1: \documentclass[showpacs,preprintnumbers,amsmath,amssymb,
2: eqsecnum,
3: twocolumn, tightenlines,
4: %preprint
5: ]{revtex4}
6:
7:
8:
9: \usepackage{bm}
10:
11: \usepackage{amsmath}
12:
13: \usepackage{dcolumn}
14:
15: \usepackage[dvips]{graphicx}
16: \sloppy
17:
18: % \draft
19:
20: \begin{document}
21:
22: \bibliographystyle{apsrev}
23:
24: \title {Coulomb problem for vector bosons}
25:
26: \author{M.Yu.Kuchiev}
27: \email[Email:]{kuchiev@phys.unsw.edu.au}
28: \author{V.V.Flambaum}
29: \email[Email:]{flambaum@phys.unsw.edu.au}
30: \affiliation{School of Physics, University of New South Wales,
31: Sydney 2052, Australia}
32: \affiliation{Physics Division, Argonne
33: National Laboratory, Argonne, Illinois 60439-4843, USA}
34:
35: \date{\today}
36:
37: \begin{abstract}
38: The Coulomb problem for vector bosons $W^\pm$ incorporates a
39: well known difficulty; the charge of the boson localized in a
40: close vicinity of the attractive Coulomb center proves be
41: infinite. This fact contradicts the renormalizability of the
42: Standard Model, which presumes that at small distances all
43: physical quantities are well defined. The paradox is shown to be
44: resolved by the QED vacuum polarization, which brings in a
45: strong effective repulsion that eradicates the infinite charge
46: of the boson on the Coulomb center. This property allows to
47: define the Coulomb problem for vector bosons properly, making it
48: consistent with the Standard Model.
49: \end{abstract}
50:
51: \pacs{12.15.Ji, 12.15.Lk, 12.20.Ds}
52:
53:
54: \maketitle
55:
56: \section{introduction}
57: \label{intro}
58:
59: Consider a charged vector boson, which propagates in the Coulomb
60: field created by a heavy point-like charge $Z$ assuming that the
61: boson is massive, its mass being produced via the Higgs mechanism;
62: the $W^\pm$-bosons give an example. We study relativistic effects
63: in this Coulomb problem. A situation where they can be important
64: arises, for example, for small primordial charged black holes
65: since an impact of their Coulomb field on a $W$-boson prevails
66: over the gravitational field.
67:
68: It has ``always'' been known that there is a difficulty in the
69: Coulomb problem for vector bosons. Soon after Proca formulated
70: theory for vector particles \cite{proca_1936} it became clear that
71: it produces inadequate results for the Coulomb problem
72: \cite{massey-corben_1939,oppenheimer-snyder-serber_1940,tamm_1940-1-2}.
73: This fact inspired Corben and Schwinger
74: \cite{corben-schwinger_1940} to modify the Proca theory, tuning
75: the Lagrangian and equations of motion in such a way as to force
76: the hyromagnetic ratio of the vector boson to acquire a favorable
77: value $g=2$. Later on the formalism of
78: \cite{corben-schwinger_1940} was found to have a connection with
79: the non-Abelian gauge theory \cite{schwinger_1964}, which makes it
80: relevant for the present day studies. A role of the identity
81: $g=2$ was thoroughly discussed in literature, see e. g.
82: Ref.\cite{cheng_wu_1972,huang_1992}.
83:
84: Ref.\cite{corben-schwinger_1940} found a realistic discrete energy
85: spectrum for the Coulomb problem for vector bosons. However, it
86: discovered also a fundamental flaw in the problem. For two series of
87: quantum states the charge of the vector boson located on the
88: Coulomb center turns infinite, which indicates the fall of the
89: boson on the center. One of these series has the total angular
90: momentum zero, $j=0$, another one has $j=1$ (being further
91: specified by a label ``$\gamma-3/2$'', see Section \ref{j>0}).
92: This effect takes place for arbitrary small value of the Coulomb
93: charge $Z$, which is physically unacceptable. Moreover, it takes
94: place at small distances, while the renormalizability of the
95: Standard Model Ref.\cite{thooft-veltman-1972} guarantees that
96: there should be no problems of this type. All this indicates that
97: the Coulomb problem is poorly defined. Moreover, there exists a
98: contradiction; the Coulomb problem derived from the Standard Model
99: produces results, which challenge the Model itself.
100:
101: This difficulty was inspirational for several lines of research.
102: Early efforts are summarized in
103: Ref.\cite{vijayalakshmi-seetharaman-mathews_1979}. More recent
104: Refs. \cite{pomeransky-khriplovich_1998,pomeransky-se'nkov_1999,%
105: pomeransky-sen'kov-khriplovich_2000} suggested a new, refined
106: modification of the formalism for vector bosons.
107: Ref.\cite{silenko_2004} claimed that it complies with results of
108: Corben and Schwinger. Some authors considered other forms of the
109: equation governing vector bosons
110: \cite{fushchych-nikitin-susloparow_1985,fushchych-nikitin_1994,%
111: sergheyev_1997}, which produce more acceptable results for the
112: Coulomb problem, but this advantage is partially undermined by the
113: fact that it does not step from a renormalizable theory.
114:
115: However, in spite of a progress made over the years, there still
116: exists a contradiction between the difficulty in the Coulomb
117: problem for vector bosons and the renormalizability of the
118: Standard Model. We find a clear way to resolve this contradiction,
119: formulating the Coulomb problem for vector particles properly,
120: within the frames of the Standard Model. Our main observation is
121: that the polarization of the QED vacuum has a profound impact in
122: the problem forcing the density of charge of a vector boson to
123: decrease at the origin, thus making the Coulomb problem stable,
124: well defined. This decrease has an exponential character for the
125: $j=0$ state. For the $j=1$,``$\gamma-3/2$'' state the suppression is
126: of a power-type. In both these states the suppression eradicates
127: the difficulty of the Coulomb problem.
128:
129: From the first glance this result looks surprising. Presumably,
130: the vacuum polarization is meant to make the attractive Coulomb
131: field only stronger, which should result in an increase of the
132: charge density at the origin. In addition to this, the vacuum
133: polarization for spinor and scalar particles in the Coulomb field
134: is known to produce only small, perturbative effects. In
135: contrast, we claim a strong {\it reduction} of the charge density
136: for the vector particle. To grasp a physical mechanism involved it
137: is necessary to notice that the equation of motion for vector
138: particles incorporates a particular term, which explicitly depends
139: on the external current and has no counterparts for scalars and
140: spinors, see the last term in Eq.(\ref{form}). Precisely this
141: term brings in a strong effective repulsion, which stems from the
142: vacuum polarization and makes the Coulomb problem stable, well
143: defined.
144:
145: The renormalizability of the Standard Model means that if all
146: essential processes are taken care of, then the infinite charge of
147: a vector boson located at the Coulomb center is eliminated. It is
148: known that the amplitude of the photon exchange between leptons
149: or/and quarks at high transferred momenta should be considered
150: alongside exchange by the Higgs and $Z$-bosons. From this
151: perspective the catastrophic behavior of the charge density of a
152: vector boson at small distance, i. e. at large transferred
153: momenta, in the Coulomb problem could have been considered as an
154: indication that the Coulomb problem for vector bosons should
155: include the processes related to the Higgs and $Z$-bosons exchange
156: from the very beginning. In contrast to this widely spread
157: presumption we find a way to formulate the Coulomb problem for
158: vector bosons entirely in terms of the $W$ and electromagnetic
159: fields, as a pure QED problem.
160:
161: A complete Standard Model calculation, where all possible
162: processes are accounted for accurately, would require specific
163: information on the nature of a heavy particle that creates the
164: Coulomb field, i. e. on all its quantum numbers related to the
165: Standard Model. This information is not necessarily feasible. A
166: simple example give primordial black holes; it is not easy to
167: assert with certainty whether they have, or have not the weak
168: charge, and what are their other quantum numbers in the Standard
169: Model. Same questions arise in relation to other possible
170: candidates for the heavy Coulomb center. As a result, a
171: presumption that the exchange of the Higgs and $Z$-boson should
172: play a basic role in the Coulomb problem leads to complications.
173: It is fortunate therefore that the detailed information on
174: properties of the heavy particle proves be redundant, that the
175: Coulomb problem can be properly defined using the only physical
176: parameter of a heavy particle, its electric charge.
177:
178: This point of view, which is advocated in the present work keeps
179: the Coulomb problem simple and transparent. On the other hand,
180: it also allows one to include all other processes, which are left
181: outside the scope of the Coulomb problem, by means of perturbation
182: theory. Our preliminary calculations indicate that the exchange by
183: the Higgs and $Z$-bosons, as well as possible processes with
184: lepton or quark exchange, give only small corrections. The reason
185: stems from the fact that the found wave functions for vector
186: bosons are suppressed at small distances. Consequently, the
187: small-distance processes with the exchange by Higgs and $Z$-bosons
188: are also suppressed (the exchange by a lepton or quark contains
189: the vanishing at the Coulomb center fields, which describe the
190: $W$-boson).
191:
192: In section \ref{vector} the Corben-Schwinger formalism for charged
193: vector bosons is derived directly from the Standard Model. The
194: pure Coulomb problem is discussed in Sections
195: \ref{Nonrel}-\ref{catastroph} and several Appendixes. This
196: analyses follows Ref.\cite{corben-schwinger_1940}, but some
197: important details, including the non-relativistic limit (Section
198: \ref{Nonrel}) and the eigenvalue problem for $j=0$ states (Section
199: \ref{matrix}) are discussed in more detail. Sections
200: \ref{vacuum_polarization},\ref{numericals} present the main result
201: of the paper. They show that the QED vacuum polarization plays a
202: defining role in the problem, as was first noticed in our previous
203: work \cite{kuchiev-flambaum_2005}. The units $\hbar=c=1$,
204: $e^2=4\pi\alpha$ where $e<0$, are used below.
205:
206:
207:
208: \section{$W$-mesons in electromagnetic field}
209:
210: \label{vector}
211:
212: \subsection{$W$-bosons in Standard Model}
213: \label{standard}
214:
215: Consider boson fields in the electroweak part of the Lagrangian of
216: the Standard Model, see e.g. Ref.\cite{weinberg_2001},
217: \begin{eqnarray}
218: \label{gauge}
219: {\mathcal L}= -\frac{1}{4}\,
220: \left(\partial_\mu \boldsymbol{A}_\nu-\partial_\nu \boldsymbol{A}_\mu +
221: g \,\boldsymbol{A}_\mu \times \boldsymbol{A}_\nu\right)^2,
222: \\ \nonumber
223: -\frac{1}{4}\,
224: \left(\partial_\mu { B}_\nu-\partial_\nu {B}_\mu \right)^2+
225: \frac{1}{2}\,D_\mu\Phi^+ D^\mu \Phi~.
226: \end{eqnarray}
227: Here $\boldsymbol{A}_\mu$ and $B_\mu$ are the triplet of $SU(2)$
228: and the $U(1)$ gauge potentials respectively (abridged notation is
229: used here). The covariant derivative $D_\mu\Phi$ takes into
230: account that the Higgs field $\Phi$ has a hypercharge $Y=2$, which
231: describes its interaction with the $U(1)$ field, and is
232: transformed as a doublet under the $SU(2)$ gauge transformations.
233: Taking the unitary gauge one can present it via one real component
234: \begin{eqnarray}
235: \label{vacuum}
236: \Phi =\left( \begin{array} {c} 0 \\ \phi
237: \end{array} \right)~,\quad \phi=\phi^+~.
238: \end{eqnarray}
239: Assuming that the scalar field develops the vacuum expectation
240: value $\phi=\phi_0$ and the Higgs mechanism takes place, one finds
241: that the gauge field can be presented as a new $U(1)$ field
242: $A_\mu$, and a triplet of massive fields $W^\pm_\mu, \,Z_\mu$
243: \begin{eqnarray}
244: \label{Amu}
245: A_\mu &=&-\sin \theta \,A_\mu^3+\cos\theta \,B_\mu~,
246: \\ \label{Zmu}
247: Z_\mu &=& ~~\cos \theta \,A_\mu^3+\sin\theta \,B_\mu~,
248: \\ \label{Wmu}
249: W_\mu &=&~~\left(A_\mu^1-iA_\mu^2 \right)/\sqrt 2~,
250: \end{eqnarray}
251: Here $W_\mu\equiv W_\mu^-$ represents the $W$-boson with charge
252: $e=-|e|$, and $\theta$ is the Weinberg angle.
253:
254: Expanding the Lagrangian Eq.(\ref{gauge}) in the vicinity of
255: $\phi=\phi_0$ and retaining only bilinear in the fields
256: $W_\mu,W_\mu^+$ terms, including their interaction with the
257: electromagnetic field, one derives an effective Lagrangian
258: \begin{eqnarray}
259: \nonumber
260: {\mathcal L}^W &=& -\frac{1}{2}\left(\nabla_\mu W_\nu -
261: \nabla_\nu W_\mu\right)^+ \left(\nabla^\mu W^\nu-
262: \nabla^\nu W^\mu\right)
263: \\ \label{W}
264: &&+ i e \, F^{\mu\nu} \,W^+_\mu W_\nu +
265: m^2 \,W_\mu^+ W^\mu~,
266: \end{eqnarray}
267: which describes the propagation of $W$-bosons in an external
268: electromagnetic field. Here $m$ is the mass of $W$. The external
269: field is accounted for in Eq.(\ref{W}) in the derivative
270: $\nabla_\mu=\partial_\mu +i e A_\mu$ and by the term with the
271: field $F^{\mu\nu}=\partial^\mu A^\nu-\partial^\nu A^\mu$. The
272: first and the last terms in Eq.(\ref{W}) are present in the Proca
273: formalism \cite{proca_1936}, while the second one was introduced
274: by Corben and Schwinger \cite{corben-schwinger_1940}.
275:
276: From Eq.(\ref{W}) one derives the classical Lagrange equation of
277: motion for vector bosons
278: \begin{eqnarray}
279: \label{wave}
280: \left( \nabla^2+m^2\right) W^\mu
281: + 2 i e \,F^{\mu\nu}\,W_\nu-
282: \nabla^\mu \nabla^\nu \,W_\nu =0~.
283: \end{eqnarray}
284: Here an identity $[\nabla_\mu,\nabla_\nu]=ieF_{\mu\nu}$ was used.
285: Taking a covariant derivative in Eq.(\ref{wave}) one finds
286: \begin{eqnarray}
287: \label{Lgauge}
288: m^2\nabla_\mu W^\mu+ie\,j_\mu W^\mu=0~,
289: \end{eqnarray}
290: where
291: \begin{eqnarray}
292: \label{j}
293: j^\mu=\partial_\nu F^{\nu\mu}~.
