nucl-th0608059/nrf4.tex
1: \documentclass[twocolumn,showpacs,amsmath,amssymb,prc]{revtex4}
2: \bibliographystyle{unsrt}
3: 
4: \usepackage{graphicx}
5: \usepackage{dcolumn}
6: \usepackage{bm}
7: 
8: \begin{document}
9: 
10: \title{Theoretical Standard Model Rates of Proton to Neutron \\ 
11: Conversions Near Metallic Hydride Surfaces}
12: 
13: \author{A. Widom}
14: \affiliation{Physics Department, Northeastern University, Boston MA 02115}
15: \author{L. Larsen}
16: \affiliation{Lattice Energy LLC, 175 North Harbor Drive, Chicago IL 60601}
17: 
18: \begin{abstract}
19: The process of radiation induced electron capture by protons or deuterons producing new 
20: ultra low momentum neutrons and neutrinos may be theoretically described within the standard 
21: field theoretical model of electroweak interactions. For protons or deuterons  in the 
22: neighborhoods of surfaces of condensed matter metallic hydride cathodes, such conversions are 
23: determined in part by the collective plasma modes of the participating charged particles, 
24: e.g. electrons and protons or deuterons. The radiation energy required for such low energy nuclear 
25: reactions may be supplied by the applied voltage required to push a strong charged current across a 
26: metallic hydride surface employed as a cathode within a chemical cell. The electroweak rates 
27: of the resulting ultra low momentum neutron production are computed from these considerations.  
28: \end{abstract}
29: 
30: \pacs{12.15.Ji, 23.20.Nx, 23.40.Bw, 24.10.Jv, 25.30.-c}
31: 
32: \maketitle
33: 
34: \section{Introduction \label{intro}}
35: 
36: Excess heats of reaction have often been observed to be generated in  
37: the metallic hydride cathodes of certain electrolytic chemical cells. 
38: The conditions required for such observations include high electronic current 
39: densities passing through the cathode surface as well as high packing fractions 
40: of hydrogen or deuterium atoms within the metal. Also directly observed in 
41: such chemical cells are nuclear transmutations into elements {\em not} 
42: originally present prior to running a current through  and/or prior to 
43: applying a LASER light beam incident to the cathode 
44: surface\cite{Iwamura:1998,Violante:2002,Dash:2002,Miley:2005,Miley:1996,Miley:1997}. 
45: It seems {\em unlikely} that the direct cold fusion of two deuterons can be a 
46: requirement to explain at least many of such observations\cite{Pons:1989} because 
47: in many of these experiments, deuterons were initially absent.
48: For simplicity of presentation, we consider ``light water'' chemical cells 
49: in which deuterons are not to any appreciable degree present before the 
50: occurrence of heat producing nuclear transmutations.  
51: 
52: Nuclear transmutations in the work which follows are attributed to the creation and 
53: absorption of ultra low momentum neutrons as well as related production of neutrinos.  
54: Although other workers\cite{Mizuno:1998,Kozima:1998} have previously speculated on a 
55: central role for neutrons in such transmutations, they were unable to articulate a 
56: physically plausible mechanism that could explain high rates of neutron production 
57: under the stated experimental conditions. By contrast, in this work electrons are 
58: captured by protons all located in collectively oscillating "patches" on the metallic 
59: surface. Electrons are captured by protons all located in collectively oscillating 
60: ``patches'' on the metallic surface. Since the energy threshold for such a reaction is 
61: \begin{equation}
62: Q_{in}\approx 
63: \left\{M_{\rm n}-(M_{\rm p}+m)\right\}c^2 
64: \approx 0.78233\ {\rm MeV}, 
65: \label{intro1}
66: \end{equation}
67: one requires a significant amount of initial collective radiation energy to 
68: induce the proton into neutron conversion  
69: \begin{equation}
70: {\rm (radiation\ energy)}+e^- +p^+ \to n+\nu_e.
71: \label{intro2}
72: \end{equation}
73: The radiation energy may be present at least in part due the power absorbed 
74: at the surface of the cathode. If \begin{math} {\cal V}  \end{math} denotes 
75: the voltage difference between the metallic hydride and the electrolyte and 
76: if \begin{math} {\cal J}  \end{math} denotes the electrical current per unit 
77: area into the cathode from the electrolyte, then 
78: the power per unit cathode surface area dissipated into infrared heat 
79: radiation is evidently 
80: \begin{equation}
81: {\cal P}={\cal V}{\cal J}=e{\cal V}\tilde{\Phi },
82: \label{intro3}
83: \end{equation}
84: wherein \begin{math} \tilde{\Phi } \end{math} is the flux per unit area of electrons 
85: exiting the cathode into the electrolyte. Typical metallic hydride cathodes will 
86: exhibit soft surface photon radiation in much the same physical manner as a ``hot wire'' 
87: in a light bulb radiates light. For the case of chemical cell cathodes, there will be 
88: a frequency {\em upward conversion} from virtually DC cathode currents and voltages up 
89: to infrared frequency radiation. Such an upward frequency conversion requires  
90: high order electromagnetic interactions between electrons, protons and photons.  
91: 
92: \begin{figure}[bp]
93: \scalebox {0.5}{\includegraphics{ratefig1}}
94: \caption{A low order diagram for   
95: $e^- +p^+ \to n+\nu_e$ in the vacuum is exhibited. 
96: In condensed matter metallic hydrides, 
97: the amplitude must include radiative corrections to very 
98: high order in $\alpha $.}
99: \label{fig1}
100: \end{figure}
101: 
102: The purpose of this work is to estimate the total rates of the reaction 
103: Eq.(\ref{intro2}) in certain metallic hydride cathodes. The lowest order 
104: vacuum Feynman diagram for the proton to neutron conversion  
105: is shown in FIG. \ref{fig1}. For the case of the reactions in metallic 
106: hydrides, one must include radiative corrections to FIG. \ref{fig1} to 
107: very high order in the quantum electrodynamic coupling strength  
108: \begin{equation}
109: \alpha =\frac{e^2}{4\pi \hbar c}\approx 
110: 7.2973526 \times 10^{-3}.    
111: \label{intro4}
112: \end{equation}
113: The {\it W}-coupling in terms of the weak rotation angle 
114: \begin{math} \theta_W \end{math} will be taken to lowest order in 
115: \begin{equation}
116: \alpha_W =\frac{g^2}{4\pi \hbar c}=
117: \frac{\alpha }{\sin^2 \theta_W}\ .    
118: \label{intro5}
119: \end{equation}
120: Charge conversion reactions are weak due to the large mass 
121: \begin{math} M_W  \end{math} of the \begin{math} W^\pm \end{math}. 
122: The Fermi interaction constant, scaled by either the proton or 
123: electron masses, is determined\cite{Marshak:1969,Pokorski:2000} by    
124: \begin{eqnarray} 
125: G_F\approx \frac{\pi \alpha_W}{\sqrt{2}}\left(\frac{\hbar c}{M_W^2}\right),
126: \nonumber \\ 
127: \frac{G_FM_p^2}{\hbar c}\approx 1.02682\times 10^{-5},
128: \nonumber \\ 
129: \frac{G_Fm^2}{\hbar c}\approx  3.04563\times 10^{-12}.    
130: \label{intro6}
131: \end{eqnarray}  
132: In the work which follows, it will be shown how the weak proton 
133: to ultra low  momentum neutron conversions on metallic hydride surfaces 
134: may proceed at appreciable rates in spite of the small size of the Fermi weak 
135: coupling strength. 
136: 
137: An order of magnitude estimate can already be derived from a four 
138: fermion weak interaction model presuming a previously 
139: discussed\cite{Widom:2006} electron mass renormalization 
140: \begin{math} m\to \tilde{m}=\beta m  \end{math} due 
141: to strong local radiation fields. Surface electromagnetic modes excited 
142: by large cathode currents can add energy to a bare electron state 
143: \begin{math} e^- \end{math} yielding a mass renormalized heavy electron 
144: state \begin{math} \tilde{e}^- \end{math}, with  
145: \begin{equation} 
146: \tilde{m}=\beta m.
147: \label{est1} 
148: \end{equation}
149: The threshold value for the renormalized electron mass which allows 
150: for the reaction Eqs.(\ref{intro1}) and (\ref{intro2}) is   
151: \begin{equation}
152: \beta > \beta_0\approx 2.531.
153: \label{est2}
154: \end{equation}
155: For a given heavy electron-proton pair 
156: \begin{math} (\tilde{e}^- p^+) \end{math}, 
157: the transition rate into a neutron-neutrino pair may be estimated in the 
158: Fermi theory by  
159: \begin{eqnarray}
160: \Gamma_{(\tilde{e}^- p^+)\to n+\nu_e }\sim 
161: \left(\frac{G_Fm^2}{\hbar c}\right)^2\left(\frac{mc^2}{\hbar }\right)
162: (\beta-\beta_0)^2,
163: \nonumber \\ 
164: \Gamma_{(\tilde{e}^- p^+)\to n+\nu_e }\sim 
165: 9\times 10^{-24}\left(\frac{mc^2}{\hbar }\right)
166: (\beta-\beta_0)^2,
167: \nonumber \\ 
168: \Gamma_{(\tilde{e}^- p^+)\to n+\nu_e }\sim 
169: 7\times 10^{-4}\ {\rm Hz}\times (\beta-\beta_0)^2,
170: \label{est3}
171: \end{eqnarray}
172: If there are \begin{math} n_2 \sim 10^{16}/{\rm cm^2} \end{math} 
173: such \begin{math} (\tilde{e}^- p^+) \end{math} pairs per unit 
174: surface area within several atomic layers below the cathode surface, 
175: then the neutron production rate per unit surface area per unit time 
176: may be estimated by 
177: \begin{eqnarray} 
178: \varpi_2 \approx n_2\Gamma_{(\tilde{e}^- p^+)\to n+\nu_e }\ ,
179: \nonumber \\ 
180: \varpi_2\sim \left(\frac{10^{13}\ {\rm Hz}}{{\rm cm}^2}\right)
181: \times (\beta-\beta_0)^2 .
182: \label{est4}
183: \end{eqnarray}
184: Significantly above threshold, say 
185: \begin{math} \beta \sim 2\beta_0\sim 5 \end{math},  
186: the estimated rate 
187: \begin{math} \varpi_2\sim 10^{13}\ {\rm Hz}/{\rm cm}^2 \end{math} 
188: is sufficiently large so as to explain observed nuclear transmutations in 
189: chemical cells in terms of weak interaction transitions of 
190: \begin{math} (\tilde{e}^- p^+) \end{math} 
191: pairs into pairs into neutrons and neutrinos and the subsequent absorption 
192: of these ultra low momentum neutrons by local nuclei.
193: 
194: It is worthy of note that the total cross section for the scattering of neutrons 
195: with momentum \begin{math} p  \end{math} may be written 
196: \begin{math} \sigma_{tot}=(4\pi \hbar/p){\Im m}{\cal A}(p)\end{math} wherein the 
197: forward scattering amplitude for neutrons is 
198: \begin{math} {\cal A}(p) \end{math}\cite{Larsen:2005}. In the ultra-low momentum limit 
199: \begin{math} p\to 0 \end{math}, the cross section for neutron absorption associated 
200: with a complex scattering length 
201: \begin{math} b=\lim_{p\to 0} {\Im m}{\cal A}(p) \end{math} 
202: formally diverges. For a finite but large neutron wavelength 
203: \begin{math} \lambda  \end{math}, the mean free neutron absorption path 
204: length \begin{math} \Lambda  \end{math} corresponding to \begin{math} n_a \end{math} 
205: absorbers per unit volume obeys 
206: \begin{equation}
207: \Lambda ^{-1}=n_a\sigma_{tot}\approx 2n\lambda b. 
208: \label{Lambda1}
209: \end{equation}
210: Numerically, for \begin{math} n_a\sim 10^{22}/{\rm cm}^3  \end{math} neutron 
211: absorbers per unit volume with an imaginary part of the scattering length 
212: \begin{math} b\sim 10^{-13}\ {\rm cm}  \end{math}
213: and with ultra-low momentum neutrons formed with a wavelength of 
214: \begin{math} \lambda \sim 10^{-3}\ {\rm cm}  \end{math}, a neutron will move 
215: on a length scale of \begin{math} \Lambda \sim 10^{-6}\ {\rm cm} \end{math} 
216: before being absorbed. The externally detectible neutron flux into the laboratory 
217: from the cathode is thereby negligible\cite{Widom:2006}.
218: 
219: Similarly, there is a strong suppression 
220: of gamma-ray emission due to the absorption of such rays by the heavy surface 
221: electrons\cite{Larsen:2005}. The mean free path of a photon in a metallic 
222: condensed matter system is related to the conductivity 
223: \begin{math} \sigma  \end{math}; employing Maxwell's equations and the equations 
224: \begin{equation}
225: \Lambda_\gamma ^{-1}=\frac{\sigma }{c}=R_{vac}\sigma 
226: \approx 4\alpha \left(\frac{\pi }{3}\right)^{1/3}n^{2/3}\bar{l}
227: \label{Lambda2}
228: \end{equation}
229: wherein \begin{math} n \end{math} is the density per unit volume of heavy electrons 
230: on the cathode surface and \begin{math} \bar{l} \end{math} is the heavy electron 
231: mean free path. For hard photon energies in the range 
232: \begin{math} 0.5\ {\rm Megavolt}<(\hbar \omega_{\gamma }/e) <10\ {\rm Megavolt}  \end{math} 
233: we estimate \begin{math} n^{2/3}\sim 10^{15}/{\rm cm}^2 \end{math} and 
234: \begin{math} \bar{l}\sim 10^{-6}\ {\rm cm} \end{math} leading to the short gamma-ray 
235: mean free path \begin{math} L_\gamma \sim 3.4\times 10^{-8}\ {\rm cm}  \end{math}. 
236: In estimating such a small mean free path for hard gamma photon, let us remind the reader 
237: of the inadequate nature of single electron single photon scattering when making estimates
238: of photon mean free paths. For example, a single optical photon scatters off a single 
239: single non-relativistic electron with a Thompson cross section. For many electrons, 
240: if the Thompson cross section were applied then one might conclude that metals were transparent 
241: in the optical regime. The proper estimates of photon absorption rates {\em in condensed matter}  
242: employs the electrical conductivity.
243: 
244: In Sec.\ref{NS}, an exact expression is derived for the emission  
245: rate \begin{math} \varpi  \end{math} per unit time per unit volume for 
246: creating neutrinos. It is then argued, purely on the basis of 
247: conservation laws, that \begin{math} \varpi  \end{math} also represents 
248: the rate per unit time per unit volume of neutron production.
249: The rate \begin{math} \varpi  \end{math}, in Sec.\ref{CCF}, is expressed 
250: in terms of composite fields consisting of charged electrons and opposite  
251: charged {\it W}-bosons. The effective {\it W}-bosons for condensed matter 
252: systems may be written to a sufficient degree of accuracy in terms of 
253: Fermi weak interaction currents 
254: \begin{eqnarray}
255: {\cal I}^+_\mu =c\left(\bar{\psi}_n  \gamma_\mu (g_V-g_A\gamma_5)\psi_p\right),
256: \nonumber \\ 
257: {\cal I}^-_\mu =c\left(\bar{\psi}_p  \gamma_\mu (g_V-g_A\gamma_5)\psi_n\right), 
258: \label{intro7}
259: \end{eqnarray} 
260: wherein the Dirac matrices are defined in Sec.\ref{NS},  
261: \begin{math} \psi_p \end{math} and \begin{math} \psi_n \end{math} 
262: represent, respectively, the proton and neutron Dirac fields and the 
263: vector and axial vector coupling strengths are determined by 
264: \begin{eqnarray}
265: \lambda \equiv \frac{g_A}{g_V}\approx 1.2695,
266: \nonumber \\ 
267: \cos \theta_C\equiv g_V\approx 0.9742, 
268: \label{intro8}
269: \end{eqnarray} 
270: wherein \begin{math} \theta_C  \end{math} is a strong interaction 
271: quark rotation angle. In Sec.\ref{EMR}, the electron fields as  
272: renormalized by metallic hydride surface radiation are explored and   
273: the effective mass renormalization in Eq.(\ref{est1}) is 
274: established. In Sec.\ref{EPO}, we consider the coupled electron and 
275: proton oscillations near the surface of a metallic hydride.
276: 
277: \begin{figure}[tp]
278: \scalebox {0.6}{\includegraphics{ratefig2}}
279: \caption{The four fermion vertex for     
280: $e^- +p^+ \to n+\nu_e$ in the vacuum is exhibited. 
281: In the large $M_W$ limit, the Feynman diagram of 
282: FIG. \ref{fig1} collapses into the above Feynman diagram.
283: In condensed matter metallic hydrides, the resulting 
284: effective $W^\pm $ fields are defined in 
285: Eqs.(\ref{intro7}) and (\ref{CCF1W}).}
286: \label{fig2}
287: \end{figure}
288: 
289: In Sec.\ref{ISPW} the nature of the neutron production is 
290: discussed in terms of {\em isotopic spin waves}. In the limit 
291: in which the protons and neutrons are {\em non-relativistic}, 
292: one may view the proton and neutron as different isotopic spin 
293: states of a nucleon\cite{Cassin:1936} with the charged proton 
294: having an isotopic spin \begin{math} +1/2  \end{math} and with 
295: the neutron having an isotopic spin \begin{math} -1/2  \end{math}. 
296: If \begin{math} n({\bf r}) \end{math} and 
297: \begin{math} p({\bf r}) \end{math} 
298: represent, respectively, the two (real) spin component fields 
299: for non-relativistic neutrons and protons, then the operator 
300: isotopic spin density 
301: \begin{math} 
302: {\bf T}({\bf r})=
303: \big({\cal T}_1({\bf r}),{\cal T}_2({\bf r}),{\cal T}_3({\bf r}) \big) 
304: \end{math} of 
305: the many body neutron-proton states may be written 
306: \begin{eqnarray}
307: {\cal T}_1({\bf r})&=&\frac{1}{2}
308: \left(p^\dagger ({\bf r})n({\bf r})
309: +n^\dagger ({\bf r})p({\bf r})\right),
310: \nonumber \\ 
311: {\cal T}_2({\bf r})&=&\frac{i}{2}
312: \left(n^\dagger ({\bf r})p({\bf r})
313: -p^\dagger ({\bf r})n({\bf r})\right),
314: \nonumber \\ 
315: {\cal T}_3({\bf r})&=&\frac{1}{2}
316: \left(p^\dagger ({\bf r})p({\bf r})
317: -n^\dagger ({\bf r})n({\bf r})\right),
318: \nonumber \\ 
319: {\cal T}^\pm({\bf r})&=&T_1({\bf r})\pm iT_2({\bf r}).
320: \label{intro9}
321: \end{eqnarray}
322: In the non-relativistic limit, these isotopic spin operators determine 
323: the time-component of Fermi weak interaction currents in 
324: Eq.(\ref{intro7}) via 
325: \begin{equation}
326: {\cal I}^{\mp \ 0} ({\bf r})\approx cg_V{\cal T}^\pm({\bf r}).
327: \label{intro10}
328: \end{equation}
329: The remainder of the non-relativistic weak currents are of the 
330: Gamow-Teller variety\cite{Gamow:1936} and require the true spin as 
331: well as isotopic spin version of Eq.(\ref{intro10}); i.e. 
332: with \begin{math} {\bf S}={\bf \sigma}/2 \end{math} as the 
333: Fermion spin matrices, the combined 
334: spin and isotopic spin operator densities are
335: \begin{eqnarray}
336: {\cal S}_{1,j}({\bf r})&=&
337: \left(p^\dagger ({\bf r})S_j n({\bf r})
338: +n^\dagger ({\bf r})S_j p({\bf r})\right),
339: \nonumber \\ 
340: {\cal S}_{2,j}({\bf r})&=&i
341: \left(n^\dagger ({\bf r})S_jp({\bf r})
342: -p^\dagger ({\bf r})S_jn({\bf r})\right),
343: \nonumber \\ 
344: {\cal S}_{3,j}({\bf r})&=&
345: \left(p^\dagger ({\bf r})S_j p({\bf r})
346: -n^\dagger ({\bf r})S_j n({\bf r})\right),
347: \nonumber \\ 
348: {\cal S}^{\pm \ j}({\bf r})&=&{\cal S}_{1,j}({\bf r})
349: \pm i{\cal S}_{2,j}({\bf r}).
350: \label{intro11}
351: \end{eqnarray}
352: In the non-relativistic limit for the protons and neutrons, 
353: the spatial components of the weak interaction currents 
354: in Eq.(\ref{intro7}) are 
355: \begin{equation}
356: {\cal I}^{\mp \ j} ({\bf r})\approx 
357: -cg_A{\cal S}^\pm _j({\bf r}).
358: \label{intro12}
359: \end{equation}
360: Altogether, in the nucleon non-relativistic limit
361: \begin{equation}
362: {\cal I}^{\mp \ \mu} \approx 
363: c\left(-g_A{\cal S}^\pm _1,-g_A{\cal S}^\pm _2,
364: -g_A{\cal S}^\pm _3,g_V{\cal T}^\pm \right).
365: \label{intro13}
366: \end{equation}
367: The isotopic formalism describes the neutron creation 
368: as a surface isotopic spin wave.  
369: Out of many oscillating protons in a surface patch, 
370: only one of these protons will convert into a neutron. 
371: However, one must superimpose charge conversion 
372: amplitudes over all of the possibly converted protons 
373: in the patch. This describes an isotopic spin wave localized in the 
374: patch with wavelength \begin{math} k^{-1}  \end{math}. 
375: The wavelength in turn describes the ultra low momentum 
376: \begin{math} p\sim \hbar k  \end{math} of the produced 
377: neutron. Finally in the concluding Sec.\ref{conc}, further numerical  
378: estimates will be made concerning the weak interaction 
379: production rate of such neutrons.  
380: 
381: 
382: \section{Neutrino Sources \label{NS}}
383: 
384: The conventions here employed are as follows: The Lorentz metric 
385: \begin{math} \eta^{\mu \nu} \end{math} has the signature 
386: \begin{math} (+,+,+,-) \end{math} so that the Dirac matrix algebra 
387: may be written 
388: \begin{equation}
389: \gamma^\mu \gamma^\nu =-\eta^{\mu \nu}-i\sigma^{\mu \nu}
390: \ \ \ {\rm wherein}\ \ \ \sigma^{\mu \nu}=-\sigma^{\nu \mu }.
391: \label{NS1}
392: \end{equation}
393: The chiral matrix \begin{math} \gamma_5  \end{math} is 
394: defined with the antisymmetric symbol signature 
395: \begin{math} \epsilon_{1230}=+1 \end{math} employing   
396: \begin{equation}
397: \frac{1}{4!}\epsilon_{\mu \nu \lambda \sigma }\gamma^\mu \gamma^\nu 
398: \gamma^\lambda \gamma^\sigma =i\gamma_5   
399: \label{NS2}
400: \end{equation}
401: and chiral projection matrices are thereby 
402: \begin{equation}
403: P_{\pm }=\frac{1}{2}\left(1\mp \gamma_5\right).
404: \label{NS3}
405: \end{equation} 
406: Further algebraic matrix identities of use in the work below, 
407: such as 
408: \begin{eqnarray}
409: \gamma^\lambda \gamma^\mu \gamma^\sigma P_\pm =
410: \pm h^{\lambda \mu \sigma \nu}\gamma_\nu P_\pm \ ,
411: \ \ \ \ \ \ \ \ \ \ 
412: \nonumber \\ 
413: h^{\lambda \mu \sigma \nu}=i\epsilon^{\lambda \mu \sigma \nu}
414: -\eta^{\lambda \mu }\eta^{\sigma \nu }
415: + \eta^{\lambda \sigma }\eta^{\mu \nu }
416: -\eta^{\mu \sigma }\eta^{\lambda \nu } , 
417: \label{NSS}
418: \end{eqnarray}
419: all follow from Eqs.(\ref{NS1}), (\ref{NS2}) and (\ref{NS3}).
420: 
421: The average flux of left handed electron neutrinos (presumed massless) 
422: is determined by 
423: \begin{equation}
424: {\cal F}^\mu (x)=c\left<\bar{\nu }(x)\gamma^\mu P_+\nu (x)\right>.
425: \label{NS4}
426: \end{equation}  
427: Initial state averaging in Eq.(\ref{NS4}) is with respect to a chemical cell 
428: density matrix 
429: \begin{eqnarray}
430: \left<\ldots \right>\equiv Tr\ \rho \left(\ldots \right),
431: \nonumber \\ 
432: \rho =\sum_I p_I \left| I \right>\left< I \right|.
433: \label{NS4density}
434: \end{eqnarray}
435: The mean number of neutrinos created per unit time per unit volume may be 
436: computed from the four divergence of the neutrino flux; i.e.  
437: \begin{equation}
438: \varpi(x) =\partial_\mu  {\cal F}^\mu (x).
439: \label{NS5}
440: \end{equation}
441: Let us now argue, purely from standard model conservation laws, that 
442: \begin{math} \varpi  \end{math} is also the mean number of 
443: neutrons created per unit time per unit volume within the metallic hydride 
444: cathode in a chemical cell. 
445: 
446: If a neutrino is created, then {\em lepton number conservation} dictates that 
447: an electron had to be destroyed. If an electron is destroyed, then 
448: {\em charge conservation} dictates that a proton had to be destroyed. 
449: If a proton is destroyed, then {\em baryon number conservation} dictates that a neutron 
450: had to be created. Thus, the rate of neutrino creation must be equal to the rate 
451: of neutron creation. It is theoretically simpler to keep track 
452: of neutrino creation within the cathode.
453: 
454: The neutrino sinks and sources, respectively \begin{math} \bar{\eta}  \end{math} 
455: and \begin{math} \eta  \end{math}, are defined by that part of the standard model 
456: action which destroy and create neutrinos; i.e. 
457: \begin{equation}
458: S_{\rm int}=\hbar \int \left(\bar{\eta }(x)\nu (x)
459: +\bar{\nu}(x)\eta (x)\right)d^4x.
460: \label{NS6}
461: \end{equation} 
462: The neutrino field equations are thereby 
463: \begin{eqnarray}
464: -i\gamma^\mu \partial_\mu \nu (x)=\eta (x),
465: \nonumber \\ 
466: i \partial_\mu \bar{\nu }(x)\gamma^\mu =\bar{\eta }(x). 
467: \label{NS7}
468: \end{eqnarray} 
469: Eqs.(\ref{NS4}), (\ref{NS5}) and (\ref{NS7}) imply the neutrino 
470: creation rate per unit time per unit volume at space-time point 
471: \begin{math} x \end{math} has the form  
472: \begin{equation}
473: \varpi (x)=2c{\Im m}\left<\bar{\eta }(x)P_+\nu (x)\right>.
474: \label{NS8}
475: \end{equation}
476: Introducing the retarded massless Dirac Green's function,  
477: \begin{equation}
478: -i\gamma^\mu \partial_\mu S(x-y)=\delta (x-y),
479: \label{NS9}
480: \end{equation} 
481: allows us to solve the neutrino field Eqs.(\ref{NS7}) in the form 
482: \begin{equation}
483: \nu (x)=\nu_{in}(x)+\int S(x-y)\eta (y)d^4y,
484: \label{NS10}
485: \end{equation} 
486: wherein \begin{math} \nu_{in}(x) \end{math} represents the asymptotic 
487: incoming neutrino field. The assumption of {\em zero initial 
488: background neutrinos} is equivalent to the mathematical statement that  
489: the neutrino destruction operator   
490: \begin{math} \nu_{in}^+(x)\left|I\right>=0 \end{math} for the initial  
491: states in Eq.(\ref{NS4density}). 
492: In such a case, Eqs.(\ref{NS8}) and (\ref{NS10}) imply
493: \begin{equation} 
494: \varpi (x)=2c{\Im m}\int \left<\bar{\eta }(x)P_+S(x-y)\eta (y)\right>d^4y.
495: \label{NS11}
496: \end{equation}
497: 
498: The retarded massless Dirac Green's function may be found by looking 
499: for a solution of Eq.(\ref{NS9}) of the form 
500: \begin{equation}
501: S(x-y)=i\gamma^\mu \partial_\mu \Delta (x-y). 
502: \label{NS12}
503: \end{equation}
504: From Eqs.(\ref{NS9}) and (\ref{NS12}) it follows that 
505: \begin{equation}
506: -\partial_\mu \partial^\mu \Delta (x-y)=\delta (x-y).
507: \label{NS13}
508: \end{equation}
509: The retarded solution to Eq.(\ref{NS13}) requires the step function 
510: \begin{eqnarray}
511: \vartheta (x-y)=1\ \ {\rm if}\ \ x^0 > y^0,
512: \nonumber \\ 
513: \vartheta (x-y)=0\ \ {\rm if}\ \ x^0 < y^0;
514: \label{NS14}
515: \end{eqnarray}
516: In detail 
517: \begin{equation}
518: \Delta (x-y)=\frac{\vartheta (x-y)}{2\pi}
519: \ \delta \left((x-y)^2\right).
520: \label{NS15}
521: \end{equation}
522: Eqs.(\ref{NS11}) and (\ref{NS12}) imply 
523: \begin{eqnarray}
524: \varpi (x)=2c{\Re e}\int 
525: \left<\bar{\eta }(x)P_+\gamma^\mu \eta (y)\right>
526: \partial_\mu \Delta (x-y)d^4y,
527: \nonumber \\ 
528: \varpi (x)=2c{\Re e}\int \Delta (x-y)
529:  \left<\bar{\eta }(x)P_+\gamma^\mu \partial_\mu \eta (y)\right>d^4y.
530: \label{NS16}
531: \end{eqnarray}
532: The neutron production rate 
533: \begin{math} \varpi  \end{math} per unit time per unit volume 
534: can thus be computed in terms of the 
535: neutrino sinks \begin{math} \bar{\eta}  \end{math} and 
536: sources \begin{math} \eta  \end{math}.
537: 
538: 
539: 
540: \section{Composite Charged Fields \label{CCF}}
541: 
542: The neutrino sinks and sources of interest in this work 
543: can be written in terms of the composite fields of charged 
544: electrons and charged effective condensed matter 
545: \begin{math} W^\pm \end{math}-bosons; i.e. 
546: \begin{eqnarray}
547: \eta (y)=\frac{1}{\sqrt{2}}\gamma^\sigma W^+_\sigma (y) 
548: P_+\psi (y)\ ,
549: \nonumber \\ 
550: \bar{\eta }(x)=\frac{1}{\sqrt{2}}\bar{\psi }(x)
551: P_-\gamma^\lambda W^-_\lambda (x)\ ,
552: \label{CCF1}
553: \end{eqnarray}
554: in which \begin{math} \psi \end{math} and 
555: \begin{math} \bar{\psi}\end{math} are the Dirac electron fields 
556: and 
557: \begin{eqnarray}
558: W^+_\sigma (y)= \left(\frac{2\hbar G_F}{c^4}\right){\cal I}^+_\sigma (y)\ ,
559: \nonumber \\ 
560: W^-_\lambda (x)=\left(\frac{2\hbar G_F}{c^4}\right){\cal I}^-_\lambda (x)  \ .
561: \label{CCF1W}
562: \end{eqnarray}
563: 
564: The weak interaction proton-neutron charged conversion currents 
565: \begin{math} {\cal I}^\pm_\mu  \end{math} are defined in Eq.(\ref{intro7}). 
566: In Feynman diagram language, the amplitude pictured in FIG.\ref{fig1} has been 
567: replaced via a field current identity of the Fermi four field point 
568: interaction in FIG.\ref{fig2}.
569: Eqs.(\ref{NS16}) and (\ref{CCF1}) imply
570: \begin{equation}
571: \varpi (x)=c\ {\Re e}\int 
572: {\cal G}(x,y) \Delta (x-y)d^4y, 
573: \label{CCF2}
574: \end{equation}
575: wherein 
576: \begin{eqnarray}
577: {\cal G}(x,y) =
578: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
579: \nonumber \\ 
580: \left<W^-_\lambda (x)\bar{\psi }(x)\gamma^\lambda 
581: \gamma^\mu \gamma^\sigma P_+\partial_\mu \big(\psi (y)W^+_\sigma (y)\big)\right>=
582: \nonumber \\ 
583: h^{\lambda \mu \sigma \nu }\left<W^-_\lambda (x)\bar{\psi }(x)
584: \gamma_\nu \partial_\mu \big(P_+\psi (y)W^+_\sigma (y)\big)\right>.
585: \label{CCF3}
586: \end{eqnarray}
587: The neutron production rate \begin{math} \varpi \end{math} per unit time per 
588: unit volume implicit in Eqs.(\ref{NS15}), (\ref{CCF2}) and (\ref{CCF3}) may be 
589: considered to be exact. 
590: 
591: \section{Electron Mass Renormalization \label{EMR}}
592: 
593: \begin{figure}[tp]
594: \scalebox {0.6}{\includegraphics{ratefig3}}
595: \caption{In the presence of electromagnetic surface radiation, the charged 
596: particles in weak interaction must be described by wave functions which 
597: include to high order in $\alpha $ the effects of the electromagnetic 
598: fields. For the reaction at hand, both the proton and the electron react 
599: to surface radiation. The resulting mass renormalization is stronger for 
600: the electronic degrees of freedom than for the proton degrees of freedom. 
601: The density of states including radiation is computed employing 
602: Eqs.(\ref{EMR4}) and (\ref{EMR5}).}
603: \label{fig3}
604: \end{figure}
605: 
606: When the reacting charged particles  
607: \begin{math} e^- +p^+ \to n+\nu_e \end{math} 
608: are in the presence of surface plasmon radiation, then the external 
609: charged lines (incoming wave functions) must include the radiation 
610: fields to a high order in the quantum electrodynamic coupling 
611: strength\cite{Lifshitz:1997} \begin{math} \alpha  \end{math}. 
