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

Estimates of eigenvalues of an elliptic differential system in divergence form

Marcio C. Araújo Filho1  and  José N.V. Gomes2 1Departamento de Matemática, Universidade Federal de Rondônia, Campus Ji-Paraná, R. Rio Amazonas, 351, Jardim dos Migrantes, 76900-726 Ji-Paraná, Rondônia, Brazil 2Departamento de Matemática, Universidade Federal de São Carlos, Rod. Washington Luíz, Km 235, 13565-905 São Carlos, São Paulo, Brazil 1[email protected] 2[email protected]
Abstract.

In this paper, we compute universal estimates of eigenvalues of a coupled system of elliptic differential equations in divergence form on a bounded domain in Euclidean space. As an application, we show an interesting case of rigidity inequalities of the eigenvalues of the Laplacian, more precisely, we consider a countable family of bounded domains in Gaussian shrinking soliton that makes the behavior of known estimates of the eigenvalues of the Laplacian invariant by a first-order perturbation of the Laplacian. We also address the Gaussian expanding soliton case in two different settings. We finish with the special case of divergence-free tensors which is closely related to the Cheng-Yau operator.

Key words and phrases:
Eigenvalue problems, Estimate of eigenvalues, Elliptic differential system, Gaussian soliton, Rigidity results.
2010 Mathematics Subject Classification:
Primary 47A75; Secondary 47F05, 35P15, 53C24, 53C25

1. Introduction

Let n\mathbb{R}^{n} be the nn-dimensional Euclidean space with its canonical metric ,\langle,\rangle, and Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain with smooth boundary Ω\partial\Omega. Let us consider a symmetric positive definite (1,1)(1,1)-tensor TT on n\mathbb{R}^{n} and a function ηC2(n)\eta\in C^{2}(\mathbb{R}^{n}), so that we can define a second-order elliptic differential operator \mathscr{L} in the (η,T)(\eta,T)-divergence form as follows:

f:=divη(T(f))=div(T(f))η,T(f),\mathscr{L}f:=\mathrm{div}_{\eta}(T(\nabla f))=\mathrm{div}(T(\nabla f))-\langle\nabla\eta,T(\nabla f)\rangle, (1.1)

where div\mathrm{div} stands for the divergence operator and \nabla for the gradient operator. Since Ω\Omega is bounded, there exist two positive real constants ε\varepsilon and δ\delta, such that εITδI\varepsilon I\leq T\leq\delta I, where II is the (1,1)(1,1)-tensor identity on n\mathbb{R}^{n}.

The analysis of the sequence of the eigenvalues of elliptic differential operators in divergence forms in bounded domains in n\mathbb{R}^{n} is an interesting topic in both mathematics and physics. In particular, problems linking the shape of a domain to the spectrum of an operator are among the most fascinating of mathematical analysis. One of the reasons which make them so attractive is that they involve different fields of mathematics such as spectral theory, Riemannian geometry, and partial differential equations. Not only the literature about this subject is already very rich, but also it is not unlikely that operators in divergence forms may play a fundamental role in the understanding of countless physical facts.

In this paper, we address the eigenvalue problem for an operator which is a second-order perturbation of \mathscr{L}. More precisely, we compute universal estimates of the eigenvalues of the coupled system of second-order elliptic differential equations, namely:

{𝐮+α(divη𝐮)=σ𝐮in Ω,𝐮=0onΩ,\left\{\begin{array}[]{ccccc}\mathscr{L}{\bf u}+\alpha\nabla(\mathrm{div}_{\eta}{\bf u})&=&-\sigma{\bf u}&\mbox{in }&\Omega,\\ {\bf u}&=&0&\mbox{on}&\partial\Omega,\end{array}\right. (1.2)

where 𝐮=(u1,u2,,un){\bf u}=(u^{1},u^{2},\ldots,u^{n}) is a vector-valued function from Ω\Omega to n\mathbb{R}^{n}, the constant α\alpha is non-negative and 𝐮=(u1,u2,,un)\mathscr{L}{\bf u}=(\mathscr{L}u^{1},\mathscr{L}u^{2},\ldots,\mathscr{L}u^{n}).

We will see that +αdivη\mathscr{L}+\alpha\nabla\mathrm{div}_{\eta} is a formally self-adjoint operator in the Hilbert space 𝕃2(Ω,dm)\mathbb{L}^{2}(\Omega,dm) of all vector-valued functions that vanish on Ω\partial\Omega in the sense of the trace. It follows from inner product induced by Eqs. (3.3) and (3.4) in Section 3. Thus the eigenvalue problem (1.2) has a real and discrete spectrum

0<σ1σ2σk,0<\sigma_{1}\leq\sigma_{2}\leq\cdots\leq\sigma_{k}\leq\cdots\to\infty, (1.3)

where each σi\sigma_{i} is repeated according to its multiplicity.

A special case that we can obtain from Problem 1.2 occurs when TT is divergence-free, see Problem 6.2. For the sake of convenience, we address this case in Section 6. Some results from Problems 1.4 and 1.5 below are particular cases of this section. However, these two latter problems still remain prototype for us. In the next two paragraphs, we make brief comments about them.

When η\eta is a constant and TT is the identity operator II on n\mathbb{R}^{n}, Problem (1.2) becomes

{Δ𝐮+α(div𝐮)=σ𝐮inΩ,𝐮=0onΩ,\left\{\begin{array}[]{ccccc}\Delta{\bf u}+\alpha\nabla(\mathrm{div}\;{\bf u})&=&-\sigma{\bf u}&\mbox{in}&\Omega,\\ {\bf u}&=&0&\mbox{on}&\partial\Omega,\end{array}\right. (1.4)

where Δ𝐮=(Δu1,,Δun)\Delta{\bf u}=(\Delta u^{1},\ldots,\Delta u^{n}) and Δ\Delta is the Laplacian operator on C(Ω)C^{\infty}(\Omega). The operator Δ+αdiv\Delta+\alpha\nabla\mathrm{div} is known as Lamé’s operator. In the 33-dimensional case it shows up in the elasticity theory and α\alpha is determined by the positive constants of Lamé, so the assumption α0\alpha\geq 0 is justified. For further details on this issue, the interested reader can consult Pleijel [16] or Kawohl and Sweers [11]. It is worth mentioning here the works of Levine and Protter [12], Livitin and Parnovski [13], Hook [10], Cheng and Yang [5] and Chen et al. [3] in which we can find some interesting estimates of the eigenvalues of Problem (1.4). We will be more precise later when we will discuss the three latter papers.

When η\eta is not necessarily constant and T=IT=I, Problem (1.2) is rewritten as

{Δη𝐮+α(divη𝐮)=σ𝐮inΩ,𝐮=0onΩ,\left\{\begin{array}[]{ccccc}\Delta_{\eta}{\bf u}+\alpha\nabla\mbox{(div}_{\eta}{\bf u})&=&-\sigma{\bf u}&\mbox{in}&\Omega,\\ {\bf u}&=&0&\mbox{on}&\partial\Omega,\end{array}\right. (1.5)

where Δη𝐮=(Δηu1,,Δηun)\Delta_{\eta}{\bf u}=(\Delta_{\eta}u^{1},\ldots,\Delta_{\eta}u^{n}) and Δη=divη\Delta_{\eta}=\mathrm{div}_{\eta}\nabla is the drifted Laplacian operator on C(Ω)C^{\infty}(\Omega). The drifted Laplacian as well as the Bakry-Emery Ricci tensor Ric+2ηRic+\nabla^{2}\eta are the most appropriate geometric objects to study the smooth metric measure spaces (Mn,g,eηdvolg)(M^{n},g,e^{-\eta}dvol_{g}). In particular, the Bakry-Emery Ricci tensor has been especially studied in the theory of Ricci solitons, since a gradient Ricci soliton (Mn,g,η)(M^{n},g,\eta) is characterized by Ric+2η=λgRic+\nabla^{2}\eta=\lambda g, for some constant λ\lambda.

In Corollary 2.5, we show an interesting case of rigidity inequalities of eigenvalues of the Laplacian in a countable family of bounded domains in Gaussian shrinking soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}) by taking a specific isoparametric function as being the drifting function η\eta, see Remarks 2.2 and 2.3. We address the Gaussian expanding soliton case in Corollaries 2.6 and 2.7.

Throughout this paper dm=eηdΩdm=e^{-\eta}d\Omega stands for the weight volume form on Ω\Omega and |||\cdot| for the Euclidean norm. Moreover, let us define

𝐮=(u1,,un)andT(𝐮)=(T(u1),,T(un)),\displaystyle\nabla{\bf u}=(\nabla u^{1},\ldots,\nabla u^{n})\quad\mbox{and}\quad T(\nabla{\bf u})=(T(\nabla u^{1}),\ldots,T(\nabla u^{n})),

so that

T(𝐮)2=Ωj=1n|T(uj)|2dm=Ω|T(𝐮)|2𝑑m.\displaystyle\|T(\nabla{\bf u})\|^{2}=\int_{\Omega}\sum_{j=1}^{n}|T(\nabla u^{j})|^{2}dm=\int_{\Omega}|T(\nabla{\bf u})|^{2}dm. (1.6)

Henceforth, since there is no danger of confusion, we are using the same notation \|\cdot\| for the norm in (1.6) as well as for the canonical norm of a real-valued function in L2(Ω,dm)L^{2}(\Omega,dm).

Our proofs will be facilitated by analyzing the more general setting in which the function η\eta is not necessarily constant and TT is not necessarily the identity. In this case, we prove a universal quadratic estimate for the eigenvalues of Problem (1.2), which is an essential tool to obtain some of our estimates.

Theorem 1.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain and 𝐮i{\bf u}_{i} be a normalized eigenfunction corresponding to ii-th eigenvalue σi\sigma_{i} of Problem (1.2). Then, for any positive integer kk, we get

i=1k(σk+1σi)24δ(nδ+α)n2ε2i=1k(σk+1σi){[(σiαdivη𝐮i2)12+T02δ]2+C0δ},\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4\delta(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\Big{[}(\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2})^{\frac{1}{2}}+\frac{T_{0}}{2\sqrt{\delta}}\Big{]}^{2}+\frac{C_{0}}{\delta}\Big{\}},

where

C0=supΩ{12div(T2(η))14|T(η)|2}+δ2T0η0,C_{0}=\sup_{\Omega}{\Big{\{}}\frac{1}{2}\mathrm{div}(T^{2}(\nabla\eta))-\frac{1}{4}|T(\nabla\eta)|^{2}{\Big{\}}}+\frac{\delta}{2}T_{0}\eta_{0}, (1.7)

T0=supΩ|tr(T)|T_{0}=\sup_{\Omega}|\mathrm{tr}(\nabla T)| and η0=supΩ|η|\eta_{0}=\sup_{\Omega}|\nabla\eta|.

Remark 1.1.

Notice that the constant C0C_{0} in Eq. (1.7) has been appropriately defined such that [(σiαdivη𝐮i2)12+T02δ]2+C0δ>0\Big{[}(\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2})^{\frac{1}{2}}+\frac{T_{0}}{2\sqrt{\delta}}\Big{]}^{2}+\frac{C_{0}}{\delta}>0, for i=1,,ki=1,\dots,k.

We identify the quadratic estimate in Theorem 1.1 as the most appropriate inequality for the applications of our results. In particular, the constant C0C_{0} in (1.7) has a crucial importance for us.

Theorem 1.1 is an extension for +αdivη\mathscr{L}+\alpha\nabla\mbox{div}_{\eta} on vector-valued functions of the well-known Yang’s estimate of the eigenvalues of the Laplacian on real-valued functions. Its proof is motivated by the corresponding results for the Laplacian on real-valued functions case by Yang [20], for Δ+αdiv\Delta+\alpha\nabla\mathrm{div} on vector-valued functions case by Chen et al. [3, Theorem 1.1], and for \mathscr{L} on real-valued functions case by Gomes and Miranda [9].

We also prove an estimate for the sum of lower order eigenvalues in terms of the first eigenvalue and its correspondent eigenfunction.

Theorem 1.2.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, σi\sigma_{i} be the ii-th eigenvalue of Problem (1.2), for i=1,,ni=1,\ldots,n, and 𝐮1{\bf u}_{1} be a normalized eigenfunction corresponding to the first eigenvalue. Then, we get

i=1n(σi+1σ1)4δ(δ+α)ε2{[(σ1αdivη𝐮12)12+T02δ]2+C0δ}.\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq\frac{4\delta(\delta+\alpha)}{\varepsilon^{2}}\Big{\{}\Big{[}(\sigma_{1}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{1}\|^{2})^{\frac{1}{2}}+\frac{T_{0}}{2\sqrt{\delta}}\Big{]}^{2}+\frac{C_{0}}{\delta}\Big{\}}.

Theorem 1.2 is an extension for +αdivη\mathscr{L}+\alpha\nabla\mbox{div}_{\eta} on vector-valued functions of a stronger result obtained by Cheng and Yang [5, Theorem 1.2] for lower order eigenvalues of Problem (1.4). Its proof is motivated by the corresponding results for Δ+αdiv\Delta+\alpha\nabla\mathrm{div} case in [5] and for Δη+αdivη\Delta_{\eta}+\alpha\nabla\mathrm{div}_{\eta} case in [8, Theorem 1.3].

2. Applications

We begin this section by defining a known class of the functions which is closely related to our applications. A nonconstant smooth function f:nf:\mathbb{R}^{n}\to\mathbb{R} is called transnormal function if

|f|2=b(f),|\nabla f|^{2}=b(f), (2.1)

for some smooth function bb on the range of ff in \mathbb{R}. The function ff is called an isoparametric function if it moreover satisfies

Δf=a(f),\Delta f=a(f), (2.2)

for some continuous function aa on the range of ff in \mathbb{R}.

Eq. (2.1) implies that the level set hypersurfaces of ff are parallel hypersurfaces and it follows from Eq. (2.2) that these hypersurfaces have constant mean curvature. Isoparametric functions appear in the isoparametric hypersurfaces theory (i.e., has constant principal curvatures) systematically developed by Cartan [2] on space forms. Wang [18] considered the problem of extending this theory to a general Riemannian manifold and studied some properties of (2.1) and (2.2) more closely. Notice that isoparametric functions exist on a large class of spaces (e.g. symmetric spaces) other than space forms. Currently, new examples of isoparametric functions on Riemannian manifolds have been discovered, for instance, the potential function of any noncompact gradient Ricci soliton (Mn,g,η)(M^{n},g,\eta) with constant scalar curvature RR is an isoparametric function, since we can assume that η\eta (after a possible rescaling) satisfies |η|2=2ληR|\nabla\eta|^{2}=2\lambda\eta-R and Δη=λnR\Delta\eta=\lambda n-R, see, e.g., Chow et al. [7]. In particular, the potential function of the Gaussian shrinking soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}) is an isoparametric function, see Example 2.1. This latter fact and a brief analysis of the constant C0C_{0} in (1.7) were the main motivations to consider the isoparametric function η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2} to give some applications of our results. The quadratic estimate below is a basic result for it.

Corollary 2.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain and 𝐮i{\bf u}_{i} be a normalized eigenfunction corresponding to ii-th eigenvalue σi\sigma_{i} of Problem (1.5). Then, for any positive integer kk, we get

i=1k(σk+1σi)2\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2} 4(n+α)n2i=1k(σk+1σi)(σiαdivη𝐮i2+C0),\displaystyle\leq\frac{4(n+\alpha)}{n^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})(\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2}+C_{0}), (2.3)

where C0=supΩ{12Δη14|η|2}C_{0}=\sup_{\Omega}\big{\{}\frac{1}{2}\Delta\eta-\frac{1}{4}|\nabla\eta|^{2}\big{\}}. Moreover, σiαdivη𝐮i2+C0>0\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2}+C_{0}>0, for i=1,,ki=1,\ldots,k.

Proof.

In Problem (1.5) we must have T=IT=I. Then, we get ε=δ=1\varepsilon=\delta=1 and T0=0T_{0}=0. Hence, the result of the corollary follows from Theorem 1.1. ∎

The following corollary is an immediate consequence of Theorem 1.2.

Corollary 2.2.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, σi\sigma_{i} be the ii-th eigenvalue of Problem (1.5), for i=1,,ni=1,\ldots,n, and 𝐮1{\bf u}_{1} be a normalized eigenfunction corresponding to the first eigenvalue. Then, we get

i=1n(σi+1σ1)4(1+α)(σ1+D1),\displaystyle\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq 4(1+\alpha)(\sigma_{1}+D_{1}), (2.4)

where D1=αdivη𝐮12+C0D_{1}=-\alpha\|\mathrm{div}_{\eta}{\bf u}_{1}\|^{2}+C_{0} and C0=supΩ{12Δη14|η|2}C_{0}=\sup_{\Omega}\Big{\{}\frac{1}{2}\Delta\eta-\frac{1}{4}|\nabla\eta|^{2}\Big{\}}.

Notice that the appearance of the constant C0C_{0} is natural, since we did not impose any restriction on the function η\eta. We highlight that this constant has an unexpected geometric interpretation. Indeed, let us consider the warped metric g=g0+eηdθ2g=g_{0}+e^{-\eta}d\theta^{2} on the product Ω×𝕊1\Omega\times\mathbb{S}^{1}, where g0g_{0} stands for the canonical metric in the domain Ωn\Omega\subset\mathbb{R}^{n}, whereas dθ2d\theta^{2} is the canonical metric of the unit sphere 𝕊1\mathbb{S}^{1}, so that the scalar curvature of gg is given by 12Δη14|η|2\frac{1}{2}\Delta\eta-\frac{1}{4}|\nabla\eta|^{2}. Hence, we can obtain C0C_{0} as supremum of the scalar curvature on the warped product Ω×eη𝕊1\Omega\times_{e^{-\eta}}\mathbb{S}^{1}. Moreover, we ask the following natural question:

Question 1.

Under which conditions the inequalities for the eigenvalues obtained from (2.3) and (2.4) do not depend on the constant C0C_{0} for a nontrivial function η\eta?

We give an answer to this question by using a specific family of domains in Gaussian shrinking soliton. More precisely, we consider a countable family of bounded domains in n\mathbb{R}^{n} that makes the behavior of known estimates of eigenvalues of the Laplacian invariant by a first-order perturbation of the Laplacian, see Corollary 2.5.

Coming back to Corollary 2.1, we define

D0=αminj=1,,kdivη𝐮j2+C0,D_{0}=-\alpha\min_{j=1,\ldots,k}\|\mathrm{div}_{\eta}{\bf u}_{j}\|^{2}+C_{0}, (2.5)

so that, from (2.3), we get

i=1k(σk+1σi)2\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2} 4(n+α)n2i=1k(σk+1σi)(σi+D0).\displaystyle\leq\frac{4(n+\alpha)}{n^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})(\sigma_{i}+D_{0}). (2.6)

Notice that σi+D0>0.\sigma_{i}+D_{0}>0.

Now, as mentioned in the introduction, we immediately recover the following inequality:

i=1k(σk+1σi)24(n+α)n2i=1k(σk+1σi)σi,\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4(n+\alpha)}{n^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\sigma_{i}, (2.7)

which has been obtained by Chen et al. [3, Corollary 1.2] for Problem (1.4). Indeed, it follows from (2.5) and (2.6), since α0\alpha\geq 0 and we can take η\eta to be a constant. Moreover, Inequality (2.7) implies Theorem 1.1 in Cheng and Yang [5], whereas [5, Theorem 1.1] implies Theorem 10 in Hook [10]. However, we highlight that Inequality (2.6) provides an estimate for the eigenvalues of Problem (1.4) which is better than Inequality (2.7).

In the case of Problem (1.5), we can see that Inequality (2.4) is better than Inequality (1.7) in Du and Bezerra [8]; whereas Inequality (2.6) is better than Inequality (1.3) again in [8].

Besides, from Inequality (2.6) and following the steps of the proof of [9, Theorem 3], we obtain the inequalities:

Corollary 2.3.

Under the same setup as in Corollary 2.1, we have

σk+1+D0\displaystyle\sigma_{k+1}+D_{0}\leq (1+2(n+α)n2)1ki=1k(σi+D0)+[(2(n+α)n21ki=1k(σi+D0))2\displaystyle\Big{(}1+\frac{2(n+\alpha)}{n^{2}}\Big{)}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})+\Big{[}\Big{(}\frac{2(n+\alpha)}{n^{2}}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})\Big{)}^{2}
(1+4(n+α)n2)1kj=1k(σj1ki=1kσi)2]12\displaystyle-\Big{(}1+\frac{4(n+\alpha)}{n^{2}}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}} (2.8)

and

σk+1σk\displaystyle\sigma_{k+1}-\sigma_{k}\leq 2[(2(n+α)n21ki=1k(σi+D0))2\displaystyle 2\Big{[}\Big{(}\frac{2(n+\alpha)}{n^{2}}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})\Big{)}^{2}
(1+4(n+α)n2)1kj=1k(σj1ki=1kσi)2]12,\displaystyle-\Big{(}1+\frac{4(n+\alpha)}{n^{2}}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}}, (2.9)

where D0D_{0} is given by (2.5).

Proof.

By setting vi:=σi+D0v_{i}:=\sigma_{i}+D_{0} into (2.6) we get

i=1k(vk+1vi)24(n+α)n2i=1k(vk+1vi)vi.\sum_{i=1}^{k}(v_{k+1}-v_{i})^{2}\leq\frac{4(n+\alpha)}{n^{2}}\sum_{i=1}^{k}(v_{k+1}-v_{i})v_{i}. (2.10)

From (1.3), v1v2vk+1v_{1}\leq v_{2}\leq\cdots\leq v_{k+1}, and from (2.6) each vi>0v_{i}>0. Notice that (2.10) is a quadratic inequality of vk+1v_{k+1}. So, we can follow the steps of the proof of Inequalities (1.3) and (1.4) in Gomes and Miranda [9] to obtain the required inequalities of the corollary. ∎

Again from (2.6) and by applying the recursion formula of Cheng and Yang [4], we obtain the following corollary.

Corollary 2.4.

Under the same setup as in Corollary 2.1, we have

σk+1+D0(1+4(n+α)n2)k2(n+α)n2(σ1+D0),\sigma_{k+1}+D_{0}\leq\Big{(}1+\frac{4(n+\alpha)}{n^{2}}\Big{)}k^{\frac{2(n+\alpha)}{n^{2}}}(\sigma_{1}+D_{0}), (2.11)

where D0D_{0} is given by (2.5).

Proof.

Notice that the recursion formula by Cheng and Yang [4, Corollary 2.1] remains true for any positive real number, in particular, it holds for n2n+α\frac{n^{2}}{n+\alpha}, then we can apply this formula in (2.10) to obtain (2.11). ∎

From the classical Weyl’s asymptotic formula for the eigenvalues [19], we know that estimate (2.11) is optimal in the sense of the order on kk.

Remark 2.1.

If D0=0D_{0}=0, then the inequalities of eigenvalues (2.6) and (2.11) have the same behavior as the known estimates of the eigenvalues of Δ+αdiv\Delta+\alpha\nabla\mathrm{div}, see Inequality (2.7) and Chen et al. [3, Corollary 1.4], respectively. In the same way, the inequalities of eigenvalues (2.3) and (2.3) imply

σk+1(1+4(n+α)n2)1ki=1kσiandσk+1σk4(n+α)n21ki=1kσi,\sigma_{k+1}\leq\Big{(}1+\frac{4(n+\alpha)}{n^{2}}\Big{)}\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\quad\mbox{and}\quad\sigma_{k+1}-\sigma_{k}\leq\frac{4(n+\alpha)}{n^{2}}\frac{1}{k}\sum_{i=1}^{k}\sigma_{i},

which have the same behavior as the inequalities of the eigenvalue of Δ+αdiv\Delta+\alpha\nabla\mathrm{div} obtained by Chen et al. [3, Corollary 1.3].

If D1=0D_{1}=0, then the inequality of eigenvalues (2.4) has the same behavior as the known estimate of the eigenvalues of Δ+αdiv\Delta+\alpha\nabla\mathrm{div} proved by Cheng and Yang [5, Theorem 1.2].

Remark 2.2.

For α=0\alpha=0 case, if C0=0C_{0}=0 for some function η\eta (possibly radial or isoparametric), then the inequalities (2.4), (2.6), (2.3) and (2.11) have the same behavior as the known estimates of the eigenvalues of the Laplacian, see [1, Inequality (6.2)], [20, Theorem 1], [4, Inequality (1.8)] and [4, Corollary 2.1], respectively.

Example 2.1 below is a special case of C0=0C_{0}=0. To see this, let us consider an isoparametric function η(x)=λ2(x12++xk2)\eta(x)=\frac{\lambda}{2}(x_{1}^{2}+\cdots+x_{k}^{2}) on n\mathbb{R}^{n}, where λ\lambda is any nonzero real number, kk an integer with 0<kn0<k\leq n and x=(x1,,xn)nx=(x_{1},\ldots,x_{n})\in\mathbb{R}^{n}. It is easy to verify that

|η|2=2ληandΔη=λk.|\nabla\eta|^{2}=2\lambda\eta\>\>\mbox{and}\>\>\Delta\eta=\lambda k.

In particular, if k=nk=n, the function η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2} is the potential function of the Gaussian shrinking (λ>0\lambda>0) or expanding (λ<0\lambda<0) soliton on n\mathbb{R}^{n}. We now take η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2} into the equation of C0C_{0} in Corollary 2.1, so that,

C0=supΩ{λn2λ24|x|2}.C_{0}=\sup_{\Omega}\Big{\{}\frac{\lambda n}{2}-\frac{\lambda^{2}}{4}|x|^{2}\Big{\}}. (2.12)

With these considerations in mind, we write the next two examples.

Example 2.1.

Let us consider the family of bounded domains {Ωl}l=1\{\Omega_{l}\}_{l=1}^{\infty} in Gaussian shrinking or expanding soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}) given by

Ωl=𝔹(rl)𝔹¯(2n/|λ|)={xn;2n|λ|<|x|2<rl2},\Omega_{l}=\mathbb{B}(r_{l})-\bar{\mathbb{B}}(\sqrt{2n/|\lambda|})=\Big{\{}x\in\mathbb{R}^{n};\frac{2n}{|\lambda|}<|x|^{2}<r_{l}^{2}\Big{\}},

where rl>2n/|λ|r_{l}>\sqrt{2n/|\lambda|} is a rational number, and 𝔹(r)\mathbb{B}(r) stands for the open ball of radius rr centered at the origin in n\mathbb{R}^{n}. So,

minΩ¯l|x|2=2n|λ|,for alll=1,2,.\min_{\bar{\Omega}_{l}}|x|^{2}=\frac{2n}{|\lambda|},\quad\mbox{for all}\quad l=1,2,\ldots.

(a) Shrinking case:

C0=λ2supΩl{nλ2|x|2}=λ2(nλ2minΩ¯l|x|2)=0,for alll=1,2,.C_{0}=\frac{\lambda}{2}\sup_{\Omega_{l}}\Big{\{}n-\frac{\lambda}{2}|x|^{2}\Big{\}}=\frac{\lambda}{2}\Big{(}n-\frac{\lambda}{2}\min_{\bar{\Omega}_{l}}|x|^{2}\Big{)}=0,\quad\mbox{for all}\quad l=1,2,\ldots.

(b) Expanding case:

C0=λ2infΩl{nλ2|x|2}=λ2(nλ2minΩ¯l|x|2)=λn,for alll=1,2,.C_{0}=\frac{\lambda}{2}\inf_{\Omega_{l}}\Big{\{}n-\frac{\lambda}{2}|x|^{2}\Big{\}}=\frac{\lambda}{2}\Big{(}n-\frac{\lambda}{2}\min_{\bar{\Omega}_{l}}|x|^{2}\Big{)}=\lambda n,\quad\mbox{for all}\quad l=1,2,\ldots.
Example 2.2.

Let us consider the domain Ω\Omega to be the open ball 𝔹(r)\mathbb{B}(r) of radius rr centered at the origin in Gaussian shrinking or expanding soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}). From Eq. (2.12), we easily see that C0=λn/2C_{0}=\lambda n/2 for both shrinking and expanding case.

We are now in the right position to give the interesting applications that we had promised.

Corollary 2.5 (Non-dependence of η\eta).

Let us consider the family of domains {Ωl}l=1\{\Omega_{l}\}_{l=1}^{\infty} given by Example 2.1 in Gaussian shrinking soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}). Let σi\sigma_{i} be the ii-th eigenvalue of the drifted Laplacian Δη\Delta_{\eta} on real-valued functions, with drifting function η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2}, on each Ωl\Omega_{l} with Dirichlet boundary condition. Then, we get

i=1k(σk+1σi)2\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2} 4ni=1k(σk+1σi)σi,\displaystyle\leq\frac{4}{n}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\sigma_{i},
σk+1(1+2n)1ki=1kσi+[(2n1ki=1kσi)2(1+4n)1kj=1k(σj1ki=1kσi)2]12,\displaystyle\sigma_{k+1}\leq\Big{(}1+\frac{2}{n}\Big{)}\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}+\Big{[}\Big{(}\frac{2}{n}\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}-\Big{(}1+\frac{4}{n}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}},
σk+1σk2[(2n1ki=1kσi)2(1+4n)1kj=1k(σj1ki=1kσi)2]12,\displaystyle\sigma_{k+1}-\sigma_{k}\leq 2\Big{[}\Big{(}\frac{2}{n}\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}-\Big{(}1+\frac{4}{n}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}},
σk+1(1+4n)k2nσ1\sigma_{k+1}\leq\Big{(}1+\frac{4}{n}\Big{)}k^{\frac{2}{n}}\sigma_{1}

and

i=1n(σi+1σ1)4σ1.\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq 4\sigma_{1}.
Proof.

We start by taking α=0\alpha=0 as in Problem (1.5). Next, we note that the constant C0=0C_{0}=0 for the shrinking case. So, the required inequalities follow from inequalities (2.6), (2.3), (2.3), (2.11) and (2.4), respectively. ∎

Remark 2.3.

Notice that Corollary 2.5 can be regarded as rigidity inequalities (see Remark 2.2) on the family of bounded domains {Ωl}l=1\{\Omega_{l}\}_{l=1}^{\infty} given by Example 2.1 in Gaussian shrinking soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}).

Now, we will address the expanding case.

Corollary 2.6.

Let 𝔹(r)\mathbb{B}(r) be the open ball of radius rr centered at the origin in Gaussian expanding soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}). Let σ1\sigma_{1} be the first eigenvalue of the drifted Laplacian Δη\Delta_{\eta} on real-valued functions, with drifting function η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2}, on 𝔹(r)\mathbb{B}(r) with Dirichlet boundary condition. Then, we have

σ1π2n64r2λn2,\sigma_{1}\geq\frac{\pi^{2}n}{64r^{2}}-\frac{\lambda n}{2},

and the next estimate for the sum of lower order eigenvalues σi\sigma_{i} of Δη\Delta_{\eta} in terms of the first eigenvalue

i=1n(σi+1σ1)4(σ1+λn2).\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq 4(\sigma_{1}+\frac{\lambda n}{2}). (2.13)
Proof.

Let σ1\sigma_{1} and σ2\sigma_{2} be the first and second eigenvalues of the drifted Laplacian Δη\Delta_{\eta} on real-valued functions on 𝔹(r)\mathbb{B}(r) with Dirichlet boundary condition, respectively. If λ<0\lambda<0, then both η=λ2|x|2\eta=\frac{\lambda}{2}|x|^{2} and f=12Δη14|η|2f=\frac{1}{2}\Delta\eta-\frac{1}{4}|\nabla\eta|^{2} are concave functions on the closure of the convex domain 𝔹(r)\mathbb{B}(r). Thus, we can apply Theorem 3 by Ma and Liu [14] to obtain

σ2σ1π216r2.\sigma_{2}-\sigma_{1}\geq\frac{\pi^{2}}{16r^{2}}. (2.14)

On the other hand, note that we can use (2.3) or (2.11), for α=0\alpha=0, to get

σ2σ14nσ1+2λ.\sigma_{2}-\sigma_{1}\leq\frac{4}{n}\sigma_{1}+2\lambda. (2.15)

Combining (2.14) and (2.15), we conclude that

σ1π2n64r2λn2.\sigma_{1}\geq\frac{\pi^{2}n}{64r^{2}}-\frac{\lambda n}{2}.

Moreover, we can use (2.4), for α=0\alpha=0, to obtain (2.13). ∎

Corollary 2.7.

Let us consider the family of domains {Ωl}l=1\{\Omega_{l}\}_{l=1}^{\infty} given by Example 2.1 in Gaussian expanding soliton (n,δij,λ2|x|2)(\mathbb{R}^{n},\delta_{ij},\frac{\lambda}{2}|x|^{2}). Let σi\sigma_{i} be the ii-th eigenvalue of the drifted Laplacian Δη\Delta_{\eta} on real-valued functions, with drifting function η(x)=λ2|x|2\eta(x)=\frac{\lambda}{2}|x|^{2}, on each Ωl\Omega_{l} with Dirichlet boundary condition. Then, it is valid the following estimate for the sum of lower order eigenvalues of Δη\Delta_{\eta} in terms of the first eigenvalue:

i=1n(σi+1σ1)4(σ1+λn).\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq 4(\sigma_{1}+\lambda n).
Proof.

In fact, we can use (2.4), for α=0\alpha=0, to deduce the required estimate. ∎

Remark 2.4.

A final remark is in order. We observe that Corollaries 2.1 and 2.2 can be obtained from Corollaries 6.1 and 6.2, respectively. Whereas Corollaries 2.3 and 2.4 can be obtained from Corollaries 6.3 and 6.4, respectively. However, as we already mentioned before, they have been a prototype for us.

3. Preliminaries

This section is brief and serves to set the stage, introducing some basic notation and describing what is meant by the properties of a (1,1)(1,1)-tensor in a bounded domain Ωn\Omega\subset\mathbb{R}^{n} with smooth boundary Ω\partial\Omega.

Throughout the paper, we will be constantly using the identification of a (0,2)(0,2)-tensor T:𝔛(Ω)×𝔛(Ω)C(Ω)T:\mathfrak{X}(\Omega)\times\mathfrak{X}(\Omega)\to C^{\infty}(\Omega) with its associated (1,1)(1,1)-tensor T:𝔛(Ω)𝔛(Ω)T:\mathfrak{X}(\Omega)\to\mathfrak{X}(\Omega) by the equation

T(X),Y=T(X,Y).\langle T(X),Y\rangle=T(X,Y).

In particular, the tensor ,\langle,\rangle will be identified with the identity II in 𝔛(Ω)\mathfrak{X}(\Omega). From the definition of η\eta-divergence of XX (see Eq. (1.1)) and the usual properties of divergence of vector fields, one has

divη(fX)=fdivηX+fX,\mathrm{div}_{\eta}(fX)=f\mathrm{div}_{\eta}X+\nabla f\cdot X, (3.1)

for all fC(Ω)f\in C^{\infty}(\Omega).

Notice that the (η,T)(\eta,T)-divergence form of \mathscr{L} on Ω\Omega allows us to check that the divergence theorem remains true in the form

ΩdivηX𝑑m=ΩX,ν𝑑μ.\int_{\Omega}\mathrm{div}_{\eta}Xdm=\int_{\partial\Omega}\langle X,\nu\rangle d\mu. (3.2)

In particular, for X=T(f)X=T(\nabla f),

Ωf𝑑m=ΩT(f,ν)𝑑μ,\int_{\Omega}\mathscr{L}{f}dm=\int_{\partial\Omega}T(\nabla f,\nu)d\mu,

where dm=eηdΩdm=e^{-\eta}d\Omega and dμ=eηdΩd\mu=e^{-\eta}d\partial\Omega are the weight volume form on Ω\Omega and the volume form on the boundary Ω\partial\Omega induced by the outward unit normal vector ν\nu on Ω\partial\Omega, respectively. Thus, the integration by parts formula is given by

Ωf𝑑m=ΩT(,f)𝑑m+ΩT(f,ν)𝑑μ,\int_{\Omega}\ell\mathscr{L}{f}dm=-\int_{\Omega}T(\nabla\ell,\nabla f)dm+\int_{\partial\Omega}\ell T(\nabla f,\nu)d\mu, (3.3)

for all ,fC(Ω)\ell,f\in C^{\infty}(\Omega). Hence, \mathscr{L} is a formally self-adjoint operator in the space of all real-valued functions in L2(Ω,dm)L^{2}(\Omega,dm) that vanish on Ω\partial\Omega in the sense of the trace. Furthermore, from (3.1) and (3.2) we obtain

Ω𝐯(divη𝐮)dm=Ωdivη𝐮divη𝐯𝑑m,\int_{\Omega}{\bf v}\cdot\nabla(\mathrm{div}_{\eta}{\bf u})dm=-\int_{\Omega}\mathrm{div}_{\eta}{\bf u}\mathrm{div}_{\eta}{\bf v}dm, (3.4)

for all vector-valued function 𝐮=(u1,u2,,un){\bf u}=(u^{1},u^{2},\ldots,u^{n}) and 𝐯=(v1,v2,,vn){\bf v}=(v^{1},v^{2},\ldots,v^{n}) both from Ω\Omega to n\mathbb{R}^{n}, with 𝐯{\bf v} vanishing on Ω\partial\Omega.

We conclude that +αdivη\mathscr{L}+\alpha\nabla\mathrm{div}_{\eta} is a formally self-adjoint operator in the Hilbert space 𝕃2(Ω,dm)\mathbb{L}^{2}(\Omega,dm) of all vector-valued functions that vanish on Ω\partial\Omega in the sense of the trace, with inner product induced by Eqs. (3.3) and (3.4). Thus the eigenvalue problem (1.2) has a real and discrete spectrum 0<σ1σ2σk0<\sigma_{1}\leq\sigma_{2}\leq\cdots\leq\sigma_{k}\leq\cdots\to\infty, where each σi\sigma_{i} is repeated according to its multiplicity.

It is worth mentioning here the paper by Gomes and Miranda [9, Section 2] from which we know some geometric motivations to work with the operator \mathscr{L} in the (η,T)(\eta,T)-divergence form in bounded domains in Riemannian manifolds. They showed that it appears as the trace of a (1,1)(1,1)-tensor on a Riemannian manifold MM, and computed a Bochner-type formula for it. An interesting fact is that Eq.  (2.3)(2.3) of [9] relates the operator \mathscr{L} to Cheng and Yau’s operator \Box defined in [6]. In particular, if we take TT to be the Einstein tensor, then this latter relation is likely to have applications in physics.

For a vector-valued function 𝐮=(u1,u2,,un){\bf u}=(u^{1},u^{2},\ldots,u^{n}) from Ω\Omega to n\mathbb{R}^{n}, we define

𝐮=(u1,,un).\displaystyle\nabla{\bf u}=(\nabla u^{1},\ldots,\nabla u^{n}).

Now, we are considering two definitions for TT, and since there is no danger of confusion, we are using the same notation TT for both definitions. Let XX be a vector field on Ω\Omega, we define

T(𝐮)=(T(u1),,T(un))\displaystyle T(\nabla{\bf u})=(T(\nabla u^{1}),\ldots,T(\nabla u^{n}))

and

T(X,𝐮)\displaystyle T(X,\nabla{\bf u}) =(T(X),u1,,T(X),un)\displaystyle=(\langle T(X),\nabla{u}^{1}\rangle,\ldots,\langle T(X),\nabla{u}^{n}\rangle)
=(X,T(u1),,X,T(un))\displaystyle=(\langle X,T(\nabla u^{1})\rangle,\ldots,\langle X,T(\nabla u^{n})\rangle)
=(T(X,u1),,T(X,un)).\displaystyle=(T(X,\nabla u^{1}),\ldots,T(X,\nabla u^{n})). (3.5)

Moreover, for all real-valued functions f,C(Ω)f,\ell\in C^{\infty}(\Omega) it is immediate from the properties of divη\mathrm{div}_{\eta} and the symmetry of TT that

(f)=f+2T(f,)+f.\mathscr{L}(f\ell)=f\mathscr{L}\ell+2T(\nabla f,\nabla\ell)+\ell\mathscr{L}f.

So, the following equation is well understood for a vector-valued function 𝐮{\bf u} and a real-valued function fC(Ω)f\in C^{\infty}(\Omega)

(f𝐮)=((fu1),,(fun))\displaystyle\mathscr{L}(f{\bf u})=(\mathscr{L}(fu^{1}),\ldots,\mathscr{L}(fu^{n}))
=(fu1+2T(f),u1+(f)u1,,fun+2T(f),un+(f)un)\displaystyle=(f\mathscr{L}u^{1}+2\langle T(\nabla f),\nabla u^{1}\rangle+\mathscr{L}(f)u^{1},\ldots,f\mathscr{L}u^{n}+2\langle T(\nabla f),\nabla u^{n}\rangle+\mathscr{L}(f)u^{n})
=f(u1,,un)+2(T(f),un,,T(f),un)+(f)(u1,,un)\displaystyle=f(\mathscr{L}u^{1},\ldots,\mathscr{L}u^{n})+2(\langle T(\nabla f),\nabla u^{n}\rangle,\ldots,\langle T(\nabla f),\nabla u^{n}\rangle)+\mathscr{L}(f)(u^{1},\ldots,u^{n})
=f𝐮+2T(f,𝐮)+f𝐮.\displaystyle=f\mathscr{L}{\bf u}+2T(\nabla f,\nabla{\bf u})+\mathscr{L}f{\bf u}. (3.6)

Besides, we are using the classical norms: |𝐮|2=i=1n(ui)2|{\bf u}|^{2}=\sum_{i=1}^{n}(u^{i})^{2} and 𝐮2=Ω|𝐮|2𝑑m\|{\bf u}\|^{2}=\int_{\Omega}|{\bf u}|^{2}dm. We observe that εITδI\varepsilon I\leq T\leq\delta I implies

|T(X)|2δT(X),Xfor allX𝔛(n).\displaystyle|T(X)|^{2}\leq\delta\langle T(X),X\rangle\quad\mbox{for all}\quad X\in\mathfrak{X}(\mathbb{R}^{n}). (3.7)

In particular, we obtain

|T(η)|2δ2|η|2.\displaystyle|T(\nabla\eta)|^{2}\leq\delta^{2}|\nabla\eta|^{2}. (3.8)

4. Three technical lemmas

In order to prove our first theorem we will need three technical lemmas. The first one is motivated by the corresponding results for Problem (1.4) proven by Chen et al. [3, Lemma 2.1] and for Problem (1.5) proven by Du and Bezerra [8, Lemma 2.1]. Here, we follow the steps of the proof of Lemma 2.1 in [3] with appropriate adaptations for +αdivη\mathscr{L}+\alpha\nabla\mathrm{div}_{\eta}.

Lemma 4.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, σi\sigma_{i} be the i-th eigenvalue of Problem (1.2) and 𝐮i{\bf u}_{i} be a normalized vector-valued eigenfunction corresponding to σi\sigma_{i}. Then, for any fC2(Ω)C1(Ω)f\in C^{2}(\Omega)\cap C^{1}(\partial\Omega) and any positive constant BB, we obtain

i=1k(σk+1σi)2\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2} {(1B)ΩT(f,f)|𝐮i|2𝑑mBαΩ|f𝐮i|2𝑑m}\displaystyle\Big{\{}(1-B)\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm-B\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm\Big{\}}
1Bi=1k(σk+1σi)T(f,𝐮i)+12f𝐮i2.\displaystyle\leq\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\|T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}\|^{2}.
Proof.

Let 𝐮i{\bf u}_{i} be a normalized vector-valued eigenfunction corresponding to σi\sigma_{i}, i.e., it satisfies

{𝐮i+α(divη𝐮i)=σi𝐮iinΩ,𝐮i=0onΩ,Ω𝐮i𝐮j𝑑m=δijfor anyi,j.\left\{\begin{array}[]{ccccc}\mathscr{L}{\bf u}_{i}+\alpha\nabla(\mathrm{div}_{\eta}{\bf u}_{i})&=&-\sigma_{i}{\bf u}_{i}&\mbox{in}&\Omega,\\ {\bf u}_{i}&=&0&\mbox{on}&\partial\Omega,\\ \int_{\Omega}{\bf u}_{i}\cdot{\bf u}_{j}dm&=&\delta_{ij}&\mbox{for any}&i,j.\end{array}\right. (4.1)

Since σk+1\sigma_{k+1} is the minimum value of the Rayleigh quotient (see, e.g., [15, Theorem 9.43]), we must have

σk+1Ω𝐯(𝐯+α(divη𝐯))𝑑mΩ|𝐯|2𝑑m,\sigma_{k+1}\leq-\frac{\int_{\Omega}{\bf v}\cdot(\mathscr{L}{\bf v}+\alpha\nabla(\mathrm{div}_{\eta}{\bf v}))dm}{\int_{\Omega}|{\bf v}|^{2}dm}, (4.2)

for any nonzero vector-valued function 𝐯:Ωn{\bf v}:\Omega\to\mathbb{R}^{n} satisfying

𝐯|Ω=0andΩ𝐯𝐮i𝑑m=0,for any,i=1,,k.{\bf v}|_{\partial\Omega}=0\quad\mbox{and}\quad\int_{\Omega}{\bf v}\cdot{\bf u}_{i}dm=0,\quad\mbox{for any},\quad i=1,\ldots,k.

Let us denote by aij=Ωf𝐮i𝐮j𝑑m=ajia_{ij}=\int_{\Omega}f{\bf u}_{i}\cdot{\bf u}_{j}dm=a_{ji} to consider the vector-valued functions 𝐯i{\bf v}_{i} given by

𝐯i=f𝐮ij=1kaij𝐮j,{\bf v}_{i}=f{\bf u}_{i}-\sum_{j=1}^{k}a_{ij}{\bf u}_{j}, (4.3)

so that

𝐯i|Ω=0andΩ𝐮j𝐯i𝑑m=0,for anyi,j=1,,k.{\bf v}_{i}|_{\partial\Omega}=0\quad\mbox{and}\quad\int_{\Omega}{\bf u}_{j}\cdot{\bf v}_{i}dm=0,\quad\mbox{for any}\quad i,j=1,\ldots,k. (4.4)

Then, we can take 𝐯=𝐯i{\bf v}={\bf v}_{i} in (4.2) and use formula (3.4) to obtain

σk+1𝐯i2Ω(𝐯i𝐯i+α(divη𝐯i)2)𝑑m.\sigma_{k+1}\|{\bf v}_{i}\|^{2}\leq\int_{\Omega}\Big{(}-{\bf v}_{i}\cdot\mathscr{L}{\bf v}_{i}+\alpha(\mathrm{div}_{\eta}{\bf v}_{i})^{2}\Big{)}dm. (4.5)

From (4.3) and (3), we get

𝐯i=\displaystyle\mathscr{L}{\bf v}_{i}= f𝐮i+2T(f,𝐮i)+f𝐮ij=1kaij𝐮j\displaystyle f\mathscr{L}{\bf u}_{i}+2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i}-\sum_{j=1}^{k}a_{ij}\mathscr{L}{\bf u}_{j}
=\displaystyle= f(σi𝐮iα(divη𝐮i))+2T(f,𝐮i)+f𝐮i\displaystyle f(-\sigma_{i}{\bf u}_{i}-\alpha\nabla(\mathrm{div}_{\eta}{\bf u}_{i}))+2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i}
j=1kaij(σj𝐮jα(divη𝐮j)),\displaystyle-\sum_{j=1}^{k}a_{ij}(-\sigma_{j}{\bf u}_{j}-\alpha\nabla(\mathrm{div}_{\eta}{\bf u}_{j})),

i.e.,

𝐯i=\displaystyle\mathscr{L}{\bf v}_{i}= σif𝐮i+j=1kaijσj𝐮j+2T(f,𝐮i)+f𝐮i\displaystyle-\sigma_{i}f{\bf u}_{i}+\sum_{j=1}^{k}a_{ij}\sigma_{j}{\bf u}_{j}+2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i}
αf(divη𝐮i)+αj=1kaij(divη𝐮j).\displaystyle-\alpha f\nabla(\mathrm{div}_{\eta}{\bf u}_{i})+\alpha\sum_{j=1}^{k}a_{ij}\nabla(\mathrm{div}_{\eta}{\bf u}_{j}).

Therefore,

Ω𝐯i𝐯idm=\displaystyle\int_{\Omega}-{\bf v}_{i}\cdot\mathscr{L}{\bf v}_{i}dm= σi𝐯i2Ω𝐯i(2T(f,𝐮i)+f𝐮i)𝑑m\displaystyle\sigma_{i}\|{\bf v}_{i}\|^{2}-\int_{\Omega}{\bf v}_{i}\cdot(2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i})dm
+α(Ωf𝐯i(divη𝐮i)dmj=1kaijΩ𝐯𝐢(divη𝐮j)dm).\displaystyle+\alpha\Big{(}\int_{\Omega}f{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{i})dm-\sum_{j=1}^{k}a_{ij}\int_{\Omega}{\bf v_{i}}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{j})dm\Big{)}. (4.6)

From (3.1) and (3.4)

Ωf𝐯i(divη𝐮i)dm=Ωfdivη𝐮idivη𝐯i𝑑mΩdivη𝐮if𝐯idm.\int_{\Omega}f{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{i})dm=-\int_{\Omega}f\mathrm{div}_{\eta}{\bf u}_{i}\mathrm{div}_{\eta}{\bf v}_{i}dm-\int_{\Omega}\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf v}_{i}dm.

But, from (4.3)

fdivη𝐮i=divη𝐯i+f𝐮ij=1kaijdivη𝐮j,-f\mathrm{div}_{\eta}{\bf u}_{i}=-\mathrm{div}_{\eta}{\bf v}_{i}+\nabla f\cdot{\bf u}_{i}-\sum_{j=1}^{k}a_{ij}\mathrm{div}_{\eta}{\bf u}_{j},

then

Ωf𝐯i(divη𝐮i)dm=\displaystyle\int_{\Omega}f{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{i})dm= Ω(divη𝐯i)2𝑑m+Ωdivη𝐯if𝐮idm\displaystyle-\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i})^{2}dm+\int_{\Omega}\mathrm{div}_{\eta}{\bf v}_{i}\nabla f\cdot{\bf u}_{i}dm
j=1kaijΩdivη𝐮jdivη𝐯i𝑑mΩdivη𝐮if𝐯idm\displaystyle-\sum_{j=1}^{k}a_{ij}\int_{\Omega}\mathrm{div}_{\eta}{\bf u}_{j}\mathrm{div}_{\eta}{\bf v}_{i}dm-\int_{\Omega}\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf v}_{i}dm
=\displaystyle= Ω(divη𝐯i)2𝑑m+Ωdivη𝐯if𝐮idm\displaystyle-\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i})^{2}dm+\int_{\Omega}\mathrm{div}_{\eta}{\bf v}_{i}\nabla f\cdot{\bf u}_{i}dm
+j=1kaijΩ𝐯i(divη𝐮j)dmΩdivη𝐮if𝐯idm.\displaystyle+\sum_{j=1}^{k}a_{ij}\int_{\Omega}{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{j})dm-\int_{\Omega}\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf v}_{i}dm.

Thus,

Ωf𝐯i(divη𝐮i)dmj=1kaijΩ𝐯i(divη𝐮j)dm\displaystyle\int_{\Omega}f{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{i})dm-\sum_{j=1}^{k}a_{ij}\int_{\Omega}{\bf v}_{i}\cdot\nabla(\mathrm{div}_{\eta}{\bf u}_{j})dm
=Ω(divη𝐯i)2𝑑m+Ω(divη𝐯if𝐮idivη𝐮if𝐯i)𝑑m\displaystyle=-\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i})^{2}dm+\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i}\nabla f\cdot{\bf u}_{i}-\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf v}_{i})dm
=Ω(divη𝐯i)2𝑑mΩ((f𝐮i)+divη𝐮if)𝐯i𝑑m.\displaystyle=-\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i})^{2}dm-\int_{\Omega}(\nabla(\nabla f\cdot{\bf u}_{i})+\mathrm{div}_{\eta}{\bf u}_{i}\nabla f)\cdot{\bf v}_{i}dm. (4.7)

So, replacing (4) into (4), we obtain

σi𝐯i2=\displaystyle-\sigma_{i}\|{\bf v}_{i}\|^{2}= Ω𝐯i𝐯i𝑑mΩ𝐯i(2T(f,𝐮i)+f𝐮i)𝑑m\displaystyle\int_{\Omega}{\bf v}_{i}\cdot\mathscr{L}{\bf v}_{i}dm-\int_{\Omega}{\bf v}_{i}\cdot(2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i})dm
αΩ(divη𝐯i)2𝑑mαΩ((f𝐮i)+divη𝐮if)𝐯i𝑑m.\displaystyle-\alpha\int_{\Omega}(\mathrm{div}_{\eta}{\bf v}_{i})^{2}dm-\alpha\int_{\Omega}(\nabla(\nabla f\cdot{\bf u}_{i})+\mathrm{div}_{\eta}{\bf u}_{i}\nabla f)\cdot{\bf v}_{i}dm. (4.8)

Hence, from (4.5) and (4), we have

(σk+1σi)𝐯i2\displaystyle(\sigma_{k+1}-\sigma_{i})\|{\bf v}_{i}\|^{2}\leq Ω(2T(f,𝐮i)+f𝐮i)𝐯i𝑑m\displaystyle-\int_{\Omega}(2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i})\cdot{\bf v}_{i}dm
αΩ((f𝐮i)+divη𝐮if)𝐯i𝑑m.\displaystyle-\alpha\int_{\Omega}(\nabla(\nabla f\cdot{\bf u}_{i})+\mathrm{div}_{\eta}{\bf u}_{i}\nabla f)\cdot{\bf v}_{i}dm. (4.9)

Using integration by parts formula (3.3) and (4.3), we get

Ω(2T(f,𝐮i)+f𝐮i)𝐯i𝑑m=Ω|𝐮i|2T(f,f)𝑑m2j=1kaijbij,\int_{\Omega}(2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i})\cdot{\bf v}_{i}dm=-\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm-2\sum_{j=1}^{k}a_{ij}b_{ij}, (4.10)

where

bij=Ω(T(f,𝐮i)+12f𝐮i)𝐮j𝑑m=bji.b_{ij}=\int_{\Omega}{\Big{(}}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}{\Big{)}}\cdot{\bf u}_{j}dm=-b_{ji}. (4.11)

Moreover, by straightforward computation from (3.1), (3.2) and (4.3), we have

Ω((f𝐮i)+divη𝐮if)𝐯i𝑑m\displaystyle\int_{\Omega}(\nabla(\nabla f\cdot{\bf u}_{i})+\mathrm{div}_{\eta}{\bf u}_{i}\nabla f)\cdot{\bf v}_{i}dm
=j=1kaijΩ(f𝐮idivη𝐮jdivη𝐮if𝐮j)𝑑mΩ|f𝐮i|2𝑑m.\displaystyle=\sum_{j=1}^{k}a_{ij}\int_{\Omega}(\nabla f\cdot{\bf u}_{i}\mathrm{div}_{\eta}{\bf u}_{j}-\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf u}_{j})dm-\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm. (4.12)

Putting

wi=Ω(2T(f,𝐮i)+f𝐮i)𝐯i𝑑mαΩ((f𝐮i)+divη𝐮if)𝐯i𝑑m,w_{i}=-\int_{\Omega}(2T(\nabla f,\nabla{\bf u}_{i})+\mathscr{L}f{\bf u}_{i})\cdot{\bf v}_{i}dm-\alpha\int_{\Omega}{\Big{(}}\nabla(\nabla f\cdot{\bf u}_{i})+\mathrm{div}_{\eta}{\bf u}_{i}\nabla f{\Big{)}}\cdot{\bf v}_{i}dm, (4.13)

from (4.10) and (4)

wi=\displaystyle w_{i}= Ω|𝐮i|2T(f,f)𝑑m+2j=1kaijbij\displaystyle\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm+2\sum_{j=1}^{k}a_{ij}b_{ij}
αj=1kaijΩ(f𝐮idivη𝐮jdivη𝐮if𝐮j)𝑑m+αΩ|f𝐮i|2𝑑m.\displaystyle-\alpha\sum_{j=1}^{k}a_{ij}\int_{\Omega}(\nabla f\cdot{\bf u}_{i}\mathrm{div}_{\eta}{\bf u}_{j}-\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf u}_{j})dm+\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm.

By a similar computation as in [3, Eq.  (2.9)], from (4.1) and (4.11), we get

2bij=(σiσj)aij+αΩ(f𝐮idivη𝐮jdivη𝐮if𝐮j)𝑑m,2b_{ij}=(\sigma_{i}-\sigma_{j})a_{ij}+\alpha\int_{\Omega}(\nabla f\cdot{\bf u}_{i}\mathrm{div}_{\eta}{\bf u}_{j}-\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf u}_{j})dm,

then

2j=1kaijbij=j=1k(σiσj)aij2+αj=1kaijΩ(f𝐮idivη𝐮jdivη𝐮if𝐮j)𝑑m.2\sum_{j=1}^{k}a_{ij}b_{ij}=\sum_{j=1}^{k}(\sigma_{i}-\sigma_{j})a_{ij}^{2}+\alpha\sum_{j=1}^{k}a_{ij}\int_{\Omega}(\nabla f\cdot{\bf u}_{i}\mathrm{div}_{\eta}{\bf u}_{j}-\mathrm{div}_{\eta}{\bf u}_{i}\nabla f\cdot{\bf u}_{j})dm.

Thus,

wi\displaystyle w_{i} =Ω|𝐮i|2T(f,f)𝑑m+j=1k(σiσj)aij2+αΩ|f𝐮i|2𝑑m.\displaystyle=\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm+\sum_{j=1}^{k}(\sigma_{i}-\sigma_{j})a_{ij}^{2}+\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm. (4.14)

Furthermore, from (4) and (4.13), we have

(σk+1σi)𝐯i2wi.(\sigma_{k+1}-\sigma_{i})\|{\bf v}_{i}\|^{2}\leq w_{i}. (4.15)

For any constant B>0B>0, from (4.4), (4.10) and the inequality of Cauchy-Schwarz, we infer

(σk+1\displaystyle(\sigma_{k+1} σi)2(ΩT(f,f)|𝐮i|2dm+2j=1kaijbij)\displaystyle-\sigma_{i})^{2}\Big{(}\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm+2\sum_{j=1}^{k}a_{ij}b_{ij}\Big{)}
=\displaystyle= (σk+1σi)2{2Ω(T(f,𝐮i)+12f𝐮ij=1kbij𝐮j)𝐯i𝑑m}\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}\Bigg{\{}-2\int_{\Omega}\Big{(}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}-\sum_{j=1}^{k}b_{ij}{\bf u}_{j}\Big{)}\cdot{\bf v}_{i}dm\Bigg{\}}
\displaystyle\leq 2(σk+1σi)2𝐯iT(f,𝐮i)+12f𝐮ij=1kbij𝐮j\displaystyle 2(\sigma_{k+1}-\sigma_{i})^{2}\|{\bf v}_{i}\|\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}-\sum_{j=1}^{k}b_{ij}{\bf u}_{j}\Big{\|}
\displaystyle\leq (σk+1σi)3B𝐯i2+σk+1σiBT(f,𝐮i)+12f𝐮ij=1kbij𝐮j2,\displaystyle(\sigma_{k+1}-\sigma_{i})^{3}B\|{\bf v}_{i}\|^{2}+\frac{\sigma_{k+1}-\sigma_{i}}{B}\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}-\sum_{j=1}^{k}b_{ij}{\bf u}_{j}\Big{\|}^{2},

hence, using (4.14) and (4.15), we obtain

(σk+1\displaystyle(\sigma_{k+1} σi)2(ΩT(f,f)|𝐮i|2dm+2j=1kaijbij)\displaystyle-\sigma_{i})^{2}\Big{(}\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm+2\sum_{j=1}^{k}a_{ij}b_{ij}\Big{)}
\displaystyle\leq (σk+1σi)2B(Ω|𝐮i|2T(f,f)𝑑m+j=1k(σiσj)aij2+αΩ|f𝐮i|2𝑑m)\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}B\Big{(}\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm+\sum_{j=1}^{k}(\sigma_{i}-\sigma_{j})a_{ij}^{2}+\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm\Big{)}
+σk+1σiB(T(f,𝐮i)+12f𝐮i2j=1kbij2).\displaystyle+\frac{\sigma_{k+1}-\sigma_{i}}{B}\Big{(}\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}\Big{\|}^{2}-\sum_{j=1}^{k}b_{ij}^{2}\Big{)}.

Summing over ii from 11 to kk, we obtain

i=1k\displaystyle\sum_{i=1}^{k} (σk+1σi)2(ΩT(f,f)|𝐮i|2𝑑m+2j=1kaijbij)\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}\Big{(}\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm+2\sum_{j=1}^{k}a_{ij}b_{ij}\Big{)}
\displaystyle\leq i=1k(σk+1σi)2B(Ω|𝐮i|2T(f,f)𝑑m+j=1k(σiσj)aij2+αΩ|f𝐮i|2𝑑m)\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}B\Big{(}\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm+\sum_{j=1}^{k}(\sigma_{i}-\sigma_{j})a_{ij}^{2}+\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm\Big{)}
+i=1kσk+1σiB(T(f,𝐮i)+12f𝐮i2j=1kbij2).\displaystyle+\sum_{i=1}^{k}\frac{\sigma_{k+1}-\sigma_{i}}{B}\Big{(}\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}\Big{\|}^{2}-\sum_{j=1}^{k}b_{ij}^{2}\Big{)}. (4.16)

Since aij=ajia_{ij}=a_{ji} and bij=bjib_{ij}=-b_{ji}, we have

2i,j=1k(σk+1σi)2aijbij=2i,j=1k(σk+1σi)(σiσj)aijbij,\displaystyle 2\sum_{i,j=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}a_{ij}b_{ij}=-2\sum_{i,j=1}^{k}(\sigma_{k+1}-\sigma_{i})(\sigma_{i}-\sigma_{j})a_{ij}b_{ij},

and

i,j=1k(σk+1σi)2(σiσj)aij2=i,j=1k(σk+1σi)(σiσj)2aij2.\displaystyle\sum_{i,j=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}(\sigma_{i}-\sigma_{j})a_{ij}^{2}=-\sum_{i,j=1}^{k}(\sigma_{k+1}-\sigma_{i})(\sigma_{i}-\sigma_{j})^{2}a_{ij}^{2}.

Therefore, from (4), we get

i=1k\displaystyle\sum_{i=1}^{k} (σk+1σi)2ΩT(f,f)|𝐮i|2𝑑m\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm
\displaystyle\leq i=1k(σk+1σi)2B(Ω|𝐮i|2T(f,f)𝑑m+αΩ|f𝐮i|2𝑑m)\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}B\Big{(}\int_{\Omega}|{\bf u}_{i}|^{2}T(\nabla f,\nabla f)dm+\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm\Big{)}
+i=1kσk+1σiBT(f,𝐮i)+12f𝐮i2,\displaystyle+\sum_{i=1}^{k}\frac{\sigma_{k+1}-\sigma_{i}}{B}\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}\Big{\|}^{2},

whence

i=1k(σk+1σi)2((1B)ΩT(f,f)|𝐮i|2𝑑mBαΩ|f𝐮i|2𝑑m)\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\Big{(}(1-B)\int_{\Omega}T(\nabla f,\nabla f)|{\bf u}_{i}|^{2}dm-B\alpha\int_{\Omega}|\nabla f\cdot{\bf u}_{i}|^{2}dm\Big{)}
1Bi=1k(σk+1σi)T(f,𝐮i)+12f𝐮i2.\displaystyle\leq\frac{1}{B}\sum_{i=1}^{k}{(\sigma_{k+1}-\sigma_{i})}\Big{\|}T(\nabla f,\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}f{\bf u}_{i}\Big{\|}^{2}.

This finishes the proof of Lemma 4.1. ∎

The proof of the next lemma follows the steps of the proof of Proposition 2 in Gomes and Miranda [9] with appropriate adaptations for vector-valued functions from Ω\Omega to n\mathbb{R}^{n}.

Lemma 4.2.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, σi\sigma_{i} be the ii-th eigenvalue of Problem (1.2) and 𝐮i{\bf u}_{i} be a normalized vector-valued eigenfunction corresponding to σi\sigma_{i}. Then, for any positive integer kk, we get

i=1k(σk+1\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}- σi)24(nδ+α)n2ε2i=1k(σk+1σi){14Ω|𝐮i|2|tr(T)T(η)|2dm\displaystyle\sigma_{i})^{2}\leq\frac{4(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm
+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i))dm+T(𝐮i)2}.\displaystyle+\int_{\Omega}{\bf u}_{i}\cdot{\Big{(}}T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}){\Big{)}}dm+\|T(\nabla{\bf u}_{i})\|^{2}\Big{\}}.
Proof.

Let {xβ}β=1n\{x_{\beta}\}_{\beta=1}^{n} be the coordinate functions of n\mathbb{R}^{n}, then taking f=xβf=x_{\beta} in Lemma 4.1 and summing over β\beta from 1 to nn, we get

i=1k(σk+1\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}- σi)2β=1n{(1B)ΩT(xβ,xβ)|𝐮i|2dmBαΩ|xβ𝐮i|2dm}\displaystyle\sigma_{i})^{2}\sum_{\beta=1}^{n}\Bigg{\{}(1-B)\int_{\Omega}T(\nabla x_{\beta},\nabla x_{\beta})|{\bf u}_{i}|^{2}dm-B\alpha\int_{\Omega}|\nabla x_{\beta}\cdot{\bf u}_{i}|^{2}dm\Bigg{\}}
\displaystyle\leq 1Bi=1k(σk+1σi)β=1nT(xβ,𝐮i)+12xβ𝐮i2\displaystyle\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\sum_{\beta=1}^{n}\Big{\|}T(\nabla x_{\beta},\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}x_{\beta}{\bf u}_{i}\Big{\|}^{2}
=\displaystyle= 1Bi=1k(σk+1σi)Ωβ=1n|T(xβ,𝐮i)+12divη(T(xβ))𝐮i|2dm.\displaystyle\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\int_{\Omega}\sum_{\beta=1}^{n}\Big{|}T(\nabla x_{\beta},\nabla{\bf u}_{i})+\frac{1}{2}\mathrm{div}_{\eta}(T(\nabla x_{\beta})){\bf u}_{i}\Big{|}^{2}dm.

Therefore,

i=1k(σk+1\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}- σi)2β=1n{(1B)ΩT(xβ,xβ)|𝐮i|2dmBαΩ|xβ𝐮i|2dm}\displaystyle\sigma_{i})^{2}\sum_{\beta=1}^{n}\Bigg{\{}(1-B)\int_{\Omega}T(\nabla x_{\beta},\nabla x_{\beta})|{\bf u}_{i}|^{2}dm-B\alpha\int_{\Omega}|\nabla x_{\beta}\cdot{\bf u}_{i}|^{2}dm\Bigg{\}}
\displaystyle\leq 1Bi=1k(σk+1σi)Ωβ=1n{|𝐮i|24(divη(T(xβ)))2\displaystyle\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\int_{\Omega}\sum_{\beta=1}^{n}\Bigg{\{}\frac{|{\bf u}_{i}|^{2}}{4}(\mathrm{div}_{\eta}(T(\nabla x_{\beta})))^{2}
+𝐮i(divη(T(xβ))T(xβ,𝐮i))+|T(xβ,𝐮i)|2}dm.\displaystyle+{\bf u}_{i}\cdot\Big{(}\mathrm{div}_{\eta}(T(\nabla x_{\beta}))T(\nabla x_{\beta},\nabla{\bf u}_{i})\Big{)}+|T(\nabla x_{\beta},\nabla{\bf u}_{i})|^{2}{\Bigg{\}}}dm. (4.17)

By straightforward computation, we have

β=1nT(xβ,xβ)=β=1nT(eβ),eβ=tr(T)andβ=1n|xβ𝐮i|2=|𝐮i|2.\displaystyle\quad\quad\sum_{\beta=1}^{n}T(\nabla x_{\beta},\nabla x_{\beta})=\sum_{\beta=1}^{n}\langle T(e_{\beta}),e_{\beta}\rangle=\mathrm{tr}(T)\quad\mbox{and}\quad\sum_{\beta=1}^{n}|\nabla x_{\beta}\cdot{\bf u}_{i}|^{2}=|{\bf u}_{i}|^{2}.

Similarly to the calculations in [9, Eq. (3.23)], we obtain

β=1n(divη(T(xβ)))2=|tr(T)T(η)|2.\displaystyle\sum_{\beta=1}^{n}(\mathrm{div}_{\eta}(T(\nabla x_{\beta})))^{2}=|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}.

From  [9, Eq. (3.24)], for all k=1,,nk=1,\ldots,n, we get

β=1ndivη(T(xβ))T(xβ,uik)=tr(T),T(uik)T(η),T(uik),\sum_{\beta=1}^{n}\mathrm{div}_{\eta}(T(\nabla x_{\beta}))T(\nabla x_{\beta},\nabla u_{i}^{k})=\langle\mathrm{tr}(\nabla T),T(\nabla u_{i}^{k})\rangle-\langle T(\nabla\eta),T(\nabla u_{i}^{k})\rangle,

then, using (3), we obtain

β=1ndivη(T(xβ))T(xβ,𝐮i)=\displaystyle\sum_{\beta=1}^{n}\mathrm{div}_{\eta}(T(\nabla x_{\beta}))T(\nabla x_{\beta},\nabla{\bf u}_{i})=
=(β=1ndivη(T(xβ))T(xβ,ui1),,β=1ndivη(T(xβ))T(xβ,uin))\displaystyle=\Big{(}\sum_{\beta=1}^{n}\mathrm{div}_{\eta}(T(\nabla x_{\beta}))T(\nabla x_{\beta},\nabla u_{i}^{1}),\ldots,\sum_{\beta=1}^{n}\mathrm{div}_{\eta}(T(\nabla x_{\beta}))T(\nabla x_{\beta},\nabla u_{i}^{n})\Big{)}
=(tr(T),T(ui1)T(η),T(ui1),,tr(T),T(uin)T(η),T(uin))\displaystyle=(\langle\mathrm{tr}(\nabla T),T(\nabla u_{i}^{1})\rangle-\langle T(\nabla\eta),T(\nabla u_{i}^{1})\rangle,\ldots,\langle\mathrm{tr}(\nabla T),T(\nabla u_{i}^{n})\rangle-\langle T(\nabla\eta),T(\nabla u_{i}^{n})\rangle)
=(tr(T),T(ui1),,tr(T),T(uin))(T(η),T(ui1),,T(η),T(uin))\displaystyle=(\langle\mathrm{tr}(\nabla T),T(\nabla u_{i}^{1})\rangle,\ldots,\langle\mathrm{tr}(\nabla T),T(\nabla u_{i}^{n})\rangle)-(\langle T(\nabla\eta),T(\nabla u_{i}^{1})\rangle,\ldots,\langle T(\nabla\eta),T(\nabla u_{i}^{n})\rangle)
=T(tr(T),𝐮i)T(T(η),𝐮i).\displaystyle=T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}).

Moreover,

β=1n\displaystyle\sum_{\beta=1}^{n} |T(xβ,𝐮i)|2=β=1n|T(eβ,𝐮i)|2=β=1n|(eβ,T(ui1),,eβ,T(uin))|2\displaystyle|T(\nabla x_{\beta},\nabla{\bf u}_{i})|^{2}=\sum_{\beta=1}^{n}|T(e_{\beta},\nabla{\bf u}_{i})|^{2}=\sum_{\beta=1}^{n}|\big{(}\langle e_{\beta},T(\nabla u_{i}^{1})\rangle,\ldots,\langle e_{\beta},T(\nabla u_{i}^{n})\rangle\big{)}|^{2}
=β=1neβ,T(ui1)2++β=1neβ,T(uin)2=j=1n|T(uij)|2=|T(𝐮i)|2.\displaystyle=\sum_{\beta=1}^{n}\langle e_{\beta},T(\nabla u_{i}^{1})\rangle^{2}+\cdots+\sum_{\beta=1}^{n}\langle e_{\beta},T(\nabla u_{i}^{n})\rangle^{2}=\sum_{j=1}^{n}|T(\nabla u_{i}^{j})|^{2}=|T(\nabla{\bf u}_{i})|^{2}.

Substituting the previous equalities into (4) and remembering that T(𝐮i)2=Ω|T(𝐮i)|2𝑑m\|T(\nabla{\bf u}_{i})\|^{2}=\int_{\Omega}|T(\nabla{\bf u}_{i})|^{2}dm, we get

i=1k\displaystyle\sum_{i=1}^{k} (σk+1σi)2[(1B)Ωtr(T)|𝐮i|2𝑑mBα]\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}\Big{[}(1-B)\int_{\Omega}\mathrm{tr}(T)|{\bf u}_{i}|^{2}dm-B\alpha\Big{]}
\displaystyle\leq 1Bi=1k(σk+1σi){14Ω|𝐮i|2|tr(T)T(η)|2dm\displaystyle\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i}){\Bigg{\{}}\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm
+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i))dm+T(𝐮i)2}.\displaystyle+\int_{\Omega}{\bf u}_{i}\cdot{\Big{(}}T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}){\Big{)}}dm+\|T(\nabla{\bf u}_{i})\|^{2}{\Bigg{\}}}. (4.18)

Since εITδI\varepsilon I\leq T\leq\delta I, then nεtr(T)nδn\varepsilon\leq\mathrm{tr}(T)\leq n\delta. Hence from (4)

i=1k\displaystyle\sum_{i=1}^{k} (σk+1σi)2[nε(nδ+α)B]\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}[n\varepsilon-(n\delta+\alpha)B]
\displaystyle\leq 1Bi=1k(σk+1σi){14Ω|𝐮i|2|tr(T)T(η)|2dm\displaystyle\frac{1}{B}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i}){\Bigg{\{}}\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm
+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i))dm+T(𝐮i)2}.\displaystyle+\int_{\Omega}{\bf u}_{i}\cdot{\Big{(}}T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}){\Big{)}}dm+\|T(\nabla{\bf u}_{i})\|^{2}{\Bigg{\}}}. (4.19)

Furthermore, since BB is arbitrary positive constant, putting

B=\displaystyle B= {i=1k(σk+1σi)(nδ+α)i=1k(σk+1σi)2[14Ω|𝐮i|2|tr(T)T(η)|2dm\displaystyle{\Big{\{}}\frac{\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})}{{(n\delta+\alpha)\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}}}{\Big{[}}\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm
+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i))dm+T(𝐮i)2]}12\displaystyle+\int_{\Omega}{\bf u}_{i}\cdot(T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}))dm+\|T(\nabla{\bf u}_{i})\|^{2}{\Big{]}}{\Big{\}}}^{\frac{1}{2}}

into (4), we obtain the required inequality and complete the proof of Lemma 4.2. ∎

With these considerations in mind, we can rewrite the previous lemma in a more convenient way for us.

Lemma 4.3.

Under the same setup as in Lemma 4.2, we get

i=1k\displaystyle\sum_{i=1}^{k} (σk+1σi)24(nδ+α)n2ε2i=1k(σk+1σi){T(𝐮i)2+C\displaystyle(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\|T(\nabla{\bf u}_{i})\|^{2}+C
+14Ω|𝐮i|2tr(T),tr(T)2T(η)dm+Ω𝐮iT(tr(T),𝐮i)dm},\displaystyle+\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),\mathrm{tr}(\nabla T)-2T(\nabla\eta)\rangle dm+\int_{\Omega}{\bf u}_{i}\cdot T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})dm\Big{\}},

where C=supΩ{12div(T2(η))14|T(η)|2}C=\sup_{\Omega}\big{\{}\frac{1}{2}\mathrm{div}(T^{2}(\nabla\eta))-\frac{1}{4}|T(\nabla\eta)|^{2}\big{\}} has been chosen such that the term on the right-hand side must be positive.

Proof.

We make use of Lemma 4.2. For this, we must notice that

|tr(T)T(η)|2=|tr(T)|22tr(T),T(η)+|T(η)|2,\displaystyle|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}=|\mathrm{tr}(\nabla T)|^{2}-2\langle\mathrm{tr}(\nabla T),T(\nabla\eta)\rangle+|T(\nabla\eta)|^{2},

hence

14\displaystyle\frac{1}{4} Ω|𝐮i|2|tr(T)T(η)|2dm+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i)))dm\displaystyle\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm+\int_{\Omega}{\bf u}_{i}\cdot(T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i})))dm
=\displaystyle= 14Ω|𝐮i|2|T(η)|2𝑑mΩ𝐮iT(T(η),𝐮i)𝑑m+14Ω|𝐮i|2|tr(T)|2𝑑m\displaystyle\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|T(\nabla\eta)|^{2}dm-\int_{\Omega}{\bf u}_{i}\cdot T(T(\nabla\eta),\nabla{\bf u}_{i})dm+\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)|^{2}dm
12Ω|𝐮i|2tr(T),T(η)𝑑m+Ω𝐮iT(tr(T),𝐮i)𝑑m.\displaystyle-\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),T(\nabla\eta)\rangle dm+\int_{\Omega}{\bf u}_{i}\cdot T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})dm. (4.20)

Since 𝐮i|Ω=0{\bf u}_{i}|_{\partial\Omega}=0 by Eq. (3) and the divergence theorem, we have

\displaystyle- Ω𝐮iT(T(η),𝐮i)𝑑m\displaystyle\int_{\Omega}{\bf u}_{i}\cdot T(T(\nabla\eta),\nabla{\bf u}_{i})dm
=Ωui1T2(η),ui1𝑑mΩuinT2(η),uin𝑑m\displaystyle=-\int_{\Omega}u_{i}^{1}\langle T^{2}(\nabla\eta),\nabla u_{i}^{1}\rangle dm-\cdots-\int_{\Omega}u_{i}^{n}\langle T^{2}(\nabla\eta),\nabla u_{i}^{n}\rangle dm
=12ΩT2(η),(ui1)2dm12ΩT2(η),(uin)2dm\displaystyle=-\frac{1}{2}\int_{\Omega}\langle T^{2}(\nabla\eta),\nabla(u_{i}^{1})^{2}\rangle dm-\cdots-\frac{1}{2}\int_{\Omega}\langle T^{2}(\nabla\eta),\nabla(u_{i}^{n})^{2}\rangle dm
=12Ω(ui1)2divη(T2(η))𝑑m++12Ω(uin)2divη(T2(η))𝑑m\displaystyle=\frac{1}{2}\int_{\Omega}(u_{i}^{1})^{2}\mathrm{div}_{\eta}(T^{2}(\nabla\eta))dm+\cdots+\frac{1}{2}\int_{\Omega}(u_{i}^{n})^{2}\mathrm{div}_{\eta}(T^{2}(\nabla\eta))dm
=12Ω|𝐮i|2divη(T2(η))𝑑m.\displaystyle=\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}\mathrm{div}_{\eta}(T^{2}(\nabla\eta))dm.

Substituting the previous equation in Eq. (4), we get

14\displaystyle\frac{1}{4} Ω|𝐮i|2|tr(T)T(η)|2𝑑m+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i))𝑑m\displaystyle\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm+\int_{\Omega}{\bf u}_{i}\cdot(T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i}))dm
=\displaystyle= Ω|𝐮i|2(14|T(η)|2+12divη(T2(η)))𝑑m+14Ω|𝐮i|2tr(T),tr(T)𝑑m\displaystyle\int_{\Omega}|{\bf u}_{i}|^{2}{\Big{(}}\frac{1}{4}|T(\nabla\eta)|^{2}+\frac{1}{2}\mathrm{div}_{\eta}(T^{2}(\nabla\eta)){\Big{)}}dm+\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),\mathrm{tr}(\nabla T)\rangle dm
12Ω|𝐮i|2tr(T),T(η)𝑑m+Ω𝐮iT(tr(T),𝐮i)𝑑m.\displaystyle-\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),T(\nabla\eta)\rangle dm+\int_{\Omega}{\bf u}_{i}\cdot T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})dm.

By setting

C=supΩ{14|T(η)|2+12divη(T2(η))}=supΩ{12div(T2(η))14|T(η)|2}C=\sup_{\Omega}\Big{\{}\frac{1}{4}|T(\nabla\eta)|^{2}+\frac{1}{2}\mathrm{div}_{\eta}(T^{2}(\nabla\eta))\Big{\}}=\sup_{\Omega}\Big{\{}\frac{1}{2}\mathrm{div}(T^{2}(\nabla\eta))-\frac{1}{4}|T(\nabla\eta)|^{2}\Big{\}}

and by the previous equality, we have

14\displaystyle\frac{1}{4} Ω|𝐮i|2|tr(T)T(η)|2dm+Ω𝐮i(T(tr(T),𝐮i)T(T(η),𝐮i)))dm\displaystyle\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)-T(\nabla\eta)|^{2}dm+\int_{\Omega}{\bf u}_{i}\cdot(T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})-T(T(\nabla\eta),\nabla{\bf u}_{i})))dm
C+14Ω|𝐮i|2tr(T),tr(T)2T(η)𝑑m+Ω𝐮iT(tr(T),𝐮i)𝑑m.\displaystyle\leq C+\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),\mathrm{tr}(\nabla T)-2T(\nabla\eta)\rangle dm+\int_{\Omega}{\bf u}_{i}\cdot T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})dm. (4.21)

Replacing Inequality (4) into Lemma 4.2, we complete the proof of Lemma 4.3. ∎

Now, we are in a position to give the proof of the two theorems of this paper.

5. Proof of Theorems 1.1 and 1.2

5.1. Proof of Theorem 1.1

Proof.

The proof is a consequence of Lemma 4.3. We begin by computing

14Ω|𝐮i|2tr(T),tr(T)2T(η)𝑑m=\displaystyle\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),\mathrm{tr}(\nabla T)-2T(\nabla\eta)\rangle dm= 14Ω|𝐮i|2|tr(T)|2𝑑m\displaystyle\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)|^{2}dm
12Ω|𝐮i|2tr(T),T(η)𝑑m.\displaystyle-\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),T(\nabla\eta)\rangle dm.

Since T0=supΩ|tr(T)|T_{0}=\sup_{\Omega}|\mathrm{tr}(\nabla T)| and η0=supΩ|η|\eta_{0}=\sup_{\Omega}|\nabla\eta|, we have

14Ω|𝐮i|2|tr(T)|2𝑑m14T02Ω|𝐮i|2𝑑m=14T02,\displaystyle\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)|^{2}dm\leq\frac{1}{4}T_{0}^{2}\int_{\Omega}|{\bf u}_{i}|^{2}dm=\frac{1}{4}T_{0}^{2},

and using (3.8) we get

12Ω|𝐮i|2tr(T),T(η)𝑑m\displaystyle-\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),T(\nabla\eta)\rangle dm 12Ω|𝐮i|2|tr(T)||T(η)|𝑑m\displaystyle\leq\frac{1}{2}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)||T(\nabla\eta)|dm
δ2Ω|𝐮i|2|tr(T)||η|𝑑m\displaystyle\leq\frac{\delta}{2}\int_{\Omega}|{\bf u}_{i}|^{2}|\mathrm{tr}(\nabla T)||\nabla\eta|dm
δ2T0η0.\displaystyle\leq\frac{\delta}{2}T_{0}\eta_{0}.

Then,

14Ω|𝐮i|2tr(T),tr(T)2T(η)𝑑m14T02+δ2T0η0.\displaystyle\frac{1}{4}\int_{\Omega}|{\bf u}_{i}|^{2}\langle\mathrm{tr}(\nabla T),\mathrm{tr}(\nabla T)-2T(\nabla\eta)\rangle dm\leq\frac{1}{4}T_{0}^{2}+\frac{\delta}{2}T_{0}\eta_{0}. (5.1)

Furthermore,

Ω𝐮iT(tr(T),𝐮i)𝑑m\displaystyle\int_{\Omega}{\bf u}_{i}\cdot T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})dm (Ω|𝐮i|2𝑑m)12(Ω|T(tr(T),𝐮i)|2𝑑m)12\displaystyle\leq\Big{(}\int_{\Omega}|{\bf u}_{i}|^{2}dm\Big{)}^{\frac{1}{2}}\Big{(}\int_{\Omega}|T(\mathrm{tr}(\nabla T),\nabla{\bf u}_{i})|^{2}dm\Big{)}^{\frac{1}{2}}
T0(Ω|T(𝐮i)|2𝑑m)12=T0T(𝐮i).\displaystyle\leq T_{0}\Big{(}\int_{\Omega}|T(\nabla{\bf u}_{i})|^{2}dm\Big{)}^{\frac{1}{2}}=T_{0}\|T(\nabla{\bf u}_{i})\|. (5.2)

Substituting (5.1) and (5.1) into Lemma 4.3, we obtain

i=1k(σk+1σi)2\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq 4(nδ+α)n2ε2i=1k(σk+1σi){T(𝐮i)2+14T02+T0T(𝐮i)\displaystyle\frac{4(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\|T(\nabla{\bf u}_{i})\|^{2}+\frac{1}{4}T_{0}^{2}+T_{0}\|T(\nabla{\bf u}_{i})\|
+δ2T0η0+C}.\displaystyle+\frac{\delta}{2}T_{0}\eta_{0}+C\Big{\}}.

Moreover, from the proof of Lemma 4.2 and Lemma 4.3, we can see that

0<β=1nT(xβ,𝐮i)+12xβ𝐮i2\displaystyle 0<\sum_{\beta=1}^{n}\Big{\|}T(\nabla x_{\beta},\nabla{\bf u}_{i})+\frac{1}{2}\mathscr{L}x_{\beta}{\bf u}_{i}\Big{\|}^{2}\leq {T(𝐮i)2+14T02+T0T(𝐮i)\displaystyle\Big{\{}\|T(\nabla{\bf u}_{i})\|^{2}+\frac{1}{4}T_{0}^{2}+T_{0}\|T(\nabla{\bf u}_{i})\|
+δ2T0η0+C}\displaystyle+\frac{\delta}{2}T_{0}\eta_{0}+C\Big{\}}
=\displaystyle= (T(𝐮i)+12T0)2+C0,\displaystyle\Big{(}\|T(\nabla{\bf u}_{i})\|+\frac{1}{2}T_{0}\Big{)}^{2}+C_{0},

where C0=δ2T0η0+CC_{0}=\frac{\delta}{2}T_{0}\eta_{0}+C and {xβ}β=1n\{x_{\beta}\}_{\beta=1}^{n} are the canonical coordinate functions of n\mathbb{R}^{n}. Thus, we get

i=1k(σk+1σi)24(nδ+α)n2ε2i=1k(σk+1σi){(T(𝐮i)+12T0)2+C0}.\displaystyle\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\Big{(}\|T(\nabla{\bf u}_{i})\|+\frac{1}{2}T_{0}\Big{)}^{2}+C_{0}\Big{\}}. (5.3)

From (1.2), (3.3) and (3.4) we obtain

σi=ΩT(𝐮i)𝐮idm+αdivη𝐮i2.\sigma_{i}=\int_{\Omega}T(\nabla{\bf u}_{i})\cdot\nabla{\bf u}_{i}dm+\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2}.

Since there exist positive real numbers ε\varepsilon and δ\delta such that εITδI\varepsilon I\leq T\leq\delta I, from the previous inequality and (3.7), we get

T(𝐮i)2δΩT(𝐮i)𝐮idm=δ(σiαdivη𝐮i2).\|T(\nabla{\bf u}_{i})\|^{2}\leq\delta\int_{\Omega}T(\nabla{\bf u}_{i})\cdot\nabla{\bf u}_{i}dm=\delta(\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2}). (5.4)

Therefore, from (5.3) and (5.4) we obtain

i=1k(σk+1σi)24(nδ+α)n2ε2i=1k(σk+1σi){[δ(σiαdivη𝐮i2)12+T0]2+C0},\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{\{}\Big{[}\sqrt{\delta}(\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2})^{\frac{1}{2}}+T_{0}\Big{]}^{2}+C_{0}\Big{\}},

which is enough to complete the proof. ∎

5.2. Proof of Theorem 1.2

Proof.

Let {xβ}β=1n\{x_{\beta}\}_{\beta=1}^{n} be the standard coordinate functions of n\mathbb{R}^{n}. Let us consider the matrix D=(dij)n×nD=(d_{ij})_{n\times n} where

dij:=Ωxi𝐮1𝐮j+1𝑑m.d_{ij}:=\int_{\Omega}x_{i}{\bf u}_{1}\cdot{\bf u}_{j+1}dm.

Using the orthogonalization of Gram and Schmidt, we know that there exists an upper triangle matrix R=(rij)n×nR=(r_{ij})_{n\times n} and an orthogonal matrix S=(sij)n×nS=(s_{ij})_{n\times n} such that R=SDR=SD, namely

rij=k=1nsikdkj=k=1nsikΩxk𝐮1𝐮j+1𝑑m=Ω(k=1nsikxk)𝐮1𝐮j+1𝑑m=0,r_{ij}=\sum_{k=1}^{n}s_{ik}d_{kj}=\sum_{k=1}^{n}s_{ik}\int_{\Omega}x_{k}{\bf u}_{1}\cdot{\bf u}_{j+1}dm=\int_{\Omega}\Big{(}\sum_{k=1}^{n}s_{ik}x_{k}\Big{)}{\bf u}_{1}\cdot{\bf u}_{j+1}dm=0,

for 1j<in1\leq j<i\leq n. Putting yi=k=1nsikxky_{i}=\sum_{k=1}^{n}s_{ik}x_{k}, we have

Ωyi𝐮1𝐮j+1𝑑m=0for1j<in.\int_{\Omega}y_{i}{\bf u}_{1}\cdot{\bf u}_{j+1}dm=0\quad\mbox{for}\quad 1\leq j<i\leq n.

Let us denote by ai=Ωyi|𝐮1|2𝑑ma_{i}=\int_{\Omega}y_{i}|{\bf u}_{1}|^{2}dm to consider the vector-valued functions 𝐰i{\bf w}_{i} given by

𝐰i=(yiai)𝐮1,{\bf w}_{i}=(y_{i}-a_{i}){\bf u}_{1}, (5.5)

so that

𝐰i|Ω=0andΩ𝐰i𝐮j+1𝑑m=0,for anyj=1,,i1.{\bf w}_{i}|_{\partial\Omega}=0\quad\mbox{and}\quad\int_{\Omega}{\bf w}_{i}\cdot{\bf u}_{j+1}dm=0,\quad\mbox{for any}\quad j=1,\ldots,i-1.

Then, we can take 𝐯=𝐰i{\bf v}={\bf w}_{i} in (4.2) and to use formula (3.4) to obtain

σi+1𝐰i2Ω(𝐰i𝐰i+α(divη𝐰i)2)𝑑m.\sigma_{i+1}\|{\bf w}_{i}\|^{2}\leq\int_{\Omega}(-{\bf w}_{i}\cdot\mathscr{L}{\bf w}_{i}+\alpha(\mathrm{div}_{\eta}{\bf w}_{i})^{2})dm. (5.6)

Using (3) we get

Ω𝐰i𝐰i𝑑m=Ω𝐰i[(yiai)𝐮1+𝐮1yi+2T(yi,𝐮1)]𝑑m\displaystyle-\int_{\Omega}{\bf w}_{i}\cdot\mathscr{L}{\bf w}_{i}dm=-\int_{\Omega}{\bf w}_{i}\cdot[(y_{i}-a_{i})\mathscr{L}{\bf u}_{1}+{\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1})]dm
=Ω𝐰i[(yiai)(σ1𝐮1αdivη𝐮1)+𝐮1yi+2T(yi,𝐮1)]𝑑m\displaystyle=-\int_{\Omega}{\bf w}_{i}\cdot[(y_{i}-a_{i})(-\sigma_{1}{\bf u}_{1}-\alpha\nabla\mathrm{div}_{\eta}{\bf u}_{1})+{\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1})]dm
=σ1𝐰i2+αΩ(yiai)𝐰idivη𝐮1dmΩ𝐰i(𝐮1yi+2T(yi,𝐮1))𝑑m.\displaystyle=\sigma_{1}\|{\bf w}_{i}\|^{2}+\alpha\int_{\Omega}(y_{i}-a_{i}){\bf w}_{i}\cdot\nabla\mathrm{div}_{\eta}{\bf u}_{1}dm-\int_{\Omega}{\bf w}_{i}\cdot({\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1}))dm. (5.7)

Using (3.1) and (3.2), by a computation analogous to (4), we obtain

αΩ(yiai)𝐰idivη𝐮1dm=\displaystyle\alpha\int_{\Omega}(y_{i}-a_{i}){\bf w}_{i}\cdot\nabla\mathrm{div}_{\eta}{\bf u}_{1}dm= αΩ(divη𝐰i)2𝑑m\displaystyle-\alpha\int_{\Omega}(\mathrm{div}_{\eta}{\bf w}_{i})^{2}dm
αΩ((yi𝐮1)+divη𝐮1yi)𝐰i𝑑m.\displaystyle-\alpha\int_{\Omega}(\nabla(\nabla y_{i}\cdot{\bf u}_{1})+\mathrm{div}_{\eta}{\bf u}_{1}\nabla y_{i})\cdot{\bf w}_{i}dm.

Substituting the previous equality into (5.2), we get

Ω(𝐰i𝐰i+α(divη𝐰i)2)𝑑m=\displaystyle\int_{\Omega}(-{\bf w}_{i}\cdot\mathscr{L}{\bf w}_{i}+\alpha(\mathrm{div}_{\eta}{\bf w}_{i})^{2})dm= σ1𝐰i2αΩ((yi𝐮1)+divη𝐮1yi)𝐰i𝑑m\displaystyle\sigma_{1}\|{\bf w}_{i}\|^{2}-\alpha\int_{\Omega}(\nabla(\nabla y_{i}\cdot{\bf u}_{1})+\mathrm{div}_{\eta}{\bf u}_{1}\nabla y_{i})\cdot{\bf w}_{i}dm
Ω𝐰i(𝐮1yi+2T(yi,𝐮1))𝑑m.\displaystyle-\int_{\Omega}{\bf w}_{i}\cdot({\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1}))dm. (5.8)

Replacing (5.2) into (5.6), we have

(σi+1σ1)𝐰i2\displaystyle(\sigma_{i+1}-\sigma_{1})\|{\bf w}_{i}\|^{2}\leq Ω𝐰i(𝐮1yi+2T(yi,𝐮1))𝑑m\displaystyle-\int_{\Omega}{\bf w}_{i}\cdot({\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1}))dm
αΩ𝐰i((yi𝐮1)+divη𝐮1yi)𝑑m.\displaystyle-\alpha\int_{\Omega}{\bf w}_{i}\cdot(\nabla(\nabla y_{i}\cdot{\bf u}_{1})+\mathrm{div}_{\eta}{\bf u}_{1}\nabla y_{i})dm. (5.9)

By a straightforward computation, we have, from (3.1), (3.2), (3.3) and (5.5),

Ω𝐰i(𝐮1yi+2T(yi,𝐮1))𝑑m=Ω|𝐮1|2T(yi,yi)𝑑m.\displaystyle-\int_{\Omega}{\bf w}_{i}\cdot({\bf u}_{1}\mathscr{L}y_{i}+2T(\nabla y_{i},\nabla{\bf u}_{1}))dm=\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm. (5.10)
αΩ𝐰i((yi𝐮1)+divη𝐮1yi)𝑑m=αΩ|yi𝐮1|2𝑑m.\displaystyle-\alpha\int_{\Omega}{\bf w}_{i}\cdot(\nabla(\nabla y_{i}\cdot{\bf u}_{1})+\mathrm{div}_{\eta}{\bf u}_{1}\nabla y_{i})dm=\alpha\int_{\Omega}|\nabla y_{i}\cdot{\bf u}_{1}|^{2}dm. (5.11)

Therefore, substituting (5.10) and (5.11) into (5.2) we obtain

(σi+1σ1)𝐰i2Ω|𝐮1|2T(yi,yi)𝑑m+αΩ|yi𝐮1|2𝑑m.\displaystyle(\sigma_{i+1}-\sigma_{1})\|{\bf w}_{i}\|^{2}\leq\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm+\alpha\int_{\Omega}|\nabla y_{i}\cdot{\bf u}_{1}|^{2}dm. (5.12)

From (5.10), for any constant B>0B>0, we infer

(σi+1σ1)\displaystyle(\sigma_{i+1}-\sigma_{1}) Ω|𝐮1|2T(yi,yi)𝑑m\displaystyle\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm
=\displaystyle= (σi+1σ1){2Ω𝐰i(12𝐮1yi+T(yi,𝐮1))𝑑m}\displaystyle(\sigma_{i+1}-\sigma_{1})\Big{\{}-2\int_{\Omega}{\bf w}_{i}\cdot\Big{(}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{)}dm\Big{\}}
\displaystyle\leq 2(σi+1σ1)𝐰i12𝐮1yi+T(yi,𝐮1)\displaystyle 2(\sigma_{i+1}-\sigma_{1})\|{\bf w}_{i}\|\Big{\|}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{\|}
\displaystyle\leq B(σi+1σ1)2𝐰i2+1B12𝐮1yi+T(yi,𝐮1)2,\displaystyle B(\sigma_{i+1}-\sigma_{1})^{2}\|{\bf w}_{i}\|^{2}+\frac{1}{B}\Big{\|}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{\|}^{2},

hence using (5.12) and the previous inequality we get

(σi+1σ1)\displaystyle(\sigma_{i+1}-\sigma_{1}) Ω|𝐮1|2T(yi,yi)𝑑m\displaystyle\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm
\displaystyle\leq B(σi+1σ1)(Ω|𝐮1|2T(yi,yi)𝑑m+αΩ|yi𝐮1|2𝑑m)\displaystyle B(\sigma_{i+1}-\sigma_{1})\Big{(}\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm+\alpha\int_{\Omega}|\nabla y_{i}\cdot{\bf u}_{1}|^{2}dm\Big{)}
+1B12𝐮1yi+T(yi,𝐮1)2.\displaystyle+\frac{1}{B}\Big{\|}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{\|}^{2}. (5.13)

Summing over ii from 11 to nn in (5.2), we conclude that

i=1n\displaystyle\sum_{i=1}^{n} (σi+1σ1)(1B)Ω|𝐮1|2T(yi,yi)𝑑m\displaystyle(\sigma_{i+1}-\sigma_{1})(1-B)\int_{\Omega}|{\bf u}_{1}|^{2}T(\nabla y_{i},\nabla y_{i})dm
Bαi=1n(σi+1σ1)Ω|yi𝐮1|2𝑑m+1Bi=1n12𝐮1yi+T(yi,𝐮1)2.\displaystyle\leq B\alpha\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\int_{\Omega}|\nabla y_{i}\cdot{\bf u}_{1}|^{2}dm+\frac{1}{B}\sum_{i=1}^{n}\Big{\|}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{\|}^{2}. (5.14)

From the definition of yiy_{i} and the fact that SS is an orthogonal matrix, we know that {yi}i=1n\{y_{i}\}_{i=1}^{n} are also the coordinate functions in n\mathbb{R}^{n}. Then, as in the proof of Theorem 1.1, we can also get

0<i=1n12𝐮1yi+T(yi,𝐮1)2(T(𝐮1)+12T0)2+C0,0<\sum_{i=1}^{n}\Big{\|}\frac{1}{2}{\bf u}_{1}\mathscr{L}y_{i}+T(\nabla y_{i},\nabla{\bf u}_{1})\Big{\|}^{2}\leq\Big{(}\|T(\nabla{\bf u}_{1})\|+\frac{1}{2}T_{0}\Big{)}^{2}+C_{0},

where C0C_{0} is given by Eq. (1.7). Using (5.2) and εITδI\varepsilon I\leq T\leq\delta I, we obtain

i=1n(σi+1σ1)(εB(δ+α))1B{(T(𝐮1)+12T0)2+C0}.\displaystyle\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})(\varepsilon-B(\delta+\alpha))\leq\frac{1}{B}\Big{\{}(\|T(\nabla{\bf u}_{1})\|+\frac{1}{2}T_{0})^{2}+C_{0}\Big{\}}. (5.15)

Since BB is an arbitrary positive constant, we can take

B={(T(𝐮1)+12T0)2+C0(δ+α)i=1n(σi+1σ1)}12B=\Bigg{\{}\frac{(\|T(\nabla{\bf u}_{1})\|+\frac{1}{2}T_{0})^{2}+C_{0}}{(\delta+\alpha)\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})}\Bigg{\}}^{\frac{1}{2}}

into (5.15) and therefore we get

i=1n(σi+1σ1)4(δ+α)ε2{(T(𝐮1)+12T0)2+C0}.\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq\frac{4(\delta+\alpha)}{\varepsilon^{2}}\Big{\{}(\|T(\nabla{\bf u}_{1})\|+\frac{1}{2}T_{0})^{2}+C_{0}\Big{\}}. (5.16)

We can take i=1i=1 in inequality (5.4) and replace in (5.16) to obtain Theorem 1.2. ∎

6. Divergence-free tensors case

This section is a generalization of some results of Section 2. Here, we are assuming the tensor TT to be divergence-free, i.e., divT=0\mathrm{div}T=0. Divergence-free tensors often arise from physical facts. We can find some of them in fluid dynamics, for instance, in the study of: compressible gas; rarefied gas; steady/self-similar flows and relativistic gas dynamics, see e.g. Serre [17]. We highlight that Serre’s work deals with divergence-free positive definite symmetric tensors and fluid dynamics.

For divergence-free tensors, the operator \mathscr{L} can be decomposed as follows

f=fη,T(f),\mathscr{L}f=\square f-\langle\nabla\eta,T(\nabla f)\rangle, (6.1)

where \square is the operator introduced by Cheng and Yau [6], namely:

f=tr(2fT)=2f,T.\square f=\mathrm{tr}{(\nabla^{2}f\circ T)}=\langle\nabla^{2}f,T\rangle.

Cheng-Yau operator arise from the study of complete hypersurfaces of constant scalar curvature in space forms. For more details, the reader can be consult Gomes and Miranda [9].

Eq. (6.1) is a first-order perturbation of the Cheng-Yau operator, and it defines a drifted Cheng-Yau operator which we denote by η\square_{\eta} with a drifting function η\eta.

We now turn our attention to the main problem of this paper. Since TT is divergence-free, the coupled system of second-order elliptic differential equations (1.2) becomes

{η𝐮+α(divη𝐮)=σ𝐮in Ω,𝐮=0onΩ,\left\{\begin{array}[]{ccccc}\square_{\eta}{\bf u}+\alpha\nabla(\mathrm{div}_{\eta}{\bf u})&=&-\sigma{\bf u}&\mbox{in }&\Omega,\\ {\bf u}&=&0&\mbox{on}&\partial\Omega,\end{array}\right. (6.2)

where 𝐮=(u1,u2,,un){\bf u}=(u^{1},u^{2},\ldots,u^{n}) is a vector-valued function from Ω\Omega to n\mathbb{R}^{n}, the constant α\alpha is non-negative and η𝐮:=(ηu1,ηu2,,ηun)\square_{\eta}{\bf u}:=(\square_{\eta}u^{1},\square_{\eta}u^{2},\ldots,\square_{\eta}u^{n}). Moreover, we have tr(T)=0\mathrm{tr}{(\nabla T)=0}, because TT is divergence-free. Thus, the constant C0C_{0} in (1.7) becomes

C0=supΩ{12div(T2(η))14|T(η)|2}.C_{0}=\sup_{\Omega}{\Big{\{}}\frac{1}{2}\mathrm{div}(T^{2}(\nabla\eta))-\frac{1}{4}|T(\nabla\eta)|^{2}{\Big{\}}}.

Hence, from Theorems 1.1 and 1.2 we immediately obtain the next two corollaries.

Corollary 6.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, and 𝐮i{\bf u}_{i} be a normalized eigenfunction corresponding to ii-th eigenvalue σi\sigma_{i} of Problem 6.2. For any positive integer kk, we get

i=1k(σk+1σi)24δ(nδ+α)n2ε2i=1k(σk+1σi)(σiαdivη𝐮i2+C0δ).\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})^{2}\leq\frac{4\delta(n\delta+\alpha)}{n^{2}\varepsilon^{2}}\sum_{i=1}^{k}(\sigma_{k+1}-\sigma_{i})\Big{(}\sigma_{i}-\alpha\|\mathrm{div}_{\eta}{\bf u}_{i}\|^{2}+\frac{C_{0}}{\delta}\Big{)}.
Corollary 6.2.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain, σi\sigma_{i} be the ii-th eigenvalue of Problem 6.2, for i=1,,ni=1,\ldots,n, and 𝐮1{\bf u}_{1} be a normalized eigenfunction corresponding to the first eigenvalue. Then, we get

i=1n(σi+1σ1)4δ(δ+α)ε2(σ1+D1),\sum_{i=1}^{n}(\sigma_{i+1}-\sigma_{1})\leq\frac{4\delta(\delta+\alpha)}{\varepsilon^{2}}(\sigma_{1}+D_{1}),

where D1=αdivη𝐮12+C0δD_{1}=-\alpha\|\mathrm{div}_{\eta}{\bf u}_{1}\|^{2}+\frac{C_{0}}{\delta}.

Now, from Corollary 6.1 and following the same steps of the proof of Corollary 2.3, we obtain the estimates.

Corollary 6.3.

Under the same setup as in Corollary 6.1, and by defining D0=αminj=1,,kdivη𝐮j2+C0δD_{0}=-\alpha\min_{j=1,\ldots,k}\|\mathrm{div}_{\eta}{\bf u}_{j}\|^{2}+\frac{C_{0}}{\delta}, we have

σk+1+D0\displaystyle\sigma_{k+1}+D_{0}\leq (1+2δ(nδ+α)ε2n2)1ki=1k(σi+D0)+[(2δ(nδ+α)ε2n21ki=1k(σi+D0))2\displaystyle\Big{(}1+\frac{2\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}\Big{)}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})+\Big{[}\Big{(}\frac{2\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})\Big{)}^{2}
(1+4δ(nδ+α)ε2n2)1kj=1k(σj1ki=1kσi)2]12\displaystyle-\Big{(}1+\frac{4\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}}

and

σk+1σk\displaystyle\sigma_{k+1}-\sigma_{k} \displaystyle\leq 2[(2δ(nδ+α)ε2n21ki=1k(σi+D0))2\displaystyle 2\Big{[}\Big{(}\frac{2\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}\frac{1}{k}\sum_{i=1}^{k}(\sigma_{i}+D_{0})\Big{)}^{2}
(1+4δ(nδ+α)ε2n2)1kj=1k(σj1ki=1kσi)2]12.\displaystyle-\Big{(}1+\frac{4\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}\Big{)}\frac{1}{k}\sum_{j=1}^{k}\Big{(}\sigma_{j}-\frac{1}{k}\sum_{i=1}^{k}\sigma_{i}\Big{)}^{2}\Big{]}^{\frac{1}{2}}.

Again from Corollary 6.1 and by applying the recursion formula of Cheng and Yang [4], we obtain the next corollary.

Corollary 6.4.

Under the same setup as in Corollary 6.3, we have

σk+1+D0(1+4δ(δn+α)ε2n2)k2δ(nδ+α)ε2n2(σ1+D0).\sigma_{k+1}+D_{0}\leq\Big{(}1+\frac{4\delta(\delta n+\alpha)}{\varepsilon^{2}n^{2}}\Big{)}k^{\frac{2\delta(n\delta+\alpha)}{\varepsilon^{2}n^{2}}}(\sigma_{1}+D_{0}).

Acknowledgements

The authors would like to express their sincere thanks to Chang Yu Xia and Dragomir Mitkov Tsonev for useful comments, discussions and constant encouragement. The first author has been partially supported by Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) in conjunction with Fundação Rondônia de Amparo ao Desenvolvimento das Ações Científicas e Tecnológicas e à Pesquisa do Estado de Rondônia (FAPERO). The second author has been partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), of the Ministry of Science, Technology and Innovation of Brazil.

References

  • [1] M.S. Ashbaugh and R.D. Benguria, More bounds on eigenvalue ratios for Dirichlet Laplacians in n dimensions, SIAM J. Math. Anal. 24 (6) (1993) 1622-1651.
  • [2] É. Cartan, Familles de surfaces isoparametriques dans les espaces à courbure constante, Ann. Mat. Pura Appl. IV. Ser. 17 (1938) 177-191.
  • [3] D. Chen, Q.M. Cheng, Q. Wang and C. Xia, On eigenvalues of a system of elliptic equations and of the biharmonic operator, J. Math. Anal. Appl. 387 (2012) 1146-1159.
  • [4] Q.M. Cheng and H.C. Yang, Bounds on eigenvalues of Dirichlet Laplacian, Math. Ann. 337 (2007) 159-175.
  • [5] Q.M. Cheng and H.C. Yang, Universal inequalities for eigenvalues of a system of elliptic equations, Proc. Roy. Soc. Edinburgh Sect. A 139 (2009) 273–285.
  • [6] S.Y. Cheng and S.T. Yau, Hypersurfaces with constant scalar curvature, Math. Ann. 225 (1977) 195-204.
  • [7] B. Chow et al., The Ricci Flow: Techniques and Applications Part I: Geometric Aspects. Mathematical Surveys and Monographs, vol. 135, AMS, Providence, RI, 2007.
  • [8] F. Du and A.C. Bezerra, Estimates for eigenvalues of a system of elliptic equations with drift and of bi-drifting laplacian, Comm. Pure Appl. Anal. 6 (2) (2017) 475-491.
  • [9] J.N.V. Gomes and J.F.R. Miranda, Eigenvalue estimates for a class of elliptic differential operators in divergence form, Nonlinear Anal. 176 (2018) 1-19.
  • [10] S.M. Hook, Domain independent upper bounds for eigenvalues of elliptic operator, Trans. Amer. Math. Soc. 318 (1990) 615–642.
  • [11] B. Kawohl and G. Sweers, Remarks on eigenvalues and eigenfunctions of a special elliptic system, Z. Angew. Math. Phys. 38 (1987) 730-740.
  • [12] H.A. Levine and M.H. Protter, Unrestricted lower bounds for eigenvalues of elliptic equations and systems of equations with applications to problems in elasticity, Math. Methods Appl. Sci. 7 (1985) 210-222.
  • [13] M. Levitin and L. Parnovski, Commutators, spectral trance identities, and universal estimates for eigenvalues, J. Funct. Anal. 192 (2002) 425–445.
  • [14] L. Ma and B. Liu, Convexity of the first eigenfunction of the drifting Laplacian operator and its applications, New York J. Math. 14 (2008) 393-401.
  • [15] P.J. Olver, Introduction to Partial Differential Equations, Undergraduate Texts in Mathematics, Springer, DOI 10.1007/978-3-319-02099-0.
  • [16] A. Pleijel, Proprietés asymptotique des fonctions fondamentales du problems des vibrations dans un corps élastique, Ark. Mat. Astron. Fys. 26 (1939) 1-9.
  • [17] D. Serre, Divergence-free positive symmetric tensors and fluid dynamics, Ann. I.H.Poincaré - AN35 (2018) 1209-1234.
  • [18] Q.M. Wang, Isoparametric functions on Riemannian manifolds I, Math. Ann. 277 (1987) 639-646.
  • [19] H. Weyl, Über die asymptotische Verteilung der Eigenwerte, Nachr. König. Gesellschaft. Wiss. Göttingen, Mathe-Phys. Klass (1911) 110-117.
  • [20] H.C. Yang, An estimate of the difference between consecutive eigenvalues, preprint IC/91/60 of ICTP, Trieste, 1991.