294: \end{eqnarray}
295: is the external current, which creates the external field
296: $F^{\nu\mu}$. Evaluating $\nabla_\mu W^\mu$ from
297: Eq.(\ref{Lgauge}) and substituting the result back into
298: Eq.(\ref{wave}) one rewrites the latter one in a more transparent
299: form
300: \begin{eqnarray}
301: \label{form}
302: \left( \nabla^2+m^2\right) W^\mu
303: + 2 i e F^{\mu\nu}W_\nu
304: +\frac{ie}{m^2} \nabla^\mu (j_\nu W^\nu) =0.~~\quad
305: \end{eqnarray}
306: This equation of motion for vector bosons was suggested in
307: Ref.\cite{corben-schwinger_1940}. The coefficient 2 in front of
308: the second term ensures that the g-factor of the boson takes the
309: value $g=2$, see Eq.(\ref{moment}) below.
310:
311: The derivation outlined shows that Eq.(\ref{form}) represents the
312: classical equation of motion for $W$-bosons in the external
313: electromagnetic field, which is valid within the frames of the
314: Standard Model. This equation has similarities with the
315: Klein-Gordon and Dirac equations (if the latter one is written as
316: the second-order differential equation), but there is also an
317: important distinction. It is produced by the last term in
318: Eq.(\ref{form}), which explicitly contains the external current;
319: there is no similar terms for scalars and spinors. We will see how
320: important this term is, when we discuss the vacuum polarization.
321:
322: We will use below a current of vector bosons
323: $j_{\phantom{\,}\mu}^{W}$, which can be obtained by considering a
324: variation of the Lagrangian Eq.(\ref{W}) under variation of
325: $A_\mu$, which yields
326: \begin{eqnarray}
327: \label{jjj}
328: j_{\phantom{\,}\mu}^{W}&=&j_{\mu}^{(1)}+j_{\mu}^{(2)}+j_{\mu}^{(3)}~,
329: \\ \label{j1}
330: j_{\mu}^{(1)} &=& -ie\,( \,W^+_\nu \nabla_\mu W^\nu -
331: \nabla_\mu W_\nu^+W^\nu\,)
332: \\ \label{j2}
333: j_{\mu}^{(2)} &=& -ie\,( \,\nabla_\nu W_\mu^+W^\nu -
334: W_\nu^+ \nabla^\nu W_\mu \,)
335: \\ \label{j3}
336: j_{\mu}^{(3)} &=& -ie \,\partial^\nu(\, W_\mu^+ W_\nu -
337: W_\nu^+W_\mu \,)~.
338: \end{eqnarray}
339: Differentiating in Eq.(\ref{j3}) term by term and taking into
340: account Eq.(\ref{Lgauge}) one verifies that
341: \begin{eqnarray}
342: \label{easy}
343: j_{\mu}^{(3)} &=& j_{\mu}^{(2)} -ie( \,W_\mu^+\nabla_\nu W^\nu -
344: \nabla^\nu W_\nu^+ W_\mu \,)
345: \\ \nonumber
346: &=& j_{\mu}^{(2)}
347: -\frac{e^2}{m^2}(\, W_\mu^{+} W_\nu+W_\nu^{+}W_\mu\,)\,j^\nu~.
348: \end{eqnarray}
349: Using this result, the current Eq.(\ref{jjj}) can be written in a
350: compact form
351: \begin{eqnarray}
352: \nonumber
353: j_{\phantom{\,}\mu}^{W}&=&-ie\Big( \,W^+_\nu \nabla_\mu W^\nu +
354: 2 \nabla_\nu W^+_\mu W^\nu -c.c. \,\Big)
355: \\
356: \label{jw}
357: &&-\frac{e^2}{m^2}(\, W_\mu^{+} W_\nu+W_\nu^{+}W_\mu\,)\,j^\nu~.
358: \end{eqnarray}
359: Here $c.c.$ refers to two complex conjugated terms.
360:
361: \subsection{Static electric field}
362: \label{static_electric_field}
363:
364: Consider a static electric field described by the electric
365: potential $A_0=A_0(\boldsymbol{r})$ and charge density
366: $\rho=\rho(\boldsymbol{r})=-\Delta A_0$. For a stationary state of
367: the $W$-boson one can presume that
368: \begin{eqnarray}
369: \label{sta}
370: \nabla_0^2
371: W_\mu= -(\varepsilon-U)^2W_\mu~,
372: \end{eqnarray}
373: where $\varepsilon$ is the energy of the stationary state, and
374: $U=U(\boldsymbol{r})= eA_0$ is the potential energy of the
375: $W$-boson in the electric field. Eq.(\ref{Lgauge}) in this case
376: gives
377: \begin{eqnarray}
378: \label{simp}
379: {\mathsf w}
380: = ( \varepsilon-U-\Upsilon )^{-1} \boldsymbol{\nabla \cdot W} .\quad
381: \end{eqnarray}
382: The four-vector $W^\mu=(W_0,\boldsymbol{W})$ is presented here via
383: the three-vector $\boldsymbol{W}$ and the modifies
384: zeroth-component ${\mathsf w}= i W_0$. In order to simplify notation we
385: introduce also a very important for us quantity
386: $\Upsilon=\Upsilon(\boldsymbol{r})$,
387: \begin{eqnarray}
388: \label{Ups}
389: \Upsilon=\frac{e\rho}{\,m^2}=-\frac{\Delta U}{m^2}~.
390: \end{eqnarray}
391: Eqs.(\ref{Lgauge}),(\ref{sta}) show that this definition complies
392: with (\ref{simp}). The quantity $\Upsilon$ appears in the
393: equations of motion alongside the initial potential $U$, see e.g.
394: Eq.(\ref{simp}). In this sense it plays a role of an effective
395: potential energy, which is specific for vector bosons. We will
396: call it the $\Upsilon$-term, or $\Upsilon$-potential. In this
397: notation Eq.(\ref{form}) reads
398: \begin{eqnarray}
399: \label{wa}
400: \big((\varepsilon-U)^2\!-m^2\big)\boldsymbol{W}=-\Delta
401: \boldsymbol{W}-2\boldsymbol{\nabla} U
402: {\mathsf w}-\boldsymbol{\nabla}(\Upsilon {\mathsf w}),&\quad\quad&
403: \\
404: \label{0}
405: \big((\varepsilon-U)^2\!-m^2\big) {\mathsf w}=-\Delta
406: {\mathsf w}+2\boldsymbol{\nabla}U\boldsymbol{ \cdot
407: W}\quad\quad\quad~~~ &&
408: \\ \nonumber
409: + (\varepsilon-U)\Upsilon {\mathsf w}.&&
410: \end{eqnarray}
411: A relation between ${\mathsf w}$ and $\boldsymbol{W}$ given by
412: Eq.(\ref{simp}) shows that among four equations of motion
413: Eqs.(\ref{wa}),(\ref{0}) only three are independent, precisely
414: what one expects for massive vector particles.
415:
416: It will be useful to present Eq.(\ref{wa}) in a slightly
417: different form, which can be derived by combining it with
418: Eq.(\ref{simp}) and using an identity $\Delta\boldsymbol{W}
419: =\boldsymbol{\nabla \times}(\boldsymbol{ \nabla \times
420: W})-\boldsymbol{\nabla}(\boldsymbol{\nabla \cdot W})$, which
421: gives
422: \begin{eqnarray}
423: \label{tran}
424: \!\big((\varepsilon-U)^2\!-m^2\big)\boldsymbol{W}=
425: \boldsymbol{\nabla \times}(\boldsymbol{
426: \nabla \times W})
427: \\ \nonumber
428: -(\varepsilon-U)\boldsymbol{\nabla} {\mathsf w}\,
429: -\boldsymbol{\nabla}U\,{\mathsf w}~.
430: \end{eqnarray}
431: From the expression for the current of vector bosons Eq.(\ref{jw})
432: one derives the charge density
433: \begin{eqnarray}
434: \label{ro}
435: \rho^W&=& 2e\Big( (\varepsilon-U)(\boldsymbol{W}^+
436: \!\cdot\!\boldsymbol{W}+{\mathsf w}^+{\mathsf w})
437: \\ \nonumber
438: &+&\boldsymbol{W}^+\!\!\cdot \!\boldsymbol{ \nabla} {\mathsf w}
439: +
440: \boldsymbol{W}\!\cdot \!\boldsymbol{ \nabla}\,{\mathsf w}^+ -\Upsilon \,
441: {\mathsf w}^+{\mathsf w}\,\Big).
442: \end{eqnarray}
443:
444: \subsection{G-factor}
445: \label{g-factor}
446:
447: The behavior of vector bosons in the homogeneous magnetic fields
448: was studied in detail, see e.g.
449: \cite{vijayalakshmi-seetharaman-mathews_1979} and references
450: therein. The spectrum of this problem reads, see Section
451: \ref{homo},
452: \begin{eqnarray}
453: \label{e2}
454: \varepsilon^2=m^2+p_z^2+2|e|B\left(n+1/2+\sigma \right)~.
455: \end{eqnarray}
456: Here $n=0,1\dots$ specifies the Landau levels, and $\sigma=-1,0,1$
457: gives a projection of spin $S=1$ of the vector boson.
458: Eq.(\ref{e2}) shows that vector bosons possess the magnetic moment
459: \begin{eqnarray}
460: \label{moment}
461: \mbox{\boldmath $\mu$} =e\boldsymbol{ S}/m~,
462: \end{eqnarray}
463: which means that the magnetic $g$-factor is $g=2$.
464:
465:
466:
467: \section{Non-relativistic limit}
468: \label{Nonrel}
469:
470: Consider a vector boson in a static electric field with the
471: potential energy $U=eA_0(\boldsymbol{ r})$. If we presume that
472: the non-relativistic approach is valid, which needs that $|U| \ll
473: m$, then in the lowest order of the perturbation theory in powers
474: of $U/m$ one immediately finds from Eqs.(\ref{wa}),(\ref{simp})
475: \begin{eqnarray}
476: \label{nonrel}
477: E\, \boldsymbol{ W} = - \frac{1}{2m}\,\Delta\,\boldsymbol{ W}
478: +U\boldsymbol{ W} ~.
479: \end{eqnarray}
480: Here $E\simeq \varepsilon-m$ is the energy, the vector $\bf W$
481: plays a role of the wave function for the vector boson, and the
482: non-relativistic Hamiltonian on the right-hand side has a usual
483: form for a massive charged particle.
484:
485: Let us find corrections to Eq.(\ref{nonrel}) induced by
486: relativistic effects. The wave function of the massive vector
487: particle ${\mbox {\boldmath $\Phi$}}$ is well defined in the rest
488: frame. Therefore the vector $\bf W$, which describes the moving
489: vector particle, inevitably deviates from the wave function $\mbox
490: {\boldmath $\Phi$}$. A relation between $\bf W$ and $\mbox{
491: \boldmath $\Phi$}$ is easy to articulate for the free motion,
492: when it is given by the Lorentz boost, see e.g. a book \cite{LL4},
493: \begin{eqnarray}
494: \label{fw}
495: \boldsymbol{ W} =\mbox{\boldmath $\Phi$}
496: +\frac{\boldsymbol{ p}\,
497: ( \boldsymbol{ p} \cdot \mbox{\boldmath $\Phi$}) }{m(m+\varepsilon)}
498: \end{eqnarray}
499: Generically, the potential energy brings in complications, but
500: within the necessary accuracy we can neglect them, presuming also
501: that $\varepsilon \simeq m$. Then Eq.(\ref{fw}) gives
502: \begin{eqnarray}
503: \label{fwapp}
504: \boldsymbol{ W}&\simeq& \mbox{\boldmath $\Phi$}+\frac{\boldsymbol{ p}\,
505: ( \boldsymbol{ p} \cdot \mbox{\boldmath $\Phi$}) }{2m^2}~,
506: \end{eqnarray}
507: where $\boldsymbol{ p}=-i\boldsymbol{ \nabla}$. This relation
508: plays a role similar to the Foldy-Wouthuysen transformation
509: \cite{foldy-wouthuysen_50} for fermions.
510:
511:
512: Substituting Eq.(\ref{fwapp}) in Eqs.(\ref{wa}),(\ref{simp}) and
513: expanding the latter ones in powers of $U/m$ one finds the
514: following Schr\"odinger-type equation for the wave function $\mbox
515: {\boldmath $\Phi$}$ of the vector boson
516: \begin{eqnarray}
517: \label{corr}
518: E\,\mbox{\boldmath $\Phi $}_i &=& H_{ij} \mbox{\boldmath $\Phi $}_j~,
519: \\ \label{hami}
520: H_{ij}&=& \left( \frac{ \boldsymbol{ p}^2 }{2m}+U \right)
521: \delta_{ij} +\delta H_{ij}~,
522: \\ \nonumber
523: \delta H_{ij} &=& - \frac{ \boldsymbol{ p}^4 }{8m^3}\,\delta_{ij}
524: -\frac{ \boldsymbol{ F} \cdot ( \boldsymbol{ p \times S}_{ij} ) }{ 2m^2 }
525: +\frac{\Delta U}{6m^2} \,\delta_{ij}
526: \\ \label{dH}
527: &+&\frac{1}{6m^2}\left( 3 \frac{\partial^2 U}{\partial r_i
528: \partial r_j}
529: -\Delta
530: U \,\delta_{ij}\right)~.
531: \end{eqnarray}
532: Here $i,j=1,2,3$ label components of three-vectors,
533: $\boldsymbol{ S}$ is the spin, which operates on a vector
534: $\boldsymbol{ V}$ according to $\boldsymbol{ S}_{ij}
535: V_j=-i\epsilon_{ijk} V_k$.
536:
537: The relativistic correction to the Hamiltonian $\delta H$ of
538: vector particles is given in Eq.(\ref{dH}). It is instructive to
539: compare this correction with the known Darwin Hamiltonian $\delta
540: H_\mathrm{D}$, which accounts for relativistic effects for spinor
541: particles
542: \begin{eqnarray}
543: \label{D}
544: \delta H_D = - \frac{ \boldsymbol{ p}^4 }{8m^3}
545: -\frac{ \boldsymbol{ F} \cdot ( \boldsymbol{ p \times s} ) }{ 2m^2 }
546: +\frac{\Delta U}{8m^2}~.
547: \end{eqnarray}
548: Here $\boldsymbol{ s}=\mbox{\boldmath $\sigma$ }/2 $ is the
549: operator of spin for spinor particles. The three terms in the
550: first line of Eq.(\ref{dH}) resemble their counterparts in
551: Eq.(\ref{D}), the only distinction is the numerical coefficient in
552: front of the term with $\Delta U$. We conclude that these three
553: terms have conventional meaning, describing the relativistic
554: correction to the kinetic energy, the spin-orbit interaction, and
555: the contact correction to the potential. The coefficient in front
556: of the term responsible for the spin-orbit interaction in
557: Eq.(\ref{hami}) complies with the hyromagnetic ratio $g=2$ of the
558: vector boson, if one presumes that the Thomas ``one-half rule'' is
559: applicable for vector particles the same way as for spinors.
560:
561: The last, forth term in Eq.(\ref{dH}) finds no counterpart in the
562: Darwin Hamiltonian. It is instructive to write a contribution of
563: this term to the energy shift
564: \begin{eqnarray}
565: \label{dE}
566: \delta E_Q&=&\frac{1}{6}\, \int \, Q_{ij} \,
567: \frac{\partial^2 A_0}{\partial r_j
568: \partial r_i} \,d^3 r~,
569: \\ \label{Q}
570: Q_{ij}&=&\frac{e}{m^2} \left( 3
571: \mbox{\boldmath $\Phi $}_i^*
572: \mbox{\boldmath $\Phi $}_j-\delta_{ij}
573: | \mbox{\boldmath $\Phi $} |^2 \right)~.
574: \end{eqnarray}
575: Eqs.(\ref{dE}),(\ref{Q}) show that $Q_{ij}$ plays a role of the
576: density of the quadrupole moment for vector bosons. We conclude
577: that the last, forth term in Eq.(\ref{dH}) indicates that vector
578: bosons have a quadruple moment.
579:
580: From the first glance the contact and the quadrupole terms in the
581: Eq.(\ref{dH}) have similarity with the $\Upsilon$-term in
582: Eq.(\ref{Ups}). However this resemblance is coincidental, since
583: the $\Upsilon$-term does not contribute to (\ref{dH}), which takes
584: into account corrections of the order of $(Z\alpha)^2$.
585: Eq.(\ref{Ups}) allows to estimate the $\Upsilon$-potential as
586: $\Upsilon\sim (mr_0)^{-2} U\sim (Z\alpha)^2 U$, where
587: $r_0=(Z\alpha m)^{-1}$ is the Bohr radius. The
588: $\Upsilon$-potential comes into the equation of motion with the
589: factor $w$, see the last term in Eq.(\ref{wa}). Eq.(\ref{simp})
590: gives an estimate $w\sim (mr_0)^{-1}|\boldsymbol{W}|\sim Z\alpha
591: |\boldsymbol{W}|$. Overall, an estimate for the correction
592: produced by the $\Upsilon$-term in Eq.(\ref{wa}) is $\sim
593: (Z\alpha)^3$, which means that the $\Upsilon$-term is too small to
594: contribute to Eq.(\ref{dH}). Thus, the contact and quadrupole
595: interactions in Eq.(\ref{dH}) have no direst connection with the
596: $\Upsilon$-term. This fact makes a difference in coefficients in
597: front of the contact term in Eq.(\ref{dH}) and the $\Upsilon$-term
598: in Eq.(\ref{Ups}) acceptable. In particular, the fact that they
599: have opposite signs produces no contradiction.
600:
601:
602:
603: \section{Coulomb problem}
604: \label{Coulomb problem}
605:
606:
607: Consider the pure Coulomb field, presuming that it is created by a
608: point-like heavy object with charge $Z>0$. Then for $r>0$ one has
609: \begin{eqnarray}
610: \label{UF}
611: U=-\frac{Z\alpha}{r}~,\quad \Upsilon=0~.
612: \end{eqnarray}
613: The second identity here follows from Eq.(\ref{Ups}).
614:
615: \subsection{Perturbation theory}
616:
617: Let us treat the Coulomb problem using the non-relativistic
618: perturbation theory. Take the non-relativistic Eq.(\ref{nonrel})
619: as a starting point, and consider the Hamiltonian Eq.(\ref{dH}) as
620: a perturbation. Conventional calculations, see Appendix
621: \ref{appendix}, lead to the following result for the shift of the
622: energy level characterized by the main quantum number $n$, orbital
623: momentum $l$ and total angular momentum $j=l,l\pm 1$
624: \begin{eqnarray}
625: \label{nlj}
626: \delta E_{nlj}= \frac{m(Z\alpha)^4}{n^3}\left(
627: \frac{3}{8n}-\frac{1}{2j+1} \right)~.
628: \end{eqnarray}
629: This formula is similar to the one that describes the energy
630: shifts for spinor particles; the only distinction comes from
631: values of $j$ in Eq.(\ref{nlj}), which are integers for vector
632: particles and half-integers for spinors. The order of several
633: lowest levels shows the following pattern
634: \begin{eqnarray}
635: \label{pattern}
636: \begin{array}{llllll}
637: n=1 &~~1 s_1\,; & & & &
638: \\
639: n=2 &~~2p_0, &~~\{2s_1, & ~~2p_1\}, &~~2p_2\,; &
640: \\
641: n=3 &~~3p_0, &~~\{3s_1, & ~~3p_1, &~~3d_1\},&
642: \\
643: & & &~~\{3p_2, & ~~3d_2\}, &~~3d_3~.
644:
645: \end{array}
646: \end{eqnarray}
647: Here the atomic-like notation $nl_j$ is adopted, the brackets
648: combine together the degenerate energy levels.
649:
650:
651:
652: \subsection{Central field}
653: \label{central}
654:
655: Consider the static central electric field (the Coulomb problem
656: gives an important example). The conservation of the total angular
657: momentum $j$ in this field allows one to separate the angular
658: variables. We will use for this purpose the electric, longitudinal
659: and magnetic spherical vectors, $\boldsymbol{
660: Y}^{(e)}_{jm},\boldsymbol{ Y}^{(l)}_{jm}, \boldsymbol{
661: Y}^{(m)}_{jm}$ defined conventionally, see \cite{LL4} and
662: Appendix \ref{spherical}. Generically, one can present the vector
663: $\bf W$ as a linear combination of three spherical vectors with
664: the given value of $j$. It is convenient to refer to the three
665: terms in this combination as the electric, longitudinal and
666: magnetic modes (or polarizations) of a vector boson. The parity
667: conservation simplifies the problem further on. The state with the
668: magnetic polarization, which parity is different from the parity
669: of other two modes, is not coupled with these modes. Therefore the
670: magnetically polarized mode can be written in a simple form
671: \begin{eqnarray}
672: \label{linm}
673: \boldsymbol{ W}= f \,\boldsymbol{ Y}^{(m)}_{jm}~,
674: \end{eqnarray}
675: where $f=f(r)$ is the radial function. The two modes related to
676: electric and longitudinal polarizations have same parity, which
677: makes coupling between these modes possible. One needs therefore
678: to consider them on the same footing assuming that
679: \begin{eqnarray}
680: \label{linel}
681: \boldsymbol{ W}= u \,\boldsymbol{ Y}^{(e)}_{jm}+v \,\boldsymbol{ Y}^{(l)}_{jm}~,
682: \end{eqnarray}
683: where $u=u(r),\,v=v(r)$. We will refer to them as
684: electro-longitudinal modes, or polarizations.
685:
686: \subsection{Magnetic polarization, $j\ge 1$}
687: \label{magnetic}
688: For the magnetic mode the angular momentum is restricted $j\ge 1$
689: (the magnetic spherical vector is not defined for $j=0$, see
690: Eq.(\ref{Y})). Substituting Eq.(\ref{linm}) into Eq.(\ref{wa}) one
691: finds the following equation for the radial function $f$
692: \begin{eqnarray}
693: \label{wCoulomb}
694: \big( \Delta_j+(\varepsilon+
695: Z\alpha/r)^2-m^2 \big)\,f=0~.
696: \end{eqnarray}
697: Here $\Delta_j$ is
698: \begin{eqnarray}
699: \label{Dj}
700: \Delta_j= \frac{1}{r^2} \frac{d}{dr} \left(r^2\frac{d}{dr}\right)
701: -\frac{j(j+1)}{r^2} ~.
702: \end{eqnarray}
703: The form of Eq.(\ref{wCoulomb}) coincides with the Klein-Gordon
704: equation. Therefore the spectrum of the magnetic mode
705: replicates the spectrum of scalar particles, which is given by
706: the Sommerfeld formula
707: \begin{eqnarray}
708: \label{Mspe}
709: \varepsilon=m\left(1+\frac{(Z\alpha)^2}{\left(\gamma+n-j-1/2\right)^2}
710: \right)^{-1/2}~.
711: \end{eqnarray}
712: Here
713: \begin{eqnarray}
714: \label{gam}
715: \gamma=\left( \left( j+1/2\right)^2-(Z\alpha)^2 \right)^{1/2} ~.
716: \end{eqnarray}
717: In Eq.(\ref{Mspe}) $n=1,2 \dots$ plays a role of the main quantum
718: number. In the non-relativistic limit the magnetic mode corresponds
719: to the states $2p_1,3d_2,4f_3,...$.
720:
721: \subsection{Electro-longitudinal polarizations, $j\ge 1$}
722: \label{j>0}
723: Consider electro-longitudinal polarizations, when the vector $\bf
724: W$ is given by Eq.(\ref{linel}). Substituting it into
725: Eqs.(\ref{wa}),(\ref{simp}) and using the properties of the
726: spherical vectors from Appendix \ref{spherical} one finds a system
727: of coupled equations for radial functions $u,v$
728: \begin{eqnarray}
729: \label{sys1}
730: \big( \Delta_j+(\varepsilon
731: +Z\alpha/r )^2-m^2 \big)u=
732: -2\sqrt{j(j+1)}\,\,\frac{v}{r^2}&,&\quad\quad
733: \\ \label{sys2}
734: \big( \Delta_j+(\varepsilon+Z\alpha/r)^2-m^2 \big)v=
735: -2\sqrt{j(j+1)}\,\,\frac{u}{r^2}&&\quad\quad
736: \\ \nonumber
737: +\frac{2v}{r^2}-
738: \frac{2Z\alpha\,w}{r^2}~.&&
739: \end{eqnarray}
740: Here $w=w(r)$ denotes the radial part of ${\mathsf w}$. Using
741: Eq.(\ref{simp}) one finds for it
742: \begin{eqnarray}
743: \label{divW}
744: &&{\mathsf w}=w\,Y_{jm}~,
745: \\ \label{divWrad}
746: &&w =
747: \frac{1}{\varepsilon+Z\alpha/r}
748: \left(-\sqrt{j(j+1)}\,\,\frac{u}{r}+\frac{dv}{dr}+\frac{2v}{r}\right).~~\quad
749: \end{eqnarray}
750: Eqs.(\ref{sys1}),(\ref{sys2}) are sufficient to define the
751: functions $u,v$, but it is convenient to compliment them by the
752: radial form of Eq.(\ref{0}), which reads
753: \begin{eqnarray}
754: \label{W0}
755: \big( \Delta_j+(\varepsilon+Z\alpha/r)^2-m^2 \big) w =
756: \frac{2 Z\alpha v}{r^2}~.
757: \end{eqnarray}
758: Let us verify first that Eqs.(\ref{sys1}),(\ref{sys2}) describe
759: two different modes. Consider with this purpose distances so small
760: that $m \ll Z\alpha/r$, where the potential energy dominates over
761: mass. In this region Eqs.(\ref{sys1}),(\ref{sys2}) reduce to
762: \begin{eqnarray}
763: \label{sim1}
764: &&\left( \frac{d^2}{dr^2}+\frac{2}{r}\frac{d}{dr}+
765: \frac{(Z\alpha)^2-j(j+1)}{r^2}\right) u \\
766: \nonumber
767: &&
768: \quad\quad\quad\quad\quad\quad\quad\quad\quad~
769: + 2 \left( j(j+1)\right)^{1/2}\frac{v}{r^2}=0,\\
770: \label{sim2}
771: &&\left(\frac{d^2}{dr^2}+\frac{4}{r}\frac{d}{dr}+
772: \frac{(Z\alpha)^2-j(j+1)+2}{r^2}\right) v=0.\quad\quad
773: \end{eqnarray}
774: One derives from Eqs.(\ref{sim1}),(\ref{sim2}) that there exists a
775: mode, in which at small distances $v$ is small, $|v|\ll |u|$,
776: which means that in this region the polarization is predominantly
777: electric. From Eq.(\ref{sim1}) one finds that this mode satisfies
778: the following asymptotic conditions at $r\rightarrow 0$
779: \begin{eqnarray}
780: \label{uas}
781: u \rightarrow a \, r^{\gamma-1/2}~,\quad |v| \ll |u|~.
782: \end{eqnarray}
783: We will call it the ``$\gamma-1/2$'' mode below.
784:
785: In order to find the second mode let us assume the following
786: asymptotic behavior for $r\rightarrow 0$
787: \begin{eqnarray}
788: \label{anu}
789: &&u\rightarrow b\,r^\nu~,\\
790: \label{bnu}
791: &&v\rightarrow c\,r^\nu~,
792: \end{eqnarray}
793: Substituting Eqs.(\ref{anu}),(\ref{bnu}) in
794: Eqs.(\ref{sim1}),(\ref{sim2}) one finds a system of two
795: homogeneous linear equations, in which $\nu$ plays a role of
796: the eigenvalue. Solving this system one finds $\nu$ and the ratio
797: $c/b$, deriving
798: \begin{eqnarray}
799: \label{nu}
800: && u\rightarrow b\,\,r^{\gamma-3/2}~,\\
801: \label{banu}
802: && v\rightarrow b\,\frac{\gamma-1/2 }{ \sqrt{ j(j+1)} } \,\,\,
803: r^{\gamma-3/2}.~
804: \end{eqnarray}
805: This mode will be referred to as the ``$\gamma-3/2$'' mode \cite{l=jpm1}.
806:
807: Let us find now the discrete energy spectrum. Introduce a
808: function $g=g(r)$
809: \begin{eqnarray}
810: \label{chi}
811: g&=&Z\alpha u+\sqrt{j(j+1)}\,w
812: \\ \nonumber
813: &=&Z\alpha u
814: +\frac{ \sqrt{j(j+1)} }{ \varepsilon+Z\alpha/r }
815: \left(-\sqrt{j(j+1)} \,\frac{u}{r}+\frac{dv}{dr}
816: +\frac{2v}{r}\right)~.
817: \end{eqnarray}
818: Here Eq.(\ref{divWrad}) was used in the second identity. Taking
819: the corresponding linear combination of
820: Eqs.(\ref{sys1}),(\ref{W0}) one finds that $g$ satisfies the
821: Klein-Gordon equation
822: \begin{eqnarray}
823: \label{Dchi}
824: \big( \,\Delta_j+(\varepsilon+Z\alpha/r)^2-m^2\,
825: \big)
826: g=0~.
827: \end{eqnarray}
828: This result leaves only two options; either $g$ equals zero
829: identically, or, alternatively, the spectrum of
830: electro-longitudinal modes can be found from Eq.(\ref{Dchi}). The
831: first alternative takes place for $j=0$, when only the
832: longitudinal mode is present. The function $u$ in this case should
833: be taken as zero, which makes zero also the function $g$ in
834: Eq.(\ref{chi}). Thus, Eq.(\ref{Dchi}) provides no help for $j=0$
835: states.
836:
837: For $j\ge 1$ the function $g$ is nonzero, for both
838: ``$\gamma-1/2$'' and ``$\gamma-3/2$'' modes, see Appendix
839: \ref{non0}. Eq.(\ref{Dchi}) defines the spectrum, which therefore
840: satisfies the Sommerfeld formula Eq.(\ref{Dchi}). In the
841: non-relativistic limit the mixed electric-longitudinal modes
842: correspond to the following states with $j\ge 1$: $1s_1, 2p_2,
843: 3d_1, 3d_3, 4f_2 \ldots$.
