Eigenvalue distributions of high-dimensional matrix processes driven by fractional Brownian motion
Abstract
In this article, we study high-dimensional behavior of empirical spectral distributions for a class of symmetric/Hermitian random matrices, whose entries are generated from the solution of stochastic differential equation driven by fractional Brownian motion with Hurst parameter . For Wigner-type matrices, we obtain almost sure relative compactness of in following the approach in [Anderson2010]; for Wishart-type matrices, we obtain tightness of on by tightness criterions provided in Appendix B. The limit of as is also characterised.
keywords:
[class=AMS]keywords:
and and
1 Introduction
There has been increasing research activity on matrix-valued stochastic differential equations (SDEs) in recent years. A prominent example is the class of generalized Wishart processes introduced in [Graczyk2013]. This class generalizes classical examples of symmetric Brownian motion [Dyson62], the Wishart process [Bru91] and the symmetric matrix-valued process whose entries are independent Ornstein-Uhlenbeck processes [Chan1992]. Recent works on generalized Wishart processes include [Song20, Song19] and [Graczyk2014, Graczyk2019]. A common feature of these processes is that they are all driven by independent Brownian motions.
In contrast, the study of SDE matrices driven by fractional Brownian motions (fBm) has a shorter history and less literature. To our best knowledge, the first paper is [Nualart20144266], where the symmetric fractional Brownian matrix was studied. The SDE for the associated eigenvalues and conditions for non-collision of the eigenvalues were obtained when the Hurst parameter by using fractional calculus and Malliavin calculus. The convergence in distribution of the empirical spectral measure of a scaled symmetric fractional Brownian matrix was established in [Pardo2016] by using Malliavin calculus and a tightness argument. These results were generalized to centered Gaussian process in [Jaramillo2019]. Besides, [Pardo2017] obtained SDEs for the eigenvalues, conditions for non-collision of eigenvalues and the convergence in law of the empirical eigenvalue measure processes for a scaled fractional Wishart matrix (the product of an independent fractional Brownian matrix and its transpose). In the present paper, by adopting a different approach, we obtain stronger results on the convergence of empirical measures of eigenvalues for a significantly larger class of matrix-valued processes driven by fBms (see Remark 3.1).
More precisely, consider the following 1-dimensional SDE
(1.1) |
with initial value independent of . Here, is a fractional Brownian motion with Hurst parameter , and the differential is in the Stratonovich sense. It has been shown in [Lyons1994] that there is a unique solution to SDE (1.1) if the coefficient functions and have bounded derivatives which are Hölder continuous of order greater than .
Let be i.i.d. copies of and be a symmetric matrix with entries
(1.2) |
Let be the eigenvalues of and
(1.3) |
be the empirical distribution of the eigenvalues.
In this paper, we aim to study, as the dimension tends to infinity, the limiting behavior of the empirical measure-valued processes defined by (1.3) and the empirical measure-valued processes arising from other related matrix-valued processes in the space of continuous measure-valued processes, with being the space of probability measures equipped with its weak topology. A summary of the contents of the paper is as follows.
In Section 2, we study the solution of SDE (1.1). In [Hu2007], a pathwise upper bound for on the time interval was obtained under some boundedness and smoothness conditions on the coefficient functions. We adapt the techniques used in [Hu2007] to study the increments for and obtain the pathwise Hölder continuity for . Moreover, under some integrability conditions on the initial value , the Hölder norm of is shown to be -integrable. These results improve the previous results in [nr02] with sharper bounds for the increments of the process.
In Section 3, we prove the almost sure high-dimensional convergence in for the empirical spectral measure-valued process (1.3) of the Wigner-type matrix (1.2). We also obtain the PDE for the Stieltjes transform of the limiting measure process , which turns to be the complex Burgers’ equation up to a change of variable (see Remark 3.2 and Remark 3.3). This generalizes the results of [Pardo2016], where the special case of was studied (note that the convergence obtained therein is in law). A complex analogue is also studied at the end of this section.
In Section 4, we extend the model of Wigner-type matrix to the locally dependent symmetric matrix-valued stochastic process , where each is a weighted sum of i.i.d. for in a fixed and bounded neighborhood of (see (4.1) for the definition). We also establish the almost sure convergence of the empirical spectral measure-valued process in . Moreover, the limiting measure-valued process is characterized by an equation satisfied by its Stieltjes transform (see Theorem 4.1). It is worth noticing that for the proofs of the almost sure convergence for the empirical spectral measure-valued processes in both Section 3 and Section 4, we follow the strategy used in [Song20] (which was inspired by [Anderson2010]).
In Section 5, we study the high-dimensional limit of the empirical spectral measure-valued processes of the Wishart-type matrix-valued processes given by (5.1) and its complex analogue (5.9). In the special case where , we recover the convergence result in [Pardo2017]. We also obtain the PDE for the Stieltjes transform of the limit measure process (see Remarks 5.4 and 5.6).
For Wishart-type matrix-valued processes, we are not able to obtain the almost sure convergence in for the empirical spectral measure-valued processes as done in Section 3 and Section 4 since the method used there heavily relies on the independence of the upper triangular entries. Instead, we obtain the convergence in law by a tightness criterion for probability measures on (see Theorem B.2, Propositions B.2 and B.3 in Appendix B). The tightness criterion Theorem B.2 might be already well-known in the literature (we refer to, e.g., [Rogers1993, Pardo2016, Pardo2017, Jaramillo2019] for related arguments), however, we could not find a reference stating it explicitly, and we provide it in Appendix B.
2 Hölder continuity of the solution to the SDE (1.1)
Some notations are in order. The Hölder norm of a Hölder continuous function of order is
We also use
to denote the Frobenius norm (also known as the Hilbert-Schmidt norm or -Schatten norm) of a matrix .
2.1 Preliminaries on fractional calculus and fractional Brownian motion
In this subsection, we recall some basic results in fractional calculus. See [skm93] for more details. Let with and let The left-sided and right-sided fractional Riemann-Liouville integrals of of order are defined for almost all by
and
respectively, where and is the Euler gamma function.
Let (resp. ) be the image of by the operator (resp. ). If (resp. ) and , then the left-sided and right-sided fractional derivatives are defined as
(2.1) |
respectively, for almost all .
Let denote the space of -Hölder continuous functions of order on the interval . When , then we have . On the other hand, if , then for all .
The following inversion formulas hold:
Similar inversion formulas hold for the operators and as well.
We also have the following integration by parts formula.
Proposition 2.1.
If and , we have
(2.2) |
The following proposition indicates the relationship between Young’s integral and Lebesgue integral.
Proposition 2.2.
Suppose that and with . Let and . Then the Riemann-Stieltjes integral exists and it can be expressed as
(2.3) |
where .
It is known that almost all the paths of are -Hölder continuous for . By the Fernique Theorem, we have the following estimation for its Hölder norm.
Lemma 2.1.
There exists a positive constant depending on , such that
2.2 Hölder continuity
In this subsection, for the solution to (1.1), we provide some estimations for its Hölder norm, following the approach developed in [Hu2007, Theorem 2].
Theorem 2.1.
Suppose that the coefficient functions and are bounded and have bounded derivatives which are Hölder continuous of order greater than for some . Then there exists a constant that depends on only, such that for all and ,
(2.4) |
Consequently, there exists a random variable with for all such that for ,
Remark 2.1.
Note that the pathwise Hölder continuity of the process was established in [nr02, Theorem 2.1] under some Lipschitz conditions and growth conditions on the coefficient functions and . They also obtained the -integrability of the Hölder norm of the process for where is the growth rate of the function .
An upper bound for was obtained in [Hu2007, Theorem 2], and it turns out that the techniques introduced in [Hu2007] can be used to improve the estimations for the pathwise Hölder norm. This is done in Theorem 2.1 for bounded coefficient functions, and in Theorem 2.2 for unbounded coefficient functions (with linear growth). Compared with the result in [nr02], Theorems 2.1 and 2.2 provide sharper upper bounds for the increments of the processes, which allow to establish the -integrability of the Hölder norm for all . In contrast, [nr02, Theorem 2.1] obtained the -integrability for when the coefficient functions have linear growth.
Proof of Theorem 2.1.
Fix . By Proposition 2.2, for ,
(2.5) |
By (2.1) and Lemma 2.1, for , we have
(2.6) |
and
(2.7) |
where the Hölder norm is finite a.s. by [nr02, Theorem 2.1].
By (2.5), (2.2) and (2.2), we have
(2.8) |
where is the Beta function, and is a constant depending on only.
Hence, by (2.2), for ,
(2.9) |
and therefore,
(2.10) |
Choose such that
and if the denominator vanishes, we set .
For the case where the coefficient functions or are unbounded, we have the following estimation.
Theorem 2.2.
Suppose that the coefficient functions and have bounded derivatives which are Hölder continuous of order greater than for some . Moreover, if , then for all , for all ,
(2.13) |
where is a constant depending only on .
Proof.
Obviously the function is Hölder continuous of any order with the Hölder norm . Hence, by [Hu2007, Theorem 2 (i)], we have
(2.14) |
In this proof, and are generic positive constants depending only on and , respectively, and they may vary in different places.
The estimations (2.5) and (2.2) are still valid. Instead of (2.2), we have
(2.15) |
By (2.5), (2.2), (2.2) and (2.2), we have
(2.16) |
Similarly, it is easy to see that
(2.17) |
Fix a positive constant such that
3 High-dimensional limit for Wigner-type matrices
3.1 Relative compactness of empirical spectral measure-valued processes
Denote by the space of probability measures equipped with weak topology, then is a Polish space. For , let be the space of continuous -valued processes. In this subsection, under proper conditions, we obtain the almost sure relative compactness in for the set of empirical spectral measures of with entries given in (1.2) by using Lemma B.4([Anderson2010, Lemma 4.3.13]).
Theorem 3.1.
Assume one of the following hypotheses holds:
Assume that there exists a positive function with bounded derivative, such that and
for some positive constant almost surely. Then for any , the sequence is relatively compact in almost surely.
Proof.
Under hypothesis H1 (hypothesis H2, resp.), by Theorem 2.1 (Theorem 2.2, resp.), we have
where is a random variable with finite moment of any order (with finite -th order moment, resp.).
Recall that by the Arzela-Ascoli Theorem (or Lemma B.3), for any number , the set
where and are two sequences of positive real numbers going to zero as goes to infinity, is compact in .
Define
where is a positive number that will be determined later. Then by Lemma B.2, is compact in . Note that there exists a sequence of functions that it is dense in . Choose a positive integer , such that , and define
where is a sequence of positive numbers that are independent of and will be determined later. Denote
then is compact in , according to Lemma B.4 ([Anderson2010, Lemma 4.3.13]). Hence, it is enough to show
Note that are i.i.d. with mean zero and finite variance denoted by for . By (3.1) and the Markov inequality, we have
(3.2) |
If we choose , then (3.1) becomes
(3.3) |
Similarly, by (3.1) and the Markov inequality, we have
(3.4) |
Now, we choose , where is a positive real number such that . Then (3.1) becomes
(3.5) |
Hence, by the definition of , (3.1) and (3.5), we have
(3.6) |
Therefore, by the Borel-Cantelli Lemma and (3.1), we have
The proof is concluded. ∎
3.2 Limit of empirical spectral distributions
Recall that the celebrated semicircle distribution on has density function
(3.7) |
Theorem 3.2.
Proof.
First, for fixed , we prove the almost sure weak convergence of the empirical measure .
Note that by Corollary 2.1, and are continuous functions of on . Let be a symmetric matrix with entries
(3.8) |
Then is a Wigner matrix. Let be the eigenvalues of and be the empirical spectral measure. By Lemma A.2, converges weakly to the semicircle distribution almost surely for all . Hence the empirical measure of the eigenvalues of converges weakly to almost surely.
Note that by (3.8), , where is an matrix with unit entries. Then by [Tao2012, Exercise 2.4.4], we can conclude that the empirical distribution converges weakly to , almost surely, for all .
Now, we show that converges weakly to almost surely.
Let be an arbitrary convergent subsequence with the limit , then for , . Therefore, noting that Theorem 3.1 yields that the sequence is relatively compact almost surely in , any subsequence of has a convergent subsequence with the unique limit . This implies that the total sequence converges in with the limit , almost surely. ∎
Remark 3.1.
If , and , then the solution to the SDE (1.1) is the fractional Brownian motion , and Theorem 3.2 implies that the the empirical spectral measure converges weakly to the scaled semicircle distribution with density almost surely. This improves the results in [Pardo2016], where the convergence of the empirical spectral measure in law is obtained.
Remark 3.2.
The Stieltjes transform of the limiting measure is
where
is the Stieltjes transform of the semi-circle distribution. If we assume that the variance of the solution is continuously differentiable on , then we have
(3.9) |
By [Bai2010, Lemma 2.11], it is easy to get
and by (3.9),
(3.10) |
For the case with , equation (3.2) becomes
(3.11) |
Denoting , by change of variable and (3.11), one can deduce that satisfies the complex Burgers’ equation
This relationship was obtained in [Jaramillo2019].
3.3 Complex case
In this subsection, we consider the following 2-dimensional SDE for ,
(3.12) |
with initial value that is independent of . Here , are continuously differentiable functions and is a 2-dimensional fractional Brownian motion. By [Lyons1994], there exists a unique solution to SDE (3.12), if and have bounded derivatives which are Hölder continuous of order greater than .
Denote by the imaginary unit. Let be i.i.d. copies of and be a Hermitian matrix with entries
(3.13) |
Here, are i.i.d. copies of the real-valued process satisfying (1.1) and independent of the family . Let be the eigenvalues of and denote the empirical spectral measure by
Theorem 3.3.
Suppose that the coefficient functions , , and have bounded (partial) derivatives which are Hölder continuous of order greater than . Besides, assume that among the following conditions,
-
and are bounded and ;
-
, ;
-
and are bounded and ;
-
, ;
or holds and or holds. Furthermore, suppose that there exists a positive function with bounded derivative, such that and
for some positive constant almost surely.
Then for any , for , and the sequence converges to in almost surely. The limiting measure has density , where is the variance of the solution to SDE (3.12).
Proof.
The proof is similar to the real case, which is sketched below.
First of all, following the proof of Theorem 2.1 or Theorem 2.2, we can still establish the pathwise Hölder continuity for the solution to the SDE (3.12). More specifically, for the case that condition holds, we can obtain
and for the case of ,
for all , for all . Thus, by Lemma 2.1 and the moment assumption on , we have that for both cases, there exists a positive random variable with finite second moment, such that
(3.14) |
Similar to (3.1), we have
for any with bounded derivative. Then following the same approach used in the proof of Theorem 3.1, we can show that the sequence of empirical spectral measures is relatively compact in almost surely.
Following the proof of Corollary 2.1, it is easy to see that the mean and the variance are continuous functions of .
Next, we introduce a Hermitian matrix satisfying
Hence, is a Hermitian Wigner matrix for all . Finally, by [Tao2012, Exercise 2.4.3, Exercise 2.4.4], we can conclude that the almost-sure limit of the empirical distribution of the eigenvalues of coincides with that of for all . Therefore, by Lemma A.3, converges towards in with density . ∎
4 High-dimensional limit for symmetric matrices with dependent entries
Let be i.i.d. copies of , the solution of (1.1). Fix a finite index set and a family of constants . Let and be the cardinality of . Note that . Let be an real symmetric matrix with entries
(4.1) |
Let be the eigenvalues of , and
be the empirical spectral measure of .
Theorem 4.1.
Suppose that the conditions in Theorem 3.1 hold. Then for any , the sequence converges to in almost surely. The Stieltjes transform of the limit measure is given by, for ,
where is the solution to the equation
with
where for .
Proof.
where are i.i.d. copies of with . Thus, for ,
Define
Then all ’s for are distributed identically with finite second moment.
Analogous to (3.1), we have
Noting that the mutual independence among implies the independence between and if or , we have
Thus, following the proof of Theorem 3.1, we may get estimations analogous to (3.1) and (3.1) therein, and then obtain the almost sure relatively compactness of the empirical spectral measure .
Now, let be a symmetric matrix with entries
Let be the empirical spectral measure of . Then by Lemma A.6 ([Banna2015, Theorem 3]), for each , converges to a deterministic probability measure almost surely. Moreover, the Stieltjes transform of the limit measure is given by
where is the solution to the equation
with
where, for ,
Finally, by [Tao2012, Exercise 2.4.4], the empirical spectral measure of converges to the same limit almost surely. The proof is concluded. ∎
5 High-dimensional limit for Wishart-type matrices
5.1 Real case
Recall that are i.i.d. copies of which is the solution to (1.1). Let
be a matrix with entries . Here, is a positive integer that depends on . Let
(5.1) |
be a symmetric matrix with eigenvalues , and
be the empirical spectral measure of .
Theorem 5.1.
Suppose that one of the following conditions holds,
Assume that there exists a positive function with bounded derivative, such that and
for some positive constant almost surely. Furthermore, assume that there exists a positive constant , such that as .
Then for any , for , and the sequence converges in probability to in , where with given in (A.1).
Proof.
Noting that exists finitely for all , we have that has mean 0 and finite second moment . Then by Lemma A.4, for any , almost surely, the empirical distribution
(5.2) |
weakly as . Thus, it remains to obtain the tightness of in the space .
Recalled that , by Theorem 2.1 and Theorem 2.2, we have that , where are i.i.d. copies of with . Thus,
(5.3) |
Hence, by (5.1), for ,
(5.4) |
Here, is a positive random variable that has finite second moment which is given by
Let, for ,
and for ,
Then are two positive numbers depending only on .
Without loss of generality, we assume that . Recall that the entries of are independent. Using the Cauchy-Schwarz inequality twice, the mean value theorem, Lemma A.1, and (5.1), we can obtain
(5.5) |
for any with bounded derivative. Hence, by Proposition B.3 and (5.2), we can conclude that the sequence converges in law to . Finally, noting that the limit measure is deterministic, the convergence in law actually coincides with the convergence in probability.
The proof is concluded. ∎
Remark 5.1.
In contrast, the convergences of the empirical measure-valued processes obtained in Theorem 3.1 and other subsequent results in Section 3 are almost-sure convergence, which is stronger than the in-probability convergence obtained in Theorem 5.1.
In section 3, we construct a compact set in and show that the sequence is in that compact set almost surely. However, in the Wishart case, we are not able to get an estimation analogous to (3.1) which is the key ingredient to get the almost-sure convergence, due to the lack of the independence for the upper triangular entries. Instead, we obtain the tightness on for thanks to Proposition B.2, and then the convergence in law follows consequently.
Remark 5.2.
Let , and , then the solution to (1.1) is the fractional Brownian motion . Then we have the convergence in law of the empirical spectral measures towards the scaled Marchenko-Pastur law , which recovers the results obtained in [Pardo2017].
Remark 5.3.
Let . Then under the conditions in Theorem 5.1, the sequence of empirical measures of the eigenvalues of converges in probability to in . Indeed, by the Lidskii inequality in [Tao2012, Exercise 1.3.22 (ii)], we have
where is the number of the eigenvalues of that are smaller than . Noting that the rank of
is at most for all , the convergence in probability of towards implies that the empirical spectral measures of converges to the same limit in probability.
Remark 5.4.
The Stieltjes transform of the limiting measure is
where is the probability density of the Marchenko-Pastur distribution given in (A.1). Assuming that the variance of the solution is continuously differentiable on , we have
(5.6) |
5.2 Complex case
Recall that is the solution to (3.12). Let be a matrix with entries , where are i.i.d. copies of and is a positive integer depending on . Let
(5.9) |
be a symmetric matrix with eigenvalue empirical measure .
Theorem 5.2.
Suppose that the coefficient functions , have bounded derivatives which are Hölder continuous of order greater than . Besides, assume that one of the following conditions holds,
-
(a)
, .
-
(b)
and are bounded and .
Moreover, suppose that there exists a positive function with bounded derivative, such that and
for some positive constant almost surely. Furthermore, suppose that there exists a positive constant , such that as .
Then for any , , and the sequence converges in probability to in .
Proof.
From the proof of Theorem 3.3, we can obtain the finiteness of the mean and . Analogous to (5.2), by using Lemma A.5, we have the almost-sure convergence
(5.10) |
Note that the estimation (3.14) in the proof Theorem 3.3 is still valid. Similar to the estimation (5.1) and (5.1) in the proof of Theorem 5.1, we can obtain
Then following the argument at the end of the proof of Theorem 5.1, we can obtain the tightness of the sequence , which implies the convergence in distribution and hence the convergence in probability, with the deterministic limit given in (5.10). ∎
Remark 5.5.
Let . Then under the conditions in Theorem 5.2, the sequence of empirical spectral measures of converges in probability to in .
Appendix A Preliminaries on (random) matrices
The following is the Hoffman-Wielandt lemma, which can be found in [Anderson2010, Lemma 2.1.19], see also [Tao2012].
Lemma A.1 (Hoffman-Wielandt).
Let and be Hermitian matrices, with ordered eigenvalues and . Then
The next two lemmas are the famous Wigner semi-circle law for the real case and complex case respectively (see, e.g., [Tao2012]).
Lemma A.2.
Let be the top left minors of an infinite Wigner matrix , which is symmetric, the upper-triangular entries are i.i.d. real random variables with mean zero and unit variance, and the diagonal entries are i.i.d. real variables, independent of the upper-triangular entries, with bounded mean and variance. Then the empirical spectral distributions converge almost surely to the Wigner semicircular distribution
Lemma A.3.
Let be the top left minors of an infinite complex Wigner matrix , which is Hermitian, the upper-triangular entries are i.i.d. complex random variables with mean zero and unit variance, and the diagonal entries are i.i.d. real variables, independent of the upper-triangular entries, with bounded mean and variance. Then the conclusion of Lemma A.2 holds.
The next two lemmas concern the celebrated Marchenko-Pastur law which was introduced in [Bai2010].
Lemma A.4.
Let be the top left minors of an infinite random matrix, whose entries are i.i.d. real random variable with mean zero and variance . Here, is a positive integer such that as . Then the empirical distribution of the eigenvalues of the matrix
converges weakly to the Marchenko-Pastur distribution
(A.1) |
almost surely, where is the point mass at the origin.
Lemma A.5.
Let be the top left minors of an infinite random matrix, whose entries are i.i.d. complex random variable with mean zero and variance . Here, is a positive integer such that as . Then the empirical distribution of the matrix
converges almost surely to the Marchenko-Pastur distribution described in Lemma A.4.
The following result characterizes the limiting empirical spectral distribution of the symmetric random matrix with correlated entries, which is a direct corollary of [Banna2015, Theorem 3].
Lemma A.6.
Let be an array of i.i.d. real-valued random variables with finite second moment. Let be a finite subset of , be a family of constants and
Suppose that . Denote for all . Let be a symmetric matrix with entries for . Then the empirical spectral measure of converges to a nonrandom probability measure with Stieltjes transform , where is the solution to the equation
Appendix B Tightness criterions for probability measures on
In this section, we collect some lemmas used in the proofs, and then we provide two tightness criterions for probability measures on (Theorems B.1 and B.2). We also provide sufficient conditions for tightness which can be verified by computing moments (Propositions B.1, B.2 and B.3).
Note that there has been fruitful literature on tightness of probability measures on a Skorohod space , where is a completely regular topological space. We refer the interested reader to [Mitoma, jaku, EK, Dawson1993, Perkins, KX, Kouritzin] and the references therein. Theorem B.1 is a direct consequence of Jakubowski’s criterion [jaku] (see also e.g, [Dawson1993, Theorem 3.6.4] and [Sun2011] for the statement of the criterion), noting that is a closed subset of . Theorem B.2 might be also well-known in the literature of tightness criterion for probability measures, but we could not find a reference addressing this explicitly. For both Theorems B.1 and B.2, we include self-contained proofs for the reader’s convenience.
Recall that is the set of probability measures on endowed with its weak topology, and that is the space of continuous probability-measure-valued processes, both of which are Polish spaces. Denote by the set of continuous functions on vanishing at infinity, which is also a Polish space. Also the space of sub-probabilities on endowed with its vague topology is a Polish space (see, e.g., [Kallenberg, Theorem 4.2]), and so is
Let’s also recall some basic facts for probability measures on a Polish space (see, e.g., [Billingsley] for details). Denote by the set of probability measures on where is the Borel -field on the Polish space . Let be a family of probability measures on . The family is called tight if for every , there exists a compact set such that for all . The family is called relatively compact if every sequence of elements of contains a weakly convergent subsequence. The Prokhorov’s theorem guarantees the equivalence between tightness and relatively compactness. Also note that a sequence converges weakly to if and only if converges to in the Polish space .
The following lemma (see, e.g., [Durrett2019, Theorem 3.2.14]) provides a method to obtain tightness for a set of probability measures.
Lemma B.1.
Let be an index set. If there is a non-negative function so that as and
then the family of probability measures is tight.
Based on the above tightness criterion, one can construct compact subsets of :
Lemma B.2.
Proof.
By Lemma B.1 and Prokhorov’s theorem (see, e.g., [Billingsley, Theorems 5.1 and 5.2]) which claims that a subset of is tight if and only if the closure of is compact, it suffices to show that is a closed set in , which is easy to verify. ∎
By the Arzela-Ascoli Theorem, we have the following lemma to construct compact sets in .
Lemma B.3.
is compact in , where is a positive constant and and are two sequences of positive numbers going to zero as goes to infinity.
The following lemma ([Anderson2010, Lemma 4.3.13]) provides an approach to construct compact subsets in . It will be used in the proof of Theorem B.1.
Lemma B.4.
Let be a compact subset of , let be a sequence of bounded continuous functions that is dense in , and let be a family of compact subsets of . Then the set
is a compact subset of .
The following lemma, which constructs compacts subsets in , will play a critical role in the proof of Theorem B.2.
Lemma B.5.
Let be a countable dense subset of , and let be a family of compact subsets of . Then the set
is a compact subset of .
Proof.
The proof is similar to the proof of Lemma 4.3.13 in [Anderson2010], which is provided here for the reader’s convenience.
Noting that is a closed subset of which is a Polish space, it suffices to prove that is sequentially compact.
Take a sequence . Then the functions for . Let be a countable dense subset of . Note that for each , has a subsequence that converges vaguely, i.e., converges in the Polish space . Then by the diagonal procedure and the compactness of , we can find a subsequence such that converges in for all and converges in for all , as tends to infinity. Denoting for all and for all , then for and for . The vague convergence of the measures for implies that for all and . Noting that is dense in , is dense in , is continuous, one can extend the family of sub-probability measures uniquely to a sub-probability-measure-valued process such that . This shows that is sequentially compact in , and the proof is completed. ∎
Throughout the rest of the section, let be a fixed positive number, and let be a sequence of continuous probability-measure-valued stochastic processes. We assume the following conditions on , which will be used in Theorems B.1 and B.2.
-
(I)
For any , there exists a compact set in , such that for all ,
-
(I’)
For each , the family of -valued random elements is tight.
-
(II)
There exists a countable dense subset of , such that for each , is tight on .
We also list some conditions that imply the above conditions.
Let a be nonnegative function such that .
-
(A)
(B.1) -
(A’)
For each , the family of random variables is tight.
-
(A”)
For each ,
(B.2) for some ,
-
(B)
There exists a countable dense subset of , such that there exist positive constants , ,
(B.3) for all , where is a constant depending only on and .
Lemma B.6.
(A) (I); (A”) (A’); (B)(II).
Proof.
By (B.1), let . For any , choose . The set is a compact subset of by Lemma B.2. Then, by Markov inequality we have, for
Thus, (A) (I).
(A”)(A’) follows directly from Lemma B.1. Now we show (A’)(I’). Fix an arbitrary . For any , due to the tightness of , one can find a positive constant such that for all ,
This implies that for all ,
where is a compact subset of by Lemma B.2. Therefore, is tight, and hence (A’)(I’).
Finally, (B)(II) follows directly from the Kolomogorov tightness criterion (see, e.g., [Ikeda1981, Theorem 4.2 and Theorem 4.3]). ∎
The following is Jakubowski’ tightness criterion for probability measures on .
Theorem B.1.
Assume that conditions (I) and (II) are satisfied. Then the set induces a tight family of probability measures on .
Proof.
By condition (II), for any , there exist compact subsets of for , such that for each , for all ,
and hence, for all ,
(B.4) |
Remark B.1.
Following the proof of [jaku, Theorem 3.1], one can easily show that conditions (I) and (II) are also necessary conditions for tightness of probability measures on
The criterion in Theorem B.1 can be verified by computing moments:
Proposition B.1.
Assume that conditions (A) and (B) are satisfied. Then the set induces a tight family of probability measures on .
In general situations, it might not be easy to check condition (I) or (A). Below, we provide another tightness criterion which weakens condition (I).
Theorem B.2.
Assume conditions (I’) and (II) are satisfied. Then the set induces a tight family of probability measures on .
Proof.
By condition (II), we can choose the same compact subsets of for as in the proof of Theorem B.1, and hence (B.4) still holds. By Lemma B.5, the set
is compact in for any . By (B.4), we have, for all ,
This implies the tightness of on . Therefore, for any subsequence of , by Prokhorov’s theorem, there exists a subsequence which converge weakly to some which is a continuous sub-probability-measure-valued process. Thus, for each , the sequence (as -valued random elements) converges weakly to . This together with the tightness of (as -valued random elements) in condition (I’) implies that and hence . Therefore, Prokhorov’s theorem implies that is tight on . The proof is concluded. ∎
Remark B.2.
Noting that condition (I) implies condition (I’), then by Remark B.1, conditions (I’) and (II) are also necessary conditions for tightness of probability measures on .
Similarly, we can justify the criterion in Theorem B.2 by computing moments. The two Propositions below are direct consequences of Theorem B.2 and Lemma B.6.
Proposition B.2.
Assume that conditions (A’) and (B) are satisfied. Then the set induces a tight family of probability measures on .
Proposition B.3.
Assume that conditions (A”) and (B) are satisfied. Then the set induces a tight family of probability measures on .
Acknowledgment: We would like to thank Rongfeng Sun for reminding us of the Jakubowski’s tightness criterion. J. Yao is partially supported by HKSAR-RGC-Grant GRF-17307319.