Let $ M $ be an $ n $-dimensional complete Riemannian manifold with a smooth metric $ g $. The triple $ (M, g, e^{-\phi }dv) $ is called a smooth metric measure space, where $ \phi $ is a smooth function on $ M $. A smooth metric measure space can also arise as the smooth collapse limit of a sequence of manifolds with lower bounds on Ricci curvature, under convergence in the Gromov-Hausdorff sense. As an important topic, smooth metric measure space has received lots of attention (cf. [1–3]).
The drifting Laplacian associated with $ (M, g, e^{-\phi }dv) $ is defined by
where $ \triangle $ denotes the Laplacian on $ M $. If $ \phi $ is a constant, it is easy to see that the drifting Laplacian is exactly the Laplacian. In particular, when $ M $ is a self-shrinker and $ \phi =\frac{1}{2}|x|^{2} $, the drifting Laplacian becomes $ \mathfrak{L} $ operator introduced by Colding and Minicozzi [4]. Besides, it is a self-adjoint operator with respect to weighted volume density $ d\mu= e^{-\phi}dv $. Namely, it holds
where $ \Omega $ is a bounded domain of $ M $. In recent years, the estimation of eigenvalues of the drifting Laplacian has received widespread attention (see [5, 6]).
In this paper, we investigate the weighted eigenvalue problem of polynomial operator of the drifting Laplacian as follows
where $ \rho $ is a positive continuous function on $ \Omega $, $ v $ denotes the outward unit normal to the boundary $ \partial \Omega $, and $ a,b $ are two nonnegative constants. It has the following real and discrete spectrum
where each eigenvalue is repeated according to its multiplicity.
Problem (1.3) has some interesting connection with some classical problems. On the one hand, when $ \phi $ is a constant, it becomes the following weighted Dirichelet problem of quadratic polynomial operator of the Laplacian
On the other hand, when $ a=b=0 $ and $ \rho \equiv 1 $, problem (1.3) becomes the Dirichelet eigenvalue problem of the bi-drifting Laplacian
Furthermore, if $ \phi $ is a constant, problem (1.6) becomes the clamped plate problem
There have been some intereting results for problems (1.5-1.7). In 2007, Wang and Xia [7] established the following inequality for eigenvalues of problem (1.7) on a unit sphere
In 2010, Cheng, Ichikawa and Mametsuka [8] established a Yang's inequality for problem (1.7) in an $ n $-dimensional complete Riemannian manifold
In 2019, for problem (1.6) on a bounded domain of the cigar soliton $ (\mathbb{R}^2,g,\phi) $, Li and Xiong [15] obtained
and
where $ C_0=\max \limits_{x\in \Omega }|x|^{2} $ and $ C_1=\min \limits_{x\in \Omega}|x|^2 $. For more reference on problems (1.5-1.7), we refer to [9–11] and the references therein.
Ricci solitons are an important kind of complete metric measure spaces. They are corresponding to self-similar solutions of Hamilton's Ricci flow [12, 13]. $ (M, g, \phi) $ is called a gradient Ricci soliton if there is a constant $ K $, such that
The function $ \phi $ is called a potential function of the gradient Ricci soliton. For $ K>0 $, $ K=0 $ and $ K<0 $, the Ricci soliton is called shrinking, steady or expanding respectively. When the dimension is two, Hamilton discovered the first complete non-compact example of a steady Ricci soliton on $ \mathbb{R}^2 $, called the cigar soliton. The metric and potential function of the cigar soliton $ (\mathbb{R}^2, g, \phi) $ are given by
where $ |x|^2=(x^1)^2+(x^2)^2 $ and $ \phi=-{\rm log}(1+|x|^2) $. In physics, the cigar soliton $ (\mathbb{R}^2, g, \phi) $ is regarded as the Euclidean-Witten black hole under first-order Ricci flow of the world-sheet sigma model. Moreover, the cigar soliton was also studied by Witten as a target space in string theory [14]. Thus, it is of great importance both in geometry and physics.
In this paper, we obtain the following results for problem (1.3) on a bounded domain $ \Omega $ of the cigar soliton $ (\mathbb{R}^2,g,\phi) $.
Theorem 1.1 Let $ \Omega $ be a bounded domain of the cigar soliton $ (\mathbb{R}^2, g, \phi ) $. Set $ \rho_1=\min \limits_{x\in \Omega }\rho (x) $, $ \rho_2=\max \limits_{x\in \Omega }\rho (x) $ and $ \varpi_i=\frac{1}{2\rho_1}[-a+\sqrt{a^{2}+4\rho_1(\lambda _i-\frac{b}{\rho_2})}] $. Denote by $ \lambda _i $ the $ i $-th eigenvalue of problem (1.3). Then we have
where $ C_1=\min \limits_{x\in \Omega}|x|^2 $ and $ C_0=\max \limits_{x\in \Omega }|x|^{2} $.
Theorem 1.2 Under the same assumptions as Theorem 1.1, we have
Remark 1.1 It is easy to find that when $ a=b=0 $ and $ \rho =1 $, (1.13) and (1.14) respectively become (1.10) and (1.11) for problem (1.6) in [15]. Therefore, our results generalize the results in [15].
In this section, we give the proof of Theorem 1.1. For this goal, we first establish a necessary lemma which plays a key role in the proof of Theorem 1.1.
Lemma 2.1 Let $ u_i $ be the orthonormal eigenfunction corresponding to the $ i $-th eigenvalue $ \lambda_i $ of problem (1.3). Then, for any function $ h\in C^{4}(M)\cap C^3(\partial{M}) $ and any positive integer $ k $, we have
where $ \delta $ is any positive constant.
Proof Set $ \varphi_{i}=hu_i-\sum_{j=1}^{k} \alpha_{ij}u_j $, where $ \alpha_{ij}=\int_M\rho hu_iu_jd\mu. $ Then we have
Using the Rayleigh-Ritz inequality, we get
According to the definition of $ \varphi_{i} $, we have
Therefore, using (2.4) and (2.5), we get
where
Then it follows from (2.6) that
where $ \beta_{ij}=\int_\Omega \Psi_iu_j $. Hence, substituting (2.6) into (2.3), we derive
Using the divergence theorem, we deduce
Moreover, since
we have
It implies
Moreover, we have
Thus, using (2.13) and (2.14), we derive
Therefore, we obtain
Set $ \xi_{ij}=\int_\Omega u_j( \langle \nabla h,\nabla u_i \rangle +\frac{1}{2}u_iL_\phi h)d\mu $. Then it holds $ \xi_{ij}=\xi_{ji} $. Moreover, since
we obtain
Hence we derive
Multiplying both sides of (2.17) by $ (\lambda_{k+1}-\lambda _i)^{2} $, and using the Schwarz inequality, we get
where $ \delta $ is any positive constant. Summing over $ i $ from 1 to $ k $, we have
Since $ \alpha_{ij} $ is symmetric and $ \xi_{ij} $ is anti-symmetric, we deduce
Therefore, combining (2.23) and (2.23) with (2.21), we obtain
Using (2.25-2.28), we have
Substituting (2.29) into (2.24), we obtain (2.1). This completes the proof of Lemma 2.1.
Now we give the proof of Theorem 1.1 by using Lemma 1.1.
Proof of Theorem 1.1 Suppose that $ x^p $ is the $ p $-th local coordinate of $ x_0 \in \Omega \subset \mathbb{R}^2 $, where $ p = 1, 2 $. Taking $ h=x^{p} $ in Lemma 2.1, we have
It is not difficult to obtain
Using (2.31) and (2.32), we get
Taking the sum over $ p $ from $ 1 $ to $ 2 $ on (2.30), and using (2.31), (2.34) and (2.35), we have
Since
Moreover, using
It yields
This is a quadratic inequality of $ \int_\Omega |\nabla u_i|^2 d\mu $. Hence we get
Substituting (2.37) and (2.41) into (3.25), we infer
Taking
in (2.42), we obtain (1.13). The proof of Theorem 1.1 is finished.
In this section, we give the proof of Theorem 1.2. For this goal, we first prove the following lemma.
Lemma 3.1 Under the same assumptions as Lemma 2.1, for any function $ \zeta^p \in C^{4}(M)\cap C^3(\partial{M}) $ $ (p=1,2) $ and any positive integer $ k $, we have
Proof Set $ \psi^p=(\zeta^p-\gamma^p)u_1 $, where $ \gamma^p=\int_\Omega \rho \zeta^p u_1^{2}d\mu $. It implies $ \int_\Omega \rho \psi^p u_1d\mu =0 $. Noticing that $ \int_\Omega \rho \zeta^p u_1u_{q+1}d\mu =0 $, for $ 1\leq q<p $, we have
From the Rayleigh-Ritz inequality, we have
According to the definition of $ \psi^p $, we have
It follows from (3.4) and (3.5) that
Therefore, using (3.7-3.9), we have
Using (3.3), (3.6) and (3.10), we derive
Furthermore, using the similiar computation, we have
Therefore, it follows from (3.11-3.15) that
Since $ \displaystyle{\int}_\Omega \zeta^p u_1^{2} L_\phi \zeta^p d\mu=-\displaystyle{\int} _\Omega u_1^{2}|\nabla \zeta^p|^{2} d\mu -2\displaystyle{\int}_\Omega \zeta^p u_1\langle \nabla u_1,\nabla \zeta^p\rangle d\mu, $ we have
Multiplying both sides of (3.17) by ($ \lambda_{p+1}-\lambda_1)^{\frac{1}{2}} $, and using (3.16), we derive
The proof of Lemma 3.1 is ended.
Now we give the proof of Theorem 1.2 by using Lemma 3.1.
Proof of Theorem 1.2 Define a ($ 2\times 2 $)-matrix $ B=( \epsilon_{pt})_{2\times 2}, $ where $ \epsilon_{pt}=\int_\Omega \rho x^{p}u_1u_{t+1} d\mu $. Using the orthogonalization of Gram-Schmidt, we know that there exists an upper triangle matrix $ U=(\vartheta_{pt})_{2\times 2} $ and an orthogonal matrix $ P=(\varsigma_{pt})_{2\times 2} $ such that $ U=PB $. That is to say, for $ 1\leqslant t<p\leqslant 2 $, we have
Setting $ y^{p}=\sum_{s=1}^{2} \varsigma_{ps}x^{s} $, we obtain
Taking $ \zeta^p=y^p $ in (3.1), and taking sum on $ p $ from 1 to 2, we get
Since $ y^{p}=\sum_{s=1}^{2} \varsigma_{ps}x^{s} $ and $ P $ is an orthogonal matrix, we know that $ y^1 $ and $ y^2 $ are the standard coordinate functions of $ \mathbb{R}^2 $. It is not difficult to check that
Substituting it into (3.21), we have
Similar to the computation as (2.42), we obtain
Taking $ \delta =\left\{\frac{\frac{1}{\rho_1}\left[\frac{1+C_0}{1+C_1}\varpi_1+\frac{1}{\rho_2(1+C_1)}-\frac{3}{\rho_2} \right]}{ \left[2\frac{1+C_0}{1+C_1}\varpi_1+\frac{1}{\rho_2(1+C_1)}-\frac{3}{\rho_2} +\frac{a}{2\rho_1}\frac{1+C_0}{1+C_1}\right] } \right\}^{\frac{1}{2}} $ in (3.26), we obtain (1.14). The proof of Theorem 1.2 is completed.