Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Harnack inequality and Liouville-type theorems for Ornstein-Uhlenbeck and Kolmogorov operators

  • We prove, with a purely analytic technique, a one-side Liouville theorem for a class of Ornstein-Uhlenbeck operators L0 in RN, as a consequence of a Liouville theorem at "t=" for the corresponding Kolmogorov operators L0t in RN+1. In turn, this last result is proved as a corollary of a global Harnack inequality for non-negative solutions to (L0t)u=0 which seems to have an independent interest in its own right. We stress that our Liouville theorem for L0 cannot be obtained by a probabilistic approach based on recurrence if N>2. We provide a self-contained proof of a Liouville theorem involving recurrent Ornstein--Uhlenbeck stochastic processes in the Appendix.

    Citation: Alessia E. Kogoj, Ermanno Lanconelli, Enrico Priola. Harnack inequality and Liouville-type theorems for Ornstein-Uhlenbeck and Kolmogorov operators[J]. Mathematics in Engineering, 2020, 2(4): 680-697. doi: 10.3934/mine.2020031

    Related Papers:

    [1] Prashanta Garain, Kaj Nyström . On regularity and existence of weak solutions to nonlinear Kolmogorov-Fokker-Planck type equations with rough coefficients. Mathematics in Engineering, 2023, 5(2): 1-37. doi: 10.3934/mine.2023043
    [2] Serena Dipierro, Giovanni Giacomin, Enrico Valdinoci . The fractional Malmheden theorem. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023024
    [3] Petteri Harjulehto, Peter Hästö, Jonne Juusti . Bloch estimates in non-doubling generalized Orlicz spaces. Mathematics in Engineering, 2023, 5(3): 1-21. doi: 10.3934/mine.2023052
    [4] Ugo Gianazza, Sandro Salsa . On the Harnack inequality for non-divergence parabolic equations. Mathematics in Engineering, 2021, 3(3): 1-11. doi: 10.3934/mine.2021020
    [5] Daniela De Silva, Ovidiu Savin . On the boundary Harnack principle in Hölder domains. Mathematics in Engineering, 2022, 4(1): 1-12. doi: 10.3934/mine.2022004
    [6] Giacomo Ascione, Daniele Castorina, Giovanni Catino, Carlo Mantegazza . A matrix Harnack inequality for semilinear heat equations. Mathematics in Engineering, 2023, 5(1): 1-15. doi: 10.3934/mine.2023003
    [7] Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039
    [8] Marco Bramanti, Sergio Polidoro . Fundamental solutions for Kolmogorov-Fokker-Planck operators with time-depending measurable coefficients. Mathematics in Engineering, 2020, 2(4): 734-771. doi: 10.3934/mine.2020035
    [9] Tommaso Barbieri . On Kolmogorov Fokker Planck operators with linear drift and time dependent measurable coefficients. Mathematics in Engineering, 2024, 6(2): 238-260. doi: 10.3934/mine.2024011
    [10] Luz Roncal . Hardy type inequalities for the fractional relativistic operator. Mathematics in Engineering, 2022, 4(3): 1-16. doi: 10.3934/mine.2022018
  • We prove, with a purely analytic technique, a one-side Liouville theorem for a class of Ornstein-Uhlenbeck operators L0 in RN, as a consequence of a Liouville theorem at "t=" for the corresponding Kolmogorov operators L0t in RN+1. In turn, this last result is proved as a corollary of a global Harnack inequality for non-negative solutions to (L0t)u=0 which seems to have an independent interest in its own right. We stress that our Liouville theorem for L0 cannot be obtained by a probabilistic approach based on recurrence if N>2. We provide a self-contained proof of a Liouville theorem involving recurrent Ornstein--Uhlenbeck stochastic processes in the Appendix.


    Dedicato a Sandro Salsa con stima ed amicizia.

    The main "motivation" of this paper is to provide a purely analytical proof of a one-side Liouville theorem for the following Ornstein–Uhlenbeck operator in RN:

    L0:=Δ+Bx,, (1.1)

    where Δ is the Laplace operator, while   and denote, respectively, the inner product and the gradient in RN. Moreover B is a N×N real matrix which we suppose to satisfy the following condition: letting

    E(t):=exp(tB), (1.2)

    then,

    b:=suptRE(t)<. (H)

    It is not difficult to show that condition (1.3) is equivalent to the following one:

    Bisdiagonalizableoverthecomplexfieldwithalltheeigenvaluesontheimaginaryaxis.

    This condition is satisfied in particular if B=BT and if B2=IN, where IN is the N×N identity matrix.

    Our positive (one-side) Liouville Theorem for (1.1) is the following one.

    Theorem 1.1. Let v be a smooth* solution to

    L0v=0inRN.

    *L0 is hypoelliptic, so that every distributional solution to L0v=0 actually is of class C.

    If infRNv>, then v is constant.

    If we assume the solution v to be bounded both from below and from above then the conclusion of Theorem 1.1 immediately follows from a theorem due to Priola and Zabczyk [14,Theorem 3.1], which, for the operator L0 in (1.1), takes this form:

    Consider the Ornstein–Uhlenbeck operator

    L0=Δ+Bx,,

    where B is any N×N constant matrix. Then the following statements are equivalent:

    (i) L0 has the simple Liouville property, i.e.,

    L0v=0 in RN, supRN|v|<vconstant;

    (ii) the real part of every eigenvalue of the matrix B is non-positive.

    If the matrix B satisfies (1.3), its eigenvalues have real part equal to zero. Then, the aforementioned Priola and Zabczyk theorem implies that the bounded solutions to L0v=0 in RN are constant.

    Theorem 1.1 is a Corollary of the following Liouville theorem "at t= " for the evolution counterpart of L0, i.e., for the Kolmogorov operator in RN+1=RNx×Rt

    L:=Δ+Bx,t. (1.3)

    Theorem 1.2. Let u be a smooth solution to

    Lu=0inRN+1.

    If infRNu>, then

    limtu(x,t)=infRN+1uforeveryxRN.

    It easy to show that this theorem implies Theorem 1.1. Indeed, let v:RNR be a smooth and bounded below solution to L0v=0 in RN. Then, letting

    u(x,t)=v(x),xRN,tR,

    we have

    Lu=0 in RN+1 and infRN+1u=infRNv>.

    Then, by Theorem 1.2,

    infRNv=infRN+1u=limtu(x,t)=v(x)for every xRN.

    Hence, v is constant.

    From Theorem 1.2 it also follows a Liouville theorem for bounded solutions to Lu=0 (for a related result see Theorem 3.6 in [13]).

    Theorem 1.3. Let u be a bounded smooth solution to

    Lu=0inRN+1.

    Then, u is constant.

    Proof. Let

    m=infRN+1u and M=supRN+1u.

    Applying Theorem 1.2 to Mu and um, we obtain for every xRN that:

    0=infRN+1(Mu)=limt(Mu(x,t))

    and

    0=infRN+1(um)=limt(u(x,t)m).

    Hence, M=m and u is constant.

    Theorem 1.2 is, in turn, a consequence of a "global" Harnack inequality for non-negative solutions to Lu=0 in RN+1. To state this inequality we need to recall that L is left translation invariant on the Lie group K=(RN+1,) with composition law

    (x,t)(y,τ)=(y+E(τ)x,t+τ),

    see [12]. For every z0 in RN+1 we define the "paraboloid"

    P(z0)=z0P,

    where

    P={(x,t)RN+1 : t<|x|24}.

    Then, inspired by an idea used in [8] for classical parabolic operators, and exploiting Mean Value formulas for solutions to Lu=0, we establish the following Harnack inequality.

    Theorem 1.4. Let z0RN+1 and let u be a non-negative smooth solution to

    Lu=0inRN+1.

    Then, there exists a positive constant C, independent of u and z0, such that

    u(z)Cu(z0),

    for every zP(z0).

    We will prove this theorem in Section 5. Here we show how it implies Theorem 1.2 by using the following lemma (for the reader's convenience we postpone its proof to Section 3).

    Lemma 1.5. For every xRN and for every z0RN+1 there exists a real number T=T(x,z0) such that

    (x,t)P(z0)t<T.

    Proof of Theorem 1.2. Let u be a smooth bounded below solution to Lu=0 in RN+1. Define

    m=infRN+1u.

    Then, for every ε>0, there exists zεRN+1 such that

    u(zε)m<ε.

    Theorem 1.4 applies to function um, so that

    u(z)m<C(u(zε)m)<Cε, (1.4)

    for every zP(zε), where C>0 does not depend on z and on ε. Let us now fix xRN. By Lemma 1.5, there exists T=T(zε,x)R such that (x,t)P(zε) for every t<T. Then, from (1.4), we get

    0u(x,t)mCεt<T.

    This means

    limtu(x,t)=m.

    We conclude the introduction with the following remark.

    Remark 1.6. One-side Liouville theorems for a class of Ornstein–Uhlenbeck operators can be proved by a probabilistic approach based on recurrence of the corresponding Ornstein–Uhlenbeck process. We present this approach in Appendix, showing how it leads to one-side Liouville theorems also for degenerate Ornstein–Uhlenbeck operators. However, the results obtained with this probabilistic approach contain Theorem 1.1 only in the case N=2. We mention that, in this last case, Theorem 1.1 is contained in [3], where a full description of the Martin boundary for a non-degenerate two-dimensional Ornstein–Uhlenbeck operator is given.

    We also mention that under particular assumptions on the matrix B that make the operator L homogenous with respect to a group of dilations, asymptotic Liouville theorems at t= for the solutions to Lu=0 in RN+1 are known (see [10] and the references therein); as a consequence, in such cases, one-side Liouville theorems for the solutions to L0v=0 hold.

    The matrix

    E(τ)=exp(τB),τR,

    introduced in (1.2), plays a crucial rôle for the operator L. First of all, as already recalled in the Introduction, defining the composition law in RN+1 as follows:

    (x,t)(y,τ)=(y+E(τ)x,t+τ), (2.1)

    we obtain a Lie group

    K=(RN+1,),

    on which L is left translation invariant (see [12]; see also [1], Section 4.1.4).

    As already observed, assumption (1.3) implies

    σ(B):= {eigenvalues of B}iR.

    Then, since B has real entries, λσ(B) if λσ(B). As a consequence,

    trace(B)=0.

    A fundamental solution for L is given by

    Γ(z,ζ)=γ(ζ1z), (2.2)

    where,

    γ(z)=γ(x,t)={0 if t0,(4π)N2detC(t)exp(14C1(t)x,x) if t>0,

    and

    C(t)=t0E(s)E(s)T ds,

    (see [12,(1.7)], and keep in mind that trace(B)=0 since B satisfies (1.3)).

    It is noteworthy to stress that

    C(t) is symmetric and C(t)>0

    for every t>0.

    The solutions to Lu=0 in RN+1 satisfy the following Mean Value formula: for every z0RN+1, r>0 and pN,

    u(z0)=1rΩ(p)r(z0)u(z)W(p)r(z10z) dz, (2.3)

    where

    Ω(p)r(z0)={z : ϕp(z0,z)>1r},

    with

    ϕp(z0,z):=Γ(z0,z)(4π(t0t))p2,

    if z=(x,t) and z0=(x0,t0).

    Remark 2.1. If zΩ(p)r(z0), then Γ(z0,z)>0, hence t0t>0.

    Moreover,

    W(p)r(z)=ωpRpr(0,z){W(z)+p4(p+2)(Rr(0,z)t)2}, (2.4)

    where ωp denotes the Lebesgue measure of the unit ball of Rp,

    W(z)=W(x,t)=14|C1(t)x|2, (2.5)

    and

    Rr(0,z)=4(t)log(rϕp(0,z)). (2.6)

    A complete proof of the Mean Value formula (2.3) can be found in Section 5 of [2].

    Let z0=(x0,t0) and z=(x,t). Then,

    zP(z0)=z0Pz10zP(xE(tt0)x0,tt0)P.

    Hence, keeping in mind the definition of P,

    z=(x,t)P(z0)|xE(tt0)x0|24(t0t)<1. (3.1)

    On the other hand, from (H), we have

    |xE(tt0)x0|24(t0t)(|x|+b|x0|)24(t0t)0,as t.

    Therefore: for every fixed z0RN+1 and xR, there exists T=T(z0,x) s.t.

    z=(x,t)P(z0) t<T.

    The aim of this section is to prove a geometrical lemma on the level sets Ω(p)r (which we call L-"onions"), that will play a crucial rôle in the proof of the Harnack inequality in Theorem 1.4.

    First of all we resume that hypothesis (1.3) implies:

    1b2|x|2tC1(t)x,xb2|x|2, (4.1)

    for every tR and for every xRN.

    Indeed, from (H), we obtain

    b:=suptRE(t)T<.

    Since we are considering the operator norm, we have

    |E(s)Ty|b|y|=b|E(s)TE(s)Ty|b2|E(s)Ty|,

    so that

    1b|y||E(s)Ty|b|y|

    for every tR and every yRN.

    Then, since

    C(t)y,y=t0|E(s)Ty|2 ds,

    we get

    1b2|y|21tC(t)y,yb2|y|2

    for every yRN and tR{0}.

    If in these inequalities we choose

    y=(C(t))12x if t>0

    and

    y=(C(t))12x if t<0,

    we immediately obtain (4.1).

    Now, for every r>0, define

    Σr={z=(x,t) : t=r2N+p, |x|2<4t}.

    Then, the following lemma holds

    Lemma 4.1. For every pN, there exists a constant θ=θ(p)>1 such that,

    Ω(p)θr(0)Ω(p)r(z)zΣr,r>0.

    Proof. Let r>0 and zΣr. Then z=(x,t), with

    t=r2N+p and |x|2<4r2N+p.

    Let us now take ζ=(ξ,τ)Ω(p)r(z). This means

    ϕp(z,ζ)>1rC1(tτ)(xE(tτ)ξ),xE(tτ)ξ<logr(4π(tτ))N+p2. (4.2)

    Analogously,

    ζΩ(p)θr(0)C1(τ)E(τ)ξ,E(τ)ξ<logθr(4π(τ))N+p2.

    On the other hand, by (4.1) and (1.3),

    C1(τ)E(τ)ξ,E(τ)ξb4|ξ|2|τ|,

    so that, ζ=(ξ,τ)Ω(p)θr(0) if τ<0 and

    |ξ|2<1b4|τ|logθr(4π|τ|)N+p2. (4.3)

    Then, to prove our lemma, it is enough to show that inequality (4.2) implies (4.3). Now, from (4.2), using (1.3), (4.1) and the inclusion z=(x,t)Σr, we obtain (we assume b1 so that b2b4)

    |ξ|2b2|E(tτ)ξ|22b2(|E(tτ)ξx|2+|x|2)2b4((tτ)C1(tτ)(E(tτ)ξx),E(tτ)ξx+4|t|)<2b4((tτ)logr(4π(tτ))N+p2+4|t|).

    Therefore, we will obtain (4.3), and hence the lemma, if for a suitable θ>1 independent of z and ζ, the following inequality holds

    2b4((tτ)logr(4π(tτ))N+p2+4|t|)1b4|τ|logθr(4π|τ|)N+p2. (4.4)

    To simplify the notation we put

    r(4π)N+p2=ρN+p2ρ=r2N+p4π.

    Hence, since zΣr,

    |t|=4πρ,

    and inequality (4.4) can be written as follows:

    A0(tτ)logρtτ+A1ρA2|τ|logθρ|τ|, (4.5)

    and the Ai's are strictly positive constants independent of z and ζ.

    Since ζΩ(p)r(z), we have

    1r<ϕρ(z,ζ)(14π(tτ))N+p2,

    then,

    0<tτ<ρ.

    As a consequence, since

    4πρ=|t|<|τ||τt|+|t|<ρ+4πρ,

    we get

    14π+1ρ|τ|14π.

    Thus, the left hand side of (4.5) can be estimated from above as follows:

    A0(tτ)logρtτ+A1ρ=ρ(Aotτρlogρtτ+A1)ρ(A0S+A1),

    where

    S=sup{slog1s : 0<s<1}.

    Moreover, the right hand side of (4.5) can be estimated from below as follows:

    A2|τ|logθρ|τ|ρ4πA2logθ4π+1.

    Therefore, if we choose θ>0 such that

    A0S+A14πA2logθ4π+1

    inequality (4.5) is satisfied. This completes the proof.

    Since L is left translation invariant on the Lie group (K,), it is enough to prove Theorem 1.4 in the case z0=0RN+1. In particular, it is enough to prove the inequality

    u(z)Cu(z0), with z0=0, (5.1)

    for every non-negative smooth solution u to

    Lu=0 in RN+1,

    and for every z=(x,t)P={(x,t) : |x|2<4t}.

    The constant C in (5.1) has to be independent of u. To this end, taken a non-negative global solution u to Lu=0, we start with the Mean Value formula for u on the L-level set Ω(p)2θr(z0), with p>4 and with θ given by Lemma 4.1:

    u(z0)=12θrΩ(p)2θr(z0)u(ζ)W(p)2θr(z10ζ) dζ. (5.2)

    Let us arbitrarily fix z=(x,t)P. Then t<0 and |x|2<4|t|. In (5.2) we choose r>0 such that

    t=r2N+p.

    By Lemma 4.1 we have the inclusion

    Ω(p)2θr(z0)Ω(p)r(z),

    so that, since u0, from (5.2) we get

    u(z0)12θrΩ(p)r(z)u(ζ)W(p)2θr(z10ζ) dζ. (5.3)

    Let us now prove that, for a suitable positive constant C independent of u and of z, we have (z10=z0=0) :

    W(p)2θr(z10ζ)W(p)r(z1ζ)2θCζΩ(p)r(z). (5.4)

    It will follow, from (5.3),

    u(z0)1rCΩ(p)r(z)u(ζ)W(p)r(z1ζ) dζ=(again by the Mean Value formula (2.3))=1Cu(z),

    i.e., u(z)Cu(z0), which is (5.1).

    To prove (5.4) we first estimate from below W(p)2θr(z10ζ). From the very definition of this kernel, by keeping in mind that z0=0, and letting ζ=(ξ,τ), we obtain:

    W(p)2θr(z10ζ)pωp4(p+2)(R2θr(z0,ζ))p+2|τ|2=cp|τ|p+222(log(2θrϕp(z0,ζ)))p2+1 (ϕp(z0,ζ)1θr since ζΩ(p)r(ζ)Ω(p)θr(z0))cp(log(2θ))p2+1|τ|p21 (if p>2)cp|t|p21=cprp2p+N.

    Here, and in what follows, cp,cp,,cp denote strictly positive constants only depending on p. So, we have proved the following inequality

    W(p)2θr(z10ζ)cprp2p+NζΩ(p)r(z). (5.5)

    Now we estimate W(p)r(z1ζ) from above, estimating, separately

    K1(z,ζ)=Rpr(0,z1ζ)W(z1ζ) (5.6)

    and

    K2(z,ζ)=Rp+2r(z0,z1ζ)(tτ)2. (5.7)

    We have

    K1(z,ζ)=(4(tτ)log(rΓ(z,ζ)(4π(tτ))N+p2))p2W(z1ζ))2p((tτ)logr(tτ)N+p2)p2W(z1ζ). (5.8)

    Moreover, from (2.5) and (4.1), we obtain

    W(z1ζ)=14|C1(τt)(ξE(τt)x)|2b44|ξE(τt)x|2(τt)2. (5.9)

    To estimate the right hand side of this inequality we use the inclusion ζΩ(p)r(z) which implies:

    ϕp(z,ζ)>1r(1(4π(tτ)))N+p2exp(14C1(tτ)(xE(tτ)ξ),xE(tτ)ξ)>1rC1(tτ)(xE(tτ)ξ),xE(tτ)ξ<logr(4π(tτ))N+p2.

    This inequality, keeping in mind (4.1), implies

    |xE(tτ)ξ|2b2(tτ)logr(4π(tτ))N+p2.

    Then

    |ξE(τt)x|2E(τt)2|E(tτ)ξx|2b4(tτ)logr(4π(tτ))N+p2cpr2N+p4π,

    where

    cp=b4sup{slog1s:0<s<1}.

    Using this estimate in (5.9) and (5.8) we obtain:

    K1(z,ζ)cpr2N+p(tτ)p21(logr(4π(tτ))N+p2)p2cprp2N+p, (5.10)

    where, cp=cpSp, with

    Sp=sup{sp22(log1s)p2:0<s<1}.

    We stress that Sp< since p>4.

    The same estimate holds for K2. Indeed:

    K2(z,ζ)cp(tτ)p21(logr(4π(tτ))N+p2)p+22cpr2N+p(p21)=cprp2N+p, (5.11)

    where,

    cp=cpsup{sp21(log1s)p+22:0<s<1}<.

    Keeping in mind (5.6) and (5.7), and the very definition of W(p)r(z,ζ), from inequalities (5.10) and (5.11) we obtain

    W(p)r(z1ζ)cprp2p+NζΩ(p)r(z). (5.12)

    This inequality, together with (5.5), implies (5.4), and completes the proof of Theorem 1.4.

    We would like to warmly thank the anonymous referee whose criticism to the first version of the paper led us to strongly improve our results.

    The authors have been partially supported by the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).

    The authors declare no conflict of interest.

    Here we show a one-side Liouville theorem for some Ornstein–Uhlenbeck (OU) operators based on recurrence of the corresponding OU stochastic processes.

    It is a general fact from probabilistic potential theory (see in particular [9]) that recurrence of a Markov process is equivalent to the fact that all excessive functions are constants (we also mention that the equivalence between excessive functions and super harmonic functions has been established in a general setting; see [6] and the references therein). On the other hand, a characterization of recurrent OU processes is known (see [7] which extends the seminal paper [5]; see also [15] for connections between recurrence and stochastic controllability).

    We present the main steps to prove a one-side Liouville theorem in a self-contained way. Comparing with [5,7,9], we simplify some proofs; see in particular the proof of Theorem 6.6 in which we also use a result in [14]. We do not appeal to the general theory of Markov processes but we use some basic stochastic calculus. It seems to be an open problem to find a purely analytic approach to proving such result.

    Let Q be a non-negative symmetric N×N matrix and let B be a real N×N matrix. The OU operator we consider is

    K0=12 tr(QD2)+Bx,=12 div(Q)+Bx,. (6.1)

    We will always assume the well-known Kalman controllability condition:

    rank[Q,BQ,,BN1Q]=N, (6.2)

    see [4,7,11,12,14,15] and the references therein. Under this assumption K0 is hypoelliptic, see [12]. Before stating the Liouville theorem we recall that a matrix C is stable if all its eigenvalues have negative real part.

    Theorem 6.1. Assume (6.2). Let v:RNR be a non-negative C2-function such that K0v0 on RN. Then v is constant if the following condition holds:

    (HR) The real Jordan representation of B is

    (B000B1) (6.3)

    where B0 is stable and B1 is at most of dimension 2 and of the form B1=[0] or B1=(0αα0) for some αR (in this case we need N2).

    The proof of Theorem 6.1 will immediately follow by Lemma 6.4 and Theorem 6.6 below.

    Remark 6.2. Note that when N=2 the matrix B=(0100) does not satisfy (HR). On the other hand B=(0000) verifies (HR) with α=0. Moreover, an example of possibly degenerate two-dimensional OU operator for which the one-side Liouville theorem holds is

    K0=2xx+a2yy+xyyx,a0.

    Remark 6.3. It is well-known, that condition (6.2) is equivalent to the fact that

    Qt=t0exp(sB)Qexp(sBT)dsis positive definite for allt>0 (6.4)

    (cf. [4,7,12]). Note that C(t)=exp(tB)Qtexp(tBT) is used in [12] and in Section 5 of [2] with Q replaced by A.

    Let us introduce the OU stochastic process starting at xRN. It is the solution to the following linear SDE

    Xxt(ω)=x+t0BXxs(ω)ds+QWt(ω),t0,xRN,ωΩ, (6.5)

    see, for instance, [7,14]. Here W=(Wt) is a standard N-dimensional Wiener process defined a stochastic basis (Ω,F,(Ft),P) (the expectation with respect to P is denoted by E; as usual in the sequel we often do not indicate the dependence on ωΩ).

    For any non-empty open set ORN, we consider the hitting time τxO=inf{t0:XxtO} (if {} is empty we write τxO=).

    Now we recall the notion of recurrence. The OU process (Xxt)t0=Xx is recurrent if for any xRN, for any non-empty open set ORN, one has

    ϕO(x)=P(τxO<)=1. (6.6)

    Thus recurrence means that with probability one, the OU process reaches in finite time any open set starting from any initial position x.

    Lemma 6.4. Suppose that the OU process is recurrent. Let vC2(RN) be a non-negative function such that K0v0 on RN. Then v is constant.

    Proof. We will adapt an argument used in the proof of Lemma 3.2 of [9] to show that excessive functions are constant for recurrent Markov processes.

    Let us fix xRN. Applying the Itô formula and using the fact that K0v0 we get, P-a.s.,

    v(Xxt)=v(x)+t0K0v(Xxs)ds+Mtv(x)+Mt,t0,

    where we are considering the martigale M=(Mt), Mt=t0v(Xxs)QdWs.

    Let ORN be a non-empty open set and consider the hitting time τxO. We have 0v(XxtτxO)v(x)+MtτxO, t0. By the Doob optional stopping theorem we obtain

    E[v(XxtτxO)]v(x),t0.

    Hence

    v(x)E[v(XxnτxO)]E[v(XxnτxO)1{τxO<}],xRN,n1. (6.7)

    Recall that P(τxO<)=1, for any xRN. By the Fatou lemma (using also the continuity of the paths of the OU process) we infer

    E[v(XxτxO)]=E[lim infnv(XxnτxO)]v(x). (6.8)

    Now we argue by contradiction. Suppose that v is not constant. Then there exists 0<a<b, zRN such that v(z)<a and U={v>b} ={xRN:v(x)>b} which is a non-empty open set. By (6.8) with x=z we obtain

    a>v(z)E[v(XzτzU)]b

    because on the event {τzU<} we know that XzτzU{vb}. We have found the contradiction a>b. Thus v is constant.

    Recall the OU Markov semigroup (Pt)=(Pt)t0,

    Ptf(x)=(Ptf)(x)=E[f(Xxt)]=RNf(y)pt(x,y)dy,t>0, (6.9)

    where xRN, f:RNR Borel and bounded and pt(x,y)=e|Q1/2t(etBxy)|22(2π)Ndet(Qt). We set P0f=f. The associated potential of a non-negative Borel function g:RNR is

    Ug(x)=0Ptg(x)dt,xRN. (6.10)

    Clearly, in general it can also assume the value (cf. [9]).

    Remark 6.5. Let A be an empty open set and let 1A be the indicator function of A. The probabilistic interpretation of U1A is as follows. First one defines the sojourn time or occupation time of A (by the OU process starting at x) as

    JxA(ω)=01A(Xxt(ω))dt,ωΩ;

    it is the total amount of time that the sample path tXxt(ω) spends in A. Then E[JxA]=0E[1A(Xxt)]dt =U1A(x) is the average sojourn time or the expected occupation time of A.

    The next result is a reformulation of a theorem in [7] at page 822 (see also the comments before such theorem and [5]). Erickson proves some parts of the theorem and refers to [5] for the proof of the remaining parts.

    Theorem 6.6. Assume (6.2). The next conditions for the OU process are equivalent.

    (i) Condition (HR) holds.

    (ii) 11det(Qt)dt=.

    (iii) For any x,yRN,

    1pt(x,y)dt=. (6.11)

    (iv) The OU process (Xxt) is recurrent.

    We will only deal with the proofs of (i)(ii)(iii) and (i)(iv); the last implication is needed to prove the one-side Liouville theorem in Lemma 6.4.

    The proof of the recurrence (i)(iv) is different and simpler than the proof given in [5] which also [7] mentions (see the remark below for more details).

    Remark 6.7. In [5] it is proved that (iii)(iv) by showing first that (iii) implies that, for any non-empty open set O, one has U1O, and then using a quite involved Khasminskii argument (see pages 142–143 in [5]) which uses the strong Markov property, the irreducibility and strong Feller property of the OU process. Alternatively, the fact that U1O, for any non-empty open set O, is equivalent to recurrence can be obtained using a potential theoretical approach involving excessive functions as in [9] (see in particular the proof that (ii) implies (iv) in Proposition 2.4 and Lemma 3.1 in [9]).

    Proof. (i)(ii). This can be proved as in the proof of Lemma 6.1 in [5] by using the Jordan decomposition of the matrix B (see also the remarks in [7]).

    (ii)(iii) Note that QtQT (in the sense of positive symmetric matrices) if 0<tT. Hence by the Courant-Fischer min-max principle, we have λ(t)λ(T) (where λ(s) is the minimal eigenvalue of Qs). Hence, there exists M>0 such that, for t1,

    Q1t(etBxy),etBxy1λ(t)|etBxy|2Mλ(1)(|x|2+|y|2).

    Then pt(x,y)exp(M2λ(1)(|x|2+|y|2)) 1(2π)Ndet(Qt), t1, and (6.11) holds if (ii) is satisfied.

    (i)(iv) The proof of this assertion is inspired by [9] and uses also the Liouville-type theorem for bounded harmonic function proved in [14].

    Let us fix a non-empty open set ORN and consider the function ϕO=ϕ:RN[0,1] (cf. (6.6)), ϕ(x)=P(τxO<), xRN. We have to prove that ϕ is identically 1.

    Using the OU semigroup (Pt) we first check that

    Prϕ(x)ϕ(x),r0,xRN. (6.12)

    This is a known fact. We briefly recall the proof for the sake of completeness. Let us fix xRN and r>0 and note that ϕ is a Borel and bounded function. Since P(Xxt+rO,for somet0) P(XxtO,for somet0)=ϕ(x), we get (6.12) by the Markov property:

    P(Xxt+rO,for somet0)=E[E[1{Xxt+rO,for somet0}Fr]]=E[ϕ(Xxr)]=Prϕ(x).

    Now take any decreasing sequence (rn) of positive numbers converging to 0, i.e., rn0. We have {XxtO,for somet0}= n1{Xxt+rnO,for somet0} (increasing union) and so P(Xxt+rnO,for somet0) =Prnϕ(x)ϕ(x). Hence

    Psϕ(x)ϕ(x),as s0+xRN.  (6.13)

    Since ϕ0, properties (6.12) and (6.13) say that ϕ is an excessive function.

    Let us fix s>0 and introduce the non-negative function fs=(fPsϕ)s. We have

    0Ufs(x)=1ss0Ptϕ(x)dt<,xRN. (6.14)

    Indeed, for any T>s,

    01sT0Pt(ϕPsϕ)(x)dt=1sT0Ptϕ(x)dt1sT0Pt+sϕ(x)dt=1sT0Ptϕ(x)dt1sT+ssPtϕ(x)dt=1ss0Ptϕ(x)dt1sT+sTPtϕ(x)dt1ss0Ptϕ(x)dt

    (in the last passage we have used that ϕ0). Passing to the limit as T we get (6.14). Now by the Fubini theorem, for any s>0,

    >Ufs(x)=0dtRNfs(y)pt(x,y)dyRNfs(y)(1pt(x,y)dt)dy.

    Since we know (6.11) we deduce that fs=0, a.e. on RN. This means that, for any s0,

    ϕ(x)=Psϕ(x),for any xRN a.e.. (6.15)

    It follows that, for any t>0,

    Ptϕ(x)=Pt(Psϕ)(x)=Ps(Ptϕ)(x),s0, (6.16)

    holds, for any xRN (not only a.e.). Thus, for any t>0, Ptϕ is a bounded harmonic function for (Pt). By hypothesis (HR) and Theorem 3.1 in [14] we deduce that Ptϕct for some constant ct.

    Since ϕ is excessive we know that Ptϕ(x)ϕ(x) as t0+, xRN. It follows that ctc0 and ϕc0. Take zO. We have ϕ(z)=1. Hence ϕ is identically 1 and the proof is complete.



    [1] Bonfiglioli A, Lanconelli E, Uguzzoni F (2007) Stratified Lie Groups and Potential Theory for Their Sub-Laplacians, Berlin: Springer.
    [2] Cupini G, Lanconelli E (2020) On mean value formulas for solutions to second order linear PDEs. Ann Scuola Norm Sci, in press.
    [3] Cranston M, Orey S, Rösler U (1983) The Martin boundary of two-dimensional Ornstein-Uhlenbeck processes, In: Probability, Statistics and Analysis, Cambridge-New York: Cambridge University Press, 63-78.
    [4] Da Prato G, Zabczyk J (1996) Ergodicity for Infinite-Dimensional Systems, Cambridge: Cambridge University Press.
    [5] Dym H (1966) Stationary measures for the flow of a linear differential equation driven by white noise. T Am Math Soc 123: 130-164. doi: 10.1090/S0002-9947-1966-0198541-2
    [6] Dynkin EB (1965) Markov Processes Vols. I & II, Berlin-Göttingen-Heidelberg: Springer-Verlag.
    [7] Erickson RV (1971) Constant coefficient linear differential equations driven by white noise. Ann Math Statist 42: 820-823. doi: 10.1214/aoms/1177693440
    [8] Garofalo N, Lanconelli E (1989) Asymptotic behavior of fundamental solutions and potential theory of parabolic operators with variable coefficients. Math Ann 283: 211-239. doi: 10.1007/BF01446432
    [9] Getoor RK (1980) Transience and recurrence of Markov processes, In: Seminar on Probability, XIV (Paris, 1978/1979) (French), Berlin: Springer, 397-409.
    [10] Kogoj AE, Lanconelli E (2007) Liouville theorems for a class of linear second-order operators with nonnegative characteristic form. Bound Value Probl 2007: 16.
    [11] Kupcov LP (1972) The fundamental solutions of a certain class of elliptic-parabolic second order equations. Differ Uravn 8: 1649-1660.
    [12] Lanconelli E, Polidoro S (1994) On a class of hypoelliptic evolution operators. Rend Semin Mat U Pad 52: 29-63.
    [13] Priola E, Wang FY (2006) Gradient estimates for diffusion semigroups with singular coefficients. J Funct Anal 236: 244-264. doi: 10.1016/j.jfa.2005.12.010
    [14] Priola E, Zabczyk J (2004) Liouville theorems for non-local operators. J Funct Anal 216: 455-490. doi: 10.1016/j.jfa.2004.04.001
    [15] Zabczyk J (1981/82) Controllability of stochastic linear systems. Syst Control Lett 1: 25-31.
  • This article has been cited by:

    1. Marco Cirant, Alessandro Goffi, 2023, 781, 978-1-4704-7208-5, 7, 10.1090/conm/781/15707
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3815) PDF downloads(328) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog