1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amsthm,amscd,amssymb}
3: \usepackage{latexsym}
4: %\usepackage{showkeys}
5: \theoremstyle{plain}
6: \usepackage{epsf}
7: \usepackage{epsfig}
8: \arraycolsep3.0pt
9: %definition for column separation in some matrices
10: \newlength{\mycolspace}
11: \setlength{\mycolspace}{.075in}
12: \newcommand{\mcs}{\hspace{\mycolspace}}
13: \newtheorem{theorem}{Theorem}
14: \newtheorem{lemma}{Lemma}
15: \newtheorem{corollary}{Corollary}
16: \newtheorem{proposition}{Proposition}
17: \newtheorem{conjecture}{Conjecture}
18:
19: \theoremstyle{definition}
20: \newtheorem{definition}{Definition}
21: \newtheorem{condition}{Condition}
22: \newtheorem{problem}{Problem}
23: \newtheorem{example}{Example}
24:
25: \theoremstyle{remark}
26: \newtheorem{remark}{Remark}
27: \newtheorem{note}{Note}
28: \newtheorem{notation}{Notation}
29: \newtheorem{claim}{Claim}
30: \newtheorem{summary}{Summary}
31: \newtheorem{acknowledgment}{Acknowledgment}
32: \newtheorem{case}{Case}
33: \newtheorem{conclusion}{Conclusion}
34: \newcommand{\abs}[1]{|#1|}
35: \newcommand{\norm}[1]{\|#1\|}
36: \newcommand{\calA}{{\cal{A}}}
37: \newcommand{\calB}{{\cal{B}}}
38: \newcommand{\calC}{{\cal{C}}}
39: \newcommand{\calD}{{\cal{D}}}
40: \newcommand{\calE}{{\cal{E}}}
41: \newcommand{\calF}{{\cal{F}}}
42: \newcommand{\calG}{{\cal{G}}}
43: \newcommand{\calH}{{\cal{H}}}
44: \newcommand{\calI}{{\cal{I}}}
45: \newcommand{\calK}{{\cal{K}}}
46: \newcommand{\calL}{{\cal{L}}}
47: \newcommand{\calM}{{\cal{M}}}
48: \newcommand{\calN}{{\cal{N}}}
49: \newcommand{\calO}{{\cal{O}}}
50: \newcommand{\calP}{{\cal{P}}}
51: \newcommand{\calQ}{{\cal{Q}}}
52: \newcommand{\calR}{{\cal{R}}}
53: \newcommand{\calS}{{\cal{S}}}
54: \newcommand{\calT}{{\cal{T}}}
55: \newcommand{\calU}{{\cal{U}}}
56: \newcommand{\calV}{{\cal{V}}}
57: \newcommand{\calZ}{{\cal{Z}}}
58: \newcommand{\vp}{{\varphi_0}}
59: \newcommand{\bbR}{{\mathbb{R}}}
60: \newcommand{\bbC}{{\mathbb{C}}}
61: \def\re{\mathop{\rm Re}}
62:
63:
64:
65: \begin{document}
66: \title{Acceleration-induced nonlocality: kinetic memory versus dynamic memory}
67: \author{C. Chicone\\Department of Mathematics\\University of
68: Missouri-Columbia\\Columbia, Missouri 65211, USA
69: \and B. Mashhoon\thanks{Corresponding author. E-mail:
70: mashhoonb@missouri.edu (B. Mashhoon).\newline Phone: (573) 882-6526;\;
71: FAX: (573) 882-4195.} \\Department of Physics and
72: Astronomy\\University of Missouri-Columbia\\Columbia, Missouri 65211, USA}
73: \maketitle
74: \begin{abstract}
75: The characteristics of the memory of accelerated motion in Min\-kow\-ski
76: spacetime are discussed within the framework of
77: the nonlocal theory of accelerated observers.
78: Two types of memory are distinguished: kinetic
79: and dynamic. We show that only kinetic memory is acceptable, since
80: dynamic memory leads to divergences for nonuniform accelerated motion.
81: \end{abstract}
82: \noindent PACS numbers: 03.30.+p, 11.10.Lm; Keywords: relativity, nonlocality
83: \section{Introduction}
84:
85: The special theory of relativity is based on two basic postulates:
86: Lorentz invariance and the hypothesis of locality. Lorentz invariance
87: refers to a fundamental
88: symmetry principle, namely, the invariance of basic physical laws
89: under inhomogeneous Lorentz transformations. In practice these laws
90: of nature involve physical
91: quantities measured by inertial observers in Minkowski spacetime.
92: An inertial observer always moves uniformly and refers
93: its observations to the fixed spatial axes
94: of an inertial frame; it can be depicted by a straight line in the Minkowski
95: diagram and represents an
96: ideal; in fact, physical observers are all effectively accelerated.
97: For instance, one can imagine the influence of radiation pressure on
98: the path of a cosmic particle.
99: In general, the acceleration of an observer consists of the translational
100: acceleration of its path as well as the rotation of its spatial frame.
101: Observers with translational acceleration
102: are therefore
103: represented by curved lines in the Minkowski diagram.
104: As an example of a rotating observer, consider a uniformly moving
105: observer that refers its observations to spatial axes that rotate with
106: respect to the spatial frame of the underlying inertial coordinate system.
107: The hypothesis of locality refers to the
108: measurements of realistic (i.e. accelerated) observers: such an
109: observer is postulated to be equivalent, at each event along its
110: worldline, to a momentarily comoving
111: inertial observer. The origin of this assumption can be traced back
112: to the work of Lorentz in the context of his classical electron
113: theory~\cite{1}; later, it was
114: simply adopted as a general rule in relativity theory~\cite{2}.
115:
116: Along its worldline, the accelerated observer passes through a
117: continuous infinity of hypothetical momentarily comoving inertial
118: observers. Stated mathematically, the translationally accelerated
119: observer's curved worldline
120: is the \emph{envelope} of the straight worldlines of this class
121: of hypothetical inertial observers. Therefore, the
122: hypothesis of locality has two components: (i) the assumption that
123: the measurements of the accelerated observer must be somehow
124: connected to the measurements of
125: the hypothetical class of momentarily comoving inertial observers
126: along its worldline and (ii) that this connection is
127: postulated to be the pointwise
128: equivalence of the accelerated observer and the momentarily comoving
129: inertial observer. The latter means that the acceleration of the
130: observer does not {\it
131: directly} affect the result of its measurement; devices that obey
132: this rule are called ``standard". Thus the hypothesis of locality is
133: a simple generalization of
134: the assumption that the rods and clocks of special relativity theory
135: are not directly affected by acceleration
136: \cite{2}.
137:
138: What is the physical basis for the hypothesis of locality? It is
139: difficult to argue with part (i) of this hypothesis, since the
140: fundamental laws of
141: (nongravitational) physics have been formulated with respect to
142: inertial observers and hence the measurements of accelerated
143: observers should be in some way
144: related to those of inertial observers.
145: On the other hand, part (ii) can only
146: be valid if the measurement process occurs instantaneously and
147: in a pointwise manner.
148: That is, (ii) is appropriate
149: for phenomena involving coincidences of classical point particles
150: and null rays.
151: Classical waves,
152: on the other hand, are extended in time and space with a
153: characteristic wave period $T$ and a
154: corresponding wavelength $\lambda$, respectively. Imagine, for
155: example, the measurement of the frequency of an incident
156: electromagnetic wave by an accelerated
157: observer; at least a few periods of the wave must be received by the
158: observer before an adequate determination of the frequency would
159: become possible. Thus this measurement
160: process is nonlocal and extends over the worldline of the observer.
161: The observer's acceleration can be characterized by certain
162: {\it acceleration lengths} $L$ given by
163: $c^2/g$ and $c/\Omega$ for translational acceleration $g$ and
164: rotational frequency $\Omega$, respectively.
165: The nonlocality of the external radiation is thus expected to couple
166: with the intrinsic scales associated with the acceleration of the observer.
167:
168: Classical wave phenomena are expected to violate the hypothesis of
169: locality. The scale of such violation would be given by $\lambda
170: /L=T/(L/c)$, where $L/c$ is the
171: {\it acceleration time}. The hypothesis of locality will hold if
172: $\lambda$ is so small that the incident radiation behaves like a ray,
173: i.e. in the eikonal (or JWKB)
174: limit such that
175: $\lambda /L\to 0$; alternatively, $L$ can be so large that
176: deviations of the form $\lambda /L$ would be below the sensitivity
177: threshold of the detectors available
178: at present. Consider, e.g., laboratory experiments on the Earth;
179: typical acceleration lengths would be $c^2/g_\oplus \simeq 1\,{\rm lyr}$ and
180: $c/\Omega _\oplus \simeq 28\, {\rm AU}$,
181: so that for essentially all practical purposes one can ignore
182: any possible deviations from locality at the present time. In this
183: way, we can account for the
184: fact that the standard theory of relativity is in agreement with all
185: observational data available at present. As a matter of principle,
186: however, it is necessary to
187: contemplate generalizations of the hypothesis of locality in order to
188: take due account of intrinsic wave phenomena for realistic
189: (accelerated) observers.
190:
191: All of our considerations in this paper are within the framework of
192: classical field theory; nevertheless, it is necessary to remark that
193: quantum theory is based on
194: the notion of wave-particle duality, and so an adequate treatment of
195: classical wave phenomena is a necessary prelude to a satisfactory
196: quantum theory.
197:
198: To proceed, we consider the most general extension of the hypothesis
199: of locality that is consistent with causality and the superposition
200: principle. A nonlocal Lorentz-invariant
201: theory of accelerated observers has been developed along these
202: lines~\cite{3,3a, 3b,3c}
203: and is presented in Section~2. The theory involves a kernel
204: that depends primarily on the
205: acceleration of the observer; that is, the measurements of the
206: observer depend on its past history of acceleration.
207: The main physical principle that is employed in the nonlocal theory
208: for the determination of the kernel is the assumption that
209: an intrinsic radiation field can never stand completely still with
210: respect to an accelerated observer;
211: this statement involves a simple generalization of a property of inertial
212: observers to all observers.
213: Thus the
214: accelerated observer is endowed with
215: {\it memory}, and the past affects the present through an averaging
216: process, where the weight function is proportional to the kernel
217: $K(\tau ,\tau')$. It turns out
218: that the kernel $K$ cannot be completely determined by the theory
219: presented in Section~2. An additional simplifying assumption is
220: therefore introduced in Section~3:
221: $K(\tau ,\tau ')$ must be a function of a single variable. Two
222: cases are then considered:
223: (1) $K(\tau ,\tau ')=k_0(\tau ')$
224: and
225: (2) $K(\tau ,\tau ')=k(\tau-\tau').$
226: We show that case~(1)---i.e. the
227: \emph{kinetic} memory case---has acceptable
228: properties that are described in Section 3. Case~(2), i.e.
229: the \emph{dynamic} memory case, is treated in detail in Sections 3
230: and 4, where it is shown that the kernel function $k$ can be unbounded
231: even if the observer's past history has constant velocity except
232: for one episode of smooth translational acceleration with finite duration.
233: Specifically, we study the measurement of electromagnetic
234: radiation fields by an
235: observer that undergoes translational or rotational acceleration
236: that lasts for only a finite interval of its proper time.
237: After the acceleration is turned off,
238: the observer measures in addition to the regular field
239: a residual field that contains the
240: memory of its past acceleration. This leftover piece is a finite
241: \emph{constant} field (\emph{kinetic} memory) in case (1); however,
242: it is time dependent (\emph{dynamic} memory) in case (2).
243: We rule out the latter case, since we prove that the measured field could
244: diverge under certain reasonable circumstances. We are thus left
245: with a unique theory that involves kinetic memory.
246: An important aspect of our nonlocal ansatz is that the kernel
247: induced by the acceleration of the observer
248: depends on the spin of the radiation field under consideration.
249: In particular, the kernel vanishes for an intrinsic scalar field,
250: i.e. such a field is always \emph{local}.
251: As discussed in Section~5,
252: our theory therefore rules out the possibility that a pure
253: scalar (or pseudoscalar) field exists in nature.
254: This conclusion is in agreement with available
255: experimental data.
256: The nonlocal theory therefore predicts that any scalar particle would
257: have to be a composite.
258: Section 5 contains a brief discussion and our
259: conclusions.
260: A detailed discussion of the observational consequences of the nonlocal
261: theory is beyond the scope of this work.
262: In the following, we use units such that
263: $c=1$, i.e. the speed of light in vacuum is unity.
264:
265: \section{Accelerated observers and nonlocality}
266:
267:
268: The measurement of length by accelerated observers involves subtle
269: issues in relativity theory that have been investigated in detail
270: \cite{4,4a,4b}; for our present
271: purpose, the main result of such studies is that an accelerated frame
272: of reference, i.e. an extended coordinate system set up in the
273: neighborhood of an accelerated
274: observer, is of rather limited theoretical significance. We shall
275: therefore refer all measurements to an inertial reference frame in
276: Minkowski spacetime.
277:
278: Imagine a global inertial frame with coordinates $x=(t,\mathbf{ x})$
279: and the standard class of static inertial observers with their
280: orthonormal tetrad frame
281: $\lambda^\mu_{(\alpha)}=\delta ^\mu _\alpha$, where $\lambda^\mu
282: _{(0)}$ is the temporal direction at each event and $\lambda^\mu
283: _{(i)}$, $i=1,2,3$, are the spatial
284: directions. The hypothesis of locality implies that an accelerated
285: observer is also endowed with a tetrad frame
286: $\hat{\lambda}^\mu_{(\alpha)}(\tau )$, where $\tau
287: $ is the proper time along its worldline. For each $\tau$,
288: $\hat{\lambda}^\mu_{(\alpha)}(\tau)$ coincides with the constant
289: tetrad frame (related to $\lambda^\mu
290: _{(\alpha)}$ by a Lorentz transformation) of the momentarily comoving
291: inertial observer. We note that
292: $
293: d\hat{\lambda}^\mu_{(\alpha)}/d\tau
294: =\phi_\alpha ^{\hspace{0.075in}\beta}\hat{\lambda}^\mu _{(\beta)}
295: $,
296: where $\phi_{\alpha
297: \beta}=-\phi_{\beta \alpha}$ is a tensor such that
298: $\phi_{0 i}=(\mathbf{g})_i$ and
299: $\phi_{ij}=\epsilon_{ijk}($\boldmath$\Omega$\unboldmath$)_k$. Here
300: $\mathbf{g}(\tau )$ is the translational acceleration of the observer
301: and
302: \boldmath$\Omega$\unboldmath$(\tau )$ is the rotational frequency of
303: its spatial frame.
304: Each element of the acceleration tensor $\phi_{\alpha\beta}$
305: is a scalar under the inhomogeneous Lorentz transformations
306: of the background spacetime.
307: We assume throughout that the acceleration is
308: turned on at $\tau =\tau_0$
309: and will in general be turned off at $\tau_1>\tau_0$.
310:
311: Let $f_{\mu \nu}$ represent an electromagnetic radiation field as
312: measured by the standard set of static inertial observers. According
313: to the hypothesis of
314: locality $\hat{f}_{\alpha \beta}=f_{\mu \nu}\hat{\lambda
315: }^\mu_{(\alpha)}\hat{\lambda}^\nu _{(\beta)}$, i.e. the projection of
316: the field on the instantaneous
317: tetrad frame, would be the field measured by the
318: accelerated observer. On the other hand, let $F_{\alpha \beta }(\tau
319: )$ be the true result of such a
320: measurement. Taking causality into account, the most general linear
321: relationship between $F_{\alpha \beta}(\tau )$ and $\hat{f}_{\alpha
322: \beta}(\tau )$ is
323: \begin{equation}\label{eq1}
324: F_{\alpha \beta}(\tau )
325: =\hat{f}_{\alpha \beta}(\tau )
326: +\int^\tau_{\tau_0}K_{\alpha \beta \gamma\delta }(\tau ,\tau ')
327: \hat{f}^{\gamma\delta}(\tau')\, d\tau '.
328: \end{equation}
329: This relation refers to quantities that are all scalars under
330: the Poincar\'e group of spacetime transformations of the underlying
331: inertial coordinate system.
332: We note that the magnitude of the nonlocal part of equation
333: \eqref{eq1} is of the form $\lambda /L$ if the kernel is proportional
334: to the acceleration of the
335: observer. It follows from Volterra's theorem that in the space of
336: continuous functions the relationship between $F$ and $f$ is
337: unique~\cite{5,5a};
338: this theorem has been
339: extended to the Hilbert space of square-integrable functions
340: by Tricomi~\cite{5b}.
341:
342: The basic ansatz~\eqref{eq1} is consistent with an observation originally
343: put forward by Bohr and Rosenfeld that the electromagnetic field cannot
344: be measured at a spacetime \emph{point}; in fact, an averaging process
345: is necessary over a spacetime neighborhood~\cite{br,bra}. In the case of
346: measurements by \emph{inertial} observers envisaged by
347: Bohr and Rosenfeld~\cite{br,bra}, there is no intrinsic temporal or
348: spatial scale
349: associated with the inertial observers;
350: therefore, one can effectively pass to the
351: limiting case of a point with no difficulty as
352: the dimensions of the spacetime neighborhood can be shrunk to zero
353: without any obstruction. For an accelerated observer, however, the
354: intrinsic acceleration time and length need to be properly taken into
355: account. Hence the nonlocal ansatz~\eqref{eq1} may be interpreted in
356: terms of a certain averaging process over the past worldline of the
357: accelerated observer.
358:
359: To determine the kernel $K$, let us first mention a basic consequence
360: of the hypothesis of locality for a radiation field. Imagine plane
361: monochromatic
362: electromagnetic waves of frequency $\omega$ propagating along the
363: $z$-axis and an observer rotating uniformly about this axis with
364: frequency $\Omega_0$ in the
365: $(x,y)$-plane on a circle of radius $\rho$ in the underlying
366: inertial reference frame. We find from
367: $\hat{f}_{\alpha \beta}=f_{\mu \nu}\hat{\lambda}^\mu_{(\alpha
368: )}\hat{\lambda}^\nu _{(\beta)}$ that
369: according to the rotating observer the frequency of the wave is
370: $\hat{\omega}=\gamma (\omega \mp \Omega_0)$, where $\gamma$ is the
371: Lorentz factor corresponding to
372: the speed $\rho \Omega_0$ of the observer and the upper (lower) sign
373: refers to incident positive (negative) helicity radiation. This result
374: has a simple intuitive
375: interpretation: In an incident positive (negative) helicity wave the
376: electric and magnetic field vectors rotate with frequency $\omega
377: (-\omega)$ about the
378: direction of propagation of the wave. As seen by the rotating
379: observer, the field vectors rotate with frequency
380: $\omega-\Omega_0\,(-\omega-\Omega_0)$ with respect to
381: the inertial temporal coordinate $t$; moreover, the Lorentz
382: factor simply accounts for time dilation $dt=\gamma d\tau$. It
383: follows that a positive helicity incident wave can stand completely
384: still with respect to all observers rotating uniformly
385: with frequency $\Omega_0=\omega$. In terms of energy, we
386: have
387: $\hat{E}=\gamma(E-$\boldmath$\sigma$\unboldmath$\cdot$
388: \boldmath$\Omega$\unboldmath$_0)$,
389: where
390: \boldmath$\sigma$\unboldmath\; is the spin of the incident photon. More
391: generally, for oblique incidence $\hat{E}=\gamma (E- \hbar
392: M\Omega_0)$, where $M$ is the
393: multipole parameter such that $\hbar M$ is the component of the total
394: (orbital plus spin) angular momentum along the $z$-axis.
395: This is an example of the
396: general phenomenon of
397: spin-rotation coupling; various aspects of this effect and the
398: available observational evidence are discussed in~\cite{6,6a,6b,6c,6d,6e}.
399: Again,
400: the incident wave can
401: theoretically stand completely still for all observers rotating with
402: frequency $\Omega_0$ such that $\omega =M\Omega_0$. Let us recall
403: here a fundamental
404: consequence of Lorentz invariance, namely that a radiation field can
405: never stand completely still with respect to an inertial observer.
406: That is, an inertial observer can
407: move along the direction of propagation of a wave so fast that the
408: frequency $\hat{\omega}=\gamma \omega (1-\beta)$ can approach zero
409: but the mathematical limit of
410: $\hat{\omega}=0$ is never physically achieved, since the observer's
411: speed cannot reach the speed of light in vacuum $(\beta <1)$.
412: Therefore, for an inertial observer $\hat{\omega}=0$
413: implies that $\omega=0$. On the other hand, while we find that the
414: hypothesis of locality predicts that a circularly polarized wave can
415: stand completely still with
416: respect to a uniformly rotating observer, this possibility
417: can be avoided in the nonlocal theory by an appropriate choice of the
418: kernel.
419:
420: To implement the requirement that a radiation field can never stand
421: completely still with respect to any observer, we assume that if
422: $F_{\alpha \beta}(\tau )$
423: turns out to be constant in equation~\eqref{eq1}, then $f_{\mu \nu}$
424: must have been originally constant just as in the case of
425: inertial observers in the standard theory of relativity.
426: It is convenient to replace the tensor
427: $f_{\mu \nu}$ by a six-vector
428: $f$, with electric and magnetic fields as components, and introduce
429: the ``Lorentz'' matrix $\Lambda $ such that $\hat{f}=\Lambda f$. Then for
430: constant fields $f$ and $F$,
431: equation~\eqref{eq1} takes the form
432: \begin{equation}\label{eq2} F=\Lambda(\tau) f+\int^\tau_{\tau_0}K(\tau
433: ,\tau ')\Lambda (\tau')f\,d\tau',\end{equation}
434: where for $\tau=\tau_0$, the matrix
435: $\Lambda_0:=\Lambda(\tau_0)$ is constant and
436: $F=\Lambda_0 f$. Thus in the nonlocal
437: theory the kernel $K$ should be determined from the Volterra integral
438: equation
439: \begin{equation}\label{eq3} \Lambda_0=\Lambda (\tau
440: )+\int^\tau _{\tau_0}K(\tau ,\tau')\Lambda (\tau ')\,d\tau'.
441: \end{equation}
442: It follows from Volterra's theory (see Appendix A) that to every
443: kernel $K$ corresponds a unique {\it resolvent} kernel $R(\tau
444: ,\tau')$ such that
445: \begin{equation}\label{eq4} \Lambda (\tau )=\Lambda_0
446: +\int^\tau_{\tau_0}R(\tau ,\tau')\Lambda_0\,d\tau
447: '.\end{equation}
448: Therefore, only the integral of the resolvent kernel is
449: completely determined by our physical requirement
450: \begin{equation}\label{eq5}
451: \int^\tau_{\tau_0} R(\tau ,\tau')\,d\tau '
452: =\Lambda (\tau )\Lambda_0^{-1}-I,\end{equation}
453: where $I$ is the unit matrix. It is clear at this point that given
454: $\Lambda (\tau )$, relations \eqref{eq3}--\eqref{eq5} are not
455: sufficient to determine the kernel
456: $K$ uniquely. To proceed further, other simplifying
457: restrictions are necessary on $K$ or $R$,
458: \begin{equation}\label{eq6} \hat{f}(\tau )
459: =F(\tau)+\int^\tau_{\tau_0}R(\tau ,\tau')F(\tau ')\,d\tau '.
460: \end{equation}
461: This must be done in such a way as to preserve time translation
462: invariance in the underlying inertial coordinate system.
463:
464: Let us finally remark that for a scalar field, $\Lambda (\tau )=1$ and
465: equations \eqref{eq3}--\eqref{eq5} simply reduce to the requirement
466: that $K(\tau ,\tau')$ must
467: have a vanishing integral over $\tau':\tau_0\to \tau$. That is, the
468: connection between the kernel and the acceleration of the observer
469: disappears. This
470: circumstance is further discussed in Section 5.
471:
472: \section{Memory}
473:
474: It is necessary to introduce simplifying assumptions in order to
475: find a unique kernel $K$. We therefore tentatively postulate that $K$
476: is a function of a single
477: variable. There are two reasonable possibilities:
478: \begin{align*} \tag{case 1} K(\tau ,\tau')=k_0(\tau ')\\
479: \intertext{and}
480: \tag{case 2} K(\tau ,\tau')=k(\tau -\tau');
481: \end{align*}
482: in either case, the basic requirement of time translation invariance
483: in the background global inertial frame is satisfied.
484: \subsection{Kinetic memory}
485: In case~(1), the kernel $k_0$ corresponds to a simple weight
486: function that can be determined by differentiating
487: equation~\eqref{eq3},
488: \begin{equation}\label{eq7} k_0(\tau
489: )=-\frac{d\Lambda}{d\tau}\Lambda^{-1}(\tau )=\Lambda (\tau
490: )\frac{d\Lambda^{-1}}{d\tau }.\end{equation}
491: The kernel $k_0$ is thus directly proportional to the acceleration of
492: the observer. A significant feature of this kernel is that once the
493: acceleration is turned
494: off at $\tau =\tau_1$, then for $\tau >\tau_1$,
495: \begin{equation}\label{eq8} F(\tau )=\hat{f}(\tau
496: )+\int^{\tau_1}_{\tau _0}k_0(\tau ')\hat{f}(\tau ')\,d\tau'.
497: \end{equation}
498: There is therefore a {\it constant} memory of past acceleration and
499: the field $F$ satisfies the standard field equations in the inertial
500: frame. That is, the field
501: equations are linear differential equations and the addition of a
502: constant solution is always permissible but subject to boundary
503: conditions. In terms of actual
504: laboratory devices that have experienced accelerations in the past,
505: such constant fields as in equation~\eqref{eq8} would be canceled
506: once the devices are reset.
507: Thus case~(1) involves simple ``nonpersistent'' memory of past
508: acceleration; therefore, we call $k_0$ the {\it kinetic memory}
509: kernel.
510:
511: It is interesting to note that our basic integral
512: equation~\eqref{eq2} together with the kinetic memory kernel~\eqref{eq7}
513: and an integration by parts takes the form
514: \[
515: F(\tau)=F(\tau_0)+\int_{[\tau_0,\tau]} \Lambda df,
516: \]
517: so that $dF=\Lambda df$ along the worldline of the accelerated observer.
518: \subsection{Dynamic memory}
519: The second case involves a convolution type kernel
520: $K=k(\tau -\tau ')$. It follows
521: (see Appendix A) that in this case the resolvent
522: kernel is of convolution type
523: as well, $R=r(\tau -\tau')$. Thus equation~\eqref{eq5} can be
524: written, after expressing the left side as the area under the
525: graph of the function $r$ from the origin to
526: $\tau-\tau_0=t$, as
527: \begin{equation}
528: \label{eq9}r(t)=\frac{d\Lambda
529: (t+\tau_0)}{dt}\Lambda_0^{-1}.
530: \end{equation}
531: The kernel $k$ is then given by (cf. Appendix A)
532: \begin{equation}\label{eq10} k(t)=-r (t)+r\ast r(t)-r\ast r\ast
533: r(t)+\cdots ,\end{equation}
534: where a star denotes the convolution operation. We note that in this
535: case the {\it resolvent} kernel is directly proportional to
536: acceleration, so that $r=0$ and,
537: by equation~\eqref{eq10}, $k=0$ for $t<0$ or $\tau <\tau_0$, i.e.
538: before the acceleration is turned on. However, the character of
539: memory that is indicated by $k$,
540: \begin{equation}\begin{split}\label{eq11} F(\tau )&=\hat{f}(\tau
541: )+\int^\tau_{\tau_0}k(\tau -\tau ')\hat{f}(\tau ')\,d\tau '\\
542: &=\hat{f}(\tau)+\int^{\tau-\tau_0}_0k(t)\hat{f}(\tau
543: -t)\,dt,\end{split}\end{equation}
544: is more complicated than in case~(1)
545: due to the intricate relationship between $r(t)$
546: and $k(t)$ in equation~\eqref{eq10}. Even if the acceleration is
547: turned off at $\tau =\tau_1$,
548: it turns out that $k$ does not vanish in general for $\tau>\tau_1$
549: and could even be divergent; in fact, proving the latter point is the
550: main purpose of this paper.
551:
552: Imagine, for instance, that $k(t)$ is finite everywhere and decays
553: exponentially to zero for $t\to \infty$. Then in
554: equation~\eqref{eq11}, as $\tau \to \infty$
555: long after the acceleration has been turned off at $\tau =\tau_1$,
556: the contribution of the nonlocal term in \eqref{eq11} rapidly
557: approaches a constant and we essentially
558: recover the ``nonpersistent'' kinetic memory familiar from
559: case~(1). It turns out, however, that in general
560: case~(2) involves situations with persistent
561: or {\it dynamic} memory such that under certain conditions $k(t)$
562: could diverge resulting in an asymptotically divergent $F(\tau )$.
563:
564: The convolution (Faltung) type kernel is generally employed in many
565: branches of physics and mathematics. As in equation~\eqref{eq11}, to
566: produce the nonlocal part
567: of the output $F(\tau )$, an input signal $\hat{f}(\tau-t)$ is
568: linearly folded, starting from $\tau$ and going backwards in proper
569: time until $\tau_0$, with a
570: weight function $k(t)$ that is the impulse response of the system.
571: The use of convolution type kernels is standard practice in
572: phenomenological treatments of the
573: electrodynamics of media~\cite{7,7a,7b}, feedback control systems~\cite{8},
574: etc. We find, however, that for the pure vacuum case the convolution
575: kernel due to \emph{nonuniform acceleration}
576: in general leads to instability and is
577: therefore unacceptable. This proposition is proved in the following
578: section for the translational and rotational
579: accelerations of the observer.
580:
581: The simplicity of the kinetic memory versus dynamic memory has been
582: particularly stressed by Hehl and Obukhov in their investigations of
583: nonlocal electrodynamics~\cite{ho,hoa}; moreover, their work has led
584: to the question of the ultimate physical significance of the convolution type
585: kernel in the nonlocal theory of accelerated systems~\cite{ho,hoa}.
586: This question is settled in the present paper in favor of the
587: kinetic memory kernel.
588: \section{Dynamic memory of accelerated motion}
589: \subsection{Linear acceleration}
590:
591: Imagine an observer at rest on the $z$-axis for $-\infty <\tau
592: <\tau_0$. At $\tau =\tau _0$, the observer accelerates along the
593: positive $z$-direction with
594: acceleration $g(\tau )>0$. For $\tau\ge \tau_0$, we set
595: \begin{equation}\label{eq12}
596: \theta (\tau )=\int^\tau_{\tau_0}g(\tau ')\,d\tau ',\end{equation}
597: $C=\cosh \theta$ and $S=\sinh \theta$. The natural nonrotating
598: orthonormal tetrad frame of the observer along its worldline is given
599: by
600: \begin{equation}\begin{array}{ll} \hat{\lambda}^\mu_{(0)} = (C,0,0,S), &
601: \hat{\lambda}^\mu_{(1)}=(0,1,0,0),\\
602: \hat{\lambda}^\mu_{(2)}=(0,0,1,0), &
603: \hat{\lambda}^\mu_{(3)}=(S,0,0,C).\end{array}\end{equation}
604: In this case $\Lambda (\tau )$ is given by
605: \begin{equation}\label{eq14} \Lambda =\begin{bmatrix} U & V\\-V
606: &U\end{bmatrix},\quad U=\begin{bmatrix} C & 0 & 0\\ 0 & C & 0\\0 & 0
607: & 1\end{bmatrix},\quad
608: V=SI_3,\end{equation}
609: where $I_i$, $(I_i)_{jk}=-\epsilon_{ijk}$, is a $3\times 3$ matrix
610: proportional to the operator of infinitesimal rotations about the
611: $x^i$-axis.
612:
613: Let us first consider case~(1), for which the kernel can be
614: easily computed using equation~\eqref{eq7},
615: \begin{equation}\label{eq15} k_0(\tau )=-g(\tau )\begin{bmatrix}0 &
616: I_3\\ -I_3 & 0\end{bmatrix},\end{equation}
617: so that when the acceleration is turned off at $\tau =\tau_1$ the
618: kernel $k_0$ vanishes with the acceleration for $\tau \geq \tau_1$.
619: On the other hand, $k_0$ is
620: simply constant for uniform acceleration (i.e. hyperbolic motion)
621: with $g(\tau )=g_0$ for $\tau \geq \tau_0$. In the rest of this
622: section, we focus attention on
623: case~(2) involving the convolution kernel.
624:
625: For the convolution kernel, the resolvent kernel is given, via
626: equation~\eqref{eq9}, by
627: \begin{equation}\label{eq16} r(\tau -\tau_0)=g(\tau )\begin{bmatrix}
628: SJ_3 & CI_3\\-CI_3 & SJ_3\end{bmatrix},\end{equation}
629: where $(J_k)_{ij}=\delta_{ij}-\delta_{ik}\delta_{jk}$. In principle,
630: the convolution kernel can be computed via the substitution of
631: equation~\eqref{eq16} in
632: equation~\eqref{eq10}; however, this turns out to be a daunting task in
633: practice. Imagine, for instance, that the acceleration is turned off
634: at $\tau=\tau_1$, so
635: that the resolvent kernel \eqref{eq16} has compact support over a
636: time interval of length $\alpha=\tau_1-\tau_0$ and vanishes
637: otherwise. It then follows that the
638: $r^{\ast\, n}$ term in the expansion \eqref{eq10} has compact support over
639: a time interval of length $n\alpha$. The summation of series
640: \eqref{eq10} turns out to be
641: rather complicated, except for the case of {\it uniform}
642: acceleration, i.e. $g(\tau )=g_0$ for $\tau \geq \tau_0$, and the
643: result is
644: \begin{equation}\label{eq17} k=-g_0\begin{bmatrix} 0 & I_3\\ -I_3 &
645: 0\end{bmatrix}.\end{equation}
646: It is interesting to note that equation \eqref{eq17} is the \emph{same} as
647: the result of case~(1), equation~\eqref{eq15}, for \emph{uniform}
648: acceleration.
649:
650: In view of the difficulty of summing the series \eqref{eq10}
651: directly, we find it advantageous to use Laplace transforms, which we
652: denote by an overbar, i.e.
653: $\mathcal{L} \{ k(t)\}=\bar{k}(s)$, where
654: \begin{equation}\label{eq18} \bar{k} (s)=\int^\infty_0
655: e^{-st}k(t)\,dt;\end{equation}
656: then, taking the Laplace transform of equation~\eqref{eq10} and using
657: the convolution (Faltung) theorem repeatedly, we arrive at
658: \begin{equation}\label{eq19} \bar{k} (s)=[I+\bar{r} (s)]^{-1}-I,\end{equation}
659: which is consistent with the reciprocity between $k$ and $r$.
660:
661: \subsection{Stepwise acceleration}\label{sec:sa}
662: \begin{figure}[tb]
663: \centerline{\psfig{file=fig1.eps, width=25pc}}
664: \caption{\label{fig1}
665: The linear acceleration of an observer that undergoes uniform
666: acceleration $g_0$ during a period $\alpha=\tau_1-\tau_2$ of its
667: proper time. If the area under the graph exceeds a critical value given
668: by $\beta_0\approx 1.2931$,
669: then the convolution kernel leads to divergences.}
670: \end{figure}
671: Let us specialize to a simple case of stepwise
672: uniform acceleration, namely, we let $g(\tau)=g_0$ for $\tau_0\leq
673: \tau \leq \tau_1$ and zero
674: otherwise (see Figure~\ref{fig1}). In this case,
675: \begin{equation} \label{eq20} \bar{r} (s)=\begin{bmatrix} \bar{r}_1 &
676: \bar{r}_2\\ -\bar{r}_2 & \bar{r}_1\end{bmatrix},\end{equation}
677: where $\bar{r}_1(s)=q(s)J_3$ and $\bar{r}_2(s)=p(s)I_3$. Here
678: $p(s)=\mathcal{L}\{gC\}$ and $q(s)=\mathcal{L}\{ gS\}$. All $6\times
679: 6$ matrices that we consider in
680: this paper have the general form \eqref{eq20}, i.e. each is
681: completely determined by two $3\times 3$ matrices just as $\bar{r}_1$
682: and $\bar{r}_2$ characterize
683: $\bar{r}$ in equation~\eqref{eq20}; we therefore write $\bar{r}\to
684: [\bar{r}_1;\bar{r}_2]$ to express this decomposition as
685: in equation~\eqref{eq20}. To find the Laplace
686: transforms of $gC$ and $gS$, we
687: note that in equation~\eqref{eq12}, $\theta=g_0(\tau-\tau_0)$ for
688: $\tau \leq \tau_1$ and $\theta=\beta_0=g_0(\tau _1-\tau_0)$ for
689: $\tau \geq \tau_1$; therefore,
690: \begin{equation}\label{eq21} p(s)\pm q(s)=\frac{g_0}{s\mp
691: g_0}[1-e^{-(s\mp g_0)\alpha}],\end{equation}
692: where $\alpha = \tau_1-\tau_0=\beta_0/g_0$ is the acceleration time
693: interval. Using the results of Appendix B, we find from
694: equation~\eqref{eq19} that $\bar{k}(s)$
695: can be expressed as
696: \[\bar{k}(s)\to [\beta_0Q(s)J_3;\beta_0P(s)I_3],\] where
697: \begin{align}\label{eq22}
698: P(s)&=\frac{e^w}{D}[w(-e^w+\cosh \beta_0)+\beta_0\sinh \beta_0],\\
699: \label{eq23} Q(s) &= \frac{1}{D} [e^w(w\sinh \beta_0-\beta_0\cosh
700: \beta_0)+\beta_0].
701: \end{align}
702: Here $w:=s\alpha$ and the denominator $D$ can be factorized as
703: \begin{equation}\label{eq24}
704: D=(we^w-\beta_0e^{\beta_0})(we^w+\beta_0e^{-\beta_0}).
705: \end{equation}
706:
707: It is useful to recall that the kernel $k\to [k_1;k_2]$
708: refers to a system at rest
709: on the $z$-axis for $\tau \leq \tau_0$ that is uniformly accelerated
710: at $\tau=\tau_0$
711: with acceleration $g_0$
712: until $\tau_0+\alpha =\tau_1$, and then continues with uniform speed
713: $\tanh \beta_0$ along the positive $z$-direction for $\tau \geq
714: \tau_1$. Under certain
715: conditions, it is possible to obtain series representations for $k_1$
716: and $k_2$ (see Appendix C); however, to gain insight into the
717: asymptotic behavior of $k_1$
718: and $k_2$ it proves more fruitful to proceed with an investigation of
719: the singularities of $\bar{k}_1(s)=\beta_0 Q(s) J_3$ and
720: $\bar{k}_2(s)=\beta_0 P(s)I_3$ in the complex
721: $s$-plane.
722: This is due to a simple property of the Laplace transformation in
723: equation~\eqref{eq18} extended to the complex $s$-plane: let us
724: suppose that the convolution kernel $k(t)$ is a bounded function
725: for all $t=\tau-\tau_0>0$ as one naturally expects of a function
726: that represents memory; then, for any $s$ in the complex plane with
727: positive real part, i.e. $\re(s)>0$, equation~\eqref{eq18} implies
728: that the absolute magnitude of $\bar k(s)$ should be finite,
729: i.e. $\bar k(s)$ cannot be singular. Therefore, if we could show that
730: $\bar k(s)$ has in fact pole singularities at complex values of $s$
731: with $\re(s)>0$, then it would simply follow that
732: $k(t)$ cannot be bounded for all $t>0$ and would thus be unsuitable
733: to represent the memory of finite accelerated motion.
734:
735: We will prove the following result: If $\beta_0 \exp(\beta_0)> 3\pi/2$,
736: then the corresponding function $k$ is unbounded for $t\ge 0$.
737: It suffices to show that $\bar k$ has a pole in the right half of
738: the complex $s$-plane. In fact, let us suppose that $\bar k$ has a pole
739: at $s=s_0$, where $\re (s_0)>0$,
740: but $\norm{k}:=\sup_{t\ge 0} \abs{k(t)}<\infty$.
741: In this case, $\bar k$ has a pole in the half-plane
742: $\calH$ consisting of all complex numbers $s$ such that
743: $\re(s)\ge \frac{1}{2} \re(s_0)$, and therefore $\abs {\bar k}$ is not
744: bounded on $\calH$. On the other hand, for $s\in \calH$,
745: we have that
746: \[
747: \abs{\bar k(s)}\le \int_0^\infty e^{-\re(s) t}\abs{k(t)}\,dt
748: \le \norm{k}\int_0^\infty e^{-\re(s_0) t/2}\,dt<\infty,
749: \]
750: in contradiction. Thus the rest of this subsection is devoted
751: to the
752: determination of the poles of $\bar k(s)$ in the right half-plane.
753:
754:
755: The poles of $\bar k$ are elements of the zero set of $D$ with
756: $w=s \alpha$ and $\alpha>0$. Note, however,
757: that the (real) zeros $w=\pm\beta_0$ are removable singularities.
758: Poles in the right half-plane are the zeros of $D$ with nonzero
759: imaginary parts. In view of the definition of $D$, let us consider
760: the complex roots in the right half-plane of the
761: equation $w\exp(w)=b$,
762: where $b$ is one of the real numbers $\pm\beta_0\exp(\pm \beta_0)$.
763: Because the zero set of this relation is symmetric with respect
764: to the real axis, it suffices to consider only roots in the
765: first quadrant of the complex $w$-plane.
766:
767: We set $w=\xi+i\eta$, where $\xi\ge 0$ and $\eta\ge 0$ are real variables,
768: and note that $w\exp(w)=b$ if and only if
769: \[
770: \xi e^\xi=b\cos\eta,\qquad \eta e^\xi=-b \sin\eta.
771: \]
772: If this system of equations has a solution,
773: then, by squaring, adding and rearranging,
774: we have that
775: $
776: \eta^2=b^2 \exp(-2\xi)-\xi^2
777: $
778: or, since $\eta\ge 0$, $\eta=\sqrt{b^2 \exp(-2\xi)-\xi^2}$.
779:
780:
781: There are several cases. For example,
782: for $b>0$, there is a pole in the right half-plane
783: if the system of equations
784: \[
785: \eta=\sqrt{b^2 e^{-2\xi}-\xi^2},\qquad \xi e^\xi=b\cos\eta
786: \]
787: has a solution with $\xi>0$ and $\eta\bmod 2\pi\in (3\pi/2,2\pi)$.
788: Similarly, for $b<0$, there is a pole in the right half-plane
789: if the system of equations
790: \[
791: \eta=\sqrt{b^2 e^{-2\xi}-\xi^2},\qquad \xi e^\xi=b\cos\eta
792: \]
793: has a solution with $\xi>0$ and $\eta\bmod 2\pi\in (\pi/2,\pi)$.
794:
795: A necessary condition for the relation
796: $\eta=\sqrt{b^2 \exp(-2\xi)-\xi^2}$ to have a solution $(\xi,\eta)$
797: is that $\xi \exp(\xi)<\abs{b}$.
798: For $b<0$, we must have $\xi \exp(\xi)<\beta_0\exp(-\beta_0)$; hence,
799: there is a unique real number $\xi_0$ such that the necessary condition
800: is met whenever $\xi\le \xi_0$.
801: On the other hand, for $b>0$, the necessary condition,
802: $\xi \exp(\xi)<\beta_0\exp(\beta_0)$, is met if and only if $\xi<\xi_0=\beta_0$.
803:
804: Let us view $\eta$ as a function of $\xi$ and note that
805: $\eta(0)=\abs{b}$, $\eta(\xi_0)=0$, and
806: \[
807: \eta\frac{d\eta}{d\xi}=-e^{-2\xi}b^2-\xi<0
808: \]
809: for $\xi\ge 0$. In particular, $\eta$ decreases monotonically for
810: $0\le \xi\le \xi_0$.
811:
812: Consider the relation $\xi \exp(\xi)=b\cos\eta$.
813: At $\xi=0$, we have $\cos\eta=0$; therefore, the implicitly
814: defined function $\eta$ is such that
815: $\eta(0)$ is an odd integer multiple of $\pi/2$.
816: At $\xi_0$, we have $\cos\eta=\pm 1$ according to the sign
817: of $b$. In fact, $\eta(\xi_0)$ is an even multiple of
818: $\pi$ for $b>0$ and an odd multiple of $\pi$ for $b<0$.
819: Also, let us note that
820: \[
821: (\xi+1)e^\xi=-b\frac{d\eta}{d\xi}\sin\eta .
822: \]
823:
824: \begin{figure}[tb]
825: \centerline{\psfig{file=fig2.eps, width=25pc}}
826: \caption{\label{fig2}The real branches of $\xi \exp(\xi)=b\cos\eta$
827: for $b>0$.}
828: \end{figure}
829: Suppose that $b>0$. We will determine the positions of the
830: real branches of the curve defined
831: by $\xi \exp(\xi)=b\cos\eta$. For $0\le \eta\le \pi/2$, we have
832: $\sin\eta>0$ and $d\eta/d\xi<0$, so there is a real branch
833: connecting the points $(0,\pi/2)$ and $(\xi_0,0)$ in the
834: $(\xi,\eta)$-plane. For $\pi/2<\eta< 3\pi/2$, we have
835: $\cos\eta<0$; thus, there is no real branch in this region.
836: There is a real branch connecting $(0,3\pi/2)$ and $(\xi_0,2\pi)$
837: with $d\eta/d\xi>0$.
838: This pattern continues as depicted in Figure~\ref{fig2}.
839: Note, however, that only the ``increasing'' branches correspond to
840: poles in the right half-plane. Indeed, for $b>0$, it is necessary that
841: $\eta\bmod 2\pi$ be in the interval $(3\pi/2,2\pi)$. In particular,
842: the ``lowest'' branch corresponding to a pole connects
843: the points $(0,3\pi/2)$ and $(\xi_0,2\pi)$. It is now clear that
844: the curve defined by $\eta=\sqrt{b^2 \exp(-2\xi)-\xi^2}$ intersects
845: an increasing branch with $\xi>0$ if and only if $b>3\pi/2$.
846: The number of poles in the right half-plane increases by one
847: as $b$ increases past an odd multiple of $\pi/2$.
848:
849: \begin{figure}[tb]
850: \centerline{\psfig{file=fig3.eps, width=25pc}}
851: \caption{\label{fig3}The real branches of $\xi \exp(\xi)=b\cos\eta$
852: for $b<0$.}
853: \end{figure}
854: Suppose that $b<0$. In this case, the real branches of
855: $\xi \exp(\xi)=b\cos\eta$
856: exist only if $\cos\eta<0$ as in Figure~\ref{fig3} and
857: a corresponding
858: pole in the open right half-plane does not exist unless $\abs{b}>\pi/2$.
859: Using the definition of $b$,
860: this condition is equivalent to the requirement
861: that $\beta_0\exp(-\beta_0)>\pi/2$. But, the maximum value
862: of $\beta_0\exp(-\beta_0)$ is $1/e<\pi/2$. Hence, negative values of
863: $b$ do not correspond to poles in the right half-plane.
864:
865: We conclude that the dynamic memory kernel $k$ for
866: stepwise uniform linear acceleration is unbounded for
867: $\beta_0=g_0\alpha>1.3$.
868:
869:
870: \subsection{Rotation}
871: \begin{figure}[tb]
872: \centerline{\psfig{file=fig4.eps, width=25pc}}
873: \caption{\label{fig4} Schematic plot of the motion of the observer
874: that undergoes stepwise uniform rotation of frequency $\Omega_0$ during
875: a period $\alpha=\tau_1-\tau_2$ of its proper time such that
876: $\vp=\gamma \Omega_0\alpha$. If $\vp$ exceeds $\pi/2$,
877: then the convolution kernel leads to divergences.}
878: \end{figure}
879: Imagine next an observer that is initially moving uniformly with speed $v$
880: in the $(x,y)$-plane along a line parallel to the $y$-axis at
881: $x=\rho_0$. At $t=0$,
882: $x=\rho_0$ and $y=0$, the observer starts rotating on a circle of
883: radius $\rho_0$ with uniform frequency $\Omega_0=v/\rho_0$ in the
884: positive sense around the
885: $z$-axis. Though the motion is continuous, there is no acceleration
886: for $t<0$ and uniform circular acceleration for $t>0$. The natural
887: orthonormal tetrad frame of
888: the uniformly rotating observer is given by
889: \begin{equation}\begin{split} \hat{\lambda}^\mu_{(0)}&=\gamma
890: (1,-v\sin \varphi ,v\cos \varphi ,0),\\
891: \hat{\lambda}^\mu _{(1)}&= (0,\cos \varphi ,\sin \varphi ,0),\\
892: \hat{\lambda}^\mu_{(2)}&= \gamma (v,-\sin \varphi ,\cos \varphi ,0),\\
893: \hat{\lambda}^\mu _{(3)}&= (0,0,0,1),\label{eq25}\end{split}\end{equation}
894: where $\gamma =(1-v^2)^{-\frac{1}{2}}$ is the Lorentz factor and
895: $\varphi =\Omega_0t=\gamma \Omega_0\tau$, so that we have set
896: $\tau_0=0$ in this case. Computing
897: $\phi_{\alpha \beta}$ for the tetrad frame \eqref{eq25}, we find as
898: expected that the translational acceleration has only a radial
899: component
900: $g_1=-v\gamma^2\Omega_0$ and the rotational frequency is along the
901: $z$-direction with magnitude $\Omega_3=\gamma^2\Omega_0$. Thus
902: $\hat{f}=\Lambda f$, where
903: $\Lambda \to [\Lambda_1;\Lambda_2]$ is given by
904: \begin{equation}\label{eq26} \Lambda_1 =\begin{bmatrix} \gamma \cos
905: \varphi & \gamma \sin \varphi & 0\\ -\sin \varphi & \cos \varphi &0\\
906: 0 & 0 &\gamma
907: \end{bmatrix},\quad \Lambda_2=v\gamma \begin{bmatrix} 0 & 0 &1\\ 0 &
908: 0 & 0\\ -\cos \varphi & -\sin \varphi & 0\end{bmatrix} .\end{equation}
909:
910: Let us first consider case~(1); the kinetic memory kernel
911: $k_0$ can be easily computed using the fact that for $\Lambda$
912: given by equation~\eqref{eq26} we have $\Lambda^{-1}\to
913: [\Lambda^T_1;\Lambda_2^T]$. Then
914: we find that $k_0\to
915: [$\boldmath$\Omega$\unboldmath$\cdot$\boldmath$I$\unboldmath$;-\mathbf{g}\cdot
916: \mathbf{I}]$, where
917: \boldmath$\Omega$\unboldmath$=(0,0,\gamma^2\Omega_0)$ and $\mathbf{g}
918: =(-v\gamma^2\Omega_0,0,0)$ with respect to the orthonormal tetrad
919: frame \eqref{eq25}. Thus
920: $k_0$ is a constant kernel so long as the observer rotates uniformly;
921: for instance, if the acceleration is turned off at $\tau_1=\alpha$
922: corresponding to
923: $\vp=\gamma \Omega _0\alpha$, then the observer will have
924: uniform linear motion again with speed $v$ for $\tau >\tau_1$ and the
925: kernel $k_0$ will vanish (see Figure~\ref{fig4}).
926:
927: Let us now consider case~(2); the dynamic memory kernel is
928: given by the series \eqref{eq10} in terms of the resolvent kernel.
929: This is given by
930: equation~\eqref{eq9}, $r\to [r_1;r_2]$, where
931: \begin{equation}\begin{split}\label{eq27} r_1&=\gamma
932: \Omega_0\begin{bmatrix} -\gamma^2 \sin \varphi & \gamma \cos \varphi
933: &0\\ -\gamma \cos \varphi & -\sin \varphi
934: & 0\\ 0& 0 & v^2\gamma^2\sin \varphi \end{bmatrix},\\ r_2&=
935: v\gamma^2\Omega_0\begin{bmatrix} 0 & 0 & \gamma \sin \varphi\\ 0 & 0
936: & \cos \varphi \\ \gamma \sin
937: \varphi & -\cos
938: \varphi & 0\end{bmatrix}.\end{split}\end{equation}
939: The explicit calculation of $k$ using the series \eqref{eq10} for the
940: general case of stepwise uniform rotation from $\tau =0$ to
941: $\tau_1=\alpha$ is rather
942: complicated; however, for $\tau_1\to \infty$ the calculation can be
943: carried through and the result is a constant kernel given by $k\to
944: [$\boldmath$\Omega$\unboldmath$\cdot \mathbf{I};-\mathbf{g}\cdot
945: \mathbf{I}]$. Just as in the case of uniform translational
946: acceleration (cf. Section 4), we have
947: $k_0=k$ for uniform rotation as well.
948:
949: To calculate $k$ for the stepwise uniform rotation of duration
950: $\tau_1-\tau_0=\alpha >0$, we use Laplace transforms as in the
951: previous section (see Figure~\ref{fig4}). Let
952: $C'=\alpha^{-1}\mathcal{L}\{\cos \varphi\}$ and $S'=\alpha
953: ^{-1}\mathcal{L} \{ \sin \varphi \}$; then, with $w=s\alpha$ we find
954: \begin{equation}\label{eq28}
955: C'\pm i S'=
956: \frac{1-e^{-(w\mp i\vp)}}{w\mp i\vp},
957: \end{equation}
958: and hence the Laplace transform of the resolvent kernel is given by
959: $\bar{r}\to [\bar{r}_1;\bar{r}_2]$, where
960: \begin{equation}\label{eq30} \bar{r}_1=\vp\begin{bmatrix}
961: -\gamma^2S' & \gamma C' & 0\\ -\gamma C' & -S' & 0\\ 0 & 0 &
962: v^2\gamma^2S'\end{bmatrix},\quad
963: \bar{r}_2=v\gamma \vp\begin{bmatrix} 0 & 0 & \gamma S'\\ 0 & 0
964: & C'\\ \gamma S' & -C' & 0\end{bmatrix}.\end{equation}
965: Using methods given in Appendix B, equation~\eqref{eq19} leads to
966: \[\bar{k}(s)\to [\bar{k}_1(s);\bar{k}_2(s)],\]
967: where
968: \begin{eqnarray}
969: \label{eq31}
970: \bar{k}_1(s)&=&\vp
971: \begin{bmatrix}
972: \gamma^2 \mathcal{Q} & \gamma \mathcal{P} & 0\\
973: -\gamma \mathcal{P} & \mathcal{Q} & 0\\
974: 0 & 0 & -v^2\gamma^2\mathcal{Q}
975: \end{bmatrix},\\
976: \label{eq31b}\bar{k}_2(s)&=&v\gamma \vp
977: \begin{bmatrix}
978: 0 & 0 & -\gamma \mathcal{Q}\\
979: 0 & 0 & \mathcal{P}\\
980: -\gamma \mathcal{Q} & -\mathcal{P} & 0
981: \end{bmatrix}.
982: \end{eqnarray}
983: Here $\mathcal{P}$ and $\mathcal{Q}$ are given by
984: \begin{align}\label{eq32} \mathcal{P} &=
985: \frac{e^w}{\mathcal{D}}[w(-e^w+\cos \vp)-\vp\sin
986: \vp],\\
987: \label{eq33} \mathcal{Q} &= \frac{1}{\mathcal{D}} [e^w(-w\sin
988: \vp+\vp\cos \vp)-\vp],\end{align}
989: and the denominator $\mathcal{D}$ is given by
990: \begin{equation}\label{eq34}
991: \mathcal{D}=(we^w-i\vp e^{i\vp})
992: (we^w+i\vp e^{-i \vp}).
993: \end{equation}
994: It is interesting to note that if we formally substitute $\beta_0$
995: for $i\vp$ in equations~\eqref{eq32}--\eqref{eq34}, we obtain
996: results familiar from the previous subsection;
997: specifically, under $i\vp\to \beta_0$, $\mathcal{P}\to P$,
998: $\mathcal{Q}\to iQ$ and $\mathcal{D}\to D$, where $P,Q$ and $D$ are
999: given in
1000: equations~\eqref{eq22}--\eqref{eq24}. Therefore, the main results of
1001: the previous subsection
1002: can also be used in the analysis of stepwise uniform rotation;
1003: for instance, with appropriate
1004: modifications the explicit
1005: expressions given in Appendix C for the convolution kernel in a
1006: special case can be employed here as well. However, since
1007: $\vp>0$, the singularities of
1008: $\mathcal{P}$ and $\mathcal{Q}$ are in general different from those
1009: in the previous subsection.
1010:
1011: To determine the pole singularities of $\bar k(s)$ in
1012: the right half-plane in the case of stepwise rotation
1013: it suffices to consider the equation
1014: \begin{equation}\label{eq:our}
1015: we^w= i\vp e^{i \vp}
1016: \end{equation}
1017: with $\vp>0$.
1018: Indeed, note that if $w$ is a solution of this equation,
1019: then the complex conjugate of $w$ is a solution of
1020: $w\exp(w)= -i\vp \exp(-i \vp)$.
1021:
1022:
1023: As before, let us set
1024: $w=\xi+i\eta$ and note that equation~\eqref{eq:our}
1025: is equivalent to the system
1026: of real equations given by
1027: \begin{equation}\label{eq:psign}
1028: \xi e^\xi=\vp\sin(\eta-\vp),\qquad \eta e^\xi=\vp\cos(\eta-\vp),
1029: \end{equation}
1030: where $\xi\ge 0$ and $\vp>0$.
1031: We recall here that the solution $w=i\vp$, i.e. $\xi=0$
1032: and $\eta=\vp$, of equations~\eqref{eq:our} and~\eqref{eq:psign}
1033: corresponds to a removable singularity.
1034: A necessary condition for system~\eqref{eq:psign}
1035: to have a solution with $\xi>0$ is that $\sin(\eta-\vp)>0$ and
1036: $\eta\cos(\eta-\vp)>0$; the latter condition means that
1037: $\cos(\eta-\vp)$ and $\eta$ must have the
1038: same sign.
1039:
1040: Consider the system
1041: \begin{equation}\label{eq:nsys}
1042: (\xi^2+\eta^2) e^{2 \xi}=\vp^2,\qquad \xi e^\xi=\vp\sin(\eta-\vp).
1043: \end{equation}
1044: If it has a solution $(\xi,\eta)$,
1045: then it follows from display~\eqref{eq:nsys} that
1046: \[
1047: \eta^2 e^{2\xi}=\vp^2\cos^2(\eta-\vp),
1048: \]
1049: and therefore $\eta \exp(\xi)=\pm \vp\cos(\eta-\vp)$.
1050: Comparing this result with system~\eqref{eq:psign}, we conclude
1051: that we can use system~\eqref{eq:nsys} for finding the poles if we keep
1052: in mind that $\eta$ and
1053: $\cos(\eta-\vp)$ must have the same sign.
1054:
1055: \begin{figure}[tb]
1056: \centerline{\psfig{file=fig5.eps, width=25pc}}
1057: \caption{\label{fig5} The graph of $\eta^2=\vp^2\exp(-2\xi)-\xi^2$.}
1058: \end{figure}
1059: The first equation in display~\eqref{eq:nsys} is equivalent to
1060: $\eta^2=\vp^2\exp(-2\xi)-\xi^2$. Its graph
1061: in the right half-plane has the form depicted in Figure~\ref{fig5},
1062: where $\xi_0$ is the unique real solution of the equation
1063: $\xi \exp(\xi)=\vp$.
1064:
1065: The poles we seek correspond to the intersections of the graph in
1066: Figure~\ref{fig5}
1067: with the real branches of the second curve in display~\eqref{eq:nsys}.
1068: The intercepts of these branches with the $\eta$-axis are given
1069: by the solutions of the equation $\sin(\eta-\vp)=0$; that is,
1070: $\eta$ is equal to $\vp$ plus an integer multiple of $\pi$. Along the
1071: line given by $\xi=\xi_0$, the intercepts are given by
1072: $\xi_0 \exp(\xi_0)=\vp \sin(\eta-\vp)$. Because $\xi_0 \exp(\xi_0)=\vp$,
1073: these intercepts
1074: are the solutions of $\sin(\eta-\vp)=1$; that is,
1075: $\eta$ is $\vp+\pi/2$ plus an integer multiple of $2\pi$.
1076: The shape of the branches connecting points on the two vertical lines
1077: (at $\xi=0$ and $\xi=\xi_0$) is determined by the sign of $\cos(\eta-\vp)$ along the branch.
1078: Indeed, we have already established that poles occur only
1079: at points where $\eta$ and $\cos(\eta-\vp)$ have the same sign.
1080: Note that
1081: \[
1082: (\xi+1)e^\xi=\vp\frac{d\eta}{d\xi}\cos(\eta-\vp),
1083: \]
1084: and therefore the slope of the branch has the same sign as $\cos(\eta-\vp)$.
1085: Moreover, only the branches with $\sin(\eta-\vp)>0$ correspond
1086: to poles in the right half-plane.
1087:
1088: \begin{figure}[tb]
1089: \centerline{\psfig{file=fig6.eps, width=25pc}}
1090: \caption{\label{fig6} The graph of $\xi \exp(\xi)=\vp \sin(\eta-\vp)$
1091: for $0<\vp<\pi/2$.}
1092: \end{figure}
1093: There are several cases depending on the size of $\vp$.
1094: For $0<\vp<\pi/2$, it is easy to see that the important branches
1095: are as depicted in Figure~\ref{fig6}. These would not
1096: intersect the graph in Figure~\ref{fig5}; hence, there are no poles in
1097: the right half-plane.
1098:
1099:
1100: We will next show that if $\vp>\pi/2$, then there is at least one pole in
1101: the right half-plane. For $\vp$ in this range, there is
1102: an integer $j\ge 1$ such that $j\pi/2\le \vp< (j+1)\pi/2$. In particular,
1103: we have that
1104: $\vp-j\pi/2\ge 0$ and $\vp-(j+1)\pi/2<0$.
1105: There are four cases.
1106: (1) Suppose that $j$ is even and $\cos (j\pi/2)=1$. The branch of the
1107: curve $\xi \exp(\xi)=\vp \sin(\eta-\vp)$ with $\eta$-intercept
1108: $\vp-j\pi/2\ge 0$ has positive slope (like the upper
1109: branch in Figure~\ref{fig6}). Because $\vp-j\pi/2<\vp$,
1110: this branch intersects the curve depicted in Figure~\ref{fig5}
1111: in the upper half-plane. This point corresponds to a pole. Indeed,
1112: at the point of intersection $\sin(\eta-\vp)>0$ and
1113: $\eta\cos(\eta-\vp)>0$.
1114: (2) Suppose that $j$ is even and $\cos (j\pi/2)=-1$.
1115: The branch with $\eta$-intercept $\vp-j\pi/2\ge 0$ has negative
1116: slope and meets the line $\xi=\xi_0$ with ordinate $\vp-(j+1)\pi/2<0$.
1117: Hence, this branch intersects the curve depicted in
1118: Figure~\ref{fig5}
1119: in the lower half-plane. This point corresponds to a pole.
1120: (3) Suppose that $j$ is odd and $\cos ((j+1)\pi/2)=1$. The
1121: branch of the curve with $\eta$-intercept $\vp-(j+1)\pi/2$ has
1122: positive slope and it meets the curve depicted in Figure~\ref{fig5}
1123: in the upper half-plane where the intersection point corresponds to a pole.
1124: For the subcase where $j=3$ and $\vp=3\pi/2$,
1125: it is interesting to note that
1126: $\eta=0$ and $\xi_0$, such that
1127: $\xi_0 \exp(\xi_0)=3\pi/2$, is the pole.
1128: (4) Suppose that $j$ is odd and $\cos ((j+1)\pi/2)=-1$.
1129: The curve with $\eta$-intercept $\vp-(j+1)\pi/2$ has
1130: negative slope and $-\vp\le \vp-(j+1)\pi/2$. Hence, this branch meets
1131: the curve depicted in Figure~\ref{fig5} in the lower half-plane where
1132: the intersection corresponds to a pole.
1133:
1134: We conclude that the dynamic memory kernel $k$ for stepwise uniform
1135: rotation is unbounded for $\vp=\gamma\Omega_0\alpha>\pi/2$.
1136: \subsection{Smooth acceleration}
1137: We have demonstrated
1138: that the convolution kernel $k$ is unbounded for certain
1139: stepwise translational and rotational accelerations.
1140: Could this result be due to the discontinuities of these accelerations
1141: at $\tau_0$ and $\tau_1$? To prove that this is \emph{not} the case,
1142: we are interested here instead in smooth accelerations that
1143: closely approximate the stepwise ones already studied. The
1144: translational and rotational cases are in fact closely related as
1145: we have demonstrated; therefore
1146: in this subsection we show the same result for the simpler case of
1147: \emph{smooth} translational acceleration.
1148:
1149: Let us consider an acceleration $g$ with compact
1150: support in the interval $[\tau_0,\tau_1]$.
1151: By the definition of $\Lambda$ and the choice of $g$,
1152: the matrix $\Lambda(\tau_0)=\Lambda_0$ is the $6\times 6$ identity
1153: matrix. Using this fact and
1154: equations~\eqref{eq9} and~\eqref{eq14}, we find that
1155: $ r(t)=[g(\tau)S(\tau) J_3;g(\tau)C(\tau) I_3]$,
1156: where $t=\tau-\tau_0$, $S(\tau)=\sinh\theta$, $C(\tau)=\cosh\theta$
1157: and $\theta(\tau)$ is given by equation~\eqref{eq12}.
1158: It follows from equation~\eqref{eq18} that
1159: $ \bar r(s)=[\calS(s) J_3;\calC(s) I_3]$,
1160: where
1161: \[
1162: \calC(s)\pm\calS(s)
1163: =\int_0^\infty e^{-st}g(t+\tau_0)e^{\pm\theta(t+\tau_0)}\,dt.
1164: \]
1165: Using equation~\eqref{eq19} and the results of Appendix B, we find that
1166: the Laplace
1167: transform of the convolution kernel $k$ is given by
1168: $\bar k(s)=[\calH_1(s)J_3;\calH_2(s)I_3]$, where
1169: \[
1170: \calH_1(s)=\frac{1+\calS}{(1+\calS)^2-\calC^2}-1,\qquad
1171: \calH_2(s)=- \frac{\calC}{(1+\calS)^2-\calC^2}.
1172: \]
1173:
1174: We are interested in the zeros of the denominator
1175: \[(1+\calS)^2-\calC^2=(1+\calS+\calC)(1+\calS-\calC).\]
1176: It suffices to demonstrate that $1+\calS+\calC$ has a zero in
1177: the right half of the complex $s$-plane.
1178: Because $g$ has compact support in the interval $[\tau_0,\tau_1]$,
1179: the function $1+\calS+\calC$ is given by
1180: \[
1181: s\mapsto 1+\int_0^\alpha
1182: e^{-s t} e^{\int_0^t g(\sigma+\tau_0)\,d\sigma} g(t+\tau_0)\,dt,
1183: \]
1184: where $\alpha=\tau_1-\tau_0$.
1185: If $g$ is the stepwise uniform acceleration
1186: considered previously, then this function reduces to
1187: \[ s\mapsto \frac{e^{-\alpha s}}{s-g_0}
1188: ( se^{\alpha s}- g_0 e^{\alpha g_0});
1189: \]
1190: and, by the results in Subsection~\ref{sec:sa} for equation~\eqref{eq24},
1191: if $\beta_0 \exp(\beta_0)>3\pi/2$, it
1192: has a zero in the right half of the complex $s$-plane corresponding
1193: to a pole of $\bar k$.
1194: We will show that such a pole persists for a smooth acceleration
1195: that is sufficiently close to the stepwise acceleration.
1196:
1197: For an arbitrary acceleration $g$ with support in
1198: the interval $[\tau_0,\tau_1]$, we define
1199: the associated real-valued function $\zeta$ on
1200: the interval $[0,\alpha]$ given by $\zeta(t)=g(t+\tau_0)$.
1201: Also, recall that the $L^1$-norm of a real-valued function $\upsilon$
1202: defined on the interval
1203: $[0,\alpha]$ is given by
1204: \[\norm{\upsilon}_1:=\int_0^\alpha\abs{\upsilon(t)}\,dt.\]
1205:
1206: Suppose that $\zeta$ and $\upsilon$ are real-valued functions
1207: defined on the interval $[0,\alpha]$ such that $\norm{\zeta}<\infty$ and
1208: $\norm{\upsilon}_1<\infty$, and consider the
1209: complex-valued analytic functions $Z$ and $\Upsilon$
1210: of the complex variable $s$ given by
1211: \begin{eqnarray*}
1212: Z(s)&=&
1213: 1+\int_0^\alpha
1214: e^{-s t} e^{\int_0^t \zeta(\sigma)\,d\sigma} \zeta(t)\,dt,\\
1215: \Upsilon(s)&=&
1216: 1+\int_0^\alpha
1217: e^{-s t} e^{\int_0^t \upsilon(\sigma)\,d\sigma} \upsilon(t)\,dt.
1218: \end{eqnarray*}
1219: We will prove the following proposition.
1220: \emph{If $Z$ has a zero in the open right-half of the complex
1221: $s$-plane
1222: and $\norm{\upsilon-\zeta}_1$ is sufficiently small,
1223: then $\Upsilon$ has a zero in the open right-half of
1224: the complex $s$-plane.}
1225: By a standard result from mathematical analysis (see~\cite{ll}),
1226: if $\zeta$
1227: is an $L^1$ function
1228: (for example, if $\zeta(t)=g(t+\tau_0)$ for
1229: the stepwise acceleration $g$), then $\norm{\zeta-\upsilon}_1$
1230: can be made as small as desired for
1231: a $C^\infty$ function $\upsilon$. Hence, by the proposition,
1232: there is a smooth acceleration with compact support such that
1233: its associated
1234: convolution kernel is unbounded.
1235:
1236: Our proof begins with two estimates.
1237: For notational convenience, let us define
1238: \[
1239: \hat{\zeta}(t)=e^{\int_0^t \zeta(\sigma)\,d\sigma} \zeta(t),\qquad
1240: \hat{\upsilon}(t)=e^{\int_0^t \upsilon(\sigma)\,d\sigma} \upsilon(t).
1241: \]
1242: The first estimate is
1243: \begin{equation}\label{est:1}
1244: \abs{\Upsilon(s)-Z(s)}\le \norm{\hat\upsilon-\hat\zeta}_1
1245: \end{equation}
1246: for all $s$ such that $\re (s)\ge 0$.
1247: To prove it, note that
1248: \[
1249: \abs{\Upsilon(s)-Z(s)}\le
1250: \int_0^\alpha \abs{e^{-s t}}\abs{\hat\upsilon(t)-\hat\zeta(t)}\,dt.
1251: \]
1252: Because $\abs{\exp(-s t)}\le 1$ for $\re(s)\ge 0$, we have the
1253: inequality
1254: \[\abs{\Upsilon(s)-Z(s)}\le \norm{\hat\upsilon-\hat\zeta}_1\]
1255: for all $s$ in the closed right half-plane.
1256: The second estimate is
1257: \begin{equation}\label{est:2}
1258: \norm{\hat\upsilon-\hat\zeta}_1\le
1259: e^{\norm{\zeta}_1}(1+\alpha\norm{\zeta}) e^{\norm{\upsilon-\zeta}_1}
1260: \norm{\upsilon-\zeta}_1.
1261: \end{equation}
1262: To prove it, we have the triangle law estimate
1263: \begin{eqnarray}\label{est:tl}
1264: \nonumber\abs{\hat\upsilon(t)-\hat\zeta(t)}&\le&
1265: \abs {e^{\int_0^t \upsilon(\sigma)\,d\sigma}\upsilon(t)
1266: -e^{\int_0^t \upsilon(\sigma)\,d\sigma}\zeta(t)}\\
1267: \nonumber &&{}+\abs {e^{\int_0^t \upsilon(\sigma)\,d\sigma}\zeta(t)
1268: -e^{\int_0^t \zeta(\sigma)\,d\sigma}\zeta(t)}\\
1269: &\le & e^{\int_0^t \abs{\upsilon(\sigma)}\,d\sigma}\abs{\upsilon-\zeta}
1270: +\abs{\zeta}
1271: \abs{e^{\int_0^t \upsilon(\sigma)\,d\sigma}-e^{\int_0^t \zeta(\sigma)\,d\sigma}}
1272: \end{eqnarray}
1273: and, by the mean value theorem (applied to the exponential function),
1274: the inequality
1275: \[
1276: \abs{e^{\int_0^t \upsilon(\sigma)\,d\sigma}-e^{\int_0^t \zeta(\sigma)\,d\sigma}}
1277: \le e^\varsigma
1278: \big|\int_0^t \upsilon(\sigma)\,d\sigma-\int_0^t \zeta(\sigma)\,d\sigma\big|,
1279: \]
1280: where $\varsigma$ is some number between
1281: $\int_0^t \upsilon(\sigma)\,d\sigma$ and $\int_0^t \zeta(\sigma)\,d\sigma$.
1282: If $\varsigma\le 0$, then $\exp(\varsigma)<1$; and if
1283: $\varsigma>0$, then $\varsigma<\max\{\norm{\upsilon}_1,\norm{\zeta}_1\}$.
1284: Hence,
1285: \[e^\varsigma\le e^{\max\{\norm{\upsilon}_1,\norm{\zeta}_1\}}\]
1286: and, because
1287: $\norm{\upsilon}_1\le \norm{\upsilon-\zeta}_1+ \norm{\zeta}_1$,
1288: we have that
1289: \[e^\varsigma\le e^{\norm{\zeta}_1}e^{\norm{\upsilon-\zeta}_1}.\]
1290: Using this result and the estimate~\eqref{est:tl}, it follows that
1291: \begin{eqnarray*}
1292: \abs{\hat\upsilon(t)-\hat\zeta(t)}
1293: &\le&
1294: e^{\norm{\upsilon}_1}\abs{\upsilon(t)-\zeta(t)}
1295: +\norm{\zeta}
1296: e^{\norm{\zeta}_1}e^{\norm{\upsilon-\zeta}_1}
1297: \int_0^\alpha\abs{\upsilon(\sigma)-\zeta(\sigma)}\,d\sigma\\
1298: &\le& e^{\norm{\zeta}_1}e^{\norm{\upsilon-\zeta}_1}
1299: (\abs{\upsilon(t)-\zeta(t)}+\norm{\zeta}\norm{\upsilon-\zeta}_1).
1300: \end{eqnarray*}
1301: Therefore,
1302: \begin{eqnarray*}
1303: \norm{\hat\upsilon-\hat\zeta}_1
1304: &=&\int_0^\alpha\abs{\hat\upsilon-\hat\zeta}\,dt\\
1305: &\le& e^{\norm{\zeta}_1}e^{\norm{\upsilon-\zeta}_1}
1306: (\norm{\upsilon-\zeta}_1+\alpha\norm{\zeta}\norm{\upsilon-\zeta}_1)
1307: \end{eqnarray*}
1308: and
1309: a rearrangement of the right-hand side of the last inequality
1310: gives the desired result.
1311:
1312: In the rest of this section, we let $\zeta$ represent the
1313: stepwise uniform linear acceleration and $\upsilon$ the
1314: smooth linear acceleration that approximates it sufficiently closely.
1315: We then
1316: choose a circle centered at a zero of $Z$ in the open right half
1317: of the complex $s$-plane such that the circle does not pass through
1318: a zero of $Z$ and such that the circle is contained in
1319: the open right half-plane. Let $\kappa$, a complex-valued function
1320: defined on the interval $[0,2\pi]$, be a continuous
1321: parametrization of this circle and define two new functions
1322: $\kappa_Z$ and $\kappa_\Upsilon$ on this interval by
1323: \[\kappa_Z(\vartheta)=Z(\kappa(\vartheta)),\qquad
1324: \kappa_\Upsilon(\vartheta)=\Upsilon(\kappa(\vartheta)).
1325: \]
1326: The images of these functions are closed curves in the complex
1327: $s$-plane.
1328: In complex analysis, the principle of the argument theorem~\cite{hen}
1329: for an analytic function $\Delta$ relates the winding number of the
1330: image of $\kappa_\Delta$ with respect to the origin to
1331: the number of zeros of the function $\Delta$ inside the circle,
1332: provided that the circle does not pass through any zero of $\Delta$.
1333: If we show that $\kappa_Z$ and $\kappa_\Upsilon$ are homotopic and therefore
1334: have the same winding number with respect to the origin and that the circle
1335: does not pass through a zero of $\Upsilon$, then $Z$ and $\Upsilon$ must
1336: have the same number of zeros inside the circle.
1337:
1338: We claim that if $\norm{\upsilon-\zeta}_1$ is
1339: sufficiently small, then the image of $\kappa$ does not pass through a zero
1340: of $\Upsilon$. To prove the claim, note that
1341: \[ m:= \min\{\abs{\kappa_Z(\vartheta)}:0\le \vartheta\le 2\pi\}>0\]
1342: (because $\kappa$ does not pass through a zero of $Z$)
1343: and using the triangle inequality
1344: \[
1345: 0<m\le \abs{\kappa_\Upsilon(\vartheta)}+\norm{\kappa_\Upsilon-\kappa_Z},
1346: \]
1347: where $\norm{\kappa_\Upsilon-\kappa_Z}$ is the supremum
1348: of $\abs{\kappa_\Upsilon(\vartheta)-\kappa_Z(\vartheta)}$ for
1349: $0\le \vartheta\le 2 \pi$.
1350: Using the estimates~\eqref{est:1} and~\eqref{est:2},
1351: we have that
1352: \begin{equation}\label{est:kk}
1353: \abs{\kappa_\Upsilon(\vartheta)-\kappa_Z(\vartheta)}
1354: \le e^{\norm{\zeta}_1}e^{\norm{\upsilon-\zeta}_1}
1355: (1+\alpha\norm{\zeta})\norm{\upsilon-\zeta}_1.
1356: \end{equation}
1357: By the estimate~\eqref{est:kk}, $\norm{\kappa_\Upsilon-\kappa_Z}$
1358: can be made small, say less than
1359: $m/2$,
1360: by taking $\norm{\upsilon-\zeta}_1$ sufficiently small.
1361: For all
1362: $\upsilon$ satisfying this requirement, which we impose for
1363: the remainder of the proof, we have that
1364: $\abs{\kappa_\Upsilon(\vartheta)}>0$; that is, $\kappa$ does not pass through a zero
1365: of $\Upsilon$.
1366:
1367: It remains to show that
1368: $\kappa_\Upsilon$ is homotopic to $\kappa_Z$. Assuming this
1369: homotopy relation, the image
1370: curves of $\kappa_\Upsilon$ and $\kappa_Z$ would
1371: have the same winding number with respect to the origin.
1372: By the choice of $\kappa$ and the argument principle (see~\cite{hen}),
1373: the curve
1374: $\kappa_Z$ has a nonzero winding number. Hence,
1375: $\kappa_\Upsilon$ would have the same nonzero winding number. Again,
1376: by the argument principle, $\Upsilon$ must then have a zero in
1377: the disk bounded by the circle parametrized by $\kappa$, which
1378: is the desired result.
1379:
1380: To complete the proof we need to show that
1381: $\kappa_\Upsilon$ and $\kappa_Z$
1382: are indeed homotopic. Let $\bbC$ denote the complex numbers.
1383: We will show that
1384: $H:[0,1]\times[0,2\pi]\to \bbC\setminus\{0\}$ given by
1385: \[
1386: H(\sigma, \vartheta)=\kappa_Z(\vartheta)-
1387: \sigma (\kappa_Z(\vartheta)-\kappa_\Upsilon(\vartheta))
1388: \]
1389: is the required homotopy. By inspection, $H$ is continuous,
1390: $H(0,\vartheta)=\kappa_Z(\vartheta)$ and
1391: $H(1,\vartheta)=\kappa_\Upsilon(\vartheta)$. Hence, it suffices
1392: to show that $H(\sigma,\vartheta)\ne 0$ for all
1393: $(\sigma,\vartheta)\in [0,1]\times[0,2\pi]$.
1394: By our choice of $\upsilon$, we have that
1395: $\norm{\kappa_\Upsilon-\kappa_Z}<m/2$; therefore
1396: \[
1397: \abs{H(\sigma, \vartheta)}
1398: \ge \abs{\kappa_Z(\vartheta)}-
1399: \abs{\sigma}\abs{\kappa_\Upsilon(\vartheta)-\kappa_Z(\vartheta)}
1400: \ge m-\norm{\kappa_\Upsilon-\kappa_Z}> m/2,
1401: \]
1402: as required.
1403:
1404: We conclude that the dynamic memory kernel $k$ for the smooth
1405: linear acceleration that closely approximates the stepwise
1406: acceleration is unbounded if the area under the graph of $g(\tau)$
1407: exceeds a critical value $\sim 1$.
1408:
1409:
1410: \section{Discussion}
1411:
1412: We have investigated the properties of the nonlocal kernel that is
1413: induced by accelerated motion in Minkowski spacetime. The physical
1414: principles outlined in this
1415: paper do not completely determine the kernel; therefore, simplifying
1416: mathematical assumptions need to be introduced in order to identify a
1417: unique kernel. Two
1418: possibilities have been explored in this work corresponding to
1419: kinetic memory $(k_0)$ and dynamic memory $(k)$. We show that for
1420: accelerated motion that is uniform
1421: (linear or circular), the two kernels give the same constant result
1422: $k_0=k$. They differ, however, if the acceleration is turned off at a
1423: certain moment. We have
1424: therefore studied piecewise uniform acceleration (linear and circular)
1425: and have demonstrated that the dynamic memory (convolution) kernel
1426: could be divergent and is
1427: therefore ruled out. Furthermore,
1428: this conclusion is shown to be
1429: independent of the stepwise character of the linear acceleration considered.
1430:
1431: The use of convolution kernels is standard practice in the nonlocal
1432: electrodynamics of continuous media, where it is assumed phenomenologically
1433: that memory always fades. In our treatment of acceleration-induced
1434: nonlocality in vacuum, however, the behavior of memory must be determined
1435: from first principles. In this connection, the possible advantage of kinetic
1436: memory in terms of its simplicity was first emphasized by Hehl and
1437: Obukhov~\cite{ho,hoa}.
1438:
1439: The theory developed here is applicable to any basic field; however,
1440: for the sake of concreteness and in view of possible observational
1441: consequences, we employ
1442: electromagnetic radiation fields throughout. A basic consequence of
1443: the nonlocal theory of accelerated systems is that it is incompatible
1444: with the existence of a
1445: basic scalar field; that is, in this case $\Lambda (\tau )=1$, $k_0=0$
1446: and the nonlocality disappears so that a basic scalar radiation field
1447: can stay completely at rest with respect to a rotating observer
1448: in contradiction with our fundamental
1449: physical assumption. This prediction of the nonlocal theory is in
1450: agreement with present
1451: experimental data. Further confrontation of the nonlocal theory with
1452: observation is urgently needed.
1453:
1454: \appendix
1455:
1456: \section*{Appendix A}
1457:
1458: Consider an integral equation of the form
1459: \begin{equation}\label{A1} \phi (x)=\psi (x)+\epsilon
1460: \int^x_a K(x,y)\phi (y)\,dy,\tag{A1}\end{equation}
1461: where $\psi$ is a continuous function, the kernel
1462: $K$ is continuous and $\epsilon$ is a constant parameter.
1463: There is a unique continuous resolvent
1464: kernel $R$ such that
1465: \begin{equation}\label{A2}\tag{A2} \psi (x)=\phi (x)+\epsilon
1466: \int^x_aR(x,y)\psi (y)\,dy.\end{equation}
1467: In turn, $K$ can be
1468: thought of as the resolvent kernel for $R$; this follows from the
1469: complete reciprocity between $K$
1470: and $R$.
1471:
1472: The proof of the existence and uniqueness of the resolvent
1473: kernel is by successive approximation. In fact, the solution $\phi$
1474: can be obtained as the uniform limit of the sequence of
1475: continuous functions
1476: $\{\phi_n\}_{n=0}^\infty$ defined as follows:
1477: $\phi_0 (x)=\psi(x)$ and
1478: \begin{equation}\label{A3}\tag{A3} \phi_{n+1}(x)=\psi (x)+\epsilon
1479: \int^x_a K(x,y)\phi_n(y)\,dy.\end{equation}
1480: Thus
1481: \begin{align}\label{A4}\tag{A4}
1482: \phi_1(x)=&\; \psi (x)+\epsilon \int^x_aK(x,y)\psi (y)\, dy,\\
1483: \nonumber \phi_2 (x)=&\;\psi (x)+\epsilon
1484: \int^x_aK(x,y)[\psi (y)+\epsilon \int^y_a K(y,z)\psi (z)\,dz]\,dy\\
1485: \label{A5}
1486: \tag{A5}
1487: =&\;\phi_1(x)+\epsilon^2\int^x_aK(x,y)\int^y_a K(y,z)\psi (z)\, dz dy.
1488: \end{align}
1489: The integration in \eqref{A5} is over a triangular domain in the
1490: $(y,z)$-plane defined by the vertices $(a,a)$, $(x,a)$ and $(x,x)$.
1491: Changing the order of the
1492: integration in \eqref{A5} results in the equality
1493: \begin{equation*}\label{A6}\tag{A6}\begin{split}
1494: &\int^x_aK(x,y)\left[ \int^y_aK(y,z)\psi (z) dz\right]\, dy\\
1495: &\quad =\int^x_a\left[ \int^x_z K (x,y) K(y,z)\,dy\right] \psi
1496: (z)\,dz.\end{split}\end{equation*} Let us define the successive
1497: iterated kernels of $K$ by $K_1(x,z)=K(x,z)$ and
1498: \begin{equation}\label{A7}\tag{A7} K_{n+1} (x,z) =\int^x_z K(x,y) K_n
1499: (y,z)\, dy.\end{equation}
1500: Then we can write \eqref{A5} as
1501: \begin{equation}\label{A8}\tag{A8} \phi_2
1502: (x)=\phi_1(x)+\epsilon^2\int^x_a K_2 (x,z) \psi (z)\,dz,\end{equation}
1503: and similarly
1504: \begin{equation}\label{A9}\tag{A9} \phi_3 (x)=\phi_2 (x) +\epsilon^3
1505: \int^x_a K_3 (x,z)\psi (z)\,dz,\end{equation}
1506: etc., such that in general
1507: \begin{equation}\label{A10}\tag{A10} \phi_m (x)=\phi_{m-1}
1508: (x)+\epsilon^m\int^x_a K_m(x,z)\psi (z)\,dz.\end{equation}
1509:
1510: Iterating \eqref{A10} for $m=1,2,3,\ldots ,n$ and summing the
1511: equations results in
1512: \begin{equation}\label{A11}\tag{A11} \phi_n(x)=\psi (x)+\int^x_a
1513: \left[ \sum^n_{m=1}\epsilon^mK_m(x,z)\right] \psi (z)\,dz,\end{equation}
1514: which can be rewritten as
1515: \begin{equation}\label{A12}\tag{A12} \psi (x)=\phi_n(x)+\epsilon
1516: \int^x_a\left[ -\sum^n_{m=1}\epsilon^{m-1}K_m(x,y)\right] \psi
1517: (y)\,dy.\end{equation}
1518: It can be shown that the uniform limit as $n\to \infty$ exists
1519: (see~\cite{5,5a,5b}).
1520: Thus,
1521: we obtain equation~\eqref{A2} with
1522: \begin{equation}\label{A13}\tag{A13}
1523: R(x,y)=-\sum^\infty_{n=1}\epsilon^{n-1} K_n(x,y).
1524: \end{equation}
1525:
1526: In case (1), $K(x,y)=k_0(y)$, the iterated kernels $K_n$
1527: for $n>1$ and the resolvent kernel $R$ are in general functions of
1528: both $x$ and $y$.
1529:
1530: In case (2),
1531: $K(x,y)=k(x-y)$, i.e.\ the kernel is of the convolution (Faltung)
1532: type, it follows from \eqref{A7} that
1533: \begin{equation}\label{A14}\tag{A14}
1534: k_{n+1}(t)=\int^t_0k(u)k_n(t-u)\,du,\end{equation}
1535: where $x-z=t$ and $x-y=u$; therefore, all of the iterated kernels are
1536: of the convolution type and can be obtained by successive
1537: convolutions of $k$ with itself. More precisely,
1538: let a star denote the Faltung operation,
1539: \begin{equation}\label{A15}\tag{A15} \phi \ast \chi (t)
1540: =\int^t_0\phi (t)\chi (t-u)\,du=\chi \ast \phi (t),\end{equation}
1541: and write $\phi^{\ast\, 2} = \phi \ast \phi$, etc. Then, the resolvent
1542: kernel \eqref{A13} can be expressed as $R(x,y)=r(x-y)$, where
1543: \begin{equation}\label{A16}\tag{A16}
1544: r(t)=-\sum^\infty_{n=1} \epsilon^{n-1}k^{\ast\, n}(t).\end{equation}
1545:
1546: \section*{Appendix B}
1547:
1548: In this paper, we deal with $6\times 6$ matrices of the form
1549: \begin{equation}\label{B1}\tag{B1} \mathcal{M}=\begin{bmatrix} A &
1550: B\\ -B & A\end{bmatrix},\end{equation}
1551: where $\det A\neq 0$ and $\det B=0$. The inverse of the matrix
1552: $\mathcal{M}$ is given by
1553: \begin{equation}\label{B2}\tag{B2} \mathcal{M}^{-1}=\begin{bmatrix} G
1554: & H\\ -H & G\end{bmatrix},\end{equation}
1555: where
1556: \begin{equation}\label{B3}\tag{B3} G=(A+BA^{-1}B)^{-1},\quad
1557: H=-GBA^{-1}=-A^{-1}BG.\end{equation}
1558:
1559: \section*{Appendix C}
1560:
1561: Let us rewrite $P(s)$ and $Q(s)$ given by equations~\eqref{eq22} and
1562: \eqref{eq23} in the form
1563: \begin{align*}\label{C1}\tag{C1}
1564: 2\beta_0\, P(s) & =
1565: \frac{1-\frac{\beta_0}{w}}{1-\zeta_+}-\frac{1+\frac{\beta_0}{w}}{1+\zeta_-},\\
1566: \label{C2}\tag{C2} 2\beta_0\, Q(s)
1567: &=-2+\frac{1-\frac{\beta_0}{w}}{1-\zeta_+}
1568: +\frac{1+\frac{\beta_0}{w}}{1+\zeta_-},\end{align*}
1569: where $w=s\alpha$, $\beta_0=g_0\alpha$ and $\zeta_\pm$ are given by
1570: \begin{equation}\label{C3}\tag{C3} \zeta_\pm =\frac{\beta_0}{w}\exp
1571: (-w\pm \beta_0).\end{equation}
1572: If we assume that $\re (s)>g_0$, then $|\zeta_\pm |<1$. We can
1573: therefore expand $(1\mp \zeta_\pm)^{-1}$ in powers of $\zeta_\pm$ and
1574: use the relation
1575: \begin{equation}\label{C4}\tag{C4} \mathcal{L} \left\{
1576: u_{n\alpha}(t)\frac{(t-n\alpha)^{\ell-1}}{(\ell-1)!}\right\}
1577: =\frac{e^{-n\alpha s}}{s^\ell}\end{equation}
1578: for integers $n\ge 0$ and $\ell >1$ to find $k(t)\to [k_1(t);k_2(t)]$.
1579: Here
1580: we use unit step functions such that $u_{n\alpha}(t)=u_0(t-n\alpha )$
1581: and $u_0(t)$ is the
1582: standard unit step function, i.e. $u_0(t)=1$ for $t\geq 0$ and
1583: $u_0(t)=0$ for $t<0$.
1584:
1585: We find that $k_1=\tilde k_1J_3$, $k_2=\tilde k_2I_3$ and
1586: \begin{align*}\label{C5}\tag{C5}
1587: g^{-1}_0\tilde k_1(t)&=\mathcal{S}_1u_\alpha (t)+\mathcal{C}_2
1588: u_{2\alpha}(t)+\mathcal{S}_3 u_{3\alpha}
1589: (t)+\mathcal{C}_4u_{4\alpha}(t)+\cdots ,\\
1590: \label{C6}\tag{C6} g^{-1}_0\tilde k_2(t)+u_0(t)&=\mathcal{C}_1u_\alpha
1591: (t)+\mathcal{S}_2 u_{2\alpha}(t) +\mathcal{C}_3 u_{3\alpha}
1592: (t)+\mathcal{S}_4 u_{4\alpha}(t)+\cdots
1593: ,\end{align*}
1594: where
1595: \begin{equation}\label{C7}\tag{C7} \mathcal{C}_n\pm
1596: \mathcal{S}_n=e^{\pm n\beta_0}\left[
1597: \frac{(g_0t-n\beta_0)^{n-1}}{(n-1)!}\mp
1598: \frac{(g_0t-n\beta_0)^n}{n!}\right]
1599: .\end{equation}
1600: Note that for any fixed value of $t$, only a finite number of terms
1601: contribute to the kernel $k(t)$.
1602:
1603: \subsection*{Acknowledgements}
1604:
1605: One of us (B.M.) is very grateful to Friedrich Hehl and Yuri Obukhov
1606: for many stimulating discussions regarding the nature of memory in
1607: nonlocal electrodynamics.
1608:
1609: \begin{thebibliography}{xxxx}
1610: \bibitem{1} H.A. Lorentz, The Theory of Electrons (Dover, New York,
1611: 1952), ch. V, \S 183, pp. 215--217.
1612:
1613: \bibitem{2} A. Einstein, The Meaning of Relativity (Princeton
1614: University Press, Princeton, 1950), p. 60.
1615:
1616: \bibitem{3} B. Mashhoon, Phys. Rev. A 47 (1993) 4498.
1617: \bibitem{3a} B. Mashhoon, ``Nonlocal
1618: Electrodynamics", in: Cosmology and Gravitation, edited by M. Novello
1619: (Editions Fronti\`eres,
1620: Gif-sur-Yvette, 1994), pp. 245--295.
1621: \bibitem{3b} U. Muench, F.W. Hehl and B.
1622: Mashhoon, Phys. Lett. A 271 (2000) 8.
1623: \bibitem{3c} B. Mashhoon, ``Relativity and
1624: Nonlocality", in: Reference
1625: Frames and Gravitomagnetism, edited by J.-F. Pascual-Sanchez, L.
1626: Floria, A. San Miguel and F. Vicente (World Scientific, Singapore,
1627: 2001), pp. 133--144.
1628:
1629: \bibitem{4} B. Mashhoon, Phys. Lett. A 143 (1990) 176.
1630: \bibitem{4a} B. Mashhoon, Phys. Lett. A 145 (1990) 147.
1631: \bibitem{4b} B. Mashhoon,
1632: ``Measurement Theory and General Relativity", in: Black Holes:
1633: Theory and Observation,
1634: edited by F.W. Hehl, C. Kiefer and R. Metzler (Springer, Berlin,
1635: 1998), pp. 269--284.
1636:
1637: \bibitem{5} V. Volterra, Theory of Functionals and of Integral and
1638: Integro-Differential Equations (Dover, New York, 1959).
1639: \bibitem{5a} H.T. Davis,
1640: The Theory of the Volterra
1641: Integral Equation of Second Kind (Indiana University Studies, 17,
1642: 1930).
1643: \bibitem{5b} F.G. Tricomi, Integral Equations (Interscience, New York,
1644: 1957).
1645:
1646: \bibitem{br} N. Bohr and L. Rosenfeld,
1647: Det. Kgl. dansk. Vid. Selskab. 12 (1933) no. 8, translated in:
1648: Quantum Theory and Measurement, edited by J. A. Wheeler and W. H. Zurek
1649: (Princeton University Press, Princeton, 1983).
1650: \bibitem{bra} N. Bohr and L. Rosenfeld, Phys. Rev. 78 (1950) 794.
1651:
1652: \bibitem{6} B. Mashhoon, Phys. Rev. Lett. 61 (1988) 2639.
1653: \bibitem{6a} B. Mashhoon, Phys. Rev. Lett. 68 (1992) 3812.
1654: \bibitem{6b} B. Mashhoon, Phys. Lett. A 198 (1995) 9.
1655: \bibitem{6c} B. Mashhoon, Gen. Rel. Grav. 31 (1999) 681.
1656: \bibitem{6d} B. Mashhoon, Class. Quantum Grav. 17 (2000) 2399.
1657: \bibitem{6e} B. Mashhoon, R. Neutze, M. Hannam and G.E. Stedman, Phys. Lett.
1658: A 249 (1998) 161.
1659:
1660: \bibitem{7} L.D. Landau and E.M. Lifshitz, Electrodynamics of
1661: Continuous Media (Pergamon, Oxford, 1960), p. 249.
1662: \bibitem{7a} A.C. Eringen, J. Math. Phys. 25 (1984) 3235.
1663: \bibitem{7b}
1664: G. Bertotti, Hysteresis in Magnetism (Academic Press, San Diego, 1998).
1665:
1666: \bibitem{8} J.K. Roberge, in: Methods of Experimental Physics:
1667: Electronic Methods, vol. 2, 2nd ed., edited by E. Bleuler and R.O.
1668: Haxby (Academic Press, New York,
1669: 1975), ch. 12.
1670:
1671: \bibitem{ho} F.W. Hehl and Y.N. Obukhov, in:
1672: Gyros, Clocks, Interferometers $\cdots:$ Testing Relativistic Gravity in
1673: Space, edited by C. L\"ammerzahl, C.W.F. Everitt and F.W. Hehl,
1674: Lecture Notes in Physics 562 (Springer, Berlin, 2001), pp. 479--504.
1675: \bibitem{hoa} F.W. Hehl and Y.N. Obukhov,
1676: Foundations of Classical Electrodynamics, in press (Birkh\"auser, Boston,
1677: 2002).
1678:
1679: \bibitem{ll} E.H. Lieb and M. Loss, Analysis, 2nd ed.
1680: (Amer. Math. Soc., Providence, 2001), p. 64.
1681:
1682: \bibitem{hen} P. Henrici, Applied and Computational Complex Analysis
1683: (Wiley, New York, 1986).
1684: \end{thebibliography}
1685:
1686: \end{document}
1687: