1: \documentclass{article}
2: \usepackage{epsfig}
3: \textwidth 5.5in
4:
5: \begin{document}
6:
7: \title{Zeno meets modern science}
8: \author{Z.K. Silagadze \\ Budker Institute of Nuclear Physics,
9: 630 090, Novosibirsk, Russia}
10:
11: \date{}
12:
13: \maketitle
14:
15: \begin{abstract}
16: ``No one has ever touched Zeno without refuting him''. We will not refute
17: Zeno in this paper. Instead we review some unexpected encounters of Zeno with
18: modern science. The paper begins with a brief biography of Zeno of Elea
19: followed by his famous paradoxes of motion. Reflections on continuity of
20: space and time lead us to Banach and Tarski and to their celebrated paradox,
21: which is in fact not a paradox at all but a strict mathematical theorem,
22: although very counterintuitive. Quantum mechanics brings another flavour in
23: Zeno paradoxes. Quantum Zeno and anti-Zeno effects are really paradoxical
24: but now experimental facts. Then we discuss supertasks and bifurcated
25: supertasks. The concept of localization leads us to Newton and Wigner and to
26: interesting phenomenon of quantum revivals. At last we note that the
27: paradoxical idea of timeless universe, defended by Zeno and Parmenides
28: at ancient times, is still alive in quantum gravity. The list of references
29: that follows is necessarily incomplete but we hope it will assist interested
30: reader to fill in details.
31:
32: \end{abstract}
33:
34: \section{Introduction}
35: Concepts of localization, motion and change seems so familiar to our
36: classical intuition:
37: everything happens in some place and everything moves from one place to
38: another in everyday life. Nevertheless it becomes a rather thorny issue
39: then subjected to critical analysis as witnessed long ago by Zeno's
40: paradoxes of motion. One can superficially think that the resolution of the
41: paradoxes was provided by calculus centuries ago by pointing out now the
42: trivial fact that an infinite series can have a finite sum. But on the
43: second thought we realize that this "resolution" assumes infinite
44: divisibility of space and time and we still do not know whether the physical
45: reality still corresponds to the continuous space and time at very small
46: (Plankian) scales. Even in pure mathematics the infinite divisibility leads
47: to paradoxical results like Banach-Tarski paradox which are hard to swallow
48: despite their irrefutable mathematical correctness.
49:
50: More subtle under the surface truth about Zeno's paradoxes is that even if one
51: assumes the infinitely divisible space and time calculus does not really
52: resolves the paradoxes but instead makes them even more paradoxical and leads
53: to conclusion that things cannot be localized arbitrarily sharply. Of course
54: the latter is just what we expect from basic principles of quantum mechanics
55: and special relativity. But it is certainly amazing to find roots of these
56: pillars of the modern physics at Zeno's times!
57:
58: \section{Zeno of Elea}
59: ``No one has ever touched Zeno without refuting him, and every century thinks
60: it worthwhile to refute him'' \cite{1}. Therefore it seems that refuting Zeno
61: is eternal and unchanging affair in complete accord with the Eleatic
62: philosophy. According to this philosophy all appearances of multiplicity,
63: change, and motion are mere illusions. Interestingly the foundation of the
64: Eleatic philosophical school was preceded by turbulent events in drastic
65: contrast with its teaching of the unique, eternal, and unchanging universe
66: \cite{2}. The school was founded by Xenophanes (born circa 570 BC),
67: a wandering exile from his native city of Colophone in Ionia. Before finally
68: joining the colony at Elea, he lived in Sicily and then in Catana. The Elea
69: colony itself was founded by a group of Ionian Greeks which seize the site
70: from the native Oenotrians. Earlier these Ionian Greeks were expelled from
71: their native city of Phocaea by an invading Persian army. Having lost their
72: homes, they sailed to the Corsica island and invaded it after a awful
73: sea battle with the Carthaginians and Etruscans, just to drive once again into
74: the sea as refugees after ten years later (in 545 BC) their rivals regained
75: the island. We can just wonder about psychological influence of these events
76: on the Eleatic school's belief in permanent and unalterable universe\cite{2}.
77:
78: Zeno himself had experienced all treacherous vicissitudes of life. Diogenes
79: Laertius describes him \cite{3} as the very courageous man:
80:
81: ``He, wishing to put an end to the power of Nearches, the tyrant (some,
82: however, call the tyrant Diomedon), was arrested, as we are informed by
83: Heraclides, in his abridgment of Satyrus. And when he was examined, as to his
84: accomplices, and as to the arms which he was taking to Lipara, he named all
85: the friends of the tyrant as his accomplices, wishing to make him feel himself
86: alone. And then, after he had mentioned some names, he said that he wished to
87: whisper something privately to the tyrant; and when he came near him he bit
88: him, and would not leave his hold till he was stabbed. And the same thing
89: happened to Aristogiton, the tyrant slayer. But Demetrius, in his treatise on
90: People of the same Name, says that it was his nose that he bit off.
91:
92: Moreover, Antisthenes, in his Successions, says that after he had given him
93: information against his friends, he was asked by the tyrant if there was any
94: one else. And he replied, "Yes, you, the destruction of the city." And that he
95: also said to the bystanders, "I marvel at your cowardice, if you submit to be
96: slaves to the tyrant out of fear of such pains as I am now enduring." And at
97: last he bit off his tongue and spit it at him; and the citizens immediately
98: rushed forward, and slew the tyrant with stones. And this is the account that
99: is given by almost every one''.
100:
101: Although this account of Zeno's heroic deeds and torture at the hands of the
102: tyrant is generally considered as unreliable \cite{4,5,6}, Zeno after all is
103: famous not for his brevity but for his paradoxes \cite{7,8,9,10,11}.
104:
105: \section{Zeno's paradoxes of motion}
106: Zeno was a disciple of Parmenides, the most illustrious representative of the
107: Eleatic philosophy. According to Parmenides, many things taken for granted,
108: such as motion, change, and plurality, are simply illusions and the
109: reality is in fact an absolute, unchanging oneness. Of course, nothing
110: contradicts more to our common sense experience than this belief. It is not
111: surprising, therefore, that Parmenides' views were ridiculed by contemporaries
112: (and not only). In his ``youthful effort'' Zeno elaborated a number of
113: paradoxes in order to defend the system of Parmenides and attack the common
114: conceptions of things. The four most famous of these paradoxes deny the
115: reality of motion. The Dichotomy paradox, for example, states that it is
116: impossible to cover any distance \cite{2}:
117:
118: \begin{itemize}
119: \item
120: {\it There is no motion, because that which is moved must arrive at the
121: middle before it arrives at the end, and so on ad infinitum.}
122: \end{itemize}
123:
124: According to Simplicius, Diogenes the Cynic after hearing this argument from
125: Zeno's followers silently stood up and walked, so pointing out that it is a
126: matter of the most common experience that things in fact do move \cite{11}.
127: This answer, very clever and effective perhaps, is unfortunately completely
128: misleading, because it is not the apparent motion what Zeno questions but
129: how this motion is logically possible. And the Diogenes's answer does not
130: enlighten us at all in this respect \cite{12}:
131: \begin{verbatim}
132: A bearded sage once said that there's no motion.
133: His silent colleague simply strolled before him, --
134: How could he answer better?! -- all adored him!
135: And praised his wise reply with great devotion.
136: But men, this is enchanting! -- let me interject,
137: For me, another grand occurrence comes to play:
138: The sun rotates around us every single day,
139: And yet, the headstrong Galileo was correct.
140: \end{verbatim}
141:
142: But what is paradoxical in Zeno's arguments? Let us take a closer look. He
143: says that any movement can be subdivided into infinite number of ever
144: decreasing steps. This is not by itself paradoxical, if we assume infinite
145: divisibility of space and time. What is paradoxical is an ability to perform
146: infinite number of subtasks in a finite time -- to perform a supertask. Any
147: movement seems to be a supertask according to Zeno and it is by no means
148: obvious that it is ever possible to perform infinite number of actions in
149: a finite time. Our intuition tells us just the contrary -- that it is a clear
150: impossibility for finite beings to manage any supertask. In the case of
151: Dichotomy it is even not clear how the movement can begin at all because there
152: is no first step to be taken.
153:
154: Aristotle tried to resolve this situation by distinguishing potential and
155: actual infinities \cite{13}: ``To the question whether it is possible to pass
156: through an infinite number of units either of time or of distance we must
157: reply that in a sense it is and in a sense it is not. If the units are actual,
158: it is not possible; if they are potential, it is possible''. But Aristotle's
159: answer is not much better than Diogenes'. It is incomplete. In fact doubly
160: incomplete. According to it Zeno's infinite subdivision of a motion is purely
161: mathematical, just an action of imagination. But even if we accept Aristotle's
162: position it is desirable to show that in mathematics we have tools to handle
163: infinities in a logically coherent way. In fact no such tools were at
164: Aristotle's disposal and they were only germinated after two thousand years
165: when the notion of limit emerged, Calculus was developed and Georg Cantor
166: created his set theory. We can say that Dichotomy is not mathematically
167: paradoxical today. Either classical or non-standard \cite{14,15} analysis can
168: supply sufficient machinery to deal with both the Dichotomy and its more
169: famous counterpart, the Achilles and the tortoise paradox \cite{2}:
170: \begin{itemize}
171: \item
172: {\it The slower will never be overtaken by the quicker, for that which is
173: pursuing must first reach the point from which that which is fleeing started,
174: so that the slower must always be some distance ahead.}
175: \end{itemize}
176: This latter paradox is more impressively formulated in terms of two bodies
177: but in fact it is a symmetric counterpart of the Dichotomy and has a variant
178: involving only one moving body \cite{16}: ``To reach a given point, a body
179: in motion must first traverse half of the distance, then half of what remains,
180: half of this latter, and so on ad infinitum, and again the goal can never be
181: reached''. Therefore if the Dichotomy wonders how the motion can begin as
182: there is no first step, the Achilles makes it equally problematic the end of
183: the motion because there is no last step. Modern mathematics partly completes
184: Aristotle's argument and provides a coherent mathematical picture of motion.
185: But all this mathematical developments, although very wonderful, do not
186: answer the main question implicit in Aristotle's rebuttal of Zeno: how the
187: real motion actually takes place and whether its present day mathematical
188: image still corresponds to reality at the most fundamental level.
189:
190: \section{Zeno meets Banach and Tarski}
191: Let us take, for example, infinite divisibility of space and time. This
192: infinite divisibility is in fact paradoxical, even though the modern
193: mathematics have no trouble to deal with this infinite divisibility. Let us
194: explain what we have in mind.
195:
196: Zeno's argument shows that any spatial or temporal interval contains
197: uncountably many points. Nevertheless a moving body manages to traverse all
198: these points in a finite time. Let us consider any division of the interval
199: into non-empty pairwise mutually exclusive subintervals (that is any pair of
200: them have no common points). Then there exists at least one set N that
201: contains one and only one point from each of the subintervals. indeed, a
202: moving point body enters into a given subinterval sooner or later while
203: traversing the initial interval and will remain into this subinterval for some
204: amount of time. We can take any point the moving body occupies during this
205: time interval as an element of N. All this seems very natural and
206: self-evident, and so does its natural generalization, the axiom of choice
207: \cite{17}:
208: \begin{itemize}
209: \item
210: {\it If M is any collection of pairwise mutually exclusive, non-empty sets P,
211: there exists at least one set N that contains one and only one element from
212: each of the sets P of the collection M.}
213: \end{itemize}
214: \noindent If one may choose an element from each of the sets P of M, the set
215: N can evidently be formed -- hence the name of the axiom.
216:
217: Now this innocently looking ``self-evident'' axiom leads to the most
218: paradoxical result in the mathematics, the Banach-Tarski theorem, which is so
219: contrary to our intuition that is better known as the Banach-Tarski paradox.
220: The most artistic presentation of this paradox can be found in the Bible
221: \cite{18}:
222: ``As he went ashore he saw a great throng; and he had compassion on them, and
223: healed their sick. When it was evening, the disciples came to him and said:
224: "This is a lonely place, and the day is now over; send the crowds away to go
225: into the villages and buy food for themselves". Jesus said: "They need not go
226: away; you give them something to eat". They said to him: "We have only five
227: loaves here and two fish". And he said: "Bring them here to me". Then he
228: ordered the crowds to sit down on the grass; and taking the five loaves and
229: the two fish he looked up to heaven, and blessed, and broke and gave the
230: loaves to the disciples, and the disciples gave them to the crowds. And they
231: all ate and were satisfied. And they took up twelve baskets full of the broken
232: pieces left over. And those who ate were about five thousand men, besides women
233: and children''.
234:
235: In the more formal language the Banach-Tarski theorem states that \cite{17}
236: \begin{itemize}
237: \item
238: {\it in any euclidean space of dimension $n>2$, any two arbitrary bounded sets
239: are equivalent by finite decomposition provided they contain interior points.}
240: \end{itemize}
241: \noindent This theorem opens, for example, the door to the following
242: mathematical alchemy \cite{17}: a ball of the size of orange can be
243: divided into a finite number of pieces which can be reassembled by using
244: merely translations and rotations to yield a solid ball whose diameter is
245: larger than the size of the solar system. Of course a real orange cannot be
246: chopped in such a way because atoms and molecules constitute a limit of
247: divisibility of any chemical substance and the pieces required in the
248: Banach-Tarski theorem are so irregular that they are nonmeasurable and the
249: concept of volume (Lebesgue measure) does not make sense for them.
250: But does space-time itself also have a limit of divisibility? It is yet an
251: open question.
252:
253: The comprehensive discussion of the Banach-Tarski theorem is given in
254: \cite{19} and for an elementary approach with a full proof of the theorem see
255: \cite{20}. One can question whether paradoxical counter-intuitive
256: decompositions like the ones implied by the Banach-Tarski theorem are of any
257: use in physics. Surprisingly, there were several attempts in this direction.
258: Pitowsky was the first (to our knowledge) to consider a certain extension of
259: the concept of probability to nonmeasurable sets in connection with the
260: Einstein-Podolsky-Rosen paradox and Bell's inequalities \cite{21,22}. Another
261: examples can be traced trough \cite{23,24}.
262:
263: One cannot blame the axiom of choice as the only culprit of such paradoxical
264: mathematical results. Even without the use of this axiom one can argue that
265: there is some truth in the proverb that the world is small, because the
266: results proved in \cite{25} entirely constructively, without the axiom of
267: choice, imply that there is a finite collection of disjoint open subsets of
268: the sun that fill the whole sun without holes of positive radius and that
269: nevertheless can be rearranged by rigid motions to fit inside a pea and remain
270: disjoint.
271:
272: Maybe the the Banach-Tarski theorem and analogous paradoxical decompositions
273: will appear a bit less paradoxical if we realize that the Achilles and the
274: tortoise paradox illustrates that any two intervals contain the same number of
275: points regardless their length. Indeed, during their race Achilles and the
276: tortoise cover desperately different intervals. Nevertheless one can arrange
277: a one-to-one correspondence between points of these intervals because for
278: every point A from the Achilles' track there is only one point B on the track
279: of tortoise which the tortoise occupied at the same instant of time when
280: Achilles occupied A. In fact, as Cantor proved in 1877, there is a one-to-one
281: correspondence of points on the interval [0, 1] and points in a square, or
282: points in any n-dimensional space. Cantor himself was surprised at his own
283: discovery and wrote to Dedekind \cite{26} ``I see it, but I don't believe
284: it!''
285:
286: \section{Zeno meets quantum mechanics}
287: Despite some paradoxical flavour, the infinite divisibility of real space-time,
288: although unwarranted at yet, is mathematically coherent. But Zeno's paradoxes
289: contain some other physical premises also that deserve careful consideration.
290: The Achilles and the tortoise paradox, for example, assumes some observation
291: procedure:
292: \begin{itemize}
293: \item check the positions of the contenders in the race.
294: \item check again when Achilles reach the position the tortoise occupied
295: at previous step.
296: \item repeat the previous instruction until Achilles catch the tortoise
297: (and this is an infinite loop because he never does).
298: \end{itemize}
299: \noindent Calculus teach us that the above observational process covers only
300: finite interval of time in spite of its infinitely many steps. And during
301: this time interval the tortoise will be indeed always ahead of Achilles.
302: The observational procedure Zeno is offering simply does not allow us to check
303: the contenders positions later when Achilles overtake the tortoise. So the
304: paradox is solved? Not at all. Zeno's procedure implicitly assumes an ability
305: to perform position measurements. Therefore two questions remain: whether it
306: is possible to perform infinitely frequent measurements taken for granted by
307: Zeno, and how the race will be effected by back-reaction from these
308: measurements. The world is quantum mechanical after all and the measurement
309: process is rather subtle thing in quantum mechanics.
310:
311: Simple arguments \cite{27} show that something interesting is going on if the
312: observational procedure of Zeno is considered from the quantum mechanical
313: perspective.
314: Let $|\Phi,0>$ denote the initial state vector of the system (Achilles and
315: the tortoise in uniform motions). After a short time $t$ the state vector will
316: evolve into $$|\Phi,t>=\exp{\left (-\frac{i}{\hbar}Ht\right )}|\Phi,0>\approx
317: \left (1-\frac{i}{\hbar}Ht -\frac{1}{2\hbar^2} H^2t^2\right )|\Phi,0>\; ,$$
318: where $H$ is the Hamiltonian of the system assumed to be time independent.
319: If now the position measurements of the competitors are performed we find that
320: the initial state have still not changed with the probability
321: $$|<\Phi,0|\Phi,t>|^2\approx 1-\frac{(\Delta E)^2}{\hbar^2}t^2\; ,$$ where
322: $$(\Delta E)^2=<\Phi,0|H^2|\Phi,0>-<\Phi,0|H|\Phi,0>^2$$ is positive (there
323: should be some energy spread in the initial state because we assume good
324: enough localizations for Achilles and the tortoise). If these measurements
325: are performed $n$-times, at intervals $t/n$, there is a probability
326: $$\left ( 1-\frac{(\Delta E)^2}{\hbar^2}\frac{t^2}{n^2}\right )^n$$ that at
327: all times the system will be found in the initial state. But this probability
328: tends to unity when $n\rightarrow \infty$, because in this limit
329: $$\ln{\left ( 1-\frac{(\Delta E)^2}{\hbar^2}\frac{t^2}{n^2}\right )^n}\approx
330: -\frac{(\Delta E)^2}{\hbar^2}\frac{t^2}{n}\rightarrow 0.$$
331: Therefore, if the observations are infinitely frequent the initial state does
332: not change at all. Zeno was right after all: Achilles will never catch
333: the tortoise under proposed observational scheme! This scheme implicitly
334: assumes a continues monitoring of Achilles' position and therefore he
335: will fail even to start the race.
336:
337: Matters are not as simple however. Repeated measurement of a system effects
338: its dynamics much more complex and delicate way than just slowing the
339: evolution \cite{28}. The above described Quantum
340: Zeno Effect became popular after seminal paper of Misra and Sudarshan
341: \cite{29}, although it dates back to Alan Turing (see \cite{30,31} and
342: references wherein) and was known earlier as ``Turing's paradox''. The initial
343: time $t^2$ dependence of quantum mechanical evolution, from which the Quantum
344: Zeno Effect follows most simply, is quite general though not universal. The
345: experimental difficulty resides in the fact that the $t^2$ dependence takes
346: place usually at very short times for natural unstable systems. For example,
347: the "Zeno" time of the 2P-1S transition of the hydrogen atom is estimated
348: to be approximately $3.6 \cdot 10^{-15}~{\mathrm s}$ \cite{32}.
349: Nevertheless modern experimental techniques enable to prepare artificial
350: unstable systems with long enough Zeno time. In beautiful experiment \cite{33}
351: ultra-cold sodium atoms were trapped in a periodic optical potential created
352: by a accelerated standing wave of light. Classically atoms can be trapped
353: inside the potential wells but they will escape by quantum tunneling.
354: The number that remain is measured as a function of duration of the
355: tunneling. The results are shown in the Fig.\ref{qz}.
356: \begin{figure}[htb]
357: \begin{center}
358: \mbox{\epsfig{figure=qz.eps}}
359: \end{center}
360: \caption {Observation of the Quantum Zeno Effect in the experiment \cite{33}.}
361: \label{qz}
362: \end{figure}
363: Hollow squares in this figure show the probability of survival in the
364: accelerated potential as a function of duration of the tunneling acceleration.
365: The solid line represents what is expected according to quantum mechanics
366: and we see a very good agreement with the experimental data. The Zeno time for
367: this unstable system is of the order of about $\mu$s and during this short
368: time period the survival probability exhibits a $t^2$ drop. For longer times
369: we see a gradual transition from the $t^2$ dependence to linear $t$ dependence,
370: which corresponds to the usual exponential decay law. Such behaviour can be
371: simply understood in the framework of time-dependent perturbation
372: theory \cite{34}.
373: The probability of decay of some excited state $|i>$ under the influence of
374: time independent small perturbation $V$ is given by the formula
375: \begin{equation}
376: Q(t)=\frac{1}{\hbar^2}\int\limits_{-\infty}^\infty |<i|V|k>|^2\rho_k\sin^2{
377: \left (\frac{(E_k-E_i)t}{2\hbar}\right )}\left (\frac{2\hbar}{E_k-E_i}
378: \right )^2 dE_k,
379: \label{GR}
380: \end{equation}
381: where $|i>,\; |k>$ are eigenstates of unperturbed Hamiltonian and $\rho_k$
382: is the density of states $|k>$. For very short times one has clearly a $t^2$
383: dependence:
384: $$Q(t)\approx\frac{1}{\hbar^2} \left ( \int\limits_{-\infty}^\infty
385: |<i|V|k>|^2 \rho_k dE_k \right )t^2.$$
386: For longer times, when $$\frac{(E_k-E_i)t}{2\hbar}$$ is not small,
387: one cannot replace the sine function by the first term of its Taylor
388: expansion. However we can expect that only states with small $|E_k-E_i|$
389: contribute significantly in the integral, because if
390: $$z=\frac{(E_k-E_i)t}{2\hbar}\gg 1$$ the integrand oscillates quickly. But
391: then it can be assumed that $|<i|V|k>|^2$ and $\rho_k$ are constant and by
392: using $$\int\limits_{-\infty}^\infty \frac {\sin^2{z}}{z^2} dz=\pi ,$$
393: one obtains linear $t$ dependence
394: $$Q(t)\approx \frac{2\pi}{\hbar}|<i|V|k>|^2\rho_k t.$$
395:
396: In the experiment \cite{33} the number of atoms remaining in the potential
397: well after some time of tunneling was measured by suddenly interrupting the
398: tunneling by a period of reduced acceleration. For the reduced acceleration
399: tunneling was negligible and the atoms increased their velocity without being
400: lost out of the well. Therefore the remaining atoms and the ones having
401: tunneled out up to the point of interruption become separated in velocity
402: space enabling the experimenters to distinguish them. This measurement
403: of the number of remaining atoms projects the system in a new initial state
404: when the acceleration is switched back and the system returns to its unstable
405: state. The evolution must therefore start again with the non-exponential
406: initial segment and one expects the Zeno impeding of the evolution under
407: frequent measurements.
408:
409: \begin{figure}[htb]
410: \begin{center}
411: \mbox{\epsfig{figure=aqz.eps}}
412: \end{center}
413: \caption {Observation of the Quantum Anti-Zeno Effect in the experiment
414: \cite{33}.}
415: \label{aqz}
416: \end{figure}
417: Fig.\ref{qz} really shows the Zeno effect in a rather dramatic way. The solid
418: circles in this figure correspond to the measurement of the survival
419: probability when after each tunneling segment of 1 $\mu$s an interruption of
420: 50 $\mu$s duration was inserted. One clearly sees a much slower decay trend
421: compared to the measurements without frequent interruptions (hollow squares).
422: Some disagreement with the theoretical expectation depicted by the solid line
423: is attributed by the authors of \cite{33} to the under-estimate of the actual
424: tunneling time, so that in reality the decay might be slowed down even at
425: higher degree.
426:
427: Uninterrupted decay curve shows damped oscillatory transition region between
428: initial period of slow decay and the exponential decay at longer times. At
429: that a steep drop in the survival probability is observed immediately after
430: the Zeno time. It is expected, therefore, that the decay will be not slowed
431: down but accelerated if frequent interruptions take place during the steep drop
432: period of evolution forcing the system to repeat the initial period of fast
433: decay again and again after every measurement. This so called Quantum
434: Anti-Zeno effect \cite{35,36,37} is experimentally demonstrated in
435: Fig.\ref{aqz}. The solid circles in this figure show the evolution of the
436: unstable system when after every 5 $\mu$s of tunneling the decay was
437: interrupted by a slow acceleration segment.
438:
439: \section{Supertasks}
440: The Zeno and Anti-Zeno effects in quantum theory are of course very
441: interesting phenomena and even some practical application of the Zeno effect
442: in quantum computing is foreseeable \cite{38}. But in the context of Zeno
443: paradoxes we are more interested in the limit of infinitely frequent
444: measurements with complete inhibition of evolution. Although such limit leads
445: to interesting mathematics \cite{39}, its physical realizability is dubious.
446: The Calculus argument that it is possible for infinite sum to converge to
447: a finite number is not sufficient to ensure a possibility to perform a
448: supertask. This becomes obvious if we somewhat modify Zeno's arguments to
449: stress the role supertasks play in them. The resulting paradox, Zeno Zigzag,
450: goes as follows \cite{2}. A light ray is bouncing between an infinite
451: sequence of mirrors as illustrated in the Fig.\ref{zzg}.
452: \begin{figure}[htb]
453: \begin{center}
454: \mbox{\epsfig{figure=zzg.eps}}
455: \end{center}
456: \caption {The Zeno Zigzag.}
457: \label{zzg}
458: \end{figure}
459: The sizes of mirrors and their separations decrease by a factor of two on
460: each step. The total length of the photon's zigzag path is finite (because
461: the geometric series 1+1/2+1/4+...converges), as well as the envelope box
462: size around the mirrors. Therefore the photon is expected to perform infinite
463: number of reflections in finite time and emerge on the other side of our
464: mirror box. But the absence of the last mirror the photon hit is obviously
465: troublesome now: there is no logical way for the photon to decide at what
466: direction to emerge from the box.
467:
468: Of course, in reality it is impossible to realize the Zeno Zigzag for a number
469: of reasons. One cannot make arbitrarily small ``mirrors'', for example,
470: because sharp localization leads to a significant momentum spread according
471: to uncertainty relations and then the relativity makes possible a pair
472: production which will smear the ``mirror'' position. The wave-like behaviour
473: of the photon (or any other particle) will anyway make impossible to maintain
474: definite direction of the reflected photon if the mirror size is less than
475: the photon wavelength.
476:
477: The question, however, naturally arises whether supertasks are logically
478: impossible irrespective of the nature of physical reality which may restrict
479: their practical realization. To support the opinion that the very notion of
480: completing an infinite sequence of acts in a finite time is logically
481: contradictory, Thomson suggested the following supertask \cite{40}. A lamp is
482: switched ON and OFF more and more rapidly so that at the end of the two
483: minutes a supertask of infinite switching of the lamp is over. The question
484: now is whether the lamp is in the ON state or in the OFF state after this two
485: minutes. Clearly the lamp must be in one of these states but both seem equally
486: impossible. The lamp cannot be in the ON state because we never turned it on
487: without immediately turning it off. But the lamp cannot be in the OFF state
488: either because we never turned it off without at once turning it on. It seems
489: impossible to answer the question and avoid a contradiction.
490:
491: \begin{figure}[htb]
492: \begin{center}
493: \mbox{\epsfig{figure=Tlamp.eps}}
494: \end{center}
495: \caption {Alternative switching mechanisms for Thomson's lamp from \cite{42}.}
496: \label{Tlamp}
497: \end{figure}
498: The Thomson lamp argument is seductive but fallacious \cite{41,42}.
499: Surprisingly there is a coherent answer to the Thomson's question without any
500: contradiction. To come to this answer, it is instructive to consider another
501: supertask \cite{42} which is not paradoxical in any obvious way. A ball
502: bounces on a hard surface so that on each rebound it loses $1-k$-fraction of
503: its velocity prior to the bounce, where $0<k<1$. The ball will perform
504: infinitely many bounces in a finite time because, in classical mechanics,
505: the time between bounces is directly proportional to the initial velocity of
506: the ball and the geometric series $1,\;k,\;k^2,\;k^3,...$ has finite sum
507: $1/(1-k)$. Now let us use this bouncing ball as a switching mechanism for the
508: Thomson's lamp. Then it is immediately obvious that depending on the
509: organization of the circuit the lamp can be in either state (ON or OFF) after
510: the supertask is completed, see Fig.\ref{Tlamp}.
511:
512: Logic once again demonstrates its flexibility. Note that even such a weird
513: notion as the lamp being in a superposition of the ON and OFF states makes
514: perfect sense in quantum mechanics. Although, as was indicated above, we can
515: make the Thomson's lamp logically consistent without any such weirdness.
516: But other surprises with supertasks are lurking ahead.
517:
518: In \cite{43} P\'{e}rez Laraudogoitia constructed a beautifully simple
519: supertask which demonstrates some weird things even in the context of
520: classical mechanics. Fig.\ref{Balls} shows an infinite set of identical
521: particles arranged in a straight line. The distance between the particles
522: and their sizes decrease so that the whole system occupies an interval of
523: unit length. Some other particle of the same mass approaches the system from
524: the right with unit velocity. In elastic collision with identical particles
525: the velocities are exchanged after the collision. Therefore a wave of elastic
526: collisions goes through the system in unit time. And what then? Any particle
527: of the system and the projectile particle comes to rest after colliding its
528: left closest neighbor. Therefore all particles are at rest after the collision
529: supertask is over and we are left with paradoxical conclusion \cite{43} that
530: the total initial energy (and momentum) of the system of particles can
531: disappear by means of an infinitely denumerable number of elastic collisions,
532: in each one of which the energy (and momentum) is conserved!
533: \begin{figure}[htb]
534: \begin{center}
535: \mbox{\epsfig{figure=balls.eps}}
536: \end{center}
537: \caption {P\'{e}rez Laraudogoitia's supertask.}
538: \label{Balls}
539: \end{figure}
540:
541: If you fill uneasy about this energy-momentum nonconservation, here is the same
542: story in more entertaining incarnation \cite{44}.
543:
544: Suppose you have some amount of one dollar bills and the Devil approaches you
545: in a nefarious underground bar. He says that he has a hobby of collecting
546: one dollar bills of particular serial numbers and it happened that you do have
547: one such bill. So he is offering you a bargain: he will give you ten one
548: dollar bills for this particular one. Should you accept the bargain? Why not,
549: it seems so profitable. You agreed and the bargain is done. After half an hour
550: the Devil appears again with the similar offer. Then after a quarter-hour and
551: so on. The time interval between his appearances decreases by a factor of two
552: each time. The amount of your money increases quickly. After an hour infinite
553: number of bargains are done and how much money will you have? You would have
554: a lot if the Devil wanted you to succeed. But he tries to perish your soul not
555: to save it and during the bargains is indeed very selective about bill serial
556: numbers: he always takes the bill with smallest serial number you posses and
557: instead gives the bills with serial numbers greater than any your bill's
558: at that moment. Under such arrangement any bill you posses will end its path
559: into Devil's pocket. Indeed, for any particular bill at your disposal at some
560: instant you will have only finite amount of bills with less serial numbers and
561: you never will get additional bills with smaller serial numbers. Therefore
562: sooner or later this particular bill will become the bill with smallest serial
563: number among your bills and hence subject of the exchange. We come again to
564: the strange conclusion that in spite of your money's continuous growth during
565: the transaction you will have no bill at all left after the infinite
566: transaction is over!
567:
568: Of course, this example does not enlighten us much about how the initial energy
569: disappears in the P\'{e}rez Laraudogoitia's supertask. However it clearly
570: shows a somewhat infernal flavour the supertasks have. And not only the
571: energy-momentum conservation is on stake. Classical dynamics is time reversal
572: invariant. Therefore the following process, which is the time reversal of the
573: P\'{e}rez Laraudogoitia's supertask, is also possible \cite{43}:
574: a spontaneous self-excitation can propagate through the infinite system of
575: balls at rest causing the first ball to be ejected with some nonzero velocity.
576: The system displays not only the energy-momentum non-conservation but also
577: indeterminism \cite{43,45,46}.
578:
579: The locus of the difficulty is the same as in the Zeno paradoxes: there is no
580: last member in a sequence of acts (collisions) and therefore there is no last
581: ball to carry off the velocity \cite{47}. The supertask illustrate the
582: indeterminism of classical Newtonian or even relativistic dynamics as far as
583: infinite localization of bodies is admissible. In quantum mechanics balls
584: cannot be simultaneously at rest and infinitely localized thanks to
585: uncertainty relations. Therefore P\'{e}rez Laraudogoitia's supertask will not
586: persist in quantum theory. However, it was shown \cite{48} that
587: a (nonrelativistic) quantum mechanical supertask can be envisaged in which
588: the deterministic time development of the wave function is lost and
589: spontaneous self-excitation of the ground state is allowed. Yet
590: pathologies disappear if one demands normalizability of the state vector.
591: In this sense quantum mechanical supertasks are better behaved than their
592: classical counterparts \cite{48}.
593:
594: Surprisingly and contrary to common wisdom, classical Newtonian physics is
595: more hostile and unfriendly to determinism than either quantum mechanics or
596: special relativity \cite{49}. Another example of esoteric behaviour of
597: seemingly benign Newtonian system was given by Xia \cite{50} while solving
598: the century-old intriguing problem of noncollisional singularities. Xia's
599: construction constitutes in fact a supertask and involves only five bodies
600: interacting via familiar Newtonian inverse square force law. The essentials
601: of this supertask can be explained as follows \cite{51}. Imagine a system of
602: two equal masses $M$ moving in the $x-y$ plane under Newton's law of
603: attraction with center of mass at rest, and the third mass $m\ll M$ (so that
604: it does not disturb the motion of the first two) on the $z$-axis which goes
605: through the center of mass of the planar binary system. It is clear from the
606: symmetry that the total force of gravitational attraction acting on the mass
607: $m$ has only $z$-component and therefore the third mass will experience a one
608: dimensional motion along the $z$-axis. When the mass $m$ passes through and
609: is lightly above the binary plane, we can arrange extremely powerful downward
610: pull on $m$ if the highly elliptical binary has its closest approach at that
611: moment. In fact, assuming an ideal case of point masses, the downward force
612: acting on $m$ can be made arbitrarily strong by just adjusting the separating
613: distances among the bodies. As a result the third mass is jolted down with
614: high velocity while the binary starts separating and loses significantly its
615: braking effect on $m$.
616:
617: The mirror replica of the binary, see Fig.\ref{Xia}, is placed at large
618: distance (not to disturb the first binary) further down to prevent the mass
619: $m$ from being expelled to infinity. In case of proper timing, the second
620: binary will break $m$'s downward motion and will thrust it upwards with even
621: higher velocity.
622: \begin{figure}[htb]
623: \begin{center}
624: \mbox{\epsfig{figure=xia.eps}}
625: \end{center}
626: \caption {Xia's five body supertask.}
627: \label{Xia}
628: \end{figure}
629:
630: Xia was able to show that there exists a Cantor set of the initial conditions
631: allowing to repeat this behaviour infinitely often in a finite time. As
632: a result the four of five bodies from the Xia's construction will escape
633: to spatial infinity in a finite time, while the fifth will oscillate back and
634: forth among the other four with ever increasing speed.
635:
636: Xia's supertask does not violate energy conservation, the bodies drawing out
637: their energies from the infinitely deep $1/r$ potential well. However it
638: implies indeterminacy in idealized Newtonian world. The time reverse of this
639: supertask is an example of ``space invaders'' \cite{49}, particles appearing
640: from spatial infinity in a surprise attack. Note that \cite{49} ``the
641: prospects for determinism brighten considerably when we leave classical
642: spacetimes for Minkowski spacetime, the spacetime setting for special
643: relativistic theories''. Because with no superluminal propagation there are
644: no space invaders.
645:
646: Leaving aside interesting philosophical questions raised by Xia's supertask,
647: it seems otherwise completely artificial and far from reality. Curiously, some
648: ideas involved in this construction can be used by mankind, admittedly in the
649: very long run, to transfer orbital energy from Jupiter to the Earth by
650: a suitable intermediate minor space body, causing the Earth's orbit to expand
651: and avoid an excessive heating from enlarging and brightening Sun at the last
652: stage of its main sequence life \cite{52}.
653:
654: \section{Bifurcated supertasks}
655: Let us return to Zeno. Hermann Weyl pointed out \cite{53} another possibility
656: mankind can benefit from Zeno supertask: ``if the segment of length 1 really
657: consists of infinitely many subsegments of length 1/2, 1/4, 1/8, . . ., as of
658: ``chopped-off'' wholes, then it is incompatible with the character of the
659: infinite as the ``incompletable'' that Achilles should have been able to
660: traverse them all. If one admits this possibility, then there is no reason
661: why a machine should not be capable of completing an infinite sequence of
662: distinct acts of decision within a finite amount of time; say, by supplying
663: the first result after 1/2 minute, the second after another 1/4 minute, the
664: third 1/8 minute later than the second, etc. In this way it would be possible,
665: provided the receptive power of the brain would function similarly, to
666: achieve a traversal of all natural numbers and thereby a sure yes-or-no
667: decision regarding any existential question about natural numbers!''
668:
669: Relativity brings additional flaviour in discussion of Weyl's infinity
670: machines. One can imagine, for example, bifurcated supertasks \cite{42,54}.
671: To check the Goldbach's conjecture whether every even number greater than two
672: can be written as the sum of two primes, the Master organizes two space
673: missions. In one of them a computer, the Servant, is sent with constant
674: acceleration which examines all even numbers one case after the other. In
675: another space mission the Master himself contrives to accelerate in such a way
676: to keep the Servant within his causal horizon. It is possible to arrange
677: Master's acceleration so that the Servant disappears from his causal shadow
678: only after spending infinite amount of time on the Goldbach's conjecture,
679: while in the Master's frame the amount of time passed remains finite. If the
680: Servant finds a counterexample to the Goldbach's conjecture it sends
681: a message to the Master and the latter will know that the conjecture is false.
682: If no message is received, however, the Master will know in a finite time that
683: Goldbach was wright.
684:
685: This looks too good to be true and indeed the realization of such Pitowsky
686: \cite{55} infinite machine is suspicious for a number of reasons \cite{42,54}.
687: The Servant moves with constant acceleration during infinite proper time.
688: Therefore it needs an infinite fuel supply. Moreover, the Master's
689: acceleration increases without limit and eventually he would be crushed to
690: death by artificial gravity. There is a conceptual problem also. At no point
691: on his world line does the Master have a causal access to all events on the
692: Servant's world line. This means there is no Moment of Truth, if the
693: Goldbach's conjecture is true, at which the Master attained that knowledge.
694: Emergence of the knowledge about validity of the Goldbach's conjecture will
695: be as mysterious in Pitowsky bifurcated supertask as is the disappearance
696: of energy in P\'{e}rez Laraudogoitia's supertask.
697:
698: But general relativity allows to improve the above construction by admitting
699: the so called Malament-Hogarth spacetimes \cite{54,56}. In such a spacetime
700: there is a point (the Malament-Hogarth point) at Master's world line such that
701: entire infinite history of the Servant's world line is contained in this
702: point's causal shadow. From that event on the Master will be enlightened about
703: validity of that particular number theoretic problem through the Servant's
704: infinite labors.
705:
706: Some Malament-Hogarth spacetimes seem quite reasonable physically. Among them
707: are such well known spacetimes as the anti-de Sitter spacetime,
708: Reissner-Nordstr\"{o}m spacetime and Kerr-Newman spacetime \cite{57}, the
709: latter being the natural outcome of the late-time evolution of a collapsed
710: rotating star. Therefore it is not excluded that ``if the Creator had
711: a taste for the bizarre we might find that we are inhabiting one of them''
712: \cite{42}. Even if this proves to be the case, the practical realization of
713: infinite computation is not at all guaranteed. Physics beyond the classical
714: general relativity, for example proton decay and other issues related to the
715: long term fate of the universe \cite{58}, can prohibit the infinite
716: calculations. There is still another reason that can make a practical
717: realization of bifurcated supertasks dubious \cite{55}: the Servant might
718: have infinite time to accomplish its labors but not enough computation space.
719: the argument goes as follows \cite{59,60}.
720:
721: Any system which performs computation as an irreversible process will
722: dissipate energy. Many-to-one logical operations such as AND or ERASE are not
723: reversible and require dissipation of at least $kT\ln{2}$ energy per bit of
724: information lost when performed in a computer at temperature $T$. While
725: one-to-one logical operations such as NOT are reversible and in theory can be
726: performed without dissipation \cite{61}.
727:
728: That erasure of one bit information has an energy cost $kT\ln{2}$ (the
729: Landauer's principle) can be demonstrated by Maxwell's demon paradox
730: originally due to Leo Szilard \cite{62,63}.
731: \begin{figure}[htb]
732: \begin{center}
733: \mbox{\epsfig{figure=demon.eps}}
734: \end{center}
735: \caption {Szilard's demon engine.}
736: \label{demon}
737: \end{figure}
738:
739: A schematic view of the Szilard's demon engine is shown in the
740: Fig.\ref{demon}. Initially the entire volume of a cylinder is available to
741: the one-atom working gas (step (a)). At step (b) the demon measures the
742: position of the atom and if it is found in the right half of the cylinder
743: inserts a piston. At step (c) the one-atom gas expands isothermally by
744: extracting necessary heat from the reservoir and lifts the load. At step
745: (d) the piston is removed and the system is returned to its initial state ready
746: to repeat the cycle.
747:
748: It seems the Szilard's demon engine defeats the second law of thermodynamics:
749: the heat bath, which has transferred energy to the gas, has lowered its
750: entropy while the engine has not changed its entropy because it returned to
751: the initial state. But this reasoning is fallacious because it misses an
752: important point: the demon has not returned to his initial state. He still
753: possesses one bit information about left-right position of the atom he
754: recorded. The system truly to return to its initial state it is necessary to
755: erase this information from the demon's memory. According to the
756: Landauer's principle, one has to pay $kT\ln{2}$ energy cost for this erasure.
757: On the other hand work done during isothermal expansion equals
758: $$\int\limits_{V/2}^V pdV=\int\limits_{V/2}^V \frac{m}{\mu}RT \frac{dV}{V}=
759: \frac{m}{\mu}RT \ln{2}$$
760: and normalized to one atom this is just $(R/N_A)T \ln{2}=kT\ln{2}$. Therefore
761: all extracted work is paid back for erasure of the demon's memory and the net
762: effect of the circle is zero. The second law defeats the demon.
763:
764: In principle all computations could be performed using reversible logical
765: operations and hence without energy cost \cite{61}. Interestingly enough, some
766: important stages of biomolecular information processing, such as
767: transcription of DNA to RNA, appear to be accomplished by reversible chemical
768: reactions \cite{64}. Real computers, however, are subject to thermal
769: fluctuations that cause errors. To perform reliable computations, therefore,
770: some error-correcting codes must be used to detect inevitable errors and
771: reject them to the environment at the cost of energy dissipation \cite{61}.
772:
773: The Servant computer from bifurcated supertask needs to consume energy from
774: surrounding universe to perform its task. Different energy mining strategies
775: were considered in \cite{59} and it was shown that none of them allows to
776: gather infinite amount of energy even in infinite time irrespective assumed
777: cosmological model. To perform infinite number of computations with finite
778: available energy the Servant should be able to continuously decrease its
779: operating temperature to reduce the energy cost of computations. But the
780: Servant is not completely free to choose its temperature: the waste heat
781: produced while performing computations must be radiated away to avoid
782: overheating. But physical laws place limits on the rate at which the waste
783: heat can be radiated. Assuming that the electromagnetic dipole radiation by
784: electrons is the most efficient way to get rid of the heat, Dyson argued
785: \cite{65} that there is a lower limit on the operating temperature
786: $$T>\frac{Q}{N_e}10^{-12}K,$$
787: where $Q$ is a measure of the complexity of the computing device ($Q\sim
788: 10^{23}$ for humans) and $N_e$ is the number of electrons in the computer.
789: Since $Q/N_e$ cannot be made arbitrarily small, it seems infinite
790: computations are impossible.
791:
792: But where is a way out \cite{65}: hibernation. The Servant can compute
793: intermittently while continuing to radiate waste heat into space during its
794: periods of hibernation. This strategy will allow to operate below the Dyson
795: limiting temperature. But eventually the wavelength of thermal radiation will
796: become very large compared to the characteristic size of the computer, the
797: thermal energies will be small compared to the characteristic quantized
798: energy levels of the system and radiation will be suppressed by an exponential
799: factor compared to the estimates of Dyson \cite{59}. Therefore the Servant
800: must increase its size as time goes by. It also needs more and more memory to
801: store digital codes of ever increasing even and prime numbers. This is another
802: reason it to increase in size. But as was mentioned above there is only finite
803: amount of energy, and hence material, available. Therefore it appears the
804: servant will not be able to accomplish its infinite labors.
805:
806: \section{Zeno meets Newton and Wigner}
807: Our discussion of Zeno's paradoxes indicates a rather subtle role the concept
808: of localization plays in description of motion.
809: It is not surprising therefore that the notion of localization in
810: relativistic quantum mechanics was intensively examined. Many concepts of
811: localization have been proposed but we focus on the one known as the
812: Newton-Wigner localization \cite{66,67}.
813: Initially this concept of localization and the
814: corresponding position operators were suggested in the context of the single
815: particle relativistic quantum mechanics. But later their result was
816: reformulated in quantum field theory also, where the concept of local
817: observables is the central concept leading to numerous troubles (infinities)
818: which are swept under carpet with great artistic skill by renormalization.
819: Many troubles related to localization in relativistic quantum field theory have
820: their formal root in the Reeh-Schlieder theorem which in the non-formal
821: artistic formulation of Hans Halvorson \cite{68}
822: states that the vacuum is seething
823: with activity at the local level: any local event has the nonzero
824: probability to occur in the vacuum state for the standard localization scheme.
825: The Newton-Wigner localization scheme avoids some consequences of the
826: Reeh-Schlieder theorem and leads to a mathematical structure which seems
827: more comfortable for our a priori physical intuition about localization
828: \cite{69}. But the story is not yet over. The suggested generalizations of the
829: Newton-Wigner localization are still not immune against the full strength of
830: the Reeh-Schlieder theorem and have their own counterintuitive features
831: \cite{68,70}.
832:
833: Let us sketch the derivation of the Newton-Wigner position operator for Dirac
834: particles \cite{71}. The Newton-Wigner operator $\hat Q^k$ can be defined as
835: the operator whose eigenvalues are the most localized wave-packets formed
836: from only positive-energy solutions of the Dirac equation. Let $\psi^{(s)}
837: _{(\vec{y})}(\vec{x})$ be such a wave-function describing an electron with
838: spin projection $s$ localized at the point $\vec{y}$ at the time $t=0$.
839: Defining the scalar product by
840: $$(\psi,\phi)=\int \psi^+(\vec{x})\phi(\vec{x})\; d\vec{x}=
841: \int \psi^+(\vec{p})\phi(\vec{p})\; d\vec{p},$$
842: the natural normalization condition for these states is
843: \begin{equation}
844: \left (\psi^{(s)}_{(\vec{y})},\psi^{(r)}_{(\vec{z})}\right )=
845: \delta_{rs}\delta(\vec{z}-\vec{y}).
846: \label{norm}
847: \end{equation}
848: Translational invariance implies
849: $$\psi^{(s)}_{(\vec{y})}(\vec{x})=\psi^{(s)}_{(\vec{y}+\vec{a})}(\vec{x}+
850: \vec{a}),$$
851: or in the momentum space
852: \begin{equation}
853: \psi^{(s)}_{(\vec{y})}(\vec{p})=e^{i\vec{p}\cdot \vec{a}}
854: \psi^{(s)}_{(\vec{y}+\vec{a})}(\vec{p}),
855: \label{trans}
856: \end{equation}
857: where momentum space wave-functions are defined through
858: $$\psi(\vec{p})=\frac{1}{(2\pi)^{3/2}}\int \psi(\vec{x})
859: e^{-i\vec{p}\cdot \vec{x}}\; d\vec{x}. $$
860: Equations (\ref{norm}) and (\ref{trans}) imply
861: \begin{equation}
862: \psi^{(s)+}_{(\vec{y})}(\vec{p})\;\psi^{(r)}_{(\vec{y})}(\vec{p})=
863: \frac{\delta_{rs}}{(2\pi)^3}.
864: \label{eq4}
865: \end{equation}
866: On the other hand
867: \begin{equation}
868: \psi^{(s)}_{(\vec{y})}(\vec{p})=f_{(\vec{y})}(\vec{p})\;u(\vec{p},s),
869: \label{eq5}
870: \end{equation}
871: where $u(\vec{p},1/2)$ and $u(\vec{p},-1/2)$ are two independent
872: positive-energy solutions of the Dirac equation, which can be
873: taken in the form
874: $$u(\vec{p},s)=\Lambda_+(\vec{p})\;u(\vec{0},s),$$
875: where the rest state four-component spinors are
876: $$u(\vec{0},1/2)=(1,0,0,0)^T,\;\;\; u(\vec{0},-1/2)=(0,1,0,0)^T,$$
877: and $\Lambda_+(\vec{p})$ is a positive-energy projection operator
878: for a Dirac particle of mass $m$:
879: $$\Lambda_+(\vec{p})=\frac{1}{2}\left ( 1+\frac{\vec{\alpha}\cdot\vec{p}+
880: \beta m}{p_0}\right)=\frac{(\hat p+m)\gamma_0}{2p_0},\;\;p_0=\sqrt{\vec{p}^2+
881: m^2}.$$
882: The normalization of $u(\vec{p},s)$ is
883: \begin{equation}
884: u^{+}(\vec{p},s)u(\vec{p},s^\prime)=u^{+}(\vec{0},s)\Lambda_+(\vec{p})
885: u(\vec{0},s^\prime)=\frac{p_0+m}{2p_0}\delta_{ss^\prime}.
886: \label{eq6}
887: \end{equation}
888: From equations (\ref{eq4}-\ref{eq6}) we get
889: $$|f_{(\vec{y})}(\vec{p})|^2=(2\pi)^{-3}\frac{2p_0}{p_0+m}.$$
890: We fix the phase of $f_{(\vec{y})}(\vec{p})$ by assuming
891: $$f^*_{(\vec{0})}(\vec{p})=f_{(\vec{0})}(\vec{p}).$$
892: Then (\ref{trans}) indicates that the momentum space wave-function for the
893: electron localized at $\vec{y}$ at the time $t=0$ has the form
894: \begin{equation}
895: \psi^{(s)}_{(\vec{y})}(\vec{p})=\frac{1}{(2\pi)^{3/2}}\sqrt{\frac{2p_0}{p_0+m}}
896: e^{-i\vec{p}\cdot\vec{y}}\;u(\vec{p},s).
897: \label{eq7}
898: \end{equation}
899: Now we construct the Newton-Wigner position operator $\hat Q^k$ for which
900: (\ref{eq7}) is the eigenfunction with eigenvalue $y^k$:
901: $$\hat Q^k\psi^{(s)}_{(\vec{y})}(\vec{p})=y^k \psi^{(s)}_{(\vec{y})}
902: (\vec{p}).$$
903: Assuming that $\psi^{(s)}_{(\vec{y})}$ eigenfunctions form a complete system
904: for positive-energy solutions, we get for any positive-energy wave-function
905: $\psi (\vec{p})$
906: $$\hat Q^k\psi (\vec{p})=\hat Q^k \sum_s\int d\vec{y} \left (\psi^{(s)}_
907: {(\vec{y})},\psi\right)\psi^{(s)}_{(\vec{y})}(\vec{p})=
908: \sum_s\int d\vec{y} y^k \left (\psi^{(s)}_{(\vec{y})},\psi\right)\psi^{(s)}_
909: {(\vec{y})}(\vec{p}). $$
910: Substituting (\ref{eq7}) and performing $y$-integration we get
911: $$\hat Q^k\psi (\vec{p})=\int d\vec{q} \frac{2\sqrt{q_0p_0}}{(q_0+m)(p_0+m)}
912: \sum_s u(\vec{p},s)\,u^+(\vec{q},s)\left( -i\frac{\partial}{\partial q^k}
913: \delta(\vec{q}-\vec{p})\right ) \psi(\vec{q}).$$
914: But
915: $$\sum_s u(\vec{p},s)\,u^+(\vec{q},s)=\Lambda_+(\vec{p})\left (\sum_s
916: u(\vec{0},s)\,u^+(\vec{0},s)\right )\Lambda_+(\vec{q})=\frac{1}{2}
917: \Lambda_+(\vec{p})(1+\gamma_0)\Lambda_+(\vec{q}).$$
918: Therefore
919: $$\hat Q^k\psi (\vec{p})=\Lambda_+(\vec{p})(1+\gamma_0)\sqrt{\frac{p_0}
920: {p_0+m}}\left( i\frac{\partial}{\partial p^k}\right )\sqrt{\frac{p_0}
921: {p_0+m}}\Lambda_+(\vec{p})\psi (\vec{p}).$$
922: From this equation it is clear that
923: \begin{equation}
924: \hat Q^k=\Lambda_+(\vec{p})(1+\gamma_0)\sqrt{\frac{p_0}
925: {p_0+m}}\left( i\frac{\partial}{\partial p^k}\right )\sqrt{\frac{p_0}
926: {p_0+m}}\Lambda_+(\vec{p}).
927: \label{eq8}
928: \end{equation}
929: Newton-Wigner position operator simplifies in the Foldy-Wouthuesen
930: representation. In this representation its eigenfunctions are obtained
931: via unitary transformation \cite{72,73,74}:
932: $$\phi^{(s)}_{(\vec{y})}(\vec{p})=e^{iW}\psi^{(s)}_{(\vec{y})}(\vec{p}),$$
933: where the Foldy-Wouthuesen unitary operator is
934: $$e^{iW}=\sqrt{\frac{2p_0}{p_0+m}}\left\{\frac{1}{2}(1+\gamma_0)
935: \Lambda_+(\vec{p})+\frac{1}{2}(1-\gamma_0)\Lambda_-(\vec{p})\right\}.$$
936: Using the identity
937: $$\frac{1}{2}(1+\gamma_0)\Lambda_+(\vec{p})\,u(\vec{0},s)=
938: \frac{1}{2}(1+\gamma_0)\Lambda_+(\vec{p})\frac{1}{2}(1+\gamma_0)
939: \,u(\vec{0},s)=\frac{p_0+m}{2p_0}u(\vec{0},s)$$
940: we get
941: $$\phi^{(s)}_{(\vec{y})}(\vec{p})=\frac{1}{(2\pi)^{3/2}}e^{-i\vec{p}\cdot
942: \vec{y}}u(\vec{0},s).$$
943: These are Newton-Wigner position operator eigenstates in the Foldy-Wouthuesen
944: representation. The corresponding position operator can be derived from them
945: in the above described manner. The result is (in the momentum space)
946: $$\hat {\vec{Q}}_{(W)}=\frac{1}{2}(1+\gamma_0)i\frac{\partial}
947: {\partial \vec{p}}\;,$$
948: or in configuration space
949: $$\hat {\vec{Q}}_{(W)}=\frac{1}{2}(1+\gamma_0)\vec{x}=\frac{1}{2}
950: (1+\gamma_0)\vec{x}\frac{1}{2}(1+\gamma_0).$$
951: Therefore the Newton-Wigner position operator appears to be just the
952: positive-energy projection of the Foldy-Wouthuesen ``mean position operator''
953: \cite{72,73,74} and hence (\ref{eq8}) is equivalent to
954: \begin{equation}
955: \hat{\vec{Q}}=\Lambda_+(\vec{p})\left\{i\frac{\partial}{\partial \vec{p}}+
956: i\frac{\beta \vec{\alpha}}{2p_0}-\frac{i\beta(\vec{\alpha}\cdot\vec{p})\vec{p}
957: +(\vec{\sigma}\times\vec{p})p_0}{2p_0^2(p_0+m)}\right\}\Lambda_+(\vec{p}),
958: \label{FWmp}
959: \end{equation}
960: where $\vec{\sigma}=(1/2i)\vec{\alpha}\times\vec{\alpha}.$
961:
962: Let us compare time evolutions of the conventional (Dirac) and
963: Newton-Wigner position operators in the Heisenberg picture \cite{75,76}.
964: The Dirac position $\vec{x}(t)=e^{i\hat Ht}\vec{x}e^{-i\hat Ht}$
965: satisfies the Heisenberg equation of motion
966: $$\frac{d\vec{x}(t)}{dt}=i[\hat H,\vec{x}(t)]=e^{i\hat Ht}\vec{\alpha}
967: e^{-i\hat Ht}=\vec{\alpha}(t),$$
968: where we have used $\hat H=\vec{\alpha}\cdot\hat{\vec{p}}+\beta m$ and the
969: canonical commutation relations (we are using $\hbar=c=1$ convention
970: throughout this section)
971: $$[x_i,\hat p_j]=i\delta_{ij}.$$
972: On the other hand by using
973: $$[\alpha_i,\alpha_j]=2(\delta_{ij}-\alpha_j\alpha_i)=2(\alpha_i\alpha_j-
974: \delta_{ij}),\;\;[\beta,\vec\alpha]=2\beta\vec\alpha=-2\vec\alpha\beta,$$
975: we get
976: $$\frac{d\vec{\alpha}(t)}{dt}=i[\hat H,\vec{\alpha}(t)]=
977: ie^{i\hat Ht}[\hat H,\vec{\alpha}]e^{-i\hat Ht}=2i[\hat{\vec{p}}-\vec{\alpha}
978: (t)\hat H]=2i[\hat H \vec{\alpha}(t)-\hat{\vec{p}}].$$
979: Differentiating once more, we obtain
980: \begin{equation}
981: \frac{d}{dt}\dot{ \vec{\alpha}}(t)=-2i\dot{ \vec{\alpha}}(t)\hat H=
982: 2i\hat H\dot{ \vec{\alpha}}(t).
983: \label{dadot}
984: \end{equation}
985: Therefore
986: $$\dot{ \vec{\alpha}}(t)\hat H=-\hat H \dot{\vec{\alpha}}(t)$$
987: and the solution of (\ref{dadot}) is
988: \begin{equation}
989: \dot{ \vec{\alpha}}(t)\equiv \frac{d\vec{\alpha}(t)}{dt}=
990: \dot{ \vec{\alpha}}(0)e^{-2i\hat H t}=2i(\hat{\vec{p}}-\vec{\alpha}
991: \hat H)e^{-2i\hat H t}.
992: \label{adot}
993: \end{equation}
994: One can integrate (\ref{adot}) by using
995: $$\int\limits_0^t e^{-2i\hat H \tau} d\tau=\frac{i}{2}\hat{H}^{-1}\left (
996: e^{-2i\hat H t}-1\right )$$
997: and the result is
998: $$\vec{\alpha}(t)=\hat{\vec{p}}\hat{H}^{-1}+(\vec{\alpha}-\hat{\vec{p}}
999: \hat{H}^{-1})e^{-2i\hat H t}.$$
1000: Inserting this into
1001: $$\frac{d\vec{x}(t)}{dt}=\vec{\alpha}(t)$$
1002: and integrating the resulting equation, we get the Dirac position operator
1003: in the Heisenberg picture
1004: \begin{equation}
1005: \vec{x}(t)=\vec{x}+\hat{\vec{p}}\hat{H}^{-1}t+\frac{i}{2}(\vec{\alpha}-
1006: \hat{\vec{p}}\hat{H}^{-1})\hat{H}^{-1}\left (e^{-2i\hat H t}-1\right ).
1007: \label{xdirac}
1008: \end{equation}
1009: This equation shows that a free electron surprisingly performs a very
1010: complicated oscillatory motion. The first two terms are just what is expected,
1011: because
1012: $$\vec{p}H^{-1}=\frac{\vec{p}}{p_0}\frac{H}{p_0}=\frac{\vec{p}}{p_0}\left (
1013: \Lambda_+(\vec{p})-\Lambda_-(\vec{p})\right )$$
1014: is essentially the conventional relativistic velocity for positive-energy
1015: wave functions. But the last term in (\ref{xdirac}) is at odds with our
1016: classical intuition. According to it the Dirac electron executes rapid
1017: oscillatory motion, which Schr\"{o}dinger called ``Zitterbewegung''.
1018: As $\vec\alpha(t)\cdot\vec\alpha(t)=1$, the instantaneous value of electron's
1019: velocity during this trembling motion is always 1 (that is the velocity of
1020: light). The Zitterbewegung is a result of an interference between positive
1021: and negative frequency Fourier components of the particle wave-packet. The way
1022: the Zitterbewegung shows itself depends on the character of the particle
1023: wave-packet \cite{75}. In the case of plane wave (not localized particles)
1024: one has a steady-state violent oscillations with amplitude $\sim 1/m$ and
1025: angular frequency $\sim 2m$ ($\hbar/mc$ and $2mc^2/\hbar$ if $\hbar$ and $c$
1026: are restored. For the electron $\hbar/mc=3.85\cdot 10^{-3}$~{\rm \AA} and
1027: $2mc^2/\hbar=1.55\cdot 10^{21}~\mathrm{Hz}$).
1028:
1029: For the Newton-Wigner position operator we will have
1030: $$\frac{d\hat{\vec{Q}}(t)}{dt}=i[\hat H,\hat{\vec{Q}}(t)]=
1031: ie^{-iW}e^{i\hat{H}_{(W)}t}[\hat{H}_{(W)},\hat{\vec{Q}}_{(W)}]
1032: e^{-i\hat{H}_{(W)}t}e^{iW}.$$
1033: But in the Foldy-Wouthuesen representation
1034: $$\hat{H}_{(W)}=\beta p_0,\;\; \hat{\vec{Q}}_{(W)}=\frac{1}{2}(1+\beta)
1035: i\frac{\partial}{\partial \vec{p}}$$
1036: and
1037: $$[\hat{H}_{(W)},\hat{\vec{Q}}_{(W)}]=-\frac{i}{2}(1+\beta)\frac{\vec{p}}
1038: {p_0}.$$
1039: Therefore
1040: $$\frac{d\hat{\vec{Q}}(t)}{dt}=ie^{-iW}\frac{1}{2}(1+\beta)\frac{\vec{p}}
1041: {p_0}\,e^{iW}=\frac{\vec{p}}{p_0}\Lambda_+(\vec{p})$$
1042: and
1043: \begin{equation}
1044: \hat{\vec{Q}}(t)=\hat{\vec{Q}}+\frac{\vec{p}}{p_0}\Lambda_+(\vec{p})t,
1045: \label{QNW}
1046: \end{equation}
1047: where $\hat{\vec{Q}}\equiv \hat{\vec{Q}}(0)$ is the initial value of the
1048: Newton-Wigner position operator given by (\ref{FWmp}).
1049: As we see, the Zitterbewegung is absent in the time development of the
1050: Heisenberg picture Newton-Wigner position operator. This is not surprising
1051: because only positive frequencies are present in the Fourier components
1052: of wave packets that are localized in the Newton-Wigner sense. However this
1053: absence of trembling has a price: Newton-Wigner wave-packets cannot be
1054: localized sharper than $1/m$. To see this, let us consider the Newton-Wigner
1055: eigenfunction (\ref{eq7}) in configuration space.
1056: $$\psi^{(s)}_{(\vec{y})}(\vec{x})=\frac{1}{(2\pi)^3}\int e^{i\vec{p}\cdot
1057: (\vec{x}-\vec{y})}\sqrt{\frac{2p_0}{p_0+m}}\Lambda_+(\vec{p})\,u(\vec{0},s)\,
1058: d\vec{p}.$$
1059: By using
1060: $$u(\vec{0},s)=\frac{1}{2}(1+\gamma_0)u(\vec{0},s),\;\;
1061: \Lambda_+(\vec{p})\frac{1}{2}(1+\gamma_0)=\frac{1}{2p_0}
1062: \left ( \begin{array}{cc} p_0+m & 0 \\ \vec{p}\cdot\vec{\sigma} & 0
1063: \end{array} \right ),$$
1064: we get
1065: $$\psi^{(s)}_{(\vec{y})}(\vec{x})=\frac{1}{(2\pi)^3}\int e^{i\vec{p}\cdot
1066: (\vec{x}-\vec{y})}\frac{1}{\sqrt{2p_0(p_0+m)}}\left ( \begin{array}{cc}
1067: p_0+m & 0 \\ \vec{p}\cdot\vec{\sigma} & 0 \end{array} \right )u(\vec{0},s)\,
1068: d\vec{p},$$
1069: or
1070: $$\psi^{(s)}_{(\vec{y})}(\vec{x})=\left ( \begin{array}{cc} A(\vec{z};m) & 0
1071: \\ -2i\vec{\sigma}\cdot\frac{\partial^2 A(\vec{z};m)}{\partial \vec{z}
1072: \partial m} & 0 \end{array} \right )u(\vec{0},s),$$
1073: where $\vec{z}=\vec{x}-\vec{y}$ and
1074: $$A(\vec{z};m)=\frac{1}{(2\pi)^3}\int e^{i\vec{p}\cdot\vec{z}}\sqrt{\frac{
1075: p_0+m}{2p_0}}\,d\vec{p}.$$
1076: For evaluation of the latter function, it is convenient to decompose
1077: $$\sqrt{1+\frac{m}{p_0}}=1+\sum\limits_{n=1}^\infty a_n \left (\frac{m}{p_0}
1078: \right )^n.$$
1079: Then
1080: \begin{equation}
1081: A(\vec{z};m)=\frac{\delta(\vec{z})}{\sqrt{2}}+\frac{\sqrt{2}}{(2\pi)^2}
1082: \sum\limits_{n=1}^\infty a_n\int\limits_0^\infty p \left (\frac{m}{p_0}
1083: \right )^n \frac{\sin{(pz)}}{z}\, dp,
1084: \label{Az}
1085: \end{equation}
1086: where $p=|\vec{p}|$ and $z=|\vec{z}|$.
1087: But
1088: $$I=m^n \int\limits_0^\infty \frac{p}{p_0^n} \frac{\sin{(pz)}}{z} dp=
1089: -\frac{m^n}{z}\frac{d}{dz}\int\limits_0^\infty \frac{\cos{(pz)}}{(p^2+m^2)^
1090: \frac{n}{2}}\, dp.$$
1091: This latter integral can be evaluated by using \cite{77}
1092: $$K_\nu (mz)=\frac{\Gamma(\nu+\frac{1}{2}) (2m)^\nu}{\sqrt{\pi}z^\nu}
1093: \int\limits_0^\infty \frac{\cos{(pz)}}{(p^2+m^2)^{\nu+\frac{1}{2}}}\,dp$$
1094: and
1095: $$\frac{1}{z}\frac{d}{dz}\left [ z^\nu K_\nu (z)\right ]=-z^{\nu-1}
1096: K_{\nu-1} (z),$$
1097: where $K_\nu (z)$ is the Macdonald function. The result is
1098: $$I=\frac{\sqrt{\pi}m^3}{\Gamma(\frac{n}{2})2^{\frac{n-1}{2}}} (mz)^{\frac
1099: {n-3}{2}} K_{\frac{n-3}{2}} (mz).$$
1100: Substituting this into (\ref{Az}), we get
1101: $$A(\vec{z};m)=\frac{1}{\sqrt{2}}\left [ \delta(\vec{z})+\frac{\sqrt{\pi}m^3}
1102: {(2\pi)^2}\sum\limits_{n=1}^\infty \frac{a_n}{\Gamma(\frac{n}{2})}\left (
1103: \frac{mz}{2}\right )^{\frac{n-3}{2}}K_{\frac{n-3}{2}} (mz)\right ].$$
1104: But for large arguments \cite{77}
1105: $$K_\nu(z)\approx \sqrt{\frac{\pi}{2z}}e^{-z}\left \{1+\frac{4\nu^2-1}
1106: {8z}\right \}.$$
1107: Therefore the function $A(z;m)$ decays for large $|\vec{x}-\vec{y}|$ as
1108: $e^{-m|\vec{x}-\vec{y}|}$ indicating that the localization provided by the
1109: $\psi^{(s)}_{(\vec{y})}(\vec{x})$ wave-function is no better than $1/m$.
1110:
1111: It seems therefore that the quantum mechanics and relativity create conceptual
1112: problems for Zeno's prescription to observe Achilles and the tortoise
1113: race (check the tortoise position when Achilles reach the position tortoise
1114: occupied at previous step). If we use the Dirac position for this goal, we will
1115: face Zitterbewegung, and if we use the Newton-Wigner one, the contenders will
1116: be not sharply localized. The notion of object's localization, which seems so
1117: benignly obvious in classical mechanics, undergoes a radical change when
1118: distances are smaller than the Compton wavelength of the object. At first sight
1119: the notion of localization at Compton scales appears as purely academic concept
1120: because even for electron the Compton wavelength is very small and the
1121: Zitterbewegung frequency very high, far beyond the present day experimental
1122: accessibility. But this situation can be changed in near future thanks to
1123: narrow-gap semiconductors \cite{78}.
1124:
1125: In such semiconductors the dispersion relation between the energy and the
1126: wavenumber for electrons is analogous to that of relativistic electrons in
1127: vacuum and as a result effective relativity arises with maximum velocity
1128: of about $10^8~\mathrm{cm/sec}$ instead of the velocity of light \cite{78,79}.
1129: Such systems can be used to model various relativistic phenomena the
1130: Zitterbewegung included. The amplitude of the corresponding trembling motion
1131: in semiconductors can be quite large, as much as 64~{\rm \AA} for
1132: InSb \cite{78}, and can be experimentally observed using high-resolution
1133: scanning-probe microscopy imaging techniques \cite{80}.
1134:
1135: \section{Quantum revivals of Zeno}
1136: Achilles is localized not only at the initial instant but remains so during
1137: the whole race. Maybe for Zeno this facet of classical world was not
1138: paradoxical at all but it appears not so trivial at modern times, after the
1139: quantum revolution. In quantum mechanics wave-packets spread as time goes by
1140: and it is not after all obvious how the classical reality with its definiteness
1141: arises from the weird quantum world. It is assumed, usually, that classical
1142: behaviour of macroscopic objects, like Achilles and tortoise, is something
1143: obvious and always guaranteed. But this is not correct. For example, a
1144: cryogenic bar gravity-wave detector must be treated as a quantum harmonic
1145: oscillator even though it may weigh several tons \cite{81}. Superconductivity
1146: and superfluidity provide another examples of quantum behaviour at
1147: macroscopic scales.
1148:
1149: To illustrate surprises that a quantum particle on a racetrack can offer, let
1150: us consider a wave packet of an electron in a hydrogen atom constructed from
1151: a superposition of highly excited stationary states centered at a large
1152: principal quantum number $\bar n =320$ \cite{82}. Initially the wave packet is
1153: well-localized near a point on the electron's classical circular orbit. Then
1154: it propagates around the orbit in accordance with the correspondence principle.
1155: During the propagation the wave packet spreads along the orbit and after
1156: a dozen classical periods $T_{Kepler}$ appears as nearly uniformly distributed
1157: around the orbit (Fig.\ref{kepler})
1158: \begin{figure}[htb]
1159: \begin{center}
1160: \mbox{\epsfig{figure=kepler.eps}}
1161: \end{center}
1162: \caption {Circular-orbit wave packet at initial stages of its time evolution
1163: \cite{82}.
1164: }
1165: \label{kepler}
1166: \end{figure}
1167:
1168: If we wait long enough something extraordinary will happen: after some time
1169: $T_{rev}$ the wave packet contracts and reconstructs its initial form
1170: (Fig.\ref{revival}). This resurrection of the wave packet from the dead is
1171: called a quantum revival and it is closely related to the Talbot effect in
1172: optics \cite{83}. In many circumstances the revivals are almost perfect and
1173: repeat as time goes on. At times $(k_1/k_2)T_{rev}$, where $k_1$ and
1174: $k_2$ are two mutually prime numbers, fractional revivals happen and the
1175: wave packet consists of several high-correlated smaller clones of the original
1176: packet.
1177: \begin{figure}[htb]
1178: \begin{center}
1179: \mbox{\epsfig{figure=revival.eps}}
1180: \end{center}
1181: \caption {Wave-packet revival and fractional revivals \cite{82}.}
1182: \label{revival}
1183: \end{figure}
1184:
1185: To explain the origin of quantum revivals, let us consider a particle of
1186: mass $m$ in an infinite square well of width $L$ \cite{84,85}. The energy
1187: spectrum of the system is given by
1188: $$E_n=\frac{1}{2m}\left (\frac{\pi\hbar}{L}\right)^2 n^2.$$
1189: Suppose the initial wave function is
1190: $$\psi(x;0)=\sum\limits_{n=1}^\infty a_n \phi_n(x),$$
1191: where $\phi_n(x)$ are the energy eigenfunctions. Time evolution of this wave
1192: function is described by
1193: $$\psi(x;t)=\sum\limits_{n=1}^\infty a_n e^{-\frac{i}{\hbar}E_n t}\phi_n(x).$$
1194: Therefore, if there exists such a revival time $T_{rev}$ that
1195: \begin{equation}
1196: \frac{E_n}{\hbar}T_{rev}=2\pi N_n +\varphi
1197: \label{Trev}
1198: \end{equation}
1199: for all nonzero $a_n$, where $N_n$ is an integer that can depend on $n$
1200: and $\varphi$ does not depend on $n$, then $\psi(x;T_{rev})$ will describe
1201: exactly the same state as $\psi(x;0)$.
1202:
1203: But (\ref{Trev}) is fulfilled if
1204: $$T_{rev}=\frac{4mL^2}{\pi\hbar}$$
1205: and $\varphi=0$. Therefore any quantum state in an infinite square well will
1206: be exactly revived after a time $T_{rev}$. Note that the classical period of
1207: bouncing back and forth between the walls is
1208: $$T_{cl}=\sqrt{\frac{2m}{E}}L$$
1209: and for highly excited states ($n\gg 1$)
1210: $$T_{cl}=\frac{2mL^2}{n\pi\hbar}\ll T_{rev}.$$
1211:
1212: Now let us consider the quantum state $\psi(x;t)$ at half of revival time
1213: $$\psi(x;T_{rev}/2)=\sum\limits_{n=1}^\infty a_n e^{-i\pi n^2}\phi_n(x)=
1214: \sum\limits_{n=1}^\infty (-1)^n a_n \phi_n(x),$$
1215: where we have used $e^{-i\pi n^2}=(-1)^n$ identity. But
1216: $$\phi_n(x)=\sqrt{\frac{2}{L}}\sin\frac{n\pi x}{L}=-(-1)^n\phi_n(L-x).$$
1217: Therefore
1218: $$\psi(x;T_{rev}/2)=-\sum\limits_{n=1}^\infty a_n \phi_n(L-x)=
1219: -\psi(L-x;0)$$
1220: and we have the perfect revival of the initial quantum state but at a location
1221: $L-x$ which mirrors the initial position about the center of the well.
1222:
1223: At one quarter of the revival time
1224: $$\psi(x;T_{rev}/4)=\sum\limits_{n=1}^\infty a_n e^{-i\pi n^2/2}\phi_n(x).$$
1225: But
1226: $$e^{-i\pi n^2/2}=\cos{\frac{n^2\pi}{2}}-i\sin{\frac{n^2\pi}{2}}=
1227: \left \{\begin{array}{c} 1,\;\; {\mathrm if} \;\; n \;\;{\mathrm even} \\
1228: -i,\;\; {\mathrm if}\;\; n \;\;{\mathrm odd} \end{array} \right .$$
1229: Therefore
1230: $$\psi(x;T_{rev}/4)=\sum\limits_{n\;even} a_n \phi_n(x)-i
1231: \sum\limits_{n\;odd} a_n \phi_n(x).$$
1232: Comparing this expression to
1233: $$\psi(x;0)=\sum\limits_{n\;even} a_n \phi_n(x)+
1234: \sum\limits_{n\;odd} a_n \phi_n(x)$$
1235: and
1236: $$\psi(L-x;0)=-\sum\limits_{n\;even} a_n \phi_n(x)+
1237: \sum\limits_{n\;odd} a_n \phi_n(x),$$
1238: we deduce that \cite{85}
1239: $$\psi(x;T_{rev}/4)=\frac{1-i}{2}\psi(x;0)-\frac{1+i}{2}\psi(L-x;0).$$
1240: Therefore we have the perfect fractional revival at a time $T_{rev}/4$
1241: when two smaller copies of the initial wave packet appear at locations
1242: $x$ and $L-x$.
1243:
1244: Space-time structure of the probability density $|\psi(x;t)|^2$ is also
1245: very interesting. When plotted over long time periods (of the order of
1246: $T_{rev}$) it exhibits fine interference patterns known as quantum carpets
1247: \cite{86,87}. Some examples are shown in Fig.\ref{carpet}.
1248: \begin{figure}[htb]
1249: \begin{center}
1250: \mbox{\epsfig{figure=carpet.eps}}
1251: \end{center}
1252: \caption {The quantum carpets for the P\"{o}schel-Teller and Rosen-Morse
1253: potentials \cite{85}.}
1254: \label{carpet}
1255: \end{figure}
1256:
1257: Contemporary experimental technique allows to investigate quantum revivals and
1258: carpets and many theoretical results have been confirmed by experiments
1259: \cite{83,88}. Of course, nothing remotely similar to this weird phenomena
1260: happens during the Achilles and the tortoise race. And we come to one more
1261: paradox: why Zeno's paradoxes, formulated purely in classical terms, make
1262: complete sense for us? The answer may sound like this \cite{81}: ``The
1263: environment surrounding a quantum system can, in effect, monitor some of the
1264: systems observables. As a result, the eigenstates of these observables
1265: continuously decohere and can behave like classical states''. Objects have no
1266: a priori classical properties. These properties are emergent phenomenon and
1267: come into being only through the very weak interaction with the ubiquitous
1268: degrees of freedom of the environment. Amazingly, the emergence of the
1269: classical world seems to be just another side of the quantum Zeno effect.
1270: As an example one can consider a large chiral molecule (like sugar) which can
1271: have both left-handed and right-handed classical spatial structures
1272: \cite{89,90}. For symmetry reasons, the ground state is equal mixture of both
1273: chiral states. Chiral molecules are never found in energy eigenstates and this
1274: is probably not surprising because such states are examples of Schr\"{o}dinger
1275: cat states (like a superposition of a dead and an alive cat) which look truly
1276: absurd from classical viewpoint. But the real reason why nonclassical states
1277: of the Chiral molecules are not observed is that it is chirality (not parity)
1278: that is recognized by the environment, for example by scattered air molecules.
1279: The chirality of the molecule is thus continuously ``observed'' by the
1280: environment and therefore cannot change because of quantum Zeno effect.
1281:
1282: \section{Zeno meets quantum gravity}
1283: Another twist to the story of localization is added when one tries to
1284: incorporate gravity into a quantum theory. Any sharp localization of the
1285: system creates a significant local energy density due to uncertainty
1286: relations and therefore changes the space-time metric according to the
1287: philosophy of general relativity. This can effect another localization
1288: effort nearby in an unavoidable manner \cite{91} and as a result it will
1289: matter whether x-position measurement is carried before or after y-position
1290: measurement. Moreover, localization sharper than the Plankian scale creates a
1291: singularity in the space-time metric and therefore is problematic. We expect
1292: consequently that any coherent quantum gravity theory will not only bring the
1293: fundamental Plankian scale as the limit of space-time divisibility with it,
1294: but also a non-commutative space-time geometry.
1295:
1296: Not only the concept of localization of material objects but also operational
1297: meaning of the space-time itself is expected to be lost at Plankian scales.
1298: The principles of quantum mechanics and general relativity limit the accuracy
1299: of space-time distance measurements. The argument goes as follows \cite{92}.
1300:
1301: Suppose we want to measure the initial distance between Achilles and the
1302: tortoise. We can attach a small mirror to the tortoise while a clock with
1303: light-emitter and receiver to Achilles. When the clock reads zero a light
1304: signal is sent to be reflected by the mirror. If the reflected signal arrives
1305: back at time $t$ then the distance is $l=ct/2$. Note that this is a quite
1306: realistic scheme used, for example, in Lunar Laser Ranging experiments
1307: \cite{93}. Quantum mechanics sets some limits on ultimate precision which
1308: can be reached in such distance measurements. Let, for example, the initial
1309: uncertainty in the clock's position be $\Delta x$. Then according to
1310: uncertainty relation its speed is also uncertain with the spread
1311: $$\Delta v\ge \frac{1}{2}\frac{\hbar}{m\Delta x}.$$
1312: where $m$ is the mass of the clock. At time $t$ the uncertainty of the
1313: clock's position will be larger
1314: $$\Delta x(t)= \Delta(x(0)+vt)=\sqrt{(\Delta x)^2+(t\Delta v)^2}=
1315: \sqrt{(\Delta x)^2+\frac{1}{4}\frac{t^2\hbar^2}{m^2(\Delta x)^2}}.$$
1316: The optimal uncertainty at initial time that minimizes the uncertainty at time
1317: $t$ is given by
1318: $$(\Delta x)^2=\frac{1}{2}\frac{t\hbar}{m}.$$
1319: Therefore the minimal uncertainty at time $t$ is
1320: $$\Delta x(t)=\sqrt{\frac{t\hbar}{m}}=\sqrt{\frac{2l\hbar}{mc}}
1321: =\lambda_C\sqrt{\frac{2l}{\lambda_C}},$$
1322: where $\lambda_C=\hbar/mc$ is the Compton wavelength of the clock. The mirror
1323: contributes similarly to the overall uncertainty of the distance $l$ to be
1324: measured. Therefore, ignoring small factors of the order of 2, we obtain an
1325: order of magnitude estimate \cite{92,94} (the argument actually goes back to
1326: Wigner \cite{95,96})
1327: \begin{equation}
1328: \Delta l \ge \lambda_C\sqrt{\frac{l}{\lambda_C}}.
1329: \label{Deltal}
1330: \end{equation}
1331: Hence we need massive clock to reduce the uncertainty. But the mass of the
1332: clock cannot be increased indefinitely because the distance $l$ should be
1333: greater than the clock's Schwarzschild radius. Otherwise the clock will
1334: collapse into a black hole as it is assumed that its size is smaller than $l$.
1335: Therefore
1336: $$l\ge \frac{Gm}{c^2}$$
1337: and inserting this into (\ref{Deltal}) we get
1338: \begin{equation}
1339: \Delta l \ge l_P,
1340: \label{DeltaP}
1341: \end{equation}
1342: where
1343: $$l_P=\sqrt{\frac{\hbar G}{c^3}}$$
1344: is the Planck length.
1345:
1346: Ng and Van Dam argue even for a more restrictive bound \cite{92}. suppose
1347: the clock consists of two parallel mirrors a distance $d$ apart. Then its one
1348: tick cannot be less than $d/c$ and this implies $\Delta l \ge d$ if the clock
1349: is used for timing in distance measurements. But $d$ should be greater than
1350: the clock's Schwarzschild radius. Therefore
1351: \begin{equation}
1352: \Delta l \ge \frac{Gm}{c^2}.
1353: \label{DeltaG}
1354: \end{equation}
1355: Squaring (\ref{Deltal}) and multiplying the result by (\ref{DeltaG}), we
1356: eliminate the clock's mass and obtain $(\Delta l)^3\ge l l_P^2$. Therefore
1357: \begin{equation}
1358: \Delta l \ge l_P\left (\frac{l}{l_P}\right)^{1/3}.
1359: \label{DeltalP}
1360: \end{equation}
1361:
1362: It seems there is a common consensus about the validity of (\ref{DeltaP}).
1363: While the more stronger bound (\ref{DeltalP}) is still under debate (see,
1364: for example, \cite{94}). Nevertheless this latter bound is consistent with the
1365: holographic principle \cite{97} which states that the information content of
1366: any region of space cannot exceed its surface area in Planck units. Indeed
1367: \cite{98}, suppose we have a cube of dimension $l\times l\times l$ and every
1368: cell $\Delta l\times \Delta l\times \Delta l$ of this cube can be used to
1369: store one bit of information. $\Delta l$ cannot be made less than dictated
1370: by the bound (\ref{DeltalP}), otherwise it will be possible to measure the
1371: size of the cube in $\Delta l$ units and reach the precision superior to the
1372: limit (\ref{DeltalP}). Therefore the information content of the cube
1373: $$N=\frac{l^3}{(\Delta l)^3}\le \frac{l^2}{l_P^2}.$$
1374:
1375: In any case, we can conclude that basic principles of quantum mechanics and
1376: general relativity strongly suggest discreteness of space-time at some
1377: fundamental scale. Zeno anticipated such possibility and attacked it with
1378: another couple of paradoxes. The first one, the Arrow, states that \cite{2}
1379: \begin{itemize}
1380: \item
1381: {\it If everything is either at rest or moving when it occupies a space equal
1382: to itself, while the object moved is always in the instant, a moving arrow is
1383: unmoved.}
1384: \end{itemize}
1385: Stated differently, if a particle exists only at a sequence of discrete
1386: instants of time, what is the instantaneous physical properties of a moving
1387: particle which distinguishes it from the not moving one? And if there is no
1388: such properties (well, the notion of instantaneous velocity requires the
1389: concept of limit and thus is inappropriate in discrete space-time) how the
1390: motion is possible?
1391:
1392: Modern physics changed our perspective of particles and motion and the Arrow
1393: seems not so disturbing today. For example we can defy it by stating that the
1394: arrow cannot be at rest at definite position according to the uncertainty
1395: principle. Alternatively, we can evoke special relativity and say that there
1396: is a difference how the world looks for a moving arrow and for an arrow at
1397: rest \cite{2}: they have different planes of simultaneity. Special relativity
1398: can cope also with the second Zeno paradox against space-time discreteness,
1399: the Stadium:
1400: \begin{itemize}
1401: \item
1402: {\it Consider two rows of bodies, each composed of an equal number of bodies
1403: of equal size. They pass each other as they travel with equal velocity in
1404: opposite directions. Thus, half a time is equal to the whole time.}
1405: \end{itemize}
1406: If motion takes place in discrete quantum jumps then there should be an
1407: absolute upper bound on velocity. The maximum velocity is achieved
1408: then all jumps are in the same direction. The Stadium is intended to show
1409: logical impossibility of the maximum relative velocity. Suppose the rows
1410: of bodies from the paradox move at maximum or nearly maximum speed. Then
1411: in the rest frame of the first row other bodies are approaching at twice or
1412: so the maximum possible speed. Now we know that the latter inference is
1413: not sound. But Zeno was right that radical change of classical concepts of
1414: space and time is necessary to assimilate an observer-independent maximum
1415: velocity. ``Space by itself and time by itself are to sink fully into shadows
1416: and only a kind of union of the two should yet preserve autonomy'' - to quote
1417: Minkowski from his famous Cologne lecture in 1908.
1418:
1419: In fact space-time discreteness coupled with the relativity principle assumes
1420: two invariant scales: not only the maximum velocity but also the minimum
1421: length. In this respect the special relativity refutation of Zeno is not
1422: complete. Only recently a significant effort was invested in developing
1423: Doubly Special Relativity \cite{99,100} - relativity theories with two
1424: observer-independent scales. These developments, although interesting, still
1425: are lacking experimental confirmation.
1426:
1427: Doubly Special Relativity, if correct, should be a limiting case of quantum
1428: gravity - an ultimate theory of quantum space-time and a major challenge of
1429: contemporary physics to combine general relativity with quantum mechanics.
1430: This synthesis is not achieved yet but curiously enough its outcome may
1431: turn out to vindicate Parmenidean view that time and change are not
1432: fundamental reality. In any case the problem of time seems to be central in
1433: quantum gravity because time plays conceptually different roles in quantum
1434: mechanics and general relativity not easy to reconcile \cite{101,102,103,104,
1435: 105}.
1436:
1437: In quantum mechanics time is an external parameter, not an observable in the
1438: usual sense -- it is not represented by an operator. Indeed, suppose there is
1439: an operator $\hat T$, representing a perfect clock, such that in the
1440: Heisenberg picture $\hat T(t)=e^{i\hat Ht}\hat T e^{-i\hat Ht}=t$. Then
1441: (we are using again $\hbar=1$ convention)
1442: $$i[\hat H,\hat T(t)]=\frac{d}{dt}\hat T(t)=1.$$
1443: Therefore $[\hat T(t),\hat H]=i$ and $[\hat H,e^{i\alpha\hat T}]=\alpha
1444: e^{i\alpha\hat T}$. If $\psi$ is an energy eigenstate with energy $E$ then
1445: $$\hat H e^{i\alpha\hat T}\psi=\left \{ [\hat H, e^{i\alpha\hat T}]+
1446: e^{i\alpha\hat T} \hat H\right \}\psi=(E+\alpha)e^{i\alpha\hat T}\psi.$$
1447: Therefore $e^{i\alpha\hat T}\psi$ is also the energy eigenstate with energy
1448: $E+\alpha$ and the Hamiltonian spectrum cannot be bounded from below as it
1449: usually is. This argument goes back to Pauli \cite{106}. Alternatively we can
1450: resort to the Stone-von Neumann theorem \cite{107} and argue that the
1451: canonically conjugate $\hat T$ and $\hat H$ are just the disguised versions of
1452: the position and momentum operators and therefore must have unbounded spectra.
1453:
1454: However, there are subtleties in both the Pauli's argument \cite{106}
1455: as well as
1456: in the Stone-von Neumann theorem \cite{108} and various time operators are
1457: suggested occasionally. Note that many useful concepts in physics are
1458: ambiguous or even incorrect from the mathematical point of view. In an
1459: extremal manner this viewpoint was expressed by Dieudonn\'{e} \cite{109}:
1460: "When one gets to the mathematical theories which are at the basis of quantum
1461: mechanics, one realizes that the attitude of certain physicists in the
1462: handling of these theories truly borders on the delirium''. Of course, this
1463: is an exaggeration but sometimes mathematical refinement leads to new physical
1464: insights and it is not excluded that the last word about the time operator is
1465: not said yet. In any case, subtle is the time in quantum mechanics!
1466:
1467: The idea of an event happening at a given time plays a crucial role in quantum
1468: theory \cite{110} and at first sight introduces unsurmountable difference
1469: between space and time. For example \cite{101}, if $\Psi(\vec{x},t)$ is a
1470: normalized wave function then $\int |\Psi(\vec{x},t)|^2 d\vec{x}=1$ for all
1471: times, because the particle must be somewhere in space at any given instant of
1472: time. While $\int |\Psi(\vec{x},t)|^2 dt$ can fluctuate wildly for various
1473: points of space.
1474:
1475: Nevertheless, the conceptual foundation of quantum theory is compatible with
1476: special relativity. Absolute Newtonian time is simply replaced by Minkowski
1477: spacetime fixed background and unitary representations of the Poincar\'{e}
1478: group can be used to develop quantum theory. Situation changes dramatically
1479: in general relativity. ``There is hardly any common ground between the general
1480: theory of relativity and quantum mechanics'' \cite{95}. The central problem is
1481: that spacetime itself becomes a dynamical object in general relativity. Not
1482: only matter is influenced by the structure of spacetime but the metric
1483: structure of spacetime depends on the state of ambient matter. As a result,
1484: the spatial coordinate $\vec{x}$ and the temporal coordinate $t$ lose any
1485: physical meaning whatsoever in general relativity. If in
1486: non-general-relativistic physics (including special relativity) the
1487: coordinates correspond to readings on rods and clocks, in general relativity
1488: they correspond to nothing at all and are only auxiliary quantities which
1489: can be given arbitrary values for every event \cite{95,111}.
1490:
1491: Already at classical level, general relativity is a great deal Parmenidean
1492: and usually some tacit assumptions which fix the coordinate system is needed
1493: to talk meaningfully about time and time-evolution. To quote Wigner \cite{95}
1494: ``Evidently, the usual statements about future positions of particles, as
1495: specified by their coordinates, are not meaningful statements in general
1496: relativity. This is a point which cannot be emphasized strongly enough and is
1497: the basis of a much deeper dilemma than the more technical question of the
1498: Lorentz invariance of the quantum field equations. It pervades all the general
1499: theory, and to some degree we mislead both our students and ourselves when we
1500: calculate, for instance, the mercury perihelion motion without explaining how
1501: our coordinate system is fixed in space, what defines it in such a way that it
1502: cannot be rotated, by a few seconds a year, to follow the perihelion's
1503: apparent motion''.
1504:
1505: In quantum theory the situation only worsens. The Hamiltonian generates the
1506: time evolution of quantum system. But the equations of motion of general
1507: relativity are invariant under time reparametrization. Therefore the time
1508: evolution is in fact unobservable - it is a gauge. The Hamiltonian vanishes
1509: and the Schr\"{o}dinger equation in cosmology - the Wheeler-DeWitt equation
1510: does not contain time. ``General relativity does not describe evolution in
1511: time: it describes the relative evolution of many variables with respect to
1512: each other'' \cite{111}. Zeno and Parmenides with their strange idea that
1513: time and change are some kind of illusion still have chance in quantum
1514: gravity!
1515:
1516: \section{Conclusion}
1517: The main conclusion of this paper is that physics is beautiful. Questions
1518: aroused two and half millennium ago and scrutinized many times are still not
1519: exhausted. Zeno's paradoxes deal with fundamental aspects of reality like
1520: localization, motion, space and time. New and unexpected facets of these
1521: notions come into sight from time to time and every century finds it
1522: worthwhile to return to Zeno over and over. The process of approaching to
1523: the ultimate resolution of Zeno's paradoxes seems endless and our
1524: understanding of the surrounding world is still incomplete and fragmentary.
1525:
1526: ``Nevertheless, I believe that there is something great in astronomy, in
1527: physics, in all the natural sciences that allows the human being to look
1528: beyond its present place and to arrive at some understanding of what goes on
1529: beyond the insignificant meanness of spirit that so often pervades our
1530: existence. There is a Nature; there is a Cosmos; and we walk towards the
1531: understanding of it all. Is it not wonderful? There are many charms in the
1532: profession; as many charms as in love provided, of course, that they are not
1533: in the service of mercantile aims'' \cite{112}.
1534:
1535:
1536: \begin{thebibliography}{99}
1537: \bibitem{1}
1538: Whitehead, A.N. {\it Essays in Science and Philosophy};
1539: Philosophical Library: New York, 1948; p. 87.
1540: \bibitem{2}
1541: Brown, K., {\it Reflections on Relativity}; pp. 210 - 216.
1542: \newline http://www.mathpages.com/rr/rrtoc.htm
1543: \bibitem{3}
1544: Diogenes Laertius, {\it The Lives and Opinions of Eminent
1545: Philosophers}, translated by C.D. Yonge; Henry G. Bohn: London, 1853;
1546: \newline http://classicpersuasion.org/pw/diogenes/dlzeno-eleatic.htm
1547: \bibitem{4}
1548: O'Connor, J.~J., \& Robertson, E.~F. {\it Zeno of Elea};\newline
1549: http://www-groups.dcs.st-and.ac.uk/$\sim$history/Mathematicians/
1550: Zeno\_of\_Elea.html
1551: \bibitem{5}
1552: Burnet, J. {\it Early Greek Philosophy}; Black: London, 1930.
1553: \bibitem{6}
1554: Makin, S. {\it Zeno of Elea}; Routledge Encyclopedia of Philosophy; Routledge:
1555: London, 1998; Vol. 9, pp. 843-853.
1556: \bibitem{7}
1557: Fairbanks, A. {\it The First Philosophers of Greece};
1558: K. Paul, Trench, Trubner: London, 1898; p. 112. \newline
1559: http://history.hanover.edu/texts/presoc/zeno.htm
1560: \bibitem{8}
1561: Cajori, F. {\it History of Zeno's Arguments on Motion}; Amer. Math. Monthly
1562: 1915, Vol. 22, pp. 1-6; 39-47; 77-82; 109-115; 143-149; 179-186; 215-220;
1563: 253-258; 292-297.
1564: %%CITATION = AMMYA,22,1;%%
1565: %%CITATION = AMMYA,22,39;%%
1566: %%CITATION = AMMYA,22,77;%%
1567: %%CITATION = AMMYA,22,109;%%
1568: %%CITATION = AMMYA,22,143;%%
1569: %%CITATION = AMMYA,22,179;%%
1570: %%CITATION = AMMYA,22,215;%%
1571: %%CITATION = AMMYA,22,253;%%
1572: %%CITATION = AMMYA,22,292;%%
1573: \bibitem{9}
1574: Grunbaum, A. {\it Modern Science and Zeno's Paradoxes};
1575: George Allen and Unwin Ltd.: London, 1968.
1576: \bibitem{10}
1577: Salmon, W.~C. {\it Zeno's Paradoxes}; Bobbs Merrill: New York, 1970.
1578: \bibitem{11}
1579: Huggett, N. {\it Zeno's Paradoxes}; The Stanford Encyclopedia of Philosophy
1580: (Summer 2004 Edition), Edward N.~Zalta (ed.); \newline
1581: http://plato.stanford.edu/archives/sum2004/entries/paradox-zeno/
1582: \bibitem{12}
1583: Alexander Pushkin {\it Speaking In Tongues Guided by Voices}; Translated by
1584: Andrey Kneller; \newline
1585: http://spintongues.vladivostok.com/pooshkin3.htm
1586: \bibitem{13}
1587: Aristotle {\it Physics}; Book 8; Translated by R. P. Hardie and R. K. Gaye;
1588: \newline http://www.abu.nb.ca/Courses/GrPhil/Physics.htm
1589: \bibitem{14}
1590: McLaughlin, W.~I., \& Miller, S.~L. {\it An Epistemological Use of
1591: Nonstandard Analysis to Answer Zeno's Objections against Motion}; Synthese
1592: 1992, Vol. 92, pp.371-384.
1593: %%CITATION = SYNTA,92,371;%%
1594: \bibitem{15}
1595: McLaughlin, W. I. {\it Resolving Zeno's Paradoxes}; Scientific American 1994,
1596: Vol. 271, November issue, pp. 84-89.
1597: %%CITATION = SCAMA,271N5,84;%%
1598: \bibitem{16}
1599: Bell, J.~L. {\it The Art of the Intelligible: An Elementary Survey of
1600: Mathematics in Its Conceptual Development}; Kluwer Academic Pub.:
1601: Dordrecht, The Netherland, 1999; p. 175.
1602: \bibitem{17}
1603: Blumenthal, L.~M. {\it A Paradox, a Paradox, a Most Ingenious Paradox};
1604: Amer. Math. Monthly 1940, Vol. 47, pp. 346-353.
1605: %%CITATION = AMMYA,47,346;%%
1606: \bibitem{18}
1607: Runde, V. {\it The Banach-Tarski paradox or What mathematics and miracles
1608: have in common}; $\pi$ in the Sky 2000, Vol. 2, pp. 13-15.
1609: \newline http://ru.arxiv.org/abs/math.GM/0202309
1610: \bibitem{19}
1611: Wagon, S. {\it The Banach-Tarski Paradox}; Cambridge University Press:
1612: New York, 1993.
1613: \bibitem{20}
1614: Stromberg, K. {\it The Banach-Tarski paradox}; Amer. Math. Monthly 1979,
1615: Vol. 86, pp. 151-161.
1616: %%CITATION = AMMYA,86,151;%%
1617: \bibitem{21}
1618: Pitowsky, I. {\it Resolution of the Einstein-Podolsky-Rosen and Bell
1619: Paradoxes}; Phys. Rev. Lett. 1982, Vol. 48, pp. 1299-1302; 1768;
1620: Vol. 49, p. 1216.
1621: %%CITATION = PRLTA,48,1299;%%
1622: %%CITATION = PRLTA,49,1216;%%
1623: \bibitem{22}
1624: Pitowsky, I. {\it Deterministic model of spin and statistics};
1625: Phys. Rev. D 1983, Vol. 27, pp. 2316-2326.
1626: %%CITATION = PHRVA,D27,2316;%%
1627: \bibitem{23}
1628: Svozil, K., \& Neufeld, N. {\it 'Linear' chaos via paradoxical set
1629: decompositions}; Chaos, Solitons \& Fractals 1996, Vol. 7(5), pp. 785-793.
1630: %%CITATION = CSFOE,7,785;%%
1631: \bibitem{24}
1632: Augenstein, B.~W. {\it Links Between Physics and Set Theory};
1633: Chaos, Solitons \& Fractals 1996, Vol. 7(11), pp. 1761-1798.
1634: %%CITATION = CSFOE,7,1761;%%
1635: \bibitem{25}
1636: Dougherty, R., \& Foreman, M. {\it Banach-Tarski Decompositions Using Sets
1637: with the Property of Baire}; Journal of the American Mathematical
1638: Society 1994, Vol.7(1), pp.75-124.
1639: %%CITATION = 00052,7N1,75;%%
1640: \bibitem{26}
1641: O'Connor, J.~J., \& Robertson, E.~F. {\it Georg Ferdinand Ludwig Philipp
1642: Cantor}; \newline
1643: http://www-groups.dcs.st-andrews.ac.uk/~history/Mathematicians/Cantor.html
1644: \bibitem{27}
1645: Peres, A. {\it Zeno paradox in quantum theory}; Am. J. Phys. 1980,
1646: Vol. 48, pp. 931-932.
1647: %%CITATION = AJPIA,48,931;%%
1648: \bibitem{28}
1649: Wallace, D. {\it Simple computer model for the quantum Zeno effect};
1650: Phys. Rev. A 2001, Vol. 63, 022109.
1651: %%CITATION = PHRVA,A63,022109;%%
1652: \bibitem{29}
1653: Misra, B., \& Sudarshan, E.~C.~G. {\it The Zeno's paradox in quantum theory};
1654: J. Math. Phys. 1977, Vol. 18, pp. 756763.
1655: %%CITATION = JMAPA,18,756;%%
1656: \bibitem{30}
1657: Home, D., \& Whitaker, M.~A.~B. {\it A Conceptual Analysis of Quantum Zeno;
1658: Paradox, Measurement, and Experiment}; Annals of Physics 1997, Vol. 258,
1659: pp. 237-285.
1660: %%CITATION = APNYA,258,237;%%
1661: \bibitem{31}
1662: Bokulich, A. {\it Quantum measurements and supertasks}; International Studies
1663: in the Philosophy of Science 2003, Vol. 17, pp. 127-136.
1664: %%CITATION = 00249,17,127;%%
1665: \bibitem{32}
1666: Facchi, P., \& Pascazio, S. {\it Temporal behavior and quantum Zeno time of
1667: an excited state of the hydrogen atom}; Phys. Lett. A 1998, Vol. 241,
1668: pp. 139-144.
1669: %%CITATION = QUANT-PH 9905017;%%
1670: \bibitem{33}
1671: Fischer, M.~C., Guti\'{e}rrez-Medina, B., \& M. G. Raizen, M.~G.
1672: {\it Observation of the Quantum Zeno and Anti-Zeno Effects in an Unstable
1673: System}; Phys. Rev. Lett. 2001, Vol. 87, 040402.
1674: %%CITATION = PRLTA,87,040402;%%
1675: \bibitem{34}
1676: Home, D., \& Whitaker, M.~A.~B. {\it Reflections on the quantum Zeno paradox};
1677: J. Phys. A: Math. Gen. 1986, Vol. 19, pp. 1847-1854.
1678: %%CITATION = JPAGB,A19,1847;%%
1679: \bibitem{35}
1680: Kaulakys, B., \& Gontis, V. {\it Quantum anti-Zeno effect};
1681: Phys. Rev. A 1997, Vol. 56, pp. 1138-1141.
1682: %%CITATION = PHRVA,A56,1138;%%
1683: \bibitem{36}
1684: Kofman, A.~G., \& Kurizki, G. {\it Quantum Zeno effect on atomic excitation
1685: decay in resonators}; Phys. Rev. A 1996, Vol. 54, pp. R3750-R3753.
1686: %%CITATION = PHRVA,A54,R3750;%%
1687: \bibitem{37}
1688: Balachandran, A.~P., \& Roy, S.~M. {\it Quantum anti Zeno paradox};
1689: Phys. Rev. Lett. 2000, Vol. 84, pp. 4019-4022.
1690: %%CITATION = PRLTA,84,4019;%%
1691: \bibitem{38}
1692: Franson, J.~D., Jacobs, B.~C., \& Pittman, T.~B. {\it Quantum computing using
1693: single photons and the Zeno effect}; Phys. Rev. A 2004, Vol. 70, 062302.
1694: %%CITATION = PHRVA,A70,062302;%%
1695: \bibitem{39}
1696: Schmidt, A.~U. {\it Mathematics of the Quantum Zeno Effect}; \newline
1697: http://arxiv.org/abs/math-ph/0307044
1698: \bibitem{40}
1699: Thomson, J. {\it Tasks and super-tasks}; Analysis 1954, Vol.15, pp. 1-13.
1700: \bibitem{41}
1701: Benacerraf, P. {\it Tasks, super-tasks, and the modern eleatics}; Journal of
1702: Philosophy 1962, Vol.59, pp. 765-784.
1703: %%CITATION = JPHIB,59,765;%%
1704: \bibitem{42}
1705: Earman, J., \& Norton, J.~D. {\it Infinite pains: the trouble with
1706: supertasks}; in: Adam Morton and Stephen P. Stich (Eds.) {\it Benacerraf
1707: and his Critics}; Blackwell: Oxford, 1996; pp. 231-261.
1708: \bibitem{43}
1709: P\'{e}rez Laraudogoitia, J. {\it A Beautiful Supertask};
1710: Mind 1996, Vol. 105, pp. 81-83.
1711: %%CITATION = JMNDDA,105,81;%%
1712: \bibitem{44}
1713: Hamkins, J.~D. {\it Supertask Computation}; \newline
1714: http://arxiv.org/abs/math.LO/0212049
1715: \bibitem{45}
1716: P\'{e}rez Laaraudogoitia J., Bridger M., \& Alper J.~S. {it Two Ways Of
1717: Looking At A Newtonian Supertask}; Synthese 2002, Vol. 131, pp. 173-189.
1718: %%CITATION = SYNTA,131,173;%%
1719: \bibitem{46}
1720: Alper, J.~S., Bridger, M., Earman, J., \& Norton, J.~D. {\it What is a
1721: Newtonian System? The Failure of Energy Conservation and Determinism in
1722: Supertasks}; Synthese 2000, Vol. 124, pp. 281-293.
1723: %%CITATION = SYNTA,124,281;%%
1724: \bibitem{47}
1725: Earman, J., \& Norton, J.~D. {\it Comments on Laraudogoitia's 'Classical
1726: Dynamics, Indeterminism and a Supertask'}; British Journal for the Philosophy
1727: of Science 1998, Vol. 49, pp.123-133.
1728: %%CITATION = BJPIA,49,123;%%
1729: \bibitem{48}
1730: Norton, J.~D. {\it A Quantum Mechanical Supertask};
1731: Foundations of Physics 1999, Vol. 29, pp. 1265-1302.
1732: %%CITATION = FNDPA,29,1265;%%
1733: \bibitem{49}
1734: Earman, J. {\it Determinism: What We Have Learned and What We Still Don't
1735: Know}; in: Joseph Keim Campbell, Michael O'Rourke and David Shier (Eds.)
1736: {\it Freedom and Determinism}; MIT Press: Cambridge, 2004; pp. 21-46.
1737: \bibitem{50}
1738: Xia, Z. {\it The Existence of Noncollision Singularities in Newtonian
1739: Systems}; Ann. of Math. 1992, Vol. 135, pp. 411-468.
1740: %%CITATION = ANMAA,135,411;%%
1741: \bibitem{51}
1742: Saari, D.~G., \& Xia, Z. {\it Off to infinity in finite time};
1743: AMS Notices 1995, Vol. 42, pp. 538-546.
1744: \bibitem{52}
1745: Korycansky,D.~G., Laughlin, G., \& Adams, F.~C. {\it Astronomical engineering:
1746: a strategy for modifying planetary orbits};
1747: Astrophys. Space Sci. 2001, Vol. 275, pp. 349-366.
1748: %%CITATION = ASTRO-PH 0102126;%%
1749: \bibitem{53}
1750: Weyl, H. {\it Philosophy of Mathematics and Natural Science};
1751: Princeton University Press: Princeton, 1949; p. 42.
1752: \bibitem{54}
1753: Earman, J., \& Norton, J.~D. {\it Forever is a Day: Supertasks in Pitowsky and
1754: Malament-Hogarth Spacetimes}; Philosophy of Science 1993, Vol. 60, pp. 22-42.
1755: %%CITATION = PHSCA,60,22;%%
1756: \bibitem{55}
1757: Pitowsky, I. {\it The Physical Church's Thesis and Physical Computational
1758: Complexity}; Iyun 1990, Vol. 39, pp. 81-99.
1759: \bibitem{56}
1760: Hogarth, M. {\it Does General Relativity Allow an Observer to View an Eternity
1761: in a Finite Time?} Foundations of Physics Letters 1992, Vol. 5, pp. 173-181.
1762: %%CITATION = FPLEE,5,173;%%
1763: \bibitem{57}
1764: Etesi, G., \& Nemeti, I. {\it Non-Turing computations via Malament-Hogarth
1765: space-times}; Int. J. Theor. Phys. 2002, Vol. 41, pp. 341-370.
1766: %%CITATION = IJTPB,41,341;%%
1767: \bibitem{58}
1768: Adams, F.~C., \& Laughlin, G. {\it A Dying universe: The Long term fate and
1769: evolution of astrophysical objects}; Rev. Mod. Phys. 1997, Vol. 69, 337-372.
1770: %%CITATION = ASTRO-PH 9701131;%%
1771: \bibitem{59}
1772: Krauss, L.~M., \& Starkman, G.~D. {\it Life, The Universe, and Nothing: Life
1773: and Death in an Ever-Expanding Universe};
1774: Astrophys. J. 2000, Vol. 531, pp. 22-30.
1775: %%CITATION = ASTRO-PH 9902189;%%
1776: \bibitem{60}
1777: Freese, K., \& Kinney, W.~H. {\it The ultimate fate of life in an accelerating
1778: universe}; Phys. Lett. B 2003, Vol. 558, pp. 1-8.
1779: %%CITATION = ASTRO-PH 0205279;%%
1780: \bibitem{61}
1781: Lloyd, S.{\it Ultimate physical limits to computation};
1782: Nature 2000, Vol. 406, pp. 1047-1054.
1783: %%CITATION = NATUA,406,1047;%%
1784: \bibitem{62}
1785: Plenio, M.~B., Vitelli, V. {\it The physics of forgetting: Landauer's erasure
1786: principle and information theory};
1787: Contemporary Physics 2001, Vol. 42, pp. 25-60.
1788: %%CITATION = CTPHA,42,25;%%
1789: \bibitem{63}
1790: Leff, H.~S., \& Rex, A.~F. (Eds.) {\it Maxwell's Demon 2: Entropy, Classical
1791: and Quantum Information, Computing}; IOP Publishing: Bristol, 2003.
1792: \bibitem{64}
1793: Bennett, C.~H. {\it Notes on Landauer's Principle, Reversible Computation, and
1794: Maxwell's Demon}; Studies in History and Philosophy of Modern Physics 2003,
1795: Vol. 34, pp. 501-510.
1796: %%CITATION = 00218,34,501;%%
1797: \bibitem{65}
1798: Dyson, F.~J. {\it Time without end: Physics and biology in an open universe};
1799: Rev. Mod. Phys. 1979, Vol. 51, pp. 447-460.
1800: %%CITATION = RMPHA,51,447;%%
1801: \bibitem{66}
1802: Newton, T.~D., \& Wigner, E.~P. {\it Localized States For Elementary Systems};
1803: Rev. Mod. Phys. 1949, Vol. 21, pp. 400-406.
1804: %%CITATION = RMPHA,21,400;%%
1805: \bibitem{67}
1806: Wightman, A.~S {\it On The Localizability Of Quantum Mechanical Systems};
1807: Rev. Mod. Phys. 1962, Vol. 34, 845-872.
1808: %%CITATION = RMPHA,34,845;%%
1809: \bibitem{68}
1810: Halvorson, H. {Reeh-Schlieder Defeats Newton-Wigner: On alternative
1811: localization schemes in relativistic quantum field theory};
1812: Philosophy of Science 2001, Vol. 68, pp. 111-133.
1813: %CITATION = PHSCA,68,111;%%
1814: \bibitem{69}
1815: Fleming, G.~N. {\it Reeh-Schlieder meets Newton-Wigner};
1816: Philosophy of Science 2000, Vol. 67, pp. S495-S515.
1817: %CITATION = PHSCA,67,S495;%%
1818: \bibitem{70}
1819: Halvorson, H. {\it Locality, Localization, and the Particle Concept: Topics
1820: in the Foundations of Quantum Field Theory}; \newline
1821: http://philsci-archive.pitt.edu/archive/00000346/
1822: \bibitem{71}
1823: Silagadze, Z.~K. {\it The Newton-Wigner position operator and the domain of
1824: validity of one particle relativistic theory};
1825: Preprint SLAC-PUB-5754-REV, 1993.
1826: \bibitem{72}
1827: Foldy, L.~L., \& Wouthuysen, S.~A. {\it On The Dirac Theory Of Spin 1/2
1828: Particle And Its Nonrelativistic Limit};
1829: Phys. Rev. 1950, Vol. 78, pp. 29-36.
1830: %%CITATION = PHRVA,78,29;%%
1831: \bibitem{73}
1832: Feshbach, H., \& Villars, F. {\it Elementary Relativistic Wave Mechanics Of
1833: Spin 0 And Spin 1/2 Particles}; Rev. Mod. Phys. 1958, Vol. 30, pp. 24-45.
1834: %%CITATION = RMPHA,30,24;%%
1835: \bibitem{74}
1836: de Vries E. {it Foldy-Wouthuysen transformations and related Problems};
1837: Fortsch. der Physik 1970, Vol. 18, pp. 149-182.
1838: %%CITATION = FPYKA,18,149;%%
1839: \bibitem{75}
1840: Lock, J.~A. {\it The Zitterbewegung Of A Free Localized Dirac Particle};
1841: Am. J. Phys. 1979, Vol. 47, 797-802.
1842: %%CITATION = AJPIA,47,797;%%
1843: \bibitem{76}
1844: Barut, A.~O., \& Cruz, M.~G. {\it On the Zitterbewegung of the relativistic
1845: electron}; Eur. J. Phys. 1994, Vol. 15, pp. 119-120.
1846: %%CITATION = EJPHD,15,119;%%
1847: \bibitem{77}
1848: Abramowitz, M., \& Stegun, I.~A. (Eds.) {\it Handbook of Mathematical
1849: Functions With Formulas, Graphs, and Mathematical Tables};
1850: U.S. Government Printing Office: Washington, 1972; pp. 374-378.
1851: \bibitem{78}
1852: Zawadzki, W. {\it Zitterbewegung and its effects on electrons in
1853: semiconductors}; \newline
1854: http://arxiv.org/abs/cond-mat/0411488
1855: %%CITATION = COND-MAT 0411488;%%
1856: \bibitem{79}
1857: Zawadzki, W., Klahn, S., \& Merkt, U. {\it Semirelativistic Behavior of
1858: Electrons in InSb in Crossed Magnetic and Electric Fields};
1859: Phys. Rev. Lett. 1985, Vol. 55, pp. 983-986.
1860: %%CITATION = PRLTA,55,983;%%
1861: \bibitem{80}
1862: Schliemann, J., Loss, D.,\& Westervelt, R.~W. {\it Zitterbewegung of
1863: electronic wave packets in semiconductor nanostructures}; \newline
1864: http://arxiv.org/abs/cond-mat/0410321
1865: %%CITATION = COND-MAT 0410321;%%
1866: \bibitem{81}
1867: Zurek, W.~H. {\it Decoherence and the transition from quantum to classical};
1868: Phys. Today 1991, Vol. 44, pp. 36-44. \newline
1869: http://xxx.lanl.gov/abs/quant-ph/0306072
1870: %%CITATION = PHTOA,44,36;%%
1871: %%CITATION = QUANT-PH 0306072;%%
1872: \bibitem{82}
1873: Gaeta, Z.~D., \& Stroud, C.~R. (Jr.) {\it Classical and quantum-mechanical
1874: dynamics of a quasiclassical state of the hydrogen atom};
1875: Phys. Rev. A 1990, Vol. 42, pp. 6308-6313.
1876: %%CITATION = PHRVA,A42,6308;%%
1877: \bibitem{83}
1878: Berry, M.~V, Marzoli, I., \& Schleich, W {\it Quantum carpets, carpets of
1879: light}; Physics World 2001, Vol. 14N6, pp. 39-44.
1880: %%CITATION = PHWOE,14N6,39;%%
1881: \bibitem{84}
1882: Styer, D.~F. {\it Quantum Revivals versus Classical Periodicity in the
1883: Infinite Square Well}; Am. J. Phys. 2001, Vol. 69, pp. 5662.
1884: %%CITATION = AJPIA,69,56;%%
1885: \bibitem{85}
1886: Loinaz, W., \& Newman, T.~J. {\it Quantum revivals and carpets in some
1887: exactly solvable systems}; J. Phys. A 1999, Vol. 32, pp. 8889-8895.
1888: %%CITATION = QUANT-PH 9902039;%%
1889: \bibitem{86}
1890: Kinzel, W. {\it Bilder elementarer Quantenmechanik};
1891: Physikalische Bl\"{a}tter 1995, Vol. 51, pp. 1190-1191.
1892: %%CITATION = PHBLA,51,1190;%%
1893: \bibitem{87}
1894: Berry, M.~V. {\it Quantum fractals in boxes};
1895: J. Phys. A 1996, Vol. 29, pp. 6617-6629.
1896: %%CITATION = JPAGB,A29,6617;%%
1897: \bibitem{88}
1898: Robinett, R.~W. {\it Quantum wave packet revivals}; \newline
1899: http://xxx.lanl.gov/abs/quant-ph/0401031
1900: %%CITATION = QUANT-PH 0401031;%%
1901: \bibitem{89}
1902: Kiefer, C., \& Joos, E. {\it Decoherence: Concepts and examples}; \newline
1903: http://arxiv.org/abs/quant-ph/9803052
1904: %%CITATION = QUANT-PH 9803052;%%
1905: \bibitem{90}
1906: Zeh, H.~D. {\it The Meaning of Decoherence};
1907: Lect. Notes Phys. 2000, Vol. 538, pp. 19-42.
1908: %%CITATION = QUANT-PH 9905004;%%
1909: \bibitem{91}
1910: Ahluwalia, D.~V. {\it Quantum measurements, gravitation, and locality};
1911: Phys. Lett. B 1994, Vol. 339, pp. 301-303.
1912: %%CITATION = GR-QC 9308007;%%
1913: \bibitem{92}
1914: Ng, Y.~J., \& Van Dam, H. {\it Limit to space-time measurement};
1915: Mod. Phys. Lett. A 1994, Vol. 9, pp. 335-340.
1916: %%CITATION = MPLAE,A9,335;%%
1917: \bibitem{93}
1918: Williams, J.~G., Boggs, D.~H., Turyshev S.~G., \& Ratcliff, J.~T.
1919: {\it Lunar Laser Ranging Science}; \newline
1920: http://arxiv.org/abs/gr-qc/0411095
1921: %%CITATION = GR-QC 0411095;%%
1922: \bibitem{94}
1923: Baez, J.~C., \& Olson, S.~J {\it Uncertainty in Measurements of Distance};
1924: Class. Quant. Grav. 2002, Vol. 19, pp. L121-L126.
1925: %%CITATION = GR-QC 0201030;%%
1926: \bibitem{95}
1927: Wigner, E.~P. {\it Relativistic invariance and quantum phenomena};
1928: Rev. Mod. Phys. 1957, Vol. 29, pp. 255-268.
1929: %%CITATION = RMPHA,29,255;%%
1930: \bibitem{96}
1931: H. Salecker, H., \& Wigner, E.~P. {\it Quantum limitations of the
1932: measurement of space-time distances}; Phys. Rev. 1958, Vol. 109, pp. 571-577.
1933: %CITATION = PHRVA,109,571;%%
1934: \bibitem{97}
1935: Bousso, R. {\it The holographic principle};
1936: Rev. Mod. Phys. 2002, Vol. 74, pp. 825-874.
1937: %%CITATION = HEP-TH 0203101;%%
1938: \bibitem{98}
1939: Ng, Y.~J., \& Van Dam, H. {\it Comment on Uncertainty in measurements of
1940: distance}; Class. Quant. Grav. 2003, Vol. 20, pp. 393-396.
1941: %%CITATION = GR-QC 0209021;%%
1942: \bibitem{99}
1943: Amelino-Camelia, G. {\it Doubly special relativity};
1944: Nature 2002, Vol. 418, pp. 34-35.
1945: %%CITATION = GR-QC 0207049;%%
1946: \bibitem{100}
1947: Amelino-Camelia, G. {\it Doubly-Special Relativity: First Results and Key
1948: Open Problems}; Int. J. Mod. Phys. D 2002, Vol. 11, pp. 1643-1669.
1949: %%CITATION = GR-QC 0210063;%%
1950: \bibitem{101}
1951: Belot, G., \& Earman, J. {\it PreSocratic Quantum Gravity}; in: Callender, C.
1952: \& Huggett, N. (Eds.) {\it Physics Meets Philosophy at the Planck Scale};
1953: Cambridge University Press: Cambridge, 2001, pp. 213-255.
1954: \bibitem{102}
1955: Isham, C.~J. {\it Canonical quantum gravity and the problem of time};
1956: \newline http://arxiv.org/abs/gr-qc/9210011
1957: %%CITATION = GR-QC 9210011;%%
1958: \bibitem{103}
1959: Rovelli, C. {\it Quantum Gravity};
1960: Cambridge University Press: Cambridge, 2004.
1961: \bibitem{104}
1962: Unruh, W.~G. {\it Time, gravity, and quantum mechanics};
1963: \newline http://arxiv.org/abs/gr-qc/9312027
1964: %%CITATION = GR-QC 9312027;%%
1965: \bibitem{105}
1966: Kuchar, K.~V. {\it Time and interpretations of quantum gravity};
1967: In Kunstatter, G., Vincent, D.~E., \& Williams, J. G.(Eds.) {\it Proceedings
1968: of the 4th Canadian Conference on General Relativity and Relativistic
1969: Astrophysics}; World Scientific: Singapore, 1992, pp. 211-314.
1970: \bibitem{106}
1971: Wang, Z.~Y., \& Chen, B. {\it Time in quantum mechanics and quantum field
1972: theory}; J. Phys. A 2003, Vol. 36, pp. 5135-5148.
1973: %%CITATION = QUANT-PH 0211047;%%
1974: \bibitem{107}
1975: Rosenberg, J. {\it A selective history of the Stone-von Neumann theorem};
1976: \newline http://www.math.umd.edu/~jmr/StoneVNart.pdf
1977: \bibitem{108}
1978: Summers, S.~J. {\it On the Stone-von Neumann uniqueness theorem and its
1979: ramifications}; in R\'{e}dei, M., \& St\"{o}ltzner, M. (Eds.)
1980: {\it John von Neumann and the Foundations of Quantum Physics};
1981: Kluwer: Dordrecht, 2001, pp. 135152.
1982: \bibitem{109}
1983: Gieres, F. {\it Mathematical surprises and Dirac's formalism in quantum
1984: mechanics}; Rep. Prog. Phys. 2000, Vol. 63, pp. 1893-1931.
1985: %%CITATION = QUANT-PH 9907069;%%
1986: \bibitem{110}
1987: Butterfield, J., \& Isham, C.~J. {\it Spacetime and the philosophical
1988: challenge of quantum gravity}; in: Callender, C. \& Huggett, N. (Eds.)
1989: {\it Physics Meets Philosophy at the Planck Scale};
1990: Cambridge University Press: Cambridge, 2001, pp. 33-89.
1991: %%CITATION = GR-QC 9903072;%%
1992: \bibitem{111}
1993: Rovelli, C. {\it Quantum space-time: What do we know?}
1994: in: Callender, C. \& Huggett, N. (Eds.)
1995: {\it Physics Meets Philosophy at the Planck Scale};
1996: Cambridge University Press: Cambridge, 2001, pp. 101-124.
1997: %%CITATION = GR-QC 9903045;%%
1998: \bibitem{112}
1999: Lopez-Corredoira, M. {\it What do astrophysics and the world's oldest
2000: profession have in common?} \newline
2001: http://arxiv.org/abs/astro-ph/0310368
2002: %%CITATION = ASTRO-PH 0310368;%%
2003:
2004: \end{thebibliography}
2005:
2006: \end{document}
2007: