Flattenability in some -dimensional normed vector space is a property of graphs which states that any embedding, or drawing, of the graph in some high dimension can be "flattened" down to live in -dimensions, such that the distances between pairs of points connected by edges are preserved. A graph is -flattenable if every distance constraint system (DCS) with as its constraint graph has a -dimensional framework. Flattenability was first called realizability,[1] but the name was changed to avoid confusion with a graph having some DCS with a -dimensional framework.[2]

Flattenability has connections to structural rigidity, tensegrities, Cayley configuration spaces, and a variant of the graph realization problem.

Definitions edit

A distance constraint system  , where   is a graph and   is an assignment of distances onto the edges of  , is  -flattenable in some normed vector space   if there exists a framework of   in  -dimensions.

A graph   is  -flattenable in   if every distance constraint system with   as its constraint graph is  -flattenable.

Flattenability can also be defined in terms of Cayley configuration spaces; see connection to Cayley configuration spaces below.

Properties edit

Closure under subgraphs. Flattenability is closed under taking subgraphs.[1] To see this, observe that for some graph  , all possible embeddings of a subgraph   of   are contained in the set of all embeddings of  .

Minor-closed. Flattenability is a minor-closed property by a similar argument as above.[1]

Flattening dimension. The flattening dimension of a graph   in some normed vector space is the lowest dimension   such that   is  -flattenable. The flattening dimension of a graph is closely related to its gram dimension.[3] The following is an upper-bound on the flattening dimension of an arbitrary graph under the  -norm.

Theorem. [4] The flattening dimension of a graph   under the  -norm is at most  .

For a detailed treatment of this topic, see Chapter 11.2 of Deza & Laurent.[5]

Euclidean flattenability edit

This section concerns flattenability results in Euclidean space, where distance is measured using the   norm, also called the Euclidean norm.

1-flattenable graphs edit

The following theorem is folklore and shows that the only forbidden minor for 1-flattenability is the complete graph  .

Theorem. A graph is 1-flattenable if and only if it is a forest.

 
Figure 1. A 2-dimensional framework of a DCS whose graph is a tree with six vertices is shown in blue. A 1-dimensional framework of the same DCS is shown in red.

Proof. A proof can be found in Belk & Connelly.[1] For one direction, a forest is a collection of trees, and any distance constraint system whose graph is a tree can be realized in 1-dimension. For the other direction, if a graph   is not a forest, then it has the complete graph   as a subgraph. Consider the DCS that assigns the distance 1 to the edges of the   subgraph and the distance 0 to all other edges. This DCS has a realization in 2-dimensions as the 1-skeleton of a triangle, but it has no realization in 1-dimension.

This proof allowed for distances on edges to be 0, but the argument holds even when this is not allowed. See Belk & Connelly[1] for a detailed explanation.

 
Figure 2. For certain linkages, this graph has a 1-dimensional realization (e.g. the assignment of 1 to every edge). However,   can be obtained by contracting the edge  , so the graph is not 1-flattenable.

2-flattenable graphs edit

The following theorem is folklore and shows that the only forbidden minor for 2-flattenability is the complete graph  .

Theorem. A graph is 2-flattenable if and only if it is a partial 2-tree.

Proof. A proof can be found in Belk & Connelly.[1] For one direction, since flattenability is closed under taking subgraphs, it is sufficient to show that 2-trees are 2-flattenable. A 2-tree with   vertices can be constructed recursively by taking a 2-tree with   vertices and connecting a new vertex to the vertices of an existing edge. The base case is the  . Proceed by induction on the number of vertices  . When  , consider any distance assignment   on the edges  . Note that if   does not obey the triangle inequality, then this DCS does not have a realization in any dimension. Without loss of generality, place the first vertex   at the origin and the second vertex   along the x-axis such that   is satisfied. The third vertex   can be placed at an intersection of the circles with centers   and   and radii   and   respectively. This method of placement is called a ruler and compass construction. Hence,   is 2-flattenable.

Now, assume a 2-tree with   vertices is 2-flattenable. By definition, a 2-tree with   vertices is a 2-tree with   vertices, say  , and an additional vertex   connected to the vertices of an existing edge in  . By the inductive hypothesis,   is 2-flattenable. Finally, by a similar ruler and compass construction argument as in the base case,   can be placed such that it lies in the plane. Thus, 2-trees are 2-flattenable by induction.

If a graph   is not a partial 2-tree, then it contains   as a minor. Assigning the distance of 1 to the edges of the   minor and the distance of 0 to all other edges yields a DCS with a realization in 3-dimensions as the 1-skeleton of a tetrahedra. However, this DCS has no realization in 2-dimensions: when attempting to place the fourth vertex using a ruler and compass construction, the three circles defined by the fourth vertex do not all intersect.

Example. Consider the graph in figure 2. Adding the edge   turns it into a 2-tree; hence, it is a partial 2-tree. Thus, it is 2-flattenable.

Example. The wheel graph   contains   as a minor. Thus, it is not 2-flattenable.

3-flattenable graphs edit

The class of 3-flattenable graphs strictly contains the class of partial 3-trees.[1] More precisely, the forbidden minors for partial 3-trees are the complete graph  , the 1-skeleton of the octahedron  ,  , and  , but  , and   are 3-flattenable.[6] These graphs are shown in Figure 3. Furthermore, the following theorem from Belk & Connelly[1] shows that the only forbidden minors for 3-flattenability are   and  .

 
Figure 3. The graphs of interest for 3-flattenability.

Theorem. [1] A graph is 3-flattenable if and only if it does not have   or   as a minor.

Proof Idea: The proof given in Belk & Connelly[1] assumes that  , and   are 3-realizable. This is proven in the same article using mathematical tools from rigidity theory, specifically those concerning tensegrities. The complete graph   is not 3-flattenable, and the same argument that shows   is not 2-flattenable and   is not 1-flattenable works here: assigning the distance 1 to the edges of   yields a DCS with no realization in 3-dimensions. Figure 4 gives a visual proof that the graph   is not 3-flattenable. Vertices 1, 2, and 3 form a degenerate triangle. For the edges between vertices 1- 5, edges   and   are assigned the distance   and all other edges are assigned the distance 1. Vertices 1- 5 have unique placements in 3-dimensions, up to congruence. Vertex 6 has 2 possible placements in 3-dimensions: 1 on each side of the plane   defined by vertices 1, 2, and 4. Hence, the edge   has two distance values that can be realized in 3-dimensions. However, vertex 6 can revolve around the plane   in 4-dimensions while satisfying all constraints, so the edge   has infinitely many distance values that can only be realized in 4-dimensions or higher. Thus,   is not 3-flattenable. The fact that these graphs are not 3-flattenable proves that any graph with either   or   as a minor is not 3-flattenable.

 
Figure 4. Construction steps to show the 1-skeleton of an octahedron is not 3-flattenable.[1]

The other direction shows that if a graph   does not have   or   as a minor, then   can be constructed from partial 3-trees,  , and   via 1-sums, 2-sums, and 3-sums. These graphs are all 3-flattenable and these operations preserve 3-flattenability, so   is 3-flattenable.

The techniques in this proof yield the following result from Belk & Connelly.[1]

Theorem. [1] Every 3-realizable graph is a subgraph of a graph that can be obtained by a sequence of 1-sums, 2-sums, and 3-sums of the graphs  ,  , and  .

Example. The previous theorem can be applied to show that the 1-skeleton of a cube is 3-flattenable. Start with the graph  , which is the 1-skeleton of a tetrahedron. On each face of the tetrahedron, perform a 3-sum with another   graph, i.e. glue two tetrahedra together on their faces. The resulting graph contains the cube as a subgraph and is 3-flattenable.

In higher dimensions edit

Giving a forbidden minor characterization of  -flattenable graphs, for dimension  , is an open problem. For any dimension  ,   and the 1-skeleton of the  -dimensional analogue of an octahedron are forbidden minors for  -flattenability.[1] It is conjectured that the number of forbidden minors for  -flattenability grows asymptotically to the number of forbidden minors for partial  -trees, and there are over   forbidden minors for partial 4-trees.[1]

An alternative characterization of  -flattenable graphs relates flattenability to Cayley configuration spaces.[2][7] See the section on the connection to Cayley configuration spaces.

Connection to the graph realization problem edit

Given a distance constraint system and a dimension  , the graph realization problem asks for a  -dimensional framework of the DCS, if one exists. There are algorithms to realize  -flattenable graphs in  -dimensions, for  , that run in polynomial time in the size of the graph. For  , realizing each tree in a forest in 1-dimension is trivial to accomplish in polynomial time. An algorithm for   is mentioned in Belk & Connelly.[1] For  , the algorithm in So & Ye[8] obtains a framework   of a DCS using semidefinite programming techniques and then utilizes the "folding" method described in Belk[6] to transform   into a 3-dimensional framework.

Flattenability under p-norms edit

This section concerns flattenability results for graphs under general  -norms, for  .

Connection to algebraic geometry edit

Determining the flattenability of a graph under a general  -norm can be accomplished using methods in algebraic geometry, as suggested in Belk & Connelly.[1] The question of whether a graph   is  -flattenable is equivalent to determining if two semi-algebraic sets are equal. One algorithm to compare two semi-algebraic sets takes   time.[9]

Connection to Cayley configuration spaces edit

For general  -norms, there is a close relationship between flattenability and Cayley configuration spaces.[2][7] The following theorem and its corollary are found in Sitharam & Willoughby.[2]

Theorem. [2] A graph   is  -flattenable if and only if for every subgraph   of   resulting from removing a set of edges   from   and any  -distance vector   such that the DCS   has a realization, the  -dimensional Cayley configuration space of   over   is convex.

Corollary. A graph   is not  -flattenable if there exists some subgraph   of   and some  -distance vector   such that the  -dimensional Cayley configuration space of   over   is not convex.

2-Flattenability under the l1 and l norms edit

The   and   norms are equivalent up to rotating axes in 2-dimensions,[5] so 2-flattenability results for either norm hold for both. This section uses the  -norm. The complete graph   is 2-flattenable under the  -norm and   is 3-flattenable, but not 2-flattenable.[10] These facts contribute to the following results on 2-flattenability under the  -norm found in Sitharam & Willoughby.[2]

Observation. [2] The set of 2-flattenable graphs under the  -norm (and  -norm) strictly contains the set of 2-flattenable graphs under the  -norm.

Theorem. [2] A 2-sum of 2-flattenable graphs is 2-flattenable if and only if at most one graph has a   minor.

The fact that   is 2-flattenable but   is not has implications on the forbidden minor characterization for 2-flattenable graphs under the  -norm. Specifically, the minors of   could be forbidden minors for 2-flattenability. The following results explore these possibilities and give the complete set of forbidden minors.

Theorem. [2] The banana graph, or   with one edge removed, is not 2-flattenable.

Observation. [2] The graph obtained by removing two edges that are incident to the same vertex from   is 2-flattenable.

Observation. [2] Connected graphs on 5 vertices with 7 edges are 2-flattenable.

The only minor of   left is the wheel graph  , and the following result shows that this is one of the forbidden minors.

Theorem. [11] A graph is 2-flattenable under the  - or  -norm if and only if it does not have either of the following graphs as minors:

  • the wheel graph   or
  • the graph obtained by taking the 2-sum of two copies of   and removing the shared edge.

Connection to structural rigidity edit

This section relates flattenability to concepts in structural (combinatorial) rigidity theory, such as the rigidity matroid. The following results concern the  -distance cone  , i.e., the set of all  -distance vectors that can be realized as a configuration of   points in some dimension. A proof that this set is a cone can be found in Ball.[12] The  -stratum of this cone   are the vectors that can be realized as a configuration of   points in  -dimensions. The projection of   or   onto the edges of a graph   is the set of   distance vectors that can be realized as the edge-lengths of some embedding of  .

A generic property of a graph   is one that almost all frameworks of distance constraint systems, whose graph is  , have. A framework of a DCS   under an  -norm is a generic framework (with respect to  -flattenability) if the following two conditions hold:

  1. there is an open neighborhood   of   in the interior of the cone  , and
  2. the framework is  -flattenable if and only if all frameworks in   are  -flattenable.

Condition (1) ensures that the neighborhood   has full rank. In other words,   has dimension equal to the flattening dimension of the complete graph   under the  -norm. See Kitson[13] for a more detailed discussion of generic framework for  -norms. The following results are found in Sitharam & Willoughby.[2]

Theorem. [2] A graph   is  -flattenable if and only if every generic framework of   is  -flattenable.

Theorem. [2]  -flattenability is not a generic property of graphs, even for the  -norm.

Theorem. [2] A generic  -flattenable framework of a graph   exists if and only if   is independent in the generic  -dimensional rigidity matroid.

Corollary. [2] A graph   is  -flattenable only if   is independent in the  -dimensional rigidity matroid.

Theorem. [2] For general  -norms, a graph   is

  1. independent in the generic  -dimensional rigidity matroid if and only if the projection of the  -stratum   onto the edges of   has dimension equal to the number of edges of  ;
  2. maximally independent in the generic  -dimensional rigidity matroid if and only if projecting the  -stratum   onto the edges of   preserves its dimension and this dimension is equal to the number of edges of  ;
  3. rigid in  -dimensions if and only if projecting the  -stratum   onto the edges of   preserves its dimension;
  4. not independent in the generic  -dimensional rigidity matroid if and only if the dimension of the projection of the  -stratum   onto the edges of   is strictly less than the minimum of the dimension of   and the number of edges of  .

References edit

  1. ^ a b c d e f g h i j k l m n o p q Belk, Maria; Connelly, Robert (2007). "Realizability of Graphs". Discrete & Computational Geometry. 37 (2): 125–137. doi:10.1007/s00454-006-1284-5. S2CID 12755057.
  2. ^ a b c d e f g h i j k l m n o p q Sitharam, M.; Willoughby, J. (2014). "On Flattenability of Graphs". Automated Deduction in Geometry. Lecture Notes in Computer Science. 9201: 129–148. doi:10.1007/978-3-319-21362-0_9. ISBN 978-3-319-21361-3. S2CID 23208.
  3. ^ Laurent, Monique; Varvitsiotis, Antonios (2012). "The Gram Dimension of a Graph". Combinatorial Optimization. Lecture Notes in Computer Science. Vol. 7422. pp. 356–367. doi:10.1007/978-3-642-32147-4_32. ISBN 978-3-642-32146-7. S2CID 18567150.
  4. ^ Barvinok, A. (1995). "Problems of distance geometry and convex properties of quadratic maps". Discrete & Computational Geometry. 13 (2): 189–202. doi:10.1007/BF02574037. S2CID 20628306.
  5. ^ a b Deza, Michel; Laurent, Monique (1997). Geometry of Cuts and Metrics. Springer-Verlag Berlin Heidelberg. doi:10.1007/978-3-642-04295-9. ISBN 978-3-540-61611-5.
  6. ^ a b Belk, Maria (2007). "Realizability of Graphs in Three Dimensions". Discrete & Computational Geometry. 37 (2): 139–162. doi:10.1007/s00454-006-1285-4. S2CID 20238879.
  7. ^ a b Sitharam, Meera; Gao, Heping (2010). "Characterizing Graphs with Convex and Connected Cayley Configuration Spaces". Discrete & Computational Geometry. 43 (3): 594–625. doi:10.1007/s00454-009-9160-8. S2CID 35819450.
  8. ^ So, Anthony Man-Cho; Ye, Yinyu (2006). "A semidefinite programming approach to tensegrity theory and realizability of graphs". Proceedings of the seventeenth annual ACM-SIAM symposium on Discrete algorithm - SODA '06. pp. 766–775. doi:10.1145/1109557.1109641. ISBN 0898716055. S2CID 10134807.
  9. ^ Basu, Saugata; Pollack, Richard; Marie-Francoise, Roy (2003). "Existential Theory of the Reals". Algorithms in Real Algebraic Geometry. Algorithms and Computation in Mathematics. Vol. 10. Springer, Berlin, Heidelberg. doi:10.1007/3-540-33099-2_14. ISBN 978-3-540-33098-1.
  10. ^ Witsenhausen, Hans S. (1986). "Minimum dimension embedding of finite metric spaces". Journal of Combinatorial Theory. Series A. 42 (2): 184–199. doi:10.1016/0097-3165(86)90089-0.
  11. ^ Fiorini, Samuel; Huynh, Tony; Joret, Gwenaël; Varvitsiotis, Antonios (2017-01-01). "The Excluded Minors for Isometric Realizability in the Plane". SIAM Journal on Discrete Mathematics. 31 (1): 438–453. arXiv:1511.08054. doi:10.1137/16M1064775. hdl:10356/81454. ISSN 0895-4801. S2CID 2579286.
  12. ^ Ball, Keith (1990). "Isometric Embedding in lp-spaces". European Journal of Combinatorics. 11 (4): 305–311. doi:10.1016/S0195-6698(13)80131-X.
  13. ^ Kitson, Derek (2015). "Finite and Infinitesimal Rigidity with Polyhedral Norms". Discrete & Computational Geometry. 54 (2): 390–411. arXiv:1401.1336. doi:10.1007/s00454-015-9706-x. S2CID 15520982.