1: % Please use the skeleton file you have received in the
2: % invitation-to-submit email, where your data are already
3: % filled in. Otherwise please make sure you insert your
4: % data according to the instructions in PoSauthmanual.pdf
5: \documentclass{PoS}
6: \newcommand{\bb}{\begin{equation}}
7: \newcommand{\ee}{\end{equation}}
8: \newcommand{\eqb}{\begin{eqnarray}}
9: \newcommand{\eqf}{\end{eqnarray}}
10: \frenchspacing
11: \newcommand{\hs}{/\kern-.52em h}
12: \newcommand{\D}{{\cal D}}
13: \newcommand{\f}{F^{\mu\nu}}
14: \newcommand{\DB}{\delta_{BREST}}
15: \newcommand{\QB}{Q_{BREST}}
16: \newcommand{\GG}{\Gamma^{2n}}
17: \newcommand{\GGG}{\Gamma^{*2(n-m)}}
18: \newcommand{\med}{\D \bar \psi \D \psi}
19: \newcommand{\medp}{\D \bar \psi' \D \psi'}
20: \newcommand{\id}
21: {i\kern.06em\hbox{\raise.25ex\hbox{$/$}\kern-.60em$\partial$}}
22: \newcommand{\as}{/\kern-.68em A}
23: \newcommand{\Ds}{/\kern-.69em D}
24: \newcommand{\vs}{\varphi^{0}_n}
25: \newcommand{\ks}{/\kern-.67em k}
26: \newcommand{\Ps}{/\kern-.65em p}
27: \newcommand{\uh} {\frac{1}{\hbar}}
28: \newcommand{\BBs}{\!\not\!\! B}
29: \newcommand{\rD}{\!\not\!\! D}
30: \newcommand{\bs}{/\kern-.52em b}
31: \newcommand{\qs}{/\kern-.52em s}
32: \newcommand{\dv}{\!d^3\!x\,}
33: \newcommand{\Z}{{\cal Z}}
34: \def\p{{\bf p}}
35: \def\x{{\bf x}}
36: \def\bn{{\boldsymbol{ \nabla}}}
37: \def\bth{{\boldsymbol{ \theta}}}
38: \usepackage{amsmath}
39: \def\B{{\bf{ B}}}
40: \frenchspacing
41:
42:
43: \title{CPT/Lorentz Invariance Violation and Quantum Field Theory}
44:
45: \ShortTitle{Noncommutative fields}
46:
47: \author{\speaker{Jorge Gamboa}\\
48: Departamento de F\'{\i}sica, Universidad de Santiago de Chile\\
49: Casilla 307, Santiago 2, Chile\\
50: E-mail: \email{jgamboa@lauca.usach.cl}}
51:
52: \author{Paola Arias\\
53: Departamento de F\'{\i}sica, Universidad de Santiago de Chile\\
54: Casilla 307, Santiago 2, Chile\\
55: E-mail: \email{paola.arias@gmail.com}}
56:
57: \author{Ashok Das\\
58: Department of Physics and Astronomy, University of
59: Rochester, Rochester, NY 14627-0171, USA \\
60: and
61: \\
62: Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Calcutta 700064, India.
63: \\
64: E-mail: \email{das@pas.rochester.edu}}
65:
66: \author{Justo Lopez-Sarrion\\
67: Department of Physics, City College of CUNY\\
68: New York, NY 10031, USA\\
69: E-mail: \email{justinux75@gmail.com}}
70:
71: \author{Fernando Mendez\\
72: Departamento de F\'{\i}sica, Universidad de Santiago de Chile\\
73: Casilla 307, Santiago 2, Chile\\
74: E-mail: \email{ fmendez@lauca.usach.cl}}
75:
76: \abstract{Analogies between the noncommutative harmonic oscillator and
77: noncommutative fields are analyzed. Following this analogy we
78: construct examples of quantum fields theories with explicit CPT and
79: Lorentz symmetry breaking. Some applications to baryogenesis and
80: neutrino oscillation are also discussed.
81: }
82:
83: \FullConference{Fifth International Conference on Mathematical Methods in Physics --- IC2006\\
84: April 24-28 2006\\
85: Centro Brasilerio de Pesquisas Fisicas, Rio de Janeiro, Brazil}
86:
87: \begin{document}
88:
89: \section{Introduction}
90:
91: The Lorentz and CPT symmetries are fundamental cornerstones in the
92: twentieth century for which there are not clear indications if
93: persists at very high energies. By other hand, from effective
94: theories point of view, it seems natural to think that the presently
95: quantum field theories (QFT) and the symmetry principles they are
96: based on, remain valid just for a given range of energies beyond
97: which, possibly new and unexpected phenomena could emerge. For
98: example, the relativistic invariance itself could be broken or
99: deformed \cite{effective}.
100:
101: If QFT describes fundamental interactions for any energy range,
102: then it seems natural to think that any QFT --seen as effective
103: theory-- must incorporate two energy scales, namely, the infrared and
104: ultraviolet ones. Both scales might give rise to non conventional
105: implications.
106:
107: There are at least two examples of the previous idea, where the
108: infrared scale can be very important. The first one, is the physics
109: involved in the infrared sector of QED where still several technical
110: aspects need to be understood as well as many conceptual
111: problems still remain open \cite{yennie}.
112:
113: The infrared sector of QED is the natural link between quantum field
114: theory and quantum mechanics and then we ask which are the problems and how
115: can we understand the physics in this interface?, what are the most
116: convenient approximation criteria? In spite of many efforts
117: performed during the fifties and sixties this problem still have not
118: been clarified.
119:
120: In the same context, another important example is QCD where the physical picture
121: in the infrared limit is nontrivial because, at low
122: energy, the theory is nonperturbative and phenomena such as
123: confinement or hadronization should be solved using new methods beyond
124: perturbation theory.
125:
126: In this paper we would like to report our previous results on the idea
127: of how the Lorentz symmetry could be broken in a QFT and also to point
128: out new progress in the application of this to neutrino physics.
129:
130: In order to expose our results, we will start in section II describing
131: noncommutative mechanics and the harmonic oscillator in order to introduce, in the
132: same section, the notion of noncommutative scalar and gague fields. In section III we
133: we will report our recent progress in neutrino physics when noncommutative
134: fermion fields are introduced. Section IV is devoted to resume the results of
135: previous ideas when they are applied to the study of early universe physics. In the last
136: section, conclusion and outlook is presented.
137:
138: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
140: \section{Noncommutative quantum mechanics and noncommutative
141: fields}
142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
144: In this section we will study
145: the non commutative quantum harmonic
146: oscillator and how it can be used to define a non commutative field
147: theory. As a warm up exercise let us consider first the standard
148: bidimensional quantum harmonic oscillator described by the following
149: Hamiltonian operator
150: \begin{equation}
151: H=\frac{\omega}2[p_1^2 +p_2^2+q_1^2+q_2^2], \label{ha}
152: \end{equation}
153: with standard commutation relations ($i,j=1,2$)
154: \begin{eqnarray}
155: \left[q_i,q_j\right]&=&\left[p_i,p_j \right]=0, \nonumber
156: \\
157: \left[q_i,p_j\right] &=& i \delta_{ij}. \label{us}
158: \end{eqnarray}
159: Note that variables $\{q_i,p_j\}$ are dimensionless and are related
160: with usual ones $\{Q_i,P_j\}$ by $q_i=\sqrt{m\omega}Q_i$ and
161: $p_i=(m\omega)^{-1/2} P_i$.
162:
163: The dynamics of this system is described by the Heisenberg equations
164: \begin{equation}
165: i\,\frac {d{\cal O}}{dt}=[{\cal O}, H]
166: \end{equation}
167: which, specified to (\ref{ha}) and (\ref{us}) give rise to
168: \begin{subequations}
169: \label{mov}
170: \begin{eqnarray}
171: \dot{q_i}&=&\omega p_i, \label{eq1}
172: \\
173: \dot{p_i}&=& -\omega q_i. \label{eq2}
174: \end{eqnarray}
175: \end{subequations}
176: \noindent The system (\ref{mov}) is equivalent to the very well known
177: second order differential equation
178: \begin{equation}
179: \ddot{q_i}= - \omega^2 \,q_i. \label{sec}
180: \end{equation}
181: Therefore, the solution of (\ref{mov}) turn out to be
182: \begin{eqnarray}
183: \label{solu2}
184: q_i (t) &=& A_i ~ e^{i\omega t} + B_i ~e^{-i \omega t},\nonumber
185: \\
186: p_i(t) &=& i A_i ~e^{i \omega t} - i ~ B_i ~e^{-i \omega t}.
187: \end{eqnarray}
188:
189: The algebra of operators $A_i$ and $B_i$ can be fixed using the
190: canonical algebra (\ref{us}). Indeed, replacing (\ref{solu2}) in
191: (\ref{us}) we find that
192: \begin{eqnarray}
193: \left[A_i, A_j \right] &=&\left[ B_i, B_j \right] = 0, \nonumber
194: \\
195: \left[ A_i,B_j \right] &=& -\frac{1}{2}\delta_{ij}. \label{algeb}
196: \end{eqnarray}
197: Then we can identify
198: \begin{equation}
199: \sqrt{2}A_i \rightarrow a^{\dagger}_i, \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
200: \sqrt{2}B_i \rightarrow a_i,
201: \label{sealge}
202: \end{equation}
203: and the algebra (\ref{algeb}) becomes
204: \begin{eqnarray}
205: \left[a_i, a_j \right] &=&\left[ a^{\dagger}_i, a^{\dagger}_j \right]
206: = 0, \nonumber
207: \\
208: \left[ a_i,a^{\dagger}_j \right] &=& \delta_{ij} . \label{algeb1}
209: \end{eqnarray}
210:
211: In terms of $a_i^{\dagger}$ and $a_i$, as is well known, one find that
212: the Hamiltonian becomes
213: \begin{equation}
214: H = \omega ( a^{\dagger}_1 a_1+ a^{\dagger}_2 a_2 + 1), \label{haa}
215: \end{equation}
216: The construction of the Hilbert space is straightforward.
217:
218: Using these results let us solve the noncommutative harmonic
219: oscillator described by the Hamiltonian (\ref{ha}) but commutation
220: relations deformed as follows:
221: \begin{equation}
222: \label{conmus}
223: [q_i,q_j]=i\theta \epsilon_{ij},\,\,\,\,\,\,\,[p_i,p_j]=i{\cal
224: B}, \epsilon_{ij}\,\,\,\,\,\,\,[q_i,p_j]=i\delta _{ij},
225: \end{equation}
226: where $\theta$ and ${\cal B}$ are parameters \lq \lq measuring"
227: noncommutativity in $q$ and $p$ respectively\footnote{Note that this
228: parameters are dimensionless in our notation, but they actually have
229: dimensions in the standard variables.}.
230:
231: Using the Hamiltonian (\ref{ha}) and (\ref{conmus}) one find that
232: equations of motion are
233: \begin{subequations}
234: \label{ncq}
235: \begin{eqnarray}
236: \dot{q_i}&=&\omega (p_i +\theta\,\epsilon_{ij} q_j),
237: \\
238: \dot{p_i}&=&\omega (-q_i +{\cal B}\,\epsilon_{ij} p_j). \label{la}
239: \end{eqnarray}
240: \end{subequations}
241:
242: Following the same procedure previously discussed, one find that the
243: analogous to (\ref{sec}) turn out to be now
244: \begin{equation}
245: \ddot{q_i}=\omega(\theta + {\cal
246: B})\,\epsilon_{ij}\,\dot{q_j}+\omega^2({\cal B}\theta
247: -1)\,q_i. \label{senc}
248: \end{equation}
249: Although this set of equations are coupled ones, one decouple these
250: equations using the following trick: let us define the complex
251: variable $Z= q_1 + i q_2 $, then (\ref{senc}) can be written as
252: \begin{equation}
253: \ddot{Z}=-i\,\omega(\theta + {\cal B})\,\dot{Z}+\omega^2({\cal
254: B}\theta -1)\,Z, \label{senc1}
255: \end{equation}
256: and a similar equation for the conjugate $Z^\dag$.
257: This problem is formally is equivalent to the damped harmonic oscillator.
258:
259: As in the standard case, we look now for a solution with the shape
260: \[
261: Z(t) = e^{ \alpha t},
262: \]
263: which imply that
264: \[
265: \alpha^2 + i \omega(\theta + {\cal B})\,\alpha - \omega^2({\cal
266: B}\theta -1)=0 ,
267: \]
268: and as a consequence the possible values of $\alpha$ are
269: \begin{equation}
270: \label{solal}
271: \frac{\alpha_\pm}{\omega}=i\,\left(-\frac{\theta +{\cal B}}{2} \pm
272: \sqrt{1+\left(\frac{\theta -{\cal B}}{2}\right)^2}\right).
273: \end{equation}
274:
275: Thus, the general solution of the noncommutative harmonic oscillator is
276: \begin{equation}
277: \label{z}
278: Z(t)=A_+\,e^{i\alpha_+ t} + A_-\,e^{i\alpha_- t},
279: \end{equation}
280: where $A_\pm$ are complex operators. Note that there is a redefinition
281: of $\alpha_\pm$ since we have factorized the $i$ in (\ref{solal}). It
282: is interesting to note also that the solution (\ref{z}) is a
283: superposition of two oscillation modes, one positive (because
284: $\alpha_+>0$ ) and other negative (because
285: $\alpha_-<0$). Therefore, the solution has the same structure of the
286: standard case, but there is an asymmetry due to the fact that
287: $|\alpha_-|\neq \alpha_+$.
288:
289: Of course from (\ref{z}) it is possible to compute $q_i (t)$, and from
290: the equation of motion we obtain $p_i$
291: \begin{subequations}
292: \label{qa}
293: \begin{eqnarray}
294: q_j&=&\frac{(-i)^{j-1}}{2}\bigg[ a+b-(-)^{j}(a^\dag + b^\dag)
295: \bigg],\\
296: p_j&=&-\frac{(-i)^{j}}{2}\bigg[\lambda_+
297: (a+(-)^{j}a^\dag)+\lambda_-(b+(-)^{j}b^\dag) \bigg].
298: \end{eqnarray}
299: \end{subequations}
300: with $a=A_+e^{i\alpha_+t}$, $b=A_-e^{i\alpha_-t}$ and
301: $\lambda_\pm=\theta+ \alpha_\pm/\omega$
302:
303: Following the example of the commutative case, we must find the
304: commutation relation between operators $A_\pm$ and $A_{\pm}^\dag$ from
305: the known relations (\ref{conmus}).
306:
307: Since the result of this commutators do not depend on time, the
308: following condition fulfills
309: $$
310: [A_+,A_-]=0=[A_+,A_{-}^\dag].
311: $$
312: Note that the remaining commutators, $[A_\pm,A_{\pm}^\dag]$ are
313: obtained from the conditions $[q_1,q_2]=i\theta$ and $[p_1,p_2]=iB$
314: and then, the condition $[q_1,p_1]=i$ is just a consistency
315: check. After a straightforward calculation we obtain
316: \begin{equation}
317: \label{aes}
318: {[}A_\pm,A_{\pm}^\dag{]}=\mp 2 \frac{1+\theta\lambda_\mp}
319: {\lambda_+ -\lambda_-}.
320: \end{equation}
321:
322:
323: This equation shows that non commutative harmonic oscillator in two
324: dimensions is equivalent to two one-dimensional harmonic
325: oscillator.
326:
327: In fact, by a rescaling of $A_\pm$ operators
328: $$
329: \tilde{A}_+=\left(\frac{\lambda_+ -\lambda_-}
330: {2(1+\theta\lambda_-)}\right)^{1/2}~A_+^\dag,~~~~~~~~
331: \tilde{A}_-=\left(\frac{\lambda_+ -\lambda_-}
332: {2(1+\theta\lambda_+)}\right)^{1/2}~A_-,
333: $$
334: and similar relations for $\tilde{A\pm}^\dag$, we find
335: $$
336: [\tilde{A}_+,\tilde{A}_+^\dag]=1=[\tilde{A}_-,\tilde{A}_-^\dag].
337: $$
338: Operators $\tilde{A}_\pm$ play the role of standard lowering
339: and rising operator as in the commutative case. The Hamiltonian of
340: this two dimensional non commutative harmonic oscillator is
341: \begin{equation}
342: \label{hamil}
343: H=\omega_+(\tilde{A}_+\tilde{A}_+^\dag+1/2)+\omega_-(\tilde{A}_-
344: \tilde{A}_-^\dag+1/2).
345: \end{equation}
346: with $\omega_\pm=\sqrt{1+\left(\frac{\theta - B}{2}\right)^2}\pm
347: \left(\frac{\theta+B}{2}\right)$.
348:
349:
350: Previous result is the starting point for constructing noncommutative
351: complex scalar field theory \cite{nos2a}. Consider
352: the standard relativistic Hamiltonian density for a complex scalar
353: field
354: \begin{equation}
355: \label{hamrel}
356: {\cal H}=\Pi^\dag\Pi+{\boldsymbol {\nabla}}\Phi^\dag
357: {\boldsymbol{ \nabla}} \Phi +m^2\Phi^\dag\Phi
358: \end{equation}
359: plus non standard commutation relations
360: \begin{subequations}
361: \label{defcons}
362: \begin{eqnarray}
363: {[}\Phi(\x),\Phi^\dag({\x}'){]}&=&\theta\delta^3(\x-{\x}'),
364: \\
365: {[}\Pi(\x),\Pi^\dag({\x}'){]}&=&B\delta^3(\x-{\x}'),
366: \\
367: {[}\Phi(\x),\Pi({\x}'){]}&=&\delta^3(\x-{\x}'),
368: \end{eqnarray}
369: \end{subequations}
370: where $\theta$ and $B$ parameterizes the non commutativity in the
371: field space and have dimensions of energy$^{-1}$ and energy,
372: respectively.
373:
374: In the standard case $(\theta=0=B)$, quantized fields are a
375: superposition of quantum harmonic oscillators with frequency
376: $\omega(\p)=\sqrt{\p^2+m^2}$, one for each value of
377: momenta $\p$. The structure of these linear superpositions are
378: given by (\ref{solu2}).
379:
380: For non commutative fields, the constructions proceeds in a
381: similar way. The analog of (\ref{solu2}) is given by (\ref{qa}) once
382: they are expressed in terms of operators $\tilde{A}_\pm$.
383: Therefore we consider now a linear superposition with
384: \begin{subequations}
385: \label{ncfields}
386: \begin{eqnarray}
387: \Phi(\x)&=&\int \frac{d^3p}{(2\pi)^3}\frac{1}{\sqrt{\omega}}\left[
388: \eta\epsilon_ 1 a_\p~e^{i\p\x} + \epsilon_2
389: b_\p^\dag~e^{-i\p\x}\right],
390: \\
391: \Pi(\x)&=&i\int \frac{d^3p}{(2\pi)^3}\sqrt{\omega}\left[
392: -\epsilon_ 1 a_\p~e^{i\p\x} +\eta \epsilon_2
393: b_\p^\dag~e^{-i\p\x}\right],
394: \end{eqnarray}
395: \end{subequations}
396: where coefficients $\eta$ and $\epsilon_i$ are those appearing in
397: the previously mentioned linear superposition, but with $\omega$
398: replaced by $\omega(\p)$. We are following notation of
399: \cite{nos2a} where $\eta=\lambda_+$ and
400: $\epsilon_1^2=(\lambda_++B)(\lambda_+^2+1)^{-1}$ and
401: $\epsilon_1^2=(\lambda_+-\theta)(\lambda_+^2+1)^{-1}$. Note that
402: dependences on ${\bf p}$ are inherited from $\omega(\p)$.
403:
404: Operators $a,b$, by other hand, satisfy the canonical algebra
405: $$
406: [a_\p,a_{{\p}'}^\dag] =(2\pi)^3\delta^3(\p-\p'),
407: \,\,\,\,\,\,\,\,\,
408: [b_\p,b_{{\p}'}^\dag] =(2\pi)^3\delta^3(\p-\p'),
409: $$
410: and are in correspondence with $\tilde{A}_\pm$. Is straightforward
411: to prove that fields constructed in this way satisfy the
412: commutation relations (\ref{defcons}).
413:
414: The Hamiltonian of this theory
415: $$
416: H=\int d^3x{\cal H}(x),
417: $$
418: with the density (\ref{hamrel}) expressed in terms of the non
419: commutative fields (\ref{ncfields}) is a superposition of two
420: anisotropic oscillators (\ref{hamil}) where frequencies are now
421: $\omega(\p)$. That is
422: \begin{eqnarray}
423: \label{nchamsf}
424: H&=&\int
425: \frac{d^3p}{(2\pi)^3}\left[E_+(\p)\left(a_{{\p}}^\dag
426: a_\p+\frac{1}{2}\right)+E_-(\p)\left(b_{{\p}}^\dag
427: b_\p+\frac{1}{2}\right)\right],
428: \end{eqnarray}
429: where energies are
430: \begin{eqnarray}
431: \label{scalener}
432: E_\pm&=&\sqrt{\omega^2(\p)+\frac{1}{4}\bigg[B-\theta \omega^ 2
433: (\p)\bigg]^2}\pm\frac{1}2\left(B+\theta\omega^2(\p)\right ) .
434: \end{eqnarray}
435:
436: This shows that the free non commutative complex scalar fields is a
437: theory with two types of particles with different dispersion
438: relation. This asymmetry can be interpreted as a particle-antiparticle
439: asymmetry and their phenomenological consequences were
440: explored in \cite{nos2,nos5}.
441:
442:
443: A natural question raised by this approach is what happens with other
444: fields as gauge and fermionic fields. Let us discuss gauge fields in next subsection and
445: postpone fermionic fields and its phenomenology to subsequent sections.
446:
447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
448: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
449: \subsection{Non commutative Gauge Fields}
450: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452: We start the discussion with the $U(1)$ gauge field. As in the
453: previous case, the theory is defined by the standard Hamiltonian \cite{gl} (for a previous approach see also\cite{jackiw}).
454:
455: \begin{equation}
456: \label{5}
457: H = \int d^3x \left( \frac{1}{2}{\vec \pi}^2 + \frac{1}{2}{\vec
458: B}^2+A_0\,\vec\nabla\cdot\vec\pi\right),
459: \end{equation}
460: plus a set of deformed Poisson bracket structure
461: \begin{subequations}
462: \label{gaupb}
463: \begin{eqnarray}
464: {[} A_i ({\vec x}),A_j ({\vec y}){]}_{PB} &=&0, \nonumber
465: \\
466: {[} A_i ({\vec x}),\pi_j ({\vec y}){]}_{PB} &=& \delta_{ij}
467: \delta ({\vec x}-{\vec y}), \label{}
468: \\
469: {[}\pi_i ({\vec x}),\pi_j ({\vec y}){]}_{PB}&=&\theta_{ij}\delta
470: ({\vec x}-{\vec y}), \nonumber
471: \end{eqnarray}
472: \end{subequations}
473: where $\theta_{ij}$ is the most general antisymmetric three
474: dimensional matrix $\theta_{ij}=\epsilon_{ijk}\theta_k$. Note that
475: this term modifies the infrared sector of the theory due to the
476: dimension of $\theta_k$.
477:
478: Poisson brackets (\ref{gaupb}) break Lorentz invariance and also the
479: gauge symmetry (GS). In fact, gauge transformation is generated by
480: the Gauss law, which is just the condition that primary constrain
481: preserves in time. The primary constrain is not modified, but its
482: time preservation indeed changes because of extra terms coming from
483: the modified Poisson bracket structure.
484:
485: In order to study just LIV, preserving gauge symmetry, a modified
486: Gauss law must be introduced. A direct calculation shows that
487: \begin{equation}
488: \label{ngl}
489: \chi=\bn\cdot{\boldsymbol \pi} -\bth\cdot\B,
490: \end{equation}
491: with $\bth\cdot\B=\theta_i\epsilon_{ijk}\partial_jA_k$ is the
492: modified generator for the gauge symmetry. That is, if $\alpha(\x)$
493: is an arbitrary and real function, then
494: \begin{subequations}
495: \label{gf}
496: \begin{eqnarray}
497: \delta A_i(\x)&=&{[}A_i(\x),\Delta_\alpha{]}_{PB}
498: \nonumber
499: \\
500: &=&\partial_i\alpha(\x),
501: \\
502: \delta \pi_i(\x)&=&{[}\pi_i(\x),\Delta_\alpha{]}_{PB}
503: \nonumber
504: \\
505: &=&0.
506: \end{eqnarray}
507: \end{subequations}
508: where the gauge transform operator $\Delta_\alpha$ is
509: \begin{equation}
510: \Delta_\alpha=\int d^3y~\alpha({\bf y })\bn[ {\boldsymbol \pi}
511: ({\bf y}) + \bth\times {\bf A}({\bf y})].
512: \end{equation}
513:
514: This last relation allows to write the modified Hamiltonian which
515: includes now this new gauge symmetry generator, namely
516: \begin{equation}
517: \label{modham}
518: H=\int d^3x \left(\frac{1}{2}{\boldsymbol \pi}^2 + \frac{1}{2}
519: \B^2 +A_0 \bn \bigg[{\boldsymbol \pi} +
520: \bth\times {\bf A}\bigg]\right).
521: \end{equation}
522:
523: The Hamiltonian (\ref{modham}) with the Poisson bracket structure
524: (\ref{gaupb}) defines a $U(1)$ gauge field theory which breaks
525: Lorentz symmetry.
526:
527: This model originates modified Maxwell equations
528: \begin{eqnarray}
529: \dot{A}_i&=&\pi_i - \partial_iA_0,
530: \\
531: \dot{\pi}_i&=&({\boldsymbol \pi}\times\bth)_i -(\bn\times\B)_i.
532: \end {eqnarray}
533: First equation is the electric field definition
534: $$
535: -E_i\equiv \dot{A}_i+\partial_i A_0,
536: $$
537: which allow to write the remaining equations in the usual form.
538: Including the modified Gauss law they read
539: \begin{subequations}
540: \label{mxw}
541: \begin{eqnarray}
542: \label{mxw1}
543: \bn\cdot{\bf E}&=&-{\bth}\cdot\B,
544: \\
545: \label{mxw2}
546: \bn\cdot\B&=&0,
547: \\
548: \label{mxw3}
549: \bn\times{\bf E}+\frac{\partial\B}{\partial t}&=&0.
550: \\
551: \label{mxw4}
552: \bn\times\B -\frac{\partial {\bf E}}{\partial t} &=&
553: \bth\times {\bf E}.
554: \end{eqnarray}
555: \end{subequations}
556:
557: In this equation $\theta$ plays the role of a \lq\lq source \rq\rq~
558: term that can be interpreted as a polarization charge and induced
559: current in a medium, in a similar way to the standard electromagnetic
560: theory. In section IV we will discuss the physical
561: implications of this fact.
562:
563: Now we would like to discuss other issue, related with the Lagrangian
564: formulation of our approach. From here, the generalization for other
565: gauge groups will be straightforward \cite{nos3}.
566:
567: The set of equations (\ref{mxw}) can be obtained from a Lagrangian
568: which is constructed from the Hamiltonian (\ref{modham}) in the
569: standard way only if dynamical variables are canonical. Then, we
570: need to find a transformation from variables $\{\pi_i,A_j\}$ to
571: variables $\{\tilde{\pi}_i,\tilde{A}_j\}$ such that the Poison bracket
572: structure (\ref{gaupb}) maps to the canonical one.
573:
574: This procedure is completely analog to the change of variables, in
575: previous section, which takes the non commutative
576: phase space variables to a set of standard rising and lowering
577: operators. In the present case, transformations read
578: \begin{equation}
579: \label{ptop}
580: \tilde{\pi}_i=\pi_i+\frac{1}{2}(\bth\times{\bf A})_i,\,
581: \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
582: \tilde{ A}_i=A_i.
583: \end{equation}
584: and the Lagrangian turn out to be
585: \begin{eqnarray}
586: \label{lagmxw}
587: \label{lag}
588: L &=& \int d^3 x ~(\tilde{\pi}_i {\dot A}_i - H), \nonumber
589: \\
590: &=& \int d^3x \left( {\bf E}^2 - {\B}^2 + \frac{1}{2}A_0
591: {\bth}\cdot{\B} - \frac{1}{2}{\bf A}\cdot{\bth} \times
592: {\bf E}) \right).
593: \end{eqnarray}
594:
595: Using the standard definition for $F_{\mu\nu}$ and ${\tilde F}^{\mu
596: \nu}= \frac{1}{2}\epsilon^{\mu\nu\lambda\rho}F_{\lambda\rho}$, ($\mu,\nu,..=0,1,...3$) one
597: finds that
598: \begin{equation}
599: L = \int \left( -\frac{1}{4} F_{\mu \nu} F^{\mu \nu} + \frac{1}{2}
600: \theta_\mu {\tilde F}^{\mu \nu}A_\nu\right)d^3x. \label{21}
601: \end{equation}
602:
603: This approach can be generalized to others gauge groups. For
604: instance in \cite{nos3}, the $SU(2)$ gauge group was studied and it
605: was shown that the Lagrangian density that generalizes (\ref{21}) is
606: \begin{equation}
607: L = - \frac{1}{2} {\mbox{tr}} \left\{F_{\mu \nu}
608: F^{\mu\nu}\right\} + 2\,\theta^\mu \epsilon_{\mu \nu \rho \sigma}
609: \mbox{tr}\left(A^\nu F^{\rho \sigma} + \frac{2}{3}g A^\nu A^\rho
610: A^\sigma \right), \label{lag2}
611: \end{equation}
612:
613: Finally, let us point out that the Chern-Simons term appearing in
614: this formulation is not a perturbative contribution, it
615: appears indeed at the same footing as $F^2$ in the $g$ expansion.
616:
617: We would like now to call the attention on the fact that here we have
618: considered commutator deformations only in the momenta, although we
619: start the discussion with the complex non commutative scalar field
620: where deformations in fields commutators also appears.
621:
622: It is natural to ask, therefore, what kind of modifications suffers
623: Maxwell theory if we consider a Poisson bracket structure modified as
624: \begin{equation}
625: \label{uvmaxmod}
626: [A_i(x),A_j(x)]=\epsilon_{ijk}\theta_k\delta(x-y),
627: \end{equation}
628: with $\theta_k$ a Lorentz violating vector which plays a role of
629: ultraviolet energy scale. The rest Poisson brackets are canonical.
630:
631: This theory was considered in \cite{glp} and we will not give details
632: on its construction. We would like just to say that, as cases
633: presented before, is possible to restore the gauge symmetry.
634:
635: The main phenomenological feature of this approach is that it
636: presents a birefringence effect with polarization planes shifted by
637: an amount proportional to
638: $$
639: \omega^2\theta\cos \alpha,
640: $$
641: where $\theta$ is the modulus of $\theta_k$ and $\alpha$, the angle
642: between the wave vector $\vec{k}$ and $\theta_k$. This frequency
643: dependent behavior is distinctive and it does not appear when
644: noncommutativity in space-time is considered. It is a pure field
645: theory result.
646:
647: Our last example of non commutative fields is the fermionic case. The
648: next section is devoted to this issue.
649:
650: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
651: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
652: \section{Neutrino physics and noncommutative fermionic fields}
653: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
655:
656: A natural question, considering all the previous discussion, is what
657: happens with the fermions fields under a deformation of Poisson
658: structure.
659:
660: From the phenomenological point of view, our approach is a
661: mechanism that offers an alternative way out to the problem of
662: neutrino oscillations. Let us emphasize that several papers have
663: already dealt with effects of Lorentz \cite{8} and CPT violation
664: \cite{9} in this scenario, however the model we present here has the
665: advantage that depends on a few parameters, as was discussed in
666: previous sections.
667:
668: Since we are mainly interested in the neutrino sector, and
669: particularly in the problem of oscillations, let us briefly review
670: the situation \cite{neutrino}. Neutrino oscillation is a
671: phenomenological model proposed to explain the deficit of solar and
672: atmospheric neutrinos in fluxes measured on earth \cite{more}.
673:
674: The key idea of this mechanism is to assume that neutrino are massive
675: particles which, upon propagation, oscillates between different
676: flavor eigenstates. In its simplest form, the oscillation between two
677: flavors $i,j$ is considered and it is shown that the probability
678: $P_{i\to j}(t)$ for specie $i$ to oscillates to $j$ after a time $t$
679: is
680: \begin{equation}
681: \label{Pij}
682: P_{i\to j}(t)=\sin^2(2\theta_{ij})\sin^2\left(\frac{(E_i - E_j)t}{2}
683: \right),
684: \end{equation}
685: where $\theta_{ij}$ is the mixing angle\footnote{This angle is
686: introduced to take into account the fact that what propagates is a
687: linear superposition of mass eigenstates.} and $E_{i(j)}$ is the
688: energy of $i(j)$ species.
689:
690: It is interesting to note that non vanishing oscillation probability
691: occurs in free space only if there is a non zero $\Delta
692: E_{ij}=E_i-E_j$ and
693: $\theta_{ij}\neq 0$. Then, if neutrino have equal masses or they are
694: zero, oscillation does not come out.
695:
696: The standard scenario assumes that neutrino species have small masses
697: and therefore, the probability for oscillation of two neutrino
698: $\nu_i$, $\nu_j$, in traversing a path length $L$ turns out to be
699: \begin{equation}
700: \label{standprob}
701: P_{\nu_i \to \nu_j}(L) = \sin^2(2\theta_{ij})
702: \sin^2\left( \frac{1.27\Delta m^2_{ij} L}{E} \right),
703: \end{equation}
704: where $\Delta m^2_{ij} =m^2_i -m^2_j$ is taken in (eV)$^2$, the
705: neutrino energy $E$ in MeV and $L$ in meters.
706:
707: Clearly, with three families of neutrino there can be only two
708: independent combinations of squared mass differences, lets say
709: $\Delta m^2_{12},\Delta m^2_{23}$ from which a solution for solar
710: neutrino as well as atmospheric neutrino puzzles is found. The
711: bounds for this masses in this scenario are \cite{neutrino}
712: \begin{equation}
713: \label{masses}
714: \Delta m^2_{12}\leq 10^{-4}\mbox{eV}^2,\,\,\,\,\,\,\,\,\,\,
715: 10^{-3}\leq \mbox{eV}^2 \Delta m^2_{23}\leq 10^{-2}\mbox{eV}^2.
716: \end{equation}
717: From here, the bound for $\Delta m^2_{13}$ is fixed.
718:
719: LSND (Liquid Scintillator Neutrino Detector at Los Alamos)
720: \cite{lsnd} is one of the several experiments that have looked for
721: neutrino oscillations. It has used muon sources from the decay
722: $\pi^+\to \mu^+ +\nu_\mu$. These muons decay through $\mu^+\to
723: e^++\nu_e+\bar{\nu}_\mu$ and, after a $30$ meters long path, the
724: experiment finds the oscillation channel $\bar{\nu}_\mu\to
725: \bar{\nu}_e$ at $20~\mbox{MeV}\leq E_{\nu_\mu}\leq 58.2~\mbox{MeV}
726: $ with probability of $0.26\%$. According to (\ref{standprob}), the
727: mass difference involved in this process should be
728: \begin{equation}
729: \label{mlsnd}
730: \Delta m^2<1~\mbox{eV}.
731: \end{equation}
732:
733: Analysis of the MiniBooNE experiment \cite{mb} will confirm or
734: discard this result. However, if it is true, we must face a puzzle
735: within the context of three families neutrinos. An explanation of
736: this anomaly, compatible with the standard model, may require the
737: existence of sterile neutrinos \cite{steril}.
738:
739: There is a different approach that might solve the puzzle. As was
740: pointed out first by Coleman and Glashow \cite{colgla} and
741: more extensively developed by Kostelecky and collaborators
742: \cite{kostel}, a departure from Lorentz and/or CPT symmetry in the
743: neutrino sector, gives raise to oscillations as neutrino propagates
744: in free space, even for massless neutrinos.
745:
746: From (\ref{Pij}) is clear that the oscillation probability does not
747: vanish if the mixing angle is not zero and the energy of specie $1$
748: is different from specie $2$. This condition satisfies if neutrino
749: masses are different in a theory respecting Lorentz and CPT symmetry
750: (the standard case), but if one of these symmetries is broken, a non
751: zero probability can be obtained.
752:
753: Note that it is enough to have different dispersion relations for
754: different neutrino species in order to fulfill previous requirements.
755: But this is indeed the case for noncommutative fields approach if we
756: choose non standard commutators among different particle species. In
757: fact, noncommutative scalar field is the first example of this kind,
758: since there field $\phi$ and its conjugate $\phi^\dag$ have a non
759: standard Poisson bracket.
760:
761: The final result is that, due to non standard anticommutators for
762: massless fermionic fields, dispersion relations are species
763: dependent and a non vanishing oscillation probability is obtained. In
764: what follows we will discuss the technicalities of this approach
765: \cite{last}.
766:
767: In the chiral basis, which is more convenient for
768: our analysis, the Hamiltonian density has the form
769: \begin{equation}
770: {\cal H} = i\left( \psi_L^{i\dagger} {\vec \sigma} \cdot~ \vec{\nabla}
771: \psi^{i}_L -
772: \psi_R^{i\dagger} {\vec \sigma} \cdot~ \vec{\nabla} \psi^{i}_R
773: \right), \label{dh1}
774: \end{equation}
775: where the superscript $i=\{1,2\}$ runs over the flavor quantum number
776: (sum over repeated indexes).
777:
778: The non-commutative theory is obtained by
779: deforming the canonical anti-commutation relations while maintaining
780: the form of the Hamiltonian density \eqref{dh1}. We postulate the
781: deformed equal-time anti-commutation relations to have the form (with all others
782: vanishing)
783: \begin{eqnarray}
784: \{ \psi^i_L({\bf x}),\psi^{j\dagger}_L ({\bf y})\} &=& A^{ij}
785: ~\delta^{(3)} ({\bf x} - {\bf y}),
786: \label{ant1}
787: \\
788: \{ \psi^i_R ({\bf x}),\psi^{j\dagger}_R({\bf y})\} &=& B^{ij}
789: ~\delta^{(3)} ({\bf x} - {\bf y}),
790: \label{ant2}
791: \end{eqnarray}
792: where $A^{ij}$ and $B^{ij}$ are $2\times 2$ matrices with constant,
793: complex elements in general, but if we want to maintain rotational
794: invariance, they can be chosen to have the forms
795: \begin{equation}
796: \label{abmatri}
797: A = \left(\begin{array}{cc} 1 & \alpha \\
798: \alpha^\ast & 1 \end{array}\right),~~~~~~~~~
799: B = \left(\begin{array}{cc} 1 & \beta \\
800: \beta^\ast & 1 \end{array}\right),
801: \end{equation}
802: so that the complex parameters $\alpha,\beta$ can be thought of as the
803: parameters of deformation. Clearly, these deformed anti-commutation
804: relations reduce to the conventional ones when the parameters of
805: deformation vanish.
806:
807: Equation of motions in the momentum space read
808: \begin{eqnarray}
809: E\psi^i_L &=& -A^{ij} \left( {\vec \sigma}\cdot {\vec p}\
810: \psi^j_L\right),
811: \label{mom1}
812: \\
813: E\psi^i_R&=& B^{ij}\left( {\vec \sigma}\cdot {\vec p}\
814: \psi^j_R\right). \label{mom2}
815: \end{eqnarray}
816:
817: The energy spectrum can be find independently for \eqref{mom1} and
818: \eqref{mom2}. If we take the first one, it is straightforward to find
819: a diagonalization matrix for $A$
820: \begin{equation}
821: D= \frac{1}{\sqrt{2}} \left(\begin{array}{cc} \frac{|\alpha|}{\alpha}
822: & 1 \\
823: -\frac{|\alpha|}{\alpha} &1\end{array}\right),\
824: \end{equation}
825: and then
826: \begin{eqnarray}
827: E^{1}_\pm &=& \pm \left( 1 +|\alpha| \right) |{\vec p}|,
828: \nonumber
829: \\
830: E^{2}_\pm &=& \pm \left( 1 -|\alpha| \right) |{\vec
831: p}|,
832: \label{disp2}
833: \end{eqnarray}
834: are the eigenvalues in \eqref{mom1}.
835:
836: A similar analysis for \eqref{mom2} gives
837: \begin{eqnarray}
838: E^{1}_\pm &=& \pm \left( 1 +|\beta| \right) |{\vec p}|,
839: \nonumber
840: \\
841: E^{2}_\pm &=& \pm \left( 1 -|\beta| \right)
842: |{\vec p}|.
843: \label{disp22}
844: \end{eqnarray}
845:
846: Since $\psi^i_{L}$ does not diagonalize the Hamiltonian --eigenvectors
847: are $D\psi^i_L$, which are a linear combination of $\psi^1_L$ and
848: $\psi^2_L$-- the time evolution of this field gives rise to a linear
849: combination of species 1 and 2, namely
850: a neutrino initially in the state $\psi_{L}^{1}$ would evolve in
851: time as
852: \begin{eqnarray}
853: \psi^1_L (t) &=& \cos \theta_{12} ~{\tilde \psi}^1_L (t) -
854: \sin \theta_{12} ~{\tilde \psi}^2_L (t) \nonumber
855: \\
856: &=& \biggl[ \left( \cos^2 \theta_{12}~e^{-iE_+^{1} t} +
857: \sin ^2\theta_{12}~ e^{-iE_+^{2} t}\right) \psi^1_L(0)
858: \nonumber
859: \\
860: &+& \frac{1}{2} \sin2 \theta_{12} \left( e^{-iE_+^{1} t} -
861: e^{-iE_+^{2}
862: t}\right) \psi_L^2 (0)\biggr]e^{i{\vec p}\cdot{\vec x}}. \nonumber
863: \end{eqnarray}
864: Therefore, after a path of length $L$, the probability of finding the
865: state $\psi^2_L$ in the beam is given by
866: \begin{equation}
867: P_{\nu_1 \rightarrow \nu_2}= \sin^2 \left(2\theta_{12} \right) ~
868: \sin^2
869: \left( \vert \alpha \vert~E ~L\right),\label{prob3}
870: \end{equation}
871:
872:
873: \noindent where we have used the fact that for $|\alpha|\ll 1,
874: E\approx |\vec{p}|$.
875:
876: Same arguments demonstrate that, for antineutrino propagation, the
877: probability for oscillations is
878: \begin{equation}
879: P_{{{\bar \nu}_1 \rightarrow {\bar \nu}_2}}= \sin^2
880: \left(2\theta_{12}
881: \right) ~
882: \sin^2 \left( \vert \beta \vert~E~ L\right). \label{prob4}
883: \end{equation}
884: The important thing to note here is that
885: \begin{equation}
886: P_{\nu_1 \rightarrow \nu_2} \neq P_{{\bar \nu}_1 \rightarrow
887: {\bar \nu}_2 }, \label{000}
888: \end{equation}
889: which is a consequence of $CPT$ and Lorentz invariance violation.
890:
891: Two comments are in order here. First, noncommutative fermionic fields
892: have a mixing angle equals $\pi/4$, and then is consistent with the so
893: called Large Mixing Angle (LMA) scenario. However, we have introduced
894: in all the discussion a $\theta_{ij}$, in order to mimic that
895: standard case, but it does not appear in a pure noncommutative
896: fermionic theory. Secondly, our description has a linear dependence
897: on the energy, what could be disturbing, but is a natural consequence of the
898: fact that we have incorporated two dimensionless parameters on the dispersion relation.
899:
900: From previous results we can compute bounds for deformation
901: parameters as follow. Assuming that flavor oscillations involve only
902: pairs of neutrinos, when dealing with three families we must
903: generalize the parameter $\alpha$ to $\alpha_{ij}$ (and the same for
904: antineutrinos). Therefore, flavor oscillation probabilities are
905: \begin{eqnarray}
906: P_{\nu_{i}\rightarrow \nu_{j}}&=& \sin^{2}
907: \bigg(2\theta_{ij}\bigg) \sin^{2}
908: \bigg(|\alpha_{ij}| EL\bigg),\label{probij}
909: \\
910: P_{\bar{\nu}_{i}\rightarrow \bar{\nu}_{j}} &=& \sin^{2}
911: \bigg(2\theta_{ij}\bigg) \sin^{2}
912: \bigg(|\beta_{ij}| EL\bigg).\label{probijbar}
913: \end{eqnarray}
914:
915: From the solar neutrino experiments, oscillations of the
916: flavors $1\rightarrow 2$ are involved, while the channel
917: $2\rightarrow 3$ is related with atmospheric neutrino oscillations.
918: Therefore, from \eqref{probij} next bounds on $\alpha$ parameter is
919: found
920: \begin{eqnarray}
921: |\alpha_{12}| &\leq& 10^{-17}.\label{result1}
922: \\
923: |\alpha_{23}| &\leq& 10^{-22}.\label{result2}
924: \end{eqnarray}
925:
926: LSND, by other hand, due to the fact that involves antineutrinos,
927: gives a bound for $\beta$ parameter
928: \begin{equation}
929: |\beta_{12}|\leq 10^{-16}.\label{result3}
930: \end{equation}
931: It is clear that within this scenario, all the experimental results
932: can be naturally explained.
933:
934: Finally we would like to call the attention on the fact that here
935: neutrino are massless particles and, looking at dispersion relations,
936: we realize that the origin of the energy difference between species
937: can be understood also as a difference in the propagation velocities.
938: Since particles are massless, we would say that particles propagates
939: with different speed of light.
940:
941:
942:
943: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
945: \section{Phenomenological consequences of noncommutative fields in
946: the early universe.}
947:
948:
949: This section is devoted to analyze possible phenomenological
950: consequences derived from models based on noncommutative fields.
951:
952: A consequences of this model which is common for scalars and massless
953: fermions is the asymmetry of the dispersion relation for
954: particles and antiparticles.
955:
956: For the scalar field, situation is rather simple and the asymmetry
957: can be checked at the level of quantum Hamiltonian
958: \eqref{nchamsf} or at the level dispersion relation
959: \eqref{scalener}. This theory has two scales, one infrared $B$ and other ultraviolet
960: $\theta$. At momentum ${\bf p}$ such that
961: $B<<\omega({\bf p})<<\theta^{-1}$ energy \eqref{scalener} satisfy
962: $ E_+\sim E_-\sim \omega({\bf p})$ and we are in the Lorentz invariant
963: region.
964:
965: Consider a system with this two types of particles in thermodynamical
966: equilibrium at temperature $T$. The density $n/V$ of each specie
967: contained in a volume $V$ (with zero chemical potential) is
968: \cite{nos5}
969: \begin{eqnarray}
970: \label{nmas}
971: n_+=4\pi\int_0^\infty\frac{p^2dp}{e^{{E_+}/T}-1},
972: \\
973: \label{nmenos}
974: n_-=4\pi\int_0^\infty\frac{p^2dp}{e^{{E_-}/T}-1}.
975: \end{eqnarray}
976:
977: For a temperature $T$ such that $\theta T<<B/T<<m/T<<1$, there is a
978: tiny asymmetry in the dispersion relation due to the infrared scale
979: and then, a tiny asymmetry in the content of baryonic
980: matter-antimatter content in the volume $V$. In fact
981: \begin{equation}
982: \label{barasymm}
983: \frac{n_+ - n_-}{n_-}\sim \frac{B}{T},
984: \end{equation}
985: as it is expected if CPT violating effects are tiny.
986:
987: This example shows that a baryon asymmetry can be generated without
988: departure from thermal equilibrium and it suggests a critical
989: reevaluation of the third criterion of Sakharov for baryogenesis
990: \cite{sakha}.
991:
992: Since this effect is related only with the asymmetry on the
993: dispersion relation for particles and antiparticles, one could wonder
994: what happens with fermions.
995:
996: Situation is similar to the previous one. There is an asymmetry
997: due to the different dispersion relations for neutrinos and
998: antineutrinos\footnote{In this case, however, the distinction
999: between particles and antiparticles is more subtle, but the
1000: quantum Hamiltonian of the theory can also be written as two types
1001: of particles with different frequencies\cite{lastt}} and therefore
1002: a the ratio of the neutrino density to the antineutrino density,
1003: in equilibrium at certain temperature is different from one.
1004:
1005: Calculation in this case is more involved due to the presence of
1006: different flavors and to the oscillations between them. For two
1007: species, however, this problem is formally equivalent to a quantum
1008: mechanical two level system with a Hamiltonian which is responsible
1009: for the for inducing transition between levels.
1010:
1011: The crucial step is to identify this Hamiltonian responsible for the
1012: transitions, lets say $\nu_e\to\nu_\mu$. According to Stodolsky
1013: \cite{sto} and others \cite{others} this Hamiltonian is
1014: \begin{equation}
1015: \label{stodol}
1016: H=\vec{\sigma}\cdot\vec{V},
1017: \end{equation}
1018: with $|\vec{V}| =|E_1-E_2|$.
1019:
1020: Following the results of previous section, we find that the
1021: Hamiltonian for a system violating CPT and Lorentz symmetries can be
1022: written in this two dimensional space as
1023: \begin{eqnarray}
1024: \label{hmas}
1025: H_+=2|\alpha|\vec{\sigma}\cdot\vec{p},
1026: \\
1027: H_-=2|\beta|\vec{\sigma}\cdot\vec{p}.
1028: \end{eqnarray}
1029:
1030: Then, if there is a CPT and Lorentz invariance violation,
1031: oscillations neutrino-neutrino and antineutrino-antineutrino take
1032: place with different probabilities leading to a neutrino asymmetry.
1033:
1034: With present data on neutrino experiments, this asymmetry could be
1035: evaluated, however, it seems to premature to carry this analysis at
1036: this stage, when still LSND experiment requires a confirmation.
1037:
1038: We will finish this section with some comments on the
1039: cosmological implications of noncommutative $U(1)$ gauge field. As
1040: we mentioned in section II.A, modified Maxwell equations
1041: \eqref{mxw} mimics standard Maxwell equations with sources, but there
1042: is an important difference:electrostatic and magnetostatic appears
1043: mixed and then the presence of polarization implies a magnetization
1044: and viceversa.
1045:
1046: Remarkably, is this structure which offers an alternative to the
1047: dynamo mechanism to generate the so called {\it primordial magnetic
1048: field}. In fact, it was shown in \cite{nos3} that modified equations
1049: admits a solution with the shape
1050: \begin{eqnarray}
1051: {\bf B}&=&\B^{(0)} +\B^{(2)}+\B^{(4)}+\ldots\B^{(2n)}+\ldots,
1052: \\
1053: {\bf E}&=&{\bf E}^{(0)} +{\bf
1054: E}^{(1)} + {\bf E}^{(3)}+\ldots\B^{(2n+1)}+\ldots,
1055: \end{eqnarray}
1056: where superindices stand for the exponent in $\theta$ in the series
1057: expansion. In accordance with experiments, $E$ is always a lower
1058: order of magnitude that the magnetic field.
1059:
1060: Similarly to a ferromagnetic media, the system might evolve to a
1061: stable state with permanent magnetic and/or electric field because
1062: the previous expansion not necessarily converges.
1063:
1064: Previous mechanism, therefore, is a possible candidate for an
1065: alternative explanation to the dynamo mechanism of the primordial
1066: magnetic field observed in universe.
1067:
1068: \section{Conclusions and outlook}
1069:
1070: We have explored phenomenological consequences of Lorentz and CPT
1071: symmetries violation through the so called non commutative field
1072: theory.
1073:
1074: For complex scalar field we have shown how quantum noncommutative
1075: field theory can be constructed as a superposition of anisotropic
1076: harmonic oscillators. Quantum fermionic fields can be constructed in
1077: a similar way and will be reported in a forthcoming paper
1078: \cite{lastt}.
1079:
1080: The common feature of these approaches is their particle-antiparticle
1081: asymmetry manifested in dispersion relations. We have explored
1082: the possibility of using this as a possible mechanism to generate
1083: baryon-antibaryon asymmetry as well as neutrino anti-neutrino. In the
1084: firs case, we have shown how that can be compatible with thermal
1085: equilibrium scenario.
1086:
1087: For neutrinos, this asymmetry is also the mechanism that allow
1088: flavor oscillations and then, two apparently disconnected problems
1089: could be explained by the same mechanism. In other words, if CPT and
1090: Lorentz are violated, would be possible, in principle, to calculate
1091: the excess of neutrinos respect antineutrinos in the universe.
1092:
1093: In the gauge field sector, noncommutative fields offers an
1094: alternative process to the dynamo mechanism in order to explain
1095: interstellar magnetic fields.
1096:
1097: In conclusion, a tiny Lorentz symmetry violation open doors to
1098: explain observations which can not be accommodate in the conventional
1099: physics. Noncommutative fields, by other hand, is a description that
1100: incorporate this properties in economical way, that is, it depends on
1101: a small number of parameters and, just at level of free theory
1102: exhibits features that might explain these phenomena.
1103:
1104:
1105: \acknowledgments
1106:
1107: This work was supported in part by US DOE Grant number
1108: DE-FG-02-91ER40685, FONDECYT-Chile grants 1050114, 1060079 and
1109: 2105016 and a FULBRIGHT grant (JL).
1110:
1111:
1112:
1113:
1114: \begin{thebibliography}{99}
1115: \bibitem{effective} For a review see {\it e.g.} J. Kogut and K. Wilson, {\it Phys. Rept.} {\bf 12}, 75 (1974).
1116: \bibitem{yennie} See for example, C. Itzikson and J. B. Zuber, {\it Quantum Field Theory}, (1980)
1117:
1118: \bibitem{nos2a}J.M. Carmona, J.L. Cortes, J. Gamboa
1119: and F. Mendez; {\it JHEP} {\bf 0303}, 058 (2003).
1120:
1121:
1122: \bibitem{nos2}J.M. Carmona, J.L. Cortes, J. Gamboa
1123: and F. Mendez {\it Phys. Lett. } {\bf B565}, 222 (2003).
1124:
1125: \bibitem{nos5}J. M. Carmona, J. L. Cortes, A. Das, J.
1126: Gamboa and F. Mendez, {\it Mod. Phys. Lett. }{\bf A21}, 883 (2006).
1127:
1128:
1129:
1130: \bibitem{gl} J. Gamboa and J. Lopez-Sarrion, {\it Phys.Rev.}{\bf D71}, 067702 (2005).
1131: \bibitem{jackiw} S. Carroll, G. B. Field and R. Jackiw, {\it Phys. Rev.} {\bf D41}, 1231 (1990); A. A. Andrianov, P. Giacconi and
1132: R. Soldati, {\bf JHEP} {\bf 0202}, 030 (2002).
1133:
1134:
1135:
1136: \bibitem{nos3} H. Falomir, J. Gamboa, J. Lopez-Sarrion, F. Mendez and A. J. da Silva, {\it Phys. Lett.} {\bf B632}, 740 (2006)
1137: and ibid {\it Phys. Rev.} {\bf D74}, 047701 (2006).
1138: \bibitem{glp}J. Gamboa, J. Lopez-Sarrion and A. P.
1139: Polychronakos {\it Phys.Lett.} {\bf B634}, 471 (2006).
1140:
1141: \bibitem{8}M. Gasperini, {\it Phys. Rev.} {\bf D38},
1142: 2635; S. L. Glashow, A. Halprin, P. I. Krastev, C. N. Leung and J.
1143: Pantaleone, {\it Phys. Rev.} {\bf D56}, 2433 (1997); R. Foot, C. N.
1144: Leung and O. Yasuda, {\it Phys. Lett.} {\bf B443}, 185 (1998).
1145:
1146: \bibitem{9} S. Coleman and S. L. Glashow, {\it Phys.
1147: Lett.} {\bf B405}, 249 (1997).
1148:
1149: \bibitem{neutrino}For a
1150: recent review on neutrino data and experiments, see
1151: for example A. Strumia and F. Vissani , {\it hep-ph/0606054} and
1152: references therein.
1153:
1154: \bibitem{more} G. L. Fogli, E. Lisi, A. Marrone and G. Scioscia, {\it
1155: Phys. Rev.} {\bf D60}, 053006 (1999); B. Pontecorvo, {\it J. Exp.
1156: Theor. Phys. (USSR)}, {\bf 34}, 247 (1958); L. Wolfenstein, {\it
1157: Phys. Rev.} {\bf D17 }, 2369 (1978); S. P. Mikheyev and A. Yu.
1158: Smirnov, {\it Sov. J. Nucl. Phys.} {\bf 42}, 913 (1985).
1159:
1160: \bibitem{lsnd} LSND collaboration, C. Athanassopoulos et al., {\it
1161: Phys. Rev. Lett.} {\bf 81}, 1774 (2003); LSND collaboration, A.
1162: Aguilar et al. {\it Phys.
1163: Rev.} { \bf D64}, 112007 (2001).
1164:
1165: \bibitem{mb} MiniBooNE collaboration homepage:www-boone.fnal.gov
1166:
1167: \bibitem{steril}See S. L. Glashow, {\it Phys. Lett.} {\bf B256}, 255
1168: (1991); J. W. F. Valle
1169: and D. Tommasini and J. T. Peltoniemi, {\it Phys. Lett.} {\bf B298},
1170: 383 (1993).
1171:
1172: \bibitem{colgla}S. Coleman and S. L. Glashow, {\it Phys. Rev. } {\bf
1173: D59}, 116008 (1999).
1174:
1175: \bibitem{kostel}V. A. Kostelecky and M. Mewes in {\it Phys. Rev.}
1176: {\bf
1177: D69}, 016005 (2004), T. Katori, A. V. Kostelecky and R. Tayloe,
1178: hep-ph/0606154.
1179:
1180: \bibitem{last} P. Arias, Ashok Das, J. Gamboa, J. Lopez-Sarrion and
1181: F. Mendez, {\it e-Print Archive: hep-ph/0608007 }
1182:
1183:
1184:
1185: \bibitem{sakha}A. D. Sakharov, {\it JETP Lett} {\bf 6}, 24 (1967).
1186:
1187:
1188: \bibitem{lastt}P. Arias, Ashok Das, J. Gamboa, J.
1189: Lopez-Sarrion and F. Mendez, {\it in preparation}
1190:
1191: \bibitem{sto} L. Stodolsky, {\it Phys. Rev.} {\bf D36}, 2273 (1987);
1192: L. Stodolsky,
1193: preprint MPI-PAE/PTh 33/88.
1194:
1195: \bibitem{others}V. A. Kostelecky and M. Mewes in {\it Phys. Rev.}
1196: {\bf
1197: D69}, 016005 (2004), T. Katori, A. V. Kostelecky and R. Tayloe,
1198: hep-ph/0606154.
1199:
1200:
1201:
1202:
1203: \end{thebibliography}
1204:
1205:
1206:
1207:
1208: \end{document}
1209:
1210:
1211:
1212:
1213:
1214:
1215: \begin{thebibliography}{99}
1216: \bibitem{...}
1217: ....
1218:
1219: \end{thebibliography}
1220:
1221: \end{document}
1222:
1223: