A universal threshold for geometric embeddings of trees
Abstract.
A graph is geometrically embeddable into a normed space when there is a mapping such that if and only if , for all distinct . Our result is the following universal threshold for the embeddability of trees. Let , and let be sufficiently large in terms of . Every –vertex tree of maximal degree at most is embeddable into any normed space of dimension at least , and complete trees are non-embeddable into any normed space of dimension less than . In striking contrast, spectral expanders and random graphs are known to be non-embeddable in sublogarithmic dimension. Our result is based on a randomized embedding whose analysis utilizes the recent breakthroughs on Bourgain’s slicing problem.
1. Introduction
Given a normed space and a graph , a map is a geometrical embedding of into if
Say that is non-embeddable into if no such embedding exists. Note that the mapping exists if and only if is isomorphic to an intersection graph in generated by parallel translates of the unit ball .
The problem of identifying geometrically embeddable graphs has attracted significant attention, especially in the Euclidean setting. It was proved in [FM88] (see, also, [RRS89] and [RRS92]) that every –vertex graph of maximal degree at most admits a geometrical embedding into a Euclidean space of dimension , where is a universal constant. On the other hand, the Poincaré inequality implies that any spectral expander graph is non-embeddable into Euclidean space of sublogarithmic dimension [AT25]. More generally, spectral expanders also do not embed into any spaces of sublogarithmic dimension satisfying a so-called nonlinear Poincaré inequality (see, for example, [Na14, MN14, Esk22, ADTT24] for a discussion of such inequalities).
The smallest such that a graph can be embedded into an –dimensional Euclidean space is called the sphericity of , and denoted by . The aforementioned results then imply that, suppressing dependence on , the sphericity of every –vertex expander graph is of order . We refer to [MP93, Section 2] and [BL05, Section 3] for a review of results on sphericity of various families of graphs. In the non-Euclidean setting, it was recently shown in [AT25] that with high probability, a random –regular graph on vertices cannot be geometrically embedded into any normed space of dimension . It remains an interesting open question whether this dimension bound is asymptotically optimal across all normed spaces.
The purpose of this note is to study geometrical embeddings of trees into arbitrary normed spaces. The setting of trees is fundamentally different from embedding random graphs despite the fact that sparse random graphs are locally tree-like, i.e., have few short cycles. As an illustration, in [FM86] it is shown that the complete rooted binary tree on vertices admits a geometrical embedding into of dimension at most . (Recall that, in contrast, a random –regular graph on vertices does not embed geometrically into a Euclidean space of dimension .) On the other hand, a slight adaptation of a volumetric argument from the same paper [FM86] (which we reproduce in Section 4 for the reader’s convenience) yields that is not geometrically embeddable into any normed space of dimension less than , for a universal constant .
It is therefore natural to ask if the “transition window” for complete binary trees is universal across all norms. We completely resolve this problem and, moreover, provide embedding results for arbitrary bounded degree trees.
Theorem 1.1(Universal threshold for embedding of trees).
There exist absolute positive constants , and for every integer a positive integer depending only on with the following property. Let be a finite-dimensional normed space, and let be an integer.
If , then there exists a tree on vertices with maximum degree at most that is not geometrically embeddable into .
On the other hand, if , then every tree on vertices with maximum degree at most is geometrically embeddable into .
Remark 1.2.
The proof of Theorem 1.1 yields the estimate .
Let us discuss the result in the context of intersection graphs. Given a collection of parallel translates of a convex body in , the intersection graph is formed by placing an edge between and for all pairs of translates and that have a nonempty intersection. Our result implies that for any centrally-symmetric convex body , every bounded degree –vertex tree can be realized as an intersection graph in an –dimensional vector space, using translates of .
Corollary 1.3.
Let be a positive integer, and let and be as in Theorem 1.1. Also let be an integer, let be a tree on vertices with maximum degree at most , and let be any centrally-symmetric convex body in a Euclidean space of dimension . Then there exists an intersection graph for translates of which is isometric to .
Remark 1.4.
The geometric embedding obtained in the proof of part of Theorem 1.1 is -Lipschitz when the tree is equipped with the shortest-part distance , and it satisfies for all . That is, the inverse of is . This should be compared with a well-known result of Bourgain [Bou86] (see, also, [Ma99]) asserting that any embedding of the complete rooted binary tree on vertices into a Hilbert space111In fact, binary trees do not embed with bounded bi-Lipschitz distortion into any super-reflexive spaces (e.g., all spaces for .) necessarily has bi-Lipschitz distortion at least .
We conclude with some open research questions, followed by a proof overview.
Problem 1.5(Optimal constants in Theorem 1.1).
Determine optimal values of the constants and in Theorem 1.1. Are they equal up to additive term?
Problem 1.6(Trees of growing maximal degree).
Determine an explicit function and some universal constants and so that the following holds. Any tree on vertices with maximum degree embeds into any normed space of dimension at least . And, no complete tree of degree embeds into any space of dimension less than .
Proof overview
As already mentioned, part of Theorem 1.1 is known, and it appears (in a slightly different form) in the work of Frankl and Maehara [FM86]. We also note that a tree certifying the first part of the theorem can be constructed explicitly: if for an integer , then can be taken to be the complete rooted -regular tree of height .
Part is the bulk of the proof of Theorem 1.1. Given a normed space and a tree , we attach to each edge of an independent vector drawn uniformly from the unit ball . After arbitrarily choosing a root , each vertex is embedded by summing the vectors on the edges along the path from to , plus an independent “regularizing” Gaussian. Our goal is to show that this embedding is geometric with positive probability.
The main challenge is to verify that, with a non-zero probability, all pairs of non-adjacent vertices are at distance greater than one under the embedding. Note that the spatial distance between the images of two vertices with graph distance is given by a sum of independent vectors, plus some Gaussian regularization. To control the distance, we utilize small ball probability estimates for log-concave distributions; in particular, we apply thin-shell estimates (for small ), or density bounds on isotropic log-concave vectors that arise in the context of the recent resolution of the slicing conjecture [KL24b]. In either case, these probability estimates are not strong enough to survive the trivial union bound for arbitrary –vertex trees. To deal with this problem, we apply a combination of a conditioning argument (using the aforementioned Gaussian regularization) and the asymmetric Lovász local lemma.
2. Background material
2.1. Notation
All graphs in this paper have no multiple edges and no self-loops. Given a graph , we denote its vertex set by and its edge set by . For every , the neighbourhood of in is denoted by , and denotes the shortest-path graph distance on . Recall that a tree is a connected graph with no cycles. For a normed space , denotes the closed unit ball.
Definition 2.1.
Let be a tree, and let . We say that a set connects and if there exist , where , such that , and .
If , then denotes the smallest, under inclusion, set of edges that connects and ; by convention, we set if . We shall refer to sets of the form as paths.
2.2. Isotropic random vectors, and normed spaces in isotropic position
Recall that a random vector in is isotropic if its components are mean-zero random variables, and its covariance matrix is the identity.
Now, let be an -dimensional normed space, and let be a random vector in uniformly distributed in . We say that is in isotropic position if the random vector is isotropic; we emphasize that in asymptotic geometric analysis literature, an alternative definition involving normalization of the Lebesgue volume of is used occasionally. It is a standard fact that any finite-dimensional normed space can be placed in isotropic position by applying an invertible linear transformation (see, e.g., [KL24a]).
2.3. Asymptotic geometric analysis
We will need two deep results from asymptotic geometric analysis. The first one the following thin-shell estimate due to Klartag [Kl07].
Theorem 2.2(Thin-shell concentration for log-concave measures).
There exist absolute constants and with the following property. If is an isotropic random vector in with a log-concave222A function is log-concave if the set is convex and, additionally, the function is concave. distribution, then
(2.1) |
We will also need the following very recent result of Klartag and Lehec [KL24b] that yields an affirmative solution to the celebrated Bourgain’s slicing conjecture.
Theorem 2.3(Anticoncentration for log-concave measures via slicing).
There exists an absolute constant with the following property. If is an isotropic random vector in with a log-concave density , then for any ,
(2.2) |
Remark 2.4.
Thin-shell estimates for log-concave distributions have been actively studied in the asymptotic geometric analysis literature, leading to strengthening of the result in [Kl07], as a consequence of the tremendous recent progress on the KLS conjecture and the aforementioned resolution of the slicing conjecture. We do not attempt to give here a thorough overview of the history of these problems, and refer instead to recent works [Ch21, KL24a, Gu24, KL24b] for further references. Our proof of the main result is not sensitive to the value of the constant in Theorem 2.2.
2.4. Gaussian concentration
We will need the following simple consequence of gaussian concentration for Lipschitz functions.
Fact 2.5.
If is a random vector in with i.i.d. standard normal entries, then for any ,
(2.3) |
Proof.
Notice that the map defined by , is -Lipschitz, and it satisfies . Using these observations, the result follows from the standard gaussian concentration for Lipschitz functions (see, e.g., [Le01]). ∎
2.5. John’s theorem
Recall that if and are two normed spaces of the same dimension, then their Banach–Mazur distance is defined by
Also recall that, by a classical result due to John (see, e.g., [JL01]), we have for any -dimensional normed space . We will use this estimate in the following form.
Fact 2.6.
Let be an -dimensional normed space. Then there exist vectors such that for any ,
(2.4) |
2.6. Lovász local lemma
The last result we will need is the so-called asymmetric version of Lovász local lemma (see, e.g., [AS16]).
Lemma 2.7(Asymmetric Lovász local lemma).
Let be a graph, and let be a collection of measurable events in some probability space such that for every , the event is independent of the algebra generated by the events . Assume that there exists such that for any ,
(2.5) |
Then,
(2.6) |
3. Preparatory lemmas
In this section we shall present three preparatory lemmas that are used in the proof of Theorem 1.1. We start with the following volumetric estimate.
Lemma 3.1.
There exists an absolute constant such that for any -dimensional normed space in isotropic position, we have , where denotes the Lebesgue measure in .
Proof.
Let be a random vector uniformly distributed in . Then is an isotropic, log-concave, random vector in with density function . By Theorem 2.2, there exists an absolute constant such that
Therefore,
Lemma 3.2.
Proof.
The next lemma complements Lemma 3.2, and it will used when is smaller than the threshold appearing in the right-hand-side of (3.1).
Lemma 3.3.
There exists a positive integer with the following property. Let be an integer, let be a normed space in isotropic position, let be an integer, and let be i.i.d. random vectors uniformly distributed in . Then,
(3.4) |
where is as in Theorem 2.2.
Proof.
Set , , and . Notice that
(3.5) |
As in the proof of Lemma 3.2, we observe that the random vector is isotropic and log-concave. Therefore, by Theorem 2.2, if is sufficiently large, we have
(3.6) |
Notice that . Hence, by Theorem 2.2 and assuming that is sufficiently large, we get that
It follows immediately that, conditioned on any realization of , the probability that belongs to , is bounded above by
where we have used again the fact that can be taken sufficiently large. Combining the last estimate with (3.5)–(3.6), we get the result. ∎
4. Proof of Theorem 1.1
As mentioned, part of Theorem 1.1 is due to Frankl and Maehara [FM86]. For the convenience of the reader, we recall their argument. We set
(4.1) |
Fix an integer , and let be an arbitrary positive integer that is sufficiently large in terms of . We define a tree on vertices with maximum degree as follows. There exists a root such that, setting , the following hold.
For every , we have .
For every with , we have .
Notice, in particular, that if , then is just the -regular tree of height . Let be a normed space for which there exists a geometric embedding . Since is a bipartite graph (being a tree), there exists an independent set with . Using the fact that every node of has distance from at most , we see that
(4.2) |
On the other hand, since is a geometric embedding, for every with we have that . Thus, denoting by the Lebesgue measure in , by (4.2),
which is easily seen to imply, if is sufficiently large, that , as desired.
We proceed to the proof of part . We shall construct a random embedding of a given tree on vertices with maximum degree in the normed space , and we shall show that with positive probability this is a geometric embedding. To that end we start by setting
(4.3) |
where is as in Theorem 2.2. The parameter will be determined in the course of the proof. For notational convenience, we shall denote by the vertex set of , and by its edge set. Clearly, we may assume that and that is in isotropic position.
4.1. Construction of the random embedding
Let be a collection of i.i.d. random vectors uniformly distributed in . Also let be a collection of i.i.d. scalar standard normal random variables that are independent of . (As usual, denotes the discrete interval of length .) Let be as in Fact 2.6, and define for every ,
(4.4) |
Fix a distinguished vertex that we view as the root of , and set
(4.5) |
We define the random embedding by
with the convention that the sum over the empty set is equal to the zero vector. (Here, is as in Definition 2.1.) For every with , define the “good” event
Part of Theorem 1.1 will follow once we show that
(4.6) |
4.2. Auxiliary events
Set
(4.7) |
where is as in Lemma 3.1, and is as in Theorem 2.3. Taking sufficiently large, we may assume that .
Next, for every with , define the event
where is as in Theorem 2.2. Our goal in this subsection is to show that
(4.8) |
4.2.1. The “path” graph
Define the “path” graph , which will act as the dependency graph in our application of the Lovász local lemma, by setting
Moreover, for every set .
Claim 4.1.
Let , and let . Then,
Proof.
Clearly, is upper bounded by . Thus, it suffices to show that
(4.9) |
Let be arbitrary, and set .
4.2.2. Application of asymmetric Lovász local lemma
Set
since can be taken sufficiently large, we may assume that for every we have
(4.15) |
For every set
and define by . By the choice of and (4.15), we see that . Hence, by Claim 4.1, for every ,
(4.16) |
Claim 4.2.
For every , we have .
Proof.
Fix , and set . We distinguish three cases.
First, assume that . Since the random vectors and have the same distribution, then provided that is sufficiently large, by Lemma 3.3,
(4.17) |
On the other hand, by the choice of in (4.7),
(4.18) |
Since and using the fact that can be taken sufficiently large, the claim follows in this case by (4.17) and (4.18).
Next, assume that . Using again the fact that and have the same distribution, by Lemma 3.2, the choice of in (4.7) and the fact that , we obtain that
(4.19) |
Since the function is increasing for and , we see that
Taking sufficiently large and observing that and , we conclude that
(4.20) |
Finally, assume that . As before, since and have the same distribution, by Lemma 3.2, the choice of in (4.7) and the fact that and , we have
(4.21) |
Taking sufficiently large, we may assume that . Thus, by (4.21) and the choice of in (4.3), we have that
(4.22) |
Since , taking sufficiently large, we see that
(4.23) | ||||
Thus, this final case the claim follows from (4.22) and (4.23). ∎
4.3. Completion of the proof of part of Theorem 1.1
Let be the vectors selected in Subsection 4.1. It is convenient to introduce an auxiliary norm in as follows. Given written as , we set . By (2.4), we see that for every ,
(4.24) |
Moreover, recalling (4.4) and Fact 2.5, for every subset of the edge-set of and every ,
(4.25) |
Now, set
Note that in order to show (4.6), it suffices to prove that for every with ,
(4.26) |
To that end, let with and set . Notice that
(4.27) |
moreover, due to the independence of and , for every measurable ,
(4.28) |
We consider four cases according to the value of .
Case 1:
Case 2:
Case 3:
Case 4:
By (4.27), (4.28) and (4.24) and the fact that is symmetric for every , we have
(4.32) | ||||
Next observe that, by the choices of and ,
(4.33) |
Moreover, the random variable has the same distribution with , where denotes a random vector in with i.i.d. standard normal entries. Thus, by (4.32),
(4.34) |
Since the density function of is pointwise bounded by , this implies
(4.35) |
where is as in Lemma 3.1. Finally, taking sufficiently large, we conclude that
The above cases are exhaustive, and so the entire proof of Theorem 1.1 is completed.
Acknowledgments
The research was supported in the framework of H.F.R.I call “Basic research Financing (Horizontal support of all Sciences)” under the National Recovery and Resilience Plan “Greece 2.0” funded by the European Union–NextGenerationEU (H.F.R.I. Project Number: 15866). The third named author (K.T.) is partially supported by the NSF grant DMS 2331037.
References
- [AT25] D. J. Altschuler and K. Tikhomirov, Universal geometric non-embedding of random regular graphs, preprint (2025), available at https://arxiv.org/abs/2501.09142.
- [ADTT24] D. J. Altschuler, P. Dodos, K. Tikhomirov and K. Tyros, A combinatorial approach to nonlinear spectral gaps, preprint (2024), available at https://arxiv.org/abs/2410.04394.
- [AS16] N. Alon and J. H. Spencer, The Probabilistic Method, Wiley-Interscience Series in Discrete Mathematics and Optimization, John Wiley & Sons, 2016.
- [BL05] Y. Bilu and N. Linial, Monotone maps, sphericity and bounded second eigenvalue, J. Combin. Theory Ser. B 95 (2005), 283–299.
- [Bou86] J. Bourgain, The metrical interpretation of superreflexivity in Banach spaces, Israel J. Math. 56 (1986), 222–230.
- [Ch21] Y. Chen, An almost constant lower bound of the isoperimetric coefficient in the KLS conjecture, Geom. Funct. Anal. 31 (2021), 34–61.
- [Esk22] A. Eskenazis, Average distortion embeddings, nonlinear spectral gaps, and a metric John theorem [after Assaf Naor], Astérisque No. 438, Séminaire Bourbaki. Vol. 2021/2022. Exposés 1181–1196 (2022), Exp. No. 1188, 295–333.
- [FM86] P. Frankl and H. Maehara, Embedding the -cube in lower dimensions, European J. Combin. 7 (1986), 221–225.
- [FM88] P. Frankl and H. Maehara, The Johnson–Lindenstrauss lemma and the sphericity of some graphs, J. Combin. Theory Ser. B 44 (1988), 355–362.
- [Gu24] Q. Guan, A note on Bourgain’s slicing problem, preprint (2024), available at https://arxiv.org/abs/2412.09075.
- [JL01] W. B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, in “Handbook of the Geometry of Banach Spaces. Volume 1”, Elsevier, 2001, 1–84.
- [Kl07] B. Klartag, Power-law estimates for the central limit theorem for convex sets, J. Funct. Anal. 245 (2007), 284–310.
- [KL24a] B. Klartag and J. Lehec, Isoperimetric inequalities in high-dimensional convex sets, preprint (2024), available at https://arxiv.org/abs/2406.01324.
- [KL24b] B. Klartag and J. Lehec, Affirmative resolution of Bourgain’s slicing problem using Guan’s bound, preprint (2024), available at https://arxiv.org/abs/2412.15044.
- [Le01] M. Ledoux, The Concentration of Measure Phenomenon, Mathematical Surveys and Monographs, Vol. 89, American Mathematical Society, 2001.
- [Ma99] J. Matoušek, On embedding trees into uniformly convex Banach spaces, Israel J. Math. 114 (1999), 221–237.
- [MN14] M. Mendel and A. Naor, Nonlinear spectral calculus and super-expanders, Publ. Math. Inst. Hautes Étud. Sci. 119 (2014), 1–95.
- [MP93] W. Moser and J. Pach., Recent developments in combinatorial geometry, New trends in discrete and computational geometry (1993), 281–302.
- [Na14] A. Naor, Comparison of metric spectral gaps, Anal. Geom. Metr. Spaces 2 (2014), 1–52.
- [RRS89] J. Reiterman, V. Rödl and E. Šiňajová, Geometrical embeddings of graphs, Discrete Math. 74 (1989), 291–319.
- [RRS92] J. Reiterman, V. Rödl and E. Šiňajová, On embedding of graphs into Euclidean spaces of small dimension, J. Combin. Theory Ser. B 56 (1992), 1–8.