1: \def\Version{5.2} % no changes so far from 5.1 but commented out
2: % redundant macro defs
3: %%
4: %% To turn off version number comment out "\PrintVersionNumber" below
5:
6: % Outlining is at end!
7:
8: % Time-stamp{2007-Mar-08 03:32:06 17903.51718 at umoja.phy.syr.edu}
9:
10: %: Page Setup for 8.5 x 11 inch paper
11:
12: \message{<< Assuming 8.5" x 11" paper >>}
13:
14: \magnification=\magstep1 % \magstep1=1200
15: \raggedbottom
16: \parskip=9pt
17: %% \overfullrule=0pt
18:
19: \def\singlespace{\baselineskip=12pt} % spacing for stuff like abstract
20: \def\sesquispace{\baselineskip=16pt} % spacing for main text
21:
22: %%% \def\sesquispace{\baselineskip=20pt} % draft spacing
23:
24: %: Load or define some TeX macros
25:
26: \input mathmacros
27: \input mathmacros.greekbold
28: \input msmacros
29:
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31: %% make a local def of box operator
32: %% (see texfiles/developing.macros/dalembertian.tex for other versions)
33:
34: \def\sqr#1#2{\vcenter{
35: \hrule height.#2pt
36: \hbox{\vrule width.#2pt height#1pt
37: \kern#1pt
38: \vrule width.#2pt}
39: \hrule height.#2pt}}
40:
41: \def\dal{\mathop{\,\sqr{7}{5}\,}}
42: \def\block{\dal}
43: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
44:
45: %% the phantom here doesn't help!
46: \def\boxK{\dal\limits_{K} \phantom{}}
47:
48: \def\region{\Buchstabe{X}}
49: \def\interval#1#2{\langle #1, \, #2 \rangle}
50: \def\Exp{\Euclid}
51: \def\Bbar{\bar{B}}
52: \def\f{\phi}
53:
54: \def\IR{R}
55: \def\UV{l}
56: \def\meso{\lambda_0}
57: \def\Mink{\Minkowski}
58:
59:
60: \def\longto{\mathop\longrightarrow}
61: \def\O{{\cal O}}
62:
63: %% Working on figures
64:
65: \input epsf
66: \epsfverbosetrue
67:
68: \def\FigureNumberCaption#1#2#3{
69: \singlespace
70: \vbox{
71: \centerline{\vbox{\epsfbox{#1}}} % the figure
72: \leftskip=1.5truecm\rightskip=1.5truecm % begin indentation
73: \vskip 15pt
74: \noindent{\it Figure #2}. #3 % the caption
75: \vskip .25in
76: \leftskip=0truecm\rightskip=0truecm} % end indentation
77: \vfill
78: \sesquispace}
79: %
80: % \hbox seems to act similarly with \epsfbox
81:
82:
83: %: version number? preprint number(s)?
84:
85: \phantom{}
86:
87: \def\versionnumber#1{
88: \vskip -1 true in \medskip
89: \rightline{version #1}
90: \vskip 0.3 true in \bigskip \bigskip}
91:
92: \PrintVersionNumber
93:
94: %: Title
95: \sesquispace
96: \centerline{
97: {\titlefont Does Locality Fail at Intermediate Length-Scales?}%
98: \footnote{$^{^{\displaystyle\star}}$}%
99: %
100: {To appear in
101: Daniele Oriti (ed.),
102: {\it Towards Quantum Gravity}
103: (Cambridge University Press, 2007)}}
104:
105: \bigskip
106:
107: %: Authors
108:
109: \singlespace % (spacing for addresses etc.)
110:
111: \author{Rafael D. Sorkin}
112: \address
113: {Perimeter Institute, 31 Caroline Street North, Waterloo, ON N2L 2Y5, Canada}
114: \furtheraddress
115: {Department of Physics, Syracuse University, Syracuse, NY 13244-1130, U.S.A.}
116: \email{sorkin@physics.syr.edu}
117:
118: \AbstractBegins
119: %
120: If quantum gravity implies a fundamental spatiotemporal discreteness,
121: and if its ``laws of motion'' are compatible with the Lorentz
122: transformations, then physics cannot remain local. One might expect
123: this nonlocality to be confined to the fundamental discreteness
124: scale, but I will present evidence that it survives at much lower
125: energies, yielding for example a nonlocal equation of motion for a
126: scalar field propagating on an underlying causal set.
127: %
128: \AbstractEnds
129:
130: %: Turn on spacing for body of paper
131:
132: \sesquispace
133:
134: %: Untitled block of main text
135:
136: \noindent
137: Assuming that ``quantum spacetime'' is fundamentally discrete, how
138: might this discreteness show itself?
139: Some of its potential
140: effects are more evident, others less so. The atomic and molecular
141: structure of ordinary matter influences the propagation of both waves
142: and particles in a material medium. Classically, particles can be
143: deflected by collisions and also retarded in their motion, giving rise in
144: particular to viscosity and Brownian motion. In the case of
145: spatio-temporal discreteness,
146: viscosity is excluded by
147: Lorentz symmetry, but fluctuating deviations from rectilinear
148: motion are still possible. Such ``swerves'' have been described in
149: [1] and [2]. They depend (for a massive particle)
150: on a single phenomenological parameter, essentially a diffusion constant
151: in velocity space.
152: As far as I know, the corresponding analysis for a
153: quantal particle with mass has not been carried out yet,
154: but for massless quanta such as photons
155: the diffusion equation of [1] can be adapted to
156: say something, and it then describes
157: fluctuations of both energy and polarization
158: (but not of direction),
159: as well as a secular ``reddening'' (or its opposite).
160: A more complete quantal
161: story, however, would require that particles be treated as wave
162: packets,
163: raising the general question of how spatiotemporal
164: discreteness affects the propagation of {\it waves}.
165: Here, the analogy with
166: a material medium
167: suggests
168: effects such as scattering and extinction,
169: as well as possible nonlinear effects.
170: Further generalization to a ``second-quantized field'' might have more
171: dramatic, if less obvious, consequences.
172: In connection with cosmology,
173: for example,
174: people have wondered how discreteness
175: would affect the hypothetical inflaton field.
176:
177: So far, I have been assuming that, although the deep structure of
178: spacetime is discrete, it continues to respect the Lorentz
179: transformations.
180: That this is logically possible is demonstrated [3] by the
181: example of causal set (causet) theory [4].
182: %% but the latter is the only known scheme combining discreteness with
183: %% Lorentz invariance.
184: With
185: approaches such as loop quantum gravity,
186: on the other hand,
187: the status of local Lorentz
188: invariance seems to be controversial. Some people have hypothesized
189: that it would be broken or at least ``deformed'' in such a way that the
190: dispersion relations for light would cease to be those of a massless
191: field. Were this the case, empty space could also resist the passage of
192: particles (a viscosity of the vacuum), since there would now be a state
193: of absolute rest. Moreover, reference [5] has argued
194: convincingly that it would be difficult to avoid $O(1)$
195: renormalization effects that would lead to different quantum fields
196: possessing different effective light cones. Along these lines, one might
197: end up with altogether more phenomenology than one had bargained for.
198:
199: As already mentioned, the causal set hypothesis avoids such
200: difficulties, but in order to do so, it has to posit a kinematic
201: randomness, in the sense that a spacetime\footnote{$^\star$}
202: %
203: {In this article, ``spacetime'' will always mean Lorentzian
204: manifold, in particular a continuum.}
205: %
206: $M$
207: may
208: properly correspond
209: only to causets $C$ that could have been produced by
210: a
211: {\it Poisson process} in $M$. With respect to
212: an approximating spacetime $M$,
213: the causet thus functions as
214: a kind of ``random lattice''.
215: Moreover,
216: the infinite volume of
217: the Lorentz group implies that such a ``lattice'' cannot be home to a
218: local dynamics. Rather the ``couplings'' or ``interactions'' that
219: describe physical processes occurring in the causet are
220: --- of necessity ---
221: radically nonlocal.
222:
223: To appreciate why this must be, let us refer to the process that will be
224: the subject of much of the rest of this paper: propagation of a scalar
225: field $\f$
226: on a background causet $C$
227: that is well approximated by a Minkowski
228: spacetime $M=\Minkowski^d$. To describe
229: such a
230: dynamics,
231: one needs
232: to reproduce within $C$ something like the d'Alembertian operator
233: $\block$, the Lorentzian counterpart of the Laplacian operator
234: $\nabla^2$
235: of
236: Euclidean space $\Euclid^3$. Locality in the discrete context, if it
237: meant anything at all, would imply that
238: the action of $\dal$
239: would be built up
240: in terms of ``nearest neighbor couplings''
241: (as in fact $\nabla^2$ can be built up,
242: on either a crystalline or random lattice in $\Euclid^3$).
243: But Lorentz invariance
244: contradicts this sort of locality
245: because it implies that, no
246: matter how one chooses to define nearest neighbor, any given causet
247: element $e\in C$ will possess an immense number of them
248: extending throughout
249: the region of $C$ corresponding to the light cone of
250: $e$ in $M$.
251: In terms of a Poisson process in $M$ we can
252: express this more precisely by saying
253: %% claim rigorously
254: that the {\it probability} of any given element $e$ possessing a
255: limited
256: number of
257: nearest neighbors is vanishingly small. Thus, the other
258: elements to which $e$ must be ``coupled'' by our box operator
259: will be large
260: in number (in the limit infinite), and in any given frame of reference,
261: the vast majority of them
262: will be
263: remote
264: from $e$.
265: The resulting ``action at a distance'' epitomizes the maxim
266: that discreteness plus Lorentz invariance entails
267: nonlocality.
268:
269: If this reasoning is correct, it implies that physics at the Planck
270: scale must be radically nonlocal.
271: (By Planck scale I just mean the fundamental length- or volume-scale
272: associated with the causet or other discrete substratum.)
273: Were it to be confined to the Planck scale,
274: however,
275: this nonlocality would be of limited phenomenological interest
276: despite its deep significance for the underlying theory.
277: But a little thought
278: indicates that things might not be so simple.
279: On the contrary,
280: it is far from obvious that
281: the kind of nonlocality in question
282: can be confined to any scale,
283: because
284: for any given configuration of the field $\phi$, the ``local
285: couplings'' will be vastly outnumbered by the ``nonlocal'' ones. How
286: then could the latter conspire to cancel out so that the former could
287: produce a good approximation to $\block\phi$,
288: even for a slowly varying $\phi$?
289:
290: When posed like this, the
291: question looks almost hopeless, but I will try to convince you that
292: there is in fact an answer.
293: What the answer seems to say,
294: though,
295: is that
296: one can reinstate locality only conditionally and to a limited extent.
297: At any finite scale $\lambda$,
298: some nonlocality will naturally
299: persist,
300: but the scale $\meso$
301: at which it begins to disappear
302: seems to
303: reflect
304: not
305: only the ultraviolet scale $\UV$
306: but also an infrared scale $\IR$,
307: which we may identify with the age of the cosmos,
308: and which
309: (in a kind of quantum-gravitational echo of Olber's paradox)
310: seems to be needed in order that locality
311: be recovered at all.
312: On the other hand, (the) spacetime (continuum) as such can make sense
313: almost down to $\lambda=\UV$.
314: We may thus anticipate
315: that,
316: as we coarse-grain up from $\UV$ to larger and larger sizes $\lambda$,
317: we will reach
318: a stratum of reality in which
319: discontinuity has faded out and
320: spacetime has emerged,
321: but physics continues to be nonlocal.
322: One would
323: expect
324: the best description
325: of this stratum to be some type of nonlocal field theory;
326: and this would be a
327: new sort of manifestation of discreteness:
328: not as a source of
329: fluctuations, but as a source of nonlocal phenomena.
330:
331: Under still further coarse-graining, this nonlocality should disappear as
332: well, and one might think that one would land for good in the realm of
333: ordinary quantum field theory (and its further coarse-grainings).
334: However, there is reason to believe that locality would fail once again
335: when cosmic dimensions were reached;
336: in fact,
337: the non-zero cosmological constant predicted on the basis of causet theory
338: is very much a
339: nonlocal reflection,
340: on the largest scales,
341: of the underlying discreteness.
342: It
343: is a strictly quantal effect,
344: however,
345: and would be a very
346: different sort of residue of microscopic discreteness
347: than what I'll be discussing here.
348:
349: These introductory remarks express in a general way most of what I want to
350: convey in this paper, but before getting to the technical
351: underpinnings, let me just (for shortage of space) list some other
352: reasons why people have wanted to give up locality as a fundamental
353: principle of spacetime physics:
354: %
355: to cure the divergences of quantum field theory (e.g. [6]);
356: %
357: to obtain particle-like excitations of a spin-network or related graph
358: [7];
359: %
360: to give a realistic and deterministic account of quantum mechanics
361: (the Bohmian interpretation is both nonlocal and acausal, for example);
362: %
363: to let information escape from inside a black hole (e.g. [8]);
364: %
365: to describe the effects of hidden dimensions in ``brane world'' scenarios;
366: %
367: to reduce quantum gravity to a flat-space quantum field theory via the
368: so called AdS-CFT correspondence;
369: %
370: to make room for non-commuting spacetime coordinates.
371: (This ``non-commutative geometry'' reason is perhaps the most
372: suggestive in the present context, because it entails a hierarchy
373: of scales analogous to the scales $\UV$, $\meso$ and $\IR$. On the ``fuzzy
374: sphere'' in particular, the non-commutativity scale $\meso$ is the
375: geometric mean between the
376: effective ultraviolet cutoff $\UV$ and the sphere's radius $\IR$.)
377:
378:
379: %% \section{A model problem: the classical scalar field on a causet}
380:
381:
382: \section{Three D'Alembertians for two-dimensional causets}
383: %
384: The scalar field on a causet offers a simple model for the questions we
385: are considering. Kinematically, we may realize such a field simply as a
386: mapping $\phi$ of the causet into the real or complex numbers,
387: while in the continuum
388: its equations of motion take
389: --- at the classical level ---
390: the simple form $\block\phi=0$,
391: assuming (as we will) that the mass vanishes. In order to make sense of
392: this equation in the causet,
393: we ``merely'' need to give a
394: meaning to the D'Alembertian operator $\block$. This is not an easy
395: task, but it seems less difficult than giving meaning to,
396: for example, the gradient of $\phi$ (which for its accomplishment would
397: demand that we define a concept of vectorfield on a causet). Of
398: course, one wants ultimately to treat the quantum case, but one would
399: expect a definition of $\block$ to play a basic role there as well, so
400: in seeking such a definition we are preparing equally for the classical
401: and quantal cases.
402:
403: If we assume that $\block$ should act linearly on $\phi$
404: (not as obvious as one might think!),
405: then our task reduces to the finding of a suitable matrix $B_{xy}$
406: to play the role of $\dal$,
407: where the indices
408: $x$, $y$
409: range over the elements of the causet $C$.
410: We will also require that $B$ be ``retarded'' or ``causal'' in the sense
411: that $B_{xy}=0$ whenever $x$ is spacelike to, or causally precedes $y$.
412: In the first place, this is helpful classically, since it allows one to
413: propagate a solution $\phi$ forward iteratively, element by element
414: (assuming that the diagonal elements $B_{xx}$ do not vanish).
415: It might similarly be advantageous quantally, if the path integration
416: is to be conducted in the Schwinger-Kel'dysh manner.
417:
418: \subsection {First approach through the Green function}
419: %
420: I argued above that no matrix $B$ that (approximately) respects the Lorentz
421: transformations can reproduce a local expression like the
422: D'Alembertian unless the majority of terms cancel miraculously
423: in the sum,
424: \ $\sum\limits_{y} B_{xy}\phi_y =: (B\phi)_x$~,
425: that corresponds to $\dal\phi(x)$.
426:
427:
428: Simulations by Alan Daughton [9],
429: continued by Rob Salgado [10],
430: provided the first
431: evidence that the required cancellations can actually be
432: arranged for
433: without
434: appealing to anything other than the intrinsic order-structure of the
435: causet. In this approach one notices that, although in the natural
436: order of things one begins with the D'Alembertian and ``inverts'' it to
437: obtain its Green function $G$, the result in $1+1$-dimensions is so
438: simple that the procedure can be reversed. In fact, the {\it retarded}
439: Green function $G(x,y)=G(x-y)$ in $\Minkowski^2$ is
440: (with the sign convention $\block=-\ptl^2/\ptl{t}^2+\ptl^2/\ptl{x}^2$)
441: just the step function with magnitude $-1/2$ and support the future
442: of the origin (the future light cone together with its interior).
443: Moreover, thanks to the conformal invariance of $\block$
444: in $\Mink^2$,
445: the same expression remains valid in the presence of spacetime
446: curvature.
447:
448: Not only is this continuum expression very simple,
449: but it has an obvious counterpart in the causal set,
450: since it depends on nothing more than the causal relation between the two
451: spacetime points $x$ and $y$.
452: Letting the symbol $<$ denote (strict) causal precedence in the usual
453: way, we can represent the causet $C$ as a matrix whose elements
454: $C_{xy}$ take the value $1$ when $x<y$ and $0$ otherwise.
455: The two-dimensional analog $G$ of the retarded Green function is then
456: just $-1/2$ times (the transpose of) this matrix.
457:
458: From these ingredients, one can concoct some obvious candidates for the
459: matrix $B$.
460: The one that so far has worked best is obtained
461: by symmetrizing $G_{xy}$ and then inverting it.
462: More precisely, what has been done is the following:
463: begin with a specific region $R\subset\Minkowski^2$
464: (usually chosen to be an order-interval, the diamond-shaped region lying
465: causally between a timelike pair of points);
466: randomly sprinkle $N$ points $x_i,\ i=1\dots N$ into $R$;
467: let $C$ be the causet with these points as substratum and the
468: order-relation $<$
469: induced from $\Minkowski^2$;
470: for any
471: ``test''
472: scalar field $\phi$ on $R$,
473: let $\phi_i=\phi(x_i)$ be the induced ``field'' on $C$;
474: build the $N\times N$ matrix $G$ and then symmetrize and invert to get
475: $B$, as described above;
476: evaluate $B(\phi,\psi)=\sum_{ij}B_{ij}\phi_i\psi_j$
477: for $\phi$ and $\psi$ drawn from a suite of test functions
478: on $R$; compare with the continuum values,
479: $\int d^2x \, \phi(x)\block\psi(y) \, d^2y$.
480:
481: For test functions that vanish to first order on the boundary
482: $\partial{R}$ of $R$, and that vary slowly on the scale set by the
483: sprinkling density, the results so far exhibit full agreement between
484: the discrete and continuum values [9][10].
485: Better agreement than this, one could not have hoped for in either
486: respect:
487: Concerning boundary terms, the heuristic reasoning that
488: leads one to expect that inverting a Green function will reproduce
489: a discretized version of
490: $\block$ leaves open
491: its behavior on $\partial R$.
492: Indeed, one doesn't really know what
493: continuum expression to compare with:
494: If our fields don't vanish on $\partial R$,
495: should we expect to obtain an approximation to
496: $\int dx dy \phi(x)\block\psi(y)$ or
497: $\int dx dy (\grad\phi(x),\grad\psi(y))$ or \dots?
498: Concerning rapidly varying functions, it goes without saying that, just
499: as a crystal cannot support a sound wave shorter than the interatomic
500: spacing, a causet cannot support a wavelength shorter than $\UV$. But
501: unlike with crystals, this statement requires some qualification because
502: the notion of wavelength is frame-dependent. What is red light for
503: one inertial observer is blue light for another. Given that the
504: causet can support the red wave, it must be able to support the blue one
505: as well,
506: assuming Lorentz invariance in a suitable sense.
507: Conversely, such
508: paired fields can be used to test the Lorentz invariance of
509: $B$. To the limited extent that this important test has been done,
510: the results have also been favorable.
511:
512: On balance, then, the work done on the Green function approach gives
513: cause for optimism that ``miracles do happen''.
514: However, the simulations have been limited to the flat case, and, more
515: importantly, they do not suffice (as of yet) to establish that the
516: discrete D'Alembertian $B$ is truly frame independent.
517: The point is
518: that although $G$ itself clearly is Lorentz invariant in this sense, its
519: inverse (or rather the inverse of the symmetrized $G$)
520: will in general depend on the region $R$ in which one works.
521: Because this region is not itself invariant under boosts, it defines a
522: global frame that could find its way into the resulting matrix $B$.
523: Short of a better analytic understanding, one is unable
524: to rule out this subtle sort of frame dependence,
525: although the
526: aforementioned limited tests provide evidence against it.
527:
528: Moreover, the Green function prescription itself is of limited
529: application. In addition to two dimensions, the only other case where a
530: similar prescription is known is that of four dimensions {\it without}
531: curvature, where one can take for $G$ the ``link matrix'' instead of the
532: ``causal matrix''.
533:
534: Interestingly enough, the potential for Lorentz-breaking by
535: the region $R$ does not arise if one works exclusively with retarded
536: functions, that is, if one forms $B$ from the original retarded
537: matrix $G$, rather than its
538: symmetrization.\footnote{$^\dagger$}
539: %
540: {One needs to specify a nonzero diagonal for $F$.}
541: %
542: Unfortunately,
543: computer tests with the retarded Green function
544: have so far been
545: discouraging
546: on the whole
547: (with some very recent exceptions).
548: Since, for
549: quite
550: different reasons, it would be
551: desirable to find a retarded representation of $\block$,
552: this suggests
553: that we try something different.
554:
555:
556: \subsection {Retarded couplings along causal links}
557: %
558: Before taking leave of the Green function scheme just described, we can
559: turn to it for one more bit of insight. If one examines the individual
560: matrix elements $B_{xy}$
561: for a typical sprinkling, one notices
562: first of all that they seem to be equally distributed among positive and
563: negative values, and second of all that the larger magnitudes among them
564: are concentrated ``along the light cone''; that is, $B_{xy}$ tends to be
565: small unless the proper distance between $x$ and $y$ is near zero.
566: The latter observation may remind us of a collection of ``nearest
567: neighbor couplings'', here taken in the only possible Lorentz invariant
568: sense: that of small proper distance. The former observation suggests
569: that a recourse to oscillating signs might be the way to effect the
570: ``miraculous cancellations'' we are seeking.
571:
572: The suggestion of oscillating signs is in itself rather vague, but two
573: further observations will lead to a more quantitative idea,
574: as illustrated in figure 1.
575: Let $a$ be
576: some point in $\Minkowski^2$, let $b$ and $c$ be points on
577: the right and
578: left halves of its past lightcone
579: (a ``cone'' in $\Mink^2$ being just a pair of null rays),
580: and
581: let $d$ be the fourth point needed to complete the rectangle. If (with
582: respect to a given frame) all four points are chosen to make a small
583: square, and if $\phi$ is slowly varying (in the same frame), then the
584: combination $\phi(a)+\phi(d)-\phi(b)-\phi(c)$ converges, after suitable
585: normalization,
586: to \ $-\block\phi(a)$
587: as the square
588: shrinks to zero size.
589: (By
590: Lorentz invariance, the same would have happened even if we had started
591: with a rectangle rather than a square.)
592: On the other hand, four other
593: points obtained from the originals by a large boost will form a long
594: skinny rectangle, in which the points $a$ and $b$ (say)
595: are very close
596: together, as are $c$ and $d$. Thanks to the profound identity,
597: $\phi(a)+\phi(d)-\phi(b)-\phi(c)=\phi(a)-\phi(b)+\phi(d)-\phi(c)$, we
598: will obtain only a tiny contribution from this rectangle --- exactly the
599: sort of cancellation we were seeking!
600: By including all the boosts of
601: the original square,
602: we might thus hope to do justice to the Lorentz
603: group without bringing in the unwanted contributions we have been
604: worrying about.
605:
606: %% [[should we insert the figures here?]]
607:
608: Comparison with
609: the D'Alembertian in
610: one dimension leads to a
611: similar idea,
612: which in addition works a bit better in the causet, where
613: elements corresponding to the type of ``null rectangles'' just discussed
614: don't really exist. In $\Minkowski^1$, which is just the real line,
615: $\block\phi$ reduces (up to sign) to $\ptl^2\phi/\ptl{t}^2$, for which
616: a well known discretization is $\phi(a)-2\phi(b)+\phi(c)$, $a$, $b$
617: and $c$ being three evenly spaced points.
618: Such a configuration {\it does} find correspondents in the causet,
619: for example 3-chains $x<y<z$ such that no element other than $y$ lies
620: causally between $x$ and $z$
621: %
622: Once again, any single one of these chains (partly) determines a frame,
623: but the collection of all of them does not.
624: Although these examples should not be taken too seriously
625: (compare the sign in equation (1) below),
626: they bring us very close
627: to the following scheme.\footnote{$^\flat$}
628: %
629: {A very similar idea was suggested once by Steve Carlip}
630:
631: Imagine a causet $C$ consisting of points sprinkled into a region of
632: $\Mink^2$, and fix an element $x\in C$ at which we would like to know
633: the value of $\block\phi$. We can divide the ancestors of $x$ (those
634: elements that causally precede it) into ``layers'' according to their
635: ``distance from $x$'', as measured by the number of intervening
636: elements. Thus layer 1 comprises those $y$ which are
637: {\it linked}
638: to $x$ in the sense that $y<x$ with no intervening elements,
639: layer 2 comprises those $y<x$ with only a single element
640: $z$ such that $y<z<x$, etc.
641: (Figure 2 illustrates the definition of the layers.)
642: Our prescription for $\block\phi(x)$ is
643: then to take some combination, with alternating signs, of the first few
644: layers, the specific coefficients to be chosen so that the correct
645: answers are obtained from suitably simple test functions.
646: Perhaps the simplest combination of
647: this sort is
648: $$
649: B\phi(x)
650: =
651: {4\over l^2}
652: \left(
653: -\half\phi(x) + \left( \sum_1 - 2 \sum_2 + \sum_3 \right) \phi(y)
654: \right)
655: \eqno(1)
656: $$
657: % $$
658: % -\half\phi(x) + \left(\sum_1 - 2 \sum_2 + \sum_3\right) \phi(y)
659: % $$
660: where the three sums $\sum$ extend over the first three layers
661: as just defined, and $l$ is the fundamental length-scale associated with
662: the sprinkling, normalized so that each sprinkled point occupies, on
663: average, an area of $l^2$.
664: %
665: The prescription (1) yields a candidate for the
666: ``discrete D'Alembertian''
667: $B$
668: which
669: is {\it retarded},
670: unlike
671: our earlier
672: candidate based on the symmetrized Green function.
673: In order to express this new $B$ explicitly as a matrix,
674: let $n(x,y)$ denote the cardinality of
675: the order-interval $\interval{y}{x}=\braces{z\in C| y < z < x}$,
676: or in
677: other words the number of elements of $C$ causally between $y$ and $x$.
678: Then,
679: assuming that $x\ge y$,
680: we have from (1),
681: \def\AA{{l^2 \over 4}B_{xy}}
682: \def\BB{-1/2 \quad {\rm\ for\ } x=y}
683: \def\CC{1, -2, 1,
684: \ {\rm according \ as\ } n(x,y) {\rm \ is\ } 0, 1, 2, {\rm\ respectively} ,
685: \quad {\rm\ for\ } x\not=y}
686: \def\DD{\ \ 0 \qquad {\rm\ otherwise}}
687: $$
688: \AA = \left\{ \hbox{$\eqalign{&\BB \cr &\CC \cr &\DD \cr}$} \right.
689: \eqno(2)
690: $$
691:
692:
693: Now let $\phi$ be a fixed test function of compact support on $\Mink^2$,
694: and let $x$
695: (which we will always take to be included in $C$)
696: be a fixed point of $\Mink^2$.
697: If we apply $B$ to $\phi$ we will of course obtain a random
698: answer depending on the random sprinkling of $\Mink^2$.
699: However,
700: one can prove that the {\it mean}
701: of this random variable,
702: $\Exp B\phi(x)$,
703: converges to
704: $\block\phi(x)$ in the continuum limit $l\to0$:
705: $$
706: \Exp \ \sum_y B_{xy}\phi_y
707: \quad
708: \to\limits_{l\to0}
709: \quad
710: \dal\phi(x)
711: \ ,
712: \eqno(3)
713: $$
714: where $\Exp$ denotes expectation with respect to the Poisson process
715: that generates the sprinkled causet $C$.
716: %
717: [The proof rests on the following facts. Let us limit the sprinkling to
718: an ``interval'' (or ``causal diamond'') $\region$ with $x$ as its top
719: vertex. For test functions that are polynomials of low degree, one can
720: evaluate the mean in terms of simple integrals over $\region$ ---
721: for example the integral
722: $\int {dudv/l^2} \exp\braces{-uv/l^2} \ \phi(u,v)$ ---
723: and the
724: results agree with $\block\phi(x)$, up to corrections that vanish like
725: powers of $l$ or faster.]
726:
727:
728: In a sense, then, we have successfully reproduced the D'Alembertian in
729: terms of a causet expression that is {\it fully intrinsic} and therefore
730: automatically {\it frame-independent}. Moreover, the matrix $B$,
731: although it introduces nonlocal couplings, does so only on Planckian
732: scales, which is to say, on scales no greater than demanded by the
733: discreteness itself.\footnote{$^\star$}
734: %
735: {It is not difficult to convince oneself that
736: the limit in (3) sets in
737: when $\UV$ shrinks below the characteristic length associated with the
738: function $\phi$;
739: or vice versa, if we think of $\UV$ as fixed,
740: $B\phi$ will be a good approximation to $\dal\phi$ when the
741: characteristic length-scale $\lambda$ over which $\phi$ varies exceeds
742: $\UV$: $\lambda \gg \UV$.
743: But this means in turn that
744: $(B\phi)(x)$ can be sampling $\phi$
745: {\it in effect}
746: only in a neighborhood of $x$
747: of characteristic size $\UV$.
748: Although $B$ is thoroughly nonlocal at a fundamental level,
749: the scale of its effective nonlocality in application to slowly varying
750: test functions is thus no greater than $\UV$.}
751:
752: But is our ``discrete D'Alembertian'' $B$ really a satisfactory
753: tool for building a field theory on a causet? The potential problem
754: that suggests the opposite
755: conclusion
756: concerns the fluctuations in (1),
757: which grow with $N$ rather than dying away. (This growth is indicated
758: by theoretical estimates and confirmed by numerical simulations.)
759: Whether this problem is fatal or not is hard to say. For example, in
760: propagating a classical solution $\phi$ forward in time through the
761: causet, it might be that the fluctuations in $\phi$ induced by those in
762: (1) would remain small when averaged over
763: many
764: Planck lengths, so
765: that the coarse-grained field would not see them. But if this is true,
766: it remains to be demonstrated. And in any case, the fluctuations would
767: be bound to affect even the coarse-grained field when they became big
768: enough. For the remainder of this paper, I will assume that large
769: fluctuations are not acceptable, and that one
770: consequently
771: needs a
772: different
773: $B$ that will yield the desired answer not only on average,
774: but (with high probability) in each given case. For that purpose,
775: we will have to make more complicated
776: the
777: remarkably simple ansatz (2) that we arrived at above.
778:
779:
780: \subsection {Damping the fluctuations}
781: %
782: To that end,
783: let us return to
784: equation
785: (3)
786: and notice that
787: $\Exp(B\phi)=(\Exp{B})\phi$,
788: where
789: what I have
790: just
791: called
792: $\Exp{B}$
793: is effectively a continuum integral-kernel $\bar{B}$
794: in $\Mink^2$.
795: That is to say,
796: when we average over all sprinklings
797: to get
798: $\Exp B\phi(x)$,
799: the sums in (1)
800: turn into
801: integrals and there results an expression of the form
802: $\int\bar{B}(x-y)\phi(y)d^2y$,
803: where $\bar{B}$ is a retarded, continuous function
804: that can be computed explicitly.
805: Incorporating into $\Bbar$
806: the $\delta$-function answering to $\phi(x)$ in
807: (1),
808: we get for our kernel (when $x>y$),
809: $$
810: \Bbar (x-y) =
811: {4\over l^4} \, p(\xi)e^{-\xi} \; - \; {2\over l^2} \delta^{(2)}(x-y) \ ,
812: \eqno(4)
813: $$
814: where
815: $p(\xi)=1-2\xi+\half\xi^2$,
816: $\xi=v/l^2$
817: and
818: $v=\half||x-y||^2$
819: is the volume (i.e. area) of the order-interval in $\Mink^2$ delimited
820: by $x$ and $y$.
821: The convergence result (3) then states that,
822: for $\phi$ of compact support,
823: $$
824: \int \Bbar(x-y)\phi(y) d^2y
825: \quad
826: \to\limits_{l\to0}
827: \quad
828: \dal\phi(x)
829: \eqno(5)
830: $$
831: Notice that, as had to happen,
832: $\Bbar$ is Lorentz-invariant,
833: since it depends only on the invariant interval
834: $||x-y||^2=|(x-y)\cdot(x-y)|$.\footnote{$^\dagger$}
835: %
836: {The existence of a Lorentz-invariant kernel $\bar{B}(x)$ that yields
837: (approximately) $\dal\phi$ might seem paradoxical,
838: because one could take
839: the
840: function $\phi$ itself to be Lorentz invariant (about the origin $x=0$, say),
841: and
842: for such a $\phi$ the integrand in (5)
843: %% the equation, $\Bbar\f(0)=\int\limits_{x<0}\Bbar(x)\phi(x)d^2x$,
844: would also be invariant,
845: whence the integral
846: would apparently have to diverge.
847: This divergence
848: is avoided for compactly supported $\phi$,
849: of course, because the potential divergence is cut off where the
850: integrand goes to zero.
851: But
852: what is truly remarkable in the face of the
853: counter-argument just given,
854: is that
855: the answer is insensitive to the size of the supporting region.
856: With any reasonable cutoff
857: and reasonably well behaved test functions,
858: the integral still manages to converge to the correct answer
859: as the cutoff is taken to infinity.
860: Nevertheless, this risk of divergence hints at the need
861: we will soon encounter for some sort of infrared cutoff-scale.}
862:
863: Observe, now, that the fundamental discreteness-length has all but
864: disappeared from our story. It remains only in the form of a parameter
865: entering into the definition (4) of the integral kernel
866: $\Bbar$. As things stand, this parameter reflects the scale of
867: microscopic physics from which $\Bbar$ has emerged (much as the diffusion
868: constants of hydrodynamics reflect
869: atomic dimensions).
870: But nothing in the definition of $\Bbar$ per se forces us to this
871: identification.
872: If in (4) we replace $l$ by
873: a freely variable length,
874: and if we then follow the Jacobian dictum,
875: ``Man muss immer umkehren''\footnote{$^\flat$}
876: %
877: {``One must always reverse direction.''}
878: %
879: we can arrive at a modification of the discrete D'Alembertian $B$ for
880: which the unwanted fluctuations are damped out by the law of large
881: numbers.
882:
883: Carrying out the first step, let us replace $1/\UV^2$ in (4)
884: by
885: a new
886: parameter $K$. We obtain
887: a new continuum approximation
888: to $\dal$,
889: $$
890: \Bbar_K (x-y) =
891: {4 K^2} \, p(\xi)e^{-\xi} \; - \; {2K} \delta^{(2)}(x-y) \ ,
892: \eqno(6)
893: $$
894: whose associated nonlocality-scale is not $\UV$ but the length
895: $K^{-1/2}$, which we
896: can
897: take to be much larger than $\UV$.
898: % (Instead of a length as such, it is slightly more convenient to
899: % have introduced an inverse squared length.)
900: %
901: Retracing the steps that led from the discrete matrix
902: (2) to the continuous kernel (4)
903: then
904: brings us to the following
905: causet
906: expression that yields
907: (6) when its
908: sprinkling-average is taken:
909: $$
910: B_K\phi(x)
911: =
912: {4\eps\over l^2}
913: \left(
914: - \half \phi(x)
915: + \eps \sum_{y < x} f(n(x,y),\eps) \ \phi(y)
916: \right) \ ,
917: \eqno(7)
918: $$
919: where
920: $\eps=l^2K$,
921: and
922: $$
923: f(n,\eps) = (1-\eps)^n
924: \left(
925: 1 - {2\eps n\over 1-\eps} + {\eps^2 n (n-1)\over 2 (1-\eps)^2}
926: \right)
927: \ .
928: \eqno(7a)
929: $$
930: For $K=1/l^2$ we recover (1).
931: In the limit where $\eps\to0$ and $n\to\infty$,
932: $f(n,\eps)$
933: reduces to
934: the now familiar form
935: $
936: p(\xi) e^{-\xi}
937: $
938: with $\xi=n\eps$.
939: That is, we obtain in this limit
940: the Montecarlo approximation
941: to the integral $\Bbar_K\f$
942: induced by the sprinkled points.
943: (Conversely,
944: $p(\xi)e^{-\xi}$
945: can serve as a lazybones' alternative
946: to (7a)).
947:
948: %% Once again,
949: Computer
950: simulations show that $B_K\phi(x)$ furnishes a good approximation to
951: $\dal\phi(x)$ for simple test functions, but this time
952: one finds
953: that
954: the fluctuations {\it also} go to zero with $l$, assuming the
955: physical nonlocality scale $K$ remains fixed as $l$ varies.
956: For example, with $N=2^9$ points sprinkled into the interval in
957: $\Mink^2$ delimited by $(t,x)=(\pm1,0)$,
958: and with the test functions
959: $\f=1, t, x, t^2, x^2, tx$,
960: the fluctuations
961: in $B_K\f(t=1,x=0)$
962: %% in $\dal\f(t=1,x=0)$
963: for $\eps=1/64$
964: range
965: from a standard deviation
966: of $0.53$ (for $\f=x^2$)
967: to $1.32$ (for $\f=1$);
968: %
969: % eps in the simulations is called k
970: %
971: and they die out roughly like $N^{-1/2}$
972: (as one might have expected)
973: when $K$ is held fixed as $N$ increases.
974: The means are accurate by construction,
975: in the sense that they
976: exactly\footnote{$^\star$}
977: %
978: {Strictly speaking, this assumes that the number of sprinkled points is Poisson
979: distributed, rather than fixed.}
980: %
981: reproduce the continuum expression $\Bbar_K\f$
982: (which in turn reproduces $\dal\f$ to an accuracy of around 1\%
983: for $K\gto 200$.)
984: (It should also be possible to estimate the fluctuations analytically,
985: but I have not tried to do so.)
986:
987: In any case,
988: we can conclude that ``discretized D'Alembertians''
989: suitable for causal sets do exist, a fairly simple one-parameter family
990: of them being given by (7).
991: The parameter $\eps$ in that expression determines the
992: scale of the nonlocality
993: via $\eps=Kl^2$, and it must be $\ll1$ if we want the fluctuations in
994: $B\f$ to be small.
995: In other words, we need a significant separation between the two
996: length-scales $l$ and
997: $\meso=K^{-1/2}=l/\sqrt{\eps}$.
998:
999: \section {Higher dimensions}
1000: %
1001: So far, we have been concerned primarily with two-dimensional causets
1002: (ones that are well approximated by two-dimensional spacetimes).
1003: Moreover, the quoted result,
1004: (3) cum (6),
1005: has been proved
1006: only under the
1007: additional assumption of flatness,
1008: although it seems likely that it could be extended to the
1009: curved case. More important, however, is finding D'Alembertian
1010: operators/matrices for four and other dimensions.
1011: It turns out that one can do this systematically in a way that
1012: generalizes what we did in two dimensions.
1013:
1014: Let me illustrate the underlying ideas in the case of
1015: four dimensional Minkowski space $\Mink^4$.
1016: In $\Mink^2$ we began with the D'Alembertian matrix $B_{xy}$, averaged
1017: over sprinklings to get
1018: $\Bbar(x-y)$, and ``discretized'' a rescaled $\Bbar$ to get the matrix
1019: $(B_K)_{xy}$.
1020: It turns out that this same procedure works in 4-dimensions if we begin
1021: with the coefficient pattern $1$~$-3$~$3$~$-1$ instead of $1$~$-2$~$1$.
1022:
1023:
1024: To see why it all works, however, it is better to start with the
1025: integral kernel and not the matrix
1026: (now that we know how to pass between them).
1027: In $\Mink^2$ we found $\Bbar$ in the form of a delta-function plus a term
1028: in $p(\xi)\exp(-\xi)$, where
1029: $\xi=Kv(x,y)$,
1030: and $v(x,y)$ was the volume of the order-interval $\interval{y}{x}$,
1031: or equivalently --- in $\Mink^2$ --- Synge's ``world
1032: function''.
1033: In other dimensions this equivalence breaks down and we can imagine using
1034: either the world function or the volume (one being a simple power of the
1035: other, up to a multiplicative constant).
1036: Whichever one chooses, the real task is to find the polynomial $p(\xi)$
1037: (together with the coefficient of the companion delta-function term).
1038:
1039: To that end, notice that the combination $p(\xi)\exp(-\xi)$ can always
1040: be expressed as the result of a differential operator $\O$ in
1041: $\ptl/\ptl K$
1042: acting on $\exp(-\xi)$.
1043: But then,
1044: $$
1045: \int p(\xi)\exp(-\xi) \phi(x) dx =
1046: \int \O \exp(-\xi) \phi(x) dx =
1047: \O \int \exp(-\xi) \phi(x) dx
1048: \ideq \O J
1049: \ .
1050: $$
1051: We want to choose $\O$ so that this last expression yields the desired
1052: results for test functions that are polynomials in
1053: the coordinates $x^\mu$ of degree two
1054: or less.
1055: But the integral $J$ has a very simple form for such $\f$. Up to
1056: contributions that are negligible for large $K$,
1057: it is just
1058: a linear combination of terms of the form
1059: $1/K^n$ or $\log{K}/K^n$. Moreover the only monomials that yield
1060: logarithmic terms are (in $\Mink^2$) $\f=t^2$, $\f=x^2$, and $\f=1$.
1061: In particular the monomials whose D'Alembertian vanishes produce only
1062: $1/K$, $1/K^2$ or $O(1/K^3)$,
1063: with the exception of $\f=1$,
1064: which produces
1065: a term in
1066: $\log{K}/K$.
1067: These are the monomials that
1068: we
1069: don't want to
1070: survive in $\O{J}$. On the other hand $\f=t^2$ and $\f=x^2$ both
1071: produce the logarithmic term $\log{K}/K^2$, and we do want them to
1072: survive.
1073: Notice further, that the survival of {\it any} logarithmic terms would
1074: be bad, because, for dimensional reasons, they would necessarily bring
1075: in an ``infrared'' dependence on the overall size of the region of
1076: integration. Taking all this into consideration, what we need from the
1077: operator $\O$ is that it remove the logarithms and annihilate the
1078: terms $1/K^n$. Such an operator is
1079: $$
1080: \O = \half (H+1)(H+2) \qquad {\rm where} \quad H=K {\ptl\over\ptl K}
1081: $$
1082: is the homogeneity operator.
1083: Applying this to $\exp(-\xi)$ turns out to yield precisely the
1084: polynomial $p(\xi)$ that we were led to above in another manner,
1085: explaining
1086: in a sense
1087: why this particular polynomial arises.
1088: (The relation to
1089: the binomial coefficients, traces back to an identity, proved by Joe
1090: Henson, that expresses $(H+1)(H+2)...(H+n)\exp(-K)$ in terms of binomial
1091: coefficients.) Notice finally that $(H+1)(H+2)$ does {\it not}
1092: annihilate $\log{K}/K$; but it converts it into $1/K$,
1093: which can be canceled by adding
1094: a delta-function to the integral kernel, as in fact we
1095: did. (It could also have been removed by a further factor of $(H+1)$.)
1096:
1097: The situation for $\Mink^4$ is very similar to that for $\Mink^2$.
1098: The low degree monomials again produce terms in $1/K^n$ or
1099: $\log{K}/K^n$,
1100: but everything has an extra factor of $1/K$.
1101: Therefore
1102: $\O={1\over6}(H+1)(H+2)(H+3)$
1103: is a natural choice and leads to a polynomial based on the binomial
1104: coefficients of $(1-1)^3$ instead of $(1-1)^2$.
1105: From it we can derive both a causet D'Alembertian and a
1106: nonlocal, retarded deformation of the continuum D'Alembertian, as before.
1107: It remains to be confirmed, however, that
1108: these expressions
1109: enjoy all the advantages
1110: of the two-dimensional operators discussed above.
1111: It also remains to be confirmed that these advantages persist in the
1112: presence of curvature (but not, of course, curvature large compared to
1113: the nonlocality scale $K$ that one is working with).
1114:
1115: It seems likely that the same procedure would yield candidates
1116: for retarded D'Alem\-bertians in all other spacetime dimensions.
1117:
1118:
1119: \section{Continuous nonlocality, Fourier transforms and Stability}
1120: %
1121: In the course of the above reflexions, we have encountered some
1122: D'Alembertian matrices for the causet and we have seen that
1123: the most promising among them
1124: contain
1125: a free parameter $K$ representing an effective nonlocality scale or
1126: ``meso-scale'', as I will sometimes call it. For processes occurring on
1127: this scale (assuming it is much larger than the ultraviolet scale $\UV$
1128: so that a continuum approximation makes sense) one would expect to
1129: recognize an effective nonlocal theory corresponding to the retarded
1130: two-point function $\Bbar_K(x,y)$.
1131: For clarity of notation, I will call the corresponding operator on
1132: scalar fields ${\boxK}$, rather than $\Bbar_K$.
1133:
1134: Although
1135: its
1136: nonlocality
1137: stems from the discreteness of
1138: the underlying causet,
1139: ${\boxK}$
1140: is a perfectly well defined operator in the
1141: continuum, which can be studied for its own sake. At the same time, it
1142: can help shed light on some questions that arise naturally in relation
1143: to its causet cousin $B_{xy}$.
1144:
1145: One such question (put to me by Ted Jacobson) asks whether the evolution
1146: defined in the causet by $B_{xy}$ is stable or not. This seems
1147: difficult to address
1148: as such
1149: except by computer simulations, but if we transpose it to a
1150: continuum question about $\boxK$, we can come near to a full
1151: answer.
1152: Normally, one expects that if there were an instability then $\boxK$
1153: would possess an ``unstable mode'' (quasinormal mode), that is, a
1154: spacetime function $\f$ of the form
1155: $\f(x)=\exp(ik\cdot x)$ satisfying $\boxK\f=0$, with the imaginary
1156: part of the wave-vector $k$ being future-timelike.\footnote{$^\dagger$}
1157: %
1158: {One might question whether $\boxK\f$ is defined at all for a general mode since
1159: the integral that enters into its definition might diverge, but for a
1160: putative unstable mode, this should not be a problem because the
1161: integral has its support precisely where the mode dies out: toward the
1162: past.}
1163:
1164:
1165:
1166: Now by Lorentz invariance, $\boxK\f$ must be expressible in terms of
1167: $z=k\cdot k$, and it is not too difficult
1168: to reduce it to an ``Exponential integral'' Ei in $z$.
1169: This being done,
1170: some exploration in Maple strongly suggests
1171: that the only solution of $\boxK\exp(ik\cdot x)=0$
1172: is $z=0$, which would mean
1173: the dispersion relation was unchanged from the usual one,
1174: $\omega^2=k^2$.
1175: If this is so,
1176: then
1177: no instabilities can result from
1178: the introduction of our nonlocality scale $K$,
1179: since the solutions of $\boxK\phi=0$ are
1180: precisely those belonging to the usual D'Alembertian.
1181: The distinction between
1182: propagation based on the latter and
1183: propagation based on $\boxK$
1184: would repose only on
1185: the different relationship that $\f$ would bear to its sources;
1186: propagation in empty (and flat) space would show no differences at all.
1187: (The massive case might tell a different story, though.)
1188:
1189:
1190: \subsection {Fourier transform methods more generally}
1191:
1192: What we've just said is essentially that the Fourier transform of
1193: $\boxK$ vanishes nowhere in the complex $z$-plane ($z\ideq k\cdot k$),
1194: except at the origin.
1195: But this draws our attention to the Fourier transform
1196: as yet another route for arriving at a nonlocal
1197: D'Alembertian.
1198: Indeed, most people investigating deformations of $\dal$ seem to have
1199: thought of them in this way, including for example [6].
1200: They
1201: have written down expressions like
1202: $\dal\exp(\dal/K)$, but without seeming to pay much attention to whether
1203: such an expression makes sense in a spacetime whose signature has not
1204: been Wick rotated to $(++++)$.
1205: In contrast, the operator $\boxK$ of this paper was defined directly in
1206: ``position space''
1207: as an integral kernel, not
1208: as
1209: a formal function of $\dal$.
1210: Moreover, because it is retarded,
1211: its Fourier transform is rather special \dots.
1212: By continuing in this vein, one can come up with a third derivation of
1213: $\boxK$ as
1214: (apparently)
1215: the simplest operator whose Fourier
1216: transform obeys the analyticity and boundedness conditions required in
1217: order that $\boxK$ be well-defined and retarded.
1218:
1219: The Fourier transform itself can be given in many forms, but the
1220: following is among the simplest:
1221: $$
1222: \boxK e^{ik\cdot x}|_{x=0}
1223: \ = \
1224: {2z\over i}
1225: \int_0^\infty dt \; {e^{itz/K} \over (t-i)^2}
1226: \eqno(8)
1227: $$
1228: where
1229: here,
1230: $z=-k\cdot k/2$.
1231:
1232: It would be interesting to learn what operator would result if one
1233: imposed ``Feynman boundary conditions'' on the inverse Fourier transform
1234: of this function, instead of ``causal'' ones.
1235:
1236:
1237: \section{What next?}
1238: %
1239: Equations (7) and (6)
1240: offer us
1241: two distinct, but closely related, versions of $\dal$,
1242: one suited to a causet and the other being an effective
1243: continuum operator arising as an average or limit of the first.
1244: Both are retarded and each is Lorentz invariant in the relevant sense.
1245: How can we use them?
1246: First of all,
1247: we can
1248: take up the
1249: questions about wave-propagation raised
1250: in the introduction,
1251: looking in particular for deviations from the
1252: simplified model of [11]
1253: based on ``direct transmission'' from source to sink
1254: (a model that has much in common with the approach discussed above under
1255: the heading ``Approach through the Green function'').
1256: Equation (7) in particular, would let us propagate a
1257: wave-packet through the causet and look for some of the effects
1258: indicated in the introduction, like ``swerves'', scattering and
1259: extinction.
1260: These of course
1261: hark back
1262: directly to the granularity of the causet, but
1263: even in
1264: the continuum
1265: limit
1266: the nonlocality associated with
1267: (6) might modify the field emitted by a
1268: given
1269: source in
1270: an interesting manner;
1271: and this would be relatively easy to analyze.
1272:
1273: Also relatively easy to study would be the effect of the nonlocality on
1274: free propagation in a curved background.
1275: % To begin with, one could investigate the equation $\boxK \phi=0$ in
1276: % simple curved spacetimes like de Sitter.
1277: Here one {\it would} expect some change to the
1278: propagation law.
1279: % dispersion relations.
1280: Because of the retarded character of $\boxK$, one
1281: might also expect to see some sort of induced CPT violation in an expanding
1282: cosmos. Because (in a quantal context) this would disrupt the equality
1283: between the masses of particles and antiparticles, it would
1284: be a potential source of baryon-anti-baryon asymmetry not resting on any
1285: departure from thermal equilibrium.
1286:
1287: When discreteness combines with spacetime curvature, new issues arise.
1288: Thus, propagation of wave-packets in an expanding universe and in a
1289: black hole background both raise puzzles having to do with the extreme
1290: red shifts that occur in both situations
1291: (so-called transplanckian puzzles).
1292: In the black hole context, the red shifts are of course responsible
1293: for Hawking radiation, but their analysis in the continuum seems to
1294: assign a role to
1295: modes of exponentially high frequency
1296: that arguably
1297: should be eschewed if one posits a minimum length.
1298: Equation (7) offers a framework in which
1299: such questions
1300: can be
1301: addressed without infringing on Lorentz invariance.
1302: The same holds for questions about what happens to wave-packets in (say)
1303: a de Sitter spacetime when they are traced backward toward the past far
1304: enough so that their frequency (with respect to some cosmic rest frame)
1305: exceeds Planckian values.
1306: Of course, such questions will not be resolved fully
1307: on the basis of classical equations of motion.
1308: Rather one will have to formulate
1309: quantum field theory on a causet,
1310: or possibly one will have to go all the way to
1311: a quantal field on a quantal causet
1312: (i.e. to quantum gravity).
1313: Nevertheless, a better understanding of the classical case is
1314: likely
1315: to be relevant.
1316:
1317: I will not try to discuss here how to do quantum field theory on a
1318: causet,
1319: or even in Minkowski spacetime with a nonlocal D'Alembertian.
1320: That would raise a whole set of new issues,
1321: path-integral vs. operator methods and the roles of unitarity and causality
1322: being just some of them.\footnote{$^\flat$}
1323: %
1324: {I will however echo a comment made earlier: I suspect that one should
1325: not try to formulate a path-integral propagator as such; rather one
1326: will work with Schwinger-Kel'dysh paths. }
1327: %
1328: But it does seem in harmony with the aim of this paper to comment
1329: briefly on the role of nonlocality in this connection.
1330: As we have seen, the ansatz (6) embodies a nonlocal
1331: interaction that has survived in the continuum limit, and thus might be
1332: made the basis of a nonlocal field theory of the sort that people have
1333: long been speculating about.
1334:
1335: What is
1336: especially
1337: interesting from this point of view is the potential for a new
1338: approach to renormalization theory (say in flat spacetime $\Mink^d$).
1339: People have sometimes hoped that nonlocality would eliminate the
1340: divergences of quantum field theory,
1341: but as far as I can see, the opposite is true,
1342: at least for the specific sort of nonlocality embodied in (6).
1343: In saying this, I'm assuming that the divergences can all be traced to
1344: divergences of the Green function $G(x-y)$
1345: in the coincidence limit $x=y$.
1346: If this is correct then one would need to soften
1347: the high frequency behavior of $G$,
1348: in order to eliminate them.
1349: But a glance at (8)
1350: reveals that $\boxK$ has a milder ultraviolet behavior than $\dal$,
1351: since its Fourier transform goes to a constant at $z=\infty$, rather
1352: than blowing up linearly.
1353: Correspondingly,
1354: one would expect its Green
1355: function to be
1356: more singular
1357: than that of the local operator $\dal$,
1358: making the divergences worse, not better.
1359: If so, then
1360: one must look to the discreteness itself to cure the divergences;
1361: its associated nonlocality will not do the job.
1362:
1363: But if nonlocality alone
1364: cannot remove the need for renormalization altogether,
1365: it might
1366: nevertheless
1367: open up a new and more symmetrical way to arrive at finite answers.
1368: The point is that (8) behaves at $z=\infty$ like $1+O(1/z)$,
1369: an expression whose reciprocal has exactly the same behavior!
1370: The resulting Green function should therefore also be the sum of a
1371: delta-function with a regular\footnote{$^\star$}
1372: %
1373: {At worst, it might diverge logarithmically on the light cone, but in
1374: that case, the residual divergence could be removed by adjusting the
1375: Fourier transform to behave like $1 + O(1/z^2)$.}
1376: %
1377: function
1378: (and the same reasoning would apply in four dimensions).
1379: The resulting
1380: Feynman diagrams would be finite
1381: {\it except for} contributions from the delta-functions.
1382: But these could be removed by hand
1383: (``renormalized away'').
1384: If this idea worked out, it could provide a new approach to renormalization
1385: based on a new type of Lorentz invariant regularization.
1386: (Notice that this all makes sense in real space, without the need for
1387: Wick rotation.)
1388:
1389:
1390:
1391: \section{How big is $\meso$?}
1392: %
1393: From a phenomenological perspective, the most burning question is one
1394: that I cannot really answer here: Assuming there are nonlocal effects
1395: of the sort considered in the preceding lines, on what length-scales
1396: would they be expected to show up?
1397: In other words, what is the value of $\meso=K^{-1/2}$~?
1398: Although I don't know how to answer this question theoretically,\footnote{$^\dagger$}
1399: %
1400: {The question of why $\UV$ and $\meso$ are not comparable joins the
1401: other ``large number'' (or ``hierarchy'') puzzles. Perhaps their ratio
1402: is set dynamically (e.g. ``historically'' in relation to the large age
1403: of the cosmos), along with the small size of the cosmological constant
1404: and the large size of the cosmic diameter. Such a mechanism could be
1405: either complementary to that suggested in [12] (explaining why the value
1406: about which Lambda fluctuates is so close to zero) or alternative to
1407: it.}
1408: %
1409: it is possible to deduce bounds on $\meso$ if we assume that the
1410: fluctuations in individual values of
1411: $\dal\phi(causet)=B_K\phi$ are small, as
1412: discussed above.
1413: Whether such an assumption will still seem necessary at the end of the
1414: day is of course very much an open question. Not only could a sum over
1415: individual elements of the causet counteract the fluctuations (as
1416: already mentioned), but the same thing could result from the sum over
1417: causets implicit in quantum gravity.
1418: This would be a sum of exponentially more terms, and as such it could
1419: potentially remove the need for any intermediate nonlocality-scale
1420: altogether.
1421:
1422: In any case, if we do demand that the fluctuations be elementwise
1423: small,
1424: then
1425: $\meso$ is bounded from below by this requirement. (It is of course
1426: bounded above by the fact that --- presumably --- we have not seen it
1427: yet.) Although this bound is not easy to analyze, a very crude estimate
1428: that I will not reproduce here suggests that we make a small fractional
1429: error in $\dal\phi$ when (in dimension four)
1430: $$
1431: \lambda^2 \UV^2 \IR \ll \meso^5 \ ,
1432: $$
1433: where $\lambda$ is the characteristic length-scale associated with the
1434: scalar field.
1435: On the other hand, even the limiting continuum expression
1436: $\boxK\phi$ will be a bad approximation unless
1437: $\lambda\gg\meso$.
1438: Combining these inequalities yields
1439: $\lambda^2\UV^2\IR\ll\meso^5\ll\lambda^5$,
1440: or $\UV^2\IR\ll\lambda^3$.
1441: For smaller $\lambda$, accurate approximation to
1442: $\dal\phi$ is incompatible with small fluctuations.
1443: Inserting for $\UV$ the Planck length\footnote{$^\flat$}
1444: %
1445: {This could be an underestimate if a significant amount of
1446: coarse-graining of the causet were required for spacetime to emerge.}
1447: %
1448: of $10^{-32}cm$ and for $\IR$ the Hubble radius, yields
1449: $\lambda\sim10^{-12}cm$
1450: as the smallest wavelength that would be immune to
1451: the nonlocality.
1452: That this is not an extremely small length,
1453: poses the question
1454: whether observations already exist that could rule out nonlocality on
1455: this scale.\footnote{$^\star$}
1456: %
1457: {Compare the interesting observations (concerning ``swerves'') in
1458: [13]}
1459:
1460:
1461:
1462: %: Acknowledgments including grant citation
1463:
1464: \bigskip
1465: \noindent
1466: It's a pleasure to thank Fay Dowker and Joe Henson for extensive
1467: discussions and help on these matters, during their visits to Perimeter
1468: Institute.
1469: Research at Perimeter Institute for Theoretical Physics is supported in
1470: part by the Government of Canada through NSERC and by the Province of
1471: Ontario through MRI.
1472: This research was partly supported
1473: by NSF grant PHY-0404646.
1474:
1475:
1476: %: Here put figures 1 and 2 on their own pages
1477:
1478: \pagebreak
1479: \phantom{blah blah}
1480: \vskip 2 truein
1481: \epsfxsize=6.2in
1482: \FigureNumberCaption {figure.a.mars.eps} {1} {How a miracle might happen}
1483:
1484: %\pagebreak
1485: %\phantom{blah blah}
1486: \vskip 2 truein
1487:
1488: \epsfxsize=3.6in
1489: \FigureNumberCaption{figure.b.eps}{2}%
1490: {Illustration of the definition of the ``proximity measure'' $n(a,b)$.
1491: For the causet shown in the figure, $n(x,y)=3$
1492: % (not 2)
1493: because three elements intervene causally between $y$ and $x$.
1494: The first layer below element $x$ contains a single element, while the
1495: second, third and fourth, contain 2, 1, and 2, elements respectively.}
1496:
1497: \pagebreak
1498:
1499: \ReferencesBegin
1500:
1501: \ref [1] % swerves paper
1502: Fay Dowker, Joe Henson and Rafael D.~Sorkin,
1503: ``Quantum Gravity Phenomenology, Lorentz Invariance and Discreteness'',
1504: \journaldata {Modern Physics Letters~A} {19} {1829-1840} {2004}
1505: \eprint{gr-qc/0311055}
1506:
1507:
1508: \ref [2] % another paper on Lor-invar diffusion
1509: R.M Dudley,
1510: ``Lorentz-invariant Markov processes in relativistic phase space'',
1511: \journaldata{Arkiv f{\"o}r Matematik}{6(14)}{241-268}{1965}
1512:
1513:
1514: \ref [3] % proving that Poisson is in fact Lor invariant
1515: %
1516: Luca Bombelli, Joe Henson and Rafael D. Sorkin,
1517: ``Discreteness without symmetry breaking: a theorem''
1518: (in preparation)
1519:
1520: % [[update this reference asap]]
1521:
1522:
1523: \ref [4]
1524: See the article by Joe Henson in this volume (\eprint{gr-qc/0601121}).
1525: Some other references of a general nature are: \linebreak
1526: %
1527: Luca Bombelli, Joohan Lee, David Meyer and Rafael D.~Sorkin,
1528: ``Spacetime as a Causal Set'',
1529: \journaldata {Phys. Rev. Lett.}{59}{521-524}{1987}; \lbr
1530: %
1531: Rafael D.~Sorkin,
1532: ``Causal Sets: Discrete Gravity (Notes for the Valdivia Summer School)'',
1533: in {\it Lectures on Quantum Gravity}
1534: (Series of the Centro De Estudios Cient{\'\i}ficos),
1535: proceedings of the Valdivia Summer School,
1536: held January 2002 in Valdivia, Chile,
1537: edited by Andr{\'e}s Gomberoff and Don Marolf
1538: (Plenum, 2005)
1539: \eprint{gr-qc/0309009}; and \linebreak
1540: %
1541: Fay Dowker, ``Causal sets and the deep structure of Spacetime'',
1542: in
1543: {\it 100 Years of Relativity - Space-time Structure: Einstein and Beyond}"
1544: ed Abhay Ashtekar
1545: (World Scientific, to appear)
1546: \eprint{gr-qc/0508109}
1547:
1548:
1549: \ref [5] % sudarsky et al contra regular lattice
1550: John Collins, Alejandro Perez, Daniel Sudarsky, Luis Urrutia, and He'ctor Vucetich,
1551: ``Lorentz invariance and quantum gravity: an additional fine-tuning problem?''
1552: gr-qc/0403053
1553:
1554:
1555: \ref [6] % namsrai, moffat articles
1556: K. Namsrai,
1557: {\it Nonlocal Quantum Field Theory and Stochastic Quantum Mechanics}
1558: (D. Reidel, 1986);
1559: %%% \lbr
1560: J.W. Moffat, ``Finite nonlocal gauge field theory'',
1561: \journaldata{Phys. Rev. D}{41}{1177-1184}{1990}
1562:
1563:
1564: \ref [7] % particles from nonlocality
1565: See the article by Fotini Markopoulou in this volume;
1566: also
1567: Sundance O. Bilson-Thompson, Fotini Markopoulou and Lee Smolin,
1568: ``Quantum gravity and the standard model'',
1569: \eprint{hep-th/0603022}
1570:
1571:
1572: \ref [8] % antony on getting info out from black hole
1573: %
1574: Antony Valentini,
1575: ``Black Holes, Information Loss, and Hidden Variables''
1576: \lbr
1577: \eprint{hep-th/0407032}
1578:
1579: \ref [9] % daughton thesis
1580: Alan R. Daughton, % doctoral thesis
1581: {\it The Recovery of Locality for Causal Sets and Related Topics},
1582: Ph.D. dissertation (Syracuse University, 1993)
1583:
1584:
1585: \ref [10] % rob salgado thesis
1586: Roberto Salgado, % doctoral thesis
1587: ``Toward a Quantum Dynamics for Causal Sets''
1588: (Ph.D. dissertation, Syracuse University, in preparation)
1589:
1590:
1591: \ref [11] % direct transmission paper (in preparation)
1592: Fay Dowker, Joe Henson and Rafael D.~Sorkin,
1593: ``Wave propagation on a causet I: direct transmission along causal links''
1594: (in preparation)
1595:
1596:
1597: \ref [12]
1598: Maqbool Ahmed, Scott Dodelson, Patrick Greene and Rafael D.~Sorkin,
1599: ``Everpresent $\Lambda$'',
1600: \journaldata {Phys. Rev.~D} {69} {103523} {2004}
1601: \eprint{astro-ph/0209274}
1602:
1603:
1604: \ref [13]
1605: Nemanja Kaloper and David Mattingly,
1606: ``Low energy bounds on Poincar{\'e} violation in causal set theory'',
1607: \eprint{astro-ph/0607485}
1608:
1609:
1610: \end
1611:
1612: %: Outline mode stuff (put here so doesn't need to be commented out)
1613:
1614: (prog1 'now-outlining
1615: (Outline
1616: "\f......"
1617: "%------"
1618: "%: ....."
1619: "%:: ....."
1620: ;; "\\\\message"
1621: "\\\\Abstrac"
1622: "\\\\section"
1623: "\\\\subsectio"
1624: "\\\\appendi"
1625: "\\\\Referen"
1626: "\\\\ref....[^|]"
1627: ;"\\\\ref....."
1628: "\\\\end...."))
1629: