1: \documentclass[12pt]{article}
2: \usepackage[dvips]{epsfig}
3: \usepackage[dvips]{graphics}
4: \usepackage{a4}
5: \usepackage{amssymb}
6:
7: \begin{document}
8: \setlength{\unitlength}{1mm}
9:
10: \title{On the origin of divergences in massless $QED_2$}
11: \author{Rodolfo Casana$^{1\footnote{casana@cbpf.br}}$,
12: Sebasti\~ao A. Dias$^{1,2\footnote{tiao@cbpf.br}}$\\
13: \it{\small $^1$Centro Brasileiro de Pesquisas F\'{\i}sicas}\\
14: \it{ \small Departamento de Campos e Part\'{\i}culas}\\ \it{\small
15: Rua Xavier Sigaud, 150, 22290-180, Rio de Janeiro, Brazil}\\
16: \it{\small $^2$Pontif\'\i cia Universidade Cat\'olica do Rio de
17: Janeiro}\\ {\it \small Departamento de F\'\i sica}\\ {\small \it
18: Rua Marqu\^es de S\~ao Vicente, 225, 22543-900, Rio de Janeiro,
19: Brazil}}
20: \date{ }
21: \maketitle
22:
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24:
25: \begin{abstract}
26: We show that ultraviolet divergences found in fermionic Green's
27: functions of massless $QED_2$ have an essentially non-perturbative
28: nature. We investigate their origin both in gauge invariant
29: formalism (the one where we introduce Wess-Zumino fields to
30: restore quantum gauge invariance) and in gauge non-invariant
31: formalism, mapping two different but equivalent mechanisms
32: responsible for their appearance. We find the same results in both
33: approaches, what contradicts a previous work of Jian-Ge, Qing-Hai
34: and Yao-Yang, that found no divergences in the chiral Schwinger
35: model considered in the gauge invariant formalism.
36: \end{abstract}
37:
38: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
39:
40: \section{Introduction}
41:
42: Gauge theories are nowadays responsible for the description of
43: elementary interactions \cite{Abers-Lee}. One of the main
44: requirements of the standard model is that of anomaly
45: cancellation, without which it is not known how to perform
46: perturbative calculations \cite{t'Hooft}. The phenomenologically
47: achieved equilibrium between the number of families of quarks and
48: leptons guarantees this cancellation. However, in practice,
49: nothing prevents the discovery of new kinds of quarks or leptons,
50: as higher energies are reached. This could threaten the
51: theoretical consistency of the model and raise questions about the
52: correctness of the gauge approach to elementary interactions.
53:
54: However, some features of gauge theories were discovered in the
55: 80's that could put the questions above under a more comfortable
56: perspective. It was realized that, at least in 2 dimensional gauge
57: theories, quantum consistency could be reached even for anomalous
58: theories \cite{Jackiw-Rajaraman} (the anomaly being understood as
59: occurring in the gauge symmetry). A mechanism of symmetry
60: restoration seemed to be operating in the background, becoming
61: explicit through the natural introduction of a new set of degrees
62: of freedom, available only at quantum level, the so-called
63: {\it{Wess-Zumino fields}} \cite{FaddeevS,SchapoV,HaradaT}. Since
64: then, this mechanism has been intensively studied \cite{Abdalla}
65: although the achievements have been little in four dimensions
66: (see, however, \cite{TDKieu}).
67:
68: In two dimensions, two strategies have been mainly followed
69: (taking advantage of the fact that, for this number of dimensions,
70: exactly soluble models are well known). The first, already
71: mentioned above, consists in studying the dynamics of the theory
72: that emerges when one considers explicitly the Wess-Zumino fields,
73: and is called {\it{gauge invariant formalism}} (GIF). The other
74: \cite{LinharesR}, takes into account explicitly the dynamics of
75: the longitudinal part of the gauge field, given by the anomaly. It
76: is called {\it{gauge non-invariant formalism}} (GNIF). In both
77: formalisms one ends with a gauge invariant theory, whether one
78: integrates over the fermions and the Wess-Zumino fields (in GIF)
79: or over the fermions and the longitudinal part of the gauge field
80: (in GNIF). This is achieved regardless of the regularization
81: method employed, which can preserve or not gauge invariance in
82: intermediate computations.
83:
84: This fact suggests that one should consider gauge theories in
85: schemes wider than usual. As this mechanism of ``restoration" of
86: gauge symmetry is acting, there is {\it a priori} no reason to
87: consider a fixed value for the Jackiw-Rajaraman parameter (that
88: value which preserves gauge invariance in intermediate
89: calculations) which appears precisely as a manifestation of
90: regularization ambiguities. In fact, for theories involving chiral
91: fermions, there is no value for this parameter that can preserve
92: intermediate gauge invariance. However one ends up with an
93: effective action which is explicitly gauge invariant
94: \cite{HaradaT}.
95:
96: Another well-known fact is that fermionic correlation functions
97: are divergent if one considers a gauge non-invariant scheme for
98: regularizing the theory \cite{GirottiRR,Boianovsky,Chineses}. A
99: fermion wave function renormalization is enough to render the
100: theory finite, but several subtleties appear that make it very
101: hard to be done exactly, making that the label ``exactly soluble"
102: be dependent on technical advances \cite{tiao-rodolfo}. The main
103: problems are: 1) to identify precisely the origin of the
104: divergences and 2) to learn how to deal with them. The second
105: problem is treated in \cite{tiao-rodolfo}. This paper addresses
106: itself to the detailed examination of the origin of the
107: divergences. We tried to localize this origin in both GIF and
108: GNIF, as a means to check {\it a posteriori} the self-consistency
109: of the two approaches and to have a more clear picture of what is
110: happening.
111:
112: In particular, we found results that are in explicit contradiction
113: to the ones found by Jian-Ge, Qing-Hai and Yao-Yang in
114: \cite{chineses-JP}. In their paper, they found no divergences in
115: the chiral Schwinger model, when considered in the gauge invariant
116: formalism. At the conclusions, we briefly comment their paper and
117: the reasons that conducted them to this mistake.
118:
119: This paper is organized as follows: in section 2 we review the GIF
120: and the GNIF, applying the results to study the Schwinger model
121: (in section 3) and the chiral Schwinger model (in section 4)
122: perturbatively in both formalisms. In these sections we show that
123: the divergences in fermionic Green's functions have a
124: non-perturbative nature, in both models and in both formalisms. We
125: present our conclusions in section 5.
126:
127: \section{GIF and GNIF}
128:
129: The models to be studied are defined by the Lagrangian densities,
130: \begin{eqnarray} \label{lag} {\cal
131: L}[\psi,\overline\psi,A]=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu} +
132: \overline\psi(\,i\partial\!\!\!\slash +e\,A\!\!\!\slash\,P\,)\psi
133: ,
134: \end{eqnarray} where
135: \begin{eqnarray} \label{projetor} P=\left\{\begin{array}{l}
136: {\mathbf 1} \;,\;\mbox{Schwinger model}\\ P_\pm\;,\;\mbox{chiral
137: Schwinger model}
138: \end{array}\right.\nonumber
139: \end{eqnarray} The fermionic determinant is computed exactly in both
140: cases,
141: \begin{eqnarray}\label{detf}
142: \det(i\partial\!\!\!\slash+eA\!\!\!\slash\,P)=\exp\left(iW[A_\mu]
143: \right)=\int\!\!
144: d\psi d\overline\psi \exp\left(i\int\!\!
145: dx\;\overline\psi(i\partial\!\!\!\slash+eA\!\!\!\slash\,P)\psi\right).
146: \end{eqnarray} In general, the determinant has to be calculated
147: through a regularization prescription. For the Schwinger model it
148: is possible to do it in an gauge invariant way, but this is not
149: the case for the chiral coupling. We will calculate the
150: determinant using gauge non-invariant prescriptions, even though,
151: for the Schwinger model, no authentic gauge anomalies are obtained
152: in this way \cite{Zinn-Justin}.
153:
154: To define a free propagator for the field $A_\mu$ from
155: (\ref{lag}), it is usual to introduce a gauge fixing condition.
156: However, the situation here is not the same as in usual gauge
157: theories. As the theory after quantizing fermion fields is gauge
158: non-invariant, the results of its quantization will depend,
159: potentially, on the gauge fixing condition used. This prevents us
160: from using a gauge fixing condition in (\ref{lag}). As we will see
161: below, this problem can be easily bypassed, both in GIF and in
162: GNIF.
163:
164: Let us consider the functional integration over the fermion fields
165: in (\ref{detf}), performing the following change in fermionic
166: variables \begin{eqnarray} \label{psipsi0}
167: \begin{array}{lcl}
168: \psi &\rightarrow & \psi=g\psi^g \\ \overline\psi &\rightarrow &
169: \overline\psi=\overline{{\psi}^g} g^\dag,
170: \end{array}
171: \end{eqnarray} where $g$ belongs to the gauge group under which the
172: fermions transform. In general, we expect that the fermionic
173: measure would not be invariant under this transformation, unless
174: we use explicitly a gauge invariant prescription to define it
175: (which is impossible, in the chiral Schwinger model). Then, we
176: have in general, \begin{eqnarray} \label{mednicf} d\psi
177: d\overline\psi =J[A_\mu,g]d\psi^g d\overline{\psi^g} ,
178: \end{eqnarray} where $J[g, A_\mu]$ is the Jacobian of the
179: transformation. Having computed the fermion determinant, it is
180: possible to obtain this Jacobian easily \cite{HaradaT},
181: \begin{eqnarray}\label{mednicf1}
182: J[A_\mu,g]&=&e^{i\left(W[A_\mu]-W[A^{g}_\mu]\right)},
183: \end{eqnarray} where,
184: \begin{eqnarray} A^{g}_\mu&=&g^{-1}A_\mu g+\frac{i}{e}g^{-1}
185: \partial_\mu g. \end{eqnarray}
186: We observe that, if we use a prescription which preserves the
187: gauge symmetry of $W[A_\mu]$, we obtain $J[g, A_\mu]=1$, according
188: to our expectations.
189:
190: These are the basic facts that lie below the formalisms that we
191: are going to review in the next sections.
192:
193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194:
195: \subsection{The gauge invariant formalism}
196:
197: The generating functional of the theory~(\ref{lag}) is given by
198: the following definition \begin{eqnarray} \label{gerf1}
199: Z[\eta,\overline\eta,J]=N\int\!\! dA_\mu d\psi
200: d\overline\psi\;\exp\left[\;i\int\!\! dx\left(\;{\cal
201: L}[\psi,\overline\psi,A]+\overline\eta\psi+\overline\psi\eta
202: +J{\cdotp\!}A \;\right)\right]. \end{eqnarray} We return to the
203: problem of defining a free propagator for the field $A_\mu$. We
204: notice that, if the theory were gauge invariant at quantum level,
205: we should use Faddeev-Popov's technique \cite{Zinn-Justin} to
206: obtain a well defined functional integration. Harada and Tsutsui
207: ~\cite{HaradaT}, and Babelon, Schaposnik e Viallet~\cite{SchapoV},
208: observed that this is not necessary (in fact, it is redundant)
209: when the theory is not gauge invariant at quantum level because,
210: in this case, different gauge orbits of $A_\mu$ give different
211: contributions to the effective action. However, Faddeev-Popov's
212: technique can still be applied, as it consists of multiplication
213: by $1$, expressed as \begin{eqnarray} \label{f-p}
214: 1=\Delta_f[A_\mu]\int\!\! dg\;\delta(f[A^g_\mu]). \end{eqnarray}
215: In the above formula, $dg$ represents the invariant measure over
216: the gauge group $\mathbf G$, $g\in \mathbf G$ and $f[A_\mu]$ is
217: the gauge fixing condition. Thus, as usual, we insert~(\ref{f-p})
218: in equation (\ref{gerf1}) and change integration variables in the
219: bosonic sector $A_\mu \rightarrow A^{g^{-1}}_\mu$ ($dA_\mu$ and
220: $\Delta_f[A_\mu]$ are gauge invariant by construction). The
221: generating functional (\ref{gerf1}) becomes \begin{eqnarray}
222: \label{gerf3} Z[\eta,\overline\eta,J]&=&N\int\!\! dA_\mu d\psi
223: d\overline\psi dg\;\Delta_f[A_\mu]\,\delta(f[A_\mu]) \\ & &
224: \times\exp\left[\;i\int\!\! dx\left(\;{\cal L}
225: [\psi,\overline\psi,A^{g^{-1}}]+ \overline\eta\psi+
226: \overline\psi\eta +J^\mu A^{g^{-1}}_\mu \;\right)\right].\nonumber
227: \end{eqnarray} Now, we redefine the fermionic fields according to
228: the rule
229: \begin{eqnarray} \label{changfer}
230: \begin{array}{lcl}
231: \psi &\rightarrow & \psi=g\psi^g \label{psipsi},\\ \overline\psi
232: &\rightarrow & \overline\psi= \overline{\psi^g}g^{\dag},
233: \end{array}
234: \end{eqnarray} and we see that the Lagrangian returns to its original
235: form.
236: However, as the measure has not necessarily been defined in a
237: gauge invariant way, it is not invariant under the
238: transformation~(\ref{changfer}), but changes as we saw in
239: (\ref{mednicf1}), \begin{eqnarray} d\psi d\overline\psi =d\psi^g d
240: \overline{\psi^g} e^{i\alpha[A_\mu,g^{-1}]}, \end{eqnarray} where
241: $\alpha[A_\mu,g^{-1}]=W[A^{g^{-1}}_\mu]-W[A_\mu]$ is the
242: Wess-Zumino action.
243: Thus, we obtain the following expression for the generating functional,
244: \begin{eqnarray} \label{gerf4} Z[\eta,\overline\eta,J] & = & N\int\!\!
245: dA_\mu d\psi d\overline\psi dg\;
246: \Delta_f[A_\mu]\delta(f[A_\mu])\;\exp\left(i\alpha[A_\mu,g^{-1}]\right)
247: \\
248: & &\hspace{-0.5cm}\times\exp\left[\,i\int\!\! dx\left({\cal
249: L}[\psi,\overline\psi,A]+\overline\eta g\psi+\overline\psi
250: g^\dag\eta + J{\cdotp\!}A^{g^{-1}}\right)\right].\nonumber
251: \end{eqnarray} Now we can define a free propagator for the $A_\mu$
252: field, by exponentiating the $\delta-$function as in the ordinary
253: situation. The quantization of the theory is, thus, independent of
254: the choice of the gauge fixing condition, as in the usual
255: formulation of Faddeev-Popov. We will use the Lorentz gauge fixing
256: condition, $f[A_\mu]=\frac{1}{\sqrt \xi}\partial\cdotp\!A\;$ and
257: we will absorb $\Delta_f[A]$ in the normalization constant
258: (because the theories are Abelian and the Faddeev-Popov's ghost
259: fields decouple). Doing this, we arrive at \begin{eqnarray}
260: \label{gfunt} Z[\eta,\overline\eta,J] & = & N' \int\!\! dA_\mu
261: d\psi d\overline\psi dg\,\exp\left(i\alpha[A_\mu,g^{-1}]\right)\\
262: & &\times\exp\left[i\int\!\! dx\left({\cal
263: L}[\psi,\overline\psi,A]-\frac{1}{2\xi}
264: (\partial{\cdotp}A)^2+\overline\eta g\psi+\overline\psi g^\dag
265: \eta + J{\cdotp\!}A^{g^{-1}}\right)\right].\nonumber
266: \end{eqnarray} This equation will be the starting point for the
267: perturbative analysis of the theory defined by (\ref{lag}). As we
268: are in a gauge non-invariant theory, the source $J^\mu$ will have
269: its divergence $\partial_\mu J^\mu \neq 0$, in general. Another
270: thing that must be indicated is that, in the process of defining
271: the free propagator of $A_\mu$, the theory acquires an additional
272: degree of freedom, given by the Wess-Zumino field, which, in the
273: end, interacts with the fermion fields through the fermionic
274: sources. This interaction is very complicated, and prevents the
275: exact calculation of $Z[\eta,\overline\eta,J]$. However, the exact
276: calculation of an arbitrary correlation function is possible, at
277: least in principle (once one renormalizes the divergences to be
278: found in the next sections). The possibility of defining correctly
279: a free propagator for $A_\mu$ is essential to perform a
280: perturbative analysis of the theory, and thus, to be able to see
281: if the ultraviolet divergences that appear in the fermionic
282: Green's functions have a perturbative origin or not.
283:
284:
285: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
286:
287: \subsection{The gauge non-invariant formalism}
288:
289: Another approach to the perturbative problem is commonly called
290: gauge non-invariant formalism~\cite{LinharesR,HaradaN}. In this
291: context, we use the fact that the classical decoupling of the
292: longitudinal part of $A_\mu$ (that can be obtained with a gauge
293: transformation of the fermion fields) does not keep the fermionic
294: measure invariant, in general. This fact can be exploited to
295: obtain a perturbative description of the theory, as we will see
296: below.
297:
298: Let us separate the field $A_\mu$ in its longitudinal and
299: transverse parts, as usual, \begin{eqnarray} eA_\mu&=&
300: \partial_\mu \rho-\tilde\partial_\mu\phi, \end{eqnarray} and substitute
301: the expression above into the generating functional~(\ref{gerf1})
302: \begin{eqnarray} \label{ger2} Z[\eta,\overline\eta,J]&\!\!\!=&\!\!\!
303: N\int\!\! d\rho d\phi d\psi d\overline\psi\;\exp\left(i\int\!\!
304: dx\;{\cal L}[\psi,\overline\psi,\frac{1}{e} \partial_\mu
305: \rho-\frac{1}{e}\tilde\partial_\mu\phi] \right) \nonumber \\ &
306: &\times\exp\left[i\int\!\!
307: dx\left(\overline\eta\psi+\overline\psi\eta +\frac{1}{e}J_\mu
308: \partial^\mu\rho-\frac{1}{e}J_\mu
309: \tilde\partial^\mu\phi\;\right)\right]. \end{eqnarray} We see
310: that, as the classical action is gauge invariant, the longitudinal
311: part of the field $A_\mu$ (the field $\rho$) does not have a
312: kinetical term, apparently appearing as an auxiliary field.
313: Classically it is possible to remove completely this field from
314: the sourceless part of the action through the following
315: transformation
316: \begin{eqnarray}\label{mudfer}
317: \begin{array}{lcl}
318: \psi&\rightarrow&\psi=g\psi'\\
319: \overline\psi&\rightarrow&\overline\psi=\overline{\psi'} g^\dag.
320: \end{array}
321: \end{eqnarray} with $g=e^{i\rho P}$. If the measure were invariant under
322: this transformation, we would have a linear dependence on $\rho$,
323: that would render its integration undefined (in the non-anomalous
324: case, that is why we have to use Faddeev-Popov's method: to
325: generate a kinetical term for $\rho$). However, the fermionic
326: measure, as we saw, is not invariant under (\ref{mudfer}), but
327: changes as
328: \begin{eqnarray} d\psi d\overline\psi = d\psi' d\overline{\psi'}
329: e^{i\alpha[\rho,\phi]}\; , \end{eqnarray} where
330: $\alpha[\rho,\phi]=W[\frac{1}{e}\partial_\mu\rho-\frac{1}{e}
331: \tilde\partial_\mu\phi]-W[-\frac{1}{e}\tilde\partial_\mu\phi]$.
332:
333: The generating functional~(\ref{ger2}) then acquires the following
334: form\begin{eqnarray} \label{ger333}
335: Z[\eta,\overline\eta,J]&=&N\int\!\! d\rho d\phi d\psi
336: d\overline\psi\;\exp\left(i\alpha[\rho,\phi]+i \int\!\! dx\; {\cal
337: L}[\psi,\overline\psi,-\frac{1}{e}\tilde\partial_\mu\phi]\right)\\
338: & &\times\exp\left[i\int\!\! dx\left(\overline\eta
339: g\psi+\overline\psi g^\dag \eta +\frac{1}{e}J_\mu\partial^\mu\rho
340: -\frac{1}{e}J_\mu\tilde\partial^\mu \phi\right)\right].\nonumber
341: \end{eqnarray} As we are going to see in the next two sections,
342: the $\alpha$ term contains a kinetical term for the $\rho$ field
343: which allows us to treat the theory perturbatively. We notice that
344: the coupling of $\rho$ to the fermion fields is done through the
345: fermionic sources and is not minimal anymore. However, the theory
346: in (\ref{ger333}) now admits a perturbative analysis, as we will
347: explicitly show for the two models mentioned in the beginning.
348:
349: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
350:
351: \section{Schwinger model}
352:
353: \subsection{Perturbative analysis in GIF}
354:
355: The Schwinger model is defined by setting $P={\mathbf 1}$ (see
356: equation (\ref{projetor})). A typical gauge group element can be
357: parameterized by a field $\theta$ as $g=e^{i\theta}$. The
358: Wess-Zumino action is
359: \cite{Dias-Linhares}\begin{eqnarray}\label{wzr}
360: \alpha(A_\mu,\theta)&=&\frac{(a-1)}{2\pi}\int\!\!
361: dx\left(\;\frac{1}{2}\partial_\mu \theta \partial^\mu \theta -
362: e\theta\partial_\mu A^\mu\;\right). \end{eqnarray} Notice that,
363: for $a=1$, the Wess-Zumino action is zero and we have a Jacobian
364: equal to one, which characterizes quantum gauge invariance. We are
365: going to compute bosonic and fermionic Green's functions
366: perturbatively, looking for the appearance of divergences.
367: \subsubsection{Photon Propagator}
368: If we take two functional derivatives with respect to $J_\mu(x)$
369: and $J_\nu(y)$ of (\ref{gfunt}) we obtain, \begin{eqnarray}
370: \label{profot22} \langle
371: 0|TA_\mu(x)A_\nu(y)|0\rangle&\!\!\!\equiv&\!\!\! _\theta\langle
372: 0|TA_\mu(x) A_\nu(y)|0\rangle_\theta+\frac{1}{e} \;_\theta\langle
373: 0|TA_\mu(x)\partial^y_\nu \theta(y)|0\rangle_\theta\;+\\ &
374: &+\;\frac{1}{e}\;_\theta\langle0|T\partial^x_\mu\theta(x)A_\nu(y)|0
375: \rangle_\theta +\frac{1}{e^2}\;_\theta\langle
376: 0|T\partial^x_\mu\theta(x)\partial^y_\nu
377: \theta(y)|0\rangle_\theta\;.\nonumber \end{eqnarray} In the
378: equation above we see that the photon propagator in the original
379: theory is expressed as a sum of propagators referred to another
380: theory given by an action $S_{\theta}[\psi,\overline\psi,
381: A_\mu,\theta]$ that includes the $\theta$ field,
382:
383: \vskip 1cm
384: \begin{eqnarray}
385: S_{\theta}[\psi,\overline\psi,A_\mu,\theta]&\!\!\!=&\!\!\!\!\int\!\!
386: dx\left[-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}-\frac{1}{2\xi}(\partial
387: {\cdotp}A)^2+
388: \overline\psi(i\partial\!\!\!\slash+eA\!\!\!\slash)\psi\,+\right.\\
389: & & \left.\hspace{2.5cm}+\frac{(a-1)}{4\pi}\partial_\mu \theta
390: \partial^\mu \theta - \frac{(a-1)}{2\pi}e\theta\partial_\mu
391: A^\mu\right]\nonumber \end{eqnarray} We will denote the
392: expectation values in this theory by
393: $_\theta\langle\;\;\rangle_\theta$. From
394: $S_{\theta}[\psi,\overline\psi,A_\mu,\theta]$ we easily obtain
395: Feynman rules and compute the necessary expectation values. In the
396: case that we are considering here, it will be possible to add up
397: the perturbative series and compare with the exact result. In the
398: following we exhibit the relevant Feynman diagrams and their
399: result after computation:
400:
401: a) $A_\mu-$propagator, $\;_\theta\langle
402: 0|TA_\mu(x)A_\nu(y)|0\rangle_\theta$ :
403: \begin{eqnarray}\label{serieic}
404: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,550][300,650]
405: {w2fig2.ps}}
406: \end{eqnarray}
407: %\setcounter{equation}{22}
408: The fermionic loop in~(\ref{serieic}) has to be computed using
409: some regularization, which may be of the Pauli-Villars kind, for
410: example (see also \cite{Dias-Linhares}). However, once this
411: regularization is employed, as is well known, the result is
412: finite. Adding the contributions up to the third graphic, we get a
413: result valid at order $e^2$
414: \begin{eqnarray} -\frac{i\,e^2(a+1)}{2\pi
415: p^4}\left(g_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\right).\nonumber
416: \end{eqnarray} Noticing that (\ref{serieic}) is a geometrical
417: series with $n-$th term ($n\geqslant 1$) \begin{eqnarray}
418: \frac{-i}{p^2}\left(\frac{e^2(a+1)}{2\pi p^2}\right)^n
419: \left(g_{\mu\nu}- \frac{p_\mu p_\nu}{p^2}\right),\nonumber
420: \end{eqnarray} we can easily add the series~(\ref{serieic}) to
421: obtain the complete $A_\mu-$propagator in momentum space,
422: \begin{eqnarray} _\theta\langle
423: 0|TA_\mu(p)A_\nu(-p)|0\rangle_\theta&=&\frac{-i}{p^2-
424: \frac{e^2(a+1)}{2\pi}}\left(\,g_{\mu\nu}-\frac{p_\mu
425: p_\nu}{p^2}\,\right) -\frac{i\xi p_\mu p_\nu}{p^4} \end{eqnarray}
426:
427: b) $\theta-$propagator, $\;_\theta\langle
428: 0|T\theta(x)\theta(y)|0\rangle_\theta$ : \begin{eqnarray}
429: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
430: {w2fig3.ps}}\nonumber \end{eqnarray} Computing these two
431: contributions, we get
432: \begin{eqnarray} _\theta\langle
433: 0|T\theta(p)\theta(-p)|0\rangle_\theta &=&
434: i\left(\frac{2\pi}{a-1}\frac{1}{p^2}-\frac{\xi e^2}{p^4}\right).
435: \end{eqnarray}
436:
437: %\newpage
438: c) Mixing terms, $\;_\theta\langle
439: 0|T\theta(x)A_\mu(y)|0\rangle_\theta$ and $_\theta\langle
440: 0|TA_\mu(x)\theta(y)|0\rangle_\theta$ :
441:
442: The two terms that contribute are: \begin{eqnarray}
443: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
444: {w2fig4.ps}}\nonumber
445: \end{eqnarray} \begin{eqnarray} _\theta\langle
446: 0|T\theta(p)A_\mu(-p)|0\rangle_\theta&=&-\frac{\xi e p_\mu}{p^4}
447: \end{eqnarray} \begin{eqnarray}
448: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
449: {w2fig5.ps}}\nonumber
450: \end{eqnarray} \begin{eqnarray} _\theta\langle
451: 0|TA_\mu(p)\theta(-p)|0\rangle_\theta&=&\frac{\xi e p_\mu}{p^4}
452: \end{eqnarray} Adding all results as in (\ref{profot22}), we obtain
453: the full photon propagator for the theory,
454: \begin{eqnarray}\label{propfotex}
455: i\langle0|TA_\mu(p)A_\nu(-p)|0\rangle=\frac{1}{p^2-
456: \frac{e^2}{2\pi}(a+1)}
457: \left(g_{\mu\nu}-\frac{p_\mu
458: p_\nu}{p^2}\right)-\frac{2\pi}{e^2(a-1)} \frac{p_\mu p_\nu}{p^2}.
459: \end{eqnarray} This result agrees exactly with which is obtained by
460: non-perturbative methods \cite{Mitra-Rahaman} (taking into account
461: that the $\overline{a}$ parameter there is related to ours as
462: $\overline{a}=(a-1)/2$).
463:
464: \subsubsection{Fermion propagator}
465:
466: From~(\ref{gfunt}), we can take functional derivatives with
467: respect to the fermionic sources $\overline\eta(x)$ e $\eta(y)$
468: and compute the fermion propagator as \begin{eqnarray} \langle
469: 0|T\psi(x)\overline\psi(y)|0\rangle &=& N'\!\int\!\! dA_\mu
470: d\theta d\psi d\overline\psi \;\;\psi(x)\overline\psi(y)\nonumber
471: \\ &
472: &\times\exp\left(i\,S_{\theta}[\psi,\overline\psi,A_\mu,\theta]+
473: \!\int\!\! dz\;\theta(z)j(z,x,y)\right) \end{eqnarray} where
474: $j(z,x,y)=\delta(z-x)-\delta(z-y)$. Integrating over the $\theta$
475: field, we obtain \begin{eqnarray} \label{petfer22} \langle
476: 0|T\psi(x)\overline\psi(y)|0\rangle =
477: \exp\left\{-\frac{2\pi\,i}{a-1} \int\!\!
478: \frac{dk}{(2\pi)^2}\;\frac{1-e^{-ik{\cdotp}(x-y)}}{k^2}\right\}\;
479: G_p(x-y),
480: \end{eqnarray} where the exponential contains a divergence already
481: found elsewhere \cite{tiaomarcelo,Mitra-Rahaman} which is
482: generated by the integration of the field $\theta$. This
483: divergence is not cancelled by the normalization factor $N'$, as
484: it is induced by the presence of $j(z,x,y)$ (which, in turn, is
485: generated by the functional derivations, absent in the
486: normalization factor). $G_p(x-y)$ is defined from the remaining
487: functional integration in terms of the fields $A_\mu$, $\psi$ and
488: $\overline\psi$,
489: \begin{eqnarray}\label{petfer2} G_p(x-y)&=&N''\int\!\! dA_\mu d\psi
490: d\overline\psi
491: \;\;\psi(x) \overline\psi(y)\\ & & \times\exp\left\{i\int\!\!
492: dz\left(\;\frac{1}{2}A_\mu H^{\mu\nu}_\xi
493: A_\nu+\overline\psi(i\partial\!\!\!\slash
494: +eA\!\!\!\slash)\psi+eA_\mu l^\mu(z,x,y)\right)\right\}\nonumber
495: \end{eqnarray} where
496: \begin{eqnarray}
497: l^\mu(z,x,y)=\partial^\mu_z[\,D_F(z-x)-D_F(z-y)\,], \nonumber
498: \end{eqnarray} and
499: \begin{eqnarray} \label{HHH}
500: H^{\mu\nu}_\xi&=&g^{\mu\nu}\square+(\frac{1}{\xi}-1)\partial^\mu
501: \partial^\nu
502: -\frac{e^2}{2\pi}(a-1)\frac{\partial^\mu\partial^\nu}{\square}.
503: \end{eqnarray} From (\ref{HHH}), we obtain an effective free propagator
504: for $A_\mu$, that we call $h^\xi_{\mu\nu}$ \begin{eqnarray}
505: h^\xi_{\mu\nu}(k)=-\frac{i}{k^2}\left(g_{\mu\nu}-\frac{k_\mu
506: k_\nu}{k^2}\right)-\frac{i\xi k_\mu k_\nu }{k^2[k^2+\frac{\xi
507: e^2}{2\pi}(a-1)]}.\nonumber \end{eqnarray} Its ultraviolet
508: behavior is of the form $k^{-2}$. The Feynman rules to calculate
509: the fermion self-energy are now given by the action appearing in
510: the functional integration (\ref{petfer2}). It is now easy to
511: calculate $G_p(x-y)$ to any desired loop order. We limit ourselves
512: to the 1-loop contribution to the fermion self-energy,
513: \begin{eqnarray} \label{selffer1}
514: -i\Sigma_p(p)&=&i\,p\!\!\!\slash\left[\frac{e^2}{4\pi
515: p^2}-\frac{1}{2(a-1)} \ln\left(1+\frac{\xi e^2}{2\pi
516: p^2}(a-1)\right)\right],
517: \end{eqnarray} that is finite, as well as all the other diagrams
518: that enter in the computation of $G_p(x-y)$. So, the only source
519: of divergences in (\ref{petfer22}) is the integration over the
520: $\theta$ field, which is done exactly, outside the perturbative
521: level.
522:
523: A little further reflection shows quickly that the same is true
524: for all fermionic Green's functions: they all exhibit a
525: divergence, originating in the integration over the Wess-Zumino
526: field, being finite {\it modulo} this problem.
527:
528:
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530:
531: \subsection{Perturbative analysis in GNIF}
532:
533: Now we start from (\ref{ger333}), where \begin{eqnarray} {\cal
534: L}[\psi,\overline\psi,-\frac{1}{e}\tilde\partial_\mu\phi]&=&
535: \frac{1}{2e^2}\phi\,\square^2\phi+\overline\psi(i\partial\!\!\!\slash-
536: \tilde\partial\!\!\!\slash\phi)\psi, \end{eqnarray} having
537: $g=e^{i\rho}$ and a Jacobian given by \begin{eqnarray}
538: \label{aaapha}
539: \alpha(\rho,\phi)=\exp\left(\frac{i\,(a-1)}{4\pi}\int\!\!
540: dx\;\partial_\mu\rho
541: \partial^\mu\rho\right).
542: \end{eqnarray}
543: \subsubsection{Photon propagator}
544: The propagator for the $A_\mu$ field has to be expressed in terms
545: of propagators for the $\rho$ and $\phi$ fields. In (\ref{ger333})
546: we take two functional derivatives with respect to $J_\mu$ and
547: $J_\nu$, put all the sources to zero and we are left with
548: \begin{eqnarray} \label{pertfoton11} \langle
549: 0|TA_\mu(x)A_\nu(y)|0\rangle&\!\!\!=&\!\!\!\frac{1}{e^2}
550: \left(_l\langle
551: 0|T\partial^x_\mu\rho(x)\partial^y_\nu\rho(y)|0\rangle_l\,+
552: \,_l\langle0|T\tilde\partial^x_\mu\phi(x)\tilde\partial^y_\nu
553: \phi(y) |0 \rangle_l\right).\nonumber \\ \end{eqnarray}
554: $_l\langle\,|\,\rangle_l$ refers to expectation values calculated
555: using the effective action
556: \begin{eqnarray}\label{sefef} S_{l}[\psi,\overline\psi,\rho,\phi]
557: &=&\!\!\!\int\!\!
558: dx\left(\frac{1}{2e^2} \phi \square^2 \phi +
559: \frac{(a-1)}{4\pi}\partial_\mu\rho
560: \partial^\mu\rho+\overline\psi(i \partial\!\!\!\slash -
561: \tilde\partial\!\!\!\slash\phi
562: )\psi\right). \end{eqnarray} The photon propagator is split into a
563: sum of two propagators of the fields $\rho$ and $\phi$, whose
564: dynamics is described by the action above. We see in
565: $S_{l}[\psi,\overline\psi,\rho,\phi]$ that $\rho$ is a free field,
566: which implies that the mixed propagators $\;_l\langle
567: 0|T\partial^x_\mu\rho(x)\tilde\partial^y_\nu\phi(y)|0\rangle_l\;$
568: and $_l\langle
569: 0|T\tilde\partial^x_\mu\phi(x)\partial^y_\nu\rho(y)|0\rangle_l\;$
570: are null. With the Feynman rules generated from
571: $S_{l}[\psi,\overline\psi,\rho,\phi]$, we can calculate the photon
572: propagator for the theory in a perturbative way. We show these
573: results below:
574:
575: a) $\phi-$propagator, $\;_l\langle 0|T\phi(x)\phi(y)|0\rangle_l$:
576: \begin{eqnarray}\label{series2}
577: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,545][300,650]
578: {w2fig7.ps}}
579: \end{eqnarray}
580: %\setcounter{equation}{41}
581: The fermionic loop (\ref{series2}) is calculated with a
582: Pauli-Villars regularization prescription, as before. Its
583: contribution, in this case, is \begin{eqnarray}
584: -\frac{i(a+1)}{2\pi}p^2.\nonumber \end{eqnarray} Adding the series
585: (\ref{series2}), we get the propagator of the $\phi$ field
586: \begin{eqnarray} _l\langle
587: 0|T\phi(p)\phi(-p)|0\rangle_l=\frac{ie^2}{p^2(p^2-
588: \frac{e^2(a+1)}{2\pi})} \end{eqnarray}
589:
590: b) $\rho-$propagator, $\;_l\langle 0|T\rho(x)\rho(y)|0\rangle_l$:
591: as the $\rho$ field in (\ref{sefef}) is a free field, its
592: propagator is calculated directly, \begin{eqnarray} _l\langle
593: 0|T\rho(p)\rho(-p)|0\rangle_l&=&\frac{2\pi i}{(a-1)\;p^2}.
594: \end{eqnarray}
595:
596: The photon propagator is then obtained from equation
597: (\ref{pertfoton11}), \begin{eqnarray} i\langle
598: 0|TA_\mu(p)A_\nu(-p)|0\rangle=\frac{1}{k^2-
599: \frac{e^2(a+1)}{2\pi}}\left(g_{\mu\nu}-\frac{k_\mu
600: k_\nu}{k^2}\right)-\frac{2\pi}{e^2(a-1)}\frac{k_\mu k_\nu}{k^2}.
601: \end{eqnarray} It coincides with the propagator computed by non
602: perturbative calculation, and with the one calculated previously
603: (\ref{propfotex}) in the gauge invariant formalism.
604:
605: \subsubsection{Fermion propagator}
606:
607: From (\ref{ger333}), we take functional derivatives with respect
608: to the fermion sources and we obtain, \begin{eqnarray}
609: \label{pertfermion} \langle
610: 0|T\psi(x)\overline\psi(y)|0\rangle&=&N\int\!\! d\rho d\phi d\psi
611: d\overline\psi\;\;\psi(x)\overline\psi(y)\\ & &
612: \times\exp\left(iS_{l}[\psi,\overline\psi,\rho,\phi]+i\int\!\!
613: dz\; \rho(z) j(z,x,y)\right).\nonumber \end{eqnarray} The term
614: involving $j(z,x,y)=\delta(z-x)-\delta(z-y)$ is generated when we
615: perform the above mentioned functional derivatives. The
616: integration over the $\rho$ field factorizes, but the presence of
617: $j(z,x,y)$ prevents the absorption of this integration in the
618: normalization factor $N$. Then (\ref{pertfermion}) becomes, after
619: $\rho$ integration, \begin{eqnarray} \langle
620: 0|T\psi(x)\overline\psi(y)|0\rangle&=&\exp\left\{-\frac{2\pi\,i}{a-1}
621: \int\!\!
622: \frac{dk}{(2\pi)^2}\;\frac{1-e^{-ik{\cdotp}(x-y)}}{k^2}\right\}\;
623: G_p(x-y)\,.
624: \end{eqnarray} We observe the presence of an ultraviolet divergence in
625: the exponential, that do not have perturbative origin (as it comes
626: from the $\rho$ integration) and coincides with the one calculated
627: previously in the gauge invariant formalism (see equation
628: (\ref{petfer22})). The remaining functional integration in
629: $G_p(x-y)$, given by \begin{eqnarray} \label{GPPP}
630: G_p(x-y)&\!\!\!=&\!\!\! N'\!\int\!\! d\phi d\psi
631: d\overline\psi\;\psi(x)\overline\psi(y)\;\exp\left[i\int\!\!
632: dx\left(\frac{1}{2e^2} \phi \square^2 \phi +\overline\psi(i
633: \partial\!\!\!\slash - \tilde\partial\!\!\!\slash\phi )\psi\right)
634: \right],\nonumber \\ \end{eqnarray} is
635: finite, that is, the Feynman diagrams generated from it do not
636: show ultraviolet divergences. From (\ref{GPPP}), the 1-loop
637: contribution to the fermion self-energy is given by
638: \begin{eqnarray} \label{rficfnic2} -i\Sigma_p(p)=\frac{ie^2
639: p\!\!\!\slash}{4\pi p^2}. \end{eqnarray} Since it is finite, the
640: contributions of higher loop order to the self-energy are also
641: finite.
642:
643: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
644:
645: \section{Chiral Schwinger model}
646:
647: \subsection{Perturbative analysis in GIF}
648:
649: We will perform the same analysis for the chiral Schwinger model,
650: defined by $P=P_\pm$ (see, again, equation (\ref{projetor})). As
651: before, $g=e^{i\theta}$, but now the Wess-Zumino action is given
652: by \cite{Dias-Linhares} \begin{eqnarray}\label{aaqq}
653: \alpha(A,\theta)=\int\!\! dx\left\{\frac{(a-1)}{8\pi}\partial_\mu
654: \theta \partial^\mu \theta
655: -\frac{e\theta}{4\pi}\left[\,(a-1)\partial^\mu A_\mu
656: -\tilde\partial^\mu A_\mu\,\right] \right\} \end{eqnarray} As
657: opposed to the case of the Schwinger model, there is no value of
658: $a$ which can turn this action to zero, and this is the
659: distinctive sign of the gauge anomaly. Again, we will compute
660: Green's functions perturbatively, observing similarities and
661: differences in comparison to the vectorial coupling case.
662:
663: \subsubsection{Photon propagator}
664:
665: From (\ref{lag}), (\ref{gfunt}) and (\ref{aaqq}), we obtain the
666: following expression for the photon propagator of the theory
667: \begin{eqnarray} \label{fotqui} \langle 0|TA_\mu(x)A_\nu(y)|0
668: \rangle &\equiv&{_\theta\langle} 0|TA_\mu(x)A_\nu(y)|0
669: \rangle_\theta+\frac{1}{e}{_\theta\langle}
670: 0|TA_\mu(x)\partial^y_\nu\theta(y)|0 \rangle_\theta+\\ & &
671: +\frac{1}{e}{_\theta\langle} 0|T\partial^x_\mu\theta(x)A_\nu(y)|0
672: \rangle_\theta+ \frac{1}{e^2}{_\theta\langle}
673: 0|T\partial^x_\mu\theta(x)\partial^y_\nu\theta(y)|0\rangle_\theta
674: \nonumber \end{eqnarray} The propagator of the original photon is
675: again a sum of propagators which are referred to an effective
676: theory $S_{\theta}[\psi,\overline\psi, A_\mu,\theta]$ that
677: includes the $\theta$ field. We denote the expectation values in
678: this theory by $_\theta\langle\,|\,\rangle_\theta$. This effective
679: action $S_{\theta}[\psi,\overline\psi,A_\mu,\theta]$ is given by
680: \begin{eqnarray}
681: S_{\theta}[\psi,\overline\psi,A_\mu,\theta]&\!\!\!=&\!\!\!\!\int\!\!
682: dx\left[-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}-\frac{1}{2\xi}
683: (\partial{\cdotp}A)^2+
684: \overline\psi(i\partial\!\!\!\slash+eA\!\!\!\slash\,P_+)\psi\,+\right.
685: \\
686: & & \left.+\frac{(a-1)}{8\pi}\partial_\mu \theta \partial^\mu
687: \theta - \frac{(a-1)}{4\pi}e\theta\partial_\mu A^\mu+
688: \frac{1}{4\pi}e\theta\tilde\partial_\mu A^\mu\right]\nonumber
689: \end{eqnarray} With Feynman rules obtained from
690: $S_{\theta}[\psi,\overline\psi,A_\mu,\theta]$, we will show, in
691: what follows, the perturbative calculation of the propagators that
692: appear in (\ref{fotqui}).
693:
694: a) $A_\mu-$propagator \begin{eqnarray}\label{serieq}
695: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,550][300,650]
696: {w3fig2.ps}}
697: \end{eqnarray} The fermionic loop in (\ref{serieq}) is given by
698: \begin{eqnarray}
699: -i\Pi_{\mu\nu}(p)=\frac{ie^2}{4\pi}\left[(a+1)g_{\mu\nu}-\frac{2
700: p_\mu p_\nu}{p^2}-\frac{p_\mu\tilde p_\nu+\tilde p_\mu
701: p_\nu}{p^2}\right],\nonumber \end{eqnarray} while the third
702: graphic in (\ref{serieq}) contributes as \begin{eqnarray}
703: \frac{ie^2}{4\pi}\left[\frac{g_{\mu\nu}}{a-1}-\frac{1+(a-1)^2}{a-1}
704: \frac{p_\mu p_\nu}{p^2}+\frac{p_\mu\tilde p_\nu+\tilde p_\mu
705: p_\nu}{p^2}\right].\nonumber \end{eqnarray} Adding both
706: contributions we obtain, to order $e^2$ \begin{eqnarray}
707: \frac{i\,e^2\;a^2}{4\pi\,(a-1) }\left[g_{\mu\nu}-\frac{p_\mu
708: p_\nu}{p^2}\right]. \end{eqnarray} It is easy to see that
709: (\ref{serieq}) is a geometrical series with an order $n$ term
710: ($n\geqslant 1$)
711: \begin{eqnarray}
712: \frac{-i}{p^2}\left(\frac{e^2\;a^2}{4\pi\,(a-1)p^2}\right)^n
713: \left(\;g_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\;\right).\nonumber
714: \end{eqnarray} Adding this series (\ref{serieq}), we get
715: \begin{eqnarray} _\theta\langle
716: 0|TA_\mu(p)A_\nu(-p)|0\rangle_\theta=\frac{-i}{p^2-
717: \frac{e^2\;a^2}{4\pi\,(a-1)
718: }} \left(\;g_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\;\right)-\frac{i\xi
719: p_\mu p_\nu}{p^4}. \end{eqnarray}
720:
721: b) $\theta-$propagator \begin{eqnarray}
722: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
723: {w3fig3.ps}}\nonumber
724: \end{eqnarray} \begin{eqnarray} _\theta\langle
725: 0|T\theta(p)\theta(-p)|0\rangle_\theta&=&\frac{4\pi}{a-1}
726: \frac{i}{p^2}-\frac{i\xi
727: e^2}{p^4}+\frac{ie^2}{p^2(p^2-\frac{e^2\;a^2}{4\pi\,(a-1)})}
728: \end{eqnarray}
729:
730: c) Mixed terms $A_\mu-\theta$ \begin{eqnarray}
731: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
732: {w3fig4.ps}}\nonumber
733: \end{eqnarray} \begin{eqnarray} _\theta\langle
734: 0|T\theta(p)A_\nu(-p)|0\rangle_\theta&=& -\frac{\xi e
735: p_\nu}{p^4}+\frac{e}{(a-1)}\;\frac{\tilde
736: p_\nu}{p^2(p^2-\frac{e^2\;a^2}{4\pi\,(a-1)})} \end{eqnarray}
737:
738: \begin{eqnarray}
739: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,618][300,650]
740: {w3fig5.ps}}\nonumber
741: \end{eqnarray} \begin{eqnarray} _\theta\langle
742: 0|TA_\mu(p)\theta(-p)|0\rangle_\theta&=&\frac{\xi e p_\mu}{p^4}-
743: \frac{e}{(a-1)}\;\frac{\tilde
744: p_\mu}{p^2(p^2-\frac{e^2\;a^2}{4\pi\,(a-1)})} \end{eqnarray}
745:
746: Adding all contributions in (\ref{fotqui}), we obtain the photon
747: propagator of the theory \begin{eqnarray} \label{fotonexato1q}
748: i\langle 0|TA_\mu(k)A_\nu(-k)|0 \rangle &=&\\ &
749: &\hspace{-2.5cm}=\;\frac{1}{k^2-\frac{e^2\;a^2}{4\pi\,(a-1) }}
750: \left[g_{\mu\nu}-\frac{k_\mu
751: k_\nu}{a-1}\left(\frac{4\pi}{e^2}-\frac{2}{k^2} \right)
752: +\frac{k_\mu\tilde k_\nu+\tilde k_\mu
753: k_\nu}{(a-1)\;k^2}\right]\nonumber
754: \end{eqnarray}
755:
756: This is equal to the well known results in the literature
757: \cite{Jackiw-Rajaraman}. Again we see that, apart from the
758: regularization of the fermionic loop, there are no perturbatively
759: induced divergences in this propagator. As in the vectorial case,
760: only this regularization is enough to furnish a finite result for
761: the photon propagator.
762:
763: \subsubsection{Fermion propagator}
764: From (\ref{lag}), (\ref{gfunt}) and (\ref{aaqq}), we arrive at the
765: following expression for the fermion propagator \begin{eqnarray}
766: \langle 0|T\psi(x)\overline\psi(y)|0\rangle=P_-G_F(x-y)+\langle
767: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle, \end{eqnarray} with
768: $\psi_+=P_+\psi$. The left-handed fermion propagates freely, but
769: the right-handed one interacts with the vector field $A_\mu$ as
770: \begin{eqnarray} \label{profermq} \langle
771: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle &=&N\int\!\! dA_\mu d\psi
772: d\overline\psi d\theta \; \psi_+(x)\overline\psi_+(y)\\ &
773: &\times\exp\left(i\,S_{\theta}[\psi,\overline\psi,A_\mu,\theta]+
774: \!\int\!\! dz\;\theta(z)j(z,x,y)\right),\nonumber \end{eqnarray}
775: with $S_{\theta}[\psi,\overline\psi,A_\mu,\theta]$ given in the
776: previous section and $j(z,x,y)=\delta(z-x)-\delta(z-y)$.
777: Integrating over the $\theta$ field, we are left with
778: \begin{eqnarray}\label{divqic} \langle
779: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle=\exp\left\{-\frac{4\pi\,i}{a-1}
780: \int\!\!
781: \frac{dk}{(2\pi)^2}\;\frac{1-e^{-ik{\cdotp}(x-y)}}{k^2}\right\}\;
782: G^{\prime}_+(x-y).\quad \end{eqnarray} We observe a logarithmic
783: ultraviolet divergence in the exponential, as in the case of the
784: anomalous Schwinger model. The remaining functional integration
785: $G^{\prime}_+(x-y)$ is finite, \begin{eqnarray}\label{fqpfer}
786: G^{\prime}_+(x-y)&=&N'\int\!\! dA_\mu d\psi d\overline\psi \;
787: \psi_+(x)\overline\psi_+(y)\\ &
788: &\hspace{-1cm}\times\exp\left[i\int\!\! dz \left(\frac{1}{2}A_\mu
789: H^{\mu\nu}
790: A_\nu+\overline\psi(i\partial\!\!\!\slash+eA\!\!\!\slash\,P_+)
791: \psi+eA_\mu
792: l^\mu(z,x,y)\right)\right],\nonumber \end{eqnarray} where
793: \begin{eqnarray}
794: H^{\mu\nu}&=&g^{\mu\nu}\left(\square+\frac{e^2}{4\pi(a-1)}\right)+
795: \left(\frac{1}{\xi}-1\right)\partial^\mu\partial^\nu+\\ & &
796: -\frac{e^2[(a-1)^2+1]}{4\pi(a-1)}
797: \frac{\partial^\mu\partial^\nu}{\square}
798: +\frac{e^2}{4\pi}\frac{\partial^\mu\tilde\partial^\nu+
799: \tilde\partial^\mu\partial^\nu}{\square},\nonumber \end{eqnarray}
800: and
801: \begin{eqnarray} l^\mu(z,x,y)&=&\left(\partial^\mu_z
802: -\frac{\tilde\partial^\mu_z}{a-1} \right)[D_F(z-x)-D_F(z-y)].
803: \end{eqnarray} The $A_\mu-$propagator, which enters in
804: (\ref{fqpfer}), has an ultraviolet behavior as $k^{-2}$. Then, the
805: 1-loop contribution to the fermion self-energy is finite. This
806: persists to all loop orders.
807:
808: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
809:
810: \subsection{Perturbative analysis in GNIF}
811: Here we start from (\ref{ger333}), where
812: \begin{eqnarray}\label{lagmsqni} {\cal
813: L}[\psi,\overline\psi,-\frac{1}{e}\tilde\partial_\mu\phi]&=&
814: \frac{1}{2e^2}\phi\,\square^2\phi+\overline\psi(i\partial\!\!\!\slash-
815: \tilde\partial\!\!\!\slash\phi\,P_+)\psi. \end{eqnarray} Putting
816: $g=e^{i\rho\,P_+}$, we obtain the $\alpha$ term from the Jacobian
817: of this gauge transformation \begin{eqnarray} \label{aaaphaq}
818: \alpha(\rho,\phi)=\exp\left[i\int\!\! dx
819: \left(\frac{(a-1)}{8\pi}\partial_\mu\rho
820: \partial^\mu\rho-\frac{1}{4\pi}\partial_\mu\rho\partial^\mu\phi\right)
821: \right]. \end{eqnarray}
822:
823: \subsubsection{Photon Propagator}
824: From equation (\ref{ger333}), considering (\ref{lagmsqni}) and
825: (\ref{aaaphaq}), we get the photon propagator
826: \begin{eqnarray}\label{photproq} \langle
827: 0|TA_\mu(x)A_\nu(y)|0\rangle&\!\!\!=&\!\!\!
828: \frac{1}{e^2}\left({_l\langle}0|T\partial^x_\mu\rho(x)
829: \partial^y_\nu\rho(y)|0{\rangle_l}
830: - {_l\langle}
831: 0|T\partial^x_\mu\rho(x)\tilde\partial^y_\nu\phi(y)|0{\rangle_l}+
832: \right.\nonumber
833: \\ & &\!\!\!\!\!\!\left.-\, {_l\langle}
834: 0|T\tilde\partial^x_\mu\phi(x)\partial^y_\nu\rho(y) |0{\rangle_l}+
835: {_l\langle} 0|T\tilde\partial^x_\mu\phi(x)\tilde\partial^y_\nu
836: \phi(y)|0{\rangle_l}\right). \end{eqnarray} The dynamics is
837: governed by the effective action
838: $S_{l}[\psi,\overline\psi,\rho,\phi]$, given by
839: \begin{eqnarray}\label{sefefq}
840: S_{l}[\psi,\overline\psi,\rho,\phi]&\!\!\!=&\!\!\!\!\!\!\int\!\!
841: dx\!\left(\frac{1}{2e^2} \phi \square^2 \phi +\overline\psi(i
842: \partial\!\!\!\slash - \tilde\partial\!\!\!\slash\phi )\psi+
843: \frac{(a-1)}{8\pi}\partial_\mu\rho
844: \partial^\mu\rho- \frac{1}{4\pi}\partial_\mu\rho
845: \partial^\mu\phi\right).\nonumber \\
846: & & \end{eqnarray} and $_l\langle\,|\,\rangle_l$ refers to
847: expectation values calculated in this dynamics. We proceed to the
848: perturbative calculation of the relevant graphs.
849:
850: a) $\phi-$propagator \begin{eqnarray}\label{phiprop}
851: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,550][300,650]
852: {w3fig7.ps}}
853: \end{eqnarray} The fermionic loop gives \begin{eqnarray}
854: -\frac{ip^2}{4\pi}(a+1),\nonumber \end{eqnarray}
855: and the third graphic contribution is
856: \begin{eqnarray}
857: -\frac{ip^2}{4\pi(a-1)}. \nonumber
858: \end{eqnarray} The $\phi-$self-energy is \begin{eqnarray}
859: -i\Sigma_\phi(p)=-\frac{i\,a^2\,p^2}{4\pi\,(a-1)}. \end{eqnarray}
860: Now, adding the series (\ref{phiprop}), we obtain \begin{eqnarray}
861: _l\langle 0|T\phi(p)\phi(-p)|0\rangle_l&=&\frac{i\,e^2}{p^2(p^2+
862: \frac{e^2\;a^2}{4\pi\,(a-1)})} \end{eqnarray}
863:
864: b) $\rho-$propagator \begin{eqnarray}
865: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,635][300,650]
866: {w3fig8.ps}}\nonumber
867: \end{eqnarray}
868:
869: \begin{eqnarray} _l\langle
870: 0|T\rho(x)\rho(y)|0\rangle_l&=&\frac{i\,4\pi}{(a-1)p^2}
871: +\frac{1}{(a-1)^2}\,\frac{i\,e^2}{p^2(p^2-
872: \frac{e^2\;a^2}{4\pi\,(a-1)})}\quad
873: \end{eqnarray}
874:
875: c) Mixed terms $\rho-\phi$ \begin{eqnarray}
876: \hspace{-6.24cm}\scalebox{1}{\includegraphics[86,540][300,650]
877: {w3fig9.ps}}\nonumber
878: \end{eqnarray} \begin{eqnarray} \left.\begin{array}{l} {_l\langle}
879: 0|T\rho(p)\phi(-p)|0{\rangle_l}\\ \\ {_l\langle}
880: 0|T\phi(p)\rho(-p)|0{\rangle_l}\end{array} \right\}=
881: \frac{1}{a-1}\frac{i\,e^2}{p^2(p^2-\frac{e^2\;a^2}{4\pi\,(a-1)})}
882: \end{eqnarray}
883: \[ \]
884: Adding all the contributions we obtain \begin{eqnarray}
885: \;\;\langle 0|TA_\mu(k)A_\nu(-k)|0\rangle&=&\\ & &\hspace{-2cm}
886: =\;\frac{-i}{k^2-\frac{e^2\;a^2}{4\pi\,(a-1)}}\left[g_{\mu\nu}-
887: \frac{k_\mu
888: k_\nu}{a-1}\left(\frac{4\pi}{e^2}-\frac{2}{k^2}\right)+
889: \frac{k_\mu\tilde k_\nu+\tilde k_\mu
890: k_\nu}{(a-1)\;k^2}\right]\nonumber
891: \end{eqnarray}
892:
893: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
894:
895: \subsubsection{Fermion propagator}
896: From (\ref{ger333}), we obtain the following expression for the
897: fermionic propagator \begin{eqnarray} \label{profermq1} \langle
898: 0|T\psi(x)\overline\psi(y)|0\rangle &=& i\,P_-G_F(x-y)+\langle
899: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle \end{eqnarray} where
900: \begin{eqnarray} \label{pertfermionq} \langle
901: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle&=&N\int\!\! d\rho d\phi
902: d\psi d\overline\psi\;\;\psi_+(x)\overline\psi_+(y)\\ & &
903: \times\exp\left(iS_{l}[\psi,\overline\psi,\rho,\phi]+i\int\!\!
904: dz\; \rho(z) j(z,x,y)\right).\nonumber \end{eqnarray} After
905: integration over the $\rho$-field, we find the same logarithmic
906: ultraviolet divergence already found in (\ref{divqic}),
907: \begin{eqnarray}\label{divqinc} \langle
908: 0|T\psi_+(x)\overline\psi_+(y)|0\rangle=\exp\left\{-\frac{4\pi\,i}{a-1}
909: \int\!\!
910: \frac{dk}{(2\pi)^2}\;\frac{1-e^{-ik{\cdotp}(x-y)}}{k^2}\right\}\;
911: G^{\prime}_+(x-y).\quad \end{eqnarray} The remaining functional
912: integration in $G^{\prime}_+(x-y)$ involves the following terms
913: \begin{eqnarray}\label{selfqq} G^{\prime}_+(x-y)& =&\!\!\! N'\int\!\!
914: d\phi \,d\psi
915: d\overline\psi\;\;\psi_+(x)\overline\psi_+(y)\;\exp\left[i\int\!\!
916: dz\;\overline\psi(i\partial\!\!\!\slash-\tilde
917: \partial\!\!\!\slash\phi\,P_+)\psi\right]\\ & &
918: \hspace{-1cm}\times\exp\left\{i\int\!\! dz\left[
919: \frac{1}{2e^2}\phi\square(\square+\frac{e^2}{4\pi(a-1)})\phi+
920: \frac{1}{a-1}\phi(z)j(z,x,y)\right]\right\}.\nonumber
921: \end{eqnarray} From (\ref{selfqq}) we easily see that the 1-loop
922: contribution to the fermion self-energy is finite, as well as the
923: contribution of the other loops. The ultraviolet divergence is
924: entirely due to the longitudinal component of the photon, the
925: $\rho$ field.
926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
927: \section{Conclusions}
928:
929: We demonstrated the completely non-perturbative origin of
930: divergences that occur in fermionic correlation functions in two
931: dimensional massless Quantum Electrodynamics. This has been done
932: explicitly, either by summing (wherever it was possible to do it)
933: the perturbative series or by giving arguments that showed the
934: finiteness of individual terms. It resulted clear that the
935: divergences are a consequence of the lack of gauge invariance (at
936: least in intermediate steps) and that their structure is largely
937: independent of the fact that the anomaly is a genuine one (as is
938: the case for the chiral Schwinger model) or an artifact of
939: regularization (as in the Schwinger model considered under a
940: general regularization, not necessarily preserving gauge
941: invariance). The only difference between the two cases is that, in
942: the second case, the divergences could be circumvented by choosing
943: a regularization that conducted to $a=1$ (preservation of gauge
944: invariance). Apart from this fact (which has its justification
945: only on simplicity, not reflecting any fundamental principle of
946: quantum field theory) there is no reason for choosing one or
947: another value for $a$ as, in the end, the effective action (the
948: one obtained after integration over the fermions and the
949: Wess-Zumino fields (GIF) or the longitudinal part of the gauge
950: field (GNIF)) is gauge invariant \cite{HaradaT}.
951:
952: Moreover, we performed this demonstration both in GIF and GNIF and
953: found equal results. Although this may seem to be no surprise, as
954: we were merely effecting the same integral by different means, it
955: contradicts what is said by Jian-Ge, Qing-Hai and Yao-Yang in
956: \cite{chineses-JP}. In their paper, the authors find that, when
957: they add the Wess-Zumino term to the original action, there is no
958: divergence in fermionic Green functions, as opposed to what they
959: obtain in its absence, and they justify this exhibiting a
960: different ultraviolet behaviour of the photon propagator in the
961: two approaches. In fact, they missed a crucial point in their
962: paper: they added the Wess-Zumino term {\it before} the
963: introduction of the external sources, what is {\it wrong}. If one
964: does this, one does not obtain the coupling between the fermions
965: and the Wess-Zumino fields, and it is no surprise if no divergence
966: appears. Also, they obtained two different photon propagators, one
967: in the GIF and other in the GNIF, because they lost an additional
968: coupling term between the photon and the Wess-Zumino field, that
969: comes from the Faddeev-Popov procedure, and that is crucial for
970: the final expressions for the photon propagators. The correct
971: expression for the generating functional in GIF is given by our
972: equation (\ref{gfunt}).
973:
974: We would like to remember that the generating functional has to be
975: a functional of something, so the starting expression has to
976: include the external sources. This is just one instance where this
977: kind of mistake can conduct one to completely wrong results, and
978: it is not usually noticed (for an example of the crucial role
979: played by external sources see \cite{Tiao-Teresa}). If the
980: external sources are carefully considered from the beginning, one
981: finds exactly the same results in both formalisms.
982:
983: The physical interpretation of this new type of divergence is
984: still unknown for us. The origin of conventional ultraviolet
985: divergences can be traced back to the requirement of relativistic
986: covariance and {\it non-triviality} of the field theory under
987: investigation \cite{strocchi-LNP}. This prevents a good definition
988: for the field as an operator for all points of the support,
989: introducing divergences when its powers appear. They manifest
990: themselves in the diagonal of Green's functions, thus allowing
991: themselves to be renormalized through the well known ambiguity of
992: these diagonals under local integrated polinomials in the fields
993: (counterterms). In the new situation, although the cure may be
994: similar (but significantly different \cite{tiao-rodolfo}), the
995: disease may not be the same. The new divergence that appeared
996: multiplies an effective (and finite) two-point fermionic function.
997: We called it {\it ultraviolet} just because of its form (look at
998: the divergent integral appearing in (\ref{petfer22}), for example)
999: but not because of its origin. After all, it has its origins in an
1000: integration over a quadratic portion of the action, which would
1001: contribute with linear terms in the equations of motion, not
1002: usually associated to divergences (they do not involve products of
1003: operators in the same point). There are evident connections
1004: between this divergence and the lack of gauge invariance, but they
1005: do not help to clarify the situation, as opposed to what was said
1006: above, about conventional UV divergences. Further investigation on
1007: this question is being done, and will be reported elsewhere.
1008: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1009:
1010: \begin{thebibliography}{99}
1011:
1012: \bibitem{Abers-Lee} E. S. Abers and B. W. Lee, {\it Phys. Rep.},
1013: {\bf C9} (1973), 1.
1014:
1015: \bibitem{t'Hooft} G. 't Hooft, {\it Nucl. Phys.}, {\bf B33}
1016: (1971), 173; {\bf B35} (1971), 167.
1017:
1018: \bibitem{Jackiw-Rajaraman} R. Jackiw and R. Rajaraman,
1019: {\it Phys. Rev. Lett.}, {\bf 54} (1985), 1219.
1020:
1021: \bibitem{FaddeevS} L. D. Faddeev and S. L. Shatashvili,
1022: {\it Phys. Lett.} {\bf B167}, (1986) 225.
1023:
1024: \bibitem{SchapoV} O. Babelon, F.A. Schaposnik and C.M. Viallet
1025: {\it Phys. Lett.} {\bf B177} (1986), 385.
1026:
1027: \bibitem{HaradaT} K. Harada and I. Tsutsui, {\it Phys. Lett.}
1028: {\bf B183}, (1987), 311.
1029:
1030: \bibitem{Abdalla} E. Abdalla, M. C. B. Abdalla and K. D. Rothe,
1031: {\it Non-Perturbative Methods in 2 Dimensional Quantum Field
1032: Theory}, World Scientific, 1991.
1033:
1034: \bibitem{TDKieu} T. D. Kieu, hep-th/9409198; UM-P-94/58; {\it Mod.
1035: Phys. Lett.} {\bf A5}, (1990), 175; {\it Phys. Lett.} {\bf B223},
1036: (1989), 72; {\it Phys. Lett.} {\bf B218}, (1989), 221;
1037:
1038: \bibitem{LinharesR} C. A. Linhares, K. D. Rothe and H. D. Rothe,
1039: {\it Phys. Rev.} {\bf D35}, (1987) 2501.
1040:
1041: \bibitem{GirottiRR} H. O. Girotti, H. J. Rothe and K. D. Rothe,
1042: {\it Phys. Rev.} {\bf D34}, (1986), 592.
1043:
1044: \bibitem{Boianovsky} D. Boyanovisky, {\it Nucl. Phys.} {\bf B294},
1045: (1987), 223.
1046:
1047: \bibitem{Chineses} Z. Jian-Ge, D. Qing-Hai and L. Yao-Yang,
1048: {\it Phys. Rev.} {\bf D43} (1991), 613.
1049:
1050: \bibitem{tiao-rodolfo} R. Casana and S. A. Dias, {\it Int. Jour. Mod. Phys.}
1051: {\bf A}, Vol. 15, No. 29 (2000), 4603.
1052:
1053: \bibitem{chineses-JP} Z. Jian-Ge, D. Qing-Hai and L. Yao-Yang,
1054: {\it J. Phys.} {\bf G17} (1991), L7.
1055:
1056: \bibitem{Zinn-Justin} J. Zinn-Justin, {\it Quantum Field Theory and
1057: Critical Phenomena}, second edition, Oxford Science Pub., 1993.
1058:
1059: \bibitem{HaradaN} K. Harada, {\it Nucl.Phys.} {\bf B329} (1990), 723.
1060:
1061: \bibitem{Dias-Linhares} S. A. Dias and C. A. Linhares {\it Phys. Rev.}
1062: {\bf D45} (1992), 2162.
1063:
1064: \bibitem{Mitra-Rahaman} P. Mitra and A. Rahaman, {\it Ann. Phys.}
1065: {\bf 249} (1996), 34.
1066:
1067: \bibitem{tiaomarcelo} S. A. Dias and M. B. Silva Neto,
1068: {\it Notas de F\'{\i}sica} CBPF NF-007/96; hep-th 9602092.
1069:
1070: \bibitem{Tiao-Teresa} S. A. Dias and M. T. Thomaz, {\it Phys. Rev.}
1071: {\bf D44} (1991), 1811.
1072:
1073: \bibitem{strocchi-LNP} F. Strocchi, {\it Selected Topics on the
1074: General Properties of Quantum Field Theory}, Lecture Notes in
1075: Physics, vol. 51, 1993, World Scientific Publishing Company,
1076: Singapore.
1077:
1078: \end{thebibliography}
1079: \end{document}
1080: