hep-ph0309025/nu.tex
1: \documentclass[twocolumn,amsmath,amssymb,prd,showpacs]{revtex4}
2: \usepackage{psfig}
3: 
4: \def\al{\alpha}
5: \def\be{\beta}
6: \def\ga{\gamma}
7: \def\de{\delta}
8: \def\ep{\epsilon}
9: \def\ve{\varepsilon}
10: \def\ze{\zeta}
11: \def\et{\eta}
12: \def\th{\theta}
13: \def\vt{\vartheta}
14: \def\io{\iota}
15: \def\ka{\kappa}
16: \def\la{\lambda}
17: \def\vpi{\varpi}
18: \def\rh{\rho}
19: \def\vr{\varrho}
20: \def\si{\sigma}
21: \def\vs{\varsigma}
22: \def\ta{\tau}
23: \def\up{\upsilon}
24: \def\ph{\phi}
25: \def\vp{\varphi}
26: \def\ch{\chi}
27: \def\ps{\psi}
28: \def\om{\omega}
29: \def\Ga{\Gamma}
30: \def\De{\Delta}
31: \def\Th{\Theta}
32: \def\La{\Lambda}
33: \def\Si{\Sigma}
34: \def\Up{\Upsilon}
35: \def\Ph{\Phi}
36: \def\Ps{\Psi}
37: \def\Om{\Omega}
38: \def\mn{{\mu\nu}}
39: \def\cA{{\cal A}}
40: \def\cl{{\cal L}}
41: \def\cE{{\cal E}}
42: \def\cI{{\cal I}}
43: \def\cN{{\cal N}}
44: \def\cS{{\cal S}}
45: \def\cO{{\cal O}}
46: \def\fr#1#2{{{#1} \over {#2}}}
47: \def\frac#1#2{\textstyle{{{#1} \over {#2}}}}
48: \def\pt#1{\phantom{#1}}
49: \def\prt{\partial}
50: \def\vev#1{\langle {#1}\rangle}
51: \def\ket#1{|{#1}\rangle}
52: \def\bra#1{\langle{#1}|}
53: \def\amp#1#2{\langle {#1}|{#2} \rangle}
54: \def\half{{\textstyle{1\over 2}}}
55: \def\lsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
56:     \raise1pt\hbox{$<$}}}
57: \def\gsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
58:     \raise1pt\hbox{$>$}}}
59: \def\Re{\hbox{Re}\,}
60: \def\Im{\hbox{Im}\,}
61: \def\Arg{\hbox{Arg}\,}
62: \def\etal {{\it et al.}}
63: \newcommand{\beq}{\begin{equation}}
64: \newcommand{\eeq}{\end{equation}}
65: \newcommand{\bea}{\begin{eqnarray}}
66: \newcommand{\eea}{\end{eqnarray}}
67: \newcommand{\bse}{\begin{subequations}}
68: \newcommand{\ese}{\end{subequations}}
69: \newcommand{\rf}[1]{(\ref{#1})}
70: 
71: \def\to{\rightarrow}
72: \def\from{\leftarrow}
73: \def\mix{\leftrightarrow}
74: \def\tofrom{\rightleftarrows}
75: \def\nub{\bar\nu}
76: \def\vp{\vec p}
77: \def\cmat{{\cal C}}
78: \def\cH{{\cal H}}
79: \def\heff{h_{\rm eff}}
80: \def\Ueff{U_{\rm eff}}
81: \def\tu{\widetilde U}
82: \def\bu{\overline U}
83: \def\ml{m_l}
84: \def\mt{\widetilde m^2}
85: \def\gt{\tilde g} 
86: \def\Ht{\widetilde H} 
87: \def\AA{{A'}}
88: \def\BB{{B'}}
89: \def\CC{{C'}}
90: \def\aa{{a'}}
91: \def\bb{{b'}}
92: \def\cc{{c'}}
93: \def\aaa{{\hat a'}}
94: \def\bbb{{\hat b'}}
95: \def\ring#1{{\mathaccent'27 #1}}
96: \def\cri{\ring{c}}
97: \def\ari{\ring{a}}
98: \def\mri{\ring{m}}
99: \def\mem{\mri}
100: \def\aem{\ari}
101: \def\amt{\ari'}
102: \def\cem{\cri}
103: \def\cmt{\cri}
104: \def\Gc{\check{g}}
105: \def\a3em{\check{a}}
106: \def\cee{\cri}
107: \def\Dtm{\De m^2_{\Th}}
108: \def\Dtmz{\De m^2_{0^\circ}}
109: 
110: \begin{document}
111: \title{Lorentz and CPT violation in neutrinos}
112: \author{V.\ Alan Kosteleck\'y and Matthew Mewes}
113: \affiliation{Physics Department, Indiana University, 
114:          Bloomington, IN 47405, U.S.A.}
115: \date{IUHET 459, August 2003} 
116: 
117: \begin{abstract}
118: A general formalism is presented for violations 
119: of Lorentz and CPT symmetry in the neutrino sector.
120: The effective hamiltonian for neutrino propagation 
121: in the presence of Lorentz and CPT violation is derived,
122: and its properties are studied.
123: Possible definitive signals in existing and future 
124: neutrino-oscillation experiments are discussed.
125: Among the predictions are direction-dependent effects,
126: including neutrino-antineutrino mixing,
127: sidereal and annual variations, 
128: and compass asymmetries.
129: Other consequences of Lorentz and CPT violation
130: involve unconventional energy dependences
131: in oscillation lengths and mixing angles.
132: A variety of simple models both with and without neutrino masses 
133: are developed to illustrate key physical effects.
134: The attainable sensitivities 
135: to coefficients for Lorentz violation in the Standard-Model Extension
136: are estimated for various types of experiments.
137: Many experiments have potential sensitivity
138: to Planck-suppressed effects,
139: comparable to the best tests in other sectors.
140: The lack of existing experimental constraints,
141: the wide range of available coefficient space,
142: and the variety of novel effects
143: imply that some or perhaps even all of the existing data 
144: on neutrino oscillations might be due to Lorentz and CPT violation.
145: 
146: \end{abstract}
147: 
148: \pacs{11.30.Cp, 14.60.Pq}
149: 
150: \maketitle
151: 
152: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153: 
154: \section{Introduction}
155: 
156: The minimal Standard Model (SM) of particle physics
157: offers a successful description of most processes in Nature
158: but leaves unresolved several experimental and theoretical issues.
159: On the experimental front,
160: observations of neutrino oscillations
161: have accumulated convincing evidence
162: that the description of physical properties of neutrinos
163: requires modification of the neutrino sector 
164: in the minimal SM.
165: Most experimental results to date can be described theoretically
166: by adding neutrino masses to the minimal SM,
167: but a complete understanding of the existing data
168: awaits further experimentation.
169: On the theoretical front,
170: the SM is expected to be the low-energy limit
171: of a more fundamental theory
172: that unifies quantum physics and gravity at the Planck scale,
173: $m_P \simeq 10^{19}$ GeV.
174: Direct measurements at this energy scale are impractical,
175: but suppressed low-energy signatures
176: from the anticipated new physics
177: might be detectable in sensitive existing experiments.
178: 
179: In this work,
180: we address both these topics by studying effects on the neutrino sector
181: of relativity violations, 
182: a promising class of Planck-scale signals.
183: These violations might arise through the breaking of Lorentz symmetry
184: and perhaps also the breaking of CPT symmetry
185: \cite{cpt01}.
186: Since the SM is known to provide a successful description
187: of most physics at low energies compared to the Planck scale,
188: any such signals must appear at low energies 
189: in the form of an effective quantum field theory
190: containing the SM.
191: The general effective quantum field theory constructed from the SM
192: and allowing arbitrary coordinate-independent Lorentz violation
193: is called the Standard-Model Extension (SME)
194: \cite{ck}.
195: It provides a link to the Planck scale 
196: through operators of nonrenormalizable dimension
197: \cite{kpo,kle}.
198: Since CPT violation implies Lorentz violation
199: \cite{owg},
200: this theory also allows for general CPT breaking.
201: The SME therefore provides a realistic theoretical basis
202: for studies of Lorentz violation,
203: with or without CPT breaking.
204: 
205: The lagrangian of the SME consists of the usual SM lagrangian
206: supplemented by all possible terms that can be constructed
207: with SM fields and that introduce violations of Lorentz symmetry.
208: The additional terms
209: have the form of Lorentz-violating operators
210: coupled to coefficients with Lorentz indices,
211: and they could arise in a variety of ways.
212: One generic and elegant mechanism is spontaneous Lorentz violation,
213: proposed first in string theory and field theories with gravity
214: \cite{ks}
215: and then generalized to include CPT violation
216: \cite{kp}.
217: Another popular framework for Lorentz violation is
218: noncommutative field theory,
219: in which realistic models form a subset of the SME
220: involving operators of nonrenormalizable dimension
221: \cite{ncqed}.
222: Other proposed sources of Lorentz and CPT violation include
223: various non-string approaches to quantum gravity
224: \cite{qg},
225: random dynamics
226: \cite{fn}
227: and multiverses
228: \cite{bj}.
229: Planck-scale sensitivity
230: to the coefficients for Lorentz violation in the SME
231: has been achieved in various experiments,
232: including ones with
233: mesons \cite{hadronexpt,kpo,hadronth},
234: baryons \cite{ccexpt,spaceexpt,cane},
235: electrons \cite{eexpt,eexpt2},
236: photons \cite{photonexpt,photonth,cavexpt,km},
237: and muons \cite{muons}.
238: However,
239: no experiments to date have measured 
240: neutrino-sector coefficients for Lorentz violation.
241: 
242: Here,
243: we explore neutrino behavior 
244: in the presence of Lorentz and CPT violation 
245: using the SME framework.
246: The original proposal for Lorentz and CPT violation in neutrinos
247: \cite{ck}
248: has since been followed by several theoretical investigations
249: within the context of the SME
250: \cite{fc1,fc2,fc3,fc4,fc5,nu},
251: most of which have chosen to restrict attention 
252: to a small number of coefficients.
253: A comprehensive theoretical study of Lorentz and CPT violation
254: in neutrinos has been lacking.
255: The present work partially fills this gap by applying 
256: the ideas of the SME to a general neutrino sector
257: with all possible couplings of left- and right-handed neutrinos
258: and with sterile neutrinos.
259: We concentrate mostly on Lorentz-violating operators 
260: of renormalizable dimension,
261: which dominate the low-energy physics in typical theories,
262: but some generic consequences of Lorentz-violating operators 
263: of nonrenormalizable dimension are also considered
264: \cite{kpo,kle,bef}. 
265: The effective hamiltonian describing free neutrino propagation 
266: is obtained,
267: and its implications are studied.
268: The formalism presented in this work
269: thereby provides a general theoretical basis
270: for future studies of Lorentz and CPT violation in neutrinos.
271: We also illustrate various key physical ideas 
272: of Lorentz and CPT violation through simple models, 
273: and we discuss experimental signals.
274: Our primary focus here is on oscillation data
275: \cite{pdg},
276: but the formalism is applicable 
277: also to other types of experiments including 
278: direct mass searches
279: \cite{katrin},
280: neutrinoless double-beta decay
281: \cite{bbdecay},
282: and supernova neutrinos
283: \cite{sn1987a}.
284: 
285: Several features of Lorentz and CPT violation that we uncover 
286: are common to other sectors of the SME,
287: including unconventional energy dependence
288: and dependence on the direction of propagation.
289: We also find that Lorentz-violating neutrino-antineutrino mixing
290: with lepton-number violation 
291: naturally arises from Majorana-like couplings.
292: These features lead to several unique signals 
293: for Lorentz and CPT violation.
294: For example,
295: the direction dependence potentially generates 
296: sidereal variations in terrestrial experiments as the Earth rotates, 
297: annual variations in solar-neutrino properties,
298: and intrinsic differences in neutrino flux 
299: from different points on the compass or different angular heights
300: at the location of the detector.
301: The unconventional energy dependence produces a variety
302: of interesting potential signals, 
303: including resonances in the vacuum
304: \cite{fc2,nu} 
305: as well as the usual MSW resonances in matter
306: \cite{msw}.
307: 
308: Experiments producing evidence for neutrino oscillations 
309: to date include 
310: atmospheric-neutrino experiments
311: \cite{sk},
312: solar-neutrino experiments
313: \cite{homestake,gallex,gno,sage,sksol,sno},
314: reactor experiments
315: \cite{kamland},
316: and accelerator-based experiments
317: \cite{lsnd,k2k}.
318: Most current data are consistent 
319: with the introduction of three massive-neutrino states,
320: usually attributed to GUT-scale physics.
321: However, 
322: as we demonstrate in this work,
323: the possibility remains that the observed neutrino oscillations 
324: may be due at least in part and conceivably even entirely
325: to Lorentz and CPT violation from the Planck scale.
326: In any event,
327: experiments designed to test neutrino mass 
328: are also well suited for tests of Lorentz and CPT invariance,
329: and they have the potential to produce the first measurements 
330: of violations of these fundamental symmetries,
331: signaling possible Planck-scale physics.
332: 
333: The organization of this paper is as follows.
334: Section \ref{theory} presents the basic theory and definitions,
335: obtaining the effective hamiltonian for neutrino propagation 
336: and discussing its properties.
337: Issues of experimental sensitivities 
338: and possible constraints from experiments in other sectors 
339: are considered in Section \ref{sens}.
340: Certain key features of neutrino behavior in the presence of 
341: Lorentz and CPT violation are illustrated 
342: in the sample models of Section \ref{models}.
343: Some remarks about both generic and experiment-specific predictions
344: are provided in Section \ref{disc}.
345: Throughout,
346: we follow the notation and conventions 
347: of Refs.\ \cite{ck,kle}.
348: 
349: 
350: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
351: 
352: \section{Theory}\label{theory}
353: 
354: \subsection{Basics}\label{basics}
355: 
356: Our starting point is a general theory describing $N$ neutrino species.
357: The theory is assumed to include all possible 
358: Majorana- and Dirac-type couplings of left- and right-handed neutrinos,
359: including Lorentz- and CPT-violating ones. 
360: The neutrino sector of the minimal SME is therefore included,
361: along with other terms such as those involving right-handed neutrinos.
362: 
363: We denote the neutrino fields by the set of Dirac spinors
364: \{$\nu_e$, $\nu_\mu$, $\nu_\ta$,\ldots \}
365: and their charge conjugates by
366: \{$\nu_{e^C}\equiv \nu^C_e$,
367: $\nu_{\mu^C}\equiv \nu^C_\mu$,
368: $\nu_{\ta^C}\equiv \nu^C_\ta$,\ldots \},
369: where charge conjugation of a Dirac spinor
370: is defined as usual:
371: $\nu^C_a \equiv C\nub^T_a$.
372: By definition,
373: active neutrinos are detected via
374: weak interactions with left-handed
375: components of \{$\nu_e$, $\nu_\mu$, $\nu_\ta$\}.
376: Complications may arise in the full SME,
377: where Lorentz-violating terms alter these interactions
378: and can modify the detection process.
379: However, 
380: such modifications are expected to be tiny and
381: well beyond the sensitivity of current experiments.
382: In contrast, 
383: propagation effects can become appreciable for large baselines.
384: We therefore focus in this work
385: on solutions to the Lorentz-violating equations of motion
386: that describe free propagation of the $N$ neutrino species.
387: 
388: It is convenient to place all the fields and their conjugates
389: into a single object $\nu_A$,
390: where the index $A$ ranges over the $2N$ possibilities
391: $\{e,\mu,\ta,\ldots, e^C,\mu^C,\ta^C,\ldots \}$.
392: This setup allows us to write the equations of motion
393: in a form analogous to the Lorentz-violating QED extension 
394: \cite{ck,kle},
395: and it can readily accommodate Dirac, Majorana, 
396: or more general types of neutrinos.
397: Our explicit analysis in this section is performed 
398: under the assumption that Lorentz-violating operators 
399: of renormalizable dimension dominate the low-energy physics.
400: Then,
401: the general equations of motion for free propagation
402: can be written as a first-order differential operator acting on
403: the object $\nu_A$:
404: \beq
405: (i\Ga^\nu_{AB}\prt_\nu-M_{AB})\nu_B=0 .
406: \label{de}
407: \eeq
408: Here,
409: each constant quantity $\Ga^\mu_{AB}$, $M_{AB}$
410: is also a $4 \times 4$ matrix in spinor space.
411: Note that the usual equations of motion 
412: for Dirac and Majorana neutrinos 
413: are special cases of this equation.
414: 
415: The matrices $\Ga^\mu_{AB}$ and $M_{AB}$ can be decomposed
416: using the basis of $\ga$ matrices.
417: We define 
418: \bea
419: \Ga^\nu_{AB} &\equiv&
420: \ga^\nu \de_{AB}
421: + c^\mn_{AB}\ga_\mu
422: + d^\mn_{AB}\ga_5\ga_\mu
423: \nonumber \\
424: && + e^\nu_{AB}
425: + if^\nu_{AB}\ga_5
426: + \half g^{\la\mn}_{AB}\si_{\la\mu} ,
427: \nonumber\\
428: M_{AB} &\equiv&
429: m_{AB}
430: +im_{5AB}\ga_5
431: \nonumber \\
432: && + a^\mu_{AB}\ga_\mu
433: + b^\mu_{AB}\ga_5\ga_\mu
434: + \half H^\mn_{AB}\si_\mn .
435: \label{GaM}\eea
436: In these equations,
437: the masses $m$ and $m_5$ are Lorentz and CPT conserving.
438: The coefficients $c$, $d$, $H$
439: are CPT conserving but Lorentz violating,
440: while $a$, $b$, $e$, $f$, $g$
441: are both CPT and Lorentz violating.
442: Requiring hermiticity of the theory imposes the conditions 
443: $\Ga^\nu_{AB} = \ga^0(\Ga^\nu_{BA})^\dag\ga^0$
444: and $M_{AB} = \ga^0(M_{BA})^\dag\ga^0$,
445: which implies all coefficients are hermitian in generation space.
446: 
447: The above construction carries some redundancies 
448: that stem from the interdependence of $\nu$ and $\nu^C$.
449: This implies certain symmetries for $\Ga^\nu$ and $M$.
450: Note first that charge conjugation can be written
451: as a linear transformation on $\nu_A$:
452: $\nu^C_A = \cmat_{AB} \nu_B$,
453: where $\cal C$ is the symmetric matrix
454: with nonzero elements
455: $\cmat_{e e^C}
456: =\cmat_{\mu \mu^C}
457: =\cmat_{\ta \ta^C}=\cdots=1$.
458: Then,
459: in terms of $\cal C$ and the spinor matrix $C$,
460: the interdependence of $\nu$ and $\nu^C$ implies the relations
461: \bea
462: \Ga^\nu_{AB}&=&
463: -\cmat_{AC}\cmat_{BD}C(\Ga^\nu_{DC})^TC^{-1} ,
464: \nonumber\\
465: M_{AB}&=&\cmat_{AC}\cmat_{BD}C(M_{DC})^TC^{-1} ,
466: \label{GaMC}\eea
467: where the transpose $T$ acts in spinor space.
468: Suppressing generation indices,
469: this translates to
470: \beq
471: \begin{array}{rclrcl}
472: c^\mn&=&\cmat(c^\mn)^T\cmat , &
473: m&=&\cmat (m)^T \cmat , \\
474: d^\mn&=&-\cmat (d^\mn)^T \cmat , &
475: m_{5}&=& \cmat (m_5)^T   \cmat , \\
476: e^\nu&=&-\cmat (e^\nu)^T \cmat , &
477: a^\nu&=&-\cmat (a^\nu)^T \cmat , \\
478: f^\nu&=&-\cmat (f^\nu)^T \cmat  , &
479: b^\nu&=& \cmat (b^\nu)^T \cmat , \\
480: g^{\la\mn}&=&\cmat (g^{\la\mn})^T \cmat , &
481: H^\mn&=&-\cmat (H^\mn)^T \cmat ,
482: \end{array}
483: \label{C}
484: \eeq
485: where now the transpose $T$ acts in generation space.
486: Note that the overall sign in the above equations are chosen
487: to match their derivation within the conventional
488: lagrangian formalism involving anticommuting fermion fields.
489: 
490: Equation \rf{de} provides a basis
491: for a general Lorentz- and CPT-violating
492: relativistic quantum mechanics of freely propagating neutrinos.
493: However,
494: the unconventional time-derivative term complicates
495: the construction of the corresponding hamiltonian.
496: This difficulty also arises in the minimal QED extension,
497: but it may be overcome 
498: \cite{kle}
499: if there exists 
500: a nonsingular matrix $A$ satisfying the relationship
501: $A^\dag\ga^0\Ga^0A=1$.
502: The field redefinition
503: $\nu_A = A_{AB} \ch_B$
504: then allows the equations of motion \rf{de} 
505: to be written as
506: $(i\de_{AB}\prt_0-\cH_{AB}) \ch_B = 0$,
507: where the hamiltonian is given by
508: $\cH=-A^\dag\ga^0(i\Ga^j\prt_j-M)A$.
509: 
510: Denoting $\de\Ga^\nu$ and $\de M$
511: as the Lorentz-violating portions
512: of $\Ga^\nu$ and $M$,
513: and under the reasonable assumption that $|\de\Ga^0| < 1$,
514: a satisfactory field redefinition is
515: given by the power series
516: $A=(1+\ga^0\de\Ga^0)^{-1/2}=1-\half\ga^0\de\Ga^0+\cdots$.
517: Separating the hamiltonian $\cH$ into a
518: Lorentz-conserving part $\cH_0$
519: and a Lorentz-violating part $\de \cH$,
520: which we assume is small relative to $\cH_0$,
521: we can use the above expression for $A$ 
522: to obtain an expansion of $\de \cH$
523: in terms of $\cH_0$ and coefficients for Lorentz violation.
524: Explicitly,
525: at leading order in coefficients for Lorentz violation,
526: we obtain 
527: \beq
528: \de\cH = -\half(\ga^0\de\Ga^0\cH_0+\cH_0\ga^0\de\Ga^0)
529: -\ga^0(i\de\Ga^j\prt_j-\de M) . 
530: \label{dH}
531: \eeq
532: This expression is therefore the basis for
533: a general study of leading-order Lorentz and CPT violation 
534: in the neutrino sector. 
535: 
536: At this stage,
537: prior to beginning our study of Eq.\ \rf{dH}, 
538: it is useful to review the properties
539: of the Lorentz-conserving hamiltonian
540: \cite{nuphysics,nuphysics2}
541: \beq
542: \cH_0=-\ga^0(i\ga^j\prt_j-M_0).
543: \eeq
544: The Lorentz-conserving dynamics
545: is completely determined by the mass matrix $M_0$, 
546: which in its general form can be written
547: \beq
548: M_0=m+im_5\ga_5 =m_L P_L+m_R P_R\ ,
549: \eeq
550: with $m_R=(m_L)^\dag=m+im_5$ and
551: $P_L=\half(1-\ga_5), P_R=\half(1+\ga_5)$.
552: The components of the matrix $m_R=m^\dag_L$
553: can be identified with Dirac- or
554: Majorana-type masses by separating $m_R$ 
555: into four $N \times N$ submatrices.
556: It is often encountered in
557: the form of the symmetric matrix
558: \beq
559: m_R \cmat=
560: \left(\begin{array}{cc}
561: L & D \\
562: D^T & R
563: \end{array}\right) .
564: \label{usumass}
565: \eeq
566: The matrices $R$ and $L$ are the
567: right- and left-handed Majorana-mass matrices,
568: while $D$ is the Dirac-mass matrix.
569: In general, $R$, $L$ and $D$
570: are complex matrices
571: restricted only by the requirement that
572: $R$ and $L$ are symmetric.
573: Note that a left-handed Majorana coupling
574: is incompatible with electroweak-gauge invariance.
575: In contrast, Dirac and right-handed Majorana
576: couplings can preserve the usual gauge invariance.
577: 
578: It is always possible to find a basis
579: in which the mass matrix $M_0$ is diagonal.
580: Labeling the fields in this basis by $\ch_\AA$,
581: where $\AA  = 1,\ldots,2N$,
582: then the unitary transformation
583: relating the two bases can be written as 
584: \beq
585: U_{\AA A}=V_{\AA A} P_L
586: +(V\cmat)^*_{\AA A} P_R,
587: \label{U}
588: \eeq
589: where $V$ is a $2N \times 2N$
590: unitary matrix.
591: Here, 
592: it is understood that
593: $U_{\AA A}$ carries spinor indices that
594: have been suppressed.
595: In the new basis, the mass matrix
596: $m_{L\AA \BB}
597: =m_{R\AA \BB}
598: =m_{(\AA )}\de_{\AA \BB}$
599: is diagonal with real nonnegative entries.
600: The neutrinos
601: $\ch_\AA=\ch^C_\AA
602: =V_{\AA A} P_L \ch_A
603: +V^*_{\AA A} P_R \ch^C_A$
604: are Majorana particles,
605: regardless of the form of $M_0$.
606: 
607: \subsection{Effective hamiltonian}\label{phenom}
608: 
609: The discussion above applies to an arbitrary number 
610: of neutrino species and an arbitrary mass spectrum.
611: Since a general treatment is rather cumbersome,
612: we restrict attention in what follows 
613: to the minimal physically reasonable extension with $N=3$.
614: For definiteness,
615: we also assume a standard seesaw mechanism
616: \cite{seesaw}
617: with the components of $R$ much larger than those of $D$ or $L$.
618: This mechanism suppresses the propagation of
619: right-handed neutrinos,
620: so the analysis below also contains 
621: other Lorentz- and CPT-violating scenarios dominated 
622: by light or massless left-handed neutrinos,
623: including the minimal SME.
624: 
625: Ordering the masses $m_{(\AA )}$ from smallest to largest,
626: we assume that $m_{(1)}$, $m_{(2)}$, $m_{(3)}$
627: are small compared to the neutrino energies and possibly zero,
628: and that the remaining masses $m_{(4)}$, $m_{(5)}$, $m_{(6)}$
629: are large with the corresponding energy eigenstates 
630: kinematically forbidden.
631: In this situation the submatrix $V_{\aa a}$, 
632: where $a=e,\mu,\ta$ and $\aa =1,2,3$, 
633: is approximately unitary.
634: 
635: To aid in solving the equations of motion, 
636: we define
637: \bea
638: \ch_A(t;\vec x)
639: &=&\int \fr{d^3p}{(2\pi)^3}\ch_A(t;\vp)
640: e^{i\vp\cdot\vec x} ,
641: \nonumber \\
642: \ch_A(t;\vp)
643: &=&b_A(t;\vp)u_L(\vp)
644: +(\cmat d)_A(t;\vp)u_R(\vp)
645: \nonumber \\
646: &&+(\cmat b)_A^*(t;-\vp)v_R(-\vp)
647: +d_A^*(t;-\vp)v_L(-\vp) .
648: \nonumber \\
649: \label{nup}\eea
650: This is chosen to satisfy explicitly 
651: the charge-conjugation condition $\ch^C_A=\cmat_{AB}\ch_B$.
652: The spinor basis
653: $\{u_L(\vp), u_R(\vp),v_R(-\vp), v_L(-\vp)\}$
654: obeys the usual relations for massless fermions,
655: with $v_{R,L}(\vp)=C\bar u^T_{L,R}(\vp)$.
656: It has eigenvalues of the helicity operator
657: $\ga_5\ga^0\vec\ga\cdot\vp/|\vp|$
658: given by $\{-,+,-,+\}$
659: and eigenvalues of the chirality operator $\ga_5$
660: given by $\{-,+,+,-\}$.
661: For simplicity, we normalize with
662: $u_\al^\dag u_\be = v_\al^\dag v_\be =\de_{\al\be}$
663: for $\al,\be = L,R$.
664: The definition \rf{nup} implies that the amplitudes
665: $b_{e, \mu, \ta}$
666: may be approximately identified with active neutrinos
667: and $d_{e, \mu, \ta}$ with active antineutrinos.
668: The remaining amplitudes
669: $b_{e^C,\mu^C,\ta^C}$ and $d_{e^C,\mu^C,\ta^C}$
670: cover the space of sterile right-handed neutrinos,
671: but a simple identification with flavor neutrinos and antineutrinos
672: would be inappropriate in view of their large mass.
673: 
674: In the mass-diagonal Majorana basis,
675: we restrict attention to the propagating states 
676: consisting of the light neutrinos.
677: Taking the hamiltonian in this basis,
678: \beq
679: \cH_{\aa \bb}(\vp)=
680: \ga^0(\vec\ga\cdot\vp  +m_{(\aa )})
681: \de_{\aa \bb}
682: +\de \cH_{\aa \bb}(\vp),
683: \eeq
684: and applying it to
685: $\ch_{\bb}(t;\vp)
686: =U_{\bb B}\ch_B(t;\vp)$
687: yields the equations of motion
688: in terms of the amplitudes $b$ and $d$.
689: The result takes the form of the matrix equation
690: \beq
691: [i\de_{\aa \bb}\prt_0
692: -H_{\aa \bb}(\vp)]
693: \left(
694: \begin{array}{c}
695: b_\bb(t;\vp) \\
696: d_\bb(t;\vp) \\
697: b_\bb^*(t;-\vp) \\
698: d_\bb^*(t;-\vp)
699: \end{array}
700: \right)=0 ,
701: \label{we}
702: \eeq
703: where for convenience we have defined 
704: $b_\bb=V_{\bb B}b_B$ and
705: $d_\bb=V_{\bb B}^*d_B$,
706: and where $H_{\aa \bb}$ 
707: is the spinor-decomposed form of 
708: $\cH_{\aa \bb}$.
709: 
710: The propagation of kinematically allowed states 
711: is completely determined by the amplitudes $b_\aa$ and $d_\aa$.
712: However, 
713: for purposes of comparison with experiment
714: it is convenient to express the result using
715: the amplitudes associated with active neutrinos,
716: $b_{e,\mu,\ta}$ and $d_{e,\mu,\ta}$.
717: The relevant calculation is somewhat lengthy
718: and is deferred to Appendix \ref{hcalc}.
719: It assumes that the submatrix $V_{\aa a}$ is unitary,
720: and it neglects terms that enter 
721: as small masses $m_{(\aa)}$ multiplied by 
722: coefficients for Lorentz violation,
723: since these are typically suppressed.
724: The calculation reveals that the time evolution 
725: of the active-neutrino amplitudes
726: is given by the equation
727: \beq
728: \left(
729: \begin{array}{c}
730: b_a(t;\vp)\\
731: d_a(t;\vp)
732: \end{array}
733: \right)
734: =\exp(-i\heff t)_{ab}
735: \left(
736: \begin{array}{c}
737: b_b(0;\vp)\\
738: d_b(0;\vp)
739: \end{array}
740: \right) ,
741: \label{Ut}
742: \eeq
743: where $\heff$ is the effective hamiltonian
744: describing flavor neutrino propagation.
745: To leading order,
746: it is given by
747: \begin{widetext}
748: \bea
749: (\heff)_{ab}&=&
750: |\vp|\de_{ab}
751: \left(\begin{array}{cc}
752: 1 & 0 \\
753: 0 & 1
754: \end{array}\right)
755: +\fr{1}{2|\vp|}
756: \left(\begin{array}{cc}
757: (\mt)_{ab} & 0 \\
758: 0& (\mt)^*_{ab}
759: \end{array}\right)
760: \nonumber \\
761: &&\quad+\fr{1}{|\vp|}
762: \left(\begin{array}{cc}
763: [(a_L)^\mu p_\mu-(c_L)^\mn p_\mu p_\nu]_{ab} &
764: -i\sqrt{2} p_\mu (\ep_+)_\nu
765: [(g^{\mn\si}p_\si-H^\mn)\cmat]_{ab} \\
766: i\sqrt{2} p_\mu (\ep_+)^*_\nu
767: [(g^{\mn\si}p_\si+H^\mn)\cmat]^*_{ab} &
768: [-(a_L)^\mu p_\mu-(c_L)^\mn p_\mu p_\nu]^*_{ab}
769: \end{array}\right) ,
770: \label{heff}
771: \eea
772: \end{widetext}
773: where we have defined
774: $(c_L)^\mn_{ab}\equiv(c+d)^\mn_{ab}$ and
775: $(a_L)^\mu_{ab}\equiv(a+b)^\mu_{ab}$
776: for reasons explained below.
777: The approximate four momentum $p_\mu$ 
778: may be taken as $p_\mu=(|\vp|;-\vp)$
779: at leading order.
780: The Lorentz-conserving mass term 
781: results from the usual seesaw mechanism 
782: with $\mt\equiv\ml \ml^\dag$,
783: where $\ml$ is the light-mass matrix
784: $\ml= L - DR^{-1}D^T$.
785: The complex vector $(\ep_+)_\mu$ satisfies
786: the conditions
787: \bea
788: p^\mu(\ep_+)^\nu-p^\nu(\ep_+)^\mu
789: &=&i\ep^{\mn\rh\si}
790: p_\rh(\ep_+)_\si ,\nonumber \\
791: (\ep_+)^\nu(\ep_+)^*_\nu&=&-1 .
792: \label{ep}
793: \eea
794: A suitable choice is
795: $(\ep_+)^\nu=\frac{1}{\sqrt{2}}(0;\hat\ep_1+i\hat\ep_2)$,
796: where $\hat\ep_1$, $\hat\ep_2$ are real
797: and $\{ \vp/|\vp|, \hat\ep_1, \hat\ep_2 \}$
798: form a right-handed orthonormal triad.
799: Note that $(\ep_+)^\nu$ and
800: $(\ep_-)^\nu\equiv(\ep_+)^{\nu*}$
801: is analogous to the usual photon helicity basis.
802: The appearance of these vectors
803: reflects the near-definite helicity of active neutrinos.
804: The vectors $\hat\ep_1$ and $\hat\ep_2$
805: can be arbitrarily set by rotations
806: or equivalently by multiplying $(\ep_+)^\nu$ by a phase,
807: which turns out to be equivalent to changing the relative phase 
808: between the basis spinors $u_L$ and $u_R$.
809: 
810: Only the diagonal kinetic term in $\heff$
811: arises in the minimal SM.
812: The term involving $(\mt)_{ab}$ 
813: encompasses the usual massive-neutrino case 
814: without sterile neutrinos.
815: The leading-order Lorentz-violating contributions 
816: to neutrino-neutrino mixing are controlled by 
817: the coefficient combinations 
818: $(a+b)^\mu_{ab}$ and $(c+d)^\mn_{ab}$.
819: These combinations conserve 
820: the usual SU(3)$\times$SU(2)$\times$U(1) gauge symmetry 
821: and correspond to the coefficients
822: $(a_L)^\mu_{ab}$ and $(c_L)^\mn_{ab}$
823: in the minimal SME.
824: Note that the orthogonal combinations
825: $(a-b)^\mu_{ab}$ and $(c-d)^\mn_{ab}$
826: also conserve the usual gauge symmetry,
827: but they correspond to self-couplings of right-handed neutrinos 
828: and are therefore irrelevant for leading-order processes
829: involving active neutrinos.
830: The remaining coefficients,
831: $(g^{\mn\si}\cmat)_{ab}$ and $(H^\mn\cmat)_{ab}$,
832: appear in $\heff$ through Majorana-like couplings 
833: that violate SU(3)$\times$SU(2)$\times$U(1) gauge invariance
834: and lepton-number conservation.
835: They generate Lorentz-violating neutrino-antineutrino mixing.
836: 
837: Some combinations of coefficients may be unobservable,
838: either due to symmetries 
839: or because they can be removed through field redefinitions
840: \cite{ck,kle,cm,bek}.
841: For example, 
842: the trace component $\et_\mn(c_L)^\mn$ is Lorentz invariant
843: and can be absorbed into the usual kinetic term,
844: so it may be assumed zero for convenience.
845: In fact,
846: even if this combination is initially nonzero,
847: it remains absent from the leading-order effective hamiltonian
848: because the trace of $p_\mu p_\nu$ vanishes.
849: Other examples of unobservable coefficients include 
850: certain combinations of $g^{\mn\si}$ and $H^\mn$.
851: The antisymmetry properties
852: $g^{\mn\si}=-g^{\nu\mu\si}$, $H^\mn=-H^{\nu\mu}$
853: and the properties of $(\ep_+)_\nu$ can be combined 
854: to prove that the physically significant combinations 
855: of $g^{\mn\si}$ and $H^\mn$ are given by the relations
856: \bea
857: p_\mu (\ep_+)_\nu g^{\mn\si}
858: =|\vp|(\ep_+)_\nu \gt^{\nu\si},
859: \nonumber \\
860: p_\mu (\ep_+)_\nu H^\mn
861: =|\vp|(\ep_+)_\nu \Ht^\nu, 
862: \eea
863: where we have defined 
864: \bea
865: \gt^{\nu\si}\equiv
866: g^{0\nu\si}
867: +\frac{i}{2}{\ep^{0\nu}}_{\ga\rh}g^{\ga\rh\si} ,
868: \nonumber \\
869: \Ht^\nu\equiv
870: H^{0\nu}
871: +\frac{i}{2}{\ep^{0\nu}}_{\ga\rh}H^{\ga\rh} .
872: \label{gtHt}
873: \eea
874: Only these combinations appear in $\heff$
875: and are relevant to neutrino oscillations.
876: 
877: In deriving Eq.\ \rf{heff},
878: we have focused on operators of renormalizable dimension,
879: which involve linear derivatives in the equations of motion
880: and a single power of momentum in the hamiltonian.
881: Operators of nonrenormalizable mass dimension $n>4$ are also of 
882: potential importance
883: \cite{kpo,kle}.
884: They appear as higher-derivative terms in the action,
885: along with corresponding complications in the equations of motion
886: and in the construction of the hamiltonian. 
887: An operator of dimension $n$ is associated 
888: with a term in the action involving $d=n-3$ derivatives,
889: and the associated terms in the effective hamiltonian
890: involve $d$ powers of the momentum. 
891: The corresponding coefficient for Lorentz violation 
892: carries $d+2$ or fewer Lorentz indices,
893: depending on the spinor structure of the coupling
894: and the number of momentum contractions occurring.
895: For the case $n>4$,
896: we generically denote the coefficients by $(k_d)^{\la\ldots}$.
897: These coefficients have mass dimension $1-d$.
898: Note that,
899: depending on the theory considered,
900: the mechanism for Lorentz and CPT violation
901: can cause them to be suppressed
902: by $d$-dependent powers of the Planck scale
903: \cite{kpo,kle}.
904: Some effects of operators with $d=2$ have been considered
905: in the context of quantum gravity in Ref.\ \cite{bef}. 
906: 
907: The mixing described by Eq.\ \rf{heff}
908: or its generalization to operators of dimension $n>4$
909: can be strongly energy dependent.
910: For example,
911: any nonzero mass-squared differences 
912: dominate the hamiltonian at some low-energy scale.
913: However,
914: while mass effects decrease with energy,
915: Lorentz-violating effects 
916: involving operators of renormalizable dimension 
917: remain constant or grow linearly with energy $E$
918: and so always dominate at high energies.
919: For instance, 
920: the contributions from a mass of $0.1$ eV
921: and a dimensionless coefficient of $10^{-17}$
922: are roughly comparable at an energy
923: determined by $E^2\sim (0.1$ eV$)^2/(10^{-17})$,
924: or $E\sim 30$ MeV.
925: Below this energy the mass term dominates, 
926: while above it the Lorentz-violating term does.
927: Similarly, 
928: a dimension-one coefficient of $10^{-15}$ GeV
929: has a transition energy $E\sim10$ keV.
930: More generally,
931: effects controlled by the coefficients $(k_d)^{\la\ldots}$
932: for Lorentz violation involving operators of dimension $n=d+3$
933: grow as $E^d$. 
934: 
935: Although the perturbative diagonalization leading to Eq.\ \rf{heff} 
936: is valid for dimensionless coefficients much smaller than one
937: and for energies much greater than any masses 
938: or coefficients of dimension one,
939: at sufficiently high energies
940: issues of stability and causality 
941: may require the inclusion of Lorentz-violating terms 
942: of nonrenormalizable dimension in the theory. 
943: In the context of the single-fermion QED extension,
944: for example,
945: a dimensionless $c^{00}$ coefficient
946: can lead to issues with causality and stability 
947: at energies $\sim m_{\rm fermion}/\sqrt{c^{00}}$
948: unless the effects of operators of nonrenormalizable dimension 
949: are incorporated
950: \cite{kle}.
951: A complete resolution of this issue would be of interest
952: but lies beyond our present scope.
953: It is likely to depend on the underlying mechanisms
954: leading to mass and Lorentz violation,
955: and it may be complicated further by the presence 
956: of multiple generations and the sterile neutrino sector. 
957: We limit our remarks here to noting that  
958: the values of the coefficients for Lorentz violation
959: considered in all the models in this work
960: are sufficiently small that
961: issues of stability and causality can be arranged 
962: to appear only beyond experimentally relevant energies.
963: In any case, 
964: the renormalizable sector 
965: provides a solid foundation for the basic treatment
966: of Lorentz and CPT violation in neutrinos. 
967: 
968: \subsection{Neutrinos in matter}\label{matter}
969: 
970: In many situations,
971: neutrinos traverse a significant volume of ordinary matter 
972: before detection.
973: The resulting forward scattering with electrons, protons, and neutrons 
974: can have dramatic consequences on neutrino oscillations
975: \cite{kuop}.
976: These matter interactions can readily be incorporated
977: into our general formalism.
978: Since the effective lagrangian in normal matter is given by 
979: $\De\cl_{\rm matter}=
980: -\sqrt{2}G_Fn_e\nub_e\ga^0P_L\nu_e
981: +(G_Fn_n/\sqrt{2})\nub_a\ga^0P_L\nu_a$,
982: matter effects are equivalent to contributions
983: from CPT-odd coefficients
984: \bea
985: (a_{L,\rm eff})^0_{ee}&=&G_F(2n_e-n_n)/\sqrt{2} ,
986: \nonumber \\
987: (a_{L,\rm eff})^0_{\mu\mu}&=&(a_{L,\rm eff})^0_{\ta\ta}=
988: -G_Fn_n/\sqrt{2} ,
989: \eea
990: where $n_e$ and $n_n$ are the number
991: densities of electrons and neutrons.
992: Adding these terms to the effective hamiltonian \rf{heff}
993: therefore incorporates the effects of matter.
994: 
995: For some of the analyses of Lorentz violation below,
996: it is useful to review the treatment of matter effects
997: in solar and atmospheric neutrinos. 
998: Consider first solar neutrinos.
999: These are produced in several processes
1000: that generate distinct, well-understood $\nu_e$ spectra.
1001: The most notable are the pp spectrum 
1002: with a maximum energy of about 0.4 MeV,
1003: and the $^8$B spectrum
1004: with a maximum of about 16 MeV
1005: \cite{ssm}.
1006: For $\nu_a\mix\nu_b$ mixing scenarios,
1007: the contribution from $n_n$ is the same for all species
1008: and therefore can be ignored.
1009: However,  
1010: $n_n$ may be important for $\nu_a\mix\nub_b$ mixing,
1011: such as that generated by the coefficients
1012: $(g^{\mn\si}\cmat)_{ab}$ and $(H^\mn\cmat)_{ab}$ in $\heff$.
1013: An analytic approximation to the electron number density 
1014: inside the Sun is given by \cite{ssm}
1015: $n_e/N_A=245e^{-10.54R/R_\odot}$.
1016: It is useful to define $n_s=n_e-\half n_n$,
1017: a combination that often appears in sterile-neutrino searches.
1018: This number density has a similar approximation,
1019: $n_s/N_A=223e^{-10.54R/R_\odot}$.
1020: The two linearly independent combinations can therefore be taken as
1021: $G_Fn_e\simeq 1.32\times10^{-20}e^{-10.54R/R_\odot}$ GeV
1022: and
1023: $G_Fn_s\simeq 1.20\times10^{-20}e^{-10.54R/R_\odot}$ GeV,
1024: corresponding to a neutron contribution of
1025: $G_Fn_n=2G_F(n_e-n_s)\simeq 0.24\times10^{-20}e^{-10.54R/R_\odot}$ GeV
1026: to the effective hamiltonian.
1027: These quantities set the scale for matter effects in the Sun.
1028: 
1029: Next, consider the detection of atmospheric neutrinos.
1030: Upward-going neutrinos pass through the Earth 
1031: and therefore experience higher matter potentials
1032: than the downward-going neutrinos,
1033: which traverse the less dense atmosphere and a small amount
1034: of bedrock on their way to the detector.
1035: A crude estimate of the matter potential in this case 
1036: can be obtained by assuming that the Earth
1037: consists of roughly equal numbers of 
1038: protons, neutrons, and electrons.
1039: Using the average number density then
1040: yields the approximate value
1041: $G_Fn_e\simeq G_Fn_n\simeq 1.5\times10^{-22}$ GeV.
1042: This produces a matter potential similar to that 
1043: from the Sun at $R/R_\odot\sim 2/5$.
1044: 
1045: Overall,
1046: the contribution to $\heff$ from matter ranges from about
1047: $10^{-20}$ GeV to $10^{-25}$ GeV.
1048: This means that matter effects must be incorporated
1049: when the contributions from mass or Lorentz violation 
1050: lie near these values.
1051: This range is comparable to the scale 
1052: of coefficients for Lorentz violation 
1053: that originate as suppressed effects from the Planck scale. 
1054: Note also that most terrestrial experiments
1055: involve neutrinos that traverse at least some amount of bedrock 
1056: or other shielding materials,
1057: which can result in substantially different conventional 
1058: or Lorentz-violating dynamics 
1059: relative to the vacuum-oscillation case
1060: \cite{mattercpt}.
1061: 
1062: 
1063: \subsection{Neutrino oscillations}\label{nuosc}
1064: 
1065: The analysis of neutrino mixing proceeds along the usual lines.
1066: The effective hamiltonian can be diagonalized 
1067: with a $6\times 6$ unitary matrix $\Ueff$:
1068: \beq
1069: \heff =  \Ueff^\dag E_{\rm eff} \Ueff ,
1070: \label{diaheff}
1071: \eeq
1072: where $E_{\rm eff}$ is a $6\times 6$ diagonal matrix.
1073: In contrast to the Lorentz-covariant case,
1074: where mixing without sterile neutrinos involves only three propagating states,
1075: here mixing without sterile neutrinos may occur with six states.
1076: This means that there can be 
1077: up to five energy-dependent eigenvalue differences
1078: for Lorentz-violating mixing,
1079: resulting in five independent oscillation lengths
1080: instead of the usual two.
1081: 
1082: Denoting the six propagation states 
1083: by the amplitudes $B_J(t;\vp)$
1084: with $J=1,\ldots ,6$, 
1085: we can write 
1086: $B_J(t;\vp)=\tu_{Ja}b_a(t;\vp)+\bu_{Ja}d_a(t;\vp)$,
1087: where we have split $\Ueff$ into $6\times 3$ matrices
1088: $\Ueff=(\tu, \bu)$.
1089: The time evolution operator may then be written as
1090: \bea
1091: S_{ab}(t)
1092: &=& (\Ueff^\dag e^{-iE_{\rm eff}t} \Ueff)_{ab}
1093: \nonumber \\
1094: &=& \left(\begin{array}{cc}
1095: S_{\nu_a\nu_b}(t) & S_{\nu_a\nub_b}(t)\\
1096: S_{\nub_a\nu_b}(t) & S_{\nub_a\nub_b}(t)
1097: \end{array}\right)
1098: \nonumber \\
1099: &=&\sum_{J} e^{-itE_{(J)}}
1100: \left(\begin{array}{cc}
1101: \tu^*_{Ja}\tu_{Jb} & \tu^*_{Ja}\bu_{Jb} \\
1102: \bu^*_{Ja}\tu_{Jb} & \bu^*_{Ja}\bu_{Jb}
1103: \end{array}\right) ,
1104: \label{Ut2}\eea
1105: where $E_{(J)}$ are the diagonal values of $E_{\rm eff}$.
1106: 
1107: The probabilities for a neutrino of type $b$
1108: oscillating into a neutrino or antineutrino 
1109: of type $a$ in time $t$ are therefore 
1110: $P_{\nu_b\to\nu_a}(t)=|S_{\nu_a\nu_b}(t)|^2$
1111: or 
1112: $P_{\nu_b\to\nub_a}(t)=|S_{\nub_a\nu_b}(t)|^2$,
1113: respectively.
1114: Similarly, 
1115: for antineutrinos we have
1116: $P_{\nub_b\to\nu_a}(t)=|S_{\nu_a\nub_b}(t)|^2$
1117: or 
1118: $P_{\nub_b\to\nub_a}(t)=|S_{\nub_a\nub_b}(t)|^2$.
1119: In terms of the matrices
1120: $\tu$ and $\bu$,
1121: the probabilities are
1122: \begin{widetext}
1123: \bse\bea
1124: P_{\nu_b\to\nu_a}(t)&=&\de_{ab}
1125: -4\sum_{J>K}\Re(\tu^*_{Ja}\tu_{Jb}\tu_{Ka}\tu^*_{Kb})
1126: \sin^2\fr{\De_{JK}t}2
1127: +2\sum_{J>K}\Im(\tu^*_{Ja}\tu_{Jb}\tu_{Ka}\tu^*_{Kb})
1128: \sin\De_{JK}t\ ,
1129: \label{pnn}\\
1130: %
1131: P_{\nub_b\to\nub_a}(t)&=&\de_{ab}
1132: -4\sum_{J>K}\Re(\bu^*_{Ja}\bu_{Jb}\bu_{Ka}\bu^*_{Kb})
1133: \sin^2\fr{\De_{JK}t}2
1134: +2\sum_{J>K}\Im(\bu^*_{Ja}\bu_{Jb}\bu_{Ka}\bu^*_{Kb})
1135: \sin\De_{JK}t\ ,
1136: \label{pbb}\\
1137: %
1138: P_{\nu_b\to\nub_a}(t)&=&
1139: -4\sum_{J>K}\Re(\bu^*_{Ja}\tu_{Jb}\bu_{Ka}\tu^*_{Kb})
1140: \sin^2\fr{\De_{JK}t}2
1141: +2\sum_{J>K}\Im(\bu^*_{Ja}\tu_{Jb}\bu_{Ka}\tu^*_{Kb})
1142: \sin\De_{JK}t\ ,
1143: \label{pnb}\\
1144: %
1145: P_{\nub_b\to\nu_a}(t)&=&
1146: -4\sum_{J>K}\Re(\tu^*_{Ja}\bu_{Jb}\tu_{Ka}\bu^*_{Kb})
1147: \sin^2\fr{\De_{JK}t}2
1148: +2\sum_{J>K}\Im(\tu^*_{Ja}\bu_{Jb}\tu_{Ka}\bu^*_{Kb})
1149: \sin\De_{JK}t\ ,
1150: \label{pbn}
1151: \eea\label{probs}\ese
1152: \end{widetext}
1153: where the effective-energy difference
1154: is denoted by $\De_{JK}=E_{(J)}-E_{(K)}$.
1155: 
1156: 
1157: \subsection{CPT properties}\label{cptprop}
1158: 
1159: With a conveniently chosen phase,
1160: CPT may be implemented by the transformation
1161: \beq
1162: \left(\begin{array}{c}
1163: b^{\rm CPT}_a(t;\vp)\\
1164: d^{\rm CPT}_a(t;\vp)
1165: \end{array}\right)
1166: =i\left(\begin{array}{c}
1167: -d^*_a(-t;\vp)\\
1168: b^*_a(-t;\vp)
1169: \end{array}\right)
1170: \equiv \si^2
1171: \left(\begin{array}{c}
1172: b^*_a(-t;\vp)\\
1173: d^*_a(-t;\vp)
1174: \end{array}\right) .
1175: \label{cpt}
1176: \eeq
1177: This yields precisely the expected
1178: result when applied to $\heff$:
1179: the CPT-conjugate hamiltonian
1180: $h^{\rm CPT}_{\rm eff}=\si^2\heff^*\si^2$
1181: can be obtained from Eq.\ \rf{heff}
1182: by changing the sign of the CPT-odd
1183: $a_L$ and $g$ coefficients.
