Multiple solutions of Kazdan-Warner equation on graphs in the negative case
Abstract
Let be a finite connected graph, and let be a function such that . We consider the following Kazdan-Warner equation on :
where and is a non-constant function satisfying and . By a variational method, we prove that there exists a such that when the above equation has solutions, and has no solution when . In particular, it has only one solution if ; at least two distinct solutions if ; at least one solution if . This result complements earlier work of Grigor’yan-Lin-Yang [7], and is viewed as a discrete analog of that of Ding-Liu [4] and Yang-Zhu [17] on manifolds.
keywords:
Kazdan-Warnar problem on graph , variation problem on graphMSC:
[2010] 35R02, 34B451 Introduction
Variational method is always a powerful tool in partial differential equations and geometric analysis. Recently, using this tool, Grigor’yan-Lin-Yang [7, 8, 9] obtained existence results for solutions to various partial differential equations on graphs. In particular, Kazdan-warnar equation was proposed on graphs in [7]. The Kazdan-Warner equation arises from the basic geometric problem on prescribing Gaussian curvature of Riemann surface, which systematically studied by Kazdan-Warner [12, 13]. On a closed Riemann surface with the Gaussian curvature , let be a smooth metric conformal to and be the Gaussian curvature with respect to . Then satisfies the equation
(1) |
where denotes the Laplace-Beltrami operator with respect to the metric . Let be a solution to and , where is the averaged integral of . Then the above equation is transformed to
Hence, one can free (1) from the geometric situation, and just studies the equation
(2) |
where is a constant and is a function. On graphs, it seems to be out of reach to resemble this topic in terms of Gaussian curvature. Therefore, in [7], the authors focused on the equation similar to the form of (2), namely the Kazdan-Warner equation on graph, and obtained the following: when , it has a solution if and only if changes sign and the integral of is negative; when , it has a solution if and only if is positive somewhere; when , there is a threshold such that it has a solution if , but it has no solution for any . Later, Ge [5] found a solution in the critical case . More recently Ge-Jiang [6] studied the Kazdan-Warner equation on infinite graphs and Keller-Schwarz [11] on canonically compactifiable graphs; Camilli-Marchi [3] extended the Kazdan-Warner equation on network; for other related works, we refer the readers to [10, 14].
Let us come back to a closed Riemann surface , whose Euler characteristic is negative, or equivalently . Replacing by in (1) with , , and , Ding-Liu [4] obtained the following conclusion by using a method of upper and lower solutions and a variational method: there exists a such that if , then (1) has a unique solution; if , then (1) has at least two distinct solutions; if , then (1) has at least one solution; if , then (1) has no solution. Recently, this result was partly reproved by Borer-Galimberti-Struwe [2] via a monotonicity technique due to Struwe [15, 16], and was extended to the case of conical metrics by Yang-Zhu [17].
Our aim is to extend results of Ding-Liu [4] to graphs. Let us recall some notations from graph theory. Throughout this paper, is assumed to be a finite connected graph. The edges on the graph are allowed to be weighted. Weights are given by a function , the edge from to has weight . We assume this weight function is symmetric, . Let be a positive measure on the vertices of the . Denote by the space of real functions on . and by , for any , the space of integrable functions on with respect to the measure . For , let be the set of all bounded functions. As usual, we define the norm of , by
We define the Laplacian on by
(3) |
Given the weight on , there are two typical choices of Laplacian as follows:
-
1.
for all , which is called the normalized graph Laplacian;
-
2.
for all , which is the combinatorial graph Laplacian.
In this paper, we do not restrict to the above two forms, but only require for all . Note that the Laplace operator defined in (3) is the negative usual Laplace operator. The gradient form is defined by
For the sake of simplicity, we write . Sometimes we use the notation . The length of the gradient is denoted by
From now on, we write . Define a Sobolev space with a norm on the graph by
and
respectively. Since is a finite graph, we have that is exactly , a finite dimensional linear space. This implies the following Sobolev embedding:
Lemma 1 ([7], Lemma 5).
If is a finite graph, then the Sobolev space is precompact. Namely, if is bounded in , then there exists some such that up to a subsequence, in .
The Kazdan-Warner equation we are interested in this paper reads as
(4) |
where is a function, and , , is a function. Now we are ready to state our main results.
Theorem 2.
Remark 1.
The assertion of comes from the conclusion of Step 2 in Subsection 3.3.
Remark 2.
The proof of Theorem 2 is based on the method of variation. It can be viewed as a discrete analog of the result of Ding-Liu [4]. The remaining part of this paper will be organized as follows: In Section 2, we give several preliminary lemmas for our use later; In Section 3, we finish the proof of Theorem 2.
2 Preliminaries
In this section, we provide discrete versions of the maximum principle, the Palais-Smale condition and the upper and lower solution principle. Note that is a finite connected graph.
2.1 maximum principle
To proceed, we need the following maximum principles, which are known for experts (see for examples [7, 8]). For readers’ convenience, we include the detailed proofs here.
Lemma 3 (Weak maximum principle).
For any constant , if satisfies , then on .
Proof.
Let . For any , we claim that
(5) |
from which, one has
This leads to on .
To prove this claim, we first consider the case . Therefore, and
due to for any . In the case , one has and thus
It follows that , which confirms (5) and ends the proof of the lemma. ∎
Lemma 4 (Strong maximum principle).
Suppose that , and that for some constant . If there exists such that , then on
Proof.
Let , we have
which implies
Since and for all , we obtain
Therefore, on by the connectedness of ∎
2.2 Palais-Smale condition
We define a functional by
where and are given as in the assumptions of Theorem 2, in particular . For any , denote by the Frechet derivative of the functional, by the Frechet derivative of order .
Lemma 5 (Palais-Smale condition).
Suppose that is nonempty for some . Then satisfies the condition for all , i.e. if is a sequence of functions in such that and for all as , then there exists some satisfying in
Proof.
Let be a function sequence such that and or equivalently
(6) |
(7) |
where as
We now claim that is bounded in . Suppose not, there holds . We set By the Cauchy-Schwarz inequality, one has
This together with (9) leads to
(10) |
Hence, is bounded in . In view of Lemma 1 and (10), in for some constant Here and in the sequel, we do not distinguish sequence and subsequence. Since , we have It follows from (9) that
Passing to the limit in the above inequality, we conclude that since . Therefore .
On the other hand, for any , if there exists , if such that , then , which contradicts and confirms our claim. If not, let , due to the finiteness of , we can choose a subsequence such that . Set
Then
Substituting it into (7), we have
(11) |
Since , and has finite points, we conclude
This together with (11) leads to
which is impossible since . Then our claim follows immediately.
Since is bounded in , we have is bounded in due to the finiteness of . Therefore, by Lemma 1, there exists some such that up to subsequence, in . ∎
2.3 Upper and lower solutions principle
Let be a function, and is smooth with respect to the second variable. We say that is an upper (lower) solution to the following equation
(12) |
if satisfies for any . We generalize ([7], Lemma 8) to the following:
Lemma 6.
Proof.
This is a discrete version of the argument of Kazdan-Warner ([12], Lemma 9.3), and the method of proof carries over to the setting of graphs.
Since the graph is finite, there exists a constant such that . One can find a sufficient large constant such that is increasing with respect to for any fixed . We define an operator , and is a compact operator and due to the finiteness of the graph. Hence, we can define inductively as the unique solution to
respectively. Combining with the definition of upper (lower) solution and the monotonicity of with respect to , we obtain
Then the weak maximum principle (see Lemma 3) yields that
Moreover, it turns out that and are lower and upper solution to (12) respectively. By induction, we have
Since is finite, it is easy to see that up to a subsequence, uniformly on , and or is a solution to (12) with on . ∎
3 Proof of Theorem 2
3.1 Unique solution in the case .
Claim 1.
is strictly convex on .
Proof.
We only need to show that there exists some constant such that
(13) |
Suppose not, there would be a function and a function sequence such that for all and as . From Lemma 1, there exists , such that up to a subsequence, as in . Since
and , it follows that and , which lead to for some constant , and moreover
It is easily seen that by , thus . This contradicts
Hence (13) holds. ∎
Claim 2.
For any , there exist constants such that
Proof.
By Young’s inequality, for any , there exists a constant such that
Thus, it is sufficient to find some constant such that for all
Suppose not, there would exist a sequence of functions satisfying
Clearly, is bounded in , it follows from Lemma 1 that there exists some function such that up to a subsequence, in as Due to and we have
which gets a contradiction. ∎
Proof of (1) in Theorem 2. It is a consequence of Claim 1 and Claim 2. Precisely we denote . By Claim 2, we see that is a definite real number. Take a function sequence such that as . Applying Claim 2, we have that is bounded in . Then, in view of Lemma 1, there exists a subsequence of (still denoted by ) and a function such that in . Obviously , and thus is a critical point of . We also need to explain why has only one critical point. For otherwise, we assume is another critical point of . Note that . It follows from Claim 1, particularly from (13), that
for some positive constant , where is a function lies between and . Hence we have , contradicting the fact that . This implies the uniqueness of the critical point of .
3.2 Multiplicity of solutions for .
Fixing , we will seek two different solutions of (4). One is a strict local minimum of the functional , and the other is from mountain-pass theorem. We firstly prove the existence of . Consider the case in the equation (4) as follows
(14) |
For the solution of (14), the linearized equation of (14) at
has only a trivial solution since and Indeed,
which implies when In the case , we have . Thus
It follows that is a constant function and hence By the implicit function theorem, there exists a small enough such that the equation (4) has a solution for any . Indeed, let on , we consider . It is easy to see that , and are continuous on any domain , Furthermore, unless on . Therefore, by the implicit function theorem, there exists such that and for any . In other words, is the solution of (4). Define
One can see that For otherwise, for some . Adding up the equation (4) for all , we have
which is impossible. In conclusion, we have .
Proof of (2) in Theorem 2. We separate the proof into the following three steps.
Step 1. The existence of the upper and lower solution of (4).
Take with , let be a solution of (4) at . It is easily seen that is a strict upper solution of (4) at , namely
Let be the solution to the following equation
(15) |
The existence of solution to (15) was proved in [7]. Set , where is a sufficiently large constant such that on and
since . Therefore, is a strict lower solution of (4). Let be the order interval defined by
The upper and lower-solution method (Lemma 6) asserts that (4) has a solution on .
Step 2. can be chosen as a strict local minimum of .
Let , where is sufficiently large such that is increasing in , is a constant such that on . Let . It is easy to rewrite as
It is obvious that is bounded from below on . Therefore, we denote
Taking a function sequence such that as From it, we can get that is bounded in , and thus up to subsequence, converges to some in and for any , and converges to in . Hence
As a consequence, satisfies the Euler-Lagrange equation
From it, one can conclude that
(16) |
Indeed, noting that is increasing with respect to , we have
One can conclude (16) by the strong maximum principle (Lemma 4), and the fact . For any , we define a function . There holds for sufficiently small . Since is a minimum of on , we have and Furthermore, there exists a constant such that
(17) |
which implies is strict local minimum of on . It remains to prove (17). We first denote
which is nonnegative. It is sufficient to prove , (17) follows. Suppose , we claim that there exists some with such that . To see this, let be a function sequence satisfying for all and as Up to subsequence, in from Lemma 1, and confirms our claim. To put it another way, the functional attains its minimum at , it follows that for all Hence, is a solution of the following equation
(18) |
It is easy to see that is not a constant. For otherwise (18) yields
which is impossible. Multiplying (18) by , we obtain
The last inequality is due to the fact for any satisfying . Since attains its minimum at , we have , which together with and leads to
(19) |
for small Let small enough such that , thus by (19),
which contradicts the fact that is the minimum of on . Therefore , which concludes (17).
Step 3. The second solution of (4) is given by the mountain-pass theorem.
We shall use the mountain-pass theorem due to Ambrosetti and Rabinowitz [1], which reads as follows: Let be a Banach space, , and be such that and
If satisfies the condition with , where
then is a critical value of . In our case, is a Banach space, and is a smooth functional.
Since is a strict local minimum of on , there exists a small enough number such that
(20) |
Moreover, for any has no lower bound on , namely, there exists such that
(21) |
To see this, we set for small . Note that is nonempty since and . Let be a function which equals to in and vanishes on , then
In view of (20), (21) and Lemma 5, the mountain pass theorem of Ambrosetti and Rabinowitz gives another critical point of other than . In particular,
where , and for any .
3.3 Solvability at .
For any , let be the local minimum of obtained in the previous subsection. That is, is the solution of (4), and
(22) |
Proof of (3) in Theorem 2. The crucial point in this proof is to show that is uniformly bounded in as If it is true, then up to subsequence, converges to some in , and is the solution of
Hence, we aim to prove the boundedness of To this end, we divide the proof into the following three steps.
Step 1. There exists a constant such that
(23) |
uniformly for any .
Let satisfy (15), and for . Then for sufficient large , say , is a continuous family with respect to of strict lower solution of (4) at , i.e.
It is clear that is also a strict lower solution of (4) at for any . Now, we prove that , and thus (23) holds. For otherwise, we can find for some
It follows from the strong maximum principle (Lemma 4) that , which contradicts that is strict low solution of (4) at
Step 2. The set is not empty.
From the case (1) in Theorem 2, there exists unique solution of the equation
Together with the solution to the equation (4) at , and let , we have
Multiplying the above equation by and intergrading by parts, one has
Indeed, by Green formula
since for any . Therefore
Suppose that thus . This contradicts the assumption that is not a constant.
Step 3. is uniformly bounded in as .
Fixing , we set to be a function which vanishes besides and . Consider the equation
which always has the unique solution from the case (1) in Theorem 2, set As above, the function satisfies the equation
(24) |
Multiplying (24) by and integrating by parts over gives
(25) |
Utilizing the fact that
one can estimate the first term in (25) as follows.
Inserting this estimate into (25), we obtain
(26) |
On the other hand, let in (22), yields
Together with (26), we have
(27) |
One may derive that is the uniform bound, i.e. there exists a constant such that
Indeed, if the above inequality is obvious; if as in the proof of Lemma 5, see (11), one can get the boundedness of as well. Together with (27), yields
(28) |
Next, we claim that and thus is uniformly bounded in . Suppose not, we may assume that as Let
then , and from (28). It follows that converges to a constant in . From (3.3), we have , and hence on , which contradicts and then confirms our claim. Since is bounded below by (23) in Step 1, one has the -boundedness of , and thus the -boundedness of .
3.4 No solution when .
Proof of this case is a consequence of the upper and lower solutions principle, as follows.
Proof of (4) in Theorem 2. Let be the solution of (4) at some . For any , is an upper solution of (4) at . Indeed,
From (15), it is easy to get a lower bound solution of (4) at such that . By the upper and lower solutions principle, there exists a solution of (4) at , which contradicts the definition of
References
References
- [1] A. Ambrosetti, P. Rabinowitz, Dual variational methods in critical point theory and applications, J. Funct. Anal. 14 (1973) 349-381.
- [2] F. Borer, L. Galimberti, M. Struwe, “Large” conformal metrics of prescribing Gauss curvature on surfaces of high genus, Comment. Math. Helv. 90 (2015) 407-428.
- [3] F. Camilli, C. Marchi, A note on Kazdan-Warner equation on networks, arXiv:1909.08472.
- [4] W. Ding, J. Liu, A note on the problem of prescribing Gaussian curvature on surfaces, Trans. Amer. Math. Soc. 347 (1995) 1059-1066.
- [5] H. Ge, Kazdan-Warner equation on graph in the negative case, J. Math. Anal. Appl. 453 (2017) 1022-1027.
- [6] H. Ge, W. Jiang, Kazdan-warner equation on infinite graphs, J. Korean Math. Soc. 55 (2018) 1091-1101.
- [7] A. Grigor’yan, Y. Lin, Y. Yang, Kazdan-Warner equation on graph, Calc. Var. Partial Differential Equations 55 (2016), no. 4, Art. 92, 13 pp.
- [8] A. Grigor’yan, Y. Lin, Y. Yang, Yamabe type equations on graphs. J. Differential Equations 261 (2016) 4924-4943.
- [9] A. Grigor’yan, Y. Lin, Y. Yang, Existence of positive solutions to some nonlinear equations on locally finite graphs, Sci. China Math. 60 (2017) 1311-1324.
- [10] X. Han, M. Sha, L. Zhao, Existence and convergence of solutions for nonlinear biharmonic equations on graphs, to appear in J. Differential Equations, https://doi.org/10.1016/j.jde.2019.10.007
- [11] M. Keller, M. Schwarz, The Kazdan-Warner equation on canonically compactifiable graphs, Calc. Var. Partial Differential Equations 57 (2018), no. 2, Art. 70, 18 pp.
- [12] J. Kazdan, F. Warner, Curvature functions for compact 2-manifolds, Ann. of Math. 99 (1974) 14-47.
- [13] J. Kazdan, F. Warner, Curvature functions for open 2-manifolds, Ann. of Math. 99 (1974) 203-219.
- [14] Y. Lin, Y. Wu, The existence and nonexistence of global solutions for a semilinear heat equation on graphs,Calc. Var. Partial Differential Equations 56 (2017), no. 4, Art. 102, 22 pp.
- [15] M. Struwe, Critical points of embeddings of into Orlicz spaces, Ann. Inst. H. Poincaré Anal. Non Linéaire 5 (1988) 425-464.
- [16] M. Struwe, The existence of surfaces of constant mean curvature with free boundaries, Acta Math. 160 (1988) 19-64.
- [17] Y. Yang, X. Zhu, Prescribing Gaussian curvature on closed Riemann surface with conical singularity in the negative case, Ann. Acad. Sci. Fenn. Math. 44 (2019) 167-181.