844:
845: \subsection{Longitudinal polarization, $j=0$}
846: \label{j=0}
847:
848: Consider zero angular momentum $j=0$, which corresponds to purely
849: longitudinal polarization, see Eq.(\ref{Y}). The state with $j=0$
850: is described by one radial function $v=v(r)$,
851: \begin{eqnarray}
852: \label{pres}
853: \boldsymbol{ W}= v \,\boldsymbol{n}~,\quad\quad
854: \boldsymbol{n}=\boldsymbol{r}/r~.
855: \end{eqnarray}
856: The radial function $v$ satisfies Eqs.(\ref{sys1}),(\ref{sys2}) in
857: which the function $u$ is to be put to zero (electric polarization
858: for $j=0$ is impossible). These equations therefore yield
859: \begin{eqnarray}
860: \label{v}
861: \frac{d^2v}{dr^2}+\frac{2}{r}\,\frac{dv}{dr}&+&\left(
862: \left(\varepsilon +Z\alpha/r\right)^2-m^2
863: \right)v\\
864: \nonumber
865: & =& \frac{2v}{r^2}
866: -\frac{2Z\alpha}{r^2}\,
867: \frac{1}{\varepsilon+Z\alpha/r} \left(\frac{dv}{dr}
868: +\frac{2v}{r}\right)~.
869: \end{eqnarray}
870: In order to make the physical meaning of this equation more
871: transparent let us eliminate the first derivative by means of a
872: substitution $v\rightarrow \varphi$
873: \begin{eqnarray}
874: \label{phin}
875: v =\frac{Z \alpha} {\varepsilon^2} \,\left(
876: \varepsilon+\frac{Z\alpha}{r}\right)\,
877: \frac{\varphi}{r}=
878: \frac{1+x}{x^2}\, \varphi~.
879: \end{eqnarray}
880: where it is convenient also to scale the radial variable
881: $r\rightarrow x$
882: \begin{eqnarray}
883: \label{x}
884: r=\frac{Z\alpha}{\varepsilon}\,x~,
885: \end{eqnarray}
886: assuming $\varphi=\varphi(x)$. In this notation Eq.(\ref{v}) can
887: be rewritten as a conventional Schr\"odinger-type eigenvalue
888: problem
889: \begin{eqnarray}
890: \label{schr}
891: && H\,\varphi=-\varkappa^{\,2}\varphi~, \\
892: \label{H}
893: && H=-\frac{d^2}{dx^2}
894: -\frac{2(Z\alpha)^2}{x}
895: -\frac{(Z\alpha)^2}{x^2}+\frac{2}{(x+1)^2}~,\quad\quad
896: \end{eqnarray}
897: where $-\varkappa^2$, which plays a role of an eigenvalue, is
898: related to the energy of the discrete level
899: \begin{eqnarray}
900: \label{kappa}
901: \varkappa^{\,2}=(Z\alpha)^2\frac{m^2-\varepsilon^2}{\varepsilon^2}>0~.
902: \end{eqnarray}
903: The operator $H$ in Eq.(\ref{H}) possesses three singular points,
904: $x=0,~x=\infty$ and $x=-1$. The last one lies in the non-physical
905: region, but it presents an obstacle for an analytical study
906: anyway. One can overcome this difficulty using a substitution
907: $\varphi \rightarrow \tilde \varphi$
908: \begin{eqnarray}
909: \label{for}
910: \varphi=\left( \frac{d}{dx}
911: +\left(\gamma+1/2\right)\frac{x+1}{x}-\frac{1}{x+1}
912: \right)\tilde \varphi~.
913: \end{eqnarray}
914: It can be shown that $\tilde \varphi$ satisfies an eigenvalue
915: problem
916: \begin{eqnarray}
917: \label{H2}
918: &&\tilde H\tilde \varphi=-\varkappa^2\,\tilde \varphi~,
919: \\ \label{psi1}
920: &&\tilde H=
921: -\frac{d^2}{dx^2}-2\frac{(Z\alpha)^2}{x}
922: +\Big(\gamma+\frac{1}{2}\Big)\Big(\gamma+\frac{3}{2}\Big)\frac{1}{x^2}~.\quad
923: \quad\quad
924: \end{eqnarray}
925: The main result of the transformation Eq.(\ref{for}) is that the
926: operator $\tilde H$ has only two singular points, $x=0$ and
927: $x=\infty$. An interesting method, which allows one to ``invent''
928: the substitution Eq.(\ref{for}) and derive then Eq.(\ref{H2}) is
929: presented in Appendix \ref{matrix}. It takes its origins in an
930: elegant treatment of quantum mechanics developed by the G\"otingen
931: School and known as matrix mechanics.
932:
933: A regular at $r=0$ solution of the eigenvalue problem (\ref{H2})
934: reads
935: \begin{eqnarray}
936: \label{psi2in}
937: \tilde \varphi=e^{-\varkappa x}x^{L+1}
938: F\left(L+1-\frac{(Z\alpha)^2}{\varkappa},2L+2,2\varkappa
939: x\right).\quad
940: \end{eqnarray}
941: Here $F(\alpha,\beta, z)$ is the confluent hypergeometric function
942: and $L$ is defined by
943: \begin{eqnarray}
944: \label{L}
945: L=\gamma+1/2~.
946: \end{eqnarray}
947: To make the solution given by Eq.(\ref{psi2in}) regular at
948: infinity one should assume that
949: \begin{eqnarray}
950: \label{Ry}
951: \varkappa=
952: \frac{(Z\alpha)^2}{L+n-1}=
953: \frac{(Z\alpha)^2}{\gamma+n-1/2}
954: ~,\quad n=2,3\ldots~,~~
955: \end{eqnarray}
956: The corresponding eigenfunctions are given by Eq.(\ref{psi2in}),
957: in which the hypergeometric function is reduced to a polynomial
958: \begin{eqnarray}
959: \label{psi2}
960: \tilde \varphi=e^{-\varkappa x}x^{\gamma+3/2} F(2-n,2\gamma+3,2\varkappa x
961: )~.
962: \end{eqnarray}
963: Eqs.(\ref{psi2}),(\ref{for}) give then the function $\varphi$,
964: while (\ref{phin}),(\ref{pres}) transform it into $v$ and
965: $\boldsymbol W$. The function $\varphi$ exhibits the following
966: behavior at the boundaries
967: \begin{eqnarray}
968: \label{sens1}
969: &\varphi \propto \exp(-\varkappa \,x)~,\quad &x\rightarrow \infty~,
970: \\
971: \label{sens2}
972: &\varphi \propto x^{\gamma+1/2}~, \quad &x\rightarrow 0~.
973: \end{eqnarray}
974: Eq.(\ref{Ry}) gives the spectrum
975: \begin{eqnarray}
976: \label{sati}
977: \varepsilon=m\left(1+\frac{(Z\alpha)^2}{\left(\gamma+n-1/2\right)^2}
978: \right)^{-1/2}\!\!,\quad n=2,3\ldots\quad
979: \end{eqnarray}
980: which complies with the Sommerfeld formula Eq.(\ref{Mspe}). In the
981: non-relativistic limit the longitudinal mode corresponds to the
982: following states with $j=0$: $2p_0, 3p_0, 4p_0\ldots$
983:
984: \subsection{Summary for Coulomb problem}
985:
986: \label{Spectrum}
987:
988: Our discussion of the Coulomb problem for vector particles
989: confirms that for all polarizations and all angular momenta $j$
990: the discrete energy spectrum is described by the Sommerfeld
991: formula Eq.(\ref{Mspe}), as was first found by Corben and
992: Schwinger \cite{corben-schwinger_1940}.
993:
994: For $j\ge 1$ there exist three modes. One of them is purely
995: magnetic, it has $l=j$, while two others are constructed from the
996: electric and longitudinal polarizations, each one of these two
997: modes has an admixture of $l=j+1$ and $l=j-1$ states. These two
998: modes coexist for $j\ge 1$, while for $j=0$ only one of them,
999: which in this case has a purely longitudinal polarization and
1000: $l=1$ is present.
1001:
1002: From Eq.(\ref{Mspe}) one derives that the spectrum of the Coulomb
1003: problem is degenerate; it is triply degenerate provided $n \ge
1004: j+2,~j\ge 1$, doubly degenerate for levels with $n=j+1,~j\ge 1$,
1005: while the states which have either $n=j$ or $j=0$ remain
1006: non-degenerate. This conclusion agrees with the non-relativistic
1007: expansion, see Eq.(\ref{pattern}).
1008:
1009: Interestingly, one and the same Sommerfeld formula Eq.(\ref{Mspe})
1010: describes the discrete energy spectrum in the Coulomb problem for
1011: scalar, Dirac and vector particles. The only distinction is
1012: related to the angular momentum $j$, which takes the integer
1013: values $j=0,1\ldots $ for bosons and half-integer $j=1/2,3/2\ldots
1014: $ for fermions.
1015:
1016: \section{catastrophe with charge}
1017: \label{catastroph}
1018:
1019: Consider the charge density of a vector boson for a state with
1020: $j=0$. Eqs.(\ref{ro}),(\ref{pres}) give
1021: \begin{eqnarray}
1022: \label{rode}
1023: \rho^W= 2e\left[ \left(\varepsilon+\frac{Z\alpha}{r}\right)(v^2+w^2)
1024: +2v \frac{dw}{dr}-\Upsilon\, w^2\right],\quad
1025: \end{eqnarray}
1026: where $w$ defined by Eqs.(\ref{simp}),(\ref{divW}),(\ref{divWrad}) reads
1027: \begin{eqnarray}
1028: \label{wj0}
1029: w=\frac{1}{\varepsilon+Z\alpha/r} \left(\frac{dv}{dr}+
1030: \frac{2v}{r}\right)~.
1031: \end{eqnarray}
1032: In the region of small distances $r\ll Z\alpha/m$ Eq.(\ref{sens2})
1033: shows that $\varphi \propto r^{\gamma+1/2}$. Consequently, from
1034: Eqs. (\ref{phin}),(\ref{wj0}) we find the following estimates for
1035: $v$ and $w$
1036: \begin{eqnarray}
1037: \label{estv}
1038: v &\sim& r^{\gamma-3/2}~,
1039: \\ \label{estw}
1040: w &\sim& \frac{\gamma+1/2}{Z\alpha} r^{\gamma-3/2}~,
1041: \end{eqnarray}
1042: From here one derives an estimate for the charge density
1043: (\ref{rode}) of the vector boson
1044: \begin{eqnarray}
1045: \label{cdvb}
1046: \rho^W\sim -2e\,
1047: \frac{(1-\gamma)(1+2\gamma)}{Z\alpha}\,r^{2\gamma-4}~,
1048: \quad r>0~.
1049: \end{eqnarray}
1050: It diverges at the origin so badly that the total charge $Q^W=\int
1051: \rho^W\,d^3r$ localized in any small sphere surrounding the origin
1052: is infinite.
1053:
1054: The trouble does not stop here. Remember the density $\rho=
1055: Z|e|\,\delta(\boldsymbol{r})$ of the Coulomb charge, which is
1056: located at the origin. This density results in the
1057: $\Upsilon$-term defined by Eq.(\ref{Ups})
1058: \begin{eqnarray}
1059: \label{delta}
1060: \Upsilon= \frac{e\rho}{\,m^2} =
1061: -\frac{4\pi Z\alpha}{m^2}\,\delta(\boldsymbol{r})~.
1062: \end{eqnarray}
1063: We did not consider it previously because the functions we dealt
1064: with were regular at the origin, allowing one to hope that their
1065: regular behavior makes the $\Upsilon$-term irrelevant. Since the
1066: charge density does not follow this pattern, we need to take the
1067: term given by Eq.(\ref{delta}) into account. The contribution of
1068: the $\delta$-function in Eq.(\ref{delta}) to the boson charge
1069: density is given by the last term in Eq.(\ref{rode}), which reads
1070: \begin{eqnarray}
1071: \label{ro4}
1072: (\rho^W)_{\,\Upsilon-\mathrm{term}}
1073: =e \frac{8\pi
1074: Z\alpha}{m^2}\,w^2(0)\,\delta(\boldsymbol{r}) ~.
1075: \end{eqnarray}
1076: Eq.(\ref{estw}) shows that $w(0)=\infty$, which makes the density
1077: Eq.(\ref{ro4}) infinite as well.
1078:
1079: We see that there are two closely located, though different
1080: regions, which contribute to an infinite charge of the $W$-boson
1081: in the $j=0$ state. One region is $r>0$, where the density of
1082: charge Eq.(\ref{cdvb}) behaves singularly as $r\rightarrow 0$.
1083: Another region is located strictly at the origin $r=0$, where an
1084: infinite coefficient $w^2(0)=\infty$ in front of the
1085: $\delta$-function in Eq.(\ref{ro4}) makes the charge infinite as
1086: well.
1087:
1088: The origin of Eq.(\ref{ro4}) can be traced down to the last term
1089: in Eq.(\ref{ro}). It contributes therefore to the charge density
1090: for all states. There is one more state, in which the coefficient
1091: in Eq.(\ref{ro4}) turns infinite, signaling a catastrophic
1092: behavior of the charge. This is the ``$\gamma-3/2$'' state with
1093: $j=1$, see Eqs.(\ref{nu}),(\ref{banu}). To justify this
1094: statement, note that Eqs.(\ref{divWrad}),(\ref{nu}) and (\ref{banu})
1095: imply that $w\propto r^{\gamma-3/2}$. For $j=1$ the inequality
1096: $\gamma< 3/2$ holds. Therefore for the state ``$\gamma-3/2$'',
1097: $j=1$ one finds $w(0)=\infty$, which makes the charge of the
1098: $W$-boson located strictly at the origin infinite. (There is no
1099: problem in that case with the charge in the region $r>0$.)
1100:
1101: The catastrophic behavior of the charge of the $W$-boson in $j=0$
1102: and $j=1$, ``$\gamma-3/2$'' states was discovered in
1103: \cite{corben-schwinger_1940}, forcing the authors of this work to
1104: conclude that the pure Coulomb problem for $W$-bosons is poorly
1105: defined.
1106:
1107:
1108: \section{Vacuum polarization}
1109: \label{vacuum_polarization}
1110:
1111:
1112: Consider the conventional QED vacuum polarization. The potential
1113: energy of the $W$-boson propagating in the Coulomb field acquires
1114: an additional term, let us call it $S(r)$, which describes the
1115: polarization
1116: \begin{eqnarray}
1117: \label{pot}
1118: U(r)=-\Big(\,1+S(r)\,\Big) \frac{Z\alpha}{r}~.
1119: \end{eqnarray}
1120: It suffices to consider the polarization effect in the
1121: lowest-order approximation, when it is is described by the known
1122: Uehling potential. Its small-distance asymptotic behavior is
1123: given by a simple logarithmic function, see e. g. \cite{LL4},
1124: \begin{eqnarray}
1125: \label{write}
1126: S(r)\simeq -\alpha \beta \,\ln\left(m_{Z}r\right),
1127: \quad
1128: r\rightarrow 0~.
1129: \end{eqnarray}
1130: This function is related to the logarithm responsible for the
1131: scaling of the QED coupling constant
1132: \begin{eqnarray}
1133: \label{log}
1134: \alpha^{-1}(\mu)=\alpha^{-1}(\mu_0)-\beta\,\ln(\mu/\mu_0)~.
