1 Introduction

As a fundamental problem in mathematical physics, self-adjointness of Schrödinger operators has attracted the attention of researchers over many years now, resulting in numerous sufficient conditions for this property in \(L^2(\mathbb {R}^{n})\). For reviews of the corresponding results, see, for instance, the books [14, 28].

The study of the corresponding problem in the context of a non-compact Riemannian manifold was initiated by Gaffney [15, 16] with the proof of the essential self-adjointness of the Laplacian on differential forms. About two decades later, Cordes (see Theorem 3 in  [11]) proved the essential self-adjointness of positive integer powers of the operator

$$\begin{aligned} \Delta _{M,\mu }:= \ - \ \frac{1}{\kappa }\left( \frac{\partial }{\partial x^{i}}\left( \kappa g^{ij}\frac{\partial }{\partial x^{j}}\right) \right) \end{aligned}$$
(1.1)

on an n-dimensional geodesically complete Riemannian manifold M equipped with a (smooth) metric \(g=(g_{ij})\) [here \((g^{ij})=((g_{ij})^{-1})\)] and a positive smooth measure \(\mathrm{d}\mu \) [i.e. in any local coordinates \(x^{1},\,x^{2},\dots ,x^{n}\) there exists a strictly positive \(C^{\infty }\)-density \(\kappa (x)\) such that \(\mathrm{d}\mu =\kappa (x)\,\mathrm{d}x^{1}\mathrm{d}x^{2}\dots \mathrm{d}x^{n}\)]. Theorem 1 of our paper extends this result to the operator \((D^*D+V)^{k}\) for all \(k\in \mathbb {Z}_{+}\), where D is a first order elliptic differential operator acting on sections of a Hermitian vector bundle over a geodesically complete Riemannian manifold, \(D^{*}\) is the formal adjoint of D, and V is a self-adjoint Hermitian bundle endomorphism; see Sect. 2.2 for details.

In the context of a general Riemannian manifold (not necessarily geodesically complete), Cordes (see Theorem IV.1.1 in [12], Theorem 4 in [11]) proved the essential self-adjointness of \(P^{k}\) for all \(k\in \mathbb {Z}_{+}\), where

$$\begin{aligned} Pu:=\Delta _{M,\mu }u+qu, \quad u\in C^{\infty }(M), \end{aligned}$$
(1.2)

and \(q\in C^{\infty }(M)\) is real-valued. Thanks to a Roelcke-type estimate (see Lemma 3.1 below), the technique of Cordes [12] can be applied to the operator \((D^*D+V)^{k}\) acting on sections of Hermitian vector bundles over a general Riemannian manifold. To make our exposition shorter, in Theorem 1 we consider the geodesically complete case. Our Theorem 2 concerns \((\nabla ^*\nabla +V)^{k}\), where \(\nabla \) is a metric connection on a Hermitian vector bundle over a non-compact geodesically complete Riemannian manifold. This result extends Theorem 1.1 of [13] where Cordes showed that if (Mg) is non-compact and geodesically complete and P is semi-bounded from below on \(C_{c}^{\infty }(M)\), then \(P^{k}\) is essentially self-adjoint on \(C_{c}^{\infty }(M)\), for all \(k\in \mathbb {Z}_{+}\).

For the remainder of the introduction, the notation \(D^*D+V\) is used in the same sense as described earlier in this section. In the setting of geodesically complete Riemannian manifolds, the essential self-adjointness of \(D^*D+V\) with \(V\in L^{\infty }_{\mathrm{loc}}\) was established in [20], providing a generalization of the results in [3, 26, 27, 31] concerning Schrödinger operators on functions (or differential forms). Subsequently, the operator \(D^*D+V\) with a singular potential V was considered in [5]. Recently, in the case \(V\in L^{\infty }_{\mathrm{loc}}\), the authors of [4] extended the main result of [5] to the operator \(D^*D+V\) acting on sections of infinite-dimensional bundles whose fibers are modules of finite type over a von Neumann algebra.

In the context of an incomplete Riemannian manifold, the authors of [17, 21, 22] studied the so-called Gaffney Laplacian, a self-adjoint realization of the scalar Laplacian generally different from the closure of \(\Delta _{M,d \mu }|_{C_{c}^{\infty }(M)}\). For a study of Gaffney Laplacian on differential forms, see [23].

Our Theorem 3 gives a condition on the behavior of V relative to the Cauchy boundary of M that will guarantee the essential self-adjointness of \(D^*D+V\); for details see Sect. 2.3 below. Related results can be found in  [6, 24, 25] in the context of (magnetic) Schrödinger operators on domains in \({\mathbb {R}}^n\), and in [10] concerning the magnetic Laplacian on domains in \({\mathbb {R}}^n\) and certain types of Riemannian manifolds.

Finally, let us mention that Chernoff [7] used the hyperbolic equation approach to establish the essential self-adjointness of positive integer powers of Laplace–Beltrami operator on differential forms. This approach was also applied in  [2, 8, 9, 18, 19, 30] to prove essential self-adjointness of second-order operators (acting on scalar functions or sections of Hermitian vector bundles) on Riemannian manifolds. Additionally, the authors of [18, 19] used path integral techniques.

The paper is organized as follows. The main results are stated in Sect. 2, a preliminary lemma is proven in Sect. 3, and the main results are proven in Sects. 46.

2 Main results

2.1 The setting

Let M be an n-dimensional smooth, connected Riemannian manifold without boundary. We denote the Riemannian metric on M by \(g^{\mathrm{TM}}\). We assume that M is equipped with a positive smooth measure \(\mathrm{d}\mu \), i.e. in any local coordinates \(x^{1}, \,x^{2},\dots ,x^{n}\) there exists a strictly positive \(C^{\infty }\)-density \(\kappa (x)\) such that \(\mathrm{d}\mu =\kappa (x)\,\mathrm{d}x^{1}\mathrm{d}x^{2}\dots \mathrm{d}x^{n}\). Let E be a Hermitian vector bundle over M and let \(L^2(E)\) denote the Hilbert space of square integrable sections of E with respect to the inner product

$$\begin{aligned} (u,v) \ = \ \int _{M}\, \langle u(x),v(x)\rangle _{E_{x}}\, \mathrm{d}\mu (x), \end{aligned}$$
(2.1)

where \(\langle \cdot ,\cdot \rangle _{E_{x}}\) is the fiberwise inner product. The corresponding norm in \(L^2(E)\) is denoted by \(\Vert \cdot \Vert \). In Sobolev space notations \(W^{k,2}_{\mathrm{loc}}(E)\) used in this paper, the superscript \(k\in \mathbb {Z}_{+}\) indicates the order of the highest derivative. The corresponding dual space is denoted by \(W^{-k,2}_{\mathrm{loc}}(E)\).

Let F be another Hermitian vector bundle on M. We consider a first order differential operator \(D:C_{c}^{\infty }(E)\rightarrow C_{c}^{\infty }(F)\), where \(C^\infty _c\) stands for the space of smooth compactly supported sections. In the sequel, by \(\sigma (D)\) we denote the principal symbol of D.

Assumption (A0) Assume that D is elliptic. Additionally, assume that there exists a constant \(\lambda _0>0\) such that

$$\begin{aligned} |\sigma (D)(x,\xi )| \ \le \ \lambda _0|\xi |,\quad \text {for all }x\in M, \, \xi \in T_{x}^{*}M, \end{aligned}$$
(2.2)

where \(|\xi |\) is the length of \(\xi \) induced by the metric \(g^{\mathrm{TM}}\) and \(|\sigma (D)(x,\xi )|\) is the operator norm of \(\sigma (D)(x,\xi ):E_x\rightarrow F_x\).

Remark 2.1

Assumption (A0) is satisfied if \(D=\nabla \), where \(\nabla :C^\infty (E)\rightarrow C^{\infty }(T^*M\otimes E)\) is a covariant derivative corresponding to a metric connection on a Hermitian vector bundle E over M.

2.2 Schrödinger-type operator

Let \(D^*:C_{c}^{\infty }(F)\rightarrow C_{c}^{\infty }(E)\) be the formal adjoint of D with respect to the inner product (2.1). We consider the operator

$$\begin{aligned} H \ = \ D^*D \ + \ V, \end{aligned}$$
(2.3)

where \(V\in L^{\infty }_{\mathrm{loc}}(\mathrm{End}{E})\) is a linear self-adjoint bundle endomorphism. In other words, for all \(x\in M\), the operator \(V(x):E_{x}\rightarrow E_{x}\) is self-adjoint and \(|V(x)|\in L^{\infty }_{\mathrm{loc}}(M)\), where |V(x)| is the norm of the operator \(V(x):E_{x}\rightarrow E_{x}\).

2.3 Statements of results

Theorem 1

Let M, \(g^{\mathrm{TM}}\), and \(\mathrm{d}\mu \) be as in Sect. 2.1. Assume that \((M,g^{\mathrm{TM}})\) is geodesically complete. Let E and F be Hermitian vector bundles over M, and let \(D:C_{c}^{\infty }(E)\rightarrow C_{c}^{\infty }(F)\) be a first order differential operator satisfying the Assumption (A0). Assume that \(V\in C^{\infty }(\mathrm{End}{E})\) and

$$\begin{aligned} V(x)\ge C,\quad \text {for all }x\in M, \end{aligned}$$

where C is a constant, and the inequality is understood in operator sense. Then \(H^{k}\) is essentially self-adjoint on \(C_{c}^{\infty }(E)\), for all \(k\in \mathbb {Z}_{+}\).

Remark 2.2

In the case \(V=0\), the following result related to Theorem 1 can be deduced from Chernoff (see Theorem 2.2 in [7]):

Assume that (Mg) is a geodesically complete Riemannian manifold with metric g. Let D be as in Theorem 1, and define

$$\begin{aligned} c(x) \ := \ \sup \{|\sigma (D)(x,\xi )|:\, |\xi |_{T_{x}^{*}M}=1\}. \end{aligned}$$

Fix \(x_0\in M\) and define

$$\begin{aligned} \widetilde{c}(r):=\sup _{x\in B(x_0,r)}c(x), \end{aligned}$$

where \(r>0\) and \(B(x_0,r):=\{x\in M:\mathrm{d}_{g}(x_0,x)<r\}\). Assume that

$$\begin{aligned} \int _{0}^{\infty }\frac{1}{\widetilde{c}(r)}\,\mathrm{d}r=\infty . \end{aligned}$$
(2.4)

Then the operator \((D^*D)^{k}\) is essentially self-adjoint on \(C_{c}^{\infty }(E)\) for all \(k\in \mathbb {Z}_{+}\).

At the end of this section we give an example of an operator for which Theorem 1 guarantees the essential self-adjointness of \((D^*D)^{k}\), whereas Chernoff’s result cannot be applied.

The next theorem is concerned with operators whose potential V is not necessarily semi-bounded from below.

Theorem 2

Let M, \(g^{\mathrm{TM}}\), and \(\mathrm{d}\mu \) be as in Sect. 2.1. Assume that \((M,g^{\mathrm{TM}})\) is noncompact and geodesically complete. Let E be a Hermitian vector bundle over M and let \(\nabla \) be a Hermitian connection on E. Assume that \(V\in C^{\infty }(\mathrm{End}{E})\) and

$$\begin{aligned} V(x)\ge q(x),\quad \text {for all }x\in M, \end{aligned}$$
(2.5)

where \(q\in C^{\infty }(M)\) and the inequality is understood in the sense of operators \(E_x\rightarrow E_x\). Additionally, assume that

$$\begin{aligned} ((\Delta _{M,\mu }+q)u,u)\ge C\Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(M), \end{aligned}$$
(2.6)

where \(C\in \mathbb {R}\) and \(\Delta _{M,\mu }\) is as in (1.1) with g replaced by \(g^{\mathrm{TM}}\). Then the operator \((\nabla ^*\nabla +V)^{k}\) is essentially self-adjoint on \(C_{c}^{\infty }(E)\), for all \(k\in \mathbb {Z}_{+}\).

Remark 2.3

Let us stress that non-compactness is required in the proof to ensure the existence of a positive smooth solution of an equation involving \(\Delta _{M,\mu }+q\). In the case of a compact manifold, such a solution exists under an additional assumption; see Theorem III.6.3 in [12].

In our last result we will need the notion of Cauchy boundary. Let \(d_{g^{\mathrm{TM}}}\) be the distance function corresponding to the metric \(g^{\mathrm{TM}}\). Let \((\widehat{M}, \widehat{d}_{g^{\mathrm{TM}}})\) be the metric completion of \((M, d_{g^{\mathrm{TM}}})\). We define the Cauchy boundary \(\partial _{C}M\) as follows: \(\partial _{C}M:=\widehat{M}\backslash M\). Note that \((M,d_{g^{\mathrm{TM}}})\) is metrically complete if and only if \(\partial _{C}M\) is empty. For \(x\in M\) we define

$$\begin{aligned} r(x):=\inf _{z\in \partial _{C}M}\widehat{d}_{g^{\mathrm{TM}}}(x,z). \end{aligned}$$
(2.7)

We will also need the following assumption:

Assumption (A1) Assume that \(\widehat{M}\) is a smooth manifold and that the metric \(g^{\mathrm{TM}}\) extends to \(\partial _{C}M\).

Remark 2.4

Let N be a (smooth) n-dimensional Riemannian manifold without boundary. Denote the metric on N by \(g^{TN}\) and assume that \((N,g^{TN})\) is geodesically complete. Let \(\Sigma \) be a k-dimensional closed sub-manifold of N with \(k<n\). Then \(M:=N\backslash \Sigma \) has the properties \(\widehat{M}=N\) and \(\partial _{C}M=\Sigma \). Thus, Assumption (A1) is satisfied.

Theorem 3

Let M, \(g^{\mathrm{TM}}\), and \(\mathrm{d}\mu \) be as in Sect. 2.1. Assume that (A1) is satisfied. Let E and F be Hermitian vector bundles over M, and let \(D:C_{c}^{\infty }(E)\rightarrow C_{c}^{\infty }(F)\) be a first order differential operator satisfying the Assumption (A0). Assume that \(V\in L^{\infty }_{\mathrm{loc}}(\mathrm{End}{E})\) and there exists a constant C such that

$$\begin{aligned} V(x) \ge \left( \frac{\lambda _0}{r(x)}\right) ^2-C,\quad \text { for all }x\in M, \end{aligned}$$
(2.8)

where \(\lambda _0\) is as in (2.2), the distance r(x) is as in (2.7), and the inequality is understood in the sense of linear operators \(E_{x}\rightarrow E_{x}\). Then H is essentially self-adjoint on \(C_{c}^{\infty }(E)\).

In order to describe the example mentioned in Remark 2.2, we need the following

Remark 2.5

As explained in [5], we can use a first-order elliptic operator \(D:C_{c}^{\infty }(E)\rightarrow C_{c}^{\infty }(F)\) to define a metric on M. For \(\xi ,\eta \in T^*_xM\), define

$$\begin{aligned} \langle \xi ,\eta \rangle \ = \ \frac{1}{m}\, {\text {Re}}\, {\text {Tr}}\left( \left( \sigma (D)(x,\xi )\right) ^*\sigma (D)(x,\eta )\right) , \quad m= \dim E_x, \end{aligned}$$
(2.9)

where \({\text {Tr}}\) denotes the usual trace of a linear operator. Since D is an elliptic first-order differential operator and \(\sigma (D)(x,\xi )\) is linear in \(\xi \), it is easily checked that (2.9) defines an inner product on \(T^*_xM\). Its dual defines a Riemannian metric on M. Denoting this metric by \(g^{\mathrm{TM}}\) and using elementary linear algebra, it follows that (2.2) is satisfied with \(\lambda _0=\sqrt{m}\).

Example 2.6

Let \(M=\mathbb {R}^{2}\) with the standard metric and measure, and \(V=0\). Denoting respectively by \(C_{c}^{\infty }(\mathbb {R}^{2};\mathbb {R})\) and \(C_{c}^{\infty }(\mathbb {R}^{2};\mathbb {R}^2)\) the spaces of smooth compactly supported functions \(f:\mathbb {R}^{2}\rightarrow \mathbb {R}\) and \(f:\mathbb {R}^{2}\rightarrow \mathbb {R}^2\), we define the operator \(D:C_{c}^{\infty }(\mathbb {R}^{2};\mathbb {R})\rightarrow C_{c}^{\infty }(\mathbb {R}^{2};\mathbb {R}^2)\) by

$$\begin{aligned} D \ = \ \left( \begin{array}{c}a(x,y)\frac{\partial }{\partial x}\\ b(x,y)\frac{\partial }{\partial y}\\ \end{array}\right) , \end{aligned}$$

where

$$\begin{aligned} a(x,y)&= (1-\cos (2\pi e^{x}))x^{2}+1; \\ b(x,y)&= (1-\sin (2\pi e^{y}))y^{2}+1. \end{aligned}$$

Since ab are smooth real-valued nowhere vanishing functions in \(\mathbb {R}^2\), it follows that the operator D is elliptic. We are interested in the operator

$$\begin{aligned} H \ := \ D^*D \ = \ -\frac{\partial }{\partial x} \left( a^{2}\frac{\partial }{\partial x}\right) - \frac{\partial }{\partial y} \left( b^{2}\frac{\partial }{\partial y}\right) . \end{aligned}$$

The matrix of the inner product on \(T^*M\) defined by D via (2.9) is \(\mathrm{diag}(a^2/2,b^2/2)\). The matrix of the corresponding Riemannian metric \(g^{\mathrm{TM}}\) on M is \(\mathrm{diag}(2a^{-2},2b^{-2})\), so the metric itself is \(ds^2=2a^{-2}\mathrm{d}x^2+2b^{-2}dy^2\) and it is geodesically complete (see Example 3.1 of [5]). Moreover, thanks to Remark 2.5, Assumption (A0) is satisfied. Thus, by Theorem 1 the operator \((D^*D)^{k}\) is essentially self-adjoint for all \(k\in \mathbb {Z}_{+}\). Furthermore, in Example 3.1 of [5] it was shown that for the considered operator D the condition (2.4) is not satisfied. Thus, the result stated in Remark 2.2 does not apply.

3 Roelcke-type inequality

Let M, \(\mathrm{d}\mu \), D, and \(\sigma (D)\) be as in Sect. 2.1. Set \(\widehat{D}:=-i\sigma (D)\), where \(i=\sqrt{-1}\). Then for any Lipschitz function \(\psi :M\rightarrow \mathbb {R}\) and \(u\in W^{1,2}_{\mathrm{loc}}(E)\) we have

$$\begin{aligned} D(\psi u) \ = \ \widehat{D}(d\psi )u+\psi Du, \end{aligned}$$
(3.1)

where we have suppressed x for simplicity. We also note that \(\widehat{D^*}(\xi )=-(\widehat{D}(\xi ))^{*}\), for all \(\xi \in T_{x}^{*}M\).

For a compact set \(K\subset M\), and \(u,\, v\in W^{1,2}_{\mathrm{loc}}(E)\), we define

$$\begin{aligned} (u,v)_{K}:=\int _{K}\langle u(x),v(x)\rangle \,\mathrm{d}\mu (x), \quad (Du,Dv)_{K}:=\int _{K}\langle Du(x),Dv(x)\rangle \,\mathrm{d}\mu (x). \end{aligned}$$
(3.2)

In order to prove Theorem 1 we need the following important lemma, which is an extension of Lemma IV.2.1 in [12] to operator (2.3). In the context of the scalar Laplacian on a Riemannian manifold, this kind of result is originally due to Roelcke [29].

Lemma 3.1

Let M, \(g^{\mathrm{TM}}\), and \(\mathrm{d}\mu \) be as in Sect. 2.1. Let E and F be Hermitian vector bundles over M, and let \(D:C_{c}^{\infty }(E)\rightarrow C_{c}^{\infty }(F)\) be a first order differential operator satisfying the Assumption (A0). Let \(\rho :M\rightarrow [0,\infty )\) be a function satisfying the following properties:

  1. (i)

    \(\rho (x)\) is Lipschitz continuous with respect to the distance induced by the metric \(g^{\mathrm{TM}}\);

  2. (ii)

    \(\rho (x_0)=0\), for some fixed \(x_0\in M\);

  3. (iii)

    the set \(B_{T}:=\{x\in M:\rho (x)\le T\}\) is compact, for some \(T>0\).

Then the following inequality holds for all \(u\in W^{2,2}_{\mathrm{loc}}(E)\) and \(v\in W^{2,2}_{\mathrm{loc}}(E)\):

$$\begin{aligned} \int _{0}^{T} |(Du,Dv)_{B_t}-(D^*Du,v)_{B_t}|\,\mathrm{d}t\le \lambda _0\int _{B_{T}}|\mathrm{d}\rho (x)||Du(x)||v(x)|\,\mathrm{d}\mu (x), \end{aligned}$$
(3.3)

where \(B_{t}\) is as in (iii) (with t instead of T), the constant \(\lambda _0\) is as in (2.2), and \(|\mathrm{d}\rho (x)|\) is the length of \(\mathrm{d}\rho (x)\in T_{x}^*M\) induced by \(g^{\mathrm{TM}}\).

Proof

For \(\varepsilon >0\) and \(t\in (0,T)\), we define a continuous piecewise linear function \(F_{\varepsilon ,t}\) as follows:

$$\begin{aligned} F_{\varepsilon ,t}(s)= \left\{ \begin{array}{ll} 1 &{}\quad \mathrm{for~} s<t-\varepsilon \\ (t-s)/{\varepsilon }&{} \quad \mathrm{for~} t-\varepsilon \le s <t \\ 0 &{}\quad \mathrm{for~ } s \ge t. \end{array} \right. \end{aligned}$$

The function \(f_{\varepsilon ,t}(x):=F_{\varepsilon ,t}(\rho (x))\), is Lipschitz continuous with respect to the distance induced by the metric \(g^{\mathrm{TM}}\), and \(df_{\varepsilon ,t}(x)=(F'_{\varepsilon ,t}(\rho (x)))\mathrm{d}\rho (x)\). Moreover we have \(f_{\varepsilon ,t}v\in W^{1,2}_{\mathrm{loc}}(E)\) for all \(v\in W^{1,2}_{\mathrm{loc}}(E)\), since

$$\begin{aligned} D(f_{\varepsilon ,t}v)=\widehat{D}(df_{\varepsilon ,t})v+f_{\varepsilon ,t}Dv. \end{aligned}$$

It follows from the compactness of \(B_{T}\) that \(B_{t}\) is compact for all \(t\in (0,T)\). Using integration by parts (see Lemma 8.8 in [5]), for all \(u\in W^{2,2}_{\mathrm{loc}}(E)\) and \(v\in W^{2,2}_{\mathrm{loc}}(E)\) we have

$$\begin{aligned} (D^*Du, vf_{\varepsilon ,t})_{B_{t}} = (Du, D(vf_{\varepsilon ,t}))_{B_{t}} = (Du,f_{\varepsilon ,t}Dv)_{B_{t}} + (Du,\widehat{D}(df_{\varepsilon ,t})v)_{B_{t}}, \end{aligned}$$

which, together with (2.2), gives

$$\begin{aligned}&|(Du,f_{\varepsilon ,t}Dv)_{B_{t}}-(D^*Du, vf_{\varepsilon ,t})_{B_{t}}|=|(Du,\widehat{D}(df_{\varepsilon ,t})v)_{B_{t}}|\nonumber \\&\quad \le \int _{B_{t}}|Du(x)||\widehat{D}(df_{\varepsilon ,t}(x))v(x)|\,\mathrm{d}\mu (x) \le \lambda _0 \int _{B_{t}}|Du(x)||df_{\varepsilon ,t}(x)||v(x)|\,\mathrm{d}\mu (x)\nonumber \\&\quad = \lambda _0\int _{B_{t}}|Du(x)||F'_{\varepsilon ,t}(\rho (x))||\mathrm{d}\rho (x)||v(x)|\,\mathrm{d}\mu (x)\nonumber \\&\quad \le \lambda _0\int _{B_{T}}|Du(x)||F'_{\varepsilon ,t}(\rho (x))||\mathrm{d}\rho (x)||v(x)|\,\mathrm{d}\mu (x), \end{aligned}$$
(3.4)

where \(|df_{\varepsilon ,t}(x)|\) and \(|\mathrm{d}\rho (x)|\) are the norms of \(df_{\varepsilon ,t}(x)\in T_{x}^*M\) and \(\mathrm{d}\rho (x)\in T_{x}^*M\) induced by \(g^{\mathrm{TM}}\).

Fixing \(\varepsilon >0\), integrating the leftmost and the rightmost side of (3.4) from \(t=0\) to \(t=T\), and noting that \(F'_{\varepsilon ,t}(\rho (x))\) is the only term on the rightmost side depending on t, we obtain

$$\begin{aligned}&\int _{0}^{T}|(Du,f_{\varepsilon ,t}Dv)_{B_{t}}-(D^*Du, vf_{\varepsilon ,t})_{B_{t}}|\,\mathrm{d}t\nonumber \\&\quad \le \lambda _0\int _{B_{T}}|Du(x)||\mathrm{d}\rho (x)||v(x)|I_{\varepsilon }(x)\,\mathrm{d}\mu (x), \end{aligned}$$
(3.5)

where

$$\begin{aligned} I_{\varepsilon }(x):=\int _{0}^{T}|F'_{\varepsilon ,t}(\rho (x))|\,\mathrm{d}t. \end{aligned}$$

We now let \(\varepsilon \rightarrow 0+\) in (3.5). On the left-hand side of (3.5), as \(\varepsilon \rightarrow 0+\), we have \(f_{\varepsilon ,t}(x)\rightarrow \chi _{B_t}(x)\) almost everywhere, where \(\chi _{B_t}(x)\) is the characteristic function of the set \(B_t\). Additionally, \(|f_{\varepsilon ,t}(x)|\le 1\) for all \(x\in B_t\) and all \(t\in (0,T)\); thus, by dominated convergence theorem, as \(\varepsilon \rightarrow 0+\) the left-hand side of (3.5) converges to the left-hand side of (3.3). On the right-hand side of (3.5) an easy calculation shows that \(I_{\varepsilon }(x)\rightarrow 1\), as \(\varepsilon \rightarrow 0+\). Additionally, we have \(|I_{\varepsilon }(x)|\le 1\), a.e. on \(B_{T}\); hence, by the dominated convergence theorem, as \(\varepsilon \rightarrow 0+\) the right-hand side of (3.5) converges to the right-hand side of (3.3). This establishes the inequality (3.3). \(\square \)

4 Proof of Theorem 1

We first give the definitions of minimal and maximal operators associated with the expression H in (2.3).

4.1 Minimal and maximal operators

We define \(H_{\min }u:=Hu\), with \({\text {Dom}}(H_{\min }):=C_{c}^{\infty }(E)\), and \(H_{\max }:=(H_{\min })^{*}\), where \(T^*\) denotes the adjoint of operator T. Denoting \(\mathscr {D}_{\max }:=\{u\in L^2(E):Hu\in L^2(E)\}\), we recall the following well-known property: \({\text {Dom}}(H_{\max })=\mathscr {D}_{\max }\) and \(H_{\max }u=Hu\) for all \(u\in \mathscr {D}_{\max }\).

From now on, throughout this section, we assume that the hypotheses of Theorem 1 are satisfied. Let \(x_0\in M\), and define \(\rho (x):=d_{g^{\mathrm{TM}}}(x_0,x)\), where \(d_{g^{\mathrm{TM}}}\) is the distance function corresponding to the metric \(g^{\mathrm{TM}}\). By the definition of \(\rho (x)\) and the geodesic completeness of \((M, g^{\mathrm{TM}})\), it follows that \(\rho (x)\) satisfies all hypotheses of Lemma 3.1. Using Lemma 3.1 and Proposition 4.1 below, we are able to apply the method of Cordes [11, 12] to our context. As we will see, Cordes’s technique reduces our problem to a system of ordinary differential inequalities of the same type as in Section IV.3 of [12].

Proposition 4.1

Let A be a densely defined operator with domain \(\mathscr {D}\) in a Hilbert space \(\mathscr {H}\). Assume that A is semi-bounded from below, that \(A\mathscr {D}\subseteq \mathscr {D}\), and that there exists \(c_0\in \mathbb {R}\) such that the following two properties hold:

  1. (i)

    \(((A+c_0I)u,u)_{\mathscr {H}}\ge \Vert u\Vert _{\mathscr {H}}^2\), for all \(u\in \mathscr {D}\), where I denotes the identity operator in \(\mathscr {H}\);

  2. (ii)

    \((A+c_0I)^{k}\) is essentially self-adjoint on \(\mathscr {D}\), for some \(k\in \mathbb {Z}_{+}\).

Then, \((A+cI)^{j}\) is essentially self-adjoint on \(\mathscr {D}\), for all \(j=1,2,\dots , k\) and all \(c\in \mathbb {R}\).

Remark 4.2

To prove Proposition 4.1, one may mimick the proof of Proposition IV.1.4 in [12], which was carried out for the operator P defined in (1.2) with \(\mathscr {D}=C_{c}^{\infty }(M)\), since only abstract functional analysis facts and the property \(P\mathscr {D}\subseteq \mathscr {D}\) were used.

We start the proof of Theorem 1 by noticing that the operator \(H_{\min }\) is essentially self-adjoint on \(C_{c}^{\infty }(E)\); see Corollary 2.9 in [5]. Thanks to Proposition 4.1, whithout any loss of generality we can change V(x) to \( V(x) + C\,\text {Id}(x)\) , where C is a sufficiently large constant in order to have

$$\begin{aligned} V(x)\ge \,(\lambda _0^2+1)\text {Id}(x),\quad \text {for all }x\in M, \end{aligned}$$
(4.1)

where \(\lambda _0\) is as in (2.2) and \(\text {Id}(x)\) is the identity endomorphism of \(E_{x}\). Using non-negativity of \(D^*D\) and (4.1) we have

$$\begin{aligned} (H_{\min }u,u)\ge \Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(E), \end{aligned}$$
(4.2)

which leads to

$$\begin{aligned} \Vert u\Vert ^2\le (Hu,u)\le \Vert Hu\Vert \Vert u\Vert ,\quad \text {for all }u\in C_{c}^{\infty }(E), \end{aligned}$$

and, hence, \(\Vert Hu\Vert \ge \Vert u\Vert \), for all \(u\in C_{c}^{\infty }(E)\). Therefore,

$$\begin{aligned} (H^2u,u)=(Hu,Hu)=\Vert Hu\Vert ^2\ge \Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(E), \end{aligned}$$
(4.3)

and

$$\begin{aligned} (H^3u,u)=(HHu,Hu)\ge \Vert Hu\Vert ^2\ge \Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(E). \end{aligned}$$

By (4.3) we have

$$\begin{aligned} \Vert u\Vert ^2\le (H^2u,u)\le \Vert H^2u\Vert \Vert u\Vert , \quad \text {for all }u\in C_{c}^{\infty }(E), \end{aligned}$$

and, hence, \(\Vert H^2u\Vert \ge \Vert u\Vert \), for all \(u\in C_{c}^{\infty }(E)\). This, in turn, leads to

$$\begin{aligned} (H^4u,u)=(H^2u,H^2u)=\Vert H^2u\Vert ^2\ge \Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(E). \end{aligned}$$

Continuing like this, we obtain \((H^ku,u)\ge \Vert u\Vert ^2\), for all \(u\in C_{c}^{\infty }(E)\) and all \(k\in \mathbb {Z}_{+}\). In this case, by an abstract fact (see Theorem X.26 in [28]), the essential self-adjointness of \(H^{k}\) on \(C_{c}^{\infty }(E)\) is equivalent to the following statement: if \(u\in L^2(E)\) satisfies \(H^{k}u=0\), then \(u=0\).

Let \(u\in L^2(E)\) satisfy \(H^{k}u=0\). Since \(V\in C^{\infty }(E)\), by local elliptic regularity it follows that \(u\in C^{\infty }(E)\cap L^2(E)\). Define

$$\begin{aligned} f_{j}:=H^{k-j}u,\quad j=0,\pm 1, \pm 2, \dots \end{aligned}$$
(4.4)

Here, in the case \(k-j<0\), the definition (4.4) is interpreted as \(((H_{\max })^{-1})^{j-k}\). We already noted that \(H_{\min }\) is essentially self-adjoint and positive. Furthermore, it is well known that the self-adjoint closure of \(H_{\min }\) coincides with \(H_{\max }\). Therefore \(H_{\max }\) is a positive self-adjoint operator, and \((H_{\max })^{-1}:L^2(E)\rightarrow L^2(E)\) is bounded. This, together with \(f_k=u\in L^2(E)\) explains the following property: \(f_j\in L^2(E)\), for all \(j\ge k\). Additionally, observe that \(f_j= 0\) for all \(j\le 0\) because \(f_0= 0\). Furthermore, we note that \(f_j\in C^{\infty }(E)\), for all \(j\in \mathbb {Z}\). The last assertion is obvious for \(j\le k\), and for \(j>k\) it can be seen by showing that \(H^{j}f_j=0\) in distributional sense and using \(f_j\in L^2(E)\) together with local elliptic regularity. To see this, let \(v\in C_{c}^{\infty }(E)\) be arbitrary, and note that

$$\begin{aligned} (f_j,H^{j}v)=(H^{k-j}u,H^{j}v)=(u,H^{k}v)=(H^ku,v)=0. \end{aligned}$$

Finally, observe that

$$\begin{aligned} H^{l}f_j=f_{j-l},\quad \text {for all }j\in \mathbb {Z}\text { and } l\in \mathbb {Z}_{+}\cup \{0\}. \end{aligned}$$
(4.5)

With \(f_j\) as in (4.4), define the functions \(\alpha _j\) and \(\beta _j\) on the interval \(0\le T<\infty \) by the formulas

$$\begin{aligned} \alpha _j(T):=\lambda _0^2\int _{0}^{T}(f_j,f_j)_{B_{t}}\,\mathrm{d}t,\quad \beta _j(T):=\int _{0}^{T}(D f_j,D f_j)_{B_{t}}\,\mathrm{d}t, \end{aligned}$$
(4.6)

where \(\lambda _0\) is as in (4.1) and \((\cdot ,\cdot )_{B_t}\) is as in (3.2).

In the sequel, to simplify the notations, the functions \(\alpha _j(T)\) and \(\beta _j(T)\), the inner products \((\cdot ,\cdot )_{B_{t}}\), and the corresponding norms \(\Vert \cdot \Vert _{B_t}\) appearing in (4.6) will be denoted by \(\alpha _j\), \(\beta _j\), \((\cdot ,\cdot )_{t}\), and \(\Vert \cdot \Vert _{t}\), respectively.

Note that \(\alpha _j\) and \(\beta _j\) are absolutely continuous on \([0,\infty )\). Furthermore, \(\alpha _j\) and \(\beta _j\) have a left first derivative and a right first derivative at each point. Additionally, \(\alpha _j\) and \(\beta _j\) are differentiable, except at (at most) countably many points. In the sequel, to simplify notations, we shall denote the right first derivatives of \(\alpha _j\) and \(\beta _j\) by \(\alpha _j'\) and \(\beta _j'\). Note that \(\alpha _j\), \(\beta _j\), \(\alpha _j'\) and \(\beta _j'\) are non-decreasing and non-negative functions. Note also that \(\alpha _j\) and \(\beta _j\) are convex functions. Furthermore, since \(f_j=0\) for all \(j\le 0\), it follows that \(\alpha _j\equiv 0\) and \(\beta _j\equiv 0\) for all \(j\le 0\). Finally, using (4.1) and the property \(f_j\in L^2(E)\cap C^{\infty }(E)\) for all \(j\ge k\), observe that

$$\begin{aligned} \lambda _0^2(f_j,f_j)+(Df_j,D f_j)\le (Vf_j,f_j)+(D f_j,D f_j)=(f_j,Hf_j)=(f_j,f_{j-1})<\infty , \end{aligned}$$

for all \(j>k\). Here, “integration by parts” in the first equality is justified because \(H_{\min }\) is essentially self-adjoint (i.e. \(C_{c}^{\infty }(E)\) is an operator core of \(H_{\max }\)). Hence, \(\alpha _j'\) and \(\beta _j'\) are bounded for all \(j>k\). It turns out that \(\alpha _j\) and \(\beta _j\) satisfy a system of differential inequalities, as seen in the next proposition.

Proposition 4.3

Let \(\alpha _j\) and \(\beta _j\) be as in (4.6). Then, for all \(j\ge 1\) and all \(T\ge 0\) we have

$$\begin{aligned} \alpha _j+\beta _j\le \sqrt{\alpha _j'\beta _j'}+\sum _{l=0}^{\infty }\left( \sqrt{\alpha '_{j+l+1}\beta '_{j-l-1}}+\sqrt{\alpha '_{j-l-1}\beta '_{j+l+1}}\right) \end{aligned}$$
(4.7)

and

$$\begin{aligned} \alpha _j\le \lambda _0^2\left( \sum _{l=0}^{\infty }\left( \sqrt{\alpha '_{j+l+1}\beta '_{j-l}}+\sqrt{\alpha '_{j-l}\beta '_{j+l+1}}\right) \right) , \end{aligned}$$
(4.8)

where \(\lambda _0\) is as in (4.1) and \(\alpha _i'\), \(\beta _i'\) denote the right-hand derivatives.

Remark 4.4

Note that the sums in (4.7) and (4.8) are finite since \(\alpha _{i}\equiv 0\) and \(\beta _{i}\equiv 0\) for \(i\le 0\). As our goal is to show that \(f_k=u=0\), we will only use the first k inequalities in (4.7) and the first k inequalities in (4.8).

Proof of Proposition 4.3

From (4.6) and (4.1) it follows that

$$\begin{aligned} \alpha _j+\beta _j\le \int _{0}^{T}\left( (f_j, Vf_j)_{t}+(D f_j, D f_j)_{t}\right) \,\mathrm{d}t. \end{aligned}$$
(4.9)

We start from (4.9), use (3.3), Cauchy–Schwarz inequality, and (4.5) to obtain

$$\begin{aligned} \alpha _j+\beta _j&\le \int _{0}^{T}((f_j, Vf_j)_{t}+(D f_j, D f_j)_{t})\,\mathrm{d}t\nonumber \\&=\int _{0}^{T}|(f_j, Hf_j)_{t}-(f_j, D^*D f_j)_{t}+(D f_j, D f_j)_{t}|\,\mathrm{d}t\nonumber \\&\le \lambda _0\int _{B_{T}}|D f_j(x)||f_j(x)|\,\mathrm{d}\mu (x)+\int _{0}^{T}|(f_j, Hf_j)_t|\,\mathrm{d}t\nonumber \\&\le \sqrt{\alpha _j'\beta _j'}+ \int _{0}^{T}|(Hf_{j+1}, f_{j-1})_t|\,\mathrm{d}t. \end{aligned}$$

We continue the process as follows:

$$\begin{aligned} \alpha _j+\beta _j&\le \sqrt{\alpha _j'\beta _j'}+\int _{0}^{T}|(Hf_{j+1}, f_{j-1})_t|\,\mathrm{d}t\nonumber \\&=\sqrt{\alpha _j'\beta _j'}+\int _{0}^{T}|(D^*D f_{j+1}, f_{j-1})_t+(f_{j+1}, Vf_{j-1})_{t}|\,\mathrm{d}t\nonumber \\&\le \sqrt{\alpha _j'\beta _j'}+\int _{0}^{T}|(D^*D f_{j+1}, f_{j-1})_t-(D f_{j+1}, D f_{j-1})_{t}|\,\mathrm{d}t\nonumber \\&\quad +\int _{0}^{T}|(D f_{j+1}, D f_{j-1})_{t}-(f_{j+1},D^*D f_{j-1})_{t}|\,\mathrm{d}t +\int _{0}^{T}|(f_{j+1}, Hf_{j-1})_{t}|\,\mathrm{d}t\nonumber \\&\le \sqrt{\alpha _j'\beta _j'} +\sqrt{\alpha _{j+1}'\beta _{j-1}'} +\sqrt{\alpha _{j-1}'\beta _{j+1}'}+\int _{0}^{T}|(Hf_{j+2}, f_{j-2})_{t}|\,\mathrm{d}t, \end{aligned}$$

where we used triangle inequality, (3.3), Cauchy–Schwarz inequality, and (4.5). We continue like this until the last term reaches the subscript \(j-l\le 0\), which makes the last term equal zero by properties of \(f_{i}\) discussed above. This establishes (4.7).

To show (4.8), we start from the definition of \(\alpha _j\), use (3.3), Cauchy–Schwarz inequality, and (4.5) to obtain

$$\begin{aligned} \alpha _j&= \lambda _0^2\int _{0}^{T}(f_j,f_j)_{t}\,\mathrm{d}t=\lambda _0^2\int _{0}^{T}|(f_j,Hf_{j+1})_{t}|\,\mathrm{d}t\nonumber \\&=\lambda _0^2\int _{0}^{T}|(f_j,D^*D f_{j+1})_{t}+(Vf_j,f_{j+1})_{t}|\,\mathrm{d}t\nonumber \\&\le \lambda _0^2\int _{0}^{T}|(f_j,D^*D f_{j+1})_{t}-(D f_j,D f_{j+1})_{t}|\,\mathrm{d}t\nonumber \\&\quad +\lambda _0^2\int _{0}^{T}|(D f_j,D f_{j+1})_{t}-(D^*D f_j,f_{j+1})_{t}|\,\mathrm{d}t+\lambda _0^2\int _{0}^{T}|(Hf_j,f_{j+1})_{t}|\,\mathrm{d}t\nonumber \\&\le \lambda _0^2\left( \sqrt{\alpha _{j+1}'\beta _{j}'} +\sqrt{\alpha _{j}'\beta _{j+1}'}\right) +\lambda _0^2\int _{0}^{T}|(f_{j-1},f_{j+1})_{t}|\,\mathrm{d}t. \end{aligned}$$

We continue like this until the last term reaches the subscript \(j-l\le 0\), which makes the last term equal zero by properties of \(f_{i}\) discussed above. This establishes (4.8). \(\square \)

End of the proof of Theorem 1

We will now transform the system (4.7) and (4.8) by introducing new variables:

$$\begin{aligned} \omega _j(T):=\alpha _j(T)+\beta _j(T),\quad \theta _j(T):=\alpha _j(T)-\beta _j(T)\quad T\in [0,\infty ). \end{aligned}$$
(4.10)

To carry out the transformation, observe that Cauchy–Schwarz inequality applied to vectors \(\Big \langle \sqrt{\alpha '_i}, \sqrt{\beta '_i}\Big \rangle \) and \(\Big \langle \sqrt{\beta '_p},\sqrt{\alpha '_p}\Big \rangle \) in \(\mathbb {R}^2\) gives

$$\begin{aligned} \sqrt{\alpha _i'\beta _p'}+\sqrt{\alpha _p'\beta _i'}\le \sqrt{\omega _i'\omega _p'}, \end{aligned}$$

which, together with (4.7) and (4.8) leads to

$$\begin{aligned} \omega _j\le \frac{1}{2}\sqrt{(\omega _j')^2-(\theta _j')^2}+\sum _{l=0}^{\infty }\sqrt{\omega '_{j+l+1}\omega '_{j-l-1}} \end{aligned}$$
(4.11)

and

$$\begin{aligned} \frac{1}{2}(\omega _j+\theta _j)\le \lambda _0^2\left( \sum _{l=0}^{\infty }\sqrt{\omega '_{j+l+1}\omega '_{j-l}}\right) , \end{aligned}$$
(4.12)

where \(\lambda _0\) is as in (4.1) and \(\omega _i'\), \(\theta _i'\) denote the right-hand derivatives.

The functions \(\omega _j\) and \(\theta _j\) satisfy the following properties: (i) \(\omega _j\) and \(\theta _j\) are absolutely continuous on \([0,\infty )\), and the right-hand derivatives \(\omega _j'\) and \(\theta _j'\) exist everywhere; (ii) \(\omega _j\) and \(\omega _j'\) are non-negative and non-increasing; (iii) \(\omega _j\) is convex; (iv) \(\omega _{j}'\) is bounded for all \(j\ge k\); (v) \(\omega _j(0)=\theta _j(0)=0\); and (vi) \(|\theta _j(T)|\le \omega _j(T)\) and \(|\theta _j'(T)|\le \omega _j'(T)\) for all \(T\in [0,\infty )\).

In Section IV.3 of [12] it was shown that if \(\omega _j\) and \(\theta _j\) are functions satisfying the above described properties (i)–(vi) and the system (4.11) and (4.12), then \(\omega _{j}\equiv 0\) for all \(j=1,2,\dots , k\). In particular, we have \(\omega _k(T)=0\), for all \(T\in [0,\infty )\), and hence \(f_k=0\). Going back to (4.4), we get \(u=0\), and this concludes the proof of essential self-adjointness of \(H^{k}\) on \(C_{c}^{\infty }(E)\). The essential self-adjointness of \(H^2\), \(H^{3}\), \(\dots \), and \(H^{k-1}\) on \(C_{c}^{\infty }(E)\) follows by Proposition 4.1. \(\square \)

5 Proof of Theorem 2

We adapt the proof of Theorem 1.1 in [13] to our type of operator. By assumption (2.6) it follows that

$$\begin{aligned} ((\Delta _{M,\mu }+q-C+1)u,u)\ge \Vert u\Vert ^2,\quad \text {for all }u\in C_{c}^{\infty }(M). \end{aligned}$$
(5.1)

Since (5.1) is satisfied and since M is non-compact and \(g^{\mathrm{TM}}\) is geodesically complete, a result of Agmon [1] (see also Proposition III.6.2 in [12]) guarantees the existence of a function \(\gamma \in C^{\infty }(M)\) such that \(\gamma (x)>0\) for all \(x\in M\), and

$$\begin{aligned} (\Delta _{M,\mu }+q-C+1)\gamma =\gamma . \end{aligned}$$
(5.2)

We now use the function \(\gamma \) to transform the operator \(H=\nabla ^*\nabla +V\). Let \(L_{\mu _1}^2(E)\) be the space of square integrable sections of E with inner product \((\cdot ,\cdot )_{\mu _1}\) as in (2.1), where \(\mathrm{d}\mu \) is replaced by \(\mathrm{d}\mu _1:=\gamma ^2 \mathrm{d}\mu \). For clarity, we denote \(L^2(E)\) from Sect. 2.1 by \(L_{\mu }^2(E)\). In what follows, the formal adjoints of \(\nabla \) with respect to inner products \((\cdot ,\cdot )_{\mu }\) and \((\cdot ,\cdot )_{\mu _1}\) will be denoted by \(\nabla ^{*,\mu }\) and \(\nabla ^{*,\mu _1}\), respectively. It is easy to check that the map \(T_{\gamma }:L_{\mu }^2(E)\rightarrow L_{\mu _1}^2(E)\) defined by \(Tu:=\gamma ^{-1} u\) is unitary. Furthermore, under the change of variables \(u\mapsto \gamma ^{-1} u\), the differential expression \(H=\nabla ^{*,\mu }\nabla +V\) gets transformed into \(H_1:=\gamma ^{-1}H\gamma \). Since T is unitary, the essential self-adjointness of \(H^k|_{C_{c}^{\infty }(E)}\) in \(L_{\mu }^2(E)\) is equivalent to essential self-adjointness of \((H_{1})^k|_{C_{c}^{\infty }(E)}\) in \(L_{\mu _1}^2(E)\).

In the sequel, we will show that \(H_1\) has the following form:

$$\begin{aligned} H_1=\nabla ^{*,\mu _1}\nabla +\widetilde{V}, \end{aligned}$$
(5.3)

with

$$\begin{aligned} \widetilde{V}(x):=\frac{\Delta _{M,\mu }\gamma }{\gamma }\,\text {Id}(x)+V(x). \end{aligned}$$

To see this, let \(w,\,z\in C_{c}^{\infty }(E)\) and consider

$$\begin{aligned}&(H_1w,z)_{\mu _1}=\int _{M}\langle \gamma ^{-1}H(\gamma w) ,z\rangle \,\gamma ^2\mathrm{d}\mu = \int _{M}\langle H(\gamma w) ,\gamma z\rangle \,\mathrm{d}\mu =(H(\gamma w) ,\gamma z)_{\mu }\nonumber \\&=(\nabla (\gamma w),\nabla (\gamma z))_{\mu }+(V\gamma w, \gamma z)_{\mu }=(\gamma ^2\nabla w,\nabla z)_{\mu }+(\mathrm{d}\gamma \otimes w,\mathrm{d}\gamma \otimes z)_{L^2_{\mu }(T^*M\otimes E)}\nonumber \\&\quad +(\gamma \nabla w,\mathrm{d}\gamma \otimes z)_{L^2_{\mu }(T^*M\otimes E)}+(\mathrm{d}\gamma \otimes w,\gamma \nabla z)_{L^2_{\mu }(T^*M\otimes E)}+(V\gamma w, \gamma z)_{\mu }. \end{aligned}$$
(5.4)

Setting \(\xi :=d(\gamma ^2/2)\in T^*M\) and using equation (1.34) in Appendix C of [32] we have

$$\begin{aligned}&(\gamma \nabla w,\mathrm{d}\gamma \otimes z)_{L^2_{\mu }(T^*M\otimes E)}= (\nabla w, \xi \otimes z)_{L^2_{\mu }(T^*M\otimes E)}=(\nabla _{X} w, z)_{\mu }, \end{aligned}$$
(5.5)

where X is the vector field associated with \(\xi \in T^*M\) via the metric \(g^{\mathrm{TM}}\).

Furthermore, by equation (1.35) in Appendix C of [32] we have

$$\begin{aligned} (\mathrm{d}\gamma \otimes w,\gamma \nabla z)_{L^2_{\mu }(T^*M\otimes E)}&=(\xi \otimes w,\nabla z)_{L^2_{\mu }(T^*M\otimes E)} =(\nabla ^{*,\mu }(\xi \otimes w),z)_{\mu }\nonumber \\&=-({\text {div}}_{\mu }(X)w,z)_{\mu }-(\nabla _{X}w,z)_{\mu }, \end{aligned}$$
(5.6)

where, in local coordinates \(x^{1},\,x^{2},\dots ,x^{n}\), for \(X=X^j\frac{\partial }{\partial x^{j}}\), with Einstein summation convention,

$$\begin{aligned} {\text {div}}_{\mu }(X):=\frac{1}{\kappa }\left( \frac{\partial }{\partial x^{j}}\left( \kappa X^{j}\right) \right) . \end{aligned}$$

[Recall that \(\mathrm{d}\mu =\kappa (x)\,\mathrm{d}x^{1}\mathrm{d}x^{2}\dots \mathrm{d}x^{n}\), where \(\kappa (x)\) is a positive \(C^{\infty }\)-density.] Since \(X^{j}=(g^{\mathrm{TM}})^{jl}\left( \gamma \frac{\partial \gamma }{\partial x^{l}}\right) \), we have

$$\begin{aligned} {\text {div}}_{\mu }(X)=|\mathrm{d}\gamma |^2-\gamma (\Delta _{M,\mu }\gamma ), \end{aligned}$$
(5.7)

where \(|\mathrm{d}\gamma (x)|\) is the norm of \(\mathrm{d}\gamma (x)\in T_{x}^*M\) induced by \(g^{\mathrm{TM}}\), and \(\Delta _{M,\mu }\) is as in (1.1) with metric \(g^{\mathrm{TM}}\). Combining (5.4)–(5.7) and noting that

$$\begin{aligned} (\mathrm{d}\gamma \otimes w,\mathrm{d}\gamma \otimes z)_{L^2_{\mu }(T^*M\otimes E)}=\int _{M}|\mathrm{d}\gamma |^2\langle w, z\rangle \,\mathrm{d}\mu , \end{aligned}$$

we obtain

$$\begin{aligned} (H_1w,z)_{\mu _1}&=\int _{M}\langle \nabla w,\nabla z\rangle \gamma ^2\,\mathrm{d}\mu +\int _{M}\langle V w, z\rangle \gamma ^2\,\mathrm{d}\mu +\int _{M}\gamma (\Delta _{M,\mu }\gamma )\langle w,z\rangle \,\mathrm{d}\mu \nonumber \\&=(\nabla w,\nabla z)_{L^2_{\mu _1}(T^*M\otimes E)}+(Vw,z)_{\mu _1}+(\gamma ^{-1}(\Delta _{M,\mu }\gamma )w,z)_{\mu _1}\nonumber \\&=(\nabla ^{*,\mu _1}\nabla w, z)_{\mu _1}+(Vw,z)_{\mu _1}+(\gamma ^{-1}(\Delta _{M,\mu }\gamma )w,z)_{\mu _1}, \end{aligned}$$
(5.8)

which shows (5.3).

By (2.5) and (5.2) it follows that

$$\begin{aligned} \widetilde{V}(x)=\frac{\Delta _{M,\mu }\gamma }{\gamma }\,\text {Id}(x)+V(x)\ge \,(C-1)\text {Id}(x),\quad \text {for all }x\in M, \end{aligned}$$

where C is as in (2.6). Thus, by Theorem 1 the operator \((H_1)^{k}|_{C_{c}^{\infty }(E)}\) is essentially self-adjoint in \(L^2_{\mu _1}(E)\) for all \(k\in \mathbb {Z}_{+}\). \(\square \)

6 Proof of Theorem 3

Throughout the section, we assume that the hypotheses of Theorem 3 are satisfied. In subsequent discussion, the notation \(\widehat{D}\) is as in (3.1) and the operators \(H_{\min }\) and \(H_{\max }\) are as in Sect. 4.1. We begin with the following lemma, whose proof is a direct consequence of the definition of \(H_{\max }\) and local elliptic regularity.

Lemma 6.1

Under the assumption \(V\in L^{\infty }_{\mathrm{loc}}(\mathrm{End}E)\), we have the following inclusion: \({\text {Dom}}(H_{\max })\subset W^{2,2}_{\mathrm{loc}}(E)\).

The proof of the next lemma is given in Lemma 8.10 of [5].

Lemma 6.2

For any \(u\in {\text {Dom}}(H_{\max })\) and any Lipschitz function with compact support \(\psi :M\rightarrow \mathbb {R}\), we have:

$$\begin{aligned} (D(\psi u),D(\psi u)) \ + \ (V\psi u, \psi u) \ = \ {\text {Re}}(\psi Hu,\psi u) \ + \ \Vert \widehat{D}(d\psi )u\Vert ^2. \end{aligned}$$
(6.1)

Corollary 6.3

Let H be as in (2.3), let \(u\in L^2(E)\) be a weak solution of \(Hu=0\), and let \(\psi :M\rightarrow \mathbb {R}\) be a Lipschitz function with compact support. Then

$$\begin{aligned} (\psi u, \,H (\psi u)) = \Vert \widehat{D}(d\psi )u\Vert ^2, \end{aligned}$$
(6.2)

where \((\cdot ,\cdot )\) on the left-hand side denotes the duality between \(W_{\mathrm{loc}}^{1,2}(E)\) and \(W^{-1,2}_{\mathrm{comp}}(E)\).

Proof

Since \(u\in L^2(E)\) and \(Hu=0\), we have \(u\in {\text {Dom}}(H_{\max })\subset W^{2,2}_{\mathrm{loc}}(E)\subset W^{1,2}_{\mathrm{loc}}(E)\), where the first inclusion follows by Lemma 6.1. Since \(\psi \) is a Lipschitz compactly supported function, we get \(\psi u\in W^{1,2}_{\mathrm{comp}}(E)\) and, hence, \(H(\psi u)\in W^{-1,2}_{\mathrm{comp}}(E)\). Now the equality (6.2) follows from (6.1), the assumption \(Hu=0\), and

$$\begin{aligned} (\psi u, \,H (\psi u)) = (\psi u, \, D^*D (\psi u)) \ + \ (V \psi u,\psi u) \ = \ (D(\psi u),D(\psi u)) \ + \ (V\psi u, \psi u), \end{aligned}$$

where in the second equality we used integration by parts; see Lemma 8.8 in [5]. Here, the two leftmost symbols \((\cdot ,\cdot )\) denote the duality between \(W_{\mathrm{comp}}^{1,2}(E)\) and \(W^{-1,2}_{\mathrm{loc}}(E)\), while the remaining ones stand for \(L^2\)-inner products. \(\square \)

The key ingredient in the proof of Theorem 3 is the Agmon-type estimate given in the next lemma, whose proof, inspired by an idea of [24], is based on the technique developed in [10] for magnetic Laplacians on an open set with compact boundary in \(\mathbb {R}^{n}\).

Lemma 6.4

Let \(\lambda \in \mathbb {R}\) and let \(v\in L^2(E)\) be a weak solution of \((H-\lambda )v=0\). Assume that that there exists a constant \(c_1>0\) such that, for all \(u \in W_{\mathrm{comp}}^{1,2}(E)\),

$$\begin{aligned} (u,\, (H-\lambda ) u ) \ge \lambda _0^2\int _{M}\max \left( \dfrac{1}{r(x)^2},1 \right) |u(x)|^2\,\mathrm{d}\mu (x) + c_1 \Vert u\Vert ^2, \end{aligned}$$
(6.3)

where r(x) is as in (2.7), \(\lambda _0\) is as in (2.2), the symbol \((\cdot ,\cdot )\) on the left-hand side denotes the duality between \(W_{\mathrm{comp}}^{1,2}(E)\) and \(W^{-1,2}_{\mathrm{loc}}(E)\), and \(|\cdot |\) is the norm in the fiber \(E_{x}\).

Then, the following equality holds: \(v=0\).

Proof

Let \(\rho \) and R be numbers satisfying \(0< \rho < 1/2\) and \( 1 < R < +\infty \). For any \(\varepsilon >0\), we define the function \(f_{\varepsilon }:M \rightarrow \mathbb {R}\) by \(f_{\varepsilon }(x)=F_{\varepsilon }(r(x))\), where r(x) is as in (2.7) and \(F_{\varepsilon }:[0,\infty )\rightarrow \mathbb {R}\) is the continuous piecewise affine function defined by

$$\begin{aligned} F_{\varepsilon }(s)= \left\{ \begin{array}{ll} 0 &{}\quad \mathrm{for~} s\le \varepsilon \\ \rho (s-\varepsilon )/(\rho - \varepsilon ) &{}\quad \mathrm{for~} \varepsilon \le s \le \rho \\ s &{}\quad \mathrm{for~} \rho \le s \le 1 \\ 1 &{}\quad \mathrm{for~} 1 \le s \le R \\ R+1 -s &{}\quad \mathrm{for~} R \le s \le R+1 \\ 0 &{}\quad \mathrm{for~ } s \ge R+1. \end{array} \right. \end{aligned}$$

Let us fix \(x_0\in M\). For any \(\alpha >0\), we define the function \(p_{\alpha }:M \rightarrow \mathbb {R}\) by

$$\begin{aligned} p_{\alpha }(x)=P_{\alpha }(d_{g^{\mathrm{TM}}}(x_0,x)), \end{aligned}$$

where \(P_{\alpha }:[0,\infty ) \rightarrow \mathbb {R}\) is the continuous piecewise affine function defined by

$$\begin{aligned} P_{\alpha }(s)= \left\{ \begin{array}{ll} 1 &{}\quad \mathrm{for~} s\le 1/{\alpha } \\ -{\alpha } s + 2 &{}\quad \mathrm{for~} 1/{\alpha } \le s \le 2/{\alpha } \\ 0 &{}\quad \mathrm{for~ } s \ge 2/{\alpha }. \end{array} \right. \end{aligned}$$

Since \(\widehat{d}_{g^{\mathrm{TM}}}(x_0,x)\le d_{g^{\mathrm{TM}}}(x_0,x)\), it follows that the support of \(f_{\varepsilon }p_{\alpha }\) is contained in the set \(B_{\alpha }:=\{x\in M:\widehat{d}_{g^{\mathrm{TM}}}(x_0,x)\le 2/\alpha \}\). By Assumption (A1) we know that \(\widehat{M}\) is a geodesically complete Riemannian manifold. Hence, by Hopf–Rinow Theorem the set \(B_{\alpha }\) is compact. Therefore, the support of \(f_{\varepsilon }p_{\alpha }\) is compact. Additionally, note that \(f_{\varepsilon }p_{\alpha }\) is a \(\beta \)-Lipschitz function (with respect to the distance corresponding to the metric \(g^{\mathrm{TM}}\)) with \(\beta =\frac{{\rho }}{{\rho -\varepsilon }}+\alpha \).

Since \(v\in L^2(E)\) and \((H-\lambda )v=0\), we have \(v\in {\text {Dom}}(H_{\max })\subset W^{2,2}_{\mathrm{loc}}(E)\subset W^{1,2}_{\mathrm{loc}}(E)\), where the first inclusion follows by Lemma 6.1. Since \(f_{\varepsilon }p_{\alpha }\) is a Lipschitz compactly supported function, we get \(f_{\varepsilon }p_{\alpha } v\in W^{1,2}_{\mathrm{comp}}(E)\) and, hence, \(((H-\lambda )(f_{\varepsilon }p_{\alpha } v))\in W^{-1,2}_{\mathrm{comp}}(E)\).

Using (2.2) we have

$$\begin{aligned} \Vert \widehat{D}(\mathrm{d}(f_{\varepsilon }p_{\alpha }))v\Vert ^2\le \lambda _0^2\int _{M}|\mathrm{d}(f_{\varepsilon }p_{\alpha })(x)|^2|v(x)|^2\,\mathrm{d}\mu (x), \end{aligned}$$
(6.4)

where \(|\mathrm{d}(f_{\varepsilon }p_{\alpha })(x)|\) is the norm of \(\mathrm{d}(f_{\varepsilon }p_{\alpha })(x)\in T_{x}^*M\) induced by \(g^{\mathrm{TM}}\).

By Corollary 6.3 with \(H-\lambda \) in place of H and the inequality (6.4), we get

$$\begin{aligned} (f_{\varepsilon }p_{\alpha } v,\, (H-\lambda ) (f_{\varepsilon }p_{\alpha } v)) \le \lambda _0^2\left( \frac{{\rho }}{{\rho -\varepsilon }}+\alpha \right) ^2\Vert v\Vert ^2. \end{aligned}$$
(6.5)

On the other hand, using the definitions of \(f_{\varepsilon }\) and \(p_{\alpha }\) and the assumption (6.3) we have

$$\begin{aligned} (f_{\varepsilon }p_{\alpha } v,\,(H-\lambda ) (f_{\varepsilon }p_{\alpha } v)) \ge \lambda _0^2\int _{S_{\rho ,R,\alpha }}|v(x)|^2\,\mathrm{d}\mu (x)+c_1 \Vert f_{\varepsilon }p_{\alpha } v \Vert ^2, \end{aligned}$$
(6.6)

where

$$\begin{aligned} S_{\rho ,R,\alpha }:=\{x\in M:\rho \le r(x) \le R \text { and } d_{g^{\mathrm{TM}}}(x_0,x) \le 1/\alpha \}. \end{aligned}$$

In (6.6) and (6.5), the symbol \((\cdot ,\cdot )\) stands for the duality between \(W_{\mathrm{comp}}^{1,2}(E)\) and \(W^{-1,2}_{\mathrm{loc}}(E)\). We now combine (6.6) and (6.5) to get

$$\begin{aligned} \lambda _0^2\int _{S_{\rho ,R,\alpha }}|v(x)|^2\,\mathrm{d}\mu (x) \ + \ c_1 \Vert f_{\varepsilon }p_{\alpha } v\Vert ^2\le \lambda _0^2\left( \frac{{\rho }}{{\rho -\varepsilon }}+\alpha \right) ^2\Vert v\Vert ^2. \end{aligned}$$

We fix \(\rho \), R, and \(\varepsilon \), and let \(\alpha \rightarrow 0+\). After that we let \(\varepsilon \rightarrow 0+\). The last step is to do \(\rho \rightarrow 0+ \) and \(R \rightarrow +\infty \). As a result, we get \(v=0\). \(\square \)

End of the proof of Theorem 3

Using integration by parts (see Lemma 8.8 in [5]), we have

$$\begin{aligned} (u,\, Hu) \ = \ (u,D^*Du) \ + \ (Vu,u) \ = \ (Du,Du) \\ + \ (Vu,u) \ \ge \ (Vu,u), \quad \text { for all }u\in W_{\mathrm{comp}}^{1,2}(E), \end{aligned}$$

where the two leftmost symbols \((\cdot ,\cdot )\) denote the duality between \(W_{\mathrm{comp}}^{1,2}(E)\) and \(W^{-1,2}_{\mathrm{loc}}(E)\), while the remaining ones stand for \(L^2\)-inner products. Hence, by assumption (2.8) we get:

$$\begin{aligned} (u,\, (H-\lambda ) u)&\ge \lambda _0^2\int _{M}\dfrac{1}{r(x)^2}|u(x)|^2\,\mathrm{d}\mu (x)-(\lambda +C)\Vert u\Vert ^2\nonumber \\&\ge \lambda _0^2\int _{M}\max \left( \dfrac{1}{r(x)^2},1 \right) |u(x)|^2\,\mathrm{d}\mu (x) -(\lambda +C+1)\Vert u\Vert ^2. \end{aligned}$$
(6.7)

Choosing, for instance, \(\lambda =-C-2\) in (6.7) we get the inequality (6.3) with \(c_1=1\).

Thus, \(H_{\min }-\lambda \) with \(\lambda =-C-2\) is a symmetric operator satisfying \((u,\, (H_{\min }-\lambda ) u)\ge \Vert u\Vert ^2\), for all \(u\in C_{c}^{\infty }(E)\). In this case, it is known (see Theorem X.26 in [28]) that the essential self-adjointness of \(H_{\min }-\lambda \) is equivalent to the following statement: if \(v\in L^2(E)\) satisfies \((H-\lambda )v=0\), then \(v=0\). Thus, by Lemma 6.4, the operator \((H_{\min }-\lambda )\) is essentially self-adjoint. Hence, \(H_{\min }\) is essentially self-adjoint. \(\square \)