5.2. p-adic Clocks
In this section, we introduce a p-adic model of the instrument which measure and indicates time, a p-adic clock; then, we prove that there exist only one clock, which is the same for all Little-endians and Big-endian, the universal clock.
A timekeeping element of the contemporary physical clock is a harmonic oscillator of a particular frequency, which is assumed to be a positive integer showing the number of periods per unit interval; therefore, the shortest time interval which can be measured is a reciprocal of the frequency. In order to measure the value of time elapsed, one merely counts the number of periods from one moment of time to another and represents this non-negative integer in some base, say,
p, where
p is the frequency of the oscillator. In what follows, we assume that
p is a prime as to not overload the exposition with unimportant technical details. Thus, a model of such clock can be represented using the
p-adic odometer, a dynamical system
on the space of
p-adic integers
. If the initial point is
, e.g.,
, put
, and then the base-
p expansion of
represents the time elapsed,
, where
is the
j-th digit of the base-
p expansion of
i. In loose terms, the
p-adic clock is simply a counter whose face consists of windows; at each time moment
i, each
j-th window shows
. It is convenient to assume that the number of windows is infinite to have the time elapsed be unrestricted; thus, we obtain the dynamical system
on
. Note that the initial state
may be taken arbitrarily and not necessarily as
; then, to get the base-
p representation of time elapsed since the initial moment, one has to perform subtraction
in
. The
p-adic clock is depicted in
Figure 3. To the right, the content of the registry is similar to a standard representation of time in decimal (rather than
p-ary) fractions of a second (millisecond, microsecond, nanosecond, …) with Planck time at the rightmost position; meanwhile, to the left are decimal multiples of a second (petasecond, exasecond, …).
Speaking loosely, the registry in
Figure 3 is like a face of a mechanical counter consisting of cogwheels. The period of the sequence of states of the rightmost cell of the registry (which can be judged as the rightmost cogwheel) is
p, the period of the sequence of states of the second rightmost cell is
since the figure in that cell changes once in a period of the rightmost cell, etc. The latter property is a definitive property of an ergodic transformation on
; cf. Theorem 2. Therefore, all ergodic 1-Lipschitz transformations on
should be considered to be clocks, cf. (ii) of Theorem 3, as they can be "adjusted’" one to another since they all are conjugate to the
p-adic odometer.
If the initial state of the odometer is taken to be 0 (i.e., each cell of the registry depicted by
Figure 3 is 0), then after
time units elapse, the registry will contain the base-
p expansion of the number
n since
. Let us now take any ergodic 1-Lipschitz map
, and any
and any sequence
over
which converges
p-adically to
t (such a sequence exists as
is dense in
). It turns out then that for any
, the
p-adic limit
exists; denote this limit via
, then
is a 1-Lipschitz map
which is measure-preserving with respect to
t; see [
23] [Propositions 4.87–4.88, 4.90]. Therefore,
p-adic time t is well-defined. For instance (see [
23] [Example 4.89]), given an ergodic affine map
on
, the two-variate function
is of the form
if
, and
if
. Note that if the affine map
is ergodic then
and
(see [
23] [Theorem 4.36]); thus, both
and
are well-defined
p-adic integers for every
.
The problem which immediately arises is that p-adic time t is well-defined for every , but if q is a prime number distinct from p, the p-adic time t may be meaningless for a q-adic observer, the q-adic Little-endian, not to mention the Big-endian. Fortunately, however, there is a clock (and therefore time) which is common both for all Little-endians and Big-endian. This clock/time is unique up to the direction of the time arrow. It is clear that the clock, which is common for all p-adic Little-endians and Big-endian, must be a totally consistent function. The following theorem holds:
Theorem 11. Totally consistent functions which are measure-preserving for all prime p are exactly the functions , where ; only the functions are ergodic for all prime p.
This means that the “universal clock” is a standard odometer which runs forward () or backward ().
Proof of Theorem 11. According to Theorem 5, any totally consistent function
g is a polynomial; therefore, to be measure-preserving on
,
g must be (1) bijective modulo
p and (2) its derivative
must vanish modulo
p nowhere for all prime
p; see, e.g., [
23] [Theorem 4.45]. As
g is 1-Lipschitz on
for all prime
p, and
is a polynomial, the derivative exists and takes values from
for all prime
p; hence,
. Therefore, (as
is a polynomial), condition (2) implies that
, which means that
is a constant,
. This means that
g is the affine function, namely, either
or
for some
(since
must be an integer as
due to total consistency). This proves the claim concerning measure-preservation.
The ergodicity claim follows from the ergodicity criterion for affine maps
which implies that if the map is ergodic on
then
and
; see [
23] [Theorem 4.36]. As these conditions must hold for all prime
p, we conclude that
and
. □
Interpretation 6 (Free choice of temporal ordering at the smallest of scales). The only clock that is common for “both ends of the scale” is the standard odometer which shows the time elapsed since the moment . All observers acquire the value of the time elapsed up to a nonzero error with respect to the corresponding metrics. Therefore, in a contrast to a real observer (the Big-endian) the p-adic observers (the Little-endians) generally cannot determine with the “time stamps” of events which one of the two events happened earlier and which one later since there is no order on the field of p-adic numbers which agrees with field operations.
Note 7. It is known that generally there is no ordering of events in quantum mechanics; see, e.g., [48]. 5.3. Digitalization
Initial automaton is a model of a (generally open) physical system prepared in some fixed state; the system is exposed by an experimenter to a time series of “elementary impacts” and thus produces the time series of “elementary reactions”. The impacts/reactions occurs at discrete instants of time since time is assumed to be discrete; for example, at Planck’s scale, the smallest time interval is Planck time s. Concrete values of that smallest time interval depend on the process which is modelled (e.g.,. in smart contracts of digital economy the smallest interval is usually assumed to be 24 h) and are not specified; the definitive feature of the model is that “time flow” consists of “indivisible time intervals”.
The experimenter prepares a number of identical systems in the same state and probes them by exposing them to different impacts, observing reactions and thus obtaining a number of experimental points (; ), where are measured values of components of the impact–reaction pair. The experimenter then treats any measured value as a real number up to a nonzero real error.
In order to not overload the exposition, in what follows we consider a one-dimensional case mostly when the values
and
are numbers rather than vectors. Up to normalisation, we may assume that the measured numerical values are all in the unit real interval
; thus, the experimenter obtains a number of experimental points in the real unit square
. Namely, given an automaton
, let
be its automaton function (i.e., a 1-Lipschitz map). Consider a subset
of all the following points of the Euclidean unit square
:
,
. Here,
if
is represented by its canonical form
, (
). Note that
corresponds to a
k-letter output word
of the automaton which is fed by the
k-letter input word
which corresponds to
; cf.
Figure 4.
Further, although all the word lengths k are finite, the clustering is equivalent to sending . Therefore, the clustering is equivalent to taking limit points of the closure of the set with respect to the standard topology of . We call a plot of f. Speaking very loosely, the plot is a picture the experimenter obtains as an output of the experiment which consists of a number of individual probes of a physical system which is prepared in the same state before each probe. Note that the set of cluster points of the pictures for both experimenters, the Little-endian and the Big-endian, obtained as result of the experiment look very similar for the both since Little-endian makes the word lengths as long as possible to construct the cluster points while Big-endian is only capable of obtaining the points which correspond to sufficiently long words, i.e., the points which are close to the cluster points. This fact is crucial for the future construction of wave function by the both experimenters as well as for the uncertainty relation on which the both agree.
Let us describe this procedure more formally. For
, (
), let
and
be the integral and fractional parts of
s, respectively. Recall that any complex character of additive group
of the field
of
p-adic numbers is of the form
, where
;
is a continuous group epimorphism into the group of complex roots of unity (which is isomorphic to the group
). Take
, denote
via
; given a 1-Lipschitz map
, consider the mappings
for all
. As every
maps points of the unit circle
into points of
, the pairs
constitute a set of points on the unit torus
. The unit square
is a universal cover of the torus
; this way, the points
are identified with the points
, and in what follows, we do not differ between the point sets and speak either of the points on the surface of the torus
or on the square
, whichever is more convenient.
Definition 10 (Plots of automata). Given an automaton , let be the automaton function. The closure of all the points in the square (or of all the points in the torus ), where , is called a (one-dimensional) plot of the automaton or, similarly, of the automaton function . The set of all the limit points of the plot, the derived set of the set , is called the limit plot of the automaton (of the automaton function ).
Recall that the
limit point,
accumulation point, or
cluster point is a synonymic notion of the point such that every neighbourhood of which contains points other than that point. Recall also that the derived set of a closed set is also closed; thus,
is closed. Being closed, the set
is measurable with respect to the Lebesgue measure on
; denote as
the measure of
. Respective notions for the general
n-dimensional case,
, are defined as follows: for
,
,
denote
The respective notation in this case is
,
,
, etc. We usually omit the index
n when
.
Theorem 12 (The automata 0-1 law, [
49])
. Given the arbitrary automaton , the following alternative holds: either (
equivalently, is nowhere dense in )
, or (
equivalently, ).
Note 8. Recall that nowhere dense sets can nevertheless have positive Lebesgue measures, for instance, the “fat” Cantor sets (e.g., the Smith-Volterra-Cantor set), which are also known as -Cantor sets; see e.g., [50]; however, this is not the case for the set . The Lebesgue measure of this set is 0 if and only if it is nowhere dense. Theorem 12 is true in the multidimensional case as well. We will say briefly that a 1-Lipschitz map (or respective automaton whose automaton function is f) is measure-0 in dimension n if , and measure-1 otherwise. It turns out that all polynomials over whose degree is greater than 1 are measure-1 in all dimensions. Actually, for , a much stronger result is true: if , then the distribution of points in the unit hypercube tends to uniform as , for every . Specifically, the following theorem holds:
Theorem 13 ([
11])
. Let f be a polynomial over , . Then, the sequence of random vectors weakly converges as to a random vector having a continuous uniform distribution in . Theorem 13 may be interpreted as showing another way by which chaos emerges.
Interpretation 7 (Emergence of chaos: The two ways).
1-st: Chaos emerges from infinite “chaotic sequences” such as random real numbers by iterating them via Bernoulli-shift-like mappings, logistic mappings, etc; that is, when it is assumed a priori that “chaos does exist immanently”.
2-nd: Chaos emerges from the “lack of knowledge what elementary causes happened at the very beginning”; that is, if a Big-endian observer is incapable of determining what the digits are of the input of the causal functionf if k is small enough.
Note that in the second case, the Little-endian observer is capable of determining the digits if k is “not too large”, so these digits are not hidden parameters. Nonetheless, further in the paper, we show that a specific uncertainty relation holds both for the Little-endian and Big-endian observers.
Note also that polynomials over
whose degrees are greater than 1 are automaton functions of
infinite automata; cf., Example 2. However,
any automaton function of an infinite automaton can be uniformly approximated on by automaton functions of finite automata, for instance, by the functions
. This fact, together with the finiteness assumption of
Section 2, emphasises a distinguished role the finite automata play in further considerations; thus, we now pay special attention to finite automata.
Theorem 14 (see [
23] [Section 11.1.2])
. Finite automata are measure-0 in all dimensions. Example 5. Automata may be infinite and measure-0; constants may be measure-1:
The automaton whose automaton function is , (), is infinite and measure-0. Here, is bit-by-bit logical ∨ with no carries to higher order bits; that is, if , then as is a canonical 2-adic representation of .
The automaton whose automaton function is where C is a p-adic integer whose canonical representation corresponds to a Champernowne word is a measure-1 automaton. Recall that a Champernowne word is a word obtained via concatenation of the base-p expansions of numbers 1, 2, 3, 4, 5, 6, …; for instance, the 2-adic Champernowne word is .
In short, Theorems 12 and 14 imply that plots of finite automata cannot contain “figures” but may contain “lines”. These lines are of the utmost importance in further considerations since they may naturally be treated as “experimental curves” obtained by probing a physical system both by Little-endian and Big-endian observers. It turns out that smooth lines from limit plots of finite automata are windings of torus; therefore, the lines may be treated as sine waves, so the smooth lines in the limit plot of a finite automaton constitute a collection of sine waves. Moreover, the waves are limit plots of finite affine automata. Now, we express these facts rigorously.
Recall that a
knot is a smooth embedding of a circle
into
and a
link is a smooth embedding of several disjoint circles in
; cf. [
51]. We will consider only special types of knots and links, namely, torus knots and torus links. Informally, a torus knot is a smooth closed curve without intersections which lies completely in the surface of a torus
, and a link (of torus knots) is a collection of (possibly knotted) torus knots; see, e.g., [
52] [Section 26] for formal definitions.
We also need a notion of a winding of a torus. Formally, a winding of a torus is any geodesic on a torus. Recall that geodesics on torus
are images of straight lines in
under the mapping
of
onto
; cf., e.g., [
53] [Section 5.4].
Definition 11 (Winding of the torus). A winding of the torus is an image of a straight line in under the map of the Euclidean plane onto the 2-dimensional real torus . If the line is defined by the equation , we say that a is a slope of the winding . We denote via a winding which corresponds to the line , the meridian , and say that the slope is ∞ in this case. Windings of slope 0 (i.e., the ones that correspond to straight lines ) are called parallels.
In dynamics, windings of torus
are viewed as orbits of
linear flows on the torus; that is, of dynamical systems on
defined by a pair of differential equations of the form
on
and thus by a pair of parametric equations
in Cartesian coordinates; cf., e.g., [
54] [Section 4.2.3].
Note 9. It is well known that a winding defined by the straight line is dense in if and only if and the slope is irrational; see, e.g., [54] [Proposition 4.2.8] or [53] [Section 5.4]. Theorem 15 which follows states that
-smooth lines (i.e., those which are twice differentiable and have continuous second derivatives) in
are windings of the torus
provided the automaton
is finite; cf.,
Figure 5 and
Figure 6.
Theorem 15 ([
10])
. Let be an automaton function of a finite automaton; let g be a -function with domain and range . Let the graph of the function g lie completely in . Then, there exist such that for all ; moreover, there is a winding of the torus which lies completely in and which contains the graph of the function g. There are not more than a finite number of pairwise distinct windings of the unit torus in ; all of these are images of real affine functions for under the mapping . Note 10. The -smoothness condition can be relaxed: -smoothness is sufficient to ensure the affinity; see [55]. Although Theorem 15, after proper restatement, holds for
m-variate 1-Lipschitz maps
as well, see [
10], we restrict considerations in the rest part of the paper mostly by a univariate case for simplicity.
The torus link which is a limit plot of a finite automaton affine function on is completely described by the following theorem:
Theorem 16 ([
10])
. Given a finite automaton affine function on , (i.e., such that
)
, represent as irreducible fractions: , where , . Then, the limit plot on the torus is a torus link which consists of N torus windings whose slope is a, where is a multiplicative order of p modulo , is the greatest common divisor of , and if . Every torus winding is a graph of the complex-valued function on the torus for a suitable , where , (). In cylindrical coordinates, every torus winding
of a torus that is obtained by revolving around
Z-axis of a circle that is coplanar with the axis and has radius
r and a centre at the distance
R from the origin can be represented by the following parametric equations
If
, then
a is irreducible fraction
where
and
; then, corresponding winding winds
times around the
Z-axis and
times around a circle in the interior of the torus, whereas the sign of
determines whether the rotation is clockwise or counter-clockwise. Hence, “physical meaning” that can be ascribed to the coefficient
of the affine map
, (
), which is a finite automaton function of affine automaton if and only if
, is
frequency (or, as a
wavenumber, under a proper choice of units). The choice of sign + or − depends only on what direction of rotation is assumed to be “positive” or “negative”; thus,
polarization and
spin can be ascribed to the sign of
a in relevant models.
Theorem 16 in view of representation (
15) implies that the limit plot of a finite automaton whose function is
, (where
,
z runs over
) is in one-to-one correspondence to a complex-valued function
:
It is worth noting that the function
is well-defined for all
since
p is the invertible modulo
and thus
is well defined for every
; cf., Theorem 16.
Note 11. According to Theorem 16, different affine functions may have identical limit plots. For instance, all the functions where have identical limit plots which correspond to the function . Note also that whenever a limit plot of a finite automaton is the same as that of the finite automaton whose automaton functionfis affine, ,
there exist aminimal subautomatonof (i.e., the one having no subautomata other than itself) which has exactly the same limit plot; see Figure 7 and Figure 8. A finite automaton is minimal if and only if its reduced state transition diagram is totally connected: Given two states , there is finite word w such that when the automaton in state s accepts the word w, the automaton changes its state to t. If an automaton reaches a state which belongs to its (minimal) subautomaton, the automaton will never reach a state which does not belong to the subautomaton. Example 6 (Limit plots of the automata)
. Figure 9 and Figure 10 show the limit plot of a constant function which is an automaton function of finite autonomous
automaton; autonomous automata may be judged as models of either isolated
or closed
physical systems. Parallel lines shown by Figure 9 may be ascribed to energy levels.The remaining examples are nonautonomous
automata; these can serve as models of open
physical systems. Figure 9 and Figure 10 depict limit plots produced of an autonomous automaton whose state transition diagram depicts Figure 11. Figure 12 and Figure 13 show the limit plot of an automaton having two minimal subautomata; the state transition diagram of the automaton is shown in Figure 14. Figure 15 represents a plot of a finite automaton which approximates a measure-1 (and thus infinite) automaton whose automaton function is , (). Note the pronounced straight lines in the plot; these lines constitute the limit plot of a minimal subautomaton. Figure 16 depicts a plot of a measure-0 (but infinite) automaton which has the only minimal finite affine subautomaton; the automaton function of the latter subautomaton is , (). The limit plot of the latter automaton are red lines; cf., Figure 12; the state transition diagram is the lower part of the diagram shown in Figure 14. Basically, the limit plot of a finite automaton whose minimal subautomata are affine consists of families of parallel straight lines in the unit square or, respectively, of links of the torus windings whose slopes are in ; cf., Figure 5, Figure 6, Figure 12, and Figure 13. The the minimal subautomata from the first example “exhibit nonzero phase shifts”, while for the ones from the second example, the “phase shifts” are 0. Both examples are automata having two minimal affine subautomata. The minimal subautomata from the first example (Figure 5 and Figure 6) have limit plots defined by the functions (red and green windings) and , (yellow, brown, and blue windings), respectively, . The minimal subautomata from the second example (Figure 12 and Figure 13) have limit plots defined by the respective functions (blue lines) and (red lines), . The limit plot of a finite affine automaton whose automaton function is in the unit square consists of parallel straight lines with slope ; thus, the plot may be considered not only on the torus obtained by “gluing together” opposite sides of the square but also on a cylinder obtained by “gluing together” only a pair of opposite sides of the square. This way, one obtains solenoid
rather than a torus link. This representation of a limit plot is also convenient in some cases. For instance, Figure 17 and Figure 18 depict the limit plot of the automaton whose automaton function is , where and are respectively bitwise logical “and” and bitwise logical “not” operations on base-2 expansions of numbers (with no carries), while “·” and “−” are usual multiplication and subtraction of numbers (with carries). Figure 19 represents the state transition diagram of a general automaton all whose minimal automata are finite and affine. 5.4. Wave Functions Emerging from Automata
This section discusses the main notion of quantum theory, the wave function. Our goal is to derive wave functions from causal functions; that is, from automata. Functions (
16) are building blocks of the construction of the wave function on the base of causal maps. To begin, we briefly outline the general idea of the construction.
Recall that the reduced state transition diagram of a finite automaton is a digraph in which each path ultimately reaches a minimal subautomaton. There are no outgoing paths from subautomata. By feeding the automaton with random long words, to each minimal subautomaton we assign a probability for when the automaton reaches states which belong to the subautomaton; cf.,
Figure 20. Let automaton
be such that, being fed by random long words, the automaton at some finite step reaches, with a probability 1, a state which belongs to a minimal automaton which is finite and affine. The limit plot of every such subautomaton is described by a complex-valued function of the form (
16).
To every minimal subautomaton that is finite and affine we ascribe its limit plot. There are only countably many such limit plots since there are only countably many such affine functions
that are automata functions of these subautomata: Due to the finiteness of the subautomata, coefficients of these affine functions must belong to the set
which is countable. As every two minimal subautomata have no common states due to the minimality and as to every minimal subautomaton it is assigned a probability of reaching the subautomaton, to every limit plot one assigns a probability to “observe” that limit plot in the experiment, i.e., to obtain accumulation points in the unit square which constitute that limit plot. The probability is equal to a sum of all probabilities to reach the minimal subautomata having that plot. Therefore, these probabilities constitute a distribution assigned to the automaton; a
characteristic function of that distribution is a (generally infinite) series whose terms are functions
multiplied by values of respective probabilities; cf., (
16) (there is a vast literature on characteristic functions of probability distributions; see, e.g., [
56]). We argue that this characteristic function of the distribution may be treated as a wave function.
Proceeding to a formal rigorous construction, let us review a few preliminary conventions:
A word of caution: there is a one-to-one correspondence between all paths of length
k in the state transition diagram and all numbers from
; however, to every number from
, there corresponds an infinite number of paths: Every such path has a prefix which is simply a base-
p expansion of a number and a suffix which consists of zeros only; cf.,
Section 3.1.
Given an automaton , let be its subautomaton. Let be the set of all infinite paths starting from the initial state of in a state transition diagram of which reach states of at finite steps. Note that if a path w reaches at k-th step, then all paths which correspond to infinite words having the same prefix of length k reach at the k-th step; therefore, the p-adic integers which correspond to these paths constitute a p-adic ball of radius . Therefore, all p-adic integers that correspond to infinite paths which reach the subautomaton at finite steps constitute a disjoint union of balls of nonzero radii; hence, is a μ-measurable subset of with respect to the Haar measure on which is normalised so that . This way to is assigned a probability .
Note that the set
does not depend on a concrete state transition diagram of the automaton
, but to be more definite, one may assume that the state transition diagram of the automaton is reduced; thus, given an automaton function, the reduced state transition diagram of respective automaton is unique; cf.,
Section 3.3. In this case, some care should be taken speaking of paths since some arrows in the reduced state transition diagram may actually be loops; see, e.g.,
Figure 19. The paths (which we write from left to right) that begin at the initial state
and have prefixes 0111, 01011, 010011, 0100011, … all reach the subautomaton
on the fourth, fifth, sixth, seventh,.. steps respectively, so the probability to reach the subautomaton
is
and
is a disjoint union of balls
,
,
, …,
, … where
.
Given two minimal subautomata
and
of the automaton
that are finite and affine, by virtue of the minimality one has
; thus, the probability that a random infinite path starting from the initial state reaches at a finite step some minimal subautomaton of the automaton
is the sum
taken over all minimal subautomata
which are finite and affine. We call an automaton
ultimately affine if the probability is 1. Note that if an ultimately affine automaton is infinite, then, according to König’s lemma (also known as Beth’s tree theorem) [
57], there are infinite paths that never reach states belonging to these minimal subautomata. These paths constitute a
-measurable subset in
but the measure of the subset is 0 since the subset is a complement to a countable union of balls whose measure is 1. For instance, the path
in the state transition diagram depicted by
Figure 2 never reaches a minimal subautomaton (which has only one state, namely,
), but all other paths reach the subautomaton at finite steps, so the probability to reach that minimal subautomaton is 1.
Definition 12 (Plot equivalence of automata)
. Call the finite affine automata and plot equivalent
if their respective functions defined by (16) coincide; that is, if their limit plots coincide, , i.e., if the limit plots are links of the same number of torus windings with a common slope. Given
, denote via
an automaton whose automaton function is
. Let
be the set of all minimal subautomata of
that are plot-equivalent to
. By virtue of the minimality, given
, the subautomata
and
have no common states; therefore,
; that is, the probability
is well-defined. Given
, the equivalence relation
induces an equivalence relation on the set of all pairs
which we denote by the same symbol, i.e.,
if and only if
.
Let
be the set of all equivalence classes defined by minimal subautomata of
which are finite and affine. Then, the series
converges absolutely for all
,
and therefore defines a complex-valued function
. Call the function
a
sharp wave function assigned to the automaton
.
Theorem 17 (On automata having a prescribed wave function). Given non-negative real numbers such that and finite affine automata ,(, ) which are pairwise plot-nonequivalent, there exists an ultimate affine automaton such that , , and .
To prove the theorem we require a lemma.
Lemma 2 (All discrete random variables can be modelled on ). Given convergent series of positive real numbers there exist pairwise disjoint open sets such that the normalised Haar measure μ of is , .
Proof of Lemma 2. Most likely, the lemma is known, but as the author is aware of no proper reference, a proof follows. Consider the Monna map where is a p-adic canonical expansion of . Note that , where ; that is, the Monna map mon maps p-adic balls of radii centred at onto closed subintervals of length of the unit interval ; note that where is the Haar measure on normalised so that , and is Lebesgue measure on the unit real interval , i.e., the length of the closed interval.
Split the unit interval into pairwise disjoint open intervals such that the length of the j-th interval is ; namely, let , , , …; then, is -measurable and .
For each let be a set of all balls of nonzero radii such that for every . As any two p-adic balls either disjoint or one is a subset of another one, the set is a countable disjoint union of balls of nonzero radii. Thus, is open as each p-adic ball of nonzero radius is clopen; hence, is -measurable. As every point from lies in mon-image of some ball from , we conclude that and as when by the construction. □
Proof of Theorem 17. This proof follows immediately from the proof of Lemma 2. Every , is a countable disjoint union of balls , , centred at . Let branches of a p-adic tree be , and let leafs be , . In this digraph, replace all leafs with state transition diagrams of automata . Thus, the constructed digraph is a state transition diagram of the automaton which is the ultimate affine and such that . □
Note 12. From the proof of Theorem 17 it follows that the ultimate affine automaton may be either measure-0 or measure-1. The first case occurs when, for example, the series is finite; therefore the automaton is finite and thus measure-0. The measure-1 case occurs when, for example, all coefficients constitute a dense subset in and all .
In what follows, we will need a slightly generalised version of Lemma 2:
Corollary 1 (Generalized Lemma 2). Given convergent series of positive real numbers , there exist pairwise disjoint open sets such that the normalized Haar measure μ of is , .
Proof of Corollary 1. Take instead of in the proof of Lemma 2 and modify the argument in an obvious way. □
Sharp wave functions may be considered as wave functions with respect to
discrete time since the map
is equivalent to a
k-digit
shift of the base-
p representation of
b and a reduction modulo 1 of the resulting number. As
k is the order of time elapsed (and is measured by
p-adic clock see
Section 5.2 and
Figure 3) since the moment the automaton reaches a state from its minimal affine subautomaton whose automaton function is
, a sharp wave function may be judged as the one the Little-endian can construct by observing reactions of a physical system at the smallest of scales.
We argue that a wave function with respect to
continuous time can also be constructed by using ultimate affine automata. The core idea of the construct is using the beta representations of numbers rather than the base-
p expansions. The beta representations of real numbers were first introduced by A. Rényi in 1957 and since then have attracted substantial attention in ergodic theory and symbolic dynamics; see, e.g., monograph [
21].
Recall that given real
, a
β-representation of real
is an infinite word
over the alphabet
such that
. Note that we consider
-representations of real
and not only of real
as in [
21]. Of course, in (
17), we always may assume that
; however, to assign real numbers to paths in state transition diagrams of automata we need beta representations of numbers from
which then are converted into real numbers in a way similar to what we used in
Section 5.3 by exploiting
p-adic representations.
Specifically, we first use
instead of
p. Thus, each arrow in a state transition diagram of the automaton whose input and output alphabets are
, is labelled by a pair
, where
; for an infinite path which starts from an initial state, there corresponds an infinite word
over alphabet
; for
w, we place a corresponding
-adic integer
. To construct a plot, we convert these
-adic integers into sequences of real numbers
,
,
, …, thus obtaining points
. To put it in other words, we simply use
-representations for input/output words of the automaton
when constructing a plot of the automaton, but the automaton function is still a 1-Lipschitz map from
-adic integers to
-adic integers. This way, we construct a sharp wave function
(cf., (
17)), which is a well-defined complex valued-function of
and
; then, we replace
by
in the formula, thus resulting in another complex-valued function of
and
. The crucial point is that if
, i.e., if
where
, then
. When
is small (e.g., if
s, the Planck time) then for the Big-endian observer who is incapable of performing measurements with that accuracy (which is currently only about
s),
is indistinguishable from continuous time. Thus, we obtain a
fuzzy wave function
which is ascribed to the automaton
. The function is well-defined for all
since the series converges absolutely. From this point, the
sharp wave function (which is a discrete time function) can be viewed as an approximation of a fuzzy wave function (which is a continuous time function). Note that
since , i.e., is a 2-letter alphabet, then necessarily ; see sharp wave function Formula (17).
The term “approximation” here is not rigorous (although some hint is already given by Example 4); to prove this statement with a full rigour is a separate problem which will be considered in the future. In the current paper, we only find an exact representation for
under the
finiteness assumption of
Section 2, but before doing this, we illustrate the usage of that
-representation using the analogy of film which is discussed in
Section 2. Each frame of a film contains a number of details, but to cause an illusion of motion to a viewer, only a small share of the whole number of details is changed from one frame to the next frame; the smaller the share is, the slower the motion appear to the a viewer. For a Little-endian viewer, the share is
since he uses the base-
p representation of numbers; in the case when the share is
, one has the
-representation. If
, we have the case of a Big-endian viewer.
It is important to stress that
to represent numbers from in the base β, we use only non-negative powers of β in order to guarantee the uniqueness of β-representation for each number from since if negative powers of
when
are allowed in
-representations, then every number from
has a continuum of distinct
-representations provided
[
58]. However, in such a case, the very problem of assigning a number to a finite path in a state transition diagram becomes ill-posed. Under said convention, the following theorem is true:
Theorem 18 (Finiteness assumption implies ). Let . If an automaton that performs the addition of β-representations of numbers from is finite then necessarily for some . For each , the addition of numbers from that are represented by -representations can be performed with a finite automaton.
Proof. Number 1 admits the only
-representation
in non-negative powers of
as
. A finite automaton ultimately maps periodic sequences onto ultimately periodic sequences; therefore, if a finite automaton that maps pairs of infinite words into infinite words over the alphabet
and performs
, then necessarily
where
. As the series
diverges, then all
; hence,
for suitable
,
. If
, then the right-hand side of (
19) is not equal to the left-hand side; therefore,
, and by substituting
and collecting terms of positive degrees in
we obtain the following (by binomial theorem):
where
is a polynomial of variable
x whose coefficients are in
. Hence,
, where
,
.
If is a nonzero polynomial, then ; thus, as , we must conclude that : Otherwise, the right-hand side in is strictly greater than is the left-hand side. Therefore, all and thus , i.e., .
If is a zero polynomial, then necessarily . Therefore, there must be exactly one nonzero ; hence, , where . However, since ; so we get a contradiction.
The converse statement of the theorem is obvious since the addition of numbers represented by
-expansions is an “addition with carry to the
N-th digit”; for example, when
one has
□
It is worth warning the reader that Theorem 18 is
not about the calculation of Planck time, whose value depends on the choice of units. In short,
Theorem 18 is about how much information one needs to have both worldviews, that of the Little-endian and the Big-endian, agree. Specifically, Theorem 18 implies that the fuzzy wave function is the one which corresponds to an automaton over a
-symbol alphabet; that is, to the automaton whose function is
, i.e., a
N-variate 2-adic 1-Lipschitz map; see
Section 3.3. Actually,
f is a 1-Lipschitz map
, where
is the ring of integers of the field
; we leave further discussion of theory to future papers.
We remind the reader that for multivariate p-adic 1-Lipschitz maps, most theorems that have been proven or mentioned in this paper hold true; in particular, Theorem 15 holds true. Given a real function whose domain is , by the graph of the function (on the torus ), we mean the point subset . Note that if , then stands for .
Theorem 19 ([
10])
. Let be a finite automaton over the alphabet , let have m inputs and n outputs, and let (
where , , )
be a two-times differentiable function such that all its second partial derivatives are continuous on . If is a subset in a plot of the automaton , then there exist an matrix and a vector such that , (
; )
and for all . There are not more than a finite number of and such that , (
; )
and for some ; moreover, if for some then . The theorem implies that in the multivariate case, the sharp wave function is of the following form:
Therefore,
Theorem 18 implies that a univariate fuzzy wave function is actually a multivariate sharp wave function; however, it is for a large number of dimensions. For instance, if
where
is of order of Planck time, then
; that is, the automaton function of respective automaton is a 1-Lipschitz map
. This means that the matrices
in the above formula for the sharp wave function
are
; that is, each of the matrices contains more entries than the number of atoms in the universe. An infinite-dimensional space is an adequate model for a
-dimensional space; this is why both the Big-endian and Little-endian would agree that wave functions “live” in Hilbert spaces. We postpone to a future paper more rigorous statements and proofs on how pure and fuzzy wave functions are related one to another; here, we only explain why both functions, which may be judged as “physical”, are elements of Hilbert space
of square-summable complex sequences whose terms are indexed by elements of the set
(which is countable) since a “physical” wave function must be square-summable and the sum of squares of probability amplitudes must be 1. Recall that any separable Hilbert space is metrically isomorphic to
and that the Fourier transform on the circle is such an isomorphism between the Hilbert space of square-integrable functions on
and the space
of square-summable complex sequences whose terms are enumerated by integers. It is not difficult to construct sharp wave functions which can be judged as “physical” with this meaning. Indeed, take any sequence
of positive real numbers such that
, and the series
of positive square roots converges; by using Theorem 17, construct the automaton
. Then, function
is the one we are seeking.
We finalise the subsection with the following interpretation.
Interpretation 8 (Discrete spectrum; continuous spectrum). The measure-0 ultimate affine automata may be treated as models of physical systems having discrete (energy, frequency, …) spectra, while measure-1 ultimate affine automata may be treated as models of physical systems having continuous spectra.
5.5. Uncertainty
In this subsection, we formally derive an uncertainty relation which holds for wave functions of automata. We stress, once again, that despite the Litle-endian being capable of performing observation at the smallest scale and the Big-endian not being able to do so, the uncertainty relation, which can be treated as a time-energy uncertainty, holds for both observers, i.e., for Little-endian as well as for Big-endian; thus, no hidden parameters are assumed.
The uncertainty relation we are going to deduce is an
entropic one. A number of research papers have been devoted to discussing entropic uncertainty relations; see, e.g., the expository paper [
59] and the references therein. The entropic uncertainty relation derived below is of a novel type since it relates the time during which a system reaches a “pure state" that can be ascribed to a minimal affine subautomaton and the state (i.e., an element of
) itself. Note that as the Little-endian is capable of performing measurements at the smallest of scales, the time a system reaches a state that belongs to some minimal automaton is not 0, i.e., the “wave function collapse" is not momentary, it takes some minimal time intervals (e.g., some Planck time). Note that the collapse of wave functions as a finite-time process is discussed in the literature; see, e.g., [
60].
To start with, we need to restate some results from
Section 5.4 in terms of
prefix codes since in what follows, we use some basic properties of the codes which may be found, e.g., in the book [
61].
Definition 13 (Prefix code). A nonempty set of finite nonempty words over a finite alphabet 𝒜 that consists of letters is called a prefix code if each word from is a prefix of no other word from .
Let words from the nonempty set
of finite nonempty words over 𝒜 be ordered with respect to a nondecreasing order of their lengths, and let
be the length of the
i-th word (so
).
The set is a prefix code if and only if the following Kraft–McMillan inequality holds:
Note 13. From the proof of Theorem 17, it follows that the branches of the state transition diagram constitute a prefix code since each word which corresponds to a branch of length k reaches some minimal affine subautomaton exactly at the k-th step, thus, the word cannot be a prefix of any other word which corresponds to another branch. Note that words begin from the root of the tree, and the root is the initial state in the state-transition diagram. From the construction, it follows that the Kraft–MacMillan inequality for that code is equality. However, by using Corollary 1 rather than Lemma 2 in the proof, one constructs a prefix code such that for any given . In this case, the rest infinite paths of the complete p-adic tree that lead to no minimal finite affine subautomaton constitute a set of Haar measure . The automaton having such a state transition diagram will reach minimal subautomata which are finite and affine with probability rather than exactly 1. In that case, to automaton , there corresponds a sharp wave function of the form (17) such that which therefore is normalisable. For not to overload the exposition, in what follows we mostly deal with the case when , i.e., with ultimately affine automata , cf. Section 5.4. Let
X be a random variable on the prefix code
X; we denote via
the probability that
X is equal to the word
. By definition [
61], the
entropy of the random variable X is
, whereas the
mean length of the codeword is
.
There exists a prefix code such that
for which the right-hand side inequality in (21) below holds (that right-hand side inequality is not true in general). The left-hand side inequality in (21) below holds whenever
and
, becoming an equality if and only if
[
61] [Theorem 4.3].
The time which a (both sharp and fuzzy) wave function takes to collapse can be expressed via the length of a word which reaches a state from some minimal affine subautomaton since the length of the word is the order of time expressed in the respective base; see the explanations in
Section 5.3. This is why
in what follows, we deal with the lengths of the words rather than with time itself. Note that when dealing with the lengths of the words, we may restrict considerations to the words over the alphabet
where
p is a prime since fuzzy wave functions are constructed by using words over the alphabet
; see
Section 5.4. The only difference between sharp and fuzzy wave function constructions for
are the numerical values that are assigned to words by both the observers: The Little-endian assigns numbers to words by reading the words as the base-2 expansions of numbers whereas the Big-endian reads these words as
-representations of numbers for
. We stress that in what follows, “mean time of collapse” is synonymous with “mean word length” although the
actual mean time of collapse
measured by the both observers is different due to the inevitable nonzero measurement error. For instance, to the word of length
k whose prefix is
zeros and whose suffix is 1, the Little-endian assigns the value
, whereas the Big-endian assigns the value
, which for small
and not too large
k, is indistinguishable for this observer from 1 due to the measurement error. To put it in other words, the Little-endian’s measurements of time elapsed are much more accurate than are the Big-endian’s; the time within which the wave function collapses is large for the Little-endian, whereas that time is zero for the Big-endian up to the measurement accuracy of his equipment; although both the clocks the observers use are 2-adic, according to Theorem 18, the Big-endian can observe digits in the windows that are to the left of the
-th window at the face of the clock for
N large, whereas Little-endian observes digits to the left of the lowest order position, i.e., from the rightmost window (cf.,
Figure 3). Nevertheless, we are going to show that
“time-energy” uncertainty in terms of the length of words in the state transition diagrams of automata holds for both observers.
Let
be an ultimately affine automaton, cf.
Section 5.4. Define the
automaton entropy as
For every
, the probability
is equal to the sum of all
, where
is the length of a finite word
w that reaches some state that belongs to some subautomaton from
exactly at the
-th step; the summation is over all these words. Let
be a code whose codewords are all these words
w; then,
. Note that
, the codes
are disjointed for different
, and the union of all these codes for all
is a prefix code
such that
. According to the above convention, the mean time
of wave function collapse is the mean length of a codeword of the code
:
The inequality (21) implies that
where
is the entropy of the code
. Therefore,
the mean time of collapse of the automaton wave function is equal to the entropy of the code .
For every
, let
be a set of all codewords of length
n from the code
. If
, then
is a prefix code. All of these codes are disjointed, and their union is
. Therefore,
is the probability that the wave function collapses for a time not greater than
N. Let
be the entropy of the prefix code
; that is,
. As the probability assigned to
is
and as
, then for
N not less than the length of the shortest word from
), it holds that
. Moreover, as
, then
If to minimal affine subautomata there are ascribed “energy levels” (e.g., if in the subautomata functions
, the coefficients
a are different and
) these inequalities may be judged as time-energy uncertainty relation since
if an observer measures the time which a wave function takes to collapse, he does not know for sure to which of the states the wave function has collapsed; on the other hand, if he knows to which of the states the wave function has collapsed, he does not know for sure how much time the collapse has taken.
In a general case, these inequalities cannot be sharpened. Since , then can be split arbitrarily into the disjointed union of sets , and as each of is itself a prefix code, there is an automaton such that , (). Indeed, the entropy is determined by the code only, whereas is determined completely by the partition of the code into arbitrary nonempty subsets and by the “assigning of limit plots” to each of the subsets.
The codeword lengths in can be arbitrary as well.
Theorem 20 (On maximal prefix codes [
62])
. For every non-decreasing map such that there exists a (maximal) prefix code such that , for all . That is, one can take any such code C for , split all its codewords into a partition of nonempty subsets, assign to every subset a limit plot of a finite automaton arbitrarily, and construct a respective automaton so that and all finite paths in every S lead to the .
We have that , , (as ), and nothing more definite can be said in the general case. It is possible that . For instance, let , and let . Then, the entropy may be equal to 1 if different limit plots are assigned to different balls . The entropy may be zero if the limit plots that are assigned to all these balls are equal one to another. One may split the set of all these balls into a partition of pairwise disjointed nonempty subsets and assign to each ball a limit plot so that to all balls from a subset, the same limit plot is assigned, but to balls that belong to different subsets, one assigns different limit plots. In all these cases, (as the entropy is equal to ), but the entropies are different.
Finally, consider generating series , where is the number of all words of length n in the prefix code . As , then for the radius of convergence of the series, it holds that , with . Hence, the function is differentiable at all points from , but if , then the derivative may not exist at or may go to . However, , i.e., the derivative determines the entropy .