612: The situation is shown in FIG.\ref{fig3}.
613: To see what is involved, recall how one calculates the density of states 
614: for two particles incoming and two particles outgoing:
615: \par \noindent  
616: {\it Case I: The Vacuum} The density of states may be written as the four 
617: momentum conservation law, 
618: \begin{eqnarray}
619: \int e^{i(S_{0+}+S_{0-}-S_{0n}-S_{0\nu})/\hbar}d^4x= 
620: \nonumber \\ 
621: (2\pi \hbar )^4\delta (p_++p_--p_n-p_\nu ),
622: \label{EMR1}
623: \end{eqnarray} 
624: wherein \begin{math} p_+ \end{math}, \begin{math} p_- \end{math}, 
625: \begin{math} p_n \end{math} and \begin{math} p_\nu \end{math} represent,  
626: respectively, the four momenta of the proton, electron, neutron and neutrino. 
627: For a particle of mass \begin{math} m \end{math} in the vacuum, the action 
628: \begin{math} S_0(x) \end{math} obeys the Hamilton-Jacobi equation 
629: \begin{eqnarray}
630: \partial_\mu S_0(x) \partial^\mu S_0(x)+m^2c^2=0, 
631: \nonumber \\ 
632: S_0(x)=p\cdot x\equiv p_\mu x^\mu .
633: \label{EMR2}
634: \end{eqnarray}  
635: {\it Case II: Radiation} If the reaction takes place 
636: in the presence of electromagnetic radiation, 
637: \begin{equation}
638: F_{\mu \nu }=\partial_\mu A_\nu - \partial_\nu A_\mu ,
639: \label{EMR3}
640: \end{equation}
641: then the density of states conservation of four momenta must 
642: also include the electromagnetic radiation contribution; i.e.  
643: \begin{eqnarray}
644: {\Re e}\int e^{i(S_{+}+S_{-}-S_{n}-S_{\nu})/\hbar}d^4x= 
645: \nonumber \\ 
646: (2\pi \hbar )^4\tilde{\delta }(p_+,p_-,p_n,p_\nu ),
647: \label{EMR4}
648: \end{eqnarray} 
649: wherein, for a charged particle, the Hamilton-Jacobi equation
650: reads\cite{Landau:1975} 
651: \begin{eqnarray}
652: mv_\mu (x)=\partial_\mu S(x)-\frac{e}{c}A_\mu (x), 
653: \nonumber \\  
654: v_\mu (x)v^\mu (x)+c^2=0.
655: \label{EMR5} 
656: \end{eqnarray}
657: Therefore, in the density of states Eq.(\ref{EMR4}) 
658: including radiation the full solution of the Hamilton-Jacobi 
659: equation must be solved for all of the charged particles in the 
660: interaction. This constitutes the physical difference between the 
661: diagrams in the vacuum shown in FIG.\ref{fig2} and including radiation 
662: shown in FIG.\ref{fig3}. Under a gauge transformation 
663: \begin{math} A_\mu \to A_\mu  + \partial_\mu \chi \end{math}, 
664: the Hamilton-Jacobi Eq.(\ref{EMR5}) for a charged particle 
665: implies a change in the action 
666: \begin{math} S\to S+e\chi /c  \end{math}.
667: The renormalized charged particle wave function   
668: thereby exhibits the expected gauge transformation rule 
669: \begin{math} \psi \to \psi \exp(e\chi /\hbar c ) \end{math} 
670: making the complete amplitude from FIG.\ref{fig3} gauge invariant. 
671: The electron gauge contribution to the phase is canceled 
672: by the proton gauge contribution to the phase since the two particles 
673: are oppositely charged.  
674: 
675: The mass renormalization may be understood 
676: by averaging the local momentum 
677: \begin{math} p_\mu = \partial_\mu S \end{math} 
678: over local space time regions. Presuming 
679: \begin{math} \overline{p^\mu A_\mu} =0 \end{math} 
680: we have on average that  
681: \begin{equation}
682: -\overline{p_\mu p^\mu}=m^2c^2+
683: \left(\frac{e}{c}\right)^2 \overline{A_\mu A^\mu}
684: \equiv \tilde{m}^2c^2.
685: \label{EMR5a}
686: \end{equation}
687: The mass renormalization parameter in Eq.(\ref{est1}) is then 
688: given by 
689: \begin{equation}
690: \beta=\sqrt{1+\left(\frac{e}{mc^2}\right)^2 \overline{A_\mu A^\mu}}\ .
691: \label{EMR5b}
692: \end{equation}
693: Since the electron mass is much less than the proton 
694: mass, \begin{math} m\ll M_p  \end{math}, the main effects 
695: on low energy nuclear reactions are due to 
696: the mass renormalization of the surface electrons\cite{Widom:2006}. 
697: 
698: From the viewpoint of classical physics, the Lorentz force on a charge 
699: equation of motion, 
700: \begin{eqnarray}
701: mc\frac{dv^\mu }{d\tau } = eF^{\mu \nu }v_\nu ,
702: \nonumber \\ 
703: v^\mu = \left(\frac{\bf v}{\sqrt{1-(v/c)^2}},\frac{c}{\sqrt{1-(v/c)^2}}\right),
704: \label{EMR6}
705: \end{eqnarray}
706: is reduced to first order via the Hamilton-Jacobi Eq.(\ref{EMR5}), 
707: according to
708: \begin{equation}
709: \frac{dx^\mu }{d\tau }=v^\mu (x).
710: \label{EMR7}
711: \end{equation}
712: From the viewpoint of quantum mechanics, there is a one to one correspondence 
713: between quantum solutions of the Dirac equation and the classical solutions of 
714: Hamilton-Jacobi equation. In detail, the Dirac equation in an external radiation 
715: field,
716: \begin{equation}
717: -i\hbar \gamma^\mu 
718: \left\{\partial_\mu -i\left(\frac{e}{\hbar c}\right)A_\mu (x)\right\}\psi (x)
719: +mc\psi (x)=0,
720: \label{EMR8}
721: \end{equation} 
722: may be subject to a non-linear gauge transformation employing the solution to the 
723: classical Hamilton-Jacobi equation, 
724: \begin{equation}
725: \psi(x)= e^{iS(x)/\hbar }\Psi(x).
726: \label{EMR9}
727: \end{equation} 
728: The resulting radiation renormalized wave function obeys 
729: \begin{equation}
730: \gamma^\mu \left(-i\hbar \partial_\mu +mv_\mu (x)\right)\Psi (x)
731: +mc\Psi(x)=0.
732: \label{EMR10}
733: \end{equation}
734: It is worthy of note in the quasi-classical limit 
735: \begin{math} \hbar \to 0 \end{math} that the solution to 
736: the charged particle wave Eq.(\ref{EMR10}) is reduced to algebra. 
737: 
738: \section{Proton Oscillations \label{EPO}}
739: 
740: Thermal neutron scattering from  
741: hydrogen\cite{Dreismann:2005,Cowley:2006}, metallic 
742: hydrides\cite{Kenali:1999,Kolesnikova:2002,Glagolenkol:2002,Gidopoulos:2005}  
743: and even from protons embedded in proteins\cite{Medini:2003} has 
744: been of considerable recent theoretical and 
745: experimental interest\cite{Platzman:2004,Krzystyniak:2006}.
746: Employing the total cross section per unit volume as the extinction coefficient, 
747: \begin{equation}
748: h_{tot}=\frac{\sigma_{tot}}{\cal V},
749: \label{EPO1}
750: \end{equation} 
751: the intensity of a neutron beam when passed through a sample of thickness 
752: \begin{math} d \end{math} obeys 
753: \begin{equation}
754: I(d)=I(0)\exp(-h_{tot}d).
755: \label{EPO2}
756: \end{equation} 
757: The neutron-proton cross section is quite large since there is a very shallow bound 
758: state (the deuteron). For high densities of protons, the neutron-proton scattering 
759: will normally dominate the condensed matter scattering of thermal neutrons.  
760: Let the momentum transfer and energy transfer to the neutron 
761: be defined, respectively, as 
762: \begin{equation}
763: \hbar {\bf Q}={\bf p}_i-{\bf p}_f\ \ \ {\rm and}
764: \ \ \ \hbar \omega=\epsilon_i-\epsilon_f.
765: \label{EPO3}
766: \end{equation}
767: Decomposing the neutron scattering into various momementa and energies yields 
768: \begin{equation}
769: \frac{d^2 h_{i\to f}}{d\Omega_f d\epsilon_f}=\frac{\cal{N}}{\hbar {\cal V}}\left(\frac{p_f}{p_i}\right)
770: \frac{d \sigma_{i\to f}}{d\Omega_f}S({\bf Q},\omega ),
771: \label{EPO4}
772: \end{equation}
773: wherein \begin{math} d \sigma_{i\to f} \end{math} is the differential cross section for 
774: a single neutron to scatter off a single proton and the dynamical proton form factor for 
775: \begin{math} {\cal N}  \end{math} protons is defined as 
776: \begin{eqnarray}
777: {\cal N}S({\bf Q},\omega )=\sum_{j=1}^{\cal N}\int_{-\infty }^\infty e^{i\omega t}
778: F_j({\bf Q},t)\frac{dt}{2\pi }
779: \nonumber \\ 
780: F_j({\bf Q},t)=\left<e^{-i{\bf Q\cdot R}_j(t)}e^{i{\bf Q\cdot R}_j(0)}\right>
781: \label{EPO5}
782: \end{eqnarray}
783: In what follows we shall consider the self diffusion for a fixed proton at position 
784: \begin{math} {\bf R}_j(t)\equiv  {\bf R}(t) \end{math} and write 
785: \begin{equation}
786: \left<e^{-i{\bf Q\cdot R}(t)}e^{i{\bf Q\cdot R}(0)}\right>=\int_{-\infty }^\infty 
787: S({\bf Q},\omega )e^{-i\omega t}d\omega .
788: \label{EPO6}
789: \end{equation}
790: Under the assumption that the proton motions are non-relativistic, one easily shows for 
791: the many body Hamitonian \begin{math} {\cal H} \end{math} that  
792: \begin{equation}
793: {\cal H}_{\bf Q}=e^{-i{\bf Q\cdot R}}{\cal H}e^{i{\bf Q\cdot R}}=
794: {\cal H}+\hbar {\bf V\cdot Q}+\frac{\hbar^2 Q^2}{2M},
795: \label{EQO7}
796: \end{equation} 
797: wherein the velocity of the proton is 
798: \begin{equation}
799: {\bf V}=-i\left(\frac{\hbar }{M}\right){\bf \nabla}.
800: \label{EPO8}
801: \end{equation}
802: Thus 
803: \begin{eqnarray}
804: F({\bf Q},t)&=&\left<e^{-i{\bf Q\cdot R}(t)}e^{i{\bf Q\cdot R}(0)}\right>,
805: \nonumber \\
806: F({\bf Q},t)&=&\left<e^{i{\cal H}t}e^{-i{\bf Q\cdot R}}
807: e^{-i{\cal H}t}e^{i{\bf Q\cdot R}}\right>,
808: \nonumber \\
809: F({\bf Q},t)&=&\left<e^{i{\cal H}t/\hbar}e^{-i{\cal H}_{\bf Q}t/\hbar}\right>,
810: \nonumber \\
811: F({\bf Q},t)&=&e^{-i\omega_Qt}\left<e^{-i\int_0^t {\bf Q\cdot V}(s)ds}\right>_+ ,
812: \label{EPO9}
813: \end{eqnarray}
814: wherein ``+'' indicates time ordering, 
815: \begin{equation}
816: \omega_{\bf Q}=\frac{\hbar Q^2}{2M}\ \ \ {\rm and}
817: \ \ \ {\bf V}(s)=e^{i{\cal H}s/\hbar }{\bf V}e^{-i{\cal H}s/\hbar }.
818: \label{EPO10}
819: \end{equation}
820: From Eqs.(\ref{EPO6}), (\ref{EPO9}) and (\ref{EPO10}) one derives the sum rules 
821: \begin{eqnarray}
822: \int_{-\infty}^\infty S({\bf Q},\omega )d\omega &=&1,
823: \nonumber \\
824: \int_{-\infty}^\infty \omega S({\bf Q},\omega )d\omega &=&\omega_{\bf Q},
825: \label{EPO11}
826: \end{eqnarray}
827: which represent, respectively, probability normalization and the fact that the mean 
828: recoil energy is the same as would have been computed for a single free proton.
829: We can now discuss the experimental measurements which make proton oscillations 
830: unusual. 
831: 
832: \subsection{Deep Inelastic Neutron Scattering} 
833: 
834: For energetic neutrons with relatively 
835: high momentum transfer, it was estimated that the neutron scattering event 
836: lasted on a time scale of an attosecond. It was thought that the velocity could not appreciably change in so short a time. Therefore one might try the {\em impulse} approximation wherein Eq.(\ref{EPO9}) could be estimated by the 
837: \begin{math} t\to 0  \end{math} limit   
838: \begin{eqnarray}
839: F_{\rm impulse}({\bf Q},t)=e^{-i\omega_Qt}\left<e^{-i {\bf Q\cdot V}t}\right>,
840: \nonumber \\
841: S_{\rm impulse}({\bf Q},\omega )=
842: \int \delta (\omega -{\bf Q\cdot V}-\omega_{\bf Q})p({\bf V})d^3{\bf V},
843: \label{EPO12}
844: \end{eqnarray}
845: wherein \begin{math} p({\bf V})d^3{\bf V} \end{math} is the probability that the proton 
846: velocity is in the range \begin{math} {\bf V}\in d^3{\bf V} \end{math}.
847: Eq.(\ref{EPO12}) is at the heart of the theoretical analysis of so-called ``deep inelastic'' 
848: neutron scattering. Note that the impulse approximation obeys the sum rules in 
849: Eq.(\ref{EPO11}). It came as somewhat of a shock that for protons inside liquid hydrogen, 
850: metallic hydrides and proteins, the impulse approximation does not work. At issue was the 
851: ``violation'' of the probability sum rule
852: \begin{equation}
853: \int_{\rm experimental}S(Q,\omega )d\omega <1.
854: \label{EPO13}
855: \end{equation}
856: Since Eqs.(\ref{EPO1}) and (\ref{EPO2}) were experimentally verified, the only reason for 
857: the ``experimental probability loss'' must have been that the integral over experimental 
858: high frequency data was missing a physical low frequency regime which persists and 
859: contributes to the probability integral. Such low frequency modes must be collective.
860: 
861: \subsection{Recoilless Fraction} 
862: 
863: Let us at first assume the cluster decomposition property 
864: of \begin{math} F({\bf Q},t) \end{math}; i.e. 
865: \begin{eqnarray}
866: \lim_{t\to \infty}F({\bf Q},t)=F_{\infty}({\bf Q}),
867: \nonumber \\ 
868: \lim_{t\to \infty}\left<e^{-i{\bf Q\cdot R}(t)}e^{i{\bf Q\cdot R}(0)}\right>
869: =\left|\left<e^{i{\bf Q\cdot R}}\right>\right|^2, 
870: \nonumber \\ 
871: F_{\infty}({\bf Q})=\left|\left<e^{i{\bf Q\cdot R}}\right>\right|^2. 
872: \label{EPO14}
873: \end{eqnarray}
874: If cluster decomposition holds true, then \begin{math} F_{\infty }({\bf Q}) \end{math}
875: represents the {\em recoilless fraction} in the sense that 
876: \begin{eqnarray}
877: S({\bf Q},\omega )=F_{\infty}({\bf Q})\delta (\omega )+\tilde{S} ({\bf Q},\omega )
878: \nonumber \\ 
879: \int_{-\infty}^\infty \tilde{S}({\bf Q},\omega )d\omega =1-F_{\infty}({\bf Q})<1. 
880: \label{EPO15}
881: \end{eqnarray}
882: Since the second of Eqs.(\ref{EPO15}) yield the fraction that recoils, the experimental 
883: puzzle in Eq.(\ref{EPO13}) is then resolved. 
884: 
885: The recoilless fraction \begin{math}  F_{\infty}({\bf Q}) \end{math} for the proton 
886: is the same factor as the recoilless fraction \begin{math}  F_{\infty}({\bf Q}) \end{math} 
887: that appears in the M\"ossbauer effect for the gamma ray decay in a heavy nucleus.  
888: In the case of the M\"ossbauer effect, the recoil of the gamma emission is taken 
889: up by the crystal as a whole.  For the case of deep inelastic neutron scattering, 
890: the proton recoil must be taken up by neighboring electrons and other protons in 
891: coherent oscillation. Let \begin{math} P({\bf u})d^3 {\bf u} \end{math} be the 
892: probability that the displacement of a proton from its equilibrium position
893: \begin{math} {\bf u}={\bf R}-\big<{\bf R}\big>  \end{math} is in the range 
894: \begin{math} {\bf u}\in d^3 {\bf u} \end{math}. The recoilless fraction in 
895: Eq.(\ref{EPO14}) may be written as 
896: \begin{equation}
897: F_{\infty}({\bf Q})=
898: \left|\int P({\bf u})e^{i{\bf Q\cdot u}}d^3{\bf u}\right|^2.
899: \label{EPO16}
900: \end{equation}
901: The mean square fluctuations in displacement as defined by 
902: \begin{equation}
903: \overline{\bf uu}=\int P({\bf u}) {\bf uu} d^3{\bf u},
904: \label{EPO17}
905: \end{equation}  
906: can be measured via the low momentum transfer limit 
907: \begin{equation}
908: \lim_{Q\to 0}\frac{\overline{|{\bf Q\cdot u}|^2}}{Q^2}
909: =-\lim_{Q\to 0}\frac{\ln F_{\infty}({\bf Q})}{Q^2}.
910: \label{EPO18}
911: \end{equation}  
912: For higher momentum transfers, the fractal dimensional spectator 
913: model\cite{Medini:2003} gives a fairly good representation of the 
914: recoilless fraction; i.e.  
915: \begin{equation}
916: F_{\infty}({\bf Q};D)=
917: \left[1+\frac{\overline{|{\bf Q\cdot u}|^2}}{D}\right]^{-D}.
918: \label{EPO19}
919: \end{equation}
920: The harmonic oscillation result is the Gaussian probability Debye-Waller factor  
921: \begin{equation}
922: \lim_{D\to \infty }F_{\infty}({\bf Q};D)=
923: \exp\left(-\overline{|{\bf Q\cdot u}|^2}\right).
924: \label{EPO20}
925: \end{equation}
926: At elevated temperatures, a small value of \begin{math} D \end{math} describes the 
927: fractal dimension of the probability distribution \begin{math} P({\bf u})d^3 {\bf u} \end{math}.
928: Finally, for the case of collective proton oscillations, the delta function in 
929: Eq.(\ref{EPO15}) is somewhat idealized, In experimental practice, the delta function 
930: is broadened into an infrared peak consistant in thermal equilibrium with the detailed 
931: balance condition 
932: \begin{equation}
933: S(-{\bf Q},-\omega )=e^{-\hbar \omega /k_BT}S({\bf Q},\omega ).
934: \label{EPO21}
935: \end{equation}  
936: 
937: \subsection{Proton Displacements} 
938: 
939: Neutron scattering experiments\cite{Kenali:1999} on 
940: palladium hydride at moderate momentum transfer clearly 
941: indicate a sharply defined collective oscillation peak at 
942: \begin{math} (\hbar \Omega /e)\approx 60\ {\rm millivolt} \end{math}. 
943: Such a collective proton oscillation at an infrared frequency will resonate with 
944: electronic surface plasmon oscillations of the electrons leading to the local 
945: breakdown of the Born-Oppenheimer approximation and large collective proton oscillation 
946: amplitudes. Some theoretical insights into the nature of such oscillations can be obtained 
947: from sum rules. The mobility tensor of a single proton at complex frequency may be defined 
948: via the Kubo formula 
949: \begin{equation}
950: {\sf m}(\zeta )=\frac{i}{\hbar }\int_0^\infty 
951: \left<\left[{\bf V}(t),{\bf R}(0)\right]\right>e^{i\zeta t}dt,
952: \ \ \ {\Im m}(\zeta )>0. 
953: \label{EPO22}
954: \end{equation}
955: The mobility tensor obeys the dispersion relation 
956: \begin{equation}
957: {\sf m}(\zeta )=\left(\frac{-2i\zeta }{\pi }\right)
958: \int_0^\infty \frac{\Re e{\sf m}(\omega +i0^+)d\omega }{\omega^2-\zeta^2}.
959: \label{EPO23}
960: \end{equation}
961: As the complex frequency \begin{math} |\zeta |\to \infty  \end{math} in the 
962: upper half plane, the mobility tensor obeys the asymptotic form 
963: \begin{eqnarray}
964: {\sf m}(\zeta )=
965: \left(\frac{2i}{\pi \zeta }\right)\int_0^\infty \Re e{\sf m}(\omega +i0^+)d\omega +
966: \nonumber \\ 
967: \left(\frac{2i}{\pi \zeta^3 }\right)\int_0^\infty 
968: \omega^2 \Re e{\sf m}(\omega +i0^+)d\omega +\ldots 
969: \label{EPO24}
970: \end{eqnarray}
971: The same asymptotic expansion can be obtained by continually integrating 
972: Eq.(\ref{EPO22}) by parts yielding the sum rules 
973: \begin{eqnarray}
974: \frac{2}{\pi }\int_0^\infty \Re e{\sf m}(\omega +i0^+)d\omega =
975: \frac{i}{\hbar }\left<\left[{\bf V},{\bf R}\right]\right>,
976: \nonumber \\ 
977: \frac{2}{\pi }\int_0^\infty \omega^2 \Re e{\sf m}(\omega +i0^+)d\omega =
978: \frac{i}{\hbar }\left<\left[{\bf A},{\bf V}\right]\right>,
979: \label{EPO25}
980: \end{eqnarray}
981: wherein the proton velocity is \begin{math} {\bf V}=\dot{\bf R}  \end{math} 
982: and the proton acceleration is \begin{math} {\bf A}=\dot{\bf V}  \end{math}.
983: Employing the {\em non-relativistic proton} Eqs.(\ref{EPO8}), (\ref{EPO25}) and 
984: Newton's law \begin{math} M{\bf A}=|e|{\bf E} \end{math}, we find that
985: \begin{eqnarray}
986: \frac{2}{\pi }\int_0^\infty \Re e{\sf m}(\omega +i0^+)d\omega =
987: \frac{\sf 1}{M},
988: \nonumber \\ 
989: \frac{2}{\pi }\int_0^\infty \omega^2 \Re e{\sf m}(\omega +i0^+)d\omega =
990: -\frac{|e|}{M^2}\left<{\bf \nabla E}\right>.
991: \label{EPO26}
992: \end{eqnarray}
993: Employing Gauss'law at the proton position 
994: \begin{math} \big<div{\bf E}\big>=-|e|\tilde{n} \end{math} 
995: wherein \begin{math} \tilde{n} \end{math} is the electron density right on top 
996: of the proton, 
997: \begin{equation}
998: \tilde{n}=\left<\sum_k \delta ({\bf R}-{\bf r}_k)\right>
999: =\left<\psi^\dagger ({\bf R})\psi ({\bf R})\right>,
1000: \label{EPO27}
1001: \end{equation}
1002: the sum rule estimate for the infrared frequency peak in the neutron scattering is 
1003: thereby
1004: \begin{eqnarray}
1005: \Omega^2=\frac{\int_0^\infty \omega^2 \Re e\ tr\ {\sf m}(\omega +i0^+)d\omega }
1006: {\int_0^\infty \Re e\ tr\ {\sf m}(\omega +i0^+)d\omega }\ ,
1007: \nonumber \\ 
1008: \Omega^2=\frac{e^2\tilde{n}}{3M}=
1009: \frac{4\pi }{3}\left(\frac{\hbar^2}{Mm}\right)\frac{\tilde{n}}{a}\ .
1010: \label{EPO28}
1011: \end{eqnarray}
1012: In Eq.(\ref{EPO28}), \begin{math} a \end{math} is the Bohr radius. 
1013: In order to understand more clearly our estimate of the proton oscillation 
1014: frequency \begin{math} \Omega  \end{math}, suppose that the proton was embedded 
1015: in a sphere with charge density \begin{math} -|e|\tilde{n}  \end{math}. If the 
1016: proton is pulled out by a displacement \begin{math} {\bf u} \end{math}, then an 
1017: electric field would exist  
1018: \begin{equation}
1019: |e|{\bf E}=-\left(\frac{e^2\tilde{n}}{3}\right){\bf u}=-M\Omega^2{\bf u}
1020: \label{EPO29}
1021: \end{equation}
1022: pushing the proton back to the sphere center. The equation of motion 
1023: \begin{math}M\ddot{\bf u}=-|e|{\bf E}=-M\Omega^2{\bf u} \end{math} yields 
1024: the required oscillation. One may also note the identity 
1025: \begin{equation}
1026: e^2\overline{|{\bf E}|^2}=M^2\Omega^4\overline{|{\bf u}|^2}
1027: \label{EPO30}
1028: \end{equation}
1029: relating the meansquare electric field at the proton position to 
1030: the mean field displacement fluctuation. Suppose a momentum for a 
1031: relativistic electron given by 
1032: \begin{math} {\bf p}=m{\bf v}/\sqrt{1-|{\bf v}/c|^2}  \end{math}.
1033: The energy \begin{math} K \end{math} of such an electron may 
1034: then be written 
1035: \begin{equation}
1036: \overline{K^2}=m^2c^4+\overline{|{\bf p}|^2}c^2
1037: \label{EPO31}
1038: \end{equation}
1039: equivalent to 
1040: \begin{equation}
1041: \overline{K^2}=m^2c^4+\frac{e^2\overline{|{\bf E}|^2}c^2}{\Omega^2} 
1042: =m^2c^4\left[1+\frac{\overline{|{\bf E}|^2}}{{\cal E}^2}\right].
1043: \label{EPO32}
1044: \end{equation}
1045: wherein the electric field scale has been introduced 
1046: \begin{equation}
1047: {\cal E}=\frac{mc}{\hbar}\left(\frac{\hbar \Omega}{e}\right).
1048: \label{EPO32a}
1049: \end{equation}
1050: 
1051: \begin{figure}[tp]
1052: \scalebox {0.7}{\includegraphics{ratefig4}}
1053: \caption{The predicted electron mass enhacemnt $\beta =\tilde{m}/m$ when 
1054: protons are absorbed into palladium is plotted as a function of the root 
1055: mean square proton displacement $u$ where $u^2=\overline{|{\bf u}|^2}$ and  
1056: the Bohr radius $a\approx 0.5291772108\times 10^{-8}$ cm.}
1057: \label{fig4}
1058: \end{figure}
1059: 
1060: Experimentally\cite{Kenali:1999}, for protons absorbed in Pd, 
1061: one finds the soft Boson mode with  
1062: \begin{math} (\hbar \Omega /e)\approx 6\times 10^{-2}\ {\rm volts} \end{math} 
1063: wherein 
1064: \begin{equation}
1065: {\cal E}_{\rm p \in Pd}\approx 1.55\times 10^{11}\ {\rm volts/meter}.
1066: \label{EQO32b}
1067: \end{equation}
1068: In this regard, one notes that typical atomic physics electric fields in 
1069: atoms are of that order when located a Bohr radius 
1070: \begin{math} a \end{math} from an isolated proton; i.e. with the Bohr 
1071: radius  
1072: \begin{equation}
1073: a=\frac{4\pi \hbar^2}{e^2m}=
1074: \frac{\hbar }{\alpha mc}\approx 0.5291772108\times 10^{-8}\ {\rm cm},
1075: \label{EQO36}
1076: \end{equation}
1077: atomic electric fields are of the order 
1078: \begin{equation}
1079: {\cal E}_a = \frac{|e|}{4\pi a^2}\approx 
1080: 5.142206318\times 10^{11} \ {\rm volts/meter}.
1081: \label{EQO32c}
1082: \end{equation}
1083: Employing Eqs.(\ref{EPO28}), (\ref{EPO30}) and (\ref{EQO32c})  
1084: along with the estimate 
1085: \begin{math} \tilde{n}\approx |\psi(0)|^2=1/(\pi a^3) \end{math}  
1086: yields  
1087: \begin{equation}
1088: \frac{\overline{|{\bf E}|^2}}{{\cal E}_a^2}=\frac{16}{9}
1089: \left(\frac{\overline{|{\bf u}|^2}}{a^2}\right).
1090: \label{EQO33}
1091: \end{equation}
1092: Thus, Eq.(\ref{EMR5b}) may be written\cite{Widom:2006} as 
1093: \begin{eqnarray}
1094: \beta=\sqrt{\frac{\overline{K^2}}{m^2c^4}}
1095: =\sqrt{1+\left[\frac{e^2\overline{|{\bf E}|^2}}{m^2c^2\Omega^2}\right]}\ ,
1096: \nonumber \\ 
1097: \beta =\frac{\tilde{m}}{m}\approx
1098: \sqrt{1+\frac{16{\cal E}_a^2}{{9{\cal E}}_{\rm p}^2}
1099: \left(\frac{\overline{|{\bf u}|^2}}{a^2}\right)}\ .
1100: \label{EPO33}
1101: \end{eqnarray}
1102: For example, for protons in palladium Eqs.(\ref{EQO32b}) and (\ref{EQO32c}) 
1103: imply
1104: \begin{equation}
1105: \beta_{\rm p \in Pd}\approx \sqrt{1+19.6
1106: \left(\frac{\overline{|{\bf u}|^2}}{a^2}\right)},
1107: \label{EPO34}
1108: \end{equation}
1109: which is shown in FIG.\ref{fig4}. 
1110: 
1111: \subsection{Driven Oscillations}
1112: 
1113: When an electrode has a current density passing into the cathode surface, then there 
1114: is a power per unit area \begin{math} {\cal P}  \end{math} suppled to the surface 
1115: as in Eq.(\ref{intro3}). Under such circumstance the surface temperature rises to a 
1116: high value. In detail, in a non-equilibrium steady state situation, one must introduce 
1117: a ``noise temperature'' \begin{math} T_{{\bf Q}, \omega}  \end{math} which depends on the 
1118: frequency and wave number of the excitations which form due to an Ohmic heating rate 
1119: per unit time per unit area \begin{math} {\cal P}={\cal V}{\cal J} \end{math}. The noise 
1120: temperature enters into the dynamic form factor via 
1121: \begin{equation}
1122: S(-{\bf Q},-\omega )=e^{-\hbar \omega /k_BT_{{\bf Q}, \omega}}S({\bf Q},\omega )
1123: \label{EPO35}
1124: \end{equation}
1125: which is the non-equilibrium noise temperature version of the thermal equilibrium detailed 
1126: balance Eq.(\ref{EPO21}).
1127: 
1128: Experimentally, the noise temperature is high in the infrared  
1129: \begin{math} \omega  \end{math}-regime in a small surface domain size 
1130: \begin{math} d \end{math} related to the wave number via 
1131: \begin{math} d\sim Q^{-1} \end{math}. Note that the overall temperature over the whole 
1132: electromagnetic bandwidth need not be very high, but in the infrared regime at 
1133: \begin{math} \omega \sim \Omega  \end{math} there should be a sharp peak in the noise 
1134: temperature. What has been observed\cite{Szpak:2003} for deuterons absorbed in palladium, 
1135: are surface hot spots flashing in different domains analogous to the flashing of fire-flies 
1136: during a dark evening in the country. The flashing turns on and off in apparently random 
1137: spatial and temporal patterns. The hot spots occur when and where regions of the rough 
1138: cathode surface form electromagnetic cavities of size and shape as to naturedly support the 
1139: infrared radiation. These occur randomly on rough cathode surfaces.
1140: 
1141: During these flash events the nuclear transmutations can 
1142: take place. The hot spots at any given time take up only a small fraction of the 
1143: cathode surface. Increasing the relative fraction of nuclear-active cathode surface 
1144: area which comprises hot spots would proportionately increase the measured efficiency 
1145: of neutron production on the cathode.  The higher the average density of hot spots on 
1146: a given cathode surface over time, the greater the amount of excess heat that would likely 
1147: be observed at the cathode device level using gross thermal measurement techniques such as 
1148: calorimetry.  From the nature of the surface damage on cathodes due to micron-scale hot 
1149: spots, it is evident that the metal actually melts locally during a hot flash producing 
1150: nuclear transmutations. Distinctive areas of obvious melting and explosive blow-out 
1151: cratering features that are consistent with the presence of such hot spots are very commonly 
1152: seen in post-experiment SEM images of the surfaces of various metallic cathode materials that 
1153: include gold (melting point $1554.9\ ^oC$; boiling point $2963\ ^oC$); 
1154: palladium (melting point $1554.9\ ^oC$; 
1155: boiling point $2963\ ^oC$)\cite{Toriyabe:2006,Szpak:2005}; 
1156: titanium (melting point $1064.18\ ^oC$; boiling point $2856\ ^oC$)\cite{Savvatimova:2006}, 
1157: and tungsten (melting point $3422\ ^oC$; boiling point $5555\ ^oC$)\cite{Cirillo:2006}. 
1158: Such evidence is consistent with the noise temperature at a typical value of 
1159: the peak \begin{math} T_{{\bf Q},\omega }\sim 5\times 10^3\ ^oK \end{math}
1160: corresponding to a proton displacement of \begin{math}(u/a)\sim 10 \end{math} within a domain. 
1161: In accordance with FIG.\ref{fig4}, the mass enhancement is more than sufficent for 
1162: the neutron production Eqs.(\ref{intro1}) and (\ref{intro2}).
1163: 
1164: \section{Neutrons and Isotopic Spin Waves\label{ISPW}}
1165: 
1166: The sources of the neutrinos are inhomogeneous in spatial regions near 
1167: the surfaces of cathodes. Also, the neutrinos are so weakly interacting 
1168: that after emission they are virtually unaware of the condensed matter. 
1169: The neutrino on energy shell phase space 
1170: \begin{math} Q^\nu =({\bf Q},|{\bf Q}|) \end{math}
1171: has the Lorentz invariant phase space  
1172: \begin{equation}
1173: dL_{\bf Q}=\left[\frac{d^3 {\bf Q}}{2(2\pi )^3|{\bf Q}|}\right].
1174: \label{ISPW1}
1175: \end{equation}
1176: Writing the neutrino emission part of Eqs.(\ref{CCF2}) and (\ref{CCF3}) 
1177: as the phase space integral 
1178: \begin{eqnarray}
1179: \varpi (x)=-c\ {\Im m}\int \int e^{iQ\cdot (x-y)} 
1180: h^{\lambda \mu \sigma \nu } Q_\mu \times
1181: \nonumber \\ 
1182: \left<W^-_\lambda (x)\bar{\psi }(x)
1183: \gamma_\nu  P_+\psi (y)W^+_\sigma (y)\right>d^4 y dL_{\bf Q}
1184: \nonumber \\ 
1185: =-4c\left(\frac{\hbar G_F}{c^4}\right)^2\ {\Im m}
1186: \int \int e^{iQ\cdot (x-y)} 
1187: h^{\lambda \mu \sigma \nu } Q_\mu \times
1188: \nonumber \\ 
1189: \left<{\cal I}^-_\lambda (x)\bar{\psi }(x)P_-\gamma_\nu  
1190: \psi (y){\cal I}^+_\sigma (y)\right>d^4 y dL_{\bf Q}\ .
1191: \label{ISPW2}
1192: \end{eqnarray}
1193: Under the {\em assumption} that the initial proton spins are 
1194: {\em uncorrelated} and that the free neutron density is dilute, 
1195: considerable simplification can be made in estimating the rather daunting 
1196: but rigorous Eq.(\ref{ISPW2}). The estimate for the inhomogeneous ultra 
1197: low momentum neutron production rate per unit volume is 
1198: \begin{eqnarray}
1199: \varpi (x)\approx 
1200: \left(\frac{\hbar G_F}{c^3}\right)^2
1201: \left(\frac{2mc^2}{\hbar }\right)(g_V^2+3g_A^2)\times 
1202: \nonumber \\   
1203: \ {\Re e}\int \int e^{iQ\cdot (x-y)}
1204: \left<{\cal T}^+ (x)\bar{\psi }(x)  
1205: \psi (y){\cal T}^-(y)\right>d^4 y dL_{\bf Q},\ 
1206: \label{ISPW3}
1207: \end{eqnarray}
1208: wherein Eq.(\ref{intro13}) has been invoked. 
1209: 
1210: If the neutrons are dilute, then it is sufficient to consider the creation 
1211: of a single neutron from a proton, i.e. the propagation of the 
1212: \begin{math} W^\pm  \end{math} within condensed matter. What is left of the heavy 
1213: \begin{math} W^\pm  \end{math} boson is merely an isotopic spin wave.  
1214: There is a superposition of amplitudes summed over all the possible protons 
1215: within a patch which may be converted into a neutron. The isotopic spin wave creation and 
1216: annihilation operators in the surface patch obey 
1217: \begin{equation}
1218: {\cal T}^\pm (x)\approx {\cal T}^\pm ({\bf x})e^{\mp i(c\Delta M)x^0/\hbar }
1219: \label{ISPW4}
1220: \end{equation}
1221: with the neutron-proton mass difference determining the threshold value of the 
1222: electron mass \begin{math} m \end{math} renormalization parameter 
1223: \begin{math} \beta \end{math} i.e.  
1224: \begin{equation}
1225: \Delta M=M_n-M=\beta_0 m.
1226: \label{ISPW5}
1227: \end{equation}
1228: For the creation of a single ultra low momentum neutron from non-relativistic protons
1229: \begin{eqnarray}
1230: \left<{\cal T}^+ (x)\bar{\psi }(x) \psi (y){\cal T}^-(y)\right> 
1231: \Rightarrow 
1232: \nonumber \\ 
1233: \delta ({\bf x}-{\bf y})e^{-i(c\Delta M)(x^0-y^0)/\hbar }\times  
1234: \nonumber \\ 
1235: \left<p^\dagger (x)\bar{\psi }(x) \psi (y)p(y)\right>.
1236: \label{ISPW6}
1237: \end{eqnarray}
1238: For steady state production rates, 
1239: Eq.(\ref{ISPW3}) reads  
1240: \begin{eqnarray} 
1241: \frac{\hbar \varpi ({\bf r})}{mc^2}\approx 
1242: 2 \left(\frac{\hbar G_F}{c^3}\right)^2(g_V^2+3g_A^2)\times 
1243: \nonumber \\
1244: \ {\Re e}\int \int e^{i(c^2\Delta M +\hbar c|{\bf Q}|)t/\hbar}\times 
1245: \nonumber \\ 
1246: \left<p^\dagger ({\bf r})\bar{\psi }({\bf r}) 
1247: \psi ({\bf r,t})p({\bf r,t})\right>(cdt) dL_{\bf Q}. 
1248: \label{ISPW7}
1249: \end{eqnarray}
1250: Explicitly exhibiting the neutrino energy being radiated away 
1251: in Eq.(\ref{intro2}), yields 
1252: \begin{eqnarray}
1253: \frac{\hbar \varpi ({\bf r})}{mc^2}\approx 
1254: \frac{1}{2\pi ^2 c^2} \left(\frac{\hbar G_F}{c^3}\right)^2
1255: (g_V^2+3g_A^2)\times 
1256: \nonumber \\
1257: \ {\Re e}\int_0^\infty  \int_{-\infty}^\infty 
1258: e^{i(c^2\Delta M +\hbar c|{\bf Q}|)t/\hbar}\times 
1259: \nonumber \\ 
1260: \left<p^\dagger ({\bf r})\bar{\psi }({\bf r}) 
1261: \psi ({\bf r,t})p({\bf r,t})\right>(cdt)(\omega d\omega ). 
1262: \label{ISPW8}
1263: \end{eqnarray}
1264: 
1265: The remaining correlation function in Eq.(\ref{ISPW7}) describes 
1266: how an electron which finds itself directly on top of a proton 
1267: propagates in time. The integral over time may be written 
1268: \begin{eqnarray}
1269: {\Re e}\int_{-\infty}^\infty \left<p^\dagger ({\bf r})\bar{\psi }({\bf r}) 
1270: \psi ({\bf r,t})p({\bf r,t})\right> e^{iEt/\hbar }dt
1271: \nonumber \\ 
1272: =2\pi \hbar {\cal N}({\bf r})n_e({\bf r},E),
1273: \label{ISPW9}
1274: \end{eqnarray}
1275: wherein \begin{math} {\cal N}({\bf r})  \end{math} is the mean 
1276: density per unit volume of protons and 
1277: \begin{math} n_e({\bf r},E) \end{math} is the mean collective  
1278: density per unit volume per unit energy of electrons which sit 
1279: directly on the protons. 
1280: 
1281: The steady state inhomogeneous production of 
1282: neutrons per unit time per unit volume 
1283: \begin{math} \varpi({\bf r})  \end{math}
1284: as estimated in Eq.(\ref{ISPW8}); i.e. exhibiting the radiated 
1285: neutrino energy \begin{math} \hbar \omega \end{math}, 
1286: \begin{eqnarray}
1287: \varpi({\bf r})\approx \frac{mc^2}{\pi \hbar}
1288: \left(\frac{\hbar G_F}{c^3}\right)^2(g_V^2+3g_A^2)
1289: {\cal N}({\bf r}){\cal K}({\bf r}),
1290: \nonumber \\ 
1291: {\cal K}({\bf r})=\frac{\hbar }{c}\int_0^\infty 
1292: n_e({\bf r},E=c^2\Delta M+\hbar \omega) \omega d\omega ,
1293: \label{ISPW10}
1294: \end{eqnarray}
1295: wherein \begin{math} n_e({\bf r},E)  \end{math} must be 
1296: calculated including the surface radiation energy and the driving 
1297: current through the cathode. 
1298: 
1299: The calculation of \begin{math} {\cal K} \end{math} depends on 
1300: the detailed physical properties of the cathode surface as well 
1301: as the flux \begin{math} \tilde{\Phi }  \end{math} per unit area 
1302: per unit time of electrons determining the chemical cell current 
1303: as in the power Eq.(\ref{intro3}). In the most simple model, 
1304: let us consider a smooth surface with material properties and electron 
1305: currents determining the neutron creation rate via the mass 
1306: renormalization parameter \begin{math} \beta  \end{math} as defined 
1307: in Eq.(\ref{EMR5b}). We note in passing that a smooth surface is not 
1308: likely to be the best surface for producing neutrons since rough surfaces 
1309: have patches wherein the surface plasma electromagnetic 
1310: field oscillations will be very much more intense. However, the following 
1311: smooth surface model will be employed for estimating the proper low energy 
1312: nuclear reaction rates produced by electroweak interactions.   
1313: 
1314: To compute the density of surface electron states per unit area per 
1315: unit energy, one may begin with a simple renormalized electron mass 
1316: \begin{math} \tilde{m} \end{math} model and take 
1317: \begin{equation}
1318: g_2(E )=2\int 
1319: \delta \left(E -\sqrt{c^2p^2+(\tilde{m}c^2)^2}\right)
1320: \frac{d^2{\bf p}}{(2\pi \hbar)^2}\ ,
1321: \label{ISPW11}
1322: \end{equation}
1323: i.e.   
1324: \begin{equation}
1325: g_2(E)=\frac{E}{\pi \hbar^2 c^2} \ .
1326: \label{ISPW12}
1327: \end{equation}
1328: If such surface electron states are confined to a wave function 
1329: width  \begin{math} l  \end{math} normal to the surface, then we 
1330: have within the surface region, and after integrating over the 
1331: emitted neutrino energy spectrum
1332: \begin{eqnarray}
1333: {\cal K}\approx \frac{\hbar }{lc} \int 
1334: g_2(E=\beta_0 mc^2+\hbar \omega )\omega d\omega ,
1335: \nonumber \\ 
1336: {\cal K}\approx \frac{1}{2\pi l}\left(\frac{mc}{\hbar }\right)^3
1337: (\beta -\beta_0)^2.
1338: \label{ISPW13}
1339: \end{eqnarray}
1340: 
1341: For a smooth surface, integrating the neutrino production rate over 
1342: a thin slab at the electrode surface yields the estimate for 
1343: the production rate per unit time per unit area, 
1344: \begin{equation}
1345: \varpi_2 \approx \left(\frac{g_V^2+3g_A^2}{2\pi^2}\right)
1346: n_2\left(\frac{G_Fm^2}{\hbar c}\right)^2 \frac{mc^2}{\hbar }
1347: (\beta -\beta_0)^2 .
1348: \label{ISPW14}
1349: \end{equation}
1350: The above Eq.(\ref{ISPW14}) for smooth surfaces is in agreement with the 
1351: initial order of magnitude estimate in Eqs.(\ref{est3}) and (\ref{est4}).
1352: 
1353: \section{Conclusions\label{conc}}
1354: 
1355: Electromagnetic surface plasma oscillation energies in hydrogen-loaded metal 
1356: cathodes may be combined with the normal electron-proton rest mass energies 
1357: to allow for neutron producing low energy nuclear reactions Eq.(\ref{intro2}). 
1358: The entire process of neutron production near metallic hydride surfaces may 
1359: be understood in terms of the standard model for electroweak interactions. 
1360: The produced neutrons have ultra low momentum since the wavelength is that 
1361: of a low mode isotopic spin wave spanning a surface patch. The radiation energy 
1362: required for such ultra low momentum neutron production may be supplied by the 
1363: applied voltage required to push a strong charged current across the metallic 
1364: hydride cathode surface. Alternatively, low energy nuclear reactions may be induced 
1365: directly by laser radiation energy applied to a cathode surface. They may even be 
1366: induced, albeit at comparatively low rates of reaction, simply by imposing an 
1367: adequate pressure gradient (which can be as little as one atmosphere.) of gaseous 
1368: hydrogen isotopes across a metallic membrane\cite{Iwamura:2002,Li:2003} composed of 
1369: an aggressive hydride-former such as palladium.
1370: 
1371: The electroweak rates of the resulting ultra low momentum neutron production 
1372: are computed from the above considerations. In terms of the radiation induced 
1373: mass renormalization parameter \begin{math} \beta  \end{math} in Eqs.(\ref{est1}) 
1374: and (\ref{est2}), the predicted neutron production rates per unit area per unit 
1375: time have the form 
1376: \begin{equation}
1377: \varpi_2=\nu (\beta -\beta_0)^2\ \ \ {\rm above\ threshold}
1378: \ \ \ \beta>\beta_0.
1379: \label{conc1}
1380: \end{equation}
1381: The expected range of the parameter \begin{math} \nu \end{math} 
1382: for hydrogen-loaded cathodes is approximately 
1383: \begin{equation}
1384: 10^{12}\ \frac{\rm Hz}{\rm cm^2}\  <\ \nu \ < 10^{14}\ \frac{\rm Hz}{\rm cm^2}
1385: \label{conc2}
1386: \end{equation}
1387: in satisfactory agreement with the orders of magnitude observed experimentally. 
1388: More precise theoretical estimates of \begin{math} \nu \end{math} require 
1389: specific material science information about the physical state 
1390: of cathode surfaces which must then be studied in detail. 
1391: As discussed in previous work\cite{Widom:2006}, a deuteron on certain cathode surfaces 
1392: may also capture an electron producing two ultra low momentum neutrons and a neutrino. 
1393: The neutron production rates for heavy water systems are thereby somewhat enhanced.
1394: 
1395: From a technological perspective, we note that energy must first be put 
1396: into a given metallic hydride system in order to renormalize electron masses and reach 
1397: the critical threshold values at which neutron production can occur.  Net excess energy,  
1398: actually released and observed at the physical device level, is the result of a complex 
1399: interplay between the percentage of total surface area having 
1400: micron-scale E and B field strengths high enough to create neutrons and 
1401: elemental isotopic composition of near-surface target nuclei exposed to local 
1402: fluxes of readily captured ultra low momentum neutrons. In many respects, low 
1403: temperature and pressure low energy nuclear reactions in condensed matter systems resemble 
1404: r- and s-process nucleosynthetic reactions in stars. Lastly, successful fabrication and 
1405: operation of long lasting energy producing devices with high percentages of nuclear active 
1406: surface areas will require nanoscale control over surface composition, 
1407: geometry and local field strengths.  
1408: 
1409: 
1410: 
1411: 
1412: 
1413: \begin{thebibliography}{06}
1414: 
1415: \bibitem{Iwamura:1998}
1416: Y. Iwamura, T. Itoh, N. Gotoh, and I. Toyoda, {\it Fusion Tech.} 
1417: {\bf 33}, 476 (1998).
1418: 
1419: \bibitem{Violante:2002}
1420: V. Violante, E. Castagna, C. Sibilia, S. Paoloni, and F. Sarto, 
1421: {\it Analysis of Ni-Hydride Thin Film after Surface Plasmon Generation 
1422: by Laser Technique}, Condensed Matter Nuclear Science,  
1423: Proceedings of ICCF-10, P. Hagelstein and S. Chubb, Eds.
1424: World Scientific Publishing, Singapore, 421 (2002).  
1425: 
1426: \bibitem{Dash:2002}
1427: J. Dash and D. Chicea, {\it Changes in the Radioactivity, Topography, 
1428: and Surface Composition of Uranium after Hydrogen Loading by Aqueous 
1429: Electrolysis}, Condensed Matter Nuclear Science,  Proceedings of ICCF-10, 
1430: P. Hagelstein and S. Chubb, Eds., World Scientific Publishing, 
1431: Singapore,  463  (2002).
1432: 
1433: \bibitem{Miley:2005}
1434: G.H. Miley, G. Narne, and T. Woo, 
1435: {\it J. Rad. Nuc. Chem.} {\bf 263}, 691 (2005).
1436: 
1437: \bibitem{Miley:1996}
1438: G.H. Miley and J. A. Patterson, {\it J. New Energy} {\bf 1}, 11 (1996).
1439: 
1440: \bibitem{Miley:1997}
1441: G.H. Miley, {\it J. New Energy} {\bf 2}, 6 (1997).
1442: 
1443: \bibitem{Pons:1989}
1444: S. Pons in News, {\it Nature} {\bf 338}, 681 (1989). 
1445: 
1446: \bibitem{Marshak:1969} 
1447: R.E. Marshak, Riazuddin and C.P. Ryan, 
1448: {\it Theory of Weak Interaction of Elementary Particles}, 
1449: Interscience, New York (1969).
1450: 
1451: \bibitem{Pokorski:2000}
1452: S. Pokorski, {\it Gauge Field Theories}, second edition, 
1453: Cambridge University Press, Cambridge (2000).
1454: 
1455: \bibitem{Mizuno:1998} T. Mizuno, {\it Nuclear Transmutation: The Reality of 
1456: Cold Fusion}, Infinite Energy Press, Concord (1998).
1457: 
1458: \bibitem{Kozima:1998}
1459: H. Kozima, {\it Discovery of the Cold Fusion Phenomenon}, Ohotake Shuppan, 
1460: Tokyo (1998).
1461: 
1462: \bibitem{Widom:2006}
1463: A. Widom and L. Larsen {\it Eur. Phys. J.} {\bf C 46}, 107 (2006). 
1464: 
1465: \bibitem{Larsen:2005}
1466: A. Widom and L. Larsen, arXiv:cond-mat/0509269v1, 10 Sep 2005.
1467: 
1468: \bibitem{Cassin:1936}
1469: B. Cassin and E.U. Condon, {\it Phys. Rev.} {\bf 50}, 846 (1936).
1470: 
1471: \bibitem{Gamow:1936}
1472: G. Gamow and E. Teller, {\it Phys. Rev.} {\bf 48}, 895 (1936).
1473: 
1474: \bibitem{Lifshitz:1997}
1475: V.B. Berestetskii, E.M. Lifshitz and L.P. Pitaevskii,  
1476: {\it Quantum Electrodynamics}, Sec.40, Eq.(40.15), 
1477: Butterworth Heinmann, Oxzford (1997).
1478: 
1479: \bibitem{Landau:1975}
1480: L.D. Landau and E.M. Lifshitz, {\it The Classical Theory of Fields}, 
1481: Secs.17 and 47 Prob.2, Pergamon Press, Oxford (1975).
1482: 
1483: 
1484: \bibitem{Dreismann:2005}
1485: C. A. Chatzidimitriou-Dreismann, T. Abdul-Redah, and M. Krzystyniak1,
1486: {\it Phys. Rev.} {\bf B 72}, 054123 (2005).
1487: 
1488: \bibitem{Cowley:2006}
1489: R A. Cowley and J Mayers, 
1490: {\it Phys.: Condens. Matter} {\bf 18}, 5291 (2006).
1491: 
1492: \bibitem{Kenali:1999}
1493: M. Kemali, J. E. Totolici, D. K. Ross, and I. Morrison, 
1494: {\it Phys. Rev. Lett.} {\bf 84}, 1531 (2000); 
1495: 
1496: 
1497: \bibitem{Kolesnikova:2002}
1498: A.I. Kolesnikova, V.E. Antonovb, V.K. Fedotova, G. Grossec, 
1499: A.S. Ivanovd and F.E. Wagner,
1500: {\it Physica} {\bf B}, 158 (2002).
1501: 
1502: 
1503: \bibitem{Glagolenkol:2002}
1504: I.Y. Glagolenko1, K.P. Carney1, S. Kern, E.A. Goremychkin, 
1505: T.J.Udovic, J.R.D. Copley, J.C. Cook, 
1506: {\it Appl. Phys.} {\bf A 74}, S1397 (2002).
1507: 
1508: 
1509: \bibitem{Gidopoulos:2005}
1510: Nikitas I. Gidopoulos, Phys. Rev. {\bf B 71}, 054106 (2005).
1511: 
1512: 
1513: \bibitem{Medini:2003}
1514: D. Medini and A. Widom, {\it J. Chem. Phys.} {\bf 118} 2405 (2003).
1515: {\it J. Phys.: Condens. Matter} {\bf 18} 5291 (2006).
1516: 
1517: 
1518: \bibitem{Krzystyniak:2006}
1519: M. Krzystyniak and C.A. Chatzidimitriou-Dreismann,
1520: {\it Journal of Neutron Research}, {\bf 14}, 193 (2006).
1521: 
1522: \bibitem{Platzman:2004}
1523: G. F. Reiter and P. M. Platzman, 
1524: Phys. Rev. {\bf B 71}, 054107 (2004). 
1525: 
1526: \bibitem{Szpak:2003}
1527: S. Szpak, P.A. Mosier-Boss, J. Dea, and F. Gordon, 
1528: $10^{th}$ {\it International Conference on Cold Fusion}, 
1529: World Scientific Inc. (2003).
1530: 
1531: \bibitem{Toriyabe:2006}
1532: Y. Toriyabe, T. Mizuno, T. Ohmori, and Y. Aoki, 
1533: {\it Elemental Analysis of Palladium Electrodes After Pd/Pd Light Water Critical 
1534: Electrolysis},  Eds. A. Takahashi, K. Ota, and Y. Iwamura, Proceedings of ICCF-12,
1535: page 252, World Scientific Publishing, Singapore (2006).
1536: 
1537: \bibitem{Szpak:2005}
1538: S. Szpak, P. Boss, C. Young, and F. Gordon, 
1539: {\it Naturwissenschaften}  {\bf 92}, 394 (2005).
1540: 
1541: \bibitem{Savvatimova:2006}
1542: I. Savvatimova and D. Gavritenkov, {\it Results of Analysis of Ti Foil After Glow 
1543: Discharge with Deuterium}, Proceedings of ICCF-11, Ed. J. Biberian, page 438, 
1544: World Scientific Publishing, Singapore, (2006). 
1545: 
1546: \bibitem{Cirillo:2006}
1547: D. Cirillo, and V. Iorio, {\it Transmutation of Metal at Low Energy in a Confined 
1548: Plasma in Water}, Ed. J. Biberian, page 492, Proceedings of ICCF-11, 
1549: World Scientific Publishing, Singapore, (2006). 
1550: 
1551: \bibitem{Iwamura:2002}
1552: Y. Iwamura, M. Sakano and T. Itoh, {\it Jpn. J. Appl. Phys.} {\bf 41}, 4642 (2002). 
1553: 
1554: \bibitem{Li:2003}
1555: X. Li, B. Liu, J. Tian, Q. Wei, R. Zhou and Z. Yu, 
1556: {\it J. Phys. D: Appl. Phys.}, {\bf 36},  3095 (2003).  
1557: 
1558: 
1559: 
1560: \end{thebibliography}
1561: 
1562: 
1563: \end{document}
1564: 
1565: