Graphs with nonnegative Ricci curvature and maximum degree at most 3
Abstract.
In this paper, we classify graphs with nonnegative Lin-Lu-Yau-Ollivier Ricci curvature, maximum degree at most 3 and diameter at least 6.
1. Introduction
Ricci curvature is one of the most important geometric quantities in Riemannian geometry. For discrete settings, various definitions of Ricci curvature were developed over the past few decades. In this paper, we focus on Olliver Ricci curvature and its variations. The original Ollivier Ricci curvature was developed on metric spaces and Markov chains, which can be applied on normalized graphs; see [Oll09, Oll07]. In 2011, Lin-Lu-Yau [LLY11] modified Ollivier’s definition. This version was widely used in the research of normalized graphs. Münch and Wojciechowski [MW19] extended Lin-Lu-Yau’s definition to general weighted graphs. Bakry-Émery curvature is another important Ricci curvature on graphs, readers may refer to [Sch99, LY10] for more details.
There are significant geometric and analytic results for Riemannian manifolds with nonnegative Ricci curvature. As discrete analogs, graphs with nonnegative Ricci curvature are also worthy to study. A basic method is to explore their local geometric structures. Many papers on this subject were published recently. Cushing et al. [CKL+19] classified all 3-regular graphs with nonnegative Bakry-Émery curvature. They also classified all 3-regular graphs with nonnegative original Ollivier Ricci curvature. Under the setting of Lin-Lu-Yau-Ollivier Ricci curvature, there are many works on the classification of Ricci-flat graphs, i.e. graphs whose Ricci curvature vanishes on every edge; see [CKL+18, LLY14, CP18, HLYY18, BLY21].
It is interesting to study harmonic functions on graphs. For the continuous case, Yau [Yau75] showed that positive harmonic functions on a complete, noncompact Riemannian manifold with nonnegative Ricci curvature are constant, which is known as Liouville property of harmonic functions. For the discrete case, there are also some known results. Jost, Münch and Rose [JMR19] proved that bounded harmonic functions are constant under the nonnegative Lin-Lu-Yau-Ollivier Ricci curvature condition. We also refer to [Hua19, BHL+15, HLLY19] for Liouville properties under the other curvature conditions. However, the strong Liouville property of graphs with non-negative Lin-Lu-Yau-Ollivier Ricci curvature is still unknown. As an application of the main results, we prove that it is true for those graphs with maximum degree at most three.
Let be the class of combinatorial graphs with following properties:
1) Lin-Lu-Yau-Ollivier Ricci curvatures on all edges of are nonnegative;
2) Degrees of all vertices of are at most 3;
3) Diameters of the graphs in are at least 6.
The precise definition of is given in section 2. We classify all graphs in by the following two theorems.
Theorem 1.1.
Let be a finite graph. Then is isomorphic to a finite path, a simple cycle, a prism graph, a Möbius ladder, a particular graph as shown in Figure 1, or a quasi-ladder graph as shown in Figure 2.






Theorem 1.2.
Let be an infinite graph. Then is isomorphic to one of the following graphs.










Remark 1.3.
- (1)
-
(2)
For normalized graphs, if we change the nonegative Lin-Lu-Yau-Ollivier Ricci curvature constraint to nonnegative Ollivier Ricci curvature, we also get a classification; see Theorem 2.6.
-
(3)
It is well known that manifolds with nonnegative Ricci curvature satisfy Cheeger-Gromoll splitting theorem [CG72]. As a discrete analog, the splitting theorem also holds for Graphs in , i.e. if a both-side infinite line is contained in , then is the Cartesian product of a line and a graph.
Bai, Lu and Yau [BLY21] classified all Ricci-flat graphs with maximum degree at most 4. Compared with our results, they require a more stringent curvature condition but relax the maximum degree condition from 3 to 4, which causes a huge increase in the complexity of the problem.
The proofs of the main results are shown in section 4. We give a sketch here. Let be a diameter of with . The local structure along can be characterized in Lemma 3.1 , Lemma 3.3 and Lemma 3.4. Moreover, the subgraph induced by and its neighbors is a quasi-ladder. We divide into four parts as shown in Figure 4. If is connected to via , we will get a special cycle (we call it a geodesic cycle in this paper) by Lemma 4.1, and we prove has to be a cycle, a prism or a Möbius ladder by Lemma 4.2. If is not connected to via , then and is a quasi-ladder, otherwise we will find a longer geodesic path. For infinite graphs, the proof is similar. For graphs with small diameter, say , the structures are more complicated. However, since the maximum degree is at most 3, graphs with nonnegative curvature and small diameter can be enumerated by computers, which we are not going to explore in this paper.

The paper is organized as follows: In next section, we introduce some notions on graphs and the definition of Ricci curvature. Section 3 gives the local structures of graph along a diameter, which serve as basic tools in the following discussions. In section 4, we prove the main results of this paper. In section 5, as an applications of the main results, we show that all positive harmonic functions on are constant.
2. Preliminaries
2.1. Basic notions
Throughout this paper, we always assume that is an undirected, connected, locally finite, simple graph with vertex set and edge set . Let be the number of vertices of the graph. For any vertices , we call a neighbor of if , denoted by . For notational simplicity, we write for the unordered pair . For , let . The degree of a vertex is defined via . A graph is called a subgraph of if and . A subgraph is called an induced subgraph of if , and is denoted by . For , let be the combinatorial distance between and in subgraph . If , we omit the subscript, i.e. . If or is a single vertex subset, we will omit the braces, i.e. .
A walk of length is a sequence of vertices such that for . We write for short. If , we call it a closed walk. Moreover, a closed walk is called a cycle if for all . We also regard a cycle as a subgraph with vertex set and edge set . A walk is called a path if for all . We also regard a path as a subgraph with vertex set and edge set .
Let be a subgraph of . We say a path is geodesic in if . It is easy to know that is a geodesic path of if and only if for all . We say is a geodesic subgraph if all geodesic paths in are also geodesic paths in . It is not hard to verify that is a geodesic subgraph of if and only if for all . A diameter path of is a longest geodesic path in . We let denote the length of a diameter path in , i.e. .
A graph is called rooted at if a vertex is specially appointed, denoted by . We define via . If , there always exists a such that and a geodesic path with length equals to . Given a geodesic path in , we always regard as a graph rooted at if not specified.
Let be a graph with degree at most 3, be a geodesic path. For all , has at most one extra neighbor as and are two neighbors of . We define a state function via
As figures are quite important in the discussions of this paper, we explain the drawing style here. If not specified, a horizontal segment with vertices represents a geodesic path as shown in (a) of the following figure. If has an extra neighbor and , i.e. , we draw the figure as shown in (b) such that has the same abscissa with as . Figure (c) and (d) show the drawings when and respectively.




2.2. Ricci curvature on weighted graphs
Let
be the edge weight function,
be the vertex weight function. We call the quadruple a weighted graph. Unless otherwise specified, we always regard a graph as a combinatorial graph, i.e. and . Our results also hold for normalized graphs, i.e. and for all ; see Theorem 2.5.
We define the Laplacian via
A function on is called harmonic if .
For a weighted graph , let . We define a probability measure
Particularly, for combinatorial graphs, we have
For normalized graphs, we have
Definition 2.1.
(Wasserstein distance) Let be a graph, and be two probability measures on . The Wasserstein distance is given by
where is a coupling between and , i.e. a mapping s.t.
Definition 2.2.
For any vertices , the -Ollivier-Ricci curvature is defined via
Specifically, the Ollivier Ricci curvature on a normalized graph is defined via
The Lin-Lu-Yau-Ollivier Ricci curvature is defined via
For fixed vertices , it was shown in [BCL+18] that the function is concave and piecewise linear, so the limitation exists and is well defined. There are also equivalent limit-free definitions via coupling function or Laplacian on graphs; see [BHLY20, MW19]. When we say Ricci curvature or curvature in the following, we always mean the Lin-Lu-Yau-Ollivier Ricci curvature defined above.
For combinatorial graphs, Münch and Wojciechowsk [MW19, Theorem 2.6] gave an easier way to calculate Ricci curvatures on edges as shown in the following theorem.
Theorem 2.3.
(Transport and combinatorial graphs) Let be a combinatorial graph and let be adjacent vertices. Let , and . Let . For write and . Then

We always suppose or when applying Theorem 2.3. If is an optimal map such that and , we extend by adding an extra map . It is easy to verify that the new is still an optimal map. So our hypothesis is fair. The following is a direct corollary of Theorem 2.3.
Corollary 2.4.
Let be a combinatorial graph and let be adjacent vertices. Let be the set of extra neighbors of , . If and , we have .
Let be the Ricci curvature of and the maximum degree of . The class of graphs we will classify in this paper is defined by
Cushing et al. [CKL+19] provide a powerful online tool, the Graph Curvature Calculator, to calculate various of curvatures on graphs. Readers can find it on https://www.mas.ncl.ac.uk/graph-curvature/. By using this tool, it is convenient to check that the Ricci curvatures of graphs in Figure 1, 2 and 3 are nonnegative. We also use it to prove the following theorems.
Theorem 2.5.
Let be a graph with . Let be the combinatorial graph and be the normalized graph obtained by equipping with combinatorial weight and normalized weight respectively. Then if and only if .
Proof.
For any , let be the curvature of when is equipped with combinatorial weight and be the curvature of when is equipped with normalized weight. If or , we have and by a direct calculation. If , we have by the definition of Ricci curvature. So we only need to consider the case that and , i.e. has an extra neighbor , has two extra neighbors and . Then and depend only on the distances and . By a direct calculation, we have . This completes the proof. ∎
It is noteworthy that Theorem 2.5 does not hold for graphs with maximum degree larger than 3. The minimum Ricci curvatures of the following graphs are 0, 0.5, -0.133 as normalized graphs and -1, -1, 0 as combinatorial graphs respectively. 111Graph (a) built by Bai-Lu-Yau [BLY21] is Ricci-flat as a normalized graph. Graphs (b) and (c) are pointed out to us by Florentin Münch.



Theorem 2.6.
Let be a normalized graph with nonnegative Ollivier Ricci curvature, maximum degree at most 3 and diameter at least 6. If is finite, then it is isometric to a finite path, a simple cycle, a prism graph, a Möbius ladder or a quasi-ladder as shown in the following figure. If is infinite, then it is isometric to (a), (b), (c), (d), (e), (f), (g) or (h) of Figure 3.

Proof.
For normalized graphs, for all . If , we always have , as the function is concave and piecewise linear; see [BCL+18, Theorem 1.1]. This indicates that all normalized graphs with nonnegative Ollivier Ricci curvature are also graphs with nonnegative Lin-Lu-Yau-Ollivier Ricci curvature. We check the Ollivier Ricci curvatures of graphs in Figure 1, 2, 3 by using the Graph Curvature Calculator and get the result. ∎
3. The local structures
In this section, we study the local structure on a geodesic path of a graph .
Lemma 3.1.
Let be a geodesic path of .
Case : If , then contains the following as a subgraph.

Case : If , then contains one of the followings as a subgraph.



Case : If , then contains the following as a subgraph.

Case : The situation can not exist.
Case : The situation can not exist.
Case : If , then contains the following as a subgraph.

Case : If , then contains one of the followings as a subgraph.



Case : If , then contains one of the followings as a subgraph.


Case : The situation can not exist.
Case : The situation can not exist.
Case : If , then contains one of the followings as a subgraph.





Case : If , then contains one of the followings as a subgraph.


Case : The situation can not exist.
Case : If , then contains one of the followings as a subgraph.







Case : If , then contains one of the followings as a subgraph.





Case : If , then contains one of the followings as a subgraph.


Proof.
We only show the proof of case , proofs for other cases are similar. Without loss of generality, let . Then has a subgraph as shown in the following figure with .

By applying case or case of Lemma 3.1 repeatedly, we get the following corollary.
Corollary 3.2.
Let be a geodesic path in , .
1) If there exists a such that , then for all and has a subgraph (b) or (c) of the following figure.
2) If there exists a such that , then for all and has a subgraph (e) or (f) of the following figure.






As a supplement to Lemma 3.1, we explore the local structures on long geodesic paths in the following two lemmas,
Lemma 3.3.
Proof.
If , contains a subgraph (a) of the following figure. By Corollary 3.2, contains a subgraph (b). By applying case of Lemma 3.1 on , contains a subgraph (c). By applying case of Lemma 3.1 on , contains a subgraph (d). Moreover, by Corollary 2.4. So is isometric to graph (d) of the following figure, which is isometric to (e) of Figure 1. If , it is impossible to get a graph with nonnegative Ricci curvature by a similar argument. This proves the lemma.




∎
Lemma 3.4.
Proof.
We only show (c) of Figure 13 can not be a subgraph of . Proofs for other cases are similar. We prove by contradiction. Suppose , then and contains a subgraph (a) of the following figure with by Corollary 3.2. This is impossible as and must has an extra neighbor with . Suppose . Then contains a subgraph (b) of the following figure with by Corollary 3.2. Applying case of Lemma 3.1 on , we get a contradiction. Suppose . Then contains a subgraph (c) of the following figure with by Corollary 3.2. We have by Theorem 2.3. This completes the proof.



∎
4. Classification of graphs
Let be a finite graph and be a diameter of . Then is a geodesic path. We explore all the available structures of separately by the following four cases.
Case (i): For all , . We discuss this case in Theorem 4.3.
Case (ii): There exists a such that . We discuss this case in Theorem 4.5.
Case (iii): There exists a such that . We discuss this case in Theorem 4.6.
Case (iv): There exists a such that . We discuss this case in Theorem 4.7.
We firstly introduce two lemmas.
Lemma 4.1.
Let be a finite graph. Suppose satisfying:
(1) ;
(2) The induced subgraphs and are complete graphs, i.e. for all , in , ;
(3) The induced subgraphs and are connected graphs.
Then contains a geodesic cycle with .


Proof.
Let { is a closed walk : , where , , and are walks in , , and respectively}. Let with shortest length, then it is a cycle. Moreover, we can prove that is a geodesic cycle. Suppose not, there are such that . We only prove the case that and , proofs for other cases are similar.
Let be a shortest walk between and in and be a shortest walk between and in crossing , . Note that and are disjoint paths from to and is the union of them.
Next, we claim that has one of the following forms: , , , , or . In fact, it is easy to see that is connected in and as is complete and is a shortest walk. The same conclusion holds for . So the claim holds.
Now, if , then and form a closed walk. By connecting part in and part in , we get a new closed walk with a shorter length. This contradicts to the fact that is a shortest walk. Proofs for other forms are similar. Then we proved for all , i.e. is a geodesic cycle. The inequality is obtained directly from the construction of . ∎
Lemma 4.2.
Suppose contains a geodesic cycle with , then and must be a cycle, a prism graph or a Möbius ladder.
Proof.
Suppose . We assume that there exists a vertex, say , has an extra neighbor , otherwise . Note that every path in with is geodesic in .
We claim that is a prism or Möbius ladder if . By applying local structure on in , we get that contains (a) or (b) of the following figure as a subgraph. By symmetric, has a subgraph (c) or (d) in the following figure. By doing this repeatedly, we get that has a subgraph (e). Finally we get a prism graph or a Möbius ladder.





We claim that is not a neighbor of . Suppose not, has a subgraph (a) of the following figure. Applying local structure on in , we know that has a subgraph (b) or (c) of the following figure. This is impossible by the symmetry of the geodesic cycle. So our claim is true.



The previous claim also shows . By applying local structure on in , we know that has one of the following subgraphs.



By applying local structures on as we did previously, we get the following results. If has a subgraph (c) in Figure 27, then is isomorphic to (a) of Figure 28. Otherwise must be isometric to (b), (c) or (d) of the following figure if is not a prism graph or Möbius ladder.




Note that the diameters of all the possible graphs we get are less than 6, thus they are not elements of . It follows that . We also have with similar arguments. When we do the previous steps on a geodesic cycle with , the graphs with the same forms shown in Figure 28 contain edges with negative curvature. This indicates that must be a cycle, prism graph or Möbius ladder.
∎
Theorem 4.3.
Let be a diameter of with . If for all , , then or isometric to a cycle.
Proof.
Lemma 4.4.
Let , be a diameter of . If there exists such that , then for all , and contains a subgraph shown in the following figure.

Proof.
If , it is trivial. Suppose and .
Firstly we prove . If , by applying local structure on and Corollary 3.2, we know that contains the following figure as a subgraph with . As , there is a vertex with . By applying local structure on , we get a contradiction. If , by applying local structure on and Corollary 3.2, we also get a contradiction. It follows that . By applying local structure on , we have . If , with the similar argument, we can also prove that .

By taking the previous steps repeatedly, we complete the proof. ∎
Theorem 4.5.
Let , , be a diameter of . If there exists a such that , then is a quasi-ladder or isometric to the particular graph (d) in Figure 1.
Proof.
Let . By Lemma 4.4, we have .
Case (i): . contains (a) of the following figure as a subgraph. We claim that or for all . If , by applying local structure on , we have that has a subgraph (b) of the following figure. By applying local structure on , we get a contradiction. If , we also get a contradiction by applying on . So or .


If , we have as shown in (a) of the following figure by applying local structure on . The previous argument tells us that . It is not hard to verify that neither. So and . By Lemma 4.4, has a subgraph (a) of the following figure. Moreover, , otherwise has an extra neighbor . Then by Corollary 2.4 and , which is contrary to . If , by Corollary 3.2 we have for and has a subgraph (b) of the following figure. Moreover, and is also a diameter of . We can redraw (b) as shown in (c) of the following figure, which contains (a) as a subgraph. So we only need to consider the situation that contains a subgraph (a).



As we see, the left side of is fixed, we start to study the structure of the right side. If , contains a subgraph (a) of the following figure by applying local structure on . If , we have and . With similar arguments, we classify all possible structures of the right side of as shown in the following figure.







Case (ii): . By applying on , contains a subgraph (a), (b) or (c) shown in the following figure. Obviously situation (c) is contained in (b) as in (c) is also a diameter. We only need to consider the situation that has a subgraph (a) or (b). It is easy to verify that in (a) and in (b). So the left-hand side is fixed. The right-hand side of is the same with that in case (i).



Case (iii): . If , the left side of is shown as following. The available structures of the right side are the same with that in case (i). If , has an additional available structure shown in (d) of Figure 1.

Case (iv): . It is easy to verify this case can not happen.
Case (v): . has a subgraph (a) or (b) shown in the following figure. As is also a diameter, graphs in this case have been discussed in the previous cases.


By previous discussions, we prove that must be a quasi-ladder graph with the following structures or isometric to the particular graph (d) in Figure 1.



∎
Theorem 4.6.
Let , , be a diameter of . If for all and there exists such that , then is a prism graph, a Möbius ladder or a quasi-ladder.
Proof.
Let , then .
Case (i): If , has a subgraph (a) of the following figure. By Corollary 3.2, has a subgraph (b) or (c) of the following figure. By switching and in (b), we get (b) is the same with (c). So we only consider the case that contains a subgraph (b).



Let , , and . If is connected to via , we get a geodesic cycle with by Lemma 4.1 and must be a prism graph or Möbius ladder by Lemma 4.2. In the following, we suppose that is not connected to via . With a similar argument in the proof of Theorem 4.5, the left side of has one of the following forms.


We start to deal with the right-hand side. If , we have by Corollary 2.4 and the structure of the right-hand side is (a) of the following figure. If has an extra neighbor , we have . Otherwise, has an extra neighbor and by Corollary 2.4. If , we have and the right-hand side structure is shown in (b) of the following figure. With similar arguments, we find all possible forms of the right side of as shown in the following figure.





Case (ii): . has a subgraph (a) of the following figure. The left side of must be (a) or (b) of the following figure and all possible forms of the right side have been shown in Figure 40.


Case (iii): . has a subgraph shown in the following figure which is the same with case (ii) by switching and , .

Case (iv): . contains a subgraph shown in the following figure which is the same with (a) of Figure 41 if , or (b) of Figure 41 if as is also a diameter.

By previous discussions, we prove that if is not a prism graph or a Möbius ladder, then it has one of the following forms, which are isometric to graphs in Figure 45.






∎
Theorem 4.7.
Let , , be a diameter of . Suppose that for all , , and there exists , then is a prism graph, a Möbius ladder or a quasi-ladder.
Proof.
In fact we will prove that graphs considered in this theorem have been discussed in the previous theorems. Let , then or .
Case (i): . Then and has a subgraph shown in figure (a). By applying local structure on , we get that has a subgraph (b) or (c). We have discussed those graphs by changing to .



Case (ii): . contains one of the followings as a subgraph, which have been discussed in the previous theorems by changing to .


∎
Proof of Theorem 1.2.
As is an infinite graph, we can find an infinite geodesic path which is isometric to or . By applying the arguments used for finite graphs, we get the conclusion. ∎
5. Applications
In this section, as an application of the main results, we establish the following strong Liouville property.
Theorem 5.1.
Let be an infinite graph, then all positive harmonic functions on are constant.
Proof.
By Theorem 1.2, is a line, half line, infinite ladder or infinite quasi-ladder. As line and infinite ladder are finitely generated Cayley graphs of Abelian groups, all positive harmonic functions on them are constant; see [CD60]. Besides, Hua and Münch also proved all positive harmonic functions are constant on graphs with two ends and nonnegative Ricci curvature, which provides another proof for the strong Liouville property of line and infinite ladder; see [HM21, Theorem 5.9]. For a half line, there is a stronger conclusion. In fact all harmonic functions on a half line are constant. So, we only consider the quasi-ladder case.
Let be a positive harmonic function on . If is not a constant, there exist , such that . Without loss of generality, we assume . By choosing suitable constants and such that satisfies and , we get a harmonic function which is bounded from below or above.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/3b15196d-1bf0-4716-abfc-a362310693c0/l1.png)
For any , we have
Let , . We have , . On the one hand, it is easy to check that . On the other hand, by the maximum principle. Thus , which means that for some constant . It follows that and as . We get a contradiction as is bounded from below or above. ∎
Acknowledgments
The authors are grateful to Prof. Bobo Hua for his guidance and support. The authors would also like to thank Florentin Münch and David Cushing for their helpful advice.
References
- [BCL+18] D. P. Bourne, D. Cushing, S. Liu, F. Münch, and N. Peyerimhoff. Ollivier-Ricci idleness functions of graphs. SIAM J. Discrete Math., 32(2):1408–1424, 2018.
- [BHL+15] Frank Bauer, Paul Horn, Yong Lin, Gabor Lippner, Dan Mangoubi, and Shing-Tung Yau. Li-Yau inequality on graphs. J. Differential Geom., 99(3):359–405, 2015.
- [BHLY20] Shuliang Bai, An Huang, Linyuan Lu, and Shing-Tung Yau. On the sum of ricci-curvatures for weighted graphs. arXiv preprint arXiv:2001.01776, 2020.
- [BLY21] Shuliang Bai, Linyuan Lu, and Shing-Tung Yau. Ricci-flat graphs with maximum degree at most 4. arXiv preprint arXiv:2103.00941, 2021.
- [CD60] Gustave Choquet and Jacques Deny. Sur l’équation de convolution . C. R. Acad. Sci. Paris, 250:799–801, 1960.
- [CG72] Jeff Cheeger and Detlef Gromoll. The splitting theorem for manifolds of nonnegative Ricci curvature. J. Differential Geometry, 6:119–128, 1971/72.
- [CKL+18] David Cushing, Riikka Kangaslampi, Yong Lin, Shiping Liu, Linyuan Lu, and Shing-Tung Yau. Ricci-flat cubic graphs with girth five. arXiv preprint arXiv:1802.02982, 2018.
- [CKL+19] David Cushing, Riikka Kangaslampi, Valtteri Lipiäinen, Shiping Liu, and George W Stagg. The graph curvature calculator and the curvatures of cubic graphs. Experimental Mathematics, pages 1–13, 2019.
- [CP18] Hee Je Cho and Seong-Hun Paeng. Classification of -Ricci flat graphs with girth at least five. Discrete Math., 341(10):2894–2902, 2018.
- [HLLY19] Paul Horn, Yong Lin, Shuang Liu, and Shing-Tung Yau. Volume doubling, Poincaré inequality and Gaussian heat kernel estimate for non-negatively curved graphs. J. Reine Angew. Math., 757:89–130, 2019.
- [HLYY18] Weihua He, Jun Luo, Chao Yang, and Wei Yuan. Ricci-flat graphs with girth four. arXiv preprint arXiv:1807.07253, 2018.
- [HM21] Bobo Hua and Florentin Münch. Every salami has two ends. arXiv preprint arXiv:2105.11887, 2021.
- [Hua19] Bobo Hua. Liouville theorem for bounded harmonic functions on manifolds and graphs satisfying non-negative curvature dimension condition. Calc. Var. Partial Differential Equations, 58(2):Paper No. 42, 8, 2019.
- [JMR19] Jürgen Jost, Florentin Münch, and Christian Rose. Liouville property and non-negative ollivier curvature on graphs. arXiv preprint arXiv:1903.10796, 2019.
- [LLY11] Yong Lin, Linyuan Lu, and Shing-Tung Yau. Ricci curvature of graphs. Tohoku Math. J. (2), 63(4):605–627, 2011.
- [LLY14] Yong Lin, Linyuan Lu, and S.-T. Yau. Ricci-flat graphs with girth at least five. Comm. Anal. Geom., 22(4):671–687, 2014.
- [LY10] Yong Lin and Shing-Tung Yau. Ricci curvature and eigenvalue estimate on locally finite graphs. Math. Res. Lett., 17(2):343–356, 2010.
- [MW19] Florentin Münch and Radosław K. Wojciechowski. Ollivier Ricci curvature for general graph Laplacians: heat equation, Laplacian comparison, non-explosion and diameter bounds. Adv. Math., 356:106759, 45, 2019.
- [Oll07] Yann Ollivier. Ricci curvature of metric spaces. C. R. Math. Acad. Sci. Paris, 345(11):643–646, 2007.
- [Oll09] Yann Ollivier. Ricci curvature of Markov chains on metric spaces. J. Funct. Anal., 256(3):810–864, 2009.
- [Sch99] Michael Schmuckenschläger. Curvature of nonlocal Markov generators. In Convex geometric analysis (Berkeley, CA, 1996), volume 34 of Math. Sci. Res. Inst. Publ., pages 189–197. Cambridge Univ. Press, Cambridge, 1999.
- [Yau75] Shing Tung Yau. Harmonic functions on complete Riemannian manifolds. Comm. Pure Appl. Math., 28:201–228, 1975.