1184: Then,
1185: $h^{\rm CPT}_{\rm eff}=\heff$ when $a_L$ and $g$ vanish,
1186: as expected.
1187: A notable feature here is that independent mass matrices
1188: for neutrinos and antineutrinos
1189: cannot be generated as has been proposed 
1190: \cite{mmb}.
1191: Greenberg has recently proved that this result
1192: is general 
1193: \cite{owg}.
1194: 
1195: Under CPT, 
1196: the transition amplitudes transform as
1197: \bse\bea
1198: S_{\nu_a\nu_b}(t)
1199: &\stackrel{\rm CPT}{\longleftrightarrow}&
1200: S^*_{\nub_a\nub_b}(-t),
1201: \label{Scpt1} \\
1202: S_{\nub_a\nu_b}(t)
1203: &\stackrel{\rm CPT}{\longleftrightarrow}&
1204: -S^*_{\nu_a\nub_b}(-t) .
1205: \label{Scpt2}
1206: \eea\label{Scpt}\ese
1207: These relations become equalities if CPT holds.
1208: The first relation then yields the usual result,
1209: \bse
1210: \beq
1211: \mbox{\rm CPT invariance }
1212: \implies
1213: P_{\nu_b\to\nu_a}(t) = P_{\nub_a\to\nub_b}(t) .
1214: \label{cpt1}
1215: \eeq
1216: This property has long been understood
1217: and has been identified as a potential test of CPT invariance
1218: \cite{fc2}.
1219: However, 
1220: the negation of terms in this result
1221: produces a statement that may be false in general
1222: because CPT violation need not imply 
1223: $P_{\nu_b\to\nu_a}(t) \neq 
1224: P_{\nub_a\to\nub_b}(t)$.
1225: Examples of models that violate CPT but nonetheless
1226: satisfy Eq.\ \rf{cpt1} are given in Sec.\ \ref{models}.
1227: 
1228: The above property addresses the relationship between 
1229: $\nu\mix\nu$ and $\nub\mix\nub$ mixing.
1230: There is also an analogous property
1231: associated with $\nu\mix\nub$ mixing.
1232: Thus,
1233: for CPT invariance,
1234: relation \rf{Scpt2} yields the additional result:
1235: \beq
1236: \mbox{\rm CPT invariance }
1237: \implies
1238: P_{\nu_b\rightleftarrows\nub_a}(t) = 
1239: P_{\nu_a\rightleftarrows\nub_b}(t).
1240: \label{cpt2}
1241: \eeq
1242: \ese
1243: This property may also provide opportunities 
1244: to test for Lorentz and CPT invariance.
1245: Note,
1246: however,
1247: that negation of its terms produces a statement that may be false in general,
1248: as in the previous case.
1249: 
1250: Finally,
1251: we emphasize that the presence of CPT violation increases the 
1252: number of independent oscillation lengths 
1253: without the addition of sterile neutrinos.
1254: In the general case,
1255: nonzero coefficients for CPT violation 
1256: in the effective hamiltonian \rf{heff}
1257: can generate up to six independent propagating states,
1258: rather than the usual three.
1259: 
1260: 
1261: \subsection{Reference frames} \label{frames}
1262: 
1263: The presence of Lorentz violation makes 
1264: it necessary to specify the frame 
1265: in which experimental results are reported.
1266: Coordinate invariance of the physics,
1267: in particular observer Lorentz invariance
1268: \cite{ck},
1269: ensures that the analysis and measurements of an experiment
1270: can be performed in any frame of reference.
1271: However,
1272: it is convenient to have a standard
1273: set of frames to facilitate comparisons of different experiments.
1274: In the literature,
1275: measurements are conventionally expressed 
1276: in terms of coefficients for Lorentz violation
1277: defined in a Sun-centered celestial equatorial frame 
1278: with coordinates $(T,X,Y,Z)$
1279: \cite{fn1}.
1280: For our present purposes, 
1281: it suffices to identify the $Z$ direction 
1282: as lying along the Earth's rotational axis and the
1283: $X$ direction as pointing towards the vernal equinox.
1284: The coefficients for Lorentz violation
1285: in any other inertial frame can be related 
1286: to the standard set in the Sun-centered frame 
1287: by an observer Lorentz transformation.
1288: In general, 
1289: this transformation includes both rotations and boosts,
1290: but boost effects are frequently neglected
1291: because they introduce only terms suppressed by
1292: the velocity $\be$ between frames,
1293: which is typically $\lsim 10^{-4}$.
1294: Recently, 
1295: studies of some $\be$-suppressed terms have been performed
1296: in the context of high-precision clock-comparison experiments
1297: \cite{spaceexpt,cane}
1298: and resonant cavities
1299: \cite{cavexpt,km}.
1300: 
1301: The existence of orientation-dependent effects 
1302: makes it useful to define a standard parametrization 
1303: for the direction of neutrino propagation $\hat p$ 
1304: and the corresponding $\hat\ep_1$, $\hat\ep_2$ vectors
1305: in the Sun-centered frame.
1306: A suitable set of unit vectors is given by
1307: \bea
1308: \hat p&=&(\sin\Th\cos\Ph,\sin\Th\sin\Ph,\cos\Th) ,\nonumber\\
1309: \hat\ep_1&=&(\cos\Th\cos\Ph,\cos\Th\sin\Ph,-\sin\Th) , \nonumber\\
1310: \hat\ep_2&=&(-\sin\Ph,\cos\Ph,0) ,
1311: \label{vectors}
1312: \eea
1313: where $\Th$ and $\Ph$ 
1314: are the celestial colatitude and longitude of propagation,
1315: respectively.
1316: We remark that these quantities are related 
1317: to the right ascension $r$ and declination $d$ 
1318: of the source as viewed from the detector
1319: by $\Th=90^\circ+d$ and $\Ph=180^\circ+r$.
1320: 
1321: In the remainder of this subsection,
1322: we provide some technical comments about the frame-dependence of 
1323: our choice of spinor basis in Sec.\ \ref{phenom}.
1324: This basis is normally associated with massless fermions,
1325: so the presence of mass or Lorentz violation 
1326: means that even with a covariant normalization
1327: the corresponding amplitudes 
1328: are no longer scalar functions 
1329: under observer Lorentz transformations
1330: and hence are frame dependent.
1331: However,
1332: our basis suffices for perturbative calculations
1333: in which the physically significant states are affected 
1334: only by masses and coefficients for Lorentz violation 
1335: that are small relative to $|\vp|$,
1336: while the complexity of the general Lorentz-violating case 
1337: makes the decomposition into a covariant basis impractical.
1338: Moreover, 
1339: despite the frame-dependent nature of the calculation,
1340: the probabilities \rf{probs} are frame independent at leading order.
1341: In the usual case,
1342: frame independence follows from the Lorentz-vector nature
1343: of the exact 4-momenta $(E_{(J)};\vp)$,
1344: which implies the products
1345: $E_{(J)}t-\vp\cdot\vec x$
1346: are Lorentz scalars, 
1347: and from the constancy and frame-independence 
1348: of the mixing matrix $\Ueff$.
1349: It turns out that a version of these properties 
1350: holds in the present case,
1351: as we show next.
1352: 
1353: First,
1354: we observe that
1355: the elements of the $6\times 6$ matrix 
1356: $|\vp|(\heff-|\vp|)$
1357: are scalars under observer Lorentz transformations
1358: at leading order in small quantities.
1359: Next, 
1360: note that the matrix $\Ueff$ diagonalizes
1361: $|\vp|(\heff-|\vp|)$,
1362: so its elements can be chosen to be observer Lorentz scalars as well.
1363: In turn, 
1364: this means that the diagonal elements
1365: $|\vp|(E_{(J)}-|\vp|)$
1366: are also observer Lorentz scalars,
1367: since they are functions of the elements of
1368: $|\vp|(\heff-|\vp|)$.
1369: From this result,
1370: it follows explicitly that
1371: the neutrino dispersion relations
1372: $E_{(J)}^2-\vp^{\, 2}$
1373: are observer Lorentz scalars at leading order,
1374: since
1375: \bea
1376: E_{(J)}^2-\vp^{\, 2}
1377: &=&(E_{(J)}+|\vp|)(E_{(J)}-|\vp|)
1378: \nonumber \\
1379: &\simeq&2|\vp|(E_{(J)}-|\vp|) .
1380: \label{disp}
1381: \eea
1382: The 4-momentum is therefore a vector 
1383: under observer Lorentz transformations to leading order,
1384: as desired.
1385: Combining this property with the scalar character of $\Ueff$
1386: implies that the leading-order transition amplitudes 
1387: and probabilities \rf{probs} 
1388: are covariant under observer Lorentz transformations,
1389: as claimed.
1390: 
1391: 
1392: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1393: 
1394: \section{Sensitivities}\label{sens}
1395: 
1396: \subsection{Existing constraints}
1397: 
1398: To date, 
1399: there is no compelling experimental evidence 
1400: for nonzero coefficients for Lorentz violation
1401: in any sector.
1402: Theoretical predictions of the size of the effects
1403: depend on the underlying model.
1404: However,
1405: the natural scale for a fundamental theory is the Planck mass $m_P$,
1406: which is about 17 orders of magnitude greater than the
1407: electroweak scale $m_W$ relevant to the SM 
1408: and roughly 30 orders of magnitude greater 
1409: than the scale of neutrino masses, 
1410: if they exist. 
1411: It is plausible that any observable 
1412: Lorentz- and CPT-violating effects 
1413: are suppressed by one or more powers of the dimensionless ratio 
1414: $r = m/m_P \lsim 10^{-17}$,
1415: where $m$ is the relevant low-energy scale
1416: and $m_P$ is the Planck mass  
1417: \cite{kp}.
1418: In contrast,
1419: the scale of observed neutrino oscillations is $\lsim 0.1$ eV,
1420: which enters as a squared mass $\De m^2 \lsim 10^{-20}$ GeV.
1421: At physically relevant energies,
1422: $10^{-4}$ GeV $< E < 10^3$ GeV,
1423: the oscillation physics is determined
1424: by the dimensionless ratio $r_\nu = \De m^2/E^2$. 
1425: Remarkably,
1426: the two dimensionless ratios $r$ and $r_\nu$ have a similar range,
1427: so the natural size of Lorentz- and CPT-violating effects 
1428: may be comparable to the natural size of neutrino-oscillation effects.
1429: 
1430: Certain experiments in the fermion and photon sectors 
1431: have achieved sensitivities corresponding to
1432: dimensionless suppressions of roughly $10^{-30}$.
1433: Since the coefficients for Lorentz violation 
1434: in the various sectors can be related either directly 
1435: through symmetries or indirectly through radiative corrections,
1436: it might seem that existing experimental constraints 
1437: severely restrict the possibilities for Lorentz violation in neutrinos.
1438: In fact,
1439: this expectation is incorrect,
1440: as we discuss next.
1441: 
1442: In the context of $\heff$,
1443: the relevant coefficients are
1444: $(a_L)^\mu_{ab}$ and $(c_L)^\mn_{ab}$,
1445: since these appear directly in the charged-fermion sector of the SME.
1446: A decomposition of the multi-flavor QED limit 
1447: of the charged-lepton sector 
1448: can be performed in analogy with Eq.\ \rf{GaM}.
1449: It produces the identification
1450: \bea
1451: a^\mu_{ab}&=&\half(a_L+a_R)^\mu_{ab} ,\nonumber\\
1452: b^\mu_{ab}&=&\half(a_L-a_R)^\mu_{ab} ,\nonumber\\
1453: c^\mn_{ab}&=&\half(c_L+c_R)^\mn_{ab} ,\nonumber\\
1454: d^\mn_{ab}&=&\half(c_L-c_R)^\mn_{ab} ,
1455: \eea
1456: where $(c_R)^\mn_{ab}$ and $(a_R)^\mu_{ab}$
1457: are coefficients in the SME that couple to right-handed leptons
1458: and therefore leave unaffected the active neutrinos at tree level.
1459: On this basis,
1460: it might naively appear that the charged sector is sensitive 
1461: to more combinations of coefficients for Lorentz violation 
1462: than the neutrino sector.
1463: However, 
1464: the mass hierarchy of the charged leptons
1465: $e$, $\mu$, $\ta$ 
1466: implies that only coefficients that are diagonal in flavor space 
1467: appear in leading-order perturbative calculations.
1468: As a result, 
1469: $e$, $\mu$, $\ta$ effectively decouple,
1470: resulting in three independent copies 
1471: of the fermion sector 
1472: in the Lorentz- and CPT-violating QED extension.
1473: This implies that 
1474: unsuppressed sensitivity to Lorentz violation
1475: in the charged-lepton sector 
1476: involves only flavor-diagonal components.
1477: Moreover, 
1478: the decoupling also implies that 
1479: certain coefficients 
1480: such as $a^\mu_{ee}$, $a^\mu_{\mu\mu}$, $a^\mu_{\ta\ta}$ 
1481: are physically unobservable, 
1482: further reducing the total number of coefficients 
1483: affecting charged leptons.
1484: Taken together, 
1485: these factors ensure that the CPT-odd sectors
1486: of charged leptons and neutrinos 
1487: are completely independent at tree level.
1488: Similar arguments apply to parts of the CPT-even sector as well.
1489: We therefore conclude that
1490: neutrinos are sensitive to a greater number 
1491: of coefficients for Lorentz violation than the charged leptons,
1492: and at tree level most of these coefficients are independent 
1493: from those accessible with $e$, $\mu$ or $\ta$ leptons.
1494: 
1495: Particularly stringent constraints
1496: exist on some components of the charged-lepton coefficients
1497: $b^\mu_{ee}$ and $b^\mu_{\mu\mu}$.
1498: Although these are linearly independent
1499: of neutrino-sector coefficients at tree level,
1500: it is natural to ask whether
1501: radiative corrections to these components
1502: can be used to constrain possible neutrino effects.
1503: As an example,
1504: Ref.\ \cite{fc5} explores the possibility that
1505: eV-size effects in heavy sterile neutrinos 
1506: could evade the constraints in the charged-lepton sector,
1507: finding that within a standard seesaw mechanism
1508: the existence of large $b^\mu$-type coefficients 
1509: for sterile neutrinos tends to produce $b^\mu$ coefficients 
1510: in the charged-lepton sector that conflict with observation.
1511: In this work, 
1512: we neglect seesaw-induced coefficients 
1513: because they are suppressed by the large-mass scale.
1514: However,
1515: it is of interest to ask whether radiative corrections 
1516: alter the tree-level independence of the charged- and
1517: neutral-lepton sectors.
1518: 
1519: For simplicity,
1520: we restrict our discussion 
1521: to the relevant $a^\mu$ and $b^\mu$ coefficients, 
1522: although related remarks 
1523: apply also to $c^\mn$ and $d^\mn$ coefficients.
1524: The leading-order radiative corrections are linear
1525: in the coefficients for Lorentz violation.
1526: However,
1527: loops involving weak-interactions are heavily suppressed
1528: by additional factors at the relevant energies,
1529: while strong interactions play no role. 
1530: We can therefore restrict attention 
1531: to the QED extension.
1532: In this case,
1533: general properties of the coefficients for Lorentz violation
1534: under the discrete symmetries C, P, and T
1535: imply that corrections to $b^\mu$ coefficients 
1536: involve only other $b^\mu$ type coefficients \cite{klap}.
1537: As a result,
1538: although the constraints from charged-lepton experiments 
1539: may restrict $b^\mu$ in the neutrino sector of the SME, 
1540: the $a^\mu$ coefficients are unaffected
1541: and so $a_L$ is unconstrained.
1542: Thus,
1543: the independence of the charged- and neutral-lepton sectors 
1544: remains valid for radiative corrections.
1545: 
1546: 
1547: \subsection{General features}\label{features}
1548: 
1549: In the presence of Lorentz and CPT violation,
1550: a wide range of unconventional neutrino behaviors can occur.
1551: These include unusual energy dependence,
1552: direction-dependent effects, 
1553: and neutrino-antineutrino mixing.
1554: Specific examples of these behaviors
1555: are illustrated in the examples presented 
1556: in Sec.\ \ref{models}.
1557: Here,
1558: we focus on some general features 
1559: of experimental sensitivities to Lorentz- and CPT-violating effects. 
1560: Some of these have been discussed in the context of
1561: the minimal SME in our earlier work
1562: \cite{nu},
1563: but the present discussion holds for the full theory \rf{heff}
1564: and generically for operators of nonrenormalizable dimension.
1565: 
1566: Figure 1 shows an estimate of the coverage 
1567: in baseline distance $L$ versus energy $E$
1568: of the currently published neutrino-oscillation data.
1569: Included in the evidence for oscillations are observations of
1570: solar neutrinos by 
1571: Cl- and Ga-based experiments
1572: \cite{gallex,gno,sage,homestake},
1573: Super Kamiokande (SK) \cite{sksol},
1574: and SNO \cite{sno}; 
1575: and of atmospheric neutrinos by
1576: SK \cite{sk},
1577: reactor-based KamLAND
1578: \cite{kamland},
1579: and accelerator-based
1580: LSND \cite{lsnd} 
1581: and K2K \cite{k2k}.
1582: Null results include the reactor experiments
1583: Bugey \cite{bugey},
1584: CHOOZ \cite{chooz},
1585: G\"osgen \cite{gosgen},
1586: Palo Verde \cite{pv},
1587: and various accelerator-based short-baseline experiments including,
1588: for example, 
1589: the high-energy experiments 
1590: BNL-E776 \cite{e776},
1591: CCFR \cite{ccfr},
1592: CHORUS \cite{chorus},
1593: NOMAD \cite{nomad1,nomad2},
1594: NuTeV \cite{nutev},
1595: and the low-energy
1596: KARMEN \cite{karmen}.
1597: A number of new accelerator-based experiments
1598: are likely to produce interesting results in the near future.
1599: These include the short-baseline 
1600: ($L\simeq 500$ m, $E\simeq 1$ GeV)
1601: MiniBooNE experiment 
1602: \cite{miniboone}
1603: designed to test the LSND anomaly,
1604: and the long-baseline
1605: ($L\simeq 700$ km, $E\simeq 1$ GeV)
1606: ICARUS \cite{icarus},
1607: MINOS \cite{minos},
1608: and OPERA \cite{opera}
1609: experiments,
1610: which are planned to test the atmospheric-oscillation hypothesis.
1611: Also shown on the figure are the approximate effective regions
1612: associated with the matter potentials for the Sun and the Earth.
1613: 
1614: The unusual energy dependence
1615: can be viewed as a consequence of the dimensionality
1616: of the coefficients for Lorentz violation.
1617: The standard scenario for neutrino oscillations
1618: involves mass-squared differences $\De m^2$
1619: that combine with the baseline distance $L$ and the neutrino energy $E$
1620: to yield the physically relevant dimensionless combination
1621: $\De m^2L/E$.
1622: However,
1623: Eq.\ \rf{heff} shows that Lorentz-violating oscillations 
1624: generated by the dimension-one coefficients
1625: $a^\mu$, $b^\mu$, $H^\mn$
1626: are controlled by the dimensionless combinations
1627: $a^\mu L$, $b^\mu L$, $H^\mn L$,
1628: while those generated by
1629: $c^\mn$, $d^\mn$, $g^{\mn\si}$
1630: are controlled by
1631: $c^\mn LE$, $d^\mn LE$, $g^{\mn\si} LE$.
1632: More generally,
1633: oscillations
1634: generated by a coefficient $(k_d)^{\la\ldots}$
1635: for a Lorentz-violating operator
1636: of nonrenormalizable dimension $n=d+3$
1637: are controlled by $(k_d)^{\la\ldots} LE^d$.
1638: 
1639: \begin{figure}
1640: \centerline{
1641: \hspace{2pt}\psfig{figure=expt.eps,width=1.0\hsize}\hspace{2pt}}
1642: \caption{\label{expts}
1643: Approximate sensitivities of various experiments.
1644: Lines of constant $L/E$ (solid),
1645: $L$ (dashed), and $LE$ (dotted)
1646: are shown,
1647: giving approximate sensitivities
1648: to the quantities $\{m, m_5\}$, $\{a^\mu, b^\mu, H^\mn\}$, and
1649: $\{c^\mn, d^\mn, g^{\mn\si}\}$,
1650: respectively.
1651: Also shown are the approximate effective regions 
1652: for the matter potential in the Sun and Earth.}
1653: \end{figure}
1654: 
1655: Figure 1 illustrates these various energy dependences.
1656: Lines of constant $L/E$, $L$, and $LE$ are plotted,
1657: bounding approximate regions of experimental sensitivity to 
1658: conventional mass-squared differences, 
1659: dimension-one coefficients,
1660: and dimensionless coefficients,
1661: respectively.
1662: For each nonzero coefficient in $\heff$,
1663: a bounding line on this figure exists
1664: above which the corresponding Lorentz-violating effects 
1665: become of order one.
1666: Given such a line,
1667: any experiments located near or above it 
1668: can be affected by the associated coefficient,
1669: but experiments below it have limited or no sensitivity.
1670: For example,
1671: the region of limiting sensitivity
1672: for a hypothetical dimensionless coefficient
1673: of magnitude $\sim 10^{-18}$
1674: is bounded approximately 
1675: by the dimensionless line satisfying $LE=10^{18}$,
1676: which is the dotted line running just below KamLAND.
1677: Experiments lying above this line,
1678: such as KamLAND, SNO, and SK,
1679: could be sensitive to the effects of this coefficient.
1680: Note that 
1681: approximate regions of experimental sensitivity to 
1682: coefficients $(k_d)^{\la\ldots}$ of dimension $1-d$
1683: could also be identified on the figure.
1684: They would be bounded by lines of constant $LE^d$ with $d>1$,
1685: which have negative-integer slopes. 
1686: 
1687: Figure 1 also reveals that experiments and data 
1688: allow probes well below the $10^{-17}$ Planck-suppression level.
1689: For instance,
1690: the various null results 
1691: from short-baseline reactor and accelerator experiments
1692: could be reanalysed to yield upper bounds on  
1693: certain coefficients for Lorentz violation.
1694: Thus,
1695: the high-energy experiments CHORUS and NOMAD 
1696: found no evidence of $\nu_{e,\mu}\to\nu_\ta$
1697: at energies $E\sim 100$ GeV 
1698: and at distances $L\sim 10^{18}$ GeV$^{-1}$,
1699: which suggests that reanalyses of these experiments 
1700: would yield interesting new sensitivities 
1701: of roughly $10^{-18}$ GeV to dimension-one coefficients 
1702: and roughly $10^{-20}$ to dimensionless coefficients.
1703: A similar situation holds for low-energy experiments 
1704: such as CHOOZ, Palo Verde, and KARMEN 
1705: in the $\nub_e$ sector.
1706: From Fig.\ \ref{expts} we see that,
1707: relative to CHORUS and NOMAD,
1708: CHOOZ and Palo Verde might be expected 
1709: to have comparable sensitivities 
1710: to dimension-one coefficients
1711: but reduced sensitivity to dimensionless ones,
1712: while KARMEN has comparable sensitivity 
1713: to conventional mass effects.
1714: In each case,
1715: the attainable sensitivities also depend 
1716: on various experiment-dependent factors, 
1717: so individual reanalyses are required to make definitive statements.
1718: 
1719: Another unusual effect due to Lorentz violation
1720: is direction-dependent neutrino behavior,
1721: a consequence of rotation-symmetry violation.
1722: This has consequences for comparisons of results
1723: between different terrestrial experiments
1724: or for the analysis of experiments involving multiple sources,
1725: since the orientation of the neutrino beam 
1726: or the location of the source relative to the detector
1727: can affect neutrino oscillations.
1728: Rotation-symmetry violation also implies that the
1729: daily rotation of the Earth about its axis induces 
1730: apparent periodic changes of the coefficients for Lorentz violation
1731: in the laboratory,
1732: which would be manifest as temporal variations 
1733: in neutrino oscillations. 
1734: These variations occur at multiples of the sidereal frequency
1735: $\om_\oplus\simeq 2\pi/$(23 h 56 min).
1736: Similarly,
1737: in the presence of rotation-symmetry violation,
1738: neutrinos emitted from the Sun in different directions 
1739: undergo different oscillations,
1740: which may produce observable annual variations 
1741: arising from the change in the location of the detector
1742: as the Earth orbits the Sun.
1743: All these temporal variations with appropriate periodicity 
1744: provide unique signals of Lorentz violation
1745: in neutrino oscillations.
1746: Moreover,
1747: they can also yield interesting sensitivities to certain coefficients.
1748: For instance, 
1749: SK found that the shape of the solar-neutrino flux 
1750: matches the expected value to within about 5\% over the year
1751: \cite{sksol}.
1752: The Sun-Earth distance is $L\sim 10^{27}$ GeV$^{-1}$,
1753: and $LE\sim 10^{25}$ for the SK energy range,
1754: so a reanalysis of the SK data might achieve
1755: impressive sensitivities of $\sim 10^{-28}$ GeV 
1756: to dimension-one coefficients
1757: and $\sim 10^{-26}$ to dimensionless ones,
1758: comparable to the best experimental sensitivities achieved
1759: for other sectors of the SME.
1760: 
1761: Another interesting feature of Lorentz violation
1762: involves novel resonance effects in neutrino oscillations.
1763: In the conventional case with neutrino masses,
1764: the usual MSW resonances 
1765: \cite{msw}
1766: arise when the local matter environment is such that
1767: neutrino interactions become comparable to mass effects,
1768: thereby drastically changing the character of the hamiltonian.
1769: The presence of Lorentz violation 
1770: can trigger several other types of effects,
1771: including resonances without mass or matter
1772: that involve different coefficients for Lorentz violation,  
1773: resonances involving coefficients for Lorentz violation
1774: and mass terms,  
1775: resonances involving coefficients for Lorentz violation 
1776: and matter effects,  
1777: and various combinations of the above.
1778: The earliest example 
1779: of an explicit vacuum resonance in a two-generation model 
1780: involving a mass term and a single nonzero coefficient $(a_L)^T$ 
1781: for Lorentz and CPT violation is given in Ref.\ \cite{fc2}.
1782: An example of a vacuum resonance 
1783: in a three-generation model 
1784: involving two coefficients $(a_L)^Z$ and $(c_L)^{TT}$ 
1785: for Lorentz and CPT violation 
1786: occurs in the bicycle model of Ref.\ \cite{nu}.
1787: We emphasize that resonances due to Lorentz violation
1788: can occur in the vacuum as well as in matter,
1789: and not only at particular energies 
1790: but also for particular directions of propagation.
1791: Note also that,
1792: even away from the resonance regions,
1793: matter effects may be important
1794: when considering mass terms or coefficients for Lorentz violation 
1795: that have lines of sensitivity near or above 
1796: the Sun- or Earth-potential regions shown in Figure \ref{expts}.
1797: 
1798: \subsection{The LSND anomaly}\label{lsnd}
1799: 
1800: In the LSND experiment \cite{lsnd},
1801: copious numbers of neutrinos were
1802: produced from the decay of $\pi^+$ at rest.
1803: This process is dominated by the decay
1804: $\pi^+\to\mu^+\nu_\mu$
1805: followed by $\mu^+\to e^+\nu_e\nub_\mu$.
1806: A small excess in $\nub_e$ was seen,
1807: interpreted as the oscillation
1808: $\nub_\mu\to\nub_e$
1809: with a small probability of about 0.26\%.
1810: This result is difficult to accommodate 
1811: within the context of the conventional global analysis
1812: \cite{pdg},
1813: in which two mass-squared differences
1814: are used to describe solar and atmospheric oscillation data.
1815: The solar data appear consistent 
1816: with a mass-squared difference $\de m^2 \sim 10^{-5}$ eV$^2$,
1817: while the atmospheric data
1818: suggest a second mass-squared difference 
1819: $\De m^2 \sim 10^{-3}$ eV$^2$.
1820: The regions of limiting sensitivity to 
1821: these mass-squared differences 
1822: are shown in Fig.\ \ref{expts}, 
1823: where lines of constant $L/E$ with values
1824: $L/E\sim 10^{23}$ GeV$^{-2}$ and
1825: $L/E\sim 10^{21}$ GeV$^{-2}$
1826: can be seen.
1827: Experiments lying significantly below these lines, 
1828: including LSND, 
1829: should be insensitive to oscillations 
1830: caused by $\de m^2$ and $\De m^2$.
1831: This illustrates the difficulty in explaining the LSND result
1832: within the conventional framework
1833: without introducing additional mass-squared differences.
1834: 
1835: A resolution of this LSND anomaly
1836: without the introduction of sterile neutrinos 
1837: might emerge from 
1838: the unusual energy dependence,
1839: the directional dependence,
1840: or the neutrino-antineutrino mixing
1841: introduced by Lorentz violation.
1842: For example,
1843: equal numbers of $\nu_\mu$, $\nu_e$, and $\nub_\mu$
1844: are produced in LSND,
1845: so if $\nu_e$ mix with $\nub_e$
1846: then the observed excess in $\nub_e$
1847: may be a result of $\nu_e\mix\nub_e$ mixing
1848: rather than $\nub_\mu\mix\nub_e$ mixing.
1849: We note, 
1850: however,
1851: that if the possible direction dependence is neglected
1852: then Fig.\ \ref{expts} shows that a simple solution based
1853: either on the unusual energy dependence
1854: or on $\nu\mix\nub$ mixing
1855: is likely to be hindered by existing null results
1856: in the $\nub_e$ sector,
1857: from low-energy experiments such as CHOOZ and Palo Verde
1858: or from high-energy experiments such as CHORUS, NOMAD, and NuTeV.
1859: Indeed,
1860: from this figure we see generically that 
1861: to explain the LSND result 
1862: one needs a mass-squared difference of about 
1863: $10^{-19}$ GeV$^2 = 10^{-1}$ eV$^2$,
1864: a dimension-one coefficient of about $10^{-18}$ GeV
1865: or a dimensionless coefficient of about $10^{-17}$.
1866: Note that each of these has consequences for other experiments,
1867: depending on flavor content.
1868: For example, 
1869: the upcoming MiniBooNE experiment is designed to test 
1870: the same oscillation channel 
1871: and will therefore be sensitive to all three possibilities.
1872: 
1873: 
1874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1875: 
1876: \section{Illustrative Models}\label{models}
1877: 
1878: To illustrate some of the novel behaviors
1879: of neutrino oscillations in the presence of Lorentz and CPT violation,
1880: we next consider a number of simple special cases
1881: of the theory \rf{heff}
1882: with only one or a few nonzero coefficients.
1883: For each case,
1884: some of the ways that the unusual neutrino behaviors
1885: might affect current observations are quantitatively examined.
1886: Also,
1887: we simplify expressions by adopting temporary notation
1888: for the specific nonzero coefficients for Lorentz violation
1889: within each model:
1890: quantities carrying a ring accent, 
1891: such as $\ring{c}$,
1892: denote rotation-symmetric coefficients;
1893: while those with a h\'a\v{c}ek accent,
1894: such as $\check{c}$, 
1895: denote anisotropic coefficients.
1896: 
1897: \subsection{Rotationally invariant models}\label{rimods}
1898: 
1899: The rotation-invariant restriction
1900: provides an interesting special limit of the theory \rf{heff}.
1901: While difficult to motivate without knowledge 
1902: of the underlying mechanism leading to Lorentz and CPT violation,
1903: rotation-invariant or so-called `fried-chicken' (FC) models
1904: are attractive because 
1905: rotation symmetry can significantly reduce 
1906: the complexity of calculations,
1907: thereby providing a simple context 
1908: within which to study the unusual neutrino behaviors 
1909: arising from Lorentz violation.
1910: 
1911: Restricting $\heff$ to FC terms leaves only four matrices,
1912: $(\mt)_{ab}$,
1913: $(a_L)^0_{ab}$,
1914: $(c_L)^{00}_{ab}$,
1915: and
1916: $(c_L)^{jk}_{ab}=\fr13(c_L)^{ll}_{ab}\de^{jk}$.
1917: As described in Sec.\ \ref{phenom},
1918: the trace $(c_L)^{00}_{ab}-(c_L)^{jj}_{ab}$ is
1919: unobservable and may be set to zero,
1920: so only three of these matrices are independent.
1921: Dropping the irrelevant kinetic term
1922: and assuming rotation invariance in the Sun-centered $(T,X,Y,Z)$ frame
1923: for definiteness,
1924: the $6\times6$ effective hamiltonian reduces 
1925: to the block-diagonal form
1926: \bea
1927: (\heff)^{\rm FC}_{ab}&=&{\rm diag}
1928: \big[\big(
1929: \mt/(2E)
1930: +(a_L)^T
1931: -\frac43(c_L)^{TT}E
1932: \big)_{ab}\ ,
1933: \nonumber \\
1934: &&
1935: \quad\big(
1936: \mt/(2E)
1937: -(a_L)^T
1938: -\frac43(c_L)^{TT}E
1939: \big)^*_{ab}\big] .
1940: \label{rim}
1941: \eea
1942: This hamiltonian provides a general FC model
1943: of three active neutrinos.
1944: The generalization to additional light 
1945: or massless sterile neutrinos is straightforward.
1946: 
1947: With the exception of the original proposal 
1948: for Lorentz violation in neutrinos \cite{ck}
1949: and the recent work in Ref.\ \cite{nu},
1950: which address both rotation-invariant and anisotropic effects 
1951: with and without CPT violation,
1952: existing works on the subject 
1953: \cite{fc1,fc2,fc3,fc4}
1954: involve limited special cases of the general FC model \rf{rim}.
1955: The bulk of the literature
1956: restricts attention to the two-generation special case
1957: and neglects either the $(a_L)^T$ term or the $(c_L)^{TT}$ term.
1958: A plethora of unexplored models and effects exists.
1959: 
1960: It might seem logical to impose spherical symmetry 
1961: in a special frame 
1962: such as the cosmic microwave background (CMB) frame.
1963: However,
1964: if rotation symmetry is assumed in the CMB frame
1965: then the coefficients in Eq.\ \rf{rim} 
1966: differ from $(a_L)^T$, $(c_L)^{TT}$ 
1967: in the standard Sun-centered frame,
1968: being instead $(a_L)^0$, $(c_L)^{00}$ in the CMB frame.
1969: Relating the latter to the standard Sun-centered frame 
1970: or any other experimentally attainable frame
1971: introduces direction dependence 
1972: due to the motion of our solar system in the CMB frame.
1973: The relevant hamiltonian then also involves spatial
1974: components of the coefficients,
1975: so it differs from Eq.\ \rf{rim} 
1976: and is instead an anisotropic limit of the theory \rf{heff}. 
1977: 
1978: Although the FC model \rf{rim} is rather limited 
1979: considering the wealth of possible effects 
1980: contained in the full theory \rf{heff},
1981: and although it has little theoretical motivation
1982: other than calculational convenience,
1983: further study of this model is useful 
1984: because it provides a readily workable context 
1985: within which to gain insight 
1986: about possible signals of Lorentz and CPT violation.
1987: This is illustrated in the few simple examples 
1988: discussed in this subsection.
1989: 
1990: 
1991: \subsubsection{Example: $(c_L)^{TT}_{ab} \ne 0$}\label{dcmod}
1992: 
1993: \begin{figure}
1994: \centerline{
1995: \psfig{figure=expt_dmdc.eps,width=1.0\hsize}
1996: \begin{picture}(0,0)(20,0)
1997: \put(-40,152){$\de m^2$}
1998: \put(-40,127){$\De m^2$}
1999: \put(-40,86){$\de \cri$}
2000: \put(-60,30){$\De \cri$}
2001: \end{picture}}
2002: \caption{\label{expts2}
2003: Lines of limiting sensitivity for
2004: $\de m^2 \sim 10^{-5}$ eV$^2$,
2005: $\De m^2 \sim 10^{-3}$ eV$^2$,
2006: $\de\cri \sim 10^{-22}$,
2007: $\De\cri \sim 10^{-17}$.
2008: The shaded regions are those of Fig.\ 1.}
2009: \end{figure}
2010: 
2011: A particularly simple FC model consists of a single
2012: nonzero coefficient matrix such as $(c_L)^{TT}_{ab}$.
2013: Some features of this model
2014: are similar to the conventional massive-neutrino case,
2015: but there is unusual energy dependence.
2016: Here,
2017: we take advantage of this energy dependence
2018: to illustrate one type of mechanism through which
2019: Lorentz violation might provide a solution to the LSND anomaly.
2020: 
2021: Lines of limiting sensitivity for the two mass-squared differences 
2022: $\de m^2$ and $\De m^2$ 
2023: used in the conventional global analysis 
2024: are shown in Fig.\ \ref{expts2}.
2025: The mixing angles are such that
2026: $\nu_e$ oscillations are almost completely controlled by $\de m^2$.
2027: Therefore, 
2028: one can expect to see only $\nub_e$ mixing in KamLAND,
2029: in solar neutrino experiments,
2030: and possibly in the lowest-energy atmospheric-neutrino experiments.
2031: The observed atmospheric oscillations are due to $\De m^2$,
2032: which controls $\nu_\mu\mix\nu_\ta$ mixing.
2033: Since LSND lies well below both the $\de m^2$ and the $\De m^2$ lines,
2034: no oscillations are predicted.
2035: 
2036: Replacing the mass-squared differences $\de m^2$ and $\De m^2$
2037: with a nonzero coefficient matrix $(c_L)^{TT}_{ab}$ 
2038: produces an effective hamiltonian $\heff$ that can be parametrized 
2039: as described in Appendix \ref{smeterms},
2040: using two eigenvalue differences 
2041: and CKM-like mixing angles and phases.
2042: For simplicity, 
2043: we choose here to mimic the usual solution 
2044: by taking vanishing phases and $\th^{TT}_{13}$,
2045: and we consider only the case $\th^{TT}_{23}=\pi/4$.
2046: This leaves three degrees of freedom:
2047: two eigenvalue differences,
2048: and one mixing angle $\th^{TT}_{12}\equiv\th$.
2049: It turns out to be convenient to work with 
2050: two linear combinations of the eigenvalue differences,
2051: defined by 
2052: \bea
2053: \de \cri&\equiv&\frac43\big((c_L)^{TT}_{(3)}-(c_L)^{TT}_{(2)}\big) ,
2054: \nonumber \\
2055: \De \cri&\equiv&\frac43\big((c_L)^{TT}_{(2)}-(c_L)^{TT}_{(1)}\big) .
2056: \eea
2057: The probabilities for this case are then
2058: \bea
2059: P_{\nu_e\to\nu_e}&=&
2060: 1-\sin^22\th\sin^2(\De\cri LE/2), \nonumber\\
2061: P_{\nu_\mu\to\nu_\mu}&=&P_{\nu_\ta\to\nu_\ta}
2062: =1-\frac14\sin^22\th\sin^2(\De\cri LE/2) \nonumber\\
2063: &&-\sin^2\th\sin^2((\De\cri +\de\cri)LE/2) \nonumber\\
2064: &&-\cos^2\th\sin^2(\de\cri LE/2),  \nonumber\\
2065: P_{\nu_e\mix\nu_\mu}&=&P_{\nu_e\mix\nu_\ta}
2066: =\half\sin^22\th\sin^2(\De\cri LE/2),  \nonumber\\
2067: P_{\nu_\mu\mix\nu_\ta}
2068: &=&-\frac14\sin^22\th\sin^2(\De\cri LE/2) \nonumber\\
2069: &&+\sin^2\th\sin^2((\De\cri +\de\cri)LE/2) \nonumber\\
2070: &&+\cos^2\th\sin^2(\de\cri LE/2) .
2071: \label{probsm2}
2072: \eea
2073: The corresponding antineutrino expressions
2074: are identical.
2075: 
2076: A possible approach is illustrated in the figure.
2077: The line of sensitivity
2078: for the larger difference $\De\cri$ 
2079: can be chosen to lie just above CHOOZ and LSND.
2080: This produces only a small effect in these experiments 
2081: and may provide an explanation
2082: for LSND that may not conflict with CHOOZ.
2083: The remaining difference $\de\cri$ can
2084: then be chosen to explain atmospheric data.
2085: The above situation somewhat resembles the conventional mass solution,
2086: with the role of $\de m^2/2E$ replaced by $\De\cri E$ 
2087: and that of $\De m^2/2E$ replaced by $\de\cri E$.
2088: The angle $\th$ is the analogue of
2089: the solar-neutrino mixing angle.
2090: However,
2091: the energy dependences of the two cases differ substantially,
2092: as is also evident from the figure.
2093: 
2094: To explore quantitatively how this approach might work,
2095: consider the positive LSND and KamLAND results.
2096: KamLAND detects $\nub_e$ from distant reactors
2097: and found about a $61\%$ reduction in the flux.
2098: Most reactors are 138-214 km from the detector,
2099: and the corresponding $\nub_e$ energies 
2100: fall in the range 1 MeV $\lsim E \lsim$ 10 MeV.
2101: If KamLAND lies well above the $\De\cri$ line,
2102: the relevant quantity is the average survival probability
2103: $\vev{P_{\nub_e\to\nub_e}} =1-\half\sin^22\th\simeq 61\%$,
2104: yielding a mixing angle given by $\sin^22\th \simeq 0.78$.
2105: Also,
2106: assuming LSND is in a region of small oscillation effects,
2107: then we can approximate
2108: $P_{\nub_\mu\to\nub_e}\approx
2109: \half\sin^22\th(\De\cri LE/2)^2 \simeq 0.26\%$.
2110: Then,
2111: for $E\simeq 45$ MeV and $L\simeq 30$ m
2112: we obtain $\De\cri \simeq 2.4\times 10^{-17}$.
2113: Thus, 
2114: in this simple scenario,
2115: these two experiments suggest coefficient values 
2116: near $\sin^22\th \simeq 0.78$ and
2117: $\De\cri \simeq 2.4\times 10^{-17}$,
2118: in agreement with the estimates of Sec.\ \ref{lsnd}.
2119: 
2120: The remaining coefficient $\de\cri$ can then be chosen 
2121: to match observed atmospheric-neutrino effects.
2122: The coefficient $\De\cri$ is relatively large in this region
2123: and generates rapid oscillations.
2124: Averaging over these for any value of $\de\cri$
2125: leaves a muon-neutrino survival probability of
2126: either
2127: $P_{\nu_\mu\to\nu_\mu}\simeq 0.54-0.27\sin^2(\de\cri LE/2)$
2128: or
2129: $P_{\nu_\mu\to\nu_\mu}\simeq 0.77-0.73\sin^2(\de\cri LE/2)$,
2130: depending on the solution for $\th$. 
2131: Note that the latter expression resembles 
2132: the usual maximal-mixing solution within an overall scale factor,
2133: except for the unusual energy dependence in the oscillation length.
2134: 
2135: Interestingly,
2136: atmospheric electron-neutrino oscillations
2137: are present in this model but are largely unobserved
2138: due to a compensation mechanism.
2139: The averaged $\nu_e$ survival probability is 
2140: $P_{\nu_e\to\nu_e}=61\%$,
2141: as above,
2142: and the $\nu_e\mix\nu_\mu$ mixing probability
2143: is $P_{\nu_e\mix\nu_\mu}=19.5\%$.
2144: The observed flux of atmospheric electron neutrinos
2145: is a combination of the survival flux
2146: and the appearance flux from mixing with muon neutrinos.
2147: Since the ratio of muon neutrinos to electron neutrinos 
2148: is approximately 2, 
2149: the predicted effective flux of atmospheric electron neutrinos
2150: is approximately $61\% + 2(19.5\%) \simeq 100\%$ of the flux
2151: in the absence of oscillations,
2152: in agreement with indications from existing data.
2153: Essentially,
2154: this compensation mechanism works
2155: because the disappearance probability 
2156: $1 - P_{\nu_e\to\nu_e}$
2157: of electron neutrinos given by Eq.\ \rf{probsm2}
2158: is a factor of two greater than 
2159: the appearance probability $P_{\nu_e\mix\nu_\mu}$
2160: of muon neutrinos from mixing,
2161: resulting in no net suppression 
2162: in the total observed electron-neutrino flux.
2163: 
2164: The compensation mechanism 
2165: \it per se \rm
2166: is independent of Lorentz violation
2167: and can be applied whenever 
2168: $1 - P_{\nu_e\to\nu_e}\approx 2P_{\nu_e\mix\nu_\mu}$,
2169: including in the conventional massive case.
2170: Note,
2171: however,
2172: that Monte Carlo calculations suggest the flux ratio 
2173: increases dramatically above 2 for energies 
2174: over about 10 GeV \cite{mc},
2175: so the compensation mechanism is likely to fail at higher energies.
2176: Note also that,
2177: in the case of the above Lorentz-violating model,
2178: the rapid oscillations at high energies 
2179: also help to mask $\nu_e$ oscillations.
2180: Although these rapid oscillations can change the overall flux, 
2181: they also tend to smooth away the observable $E$ and $L$ dependences
2182: that form the basis for some analyses. 
2183: 
2184: This simple model serves to illustrate a possible strategy
2185: that might remedy the conflict
2186: between LSND and reactor experiments,
2187: but it may well introduce other conflicts 
2188: between LSND and accelerator experiments testing 
2189: $\nu_e\to\nu_\ta$ and $\nu_\mu\to\nu_\ta$
2190: \cite{chorus,nomad1}
2191: or $\nu_\mu\to\nu_e$
2192: \cite{nutev,nomad2}.
2193: Note also that some work has been done 
2194: to check for unconventional energy dependences
2195: in the atmospheric data 
2196: \cite{efit},
2197: suggesting that the usual energy dependence is preferred.
2198: However, 
2199: these analyses are limited to two generations 
2200: and do not consider possible direction dependences 
2201: or $\nu\mix\nub$ mixing.
2202: A complete treatment would also need to include
2203: the effects of the Earth's matter potential,
2204: which introduces additional energy dependence.
2205: The point is that $G_Fn_e \sim 10^{-22}$ GeV for the Earth, 
2206: and at atmospheric-neutrino energies
2207: this is comparable to the contribution from $\de\cri$
2208: shown in Fig.\ \ref{expts2}.
2209: In any case,
2210: interesting sensitivities to Lorentz violation
2211: could be achieved with a complete analysis of existing data. 
2212: 
2213: 
2214: \subsubsection{Example:
2215:   $(a_L)^T_{e\mu}\ne 0$, $(c_L)^{TT}_{\mu\ta}\ne 0$}
2216: \label{acmod}
2217: 
2218: We turn next to an FC model with mixed energy dependence,
2219: incorporating only two nonzero coefficients 
2220: $(a_L)^T_{e\mu}\equiv \aem$
2221: and
2222: $\frac43(c_L)^{TT}_{\mu\ta}\equiv \cmt$
2223: and no mass terms.
2224: This case includes both Lorentz and CPT violation
2225: but remains rotation symmetric.
2226: The presence of both a dimensionless coefficient
2227: and a dimension-one coefficient
2228: leads to unusual energy behavior 
2229: in the vacuum-mixing angles as well as the oscillation lengths.
2230: This contrasts with the previous case,
2231: in which only the oscillation lengths have unconventional 
2232: energy dependence.
2233: Note that both $\aem$ and $\cmt$ are arbitrary 
2234: to an unobservable phase,
2235: and therefore they can be taken real and nonnegative 
2236: without loss of generality.
2237: 
2238: The behavior in this model can be understood qualitatively as follows.
2239: At low energies $E\ll \aem / \cmt$ 
2240: relative to the critical energy $\aem / \cmt$,
2241: the $\aem$ term dominates the effective hamiltonian.
2242: As a result, 
2243: $\nu_\ta$ decouples from $\nu_e$ and $\nu_\mu$,
2244: so only $\nu_e\mix\nu_\mu$ mixing occurs.
2245: In contrast,
2246: for high energies $E\gg \aem / \cmt$,
2247: $\cmt$ dominates and only $\nu_\mu\mix\nu_\ta$ mixing occurs.
2248: At intermediate energies $E\sim \aem / \cmt$,
2249: the two terms are comparable
2250: and produce complicated energy dependence 
2251: with mixing between all three neutrinos.
2252: 
2253: This behavior is similar to the observed energy dependence
2254: in the solar-neutrino flux.
2255: In the usual analysis with massive neutrinos,
2256: the observed energy dependence is explained through matter effects.
2257: However,
2258: the same type of behavior can appear in Lorentz-violating scenarios 
2259: even without matter.
2260: To demonstrate this,
2261: we need the probabilities for the current model:
2262: \bse\bea
2263: P_{\nu_e\to\nu_e}&=&
2264: 1-4\sin^2\th \cos^2\th\sin^2(\pi L/L_0) \nonumber \\
2265: &&-\sin^4\th\sin^2(2\pi L/L_0) , \label{pee} \\
2266: P_{\nu_\mu\to\nu_\mu}&=&
2267: 1-\sin^2(2\pi L/L_0) , \label{pmm} \\
2268: P_{\nu_\ta\to\nu_\ta}&=&
2269: 1-4\sin^2\th \cos^2\th\sin^2(\pi L/L_0) \nonumber \\
2270: &&-\cos^4\th\sin^2(2\pi L/L_0) , \label{ptt} \\
2271: P_{\nu_e\mix\nu_\mu}&=&
2272: \sin^2\th\sin^2(2\pi L/L_0) , \label{pem} \\
2273: P_{\nu_e\mix\nu_\ta}&=&
2274: \sin^2\th \cos^2\th\big(4\sin^2(\pi L/L_0)\nonumber\\
2275: &&-\sin^2(2\pi L/L_0)\big) , \label{pet} \\
2276: P_{\nu_\mu\mix\nu_\ta}&=&
2277: \cos^2\th\sin^2(2\pi L/L_0) , \label{pmt}
2278: \eea\label{p}\ese
2279: where
2280: \bea
2281: \sin^2\th&=&\aem^2/(\aem^2+\cmt^2E^2) ,
2282: \nonumber\\
2283: 2\pi/L_0&=&\sqrt{\aem^2+\cmt^2E^2} .
2284: \eea
2285: The antineutrino probabilities are
2286: again identical since the quantities
2287: $\sin^2\th$ and $L_0$ are symmetric
2288: under $\aem \to -\aem$.
2289: We remark in passing that this model serves as an example 
2290: in which CPT is violated 
2291: but the traditional test of CPT discussed in Sec.\ \ref{cptprop} 
2292: fails as an indicator of the CPT violation.
2293: 
2294: The solar-neutrino vacuum-oscillation survival probability 
2295: is given by Eq.\ \rf{pee}.
2296: As usual,
2297: depending on the size of the coefficients,
2298: matter effects can drastically alter the survival rates.
2299: Consider, 
2300: for example, 
2301: a simple matter-dominated case where the matter potential 
2302: at the point of $\nu_e$ production dominates $\heff$.
2303: Assuming adiabatic propagation,
2304: neutrinos are produced 
2305: in the highest-eigenvalue state of $\heff(R\simeq 0)$
2306: and emerge from the Sun 
2307: in the highest-eigenvalue state of $\heff(R= R_\odot)$.
2308: The overlap between this state and an electron-neutrino state
2309: is proportional to $\sin\th/\sqrt{2}$.
2310: Consequently,
2311: the average survival probability
2312: for the matter-dominated case in an adiabatic approximation is
2313: \beq
2314: \vev{P_{\nu_e\to\nu_e}}_{\rm adiabatic}
2315: =\half\sin^2\th .
2316: \label{adiabatic}
2317: \eeq
2318: In contrast, 
2319: the average for the case where matter effects can be neglected is 
2320: \beq
2321: \vev{P_{\nu_e\to\nu_e}}_{\rm vacuum}
2322: =1-2\sin^2\th+\frac32\sin^4\th.
2323: \label{vacuum}
2324: \eeq
2325: These probabilities are plotted on Fig.\ \ref{prob} 
2326: as a function of energy in units of $\aem/\cmt$.
2327: 
2328: \begin{figure}
2329: $\amt E/\mem^2$
2330: \centerline{
2331: \psfig{figure=fig_ac2.eps,width=0.9\hsize}}
2332: $\cmt E/\aem$
2333: \caption{\label{prob}
2334: Solar-neutrino survival
2335: probability assuming adiabatic
2336: propagation (solid),
2337: and average survival probability
2338: for vacuum oscillations (dashed).}
2339: \end{figure}
2340: 
2341: The observed flux is consistent with the figure,
2342: since low-energy experiments suggest 
2343: an approximate survival probability of 1/2
2344: \cite{gno,gallex,sage},
2345: while higher-energy experiments favor about 1/3
2346: \cite{homestake,sno,sksol}.
2347: Note that both cases shown in Fig.\ \ref{prob} 
2348: yield an average survival probability of 1/3
2349: at $E=\aem/\sqrt{2}\cmt$.
2350: By choosing the ratio $\aem/\cmt$ to coincide
2351: with the peak of the solar $^8$B spectrum
2352: ($E_{\rm peak}\simeq 6.4$ MeV),
2353: this simple massless Lorentz- and CPT-violating model 
2354: can be made to reproduce the gross features 
2355: of the observed solar-neutrino flux.
2356: This corresponds to imposing $\aem/\cmt\simeq 9$ MeV.
2357: 
2358: The above discussion only depends on the ratio of coefficients.
2359: To get a sense of the size of coefficients required in a realistic case,
2360: we can consider what KamLAND implies for $\aem$ and $\cmt$.
2361: Taking a representative neutrino 
2362: to have energy $E=5$ MeV and baseline $L=200$ km
2363: and assuming that it oscillates no more than once,
2364: the ratio $\aem/\cmt\simeq 9$ MeV
2365: and the survival probability $P_{\nub_e\to\nub_e}\simeq 61\%$ 
2366: can be used to extract approximate values
2367: $\aem\simeq 7\times 10^{-22}$ GeV
2368: and
2369: $\cmt\simeq 8\times 10^{-20}$.
2370: The lines of sensitivity for these values on Fig.\ \ref{expts}
2371: are approximately $L\sim 10^{21}$ GeV$^{-1}$ and $LE\sim 10^{19}$,
2372: passing just above KamLAND and intersecting in the solar-energy region,
2373: thereby producing the energy dependence seen in Fig.\ \ref{prob}.
2374: 
2375: 
2376: \subsubsection{Example: 
2377:   $(\mt)_{e\mu}\ne 0$, $(a_L)^{TT}_{\mu\ta}\ne 0$}\label{mamod}
2378: 
2379: As a variation on the above model,
2380: we next consider a special FC case with nonzero mass
2381: $(\mt)_{e\mu}\equiv\mem^2$ and coefficient
2382: $(a_L)^{TT}_{\mu\ta}\equiv \amt$ for Lorentz and CPT violation.
2383: This model has many qualitative features of the previous one.
2384: At small energies,
2385: the mass $\mem$ controls mixing between $\nu_e$ and $\nu_\mu$,
2386: while at large energies 
2387: $\amt$ dominates and produces
2388: mixing between $\nu_\mu$ and $\nu_\ta$.
2389: 
2390: The probabilities for this model are given by Eqs.\ \rf{p},
2391: \rf{adiabatic}, and \rf{vacuum},
2392: but with the definitions
2393: \bea
2394: \sin^2\th&=&\mem^4/(\mem^4+4\amt^2 E^2) ,
2395: \nonumber \\
2396: 2\pi/L_0&=&\sqrt{(\mem^2/2E)^2+\amt^2} .
2397: \eea
2398: The analysis of this model parallels the previous case.
2399: Indeed,
2400: Fig.\ \ref{prob} also holds
2401: for the solar-neutrino probabilities 
2402: in terms of $\mem$ and $\amt$,
2403: using the scale shown on the top axis.
2404: Applying the same arguments as before yields
2405: the ratio $\mem^2/\amt\simeq 18$ MeV
2406: and candidate values
2407: $\mem^2\simeq 7\times 10^{-6}$ eV$^2$ and
2408: $\amt\simeq 4\times 10^{-22}$ GeV.
2409: 
2410: A key difference between this case
2411: and the previous $\aem$-$\cmt$ model 
2412: is the asymptotic behavior of the oscillation length.
2413: In the $\aem$-$\cmt$ case,
2414: $L_0\to 2\pi/(\cmt E)$ at high energies.
2415: In contrast, 
2416: the oscillation length in the present $\mem$-$\amt$ model
2417: approaches a constant at high energies,
2418: $L_0\to 2\pi/\amt$.
2419: Consider the consequences for atmospheric neutrinos.
2420: Note that in the high-energy limit of both cases, 
2421: $\sin^2\th\to0$ and so $P_{\nu_e\to\nu_e}\to0$, 
2422: in agreement with observation.
2423: However,
2424: the first model with $\cmt\simeq 8\times 10^{-20}$
2425: gives $L_0\simeq 2\pi/(\cmt E) \simeq (15$ km GeV$)/E$,
2426: whereas the second model with $\amt\simeq 4\times 10^{-22}$ GeV
2427: yields $L_0\simeq 3100$ km.
2428: These differ from the usual massive-neutrino explanation 
2429: of the atmospheric data,
2430: which has $\De m^2\simeq 3\times 10^{-3}$ eV$^2$
2431: and results in $L_0=4\pi E/\De m^2\simeq 800E$ km/GeV.
2432: 
2433: We emphasize that both this special model and the previous one 
2434: involve only two degrees of freedom,
2435: whereas the usual massive-neutrino solution requires 
2436: two mass-squared differences and at least two mixing angles.
2437: Including additional coefficients for Lorentz violation  
2438: can only add flexibility to the analysis.
2439: For example, 
2440: one might consider a combination of the two examples above,
2441: which would have four degrees of freedom.
2442: With additional freedom,
2443: it seems likely that an appropriate simple Lorentz-violating scenario
2444: could be constructed that would reproduce most oscillation data.
2445: This also suggests that existing data analyses
2446: appear insufficient to exclude many forms of Lorentz and CPT violation,
2447: or even to distinguish between oscillations due to mass 
2448: and those due to Lorentz violation.
2449: 
2450: 
2451: 
2452: \subsection{Direction-dependent and
2453: $\nu\mix\nub$ mixing models}\label{dirmods}
2454: 
2455: Lorentz violation naturally allows directional dependence 
2456: in oscillation parameters through the violation of
2457: rotation invariance.
2458: An interesting subset of direction-dependent models are
2459: those involving $\nu\mix\nub$ mixing 
2460: via nonzero $g^{\mn\si}$ and $H^\mn$ coefficients
2461: in the theory \rf{heff}.
2462: In the general case,
2463: nonzero $\nu\mix\nub$ mixing represents one way
2464: to generate as many as five distinct oscillation lengths 
2465: without incorporating sterile neutrinos.
2466: However,
2467: we limit attention in this subsection 
2468: to a simple model that reveals some key features
2469: of $\nu\mix\nub$ mixing.
2470: For illustrative purposes, 
2471: it suffices to consider mixing in only one neutrino species,
2472: say $\nu_e\mix\nub_e$.
2473: This case may nonetheless have physical relevance,
2474: since it implies significant effects on reactor experiments 
2475: and solar neutrinos
2476: and might possibly also shed light on the LSND anomaly.
2477: 
2478: 
2479: \subsubsection{General one-species model}\label{1spmod}
2480: 
2481: The restriction to the two-dimensional $\nu_e$-$\nub_e$ subspace 
2482: radically simplifies the form of the effective hamiltonian \rf{heff}.
2483: Since the coefficients $(\mt)_{ee}$ and $(c_L)_{ee}$ are real,
2484: they lead to terms proportional to the identity
2485: that have no effect on oscillatory behavior
2486: and can therefore be ignored.
2487: Moreover,
2488: Eq.\ \rf{C} implies that $(H^\mn\cmat)_{ab}$ 
2489: is antisymmetric in generation space, 
2490: so $(H^\mn\cmat)_{ee}=H^\mn_{ee^C}=0$.
2491: Therefore,
2492: the most general single-flavor theory without mass differences
2493: is given by a $2\times 2$ effective hamiltonian
2494: containing only the coefficients
2495: $(a_L)_{ee}^\mu$ and $(g^{\mn\si}\cmat)_{ee}=g^{\mn\si}_{ee^C}$
2496: for Lorentz violation.
2497: Note that both these terms are CPT odd.
2498: 
2499: For this general single-flavor model,
2500: the probabilities are identical in form 
2501: to those of the usual two-generation mixing case:
2502: \bea
2503: P_{\nu_e\mix\nub_e}&=&
2504: 1-P_{\nu_e\to\nu_e}=
2505: 1-P_{\nub_e\to\nub_e}
2506: \nonumber \\
2507: &=&\sin^22\th\sin^2 2\pi L/L_0 .
2508: \eea
2509: However, 
2510: the mixing angle and oscillation length 
2511: can have nontrivial 4-momentum dependence.
2512: They are given by the expressions
2513: \bea
2514: \left(\fr{2\pi}{L_0}\right)^2&=&
2515: \fr{|(a_L)_{ee}^\mu p_\mu|^2}{|\vp|^2}
2516: + |\sqrt{2}(\ep_+)_\nu p_\si \gt_{ee^C}^{\nu\si}|^2 ,
2517: \nonumber \\
2518: \sin^22\th&=&
2519: \left(1+\fr{|(a_L)_{ee}^\mu p_\mu|^2}
2520: {|\vp|^2|\sqrt{2}(\ep_+)_\nu p_\si \gt_{ee^C}^{\nu\si}|^2}
2521: \right)^{-1} .
2522: \eea
2523: Note that these can also be written directly 
2524: in terms of the neutrino-propagation angles $\Th$ and $\Ph$ 
2525: defined in Eq.\ \rf{vectors}.
2526: 
2527: 
2528: \subsubsection{Example:
2529:   $\gt_{ee^C}^{ZT}\ne 0$}\label{gmod}
2530: 
2531: As an explicit example,
2532: we consider a maximal-mixing special case
2533: of the general single-flavor model
2534: for which the only nonzero coefficient 
2535: is $\gt_{ee^C}^{ZT}\equiv\Gc$.
2536: In terms of the propagation angles $\Th$ and $\Ph$,
2537: the oscillation length is found to be 
2538: \beq
2539: 2\pi/L_0=|E\sin\Th\Gc|,
2540: \eeq
2541: and the mixing angle is $\sin^22\th=1$.
2542: As in the previous examples, 
2543: this case has unconventional energy dependence,
2544: but unlike previous examples
2545: it includes neutrino-antineutrino mixing
2546: and also dependence on the direction of propagation 
2547: through the propagation angle $\Th$.
2548: 
2549: To illustrate the effects of the direction dependence,
2550: consider atmospheric neutrinos detected in the SK detector.
2551: Neutrinos that enter the detector 
2552: from the celestial north or south have $\sin\Th=0$ 
2553: and therefore do not oscillate.
2554: In contrast, 
2555: neutrinos propagating in the plane parallel 
2556: to the Earth's equatorial plane 
2557: have $\sin\Th=1$ and experience maximal mixing
2558: \cite{fn2}.
2559: Analyses of SK data often neglect 
2560: the difference between $\nu_e$ and $\nub_e$,
2561: so they may be insensitive to this effect
2562: because the total flux of electron neutrinos and antineutrinos 
2563: is unchanged.
2564: However, 
2565: the same type of directional dependence 
2566: can arise in more complicated scenarios 
2567: with $\nu_e\mix\nu_\mu\mix\nu_\ta$ mixing,
2568: and this could drastically affect 
2569: the up-down asymmetry measurements of SK.
2570: 
2571: As another example consider KamLAND,
2572: which detects neutrinos from several reactors at different locations.
2573: The total flux $\ph_{\rm total}(E)$ of $\nub_e$ can be written 
2574: \beq
2575: \ph_{\rm total}(E)
2576: =\sum_j \ph_j(E) P_{\nub_e\to\nub_e}(E,L_j,\Th_j),
2577: \eeq
2578: where the $\ph_j(E)$ are the fluxes from the
2579: individual reactors in the absence of oscillations,
2580: and $\Th_j$ are appropriate propagation angles
2581: determined by the relative positions of the reactors 
2582: and the KamLAND detector.
2583: We can approximate the positions of the reactors 
2584: as being located in the plane tangent to the surface of the Earth 
2585: at the location of the detector.
2586: It follows that neutrinos from reactors 
2587: positioned directly north and south of the detector have
2588: $\Th_j\simeq180^\circ-\ch$ and $\Th_j\simeq\ch$,
2589: where $\ch\simeq 36^\circ$
2590: is the latitude of the detector.
2591: In contrast, 
2592: neutrinos arriving from the east or west have $\Th_j\simeq 90^\circ$.
2593: This results in an approximate allowed range for the $\Th_j$
2594: given by $\sin^2\Th_j\gsim \sin^2\ch$,
2595: implying that the $\nub_e$ from every reactor 
2596: experience some degree of oscillation 
2597: on their way to the KamLAND detector.
2598: However,
2599: the net result differs from the flux
2600: in a comparable rotation-symmetric model 
2601: with a dimensionless coefficient.
2602: 
2603: For solar neutrinos,
2604: the allowed range for $\Th$ is given by
2605: $\sin^2\Th\gsim\cos^2\et\simeq0.85$
2606: because the Earth's orbital and equatorial planes differ by
2607: approximately $\et=23^\circ$.
2608: The true value of $\sin^2\Th$ oscillates between $\sin^2\Th=1$
2609: in the spring or fall
2610: and $\sin^2\Th\simeq0.85$ in the summer or winter.
2611: This simple model therefore predicts a semiannual variation 
2612: in the solar-neutrino data.
2613: 
2614: As suggested in Sec.\ \ref{lsnd},
2615: oscillations of $\nu_e$ into $\nub_e$ 
2616: may provide an alternative approach 
2617: to resolving the LSND anomaly.
2618: If the LSND result is reinterpreted
2619: as an oscillation of $\nu_e$ into $\nub_e$,
2620: then the transition probability is 
2621: likely to be comparable to the reported value of about $0.26\%$
2622: because roughly equal numbers of $\nu_e$ and $\nub_\mu$ are produced.
2623: Since mixing in this model is caused 
2624: by the dimensionless coefficient $\Gc$,
2625: a reasonable strategy here is similar to  
2626: that adopted for the $\de\cri$-$\De\cri$ model
2627: in Sec.\ \ref{dcmod},
2628: where a dimensionless coefficient is chosen 
2629: to have its line of sensitivity just above CHOOZ and LSND
2630: in Fig.\ \ref{expts}.
2631: This causes a small oscillation in LSND
2632: but avoids the null constraints from reactor experiments.
2633: Taking the energy of a typical $\nu_e$ to be about $E=35$ MeV
2634: and the distance to be $L=30$ m in LSND,
2635: and assuming that the small transition probability 
2636: is due to a small $L/L_0$,
2637: we can write 
2638: $P_{\nu_e\to\nub_e} =\sin^22\th\sin^22\pi L/L_0
2639: \simeq(\sin\Th\Gc LE)^2 \simeq 0.26\%$.
2640: For LSND, 
2641: the detector is situated approximately to the east of the source.
2642: This implies that the angle between celestial north 
2643: and the direction of propagation of the neutrinos is near $90^\circ$,
2644: which results in the estimate $|\Gc| \simeq 10^{-17}$.
2645: 
2646: In contrast, 
2647: the KARMEN detector is located roughly to the south 
2648: of the neutrino source,
2649: at latitude $\ch \simeq 51^\circ$.
2650: We can therefore approximate
2651: $\Th \simeq 180^\circ - \ch \simeq 129^\circ$.
2652: Taking $E=35$ MeV and $L=18$ m for KARMEN 
2653: yields a transition probability
2654: $P_{\nu_e\to\nub_e} =\sin^22\th\sin^22\pi L/L_0
2655: \simeq(\sin\Th\Gc LE)^2 \simeq 0.06\%$.
2656: This is more than four times smaller than the LSND probability
2657: as a consequence of the different propagation direction 
2658: and the smaller distance,
2659: confirming that direction dependence 
2660: could help reconcile the apparent conflict between KARMEN and LSND.
2661: 
2662: In the above model, 
2663: the directional dependence is rather limited
2664: because the coefficient $\Gc$ introduces only $\Th$ dependence.
2665: This causes minimal variation for any experiments
2666: with both neutrino source and detector fixed on the Earth's surface,
2667: since the angle $\Th$ is fixed as the Earth rotates
2668: and is therefore a constant experiment-dependent quantity.
2669: However,
2670: other coefficients can produce a strong dependence on $\Ph$ as well.
2671: For instance, 
2672: suppose we choose $\gt_{ee^C}^{ZX}$ 
2673: instead of $\gt_{ee^C}^{ZT}$.
2674: The result is an oscillation length given by
2675: $2\pi/L_0=|E\sin^2\Th\cos\Ph \gt_{ee^C}^{ZX}|$.
2676: The dependence on $\Ph$ can substantially 
2677: change the nature of an experiment.
2678: For purely terrestrial experiments,
2679: where the source and detector are fixed to the surface of the Earth,
2680: it follows that $\Ph = \om_\oplus (T-T_0)$,
2681: where $\om_\oplus\simeq 2\pi/$(23 h 56 min)
2682: is the Earth's sidereal frequency 
2683: and $T_0$ is an appropriately chosen experiment-dependent offset.
2684: For solar neutrinos, 
2685: $\Ph$ varies as the Earth orbits the Sun,
2686: $\Ph\approx \Om_\oplus (T-T_0)$,
2687: where $\Om_\oplus = 2\pi/$(1 year).
2688: 
2689: 
2690: \subsection{Lorentz-violating seesaw models}\label{ssmods}
2691: 
2692: The above models demonstrate 
2693: some of the striking behavior at different energy scales
2694: that can arise from Lorentz and CPT violation.
2695: Mixed energy dependence 
2696: among the coefficients for Lorentz violation in $\heff$ 
2697: can also lead to a Lorentz-violating seesaw mechanism
2698: that occurs without mass and only in particular energy regimes.
2699: This can lead to counterintuitive phenomena,
2700: such as the appearance of a pseudomass 
2701: in the bicycle model of Ref.\ \cite{nu}.
2702: In this model,
2703: an oscillation length emerges at high energies 
2704: that behaves like a mass-squared difference,
2705: even though no mass-squared differences exist in the theory.
2706: 
2707: The bicycle model has nonzero coefficients
2708: $\fr43(c_L)^{TT}_{ee} =\fr43(c_L)^{JJ}_{ee} \equiv 2\cee$
2709: and
2710: $(a_L)^Z_{e\mu}=(a_L)^Z_{e\ta}\equiv\a3em/\sqrt{2}$.
2711: The basic behavior of the oscillation lengths 
2712: $L_{ab}\equiv2\pi/\De_{ab}$
2713: and the energy-dependent mixing angle $\th$
2714: are illustrated in Fig.\ \ref{caexpts}.
2715: A key feature is that at high energies
2716: the line associated with the oscillation length $L_{32}$ 
2717: resembles that from a nonzero mass-squared difference.
2718: It turns out that the resulting high-energy dynamics
2719: reduces to two-generation maximal mixing,
2720: $P_{\nu_\mu\mix\nu_\ta}\simeq\sin^2(\De m^2_\Th L/4E)$,
2721: with a Lorentz- and CPT-violating pseudomass 
2722: $\De m^2_\Th=\a3em^2\cos^2\Th/\cee$.
2723: 
2724: Unexpected effects of this type can be expected 
2725: whenever the low- or high-energy limit of $\heff$ 
2726: contains degeneracies.
2727: Consider, for example,
2728: a $3\times 3$ hamiltonian $\heff$ for which there exists a basis,
2729: not necessarily the flavor basis,
2730: in which we can write 
2731: \beq
2732: \heff = \left(\begin{array}{ccc}
2733: 2h_1   & h_2 & h_3 \\
2734: h^*_2 & 0 & 0   \\
2735: h^*_3 & 0   & 0
2736: \end{array}\right) ,
2737: \eeq
2738: where irrelevant diagonal terms are neglected.
2739: The interesting eigenvalue difference for this case is
2740: $\De = \sqrt{(h_1)^2+|h_2|^2+|h_3|^2}-h_1$.
2741: Suppose that the mixed energy dependence introduced 
2742: by combinations of masses,
2743: dimension-one coefficients,
2744: and dimensionless coefficients enforces 
2745: $h_1\gg\sqrt{|h_2|^2+|h_3|^2}$ 
2746: at some energy scale.
2747: Expanding the eigenvalue difference then yields
2748: \cite{fn3}
2749: $\De \approx \half(|h_2|^2+|h_3|^2)/h_1+\cdots$.
2750: 
2751: \begin{figure}
2752: \centerline{\psfig{figure=ca3_all2.eps,width=\hsize}}
2753: \centerline{\psfig{figure=ca3_all1.eps,width=\hsize}}
2754: \caption{\label{caexpts}
2755: Range of oscillation parameters versus energy 
2756: in the bicycle model with 
2757: $\cee=10^{-19}$ and $\a3em=10^{-20}$ GeV.
2758: (a) Minimum ($\cos^2\Th=1$)
2759: and maximum ($\cos^2\Th=0$)
2760: of the various oscillation lengths
2761: $L_{ab}\equiv2\pi/\De_{ab}$.
2762: Note that $L_{32}$ is unbounded.
2763: (b) The allowed range of $\sin^2\th$ and $\cos^2\th$
2764: over all possible directions,
2765: $0\le\cos^2\Th\le1$,
2766: as a function of energy.}
2767: \end{figure}
2768: 
2769: In the bicycle model,
2770: $h_2$ and $h_3$ arise from a dimension-one coefficient 
2771: and are therefore constant with energy,
2772: but $h_1$ arises from a dimensionless coefficient 
2773: and therefore grows linearly with energy.
2774: As a result, 
2775: at high energies the eigenvalue difference 
2776: is proportional to $E^{-1}$,
2777: which resembles the usual mass case.
2778: Using different combinations of masses 
2779: and coefficients for Lorentz violation,
2780: it is straightforward to construct similar models 
2781: that produce $E^{-1}$, $E^{-2}$, or $E^{-3}$ dependence 
2782: at high energies,
2783: or $E^1$, $E^2$, or $E^3$ dependence at low energies.
2784: More complicated $E^n$ dependences are possible 
2785: when the full $6\times 6$ effective hamiltonian \rf{heff}
2786: with $\nu\mix\nub$ mixing is considered.
2787: 
2788: 
2789: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2790: 
2791: \section{Discussion}\label{disc}
2792: 
2793: In this paper,
2794: we have presented a general framework for the study
2795: of Lorentz violation in the neutrino sector.
2796: The key result is Eq.\ \rf{heff},
2797: which represents the general effective hamiltonian $\heff$ 
2798: for neutrino propagation in the presence
2799: of Lorentz and CPT violation.
2800: We have extracted theoretical implications of this hamiltonian
2801: and have initiated a study of experimental sensitivities
2802: to the predicted effects.
2803: The various simple models of Sec.\ \ref{models}
2804: illustrate some of the key physical features
2805: and offer numerous options for future exploration.
2806: 
2807: Our analysis shows that 
2808: the data from existing and near-future neutrino experiments
2809: could be used to attain interesting sensitivities
2810: to possible Lorentz-violating effects.
2811: Moreover,
2812: the existing analyses appear insufficient to
2813: exclude the possibility that some 
2814: or perhaps even all the established neutrino-oscillation signals 
2815: are due to Lorentz violation.
2816: 
2817: An interesting open theoretical challenge is to identify
2818: from the plethora of available choices
2819: one or more elegant models with features compatible with observed data,
2820: preferably involving only a small number of degrees of freedom.
2821: One simple candidate is the bicycle model
2822: \cite{nu},
2823: which has no mass-squared differences and only two degrees of freedom
2824: rather than the four used in the conventional massive-neutrino analysis, 
2825: but which nonetheless reproduces the major observed features of
2826: neutrino behavior. 
2827: This and similar models offer one possible path to explore,
2828: but it is likely that many other qualitatively different
2829: and interesting cases exist. 
2830: 
2831: On the experimental front,
2832: confirming or disproving these ideas
2833: would involve analysis of existing and future data
2834: to seek a `smoking-gun' signal for Lorentz violation.
2835: In the remainder of this section,
2836: we summarize some possible smoking-gun signals
2837: and then offer some remarks about experimental prospects
2838: for detection of Lorentz and CPT violation.
2839: 
2840: 
2841: \subsection{Generic predictions}\label{predictions}
2842: 
2843: The numerous options for coefficients for Lorentz and CPT violation
2844: and the size of unexplored $L$ versus $E$ space
2845: are impediments to a completely general analysis.
2846: An alternative strategy to uncover evidence of Lorentz violation
2847: is to seek model-independent features 
2848: that represent characteristic signals.
2849: We list here six classes of signal.
2850: Confirmed observation of any of them  
2851: would be evidence supporting the existence of Lorentz violation.
2852: 
2853: \begin{trivlist}
2854: \item
2855: {\it Class I:\ Spectral anomalies.}
2856: Each coefficient for Lorentz violation
2857: introduces energy dependence differing from the usual case.
2858: Detection of a vacuum oscillation length
2859: that is constant in $E$
2860: or inversely proportional to $E$ to some power
2861: would constitute a clear signal of Lorentz violation.
2862: Note that combinations of 
2863: masses, dimension-one coefficients,
2864: dimensionless coefficients, and matter potentials
2865: can produce more complicated energy dependences
2866: in both oscillation lengths and mixing angles.
2867: In general,
2868: a mixing angle is constant in energy
2869: only if all relevant coefficients for Lorentz violation,
2870: masses, and matter effects have the same dimension,
2871: which requires no more than one of these to be present.
2872: 
2873: \item
2874: {\it Class II:\ $L$--$E$ conflicts.}
2875: This class of signal refers to any null or positive measurement
2876: in a region of $L$--$E$ space that conflicts with
2877: all scenarios based on mass-squared differences.
2878: For example,
2879: consider a solid line in Fig.\ \ref{expts} 
2880: passing through CHOOZ.
2881: A measurement of substantial oscillation in the $\nub_e$ sector 
2882: in any experiment below this line would be in direct conflict 
2883: with a mass-based interpretation of the CHOOZ results.
2884: Signals in this class might best be sought 
2885: by searching for oscillation effects
2886: in each species of neutrino and antineutrino
2887: for regions of $L$--$E$ space 
2888: in which conventional oscillations are excluded.
2889: Of the six classes of signal discussed in this section, 
2890: this is the only one 
2891: for which there is presently some positive evidence,
2892: the LSND anomaly.
2893: 
2894: \item
2895: {\it Class III:\ Periodic variations.}
2896: This class involves signals for rotation-invariance violations
2897: and contains two subclasses:
2898: sidereal variations and annual variations.
2899: Consider first sidereal variations,
2900: which have been widely adopted as the basis 
2901: for Lorentz-violation searches in other sectors of the SME.
2902: In terrestrial experiments
2903: with both the detector and the source fixed on the Earth,
2904: the direction of neutrino propagation relative to 
2905: the Sun-centered frame changes during the sidereal day
2906: due to the rotation of the Earth.
2907: The induced periodic variation of observables with time 
2908: represents a signature of Lorentz violation.
2909: In the Sun-centered frame, 
2910: the neutrino-propagation angle $\Th$ is constant for a fixed source,
2911: but the angle $\Ph$ varies periodically 
2912: according to $\Ph=\om_\oplus (T-T_0)$,
2913: where $T_0$ is an experiment-dependent time 
2914: at which the detector and source both lie in a plane 
2915: parallel to the $XZ$ plane with the detector at larger values of $X$.
2916: The resulting neutrino-oscillation probabilities 
2917: exhibit periodic variations
2918: at multiples of the sidereal frequency $\om_\oplus$.
2919: The second class of periodic signals,
2920: annual variations,
2921: can also arise directly from rotation-invariance violation.
2922: For solar-neutrino experiments,
2923: the source is the Sun and the detector changes location with time  
2924: as a consequence of the orbital motion of the Earth about the Sun. 
2925: One can therefore expect variations at the Earth orbital frequency 
2926: $\Om_\oplus$ and its harmonics.
2927: In this context,
2928: note that the direction $\hat p$ of solar neutrino propagation 
2929: in the Sun-centered frame is uniquely given by
2930: $\hat p=(-\cos\Om_\oplus T, 
2931: -\cos\et\sin\Om_\oplus T,
2932: -\sin\et\sin\Om_\oplus T)$,
2933: where $\et\simeq 23.4^\circ$
2934: is the angle between the
2935: Earth's equatorial and orbital planes.
2936: We remark in passing that suppressed annual variations can also arise
2937: indirectly as boost-violating effects 
2938: \cite{cavexpt,km,spaceexpt,cane}
2939: in experiments with terrestrial 
2940: and possibly atmospheric neutrino sources,
2941: as a result of the noninertial nature of the Earth's motion
2942: around the Sun.
2943: 
2944: \item
2945: {\it Class IV:\ Compass asymmetries.}
2946: This class also results from rotation-invariance violations,
2947: but the signals are independent of time.
2948: They can be characterized as the observation
2949: of unexplained directional asymmetries
2950: at the location of the detector.
2951: For terrestrial and atmospheric experiments,
2952: averaging over time eliminates the dependence 
2953: on the neutrino-propagation angle $\Ph$,
2954: so the result depends only on energy and the angle $\Th$.
2955: Rotation-symmetry violations can therefore cause 
2956: a difference in observed properties of neutrinos 
2957: originating from different directions.
2958: Note that the east and west directions are equivalent 
2959: under the averaging process,
2960: since the $\Ph$ dependence is eliminated,
2961: but direct comparison of the north, south, and east directions
2962: would be of interest for these signals.
2963: Note also that the $\Th$ dependence 
2964: typically introduces vertical up-down effects
2965: and could include,
2966: for example,
2967: modifications in the up-down asymmetry of atmospheric neutrinos.
2968: We remark also that compass asymmetries 
2969: can carry information completely independent of 
2970: the information in periodic variations.
2971: This is seen in the example in Sec.\ \ref{gmod},
2972: which has $\Th$ dependence but no $\Ph$ dependence
2973: and consequently predicts compass asymmetries 
2974: without sidereal variations.
2975: 
2976: \item
2977: {\it Class V:\  Neutrino-antineutrino mixing.}
2978: This class of signal includes any appearance measurement 
2979: that can be traced to $\nu\mix\nub$ oscillation.
2980: Any model with nonzero coefficients of type $g$ or $H$
2981: exhibits this behavior,
2982: including the class of simple one-species models
2983: discussed in Sec.\ \ref{1spmod}.
2984: Note that this class of signal involves lepton-number violation.
2985: 
2986: \item
2987: {\it Class VI:\ Classic CPT test:
2988:   $P_{\nu_b\to\nu_a}\ne P_{\nub_a\to\nub_b}$.}
2989: This is the traditional test of CPT
2990: discussed in Sec.\ \ref{cptprop},
2991: involving violation of the result \rf{cpt1}.
2992: A related signal would be violation of the second result,
2993: Eq.\ \rf{cpt2},
2994: which also involves $\nu\mix\nub$ mixing.
2995: 
2996: \end{trivlist}
2997: 
2998: 
2999: \subsection{Experimental Prospects}\label{prospects}
3000: 
3001: \begin{table*}
3002: \renewcommand{\arraystretch}{1.5}
3003: \begin{tabular}{ccc||ccccc}
3004: \hline
3005: \hline
3006: \multicolumn{3}{c||}{Coefficients}&
3007: \multicolumn{5}{c}{Estimated sensitivities from Fig.\ \ref{expts}} \\
3008: \hline
3009: \makebox[110pt][c]{Matrix}&
3010: \makebox[30pt][c]{DOF}&
3011: \makebox[80pt][c]{Signal classes}&
3012: \makebox[50pt][c]{Solar}&
3013: \makebox[50pt][c]{Atmospheric}&
3014: \makebox[50pt][c]{Reactor}&
3015: \makebox[50pt][c]{Short base.}&
3016: \makebox[50pt][c]{Long base.}\\
3017: \hline
3018: $(a_L)^T$&8&{\it I,II,VI}&
3019: $-27$ & $-23$ & $-21$ & $-19$ & $-21$ \\
3020: $(a_L)^J$&24&{\it I,II,III,IV,VI}&
3021: $-27$ & $-23$ & $-21$ & $-19$ & $-21$ \\
3022: $(c_L)^{TT}=(c_L)^{JJ}$&8&{\it I,II}&
3023: $-25$ & $-24$ & $-19$ & $-21$ & $-22$ \\
3024: $\half(c_L)^{(TJ)}$&24&{\it I,II,III,IV}&
3025: $-25$ & $-24$ & $-19$ & $-21$ & $-22$ \\
3026: $\half(c_L)^{(JK)}-\fr13\de^{JK}(c_L)^{TT}$&40&{\it I,II,III,IV}&
3027: $-25$ & $-24$ & $-19$ & $-21$ & $-22$ \\
3028: $\gt^{JT}-\fr i2\ep^{JKL}\gt^{KL}$&36&{\it I,II,III,IV,V,VI}&
3029: $-25$ & $-24$ & $-19$ & $-21$ & $-22$ \\
3030: $\half\gt^{(JK)}-\fr13\de^{JK}\gt^{LL}$&60&{\it I,II,III,IV,V,VI}&
3031: $-25$ & $-24$ & $-19$ & $-21$ & $-22$ \\
3032: $\Ht^J$&18&{\it I,II,III,IV,V}&
3033: $-27$ & $-23$ & $-21$ & $-19$ & $-21$ \\
3034: $(k_d)^{\la\ldots}$&var.\ &{\it I,II,III,IV,V,VI}&
3035: $-27+2d$ & $-23-d$ & $-21+2d$ & $-19-2d$ & $-21-d$ \\
3036: \hline
3037: \hline
3038: \end{tabular}
3039: \caption{Experimental prospects.}
3040: \end{table*}
3041: 
3042: 
3043: We conclude with some comments 
3044: about prospects for Lorentz- and CPT-violation searches
3045: in the major types of experiments.
3046: Table I provides a summary of the present situation.
3047: The left-hand part of the table contains three columns
3048: with information about coefficients for Lorentz violation. 
3049: The first column lists 
3050: combinations of coefficient matrices 
3051: relevant to neutrino propagation, 
3052: extracted from the general hamiltonian \rf{heff}
3053: and separated according to rotation properties
3054: into timelike ($T$) and spacelike ($J$) components 
3055: in the Sun-centered frame.
3056: The second column lists the maximum number 
3057: of independent degrees of freedom (DOF) 
3058: associated with each combination of coefficient matrices.
3059: These numbers can be obtained by examining the form 
3060: of Eq.\ \rf{heff} and using the symmetry properties 
3061: in generation space listed in Eq.\ \rf{C}.
3062: In certain specific models,
3063: some of these degrees of freedom may be unobservable.
3064: The third column displays the classes of signal 
3065: that are relevant for each coefficient matrix,
3066: using the nomenclature of the previous subsection.
3067: The right-hand part of the table 
3068: contains estimated attainable sensitivities,
3069: classified according to each of five types of oscillation experiments.
3070: Each entry in the table represents the
3071: base-10 logarithm of the expected sensitivity 
3072: to the corresponding coefficient for Lorentz violation.
3073: The sensitivities shown in the table can be obtained 
3074: by examination of Fig.\ \ref{expts}.
3075: Given an experiment 
3076: with maximum $L$ coverage of $L_{\rm max}$
3077: and maximum $E$ coverage of $E_{\rm max}$,
3078: the crude sensitivity $\si$ 
3079: to a coefficient for Lorentz violation of dimension $1-d$
3080: is taken to be
3081: $\si \approx - \log L_{\rm max} - d \log E_{\rm max}$.
3082: For simplicity in the presentation,
3083: it is understood that the sensitivities 
3084: listed for the dimension-one coefficients 
3085: $a_L$, $H$ are measured in GeV.
3086: The final row of the table contains a rough estimate
3087: of sensitivities measured in GeV$^{(1-d)}$
3088: to a generic coefficient $(k_d)_{\la\ldots}$ 
3089: for a Lorentz-violating operator 
3090: of nonrenormalizable dimension $n=d+3$.
3091: Some caution is required in interpreting the latter numerical estimates
3092: because the coefficients $(k_d)_{\la\ldots}$ 
3093: are expected typically to be suppressed
3094: by $d$-dependent powers of the Planck scale. 
3095: 
3096: The table confirms that Planck-scale sensitivities 
3097: to Lorentz and CPT violation are attainable 
3098: in all classes of experiment,
3099: with the most sensitive cases 
3100: potentially rivaling the best tests in other sectors of the SME.
3101: Note that the estimated sensitivities
3102: assume order-one measurements and therefore may underestimate 
3103: the true attainable sensitivity in any specific experiment. 
3104: Note also that a variety of experimental analyses
3105: are needed to extract complete information on Lorentz and CPT violation,
3106: with no single class of experiment
3107: presently in a position to provide 
3108: measurements of a complete set of coefficients.
3109: In the remainder of this subsection,
3110: we offer a few more specific remarks 
3111: about each type of experiment.
3112: 
3113: \begin{trivlist}
3114: 
3115: \item
3116: {\it Solar-neutrino experiments.}
3117: The abundance and quality of the current solar-neutrino data 
3118: make these experiments a promising avenue  
3119: for Lorentz-violation searches.
3120: The relatively large range of solar-neutrino energies 
3121: suggests interesting information about spectral anomalies
3122: might be obtained,
3123: but complications introduced by matter effects 
3124: are likely to make this practical 
3125: only in relatively simple cases 
3126: such as the FC model \rf{rim}.
3127: Of the other classes of signals,
3128: periodic variations and neutrino-antineutrino mixing 
3129: may be the most relevant to solar neutrinos.
3130: The periodic variations in observables 
3131: would occur at multiples of $\Om_\oplus$,
3132: appearing despite compensation for the flux variation
3133: due to the eccentricity of the Earth's orbit.
3134: Direct detection of any antineutrinos originating from the Sun 
3135: would be evidence of $\nu\mix\nub$ mixing
3136: and hence of possible Lorentz violation.
3137: 
3138: \item
3139: {\it Atmospheric-neutrino experiments.}
3140: Like solar neutrinos,
3141: atmospheric neutrinos cover a relatively large region 
3142: of $L$--$E$ space,
3143: but complications from matter effects
3144: hinder a general spectral-anomaly search.
3145: However, 
3146: Fig.\ \ref{expts2} shows 
3147: that searches for atmospheric oscillations 
3148: at the highest energies and largest distances 
3149: could reveal oscillations absent in the usual solution,
3150: thereby providing evidence for $L$--$E$ conflicts.
3151: Atmospheric neutrinos originate from all directions,
3152: so they are an ideal system for directional-dependence searches.
3153: Not only are they sensitive to sidereal variations,
3154: but also the directional capabilities of detectors like SK
3155: make atmospheric neutrinos perhaps the most promising place 
3156: to search for compass asymmetries.
3157: Moreover,
3158: since atmospheric data involve both neutrinos and antineutrinos 
3159: of two species in comparable numbers,
3160: it may be possible to address both $\nu\mix\nub$ mixing 
3161: and the classic CPT tests \rf{cpt1} and \rf{cpt2}.
3162: 
3163: \item
3164: {\it Reactor experiments.}
3165: Nuclear reactors are good sources of $\nub_e$,
3166: and they are therefore well suited 
3167: to searches for $\nu\mix\nub$ mixing.
3168: Since both the sources and the detectors in all these cases are fixed,
3169: the experiments are also sensitive to sidereal variations,
3170: and some may have additional sensitivity to compass asymmetries.
3171: For example,
3172: the reactor experiment KamLAND detected neutrinos 
3173: from multiple reactors and different locations.
3174: Experiments with multiple sources like this 
3175: can analyze their data for compass asymmetries 
3176: that depend on the direction to the various neutrino sources.
3177: 
3178: \item
3179: {\it Short-baseline accelerator experiments.}
3180: LSND already seems to suggest a positive 
3181: $L$--$E$ conflict,
3182: which will be tested by the forthcoming results 
3183: of the MiniBooNE experiment.
3184: Many of these short-baseline accelerator experiments
3185: are especially interesting for signals based on $L$--$E$ conflicts
3186: because they operate in a region of $L$--$E$ space
3187: where the conventional mass scenario predicts no oscillations.
3188: Sidereal variations can readily be sought by experiments  
3189: such as CHORUS, KARMEN, MiniBooNE, NOMAD, and NuTeV,
3190: since each has a fixed source and detector.
3191: Note that the existing data from these experiments 
3192: could in principle contain a positive signal for sidereal variations
3193: because the published null results are based on an average over time.
3194: The well-defined flavor content of the sources for these experiments 
3195: may also offer sensitivity to $\nu\mix\nub$ signals 
3196: and to the classic CPT test.
3197: Some of these experiments,
3198: such as MiniBooNE and NuTeV,
3199: may be particularly sensitive to Lorentz violation
3200: because they can switch from a predominately 
3201: $\nu_\mu$ source to a predominately $\nub_\mu$ source.
3202: 
3203: \item
3204: {\it Long-baseline accelerator experiments.}
3205: Several future long-baseline accelerator-based experiments,
3206: such as ICARUS, MINOS, and OPERA,
3207: are planned to probe the GeV region of $L$--$E$ space
3208: at distances of hundreds of kilometers,
3209: and some results in this regime have already been reported by K2K.
3210: These experiments can search for oscillations in $\nu_\mu$ 
3211: obtained from meson decays,
3212: and they are designed to test the atmospheric-oscillation hypothesis.
3213: Nonetheless, 
3214: $L$--$E$ conflicts are still possible:
3215: a measurement of $\nu_\mu\to\nu_e$,
3216: for example, 
3217: would represent an $L$--$E$ conflict
3218: because this oscillation is absent at these energies and distances 
3219: in the conventional scenario with masses.
3220: The data obtained can be also analysed for sidereal variations,
3221: since in each case the source and detector are fixed.
3222: Moreover, 
3223: except for OPERA and ICARUS,
3224: which are both 
3225: part of the CERN Neutrinos to Gran Sasso (CNGS) project,
3226: the beamline for each experiment points 
3227: in a different direction.
3228: This means each is expected to respond differently 
3229: to rotation-invariance violations.
3230: These experiments may also be able to address
3231: $\nu\mix\nub$ mixing and the classic CPT signal,
3232: since the flavor content of the beams is well known.
3233: 
3234: \item
3235: {\it Other experiments.}
3236: Experiments designed to search for neutrino properties 
3237: other than oscillations can also address Lorentz violation.
3238: To some extent,
3239: most experiments are sensitive to sidereal variations
3240: and compass asymmetries.
3241: The other signals discussed in Sec.\ \ref{predictions}
3242: are more unique to neutrino oscillations,
3243: but analogous signatures are likely to arise in most cases.
3244: 
3245: One possible test of Lorentz invariance
3246: involves a direct comparison of velocities
3247: of supernova neutrinos and photons,
3248: such as those from SN1987A
3249: \cite{sn1987a,snnugamma},
3250: which could be performed 
3251: either by some of the experiments listed above 
3252: or by neutrino telescopes. 
3253: A similar method has been applied in the photon sector,
3254: where the velocities of different polarizations are compared
3255: \cite{km}.
3256: Another method that could be adapted to the neutrino case 
3257: is a simple pulse-dispersion analysis.
3258: The energy dependence 
3259: and the independent propagation of each $\heff$ eigenstate
3260: imply that different components of the neutrino pulse
3261: propagate at different velocities,
3262: causing the pulse to spread.
3263: For SN1987A, 
3264: all the observed neutrinos arrived 
3265: in a time interval of about $\de T \simeq 10$ s
3266: and had energies $E\simeq 10-20$ MeV.
3267: Since these neutrinos took roughly $T_0 \simeq 5\times 10^{12}$ s
3268: to reach the Earth,
3269: we can crudely estimate that the maximum difference in velocity 
3270: across the $\De E\simeq 10$ MeV energy spread 
3271: of the $\heff$ eigenstates
3272: is $\de v/c\simeq\de T/T_0\simeq 2\times 10^{-12}$.
3273: We can then make a simple dimensional estimate of the sensitivity 
3274: of this method to various terms in $\heff$.
3275: This suggests a sensitivity of about $\sqrt{200}$ eV to mass terms,
3276: $2\times 10^{-14}$ GeV to dimension-one coefficients,
3277: and $2\times 10^{-12}$ to dimensionless coefficients.
3278: The mass estimate agrees 
3279: with the result of a detailed analysis along these lines 
3280: \cite{snnu}.
3281: 
3282: Lorentz violation may also be relevant to 
3283: direct mass searches such as the proposed KATRIN experiment
3284: \cite{katrin},
3285: designed to measure directly the $\nu_e$ mass to better than 1 eV.
3286: Within the currently accepted solution to the oscillation data,
3287: a mass matrix with eV-scale masses but
3288: mass-squared differences of $10^{-3}$ eV$^2$ and $10^{-5}$ eV$^2$ 
3289: would be nearly degenerate.
3290: This seems unlikely in light of the charged-lepton mass hierarchies.
3291: However, 
3292: suppose that the mass matrix is nearly diagonal
3293: and that neutrino oscillations are primarily
3294: or entirely due to Lorentz violation instead.
3295: Then,
3296: individual masses of eV order or greater  
3297: may be present with little or no effect 
3298: on the existing neutrino-oscillation data,
3299: but they would produce a signal in the KATRIN experiment.
3300: 
3301: Another area of widespread interest
3302: is the search for neutrinoless double-beta decay
3303: \cite{bbdecay}.
3304: This decay mode is an indicator of lepton-number violation,
3305: which can result from Majorana-type couplings 
3306: introduced by Majorana masses
3307: or by gauge-violating coefficients for Lorentz violation.
3308: Many of the null results of searches for neutrinoless double-beta decay
3309: could therefore be reanalysed to yield constraints 
3310: on certain types of Lorentz violation.
3311: 
3312: \end{trivlist}
3313: 
3314: \acknowledgments
3315: This work was supported in part by the 
3316: United States Department of Energy
3317: under grant number DE-FG02-91ER40661
3318: and by the 
3319: National Aeronautics and Space Administration
3320: under grant number NAG8-1770.
3321: 
3322: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3323: 
3324: \appendix
3325: 
3326: \section{Effective hamiltonian}\label{hcalc}
3327: 
3328: This appendix presents some details for the derivation 
3329: of the effective hamiltonian \rf{heff}.
3330: We first perform a spinor decomposition of the hamiltonian 
3331: in the mass-diagonal Majorana basis.
3332: The result is then block diagonalized in the light-neutrino sector
3333: and transformed into the original weak-interaction basis.
3334: We remind the reader that 
3335: generation indices in the mass-diagonal basis are
3336: $\AA=1,\ldots, 6$ for $N=3$ neutrino species,
3337: while the restriction to light neutrinos in this basis 
3338: is represented by indices $\aa=1,2,3$.
3339: Also,
3340: in the flavor basis,
3341: upper-case indices take the values $A=e,\mu,\ta,e^C,\mu^C,\ta^C$,
3342: while lower-case ones span $a=e,\mu,\ta$.
3343: 
3344: \subsection{Spinor decomposition}
3345: 
3346: In this section,
3347: we project the hamiltonian onto the massless spinor basis 
3348: used in Eq.\ \rf{nup}.
3349: This corresponds to choosing a convenient
3350: $\vp$-dependent $\ga$-matrix basis
3351: that allows us to write the equations of motion 
3352: in terms of the $b$ and $d$ amplitudes.
3353: 
3354: Working in the mass-diagonal basis,
3355: the hamiltonian is given by 
3356: \bea
3357: \cH_{\AA\BB}(\vp)&=&
3358: \cH_{0\AA\BB}(\vp)+\de\cH_{\AA\BB}(\vp) ,
3359: \nonumber \\
3360: \cH_{0\AA\BB}(\vp)&=&
3361: \ga^0(\vec\ga\cdot\vp+m_{(\AA )})\de_{\AA\BB}\ ,
3362: \nonumber \\
3363: \de\cH_{\AA\BB}(\vp)
3364: &=&-\half(\ga^0\de\Ga^0\cH_0(\vp)+\cH_0(\vp)\ga^0\de\Ga^0)_{\AA\BB}
3365: \nonumber \\
3366: &&+\ga^0(\de\vec\Ga\cdot\vp+\de M)_{\AA\BB}\ .
3367: \label{cH}\eea
3368: It turns out to be useful to decompose 
3369: $\Ga^\mu_{\AA\BB}$ and $M_{\AA\BB}$
3370: in terms of $\ga$ matrices,
3371: as in Eq.\ \rf{GaM}.
3372: Therefore, we write
3373: \bea
3374: \Ga^\nu_{\AA\BB} &=&
3375: \ga^0U_{\AA A}\ga^0\Ga^\nu_{AB}(U_{\BB B})^\dag
3376: \nonumber \\
3377: &=&\ga^\nu \de_{\AA\BB}
3378: + c^\mn_{\AA\BB}\ga_\mu
3379: + d^\mn_{\AA\BB}\ga_5\ga_\mu
3380: \nonumber \\&&
3381: + e^\nu_{\AA\BB}
3382: + if^\nu_{\AA\BB}\ga_5
3383: + \half g^{\la\mn}_{\AA\BB}\si_{\la\mu}\ ,
3384: \nonumber \\
3385: M_{\AA\BB} &=&
3386: \ga^0U_{\AA A}\ga^0M_{AB}(U_{\BB B})^\dag
3387: \nonumber \\
3388: &=&m_{\AA\BB}
3389: +im_{5\AA\BB}\ga_5
3390: \nonumber \\&&
3391: + a^\mu_{\AA\BB}\ga_\mu
3392: + b^\mu_{\AA\BB}\ga_5\ga_\mu
3393: + \half H^\mn_{\AA\BB}\si_\mn\ .\quad
3394: \label{GaMp}\eea
3395: 
3396: We begin the spinor decomposition of the hamiltonian \rf{cH}
3397: by considering the Lorentz-covariant terms.
3398: The properties of the massless spinor basis imply 
3399: that the only nonzero projections of the kinetic term are 
3400: \bea
3401: \lefteqn{
3402:   u_{L,R}^\dag(\vp)(\ga^0\vec\ga\cdot\vp\ \de_{\AA\BB}) u_{L,R}(\vp)}
3403: \nonumber \\
3404: &=&-v_{R,L}^\dag(-\vp)(\ga^0\vec\ga\cdot\vp\ \de_{\AA\BB}) v_{R,L}(-\vp)
3405: \nonumber \\
3406: &=&|\vp|\de_{\AA\BB}\ ,
3407: \eea
3408: while the surviving projections of the mass term are
3409: \bea
3410: \lefteqn{
3411:   u_{L,R}^\dag(\vp)(\ga^0 m_{(\AA)} \de_{\AA\BB}) v_{R,L}(-\vp)}
3412: \nonumber \\
3413: &=&\bar u_{L,R}(\vp)v_{R,L}(-\vp) m_{(\AA)} \de_{\AA\BB}\ 
3414: \eea
3415: and conjugates.
3416: The quantities $\bar u_{L,R}(\vp)v_{R,L}(-\vp)$
3417: are phases that can be chosen arbitrarily 
3418: by changing the relative phase between
3419: $u_{L,R}(\vp)$ and $v_{R,L}(-\vp)=C\bar u^T_{L,R}(-\vp)$.
3420: 
3421: For the spinor decomposition of the Lorentz-violating terms
3422: in the hamiltonian \rf{cH},
3423: we define the $2\times 2$ matrices
3424: \bea
3425: \lefteqn{
3426:   \La_{\AA\BB}(\vp)=
3427:   \La_{\BB\AA}^\dag(\vp)}
3428: \nonumber \\
3429: &=&\left( \begin{array}{c}
3430: u_L^\dag(\vp) \\ u_R^\dag(\vp)
3431: \end{array} \right)
3432: \de\cH_{\AA \BB}(\vp)
3433: \big( u_L(\vp) , u_R(\vp) \big) ,
3434: \label{La} \\
3435: \lefteqn{
3436:   \tilde\La_{\AA \BB}(\vp)=
3437:   -\tilde\La_{\BB \AA}^T(-\vp)}
3438: \nonumber \\
3439: &=&\left(\begin{array}{c}
3440: u_L^\dag(\vp) \\ u_R^\dag(\vp)
3441: \end{array}\right)
3442: \de\cH_{\AA \BB}(\vp)
3443: \big( v_R(-\vp) , v_L(-\vp) \big) .\quad
3444: \label{tLa}
3445: \eea
3446: It can be shown that the mass-basis analogues of
3447: the relations \rf{GaMC} are
3448: $\Ga^\nu_{\AA\BB}=-C(\Ga^\nu_{\BB\AA})^TC^{-1}$
3449: and $M_{\AA\BB}=C(M_{\BB\AA})^TC^{-1}$.
3450: Note that this corresponds to $\cmat\to I$,
3451: which reflects the Majorana nature of neutrinos in this basis.
3452: These identities may then be used to
3453: show that $C^\dag\ga^0\cH_{\AA\BB}(\vp)\ga^0C
3454: =-[\cH_{\AA\BB}(-\vp)]^*$.
3455: Finally, 
3456: with the aid of the relation
3457: $v_{R,L}(\vp)=C\bar u^T_{L,R}(\vp)$,
3458: it follows that the remaining terms in the spinor decomposition 
3459: are given in terms of $\La$ and $\tilde\La$ by
3460: \bea
3461: \lefteqn{-\tilde\La_{\AA\BB}^*(-\vp)}
3462: \nonumber \\ && =
3463: \left(\begin{array}{c}
3464: v_R^\dag(-\vp) \\ v_L^\dag(-\vp)
3465: \end{array}\right)
3466: \de\cH_{\AA\BB}(\vp)
3467: \big( u_L(\vp) , u_R(\vp) \big) , \\
3468: \lefteqn{-\La_{\AA\BB}^*(-\vp)}
3469: \nonumber \\ && =
3470: \left(\begin{array}{c}
3471: v_R^\dag(-\vp) \\ v_L^\dag(-\vp)
3472: \end{array}\right)
3473: \de\cH_{\AA\BB}(\vp)
3474: \big( v_R(-\vp) , v_L(-\vp) \big) . \quad\quad
3475: \eea
3476: This implies that the $2\times 2$ matrices
3477: $\La_{\AA\BB}$, $\tilde\La_{\AA\BB}$
3478: determine the Lorentz-violating effects.
3479: 
3480: Combining the above results,
3481: we obtain the spinor-decomposed hamiltonian 
3482: appearing in Eq.\ \rf{we}:
3483: \bea
3484: \lefteqn{
3485: H_{\AA\BB}(\vp)=H_{\BB\AA}^\dag(\vp) }
3486: \nonumber \\
3487: &=&\de_{\AA\BB}\left(\begin{array}{cc}
3488: |\vp|
3489: &
3490: m_{(\AA)}\et(\vp)
3491: \\
3492: -m_{(\AA)}\et^*(-\vp)
3493: &
3494: -|\vp|
3495: \end{array}
3496: \right)
3497: \nonumber \\
3498: &&\quad
3499: +\left(\begin{array}{cc}
3500: \La_{\AA\BB}(\vp)
3501: &
3502: \tilde\La_{\AA\BB}(\vp)
3503: \\
3504: -\tilde\La_{\AA\BB}^*(-\vp)
3505: &
3506: -\La_{\AA\BB}^*(-\vp)
3507: \end{array}
3508: \right) ,
3509: \label{H}
3510: \eea
3511: where $\et$ is the $2\times 2$ diagonal matrix of phases
3512: $\et(\vp)=-\et(-\vp)
3513: ={\rm diag}[\bar u_L(\vp) v_R(-\vp),
3514: \bar u_R(\vp) v_L(-\vp)]$.
3515: 
3516: We seek an explicit expression for $\La_{\AA\BB}$.
3517: The next subsection shows that 
3518: the effects of $\tilde\La_{\AA\BB}$ are subleading order,
3519: so we concentrate here on the projections in $\La_{\AA\BB}$,
3520: which involve the spinors $u_L$ and $u_R$.
3521: It is useful first to find expressions for the quantities
3522: $\bar u_\al \{1,\ga_5,\ga^\mu,\ga_5\ga^\mu,\si^\mn\}u_\be$,
3523: where $\al,\be=L,R$.
3524: We obtain the following nonzero results:
3525: \bea
3526: \bar u_\al \ga^\mu u_\be&=&p^\mu\de_{\al\be}/|\vp| ,
3527: \nonumber \\
3528: \bar u_\al \ga_5\ga^\mu u_\be&=&S_\al p^\mu\de_{\al\be}/|\vp| ,
3529: \nonumber \\
3530: \bar u_L \si^\mn u_R&=&
3531: (\bar u_R \si^\mn u_L)^*
3532: \nonumber \\
3533: &=& i\sqrt{2}(p^\mu(\ep_+)^\nu-p^\nu(\ep_+)^\mu)/|\vp| ,\quad
3534: \label{spinpr}\eea
3535: where 
3536: $S_L=1$, 
3537: $S_R=-1$, 
3538: $p^\mu=(|\vp|;\vp)$,
3539: and $(\ep_+)^\mu$ satisfies the relations \rf{ep}.
3540: With these results and Eqs.\ \rf{cH} and \rf{GaMp},
3541: we can extract the projections of $\de\cH$
3542: onto $u_L$ and $u_R$:
3543: \begin{widetext}
3544: \beq
3545: \La_{\AA\BB}=
3546: \fr{1}{|\vp|}
3547: \left(\begin{array}{cc}
3548: [(a+b)^\mu p_\mu-(c+d)^\mn p_\mu p_\nu]_{\AA\BB} &
3549: -i\sqrt{2}p_\mu(\ep_+)_\nu[
3550: g^{\mn\si}p_\si-H^\mn]_{\AA\BB} \\
3551: i\sqrt{2}p_\mu(\ep_+)^*_\nu[
3552: g^{\mn\si}p_\si-H^\mn]_{\AA\BB} &
3553: [(a-b)^\mu p_\mu-(c-d)^\mn p_\mu p_\nu]_{\AA\BB}
3554: \end{array}\right) .
3555: \eeq
3556: \end{widetext}
3557: In this expression,
3558: we neglect off-diagonal terms entering as mass 
3559: multiplied by coefficients for Lorentz violation,
3560: since in most situations these
3561: terms are suppressed relative to those above.
3562: 
3563: 
3564: \subsection{Block diagonalization}
3565: 
3566: The above spinor decomposition of the hamiltonian 
3567: is independent of the specific neutrino mass spectrum.
3568: To make further progress,
3569: we adopt the scenario described at the beginning
3570: of Section \ref{phenom}
3571: and restrict attention to ultrarelativistic dynamics
3572: in the subspace of light neutrinos,
3573: spanned by the $\aa$ indices.
3574: The hamiltonian is then dominated by
3575: the diagonal kinetic term in Eq.\ \rf{H}.
3576: The upper and lower diagonal blocks of this term have opposite sign, 
3577: so they differ by an amount large compared to 
3578: both mass and Lorentz-violating terms.
3579: This in turn implies that standard perturbation techniques 
3580: to remove the off-diagonal blocks can be applied.
3581: As a result,
3582: terms in the off-diagonal blocks of Eq.\ \rf{H} 
3583: appear at second order in the block-diagonalized form.
3584: One consequence is that the leading-order mass contribution 
3585: appears at second order,
3586: whereas certain forms of Lorentz violation appear already at first order.
3587: This feature can ultimately be traced to the $\ga$-matrix structure 
3588: of the Lorentz-covariant portion of the theory.
3589: 
3590: Provided the conditions 
3591: $m_{(\aa )},|\La_{\aa \bb}| |\tilde\La_{\aa \bb}| \ll |\vp|$
3592: are satisfied,
3593: the block diagonalization of Eq.\ \rf{H}
3594: can proceed through the perturbative construction 
3595: of an appropriate unitary matrix ${\cal U}$.
3596: First,
3597: write ${\cal U}$ in the form
3598: ${\cal U}=I+\ep^{(1)}+\ep^{(2)}+\ldots$,
3599: where $\ep^{(n)}$ is of $n$th order in
3600: the dimensionless small quantities
3601: $m_{(\aa )}/|\vp|$,
3602: $\La_{\aa \bb}/|\vp|$, and
3603: $\tilde\La_{\aa \bb}/|\vp|$.
3604: The block-diagonal hamiltonian resulting from this transformation
3605: can be expanded in a similar fashion:
3606: \bea
3607: H_{\aaa \bbb}&=&{\cal U}_{\aaa \aa}H_{\aa \bb}{\cal U}_{\bbb \bb}^\dag
3608: \nonumber \\
3609: &=&H^{(0)}_{\aaa \bbb}+H^{(1)}_{\aaa \bbb}+H^{(2)}_{\aaa \bbb}+\cdots ,
3610: \eea
3611: where each $H^{(n)}_{\aaa \bbb}$
3612: is $n$th order in small quantities.
3613: The zeroth-order term $H^{(0)}_{\aaa \bbb}$ is the usual kinetic term,
3614: which is already block diagonal.
3615: The first-order term $H^{(1)}_{\aaa \bbb}$
3616: can be made block diagonal by an appropriate choice of $\ep^{(1)}$.
3617: A suitable leading-order transformation is
3618: \beq
3619: \ep^{(1)}_{\aaa \bb}=
3620: \fr{\de_{\aaa \aa}}{2|\vp|}
3621: \left(\begin{array}{cc}
3622: 0 & \hat\ep_{\aa\bb}(\vp) \\
3623: \hat\ep^*_{\aa\bb}(-\vp) & 0 \\
3624: \end{array}\right) ,
3625: \eeq
3626: where
3627: \beq
3628: \hat\ep_{\aa\bb}(\vp)=
3629: m_{(\aa )}\de_{\aa \bb}\et(\vp)
3630: +\tilde\La_{\aa\bb}(\vp) .
3631: \eeq
3632: Using $\ep^{(1)}$ and $H^{(2)}_{\aaa \bbb}$,
3633: which depends on both $\ep^{(1)}$ and  $\ep^{(2)}$,
3634: we can find $\ep^{(2)}$ 
3635: and then continue iteratively to arbitrary order.
3636: 
3637: Under the transformation $\cal U$,
3638: the hamiltonian restricted to light neutrinos
3639: may be written
3640: \beq
3641: H_{\aaa \bbb}=
3642: \left(\begin{array}{cc}
3643: h_{\aaa \bbb}(\vp) & 0 \\
3644: 0 & -h_{\aaa \bbb}^*(-\vp)
3645: \end{array}\right) .
3646: \eeq
3647: Calculating $\cal U$ to second order in small quantities
3648: yields the second-order hamiltonian
3649: \beq
3650: h_{\aaa \bbb}(\vp)=
3651: \de_{\aaa \aa}\de_{\bbb \bb}
3652: \left[
3653: \left(|\vp|
3654: +\frac{1}{2|\vp|}m_{(\aa)}^2
3655: \right)
3656: \de_{\aa \bb}
3657: +\La_{\aa \bb}(\vp)
3658: \right] .
3659: \label{h}
3660: \eeq
3661: This expression neglects terms 
3662: that are second order in coefficients for Lorentz violation
3663: and terms that enter as the product of $m_{(\aa)}/|\vp|$
3664: with $\tilde\La$.
3665: The latter terms constitute subleading-order corrections
3666: under the reasonable assumption that $\La$ and $\tilde\La$ are
3667: comparable in size.
3668: 
3669: While formally the two bases related by $\cal U$ are different,
3670: in practice this difference is of little consequence.
3671: Our main goal is to determine oscillation probabilities.
3672: The effects of $\cal U$ appear in the mixing matrix and
3673: therefore modify the amplitudes of oscillations.
3674: However, 
3675: since $\cal U$ is close to the identity,
3676: the basis change produces only tiny and unobservable changes 
3677: in oscillation amplitudes.
3678: It therefore suffices in practice to assume ${\cal U} = I$
3679: for purposes of the basis transformation,
3680: corresponding to ignoring the difference between the
3681: $\aa$ and $\aaa$ indices.
3682: Similar arguments apply to the field redefinition
3683: relating $\nu$ and $\ch$.
3684: This also underlies the validity of assuming 
3685: unitarity mixing matrices in the conventional case with neutrino mass,
3686: even though the submatrix $V_{\aa a}$ is only approximately unitary.
3687: In contrast,
3688: the diagonalization of $h$ in Eq.\ \rf{h} 
3689: can introduce arbitrary amounts of mixing.
3690: 
3691: The above description in the mass-diagonal basis 
3692: completely determines the neutrino dynamics,
3693: but in practical situations a description 
3694: in the weak-interaction basis is more useful.
3695: This requires the transformation of $h_{\aa\bb}$ 
3696: to the original flavor basis.
3697: 
3698: The first step in implementing the desired transformation
3699: is to determine the relation between the coefficients
3700: in Eq.\ \rf{GaM} and those in Eq.\ \rf{GaMp}.
3701: In terms of the unitary matrix $V_{\AA A}$,
3702: we find
3703: \bea
3704: c^\mn_{\AA\BB}&=&\Re  V_{\AA A}V^*_{\BB B}(c+d)^\mn_{AB} ,
3705: \nonumber \\
3706: d^\mn_{\AA\BB}&=&i\Im  V_{\AA A}V^*_{\BB B}(c+d)^\mn_{AB} ,
3707: \nonumber \\
3708: e^\nu_{\AA\BB}&=&i\Im V_{\AA A}V_{\BB B}[(e+if)^\nu\cmat]_{AB} ,
3709: \nonumber \\
3710: if^\nu_{\AA\BB}&=&\Re V_{\AA A}V_{\BB B}[(e+if)^\nu\cmat]_{AB} ,
3711: \nonumber \\
3712: \half g^{\la\mn}_{\AA\BB}&=&
3713:  \Re V_{\AA A}V_{\BB B}\half(g^{\la\mn}\cmat)_{AB}
3714: \nonumber \\
3715:  &&-\Im V_{\AA A}V_{\BB B}\frac14\ep^{\la\mu\rh\si}
3716:  ({g_{\rh\si}}^\nu\cmat)_{AB} ,
3717: \nonumber \\
3718: m_{\AA\BB}&=&\Re V_{\AA A}V_{\BB B}[(m+im_5)\cmat]_{AB}
3719:  \equiv m_{(\AA)}\de_{\AA\BB} ,
3720: \nonumber \\
3721: im_{5\AA\BB}&=&i\Im V_{\AA A}V_{\BB B}[(m+im_5)\cmat]_{AB}
3722:  \equiv0 ,
3723: \nonumber \\
3724: a^\nu_{\AA\BB}&=&i\Im V_{\AA A}V^*_{\BB B}(a+b)^\nu_{AB} ,
3725: \nonumber \\
3726: b^\nu_{\AA\BB}&=&\Re V_{\AA A}V^*_{\BB B}(a+b)^\nu_{AB} ,
3727: \nonumber \\
3728: \half H^\mn_{\AA\BB}&=&
3729:   i\Im V_{\AA A}V_{\BB B}\half(H^\mn\cmat)_{AB}
3730: \nonumber \\
3731:   &&+i\Re V_{\AA A}V_{\BB B}\frac14\ep^{\mn\rh\si}
3732:   (H_{\rh\si}\cmat)_{AB} .
3733: \label{coeffs}
3734: \eea
3735: Note that all the coefficients in the mass-diagonal basis
3736: are either pure real or pure imaginary,
3737: reflecting the Majorana nature of neutrinos in this basis.
3738: Using this equation,
3739: we obtain
3740: \bea
3741: \lefteqn{
3742:   [(a+b)^\mu p_\mu-(c+d)^\mn p_\mu p_\nu]_{\aa\bb}}
3743: \nonumber \\
3744: &&=
3745: [(a+b)^\mu p_\mu-(c+d)^\mn p_\mu p_\nu]_{ab}
3746: V_{\aa a} V^*_{\bb b}\ ,
3747: \nonumber\\
3748: \lefteqn{
3749:   [(a-b)^\mu p_\mu-(c-d)^\mn p_\mu p_\nu]_{\aa\bb}}
3750: \nonumber \\
3751: &&=[-(a+b)^\mu p_\mu-(c+d)^\mn p_\mu p_\nu]^*_{ab}
3752: V^*_{\aa a} V_{\bb b}\ ,
3753: \nonumber\\
3754: \lefteqn{
3755:   -i\sqrt{2}p_\mu(\ep_+)_\nu[
3756:   g^{\mn\si}p_\si-H^\mn]_{\aa\bb}}
3757: \nonumber \\
3758: &&=-i\sqrt{2} p_\mu (\ep_+)_\nu[
3759: (g^{\mn\si}p_\si-H^\mn)\cmat]_{ab}
3760: V_{\aa a} V_{\bb b}\ ,
3761: \nonumber\\
3762: \lefteqn{
3763:   i\sqrt{2}p_\mu(\ep_+)^*_\nu[
3764:   g^{\mn\si}p_\si-H^\mn]_{\aa\bb}}
3765: \nonumber \\
3766: &&=i\sqrt{2}p_\mu(\ep_+)^*_\nu[
3767: (g^{\mn\si}p_\si+H^\mn)\cmat]^*_{ab}
3768: V^*_{\aa a} V^*_{\bb b}\ ,\quad\quad
3769: \eea
3770: using the assumption that the submatrix
3771: $V_{\aa a}$ is unitary.
3772: 
3773: Within a standard seesaw mechanism,
3774: the right-handed Majorana-mass matrix $R$ 
3775: appearing in Eq.\ \rf{usumass} is large, 
3776: $|R| \gg |L|,|D|$.
3777: Calculating the matrix $V_{AB}$ 
3778: at leading order in small mass ratios
3779: $|L|/|R|$ and $|D|/|R|$ produces the identity
3780: \beq
3781: m_{(\aa)}\de_{\aa\bb}=V_{\aa a}V_{\bb b}(\ml)_{ab} ,
3782: \eeq
3783: where $\ml= L - DR^{-1}D^T$,
3784: and hence the relation
3785: \beq
3786: m_{(\aa)}^2\de_{\aa\bb}
3787: =V_{\aa a}V^*_{\bb b}(\ml \ml^\dag)_{ab}
3788: =V^*_{\aa a}V_{\bb b}(\ml \ml^\dag)^*_{ab} .
3789: \eeq
3790: Combining results yields the desired form,
3791: \bea
3792: \lefteqn{
3793: \left[
3794: \left(|\vp|+\frac{1}{2|\vp|}m_{(\aa)}^2\right)
3795: \de_{\aa \bb}
3796: +\La_{\aa \bb}(\vp)
3797: \right]}
3798: \nonumber \\
3799: &&=\left(\begin{array}{cc}
3800: V_{\aa a} & 0 \\ 0 & V^*_{\aa a}
3801: \end{array}\right)
3802: (\heff)_{ab}
3803: \left(\begin{array}{cc}
3804: V^*_{\bb b} & 0 \\ 0 & V_{\bb b}
3805: \end{array}\right) ,
3806: \eea
3807: where $\heff$ is given in
3808: Eq.\ \rf{heff}.
3809: 
3810: 
3811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3812: 
3813: \phantom{xxxxxxxxxxxxxxxxxxxxxxxxx}
3814: 
3815: \section{Minimal SME terms}\label{smeterms}
3816: 
3817: Restricting attention to 
3818: the coefficients $(c_L)^\mn_{ab}$, $(a_L)^\mu_{ab}$,
3819: which are contained in the minimal gauge-invariant SME,
3820: effectively decouples neutrinos and antineutrinos
3821: and produces vanishing transition probabilities 
3822: \rf{pnb} and \rf{pbn}.
3823: This appendix describes a useful parametrization 
3824: of these coefficients.
3825: 
3826: Each coefficient matrix for Lorentz violation
3827: can be parametrized with three eigenvalues 
3828: and a constant unitary matrix.
3829: We define 
3830: \beq
3831: (c_L)^\mn=
3832: (\hat U^\mn)^\dag
3833: \left(\begin{array}{ccc}
3834: (c_L)^\mn_{(1)} & 0 & 0 \\
3835: 0 & (c_L)^\mn_{(2)} & 0 \\
3836: 0 & 0 & (c_L)^\mn_{(3)}
3837: \end{array}\right)
3838: \hat U^\mn 
3839: \eeq
3840: for each coefficient matrix $(c_L)^\mn$,
3841: and
3842: \beq
3843: (a_L)^\mu=
3844: (\hat U^\mu)^\dag
3845: \left(\begin{array}{ccc}
3846: (a_L)^\mu_{(1)} & 0 & 0 \\
3847: 0 & (a_L)^\mu_{(2)} & 0 \\
3848: 0 & 0 & (a_L)^\mu_{(3)}
3849: \end{array}\right)
3850: \hat U^\mu 
3851: \eeq
3852: for each coefficient matrix $(a_L)^\mu$.
3853: The unitary diagonalizing matrices
3854: $\hat U^\mn$, $\hat U^\mu$
3855: are chosen so that 
3856: if there is only a single nonvanishing coefficient matrix
3857: then $\Ueff$ in Eq.\ \rf{diaheff}
3858: takes the block-diagonal form
3859: \beq
3860: \Ueff=
3861: \left(\begin{array}{cc}
3862: \hat U & 0 \\
3863: 0 & \hat U^* \\
3864: \end{array}\right) .
3865: \eeq
3866: The reader is warned that the above decomposition
3867: is frame dependent,
3868: so neither the eigenvalues nor the mixing matrices
3869: behave as tensors under observer Lorentz transformations.
3870: We therefore advocate restricting this type of decomposition
3871: to the standard Sun-centered celestial equatorial frame.
3872: 
3873: Adopting a CKM-like decomposition of the $\hat U$ matrices,
3874: we denote mixing angles and phases
3875: associated with each $(c_L)^\mn$ by
3876: $\th_{12}^\mn$, 
3877: $\th_{13}^\mn$, 
3878: $\th_{23}^\mn$,
3879: and 
3880: $\de^\mn$, 
3881: $\be_1^\mn$, 
3882: $\be_2^\mn$.
3883: Similarly, 
3884: for each $(a_L)^\mu$ we write 
3885: $\th_{12}^\mu$, 
3886: $\th_{13}^\mu$, 
3887: $\th_{23}^\mu$,
3888: and 
3889: $\de^\mn$, 
3890: $\be_1^\mu$, 
3891: $\be_2^\mu$.
3892: The $\hat U$ matrices may then be written explicitly 
3893: in the form
3894: \begin{widetext}
3895: \bea
3896: \hat U^\mn&=&
3897: \left[\begin{array}{ccc}
3898: c^\mn_{12}c^\mn_{13} &
3899: -s^\mn_{12}c^\mn_{23}-c^\mn_{12}s^\mn_{23}s^\mn_{13}e^{-i\de^\mn} &
3900: s^\mn_{12}s^\mn_{23}-c^\mn_{12}c^\mn_{23}s^\mn_{13}e^{-i\de^\mn} \\
3901: s^\mn_{12}c^\mn_{13} &
3902: c^\mn_{12}c^\mn_{23}-s^\mn_{12}s^\mn_{23}s^\mn_{13}e^{-i\de^\mn} &
3903: -c^\mn_{12}s^\mn_{23}-s^\mn_{12}c^\mn_{23}s^\mn_{13}e^{-i\de^\mn} \\
3904: s^\mn_{13}e^{i\de^\mn} &
3905: s^\mn_{23}c^\mn_{13} &
3906: c^\mn_{23}c^\mn_{13} \\
3907: \end{array}\right]
3908: \left[\begin{array}{ccc}
3909: 1 & 0 & 0 \\
3910: 0 & e^{i\be_1^\mn} & 0 \\
3911: 0 & 0 & e^{i\be_2^\mn} \\
3912: \end{array}\right] , \\
3913: \hat U^\mu&=&
3914: \left[\begin{array}{ccc}
3915: c^\mu_{12}c^\mu_{13} &
3916: -s^\mu_{12}c^\mu_{23}-c^\mu_{12}s^\mu_{23}s^\mu_{13}e^{-i\de^\mu} &
3917: s^\mu_{12}s^\mu_{23}-c^\mu_{12}c^\mu_{23}s^\mu_{13}e^{-i\de^\mu} \\
3918: s^\mu_{12}c^\mu_{13} &
3919: c^\mu_{12}c^\mu_{23}-s^\mu_{12}s^\mu_{23}s^\mu_{13}e^{-i\de^\mu} &
3920: -c^\mu_{12}s^\mu_{23}-s^\mu_{12}c^\mu_{23}s^\mu_{13}e^{-i\de^\mu} \\
3921: s^\mu_{13}e^{i\de^\mu} &
3922: s^\mu_{23}c^\mu_{13} &
3923: c^\mu_{23}c^\mu_{13} \\
3924: \end{array}\right]
3925: \left[\begin{array}{ccc}
3926: 1 & 0 & 0 \\
3927: 0 & e^{i\be_1^\mu} & 0 \\
3928: 0 & 0 & e^{i\be_2^\mu} \\
3929: \end{array}\right] ,
3930: \eea
3931: \end{widetext}
3932: where
3933: $s^\mn_{ab}=\sin\th^\mn_{ab}$,
3934: $c^\mn_{ab}=\cos\th^\mn_{ab}$,
3935: $s^\mu_{ab}=\sin\th^\mu_{ab}$,
3936: and
3937: $c^\mu_{ab}=\cos\th^\mu_{ab}$.
3938: 
3939: In the conventional massive-neutrino analysis,
3940: the $\be$ matrix of phases can be absorbed 
3941: into the amplitudes $b_a(t;\vp)$ and $d_a(t;\vp)$, 
3942: so these phases are normally unobservable and can be neglected.
3943: However, 
3944: in the present context,
3945: only one set of $\be$ phases may be absorbed in this fashion.
3946: The presence of multiple coefficient matrices for Lorentz violation
3947: implies that they cannot typically be neglected.
3948: 
3949: Neutrino oscillations are insensitive to terms 
3950: in the effective hamiltonian that are proportional 
3951: to the identity.
3952: Consequently, 
3953: only two eigenvalue differences 
3954: for each coefficient matrix for Lorentz violation
3955: contribute to oscillation effects.
3956: Also, 
3957: each coefficient matrix is associated with
3958: three mixing angles and three phases.
3959: It follows that
3960: the maximum number of gauge-invariant degrees of freedom
3961: that enter into neutrino oscillations in the minimal SME alone
3962: is 16$\times$8 for $c_L$ and 4$\times$8 for $a_L$,
3963: for a total of 160.
3964: However,
3965: some of these are unobservable.
3966: The 8 trace components $\et_\mn (c_L)^\mn$ are Lorentz invariant,
3967: and both these and
3968: the 6$\times$8-component antisymmetric piece of $(c_L)^\mn$
3969: are absent in the leading-order hamiltonian \rf{heff}.
3970: This leaves 104 leading-order degrees of freedom 
3971: in $a_L$ and $c_L$,
3972: in agreement with the numbers listed in Table I.
3973: For the minimal SME, 
3974: one set of $\be$ phases is also unobservable,
3975: which reduces the total number of degrees of freedom to $102$.
3976: 
3977: 
3978: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3979: 
3980: \begin{thebibliography}{99}
3981: 
3982: \bibitem{cpt01}
3983: For recent overviews of various experimental 
3984: and theoretical approaches to Lorentz and CPT violation,
3985: see, for example,
3986: V.A.\ Kosteleck\'y, ed.,
3987: {\it CPT and Lorentz Symmetry II},
3988: World Scientific, Singapore, 2002.
3989: 
3990: \bibitem{ck} 
3991: D.\ Colladay and V.A.\ Kosteleck\'y,
3992: Phys.\ Rev.\ D {\bf 55}, 6760 (1997);
3993: {\bf 58}, 116002 (1998).
3994: 
3995: \bibitem{kpo}
3996: V.A.\ Kosteleck\'y and R.\ Potting,
3997: Phys.\ Rev.\ D {\bf 51}, 3923 (1995).
3998: 
3999: \bibitem{kle}
4000: V.A.\ Kosteleck\'y and R.\ Lehnert,
4001: Phys.\ Rev.\ D {\bf 63}, 065008 (2001).
4002: 
4003: \bibitem{owg}
4004: O.W.\ Greenberg,
4005: Phys.\ Rev.\ Lett.\ {\bf 89}, 231602 (2002);
4006: Phys.\ Lett.\ B {\bf 567}, 179 (2003).
4007: 
4008: \bibitem{ks}
4009: V.A.\ Kosteleck\'y and S.\ Samuel,
4010: Phys.\ Rev.\ D {\bf 39}, 683 (1989);
4011: {\bf 40}, 1886 (1989);
4012: Phys.\ Rev.\ Lett.\ {\bf 63}, 224 (1989);
4013: {\bf 66}, 1811 (1991).
4014: 
4015: \bibitem{kp}
4016: V.A.\ Kosteleck\'y and R.\ Potting,
4017: Nucl.\ Phys.\ B {\bf 359}, 545 (1991);
4018: Phys.\ Lett.\ B {\bf 381}, 89 (1996);
4019: Phys.\ Rev.\ D {\bf 63}, 046007 (2001);
4020: V.A.\ Kosteleck\'y, M.\ Perry, and R.\ Potting,
4021: Phys.\ Rev.\ Lett.\ {\bf 84}, 4541 (2000).
4022: 
4023: \bibitem{ncqed}
4024: S.M.\ Carroll \etal,
4025: Phys.\ Rev.\ Lett.\ {\bf 87}, 141601 (2001);
4026: Z.\ Guralnik, R.\ Jackiw, S.Y.\ Pi, and A.P.\ Polychronakos,
4027: Phys.\ Lett.\ B {\bf 517}, 450 (2001);
4028: C.E.\ Carlson, C.D.\ Carone, and R.F.\ Lebed,
4029: Phys.\ Lett.\ B {\bf 518}, 201 (2001);
4030: A.\ Anisimov, T.\ Banks, M.\ Dine, and M.\ Graesser,
4031: Phys.\ Rev.\ D {\bf 65}, 085032 (2002);
4032: I.\ Mocioiu, M.\ Pospelov, and R.\ Roiban,
4033: Phys.\ Rev.\ D {\bf 65}, 107702 (2002);
4034: M.\ Chaichian, M.M.\ Sheikh-Jabbari, and A.\ Tureanu,
4035: hep-th/0212259;
4036: J.L.\ Hewett, F.J.\ Petriello, and T.G.\ Rizzo,
4037: Phys.\ Rev.\ D {\bf 66}, 036001 (2002).
4038: 
4039: 
4040: \bibitem{qg}
4041: R.\ Gambini and J.\ Pullin,
4042: in Ref.\ \cite{cpt01};
4043: J.\ Alfaro, H.A.\ Morales-T\'ecotl, L.F.\ Urrutia,
4044: Phys.\ Rev.\ {\bf D66}, 124006 (2002);
4045: D.\ Sudarsky, L.\ Urrutia, and H.\ Vucetich,
4046: Phys.\ Rev.\ Lett.\ {\bf 89}, 231301 (2002);
4047: Phys.\ Rev.\ D {\bf 68}, 024010 (2003);
4048: G.\ Amelino-Camelia,
4049: Mod.\ Phys.\ Lett.\ A {\bf 17}, 899 (2002);
4050: Y.J.\ Ng,
4051: Mod.\ Phys.\ Lett.\ {\bf A18}, 1073 (2003);
4052: R.\ Myers and M.\ Pospelov,
4053: Phys.\ Rev.\ Lett.\ {\bf 90}, 211601 (2003);
4054: N.E.\ Mavromatos,
4055: hep-ph/0305215.
4056: 
4057:  
4058: \bibitem{fn}
4059: C.D.\ Froggatt and H.B.\ Nielsen,
4060: hep-ph/0211106.
4061: 
4062: \bibitem{bj}
4063: J.D.\ Bjorken,
4064: Phys.\ Rev.\ D {\bf 67}, 043508 (2003).
4065: 
4066: \bibitem{hadronexpt}
4067: KTeV Collaboration,
4068: H.\ Nguyen, in Ref.\ \cite{cpt01};
4069: OPAL Collaboration,
4070: R.\ Ackerstaff \etal,
4071: Z.\ Phys.\ C {\bf 76}, 401 (1997);
4072: DELPHI Collaboration,
4073: M.\ Feindt \etal,
4074: preprint DELPHI 97-98 CONF 80 (1997);
4075: BELLE Collaboration,
4076: K.\ Abe \etal,
4077: Phys.\ Rev.\ Lett.\ {\bf 86}, 3228 (2001);
4078: BaBar Collaboration,
4079: B.\ Aubert
4080: \etal, 
4081: hep-ex/0303043;
4082: FOCUS Collaboration,
4083: J.M.\ Link \etal, 
4084: Phys.\ Lett.\ B {\bf 556}, 7 (2003).
4085: 
4086: \bibitem{hadronth}
4087: D.\ Colladay and V.A.\ Kosteleck\'y,
4088: Phys.\ Lett.\ B {\bf 344}, 259 (1995);
4089: Phys.\ Rev.\ D {\bf 52}, 6224 (1995);
4090: Phys.\ Lett.\ B {\bf 511}, 209 (2001);
4091: V.A.\ Kosteleck\'y and R.\ Van Kooten,
4092: Phys.\ Rev.\ D {\bf 54}, 5585 (1996);
4093: O.\ Bertolami \etal,
4094: Phys.\ Lett.\ B {\bf 395}, 178 (1997);
4095: V.A.\ Kosteleck\'y,
4096: Phys.\ Rev.\ Lett.\ {\bf 80}, 1818 (1998);
4097: Phys.\ Rev.\ D {\bf 61}, 016002 (2000);
4098: {\bf 64}, 076001 (2001);
4099: N.\ Isgur \etal,
4100: Phys.\ Lett.\ B {\bf 515}, 333 (2001).
4101: 
4102: \bibitem{ccexpt}
4103: L.R.\ Hunter \etal,
4104: in 
4105: V.A.\ Kosteleck\'y, ed.,
4106: \it CPT and Lorentz Symmetry, \rm
4107: World Scientific, Singapore, 1999;
4108: D.\ Bear \etal,
4109: Phys.\ Rev.\ Lett.\ {\bf 85}, 5038 (2000);
4110: D.F.\ Phillips \etal,
4111: Phys.\ Rev.\ D {\bf 63}, 111101 (2001);
4112: M.A.\ Humphrey \etal,
4113: physics/0103068;
4114: Phys.\ Rev.\ A {\bf 62}, 063405 (2000);
4115: V.A.\ Kosteleck\'y and C.D.\ Lane,
4116: Phys.\ Rev.\ D {\bf 60}, 116010 (1999);
4117: J.\ Math.\ Phys.\ {\bf 40}, 6245 (1999).
4118: 
4119: \bibitem{spaceexpt}
4120: R.\ Bluhm \etal,
4121: Phys.\ Rev.\ Lett.\ {\bf 88}, 090801 (2002);
4122: hep-ph/0306190.
4123: 
4124: \bibitem{cane}
4125: F.\ Can\`e \etal,
4126: in preparation. 
4127: 
4128: \bibitem{eexpt}
4129: H.\ Dehmelt \etal,
4130: Phys.\ Rev.\ Lett.\ {\bf 83}, 4694 (1999);
4131: R.\ Mittleman \etal,
4132: Phys.\ Rev.\ Lett.\ {\bf 83}, 2116 (1999);
4133: G.\ Gabrielse \etal,
4134: Phys.\ Rev.\ Lett.\ {\bf 82}, 3198 (1999);
4135: R.\ Bluhm \etal,
4136: Phys.\ Rev.\ Lett.\ {\bf 82}, 2254 (1999);
4137: Phys.\ Rev.\ Lett.\ {\bf 79}, 1432 (1997);
4138: Phys.\ Rev.\ D {\bf 57}, 3932 (1998).
4139: 
4140: \bibitem{eexpt2}
4141: B.\ Heckel,
4142: in Ref.\ \cite{cpt01};
4143: L.-S.\ Hou, W.-T.\ Ni, and Y.-C.M.\ Li,
4144: Phys.\ Rev.\ Lett.\ {\bf 90}, 201101 (2003);
4145: R.\ Bluhm and V.A.\ Kosteleck\'y,
4146: Phys.\ Rev.\ Lett.\ {\bf 84}, 1381 (2000).
4147: 
4148: \bibitem{photonexpt}
4149: S.M.\ Carroll, G.B.\ Field, and R.\ Jackiw, 
4150: Phys. Rev. D {\bf 41}, 1231 (1990);
4151: V.A.\ Kosteleck\'y and M.\ Mewes,
4152: Phys.\ Rev.\ Lett.\ {\bf 87}, 251304 (2001).
4153: 
4154: \bibitem{photonth}
4155: R.\ Jackiw and V.A.\ Kosteleck\'y,
4156: Phys.\ Rev.\ Lett.\ {\bf 82}, 3572 (1999);
4157: C.\ Adam and F.R.\ Klinkhamer,
4158: Nucl.\ Phys.\ B {\bf 657}, 214 (2003);
4159: H.\ M\"uller, C.\ Braxmaier, S.\ Herrmann, 
4160: A.\ Peters, and C.\ L\"ammerzahl,
4161: Phys. Rev. D {\bf 67}, 056006 (2003);
4162: T.\ Jacobson, S.\ Liberati, and D.\ Mattingly,
4163: Phys.\ Rev.\ D {\bf 67}, 124011 (2003);
4164: V.A.\ Kosteleck\'y, M.\ Perry, and R.\ Lehnert,
4165: astro-ph/0212003;
4166: V.A.\ Kosteleck\'y and A.G.M.\ Pickering,
4167: Phys.\ Rev.\ Lett.\ {\bf 91}, 031801 (2003);
4168: R.\ Lehnert, 
4169: Phys.\ Rev\ D, in press (gr-qc/0304013);
4170: G.M.\ Shore, gr-qc/0304059.
4171: 
4172: \bibitem{cavexpt}
4173: J.\ Lipa \etal,
4174: Phys.\ Rev.\ Lett.\ {\bf 90}, 060403 (2003);
4175: H.\ M\"uller \etal,
4176: Phys.\ Rev.\ Lett.\ {\bf 91}, 020401 (2003).
4177: 
4178: \bibitem{km}
4179: V.A.\ Kosteleck\'y and M.\ Mewes,
4180: Phys.\ Rev.\ D {\bf 66}, 056005 (2002).
4181: 
4182: \bibitem{muons} 
4183: V.W.\ Hughes \etal,
4184: Phys.\ Rev.\ Lett.\ {\bf 87}, 111804 (2001);
4185: R.\ Bluhm \etal,
4186: Phys.\ Rev.\ Lett.\ {\bf 84}, 1098 (2000).
4187: 
4188: \bibitem{fc1} 
4189: S.\ Coleman and S.\ L.\ Glashow, 
4190: Phys.\ Rev.\ D {\bf 59}, 116008 (1999).
4191: 
4192: \bibitem{fc2} 
4193: V.\ Barger, S.\ Pakvasa, T.J.\ Weiler, and K.\ Whisnant, 
4194: Phys.\ Rev.\ Lett.\ {\bf 85}, 5055 (2000).
4195: 
4196: \bibitem{fc3} 
4197: J.N.\ Bahcall, V.\ Barger, and D.\ Marfatia,
4198: Phys.\ Lett.\ B {\bf 534}, 114 (2002).
4199: 
4200: \bibitem{fc4} 
4201: A.\ de Gouv\^ea, 
4202: Phys.\ Rev.\ D {\bf 66}, 076005 (2002).
4203: 
4204: \bibitem{fc5}
4205: I.\ Mocioiu and M.\ Pospelov, 
4206: Phys.\ Lett.\ B {\bf 537}, 114 (2002).
4207: 
4208: \bibitem{nu}
4209: V.A.\ Kosteleck\'y and M.\ Mewes, hep-ph/0308300.
4210: 
4211: \bibitem{bef}
4212: R.\ Brustein, D.\ Eichler, and S.\ Foffa,
4213: Phys.\ Rev.\ D {\bf 65}, 105006 (2002).
4214: 
4215: \bibitem{pdg}
4216: A survey of the available data is provided by 
4217: The Particle Data Group, K.\ Hagiwara \etal,
4218: Phys.\ Rev.\ D {\bf 66}, 010001 (2002).
4219: 
4220: \bibitem{katrin}
4221: KATRIN Collaboration, A.\ Osipowicz \etal, hep-ex/0109033.
4222: 
4223: \bibitem{bbdecay}
4224: See, for example, 
4225: P.\ Vogel in Ref.\ \cite{pdg}.
4226: 
4227: \bibitem{sn1987a}
4228: K.\ Hirata \etal, Phys.\ Rev.\ Lett.\ {\bf 58}, 1490 (1987);
4229: R.M.\ Bionta \etal, Phys.\ Rev.\ Lett.\ {\bf 58}, 1494 (1987).
4230: 
4231: \bibitem{msw}
4232: L.\ Wolfenstein, Phys.\ Rev.\ D {\bf 17}, 2369 (1978);
4233: S.\ Mikheev and A.\ Smirnov,
4234: Sov.\ J.\ Nucl.\ Phys.\ {\bf 42}, 913 (1986);
4235: Sov.\ Phys.\ JETP {\bf 64}, 4 (1986);
4236: Nuovo Cimento {\bf 9C}, 17 (1986).
4237: 
4238: \bibitem{sk}
4239: Super-Kamiokande Collaboration, S.\ Fukuda \etal,
4240: Phys.\ Rev.\ Lett.\ {\bf 81}, 1562 (1998);
4241: {\bf 82}, 2644 (1999); {\bf 85}, 3999 (2000).
4242: 
4243: \bibitem{homestake}
4244: B.T.\ Cleveland \etal, Ap.\ J.\ {\bf 496}, 505 (1998).
4245: 
4246: \bibitem{gallex}
4247: GALLEX Callaboration, W.\ Hampel \etal,
4248: Phys.\ Lett.\ B {\bf 447}, 127 (1999).
4249: 
4250: \bibitem{gno}
4251: GNO Collaboration, M.\ Altmann \etal,
4252: Phys.\ Lett.\ B {\bf 490}, 16 (2000).
4253: 
4254: \bibitem{sage}
4255: SAGE Collaboration, J.N.\ Abdurashitov \etal,
4256: J.\ Exp.\ Theor.\ Phys.\ {\bf 95}, 181 (2002).
4257: 
4258: \bibitem{sksol}
4259: Super-Kamiokande Collaboration, S.\ Fukuda \etal,
4260: Phys.\ Rev.\ Lett.\ {\bf 86}, 5651 (2001);  {\bf 86}, 5656 (2001);
4261: Phys.\ Lett.\ B {\bf 539}, 179 (2002).
4262: 
4263: \bibitem{sno}
4264: SNO Collaboration, Q.R.\ Ahmad \etal,
4265: Phys.\ Rev.\ Lett.\ {\bf 87}, 071301 (2001);
4266: {\bf 89}, 0110301 (2002); {\bf 89}, 0110302 (2002).
4267: 
4268: \bibitem{kamland}
4269: KamLAND Collaboration, K.\ Eguchi \etal,
4270: Phys.\ Rev.\ Lett.\ {\bf 90}, 021802 (2003).
4271: 
4272: \bibitem{lsnd}
4273: LSND Collaboration, C.\ Athanassopoulos \etal,
4274: Phys.\ Rev.\ Lett.\ {\bf 81}, 1774 (1998);
4275: LSND Collaboration, A. Aguilar \etal,
4276: Phys.\ Rev.\ D {\bf 64}, 112007 (2001).
4277: 
4278: \bibitem{k2k}
4279: K2K Collaboration, M.H.\ Ahn \etal,
4280: Phys.\ Rev.\ Lett.\ {\bf 90}, 041801 (2003).
4281: 
4282: \bibitem{nuphysics}
4283: For comprehensive reviews of
4284: conventional neutrino physics see, 
4285: for example,
4286: F.\ Boehm and P.\ Vogel,
4287: {\it Physics of Massive Neutrinos, 2nd ed.},
4288: Cambridge University Press, Cambridge, 1992;
4289: R.N.\ Mohapatra and P.B.\ Pal,
4290: {\it Massive Neutrinos in Physics
4291: and Astrophysics, 3rd ed.},
4292: World Scientific, Singapore, 2002.
4293: 
4294: \bibitem{nuphysics2}
4295: For recent reviews of theoretical aspects of neutrino
4296: masses and mixings see,
4297: for example,
4298: R.\ Mohapatra,
4299: hep-ph/0211252;
4300: S.M.\ Bilenky, C.\ Giunti, J.A.\ Grifols, and E.\ Masso,
4301: Phys.\ Rep.\ {\bf 379}, 69 (2003). 
4302: 
4303: \bibitem{seesaw}
4304: M.\ Gell-Mann, P.\ Ramond, and R.\ Slansky,
4305: in P.\ van Nieuwenhuizen and D.Z.\ Freedman, ed.,
4306: \it Supergravity, \rm
4307: North Holland, Amsterdam, 1979;
4308: T.\ Yanagida, 
4309: in O.\ Sawada and A.\ Sugamoto, eds.,
4310: \it Workshop on Unified Theory and the Baryon Number of the Universe,
4311: \rm
4312: KEK, Japan, 1979;
4313: R.\ Mohapatra and G.\ Senjanovi\'c,
4314: Phys.\ Rev.\ Lett.\ {\bf 44}, 912 (1980).
4315: A recent review of the seesaw literature with discussion of 
4316: alternatives to the standard form 
4317: is given in P.\ Langacker, hep-ph/0304053. 
4318: 
4319: \bibitem{cm}
4320: D.\ Colladay and P.\ McDonald,
4321: J.\ Math.\ Phys.\ {\bf 43}, 3554 (2002).
4322: 
4323: \bibitem{bek}
4324: M.S.\ Berger and V.A.\ Kosteleck\'y,
4325: Phys.\ Rev.\ D {\bf 65}, 091701(R) (2002).
4326: 
4327: \bibitem{kuop}
4328: For a review see, for example,
4329: T.K.\ Kuo and J.\ Pantaleone,
4330: Rev.\ Mod.\ Phys.\ {\bf 61}, 937 (1989).
4331: 
4332: \bibitem{ssm}
4333: J.N.\ Bahcall, M.H.\ Pinsonneault, and S.\ Basu,
4334: Astrophys.\ J.\ {\bf 555}, 990 (2001).
4335: 
4336: \bibitem{mattercpt}
4337: For a detailed discussion in the Lorentz-conserving case, 
4338: see, for example,
4339: M.\ Jacobson and T.\ Ohlsson,
4340: hep-ph/0305064.
4341: 
4342: \bibitem{mmb}
4343: G.\ Barenboim and J.\ Lykken, Phys.\ Lett.\ B {\bf 554}, 73 (2003).
4344: 
4345: \bibitem{fn1}
4346: The precise definitions of the Sun-centered frame and
4347: of a standard Earth-based frame 
4348: appropriate for terrestrial experiments
4349: are given in Ref.\ \cite{km},
4350: along with the form of the transformation relating them.
4351: 
4352: \bibitem{klap}
4353: V.A.\ Kosteleck\'y, C.D.\ Lane, and A.G.M.\ Pickering,
4354: Phys.\ Rev.\ D {\bf 65}, 056006 (2002).
4355: 
4356: \bibitem{bugey}
4357: B.\ Achkar \etal, Phys.\ Lett.\ B {\bf 374}, 243 (1996).
4358: 
4359: \bibitem{chooz}
4360: M.\ Apollonio \etal, Phys.\ Lett.\ B {\bf 466}, 415 (1999).
4361: 
4362: \bibitem{gosgen}
4363: G.\ Zacek \etal, Phys.\ Rev.\ D {\bf 34}, 2621 (1986).
4364: 
4365: \bibitem{pv} 
4366: F.\ Boehm \etal, Phys.\ Rev.\ D {\bf 64}, 112001 (2001).
4367: 
4368: \bibitem{e776} 
4369: L.\ Borodovsky \etal,
4370: Phys.\ Rev.\ Lett.\ {\bf 68}, 274 (1992).
4371: 
4372: \bibitem{ccfr}
4373: A.\ Romosan \etal,
4374: Phys.\ Rev.\ Lett.\ {\bf 78}, 2912 (1997).
4375: 
4376: \bibitem{chorus}
4377: CHORUS Collaboration, E.\ Eskut \etal,
4378: Phys.\ Lett.\ B {\bf 497}, 8 (2001).
4379: 
4380: \bibitem{nomad1}
4381: NOMAD Collaboration, P.\ Astier \etal,
4382: Nucl.\ Phys.\ B {\bf 611}, 3 (2001).
4383: 
4384: \bibitem{nomad2}
4385: NOMAD Collaboration, P.\ Astier \etal,
4386: hep-ex/0306037.
4387: 
4388: \bibitem{nutev}
4389: S.\ Avvakumov \etal,
4390: Phys.\ Rev.\ Lett.\ {\bf 89}, 011804 (2002).
4391: 
4392: \bibitem{karmen}
4393: KARMEN Collaboration, B.\ Armbruster \etal,
4394: Phys.\ Rev.\ D {\bf 65}, 112001 (2002).
4395: 
4396: \bibitem{miniboone}
4397: E.\ Church \etal,
4398: {\it A proposal for an experiment to measure
4399:   $\nu_\mu\to\nu_e$ oscillations and $\nu_\mu$ disappearance
4400:   at the Fermilab Booster: BooNE,}
4401: Fermilab Report No.\ FERMILAB-P-0898, 1997.
4402: 
4403: \bibitem{icarus}
4404: ICARUS Collaboration,
4405: {\it The ICARUS Experiment:
4406:   A Second-Generation Proton Decay
4407:   Experiment and Neutrino Observatory
4408:   at the Gran Sasso Laboratory,}
4409: LNGS-P28/2001, LNGS-EXP 13/89 add.1/01,
4410: ICARUS-TM/2001-03, 2001.
4411: 
4412: \bibitem{minos}
4413: MINOS Collaboration,
4414: {\it MINOS Technical Design Report},
4415: Fermilab Report No.\ NuMI-L-337, 1998.
4416: 
4417: \bibitem{opera}
4418: OPERA Collaboration,
4419: {\it OPERA: An appearance experiment to search
4420: $\nu_\mu\to\nu_\ta$ oscillations
4421: in the CNGS beam,}
4422: CERN/SPSC 2000-028, SPSC/P318,
4423: LNGS P25/2000, 2000.
4424: 
4425: \bibitem{mc}
4426: See, for example, 
4427: M.\ Honda \etal, Phys.\ Rev.\ D {\bf 52}, 4985 (1995).
4428: 
4429: \bibitem{efit}
4430: G.L.\ Fogli \etal, Phys.\ Rev.\ D {\bf 60}, 053006 (1999);
4431: P.\ Lipari and M.\ Lusignoli, Phys.\ Rev.\ D {\bf 60}, 013003 (1999).
4432: 
4433: \bibitem{fn2}
4434: This example exhibits maximal mixing at $\Th=90^\circ$,
4435: but it is straightforward to generate 
4436: $\Th$-dependent examples 
4437: with maximal mixing at $\Th=0^\circ, 180^\circ$.
4438: See Ref.\ \cite{nu}.
4439: 
4440: \bibitem{fn3}
4441: Note that this behavior may also result
4442: from direction dependence or
4443: from resonances in $\heff$ at intermediate energy scales.
4444: 
4445: \bibitem{snnugamma}
4446: M.J.\ Longo,
4447: Phys.\ Rev.\ D {\bf 36}, 3276 (1987);
4448: Phys.\ Rev.\ Lett.\ {\bf 60}, 173 (1988). 
4449: 
4450: \bibitem{snnu}
4451: W.D.\ Arnett and J.L.\ Rosner,
4452: Phys.\ Rev.\ Lett.\ {\bf 58}, 1906 (1987).
4453: 
4454: \end{thebibliography}
4455: 
4456: \end{document}
4457: 
4458: 
4459: