This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

LlocpL^{p}_{loc} positivity preservation and
Liouville-type theorems

Andrea Bisterzo Università degli Studi di Milano-Bicocca
Dipartimento di Matematica e Applicazioni
Via Cozzi 55, 20126 Milano - ITALY
[email protected]
Alberto Farina Université Picardie Jules Verne, LAMFA, UMR CNRS 7352, 33 rue St. Leu, 80039 Amiens - FRANCE [email protected]  and  Stefano Pigola Università degli Studi di Milano-Bicocca
Dipartimento di Matematica e Applicazioni
Via Cozzi 55, 20126 Milano - ITALY
[email protected]
Abstract.

On a complete Riemannian manifold (M,g)(M,g), we consider LlocpL^{p}_{loc} distributional solutions of the the differential inequality Δu+λu0-\Delta u+\lambda u\geq 0 with λ>0\lambda>0 a locally bounded function that may decay to 0 at infinity. Under suitable growth conditions on the LpL^{p} norm of uu over geodesic balls, we obtain that any such solution must be nonnegative. This is a kind of generalized LpL^{p}-preservation property that can be read as a Liouville type property for nonnegative subsolutiuons of the equation Δuλu\Delta u\geq\lambda u. An application of the analytic results to LpL^{p} growth estimates of the extrinsic distance of complete minimal submanifolds is also given.

1. Introduction

In order to set our work, we first recall the notion of differential inequality in the sense of distributions. Let (M,g)(M,g) be a Riemannian manifold and λ\lambda a smooth function over MM. Given fLloc1(M)f\in L^{1}_{loc}(M), we say that a function uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λuf-\Delta u+\lambda u\geq f (respectively f\leq f) in the sense of distributions if

Mu(Δφ+λφ)dvMfφdv(resp.)\displaystyle\int_{M}u(-\Delta\varphi+\lambda\varphi)\ \textnormal{dv}\geq\int_{M}f\varphi\ \textnormal{dv}\quad(\textnormal{resp.}\ \leq)

for every 0φCc(M)0\leq\varphi\in C^{\infty}_{c}(M). Using an integration by parts, one can easily see that the notion of differential inequality in the sense of distributions is a generalization of the notion of weak differential inequality, which involves W1,1W^{1,1} functions.

Definition 1.1 (Positivity preserving property).

Given a Riemannian manifold (M,g)(M,g) and a family of function 𝒮Lloc1(M)\mathcal{S}\subseteq L^{1}_{loc}(M), we say that MM has the 𝒮\mathcal{S} positivity preserving property if any function u𝒮u\in\mathcal{S} that satisfies Δu+u0-\Delta u+u\geq 0 in the sense of distributions is nonnegative almost everywhere in MM.

Historically, the notion of positivity preserving property is motivated by the work of M. Braverman, O. Milatovic and M. Shubin, [3], where the authors conjectured that every complete Riemannian manifold is L2L^{2} positivity preserving. In particular, this conjecture stimulated the study of the correlation between completeness and LpL^{p} positivity preserving property for any p[1,+]p\in[1,+\infty].

After some partial results involving constraints on the geometry of the manifold at hand, and covering all cases p[1,+]p\in[1,+\infty], in [10] (see also [9] and [4]) the authors proved that any Riemannian manifold (M,g)(M,g) is LpL^{p} positivity preserving for every p(1,+)p\in(1,+\infty), under the only assumption that MM is complete.

For what concerns the case p=+p=+\infty, the recent work [2] points out that the LL^{\infty} positivity preservation is in a certain sense transversal to the notion of (geodesic) completeness. Indeed, the authors showed that a necessary and sufficient condition for a Riemannian manifold to satisfy the LL^{\infty} positivity preserving property is the stochastic completeness of the space.

Modeled on Definition 1.1, in what follows we will consider a notion of positivity preserving property for slightly more general differential operators. In particular, we will deal with operators of the form :=Δ+λ\mathcal{L}:=-\Delta+\lambda, where λ\lambda is a positive and locally bounded function. In this context, the present work generalizes the result of [10] and [9] for complete Riemannian manifolds providing the 𝒮p\mathcal{S}_{p} positivity preservation for any p(1,+)p\in(1,+\infty), where 𝒮p\mathcal{S}_{p} is the family of locally pp-integrable functions satisfying a certain growth condition depending on the decay rate of the potential λ\lambda at infinity. To follow, we obtain two results for the case p=1p=1 when λ\lambda is a positive constant, under the assumption that there exists a family of suitable (exhausting) cut-off functions whose laplacians have a “good” decay.

We stress that the results we obtained can be read as LpL^{p} Liouville-type theorems when one deals with nonnegative solutions to Δuλu\Delta u\geq\lambda u. In this direction we have a more direct comparison with the existing literature where, typically, one introduces a further pointwise control on the growth of the function and requires much more regularity on the solution. In the next sections we shall comment on these aspects.

The paper is organized as follows. In Section 2 we prove an integral inequality in low regularity which represents the core of the LlocpL^{p}_{loc} argument, 1<p<+1<p<+\infty. Section 3 is devoted to a generalized LlocpL^{p}_{loc} Positivity Preservation for distributional solutions of Δu+λu0-\Delta u+\lambda u\geq 0 with a possibly decaying functions λ(x)\lambda(x). The case p=1p=1 will be dealt with in Section 4 under additional curvature restrictions that guarantee the existence of the so called Laplacian cut-offs. In the final Section 5 we present an application to complete Euclidean minimal submanifolds with an extrinsic distance growth measured in integral sense. This generalizes the well known fact that complete minimal submanifolds in Euclidean space with quadratic-exponential volume growth must be unbounded; see [6] and [8].

2. Some preliminary results

In what follows, if uu is a real-valued function we denote

u+:=max{u,0}andu:=max{u,0}.\displaystyle u^{+}:=\max\{u,0\}\quad\textnormal{and}\quad u^{-}:=\max\{-u,0\}.

We start recalling the Brezis-Kato inequality in a general Riemannian setting. This result is obtained in [9] for the general inequality ΔufLloc1\Delta u\geq f\in L^{1}_{loc}.

Proposition 2.1 (Brezis-Kato inequality).

Let (M,g)(M,g) be a possibly incomplete Riemannian manifold and λ\lambda a measurable function.

If uLloc1(M)u\in L^{1}_{loc}(M) is so that λuLloc1(M)\lambda u\in L^{1}_{loc}(M) and satisfies Δu+λu0-\Delta u+\lambda u\leq 0 in the sense of distributions, then Δu++λu+0-\Delta u^{+}+\lambda u^{+}\leq 0 in the sense of distributions.

As a consequence, in the next proposition we get a refinement of the regularity result obtained in [9] for complete manifolds. The inequality (2.1) will be the key tool in the proof of the positivity preserving properties stated in Section 3.

Proposition 2.2.

Let (M,g)(M,g) be a complete Riemannian manifold and 0λLloc(M)0\leq\lambda\in L^{\infty}_{loc}(M). Assume that uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions.

Then, uLloc(M)u^{-}\in L^{\infty}_{loc}(M) and (u)p2Wloc1,2(M)(u^{-})^{\frac{p}{2}}\in W^{1,2}_{loc}(M) for every p(1,+)p\in(1,+\infty). Moreover, uu^{-} satisfies

(2.1) (p1)Mλ(u)pφ2dvM(u)p|φ|2dv\displaystyle(p-1)\int_{M}\lambda(u^{-})^{p}\varphi^{2}\ \textnormal{dv}\leq\int_{M}(u^{-})^{p}|\nabla\varphi|^{2}\ \textnormal{dv}

for every 0φCc0,1(M)0\leq\varphi\in C^{0,1}_{c}(M).

Proof.

By the Brezis-Kato inequality, the function uLloc1(M)u^{-}\in L^{1}_{loc}(M) satisfies Δuλu\Delta u^{-}\geq\lambda u^{-} in the sense of distributions. Therefore, by [10, Theorem 3.1] it follows that uLloc(M)u^{-}\in L^{\infty}_{loc}(M) and (u)p2Wloc1,2(M)(u^{-})^{\frac{p}{2}}\in W^{1,2}_{loc}(M) for every p(1,+)p\in(1,+\infty).

To prove (2.1), let δ>0\delta>0 and set vδ:=u+δLloc(M)Wloc1,2(M)v_{\delta}:=u^{-}+\delta\in L^{\infty}_{loc}(M)\cap W^{1,2}_{loc}(M). Clearly, for every q>0q>0 the function vδqv_{\delta}^{q} belongs to Lloc(M)Wloc1,2(M)L^{\infty}_{loc}(M)\cap W^{1,2}_{loc}(M) and by [10, Lemma 5.4] its weak gradient satisfies

(2.2) vδq=qvδq1vδ.\displaystyle\nabla v^{q}_{\delta}=qv^{q-1}_{\delta}\nabla v_{\delta}.

Moreover, Δvδλu\Delta v_{\delta}\geq\lambda u^{-} in the sense of distributions, implying

Mλuψdv+Mg(vδ,ψ)0\displaystyle\int_{M}\lambda u^{-}\psi\ \textnormal{dv}+\int_{M}g(\nabla v_{\delta},\nabla\psi)\leq 0

for every ψWloc1,2(M)\psi\in W^{1,2}_{loc}(M) that is nonnegative almost everywhere. In particular, choosing ψ=vδp1φ2\psi=v_{\delta}^{p-1}\varphi^{2} with φCc0,1(M)\varphi\in C^{0,1}_{c}(M) and using (2.2), we get

0\displaystyle 0\geq Mλuvδp1φ2dv+(p1)Mvδp2φ2|vδ|2dv\displaystyle\int_{M}\lambda u^{-}v_{\delta}^{p-1}\varphi^{2}\ \textnormal{dv}+(p-1)\int_{M}v_{\delta}^{p-2}\varphi^{2}|\nabla v_{\delta}|^{2}\ \textnormal{dv}
+2Mφvδp1g(vδ,φ)dv.\displaystyle+2\int_{M}\varphi v_{\delta}^{p-1}g(\nabla v_{\delta},\nabla\varphi)\ \textnormal{dv}.

By Cauchy-Schwarz inequality and Young’s inequality, for any ϵ(0,p1)\epsilon\in(0,p-1) we have

2φvδp1g(vδ,φ)\displaystyle 2\varphi v_{\delta}^{p-1}g(\nabla v_{\delta},\nabla\varphi) 2φvδp1|vδ||φ|\displaystyle\geq-2\varphi v_{\delta}^{p-1}|\nabla v_{\delta}||\nabla\varphi|
ϵφ2vδp2|vδ|2ϵ1vδp|φ|2\displaystyle\geq-\epsilon\varphi^{2}v_{\delta}^{p-2}|\nabla v_{\delta}|^{2}-\epsilon^{-1}v_{\delta}^{p}|\nabla\varphi|^{2}

and thus

0\displaystyle 0\geq Mλuvδp1φ2dv+(p1ϵ)Mvδp2φ2|vδ|2dv\displaystyle\int_{M}\lambda u^{-}v_{\delta}^{p-1}\varphi^{2}\ \textnormal{dv}+(p-1-\epsilon)\int_{M}v_{\delta}^{p-2}\varphi^{2}|\nabla v_{\delta}|^{2}\ \textnormal{dv}
ϵ1Mvδp|φ|2dv.\displaystyle-\epsilon^{-1}\int_{M}v_{\delta}^{p}|\nabla\varphi|^{2}\ \textnormal{dv}.

As ϵp1\epsilon\to p-1 we get

(p1)Mλuvδp1φ2dvMvδp|φ|2dv\displaystyle(p-1)\int_{M}\lambda u^{-}v_{\delta}^{p-1}\varphi^{2}\ \textnormal{dv}\leq\int_{M}v_{\delta}^{p}|\nabla\varphi|^{2}\ \textnormal{dv}

that, together with the fact that

λuvδp1δ0λ(u)pinLloc1(M)vδpδ0(u)pinLloc1(M)\displaystyle\begin{array}[]{rll}\lambda u^{-}v_{\delta}^{p-1}&\xrightarrow[]{\delta\to 0}\lambda(u^{-})^{p}&\textnormal{in}\ L^{1}_{loc}(M)\\ v_{\delta}^{p}&\xrightarrow[]{\delta\to 0}(u^{-})^{p}&\textnormal{in}\ L^{1}_{loc}(M)\end{array}

by Dominated Convergence Theorem, implies

(p1)Mλ(u)pφ2dvM(u)p|φ|2dv\displaystyle(p-1)\int_{M}\lambda(u^{-})^{p}\varphi^{2}\ \textnormal{dv}\leq\int_{M}(u^{-})^{p}|\nabla\varphi|^{2}\ \textnormal{dv}

obtaining the claim. ∎

3. LlocpL^{p}_{loc} positivity preserving property

In this section we face up the question of the LlocpL^{p}_{loc} positivity preserving property for p(1,+)p\in(1,+\infty), considering complete Riemannian manifolds and not requiring any curvature assumption.

Clearly, if the manifold is non-compact, we do not have any control on the growth at “infinity” of (the pp-norm of) the general function uLlocp(M)u\in L^{p}_{loc}(M), making it impossible to retrace step by step what has been done in [10] and [9] in the LpL^{p} case.

In addition, we also point out that we cannot expect to obtain a genuine positivity preserving property on the whole family of functions Llocp(M)L^{p}_{loc}(M). Indeed, if λ\lambda is a positive constant, then u(x)=eλxu(x)=-e^{\sqrt{\lambda}x} is a negative function that solves u′′+λu=0-u^{\prime\prime}+\lambda u=0 in \mathbb{R}. So the LlocpL^{p}_{loc} positivity preserving property fails in general complete Riemannian manifolds.

Taking into account what we have observed so far, it seems natural to limit ourselves to the class of LlocpL^{p}_{loc} functions whose pp-norms satisfy a suitable (sub-exponential) growth condition.

We start with the following iterative lemma.

Lemma 3.1.

Let A>0A>0 and f:[A,+)(0,+)f:[A,+\infty)\to(0,+\infty) be a nondecreasing function. Suppose there exist α>0,δ0,β1\alpha>0,\delta\geq 0,\beta\geq 1 and γ>0\gamma>0 so that

(3.1) f(r)1α(1+r)δhγ+βf(r+h)\displaystyle f(r)\leq\frac{1}{\alpha(1+r)^{-\delta}h^{\gamma}+\beta}f(r+h)

for every rAr\geq A and every h>0h>0.

Then, for every h>0h>0 the function ff satisfies

f(R)f(A)(α(1+Rh)δhγ+β)RAh1\displaystyle f(R)\geq f(A)\left(\alpha(1+R-h)^{-\delta}h^{\gamma}+\beta\right)^{\frac{R-A}{h}-1}

for every RA+hR\geq A+h.

Proof.

Fixed h>0h>0, by assumption we have f(r)(α(1+r)δhγ+β)1f(r+h)f(r)\leq\left(\alpha(1+r)^{-\delta}h^{\gamma}+\beta\right)^{-1}f(r+h) for any rAr\geq A. Iterating, for every nn\in\mathbb{N} we get

f(r)\displaystyle f(r) (α(1+r)δhγ+β)1f(r+h)\displaystyle\leq\left(\alpha(1+r)^{-\delta}h^{\gamma}+\beta\right)^{-1}f(r+h)
(α(1+r)δhγ+β)1(α(1+r+h)δhγ+β)1f(r+2h)\displaystyle\leq\left(\alpha(1+r)^{-\delta}h^{\gamma}+\beta\right)^{-1}\left(\alpha(1+r+h)^{-\delta}h^{\gamma}+\beta\right)^{-1}f(r+2h)
(α(1+r+h)δhγ+β)2f(r+2h)\displaystyle\leq\left(\alpha(1+r+h)^{-\delta}h^{\gamma}+\beta\right)^{-2}f(r+2h)
(α(1+r+(n1)h)δhγ+β)nf(r+nh)\displaystyle\leq...\leq\left(\alpha(1+r+(n-1)h)^{-\delta}h^{\gamma}+\beta\right)^{-n}f(r+nh)

for any rAr\geq A. It follows that for every R>AR>A

f(R)\displaystyle f(R) f(A+nh)\displaystyle\geq f(A+nh)
(α(1+A+(n1)h)δhγ+β)nf(A)\displaystyle\geq\left(\alpha(1+A+(n-1)h)^{-\delta}h^{\gamma}+\beta\right)^{n}f(A)
(α(1+A+(n1)h)δhγ+β)RAh1f(A),\displaystyle\geq\left(\alpha(1+A+(n-1)h)^{-\delta}h^{\gamma}+\beta\right)^{\frac{R-A}{h}-1}f(A),

where n=n(R,A,h)n=n(R,A,h) is the unique natural number satisfying A+(n+1)hRA+nhA+(n+1)h\geq R\geq A+nh. In particular, if RA+hR\geq A+h, then RAh1\frac{R-A}{h}\geq 1 obtaining

f(R)\displaystyle f(R) (α(1+A+(n1)h)δhγ+β)RAh1f(A)\displaystyle\geq\left(\alpha(1+A+(n-1)h)^{-\delta}h^{\gamma}+\beta\right)^{\frac{R-A}{h}-1}f(A)
(α(1+Rh)δhγ+β)RAh1f(A)\displaystyle\geq\left(\alpha(1+R-h)^{-\delta}h^{\gamma}+\beta\right)^{\frac{R-A}{h}-1}f(A)

since RAh1n1\frac{R-A}{h}-1\geq n-1. This concludes the proof. ∎

Combining Lemma 3.1 with Proposition 2.2 and with the choice standard family of rotationally symmetric cut-off functions, we get the following theorem.

Theorem 3.2 (Generalized LlocpL^{p}_{loc} positivity preserving property).

Let (M,g)(M,g) be a complete Riemannian manifold, λLloc(M)\lambda\in L^{\infty}_{loc}(M) a positive function and p(1,+)p\in(1,+\infty). Moreover, assume there exist oMo\in M and a constant C>0C>0 so that

λ(x)C(1+dM(x,o))2ϵxM,\displaystyle\lambda(x)\geq\frac{C}{(1+d^{M}(x,o))^{2-\epsilon}}\quad\forall x\in M,

where ϵ(0,2]\epsilon\in(0,2] and dMd^{M} is the intrinsic distance on MM.

If uLlocp(M)u\in L^{p}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and

(3.2) BR(o)(u)pdv=o(eθRϵ2)asR+,\displaystyle\int_{B_{R}(o)}(u^{-})^{p}\ \textnormal{dv}=o\left(e^{\theta R^{\frac{\epsilon}{2}}}\right)\quad as\ R\to+\infty,

where θ=(p1)Ce1\theta=\sqrt{\frac{(p-1)C}{e-1}}, then u0u\geq 0.

Remark 3.3 (A Liouville-type theorem).

It clearly follows that the unique nonpositive LlocpL^{p}_{loc} distributional solution to Δu+λu0-\Delta u+\lambda u\geq 0 that satisfies condition (3.2) is the null function. In this sense, Theorem 3.2 can be read as an LpL^{p} Liouville-type theorem.

Remark 3.4.

The case ϵ>2\epsilon>2 can be considered by reducing the problem to the case ϵ=2\epsilon=2, since

λ(x)C(1+dM(x,o))ϵ2CxM.\displaystyle\lambda(x)\geq C(1+d^{M}(x,o))^{\epsilon-2}\geq C\quad\forall x\in M.
Proof.

Let uLlocp(M)u\in L^{p}_{loc}(M) be a distributional solution to Δu+λu0-\Delta u+\lambda u\geq 0 satisfying (3.2). For any fixed a>0a>0 and b>ab>a, consider the function ηa,bC0,1([0,+))\eta_{a,b}\in C^{0,1}([0,+\infty)) so that

{ηa,b1in[0,a]ηa,b(t)=btbain[a,b]ηa,b0in[b,+).\displaystyle\left\{\begin{array}[]{ll}\eta_{a,b}\equiv 1&\textnormal{in}\ [0,a]\\ \eta_{a,b}(t)=\frac{b-t}{b-a}&\textnormal{in}\ [a,b]\\ \eta_{a,b}\equiv 0&\textnormal{in}\ [b,+\infty).\end{array}\right.

In particular, |ηa,b(t)|1ba|\eta_{a,b}^{\prime}(t)|\leq\frac{1}{b-a} almost everywhere in [0,+)[0,+\infty).

Set φa,b(x):=ηa,b(d(x,o))\varphi_{a,b}(x):=\eta_{a,b}(d(x,o)), where d(,)d(\cdot,\cdot) is the intrinsic distance on MM. Then, φa,bCc0,1(M)\varphi_{a,b}\in C^{0,1}_{c}(M) and satisfies

{φa,b0inM|φa,b(x)|1baa.e.inMφa,b0inMBb(o)¯φa,b1inBa(o).\displaystyle\left\{\begin{array}[]{ll}\varphi_{a,b}\geq 0&\textnormal{in}\ M\\ |\nabla\varphi_{a,b}(x)|\leq\frac{1}{b-a}&\textnormal{a.e.}\ \textnormal{in}\ M\\ \varphi_{a,b}\equiv 0&\textnormal{in}\ M\setminus\overline{B_{b}(o)}\\ \varphi_{a,b}\equiv 1&\textnormal{in}\ B_{a}(o).\end{array}\right.

Using φ=φa,b\varphi=\varphi_{a,b} in (2.1), we get

1(ba)2Bb(o)Ba(o)(u)pdv\displaystyle\frac{1}{(b-a)^{2}}\int_{B_{b}(o)\setminus B_{a}(o)}(u^{-})^{p}\ \textnormal{dv} (p1)Ba(o)λ(u)pdv\displaystyle\geq(p-1)\int_{B_{a}(o)}\lambda(u^{-})^{p}\ \textnormal{dv}
(p1)Ba(o)C(1+dM(,o))2ϵ(u)pdv\displaystyle\geq(p-1)\int_{B_{a}(o)}\frac{C}{(1+d^{M}(\cdot,o))^{2-\epsilon}}(u^{-})^{p}\ \textnormal{dv}
(p1)C(1+a)2ϵBa(o)(u)pdv\displaystyle\geq(p-1)\frac{C}{(1+a)^{2-\epsilon}}\int_{B_{a}(o)}(u^{-})^{p}\ \textnormal{dv}

and, by summing

1(ba)2Ba(o)(u)pdv\displaystyle\frac{1}{(b-a)^{2}}\int_{B_{a}(o)}(u^{-})^{p}\ \textnormal{dv}

to both sides of previous inequality, we obtain

((p1)C(1+a)2ϵ+1(ba)2)\displaystyle\left((p-1)\frac{C}{(1+a)^{2-\epsilon}}+\frac{1}{(b-a)^{2}}\right) Ba(o)(u)pdv1(ba)2Bb(o)(u)pdv\displaystyle\int_{B_{a}(o)}(u^{-})^{p}\ \textnormal{dv}\leq\frac{1}{(b-a)^{2}}\int_{B_{b}(o)}(u^{-})^{p}\ \textnormal{dv}

for every fixed a>0a>0 and b>ab>a. In particular, it implies that

(3.3) Ba(o)(u)pdv1(p1)C(1+a)ϵ2h2+1Ba+h(o)(u)pdv\displaystyle\begin{split}\int_{B_{a}(o)}(u^{-})^{p}\ \textnormal{dv}&\leq\frac{1}{(p-1)C(1+a)^{\epsilon-2}h^{2}+1}\int_{B_{a+h}(o)}(u^{-})^{p}\ \textnormal{dv}\end{split}

for every a>0a>0 and h>0h>0.

If we suppose that u0u^{-}\neq 0, then there exists A>0A>0 so that

BA(o)(u)pdv>0.\displaystyle\int_{B_{A}(o)}(u^{-})^{p}\ \textnormal{dv}>0.

By (3.3) we can apply Lemma 3.1 to

f:aBa(o)(u)pdv\displaystyle f:a\mapsto\int_{B_{a}(o)}(u^{-})^{p}\ \textnormal{dv}

in [A,+)[A,+\infty), with γ=ϵ\gamma=\epsilon, δ=2ϵ\delta=2-\epsilon, α=(p1)C\alpha=(p-1)C and β=1\beta=1 and we get that for any h>0h>0 and for any R>A+hR>A+h the function ff satisfies

f(R)f(A)((p1)C(1+Rh)ϵ2h2+1)RAh1.\displaystyle f(R)\geq f(A)\Big{(}(p-1)C(1+R-h)^{\epsilon-2}h^{2}+1\Big{)}^{\frac{R-A}{h}-1}.

If 0<ϵ<20<\epsilon<2 we can take h=R1ϵ2e1(p1)Ch=R^{1-\frac{\epsilon}{2}}\sqrt{\frac{e-1}{(p-1)C}}, obtaining

f(R)\displaystyle f(R) f(A)((p1)Ch2(1+Rh)2ϵ+1)RAh1\displaystyle\geq f(A)\left((p-1)C\frac{h^{2}}{(1+R-h)^{2-\epsilon}}+1\right)^{\frac{R-A}{h}-1}
f(A)((p1)Ch2(h+Rh)2ϵ+1)RAh1\displaystyle\geq f(A)\left((p-1)C\frac{h^{2}}{(h+R-h)^{2-\epsilon}}+1\right)^{\frac{R-A}{h}-1}
=f(A)((p1)Ch2R2ϵ+1)RAh1\displaystyle=f(A)\left((p-1)C\frac{h^{2}}{R^{2-\epsilon}}+1\right)^{\frac{R-A}{h}-1}
=f(A)eAh1eRh\displaystyle=f(A)e^{-\frac{A}{h}-1}e^{\frac{R}{h}}
f(A)e12eθRϵ2\displaystyle\geq\frac{f(A)e^{-1}}{2}e^{\theta R^{\frac{\epsilon}{2}}}

for every RR big enough so that

R>A+h,h1andeAh12.\displaystyle R>A+h,\quad h\geq 1\quad\textnormal{and}\quad e^{-\frac{A}{h}}\geq\frac{1}{2}.

Similarly, if ϵ=2\epsilon=2 we can choose h=e1(p1)Ch=\sqrt{\frac{e-1}{(p-1)C}}, in order to get

f(R)\displaystyle f(R) f(A)((p1)Ch2+1)RAh1\displaystyle\geq f(A)\left((p-1)Ch^{2}+1\right)^{\frac{R-A}{h}-1}
=f(A)eθA1eθR.\displaystyle=f(A)e^{-\theta A-1}e^{\theta R}.

In both cases we obtain a contradiction to (3.2), implying that u=0u^{-}=0 almost everywhere, i.e. the claim. ∎

Remark 3.5.

In the paper [7] by L. Mari, M. Rigoli and A.G. Setti, using the viewpoint of maximum principles at infinity for the φ\varphi-Laplacian, the authors proved a general a priori estimate that, in our setting, reduces as follow.

Theorem 3.6 ([7, Theorem B]).

Let (M,g)(M,g) be a complete Riemannian manifold and λC(M)\lambda\in C(M) be a positive function satisfying

λ(x)Br(x)2ϵinMBR0(o)\displaystyle\lambda(x)\geq\frac{B}{r(x)^{2-\epsilon}}\quad in\ M\setminus B_{R_{0}}(o)

for some ϵ(0,+),B>0,R0>0\epsilon\in(0,+\infty),B>0,R_{0}>0 and oMo\in M.

Let σ0\sigma\geq 0 and uC1(M)u\in C^{1}(M) be a distributional solution to

Δu+λu0inM\displaystyle-\Delta u+\lambda u\geq 0\quad in\ M

so that either u(x)=o(r(x)σ)u^{-}(x)=o(r(x)^{\sigma}) as r(x)+r(x)\to+\infty, if σ>0\sigma>0, or uu is bounded from below, if σ=0\sigma=0. Lastly, assume

lim infr+ln|Br(o)|rϵσ<+ifσ<ϵ\displaystyle\liminf_{r\to+\infty}\frac{\ln|B_{r}(o)|}{r^{\epsilon-\sigma}}<+\infty\quad if\ \sigma<\epsilon

or

lim infr+ln|Br(o)|lnr<+ifσ=ϵ.\displaystyle\liminf_{r\to+\infty}\frac{\ln|B_{r}(o)|}{\ln r}<+\infty\quad if\ \sigma=\epsilon.

Then, u0u\geq 0.

This result compares with our Theorem 3.2. Indeed, on the one hand, if we assume the pointwise control u(x)=o(rσ(x))u^{-}(x)=o(r^{\sigma}(x)), for 0<σ<ϵ0<\sigma<\epsilon, condition (3.2) is satisfied provided |BR|=O(RpσeθRϵ2)|B_{R}|=O(R^{-p\sigma}e^{\theta R^{\frac{\epsilon}{2}}}), p(1,+)p\in(1,+\infty), while Theorem 3.6 requires the volume growth |BR|=O(eRϵσ)|B_{R}|=O(e^{R^{\epsilon-\sigma}}).

On the other hand, our Theorem 3.2 improves Theorem 3.6 in two aspects. First of all, we require less regularity on the functions uu and λ\lambda. Indeed, we only need LlocpL_{loc}^{p} solutions with LlocL^{\infty}_{loc} potentials in order to use the Kato inequality and the regularity result claimed in Section 2. Secondly, we only need an LpL^{p}-bound on the asymptotic growth of uu^{-}, instead of a pointwise asymptotic control. This allows us to consider a wider class of functions, for example having a super-quadratic growth, even in the case ϵ<2\epsilon<2.

In the particular case where ϵ=2\epsilon=2, for instance when λ\lambda is a constant, we get the next version of Theorem 3.2.

Corollary 3.7.

Let (M,g)(M,g) be a complete Riemannian manifold, λLloc(M)\lambda\in L^{\infty}_{loc}(M) so that λC\lambda\geq C for a positive constant CC and p(1,+)p\in(1,+\infty).

If uLlocp(M)u\in L^{p}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and

(3.4) BR(u)pdv=o(eθR)asR+\displaystyle\int_{B_{R}}(u^{-})^{p}\ \textnormal{dv}=o(e^{\theta R})\quad as\ R\to+\infty

with θ=(p1)Ce1\theta=\sqrt{\frac{(p-1)C}{e-1}}, then u0u\geq 0 in MM.

Remark 3.8.

Corollary 3.7 improves very much one of the main results of [9] in the setting of complete manifolds. Indeed, in that paper, the LlocpL^{p}_{loc} positivity preservation is obtained under the condition BR(o)(u)pdv=o(R2)\int_{B_{R}(o)}{(u^{-})}^{p}\ \textnormal{dv}=o(R^{2}). See [9, Corollary 5.2 and Remark 5.3].

As a byproduct, by applying Corollary 3.7 to both the functions uu and u-u, we get an uniqueness statement for LlocpL^{p}_{loc} solutions to Δu+λu=0-\Delta u+\lambda u=0.

Corollary 3.9 (Uniqueness).

Let (M,g)(M,g) be a complete Riemannian manifold, λLloc(M)\lambda\in L^{\infty}_{loc}(M) so that λC\lambda\geq C for a positive constant CC and p(1,+)p\in(1,+\infty).

If uLlocp(M)u\in L^{p}_{loc}(M) satisfies Δu+λu=0-\Delta u+\lambda u=0 in the sense of distributions and

BR(u±)pdv=o(eθR)asR+\displaystyle\int_{B_{R}}(u^{\pm})^{p}\ \textnormal{dv}=o(e^{\theta R})\quad as\ R\to+\infty

with θ=(p1)λe1\theta=\sqrt{\frac{(p-1)\lambda}{e-1}}, then u=0u=0 almost everywhere in MM.

Remark 3.10.

As already observed at the beginning of this section, for every λ>0\lambda>0 the function u(x)=eλxu(x)=-e^{\sqrt{\lambda}x} provides a counterexample to the Llocp()L^{p}_{loc}(\mathbb{R}) positivity preserving property, for any p(1,+)p\in(1,+\infty). Moreover, we stress that its pp-norm has the following asymptotic growth

RR(u)p(x)dx=O(epλR)\displaystyle\int_{-R}^{R}(u^{-})^{p}(x)\ \textnormal{d}x=O(e^{p\sqrt{\lambda}R})

with pλ>(p1)λe1p\sqrt{\lambda}>\sqrt{\frac{(p-1)\lambda}{e-1}}. Therefore, Theorem 3.2 and Corollary 3.7 are not far from being sharp. It would be very interesting to understand to what extent this exponent can be refined.

4. Lloc1L^{1}_{loc} positivity preserving property

The approach used in Section 3, which is based on inequality (2.1), is clearly not applicable for p=1p=1. To overcome this problem, we resort to some special cut-off to be used as test functions in the distributional inequality satisfied by uu. The existence of these functions is ensured, for instance, by requiring certain conditions on the decay of the Ricci curvature.

4.1. Cut-off functions with decaying laplacians

The first theorem we present in this section is based on the following iterative lemma. It is an analogue of the Lemma 3.1 for the case p=1p=1.

Lemma 4.1.

Let A>0A>0 and f:[A,+)(0,+)f:[A,+\infty)\to(0,+\infty) be a nondecreasing function. Suppose there exist σ>1,γ>0,α>0\sigma>1,\gamma>0,\alpha>0 and β1\beta\geq 1 so that

(4.1) f(r)1αrγ+βf(σr)\displaystyle f(r)\leq\frac{1}{\alpha r^{\gamma}+\beta}f(\sigma r)

for every rAr\geq A. Then, ff satisfies

f(R)(RA)logσ(αAγ+β)f(A)αAγ+β\displaystyle f(R)\geq\left(\frac{R}{A}\right)^{\log_{\sigma}(\alpha A^{\gamma}+\beta)}\frac{f(A)}{\alpha A^{\gamma}+\beta}

for every R>AR>A.

Proof.

Having fixed RAR\geq A, we have

f(R)\displaystyle f(R) (αRγ+β)1f(σR)\displaystyle\leq(\alpha R^{\gamma}+\beta)^{-1}f(\sigma R)
(αRγ+β)1(α(σR)γ+β)1f(σ2R)\displaystyle\leq(\alpha R^{\gamma}+\beta)^{-1}(\alpha(\sigma R)^{\gamma}+\beta)^{-1}f(\sigma^{2}R)
(αRγ+β)2f(σ2R)\displaystyle\leq(\alpha R^{\gamma}+\beta)^{-2}f(\sigma^{2}R)

and, iterating,

f(R)(αRγ+β)nf(σnR)\displaystyle f(R)\leq(\alpha R^{\gamma}+\beta)^{-n}f(\sigma^{n}R)

for every nn\in\mathbb{N}.

Now consider nn\in\mathbb{N} so that σn+1ARσnA\sigma^{n+1}A\geq R\geq\sigma^{n}A. In particular, from

σn+1ARnlogσ(RA)1\displaystyle\sigma^{n+1}A\geq R\quad\Rightarrow\quad n\geq\log_{\sigma}\left(\frac{R}{A}\right)-1

we deduce

f(R)\displaystyle f(R) f(σnA)(αAγ+β)nf(A)\displaystyle\geq f(\sigma^{n}A)\geq(\alpha A^{\gamma}+\beta)^{n}f(A)
(αAγ+β)logσ(RA)f(A)αAγ+β\displaystyle\geq(\alpha A^{\gamma}+\beta)^{\log_{\sigma}\left(\frac{R}{A}\right)}\frac{f(A)}{\alpha A^{\gamma}+\beta}
=(RA)logσ(αAγ+β)f(A)αAγ+β\displaystyle=\left(\frac{R}{A}\right)^{\log_{\sigma}(\alpha A^{\gamma}+\beta)}\frac{f(A)}{\alpha A^{\gamma}+\beta}

as claimed ∎

As a consequence, by requiring the existence of a family {ϕR}R\{\phi_{R}\}_{R} of cut-off functions whose laplacians decay as |ΔϕR|CRγ|\Delta\phi_{R}|\leq CR^{-\gamma} for a positive constant γ\gamma, we get

Theorem 4.2 (Generalized Lloc1L^{1}_{loc} positivity preserving property).

Let (M,g)(M,g) be a complete Riemannian manifold and λ\lambda a positive constant. Assume that for a fixed oMo\in M there exist some positive constants γ\gamma and R0R_{0} and a constant σ>1\sigma>1 satisfying the following condition: for every R>R0R>R_{0} there exists ϕRCc2(M)\phi_{R}\in C^{2}_{c}(M) such that

(4.6) {0ϕR1inMϕR1inBR(o)supp(ϕR)BσR(o)|ΔϕR|CRγinM\displaystyle\left\{\begin{array}[]{ll}0\leq\phi_{R}\leq 1&\textnormal{in}\ M\\ \phi_{R}\equiv 1&\textnormal{in}\ B_{R}(o)\\ \textnormal{supp}(\phi_{R})\subset B_{\sigma R}(o)\\ |\Delta\phi_{R}|\leq\frac{C}{R^{\gamma}}&\textnormal{in}\ M\end{array}\right.

where C=C(σ)>0C=C(\sigma)>0 is a constant not depending on RR. If uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and there exists kk\in\mathbb{N} so that

(4.7) BR(o)udv=O(Rk)asR+,\displaystyle\int_{B_{R}(o)}u^{-}\ \textnormal{dv}=O(R^{k})\quad as\ R\to+\infty,

then u0u\geq 0 almost everywhere in MM.

Proof.

Fix uLloc1(M)u\in L^{1}_{loc}(M) a distributional solution to Δu+λu0-\Delta u+\lambda u\geq 0 that satisfies condition (4.7) for a certain kk\in\mathbb{N}. By Brezis-Kato inequality Δuλu\Delta u^{-}\geq\lambda u^{-} in the sense of distributions, implying

λMuϕRdvMuΔϕRdvR>R0.\displaystyle\lambda\int_{M}u^{-}\phi_{R}\ \textnormal{dv}\leq\int_{M}u^{-}\Delta\phi_{R}\ \textnormal{dv}\quad\quad\forall R>R_{0}.

Using the definition of ϕR\phi_{R}, we get

λBR(o)uϕRdvCRγBσR(o)BR(o)udvR>R0\displaystyle\lambda\int_{B_{R}(o)}u^{-}\phi_{R}\ \textnormal{dv}\leq\frac{C}{R^{\gamma}}\int_{B_{\sigma R}(o)\setminus B_{R}(o)}u^{-}\ \textnormal{dv}\quad\quad\forall R>R_{0}

and, by summing up

CRγBR(o)udv\displaystyle\frac{C}{R^{\gamma}}\int_{B_{R}(o)}u^{-}\ \textnormal{dv}

to both sides of the previous inequality, we obtain

(4.8) BR(o)udvCλRγ+CBσR(o)udv=1αRγ+1BσR(o)udvR>R0,\begin{split}\int_{B_{R}(o)}u^{-}\ \textnormal{dv}&\leq\frac{C}{\lambda R^{\gamma}+C}\int_{B_{\sigma R}(o)}u^{-}\ \textnormal{dv}\\ &=\frac{1}{\alpha R^{\gamma}+1}\int_{B_{\sigma R}(o)}u^{-}\ \textnormal{dv}\quad\quad\forall R>R_{0},\end{split}

where α=λC\alpha=\frac{\lambda}{C} depends on σ\sigma.  Similarly to what we done in Theorem 3.2, if we suppose that u0u^{-}\neq 0 almost everywhere in MM, then there exists AR0A\geq R_{0} so that

BA(o)udv>0.\displaystyle\int_{B_{A}(o)}u^{-}\ \textnormal{dv}>0.

By (4.8) we can apply Lemma 4.1 to the function f:[A,+)>0f:[A,+\infty)\to\mathbb{R}_{>0} given by

f:rBr(o)udv\displaystyle f:r\mapsto\int_{B_{r}(o)}u^{-}\ \textnormal{dv}

with β=1\beta=1, and we get

f(R)(RA)logσ(αAγ+1)f(A)αAγ+1\displaystyle f(R)\geq\left(\frac{R}{A}\right)^{\log_{\sigma}(\alpha A^{\gamma}+1)}\frac{f(A)}{\alpha A^{\gamma}+1}

for every R>AR>A. Choosing AR0A\geq R_{0} big enough so that

logσ(αAγ+1)k+1\displaystyle\log_{\sigma}(\alpha A^{\gamma}+1)\geq k+1

we have

f(R)(RA)k+1f(A)αAγ+1\displaystyle f(R)\geq\left(\frac{R}{A}\right)^{k+1}\frac{f(A)}{\alpha A^{\gamma}+1}

for every R>AR>A, thus obtaining a contradiction to (4.7). Hence u=0u^{-}=0 almost everywhere, implying the claim. ∎

As showed by D. Bianchi and A.G. Setti in [1, Corollary 2.3], a sufficient condition for the existence of a family {ϕR}R\{\phi_{R}\}_{R} satisfying (4.6) is a sub-quadratic decay of the Ricci curvature. Whence, we get the following corollary.

Corollary 4.3.

Let (M,g)(M,g) be a complete Riemannian manifold of dimension mm and λ\lambda a positive constant. Consider oMo\in M and assume that

Ricg(m1)C2(1+r2)η,\displaystyle\operatorname{Ric}_{g}\geq-(m-1)C^{2}(1+r^{2})^{\eta},

where CC is a positive constant, η[1,1)\eta\in[-1,1) and r(x):=d(x,o)r(x):=d(x,o) is the intrinsic distance from oo in MM. If uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and, for some kk\in\mathbb{N},

BR(o)udv=O(Rk)asR+,\displaystyle\int_{B_{R}(o)}u^{-}\ \textnormal{dv}=O(R^{k})\quad as\ R\to+\infty,

then u0u\geq 0 almost everywhere in MM.

4.2. Cut-off functions with equibounded laplacians

The second theorem of this section is an Lloc1L^{1}_{loc} positivity preserving property based on the existence of a family of cut-off functions with equibounded laplacians. The structure of the proof is very similar to the one adopted for Theorem 4.2 and it makes use of the following iterative lemma.

Lemma 4.4.

Let A>0A>0 and f:[A,+)(0,+)f:[A,+\infty)\to(0,+\infty) be a nondecreasing function. Suppose there exist α>1\alpha>1 and σ>1\sigma>1 so that

(4.9) f(r)1αf(σr)\displaystyle f(r)\leq\frac{1}{\alpha}f(\sigma r)

for every rAr\geq A. Then, ff satisfies

f(R)f(A)(RAσ)θ\displaystyle f(R)\geq f(A)\left(\frac{R}{A\sigma}\right)^{\theta}

for every R>AR>A, where θ=ln(α)ln(σ)>0\theta=\frac{\ln(\alpha)}{\ln(\sigma)}>0.

Proof.

Iterating (4.9), for every nn\in\mathbb{N} we get

f(r)1αnf(σnr)\displaystyle f(r)\leq\frac{1}{\alpha^{n}}f(\sigma^{n}r)

for any rAr\geq A. It follows that for any R>AR>A

f(R)f(Aσn)αnf(A)αlogσ(RAσ)f(A)=f(A)(RAσ)ln(α)ln(σ),\displaystyle f(R)\geq f(A\sigma^{n})\geq\alpha^{n}f(A)\geq\alpha^{\log_{\sigma}\left(\frac{R}{A\sigma}\right)}f(A)=f(A)\left(\frac{R}{A\sigma}\right)^{\frac{\ln(\alpha)}{\ln(\sigma)}},

where n=n(R,A,σ)n=n(R,A,\sigma) is the unique natural number satisfying σn+1RAσn\sigma^{n+1}\geq\frac{R}{A}\geq\sigma^{n}. This concludes the proof. ∎

We can now state our second main theorem that involves functions with an L1L^{1}-controlled growth.

Theorem 4.5 (Generalized Lloc1L^{1}_{loc} positivity preserving property).

Let (M,g)(M,g) be a complete Riemannian manifold and λ\lambda a positive constant. Assume that for a fixed oMo\in M there exist some positive constants CC and R0R_{0} and a constant σ>1\sigma>1 satisfying the following condition: for every R>R0R>R_{0} there exists ϕRCc2(M)\phi_{R}\in C^{2}_{c}(M) such that

(4.14) {0ϕR1inMϕR1inBR(o)supp(ϕR)BσR(o)|ΔϕR|CinM.\displaystyle\left\{\begin{array}[]{ll}0\leq\phi_{R}\leq 1&\textnormal{in}\ M\\ \phi_{R}\equiv 1&\textnormal{in}\ B_{R}(o)\\ \textnormal{supp}(\phi_{R})\subset B_{\sigma R}(o)\\ |\Delta\phi_{R}|\leq C&\textnormal{in}\ M.\end{array}\right.

If uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and

(4.15) BR(o)udv=o(Rθ)asR+\displaystyle\int_{B_{R}(o)}u^{-}\ \textnormal{dv}=o(R^{\theta})\quad as\ R\to+\infty

with θ=ln(1+λC)ln(σ)\theta=\frac{\ln\left(1+\frac{\lambda}{C}\right)}{\ln(\sigma)}, then u0u\geq 0 almost everywhere in MM.

Proof.

As in the proof of Theorem 4.2, we get

λMuϕRdvMuΔϕRR>R0\displaystyle\lambda\int_{M}u^{-}\phi_{R}\ \textnormal{dv}\leq\int_{M}u^{-}\Delta\phi_{R}\quad\quad\forall R>R_{0}

that implies

(4.16) BR(o)udvCλ+CBσR(o)udvR>R0.\displaystyle\int_{B_{R}(o)}u^{-}\ \textnormal{dv}\leq\frac{C}{\lambda+C}\int_{B_{\sigma R}(o)}u^{-}\ \textnormal{dv}\quad\quad\forall R>R_{0}.

If u0u^{-}\neq 0 almost everywhere in MM, then there exists A>0A>0 such that

BA(o)udv>0.\displaystyle\int_{B_{A}(o)}u^{-}\ \textnormal{dv}>0.

By (4.16) we can apply Lemma 4.4 with

f:rBr(o)udv\displaystyle f:r\mapsto\int_{B_{r}(o)}u^{-}\ \textnormal{dv}

and α=C+λC\alpha=\frac{C+\lambda}{C} and we deduce that f(R)C0Rln(α)ln(σ)f(R)\geq C_{0}R^{\frac{\ln(\alpha)}{\ln(\sigma)}} for every R>AR>A, where C0>0C_{0}>0. This contradicts (4.15). Hence u=0u^{-}=0 almost everywhere in MM, as required. ∎

In the proof of [5, Corollary 4.1], the authors obtained a family of cutoff functions satisfying (4.14) under the only assumption of a lower bound on the Ricci curvature. As a consequence, we obtain the following

Corollary 4.6.

Let (M,g)(M,g) be a complete Riemannian manifold and λ\lambda a positive constant. Consider oMo\in M and assume that

Ricg(x)G2(r(x))\displaystyle\operatorname{Ric}_{g}(x)\geq-G^{2}(r(x))

for every xMBR(o)x\in M\setminus B_{R}(o), where GCG\in C^{\infty} is given by

G(t)=αt0jkln[j](t)\displaystyle G(t)=\alpha t\prod_{0\leq j\leq k}\ln^{[j]}(t)

for t>1t>1, α>0\alpha>0 and kk\in\mathbb{N}. Then, there exists a constant θ=θ(λ,M,α,k)>0\theta=\theta(\lambda,M,\alpha,k)>0 such that if uLloc1(M)u\in L^{1}_{loc}(M) satisfies Δu+λu0-\Delta u+\lambda u\geq 0 in the sense of distributions and

BR(o)udv=o(Rθ)asR+,\displaystyle\int_{B_{R}(o)}u^{-}\ \textnormal{dv}=o(R^{\theta})\quad as\ R\to+\infty,

then u0u\geq 0 almost everywhere in MM. In particular, the L1L^{1} positivity preserving property holds true in the family of functions

{uLloc1(M):uL1(M)}.\displaystyle\{u\in L^{1}_{loc}(M)\ :\ u^{-}\in L^{1}(M)\}.
Remark 4.7.

In [2, Theorem B], the authors constructed a counterexample to the L1L^{1} positivity preserving property in a complete 2-manifold having Gaussian curvature with an asymptotic of the form K(x)Cr(x)2+ϵK(x)\sim-Cr(x)^{2+\epsilon}, for ϵ>0\epsilon>0. This underline that the result contained in Corollary 4.6 is sharp.

Remark 4.8.

When stated in terms of a Liouville type property, our Corollary 4.6 compares e.g. with [11, Theorem C], where the authors consider the case λ=0\lambda=0 of subharmonic functions. Their result states that a C1C^{1}, nonnegative subharmonic function with precise pointwises exponential control and a logarithmic L1L^{1} growth must be constant. They also provide a rotationally symmetric example (M,g)(M,g) with Gaussian curvature K(x)Cr(x)2K(x)\sim-Cr(x)^{2} showing that, without the pointwise control, there exists an unbounded smooth solution of Δu=1\Delta u=1 of logarithmic L1L^{1}-growth. As a consequence, keeping the curvature restriction of Corollary 4.6, in order to obtain the Liouville result under a pure L1L^{1}-growth condition, which is even faster than logarithmic, one has to assume that λ>0\lambda>0.

5. An application to minimal submanifolds

Recall that an immersed submanifold x:Σnmx:\Sigma^{n}\hookrightarrow\mathbb{R}^{m} is said to be minimal if its mean curvature vector field satisfies HΣ=0H^{\Sigma}=0. It is a standard fact that the minimality condition is equivalent to the property that the coordinate functions of the isometric immersion are harmonic, i.e.,

ΔΣxi=0i=1,,m.\displaystyle\Delta^{\Sigma}\,x_{i}=0\quad\forall i=1,...,m.

In particular, this implies that for any minimal submanifold in Euclidean space,

ΔΣ|x|2=2n.\displaystyle\Delta^{\Sigma}\,|x|^{2}=2n.

As an application of the main results in Section 3, we prove that complete minimal submanifold enjoy the following LpL^{p} extrinsic distance growth condition.

Corollary 5.1.

Let x:Σmx:\Sigma\hookrightarrow\mathbb{R}^{m} be a complete minimal submanifold and suppose there exists a positive function ξ:0>0\xi:\mathbb{R}_{\geq 0}\to\mathbb{R}_{>0} such that

(dm(x,o))2ξ(dΣ(x,o))andξ(R)=O(R2ϵ),as R+\displaystyle(d^{\mathbb{R}^{m}}(x,o))^{2}\leq\xi(d^{\Sigma}(x,o))\qquad\text{and}\qquad\xi(R)=O(R^{2-\epsilon}),\,\text{as }R\to+\infty

for some constants C>0C>0 and ϵ(0,2]\epsilon\in(0,2] and for some fixed origin oΣo\in\Sigma. Then, for every p(1,+)p\in(1,+\infty),

(5.1) lim supR+BRΣ(o)ξpdvΣeθRϵ2>0,\displaystyle\limsup_{R\to+\infty}\frac{\int_{B_{R}^{\Sigma}(o)}\xi^{p}\ \textnormal{dv}_{\Sigma}}{e^{\theta R^{\frac{\epsilon}{2}}}}>0,

where θ=(p1)Ce1\theta=\sqrt{\frac{(p-1)C}{e-1}}.

Proof.

Without loss of generality we can suppose o=0mo=0\in\mathbb{R}^{m}. Let

w(x):=dm(x,o)=|x|2w(x):=d^{\mathbb{R}^{m}}(x,o)=|x|^{2}

and define

λ(x):=2nξ(dΣ(x,o)).\lambda(x):=\frac{2n}{\xi(d^{\Sigma}(x,o))}.

Then

ΔΣw=2n=λξλw.\displaystyle\Delta^{\Sigma}\,w=2n=\lambda\xi\geq\lambda w.

By contradiction, suppose that (5.1) is not satisfied for some p(1,+)p\in(1,+\infty). Then

0=lim supR+BRΣ(o)ξpdvΣeθRϵ2lim supR+BRΣ(o)wpdvΣeθRϵ20,\displaystyle 0=\limsup_{R\to+\infty}\frac{\int_{B_{R}^{\Sigma}(o)}\xi^{p}\ \textnormal{dv}_{\Sigma}}{e^{\theta R^{\frac{\epsilon}{2}}}}\geq\limsup_{R\to+\infty}\frac{\int_{B_{R}^{\Sigma}(o)}w^{p}\ \textnormal{dv}_{\Sigma}}{e^{\theta R^{\frac{\epsilon}{2}}}}\geq 0,

showing that

BRΣ(o)wpdvΣ=o(eθRϵ2),R+.\int_{B^{\Sigma}_{R}(o)}w^{p}\ \textnormal{dv}_{\Sigma}=o(e^{\theta R^{\frac{\epsilon}{2}}}),\,R\to+\infty.

An application of Theorem 3.2, in the form of a Liouville type result, yields that w0w\equiv 0. Contradiction. ∎

Remark 5.2.

In the assumption of Corollary 5.1 we get an asymptotic estimate on the behavior of |BRΣ||B_{R}^{\Sigma}|. Indeed, since there exist two constants C>0C>0 and ϵ(0,2]\epsilon\in(0,2] such that

ξ(x)C(1+dΣ(x,o))2ϵ,\displaystyle\xi(x)\leq C\left(1+d^{\Sigma}(x,o)\right)^{2-\epsilon},

then

BRΣ(o)ξpdvCBRΣ(o)(1+dΣ(x,o))(2ϵ)pdvC(1+R)(2ϵ)p|BRΣ(o)|.\displaystyle\int_{B_{R}^{\Sigma}(o)}\xi^{p}\ \textnormal{dv}\leq C\int_{B_{R}^{\Sigma}(o)}(1+d^{\Sigma}(x,o))^{(2-\epsilon)p}\ \textnormal{dv}\leq C(1+R)^{(2-\epsilon)p}|B_{R}^{\Sigma}(o)|.

By (5.1) it follows

lim supR+(1+R)(2ϵ)p|BRΣ(o)|eθRϵ2>0.\displaystyle\limsup_{R\to+\infty}\frac{(1+R)^{(2-\epsilon)p}|B_{R}^{\Sigma}(o)|}{e^{\theta R^{\frac{\epsilon}{2}}}}>0.

Whence, we obtain the validity of the following nonexistence result.

Corollary 5.3.

There are no complete minimal submanifolds Σnm\Sigma^{n}\hookrightarrow\mathbb{R}^{m} satisfying the following conditions:

  1. a)

    the extrinsic distance from a fixed origin oΣo\in\Sigma satisfies

    (dm(x,o))2ξ(dΣ(x,o))\displaystyle\left(d^{\mathbb{R}^{m}}(x,o)\right)^{2}\leq\xi(d^{\Sigma}(x,o))

    with

    ξ(R)=O(R2ϵ)as R+\displaystyle\xi(R)=O(R^{2-\epsilon})\quad\text{as }R\to+\infty

    for some ϵ(0,2]\epsilon\in(0,2];

  2. b)

    the intrinsic geodesic balls of Σ\Sigma centered at oo satisfy the asymptotic estimate

    |BRΣ(o)|=o(R(2ϵ)peθRϵ2) as R+,|B_{R}^{\Sigma}(o)|=o\left(R^{-(2-\epsilon)p}e^{\theta R^{\frac{\epsilon}{2}}}\right)\,\text{ as }R\to+\infty,

    with θ=(p1)Ce1\theta=\sqrt{\frac{(p-1)C}{e-1}} and p(1,+)p\in(1,+\infty).

Remark 5.4.

We stress that in case ϵ=2\epsilon=2, i.e. for bounded minimal submanifolds, the volume growth we obtained is far from being optimal. Indeed, in [6] and [8] the authors achieved the rate |BRΣ(o)|=O(eCR2)|B_{R}^{\Sigma}(o)|=O(e^{CR^{2}}). This discrepancy comes from the fact that we use integral techniques and estimates.

References

  • [1] D. Bianchi and A. G. Setti. Laplacian cut-offs, porous and fast diffusion on manifolds and other applications. Calculus of Variations and Partial Differential Equations, 57(1):1–33, 2018.
  • [2] A. Bisterzo and L. Marini. The L{L}^{\infty}-positivity Preserving Property and Stochastic Completeness. Potential Analysis, pages 1–18, 2022.
  • [3] M. Braverman, O. Milatovic, and M. Shubin. Essential self-adjointness of Schrödinger-type operators on manifolds. Russian Mathematical Surveys, 57(4):641, 2002.
  • [4] B. Güneysu, S. Pigola, P. Stollmann, and G. Veronelli. A new notion of subharmonicity on locally smoothing spaces, and a conjecture by Braverman, Milatovic, Shubin. arXiv preprint arXiv:2302.09423, 2023.
  • [5] D. Impera, M. Rimoldi, and G. Veronelli. Higher order distance-like functions and Sobolev spaces. Advances in Mathematics, 396:108166, 2022.
  • [6] L. Karp. Differential inequalities on complete Riemannian manifolds and applications. Math. Ann., 272(4):449–459, 1985.
  • [7] L. Mari, M. Rigoli, and A. G. Setti. Keller–Osserman conditions for diffusion-type operators on Riemannian manifolds. Journal of Functional Analysis, 258(2):665–712, 2010.
  • [8] S. Pigola, M. Rigoli, and A. G. Setti. A remark on the maximum principle and stochastic completeness. Proc. Amer. Math. Soc., 131(4):1283–1288, 2003.
  • [9] S. Pigola, D. Valtorta, and G. Veronelli. Approximation, regularity and positivity preservation on Riemannian manifolds. arXiv preprint arXiv:2301.05159, 2023.
  • [10] S. Pigola and G. Veronelli. Lp{L}^{p} positivity preserving and a conjecture by M. Braverman, O. Milatovic and M. Shubin. arXiv preprint arXiv:2105.14847, 2021.
  • [11] M. Rigoli and A. G. Setti. Liouville type theorems for φ\varphi-subharmonic functions. Revista Matematica Iberoamericana, 17(3):471–520, 2001.