Isometric embeddings of snowflakes into finite-dimensional Banach spaces
Abstract.
We consider a general notion of snowflake of a metric space by composing the distance by a nontrivial concave function. We prove that a snowflake of a metric space isometrically embeds into some finite-dimensional normed space if and only if is finite. In the case of power functions we give a uniform bound on the cardinality of depending only on the power exponent and the dimension of the vector space.
2010 Mathematics Subject Classification:
30L05; 46B85; 54C25; 54E40; 28A80.1. Introduction
Isometric embeddings of metric spaces into infinite-dimensional Banach spaces have a long tradition. Classical results are due to Fréchet, Urysohn, Kuratowski, Banach, [Fré10, Ury27, Kur35, Ban55]. For an introduction to the subject we refer to Heinonen’s survey [Hei03]. The case of embeddings into finite-dimensional Banach spaces is much harder, even when one considers bi-Lipschitz embeddings in place of isometric embeddings. It is a wide open problem to give an intrinsic characterizations of those metric spaces which admit bi-Lipschitz embeddings into some Euclidean space. See for example [Sem99, Luo96, LP01, Seo11, LN14].
The situation is quite different for quasisymmetric maps (see [Hei01, Chapter 10-12] for an introduction to the theory of quasisymmetric embeddings). A metric space quasisymmetrically embeds into some Euclidean spaces if and only if it is doubling (see [Hei01, Theorem 12.1]). More specifically, Assouad proved the following result (see [Ass83], and also [NN12, DS13]): if is a doubling metric space and then the metric space admits a bi-Lipschitz embedding into some Euclidean space. If is the Euclidean distance and then the metric space is bi-Lipschitz equivalent to the von Koch snowflake curve, so is said to be the -snowflake of .
The Assouad Embedding Theorem is sharp in that there are examples (none of which are trivial) of doubling spaces which do not admit bi-Lipschiz embeddings into any Euclidean space, even though each of their -snowflakes do. See [Sem96a, Sem96b, Laa02, CK10]. We also stress that it has been known that snowflakes of doubling spaces in general do not isometrically embed in any Euclidean space. Indeed, the space does not, see [Hei03, Remark 3.16(b)].
The main aim of this paper is to show that if some -snowflake of a metric space isometrically embedds into a finite dimensional Banach space then the metric space in question is finite. Our main result is the following.
Theorem 1.1.
For any and there is an such that if a metric space has cardinality at least then does not admit an isometric embedding into any -dimensional normed linear space.
The techniques that we use in Section 3.3 for the proof of Theorem 1.1 can also be used to study more general notions of snowflakes. For this purpose, we introduce general snowflaking functions. We say that a function is a snowflaking function if the following hold:
-
(S1)
.
-
(S2)
is concave.
-
(S3)
, as .
-
(S4)
, as .
Let be a snowflaking function. Then function is weakly increasing and, if is a metric on a set then is also a metric on . Given a snowflaking function and a metric space we say that the metric space is the -snowflake of . If for some then is the -snowflake of . In Section 3.2 we prove the following.
Theorem 1.2.
Let be a snowflaking function and a metric space. If the -snowflake of admits an isometric embedding into some finite-dimensional Banach space then is finite.
Remark 1.3.
Note that for general snowflaking functions there may not be any bound on the number of points one can embed, see Remark 3.18. If one removes either of the requirements (S3) or (S4), then we say that is a degenerate snowflake (at zero or at infinity, respectively). Indeed, in such cases the conclusion of Theorem 1.2 does not hold, in general, see Proposition 3.14.
We conclude the introduction with few other simple observations about embeddings into Euclidean spaces. Every -snowflake of , , isometrically embeds into the Hilbert space of square summable sequences, see [Hei03, Remark 3.16(d)]. For any , there is a metric space of cardinality such that for any its -snowflake can be isometrically embedded into (just take the vertices of the standard simplex). There is a 4-point metric space which has an -snowflake which cannot be isometrically embedded into , and so cannot be isometrically embedded into any Euclidean space, (just take the vertices of the (3,1) complete bipartite graph). Every finite metric space has an -snowflake which admits an isometric embedding into some Euclidean space, see Proposition 2.2 below.
2. Corollaries
In this section we record a few small results and corollaries. First of all, note that as a corollary of Theorem 1.2, by Ascoli-Arzelà Theorem we immediately obtain:
Corollary 2.1.
Suppose that is infinite. For any snowflaking function and there is a such that does not admit a -bilipschitz embedding into any normed linear space of dimension .
Theorem 1.1 shows that there is a bound (depending on and ) on the cardinality of a metric space whose -snowflake admits an isometric embedding into . The next easy result says that any finite metric space can be isometrically embedded in some Euclidean space after some -snowflaking.
Proposition 2.2.
If is a finite metric space of cardinality , then some -snowflake of admits an isometric embedding into the Euclidean space .
Corollary 2.3.
Given a metric space , there is an -snowflake of that isometrically embedds into some finite-dimensional normed space if and only if is finite.
Proof of Proposition 2.2.
We show that admits an isometric embedding into when is sufficiently close to . In fact we show that there is an such that if is any metric on satisfying for all distinct then admits an isometric embedding into . The proposition follows as as for any distinct .
Consider the points where is the canonical basis of . Thus for all distinct . The proposition is thus a direct consequence of the following claim:
Claim A. There is an such that if , with are such that , then there exist such that for all .
To show the claim, consider the vector subspace of defined by the upper-triangular matrices, i.e., the elements of are elements of the form with and . Let be given by
Denote by the vector of vectors . Notice that .
Claim A holds if and only if lies in the interior of the image of . Thus it suffices to show that the differential has maximal rank and apply the inverse function theorem.
Denote by the -th component of an -vector . Notice that , where is the Kronecker symbol. Set Notice that span the set . Let us show that any is in the image of . Fix and so that . By direct calculation,
Evaluating at , we get
Since , we have and . Since , and , we have . Hence
This concludes the proof of Claim A. ∎
3. Proofs of the main results
In the next subsection we first give two simple geometric lemmas which allow us to prove our results for non-euclidean normed vector spaces. The proofs of Theorem 1.1 and Theorem 1.2 rely on the Ramsey-theoretic fact that given sufficiently many points in there must be a triple that forms as large angle as we want. We recall this result at the end of the subsection in Lemma 3.5. In the following two subsections we then prove Theorem 1.1 and Theorem 1.2. In the final subsection we prove that if is a degenerate snowflaking function then there exists an infinite metric space whose -snowflake isometrically embeds into -dimensional Euclidean space.
3.1. Geometric lemmas
Throughout this section is an -dimensional normed linear space with norm . For the remainder of this section we fix an inner product on for which John ellipsoid property holds. Namely, we have
(3.1) |
where is the -unit ball and is the the unit ball in the metric associated to the inner product . For we denote the length given by the inner product by .
Given two vectors let be their angle with respect to the above inner product :
Given three points we set .
Let and let be the projection of onto the line through and .
Lemma 3.2.
There exist constants depending only on such that if then
Proof.
By the John Ellipsoid Theorem, (3.1), there is an angle and a height only depending on such that for any point on the sphere with respect to . The cone at with opening and height in the direction of is inside the ball, whereas the cone at with opening and height in the direction of is outside the ball.
The inequality that we need to prove is translation and dilation invariant. So we assume that and that lies on the unit sphere. Let . We shall require for a small enough . Let be as in the picture.
If is small enough, then Setting we have
Also, if is small enough, then the point is in and so . In particular, between and there is a point such that .
We conclude using, in order, the definition of , the fact that on a line any two norms are a multiple of each other, the triangle inequality, the properties of and , and the previous bound:
where . ∎
The following bound is another easy consequence of the John Ellipsoid theorem.
Lemma 3.3.
There exists a constant depending only on such that when we have
(3.4) |
Proof.
Our arguments rely on the following Ramsey-theoretic result. Explicit bounds on the number of points that one can have in without forming an angle larger than a given bound can be found in [EF83]. A proof of Lemma 3.5 can also be found in [KS11].
Lemma 3.5.
For any and there is an such that if has cardinality at least then there are distinct such that .
3.2. Proof of Theorem 1.1
The proof of Theorem 1.1 as well as the proof of Theorem 1.2 combines two observations: snowflaking forbids the formation of large angles whereas the fact that we have many points forces such angles to exist. In the proof of Theorem 1.1 the special form of the snowflaking function allows us to directly prove a bound on the cardinality of the snowflaked space that can be embedded in the normed vector space.
Proof of Theorem 1.1.
We let be the constant in Lemma 3.5 with . We suppose towards a contradiction that has cardinality and that there is an isometric embedding of into an -dimensional normed vector space . By Lemma 3.5 there exist three isometrically embedded points such that . We declare We may assume that . We also declare and . Note that .
Let be the orthogonal projection of on the line passing through and , i.e. . Lemma 3.2 yields
(3.6) |
and
(3.7) |
We now estimate
from above. Subadditivity of yields:
Estimating the obtained terms from above using (3.7) and (3.6) we have
Now we use the fact that and Lemma 3.3 to obtain
and similarly
Again by we get . Collecting the estimates together and using the fact that we have:
Therefore
This contradicts the triangle inequality in . ∎
3.3. Proof of Theorem 1.2
In the proof of Theorem 1.2 we use the same geometric lemmas (Lemma 3.2 and Lemma 3.3) as in the proof of Theorem 1.1. However, in the proof of Theorem 1.2 the choice of a sequence of points giving the contradiction depends not only on the snowflaking function , but also on the first element of the sequence. Therefore no upper bound (depending on and ) on the number of points that can be snowflake embedded can in general be obtained in Theorem 1.2.
Proof of Theorem 1.2.
Suppose to the contrary that is infinite and that there exists an isometric embedding where is an -dimensional normed vector space. We divide our proof into two cases. An infinite bounded subset of is not discrete, so one of the following holds:
-
(i)
is unbounded;
-
(ii)
is not discrete.
If (i) holds we will arrive at a contradiction with the condition (S4) of a snowflaking function. If (ii) holds, a contradiction follows with (S3).
Case (i): Suppose is unbounded
Observe that (S4) implies that the existence of a function such
that for any and we have
(3.8) |
Combining (3.8) with (S1) and (S2) we get
(3.9) |
Now fix , . Since is unbounded, there exists a point with and for . We continue inductively. Suppose have been chosen. Now we select satisfying
(3.10) |
Let be the constants from Lemma 3.2 and Lemma 3.3. Set
(3.11) |
By Lemma 3.5 there exist in such that . By the condition (3.10), there exist with such that , , . Let be the orthogonal projection of to the line passing through and . On the one hand, we have that
On the other hand, first notice that since is concave and positive, then has to be weakly increasing. Secondly, by the definition of the function we then have that for any three with
(3.12) |
Therefore we have a contradiction.
Case (ii): Suppose is not discrete
This time we observe that (S3) implies the existence of a function such
that for all and for any and we have
(3.8), and hence (3.9), using (S1) and (S2).
Let be an accumulation point of . First we select . Next we take a radius so that for all we have both and for . Now we select a point . We continue inductively. Suppose have been chosen. Now we take a radius such that for all we have
(3.13) |
Then we select a point .
3.4. Necessity of (S3) and (S4)
We end this paper by showing that the conditions (S3) and (S4) of generalized snowflakes are indeed needed for Theorem 1.2 to hold.
Proposition 3.14.
Suppose satisfies (S1) and (S2) but fails to satisfy (S3) or (S4). Then there is an infinite metric space such that admits an isometric embeddeding into -dimensional Euclidean space.
Proof.
We only treat the case where (S4) fails. The case where (S3) fails follows in a similar way.
We construct a sequence of points in such that is a metric space. For this purpose we fix a sequence of positive angles such that . Depending on the sequence and the function we construct an increasing sequence of positive real numbers which determine the Euclidean distance between and . For notational convenience we let . Notice that by assumption as . This allows us to select for every a real number such that for all we have
(3.15) |
Now, using the sequences and we define the sequence as follows. We set , , and inductively for declare
In order to see that is a metric space we need to check that the triangle inequality holds. For this purpose let be three integers. Let be the Euclidean metric. The only nontrivial inequality that we have to verify is
Denoting and the above inequality is equivalent to
(3.16) |
By (S1) and (S2) we can estimate
(3.17) |
Since , by applying the law of cosines, (3.17) and (3.15) we obtain
(Using (3.17)) | |||
(Using (3.15)) |
and thus (3.16) holds. ∎
Remark 3.18.
The proof of Proposition 3.14 can be modified to show that there is a snowflake function (satisfying all the conditions (S1)–(S4)) such that for every there exists a metric space with cardinality so that embeds isometrically into -dimensional Euclidean space.
Indeed, suppose we are given and we have already defined on and that the slope of at is . Then we can define on an arbitrary long interval as , where . By the proof of Proposition 3.14, by taking large enough, there exist points in with the distance between any two of them between and so that they are an -snowflake of some metric space. We can define near so that it satisfies (S3) and if we take then (S4) is also satisfied. Therefore will have all the required properties.
References
- [Ass83] Patrice Assouad, Plongements lipschitziens dans , Bull. Soc. Math. France 111 (1983), no. 4, 429–448.
- [Ban55] Stefan Banach, Théorie des opérations linéaires, Chelsea Publishing Co., New York, 1955. MR 0071726
- [CK10] Jeff Cheeger and Bruce Kleiner, Differentiating maps into , and the geometry of BV functions, Ann. of Math. (2) 171 (2010), no. 2, 1347–1385.
- [DS13] Guy David and Marie Snipes, A non-probabilistic proof of the Assouad embedding theorem with bounds on the dimension, Anal. Geom. Metr. Spaces 1 (2013), 36–41. MR 3108866
- [EF83] P. Erdős and Z. Füredi, The greatest angle among points in the -dimensional Euclidean space, Combinatorial mathematics (Marseille-Luminy, 1981), North-Holland Math. Stud., vol. 75, North-Holland, Amsterdam, 1983, pp. 275–283. MR 841305
- [Fré10] Maurice Fréchet, Les dimensions d’un ensemble abstrait, Math. Ann. 68 (1910), no. 2, 145–168. MR 1511557
- [Hei01] Juha Heinonen, Lectures on analysis on metric spaces, Universitext, Springer-Verlag, New York, 2001.
- [Hei03] by same author, Geometric embeddings of metric spaces, Report. University of Jyväskylä Department of Mathematics and Statistics, vol. 90, University of Jyväskylä, Jyväskylä, 2003. MR 2014506
- [KS11] Antti Käenmäki and Ville Suomala, Nonsymmetric conical upper density and -porosity, Trans. Amer. Math. Soc. 363 (2011), no. 3, 1183–1195. MR 2737262
- [Kur35] Casimir Kuratowski, Quelques problèmes concernant les espaces métriques nonséparables, Fund. Math. 25 (1935), 534–545.
- [Laa02] Tomi J. Laakso, Plane with -weighted metric not bi-Lipschitz embeddable to , Bull. London Math. Soc. 34 (2002), no. 6, 667–676.
- [LN14] Vincent Lafforgue and Assaf Naor, A doubling subset of for that is inherently infinite dimensional, Geom. Dedicata 172 (2014), 387–398. MR 3253787
- [LP01] Urs Lang and Conrad Plaut, Bilipschitz embeddings of metric spaces into space forms, Geom. Dedicata 87 (2001), no. 1-3, 285–307.
- [Luo96] Kerkko Luosto, Ultrametric spaces bi-Lipschitz embeddable in , Fund. Math. 150 (1996), no. 1, 25–42. MR 1387955
- [NN12] Assaf Naor and Ofer Neiman, Assouad’s theorem with dimension independent of the snowflaking, Rev. Mat. Iberoam. 28 (2012), no. 4, 1123–1142. MR 2990137
- [Sem96a] Stephen Semmes, Good metric spaces without good parameterizations, Rev. Mat. Iberoamericana 12 (1996), no. 1, 187–275.
- [Sem96b] by same author, On the nonexistence of bi-Lipschitz parameterizations and geometric problems about -weights, Rev. Mat. Iberoamericana 12 (1996), no. 2, 337–410.
- [Sem99] by same author, Bilipschitz embeddings of metric spaces into Euclidean spaces, Publ. Mat. 43 (1999), no. 2, 571–653. MR 1744622
- [Seo11] Jeehyeon Seo, A characterization of bi-Lipschitz embeddable metric spaces in terms of local bi-Lipschitz embeddability, Math. Res. Lett. 18 (2011), no. 6, 1179–1202. MR 2915474
- [Ury27] Paul Urysohn, Sur un espace métrique universel, Bull. Sc. Math. 2e série 51 (1927), 43–64.