1. Introduction
Classical fractional differentiation/integration operators
,
, acting on functions
, where
is `discrete derivative’ with respect to `time’
, are defined through the binomial expansion
, viz.,
with coefficients
and
Here
denotes the gamma function:
,
, and
,
. The asymptotics
(which follows by application of Stirling’s formula to (
2)) determines the class of functions
g and summability properties of (
1).
Fractional operators in (
1) play important role in the theory of discrete-time stochastic processes, in particular, time series. See, e.g., the monographs [
5,
10,
14,
25,
31] and the references therein. The autoregressive fractionally integrated moving-average ARFIMA
process
is defined as a stationary solution of the stochastic difference equation
with white noise (a sequence of standardized uncorrelated random variables (r.v.s))
. For
the solution of (
4) is obtained by applying the inverse operator, viz.,
Since (
3) implies
(
), (
5) is a well-defined stationary process with zero mean and finite variance. ARFIMA
process is the basic parametric model in statistical inference for long memory processes (also referred to as processes with long-range dependence). It has an explicit covariance function and the spectral density
which explodes or vanishes at the origin
as
depending on the sign of
d.
In this paper, we extend fractional operators in (
1) to functions
g on
-dimensional lattice
and/or more general
T of the form
where
is a random walk on
starting at
, with (1-step) probabilities
. We assume that
, i.e. the random walk is non-degenerate at
. Clearly,
, where
are the
j-step probabilities,
. Similarly to (
4), we define fractional operators
acting on
by
with coefficients
expressed through the binomial coefficients
and random walk probabilities
.
Let us describe the content and results of this paper in more detail. The main result of
Section 2 is Theorem 1 which provides the sufficient condition
for invertibility
and square summability of fractional coefficients in (
9), in terms of the characteristic function
(the Fourier transform) of the random walk.
Section 2 also includes a discussion of the asymptotics of (
9) as
which is important in limit theorems and other applications of of fractional integrated random fields. Using classical local limit theorems, Propositions 1 and 2 obtain `isotropic’ asymptotics of (
9) for a large class of random walk
showing that
decay as
, hence,
. The last fact is interpreted as
long-range dependence [
14,
25,
30] of the fractionally integrated random field
defined as a stationary solution of the difference equation
with white noise on the r.h.s. and studied in studied in
Section 3. Corollary 1 obtains conditions for the existence of stationary solution of (
11) given by the inverse operator
which are detailed in Examples 1 and 2 for fractional Laplacian and fractional heat operators.
Section 2 and 3 also include a discussion of
tempered fractional operators
and
tempered fractional random fields solving analogous equation
which generalize the class of tempered ARFIMA processes [
29] and have
short-range dependence and a summable covariance function.
Section 4 is devoted to scaling limits of moving average random fields on
with coefficients satisfying Assumption (A)
which includes `isotropic’ fractional coefficients
as a special case. The scaling limits refer to integrals
of random field
for each
from a class of (test) functions, as scaling parameter
. The scaling limits are identified in Corollary 3 as self-similar Gaussian random fields with Hurst parameter
. We note that limit theorems for random fields with long-range dependence or negative dependence were studied in many works [
7,
8,
9,
18,
22,
23,
27,
28,
32,
33], including statistical applications [
2,
10,
11,
17].
We expect that this study can be extended in several directions, including anisotropic scaling, infinite variance random fields, and fractional operators in
. See [
1,
6,
16,
20,
21,
24] for discussion and properties of fractional random fields with continuous argument
.
Notation. In what follows, C denote generic positive constants which may be different at different locations. We write and for the weak convergence and equality of probability distributions. Denote by the absolute-value norm on , where is either or , and the Euclidean norm on . is the scalar product in . Denote by the vector in with 1 in the jth coordinate and 0’s elsewhere. For , denote by the space of functions , for which , and by the space of measurable functions , for which the p-th power of the absolute value is integrable with respect to the Lebesgue measure: , with identification of functions such that almost everywhere (a.e.). Denote by the space of measurable and functions , for which , with identification of functions such that a.e. Write for the indicator function. Write for the smallest integer greater than or equal to . .
2. Invertibility and Properties of Fractional Operators
We start with properties of binomial coefficients in (
2)
The identity
leads to
and the invertibility relation:
The following lemma gives some basic properties of fractional coefficients
in (
9).
Lemma 1.
(i) Let . Then the series in (9) converges for every and
(ii) Let . Then for every and and
Moreover, implies and
(iii) Let and . Then
Proof. (i) From (
9) and (
12) we get
since
is not possible. On the other hand, for
we have
and
in view of (
12).
(ii) Since
is obvious from (
12), it suffices to show (
15) since it implies
by (
14). We have
where exchanging the order of summation is legitimate as all summands are nonnegative. Hence, using
and (
13), we get
proving part (ii).
(iii) The convergence of the series in (
16) and the equality follow as in (
15):
Lemma 1 is proved. □
Remark 1. Let . Then the inequalities are strict: and , if for some j, i.e. is accessible from state . Moreover, if state is transient, i.e. the probability of eventual return to is strictly less that 1, which is equivalent to , then .
The main result of this section is Theorem 1, which provides necessary and sufficient conditions for square summability of fractional coefficients in (
9), in terms of the characteristic function
, see (
10). Write
for the Fourier transform of a function
. For
introduce the
tempered fractional operators
with coefficients
and the Fourier transform
Theorem 1.
For , the following conditions are equivalent:
Either of these conditions implies
Moreover, for , the above conditions (21), (22) and (23) hold with d in place of .
Proof. Let
. Firstly, we consider
in (
9). They satisfy
because of (
3) and
with
. Then
is immediate. Moreover, we have the Fourier transform
, where
satisfies
. We see that
belongs to
.
Now let us prove the implication (
21) ⇒ (
22). We use approximation by tempered fractional coefficients
in (
20) as
. We have that
a.e. as
. Next, for
,
,
, the inequality
, where
, becomes
. Using it we get the domination for all
,
,
by a function in
according to (
21). Hence, by the dominated convergence theorem (DCT),
as
holds in
. As a consequence,
,
, is a Cauchy sequence in
. By Parseval’s theorem, the inverse Fourier transforms
is a Cauchy sequence in
and so
converges in
to some
as
. This
f must be
because
as
for all
. We conclude that
, or (
22).
Let us turn to the implication (
22) ⇒ (
21). From (
22) and
for all
it follows that
as
holds in
. By Parseval’s theorem,
,
, is a Cauchy sequence in
. It follows that
for some
. We also have that
for each
such that
. Since
, see Lemma 2.3.2(a) in [
19], we conclude that
a.e., proving (
21).
The above argument also proves (
23). On one hand,
is the limit of
in
as
because
converges in
to
as
. On the other hand,
in
as
. We conclude that
a.e. Theorem 1 is proved. □
Next, we turn to asymptotics of `fractional coefficients’
in (
9). The proof uses the local limit theorem in [
19] for random walk probabilities
. Following the latter work we assume that
Conditions in (
25) imply that the random walk has zero mean:
and invertible covariance matrix
According to the classical (integral) CLT, the normalized sum
approaches a Gaussian distribution on
with density
Lemma 2.
[19] Under conditions (25) there exists such that
For `very atypical’ values
we use the following bound [
19]: for any
there exists
such that
Proposition 1.
Let satisfy (25). The coefficients in (9) are well-defined for any and satisfy
Proof. Let us prove (
31). Since
is positive-definite,
,
, is a norm. Note that it is equivalent to the Euclidean norm. Using (
9) for a large
decompose
, where
To show the first relation in (
32) use (
29). We have
, where, for each
fixed, the main term
and the remainder term
asymptotically behave when
as
and, for some constants
,
Hence, the first relation in (
32) follows using
. In view of (
3), the same argument also proves the second relation in (
32) for
.
Consider (
32) for
. Split
into two sums over
, where
and
respectively. In the sum
we also have that
and Lemma 2 entails the bound
for some constants
. Hence,
since the last integral converges for any
d. Finally, by (
30), given large enough
, there exists
such that
, which implies
. This proves (
32) and completes the proof of Proposition 1. □
Lemma 2 does not apply to the simple random walk (which is not aperiodic) in which case the local CLT takes a somewhat different form, see [
19]. The application of the latter result and the argument in the proof of Proposition 1 yields the following result.
Proposition 2.
Let , . The coefficients in (9) are well-defined for any and satisfy
Proposition 1 and as well as Lemma 2 do not apply to random walks with non-zero mean as in Example 2 below (fractional heat operator) in which case fractional coefficients exhibit an anisotropic behavior different from (
31). Such behavior is described in the following proposition. We assume that the underlying random walk factorizes into a deterministic drift by 1 in direction
and a random walk on
as in Lemma 2:
where
and
is a probability distribution concentrated on
such that
. Write
for the random walk starting at
with
j-step probabilities
,
, such that
,
. In order to apply Lemma 2, we make a similar assumption to (
25):
and denote
the respective covariance matrix. Let
be a positive function on
satisfying the homogeneity property:
. As in Example 2 fractional coefficients for
in (
36) write as
Proposition 3.
Let satisfy (37) and . Then
as and We also have that
where is a continuous function on given by
for , and equal 0 for .
Proof. Consider the following
j-step probabilities of a random walk on
starting at
:
, where
for
,
. Let us estimate them by
, where
is the covariance matrix of the 1-step distribution
,
. Note
. By Lemma 2,
Relation (
40) follows directly from (
39), (
43) and (
3). Relation (
41) writes as
The asymptotics in (
44) is immediate from (
40) for
tending to
∞ as in (
40). The general case of (
44) also follows from (
40) using the continuity of
. For
the details can be found in [
22]. □
Remark 2. The approximation in (
40) compares with the kernel
of the fractional heat operator
for all
and some
. For
,
24] has recently derived the analytic form in (
45) of the kernel from the absolute square of its Fourier transform:
which is implicit definition of this kernel in [
16]. Similarly to derivations in [
24], for
, table of integrals [
15] [3.944.5-6] gives
yielding (
46).
Finally, the tempered fractional coefficients in (
20) are summable:
for any
and
any random walk
. Assuming the existence of exponential moment
for some
, (
20) decay exponentially
for some
. Indeed, Markov’s inequality gives
for any
and large enough
. Moreover,
, proving (
47).
3. Fractionally Integrated Random Fields on
Let
be a white noise, in other words, a sequence of r.v.s with
. Given a sequence
with the above noise we can associate a moving-average random field (RF)
with zero mean and covariance
, which depends on
alone and characterizes the dependence between values of
X at distinct points
.
A moving-average RF
X in (
48) will be said to be
long-range dependent (LRD) if ;
short-range dependent (SRD) if ;
negatively dependent (ND) if .
The above classification is important in limit theorems and applications of random fields. It is not unanimous; several related but not equivalent classifications of dependence for stochastic processes can be found in [
14,
18,
25,
30] and other works.
Many RF models with discrete argument are defined through linear difference equations involving white noise [
13]. In this paper, we deal with fractionally integrated RFs
X solving fractional equations on
:
whose solutions are obtained by inverting these operator; see below.
Definition 1. Let and in (9) be well-defined. By stationary solution of Equation (49) (respectively, (50)) we mean a stationary RF X such that for each the series in (49) converges in mean square and (49) holds (respectively, the series in (50) converges in mean square and (50) holds).
Corollary 1.
(i) Let . Then
is a stationary solution of Equation (49) if condition (21) holds (for , (21) is also necessary for the existence of the above X).
(ii) Let and (21) hold. Then X in (51) is LRD. Moreover, it has nonnegative covariance function and .
(iii) Let and (21) hold. Then X in (51) is ND; moreover .
(iv) Let . Then
is a stationary solution of equation (50). Moreover, X in (52) is SRD and .
Proof. (i) Let
.
X in (
51) is well-defined if and only if (
22) holds, which is therefore a necessary condition. Let us show that
X in (
51) is a stationary solution of (
49). We use the spectral representation of white noise
where
is a random complex-valued spectral measure with zero mean and variance
. Then
writes as
see (
23). Then
follows by (
24) and absolute summability
, see (
14), (
18).
Next, let
. Then
X in (
51) is well-defined and writes as (
54) due to
. We need to show that the series in (
49) converges in mean square towards
if and only if (
21) or (
22) hold. The latter convergence writes as
From (
54),
in view of (
22). This proves part (i).
(ii) From (
12), (
9) we see that
are nonnegative and
. Thus,
and
.
(iii) As in the proof of (i) we get that
, see (
12), and
, implying
.
(iv) Using the proof is similar as above. Corollary 1 is proved. □
ARFIMA(0,
d,0) Equation (
4) is autoregressive since the best linear predictor (or conditional expectation in the Gaussian case) of
given the `past’
is a linear combination
of the `past’ observations, due to the fact that
. For spatial equations as in (
49) or (), an analogous property given the `past’
does not hold since
as a rule. This issue is important in spatial statistics and has been discussed in the literature, see [
3,
4] and the references therein, distinguishing between `simultaneous’ and `conditional autoregressive schemes’. The recent work [
12] discusses some conditional autoregressive models with LRD property.
Definition 2. Let X be an RF with for each . We say that X has:
(i) a simultaneous autoregressive representation with coefficients if for each
where the series converges in mean square and the r.v.s satisfy .
(ii) a conditional autoregressive representation with coefficients if for each
where the series converges in mean square and the r.v.s satisfy .
Corollary 2. (i) Let and X be a fractionally integrated RF in (51) and (21) holds. Then X has a simultaneous autoregressive representation with coefficients and ;
(ii) Let , X be a fractionally integrated RF in (51) and (21) holds. Then X has a conditional autoregressive representation with coefficients and where
(iii) Let and X be a (tempered) fractionally integrated RF in (52). Then X has a simultaneous autoregressive representation with and a conditional autoregressive representation with with the same as in part (ii) and
Proof. (i) is obvious from Corollary 1 and (
49),
.
(ii) By (
21),
and
are well-defined,
and
. The orthogonality relation
follows from spectral representations in (
54), (
53):
It remains to show (
56), including the convergence of the series. In view of the definition of
this amounts to showing
or, in spectral terms, to the convergence of the Fourier series
in
. Note
, where the RF
,
, results from application of the inverse operator. Since
has negative dependence, see (
54) and the proof of Corollary 1 (iii), the covariances
are absolutely summable. Therefore, the Fourier series on the l.h.s. of (
59) converges uniformly in
to
, proving (
59).
(iii) The proof is analogous (and simpler) as (i)-(ii), using . □
Example 1. Fractional Laplacian. The (lattice) Laplace operator on
is defined as
so that
, where
is the transition operator of the simple random walk
on
with equal one-step transition probabilities
to the nearest-neighbors
. For
, the fractional Laplace RF can be defined as a stationary solution of the difference equation
with weak white noise on the r.h.s., written as a moving-average RF:
We find that
,
and
for some
and
. Hence, condition (
21) for (
61) translates to
In particular, a stationary solution of the equation (
61) on
exists for all
. Finally, recall that (
21) is equivalent to the condition (
22). We could have verified the latter by using Corollary 2, which gives the asymptotics of coefficients
in (
62).
Example 2. Fractional heat operator. For a parameter
, we can extend the definition of the (lattice) heat operator on
from
in [
22] to
as follows:
Thus,
corresponds to the random walk on
with 1-step distribution
. We find that
We also find that outside the origin
for some
since
. Therefore,
and
if
. The above result agrees with [
22] for
,
, and extends it to arbitrary
,
.
Example 3. Fractionally integrated time series models (case
). As noted above, ARFIMA
process is a particular case of (
51) corresponding to backward shift
or deterministic random walk
. Another fractionally integrated time series model is given in Example 1 and corresponds to the symmetric nearest-neighbor random walk on
with probabilities 1/2. It is of interest to compare these two processes and their properties. Let
be the corresponding operators,
For
and
, processes
and
are well-defined; moreover, they are stationary solutions of respective equations
and
. The spectral densities of
and
are given by
We see that when processes and have the same 2nd order properties up to a multiplicative constant so that in the Gaussian case is a noncausal representation of ARFIMA.
4. Scaling Limits
As explained in the Introduction, isotropic scaling limits refer to the limits distribution of integrals
where
is a given stationary random field (RF), for each
from a class of (test) functions
. We choose the latter class to be
In as follows,
X is a linear or moving-average RF on
:
where
are independent identically distributed (i.i.d.) r.v.s with
and
are deterministic coefficients. Obviously, stationary solution (
51) of Equation (
49) satisfying Corollary 1 is a particular case of linear RF with
. Our limits results assume an `isotropic’ behavior of
as
detailed in as follows. Let
denote the class of all continuous functions on
.
Assumption (A) Let be a sequence of real numbers satisfying the following properties.
(i) Let
. Then
where
is not identically zero.
(ii) Let
. Then
satisfy (
68) with the same
and, moreover,
.
(iii) Let . Then and .
The class of RFs in (
67) with coefficients satisfying Assumption (A)
is related but not limited to fractionally integrated RFs in (
49)- (
50). Note that the parameter
d is no longer restricted to be in
. By easy observation, Assumption (A)
implies LRD, ND, and SRD properties of
Section 3 in respective cases
, and
. Following the terminology in time series [
14], the parameter
d in (
68) may be called the
memory parameter of the linear RF
X in (
67), except that for
the memory parameter usually is defined as
.
In particular, the covariance function
of linear RF
X in (
67) writes as
or the lattice convolution of
with itself. We will use the notation
for lattice convolution and
for continuous convolution, viz.,
which is well defined for any
(respectively, for any
,
).
Proposition 4.
Let satisfy Assumption (A) with and some Then
where the (angular) function is given by
Proof. The existence and continuity of
follow from finiteness of integrals
and
. For (
70) it suffices to show that
Let
and
, see (
68). Then
. Clearly, (
72) follows from
and
To prove (
73) rewrite
as integral and change the variable
in it. This leads to
where
where
Relation (
73) follows once we prove the uniform convergence
. Since
is a compact set and
is continuous, the last relation is implied by the sequentional convergence
for any
and any
convergent to
:
. The proof of (
76) uses the bound
which follows from boundedness of
and
with
hence
. Note
for any
and
according to (
77). Since
does not depend on
and
, Pratt’s lemma [
26] applies to the integral in (
75) resulting in (
76) and (
73). The proof of (
74) is similar and simpler and is omitted. □
The question about the asymptotics of the variance of (
65) arises assuming the power-law asymptotics of the covariance admitting a power-law behavior at large lags which is tackled in the following proposition.
Proof.
(i) For any as
(ii) Let satisfy
where and . Then for any
(iii) Let . Then for any
Proof. (i) Write
for the l.h.s. of (
78). Let first
. Then
as
. Next, let
then
, where
follows by
+
. Finally, for
we have
where
follows similarly.
(ii) The convergence of the integral in (
81) follows from that of
in part (i), with
Let
denote the integral on the l.h.s. of (
80). By change of variables,
where
for any
. Using Pratt’s lemma [
26], it suffices to prove (
80) for
. In the latter case and with
we see that
as in the proof of Proposition 4. Thus, (
80) follows from the DCT.
(iii) Let
be the same as in the proof of (ii). For a large
, write
where
and
. Here,
can be made arbitrary small uniformly in
by choosing
K large enough. Next,
By boundedness of
we see that the integral
a.e. in
, and is bounded in
. Then, since
we conclude
by the DCT. Finally,
, and we can replace the last integral by the r.h.s. of (
82) uniformly in
provided
K is large enough. □
Proposition 5 does not apply to ND covariances satisfying (
79) with negative
. This case is more delicate since it requires additional regularity conditions of test functions and the occurrence of `edge effects’. A detailed analysis of this issue in dimension
and for indicator (test) functions of rectangles in
can be found in [
33]. Below we present a result in this direction and sufficient conditions on
when the limits take a similar form to (
80). Introduce a subclass of test functions:
Proposition 6.
Let satisfy Assumption (A) with . Then for any we have that
Proof. The convergence of the integral on the r.h.s. of (
85) follows from (
83) and Minkowski’s integral inequality:
.
The proof of the convergence in (
84) resembles that of (
80). Write
for the integral on the l.h.s. of (
84). Using
we rewrite
,
, and
where the inner integrals tend to those on the r.h.s. of (
85) at each
such that
,
. The remaining details are similar as in (
80) and omitted. □
Remark 3. The restriction
in Proposition 6 is not necessary for (
85). Indeed, if
satisfies the uniform Lipschitz condition
then the integral in (
83) converges for
implying
. On the other hand, for indicator functions
of a bounded Borel set
with `regular’ boundary, we typically have
leading to
.
Relation (
68) entails the existence of the scaling limit
which is a continuous homogeneous function on
: for any
we have that
With the limit function in (
86) we associate a Gaussian RF:
where
is a real-valued Gaussian white noise with zero mean and variance
,
is the usual and
the `regularized’ convolution. For indicator test function
of a Borel set
(belonging to
) we see that the latter convolution equals
The existence of stochastic integrals in (
88) of follows from Propositions 5 and 6. Particularly, the variances
and
agree with (
81) and (
85).
Let
be the Schwartz space of all infinitely differentiable rapidly decreasing functions
. Following [
9] we say that a generalized RF
is
stationary if
and
H-self-similar (
) if
. As noted in Remark 3,
, hence (
88) are well-defined for any
and represent stationary generalized RFs on
. By scaling property in (
87) and a change of variables we see that
, hence RF
in (
88) is
-self-similar, with
The RF in (
88) appear as scaling limits in the following corollary.
Corollary 3.
Let X be a linear RF satisfying Assumption (A) and be defined in (65). Then
where .
Proof. Since (
65) writes as a linear form
in i.i.d. r.v.s, we can use the Lindeberg type condition, see also [
14]. Accordingly, it suffices to show that
holds in each case
of the corollary. The behavior of the last variance is detailed in Propositions 5 and 6 and it grows to infinity in each case of
d. On the other hand, the l.h.s. of (
91) does not exceed
which is bounded in cases
and
. Finally, in case
we see that the l.h.s. of (
91) does not exceed
and (
91) holds since
. □