = p0 ± p1, ∂± = ∂0 ± ∂1,

A± = A0 ± A1, (13)

the Lagrangian density (12) naturally decomposes into two decou-pled sectors described by

L = ψ†R(i∂+ − e A+)ψR + ψ

†L (i∂− − e A−)ψL, (14)

where ψR,ψL denote only the component spinor fields (no spinorindex left any more). While the zero temperature regularizationmixes the two sectors through the two point function (anomaly),at finite temperature we do not have divergences and we do notexpect the two sectors to mix. Therefore, we can study the finitetemperature effective action in each of the two sectors separately.

Let us consider the theory only in the sector of the right-handed fermions in (14). This is very much like the (0 + 1)-dimensional theory. However, there is one essential differencewhich makes the derivation much more difficult, namely, thefield variables depend on two coordinates (t, x) or equivalentlyon (x+, x−). We would like to emphasize here that although weuse the light-cone coordinates for simplicity, the theory is stillquantized on the equal-time surface and the propagator is definedthrough the time ordered Green’s function (namely, we do not usethe statistical mechanics of the light-front [20]). As we mentionedearlier, the finite temperature derivations become a lot simpler inthe mixed space. Thus, Fourier transforming the x− coordinate, theaction for the right handed fermions takes the form (the conjugatevariables to x− should be written as p−,k− , which we write asp,k for simplicity)

SR = 2∫

dx+ dp

2πψ

†R(x+,−p)

[i∂+ψR(x+, p)

− e

∫dk

2πA+(x+, p − k)ψR(x+,k)

]. (15)

As a result, we recognize that the equations for the propagatorwill involve a convolution. They are best described by introducingthe following operator notations for the propagator as well as thegauge potential

S(x+, x′+; p,k) = 〈p| S(x+, x′+)|k〉,A+(x+, p − k) = 〈p| A+(x+)|k〉, (16)

so that the equations for the propagator ordered along the contourtake the (operator) forms (see also (5))

(i∂+ − e A+c(x+)

)Sc(x+, x′+) = i

2δc(x+ − x′+),

Sc(x+, x′+)(i←−∂ ′+ + e A+c(x′+)

) = − i

2δc(x+ − x′+). (17)

We note from (3) and (15) that, in the present case, we can iden-tify

Sc(x+, x′+; p) =∫

dk

2π〈k + p| Sc(x+, x′+)|k〉. (18)

The solution to (17) satisfying the Ward identity as well as theappropriate anti-periodicity condition can be determined to havethe form

Sc(x+, x′+) = 1

4e− ie

2

∫dx+ sgnc(x+−x+) A+c(x+)

× (sgnc(x+ − x′+) + 1 − 2(O+ + 1)−1)

× eie2

∫dx+ sgnc(x′+−x+) A+c(x+), (19)

where

sgnc(x+ − x′+) = θc(x+ − x′+) − θc(x′+ − x+),

O+ = eie(a+(+)−a+(−))

2 eβK2 e

ie(a+(+)−a+(−))

2 , (20)

Page 4: Finite temperature effective actions

198 A. Das, J. Frenkel / Physics Letters B 680 (2009) 195–198

with K denoting the momentum operator and ((±) with theparenthesis denote the thermal indices while + without the paren-thesis represents the light-cone component of the backgroundfield)

a+(±) =∞∫

−∞dx+ A+(±)(x+). (21)

It can be checked that this complete propagator satisfies the Lipp-mann–Schwinger equation. In fact, setting x+ = x′+ and using (3)we can integrate (19) to obtain the normalized effective action inthe right-handed sector. The thermal part of the effective action iscontained in the simple form

ΓR,eff = − i

2

∫dk

2π〈k| ln

(1 + ie

2Na+

)|k〉, (22)

with

N = 1 − 2nF

(K

2

), a+ = a+(+) − a+(−). (23)

This effective action has the right (delta function) structure thathad already been observed in the perturbative calculation in theright-handed sector [16] which is a consequence of the Ward iden-tity in the theory. In fact, the expansion of this effective actionon C+ (namely, setting A+(−) = 0) agrees order by order with theperturbative result. The thermal part of the effective action for theleft-handed sector is similarly given by

ΓL,eff = − i

2

∫dk

2π〈k| ln

(1 + ie

2Na−

)|k〉, (24)

where k should be understood as the conjugate variable to x+(which should be written as k+) and we have identified

a− = a−(+) − a−(−),

a−(±) =∞∫

−∞dx− A−(±)(x−). (25)

Once again, this effective action has the right (delta function)structure as in the perturbative calculation and the thermal partagrees with the perturbative result [16] order by order whenrestricted to C+ . The finite temperature effective action for the(1 + 1)-dimensional fermion interacting with an arbitrary Abeliangauge background can, therefore, be written as

Γeff = ΓR,eff + ΓL,eff, (26)

and the thermal part of (26) leads to the correct perturbative re-sult order by order [16] on the branch C+ . However, since (26)represents the complete effective action and since it is a func-tional of (a±(+) − a±(−)) (see (22)–(24)), it can be checked as in(11) that all the retarded (advanced) amplitudes vanish in this the-ory. This should be contrasted with the fact that this had been

verified explicitly only up to the 4-point function in perturbationtheory [15].

In summary, we have proposed an alternative method for de-termining the finite temperature effective action for fermions in-teracting with an arbitrary background field. This is done by deter-mining the complete fermion propagator (in the closed time pathformalism) directly by using the anti-periodicity condition appro-priate at finite temperature. We have illustrated how our proposalworks with the examples of the (0 + 1)-dimensional QED as wellas the Schwinger model. A longer version of the results with moredetails of the calculations as well as other aspects of this analysiswill be reported separately.

Acknowledgements

This work was supported in part by US DOE Grant number DE-FG 02-91ER40685, by CNPq and FAPESP (Brazil).

References

[1] J. Schwinger, Phys. Rev. 82 (1951) 664.[2] W. Dittrich, Phys. Rev. D 19 (1978) 23;

P.H. Cox, W.S. Hellman, Ann. Phys. 154 (1984) 211;M. Loewe, J.C. Rojas, Phys. Rev. D 46 (1992) 2689;P. Elmfors, D. Persson, B.-S. Skagerstam, Phys. Rev. Lett. 71 (1993) 480;P. Elmfors, B.-S. Skagerstam, Phys. Lett. B 348 (1995) 141;H. Gies, Phys. Rev. D 60 (1999) 105002;S.P. Gavrilov, D.M. Gitman, Phys. Rev. D 78 (2008) 045017.

[3] A. Das, J. Frenkel, Phys. Rev. D 75 (2007) 025021.[4] J. Kapusta, Finite Temperature Field Theory, Cambridge University Press, Cam-

bridge, England, 1989;M. Le Bellac, Thermal Field Theory, Cambridge University Press, Cambridge,England, 1996.

[5] A. Das, Finite Temperature Field Theory, World Scientific, Singapore, 1997.[6] J. Schwinger, in: Lecture Notes of Brandeis Summer Institute in Theoretical

Physics, 1960;J. Schwinger, J. Math. Phys. 2 (1961) 407;P.M. Bakshi, K.T. Mahanthappa, J. Math. Phys. 4 (1963) 1;L.V. Keldysh, Sov. Phys. JETP 20 (1965) 1018.

[7] T. Matsubara, Prog. Theor. Phys. 14 (1954) 351.[8] T.S. Evans, Nucl. Phys. B 374 (1992) 340.[9] F.T. Brandt, A. Das, O. Espinosa, J. Frenkel, S. Perez, Phys. Rev. D 72 (2005)

085006;F.T. Brandt, A. Das, O. Espinosa, J. Frenkel, S. Perez, Phys. Rev. D 73 (2006)065010;F.T. Brandt, A. Das, O. Espinosa, J. Frenkel, S. Perez, Phys. Rev. D 73 (2006)067702.

[10] J. Schwinger, Phys. Rev. 128 (1962) 2425.[11] G. Dunne, K. Lee, C. Lu, Phys. Rev. Lett. 78 (1997) 3434.[12] A. Das, G. Dunne, Phys. Rev. D 57 (1998) 5023;

J. Barcelos-Neto, A. Das, Phys. Rev. D 58 (1998) 085022.[13] K.S. Babu, A. Das, P. Panigrahi, Phys. Rev. D 36 (1987) 3725.[14] B.A. Lippmann, J. Schwinger, Phys. Rev. 79 (1950) 469.[15] F.T. Brandt, A. Das, J. Frenkel, A.J. da Silva, Phys. Rev. D 59 (1999) 065004;

F.T. Brandt, A. Das, J. Frenkel, Phys. Rev. D 60 (1999) 105008.[16] A. Das, A.J. da Silva, Phys. Rev. D 59 (1999) 105011.[17] S. Maciel, S. Perez, Phys. Rev. D 78 (2008) 065005.[18] For quantization of massless fermion fields, see, for example, A. Das, Lectures

on Quantum Field Theory, World Scientific, Singapore, 2008.[19] For a general discussion of transformations between different coordinate sys-

tems, see, for example, A. Das, S. Perez, Phys. Rev. D 70 (2004) 065006.[20] V.S. Alves, A. Das, S. Perez, Phys. Rev. D 66 (2002) 125008.