1135: \end{eqnarray}
1136: The relation between Eqs.(\ref{write}) and (\ref{log}) is
1137: well-known, see e.g. book \cite{LL4}, which presents it for one
1138: generation of leptons. The factor $\beta$, which governs the
1139: scaling of the coupling constant and the potential in
1140: Eq.(\ref{write}) equals the lowest coefficient of the Gell-Mann
1141: - Low $\beta$-function. It is normalized here in such a way
1142: that for one generation of leptons $\beta=\beta_{e}=2/3\pi$.
1143:
1144:
1145: It is important for us that $\alpha(\mu)$ rises with the mass
1146: parameter $\mu$, i.e. $\beta$ is positive, $\beta>0$;
1147: theoretical and experimental data agree on this fact, for a
1148: brief review see e. g. Ref. \cite{eidelman-et-al_2004}, the
1149: experimental data are provided by Refs. \cite{TOPAZ,VENUS,OPAL}.
1150: An estimation of $\beta$ can be found from two reliable
1151: reference points $ \alpha^{-1}(m_\tau) = 133.498 \pm 0.017$ and
1152: $\alpha^{-1}(m_Z) = 127.918 \pm 0.018$ provided in
1153: Ref.\cite{eidelman-et-al_2004}. Using them and taking the
1154: masses $m_\tau= 1776.99 + 0.29 - 0.26$ Mev and $m_Z= 91.1876 \pm
1155: 0.0021$ Gev recommended in \cite{eidelman-et-al_2004} one
1156: derives from Eq.(\ref{log}) that
1157: \begin{eqnarray}
1158: \label{bet}
1159: \beta \simeq 1.42(1)~.
1160: \end{eqnarray}
1161: More simple estimation of $\beta$ can be done if one takes into
1162: account a contribution of all known charged fermions ``naively''
1163: (neglecting complications, related to the QCD vacuum as well as
1164: possible contribution of scalars). This estimate yields
1165: \begin{eqnarray}
1166: \label{betr}
1167: \beta_\mathrm{est} \approx \frac{2}{3\pi}\sum_i \frac{q^2_i}{e^2}
1168: =\frac{2}{3\pi}\left(1+\frac{5}{3}\right)3\simeq 1.70~.
1169: \end{eqnarray}
1170: Here summation runs over all charged fermions, $q_i$ is the
1171: charge of the fermion, the terms 1 and 3/5 in the bracket are
1172: due to the lepton and quark contribution for one generation, the
1173: factor 3 after the bracket accounts for three generations. A
1174: discrepancy between ``simple-minded'' Eq.(\ref{betr}) and more
1175: solid-based Eq.(\ref{bet}) is below 20\%. The normalization of
1176: the logarithmic function on the mass of the $Z$-boson $m_Z$
1177: adopted in Eq.(\ref{write}) presumes that the fine-structure
1178: constant $\alpha $ is taken at precisely this scale,
1179: $\alpha\equiv\alpha(m_Z)\simeq 1/128$.
1180:
1181:
1182: We are interested in high-momenta behavior in
1183: Eqs.(\ref{write}),(\ref{log}), where $\mu\sim 1/r \gg m$. An
1184: accuracy of Eqs.(\ref{bet}),(\ref{betr}), as well as any other
1185: feasible estimation, is limited in this region by a contribution
1186: of unknown heavy charged fermions and scalars. However, this
1187: uncertainty does not affect our final conclusions. For our
1188: purposes it suffices to stick to a widely accepted hypothesis
1189: that $\beta$ is a positive constant (or a slow-varying function
1190: up to the Grand Unification limit).
1191:
1192: Substituting Eqs.(\ref{pot}),(\ref{write}) into Eq.(\ref{Ups})
1193: one derives
1194: \begin{eqnarray}
1195: \label{estUps}
1196: \Upsilon(r)&\simeq&
1197: \frac{Z\alpha^2 \beta}{m^2r^3},\quad
1198: r\rightarrow 0~,
1199: \end{eqnarray}
1200: where the lowest term of the $\alpha$-expansion is retained. It
1201: is vital that for small distances, when $r\ll \sqrt{\alpha}/m$,
1202: $\Upsilon(r)$ is positive and large,
1203: \begin{eqnarray}
1204: \label{large}
1205: \Upsilon(r) \gg |U(r)|\gg m~.
1206: \end{eqnarray}
1207: Note that the direct contribution of the vacuum polarization
1208: given by the term $S(r)$ in Eq.(\ref{pot}) is not pronounced. In
1209: contrast, the $\Upsilon$-term Eq.(\ref{estUps}) becomes dominant
1210: at small distances, making the effects related to the QED vacuum
1211: polarization very important. Since this term plays a crucial
1212: role below, let us verify its sign again. Consider a positive
1213: Coulomb center, $Z>0$. Then the vacuum polarization produces
1214: negative charge density, $\rho <0 $. Since the charge of the
1215: $W^-$ meson is negative, $e<0$, we find from Eq.(\ref{Ups}) that
1216: $\Upsilon =e \rho/m^2>0$. We see that indeed, the
1217: $\Upsilon$-term is positive, in accord with Eq.(\ref{estUps}).
1218:
1219:
1220: \subsection{Longitudinal polarization, j=0}
1221: \label{longitudinal_polarization_j=0}
1222:
1223: Eq.(\ref{pres}) shows that a longitudinal state with $j=0$ is
1224: described by the single radial function $v=v(r)$. Eq.(\ref{simp})
1225: allows one to express the function $w$ via $v$
1226: \begin{eqnarray}
1227: \label{wv}
1228: w= \left(\varepsilon-U-\Upsilon\right)^{-1}\,\big(v'+2v/r\big)~.
1229: \end{eqnarray}
1230: We need now to write the classical equation of motion for $v$, in
1231: which the term $\Upsilon$ is taken into account. The simplest way
1232: is to substitute $\boldsymbol{W}$ and $w$ from
1233: Eqs.(\ref{pres}),(\ref{wv}) into Eq.(\ref{tran}), which yields
1234: \begin{eqnarray}
1235: \label{radv}
1236: && \big((\varepsilon-U)^2\!-m^2\big)v=
1237: \\ \nonumber
1238: && -(\varepsilon-U)\frac{d}{dr} \left(\,
1239: \frac{v'+2v/r }{\varepsilon-U-\Upsilon}
1240: \,\right)
1241: -U'\,\frac{v'+2v/r}{\varepsilon-U-\Upsilon}.
1242: \end{eqnarray}
1243: It is taken into account here that Eq.(\ref{pres}) ensures that
1244: $\boldsymbol{\nabla \!\times \!W}= 0$. For a purely Coulomb case,
1245: when $\Upsilon=0$ for $r>0$, Eq.(\ref{radv}) reduces to
1246: Eq.(\ref{v}). Eq.(\ref{radv}) can be rewritten in a more compact
1247: form
1248: \begin{eqnarray}
1249: \label{sec}
1250: v''+G \, v'+H \,v=0~,
1251: \end{eqnarray}
1252: where the coefficients $G=G(r)$ and $H=H(r)$ are
1253: \begin{eqnarray}
1254: \label{g}
1255: G &=& \frac{2}{r}+\frac{U'}{\varepsilon-U}+\frac{U'+\Upsilon'}
1256: {\varepsilon-U-\Upsilon}~,
1257: \\ \label{h}
1258: H &=&-\frac{2}{r^2}+\frac{2}{r}
1259: \left( \frac{U'}{\varepsilon-U}+\frac{U'+\Upsilon'}{\varepsilon-U-\Upsilon}
1260: \right)
1261: \\ \nonumber
1262: &&\quad \quad \quad +\frac{\varepsilon-U-\Upsilon}{\varepsilon-U}
1263: \big( \, (\varepsilon-U)^2-m^2 \,\big)~.
1264: \end{eqnarray}
1265: For a qualitative analyses it is convenient to eliminate the
1266: term with the first derivative by scaling the radial function
1267: $v\rightarrow\varphi=\varphi(r)$
1268: \begin{eqnarray}
1269: \label{phi}
1270: v=\frac{1}{r}\Big[ \, \big(\varepsilon-U\,\big)
1271: \big(\varepsilon-U-\Upsilon\,\big)
1272: \,\Big]^{1/2}\,\varphi~.
1273: \end{eqnarray}
1274: (This definition reduces to Eq.(\ref{phin}) when $\Upsilon=0$).
1275: The classical equation of motion for $W$-bosons takes a simple
1276: form
1277: \begin{eqnarray}
1278: \label{phi''}
1279: && -\varphi''+{\mathcal U}\,\varphi=0~,
1280: \\ \label{ugh}
1281: &&{\mathcal U}=-\,H+G^{\,2}/4+G{\,'}/2~,
1282: \end{eqnarray}
1283: where $G,H$ are defined in Eqs.(\ref{g}),(\ref{h}).
1284: Eq.(\ref{phi''}) can be looked at as a Schr\"odinger-type
1285: equation, in which ${\mathcal U}={\mathcal U}(r)$ plays the role of an
1286: effective potential energy.
1287:
1288:
1289:
1290: According to Eqs.(\ref{estUps})(\ref{large}) the $\Upsilon$-term
1291: is large and positive at small distances. This fact makes the
1292: effective potential ${\mathcal U}(r)$ in Eq.(\ref{ugh}) also large and
1293: positive when $r\rightarrow 0$
1294: \begin{eqnarray}
1295: \label{larpos}
1296: {\mathcal U}(r)\simeq -\,H(r)\simeq -\,U(r)\,\Upsilon(r)\simeq
1297: \frac{Z^2\alpha^3\beta}{m^2r^4}~.
1298: \end{eqnarray}
1299: Compare this result with the effective potential $ [ \,{\mathcal
1300: U}(r)\,]_\mathrm{C}$ for the pure Coulomb field. The latter one is a
1301: part of the Hamiltonian in Eq.(\ref{H}). For $r\rightarrow
1302: 0$ one finds from Eq.(\ref{schr}) that
1303: \begin{eqnarray}
1304: \label{UC}
1305: [ \,{\mathcal U}(r)\,]_\mathrm{C}\simeq -(Z\alpha)^2/r^2.
1306: \end{eqnarray}
1307: It is taken into account here that the variables $x$ and $r$ in
1308: Eq.(\ref{H}) are proportional, see Eq.(\ref{x}).
1309: Eq.(\ref{larpos}) shows that the vacuum polarization produces a
1310: strong repulsion in the effective potential ${\mathcal U}(r)$, in
1311: contrast with a mild attraction, which exhibits $[\,{\mathcal
1312: U}(r)\,]_\mathrm{C}$ in Eq.(\ref{UC}) for the pure Coulomb
1313: case.
1314:
1315: When the estimate Eq.(\ref{larpos}) is applicable,
1316: Eq.(\ref{phi''}) allows an analytical solution
1317: \begin{eqnarray}
1318: \label{re}
1319: \varphi(r)\propto
1320: r \,\exp\left(\!-\frac{Z
1321: \alpha\,(\alpha \beta)^{1/2}}{mr} \,\right).~~
1322: \end{eqnarray}
1323: It shows that $\varphi(r)$ exponentially decreases at small
1324: distances. According to Eqs.(\ref{wv}),(\ref{phi}) the
1325: functions $v(r),w(r)$, also decrease exponentially here;
1326: correspondingly, the charge density of the $W$-boson
1327: Eq.(\ref{rode}) decreases exponentially at the origin as well
1328: \begin{eqnarray}
1329: \label{vexp}
1330: v&\rightarrow& \frac{a}{m}\,\frac{1}{r^2}
1331: \exp\left(\!-\frac{Z
1332: \alpha\,( \alpha \beta)^{1/2}}{mr} \,\right)~,
1333: \\
1334: \label{wexp}
1335: w &\rightarrow &
1336: -\frac{a}{(\alpha \beta)^{1/2} }\,\frac{1}{r}
1337: \exp\left(\!-\frac{Z
1338: \alpha\,(\alpha \beta)^{1/2}}{mr} \,\right)~,
1339: \\
1340: \label{rhoexp}
1341: \rho^W &\rightarrow&
1342: -\frac{4 a^2 Z\alpha \,e}{m^2 }\,\frac{1}{r^5}
1343: \exp\left(\!-\frac{2Z
1344: \alpha\,(\alpha \beta)^{1/2}}{mr} \,\right).\quad
1345: \end{eqnarray}
1346: Here a constant $a$ depends on the normalization of $v$, which
1347: is specified in Eq.(\ref{norma}) below. Eq.(\ref{rhoexp}) shows
1348: that for $r>0$ in the vicinity of the origin the charge density
1349: is finite and small; which makes the charge located in this
1350: region finite as well. Eq.(\ref{wexp}) shows that $w(0)=0$,
1351: which eradicates the contribution of the $\delta$-function in
1352: Eq.(\ref{ro4}). Thus, the charge located strictly at origin
1353: $r=0$ is zero.
1354:
1355: We verified that an account of the QED vacuum polarization
1356: erases the infinite charge of a vector boson for $j=0$ state.
1357:
1358: \subsection{Electro-longitudinal polarizations, $j\ge 1$}
1359: \label{Electro-longitudinal polarizations, j>0}
1360:
1361: Eq.(\ref{large}) shows that in the region of small distances
1362: $r\ll \alpha/m$ the $\Upsilon$-term, which is related to the
1363: vacuum polarization, is large. This fact makes the function $w$
1364: in Eq.(\ref{simp}) small, $|w|\ll |\boldsymbol{W}|$. As a
1365: result the asymptotic form of the equation of motion
1366: (\ref{tran}) at small distances reads
1367: \begin{eqnarray}
1368: \label{tran1}
1369: \frac{(Z\alpha)^2\!\!}{r^2}\,\,\boldsymbol{W}=
1370: \boldsymbol{\nabla \times}(\boldsymbol{
1371: \nabla \times W})~.
1372: \end{eqnarray}
1373: Eq.(\ref{linel}) ensures that $ \boldsymbol{ \nabla \times W}$
1374: is not zero identically provided $j\ge 1$, which makes
1375: Eq.(\ref{tran1}) meaningful.
1376:
1377: Using Eq.(\ref{linel}) to represent the electro-longitudinal
1378: modes and identities Eqs.(\ref{rot}) for the spherical vectors
1379: one rewrites Eq.(\ref{tran1}) in terms of the radial functions
1380: $u,v$
1381: \begin{eqnarray}
1382: \label{rad u}
1383: (Z\alpha)^2u=-r^2 u''-r u' -u+ \sqrt{j(j+1)}\,v\,, \quad\quad\, &&
1384: \\ \label{rad v}
1385: (Z\alpha)^2 v=-\sqrt{ j(j+1)}\,( \,r u' + u-
1386: \sqrt{j(j+1)}\,v \,).&&\quad
1387: \end{eqnarray}
1388: Their solution is straightforward
1389: \begin{eqnarray}
1390: \label{sol u}
1391: u&=& b \,r^{\lambda}~,
1392: \\ \label{sol v}
1393: v&=&b \,\sqrt{
1394: j(j+1)}\,\frac{\lambda+1}{\gamma^2-1/4}\,r^{\lambda}~.
1395: \end{eqnarray}
1396: Here $b$ is a constant, and $\lambda$ can take one of the two
1397: possible values,
1398: \begin{eqnarray}
1399: \label{lpm}
1400: \lambda=\lambda_\pm=\frac{1}{2}\,\,\frac{j(j+1)\pm k}
1401: {\gamma^2-1/4}~,
1402: \end{eqnarray}
1403: where $k$ satisfies
1404: \begin{eqnarray}
1405: \label{k}
1406: k^2= j^2(j+1)^2
1407: -4(Z\alpha)^2\left(\gamma^2-\frac{5}{4}\right)\left(\gamma^2
1408: -\frac{1}{4}\right),\quad
1409: \end{eqnarray}
1410: with $\gamma$ defined in Eq.(\ref{gam}). The two available
1411: values of $\lambda_\pm$ should be attributed to the two
1412: electro-longitudinal modes.
1413:
1414: Comparing Eqs.(\ref{sol u}),(\ref{sol v}), which are valid when
1415: the vacuum polarization is taken into account, with
1416: Eqs.(\ref{uas}) and Eqs.(\ref{nu}),(\ref{banu}), which describe
1417: the purely Coulomb case, we see that the polarization changes
1418: drastically the behavior of the wave functions. One finds from
1419: Eqs.(\ref{lpm}),(\ref{k}) that $\lambda_\pm$ are positive for
1420: all $j$, $j\ge 1$, provided $Z$ is not very large, $Z\alpha\le
1421: 1/2$.
1422:
1423: From Eqs. (\ref{sol u}),(\ref{sol v}) one deduces therefore
1424: that $u(r),v(r)\rightarrow 0$, when $r\rightarrow 0$.
1425: Eq.(\ref{simp}) guarantees then that $w(0)=0$. As a result the
1426: contribution of the $\delta$-function in Eq.(\ref{ro4}) to the
1427: charge density turns zero for all electro-longitudinal modes.
1428: This differs qualitatively from the pure Coulomb case, which
1429: gives an infinite charge located at the origin for $j=1$,
1430: ``$l$''$=0$ state.
1431:
1432: We conclude that the QED vacuum polarization suppresses the wave
1433: functions of a vector boson at the origin, eradicating thus the
1434: infinite charge of the boson, which plagues the problem for the
1435: pure Coulomb field.
1436:
1437: \section{Numerical example}
1438: \label{numericals}
1439: To be more informative on the behavior of vector bosons in the
1440: Coulomb field let us solve the corresponding equations of motion
1441: numerically. Consider the $j=0$ state, describing it with the
1442: help of Eqs.(\ref{pres}) and (\ref{radv}). We need to specify
1443: the factor $S(r)$, which describes the vacuum polarization in
1444: the potential in Eq.(\ref{pot}). For small $r$, $r\ll
1445: Z\alpha/m$ this factor plays a major role, while for larger it
1446: is less important. Let us construct a simple model, which gives
1447: a correct asymptotic behavior Eq.(\ref{write}) as $r\rightarrow
1448: 0$, and is physically reasonable, though not perfect, at larger
1449: $r$. Take with this purpose the Uehling potential, see e.g.
1450: \cite{LL4}, assuming that only charged leptons and quarks
1451: contribute to it
1452: \begin{eqnarray}
1453: \label{Ueh}
1454: S(r)=\frac{2\alpha}{3\pi}\sum_i \,\frac{q_i^2}{e^2}\, F(m_ir)~.
1455: \end{eqnarray}
1456: Here
1457: \begin{eqnarray}
1458: \label{F(x)}
1459: F(x)= \int_1^\infty \!\!\exp(-2x\zeta) \left( 1+\frac{1}{2\zeta^2}\right)
1460: \frac{\sqrt{\zeta^2-1}}{\zeta^2}\,\,d\zeta~.
1461: \end{eqnarray}
1462: Summation in Eq.(\ref{Ueh}) runs over all quarks and charged
1463: leptons, their charges $q_i$ and masses $m_i$ are taken from
1464: Ref. \cite{eidelman-et-al_2004}. The model presented by
1465: Eq.(\ref{Ueh}) neglects complications related to the QCD vacuum,
1466: which may be substantial at large distances, but the role of the
1467: polarization is insignificant in this region anyway. For small
1468: distances $r\rightarrow 0$ the model Eq.(\ref{Ueh}) reproduces
1469: the correct asymptotic formula Eq.(\ref{write}), in which the
1470: coefficient $\beta$ is given by Eq.(\ref{betr}). A comparison
1471: of Eqs.(\ref{bet}) and (\ref{betr}) shows that at small
1472: distances the accuracy of the model potential can be estimated
1473: as $\sim 20$\%, which is sufficient for us.
1474: \begin{figure}
1475: \centering
1476: \includegraphics[height=5.3 cm,
1477: keepaspectratio = true,
1478: %angle = -90
1479: %%
1480: ]{vForLargeDistances}
1481: \centering
1482: \includegraphics[height=5.3 cm,
1483: keepaspectratio = true,
1484: %angle = -90
1485: %%
1486: ]{vForSmallDistances}
1487: \caption{ \label{one} The $2p_1$ radial function $v(r)$, which describes the
1488: $n=2$, $j=0$ discrete state of a $W$-boson in the Coulomb field of a
1489: charge $Z=1$. (a) Large distances $r\gg \hbar/mc$, $v(r)$ is close
1490: to conventional non-relativistic wave function $2p$; (b)
1491: ultra-relativistic region $r\ll \hbar/mc$. Solid line - numerical
1492: solution, dashed line - analytical prediction of Eq.(\ref{vexp}). }
1493: \end{figure}
1494: \noindent
1495: \begin{figure}
1496: \centering
1497: \includegraphics[height=5.3 cm,
1498: keepaspectratio = true,
1499: %angle = -90
1500: %%
1501: ]{roForLargeDistances}
1502: \centering
1503: \includegraphics[height=5.3 cm,
1504: keepaspectratio = true,
1505: %angle = -90
1506: %%
1507: ]{roForSmallDistances}
1508: \caption{ \label{two} The charge distribution $\rho^W(r)$ for the $2p_1$ state
1509: of a $W$-boson in the Coulomb field of a charge $Z=1$. (a)
1510: Non-relativistic region of large distances $r\gg \hbar/mc$; (b)
1511: ultra-relativistic region $r\ll \hbar/mc$. Solid line - numerical
1512: solution, dashed line - analytical prediction of Eq.(\ref{rhoexp}),
1513: dashed-dotted line - the pure Coulomb case Eq.(\ref{cdvb}), when the
1514: charge density diverges at the origin (shown with arbitrary
1515: normalization). }
1516: \end{figure}
1517: \noindent
1518: Using $S(r)$ from Eq.(\ref{Ueh}) one defines the potential $U(r)$
1519: Eq.(\ref{pot}) and the $\Upsilon$-potential in Eq.(\ref{Ups}).
1520: After that the solution of Eq.(\ref{radv}) is straightforward.
1521: This solution should be normalized on the total charge $e$ of the
1522: $W$-boson
1523: \begin{eqnarray}
1524: \label{norma}
1525: e=\int \rho^{W}(r)\, d^3r~,
1526: \end{eqnarray}
1527: where the charge density is defined by Eq.(\ref{rode}).
1528:
1529: Fig.\ref{one} shows the radial function $v(r)$ for the $2p_1$
1530: state ($n=2$, $j=0$) for $Z=1$. For large distances $mr\gg 1$
1531: the function $v(r)$ is close to the conventional
1532: non-relativistic wave function of the $2p$ state, as it should
1533: be, see Section \ref{Nonrel}. In the ultra-relativistic region
1534: $mr \ll 1 $ the function $v(r)$ changes sign, then shows an
1535: extremum, and decreases exponentially when $r \rightarrow 0$, in
1536: accord with an analytical estimate Eq.(\ref{vexp}), which is
1537: shown by the dashed line. The fact that the function $v(r)$ has
1538: a node in the relativistic region produces no controversy since
1539: $v(r)$ is not a proper wave function and the conventional
1540: theorem, which counts the nodes of the wave functions for
1541: discrete levels is not applicable.
1542:
1543: In our discussion we did not try to construct proper wave
1544: functions, being content with a possibility to calculate the
1545: current. As an example, Fig.\ref{two} shows the charge density
1546: for the $2p_1$ state in the Coulomb field of $Z=1$. In the
1547: non-relativistic region $mr\gg 1$ it behaves conventionally.
1548: For the ultra-relativistic case $mr\ll 1$ the density changes
1549: sign, exhibits an extremum and then decreases exponentially when
1550: $r\rightarrow 0$ in agreement with Eq.(\ref{rhoexp}). Note that
1551: the ``wrong'' sign of the charge density, i. e. the positive
1552: charge density for the negatively charged $W$-boson
1553: (Fig.\ref{two} shows $\rho^W/e=-\rho^W/|e|$) produces no
1554: contradiction with general principles. The Pauli's theorem, see
1555: e. g. Ref. \cite{LL4}, implies that the energy of a boson field
1556: is positively defined, but the sign of the charge density of the
1557: boson field remains not unequivocally determined \cite{example}.
1558:
1559: The total positive charge located at small distances proves be
1560: very small; for $2p_1$, $Z=1$ state it is
1561: \begin{eqnarray}
1562: \label{wrong}
1563: Q^W_\mathrm{+}=4\pi \int_0^{r_0} \rho^W(r)\,r^2\,dr\simeq
1564: 0.676\cdot 10^{-14}\,|e|~.
1565: \end{eqnarray}
1566: Here $r_0\simeq 1.01\cdot 10^{-3}m^{-1}$ is a node of $\rho^W
1567: (r)$.
1568:
1569: It is instructive to compare the found charge density with the
1570: one in the pure Coulomb problem, which is shown in Fig.
1571: \ref{two} (b) by the dashed-dotted line that reproduces
1572: Eq.(\ref{cdvb}). One should be content in this case with an
1573: arbitrary normalization, since for the pure Coulomb field the
1574: normalization integral in Eq.(\ref{norma}) is divergent. Fig.
1575: \ref{two} (b) illustrates the fact that the vacuum polarization reduces
1576: the charge density at the origin.
1577:
1578: The energy shift $\delta \varepsilon$ of the level $2p_1$
1579: ($\delta \varepsilon$ is a deviation of energy from the
1580: Sommerfeld formula Eq.(\ref{Mspe})) due to the vacuum
1581: polarization is found to be $\delta \varepsilon/m=-1.90 \cdot
1582: 10^{-7}$. In relative units it is much bigger than the Lamb shift
1583: in atoms $\delta \varepsilon_\mathrm{LS}/m_e\sim Z^4 \alpha^5$.
1584: The reason is obvious. The Lamb shift in atoms originates mostly
1585: from within the Compton distances $r\sim r_c= 1/m_e$, which are
1586: smaller than the Bohr radius for the electron $r_\mathrm{B}\sim
1587: 1/(Z\alpha m_e)$. For $W$-bosons the situation is different.
1588: Light fermions, which contribute to Eq.(\ref{Ueh}), allow the
1589: polarization potential to spread to large distances, as far as
1590: the Bohr radius of the $W$-boson, $r \sim 1/m_i \sim 1/(Z\alpha
1591: m)$. Therefore the energy shift due to the vacuum polarization
1592: gains substantial contribution from the non-relativistic region,
1593: where the wave function is large, which makes $\delta
1594: \varepsilon$ large as well (large compared to the Lamb shift in
1595: atoms). The accuracy of the energy shift calculations is limited
1596: by an accuracy of our model at large distances. The contribution
1597: of the QCD vacuum, which could be substantial here, is not
1598: described properly by the model based on Eq.(\ref{Ueh}).
1599: Consequently, the presented above value for the energy shift
1600: should be considered only as an estimate.
1601:
1602: Nevertheless, one can derive an important lesson from this
1603: estimate. The found energy shift is small on the absolute scale,
1604: being lower than the non-relativistic binding energy by a factor
1605: of $|\delta \varepsilon|\times 4/(Z^2\alpha^2m) \sim 1.4\cdot
1606: 10^{-2}$. Thus, the dramatic variation of the function $v(r)$ at
1607: the origin, which is produced by the vacuum polarization, makes
1608: only small impact on the spectrum. This is in contrast to a
1609: strong influence, which the vacuum polarization exercises on the
1610: charge distribution of vector bosons. The fact that the energy
1611: shift is small makes the Sommerfeld formula Eq.(\ref{Mspe}) a
1612: good approximation for discrete energy levels.
1613:
1614: \section{Discussion}
1615: \label{conclusion}
1616:
1617: We demonstrated that the conventional QED vacuum polarization
1618: plays a very important, defining role in the Coulomb problem for
1619: vector bosons. Let us summarize the reasons leading to this
1620: conclusion. The Uehling potential, which describes the vacuum
1621: polarization in the simplest approximation is known to be a
1622: weakly attractive and slowly varying function. For spinor
1623: particles it produces a small enhancement of the fermion wave
1624: functions on the Coulomb center. For vector bosons the situation
1625: is different because the equations of motion for vector
1626: particles explicitly incorporate the external current. As a
1627: result, the density of the polarized charge $\rho$ comes into
1628: the equations of motion for vector bosons. The corresponding
1629: term in the equations was called the $\Upsilon$-potential,
1630: $\Upsilon =e\rho/m^2$. The charge density $\rho$ is negative for
1631: an attractive Coulomb center, $\rho<0$ when $Z>0$, being
1632: singular on the Coulomb center, $|\rho| \sim 1/r^3$. One derives
1633: from this that the vacuum polarization produces a repulsive
1634: $\Upsilon$-potential, $\Upsilon=e\rho/m^2 \propto 1/r^3 >0$
1635: (remember, $e<0$). Since the $\Upsilon$-term is singular at the
1636: origin, it plays a dominant role at small distances.
1637:
1638: Strong effective repulsion produced by the $\Upsilon$-potential
1639: reduces the fields, which describe $W$-bosons on the Coulomb
1640: center. For $j=0$ this reduction is dramatic, exponential. For
1641: $j=1$, ''$\gamma-3/2$'' the suppression is of a more moderate
1642: power-type nature, but in both cases it is strong enough to
1643: eliminate the infinite charge, which is located at the origin in
1644: the pure Coulomb approximation.
1645:
1646: The above comments appeal to a chain of calculations. It is
1647: interesting to look at the obtained result from a more general
1648: perspective. The renormalizabity of the Standard Model implies
1649: that by renormalizing relevant physical quantities one is {\em
1650: bound} to obtain sensible physical results. The relevant
1651: quantity in question is the charge density of a vector boson.
1652: It follows from this that the important physical quantity, which
1653: should be renormalized, is the coupling constant. Its
1654: renormalization is effectively fulfilled when the vacuum
1655: polarization is taken into account. Thus, it makes sense that
1656: the account of the vacuum polarization results in acceptable
1657: physical results.
1658:
1659: A proposed approach is very straightforward, which makes the
1660: Coulomb problem for vector bosons as simple and reliable as it
1661: is for scalars and spinors. All discrete energy levels can be
1662: easily evaluated, all relevant fields can be calculated and
1663: normalized properly. Presumably, all scattering data can also
1664: be evaluated, though the scattering problem was not discussed in
1665: detail in the present work. All these quantities include the
1666: Coulomb charge $Z$ accurately, not relying on perturbation
1667: theory. Starting from this base, one can consider all other
1668: processes left outside the scope of the Coulomb problem by
1669: treating them as perturbations. This includes the conventional
1670: QED processes, such as the radiative decay, photoionization, the
1671: radiative corrections. This includes also processes related to
1672: possible exchange of Higgs and $Z$-bosons.
1673:
1674: Previous attempts to formulate the Coulomb problem for vector
1675: bosons within the framework of the Standard Model have been
1676: facing a difficulty related to an infinite charge of the boson
1677: located near an attractive Coulomb center. This work finds that
1678: the polarization of the QED vacuum eradicates the problem.
1679: Usually the QED radiative corrections produce only small
1680: perturbations. It is interesting that in the case discussed the
1681: radiative correction plays a major, defining role.
1682:
1683:
1684: % \acknowledgment
1685:
1686: This work was supported by the Australian Research Council. One
1687: of us (VF) appreciates support from Department of Energy, Office
1688: of Nuclear Physics, Contract No. W-31-109-ENG-38.
1689:
1690: \appendix
1691:
1692: \section{Homogeneous magnetic field}
1693: \label{homo}
1694: For a static homogeneous magnetic field Eq.(\ref{form}) reads
1695: \begin{eqnarray}
1696: \label{mag}
1697: \left( \varepsilon^2-m^2\right)\boldsymbol{W} =
1698: -\left( \boldsymbol{\nabla}-ie \boldsymbol{ A}\right)^2\boldsymbol{ W}-2ie
1699: \boldsymbol{B}\times \boldsymbol{W}.~~
1700: \end{eqnarray}
1701: Assuming that the magnetic field is directed along the $z$-axis
1702: and introducing the new variables $w_\sigma,~\sigma=0,\pm1$, $
1703: w_{\pm 1}=(\boldsymbol{ W}_x\pm i\boldsymbol{ W}_y)/\sqrt 2$, $
1704: w_{0}=\boldsymbol{ W}_z$ one rewrites Eq.(\ref{mag}) in a simple
1705: form
1706: \begin{eqnarray}
1707: \label{wsig}
1708: \left( \varepsilon^2-m^2\right)\,w_\sigma =
1709: -\left( \boldsymbol{ \nabla}-ie \boldsymbol{ A}\right)^2w_\sigma
1710: +
1711: 2eB\sigma\,w_\sigma~,
1712: \end{eqnarray}
1713: which looks similar to the non-relativistic Schr\"odinger equation
1714: for a particle in the homogeneous magnetic field. This similarity
1715: allows one to write the spectrum Eq.(\ref{e2}).
1716:
1717:
1718: \section{relativistic corrections to energy levels}
1719: \label{appendix}
1720:
1721: Here we present separate expectation values for four relativistic
1722: corrections in the same order as they appear in Eq.(\ref{dH}). For
1723: $l=0$
1724: \begin{eqnarray}
1725: \label{nl=0j}
1726: \delta E_{n0j}= \frac{m(Z\alpha)^4}{n^3}\left[
1727: \left(\frac{3}{8n}-1\right)+0+ \frac{2}{3}+0\right]~.
1728: \end{eqnarray}
1729: Here and below we specify the terms having zero expectation values
1730: by writing the corresponding zeros explicitly. For $l\ge 1 $
1731: \begin{eqnarray}
1732: \nonumber
1733: \delta E_{nlj}= \frac{m(Z\alpha)^4}{n^3} \left[\,
1734: \left(\frac{3}{8n}-\frac{1}{2l+1} \right)\right. +
1735: \\ \label{nl>0j}
1736: \frac{\langle ls\rangle }{l(l+1)(2l+1)}
1737: +0 \\ \nonumber
1738: \left.
1739: - \frac{6 \langle ls \rangle^2+3\langle ls
1740: \rangle-4l(l+1)}{l(l+1)(2l-1)(2l+1)(2l+3)}\,\right]~,
1741: \end{eqnarray}
1742: where
1743: \begin{equation}
1744: \label{ls}
1745: \langle ls \rangle=\frac{1}{2}\,[j(j+1)-l(l+1)-2]~.
1746: \end{equation}
1747: Both Eq.(\ref{nl=0j}) and Eq.(\ref{nl>0j}) lead to Eq.(\ref{nlj}).
1748: The total relativistic correction would look very complicated and
1749: show no degeneracy if magnetic dipole or electric quadrupole
1750: moments of a vector particle are different from those values,
1751: which follow from the gauge theory.
1752:
1753:
1754: \section{Spherical vectors}
1755: \label{spherical}
1756: The conventional definition of spherical vectors, see e.g.
1757: \cite{LL4}, reads
1758: \begin{eqnarray}
1759: \label{Y}
1760: \begin{array}{l}
1761: \boldsymbol{ Y}^{(e)}_{jm}= \boldsymbol{
1762: \nabla}_n\,Y_{jm}/\sqrt{j(j+1)}~,
1763: \\
1764: \boldsymbol{ Y}^{(l)}_{jm}= \boldsymbol{ n} \,Y_{jm}~,
1765: \\
1766: \boldsymbol{ Y}^{(m)}_{jm}= \boldsymbol{ n \times Y}^{(e)}_{jm}
1767: ~.
1768: \end{array}
1769: \end{eqnarray}
1770: Here $Y_{jm}\equiv Y_{jm}(\theta,\varphi)$ is the spherical
1771: function, $\boldsymbol{ Y}^{(e)}_{jm},\boldsymbol{ Y}^{(l)}_{jm},
1772: \boldsymbol{ Y}^{(m)}_{jm}$ are the electric, longitudinal and
1773: magnetic vectors. The symbol $\boldsymbol{ \nabla}_n$ in
1774: Eq.(\ref{Y}) indicates the angular part of the gradient, $
1775: \boldsymbol{ \nabla}F(\theta,\phi)= \boldsymbol{ \nabla}_n
1776: F(\theta,\phi)/r$, and $\boldsymbol{ n}=\boldsymbol{ r}/r$ is a
1777: unit vector along the radius vector.
1778:
1779: Definitions Eqs.(\ref{Y}) imply the following properties of the
1780: spherical vectors
1781: \begin{eqnarray}
1782: \label{div}
1783: \begin{array}{l}
1784: \boldsymbol{ \nabla }_n \cdot \boldsymbol{ Y}^{(e)}_{jm}
1785: =-\sqrt{j(j+1)}\,\,Y_{jm}\,,
1786: \\
1787: \boldsymbol{ \nabla}_n \cdot \boldsymbol{ Y}^{(l)}_{\!jm}
1788: =2\,Y_{jm}\, ,
1789: \\
1790: \boldsymbol{ \nabla}_n \cdot \boldsymbol{ Y}^{(m)}_{jm}=0\,.
1791: \end{array}
1792: \\ \label{rot}
1793: \begin{array}{l}
1794: \boldsymbol{ \nabla \times Y}^{(e)}_{jm}=
1795: \boldsymbol{ Y}^{(m)}_{jm}\, ,
1796: \\
1797: \boldsymbol{ \nabla \times Y}^{(l)}_{jm}=
1798: -\sqrt{j(j+1)}\,\boldsymbol{ Y}^{(m)}_{jm}~,
1799: \\
1800: \boldsymbol{ \nabla \times Y}^{(m)}_{jm}=
1801: -\boldsymbol{ Y}^{(e)}_{jm}-\sqrt{j(j+1)}\,\boldsymbol{ Y}^{(l)}_{jm}~.
1802: \end{array}
1803: \end{eqnarray}
1804: The formulas for the Laplace operator read
1805: \begin{eqnarray}
1806: \label{ddd}
1807: \begin{array}{l}
1808: \Delta_n\boldsymbol{ Y}^{(\,e\,)}_{jm} = -j(j+1)\,\boldsymbol{ Y}^{(e)}_{jm}+
1809: 2\sqrt{j(j+1)} \, \,\boldsymbol{ Y}^{(l)}_{jm},
1810: \\
1811: \Delta_n\boldsymbol{ Y}^{(\,l\,)}_{jm}=2\sqrt{j(j+1)}\,\,
1812: \boldsymbol{ Y}^{(e)}_{jm}
1813: - \big(j(j+1)+2\big) \boldsymbol{ Y}^{(l)}_{jm},
1814: \\
1815: \Delta_n\boldsymbol{ Y}^{(m)}_{jm} =-j(j+1)\boldsymbol{
1816: Y}^{(m)}_{jm}.
1817: \end{array}
1818: \end{eqnarray}
1819: Here $\Delta_n$ describes the angular part of the Laplacian, i.e.
1820: $\Delta F(\theta,\phi) =\Delta_nF/r^2$. The parity for electric
1821: and longitudinal polarizations equals $P=(-1)^{j}$, for magnetic
1822: polarization the parity is $P=(-1)^{j+1}$. The orbital moment $l$
1823: takes the value $l=j$ for the magnetic polarization, in agreement
1824: with the parity for this state. The electric and longitudinal
1825: polarizations are constructed as linear combinations of the two
1826: states with $l=j\pm 1$. For $j=0$ there exists only one spherical
1827: vector, which is purely longitudinal and has $l=1$.
1828:
1829: \section{Spectrum of electro-longitudinal modes for $j\ge 1$}
1830: \label{non0}
1831:
1832: Let us verify that for $j\ge 1$ the function $g$ introduced in
1833: Eq.(\ref{chi}) is nonzero. Consider first the ``$\gamma-1/2$''
1834: mode. Substituting Eq.(\ref{uas}) into Eq.(\ref{chi}) one finds
1835: \begin{eqnarray}
1836: \label{chiA}
1837: g\rightarrow a\frac{1}{Z\alpha}(1/4-\gamma^2)\,r^{\gamma-1/2}~,
1838: \end{eqnarray}
1839: which indicates that in this mode $g$ is not zero.
1840:
1841: Consider now the ''$\gamma-3/2$'' mode, which incorporates both
1842: possible polarizations at small distances. We need here the
1843: expressions for $u$ and $v$ at small distances that are more
1844: accurate, then the ones in Eq.(\ref{nu}). They can be derived by
1845: using $mr\ll 1$ as a perturbation in
1846: Eqs.(\ref{sim1}),(\ref{sim2}), and pushing calculations one step
1847: beyond the simplest approximation given by
1848: Eqs.(\ref{nu}),(\ref{banu}). The result reads
1849: \begin{eqnarray}
1850: \label{anext}
1851: && u\rightarrow b\left( 1-\frac{2+(Za)^2}{\gamma+1/2}
1852: \,\,\frac{\varepsilon r}{Z\alpha}\right)r^{\gamma-3/2}~,\\
1853: \label{bnext}
1854: && v\rightarrow \frac{ b }{ \sqrt{j(j+1)} }\left(
1855: \gamma-\frac{1}{2}
1856: -Z\alpha \varepsilon r
1857: \right)r^{\gamma-3/2}.~
1858: \end{eqnarray}
1859: Substituting Eqs.(\ref{anext}),(\ref{bnext}) into Eq.(\ref{chi})
1860: one finds that the main term $\propto r^{\gamma-3/2} $ cancels out
1861: in $g$, but the next one survives, giving
1862: \begin{eqnarray}
1863: \label{chiB}
1864: g \rightarrow b \frac{\varepsilon}{Z\alpha}
1865: \left(2\gamma-1-(Z\alpha)^2\right)\,r^{\gamma-1/2}~.
1866: \end{eqnarray}
1867: We verified that for $ j\ge1$ the function $g$ is not an identical
1868: zero for both electro-longitudinal modes.
1869:
1870: \section{Longitudinal mode $j=0$ and matrix mechanics}
1871: \label{matrix}
1872:
1873: In order to find the spectrum of the operator $H$ in Eq.(\ref{H})
1874: let us employ a method, which finds its inspiration in an elegant
1875: approach to quantum mechanics developed by the G\"otingen School
1876: and often called the {\it matrix mechanics}; the book Ref.
1877: \cite{green} gives its systematic presentation. We modify it for
1878: our purposes as follows. Assume that one needs to find discrete
1879: spectrum of some Hermitian operator ${\mathcal H}$ (in our case it is the
1880: operator ${\mathcal H}$ in Eq.(\ref{H})). Let us presume that one is able to
1881: find the operator $\theta$, which satisfies
1882: \begin{eqnarray}
1883: \label{theta}
1884: {\mathcal H}=\theta^\dag\theta+\lambda_0~,
1885: \end{eqnarray}
1886: where $\lambda_0$ is a number. Define then a new operator $\tilde {\mathcal H}$,
1887: \begin{eqnarray}
1888: \label{m+1}
1889: \tilde {\mathcal H}=\theta\,\theta^\dag+\lambda_0~.
1890: \end{eqnarray}
1891: Let us verify that the two operators ${\mathcal H},\tilde {\mathcal H}$ have very similar sets
1892: of eigenvalues. Consider an eigenfunction $\varphi$ of ${\mathcal H}$, with
1893: the eigenvalue $\lambda$
1894: \begin{eqnarray}
1895: \label{psin}
1896: {\mathcal H}\varphi=\lambda\,\varphi~.
1897: \end{eqnarray}
1898: Taking
1899: \begin{eqnarray}
1900: \label{psi2old}
1901: \tilde \varphi=\theta \varphi~,
1902: \end{eqnarray}
1903: one verifies that
1904: \begin{eqnarray}
1905: \label{psi1old}
1906: \tilde {\mathcal H}\,\tilde \varphi=(\theta\,\theta^\dag+ \lambda_0)\theta\,\varphi=
1907: \theta\,(\theta^\dag\,\theta+\lambda_0)\varphi
1908: \\ \nonumber
1909: =\theta\,{\mathcal H}\varphi=\lambda\,\theta\,\varphi=\lambda\tilde \varphi~.
1910: \end{eqnarray}
1911: This shows that either the function $\tilde \varphi$ is an eigenfunction of
1912: $\tilde {\mathcal H}$ with the eigenvalue $\lambda$, or $\tilde \varphi=0$. The first
1913: options makes $\lambda$ an eigenvalue of both operators ${\mathcal H},\tilde {\mathcal H}$.
1914: The second one implies that
1915: \begin{eqnarray}
1916: \label{impl}
1917: \theta\,\varphi=0~,
1918: \end{eqnarray}
1919: which indicates that $\lambda =\lambda_0$ is a candidate for an
1920: eigenvalue of ${\mathcal H}$ because Eq.(\ref{impl}) implies
1921: ${\mathcal H}\,\varphi=\lambda_0 \varphi$. Eq.(\ref{impl}) provides a
1922: convenient way to derive the corresponding eigenfunction. There is
1923: though a subtlety here. The found from Eq.(\ref{impl}) $\varphi$
1924: may, or may not satisfy the boundary conditions. If it does, then
1925: it represents the eigenfunction and $\lambda=\lambda_0$ is an
1926: eigenvalue. Otherwise, $\lambda_0$ does not belong to the discrete
1927: spectrum, as would be the case in an example discussed. One
1928: should also verify that an action of the operator $\theta$ in
1929: Eq.(\ref{psi2old}) (as well as the operator $\theta^\dag$ in
1930: Eq.(\ref{then}) below) does not spoil the boundary conditions. We
1931: presume here that this is the case, and verify later on that this
1932: assumption holds for a particular example discussed, see
1933: Eq.(\ref{sens1}),(\ref{sens2}).
1934:
1935: We conclude that any discrete eigenvalue of ${\mathcal H}$ is an eigenvalue
1936: of $\tilde {\mathcal H}$ as well, with one possible exception of $\lambda_0$. By
1937: reversing the argument, one derives that if $\tilde \varphi$ is an
1938: eigenfunction of $\tilde {\mathcal H}$ with the eigenvalue $\tilde\lambda$,
1939: \begin{eqnarray}
1940: \label{l'}
1941: \tilde {\mathcal H}\tilde \varphi=\tilde\lambda\tilde \varphi
1942: \end{eqnarray}
1943: then
1944: \begin{eqnarray}
1945: \label{then}
1946: \varphi=\theta^\dag\tilde \varphi
1947: \end{eqnarray}
1948: satisfies Eq.(\ref{psin}) with $\lambda=\tilde\lambda$. We see
1949: that the two sets of discrete eigenvalues of the two operators $
1950: {\mathcal H},\tilde {\mathcal H} $ are same, except for $\lambda_0$, which may, or may
1951: not be present in one, or both sets of spectra. The crucial for us
1952: point is that the operator $\tilde {\mathcal H}$ can be more simple for
1953: analyses than the initial operator ${\mathcal H}$.
1954:
1955: Taking the operator ${\mathcal H}$ from Eq.(\ref{H}), we construct the
1956: operators $\theta,\theta^\dag$ in the form
1957: \begin{eqnarray}
1958: \label{the}
1959: \theta&=&-\frac{d}{dx}+a+\frac{b}{x}+\frac{c}{x+1}~,
1960: \\ \label{the'}
1961: \theta^\dag &=&~~\,\frac{d}{dx}+a+\frac{b}{x}+\frac{c}{x+1}~,
1962: \end{eqnarray}
1963: where $a,b,c$ are real numbers. From Eqs.(\ref{the}),(\ref{the'})
1964: it follows that
1965: \begin{eqnarray}
1966: \label{thth}
1967: \theta^\dag\theta=-\frac{d^2}{dx^2}+a^2&+&\frac{ 2b(a+c) }{ x }
1968: +\frac{ b(b-1) }{x^2}\\ \nonumber
1969: &&+\frac{ 2c(a-b)}{x+1} +\frac{c(c-1)}{(x+1)^2}~.
1970: \end{eqnarray}
1971: There are four $x$-dependent rational terms in Eq.(\ref{thth}),
1972: while only three coefficients $a,b,c$ are available for tuning to
1973: make them identical to similar terms present in the operator ${\mathcal H}$.
1974: However, the coefficients in Eq.(\ref{H}) prove to be
1975: ``user-friendly'', making this procedure possible. Taking
1976: \begin{eqnarray}
1977: \label{abc}
1978: &a=b=\gamma+1/2\,,\quad \quad c=-1~,
1979: \\ \label{lambda0}
1980: &\lambda_0=-\left(\gamma+1/2\right)^2~,
1981: \end{eqnarray}
1982: one satisfies Eq.(\ref{theta}). Taking $\theta,\theta^\dag$
1983: defined in Eqs.(\ref{the}),(\ref{the'}) and (\ref{abc}) one
1984: constructs $\tilde {\mathcal H}$ Eq.(\ref{psi1old}), with the result given in
1985: Eq.(\ref{psi1}). The ``nasty'' singular at $x=-1$ term disappears
1986: from $\tilde {\mathcal H}$. The latter operator describes a conventional
1987: Coulomb-type problem with $L=\gamma+1/2 $ playing a role of an
1988: effective (non-integer) angular momentum. From Eq.(\ref{H2}) one
1989: finds that regular at $x=0$ solution of the eigenvalue problem
1990: $\tilde {\mathcal H}\tilde \varphi=-\varkappa^2\,\tilde \varphi$, satisfies
1991: Eq.(\ref{psi2in}). Eq.(\ref{Ry}), which ensures that this
1992: solution is regular at infinity, completely defines a set of
1993: discrete eigenvalues of $\tilde {\mathcal H}$.
1994:
1995: The set of eigenvalues of $\tilde {\mathcal H}$ gives the eigenvalues of the
1996: original operator ${\mathcal H}$, except for possibly one additional
1997: eigenvalue $\lambda_0$, which is discussed below. The
1998: eigenfunctions of ${\mathcal H}$ can be found from Eq.(\ref{then}). Using
1999: Eqs.(\ref{the'}),(\ref{abc}) one presents them in a form of
2000: Eq.(\ref{for}).
2001:
2002: In order to verify whether $\lambda_0$ is an eigenvalue of ${\mathcal H}$ one
2003: needs to find $\varphi$ from Eq.(\ref{impl}). Eq.(\ref{the}) gives
2004: \begin{eqnarray}
2005: \label{l0}
2006: \left(-\frac{d}{dx} +\left
2007: (\gamma+1/2\right)\frac{x+1}{x}- \frac{1}{x+1}\right)\varphi=0~,
2008: \end{eqnarray}
2009: which leads to
2010: \begin{eqnarray}
2011: \label{New-zero}
2012: \varphi=(x+1)^{-1}\,x^{\gamma+1/2}\,\exp\,[\,(\gamma+1/2)\,x\,]~.
2013: \end{eqnarray}
2014: Since this function is singular at $x=\infty$, it cannot be an
2015: eigenfunction. Consequently $\lambda_0$ is not an eigenvalue.
2016:
2017: The function $\varphi$ defined by Eq.(\ref{for}) exhibits regular
2018: behavior at both boundaries Eqs.(\ref{sens1}),(\ref{sens2}). This
2019: ensures that $\varphi$ is an eigenfunction. Note, that specifying
2020: the operators $\theta,\theta^\dag$ one had an additional option.
2021: One could have chosen in Eqs.(\ref{abc}) and all the following
2022: relevant formulas $-\gamma$ instead of $\gamma$. It this case,
2023: however, instead of Eq.(\ref{sens2}) one obtains $\varphi \propto
2024: x^{-\gamma+1/2}$ for $x\rightarrow 0$, which indicates a singular,
2025: unacceptable for an eigenfunction behavior.
2026:
2027: We conclude that the full set of all discrete eigenvalues of ${\mathcal H}$
2028: is specified by Eq.(\ref{Ry}). The corresponding eigenfunctions
2029: are given by Eqs.(\ref{for}),(\ref{psi2}).
2030:
2031:
2032: \begin{thebibliography}{99}
2033:
2034:
2035: % \bibitem{overduin-wesson_2003}
2036: %
2037: % J. M. Overduin and P S Wesson, {\it Dark Sky, Dark Matter} Series
2038: % in Astronomy and Astrophysics, Bristol, England, Institute of
2039: % Physics Press, 2003.
2040:
2041: % \bibitem{seiber-witten-1994-1}
2042: %
2043: % ELECTRIC - MAGNETIC DUALITY, MONOPOLE CONDENSATION, AND
2044: % CONFINEMENT IN N=2 SUPERSYMMETRIC YANG-MILLS THEORY. By N.
2045: % Seiberg and E. Witten, Nucl.Phys. {\bf B426}, 19 (1994);
2046: % Erratum-ibid. {\bf B430}, 485 (1994).
2047: % hep-th/9407087
2048:
2049: % MONOPOLES, DUALITY AND CHIRAL SYMMETRY BREAKING IN N=2
2050: % SUPERSYMMETRIC QCD.
2051:
2052: % \bibitem{seiber-witten-1994-2}
2053: %
2054: % N. Seiberg E. Witten, Nucl.Phys. {\bf B431}, 484 (1994).
2055: % hep-th/9408099
2056:
2057:
2058:
2059: \bibitem{proca_1936}
2060:
2061: A. Proca, Compt.Rend. {\bf 202}, 1490 (1936).
2062:
2063: \bibitem{massey-corben_1939}
2064:
2065: H. F. W. Massey and H. C. Corben, Proc.Camb.Phi.Soc. {\bf 35}, 463
2066: (1939).
2067:
2068: \bibitem{oppenheimer-snyder-serber_1940}
2069:
2070: J. R. Oppenheimer, H. Snyder and R. Serber, Phys.Rev. {\bf 57}, 75
2071: (1940).
2072:
2073: \bibitem{tamm_1940-1-2}
2074:
2075: I. E. Tamm. Phys. Rev. {\bf 58}, 952 (1940); Doklady USSR Academy of
2076: Science {\bf 8-9}, 551 (1940).
2077:
2078:
2079: \bibitem{corben-schwinger_1940}
2080:
2081: H. C. Corben and J. Schwinger, Phys.Rev. {\bf 58}, 953 (1940).
2082:
2083:
2084: \bibitem{schwinger_1964}
2085:
2086: J. Schwinger Rev. Mod. Phys. {\bf 36}, 609 (1964).
2087:
2088:
2089: \bibitem{cheng_wu_1972}
2090:
2091: H. Cheng and T. T. Wu, Phys.Rev.D {\bf 5}, 3247 (1972).
2092:
2093: \bibitem{huang_1992}
2094:
2095: K. Huang, {\it Quarks, Leptons and Gauge Fields}, 2nd edition,
2096: World Scientific, 1992.
2097:
2098:
2099:
2100: \bibitem{thooft-veltman-1972}
2101:
2102: G. 't Hooft and M.J.G. Veltman, Nucl.Phys.{\bf B44}, 189 (1972).
2103: % REGULARIZATION AND RENORMALIZATION OF GAUGE FIELDS.
2104:
2105:
2106:
2107: \bibitem{vijayalakshmi-seetharaman-mathews_1979}
2108:
2109: B. Vijayalakshmi, M. Seetharaman, and P.M. Mathews, J.Phys.A {\bf
2110: 12}, 665 (1979).
2111:
2112: \bibitem{pomeransky-khriplovich_1998}
2113:
2114: A. A. Pomeransky and I. B. Khriplovich, JETP {\bf 86}, 839 (1998).
2115:
2116:
2117: \bibitem{pomeransky-se'nkov_1999}
2118:
2119: A. A. Pomeransky and R. A. Sen'kov, Phys. Lett. B {\bf 468}, 251
2120: (1999).
2121:
2122:
2123: \bibitem{pomeransky-sen'kov-khriplovich_2000}
2124:
2125: A. A. Pomeransky, R. A. Sen'kov and I.B. Khriplovich, Phys.Usp.
2126: {\bf 43} 1055 (2000), Usp.Fiz.Nauk {\bf 43} 1129 (2000).
2127:
2128:
2129:
2130:
2131: \bibitem{silenko_2004}
2132:
2133: A. J. Silenko, Analysis of wave equations for spin 1 particles
2134: interacting with an electromagnetic field, hep-th/0404074.
2135:
2136:
2137: \bibitem{fushchych-nikitin-susloparow_1985}
2138:
2139: W. I. Fushchych, A. G. Nikitin, W. M. Susloparow. Nuovo Cimento {\bf
2140: 87}, 415 (1985).
2141:
2142:
2143: \bibitem{fushchych-nikitin_1994}
2144:
2145: W. I. Fushchych, A. G. Nikitin. {\it Symmetries of Equations of
2146: Quantum Mechanics}, N.Y. Allerton Press, 1994.
2147:
2148:
2149: \bibitem{sergheyev_1997}
2150:
2151: A. G. Sergheyev, Ukr. J. Phys. {\bf 42} (1997).
2152:
2153: \bibitem{kuchiev-flambaum_2005}
2154:
2155: M. Yu. Kuchiev and V. V. Flambaum. Coulomb problem for vector
2156: bosons versus Standard Model, hep-th/0511149.
2157:
2158: \bibitem{weinberg_2001}
2159:
2160: S. Weinberg, {\it The quantum theory of feilds, Volume II, Modern
2161: applications}, Cambridge, University Press, 2001.
2162:
2163:
2164: % \bibitem{taylor}
2165: %
2166: % J. C. Taylor et al, {\it Gauge Theories of Weak Interactions
2167: % (Cambridge Monographs on Mathematical Physics)},
2168: %% P. V. Landshoff (Series Editor), D. R. Nelson (Series Editor), D. W.
2169: %% Sciama (Series Editor), S. Weinberg (Series Editor)
2170: % Cambridge University Press, 1976.
2171:
2172:
2173: \bibitem{LL4}
2174:
2175: V. B. Berestetsky, E.M. Lifshits and L.P..Pitaevsky, {\it Quantum
2176: electrodynamics}, PergamonPress, 1982.
2177:
2178:
2179: \bibitem{foldy-wouthuysen_50}
2180:
2181: L. L. Foldy and S. A. Wouthuysen, Phys. Rev. {\bf 78}, 29
2182: (1950).
2183:
2184: \bibitem{l=jpm1}
2185:
2186: Ref.\cite{corben-schwinger_1940} uses notation ``$l$''$=j-1$ and
2187: ``$l$''$=j+1$ for the modes, which we call ``$\gamma-1/2$'' and
2188: ``$\gamma-3/2$''.
2189:
2190: % \bibitem{akhiezer}
2191: %
2192: % A. I. Akhiezer, and V. B. Berestetskii, {\it Quantum
2193: % Electrodynamics}. New York: Interscience Publishers, 1965.
2194:
2195:
2196: \bibitem{green}
2197:
2198: H. S. Green, {\it Matrix mechanics}, P.Noordhoff Ltd-Groninger-The
2199: Netherlands, 1965.
2200:
2201: \bibitem{eidelman-et-al_2004}
2202:
2203: S. Eidelman et al., Phys. Lett. B 592, 1 (2004)
2204:
2205:
2206: \bibitem{TOPAZ}
2207:
2208: I. Levine et al., Phys. Rev. Lett. {\bf 78}, 424 (1997);
2209:
2210: \bibitem{VENUS}
2211:
2212: S. Okada et al., Phys. Rev. Lett. {\bf 81}, 2428 (1998);
2213:
2214: \bibitem{OPAL}
2215:
2216: G. Abbiendi et al., Eur. Phys. J. C{\bf 13}, 553 (2000);
2217:
2218:
2219: \bibitem{example}
2220:
2221: The scalar field $\phi(r)$ propagating in a static potential
2222: $U(r)$ provides a simple example. The charge density
2223: $\rho_\phi=2e(\epsilon-U(r))\phi^*(r)\phi(r)$, where $\epsilon$
2224: is the energy, shows the ``wrong'' sign in a strongly repulsive
2225: region $\epsilon<U(r)$. It is necessarily small because the
2226: field is suppressed their; compare Eq.(\ref{wrong}), which
2227: indicates that the ``wrong'' charge of the $W$-boson in the
2228: $2p_1$ state is very small.
2229:
2230:
2231: \end{thebibliography}
2232:
2233:
2234: \end{document}
2235:
2236: