hep-th0611129/hs2.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: \usepackage{amscd}
5: \usepackage[dvips]{graphicx}
6: %==================================================================
7: \textwidth=15.5 truecm
8:   \textheight=24 truecm
9:  \hoffset=-1.5 truecm
10:  \voffset=-2 truecm
11:  \special{em:linewidth 0.4pt}
12:  \unitlength 1.10mm
13:  \linethickness{0.4pt}
14:  %========================================================================
15: \begin{document}
16: 
17: 
18: \begin{center}
19: \section*
20: {High spin particles with spin-mass coupling II}
21: \end{center}
22: \begin{center}
23: {M. DASZKIEWICZ}
24: \end{center}
25: \begin{center}
26: {Institute of Theoretical Physics, Wroc{\l}aw University, pl. Maxa
27: Borna 9\\
28: 50-206 Wroc{\l}aw, Poland \\
29: marcin@ift.uni.wroc.pl}
30: \end{center}
31: \begin{center}
32: {Z. HASIEWICZ, T. NIKICIUK and C. J. WALCZYK}
33: \end{center}
34: \begin{center}
35: {Institute of Theoretical Physics, University of Bia{\l}ystok, ul.
36: Lipowa 41\\
37: 15-424 Bia{\l}ystok, Poland} \\
38: zhas@uwb.edu.pl, niki@alpha.uwb.edu.pl, c.walczyk@alpha.uwb.edu.pl
39: \end{center}
40: 
41: 
42: \begin{abstract}
43:     The classical and quantum model of high spin particles within the manifestly
44:     covariant framework. The internal (spin) degrees of freedom are described by two
45:     ${\cal C}(3,1)$ Clifford algebra spinors. The covariant quantization leads to PCT
46:     invariant  spectrum of particles with
47:     spin dependent masses. The quantum model contains elementary particles and the
48:     cluster states generating infinite degeneracy of the mass spectrum.
49: 
50: \end{abstract}
51: 
52: 
53: \section{Introduction}
54: There are many possible descriptions of the spinning particles with
55: arbitrary spin spectrum. It seems, that the most promising ones are
56: so called spinorial models \cite{BZ}-\cite{andrzej} with the
57: internal degrees of freedom (spin) realized by classical,
58: (anti-)commuting spinors, i.e the elements of irreducible or
59: reducible representation modules of corresponding Clifford algebras,
60: most commonly ${\cal C}(3,1)$. Despite of the fact that their
61: construction relies on direct generalization of superparticle models
62: they contain the particles of arbitrarily high spins. However, the
63: masses of these particles appear to be the same once the value of
64: the mass parameter of the model is fixed. There are also
65: constructions \cite{KLS} which lead to spin dependent masses of the
66: particles but, in contrast to the models mentioned above, with spin
67: variable introduced as a free parameter. \\
68: From this point of view the classical model defined in \cite{HDS}
69: and analyzed in details in \cite{DHW} seems to be particulary
70: interesting. It's construction relies on minimal coupling of the
71: particle trajectory with  vector current build up of single
72: ${\cal C}(3,1)$ real (Majorana) spinor.
73: Due to this
74: coupling the resulting first quantized model describes the infinite
75: families of spinning particles with spin-dependent masses. As the
76: analysis at the classical level strongly indicates this is this very
77: coupling, which corresponds to, and is responsible for spinning
78: particles "zitterbewegung" phenomenon \cite{schroedinger},\cite{BB},\cite{rivas},
79: which is most clearly and efficiently described in Clifford algebra language. \\
80: Besides of the "dynamical" spin-mass coupling, the spinning particle
81: model of \cite{HDS} and \cite{DHW} has an additional advantage: the
82: non degenerate mass levels.  Unfortunately, it has also important
83: drawback. The first quantization results in the Hilbert space of
84: states, which contains orphaned\footnote{The corresponding
85: antiparticles are missing.} particles and antiparticles, i.e. it
86: breaks CPT symmetry \cite{Jost}. Consequently, it does not allow one
87: to construct the corresponding local quantum
88: field theory by direct application of the second quantization procedures \cite{Jost}.\\
89: \\
90: The main task of this paper is to present the extended spinorial
91: model which cures the above weakness of the prototype proposed in
92: \cite{DHW}. The appropriate modification consists in adding an
93: additional spinorial degree of freedom with opposite spin-mass
94: coupling constant. The model obtained in this straightforward way
95: contains however the non-linear interaction of spinors. This
96: interaction is eliminated by the substraction from the Lagrange
97: function  the quartic term in spinorial variables (see (\ref{int})).
98: As a result, one gets after first quantization, the infinitly
99: degenerate but CPT invariant
100: spectrum of particles. \\
101: As in the case of \cite{DHW} the paper contains also covariant
102: formulation of the model based on the complex polarization of the
103: second class constraints and application of Gupta-Bleuler
104: quantization procedure \cite{GB}. The polarization leads to the
105: description of the internal degrees of freedom in terms of Weyl variables i.e.
106: in the language of even subalgebra ${\cal C}(3,1)_0 \approx {\cal C}(3,0)$ spinors.
107: This approach results in a natural
108: generalization of Dirac type equations \cite{Lop} for
109: the particles with "dynamical" spin-dependent mass.\\
110: \\
111: Since the above model has CPT invariant mass spectrum, it can serve
112: as a starting point for the construction of the local quantum field
113: theory \cite{Jost} for particles lying on Regge trajectory. It would
114: allow one, in particular, to investigate the different kinds of
115: field theoretical interactions of the spinning particles with
116: arbitrary spin and with spin dependent mass in four dimensional
117: space-time. \\
118: As far as we know, the quantum field theoretical description of the
119: particles lying on Regge trajectories was proposed many years ago in
120: the series of papers
121: \cite{field1},\cite{field2},\cite{field3},\cite{field4}. Apart of
122: that, such a description was also elaborated in the framework of
123: dual model \cite{Rebbi} and string theory
124: \cite{hikko},\cite{witten}, but only for the critical dimension of
125: space-time. In the case of dimension four, the direct constructions
126: of this kind encounter serious
127: difficulties and problems related to the presence of Liouville modes \cite{Polyakov}. \\
128: \\
129: The paper is organized as follows. \\
130: In the first chapter the structure of the classical model is
131: presented in Lagrangian and Hamiltonian form. The complex
132: parametrization of the spinor degrees of freedom in terms of Weyl
133: spinors is introduced in order to make the transparent link with the
134: commonly
135: used formalism. \\
136: The next chapter is devoted to the description of particle content
137: of the corresponding quantum theory. The spectrum is divided into
138: two series: the elementary trajectories and the sector describing
139: clusters. \\
140: Finally, the results are summarized and some open questions and
141: problems are raised.
142: 
143: %-------------section one
144: 
145: \section{The classical model}
146: The classical model considered in this paper is a natural
147: generalization and extension of the one presented in \cite{DHW}. The
148: particle content of the model of \cite{DHW} appeared to be not CPT
149: invariant:  the particles and antiparticles were orphaned. \\
150: From the form of the mass spectrum obtained in \cite{DHW}: $m_j =
151: \sqrt{h^2j^2+m_0^2} - hj$ for the particles of spin $j$ and $m_j =
152: \sqrt{h^2j^2+m_0^2} + hj$ for antiparticles, one may easily conclude
153: that in order to get rid of this undesired orphancy, it is enough to
154: add  an additional spinorial degree of freedom with opposite
155: spin-mass coupling. The classical model obtained in such a way was
156: already proposed in the Conclusions of \cite{DHW}. It is defined by
157: the following Lagrange function:
158: \begin{equation}
159: \label{action}
160:     {\cal L}_{0}=\frac{1}{2}e^{-1}\dot{x}^2-\frac{1}{2}e
161:     m_{0}^{2} + \frac{h}{2} \dot{x}\cdot j_{1} - \frac{h}{2} \dot{x}\cdot j_{2} +
162:     \bar{\eta}_{1}{\dot \eta}_{1}
163:     + \bar{\eta}_{2}{\dot \eta}_{2}\;,
164: \end{equation}
165: which is a straightforward extension of the one presented in
166: \cite{DHW}. In addition to the standard terms describing the scalar
167: particle of mass $m_0$ it contains "minimal" couplings of the
168: velocity with two spinor currents
169: \begin{equation}
170: \label{currents_two}
171: j^{\mu}_{I} =
172: \bar{\eta}_{I}\gamma^{\mu}{\eta}_{I}\;,\;I=1,2\;.
173: \end{equation}
174: The currents are
175: build out of two independent Majorana spinors
176: ${\eta}_{I}$. For the Majorana spinors to exist one should
177: assume the spacetime metric in the form $g = {\rm
178: diag}(-1,+1,+1,+1)$ i.e. the Clifford algebra generators present in
179: (\ref{currents_two}) should satisfy:
180: $$
181: \gamma^\mu \gamma^\nu + \gamma^\nu\gamma^\mu = 2 g^{\mu\nu}\,.
182: $$
183: The spinorial conjugation is defined in the standard way:
184: $\bar{\eta}_{I} = {\eta}_{I}^{{\rm T}}\gamma_0$ and  it defines the
185: ${\mathbf{Spin}} (3,1)$ invariant scalar product: $\bar{\eta}\eta '
186: = {\eta}_{I}^{{\rm T}}\gamma_0\eta ' \,$. The explicite form of this
187: product is not in fact needed. The only property which is used
188: further on is that it generates $\beta_-$ antiautomorphism of
189: ${\cal C}(3,1)$ Clifford algebra: $\beta_- (\gamma^\mu) = - \gamma^\mu\,$.\\
190: %
191: The classical model defined by (\ref{action}) appears to generate
192: non-linear classical equations of motion.\\
193:     Besides of the standard algebraic equation or $1$-bein variable
194:     $e$ and momentum conservation law:
195: $$
196: \frac{d}{d\tau}\, p^\mu = 0\;,\;\;\; {\rm where}\;\;\; p^\mu =
197:     \frac{1}{e}\dot{x}^{\mu} + \frac{h}{2} j^\mu_1 - \frac{h}{2}
198:     j^\mu_2\;,
199: $$
200: it contains non-linear equations of motion for spinor variables:
201:     \begin{equation}
202:     \label{spinoreqns}
203: \frac{d}{d\tau}\, \eta_{1/2} = \frac{eh}{2} \left( \mp p^\mu -
204: \frac{h}{2} j^\mu_{2/1}\right)\gamma_{\mu}\,\eta_{1/2}
205:     \end{equation}
206: The interactions of this type are absent in single spinor model of
207: \cite{DHW} as the spin vector currents built out of single ${\cal
208: C}(3,1)$ Clifford algebra spinors are light-like. This fact can be
209: easily proved by the use of Clifford algebra structural relations
210: without any use of explicit matrix representation.
211:     This non-trivial "interaction" of spinors follows from the fact
212:     that both spinors are coupled to single particle trajectory.\\
213:     For this reason the model defined by (\ref{action}) is difficult
214:     to treat technically at both: classical and quantum levels. \\
215:     The form of the equations of motion (\ref{spinoreqns}) strongly
216:     suggests that the interaction of spinors is of potential
217:     character. It is then sensible to ask for the corresponding
218:     linearized, free model. It appears that it can be obtained by
219:     adding to (\ref{action}) the term which describes the
220:     cross interaction of spinor currents:
221: \begin{equation}
222: \label{int}
223:     {\cal L}_{\rm{int}} = -\frac{eh}{4}^2j_{1}\cdot j_{2}\;.
224: \end{equation}
225: It is not difficult to check (using the property of spinor currents
226: $j_{I}^2=0$ being null and Schwartz inequality) that (\ref{int}) is
227: non-negative. This means that the interaction present in
228: (\ref{action}) and in (\ref{spinoreqns}) is governed by non-negative
229: potential of fourth order.
230:     As it appears the term (\ref{int}) cancels the non-linear
231: interactions at the Hamiltonian level and one is left with
232: relatively simple system of constraints of mixed type. On the other
233: hand its presence leads to essential modification of the proper time
234: parameter. As it can be calculated from the equations of motion for
235: $1$-bein $e$, the corresponding density gets rescaled:
236: \begin{equation}
237: \label{einbein}
238:     ds = \sqrt{-\dot{x}^2}d\tau \;\;\rightarrow\;\;
239:     ds' = \sqrt{1 + \frac{h^2j_{1}\cdot j_{2}}{2m_0^2}}
240:     \,\sqrt{-\dot{x}^2}d\tau\;.
241: \end{equation}
242: The questions related to the above change of the internal geometry
243: of the particle trajectory are by all means interesting, but they
244: will be not pursued here. The considerations of this paper will be
245: focused on the hamiltonian formulation and the corresponding first
246: quantized theory.\\
247: \\
248:     At the canonical level the model defined by the sum of (\ref{action}) and
249: (\ref{int}) describes the constrained system  with constraints of
250: mixed type. Their structure is analogous  to that
251: of \cite{DHW}. \\
252: There are two families of spinorial constraints:
253: \begin{equation}
254: \label{secondclasscon}
255:     G_{I}^{\alpha} = \pi_{I}^{\alpha} + \eta^{\alpha}_I\;\;\;;\;\;\alpha =
256:     1,...,4\;;\;I=1,2\;,
257: \end{equation}
258: relating the canonical momenta $\pi_{I}^{\alpha}$ of spinors with
259: spinors $\eta^{\alpha}_I$ themselves\footnote{One obviously assumes
260: the canonical Poisson brackets $\{ \pi_{I}^{\alpha},\eta^{\beta}_J
261: \}\ = \delta_{IJ}C^{{\alpha}{\beta}}$. The matrix
262: $C^{{\alpha}{\beta}}$ is the inverse of the one defined in the
263: following way: $\bar{\eta}_{I}{\eta}_{I}' = {\eta}^{{\rm
264:     T}}_{I}\gamma_{0}{\eta}_{I}' = {\eta}^{\alpha}_{I} C_{\alpha \beta}{\eta
265:     '}^{\beta}_{I}$.}. \\
266: These constraints reflect the property of the system defined by
267: (\ref{action}) and (\ref{int}) that the equations of motion for
268: spinors are of first order in the evolution parameter. They are
269: obviously of second class:
270: \begin{equation}
271: \label{secondclasscon1}
272:      \{G_{I}^{\alpha},G_{J}^{\beta}\} = 2\delta_{IJ}C^{\alpha \beta}\;.
273: \end{equation}
274: There is in addition the kinematic constraint related to the
275: reparametrization invariance of the corresponding action functional.
276: It is given by canonical Dirac hamiltonian:
277: \begin{equation}
278: \label{diracham}
279:     H_{\rm D} = \frac{1}{2}(p^2 + m_0^2) -
280:     \sum_{I=1}^{2}\frac{h}{2}(-1)^{I}p_{\mu}J^{\mu}_{I}\;,
281: \end{equation}
282: where $p_{\mu}$ are the momenta canonically conjugated to the
283: positions $x^{\mu}$. The currents $J_{I}^{\mu}$ are bilinear in
284: spinors and their conjugated momenta: $J_{I}^{\mu} =
285: \pi_{I}^{\alpha}(\gamma^{\mu})_{\alpha\beta}\eta_I^{\beta}$. The
286: 1-bein variable
287: was eliminated by putting $e=1$. \\
288: The constraints (\ref{secondclasscon}) together with
289: (\ref{diracham}) constitute the closed system of mixed type as one
290: has:
291: \begin{equation}
292: \label{poissonmixed}
293:     \{H_{\rm D},G_{I}^{\alpha} \} = (-1)^{I}\frac{h}{2}p_{\;\beta}^{\alpha}G_{I}^{\beta}\;.
294: \end{equation}
295: It should be stressed that the expression (\ref{diracham}) and the
296: relation (\ref{poissonmixed}) are so simple exactly due to the
297: substraction of the fourth order term (\ref{int}) $\sim j_{1}\cdot
298: j_{2}$ from
299: the Lagrange function (\ref{action}). \\
300: \\
301: As it was noted in \cite{DHW} the system of constraints
302: (\ref{secondclasscon}) and (\ref{poissonmixed}) can be replaced by
303: an equivalent system of first class by polarization of
304: (\ref{secondclasscon1})\cite{GB}. For massive momenta $(p^2<0)$ the
305: polarization of constraints is necessarily complex. \\
306: For this reason the system is most conveniently described in terms
307: of minimal building blocks of $\mathbf{Spin}(3,1) \approx \mathbf{SL}(2;\mathbb{C})$
308: representations i.e. in terms of complex Weyl spinors. \\
309: The real space of Majorana spinors $(\eta_I^{\alpha},
310: \pi_I^{\alpha})$ decomposes into, mutually complex
311: conjugated\footnote{According to common convention $z_I^{\bar{A}} =
312: \overline{({z}_I^A)}$.}, Weyl components $(z_I^A,\mathfrak{z}_I^A
313: )_{A=1,2}$ and $(z_{I}^{\bar{A}},\mathfrak{z}_{I}^{\bar{A}})
314: _{\bar{A} = 1,2}\;$. They span the eigensubspaces of $\gamma^5$
315: matrix corresponding to $\pm
316: i$ eigenvalues. \\
317: The Poisson brackets of the canonical Weyl variables are given as
318: follows:
319: \begin{equation}
320: \label{poissoncomplex}
321:     \{\mathfrak{z}_I^A,z_J^B\} =
322: \epsilon^{AB}\delta_{IJ}\;\;,\;\;\{\mathfrak{z}_I^{\bar{A}},z_J^{\bar{B}}\}
323: = \epsilon^{{\bar{A}}{\bar{B}}}\delta_{IJ}\;,
324: \end{equation}
325: where $\epsilon^{AB}$ $(\epsilon^{{\bar{A}}{\bar{B}}})$ are
326: $\mathbf{SL}(2;\mathbb{C})$ invariant tensors. \\
327: The second class constraints of (\ref{secondclasscon}) relate the
328: canonical Weyl coordinates:
329:     $G_I^A = \mathfrak{z}_I^{{A}} + z_I^A = 0$ and $ G_I^{\bar{A}} =
330: \mathfrak{z}_I^{\bar{A}} + z_I^{\bar{A}}=0$. Their polarized
331: counterparts can be expressed in the following way:
332: \begin{equation}
333: \label{GWp}
334:     G^{A}_{I(\pm)} = p^A_{\;\bar{B}}G_{I}^{\bar{B}} \pm im(p)
335: G_{I}^{A}\;\;,\;\;G^{\bar{A}}_{I(\pm)} =
336: p^{\bar{A}}_{\;{B}}G_{I}^{{B}} \pm im(p) G_{I}^{\bar{A}}\;,
337: \end{equation}
338: where $p^A_{\;\bar{B}}$ and $p^{\bar{A}}_{\;{B}}$ are (mutually
339: complex adjoint) matrix elements of the real operator
340: $p^{\mu}\gamma_{\mu}$ in the complex basis of Weyl spinors. Due to
341: Clifford algebra relations they do satisfy:
342: $p^A_{\;\bar{B}}p^{\bar{B}}_{\; {C}} = p^2\delta^A_C\,$ and
343: $p^{\bar{A}}_{\;{B}}p^{{B}}_{\; \bar{C}} =
344: p^2\delta^{\bar{A}}_{\bar{C}}\,$.
345: \\
346: The Hamiltonian constraint rewritten in terms of Weyl variables
347: takes the form\footnote{All indices are raised and lowered by
348: $\mathbf{SL}(2;\mathbb{C})$ invariant tensors $\epsilon^{AB}$
349: $(\epsilon^{{\bar{A}}{\bar{B}}})$ and their inverses $\epsilon_{AB}$
350: $(\epsilon_{{\bar{A}}{\bar{B}}})$ respectively.}:
351: \begin{equation}
352: \label{kinetic}
353:     H_{D} = \frac{1}{2}(p^2 + m_0^2) -
354: \sum_{I=1}^{2}\frac{h}{2}(-1)^I(\mathfrak{z}_I^Ap_{A\bar{B}}z_I^{\bar{B}}
355: + \mathfrak{z}_I^{\bar{A}}p_{\bar{A}B}z_I^B)\;.
356: \end{equation}
357: The Poisson algebra of the complex constraints can be easily
358: calculated. From (\ref{poissoncomplex}) and (\ref{GWp}) it follows
359: that:
360: \begin{equation}
361: \label{zero}
362:     \{G^{A}_{I(\pm)},G^{B}_{J(\pm)}\} = 0 \;,\;\;\{G^{A}_{I(+)},G^{{B}}_{J(-)}\}=4\delta
363: _{IJ}m^2(p)\epsilon^{A {B}}\;.
364: \end{equation}
365: The functions (\ref{GWp}) are, under the Poisson bracket, the
366: mass-weighted eigenfunctions of (\ref{kinetic}):
367: \begin{equation}
368: \label{masseigen} \{H_D,G^{A}_{I(\pm)}\} = \pm \frac{ih}{2}(-1)^I
369: m(p)\,G^{A}_{I(\pm)}\;.
370: \end{equation}
371: As one should expect, the constraints $G^{A}_{I(\pm)}$ and
372: $G^{\bar{A}}_{I(\pm)}$ are not independent:
373: \begin{equation}
374: \label{gg}
375:     {G_{I(\pm)}^{\bar{A}}} =
376:     \mp\frac{i}{m(p)}p^{\bar{A}}_{\,{B}}G^{{B}}_{I(\pm)}\;.
377: \end{equation}
378: Hence, according to (\ref{zero}) and (\ref{masseigen}), the systems
379: $(H_D,G^{A}_{I(\pm)})$ constitute polarized Poisson algebras of
380: first class.  From $\overline{({G^{{A}}_{I(\pm)}})} =
381: G_{I(\mp)}^{\bar{A}}$ and (\ref{gg}) it follows that they are
382: mutually complex conjugated.
383: 
384: \section{The quantum model}
385: 
386: \subsection{The space of states}
387: The Hilbert space of states corresponding to the classical system
388: under consideration, consists in square integrable functions of
389: space-time momentum $(p^\mu)$ and Weyl spinor coordinates
390: $(z_{I}^A,z_{I}^{\bar{A}})$. In this representation space the
391: canonically conjugated variables are realized as the following
392: operators: $\;{\hat x}^{\mu} = -i{\partial}/{\partial p_{\mu}}$ and
393: $\; \hat {\mathfrak{z}}_{I}^A = i\epsilon^{AB}{\partial}/{\partial
394: z_{I}^B}$,$\;\; \hat {\mathfrak{z}}_{I}^{\bar{A}} =
395: i\epsilon^{\bar{A}\bar{B}}{\partial}/{\partial z_{I}^{\bar{B}}}\;$.
396: \\
397: It should be stressed that according to the conditions imposed
398: already at the classical level, the momenta $(p^\mu)$ are restricted
399: to the open domain $p^2<0$ supporting real massive particles. Since
400: this domain is the union of two disjoint cone interiors $p^0>0$ and
401: $p^0<0$, the Hilbert space of states decomposes into the
402: corresponding direct sum:
403: \begin{equation}
404: \label{sum}
405:     H = H^{\uparrow} \oplus H^{\downarrow}\;.
406: \end{equation}
407: The above direct components contain the functions of the
408: future-pointed
409: and past-pointed momenta respectively. \\
410: In the representation space above the operators corresponding to the
411: spinorial constraints (\ref{GWp}) are expressed by:
412: \begin{equation}
413: \label{qconstraints}
414:     G^{A}_{I(\pm)}= ip^{A {\bar{B}}}\frac{\partial}{\partial
415:     z_{I}^{\bar{B}}}
416:     \mp m(p)\epsilon^{AB}\frac{\partial}{\partial z_{I}^B} +
417:     p^A_{\;\;\bar{B}}z_{I}^{\bar{B}} \pm im(p) z_{I}^{A}\;,
418: \end{equation}
419: while the Dirac Hamiltonian (\ref{kinetic}) takes the following
420: operator form:
421: \begin{equation}
422: \label{qham}
423:      H_{D} = \frac{1}{2} \left (p^2 + m_0^2 \right) -
424:      \frac{ih}{2}\sum_{I=1}^{2} (-1)^I \left ({z}_{I}^{\bar{B}}p_{\bar{B}}^{\;\;\;A}
425:      \frac{\partial}{\partial
426:     z_{I}^{A}}
427:     + {z}_{I}^{{B}}p_{B}^{\;\;\;\bar{A}}\frac{\partial}{\partial
428:     z_{I}^{\bar{A}}} \right )\;.
429: \end{equation}
430: In order to analyze the spectrum of the model it is useful to
431: distinguish the differential part of (\ref{qham}):
432: \begin{equation}
433: \label{diff} S = -\frac{ih}{2}\sum_{I=1}^{2} (-1)^I \left
434: ({z}_{I}^{\bar{B}}p_{\bar{B}}^{\;\;\;A}\frac{\partial}{\partial
435: z_{I}^{A}}
436:     + {z}_{I}^{{B}}p_{B}^{\;\;\;\bar{A}}\frac{\partial}{\partial z_{I}^{\bar{A}}} \right)\;.
437: \end{equation}
438: As it will appear clear the operator (\ref{diff}) is responsible for
439: spin-mass correlation in the spectrum of the model. For this reason
440: it will be called
441: the spin-mass coupling operator. \\
442: \\
443: The space of physical states of the model is defined as a kernel of
444: either $G^{A}_{I(+)}$ or $G^{A}_{I(-)}$ constraints\cite{GB}
445: (off-shell
446: physical states) and $H_{D}$ kinematical constraint. \\
447: According to \cite{Lop} and the detailed analysis of the paper
448: \cite{DHW} the physical states should be looked for within the
449: vectors of the form:
450: \begin{equation}
451: \label{states}
452:     \Psi_{(\pm)}(p,z_{I},\bar{z}_{I}) = W(z_{I},\bar{z}_{I})\,\Omega_{(\pm)} (p) \;,
453: \end{equation}
454: where $W(z_{I},\bar{z}_{I})$ are the polynomials in the Weyl
455: variables with momentum dependent coefficients.
456: \\
457: The space (\ref{sum}) of the functions (\ref{states}) splits in a
458: natural way into the sum of the direct integrals:
459: \begin{equation}
460: \label{directint} {H}^{\uparrow / \downarrow} =
461: \int\limits_{V_{(\pm)}} d^4p~ {H}^{\uparrow / \downarrow}(p)\;,
462: \end{equation}
463: of the Hilbert spaces ${H}^{\uparrow}(p)$  and ${H}^{\downarrow}(p)$
464: localized at the momentum $p$, contained in $V_{(\pm)}$ - the
465: future-pointed and the past-pointed cone
466: interiors respectively. \\
467: The states $\Omega_{(\pm)} (p)$ (the spin vacua) are the Gaussian
468: solutions of spinorial constraints $G^{A}_{I(\pm)}\Omega_{(\pm)} (p)
469: =0$. It is easy to check that (up to multiplicative factors) they
470: are given by:
471: \begin{equation}
472: \label{vacuum}
473:     \Omega_{(\pm)} (p) = \exp \left (\pm \sum_{I=1}^{2}
474:     \frac{z_{I}^{\bar{A}}p_{\bar{A}B}z_{I}^B}{m(p)} \right )\;.
475: \end{equation}
476: According to the adopted conventions the momentum matrix
477: $(p_{\bar{A}B})$ is negatively defined in the future-pointed
478: $(p^0>0)$ light-cone interior and it is positive in past-pointed
479: $(p^0<0)$ domain. For this reason one should take $\Omega_{(+)}(p)$
480: spin vacuum in $H^{\uparrow}(p)$ and $\Omega_{(-)}(p)$ state in
481: $H^{\downarrow}(p)$. This choice guarantees that the states
482: (\ref{states}) belong to localized Hilbert space of the quantum
483: model, i.e. they are square integrable with respect to Weyl
484: variables. It also means, that any state from $H^{\uparrow /
485: \downarrow}(p)$ can be represented as a superposition of the
486: following vectors:
487: \begin{eqnarray}
488: \label{degspinstate}
489:     \Psi_{(j_1,j_2)}^{\uparrow / \downarrow}(p,z_{I},\bar{z}_{I})
490:     &=&
491:     \sum\limits_{n_{1}=0}^{2j_1}\sum\limits_{n_{2}=0}^{2j_2} \Psi_{A_{1}^1\ldots
492: A_{2j_1-n_{1}}^1\bar{B}_{1}^1\ldots\bar{B}_{n_{1}}^1;\,A_{1}^2\ldots
493: A_{2j_2-n_{2}}^2\bar{B}_{1}^2\ldots\bar{B}_{n_{2}}^2} (p) ~\cdot \cr
494: &~& \cr &\cdot& z_{1}^{A_{1}^1}\cdots z_{1}^{{A}^{1}_{2j_1-n_{1}}}
495: z_{1}^{\bar{B}_{1}^1}\cdots z_{1}^{\bar{B}_{n_{1}}^1}
496: z_{2}^{A_{1}^2}\cdots z_{2}^{A_{2j_2-n_{2}}^2}
497: z_{2}^{\bar{B}_{1}^2}\cdots z_{2}^{\bar{B}_{n_{2}}^2 } ~\cdot \cr
498: &\cdot&\Omega_{(\pm)} (p) \;,
499: \end{eqnarray}
500: of fixed integral degrees $(2j_1,2j_2)$ in $(z_1^{A},z_1^{\bar{A}})$
501: and
502: $(z_2^{A},z_2^{\bar{A}})$ complex variables respectively. \\
503: In the light of the above considerations one should impose
504: $G^{A}_{I(+)}$ constraints in $H^{\uparrow}$ and $G^{A}_{I(-)}$ in
505: $H^{\downarrow}$. Hence, the space $\widehat{H}$ of physical
506: off-shell states is necessarily of the form of the following direct
507: sum:
508: \begin{equation}
509: \label{sum1}
510:   \widehat{H} = \widehat{H}_{(+)}^{\uparrow} \oplus \widehat{H}_{(-)}^{\downarrow}\;,
511: \end{equation}
512: where $\widehat{H}^{\uparrow / \downarrow}_{(\pm)}$ are the kernels
513: of $G^{A}_{I(\pm)}$ operators correspondingly. \\
514: In order to find these kernels one may follow the direct method
515: presented in the paper \cite{DHW}. They are however much more
516: transparently described in terms of spectrum generating algebra.
517: This algebra contains the constraints operators (\ref{qconstraints})
518: and in addition the following ones:
519: \begin{equation}
520: \label{dconstraints}
521:     D^{A}_{I(\pm)}= ip^{A {\bar{B}}}\frac{\partial}{\partial
522:     z_{I}^{\bar{B}}}
523:     \pm m(p)\epsilon^{AB}\frac{\partial}{\partial z_{I}^B} -
524:     p^A_{\;\;\bar{B}}z_{I}^{\bar{B}} \pm im(p) z_{I}^{A}\;.
525: \end{equation}
526: The above operators commute with these of (\ref{qconstraints}). The
527: only non-zero commutators are given by:
528: \begin{equation}
529: \label{com} [\;G^{A}_{I(+)},G^{\bar{B}}_{J(-)}\;]=4\delta
530: _{IJ}m(p)p^{A \bar{B}}\;\;,\;\;
531: [\;D^{A}_{I(+)},D^{\bar{B}}_{J(-)}\;]=4\delta _{IJ}m(p)p^{A
532: \bar{B}}.
533: \end{equation}
534: Since one has $(G^{A}_{I(\pm)})^{*}=G^{\bar{A}}_{I(\mp)}$ and
535: $(D^{A}_{I(\pm)})^{*}=D^{\bar{A}}_{I(\mp)}$, the operators
536: (\ref{qconstraints}) and (\ref{dconstraints}) may serve as creation
537: and anihilation operators
538: in the space (\ref{directint}). \\
539: From the form of the generators of $\mathbf{SL}(2;\mathbb{C})$
540: transformations\footnote{They can be obtained as operator
541: counterparts of the corresponding Noether conserved quantities
542: calculated from (\ref{action}) and (\ref{int}).}:
543: \begin{equation}
544: \label{lorentz}
545:     L^{\mu\nu} = i\left(p^\mu \frac{\partial}{\partial p^\nu} - p^\nu \frac{\partial}{\partial p^\mu}
546:     \right)+ \frac{i}{2}\sum_{I=1}^{2}\left( z_{I}^A \sigma ^{(\mu\nu) B}_{\;A} \frac{\partial}{\partial z_{I}^B} +
547:     z_{I}^{\bar{A}} \sigma ^{(\mu\nu) {\bar{B}}}_{\;{\bar{A}}} \frac{\partial}{\partial
548:     z_{I}^{\bar{B}}}\right)\;,
549: \end{equation}
550: it follows that the constraints $G^{A}_{I(\pm)}$ and the operators
551: $D^{A}_{I(\pm)}$ are of spinorial character and they carry spin
552: $\frac{1}{2}$:
553: \begin{equation}
554: \label{spinorial1}
555: [\;L^{\mu\nu},G^{A}_{I(\pm)}\;]=\frac{i}{2}G^{B}_{I(\pm)}\sigma
556: ^{(\mu\nu) {{A}}}_{\;{{B}}} \;\;,\;\;
557: [\;L^{\mu\nu},D^{A}_{I(\pm)}\;]=\frac{i}{2}D^{B}_{I(\pm)}\sigma
558: ^{(\mu\nu) {{A}}}_{\;{{B}}}\;,
559: \end{equation}
560: while the kinematic constraint:
561: \begin{equation}
562: \label{hamiltonian}
563:     H_{D} =\frac{1}{2} \left (p^2 + m_0^2 \right) -
564:      \frac{h}{8m^2(p)}\sum_{I=1}^{2} (-1)^I \left (G^{A}_{I(+)}p_{A\bar{B}}
565:      G^{\bar{B}}_{I(-)} -D^{A}_{I(+)}p_{A\bar{B}}
566:      D^{\bar{B}}_{I(-)}
567:     \right )\;,
568: \end{equation}
569:  is scalar. \\
570: One may directly check that:
571: \begin{equation}
572: \label{vac} G^{A}_{I(\pm)}\Omega_{(\pm)} (p) =
573: D^{A}_{I(\pm)}\Omega_{(\pm)}(p) = 0\;,
574: \end{equation}
575: and consequently, any state (\ref{degspinstate}) from the space
576: ${H}^{\uparrow / \downarrow}(p)$ can be equivalently and more
577: transparently expressed in the Fock-type representation:
578: \begin{equation*}
579:     \Psi_{(j_1,j_2)}^{\uparrow / \downarrow}(p)
580:     =
581:     \sum\limits_{n_{1}=0}^{2j_1}\sum\limits_{n_{2}=0}^{2j_2}
582:     \Phi(p)_{A_{1}..
583: A_{2j_1-n_1}B_1 .. B_{n_1}C_1 .. C_{2j_2-n_2}D_{1}.. D_{n_2} (p)}
584: ~\cdot \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\nonumber
585: \end{equation*}
586: \begin{equation}
587: \label{opstate} \;\;\;\;\;\;\;\;\;\cdot D_{1(\mp)}^{A_{1}}..
588: D_{1(\mp)}^{{A}_{2j_1-n_{1}}} G_{1(\mp)}^{B_{1}}..
589: G_{1(\mp)}^{{B}_{n_{1}}} D_{2(\mp)}^{C_{1}}..
590: D_{2(\mp)}^{C_{2j_2-n_{2}}}G_{2(\mp)}^{{D}_{1}}..
591: G_{2(\mp)}^{{D}_{n_{2}} }\Omega_{(\pm)}(p) \;.
592: \end{equation}
593: As one has $[G^{A}_{I(+)},G^{{B}}_{J(-)}]=4i\delta
594: _{IJ}m^2(p)\epsilon^{A {B}}$ and
595: $[D^{A}_{I(+)},D^{{B}}_{J(-)}]=4i\delta _{IJ}m^2(p)\epsilon^{A
596: {B}}$, the constraints are easily solvable now. They simply mean
597: that $G_{I(\mp)}^{A}$ excitations should be absent in
598: (\ref{opstate}) and for the solutions one gets:
599: \begin{eqnarray}
600: \label{opstate1}
601:     \Psi_{(j_1,j_2)}^{\uparrow / \downarrow}(p)
602:     =
603:     \Phi(p)_{{A_{1}}..
604: A_{2j_1}C_{1}.. C_{2j_2}}D_{1(\mp)}^{A_{1}}..
605:  D_{1(\mp)}^{{A}_{2j_1}}
606: D_{2(\mp)}^{C_{1}}..D_{2(\mp)}^{C_{2j_2} } \Omega_{(\pm)} (p) \;.
607: \end{eqnarray}
608: It is worth to mention that the constraints $G_{I(\pm)}^{A}$ imposed
609: on the states in the form of (\ref{degspinstate}) become the
610: Dirac-type equations for the coefficients of the polynomials. In the
611: case of ${\widehat H}^{\uparrow}_{(+)}(p)$ they take the following
612: form:
613: \begin{eqnarray}
614: \label{equation} &-& (n_I+1)
615: p_{A_{2j_{I}-n_{I}}^{I}}^{\bar{B}_{n_{I}+1}^I}\Psi_{A_{1}^{1}\ldots
616:     A_{2j_{1}-n_{1}-
617:     \delta_{I,1}}^{1}\bar{B}_{1}^{1}\ldots\bar{B}_{n_{1}+\delta_{I,1}}^{1};\,A_{1}^{2}\ldots
618:     A_{2j_{2}-n_{2}- \delta_{I,2}}^{2}\bar{B}_{1}^{2}\ldots\bar{B}_{n_{2}+\delta_{I,2}}^{2}}(p) ~=\cr
619:     &~& \cr
620:     &=&
621:     i m(p)(2j_{I}-n_{I})\Psi_{A_{1}^{1}\ldots
622:     A_{2j_{1}-n_{1}}^{1}\bar{B}_{1}^{1}\ldots\bar{B}_{n_{1}}^{1};\,A_{1}^{2}\ldots
623:     A_{2j_{2}-n_{2}}^{2}\bar{B}_{1}^{2}\ldots\bar{B}_{n_{2}}^{2}}
624:     (p)\;.
625: \end{eqnarray}
626: where $I=1,2$ and $n_I = 0,...,2j_I-1$. \\
627: In the complementary space ${\widehat H}^{\downarrow}_{(-)}(p)$ of
628: anti-particles the above equations are completely analogous. \\
629: \\
630:     In order to find the spectrum of the model one should impose the
631: kinematical constraint $H_D$ (\ref{hamiltonian}) on the solutions of
632: (\ref{opstate1}). This constraint correlates the masses of the
633: particles with their spins. The detailed analysis of the spectrum
634: will be given in the next subsection.
635: \subsection{Spin and mass spectrum}
636: 
637: In order to find the spectrum explicitly it is convenient to use the
638: following decomposition of the off-shell spaces:
639: \begin{equation}
640: \label{decomposition} \widehat{H}_{(+)}^{\uparrow}(p) =
641: \bigoplus_{(j_1,j_2)} \widehat{H}_{(+)}^{\uparrow (j_1,j_2)}(p) \;,
642: \end{equation}
643: where $\widehat{H}_{(+)}^{\uparrow (j_1,j_2)}(p)$ contains the
644: polynomials of fixed bidegree $(2j_1,2j_2)$ of the form of
645: (\ref{opstate1}). From (\ref{hamiltonian}) and commutations
646: relations (\ref{com}) it follows that the spin-mass coupling
647: operator $S$ of (\ref{diff}) is diagonal and acts in the following
648: way:
649: \begin{equation}
650: \label{saction} S \widehat{H}_{(+)}^{\uparrow (j_1,j_2)}(p) =
651: {h}m(p) (j_1 - j_2) \widehat{H}_{(+)}^{\uparrow (j_1,j_2)}(p)\;.
652: \end{equation}
653: Hence, the kinetic constraint imposes the following mass-shell
654: condition on the particle momenta:
655: \begin{equation}
656: \label{momentum}
657:     \frac{1}{2}(p^2 + m_0^2) + {h}m(p)(j_1 - j_2) = 0\;,
658: \end{equation}
659: with the unique solution leading to normalizable states\footnote{The
660: second, negative root of (\ref{m+}) has to be rejected because it
661: makes the exponent (\ref{vacuum}) growing very fast at infinity.},
662: given by:
663: \begin{equation}
664: \label{m+} m_{(j_1,j_2)}^{\uparrow} = \sqrt{h^2(j_1-j_2)^2+m_0^2} +
665: h(j_1 - j_2) \;\;;\;\; j_1,~j_2\geq 0\;.
666: \end{equation}
667: One may analogously find the masses of the states supported by the
668: past-pointed cone interior:
669: \begin{equation}
670: \label{m-} m_{(j_1,j_2)}^{\downarrow} = \sqrt{h^2(j_1-j_2)^2+m_0^2}
671: - h(j_1 - j_2) \;\;;\;\; j_1,~j_2\geq 0\;.
672: \end{equation}
673: According to (\ref{m+}) and (\ref{m-}), for on-shell physical
674: states, the direct integrals corresponding to (\ref{directint}) are
675: restricted to the respective mass-shells. The spaces of physical
676: on-shell states will be denoted by $\widetilde{H}_{(+)}^{\uparrow}$
677: and
678: $\widetilde{H}_{(-)}^{\downarrow}$ respectively. \\
679: %
680: The comparison of the formulae (\ref{m+}) and (\ref{m-}) clearly
681: indicates that the spectrum of the model is CPT
682: invariant\cite{Jost}, i.e. it is symmetric with respect to
683: particle-antiparticle exchange. This property was missing in the
684: simple prototype model
685: presented in \cite{DHW}. \\
686: \\
687: In order to perform a more detailed analysis of the spectrum it is
688: useful to distinguish two families of subspaces of the on-shell
689: space $\widetilde{H}_{(+)}^{\uparrow}(p)$, namely:
690: \begin{equation}
691: \label{distin} \widetilde{H}_{(+)}^{\uparrow (j_1,0)}(p)\;\;\;\;
692: {\rm and}\;\;\;\; \widetilde{H}_{(+)}^{\uparrow (0,j_2)}(p) \;,
693: \end{equation}
694: where the numbers $j_1$ and $j_2$ have direct physical
695: interpretation as particle spins\cite{Lop}. As it is clearly visible
696: from (\ref{opstate1}) these spaces describe the states of single
697: particles with spins $j_1$ and $j_2$, and with the corresponding
698: masses:
699: \begin{equation}
700: \label{m+m-} m_{j_1}^{\uparrow} = \sqrt{h^2j_1^2+m_0^2} +
701: hj_1\;\;,\;\;m_{j_2}^{\uparrow} = \sqrt{h^2j_2^2+m_0^2} - hj_2\;.
702: \end{equation}
703: The antiparticles of $\widetilde{H}_{(+)}^{\uparrow (j_1,0)}(p)$ and
704: $\widetilde{H}_{(+)}^{\uparrow (0,j_2)}(p)$ are represented in the
705: spaces:
706: \begin{equation}
707: \label{distin1} \widetilde{H}_{(-)}^{\downarrow (0,j_1)}(p)\;\;\;\;
708: {\rm and}\;\;\;\; \widetilde{H}_{(-)}^{\downarrow (j_2,0)}(p)\;,
709: \end{equation}
710: respectively. \\
711: According to (\ref{directint}), as it was already mentioned, the
712: localized spaces (\ref{distin}) and (\ref{distin1}) fit together
713: into direct integrals over the corresponding mass-shells
714: ${S_{+}(j_I)}$:
715: \begin{equation}
716: \label{h1h2} \widetilde{H}_{(+)}^{\uparrow (j_1,0)} =
717: \int\limits_{S_{+}(j_1)} d\mu_{j_1}(p) \widetilde{H}_{(+)}^{\uparrow
718: (j_1,0)}(p)\;\;\;\; {\rm and}\;\;\;\; \widetilde{H}_{(+)}^{\uparrow
719: (0,j_2)} = \int\limits_{S_{+}(j_2)} d\mu_{j_2}(p)
720: \widetilde{H}_{(+)}^{\uparrow (0,j_2)}(p)\;,
721: \end{equation}
722: with respect to Lorentz invariant measures $d\mu_{j_I}$.  \\
723: Together with their antiparticles partners they describe two
724: particle-antiparticle pairs located at two diverging Regge
725: trajectories. Hence, the corresponding mass-levels are
726: non-degenerate as in the case of the prototype model of \cite{DHW}.
727: For this reason these trajectories will be called elementary. They
728: are illustrated at Fig.1A\footnote{For $p^0<0$ one obtains exactly
729: the same trajectories of antiparticles.}. \\
730: \\
731: Besides of the particle pairs located at the elementary
732: trajectories, there is a family of Hilbert spaces containing the
733: polynomials non-zero degree in both, $(z_1^{A},z_1^{\bar{A}})$ and
734: $(z_2^{A},z_2^{\bar{A}})$ Weyl variables, i.e. of the spaces:
735: \begin{equation}
736: \label{composit} \widetilde{H}_{(+)}^{\uparrow (j_1,j_2)}(p)\;\;\;\;
737: {\rm where}\;\;\;\;j_{I} > 0\;\;;\; I=1,2\;.
738: \end{equation}
739: The momenta of the states from (\ref{composit}) are obviously
740: constrained by mass-shell condition (\ref{m+}). The corresponding
741: antiparticles are contained in $\widetilde{H}_{(-)}^{\downarrow
742: (j_2,j_1)}(p)$ and according to (\ref{m-})
743: they have exactly this same masses. \\
744: In contrast to $\widetilde{H}_{(+)}^{\uparrow (j_1,0)}$ and
745: $\widetilde{H}_{(+)}^{\uparrow (0,j_2)}$ the spaces (\ref{composit})
746: are reducible with respect to $\mathbf{SL}(2;\mathbb{C})$ group
747: representation, i.e. they decay into direct sums of the subspaces
748: containing the states with fixed spins, i.e. those of elementary
749: particles. This property is clearly visible from the form of the
750: solution of $G^{A}_{I(+)}$ constraints (\ref{qconstraints}) which is
751: explicitly given in (\ref{opstate1}). For this reason it is natural
752: to call the states from (\ref{composit})
753: the clusters. \\
754: The spin content of the spaces $\widetilde{H}_{(+)}^{\uparrow
755: (j_1,j_2)}(p)$ at fixed momentum can be obtained by the standard
756: simple rule of the decomposition of $\mathbf{SU}(2)$ (the little
757: group of massive momentum $p$) representations\cite{JM}:
758: \begin{equation}
759: \label{cg} \widetilde{H}_{(+)}^{\uparrow (j_1,j_2)}(p) =
760: \bigoplus_{j=|j_1-j_2|}^{j_1+j_2}\widetilde{H}_{(+)}^{\uparrow
761: j}(p)\;,
762: \end{equation}
763: where $\widetilde{H}_{(+)}^{\uparrow j}(p)$ denote the irreducible
764: subspace of spin $j$. \\
765: From (\ref{cg}) it is clear that keeping the mass of the particle
766: fixed (fixed value of the difference $j_1-j_2$) one may obtain
767: arbitrarily high spin states. Hence, the states from
768: $\widetilde{H}_{(+)}^{\uparrow (j_1,j_2)}(p)$ generate infinite spin
769: degeneracy of the mass levels. The corresponding cluster
770: trajectories are illustrated on Fig.1B. From this diagram it is
771: clear that the cluster states contribute to elementary trajectories
772: and in addition, they generate their own ones.
773: %
774: \begin{figure}[h]
775: \begin{center}
776: \includegraphics[width=\textwidth, height=12cm]{wykres_caly.eps}
777: \end{center}
778: \caption{{\bf A} - The elementary trajectories for particles (in
779: $m_0$ units). {\bf B} - The degeneracy of the spectrum of clusters
780: (in $m_0$ units and with $j_I$ ranging from $0$ to $11/2$).}
781: \label{fig.1}
782: \end{figure}\\
783: 
784: \section*{Conclusions and outlook}
785: 
786: As it was shown the model considered in this paper has CPT invariant
787: mass spectrum. It consists of elementary (non-degenerate) Regge
788: trajectories and highly degenerate sea of cluster particles.
789: \\
790: According to the considerations of second chapter, in the case of
791: elementary trajectories, the solution (\ref{opstate1}) describes the
792: wave function of the one elementary particle with fixed spin (in the
793: sense of $\mathbf{SL}(2;\mathbb{C})$ representations). The particle
794: is
795: accompanied by the corresponding antiparticle. \\
796: In the case of cluster (\ref{cg}) the wave functions solving the
797: constraints equations (\ref{qconstraints}), (\ref{qham}) do not
798: represent the elementary objects, but the families of particles with
799: spins ranging from $|j_1-j_2|$ to $j_1+j_2$. The solution
800: (\ref{opstate1}) together with mass-shell condition
801: describe an ensemble of particles. \\
802: These objects make an impression as being untypical, however the
803: idea of cluster fields (in that case with respect to mass not spin)
804: was proposed and developed almost 50 years ago in the series of
805: papers \cite{clusters}. \\
806: \\
807: The  analysis at the classical level indicates that the infinite
808: degeneracy of the mass levels might appear as an effect of the
809: substraction of the interaction term (\ref{int}) from
810: (\ref{action}), which makes the classical dynamics of the
811: corresponding spinors linear, and almost completely independent.
812: This substraction results in the independence of the polarized
813: constraints (\ref{GWp}) for both spinors and simple form of Dirac
814: Hamiltonian. For this very reason, the corresponding wave functions
815: contain products of independent polynomials in both independent Weyl
816: spinors, which in turn amounts to the decomposition
817: given in (\ref{cg}). \\
818: The detailed analysis of the model described by (\ref{action}), i.e.
819: containing the non-linear interaction (\ref{int}) is
820: already in progress, and will be presented in the future paper.\\
821: 
822: \section*{ Acknowledgements}
823: One of the authors (M.D.) would like to thank the friends and
824: colleagues from the Institute of Theoretical Physics in Bia{\l}ystok
825: for warm hospitality. \\
826: The authors would like to thank also prof. J.Lukierski, prof. Z.Haba and  dr. A.B{\l}aut
827:  for interesting discussions. \\
828: Special thanks (from Z.H.) are due to Bo\.{z}ena Gonciarz for her
829: inspiration. \\
830: This work is partially supported by KBN grant 1P03B01828.
831: 
832: \begin{thebibliography}{99}
833: \bibitem{BZ} A.O.Barut, N.Zoughi, Phys. Rev. Lett. 52 (1984) 2009
834: \bibitem{GT} V.D.Gershun, V.I.Tkach, JETP Lett. (1979) 320
835: \bibitem{HPPT} P.S.Howe, S.Penati, M.Pernici, P.Townsend, Phys. Lett.
836: B215 (1988) 255
837: \bibitem{LF} S.Fedoruk, J.Lukierski, Phys. Lett. B632 (2006)
838: 371-378
839: %
840: \bibitem{LF1} A.Bette, J.de Azcarraga, J.Lukierski, C.Miquel-Espanya,
841: Phys. Lett. B595 (2004)
842: \bibitem{andrzej}A.Frydryszak, {\it Lagrangian models of particles
843: with spin}, Published in {\it From field theory to quantum groups},
844: Singapore World Scientific Publishing (1996)
845: \bibitem{KLS} S.M.Kuzenko, S.L.Lyakhovich, A.Yu.Segal,
846: Int. J. Mod. Phys. A10 (1995) 1529
847: \bibitem{HDS} Z.Hasiewicz, F.Defever, P.Siemion,
848: Int. J. Mod. Phys. A 7 (1992) 3979-3996
849: \bibitem{DHW} M.Daszkiewicz, Z.Hasiewicz, C.Walczyk, {\it High spin particles with spin-mass
850: coupling}, Advances in Applied Clifford Algebras (in press)
851: \bibitem{Jost}R.F.Streater, A.S.Wightman {\it PCT Spin - Statistic
852: and all that} Benjamin, New York 1964 \\
853: R.Jost {\it The General Theory of Quantized Fields} AMS, Providence,
854: Rhode Island 1965
855: \bibitem{schroedinger} E.Schroedinger, Sitzugsh. Preuss. Akad. Wiss.
856: Phys.-Math. Kl. 24 (1930) 418; 3 (1931) 1
857: \bibitem{BB} A.O.Barut, A.J.Bracken, Phys. Rev. D23 (1981) 2454
858: \bibitem{rivas} M.Rivas, {\it Classical elementary particles, spin, zitterbewegung and all that} -
859: physics/0312107
860: \bibitem{GB} S.Gupta, Proc. Roy. Soc. A63 (1950) 681 \\
861: K.Bleuler, Helv. Phys. Acta 23 (1950) 567
862: \bibitem{Lop}
863: Jan {\L}opusza\'{n}ski {\it Spinor Calculus} PWN Warsaw 1985 (in Polish) \\
864: A.O.Barut, R.R\c{a}czka {\it Theory of Group Representations and
865: Applications} PWN Warsaw 1977
866: \bibitem{JM} Jan Mozrzymas {\it Application of Group Theory in
867: Physics} PWN Warsaw 1976 (in Polish) \\
868: M.Hamermesh {\it Group Theory} PWN Warsaw 1968
869: \bibitem{field1} A.R.Swift, Phys. Rev. 176 (1968) 1848-1855
870: \bibitem{field2} W.J.Zakrzewski, Nuovo Cim. A60 (1969) 263-290
871: \bibitem{field3} A.R.Swift, R.W.Tucker, Nuovo Cim. A67 (1970) 345-355
872: \bibitem{field4} A.R.Swift, R.W.Tucker, Phys. Rev. D1 (1970) 2894-2900
873: \bibitem{Rebbi} C.Rebbi,  Phys. Rept. 12 (1974) 1-73
874: \bibitem{hikko} H.Hata, K.Itoh, T.Kugo, H.Kunitomo, K.Ogawa, Phys. Rev.
875: D34 (1986) 2360
876: \bibitem{witten} E.Witten, Nucl. Phys. B268 (1986) 253
877: \bibitem{Polyakov} A.M.Polyakov, Phys. Lett. B103 (1981) 207
878: \bibitem{clusters} O.W.Greenberg, Ann. of Phys. 16 (1961) 158 \\
879: A.L.Licht, Ann. of Phys. 34 (1965) 161 \\
880: L.Turko, Nucl. Phys. B114 (1976) 535-545
881: 
882: \end{thebibliography}
883: \end{document}