hep-ph0103285/ny.tex
1: \documentstyle[aps,prl,preprint,epsf,floats,psfig]{revtex}
2: \tightenlines
3: \parindent=20pt
4: \parskip=10pt
5: \pagestyle{plain}
6: \font\tenrm=cmr10
7: \def\square{\vcenter{\vbox{\hrule height.3pt
8:           \hbox{\vrule width.3pt height6pt
9:           \kern6pt\vrule width.3pt}\hrule height.3pt}}}
10: \def\boxx{\square}
11: 
12: \def\sumint{\hbox{$\sum$}\!\!\!\!\!\!\int}
13: \newcommand{\beq}{\begin{equation}}
14: \newcommand{\eeq}{\end{equation}}
15: \newcommand{\bqa}{\begin{eqnarray}}
16: \newcommand{\eqa}{\end{eqnarray}}
17: \newcommand{\Log}[2]{ \log{ {{#1}+{#2}\over{#1}-{#2}} } }
18: \newcommand{\LogTwo}[2]{ {\rm log}^2 {{#1}+{#2}\over{#1}-{#2}}  }
19: 
20: \begin{document}
21: 
22: \preprint{
23: \vbox{\halign{&##\hfil\cr
24: &ITF-UU-01/10\cr
25:         & hep-ph/yymmnn \cr
26: &\today\cr }}}
27: 
28: 
29: 
30: \title{Thermal Effects in Low-Temperature QED}
31: \author{Jens O. Andersen}
32: \address{Institute for Theoretical Physics, University of Utrecht,\\
33:        Leuvenlaan 4, 3584 CE Utrecht, The Netherlands}
34: 
35: \maketitle
36: 
37: 
38: 
39: \begin{abstract}
40: {\footnotesize 
41: QED is studied at low temperature ($T\ll m$, where $m$ is the electron mass) 
42: and zero chemical potential. 
43: By integrating out the electron field and the nonzero bosonic Matsubara modes,
44: we construct an effective three-dimensional field theory that is
45: valid at distances $R\gg1/T$.
46: As applications, we reproduce the ring-improved free energy
47: and calculate the Debye mass to order $e^5$.
48: }\end{abstract}
49: 
50: \newpage
51: 
52: 
53: \section{Introduction}
54: If we have a quantum field theory in equilibrium at 
55: temperature $T$, the abundance of particles of mass $m$ much larger than
56: $T$ is Boltzmann suppressed. More surprisingly, perhaps, is the fact that
57: there are additional effects that are suppressed only by powers of 
58: $T/m$~\cite{barton,finn,bj}.
59: The Boltzmann-suppressed terms can be associated with loop
60: integrations that involve distribution functions of the heavy particle.
61: On the other hand, the power-suppressed terms arise from perturbative
62: corrections involving only distribution functions of light particles, with 
63: masses on the order of $T$ or less.
64: 
65: 
66: Effective field theory ideas suggest that one integrates out the heavy
67: particle of mass $m$ to construct a low-energy effective field theory
68: that can be used for $T\ll m$~\cite{lepage,kaplan}. 
69: Such an approach was used by Kong and 
70: Ravndal~\cite{finn} to study QED at low temperature
71: (see also Refs.~[6-13] and Refs. therein
72: for various calculations in low-temperature QED). 
73: By integrating out
74: the electron field, they constructed a low-energy effective field theory
75: for photons. Since this procedure was carried out at zero temperature,
76: dimensional analysis tells us that the coefficients in the effective theory
77: are suppressed by inverse powers of the electron mass $m$.
78: These higher order interactions are induced by the coupling of the photon
79: to virtual electron-positron pairs in the vacuum. 
80: Since the momenta of the photons are on the order $T$ and thus
81: much smaller than $m$, the electron-positron pairs are off their mass shell
82: by an amount $\sim m$. Thus they can only propagate a distance $R\sim1/m$
83: and their effects can be mimicked by local interactions.
84: Using this effective field theory to calculate corrections to the 
85: Stefan-Boltzmann law for the pressure, 
86: they showed that the leading correction goes like $\alpha^2T^8/m^4$.
87: This correction was obtained by a straightforward
88: two-loop calculation in the effective theory.
89: In full QED, it would require a three-loop calculation.
90: 
91: If we are interested only in power corrections, we can determine the
92: coefficients in the effective field theory by matching at zero temperature.
93: In the case of QED, this would lead to the Euler-Heisenberg 
94: Lagrangian~\cite{uh,eh}
95: with additional higher order operators that can be written in terms of the
96: field strength and its dual~\cite{finn} (see e.g.~\cite{dicus}
97: for such higher order operators). 
98: However, it can be shown that this effective 
99: Lagrangian leads to a vanishing Debye mass to all orders in perturbation 
100: theory. We know that this is incorrect, but 
101: one can account for Debye screening and other Boltzmann-suppressed
102: effects in low-temperature QED by carrying out the
103: matching at finite temperature. Alternatively, one may reorganize the
104: usual perturbative series in QED using resummed propagators in the usual
105: way. However, such an approach is normally more cumbersome in
106: practical calculations than effective field theory.
107: 
108: If we are interested in static quantities such as the pressure or Debye
109: mass, it proves useful to construct a second effective field theory 
110: for the zero Matsubara mode and this is done by integrating out the nonzero 
111: Matsubara modes\cite{landsman,gins,BN,kaja}.
112: This effective
113: field theory is three-dimensional
114: and is valid at distances $R\gg 1/T$.
115: It can be
116: constructed in a two-step process by first integrating out the
117: electron field, and then integrating out the nonzero bosonic Matsubara modes.
118: From a calculational point of view, however, it is easier to integrate
119: out the fermions and the nonzero Matsubara modes at the same time. 
120: In this paper, we will take the latter approach.
121: 
122: The paper is organized as follows. 
123: In section II, we discuss
124: QED at low temperature
125: and the construction of the three-dimensional effective field
126: theory.
127: In section III, we determine the coefficients in the effective field theory.
128: In section IV, we apply the effective field theory to calculate the
129: the free energy to order $e^3$ and the Debye mass to order $e^5$.
130: Finally, in section V, we summarize and
131: conclude. All necessary details are collected in three appendices.
132: 
133: \section{QED at Low Temperature}
134: In the imaginary-time formalism, one can view a quantum field theory in four 
135: dimensions as a field theory in three Euclidean dimensions with an 
136: infinite tower of fields, where the Matsubara frequencies act as masses
137: in the propagators~\cite{landsman}. 
138: In low-temperature QED, this implies that the fermions
139: have masses of order $m$, while the nonzero bosonic modes have masses
140: of order $T$. The zero-frequency bosonic modes are massless.
141: Thus for distances $R\gg1/T$,
142: we can construct an effective three-dimensional field theory
143: for the zero Matsubara modes by integrating out the electron field
144: as well as the 
145: nonzero bosonic modes~\cite{landsman,gins,BN,kaja}.
146: The coefficients of this effective field theory then encode the physics 
147: at the momentum scales $m$ and $T$.
148: This procedure is briefly explained in section III.
149: 
150: The partition function of QED can be written as a path integral
151: \bqa
152: {\cal Z}=\int{\cal D}\bar{\eta}\;{\cal D}\eta\;
153: {\cal D}A_{\mu}\;{\cal D}\bar{\psi}\;{\cal D}\psi\;
154: \exp{\left[-\int_0^{\beta}d\tau\int d^3x\;{\cal L}\right]}\;,
155: \eqa
156: where the Euclidean Lagrangian is
157: \bqa
158: {\cal L}_E=\frac{1}{4}F_{\mu\nu}F_{\mu\nu}+m\overline{\psi}\psi
159: +\overline{\psi}
160: \gamma_{\mu}\Big (\partial_{\mu}-ieA_{\mu} \Big )\psi
161: +\left(\partial_{\mu}\bar{\eta}\right)\left(\partial_{\mu}\eta\right)
162: +{\cal L}_{\mbox{\footnotesize gf}}\;.
163: \eqa
164: Here, ${\cal L}_{\mbox{\footnotesize gf}}$ denotes the gauge-fixing
165: term. In the following we work in Feynman gauge, where
166: \bqa
167: {\cal L}_{\mbox{\footnotesize gf}}=
168: \frac{1}{2}(\partial_{\mu}A_{\mu})^{2}\;.
169: \eqa
170: However, we emphasize that physical quantities 
171: are independent of the gauge-fixing condition.
172: 
173: In the three-dimensional effective theory, we can write the partition function
174: as
175: \bqa
176: \label{efff}
177: {\cal Z}=e^{-f(\Lambda)V}\int{\cal D}\bar{\eta}\;{\cal D}\eta\;
178: {\cal D}\bar{A}_0\;{\cal D}\bar{A}_{i}\;
179: \exp{\left[-\int d^3x\;{\cal L}^{}_{\rm eff}\right]}\;,
180: \eqa
181: where the prefactor $f(\Lambda)$ is interpreted as the 
182: coefficient of the unit operator in the effective three-dimensional field 
183: theory. It depends on an ultraviolet cutoff $\Lambda$ that cancels
184: the cutoff-dependence in the path integral in Eq.~(\ref{efff})~\cite{BN}.
185: ${\cal L}_{\rm eff}$ is the Lagrangian of the 
186: effective three-dimensional 
187: field theory. The effective field theory 
188: consists of a gauge field $\bar{A}_i$ coupled to a real massive
189: self-interacting scalar field $\bar{A}_0$
190: \footnote{The fact that we must allow for terms such 
191: as~${1\over2}M^2(\Lambda)\bar{A}_0^2$ 
192: and ${\lambda_3(\Lambda)\over24}\bar{A}_0^4$, is a direct consequence
193: of the breakdown of Lorentz invariance at finite temperature.}. 
194: These fields 
195: can up to normalizations
196: be identified with the zero-frequency modes of the gauge field in QED.
197: We can then schematically write
198: \bqa
199: \label{eff2}
200: {\cal L}_{\rm eff}^{}&=&\frac{1}{4}F_{ij}F_{ij}+{1\over4}a_3(\Lambda)
201: F_{ij}\nabla^2F_{ij}+
202: {1\over2}(\partial_i\bar{A}_0)^2+{1\over2}M^2(\Lambda)\bar{A}_0^2
203: +{\lambda_3(\Lambda)\over24}\bar{A}_0^4+
204: \left(\partial_i\bar{\eta}\right)\left(\partial_i\eta\right) \\ \nonumber
205: &&
206: +{\cal L}_{\rm gf}+\delta{\cal L}^{}_{\rm eff}\;,
207: \eqa
208: where $\delta{\cal L}^{}_{\rm eff}$ represents all 
209: higher order local terms that can be 
210: constructed out of $\bar{A}_i$ and $\bar{A}_0$ 
211: and respect the symmetries of the theory.
212: Examples of such symmetries are three-dimensional
213: gauge invariance and rotational symmetry. This includes renormalizable
214: terms such as $F_{ij}^2$ as well as nonrenormalizable terms such as
215: $\bar{A}_0^6$. 
216: Note also that we have suppressed the $\Lambda$-dependence of the fields
217: $\bar{A}_i$ and $\bar{A}_0$ in Eq.~(\ref{eff2}).
218: Let us finally look at the power-counting rules for the effective theory
219: Eq.~(\ref{eff2})
220: The coefficients of the operators are power series in $e^2$ 
221: since we are ignoring infrared divergences and are using conventional
222: perturbation theory to determine them (see Sec.~\ref{sh}).
223: Physical quantities are expressed in powers of
224: the parameters $f(\Lambda)$, $a_3(\Lambda)$, $M(\Lambda)$,... and we must 
225: figure out
226: at which order the operators that multiply them start to contribute.
227: Each power of momentum in a loop integral in the effective theory
228: gives a factor of $M$, in particular the measure gives a factor $M^3$.
229: If we want to calculate the free energy to order $e^3$, it is necessary
230: to determine $f(\Lambda)$ to order $e^2$. Moreover, the one-loop
231: contribution to the free energy in the effective theory is proportional
232: to $M^3$, and we therefore need to know the mass parameter to
233: order $e$. 
234: (Note that the operator $F_{ij}\nabla^2F_{ij}$
235: does not contribute to the free energy
236: at order $e^2$ since it involves only massless fields
237: whose loop integral vanishes in dimensional regularization. In fact, this
238: operator can be transformed away by a field redefinition at the expense
239: of modifying the coeffcients of higher order operators.).
240: On the other hand, the operator $\bar{A}_0^4$ starts to
241: contribute first at order
242: $e^6$; its coefficient $\lambda_3(\Lambda)$ goes like $e^4$ and 
243: it gives rise to a two-loop diagram in the effective theory where each
244: loop is proportional to $M$. 
245: In this manner we can determine at which order in $e$ an operator starts to
246: contribute to a given physical quantity, and in Eq.~(\ref{eff2})
247: we have explicitly displayed those operators needed to determine the
248: free energy to order $e^3$ and the Debye mass to order $e^5$.
249: 
250: 
251: 
252: \section{Short-distance Coefficients}\label{sh}
253: In this section, we determine the short-distance coefficients in the 
254: effective Lagrangian Eq.~(\ref{eff2}). These coefficients must be tuned
255: as functions of $e$, $T$, and the ultraviolet cutoff $\Lambda$ so
256: that the effective theory reproduces correlation functions at distances much 
257: larger than $1/T$. We can carry out these 
258: calculations using conventional perturbation theory, which is an expansion
259: in powers of $e^2$. This expansion is plagued with infrared 
260: divergences due to long-range forces mediated by the massless photon.
261: These divergences are Debye screened, but can only be taken into account
262: by resummation. Although the naive perturbative expansion breaks down
263: due to these infrared divergences,
264: it can still be used to determine the short-distance coefficients~\cite{BN}.
265: As long as we treat the long-distance physics in the same incorrect way
266: using the effective theory, the infrared divergences will cancel in the
267: matching equations and the coefficients properly encode the short-distance
268: physics~\cite{BN}.
269: The Lagrangian of QED is split the usual way 
270: into free and interacting pieces 
271: \bqa
272: {\cal L}_E^{\rm free}&=&\frac{1}{4}F_{\mu\nu}F_{\mu\nu}
273: +\overline{\psi}\left(m+
274: /\!\!\!\partial 
275: \right)\psi
276: +\left(\partial_{\mu}\bar{\eta}\right)\left(\partial_{\mu}\eta\right)
277: +{\cal L}_{\mbox{\footnotesize gf}}\;,
278: \\
279: {\cal L}_E^{\rm int}&=&-ie/\!\!\!\!A 
280: \overline{\psi}\psi\;,
281: \eqa
282: while the Lagrangian Eq.~(\ref{eff2}) is split according to 
283: \bqa
284: {\cal L}_{\rm eff}^{\rm free}&=&
285: {1\over4}F_{ij}F_{ij}+{1\over2}(\partial_i\bar{A}_0)^2
286: +\left(\partial_{\mu}\bar{\eta}\right)\left(\partial_{\mu}\eta\right)
287: +{\cal L}_{\rm gf}
288: \;,
289: \\
290: {\cal L}_{\rm eff}^{\rm int}&=&
291: {1\over4}a_3(\Lambda)
292: F_{ij}\nabla^2F_{ij}
293: +{1\over2}M^2(\Lambda)
294: \bar{A}_0^2+{\lambda_3(\Lambda)\over24}
295: \bar{A}_0^4
296: +\delta{\cal L}_{\rm eff}^{}\;.
297: \eqa
298: \subsection{Coefficient of the Unit Operator}
299: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300: \begin{figure}[htb]
301: \begin{center}
302: \mbox{\psfig{figure=q1.ps}}%,width=12cm,height=11cm}}
303: \end{center}
304: \caption[One-loop vacuum diagrams in QED.]{\protect One-loop vacuum diagrams in QED.}
305: \label{q1}
306: \end{figure}
307: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
308: \begin{figure}[htb]
309: \begin{center}
310: \mbox{\psfig{figure=q2.ps}}%,width=12cm,height=11cm}}
311: \end{center}
312: \caption[Two-loop vacuum diagram in QED.]{\protect Two-loop vacuum diagram  in QED.}
313: \label{q2}
314: \end{figure}
315: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
316: The matching condition that determines the coefficient of the unit 
317: operator is~\cite{BN}
318: \bqa
319: \log{\cal Z}=-f(\Lambda)V+\log{\cal Z}_{\rm eff}^{}\;.
320: \eqa
321: The contributions to $\log{\cal Z}$ through order $e^2$ is given
322: by the Feynman graphs in Figs.~\ref{q1} and~\ref{q2}.
323: A wavy line denotes a photon, a solid line denotes a fermion, 
324: and a dotted line denotes a ghosts.
325: The expression is
326: \bqa\nonumber
327: {T\log{\cal Z}\over V}&\approx&-{1\over2}(d-1)\sumint_{P}\log P^2+
328: 2\sumint_{\{P\}}\log(P^2+m^2)
329: \\
330: &&
331: \label{two}
332: +
333: {1\over2}e^2\sumint_{\{PQ\}}\mbox{Tr}
334: \left[\gamma_{\mu}{P\!\!\!\!/-m\over P^2+m^2}
335: \gamma_{\mu}{Q\!\!\!\!/-m\over Q^2+m^2}
336: {1\over (P+Q)^2}
337: \right]\;,
338: \eqa
339: where the trace is over Dirac indices.
340: The sign $\approx$ is a reminder that an equality only holds
341: in strict perturbation theory.
342: The fermionic one-loop diagram has a pole in $\epsilon$, which is 
343: proportional to $m^4$. This pole is cancelled
344: by the one-loop vacuum counterterm $\Delta_1{\cal E}_0$. 
345: The two-loop diagram also has 
346: poles in $\epsilon$. The temperature-independent pole is cancelled
347: by the two-loop vacuum counterterm $\Delta_2{\cal E}_0$, while the
348: temperature-dependent ones are cancelled by the one-loop counterterm
349: diagrams. The two-loop sum-integrals is briefly discussed in appendix A. 
350: Our renormalization prescription is that the renormalized
351: vacuum energy vanishes at 
352: the scale $\Lambda$. Thus the fermionic one-loop diagram and the two-loop
353: diagram are given by their finite-temperature pieces
354: (See Eqs.~(\ref{fermip}) and~(\ref{fermi2}) in appendix A)
355: After renormalization,
356: the resulting expression in the low-temperature limit is
357: \bqa
358: \label{free1}
359: {T\log{\cal Z}\over V}&\approx&{\pi^2T^4\over 45}
360: +{4\over(2\pi)^{3/2}}m^{3/2}T^{5/2}e^{-m/T}
361: +{e^2\over2(2\pi)^3}m^{2}T^2e^{-2m/T}
362: \;.
363: \eqa
364: In the effective theory, all loop diagrams vanish with dimensional
365: regularization, since there is no mass scale in the 
366: corresponding integrals
367: Hence, $\log{\cal Z}_{\rm eff}$ vanishes identically in strict perturbation
368: theory.
369: The matching equation then implies 
370: \bqa
371: f(\Lambda)=-{\log{\cal Z}\over V}
372: \;,
373: \eqa
374: where the right hand side is given by Eq.~(\ref{free1}).
375: \subsection{Field Normalization Constant}
376: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
377: \begin{figure}[htb]
378: \begin{center}
379: \mbox{\psfig{figure=qs1.ps}}%,width=12cm,height=11cm}}
380: \end{center}
381: \caption[One-loop polarization tensor in QED.]{\protect One-loop polarization tensor in QED.}
382: \label{pollen}
383: \end{figure}
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: %
386: The field normalization constants for $A_i$ and $A_0$ 
387: are obtained by reading off the coefficients of
388: $\delta_{ij}k^2-k_ik_j$ and $k^2$ 
389: of $\Pi_{ij}^{(1)}(0,{\bf k})$ and $\Pi_{00}^{(1)}(0,{\bf k})$,
390: where $\Pi_{\mu\nu}^{(1)}(k_0,{\bf k})$ is the 
391: one-loop polarization tensor~(\ref{pol}).
392: These are given in Eqs.~(\ref{space}) and~(\ref{stat}) 
393: in appendix C, and we obtain
394: \bqa
395: \label{i}
396: \bar{A}_i(\Lambda)&\approx&{1\over\sqrt{T}}
397: \left[1+{4\over3}e^2\sumint_{\{P\}}{1\over(P^2+m^2)^2}\right]^{1/2}A_i
398: \;,\\ 
399: \label{0}
400: \bar{A}_0(\Lambda)&\approx&{1\over\sqrt{T}}
401: \left[1+{1\over3}e^2\sumint_{\{P\}}
402: {6\over(P^2+m^2)^2}-{8p_0^2\over(P^2+m^2)^3}\right]^{1/2}A_0\;.
403: \eqa
404: Eqs.~(\ref{i}) and~(\ref{0}) have poles in $\epsilon$ that are
405: cancelled by the wave function renormalization
406: counterterm 
407: \bqa
408: Z_{A}=1-{e^2\over12\pi^2\epsilon}\;.
409: \eqa
410: After renormalization, we obtain
411: \bqa
412: \label{redef2}
413: \bar{A}_i(\Lambda)
414: &\approx&{1\over\sqrt{T}}\left[1+{e^2\over24\pi^2}\left(L-J_2\right)\right]A_i\;, \\
415: %\label{redef1}
416: \bar{A}_0(\Lambda)&\approx&{1\over\sqrt{T}}\left[1+{e^2\over24\pi^2}
417: \left(
418: L-J_2-J_3m^2T^{-2}
419: \right)
420: \right]A_0\;.
421: \eqa
422: where $L=\log\left({\Lambda^2\over m^2}\right)$ and the integrals
423: $J_n$ are defined in appendix A.
424: Note the different normalizations of the fields $A_0$ and $A_i$. 
425: In the low-temperature limit, this reduces to
426: \bqa
427: \bar{A}_i(\Lambda)&\longrightarrow&{1\over\sqrt{T}}\left[1+{e^2\over24\pi^2}
428: \left(L-2(2\pi)^{1/2}m^{-1/2}T^{1/2}e^{-m/T}\right)\right]A_i\;, \\
429: %\label{redef1}
430: \bar{A}_0(\Lambda)&\longrightarrow&{1\over\sqrt{T}}\left[1+{e^2\over24\pi^2}
431: \left(
432: L-(2\pi)^{1/2}m^{1/2}T^{-1/2}e^{-m/T}
433: \right)
434: \right]A_0\;.
435: \eqa
436: 
437: \subsection{Mass Parameter}
438: The mass parameter can be determined in several ways. We determine it by
439: demanding that the screening mass in strict perturbation theory be the
440: same in QED and the effective theory. 
441: The screening mass $m_s$ is defined as the location of the pole in the 
442: propagator
443: for spacelike momentum~\cite{BN}:
444: \bqa
445: \label{scr}
446: k^{2}+\Pi_{00} (0,{\bf k})=0,\hspace{1cm}k^{2}=-m^{2}_{s}\;.
447: \eqa In the effective theory, we have
448: \bqa
449: k^{2}+M^{2}(\Lambda)+\Pi_{\rm eff}(k,\Lambda )=0,
450: \hspace{1cm}k^{2}=-m^{2}_{s}\;,
451: \eqa
452: where $\Pi_{\mbox{\footnotesize eff}}(k,\Lambda )$
453: is the self-energy of $\bar{A}_0$ in the effective theory. 
454: We can expand the self-energy function $\Pi_{00}(0,{\bf k})$ in powers
455: of the external momentum ${\bf k}$. To second order in the loop expansion, 
456: the solution to Eq.~(\ref{scr}) for the screening mass is~\cite{Andersen}
457: \bqa
458: m_s^2\approx\left[1-\Pi_{00}^{(1)\prime}(0,0)
459: \right]\Pi_{00}^{(1)}(0,0)+\Pi_{00}^{(2)}(0,0)
460: \;.
461: \eqa
462: Here $\Pi_{00}^{(n)}$ denotes the nth order contribution to the static
463: polarization tensor in the loop expansion,
464: and the prime denotes differentiation with respect to $k^2$.
465: The self-energy function $\Pi_{\mbox{\footnotesize eff}}(k,\Lambda )$
466: can also be expanded in a Taylor series around ${\bf k}=0$. The 
467: corresponding loop integrals are evaluated at ${\bf k}=0$ and since there
468: is no other mass scale, the self-energy function $\Pi_{\rm eff}(0,\Lambda)$
469: vanishes identically
470: in strict
471: perturbation theory. Thus the matching condition reduces to
472: \bqa
473: M^2(\Lambda)\approx m_s^2\;.
474: \eqa 
475: The one-loop contribution to $\Pi_{00}(0,0)$ is given by 
476: by the first term in Eq~(\ref{stat}), while
477: $\Pi_{00}^{\prime}(0,0)$ is given by 
478: the second term in Eq.~(\ref{stat}).
479: The two-loop contribution to $\Pi_{00}(0,0)$ is given by Eq.~(\ref{twopol}),
480: and in appendix B, we explain how one can obtain the expression for it from
481: the two-loop contribution to the free energy.
482: In the low-temperature limit, the mass parameter becomes 
483: \bqa
484: M^2(\Lambda)={4e^2\over(2\pi)^{3/2}}m^{3/2}T^{1/2}e^{-m/T}
485: -{4e^4\over3(2\pi)^{7/2}}m^{3/2}T^{1/2}Le^{-m/T}
486: +{10e^4\over3(2\pi)^3}m^2e^{-2m/T}
487: \;.
488: \eqa
489: Using the renormalization group equation for the running gauge coupling,
490: \bqa
491: \label{rg}
492: \mu{de^2\over d\mu}={e^4\over6\pi^2}\;,
493: \eqa
494: we see that the mass parameter
495: is independent of the renormalization scale 
496: $\mu$ to order $e^4$.
497: 
498: \subsection{Coupling Constants}
499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
500: \begin{figure}[htb]
501: \begin{center}
502: \mbox{\psfig{figure=four.ps}}%,width=12cm,height=11cm}}
503: \end{center}
504: \caption[Four-point function with external timelike photons at one-loop in QED.]{Four-point function with external timelike photons at one-loop in QED.}
505: \label{2lqed}
506: \end{figure}
507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
508: 
509: The one-loop Feynman diagram for the four-point
510: function $\Gamma^{(4)}$ with external
511: time-like photons is shown in
512: Fig~\ref{2lqed}. The matching equation is simply~\cite{Andersen}
513: \bqa
514: \lambda_3(\Lambda)=T\Gamma^{(4)}(0,0,0,0)\;,
515: \eqa
516: where the arguments indicate that the loop diagram is to be evaluated at
517: zero external momenta
518: The expression for the diagram is
519: \bqa\nonumber
520: \Gamma^{(4)}(0,0,0,0)&=&
521: 6e^2\sumint_{\{P\}}\mbox{Tr}
522: \Bigg[
523: \gamma_0{P\!\!\!\!/-m\over P^2+m^2}
524: \gamma_0{P\!\!\!\!/-m\over P^2+m^2}
525: \gamma_0{P\!\!\!\!/-m\over P^2+m^2}
526: \gamma_0{P\!\!\!\!/-m\over P^2+m^2}
527: \Bigg]\\
528: \label{g4}
529: &=&
530: 24e^2\sumint_{\{P\}}
531: \Bigg[
532: m^4+2m^2P^2-8m^2p_0^2+P^4-8p_0^2P^2+8p_0^4
533: \Bigg]{1\over(P^2+m^2)^4}\;.
534: \eqa
535: The particular combination of sum-integrals in Eq.~(\ref{g4}) 
536: is finite with dimensional
537: regularization and we obtain
538: \bqa
539: \label{lambda}
540: \lambda_3(\Lambda)=
541: -{8e^4\over(2\pi)^2}
542: J_4m^4T^{-3}\;.
543: \eqa
544: The fact that this coefficient vanishes identically
545: at zero temperature, simply
546: reflects the gauge invariance of QED.
547: In the low-temperature limit, Eq.~(\ref{lambda}) reduces to
548: \bqa
549: \lambda_3(\Lambda)=-{4e^4\over(2\pi)^{3/2}}m^{3/2}T^{-1/2}
550: e^{-m/T}\;.
551: \eqa
552: 
553: The coupling constant $a_3$ is not affected by the field redefinition
554: Eqs.~(\ref{redef2}) to leading order in $e$, 
555: so its value is given directly by the
556: coefficient of the last term in Eq.~(\ref{space}) 
557: \bqa
558: a_3(\Lambda)&=&{8\over15}{e^2}\sumint_{\{P\}}
559: {1\over(P^2+m^2)^3}
560: \\
561: \label{a}
562: &=&{e^2\over60\pi^2m^2}\left[1-m^2T^{-2}J_3\right]\;.
563: \eqa
564: In the low-temperature limit, this reduces to
565: \bqa
566: a_3(\Lambda)={e^2\over60\pi^2m^2}\left[1-(2\pi)^{1/2}m^{1/2}T^{-1/2}
567: e^{-m/T}\right]\;.
568: \eqa
569: The first term is the standard zero-temperature Uehling term, while 
570: the second is a new thermal correction.
571: 
572: \section{Free Energy and Debye Mass}
573: In this section, we use the effective Lagrangian Eq.~(\ref{eff2})
574: to calculate the free energy to order $e^3$ 
575: and the Debye mass to order $e^5$, respectively.
576: In order to take electric screening properly into account, we must 
577: include the effects of the mass term $M$ to all orders in perturbation theory.
578: Thus the Lagrangian is split accordingly:
579: \bqa
580: {\cal L}_{\rm eff}^{\rm free}&=&
581: {1\over4}F_{ij}F_{ij}+{1\over2}(\partial_i\bar{A}_0)^2+{1\over2}M^2(\Lambda)
582: \bar{A}_0^2
583: +\left(\partial_{\mu}\bar{\eta}\right)\left(\partial_{\mu}\eta\right)
584: +{\cal L}_{\rm gf}
585: \;,
586: \\
587: {\cal L}_{\rm eff}^{\rm int}&=&{1\over4}a_3(\Lambda)
588: F_{ij}\nabla^2F_{ij}
589: +{\lambda_3(\Lambda)\over24}\bar{A}_0^4
590: +\delta{\cal L}_{\rm eff}^{}\;.
591: \eqa
592: The $e^3$ contribution to the free energy is given by a simple 
593: one-loop calculation in the effective theory: 
594: \bqa
595: \label{free3}
596: {T\log{\cal Z}_{\rm eff}\over V}=
597: -{1\over2}T\int_{p}\log(p^2+M^2)
598: -{1\over2}(d-2)T\int_{p}\log p^2
599: +{1\over2}a_3(\Lambda)(d-2)\int_p p^2
600: \;.
601: \eqa
602: The total free energy is then given by
603: \bqa
604: {\cal F}=f(\Lambda)T-{T\log{\cal Z}_{\rm eff}\over V}\;,
605: \eqa
606: where $f(\Lambda)$ is given by Eq.(\ref{free1}).
607: Using Eq.~(\ref{3d}) in appendix B and the expression for the 
608: mass parameter $M(\Lambda)$ to leading order, Eq.~(\ref{free3}) reduces to
609: \bqa
610: \label{free33}
611: {T\log{\cal Z}_{\rm eff}\over V}=
612: {4e^3m^{9/4}T^{7/4}\over3(2\pi)^{13/4}}e^{-3m/2T}
613: \;.
614: \eqa
615: The total free energy density is minus the sum of 
616: Eqs.~(\ref{free1}) and~(\ref{free33}):
617: \bqa
618: \label{ftotal}
619: {\cal F}=-{\pi^2T^4\over45}
620: -{4\over(2\pi)^{3/2}}m^{3/2}T^{5/2}
621: e^{-m/T}
622: -{e^2\over2(2\pi)^3}m^2T^2
623: e^{-2m/T}
624: -{4e^3m^{9/4}T^{7/4}\over3(2\pi)^{13/4}}e^{-3m/2T}\;.
625: \eqa
626: Eqs.~(\ref{ftotal}) is in agreement with the result first
627: obtained by Gell-Mann and Bru\"ckner~\cite{gell} in nonrelativistic QED.
628: Note the term that is non-analytic in $e^2$. It arises from the
629: summation of an infinite number of infrared divergent loops
630: (ring diagrams).
631: 
632: The Debye mass $m_{\rm D}$ is given by the location of the pole 
633: in the propagator
634: \bqa
635: \label{pole}
636: k^2+M^2(\Lambda)
637: +\Pi_{\rm eff}({\bf k},\Lambda)=0\;,\hspace{1cm}k=im_{\rm D}\;,
638: \eqa
639: where $\Pi_{\rm eff}({\bf k},\Lambda)$
640: denotes the self-energy function of $\bar{A}_0$.
641: To leading order in $e$, the solution to Eq.~(\ref{pole})
642: is simply $m_{\rm D}^2=M^2(\Lambda)$. 
643: The one-loop approximation to the self-energy is
644: \bqa
645: \Pi^{(1)}_{\rm eff}({\bf k},\Lambda)
646: ={1\over2}\lambda_3(\Lambda)\int_{\bf p}{1\over p^2+M^2}\;.
647: \eqa
648: $\Pi^{(1)}_{\rm eff}({\bf k},\Lambda)$
649: is independent of the external momentum, so the 
650: solution to Eq.~(\ref{pole}) is simply
651: \bqa
652: m_{\rm D}^2=M^2(\Lambda)
653: +\Pi^{(1)}_{\rm eff}({\bf k},\Lambda)\;.
654: \eqa
655: Using Eq.~(\ref{bi3}) and expanding the mass parameter
656: $M(\Lambda)$ in powers of $e$, we obtain 
657: the Debye mass through order $e^5$:
658: \bqa\nonumber
659: m_{\rm D}^2
660: &=&
661: {4e^2\over(2\pi)^{3/2}}m^{3/2}T^{1/2}e^{-m/T}
662: -{4e^4\over3(2\pi)^{7/2}}m^{3/2}T^{1/2}Le^{-m/T}
663: +{10e^4\over3(2\pi)^3}m^2e^{-2m/T}
664: \\&&
665: \label{debyef}
666: +{2e^5\over(2\pi)^{13/4}}m^{9/4}T^{-1/4}e^{-3m/2T}
667: \;.
668: \eqa
669: Eq.~(\ref{debyef}) is the main result of the present paper.
670: Using the renormalization group equation (\ref{rg}) 
671: for the running gauge coupling,
672: we see that the Debye mass is independent of the renormalization scale 
673: $\Lambda$ up to corrections of order $e^6$.
674: Note also that there is no term proportional to $e^3$ in the expression
675: for the Debye mass.
676: The reason is that there are no bosonic propagators
677: in the one-loop self-energy graph in QED and fermions
678: need no resummation, since their Matsubara frequencies
679: are never zero. 
680: \section{Summary}
681: In the present work, we have studied QED at low temperature.
682: We have constructed an effective field theory in three dimensions
683: that is valid at distances $R\gg1/T$ by integrating
684: out the electron field and the nonzero Matsubara modes.
685: The three-dimensional field theory was used to calculate
686: the pressure to order $e^3$ and the Debye mass to order $e^5$. 
687: The pressure and the Debye mass
688: can be calculated either by resummation or by effective 
689: field theory. Not only does effective field theory simplify
690: these calculations, but it
691: also unravels the contributions to physical quantities 
692: from different momentum scales.
693: 
694: \section*{Acknowledgments}
695: This work was supported by the Stichting voor
696: Fundamenteel Onderzoek der Materie
697: (FOM), which is supported by the Nederlandse Organisatie voor Wetenschapplijk
698: Onderzoek (NWO). 
699: 
700: \appendix
701: \section{Sum-integrals}
702: \renewcommand{\theequation}{\thesection.\arabic{equation}}
703: \setcounter{equation}{0}
704: 
705: In the imaginary-time formalism for thermal field theory, 
706: the 4-momentum $P=(p_0,{\bf p})$ is Euclidean with $P^2=p_0^2+{\bf p}^2$. 
707: The Euclidean energy $p_0$ has discrete values:
708: $p_0=2n\pi T$ for bosons and $p_0=(2n+1)\pi T$ for fermions,
709: where $n$ is an integer. 
710: Loop diagrams involve sums over $p_0$ and integrals over ${\bf p}$. 
711: With dimensional regularization, the integral is generalized
712: to $d = 3-2 \epsilon$ spatial dimensions.
713: We define the dimensionally regularized sum-integral by
714: %
715: \bqa
716:   \hbox{$\sum$}\!\!\!\!\!\!\int_{P}& \;\equiv\; &
717:   \left(\frac{e^\gamma\mu^2}{4\pi}\right)^\epsilon\;
718:   T\sum_{p_0=2n\pi T}\:\int {d^{3-2\epsilon}p \over (2 \pi)^{3-2\epsilon}}\;,\\ 
719:   \hbox{$\sum$}\!\!\!\!\!\!\int_{\{P\}}& \;\equiv\; &
720:   \left(\frac{e^\gamma\mu^2}{4\pi}\right)^\epsilon\;
721:   T\sum_{p_0=(2n+1)\pi T}\:\int {d^{3-2\epsilon}p \over (2 \pi)^{3-2\epsilon}}\;,
722: \label{sumint-def}
723: \eqa
724: where 
725: $\mu$ is an arbitrary
726: momentum scale. 
727: The factor $(e^\gamma/4\pi)^\epsilon$
728: is introduced so that, after minimal subtraction 
729: of the poles in $\epsilon$
730: due to ultraviolet divergences, $\mu$ coincides 
731: with the renormalization
732: scale of the $\overline{\rm MS}$ renormalization scheme.
733: \subsection{One-loop Sum-integrals}
734: The specific one-loop 
735: fermionic
736: sum-integrals needed are
737: \bqa
738: \label{fermip}
739: \sumint_{\{P\}}
740: \log(P^2+m^2)&=&{1\over(4\pi)^2}
741: \left({\mu\over m}\right)^{2\epsilon}
742: \left[-
743: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon(1-\epsilon)(2-\epsilon)}
744: m^4+J_0T^4\right]
745: \;,\\
746: \label{spat1}
747: \sumint_{\{P\}}{1\over P^2+m^2}&=&{1\over(4\pi)^2}
748: \left({\mu\over m}\right)^{2\epsilon}
749: \left[-
750: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon(1-\epsilon)}
751: m^2-J_1T^2\right]
752: \;,\\
753: \sumint_{\{P\}}{1\over(P^2+m^2)^2}&=&{1\over(4\pi)^2}
754: \left({\mu\over m}\right)^{2\epsilon}
755: \left[
756: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon}
757: -J_2\right]\;,\\
758: \sumint_{\{P\}}{1\over(P^2+m^2)^3}&=&{1\over2(4\pi)^2}
759: \left({\mu\over m}\right)^{2\epsilon}
760: \left[
761: {e^{\gamma\epsilon}\Gamma(1+\epsilon)}
762: m^{-2}-J_3T^{-2}\right]
763: \;,\\
764: \sumint_{\{P\}}{1\over(P^2+m^2)^4}&=&{1\over6(4\pi)^2}
765: \left({\mu\over m}\right)^{2\epsilon}
766: \left[
767: {e^{\gamma\epsilon}\Gamma(1+\epsilon)}\left(1+\epsilon\right)
768: m^{-4}-J_4T^{-4}\right]
769: \;,\\
770: \label{spat2}
771: \sumint_{\{P\}}{p_ip_j\over(P^2+m^2)^2}&=&{\delta_{ij}\over2(4\pi)^2}
772: \left({\mu\over m}\right)^{2\epsilon}
773: \left[-
774: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon(1-\epsilon)}m^2
775: -J_1T^2
776: \right]
777: \;,\\
778: \sumint_{\{P\}}{p_0^2\over(P^2+m^2)^2}&=&{1\over2(4\pi)^2}
779: \left({\mu\over m}\right)^{2\epsilon}
780: \left[-
781: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon(1-\epsilon)}m^2
782: +(d-2)J_1T^2+2J_2m^2
783: \right]
784: \;,\\
785: \sumint_{\{P\}}{p_0^2\over(P^2+m^2)^3}&=&{1\over4(4\pi)^2}
786: \left({\mu\over m}\right)^{2\epsilon}
787: \left[
788: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon}
789: +\left(d-4\right)
790: J_2+2J_3m^2T^{-2}
791: \right]
792: \;,\\
793: \sumint_{\{P\}}{p_0^2\over(P^2+m^2)^4}&=&{1\over12(4\pi)^2}
794: \left({\mu\over m}\right)^{2\epsilon}
795: \left[
796: e^{\gamma\epsilon}\Gamma(1+\epsilon)m^{-2}
797: +\left(d-6\right)J_3T^{-2}+2J_4m^2T^{-4}
798: \right]
799: \;,\\
800: \label{p2}
801: \sumint_{\{P\}}{p_0^2p_ip_j\over(P^2+m^2)^4}&=&{\delta_{ij}\over24(4\pi)^2}
802: \left({\mu\over m}\right)^{2\epsilon}
803: \left[
804: {e^{\gamma\epsilon}\Gamma(1+\epsilon)\over\epsilon}
805: +\left(d-4\right)
806: J_2+2J_3m^2T^{-2}
807: \right]
808: \;.
809: \eqa
810: The integrals $J_n(\beta m)$ 
811: can be expressed as integrals involving the Fermi-Dirac
812: distribution function:
813: \bqa
814: J_n(\beta m)&=&{4e^{\gamma\epsilon}\Gamma(\mbox{${1\over2})$}
815: \over\Gamma(\mbox{${5\over2}$}-n-\epsilon)}
816: \beta^{4-2n}m^{2\epsilon}\int_0^{\infty}
817: dk\;{k^{4-2n-2\epsilon}\over(k^2+m^2)^{1/2}}{1\over 
818: e^{\beta(k^2+m^2)^{1/2}}+1}\;.
819: \eqa
820: These integrals are functions of $\beta m$ only and satisfy the recursion
821: relation
822: \bqa
823: m{\partial\over\partial m}J_n(\beta m)
824: =2\epsilon J_n(\beta m)-2(\beta m)^2J_{n+1}(\beta m)\;.
825: \eqa
826: We need the integrals $J_n$ for $\epsilon=0$.
827: In the low-temperature limit, these integrals reduce to
828: \bqa
829: J_0&\longrightarrow&8(2\pi)^{1/2}
830: m^{3/2}T^{-3/2}e^{-m/T}\;,\\
831: J_1&\longrightarrow&4(2\pi)^{1/2}
832: m^{1/2}T^{-1/2}e^{-m/T}\;,\\
833: J_2&\longrightarrow&2(2\pi)^{1/2}
834: m^{-1/2}T^{1/2}e^{-m/T}\;,\\
835: J_3&\longrightarrow&(2\pi)^{1/2}
836: m^{-3/2}T^{3/2}e^{-m/T}\;,\\
837: J_4&\longrightarrow&{1\over2}(2\pi)^{1/2}
838: m^{-5/2}T^{5/2}e^{-m/T}\;.
839: \eqa
840: Note that the integrals $J_3$ and $J_4$ need subtractions to 
841: remove power infrared divergences:
842: \bqa
843: J_3(\beta m)&=&-2
844: T^2\int_0^{\infty}
845: dk\;{1\over k^2}
846: \left[{1\over(k^2+m^2)^{1/2}}{1\over e^{\beta(k^2+m^2)^{1/2}}+1}
847: -{1\over m}{1\over e^{\beta(k^2+m^2)^{1/2}}+1}
848: \right]\;, \\
849: J_4(\beta m)&=&3\nonumber
850: T^4\int_0^{\infty}
851: dk\;{1\over k^4}
852: \Bigg[{1\over(k^2+m^2)^{1/2}}{1\over e^{\beta(k^2+m^2)^{1/2}}+1}
853: -{1\over m}{1\over e^{\beta(k^2+m^2)^{1/2}}+1}
854: \\
855: &&
856: \hspace{2.6cm}
857: +{k^2\over2m^3}{1\over e^{\beta(k^2+m^2)^{1/2}}+1}
858: \Bigg]\;.
859: \eqa
860: 
861: The specific one-loop bosonic sum-integral needed is
862: \bqa
863: \label{logp}
864: \sumint_{P}\log(P^2)&=&-{\pi^2T^4\over45}\left[
865: 1+{\cal O}(\epsilon)
866: \right]\;.
867: \eqa
868: 
869: \subsection{Two-loop Sum-integral}
870: We also need the value of the two-loop diagram in Eq.~(\ref{two}).
871: The two-loop sum-integral with nonzero chemical $\mu$ potential
872: was calculated in e.g.
873: Refs.~\cite{kap2,toimela,kapusta}. There are 
874: $T$-dependent and $\mu$-dependent infinities
875: in addition to the usual vacuum infinities. These are cancelled by the
876: corresponding infinities arising from the one-loop counterterm diagrams.
877: The vacuum infinity is cancelled by a vacuum counterterm in the usual way.
878: If we demand that the two-loop contribution to the vacuum energy vanishes
879: at the scale $\Lambda$, the diagram is given by its finite-temperature
880: piece.
881: The final result after renormalization is then~\cite{toimela}:
882: \bqa\nonumber
883: {1\over2}e^2
884: \sumint_{\{PQ\}}\mbox{Tr}\left[
885: \gamma_{\mu}
886: {P\!\!\!\!/-m\over P^2+m^2}
887: \gamma_{\mu}{Q\!\!\!\!/-m\over Q^2+m^2}
888: {1\over(P+Q)^2}
889: \right]&=&
890: -{e^2\over12\pi^2}T^2\int_0^{\infty}dp\;{p^2\over E_p}\left[n_p^++n_p^-\right]
891: \\ \nonumber
892: &&\hspace{-7cm}
893: -
894: {e^2\over16\pi^4}\int_0^{\infty}dp\;dq\;
895: {p^2q^2\over E_pE_q}
896: \left[
897: \left(
898: 2+{m^2\over pq}\log{E_pE_q-m^2-pq\over E_pE_q-m^2+pq}
899: \right)\left(n_p^-n_q^-+n_p^+n_q^+\right)
900: \right.
901: \\
902: &&\hspace{-7cm}
903: +
904: \left.
905: \left(
906: 2+{m^2\over pq}\log{E_pE_q+m^2+pq\over E_pE_q+m^2-pq}
907: \right)\left(n_p^-n_q^++n_p^+n_q^-\right)
908: \right]
909: \label{fermi2}
910: \;,
911: \eqa
912: where $E_p=\sqrt{p^2+m^2}$ and 
913: \bqa
914: n_p^{\pm}={1\over e^{\beta(E_p\pm\mu)}+1}\;.
915: \eqa
916: In the low-temperature limit, this reduces to
917: \bqa
918: \label{limit}
919: {1\over2}e^2
920: \sumint_{\{PQ\}}\mbox{Tr}\left[
921: \gamma_{\mu}
922: {P\!\!\!\!/-m\over P^2+m^2}
923: \gamma_{\mu}{Q\!\!\!\!/-m\over Q^2+m^2}
924: {1\over(P+Q)^2}
925: \right]&\longrightarrow&
926: {e^2\over2(2\pi)^3}m^2T^2e^{2(\mu-m)/T}
927: \;.
928: \eqa
929: Eq.~(\ref{fermi2}) gives the two-loop contribution to the pressure ${\cal P}$.
930: By applying the formula~\cite{kapusta}
931: \bqa
932: \Pi_{00}(0,0)=e^2{\partial^2{\cal P}^2\over\partial\mu^2}\;,
933: \eqa
934: to Eq.~(\ref{limit}), we obtain the two-loop contribution to the photon
935: polarization tensor at zero external momentum and vanishing 
936: chemical potential:
937: \bqa
938: \label{twopol}
939: \Pi_{00}^{(2)}(0,0)
940: &=&
941: {2e^4\over(2\pi)^3}m^2e^{-2m/T}
942: \;.
943: \eqa
944: \section{Integrals}
945: 
946: Dimensional regularization can be used to 
947: regularize both the ultraviolet divergences and infrared divergences
948: in 3-dimensional integrals over momenta. 
949: The spatial dimension is generalized to  $d = 3-2\epsilon$ dimensions.
950: Integrals are evaluated at a value of $d$ for which they converge and then
951: analytically continued to $d=3$.
952: We use the integration measure
953: %
954: \begin{equation}
955:   \int_{\bf p} \;\equiv\;
956:   \left(\frac{e^\gamma\mu^2}{4\pi}\right)^\epsilon\,
957:   \int {d^{3-2\epsilon}p \over (2 \pi)^{3-2\epsilon}}\,,
958: \end{equation}
959: %
960: where 
961: $\mu$ is an arbitrary
962: momentum scale. 
963: The factor $(e^\gamma/4\pi)^\epsilon$
964: is introduced so that, after minimal subtraction 
965: of the poles in $\epsilon$
966: due to ultraviolet divergences, $\mu$ coincides 
967: with the renormalization
968: scale of the $\overline{\rm MS}$ renormalization scheme.
969: 
970: The one-loop integrals required are
971: %
972: \begin{eqnarray}
973: \int_{\bf p} \log(p^2+m^2) & = & 
974: - {m^3\over 6\pi}\left[1+{\cal O}(\epsilon)\right]
975:   \,,
976: \label{3d}
977: \\ 
978: \int_{\bf p} {1 \over p^2+m^2} & = & 
979: - {m\over 4\pi} 
980: \left[1 + {\cal O}\left(\epsilon\right)  \right] \,.
981: \label{bi3}
982: \end{eqnarray}
983: %
984: 
985: \section{Polarization Tensor}
986: In this appendix, we calculate the polarization tensor
987: $\Pi_{\mu\nu}(\omega_n,{\bf k})$ to one loop. 
988: The Feynman diagram is shown in 
989: Fig.~\ref{pollen} and the expression for it is
990: \bqa
991: \label{pol}
992: \Pi_{\mu\nu}^{(1)}(k_0,{\bf k})=e^2\sumint_{\{P\}}\mbox{Tr}\left[
993: \gamma_{\mu}{P\!\!\!\!/-m\over P^2+m^2}
994: \gamma_{\nu}{(P\!\!\!\!/+K\!\!\!\!/)-m\over(P+K)^2+m^2}\right]\;.
995: \eqa
996: Taking the Dirac trace, 
997: and using a Feynman parameter $y$, this can be writen as
998: \bqa\nonumber
999: \Pi_{\mu\nu}^{(1)}(k_0,{\bf k})&=&e^2\int_0^1dy\;
1000: \sumint_{\{P\}}
1001: \Bigg[
1002: {8p_{\mu}p_{\nu}\over\left[P^2+m^2+K^2y(1-y)\right]^2}
1003: -{4\delta_{\mu\nu}\over\left[P^2+m^2+K^2y(1-y)\right]}
1004: \\
1005: &&
1006: +{8y(1-y)(\delta_{\mu\nu}K^2-k_{\mu}k_{\nu})
1007: \over\left[P^2+m^2+K^2y(1-y)\right]^2}
1008: \Bigg]\;.
1009: \eqa
1010: First consider $\Pi_{ij}^{(1)}(0,{\bf k})$. By virtue of 
1011: Eqs.~(\ref{spat1}) and~(\ref{spat2}), the first two terms cancel identically.
1012: The remaining term is expanded to second order 
1013: in a Taylor series around ${\bf k}=0$. 
1014: Integrating over $y$, we obtain
1015: \bqa
1016: \Pi_{ij}^{(1)}(0,{\bf k})&=&
1017: {4\over3}
1018: {e^2}(\delta_{ij}k^2-k_{i}k_{j})\sumint_{\{P\}}{1\over (P^2+m^2)^2} 
1019: \label{space}
1020: -{8\over15}{e^2}k^2(\delta_{ij}k^2-k_{i}k_{j})\sumint_{\{P\}}
1021: {1\over (P^2+m^2)^3}\;.
1022: \eqa
1023: The first term gives the one-loop correction to the field normalization
1024: constants, while the second gives the coefficient of the Uehling term.
1025: The fact that $\Pi_{ij}^{(1)}(0,{\bf k})$ vanishes 
1026: in the limit ${\bf k}\rightarrow 0$ reflects the fact that there is no 
1027: screening of static magnetic fields in QED.
1028: 
1029: Consider next $\Pi_{00}^{(1)}(0,{\bf k})$. 
1030: Expanding to second order in the external momentum ${\bf k}$
1031: and integrating over $y$, yields
1032: \bqa
1033: \label{stat}
1034: \Pi_{00}^{(1)}(0,{\bf k})
1035: &=&4e^2\sumint_{\{P\}}\Bigg[
1036: {2p_0^2\over(P^2+m^2)^2}-{1\over P^2+m^2}\Bigg]
1037: +{1\over3}e^2k^2\sumint_{\{P\}}\Bigg[
1038: {6\over(P^2+m^2)^2}-{8p_0^2\over(P^2+m^2)^3}\Bigg]
1039: \;.
1040: \eqa
1041: The first term gives the one-loop expression for the mass parameter 
1042: (which coincides with the one-loop expression for the Debye mass), while the
1043: second term gives the one-loop correction to the field normalization constant.
1044: \begin{thebibliography}{99}
1045: \bibitem{barton}G. Barton, Ann. Phys. (N.Y), {\bf 205}, 49 (1991).
1046: \bibitem{finn} X. Kong and F. Ravndal, Nucl. Phys. {\bf B526}, 627 (1998).
1047: \bibitem{bj} E. Braaten and Y. Jia, 
1048: Phys. Rev. {\bf D63}, 096009 (2001). 
1049: \bibitem{lepage}G. P. Lepage, in {\it From actions to answers},
1050: proceedings of the Advanced Study Institute in Elementary Particle Physics,
1051: Boulder, Colorado, 1989, edited by T. De--Grand and D. Toussaint, 
1052: World Scientific (1989).
1053: \bibitem{kaplan} D.B. Kaplan, nucl-th/9506035,
1054: Lectures given at 7th Summer School in 
1055: Nuclear Physics Symmetries, Seattle, WA, 18-30 Jun 1995. 
1056: \bibitem{gell} M. Gell-Mann and K. A. Br\"uckner, Phys. Rev. {\bf 106}, 
1057: 364 (1957).
1058: \bibitem{kap2} J.I Kapusta, Nucl. Phys. {\bf B148}, 461 (1979). 
1059: \bibitem{toimela}T. Toimela, Int. J. Th. Phys. {\bf 24}, (1985);
1060: Erratum-ibid. {\bf 26}, 1021 (1987). 
1061: \bibitem{wolo}R.M. Woloshyn, Phys. Rev. {\bf D27}, 1393 (1982).
1062: \bibitem{rolf1}R. Tarrach, Phys. Lett. {\bf B133}, 259 (1983).
1063: \bibitem{rolf}J.I. Latorre, P. Pascual and R. Tarrach, 
1064: Nucl. Phys. {\bf B 437}, 60 (1995).
1065: \bibitem{bosse}P. Elmfors, D. Persson, and B-S. Skagerstam,
1066: Astropart. Phys.{\bf 2}, 299 (1994).  
1067: \bibitem{holger1}H. Gies, Phys. Rev. {\bf D61}, 085021 (2000);
1068: e-print hep-ph/0010287.
1069: \bibitem{uh} E.A. Uehling, Phys. Rev. 48, 55 (1935). 
1070: \bibitem{eh} H. Euler, Ann. Phys. (Leipzig), {\bf 26}, 398 (1936);
1071: W. Heisenberg and H. Euler, Zeit. Phys. {\bf 98}, 714 (1936).
1072: \bibitem{dicus}D.A. Dicus, C. Kao and W.W. Repko, 
1073: Phys. Rev. {\bf D57}, 2443 (1998)
1074: \bibitem{landsman} N.P. Landsman. Nucl. Phys. {\bf B322}, 498 (1989).
1075: \bibitem{gins}P. Ginsparg, Nucl. Phys. {\bf B170}, 388 (1980). 
1076: \bibitem{BN}E.~Braaten and A.~Nieto, Phys.~Rev.~{\bf D51}, 6990 (1995).
1077: \bibitem{kaja}K. Kajantie, M. Laine, K. Rummukainen, M. Shaposhnikov,
1078: Nucl. Phys. {\bf B458}, 90 (1996). 
1079: \bibitem{Andersen} J.O.~Andersen, Phys.~Rev.~{\bf D53}, 7286 (1996).
1080: \bibitem{kapusta} J.I. Kapusta, {\it Finite Temperature
1081: Field Theory}, Cambridge University Press, Cambridge (1989).
1082: \end{thebibliography}
1083: \end{document}
1084: 
1085: