Sie sind hier: ICP » R. Hilfer » Publikationen

2 Foundations

2.1 Basic Desiderata for Time Evolutions

[91.2.1] The following basic requirements define a time evolution in this chapter.

  1. Semigroup
    [91.2.2] A time evolution is a pair (\{\mbox{\rm T}_{\tau}(t):0\leq t<\infty\},(B_{\tau},\|\cdot\|)) where \mbox{\rm T}_{\tau}(t)=\mbox{\rm T}(t\tau) is a semigroup of operators \{\mbox{\rm T}(t):0\leq t<\infty\} mapping the Banach space (B_{\tau}(\mathbb{R}),\|\cdot\|) of functions f_{\tau}(s)=f(s\tau) on \mathbb{R} to itself. [91.2.3] The argument t\geq 0 of \mbox{\rm T}_{\tau}(t) represents a time duration, the argument s\in\mathbb{R} of f_{\tau}(s) a time instant. [91.2.4] The index \tau>0 indicates the units (or scale) of time. [91.2.5] Below, \tau will again be frequently suppressed to simplify the notation. [91.2.6] The elements f_{\tau}(s)=f(s\tau)\in B_{\tau} represent observables or

    [page 92, §0]

    the state of a physical system as function of the time coordinate s\in\mathbb{R}. [92.0.1] The semigroup conditions require

    \displaystyle\mbox{\rm T}_{\tau}(t_{1})\mbox{\rm T}_{\tau}(t_{2})f_{\tau}(t_{0}) \displaystyle=\mbox{\rm T}_{\tau}(t_{1}+t_{2})f_{\tau}(t_{0}) (7)
    \displaystyle\mbox{\rm T}_{\tau}(0)f_{\tau}(t_{0}) \displaystyle=f_{\tau}(t_{0}) (8)

    for t_{1},t_{2}>0, t_{0}\in\mathbb{R} and f_{\tau}\in B_{\tau}. [92.0.2] The first condition may be viewed as representing the unlimited divisibility of time.

  2. Continuity
    [92.0.3] The time evolution is assumed to be strongly continuous in t by demanding

    \lim _{{t\to 0}}\|\mbox{\rm T}(t)f-f\|=0 (9)

    for all f\in B.

  3. Homogeneity
    [92.0.4] The homogeneity of the time coordinate requires commutativity with translations

    {\mathcal{T}}({t_{1}})\mbox{\rm T}(t_{2})f(t_{0})=\mbox{\rm T}(t_{2}){\mathcal{T}}({t_{1}})f(t_{0}) (10)

    for all t_{2}>0 and t_{0},t_{1}\in\mathbb{R}. [92.0.5] This postulate allows to shift the origin of time and it reflects the basic symmetry of time translation invariance.

  4. Causality
    [92.0.6] The time evolution operator should be causal in the sense that the function g(t_{0})=(\mbox{\rm T}(t)f)(t_{0}) should depend only on values of f(s) for s<t_{0}.

  5. Coarse Graining
    [92.0.7] A time evolution operator \mbox{\rm T}(t) should be obtainable from a coarse graining procedure. [92.0.8] A precise definition of coarse graining is given in Definition 2.3 below. [92.0.9] The idea is to combine a time average \frac{1}{t}\int _{{s-t}}^{s}f(t^{{\prime}})\;\mbox{\rm d}t^{{\prime}} in the limit t,s\to\infty with a rescaling of s and t.

[92.0.10] While the first four requirements are conventional the fifth requires comment. [92.0.11] Averages over long intervals may themselves be timedependent on much longer time scales. [92.0.12] An example would be the position of an atom in a glass. [92.0.13] On short time scales the position fluctuates rapidly around a well defined average position. [92.0.14] On long time scales the structural relaxation processes in the glass can change this average position. [92.0.15] The purpose of any coarse graining procedure is to connect microscopic to macroscopic scales. [92.0.16] Of course, what is microscopic

[page 93, §0]    depends on the physical situation. [93.0.1] Any microscopic time evolution may itself be viewed as macroscopic from the perspective of an underlying more microscopic theory. [93.0.2] Therefore it seems physically necessary and natural to demand that a time evolution should generally be obtainable from a coarse graining procedure.

2.2 Evolutions, Convolutions and Averages

[93.1.1] There is a close connection and mathematical similarity between the simplest time evolution \mbox{\rm T}(t)={\mathcal{T}}(t) and the operator \mbox{\rm M}(t) of time averaging defined as the mathematical mean

\mbox{\rm M}(t)f(s)=\frac{1}{t}\int^{s}_{{s-t}}f(y)\;\mbox{\rm d}y, (11)

where t>0 is the length of the averaging interval. [93.1.2] Rewriting this formally as

\mbox{\rm M}(t)f(s)=\frac{1}{t}\int^{t}_{0}f(s-y)\;\mbox{\rm d}y=\frac{1}{t}\int^{t}_{0}{\mathcal{T}}(y)f(s)\;\mbox{\rm d}y (12)

exhibits the relation between \mbox{\rm M}(t) and {\mathcal{T}}(t). [93.1.3] It shows also that \mbox{\rm M}(t) commutes with translations (see eq. (10)).

[93.2.1] A second even more suggestive relationship between \mbox{\rm M}(t) and {\mathcal{T}}(t) arises because both operators can be written as convolutions. [93.2.2] The operator \mbox{\rm M}(t) may be written as

\mbox{\rm M}(t)f(s)=\frac{1}{t}\int^{t}_{0}f(s-y)\;\mbox{\rm d}y=\int^{\infty}_{{-\infty}}f(s-y)\frac{1}{t}\chi _{{[0,1]}}\left(\frac{y}{t}\right)\;\mbox{\rm d}y=\int^{s}_{0}f(s-y)\frac{1}{t}\chi _{{[0,1]}}\left(\frac{y}{t}\right)\;\mbox{\rm d}y, (13)

where the kernel

\chi _{{[0,1]}}(x)=\begin{cases}1&\text{\ \ \ \  for }x\in[0,1]\\
0&\text{\ \ \ \  for }x\notin[0,1]\end{cases} (14)

is the characteristic function of the unit interval. [93.2.3] The Laplace convolution in the last line requires t<s. [93.2.4] The translations {\mathcal{T}}(t) on the other hand may be

[page 94, §0]    written as

{\mathcal{T}}(t)f(s)=f(s-t)=\int _{{-\infty}}^{{\infty}}f(s-y)\frac{1}{t}\delta\left(\frac{y}{t}-1\right)\;\mbox{\rm d}y=\int _{{0}}^{{s}}f(s-y)\frac{1}{t}\delta\left(\frac{y}{t}-1\right)\;\mbox{\rm d}y (15)

where again 0<t<s is required for the Laplace convolution in the last equation. [94.0.1] The similarity between eqs. (15) and (13) suggests to view the time translations {\mathcal{T}}(t) as a degenerate form of averaging f over a single point. [94.0.2] The operators \mbox{\rm M}(t) and {\mathcal{T}}(t) are both convolution operators. [94.0.3] By Lebesgues theorem \lim _{{t\to 0}}\mbox{\rm M}(t)f(s)=f(s) so that \mbox{\rm M}(0)f(t)=f(t) in analogy with eq. (8) which holds for {\mathcal{T}}(t). [94.0.4] However, while the translations {\mathcal{T}}(t) fulfill eq. (7) and form a convolution semigroup whose kernel is the Dirac measure at 1, the averaging operators \mbox{\rm M}(t) do not form a semigroup as will be seen below.

[94.1.1] The appearance of convolutions and convolution semigroups is not accidental. [94.1.2] Convolution operators arise quite generally from the symmetry requirement of eq. (10) above. [94.1.3] Let L^{p}(\mathbb{R}^{n}) denote the Lebesgue spaces of p-th power integrable functions, and let {\mathcal{S}} denote the Schwartz space of test functions for tempered distributions [27]. [94.1.4] It is well established that all bounded linear operators on L^{p}(\mathbb{R}^{n}) commuting with translations (i.e. fulfilling eq. (10)) are of convolution type [27].

Theorem 2.1

[94.1.5] Suppose the operator \mbox{\rm B}:L^{p}(\mathbb{R}^{n})\to L^{q}(\mathbb{R}^{n}), 1\leq p,q,\leq\infty is linear, bounded and commutes with translations. [94.1.6] Then there exists a unique tempered distribution g such that \mbox{\rm B}h=g*h for all h\in{\mathcal{S}}.

[94.2.1] For p=q=1 the tempered distributions in this theorem are finite Borel measures. [94.2.2] If the measure is bounded and positive this means that the operator B can be viewed as a weighted averaging operator. [94.2.3] In the following the case n=1 will be of interest. [94.2.4] A positive bounded measure \mu on \mathbb{R} is uniquely determined by its distribution function \widetilde{\mu}:\mathbb{R}\to[0,1] defined by

\widetilde{\mu}(x)=\frac{\mu(]-\infty,x[)}{\mu(\mathbb{R})}. (16)

[94.2.5] The tilde will again be omitted to simplify the notation. [94.2.6] Physically a weighted average \mbox{\rm M}(t;\mu)f(s) represents the measurement of a signal f(s) using an apparatus with response characterized by \mu and resolution t>0. [94.2.7] Note that the resolution (length of averaging interval) is a duration and cannot be negative.

[page 95, §1]   

Definition 2.1 (Averaging)

[95.1.1] Let \mu be a (probability) distribution function on \mathbb{R}, and t>0. [95.1.2] The weighted (time) average of a function f on \mathbb{R} is defined as the convolution

\mbox{\rm M}(t;\mu)f(s)=(f*\mu(\cdot/t))(s)=\int _{{-\infty}}^{\infty}f(s-s^{{\prime}})\;\mbox{\rm d}\mu(s^{{\prime}}/t)=\int _{{-\infty}}^{\infty}{\mathcal{T}}({s^{{\prime}}})f(s)\;\mbox{\rm d}\mu(s^{{\prime}}/t) (17)

whenever it exists. [95.1.3] The average is called causal if the support of \mu is in \mathbb{R}_{+}. [95.1.4] It is called degenerate if the support of \mu consists of a single point.

[95.2.1] The weight function or kernel m(x) corresponding to a distribution \mu(x) is defined as m(x)=\mbox{\rm d}\mu/\mbox{\rm d}x whenever it exists.

[95.3.1] The averaging operator \mbox{\rm M}(t) in eq. (11) corresponds to a measure with distribution function

\mu _{\chi}(x)=\begin{cases}0&\text{\ \ \ \  for }x\leq 0\\
x&\text{\ \ \ \  for }0\leq x\leq 1\\
1&\text{\ \ \ \  for }x\geq 1\end{cases} (18)

while the time translation {\mathcal{T}}(t) corresponds to the (Dirac) measure \delta(x-1) concentrated at 1 with distribution function

\mu _{\delta}(x)=\begin{cases}0&\text{\ \ \ \  for }x<1\\
1&\text{\ \ \ \  for }x\geq 1.\end{cases} (19)

[95.3.2] Both averages are causal, and the latter is degenerate.

[95.4.1] Repeated averaging leads to convolutions. [95.4.2] The convolution \kappa of two distributions \mu,\nu on \mathbb{R} is defined through

\kappa(x)=(\mu*\nu)(x)=\int _{{-\infty}}^{\infty}\mu(x-y)\mbox{\rm d}\nu(y)=\int _{{-\infty}}^{\infty}\nu(x-y)\mbox{\rm d}\mu(y). (20)

[95.4.3] The Fourier transform of a distribution is defined by

{\mathcal{F}}\left\{\mu(t)\right\}(\omega)=\widehat{\mu}(\omega)=\int _{{-\infty}}^{\infty}e^{{i\omega t}}\mbox{\rm d}\mu(t)=\int _{{-\infty}}^{\infty}e^{{i\omega t}}m(t)\;\mbox{\rm d}t (21)

where the last equation holds when the distribution admits a weight function. [95.4.4] A sequence \mu _{n}(x) of distributions is said to converge weakly to a limit \mu(x),

[page 96, §0]    written as

\lim _{{n\to\infty}}\mu _{n}=\mu, (22)

if

\lim _{{n\to\infty}}\int _{{-\infty}}^{\infty}f(x)\mbox{\rm d}\mu _{n}(x)=\int _{{-\infty}}^{\infty}f(x)\mbox{\rm d}\mu _{(}x) (23)

holds for all bounded continuous functions f.

[96.1.1] The operators \mbox{\rm M}(t) and {\mathcal{T}}(t) above have positive kernels, and preserve positivity in the sense that f\geq 0 implies \mbox{\rm M}(t)f\geq 0. [96.1.2] For such operators one has

Theorem 2.2

[96.1.3] Let T be a bounded operator on L^{p}(\mathbb{R}), 1\leq p<\infty that is translation invariant in the sense that

\displaystyle\mbox{\rm T}{\mathcal{T}}(t)f={\mathcal{T}}(t)\mbox{\rm T}f (24)

for all t\in\mathbb{R} and f\in L^{p}(\mathbb{R}), and such that f\in L^{p}(\mathbb{R}) and 0\leq f\leq 1 almost everywhere implies 0\leq\mbox{\rm T}f\leq 1 almost everywhere. [96.1.4] Then there exists a uniquely determined bounded measure \mu on \mathbb{R} with mass \mu(\mathbb{R})\leq 1 such that

\mbox{\rm T}f(t)=(\mu*f)(t)=\int _{{-\infty}}^{\infty}f(t-s)\mbox{\rm d}\mu(s) (25)
Proof.

[96.1.5] For the proof see [28]. ∎

[96.1.6] The preceding theorem suggests to represent those time evolutions that fulfill the requirements 1.– 4. of the last section in terms of convolution semigroups of measures.

Definition 2.2 (Convolution semigroup)

[96.1.7] A family \{\mu _{t}:t>0\} of positive bounded measures on \mathbb{R} with the properties that

\displaystyle\mu _{t}(\mathbb{R}) \displaystyle\leq 1\text{\ \ \ \ }{\rm for~}t>0, (26)
\displaystyle\mu _{{t+s}} \displaystyle=\mu _{t}*\mu _{s}\text{\ \ \ \ }{\rm for~}t,s>0, (27)
\displaystyle\delta \displaystyle=\lim _{{t\to 0}}\mu _{t}\text{\ \ \ \ } (28)

is called a convolution semigroup of measures on \mathbb{R}.

[page 97, §1]    [97.1.1] Here \delta is the Dirac measure at 0 and the limit is the weak limit. [97.1.2] The desired characterization of time evolutions now becomes

Corollary 2.1

[97.1.3] Let \mbox{\rm T}(t) be a strongly continuous time evolution fulfilling the conditions of homogeneity and causality, and being such that f\in L^{p}(\mathbb{R}) and 0\leq f\leq 1 almost everywhere implies 0\leq\mbox{\rm T}f\leq 1 almost everywhere. [97.1.4] Then \mbox{\rm T}(t) corresponds uniquely to a convolution semigroup of measures \mu _{t} through

\mbox{\rm T}(t)f(s)=(\mu _{t}*f)(s)=\int _{{-\infty}}^{\infty}f(s-s^{{\prime}})\mbox{\rm d}\mu _{t}(s^{{\prime}}) (29)

with \mathrm{supp}\,\mu _{t}\subset\mathbb{R}_{+} for all t\geq 0.

Proof.

[97.1.5] Follows from Theorem 2.2 and the observation that \mathrm{supp}\,\mu _{t}\cap\mathbb{R}_{-}\neq\emptyset would violate the causality condition. ∎

[97.2.1] Equation (29) establishes the basic convolution structure of the assertion in eq. (5). [97.2.2] It remains to investigate the requirement that \mbox{\rm T}(t) should arise from a coarse graining procedure, and to establish the nature of the kernel in eq. (5).

2.3 Time Averaging and Coarse Graining

[97.3.1] The purpose of this section is to motivate the definition of coarse graining. [97.3.2] A first possible candidate for a coarse grained macroscopic time evolution could be obtained by simply rescaling the time in a microscopic time evolution as

\mbox{\rm T}_{\infty}({\overline{t}})f(s)=\lim _{{\tau\to\infty}}\mbox{\rm T}_{\tau}({\overline{t}})f(s)=\lim _{{\tau\to\infty}}\mbox{\rm T}(\tau{\overline{t}})f(s)=\lim _{{\tau\to\infty}}f(s-\tau{\overline{t}}) (30)

where 0<{\overline{t}}<\infty would be macroscopic times. [97.3.3] However, apart from special cases, the limit will in general not exist. [97.3.4] Consider for example a sinusoidal f(t) oscillating around a constant. [97.3.5] Also, the infinite translation \mbox{\rm T}_{\infty} is not an average, and this conflicts with the requirement above, that coarse graining should be a smoothing operation.

[97.4.1] A second highly popular candidate for coarse graining is therefore the averaging operator \mbox{\rm M}(t). [97.4.2] If the limit t\to\infty exists and f(t) is integrable in the finite interval [s_{1},s_{2}] then the average

{\overline{f}}=\lim _{{t\to\infty}}\mbox{\rm M}(t)f(s_{1})=\lim _{{t\to\infty}}\mbox{\rm M}(t)f(s_{2}) (31)

is a number independent of the instant s_{i}. [97.4.3] Thus, if one wants to study the macroscopic time dependence of {\overline{f}}, it is necessary to consider a scaling limit in

[page 98, §0]    which also s\to\infty. [98.0.1] If the scaling limit s,t\to\infty is performed such that s/t={\overline{s}} is constant, then

\lim _{{\substack{t,s\to\infty\\
s=t{\overline{s}}}}}\mbox{\rm M}(t)f(s)=\int _{{{\overline{s}}-1}}^{{{\overline{s}}}}f_{\infty}(z)\;\mbox{\rm d}z=\mbox{\rm M}(1)f_{\infty}({\overline{s}}) (32)

becomes again an averaging operator over the infinitely rescaled observable. [98.0.2] Now \mbox{\rm M}(1) still does not qualify as a coarse grained time evolution because \mbox{\rm M}(1)\mbox{\rm M}(1)\neq\mbox{\rm M}(2) as will be shown next.

[98.1.1] Consider again the operator \mbox{\rm M}(t) defined in eq. (11). [98.1.2] It follows that

\mbox{\rm M}^{2}(t)f(s)=\left(\frac{1}{t}\chi _{{[0,1]}}\left(\frac{\cdot}{t}\right)*\frac{1}{t}\chi _{{[0,1]}}\left(\frac{\cdot}{t}\right)*f\right)(s) (33)

and

\frac{1}{t^{2}}\int^{x}_{0}\chi _{{[0,1]}}\left(\frac{x-y}{t}\right)\chi _{{[0,1]}}\left(\frac{y}{t}\right)\;\mbox{\rm d}y=\begin{cases}0&\text{\ \ \ \  for }x\leq 0\\
\frac{x}{t^{2}}&\text{\ \ \ \  for }0\leq x\leq t\\
\frac{2}{t}-\frac{x}{t^{2}}&\text{\ \ \ \  for }t\leq x\leq 2t\\
0&\text{\ \ \ \  for }x\geq 2t.\end{cases} (34)

[98.1.3] Thus twofold averaging may be written as

\mbox{\rm M}^{2}(t)f(s)=\int _{0}^{s}f(s-y)\frac{1}{t}\chi^{{(2)}}\left(\frac{y}{t}\right)\;\mbox{\rm d}y (35)

where

\chi^{{(2)}}(x)=\begin{cases}x&\text{\ \ \ \  for $0\leq x\leq 1$}\\
2-x&\text{\ \ \ \  for $1\leq x\leq 2$}\\
0&\text{\ \ \ \  otherwise}\end{cases} (36)

is the new kernel. [98.1.4] It follows that \mbox{\rm M}^{2}(t)\neq\mbox{\rm M}(2t), and hence the averaging operators \mbox{\rm M}(t) do not form a semigroup.

[98.2.1] Although \mbox{\rm M}^{2}(t)\neq\mbox{\rm M}(2t) the iterated average is again a convolution operator with support [0,2t] compared to [0,t] for \mbox{\rm M}(t). [98.2.2] Similarly, \mbox{\rm M}^{3}(t) has support [0,3t]. [98.2.3] This suggests to investigate the iterated average \mbox{\rm M}^{n}(t)f(s) in a scaling limit n,s\to\infty. [98.2.4] The limit n\to\infty smoothes the function by enlarging the

[page 99, §0]    averaging window to [0,nt], and the limit s\to\infty shifts the origin to infinity. [99.0.1] The result may be viewed as a coarse grained time evolution in the sense of a time evolution on time scales "longer than infinitely long". [99.0.2] c (This is a footnote:) c The scaling limit was called "ultralong time limit" in [10] It is therefore necessary to rescale s. [99.0.3] If the rescaling factor is called \sigma _{n}>0 one is interested in the limit n,s\to\infty with {\overline{s}}=s/\sigma _{n} fixed, and \sigma _{n}\to\infty with n\to\infty and fixed t>0

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t)^{n}f)(s)=\lim _{{n\to\infty}}(\mbox{\rm M}(t)^{n}f)(\sigma _{n}{\overline{s}}) (37)

whenever this limit exists. [99.0.4] Here {\overline{s}}>1 denotes the macroscopic time.

[99.1.1] To evaluate the limit note first that eq. (11) implies

\mbox{\rm M}(t)f(\sigma _{n}{\overline{s}})=\int^{{{\overline{s}}}}_{0}f_{{\sigma _{n}}}({\overline{s}}-z)\frac{\sigma _{n}}{t}\chi _{{[0,1]}}\left(\frac{\sigma _{n}z}{t}\right)\;\mbox{\rm d}z (38)

where f_{\tau}(t)=f(t\tau) denotes the rescaled observable with a rescaling factor \tau. [99.1.2] The n-th iterated average may now be calculated by Laplace transformation with respect to {\overline{s}}. [99.1.3] Note that

{\mathcal{L}}\left\{\frac{1}{c}\chi _{{[0,1]}}\left(\frac{x}{c}\right)\right\}(u)=\frac{1-e^{{-cu}}}{cu}=E_{{1,2}}(-cu) (39)

for all c\in\mathbb{R}, where E_{{1,2}}(x) is the generalized Mittag-Leffler function defined as

E_{{a,b}}(x)=\sum _{{k=0}}^{\infty}\frac{x^{k}}{\Gamma(ak+b)} (40)

for all a>0 and b\in\mathbb{C}. [99.1.4] Using the general relation

E_{{a,b}}(x)=\frac{1}{\Gamma(b)}+xE_{{a,a+b}}(x) (41)

gives with eqs. (37) and (38)

{\mathcal{L}}\left\{\mbox{\rm M}(t)^{n}f(\sigma _{n}{\overline{s}})\right\}({\overline{u}})=\left(1-\frac{t{\overline{u}}}{\sigma _{n}}E_{{1,3}}\left(-\frac{t{\overline{u}}}{\sigma _{n}}\right)\right)^{n}\frac{1}{\sigma _{n}}{\mathcal{L}}\left\{ f(s)\right\}\left(\frac{{\overline{u}}}{\sigma _{n}}\right) (42)

where f({\overline{u}}) is the Laplace transform of f({\overline{s}}). [99.1.5] Noting that E_{{1,3}}(0)=1/2 it becomes apparent that a limit n\to\infty will exist if the rescaling factors are

[page 100, §0]    chosen as \sigma _{n}\sim n. [100.0.1] With the choice \sigma _{n}=\sigma n/2 and \sigma>0 one finds for the first factor

\lim _{{n\to\infty}}\left(1-\frac{2t{\overline{u}}}{n\sigma}E_{{1,3}}\left(-\frac{2t{\overline{u}}}{n\sigma}\right)\right)^{n}=e^{{-t{\overline{u}}/\sigma}}. (43)

[100.0.2] Concerning the second factor assume that for each {\overline{u}} the limit

\lim _{{n\to\infty}}\frac{2}{n}{\mathcal{L}}\left\{ f(s)\right\}\left(\frac{2{\overline{u}}}{n}\right)={\overline{f}}({\overline{u}}) (44)

exists and defines a function {\overline{f}}({\overline{u}}). [100.0.3] Then

\lim _{{n\to\infty}}\frac{1}{\sigma _{n}}{\mathcal{L}}\left\{ f({\overline{s}})\right\}\left(\frac{{\overline{u}}}{\sigma _{n}}\right)=\frac{1}{\sigma}{\overline{f}}\left(\frac{{\overline{u}}}{\sigma}\right), (45)

and it follows that

\lim _{{n\to\infty}}{\mathcal{L}}\left\{\mbox{\rm M}(t)^{n}f(\sigma _{n}{\overline{s}})\right\}({\overline{u}})=e^{{-t{\overline{u}}/\sigma}}\frac{1}{\sigma}{\overline{f}}\left(\frac{{\overline{u}}}{\sigma}\right). (46)

[100.0.4] With {\overline{t}}=t/\sigma Laplace inversion yields

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t)^{n}f)(s)=\int _{{0}}^{{{\overline{s}}}}{\overline{f}}(\sigma{\overline{s}}-\sigma{\overline{y}})\delta({\overline{y}}-{\overline{t}})\;\mbox{\rm d}{\overline{y}}={\overline{f}}_{\sigma}({\overline{s}}-{\overline{t}}). (47)

[100.0.5] Using eq. (12) the result (47) may be expressed symbolically as

\lim _{{\substack{n,s\to\infty\\
s/n=\sigma{\overline{s}}/2}}}\left(\frac{1}{t}\int^{t}_{0}{\mathcal{T}}(y)\;\mbox{\rm d}y\right)^{n}f(s)={\overline{f}}_{\sigma}({\overline{s}}-{\overline{t}})=\overline{{\mathcal{T}}}({{\overline{t}}})\;{\overline{f}}_{\sigma}({\overline{s}}) (48)

with {\overline{t}}=t/\sigma. [100.0.6] This expresses the macroscopic or coarse grained time evolution \overline{{\mathcal{T}}}({{\overline{t}}}) as the scaling limit of a microscopic time evolution {\mathcal{T}}(t). [100.0.7] Note that there is some freedom in the choice of the rescaling factors \sigma _{n} expressed by the prefactor \sigma. [100.0.8] This freedom reflects the freedom to choose the time units for the coarse grained time evolution.

[100.1.1] The coarse grained time evolution \overline{{\mathcal{T}}}({{\overline{t}}}) is again a translation. [100.1.2] The coarse grained observable {\overline{f}}({\overline{s}}) corresponds to a microscopic average by virtue of the following result [29].

[page 101, §1]   

Proposition 2.1

[101.1.1] If f(x) is bounded from below and one of the limits

\lim _{{y\to\infty}}\frac{1}{y}\int _{0}^{y}f(x)\;\mbox{\rm d}x

or

\lim _{{z\to 0}}z\int _{0}^{\infty}f(x)e^{{-zx}}\;\mbox{\rm d}x

exists then the other limit exists and

\lim _{{y\to\infty}}\frac{1}{y}\int _{0}^{y}f(x)\;\mbox{\rm d}x=\lim _{{z\to 0}}z{\mathcal{L}}\left\{ f(x)\right\}(z). (49)

[101.1.2] Comparison of the last relation with eq. (44) shows that {\overline{f}}({\overline{s}}) is a microscopic average of f(s). [101.1.3] While s is a microscopic time coordinate, the time coordinate {\overline{s}} of {\overline{f}} is macroscopic.

[101.2.1] The preceding considerations justify to view the time evolution \overline{{\mathcal{T}}}({{\overline{t}}}) as a coarse grained time evolution. [101.2.2] Every observation or measurement of a physical quantity f(s) requires a minimum duration t determined by the temporal resolution of the measurement apparatus. [101.2.3] The value f(s) at the time instant s is always an average over this minimum time interval. [101.2.4] The averaging operator \mbox{\rm M}(t) with kernel \chi _{{[0,1]}} defined in equation (11) represents an idealized averaging apparatus that can be switched on and off instantaneously, and does not otherwise influence the measurement. [101.2.5] In practice one is usually confronted with finite startup and shutdown times and a nonideal response of the apparatus. [101.2.6] These imperfections are taken into account by using a weighted average with a weight function or kernel that differs from \chi _{{[0,1]}}. [101.2.7] The weight function reflects conditions of the measurement, as well as properties of the apparatus and its interaction with the system. [101.2.8] It is therefore of interest to consider causal averaging operators \mbox{\rm M}(t;\mu) defined in eq. (17) with general weight functions. [101.2.9] A general coarse graining procedure is then obtained from iterating these weighted averages.

Definition 2.3 (Coarse Graining)

[101.2.10] Let \mu be a probability distribution on \mathbb{R}, and \sigma _{n}>0, n\in\mathbb{N} a sequence of rescaling factors. A coarse graining limit is defined as

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t;\mu)^{n}f)(s) (50)

[page 102, §0]    whenever the limit exists. [102.0.1] The coarse graining limit is called causal if \mbox{\rm M}(t;\mu) is causal, i.e. if \mathrm{supp}\,\mu\subset\mathbb{R}_{+}.

2.4 Coarse Graining Limits and Stable Averages

[102.1.1] The purpose of this section is to investigate the coarse graining procedure introduced in Definition 2.3. [102.1.2] Because the coarse graining procedure is defined as a limit it is useful to recall the following well known result for limits of distribution functions [30]. [102.1.3] For the convenience of the reader its proof is reproduced in the appendix.

Proposition 2.2

[102.1.4] Let \mu _{n}(s) be a weakly convergent sequence of distribution functions. [102.1.5] If \lim _{{n\to\infty}}\mu _{n}(s)=\mu(s), where \mu(s) is nondegenerate then for any choice of a_{n}>0 and b_{n} there exist a>0 and b such that

\lim _{{n\to\infty}}\mu _{n}(a_{n}x+b_{n})=\mu(ax+b). (51)

[102.2.1] The basic result for coarse graining limits can now be formulated.

Theorem 2.3 (Coarse Graining Limit)

[102.2.2] Let f(s) be such that the limit \lim _{{a\to 0}}a\widehat{f}(a\omega)=\widehat{{\overline{f}}}(\omega) defines the Fourier transform of a function {\overline{f}}(s). [102.2.3] Then the coarse graining limit exists and defines a convolution operator

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t;\mu)^{n}f)(s)=\int _{{-\infty}}^{\infty}{\overline{f}}({\overline{s}}-{\overline{s}}^{{\prime}})\;\mbox{\rm d}\nu({\overline{s}}^{{\prime}}/t;\mu) (52)

if and only if for any a_{1},a_{2}>0 there are constants a>0 and b such that the distribution function \nu(x)=\nu(x;\mu) obeys the relation

\nu(a_{1}x)*\nu(a_{2}x)=\nu(ax+b). (53)
Proof.

[102.2.4] In the previous section the coarse graining limit was evaluated for the distribution \mu _{\chi} from eq. (18) and the corresponding \nu was found in eq. (47) to be degenerate. [102.2.5] A degenerate distribution \nu trivially obeys eq. (53). [102.2.6] Assume therefore from now on that neither \mu nor \nu are degenerate.

[102.3.1] Employing equation (17) in the form

\mbox{\rm M}(t;\mu)f(\sigma _{n}{\overline{s}})=\int _{{-\infty}}^{\infty}f(\sigma _{n}{\overline{s}}-\sigma _{n}y)\mbox{\rm d}\mu(\sigma _{n}y/t) (54)

[page 103, §0]    one computes the Fourier transformation of \mbox{\rm M}(t;\mu)^{n}f with respect to {\overline{s}}

{\mathcal{F}}\left\{\mbox{\rm M}(t;\mu)^{n}f(\sigma _{n}{\overline{s}})\right\}({\overline{\omega}})=\left[\widehat{\mu}\left(\frac{t{\overline{\omega}}}{\sigma _{n}}\right)\right]^{n}\frac{1}{\sigma _{n}}\widehat{f}\left(\frac{{\overline{\omega}}}{\sigma _{n}}\right). (55)

[103.0.1] By assumption \widehat{f}({\overline{\omega}}/\sigma _{n})/\sigma _{n} has a limit whenever \sigma _{n}\to\infty with n\to\infty. [103.0.2] Thus the coarse graining limit exists and is a convolution operator whenever [\widehat{\mu}(t{\overline{\omega}}/\sigma _{n})]^{n} converges to \widehat{\nu}({\overline{\omega}}) as n\to\infty. [103.0.3] Following [30] it will be shown that this is true if and only if the characterization (53) and \sigma _{n}\to\infty with n\to\infty apply. [103.0.4] To see that

\lim _{{n\to\infty}}\sigma _{n}=\infty (56)

holds, assume the contrary. Then there is a subsequence \sigma _{{n_{k}}} converging to a finite limit. [103.0.5] Thus

|\widehat{\mu}(t\omega/\sigma _{{n_{k}}})|^{{n_{k}}}=|\widehat{\nu}(\omega)|(1+{\it o}(1)) (57)

so that

|\widehat{\mu}(\omega)|=|\widehat{\nu}(\omega\sigma _{{n_{k}}}/t)|^{{1/n_{k}}}(1+{\it o}(1)) (58)

for all \omega. [103.0.6] As n_{k}\to\infty this leads to |\widehat{\mu}(\omega)|=1 for all \omega and hence \mu must be degenerate contrary to assumption.

[103.1.1] Next, it will be shown that

\lim _{{n\to\infty}}\frac{\sigma _{{n+1}}}{\sigma _{n}}=1. (59)

[103.1.2] From eq. (56) it follows that \lim _{{n\to\infty}}|\widehat{\mu}(\omega/\sigma _{n})|=1 and therefore

|\widehat{\mu}(t\omega/\sigma _{n})|^{n}=|\widehat{\nu}(\omega)|(1+{\it o}(1)) (60)

and

|\widehat{\mu}(t\omega/\sigma _{{n+1}})|^{{n+1}}=|\widehat{\nu}(\omega)|(1+{\it o}(1)). (61)

Substituting \omega by \sigma _{n}\omega/\sigma _{{n+1}} in eq. (60) and by \sigma _{{n+1}}\omega/\sigma _{n} in eq. (61) shows that

\lim _{{n\to\infty}}\left|\frac{\widehat{\nu}(\sigma _{{n+1}}\omega/\sigma _{n})}{\widehat{\nu}(\omega)}\right|=\lim _{{n\to\infty}}\left|\frac{\widehat{\nu}(\sigma _{n}\omega/\sigma _{{n+1}})}{\widehat{\nu}(\omega)}\right|=1. (62)

[page 104, §0]    [104.0.1] If \lim _{{n\to\infty}}\sigma _{{n+1}}/\sigma _{n}\neq 1 then there exists a subsequence of either (\sigma _{{n+1}}/\sigma _{n}) or (\sigma _{n}/\sigma _{{n+1}}) converging to a constant A<1. [104.0.2] Therefore eq. (62) implies \widehat{\nu}(\omega)=\widehat{\nu}(A\omega) which upon iteration yields

|\widehat{\nu}(\omega)|=|\widehat{\nu}(A^{n}\omega)|. (63)

[104.0.3] Taking the limit n\to\infty then gives |\widehat{\nu}(0)|=1 implying that \nu is degenerate contrary to assumption.

[104.1.1] Now let 0<a_{1}<a_{2} be two constants. [104.1.2] Because of (56) and (59) it is possible to choose for each \varepsilon>0 and sufficiently large n>n_{0}(\varepsilon) an index m(n) such that

0\leq\frac{\sigma _{m}}{\sigma _{n}}-\frac{a_{2}}{a_{1}}<\varepsilon. (64)

[104.1.3] Consider the identity

\left[\widehat{\mu}\left(\frac{a_{1}t{\overline{\omega}}}{\sigma _{n}}\right)\right]^{{n+m}}=\left[\widehat{\mu}\left(\frac{a_{1}t{\overline{\omega}}}{\sigma _{n}}\right)\right]^{n}\left[\widehat{\mu}\left(\frac{\sigma _{m}}{\sigma _{n}}\frac{a_{1}t{\overline{\omega}}}{\sigma _{m}}\right)\right]^{m}. (65)

By hypothesis the distribution functions corresponding to \left[\widehat{\mu}\left(t{\overline{\omega}}/\sigma _{n}\right)\right]^{n} converge to \nu({\overline{s}}) as n\to\infty. [104.1.4] Hence each factor on the right hand side converges and their product converges to \nu(a_{1}{\overline{s}})*\nu(a_{2}{\overline{s}}). [104.1.5] It follows that the distribution function on the left hand side must also converge. [104.1.6] By Proposition 2.2 there must exist a>0 and b such that the left hand side differs from \nu({\overline{s}}) only as \nu(a{\overline{s}}+b).

[104.2.1] Finally the converse direction that the coarse graining limit exists for \mu=\nu is seen to follow from eq. (53). [104.2.2] This concludes the proof of the theorem. ∎

[104.3.1] The theorem shows that the coarse graining limit, if it exists, is again a macroscopic weighted average \mbox{\rm M}(t;\nu). [104.3.2] The condition (53) says that this macroscopic average has a kernel that is stable under convolutions, and this motivates the

Definition 2.4 (Stable Averages)

[104.3.3] A weighted averaging operator \mbox{\rm M}(t;\mu) is called stable if for any a_{1},a_{2}>0 there are constants a>0 and b\in\mathbb{R} such that

\mu(a_{1}x)*\mu(a_{2}x)=\mu(ax+b) (66)

holds.

[104.4.1] This nomenclature emphasizes the close relation with the limit theorems of probability theory [30, 31]. [104.4.2] The next theorem provides the explicit form for distribution functions satisfying eq. (66). [104.4.3] The proof uses Bernsteins theorem and hence requires the concept of complete monotonicity.

[page 105, §1]   

Definition 2.5

[105.1.1] A C^{\infty}-function f:]0,\infty[\to\mathbb{R} is called completely monotone if

(-1)^{n}\frac{\mbox{\rm d}^{n}f}{\mbox{\rm d}x^{n}}\geq 0 (67)

for all integers n\geq 0.

[105.2.1] Bernsteins theorem [31, p. 439] states that a function is completely monotone if and only if it is the the Laplace transform (u>0)

\mu(u)={\mathcal{L}}\left\{\mu(x)\right\}(u)=\int _{0}^{\infty}e^{{-ux}}\mbox{\rm d}\mu(x)=\int _{0}^{\infty}e^{{-ux}}m(x)\;\mbox{\rm d}x (68)

of a distribution \mu or of a density m=\mbox{\rm d}\mu/\mbox{\rm d}x.

[105.3.1] In the next theorem the explicit form of stable averaging kernels is found to be a special case of the general H-function. [105.3.2] Because the H-function will reappear in other results its general definition and properties are presented separately in Section 4.

Theorem 2.4

[105.3.3] A causal average is stable if and only if its weight function is of the form

h_{\alpha}(x;b,c)=\frac{1}{b^{{1/\alpha}}}h_{\alpha}\left(\frac{x-c}{b^{{1/\alpha}}}\right)=\frac{1}{\alpha(x-c)}H_{{11}}^{{10}}\left(\frac{b^{{1/\alpha}}}{x-c}\left|\begin{array}[]{l}{(0,1)}\\
{(0,1/\alpha)}\end{array}\right.\right) (69)

where 0<\alpha\leq 1, b>0 and c\in\mathbb{R} are constants and h_{\alpha}(x)=h_{\alpha}(s;1,0).

Proof.

[105.3.4] Let c=0 without loss of generality. [105.3.5] The condition (66) together with \mathrm{supp}\,\mu\subset[0,\infty[ defines one sided stable distribution functions [31]. [105.3.6] To derive the form (69) it suffices to consider condition (66) with b=0. [105.3.7] Assume thence that for any a_{1},a_{2}>0 there exists a>0 such that

\mu(a_{1}x)*\mu(a_{2}x)=\mu(ax) (70)

where the convolution is now a Laplace convolution because of the condition \mathrm{supp}\,\subset[0,\infty[. [105.3.8] Laplace tranformation yields

\mu(u/a_{1})\mu(u/a_{2})=\mu(u/a). (71)

[105.3.9] Iterating this equation (with a_{1}=a_{2}=1) shows that there is an n-dependent constant a(n) such that

\mu(u)^{n}=\mu(u/a(n)) (72)

[page 106, §0]    and hence

\mu\left(\frac{u}{a(nm)}\right)=\mu(u)^{{nm}}=\mu\left(\frac{u}{a(n)}\right)^{m}=\mu\left(\frac{u}{(a(n)a(m)}\right). (73)

[106.0.1] Thus a(n) satisfies the functional equation

a(nm)=a(n)a(m) (74)

whose solution is a(n)=n^{{1/\gamma}} with some real constant written as 1/\gamma with hindsight. [106.0.2] Inserting a(n) into eq.(72) and substituting the function g(x)=\log\mu(x) gives

ng(u)=g(un^{{-1/\gamma}}). (75)

[106.0.3] Taking logarithms and substituting f(x)=\log g(e^{x}) this becomes

\log n+f(\log u)=f\left(\log u-\frac{\log n}{\gamma}\right). (76)

[106.0.4] The solution to this functional equation is f(x)=-\gamma x. [106.0.5] Substituting back one finds g(x)=x^{{-\gamma}} and therefore \mu(u) is of the general form \mu(u)=\exp(u^{{-\gamma}}) with \gamma\in\mathbb{R}. [106.0.6] Now \mu is also a distribution function. Its normalization requires \mu(u=0)=1 and this restricts \gamma to \gamma<0. [106.0.7] Moreover, by Bernsteins theorem \mu(u) must be completely monotone. [106.0.8] A completely monotone function is positive, decreasing and convex. [106.0.9] Therefore the power in the exponent must have a negative prefactor, and the exponent is restricted to the range -1\leq\gamma<0. [106.0.10] Summarizing, the Laplace transform \mu(u) of a distribution satisfying (70) is of the form

\mu(u)=h_{\alpha}(u;b,0)=e^{{-bu^{\alpha}}} (77)

with 0<\alpha\leq 1 and b>0. [106.0.11] Checking that h_{\alpha}(u;b,0) does indeed satisfy eq. (70) yields a^{{-\alpha}}=a_{1}^{{-\alpha}}+a_{2}^{{-\alpha}} as the relation between the constants. [106.0.12] For the proof of the general case of eq. (66) see Refs. [30, 31].

[106.1.1] To invert the Laplace transform it is convenient to use the relation

{\mathcal{M}}\left\{ m(x)\right\}(s)=\frac{{\mathcal{M}}\left\{{\mathcal{L}}\left\{ m(x)\right\}(u)\right\}(1-s)}{\Gamma(1-s)} (78)

between the Laplace transform and the Mellin transform

{\mathcal{M}}\left\{ m(x)\right\}(s)=\int _{0}^{\infty}x^{{s-1}}m(t)\;\mbox{\rm d}x (79)

[page 107, §0]    of a function m(x). [107.0.1] Using the Mellin transform [32]

{\mathcal{M}}\left\{ e^{{-bx^{\alpha}}}\right\}(s)=\frac{\Gamma(s/\alpha)}{\alpha b^{{s/\alpha}}} (80)

valid for \alpha>0 and \mathrm{Re}\, s>0 it follows that

{\mathcal{M}}\left\{ h_{\alpha}(x;b,0)\right\}(s)=\frac{1}{\alpha b^{{(1-s)/\alpha}}}\frac{\Gamma((1-s)/\alpha)}{\Gamma(1-s)}. (81)

[107.0.2] The general relation {\mathcal{M}}\left\{ x^{{-1}}f(x^{{-1}})\right\}(s)={\mathcal{M}}\left\{ f(x)\right\}(1-s) then implies

{\mathcal{M}}\left\{ x^{{-1}}h_{\alpha}(x^{{-1}};b,0)\right\}(s)=\frac{1}{\alpha b^{{s/\alpha}}}\frac{\Gamma(s/\alpha)}{\Gamma(s)} (82)

which leads to

x^{{-1}}h_{\alpha}(x^{{-1}};b,0)=\frac{1}{\alpha}H_{{11}}^{{10}}\left(b^{{1/\alpha}}x\left|\begin{array}[]{l}{(0,1)}\\
{(0,1/\alpha)}\end{array}\right.\right) (83)

by identification with eq. (153) below. [107.0.3] Restoring a shift c\neq 0 yields the result of eq. (69). ∎

[107.0.4] Note that h_{\alpha}(x)=h_{\alpha}(s;1,0) is the standardized form used in eq. (5). [107.0.5] It remains to investigate the sequence of rescaling factors \sigma _{n}. [107.0.6] For these one finds

Corollary 2.2

[107.0.7] If the coarse graining limit exists and is nondegenerate then the sequence \sigma _{n} of rescaling factors has the form

\sigma _{n}=n^{{1/\alpha}}\Lambda(n) (84)

where 0<\alpha\leq 1 and \Lambda(n) is slowly varying, i.e. \lim _{{n\to\infty}}\Lambda(bn)/\Lambda(n)=1 for all b>0 (see Chapter IX, Section 2.3).

Proof.

[33][107.0.8] Let \widehat{\mu}_{n}(\omega)=\widehat{\mu}(\omega)^{n}. [107.0.9] Then for all \omega and any fixed k

|\widehat{\mu}_{n}(\omega/\sigma _{n})|=e^{{-b|\omega|^{\alpha}}}(1+{\it o}(1))=|\widehat{\mu}_{{kn}}(\omega/\sigma _{{kn}})|. (85)

[107.0.10] On the other hand

|\widehat{\mu}_{{kn}}(\omega/\sigma _{{kn}})|=|\widehat{\mu}_{{n}}((\omega\sigma _{n}/\sigma _{{kn}})/\sigma _{n})|^{k}=e^{{-b|\omega|^{\alpha}}}(1+{\it o}(1)) (86)

where the remainder tends uniformly to zero on every finite interval. [107.0.11] Suppose that the sequence \sigma _{n}/\sigma _{{kn}} is unbounded so that there is a subsequence with \sigma _{{kn_{j}}}/\sigma _{{n_{j}}}\to~0. [107.0.12] Setting \omega=\sigma _{{kn_{j}}}/\sigma _{{n_{j}}} in eq. (86) and using eq. (85) gives

[page 108, §0]    \exp(-bk)=1 which cannot be satisfied because b,k>0. [108.0.1] Hence \sigma _{n}/\sigma _{{kn}} is bounded. [108.0.2] Now the limit n\to\infty in eqs. (85) and (86) gives

e^{{-b|\omega|^{\alpha}}}=e^{{-bk|\omega|^{\alpha}(\sigma _{n}/\sigma _{{kn}})^{\alpha}}}(1+{\it o}(1)). (87)

[108.0.3] This requires that

\lim _{{n\to\infty}}\frac{\sigma _{{kn}}}{\sigma _{n}}=k^{{1/\alpha}} (88)

implying eq. (84) by virtue of the Characterization Theorem 2.2 in Chapter IX. [108.0.4] (For more information on slow and regular variation see Chapter IX  and references therein). ∎

2.5 Macroscopic Time Evolutions

[108.1.1] The preceding results show that a coarse graining limit is characterized by the quantities (\alpha,b,c,\Lambda). [108.1.2] These quantities are determined by the coarsening weight \mu. [108.1.3] The following result, whose proof can be found in [33, p. 85], gives their relation with the coarsening weight.

Theorem 2.5 (Universality Classes of Time Evolutions)

[108.1.4] In order that a causal coarse graining limit based on \mbox{\rm M}(t;\mu) gives rise to a macroscopic average with h_{\alpha}(x;b,c) it is necessary and sufficient that \widehat{\mu}(\omega) behaves as

\log\widehat{\mu}(\omega)=ic\omega-b|\omega|^{\alpha}\Lambda(\omega) (89)

in a neighbourhood of \omega=0, and that \Lambda(\omega) is slowly varying for \omega\to 0. [108.1.5] In case 0<\alpha\leq 1 the rescaling factors can be chosen as

\sigma _{n}^{{-1}}=\inf\{\omega>0:|\omega|^{\alpha}\Lambda(\omega)=b/n\} (90)

while the case \alpha>1 reduces to the degenerate case \alpha=1.

[108.2.1] The preceding theorem characterizes the domain of attraction of a universality class of time evolutions. [108.2.2] Summarizing the results gives a characterization of macroscopic time evolutions arising from coarse graining limits.

Theorem 2.6 (Macroscopic Time Evolution)

[108.2.3] Let f(s) be such that the limit \lim _{{a\to 0}}a\widehat{f}(a\omega)=\widehat{{\overline{f}}}(\omega) defines the Fourier transform of a function {\overline{f}}(s). [108.2.4] If \mbox{\rm M}(t;\mu) is a causal average whose coarse graining limit exists with \alpha,b,c as

[page 109, §0]    in the preceding theorem then

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t;\mu)^{n}f)(s)=\int\limits _{{\overline{c}}}^{\infty}{\overline{f}}({\overline{s}}-y)h_{\alpha}\left(\frac{y}{{\overline{t}}}\right)\frac{\mbox{\rm d}y}{{\overline{t}}}=\int\limits _{{\overline{c}}}^{\infty}\overline{\mathcal{T}}_{y}{\overline{f}}({\overline{s}})h_{\alpha}\left(\frac{y}{{\overline{t}}}\right)\frac{\mbox{\rm d}y}{{\overline{t}}}=\mbox{\rm M}({\overline{t}};h_{\alpha}){\overline{f}}({\overline{s}}-{\overline{c}})={\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}){\overline{f}}({\overline{s}}-{\overline{c}}) (91)

defines a family of one parameter semigroups {\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}) with parameter {\overline{t}}=t^{\alpha}b indexed by \alpha. [109.0.1] Here \overline{\mathcal{T}}_{{\overline{t}}}{\overline{f}}({\overline{s}})={\overline{f}}({\overline{s}}-{\overline{t}}) denotes the translation semigroup, and {\overline{c}}=c/(tb)^{{1/\alpha}} is a constant.

Proof.

[109.0.2] Noting that \mathrm{supp}\, h_{\alpha}(x)\subset\mathbb{R}_{+} and combining Theorems 2.3 and 2.4 gives

\lim _{{\substack{n,s\to\infty\\
s=\sigma _{n}{\overline{s}}}}}(\mbox{\rm M}(t;\mu)^{n}f)(s)=\int _{{c}}^{\infty}{\overline{f}}({\overline{s}}-{\overline{s}}^{{\prime}})\frac{1}{tb^{{1/\alpha}}}h_{\alpha}\left(\frac{{\overline{s}}^{{\prime}}-c}{tb^{{1/\alpha}}}\right)\;\mbox{\rm d}{\overline{s}}^{{\prime}}={\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}){\overline{f}}({\overline{s}}-{\overline{c}}) (92)

where 0<\alpha\leq 1, b>0 and c\in\mathbb{R} are the constants from theorem 2.4 and the last equality defines the operators {\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}) with {\overline{t}}=t^{\alpha}b and {\overline{c}}=c/(tb)^{{1/\alpha}}. [109.0.3] Fourier transformation then yields

{\mathcal{F}}\left\{({\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}){\overline{f}})({\overline{s}}-{\overline{c}})\right\}({\overline{\omega}})=e^{{-ic{\overline{\omega}}-{\overline{t}}(i{\overline{\omega}})^{\alpha}}}, (93)

and the semigroup property (7) follows from

{\mathcal{F}}\left\{({\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}_{1}){\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}_{2}){\overline{f}})({\overline{s}}-{\overline{c}})\right\}({\overline{\omega}})=e^{{-ic{\overline{\omega}}-{\overline{t}}_{1}(i{\overline{\omega}})^{\alpha}-{\overline{t}}_{2}(i{\overline{\omega}})^{\alpha}}}={\mathcal{F}}\left\{({\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}_{1}+{\overline{t}}_{2}){\overline{f}})({\overline{s}}-{\overline{c}})\right\}({\overline{\omega}}) (94)

by Fourier inversion. [109.0.4] Condition (8) is checked similarly. ∎

[109.0.5] The family of semigroups {\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}) indexed by \alpha that can arise from coarse graining limits are called macroscopic time evolutions. [109.0.6] These semigroups are also holomorphic, strongly continuous and equibounded (see Chapter III).

[109.1.1] From a physical point of view this result emphasizes the different role played by {\overline{s}} and {\overline{t}}. [109.1.2] While {\overline{s}} is the macroscopic time coordinate whose values are {\overline{s}}\in\mathbb{R}, the duration {\overline{t}}>0 is positive. [109.1.3] If the dimension of a microscopic time duration t is [s], then the dimension of the macroscopic time duration {\overline{t}} is [s{}^{\alpha}].

[page 110, §1]   

2.6 Infinitesimal Generators

[110.1.1] The importance of the semigroups {\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}) for theoretical physics as universal attractors of coarse grained macroscopic time evolutions seems not to have been noticed thus far. [110.1.2] This is the more surprising as their mathematical importance for harmonic analysis and probability theory has long been recognized [31, 34, 35, 28]. [110.1.3] The infinitesimal generators are known to be fractional derivatives [31, 35, 36, 37]. [110.1.4] The infinitesimal generators are defined as

\mbox{\rm A}_{\alpha}{\overline{f}}({\overline{s}})=\lim _{{{\overline{t}}\to 0}}\frac{{\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}){\overline{f}}({\overline{s}})-{\overline{f}}({\overline{s}})}{{\overline{t}}}. (95)

[110.1.5] For more details on semigroups and their infinitesimal generators see Chapter III.

[110.2.1] Formally one calculates \mbox{\rm A}_{\alpha} by applying direct and inverse Laplace transformation with {\overline{c}}=0 in eq. (91) and using eq. (77)

\mbox{\rm A}_{\alpha}{\overline{f}}({\overline{s}})=\lim _{{{\overline{t}}\to 0}}\frac{1}{2\pi i}\int _{{\eta-i\infty}}^{{\eta+i\infty}}e^{{{\overline{s}}{\overline{u}}}}\left(\frac{e^{{-{\overline{t}}{\overline{u}}^{\alpha}}}-1}{{\overline{t}}}\right){\overline{f}}({\overline{u}})\;\mbox{\rm d}{\overline{u}}=\frac{1}{2\pi i}\int _{{\eta-i\infty}}^{{\eta+i\infty}}e^{{{\overline{s}}{\overline{u}}}}\lim _{{{\overline{t}}\to 0}}\left(\frac{e^{{-{\overline{t}}{\overline{u}}^{\alpha}}}-1}{{\overline{t}}}\right){\overline{f}}({\overline{u}})\;\mbox{\rm d}{\overline{u}}=-\frac{1}{2\pi i}\int _{{\eta-i\infty}}^{{\eta+i\infty}}e^{{{\overline{s}}{\overline{u}}}}{\overline{u}}^{\alpha}{\overline{f}}({\overline{u}})\;\mbox{\rm d}{\overline{u}}. (96)

[110.2.2] The result can indeed be made rigorous and one has

Theorem 2.7

[110.2.3] The infinitesimal generator \mbox{\rm A}_{\alpha} of the macroscopic time evolutions {\overline{\mbox{\rm T}}}_{\alpha}({\overline{t}}) is related to the infinitesimal generator \mbox{\rm A}=-\mbox{\rm d}/\mbox{\rm d}{\overline{t}} of \overline{\mathcal{T}}_{{\overline{t}}} through

\mbox{\rm A}_{\alpha}{\overline{f}}({\overline{s}})=-(-\mbox{\rm A})^{\alpha}{\overline{f}}({\overline{s}})=-\mbox{\rm D}^{{\alpha}}{\overline{f}}({\overline{s}})=-\frac{1}{\Gamma(-\alpha)}\int _{0}^{\infty}\frac{{\overline{f}}({\overline{s}}-y)-{\overline{f}}({\overline{s}})}{y^{{\alpha+1}}}\;\mbox{\rm d}y=-\frac{1}{\Gamma(-\alpha)}\int _{0}^{\infty}y^{{-\alpha-1}}(\overline{\mathcal{T}}_{y}-\boldsymbol{1}){\overline{f}}({\overline{s}})\;\mbox{\rm d}y. (97)
Proof.

See Chapter III. ∎

[page 111, §1]   
[111.1.1] The theorem shows that fractional derivatives of Marchaud type arise as the infinitesimal generators of coarse grained time evolutions in physics. [111.1.2] The order \alpha of the derivative lies between zero and unity, and it is determined by the decay of the averaging kernel. [111.1.3] The order \alpha gives a quantitative measure for the decay of the averaging kernel. [111.1.4] The case \alpha\neq 1 indicates that memory effects and history dependence may become important.