1: \documentstyle[11pt,epsfig]{article}
2: \def\double{\baselineskip 24pt \lineskip 10pt}
3:
4: \textheight 8.5in
5: \textwidth 6in
6: \oddsidemargin 8pt
7: \topmargin -30pt
8:
9: \def\baselinestretch{1.2}
10:
11: \parskip 0.2cm
12:
13: \def\laq{~\raise 0.4ex\hbox{$<$}\kern -0.8em\lower 0.62
14: ex\hbox{$\sim$}~}
15: \def\gaq{~\raise 0.4ex\hbox{$>$}\kern -0.7em\lower 0.62
16: ex\hbox{$\sim$}~}
17:
18:
19: \begin{document}
20:
21: \begin{titlepage}
22:
23: \begin{flushright}
24: CERN-PH-TH/2007-229
25: \end{flushright}
26: \vspace*{1.8 cm}
27:
28: \begin{center}
29:
30: \huge
31: {Magnetogenesis, \\
32: spectator fields and CMB signatures}
33:
34: \vspace{1cm}
35:
36: \large{ Massimo Giovannini}\footnote{e-mail address: massimo.giovannini@cern.ch}\\
37: \normalsize
38: \vspace{.2in}
39: {\sl Centro ``Enrico Fermi", Compendio del Viminale, Via
40: Panisperna 89/A, 00184 Rome, Italy}\\
41: \vspace{8mm}
42: {\sl Department of Physics, Theory Division, CERN, 1211 Geneva 23, Switzerland}
43:
44: \begin{abstract}
45: A viable class of magnetogenesis models can
46: be constructed by coupling the kinetic term of the hypercharge to
47: a spectator field whose dynamics does not affect the inflationary
48: evolution. The magnetic power spectrum is explicitly related to the power spectrum of
49: (adiabatic) curvature inhomogeneities when the quasi-de Sitter stage of expansion is driven by a single scalar degree of freedom.
50: Depending upon the value of the slow-roll parameters,
51: the amplitude of smoothed magnetic fields over a (comoving)
52: Mpc scale can be as large as $0.01$--$0.1$ nG at the epoch of the gravitational collapse of the
53: protogalaxy.
54: The contributions of the magnetic fields to the Sachs-Wolfe plateau and to the temperature autocorrelations in the Doppler region compare favourably
55: with the constraints imposed by galactic magnetogenesis.
56: Stimulating lessons are drawn on the interplay between magnetogenesis models and their possible CMB signatures.
57: \end{abstract}
58: \end{center}
59: \end{titlepage}
60: \newpage
61: Since the early fifties large-scale magnetic fields have
62: inspired different areas of investigation both at a theoretical and at a more phenomenological level
63: (see, as an example, Ref. \cite{zel1} for theoretical and historical
64: accounts of the subject).
65: Both elliptical and spiral galaxies have magnetic fields at the $\mu$ G level \cite{gal1}. Abell clusters possess large-scale magnetic fields (not associated with individual galaxies) with typical correlation scale which can be as large as $100$ kpc \cite{clust1}.
66: Superclusters have been also claimed
67: to have magnetic fields \cite{supclust} at the $\mu$G level even if, in this case, crucial ambiguities
68: persist on the way the magnetic field strengths are inferred from the Faraday rotation measurements. The latest analyses of the AUGER experiment
69: demonstrated a correlation between the arrival directions of cosmic rays
70: with energy above $6\times10^{19}$ eV and the positions of active galactic
71: nuclei within $75$ Mpc \cite{auger1}. At smaller energies it has been
72: convincingly demonstrated \cite{auger2} that overdensities on windows of $5$ deg radius (and for energies $10^{17.9} \mathrm{eV} < E < 10^{18.5} \mathrm{eV}$) are compatible
73: with an isotropic distribution. Thus, in the highest energy domain (i.e. energies larger than $60$ EeV),
74: cosmic rays are not appreciably deflected: within a cocoon of $70$ Mpc the intensity of the (uniform) component of the putative magnetic field should be smaller than the nG. On a theoretical ground, the existence of much larger magnetic fields (i.e. ${\mathcal O}(\mu\mathrm{G})$) cannot be justified already if
75: the correlation scale is of the order of $20$ Mpc.
76:
77: In the late sixties Harrison \cite{harrison1} suggested that
78: cosmology and astrophysics are just two complementary aspects of the origin of large-scale magnetic fields.
79: Heeding observations there is no evidence against the primeval hypothesis
80: even if the primordial origin of large-scale magnetism is not empirically compelling.
81: Compressional amplification (taking place during the gravitational collapse of the protogalaxy) allows to connect the observed magnetic field to a protogalactic field, present prior to gravitational collapse, of typical strength ranging between $0.1$ and $0.01$ nG.
82: A better understanding of the interplay between dynamo theory
83: and the global conservation laws of magnetized plasmas
84: has been recently achieved also because of the improved
85: comprehension of the solar dynamo action \cite{kanduaxel}.
86: Within the dynamo hypothesis,
87: the protogalactic field could be even much smaller than the nG and still explain some crucial properties of our magnetized Universe: astrophysical and cosmological mechanisms might really be complementary
88: rather than mutually exclusive \cite{harrison1,kanduaxel} (see also \cite{zel1}, second reference).
89:
90: The only direct way of putting the primordial hypothesis to test is represented by the observations related to the Cosmic Microwave Background \footnote{It is not excluded that
91: the study of the morphological features of galactic fields will also give indications, albeit indirect, on the primordial nature of the protogalactic field (see Ref. \cite{gal1} and discussions therein).}
92: (CMB in what follows): the possible existence of a magnetized
93: plasma prior to decoupling is germane to several CMB observables
94: like temperature autocorrelations and cross-correlations (see \cite{max2} and references therein). Recently,
95: a semi-analytical technique has been developed
96: to compute more accurately than before the magnetized temperature and polarization autocorrelations
97: as well as cross-correlations \cite{max3}: the gross logic of the method is
98: to assume a dominant adiabatic mode in the pre-equality initial conditions and to add, consistently, the effects of the magnetic fields in the Einstein-Boltzmann hierarchy and in the initial conditions.
99:
100: Large-scale magnetic fields produced
101: inside the Hubble radius after inflation will have a correlation scale
102: bounded (from above) by the Hubble radius at the moment
103: when some charge separation is produced (be it, for instance,
104: the electroweak time\footnote{This example holds under
105: the assumption the electroweak phase transition is strongly first order
106: which is, arguably, not the case.}). Since the Hubble radius, during radiation, evolves much faster than the correlation scale of the produced field, the typical scale over which the magnetic field is coherent today is
107: much shorter than the Mpc, obliterating, in this way, the
108: possibility of successfully reproducing the galactic magnetic field
109: \cite{max2}. Conversely, the physical rationale for
110: inflationary magnetogenesis resides on the possibility of achieving
111: large correlation scales: quantum fluctuations of Abelian gauge
112: fields can be amplified in the same way as zero point fluctuations of the geometry are amplified. Unlike the scalar and tensor modes of the geometry, Abelian gauge fields (like the hypercharge) in four space-time dimensions obey Weyl invariant evolution equations \cite{weyl1}.
113: Since the pumping action of the background geometry is not efficient in amplifying the fluctuations of gauge fields,
114: Weyl invariance should be broken for the viability of the whole
115: construction \cite{weyl1}.
116: The amplified gauge fields should be Abelian. The only non-screened
117: vector modes that are present at finite conductivity
118: are the ones associated with the hypercharge field \cite{max4}.
119: The non-Abelian fields develop actually a mass and they are screened as the Universe thermalizes.
120: After the electroweak phase transition the photon field remains unscreened
121: with amplitude $\cos{\theta_{\mathrm{w}}} \vec{{\mathcal Y}}$.
122: While the coupling of the hypercharge to fermions is chiral, the QED coupling is vector-like. At
123: finite conductivity, however, the descriptions of the two plasmas are similar
124: \footnote{See \cite{max4} and the equations of anomalous magnetohydrodynamics, i.e. the generalization of magnetohydrodynamics to the case where anomalous effects are included.}
125: and can be given in terms of an effective (Ohmic) current which is proportional
126: to the (hyper)electric field. The specific nature of the gauge field is often ignored in the current literature: the main endevour is to break consistently Weyl invariance (possibly maintaining gauge invariance). The
127: Abelian field arising in this case which should be thought, indeed, as a putative hypercharge field.
128:
129: In the present paper it will be argued that Weyl invariance can be broken
130: through the coupling of a spectator field to the gauge kinetic term also in the case of conventional inflationary scenarios.
131: A spectator field is defined, in the present context, as a field which does not drive the inflationary
132: evolution but which is, nonetheless, dynamical. It is not excluded, in the present context,
133: that the resulting large-scale magnetic fields are amplified to nG strength and with a
134: nearly scale-invariant spectrum.
135: The field content of the model is apparent from the total action which
136: includes, on top of the gravitational part, the contribution
137: of the inflaton $\varphi$ and of the spectator field $\psi$:
138: \begin{equation}
139: S_{\mathrm{tot}} = S_{\mathrm{gravity}} + S_{\varphi} + S_{\psi}.
140: \label{totac}
141: \end{equation}
142: The various components of the total action can be written, in explicit terms, as\footnote{The conventions on the four-dimensional metric will be mostly minus, i.e. $(+,-,-,-)$. Recall also that
143: $\overline{M}_{\mathrm{P}} = M_{\mathrm{P}}/\sqrt{8\pi}$ with $M_{\mathrm{P}} = 1.22\times
144: 10^{19} \mathrm{GeV}$.}
145: \begin{eqnarray}
146: &&S_{\mathrm{gravity}} =-\frac{\overline{M}_{\mathrm{P}}^2}{2} \int d^{4} x \sqrt{-g} R,
147: \qquad S_{\varphi} = \int d^{4} x \sqrt{-g} \biggl[\frac{1}{2} g^{\alpha\beta}\partial_{\alpha}\varphi \partial_{\beta} \varphi - V(\varphi)\biggr],
148: \label{grphiac}\\
149: && S_{\psi} = \int d^{4} x \sqrt{-g} \biggl[\frac{1}{2} g^{\alpha\beta}\partial_{\alpha}\psi \partial_{\beta} \psi - W(\psi) - \frac{\lambda(\psi)}{16 \pi} Y_{\alpha\beta}
150: Y^{\alpha\beta} \biggr],
151: \label{psiac}
152: \end{eqnarray}
153: where $V(\varphi)$ and $W(\psi)$ are, respectively, the inflaton potential
154: and the potential of the spectator field. The hypercharge field strength$Y_{\alpha\beta} = \nabla_{[\alpha} Y_{\beta]}$
155: is defined in terms of the covariant derivative with respect to the four-dimensional metric $g_{\mu\nu}$.
156: In Eq. (\ref{psiac}), $\lambda(\psi)$ denotes
157: the coupling of $\psi$ to the hypercharge field.
158:
159: In a conformally flat Friedmann-Robertson-Walker metric $g_{\mu\nu} = a^2(\tau) \eta_{\mu\nu}$ (where
160: $\eta_{\mu\nu}$ is the four-dimensional Minkowski metric),
161: the Hamiltonian constraint stemming from the equations derived from the total action
162: (\ref{totac}) is given by
163: \begin{equation}
164: \overline{M}_{\mathrm{P}}^2 {\mathcal H}^2 = \frac{1}{3}\biggl[\frac{{\varphi'}^2}{2} + V a^2\biggr]
165: + \frac{1}{3}\biggl[\frac{{\psi'}^2}{2} + W a^2\biggr] + \frac{1}{8\pi} (\vec{{\mathcal B}}^2 +
166: \vec{{\mathcal E}}^2),
167: \label{HAM1}
168: \end{equation}
169: where the prime denotes the derivation with respect to the conformal time coordinate $\tau$ and
170: ${\mathcal H} = a'/a$ is related to the Hubble parameter $H$ as ${\mathcal H} = H/a$. In Eq. (\ref{HAM1}) $\vec{{\mathcal E}} = \sqrt{\lambda} \,\vec{e}$ and $
171: \vec{{\mathcal B}} = \sqrt{\lambda} \,\vec{b}$ are, respectively, the
172: hyperelectric and the hypermagnetic fields defined, from the field strength as\footnote{The
173: rescaling of $\vec{e}$ and $\vec{b}$ through $\sqrt{\lambda}$ arises since
174: the hypercharge energy-momentum tensor contains the coupling to $\psi$ throught $\lambda$. These will not be, however, the normal modes of the system as it will be clear in a moment.}
175: $Y_{0i} = a^2 e_{i}$ and $Y_{ij} = - a^2 \epsilon_{ijk} b^{k}$.
176: The dual field strengths (appearing in the Bianchi identity) are simply $\tilde{Y}_{ij} = a^2 e^{m} \epsilon_{mij}$ and $\tilde{Y}_{0i} = a^2 b_{i}$.
177: The field $\varphi$ is the dominant source of the background geometry while
178: $\psi$ is a spectator field which is allowed to roll during inflation but which gives a negligible
179: contribution to the background geometry. Denoting by $\psi_{i}$ the initial value of $\psi$
180: at a curvature scale $H_{i}$ this requirement implies that
181: \begin{equation}
182: \psi_{\mathrm{i}}^2 < \frac{2}{3} \biggl(\frac{H_{1}}{H_{\mathrm{i}}}\biggr)^2 \overline{M}_{\mathrm{P}}^2
183: \label{cond1}
184: \end{equation}
185: where $H_{1}$ is the curvature scale at the end of inflation. When $\psi$ starts rolling at
186: $\tau_{i}$ the hyperelectric and the hypermagnetic fields are just given by
187: their corresponding quantum fluctuations and are therefore even smaller than the
188: energy density of $\psi$. The (homogeneous)
189: evolution equation for $\psi$ will therefore be given by
190: \begin{equation}
191: \psi'' + 2 {\mathcal H} \psi' + \frac{\partial W}{\partial\psi} a^2 + \frac{a^2}{8\pi} \frac{\partial \ln{\lambda}}{\partial \psi}(\vec{{\mathcal B}}^2 + \vec{{\mathcal E}}^2)=0.
192: \label{psieq}
193: \end{equation}
194: The values of the hypermagnetic (and hyperelectric) fields
195: generated from quantum fluctuations will be always smaller than the energy density of $\psi$. This implies that the back-reaction terms arising, for instance, in Eqs. (\ref{HAM1}) and (\ref{psieq}) can be safely neglected.
196: It will be assumed that $W(\psi) = m^2 (\psi -\psi_{*})/2$ with $m < H_{1}$.
197: Since $\psi$ is light during inflation, it will also be required that
198: $\psi_{*} \ll M_{\mathrm{P}}$.
199: Deep in the course of the inflationary epoch the evolution
200: equation of $\psi$ is then
201: \begin{equation}
202: \Sigma'' +[ \mu^2 - (2 - \epsilon)] a^2 H^2 \Sigma =0,\qquad
203: \Sigma = a \psi,\qquad \epsilon= - \frac{\dot{H}}{H^2} = \frac{\overline{M}_{\mathrm{P}}}{2} \biggl(\frac{V_{,\varphi}}{V}\biggr)^2
204: \label{PSI2}
205: \end{equation}
206: having introduced $\mu=m/H$ and the first slow-roll parameter $\epsilon$
207: which is related to the first derivative of the inflaton potential.
208: In the limit $\mu\ll 1$ the evolution of $\psi$ will be simply given by
209: \begin{equation}
210: \psi = \psi_{\mathrm{i}}\biggl( - \frac{\tau}{\tau_{\mathrm{i}}}\biggr)^{\beta} + \psi_{*},\qquad \beta = \frac{3 - 2 \epsilon}{1 - \epsilon}.
211: \label{PSI3}
212: \end{equation}
213: As the field $\psi$ evolves in time, the hypermagnetic and hyperelectric
214: fields can be parametrically amplified, as it follows from the
215: equations of motion easily obtainable by the appropriate functional variation of the
216: total action (\ref{totac}):
217: \begin{equation}
218: \nabla_{\mu} (\lambda Y^{\mu\nu}) = 4\pi J^{\nu}, \qquad \nabla_{\mu} (\tilde{Y}^{\mu\nu})=0,\qquad
219: \lambda(\psi) = \biggl(\frac{\psi - \psi_{*}}{\overline{M}_{\mathrm{P}}}\biggr)^{\alpha}.
220: \label{Y1}
221: \end{equation}
222: where the contribution of the (Ohmic) current $J^{\nu}$ has been
223: included for convenience. In Eq. (\ref{Y1})
224: the expression of $\lambda(\psi)$ contains the parameter
225: $\alpha$ which will eventually determine the slope of the gauge field
226: spectra and which will be constrained by phenomenological considerations.
227: In the conformally flat metric $g_{\mu\nu} = a^2(\tau) \eta_{\mu\nu}$, Eq. (\ref{Y1}) can be written, using vector notations, as:
228: \begin{eqnarray}
229: &&\vec{\nabla} \times ( a^2 \sqrt{\lambda} \vec{{\mathcal B}}) = \frac{\partial}{\partial\tau} [ a^2 \sqrt{\lambda}
230: \vec{{\mathcal E}}] + 4\pi \vec{J},\qquad \vec{\nabla}\cdot\vec{J} =0,
231: \label{Y2}\\
232: && \frac{\partial}{\partial \tau} \biggl[ \frac{a^2 \vec{\mathcal B}}{\sqrt{\lambda}}\biggr] +
233: \vec{\nabla}\times \biggl[ \frac{a^2 \vec{\mathcal E}}{\sqrt{\lambda}}\biggr] =0.
234: \label{Y3}
235: \end{eqnarray}
236: where $\vec{J} = a^3 \sigma_{\mathrm{c}} = \sigma a^2 \vec{e} = \sigma a^2 \vec{{\mathcal E}}/{\sqrt{\lambda}}$;
237: $\sigma(\tau) = \sigma_{\mathrm{c}} a(\tau)$ denotes the rescaled value of the conductivity
238: and it appears because of the choice of the conformal time coordinate as a pivot variable of the
239: system. Since $\lambda$ depends only upon $\tau$ the Ohmic current is always
240: divergence-less as it should be by definition.
241: Combining Eqs. (\ref{Y2}) and (\ref{Y3}) in the absence
242: of conductivity (i.e. during inflation) the hypermagnetic and hyperelectric fields obey the following pair
243: of (decoupled) equations:
244: \begin{eqnarray}
245: \vec{B}'' - \nabla^2 \vec{B} - \frac{(\sqrt{\lambda})''}{\sqrt{\lambda}} \vec{B} =0,\qquad \vec{E}'' - \nabla^2 \vec{E} - \sqrt{\lambda} \biggl(\frac{1}{\sqrt{\lambda}}\biggr)'' \vec{E} =0,
246: \label{hmagn}
247: \end{eqnarray}
248: where $\vec{E} = a^2 \vec{{\mathcal E}}$ and $\vec{B} = a^2 \vec{{\mathcal B}}$ are the normal modes of the system. The dual nature of the pump fields for $\vec{E}$ and $\vec{B}$ in Eq. (\ref{hmagn}) is a reflection of the strong-weak coupling duality
249: of the Abelian theory in the absence of sources (see, for instance, \cite{sdual}).
250: During inflation the gauge field fluctuations can then be quantized in the Coulomb
251: gauge (which is the appropriate one for
252: treating gauge fields in time-dependent background geometries \cite{lford}) and the vector potential
253: can be expanded in terms of the appropriate mode functions $f_{k}(\tau)$
254: \begin{equation}
255: \hat{{\mathcal Y}}_{i}(\vec{x},\tau) = \frac{1}{(2\pi)^{3/2}} \sum_{\gamma} e^{(\gamma)}_{i}
256: \int d^{3} k [ \hat{a}_{\vec{k},\gamma} f_{k}(\tau) e^{- i \vec{k}\cdot \vec{x}} +\hat{a}_{\vec{k},\gamma}^{\dagger} f_{k}^{*}(\tau)
257: e^{ i \vec{k}\cdot \vec{x}} ],
258: \label{vecpot}
259: \end{equation}
260: where $e^{(\gamma)}_{i}$ is the polarization unit vector; $\hat{a}_{k,\gamma}$ and $\hat{a}_{\vec{k},\gamma}^{\dagger}$ obey $[\hat{a}_{k,\gamma}, \hat{a}_{\vec{p},\gamma}^{\dagger}] = \delta_{\gamma\gamma'} \delta^{(3)}(\vec{k} - \vec{p})$.
261: Since $ \vec{B} = \vec{\nabla}\times \vec{{\mathcal Y}}$, the mode function $f_{k}(\tau)$ (and its complex conjugate) will satisfy the same equation obeyed by $\vec{B}$ (see Eq. (\ref{hmagn})).
262:
263: At end of inflation the Universe reheats. Thanks to the decay
264: of the inflaton and of the spectator field the quasi-de Sitter
265: background becomes effectively dominated by a fluid of ultra-relativistic
266: particles with radiative equation of state. Overall the plasma is
267: globally neutral but the conductivity becomes
268: large since charged species are copiously produced.
269: Lorentz invariance is then broken and hyperelectric fields are strongly
270: suppressed while the hypermagnetic fields survive.
271: A preferred physical frame naturally emerges,
272: i.e. the so-called plasma frame
273: where the conductivity is finite and the hyperelectric fields are dissipated.
274: Since $\psi$ decays, $\lambda$ will freeze and
275: the system of Eqs. (\ref{Y2}) and (\ref{Y3}) can be written as
276: \begin{equation}
277: \frac{\partial \vec{E}_{\mathrm{a}}}{\partial \tau} +4 \pi \sigma \vec{E}_{\mathrm{a}}
278: = \vec{\nabla}\times \vec{B}_{\mathrm{a}},\qquad \frac{\partial \vec{B}_{\mathrm{a}}}{\partial \tau} = - \vec{\nabla}\times \vec{E}_{\mathrm{a}},
279: \label{M2}
280: \end{equation}
281: where the subscript ``a" signifies that the hyperelectric and hypermagnetic fields are computed after the
282: transition to radiation.
283: Denoting with the subscript ``b" the field variables after the rise of the conductivity the
284: appropriate continuity conditions for the magnetic and the electric fields are:
285: \begin{equation}
286: \vec{B}_{\mathrm{a}} = \vec{B}_{\mathrm{b}},\qquad
287: \vec{E}_{\mathrm{a}} = \frac{\vec{\nabla}\times \vec{B}_{\mathrm{a}}}{4\pi \sigma} =
288: \frac{\vec{\nabla}\times \vec{B}_{\mathrm{b}}}{4\pi \sigma}.
289: \label{M3}
290: \end{equation}
291: Equation (\ref{M3}) stipulates that, after the transition, the electric fields
292: are suppressed by the conductivity as soon as radiation dominates.
293: Solving Eq. (\ref{hmagn}) during inflation and Eq. (\ref{M2}) during radiation
294: the boundary conditions (\ref{M3}) permit the estimate of the two point function of the hypermagnetic field operators for a generic time $\tau > \tau_{1}$ where $\tau_{1}$ denotes the epoch of the sudden rise in the conductivity:
295: \begin{equation}
296: \langle 0| \hat{B}_{i}(\vec{x},\tau) \hat{B}_{j}(\vec{y},\tau) |0 \rangle
297: = \int d \ln{k} \,\, P_{B}(k) P_{ij}(k) \frac{\sin{kr}}{kr},\qquad r =|\vec{x} - \vec{y}|,
298: \label{M15}
299: \end{equation}
300: where $P_{\mathrm{B}}(k)$ and $P_{ij}(k)$ denote, respectively, the hypermagnetic power spectrum and the traceless projector
301: \begin{equation}
302: P_{B}(k) = {\mathcal C}(\delta) H_{1}^4 \biggl(\frac{k}{k_{1}}\biggr)^{n_{\mathrm{B}-1}} e^{- 2 \frac{k^2}{k_{\sigma}^2}},\qquad P_{ij}(k)= \biggl( \delta_{ij} - \frac{k_{i} k_{j}}{k^2}\biggr).
303: \label{M16}
304: \end{equation}
305: In Eq. (\ref{M16}) ${\mathcal C}(\delta) = 2^{2\delta-1} \Gamma^2(\delta)/\pi^2$ and
306: \begin{equation}
307: \delta = \frac{3\alpha -1 + \epsilon( 1 - 2 \alpha)}{2 ( 1 - \epsilon)},\qquad n_{\mathrm{B}}
308: = \frac{7 - 3\alpha - \epsilon ( 7 - 2\alpha)}{1 - \epsilon},\qquad k_{1} = \frac{1}{\tau_{1}}.
309: \label{M16a}
310: \end{equation}
311: In Eq. (\ref{M16}) $k_{\sigma}$ is the conductivity wave-number, i.e.
312: $k_{\sigma}^{-2} = \int d\tau/(4\pi\sigma)$.
313: The wave-numbers $k_{1}$ and $k_{\sigma}$ can be also usefully expressed, within a comoving coordinate system, in
314: units of $\mathrm{Mpc}^{-1}$:
315: \begin{equation}
316: k_{1} = 1.1\times 10^{24} (\epsilon {\mathcal P}_{\mathcal R})^{1/4}\,\,\mathrm{Mpc}^{-1},\qquad
317: {\it k}_{\sigma} = 1.55 \times 10^{12}
318: \biggl( \frac{ h_{0}^2 \Omega_{\mathrm{b}0}}{0.023}\biggr)^{1/2} \biggl(\frac{h_{0}}{0.7}\biggr)^{1/2} \, \mathrm{Mpc}^{-1},
319: \label{scales}
320: \end{equation}
321: where $\epsilon$ is the slow-roll parameter already encountered in Eq. (\ref{PSI3}) and ${\mathcal P}_{\mathcal R}
322: \simeq 2.35\times 10^{-9}$ is the inflationary power spectrum of curvature perturbations evaluated at
323: the pivot scale $k_{\mathrm{p}} = 0.002\,\mathrm{Mpc}^{-1}$ and estimated according to the WMAP data
324: alone \cite{WMAP}.
325: The wave-numbers of Eq. (\ref{scales}) indeed correspond to very short wavelengths as it can be appreciated
326: by comparing them to the comoving wave-number corresponding to the Hubble radius, i.e.
327: $k_{0} = 2.33 \times 10^{-4} (h_{0}/0.7)\,\, \mathrm{Mpc}^{-1}$. The spectrum of Eq. (\ref{M16}) holds
328: for $k < k_{1}$. But since the exponential fall-off triggered by the finite value of the conductivity
329: becomes relevant already at $k\simeq k_{\sigma}$ the power-law
330: behaviour is only verified for sufficiently small wave-numbers $k < k_{\sigma}$.
331: The two-point function of Eq. (\ref{M15}) has been computed by
332: quantizing the system in the Coulomb gauge and by solving the resulting evolution equations in the Heisenberg representation. The final
333: result (\ref{M15}) can also be expressed as a statistical condition on the (classical) Fourier amplitudes
334: \begin{equation}
335: \langle B_{i} (\vec{k}) B_{j}^{*}(\vec{p})\rangle = \frac{2\pi^2}{k^3} P_{B}(k) P_{ij}(k)
336: \delta^{(3)}(\vec{k} - \vec{p}).
337: \label{M17}
338: \end{equation}
339: \begin{figure}
340: \begin{center}
341: \begin{tabular}{|c|c|}
342: \hline
343: \hbox{\epsfxsize = 6.9 cm \epsffile{FIG1.eps}} &
344: \hbox{\epsfxsize = 6.9 cm \epsffile{FIG2.eps}}\\
345: \hline
346: \end{tabular}
347: \end{center}
348: \caption{The evolution of $\lambda$ (plot at the left) and of $\sigma$ (plot at the right) is reported
349: for different parameters of the model.}
350: \label{FIGURE1}
351: \end{figure}
352: The analytical calculation will now be corroborated by the appropriate numerical treatment
353: where the transition from inflation to radiation is parametrized by
354: \begin{equation}
355: a(\tau) = a_{1} ( x + \sqrt{x^2 + 1}),\qquad g^2 \lambda(x) = \biggl(\frac{2 \sqrt{x^2 +1}}{\sqrt{x^2 + 1} + x} \biggr)^{\frac{3\alpha}{2}},\qquad x = \frac{\tau}{\tau_{1}},
356: \label{M18}
357: \end{equation}
358: where $g$ denotes the hypercharge coupling constant. The time $\tau_{1}$ controls
359: the duration of the transition regime: for $\tau \ll -\tau_{1}$, $\lambda\simeq (- \tau)^{3 \alpha}$ as implied (to leading order in the slow-roll corrections) by the third relation in Eq. (\ref{Y1}) in conjunction with
360: Eq. (\ref{PSI3}). Similarly, if $\tau \ll -\tau_{1}$ the scale factor appearing in Eq. (\ref{M18}) goes as $a(\tau) \simeq (-\tau_{1}/\tau)$ (quasi de-Sitter expansion). Conversely, if $\tau \gg \tau_{1}$, $a(\tau) \simeq (\tau/\tau_{1})$ (radiation
361: dominated evolution) and $g^2 \lambda\to 1$.
362: The evolution of $\lambda$ is graphically illustrated in Fig. \ref{FIGURE1}) (plot at the left).
363: The time evolution of the conductivity can be modeled as
364: \begin{equation}
365: \sigma_{\mathrm{c}}(x) = \frac{T_{\mathrm{rh}}}{\alpha} \theta(x),\qquad
366: \theta(x) = \frac{1}{8} \biggl( 1 + \frac{x}{\sqrt{x^2 + 1}}\biggr)^{3},
367: \label{SM5}
368: \end{equation}
369: where $\theta(x)$ is a smooth representation of the Heaviside step function. Notice
370: that the rationale for the third power stems from the fact that
371: $\sigma_{\mathrm{c}}(\tau)$ should vanish fast enough for $\tau\ll -\tau_{1}$. The graphic illustration of the evolution
372: of $\sigma$ is reported in Fig. \ref{FIGURE1} (plot at the right).
373: When the electroweak symmetry is unbroken the conductivity $\sigma_{\mathrm{c}}$
374: is of the order of $ T/\alpha$ where $\alpha= g^2/4\pi$ and $T$ is the temperature at the corresponding
375: epoch. More accurate
376: estimates of this quantity exist (see, for instance, \cite{max4} and
377: \cite{enqv}) and they agree, up to numerical factors, with the
378: figures used in the present paper. In fact, $\sigma_{\mathrm{c}}$ is anyway much larger than the Hubble rate at the corresponding epoch. By relying on the assumption that all the inflaton energy density is efficiently converted into radiation energy density and by assuming a generic number
379: $N_{\mathrm{eff}}$ of relativistic degrees of freedom $T_{\mathrm{rh}}$ can be estimated as
380: \begin{equation}
381: \frac{T_{\mathrm{rh}}}{H_{1}} = \biggl(\frac{45}{4\pi^4 N_{\mathrm{eff}}}\biggr)^{1/4} (\epsilon\,{\mathcal P}_{{\mathcal R}})^{-1/4},\qquad
382: {\mathcal P}_{\mathcal R} = \frac{8}{3} \frac{V}{\epsilon M_{\mathrm{P}}^{4}}
383: \equiv \frac{1}{24 \pi^2}\frac{V}{\epsilon\overline{M}_{\mathrm{P}}^4},
384: \label{TRH}
385: \end{equation}
386: where the slow-roll equation $ 3 H_{1}^2 \overline{M}_{\mathrm{P}}^2 \simeq V$ has been used. Even if $N_{\mathrm{eff}} =106.75$ in the standard model, a drastic variation of one order
387: of magnitude does not affect crucially $\sigma$ (see also Fig. \ref{FIGURE1}).
388: \begin{figure}
389: \begin{center}
390: \begin{tabular}{|c|c|}
391: \hline
392: \hbox{\epsfxsize = 7 cm \epsffile{FIG3.eps}} &
393: \hbox{\epsfxsize = 7 cm \epsffile{FIG4.eps}}\\
394: \hline
395: \end{tabular}
396: \end{center}
397: \caption{The numerical result for the evolution of the
398: hypermagnetic and hyperelectric power spectra
399: is illustrated on a semi-logarithmic scale.}
400: \label{FIGURE2}
401: \end{figure}
402: Recalling that $\sigma(\tau)= \sigma_{\mathrm{c}}(\tau) a(\tau)$ the evolution of the mode function in the presence of the Ohmic
403: terms
404: \begin{equation}
405: f_{k}'' + \frac{4\pi\sigma}{\lambda} f_{k}'+ \biggl\{k^2 - \biggl[ \frac{(\sqrt{\lambda})''}{\sqrt{\lambda}} +
406: \frac{4\pi \sigma}{\lambda} \frac{(\sqrt{\lambda})'}{\sqrt{\lambda}}\biggr]\biggr\} f_{k}=0,
407: \end{equation}
408: can be solved in the smooth background provided
409: by Eqs. (\ref{M18}) and (\ref{SM5}).
410: Imposing
411: quantum mechanical initial conditions on $f_{k}$ (i.e. $f_{k} = e^{- i k \tau }/\sqrt{2 k}$ for $\tau\to -\infty$)
412: the hypermagnetic and hyperelectric power spectra can be obtained and the results are summarized in Fig. \ref{FIGURE2}
413: and in the left plot of Fig. \ref{FIGURE3}.
414: According to Eqs. (\ref{M16}) and (\ref{M16a}), if
415: $\alpha= 1$ (up to slow-roll corrections) $n_{\mathrm{B}} \simeq 4$. Similarly, if $\alpha=2$, $n_{\mathrm{B}}\simeq 1$
416: and the magnetic power spectrum is nearly scale-invariant.
417: In Fig. \ref{FIGURE2} the hyperelectric and hypermagnetic
418: power spectra have been numerically computed in the case
419: $\alpha = 1$ and for different values of $\kappa = k\tau_{1}$, i.e. the comoving wave-number in units
420: of the transition time $\tau_{1}$.
421: The initial integration time $x_{i} =\tau_{i}/\tau_{1}$ depends on the mode and it is chosen in such a way that $\kappa x_{i}> 1$ at $x_{i}$ so that each mode starts its evolution
422: inside the Hubble radius.
423: By comparing the corresponding values of $\kappa$ in the left and right plots
424: of Fig. \ref{FIGURE2} the hypermagnetic power spectrum is amplified while the hyperelectric power spectrum is exponentially suppressed.
425: \begin{figure}
426: \begin{center}
427: \begin{tabular}{|c|c|}
428: \hline
429: \hbox{\epsfxsize = 7 cm \epsffile{FIG5.eps}} &
430: \hbox{\epsfxsize = 7 cm \epsffile{FIG6.eps}}\\
431: \hline
432: \end{tabular}
433: \end{center}
434: \caption{Comparison between analytical and numerical
435: results in the case $\alpha=1$ and $\epsilon =0.01$ (plot at the left). The hypermagnetic spectrum as a function of the comoving
436: wave-number in units of $\mathrm{Mpc}^{-1}$ (plot at the right).}
437: \label{FIGURE3}
438: \end{figure}
439: In Fig. \ref{FIGURE2} (plot at the left) the magnetic power spectrum is reported for different values of the
440: wave-number. The amplitude increases with $\kappa$,
441: which is exactly what we expect in the case of $\alpha=1$
442: where the magnetic power spectrum
443: should scale, approximately, as $\kappa^{n_{\mathrm{B}} -1}$ with $n_{\mathrm{B}} = 4$.
444: In Fig. \ref{FIGURE3} (plot at the left) a more detailed comparison is illustrated. The starred points correspond to results
445: of the numerical integration for different values of the $\kappa$ while the dashed line
446: corresponds to the analytical result. From Eqs. (\ref{M16}) and (\ref{M16a}), in the case $\delta = 1$, we obtain
447: \begin{equation}
448: \log{\biggl[\frac{P_{\mathrm{B}}(\kappa)}{H_{1}^4}\biggr]}= \log{2} - 2 \log{\pi} + \frac{3 - 4\epsilon}{1-\epsilon} \log{\kappa},
449: \label{FIT1}
450: \end{equation}
451: where we identified
452: $k_{1} \simeq \tau_{1}^{-1}$ so that $k/k_{1} \simeq \kappa$.
453: The dashed line in the left plot of Fig. \ref{FIGURE3}
454: is not a fit but it is the result of the analytical expectation.
455: Similar agreement is reached for different values of $\alpha$.
456: Consequently, the analytical results based on the
457: sudden approximation in conjunction with the matching
458: conditions expressed by Eq. (\ref{M3})
459: are in good agreement with the numerical integration across
460: a smooth transition of the same system of equations.
461:
462: When the hypermagnetic fields will reenter the Hubble radius (prior to equality but after neutrino decoupling, taking place around the MeV) the electroweak symmetry is already broken. The non-screened vector modes of the
463: hypercharge field will the project on the electromagnetic fields as
464: ${\mathcal A}_{i}^{\mathrm{em}} = \cos{\theta_{\mathrm{w}}} {\mathcal Y}_{i}$. The final magnetic power spectrum can then
465: be presented (see Fig. \ref{FIGURE3}, plot at the right) in units of
466: $H_{1}^4$, i.e. the fourth power of the Hubble rate at the end of inflation.
467: \begin{figure}
468: \begin{center}
469: \begin{tabular}{|c|c|}
470: \hline
471: \hbox{\epsfxsize = 7 cm \epsffile{FIG7.eps}} &
472: \hbox{\epsfxsize = 7 cm \epsffile{FIG8.eps}}\\
473: \hline
474: \end{tabular}
475: \end{center}
476: \caption{The hypermagnetic power spectrum for different choices of the parameters and as a function of the comoving wave-number in units of $\mathrm{Mpc}^{-1}$ (plot at the left). The temperature autocorrelations
477: of the CMB anisotropies computed in the nearly scale-invariant limit according to the technique described
478: in \cite{max3} (see, in particular, third reference).}
479: \label{FIGURE4}
480: \end{figure}
481: A more physical measure of the value
482: of the obtained magnetic fields is the radiation energy density. The
483: magnetic power spectrum in units of the radiation background is then
484: \begin{equation}
485: \frac{P_{\mathrm{B}}(k)}{8\pi \rho_{\gamma}} = \pi \cos^2{\theta_{\mathrm{w}}} {\cal C}(\delta) \,\epsilon\,{\mathcal P}_{\mathcal R} \biggl(
486: \frac{k}{k_{1}}\biggr)^{n_{\mathrm{B}} -1},
487: \label{magnrad}
488: \end{equation}
489: where both $n_{\mathrm{B}}$ and $\delta$ depend upon the slow-roll parameter $\epsilon$. Since \cite{WMAP} ${\mathcal P}_{{\mathcal R}} \simeq 2.35 \times 10^{-9}$, in the scale-invariant limit Eq. (\ref{magnrad}) is of the order of $10^{-10}$.
490: Consequently, the
491: present value of the magnetic field is of order $0.1$--$0.01$ nG with a theoretical error
492: that depends upon $\epsilon$ (which should be smaller than about $0.05$ according to
493: current experimental data\footnote{It is possible to obtain an upper bound on $\epsilon$ by
494: analyzing, for instance, CMB and large-scale structure data within a $\Lambda$CDM model
495: containing also a tensor component. The analysis will then lead to an upper limit on the
496: ratio between tensor and scalar power spectra which can be translated into an upper limit on $\epsilon$. The combination of WMAP \cite{WMAP} data and the data of the Sloan Digital Sky Survey \cite{SDSS} would lead, for instance to $\epsilon \laq 0.02$}).
497:
498: The results of Eq. (\ref{magnrad}) can also be illustrated by regularizing
499: the magnetic field over a typical comoving scale $L$ by means of a Gaussian window function in Fourier space
500: \cite{max3,tin}. Denoting as $B_{\mathrm{L}}$ the regularized magnetic field over the comoving scale
501: $L = 2\pi/k_{\mathrm{L}}$ we will have, in the nearly scale-invariant limit and at the time of the collapse of the protogalaxy,
502: \begin{equation}
503: \biggl(\frac{B_{\mathrm{L}}}{\mathrm{nG}}\biggr) \simeq 0.1 \biggl(\frac{\epsilon}{0.01}\biggr)^{1/2} \biggl(\frac{{\mathcal P}_{\mathcal R}}{2.35 \times 10^{-9}}\biggr)^{1/2}.
504: \label{magnreg}
505: \end{equation}
506: The magnetic energy density in units
507: of the radiation background can be expressed, in this case, as
508: \begin{equation}
509: \overline{\Omega}_{\mathrm{BL}} = \frac{B_{\mathrm{L}}^2}{8\pi \rho_{\gamma}} =
510: 7.56 \times 10^{-9} \, \biggl(\frac{B_{\mathrm{L}}}{\mathrm{nG}}\biggr)^2,
511: \label{OM1}
512: \end{equation}
513: where the pivot value of $B_{\mathrm{L}}$ has been taken at the epoch of gravitational collapse
514: of the protogalaxy.
515: It is customary
516: to require, for a successful magnetogenesis \cite{weyl1}, that\footnote{The first of these two figures stems from
517: the (overoptimistic) requirement that the galactic
518: dynamo is so efficient to amplify the protogalactic field by one e-fold for each galactic rotation. Strictly speaking this argument only applies to spiral galaxies. The second requirement
519: takes into account the possible early saturation of galactic dynamo and it is still rather optimistic (see, for instance, the second reference of \cite{zel1} for a discussion). } $\overline{\Omega}_{\mathrm{BL}} > 10^{-34}$, or, more realistically $\overline{\Omega}_{\mathrm{BL}} > 10^{-24}$.
520:
521: Equation (\ref{magnreg}) is compatible with a galactic magnetic field of the order of the $\mu$G.
522: During the process
523: of collapse the magnetic flux is frozen into the plasma element thanks to the large value of the conductivity. The mean matter
524: density increases, during collapse,
525: from its critical value (i.e. $\rho_{\mathrm{cr}} =1.05\times 10^{-5} \,h_{0}^2 \mathrm{GeV}/\mathrm{cm}^3$) to a final value $\rho_{\mathrm{f}}$ value which is 5 to 6 orders of magnitude larger than $\rho_{\mathrm{c}}$. The magnetic field after collapse will then be
526: \begin{equation}
527: B_{\mathrm{gal}} = \biggl(\frac{\rho_{\mathrm{f}}}{\rho_{\mathrm{c}}}\biggr)^{2/3} B_{\mathrm{L}} \simeq \biggl(\frac{\epsilon}{0.01}\biggr)^{1/2} \biggl(\frac{{\mathcal P}_{\mathcal R}}{2.35 \times 10^{-9}}\biggr)^{1/2}\, \mu\mathrm{G}.
528: \end{equation}
529: Over present length-scales much larger than the Mpc the magnetic fields, in this model, will be (today) smaller than the nG since these regions did not benefit of the compressional amplification. Within this lore the magnetic fields in clusters could be produced by magnetic reconnection from the ones of the galaxies but the experimental uncertainty in their correlation scale \cite{clust1} does not allow a definite statement. If
530: the spectrum of the primeval field is not nearly scale-invariant its amplitude over a comoving Mpc scale will be smaller (see right plot of Fig. \ref{FIGURE3}) and, consequently, a non negligible
531: dynamo action will be required for the phenomenological relevance of the obtained result. In this second scenario the
532: cluster magnetic field might be related to the way the dynamo is saturated.
533:
534: In a series of papers a semi-analytical technique has been developed
535: for the evaluation of the temperature autocorrelations
536: (see, in particular, the third reference in \cite{max3}). Since
537: in the present model the cross-correlation between
538: magnetic and adiabatic contribution vanishes the temperature
539: cross-correlations are given, for multipoles $\ell < 30$ by
540: the following generalization of the Sachs-Wolfe plateau:
541: \begin{eqnarray}
542: C^{(\mathrm{SW})}_{\ell} &=& \biggl[ \frac{{\cal P}_{\mathcal R}}{25} \,{\mathcal Z}_{1}(n,\ell) +
543: \frac{\epsilon^2 {\mathcal P}^2_{\mathcal R}}{400} \, R_{\gamma}^2 {\mathcal Z}_{2}(n_{\mathrm{B}},\ell) \biggr],
544: \label{SW1}\\
545: {\mathcal Z}_{1}(n,\ell) &=& \frac{\pi^2}{4} \biggl(\frac{k_0}{k_{\rm p}}\biggl)^{n-1} 2^{n} \frac{\Gamma( 3 - n) \Gamma\biggl(\ell +
546: \frac{ n -1}{2}\biggr)}{\Gamma^2\biggl( 2 - \frac{n}{2}\biggr) \Gamma\biggl( \ell + \frac{5}{2} - \frac{n}{2}\biggr)},
547: \label{Z1}\\
548: {\mathcal Z}_{2}(n_{\mathrm{B}},\ell) &=& \frac{\pi^2}{2} 2^{2(n_{\mathrm{B}} -1)} {\cal F}(n_{\mathrm{B}}) \biggl( \frac{k_{0}}{k_{1}}\biggr)^{ 2(n_{\mathrm{B}} -1)} \frac{ \Gamma( 4 - 2 n_{\mathrm{B}})
549: \Gamma(\ell + n_{\mathrm{B}} -1)}{\Gamma^2\biggl(\frac{5}{2} - n_{\mathrm{B}}\biggr)
550: \Gamma(\ell + 3 - n_{\mathrm{B}})},
551: \label{Z2}\\
552: {\mathcal F}(n_{\mathrm{B}}) &=& \frac{4 \pi^2}{27} {\mathcal C}^2(\delta)\frac{(7 - n_{\mathrm{B}}) }{(n_{\mathrm{B}} -1) ( 5 - 2 n_{\mathrm{B}})
553: },\qquad n_{\mathrm{B}}>1,
554: \label{Fnb}
555: \end{eqnarray}
556: where $n$ denotes the spectral index of the adiabatic mode\footnote{For the numerical estimate of Fig. \ref{FIGURE4}, $n$ will be taken to be $0.958$ which is the best fit value obtainable by analyzing the
557: WMAP data alone \cite{WMAP}.}. If $n_{\mathrm{B}}\simeq 1$
558: the function (\ref{Fnb}) contains the logarithm of the infra-red cut-off.
559:
560: At smaller angular scales (i.e. $\ell > 100$) the temperature
561: autocorrelations can be obtained, within the scheme of \cite{max3}
562: (third reference) by computing numerically four integrals
563: and the results for multipoles compatible with the Doppler oscillations
564: are reported in Fig. \ref{FIGURE4} (plot at the right). When $B_{\mathrm{L}} \simeq 0.1 $ nG the structure of the Doppler oscillations is not altered (see also \cite{max3} for a model-independent discussion).
565:
566: So far it has been assumed that the decay rate of the inflaton
567: and of the spectator fields were comparable, i.e. $\Gamma_{\psi} \simeq \Gamma_{\varphi}$. It can also happen that
568: the spectator field decays later than the inflaton. The predicted slopes of the magnetic power spectra will not be modified for comoving scales of the order of the Mpc. However, shorter wavelengths can be affected if the decay of $\psi$ is delayed.
569: In the latter case since $\psi_{*} < \overline{M}_{\mathrm{P}}$, the ratio between
570: the energy density of $\psi$ and the radiation background (produced by the inflaton decay) will
571: grow, after inflation. The fluctuations of the spectator field
572: may then represent a further source of curvature perturbations.
573: If the inflationary Hubble rate is much smaller than $10^{-5} \overline{M}_{\mathrm{P}}$ the fluctuations of $\psi$ will eventually become the dominant source of curvature perturbations as noticed in the context of the
574: so-called curvaton scenario \cite{curv1}. In the opposite case the two contributions will interfere.
575: In the latter case the final spectrum
576: of curvature perturbations can be computed, for different post-inflationary
577: evolutions, as a function of $\psi_{*}$. This analysis can be carried on numerically with the techniques already exploited in
578: \cite{curv2}. The
579: final result can be written in terms of the amplitude
580: of the curvature perturbations at the pivot scale:
581: \begin{equation}
582: {\mathcal P}_{\mathcal R} = \frac{1}{24\pi^2} \frac{V}{\overline{M}_{\mathrm{P}}^4}\biggl[ \frac{1}{\epsilon} + f^2(\psi_{*})\biggr],\qquad
583: f(\psi_{*}) = c_{1} \biggl(\frac{\psi_{*}}{\overline{M}_{\mathrm{P}}}\biggr)
584: + c_{2} \biggl(\frac{\overline{M}_{\mathrm{P}}}{\psi_{*}}\biggr)
585: \label{CC1}
586: \end{equation}
587: where $c_{1}= 0.13$ and $c_2= 0.25$. In the limit $f\to 0$
588: we recover the result of Eq. (\ref{TRH}). In the case $f\neq 0$
589: the curvature fluctuations induced by $\psi$ may mix, in the Sachs-Wolfe
590: plateau, with the component induced by the inflaton fluctuations.
591: In some cases there could even be a correlation term. As argued
592: in \cite{curv2} (second reference) these results strongly depend upon
593: $W(\psi)$ being quadratic and not, for instance quartic. In spite
594: of the details of the post-inflationary history Eq. (\ref{CC1}) suggests a possible violation of the consistency relation
595: which would become, in the case of Eq. (\ref{CC1}),
596: $r_{\mathrm{T}} = 16 \epsilon/[1 + 8 f^2(\psi_{*}) \epsilon]$ having
597: defined $r_{\mathrm{T}} = {\mathcal P}_{\mathrm{T}}/{\mathcal P}_{\mathcal R}$, i.e. the ratio between the tensor and the scalar
598: power spectra. This prediction would allow, in principle, to distinguish observationally
599: the situations where the spectator field decays during reheating or later. We leave for a
600: forthcoming paper the detailed analysis of this and other cases \cite{mg}.
601:
602: The main goal of the present study has been
603: to demonstrate, within conventional inflationary scenarios, the viability of a class of magnetogenesis models
604: that do not require a strong dynamo action and that
605: are compatible, at the same time, with the direct bounds stemming from the analysis of the CMB anisotropies.
606: The foreseeable improvement of the quality of CMB data stimulates
607: the effort of more accurate calculations of the impact of a magnetized plasma
608: upon the various CMB observables. As explicitly demonstrated in this paper
609: is possible to construct viable magnetogenesis models which have well defined CMB signatures.
610: Since theoretical prejudices (and diatribes)
611: are not a decisive proof for the existence (or not existence) of pre-recombination magnetic fields, it is
612: wise pursue the development of model-independent tools for accurate analyses of magnetized CMB anisotropies,
613: as suggested by the present investigation.
614: Indeed, forthcoming satellite experiments may turn some of the present speculations in more solid scientific
615: statements either in favour or against the primordial hypothesis.
616:
617: \begin{thebibliography}{99}
618: \bibitem{zel1} Ya. B. Zeldovich, A. A. Ruzmaikin, and D.D. Sokoloff,
619: {\it Magnetic Fields in Astrophysics} (Gordon and Breach Science, New York, 1983); M.~Giovannini, Int.\ J.\ Mod.\ Phys.\ D {\bf 13}, 391 (2004).
620:
621: \bibitem{gal1} R. Beck, A. Brandenburg, D. Moss, A. Skhurov, and D. Sokoloff {\it Annu. Rev. Astron. Astrophys.}{\bf 34}, 155 (1996); R. Beck, Astron.Nachr. {\bf 327}, 512 (2006);
622:
623: \bibitem{clust1} C.~L.~Carilli and G.~B.~Taylor, Ann.\ Rev.\ Astron.\ Astrophys.\ {\bf 40}, 319 (2002); T.E. Clarke, P.P. Kronberg and H. B\"ohringer,
624: Astrophys. J. {\bf 547}, L111 (2001).
625:
626: \bibitem{supclust} P.~P.~Kronberg, S.~Habib and Q.~W.~Dufton, Astrophys.\ J.\ {\bf 637}, 19 (2006); P. P. Kronberg, Rep. Prog. Phys. {\bf 57}, 325 (1994).
627:
628: \bibitem{auger1} J. Abraham {\it et al.} [Pierre Auger Collaboration],
629: Science {\bf 318}, 938 (2007).
630:
631: \bibitem{auger2} M.~Aglietta {\it et al.} [Pierre Auger Collaboration],
632: Astropart.\ Phys.\ {\bf 27}, 244 (2007)
633:
634: \bibitem{harrison1} E. Harrison, Phys. Rev. Lett. {\bf 18}, 1011 (1967);
635: Mon. Not. R. Astr. Soc. {\bf 147}, 279 (1970).
636:
637: \bibitem{kanduaxel} A.~Brandenburg and K.~Subramanian,
638: Phys.\ Rept.\ {\bf 417}, 1 (2005);
639: A. Lazarian, E. Vishniac, and J. Cho, Astrophys. J. {\bf 603}, 180
640: (2004); Lect. Notes Phys. {\bf 614}, 376 (2003).
641:
642: \bibitem{max2} M.~Giovannini, Class.\ Quant.\ Grav.\ {\bf 23}, R1 (2006);
643: K.~Subramanian, Astron.Nachr. {\bf 327}, 399 (2006); T.~Kahniashvili,
644: New Astron.\ Rev.\ {\bf 50}, 1015 (2006).
645:
646: \bibitem{max3} M.~Giovannini, Phys.\ Rev.\ D {\bf 73}, 101302 (2006);
647: Phys.\ Rev.\ D {\bf 74}, 063002 (2006); PMC Phys.\ A {\bf 1}, 5 (2007).
648:
649: \bibitem{weyl1} M.~S.~Turner and L.~M.~Widrow,
650: Phys.\ Rev.\ D {\bf 37}, 2743 (1988); B.~Ratra, Astrophys.\ J.\ {\bf 391}, L1 (1992);
651: M.~Gasperini, M.~Giovannini and G.~Veneziano,
652: Phys.\ Rev.\ Lett.\ {\bf 75}, 3796 (1995); M.~Giovannini,
653: Phys.\ Rev.\ D {\bf 62}, 123505 (2000); Phys.\ Rev.\ D {\bf 64}, 061301 (2001);
654: K.~E.~Kunze, Phys.\ Lett.\ B {\bf 623}, 1 (2005); arXiv:0710.2435 [astro-ph].
655:
656: \bibitem{max4} M.~Giovannini and M.~E.~Shaposhnikov,
657: Phys.\ Rev.\ D {\bf 57}, 2186 (1998); M.~Giovannini, Phys.\ Rev.\ D {\bf 61}, 063004 (2000).
658:
659: \bibitem{sdual} M.~K.~Gaillard and B.~Zumino, Nucl.\ Phys.\ B {\bf 193}, 221 (1981).
660:
661: \bibitem{lford} L. H. Ford, Phys.Rev. D {\bf 31}, 704 (1985).
662:
663: \bibitem{WMAP} D.~N.~Spergel {\it et al.} [WMAP Collaboration], Astrophys.\ J.\ Suppl.\ {\bf 170}, 377 (2007);
664: L.~Page {\it et al.} [WMAP Collaboration], Astrophys.\ J.\ Suppl.\ {\bf 170}, 335 (2007).
665:
666: \bibitem{enqv} J.~Ahonen and K.~Enqvist, Phys.\ Lett.\ B {\bf 382}, 40 (1996); M.~Giovannini,
667: Phys.\ Rev.\ D {\bf 61}, 063502 (2000).
668:
669: \bibitem{SDSS} D.~J.~Eisenstein {\it et al.} [SDSS Collaboration], Astrophys.\ J.\ {\bf 633}, 560 (2005);
670: M.~Tegmark {\it et al.} [SDSS Collaboration], Astrophys.\ J.\ {\bf 606}, 702 (2004).
671:
672: \bibitem{tin} A.~Kosowsky, T.~Kahniashvili, G.~Lavrelashvili and B.~Ratra,
673: Phys.\ Rev.\ D {\bf 71}, 043006 (2005); A.~Mack, T.~Kahniashvili and A.~Kosowsky,
674: Phys.\ Rev.\ D {\bf 65}, 123004 (2002).
675:
676: \bibitem{curv1} K.~Enqvist and M.~S.~Sloth, Nucl.\ Phys.\ B {\bf 626}, 395 (2002);
677: D.~H.~Lyth and D.~Wands, Phys.\ Lett.\ B {\bf 524}, 5 (2002).
678:
679: \bibitem{curv2} M.~Giovannini, Phys.\ Rev.\ D {\bf 67}, 123512 (2003);
680: V.~Bozza, M.~Gasperini, M.~Giovannini and G.~Veneziano, Phys.\ Rev.\ D {\bf 67}, 063514 (2003).
681:
682: \bibitem{mg} M. Giovannini, in preparation.
683: \end{thebibliography}
684:
685:
686: \end{document}