This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Isometric embeddings of snowflakes into finite-dimensional Banach spaces

Enrico Le Donne Tapio Rajala  and  Erik Walsberg University of Jyvaskyla
Department of Mathematics and Statistics
P.O. Box 35 (MaD)
FI-40014 University of Jyvaskyla
Finland
enrico.e.ledonne@jyu.fi tapio.m.rajala@jyu.fi Department of Mathematics
University of Illinois at Urbana-Champaign
1409 West Green Street
Urbana
Il
61802
USA
erikw@illinois.edu
(Date: September 12, 2016)
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 XX isometrically embeds into some finite-dimensional normed space if and only if XX is finite. In the case of power functions we give a uniform bound on the cardinality of XX depending only on the power exponent and the dimension of the vector space.

2010 Mathematics Subject Classification:
30L05; 46B85; 54C25; 54E40; 28A80.
E.L.D. and T.R. acknowledge the support of the Academy of Finland, projects no. 288501 and 274372. E.W acknowledges the support of the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013) / ERC Grant agreement no. 291111/ MODAG

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 (X,d)(X,d) is a doubling metric space and α(0,1)\alpha\in(0,1) then the metric space (X,dα)(X,d^{\alpha}) admits a bi-Lipschitz embedding into some Euclidean space. If dEd_{E} is the Euclidean distance and α=log2/log3\alpha=\log 2/\log 3 then the metric space ([0,1],dEα)([0,1],d_{E}^{\alpha}) is bi-Lipschitz equivalent to the von Koch snowflake curve, so (X,dα)(X,d^{\alpha}) is said to be the α\alpha-snowflake of (X,d)(X,d).

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 α\alpha-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 ([0,1],dE1/2)([0,1],d_{E}^{1/2}) does not, see [Hei03, Remark 3.16(b)].

The main aim of this paper is to show that if some α\alpha-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 nn\in\mathbb{N} and α(0,1)\alpha\in(0,1) there is an NN\in\mathbb{N} such that if a metric space (X,d)(X,d) has cardinality at least NN then (X,dα)(X,d^{\alpha}) does not admit an isometric embedding into any nn-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 h:h:\mathbb{R}_{\geq}\to\mathbb{R}_{\geq} is a snowflaking function if the following hold:

  1. (S1)

    h(0)=0h(0)=0.

  2. (S2)

    hh is concave.

  3. (S3)

    h(t)t\frac{h(t)}{t}\to\infty, as t0{t\to 0}.

  4. (S4)

    h(t)t0\frac{h(t)}{t}\to 0, as t{t\to\infty}.

Let hh be a snowflaking function. Then function hh is weakly increasing and, if dd is a metric on a set XX then hdh\circ d is also a metric on XX. Given a snowflaking function hh and a metric space (X,d)(X,d) we say that the metric space (X,hd)(X,h\circ d) is the hh-snowflake of (X,d)(X,d). If h(t)=tαh(t)=t^{\alpha} for some α(0,1)\alpha\in(0,1) then (X,hd)(X,h\circ d) is the α\alpha-snowflake of (X,d)(X,d). In Section 3.2 we prove the following.

Theorem 1.2.

Let hh be a snowflaking function and (X,d)(X,d) a metric space. If the hh-snowflake of (X,d)(X,d) admits an isometric embedding into some finite-dimensional Banach space then XX 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 (X,hd)(X,h\circ d) 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 α\alpha-snowflake of n\mathbb{R}^{n}, α(0,1)\alpha\in(0,1), isometrically embeds into the Hilbert space 2\ell^{2} of square summable sequences, see [Hei03, Remark 3.16(d)]. For any nn\in\mathbb{N}, there is a metric space of cardinality nn such that for any α(0,1)\alpha\in(0,1) its α\alpha-snowflake can be isometrically embedded into n1\mathbb{R}^{n-1} (just take the vertices of the standard simplex). There is a 4-point metric space which has an α\alpha-snowflake which cannot be isometrically embedded into 4\mathbb{R}^{4}, 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 α\alpha-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 (X,d)(X,d) is infinite. For any snowflaking function hh and nn\in\mathbb{N} there is a δ>0\delta>0 such that (X,hd)(X,h\circ d) does not admit a (1+δ)(1+\delta)-bilipschitz embedding into any normed linear space of dimension nn.

Theorem 1.1 shows that there is a bound (depending on α\alpha and nn) on the cardinality of a metric space whose α\alpha-snowflake admits an isometric embedding into n\mathbb{R}^{n}. The next easy result says that any finite metric space can be isometrically embedded in some Euclidean space after some α\alpha-snowflaking.

Proposition 2.2.

If (X,d)(X,d) is a finite metric space of cardinality nn, then some α\alpha-snowflake of (X,d)(X,d) admits an isometric embedding into the Euclidean space n\mathbb{R}^{n}.

Combining Theorem 1.1 and Proposition 2.2 we have:

Corollary 2.3.

Given a metric space (X,d)(X,d), there is an α\alpha-snowflake of (X,d)(X,d) that isometrically embedds into some finite-dimensional normed space if and only if XX is finite.

Proof of Proposition 2.2.

We show that (X,dα)(X,d^{\alpha}) admits an isometric embedding into n\mathbb{R}^{n} when α\alpha is sufficiently close to 11. In fact we show that there is an ϵ>0\epsilon>0 such that if dd^{\prime} is any metric on XX satisfying 1ϵ<d(x,y)<1+ϵ1-\epsilon<d^{\prime}(x,y)<1+\epsilon for all distinct x,yXx,y\in X then (X,d)(X,d^{\prime}) admits an isometric embedding into n\mathbb{R}^{n}. The proposition follows as d(x,y)α1d(x,y)^{\alpha}\to 1 as α0\alpha\to 0 for any distinct x,yXx,y\in X.

Consider the points pj:=12ejp_{j}:=\tfrac{1}{\sqrt{2}}e_{j} where e1,,ene_{1},\ldots,e_{n} is the canonical basis of n\mathbb{R}^{n}. Thus d(qi,qj)=1d(q_{i},q_{j})=1 for all distinct 1i,jn1\leq i,j\leq n. The proposition is thus a direct consequence of the following claim:

Claim A. There is an ϵ>0\epsilon>0 such that if ρij\rho_{ij}\in\mathbb{R}, with 1i<jn1\leq i<j\leq n are such that |ρij1|<ϵ|\rho_{ij}-1|<\epsilon, then there exist q1,,qnnq_{1},\ldots,q_{n}\in\mathbb{R}^{n} such that qiqj=ρij\left\|q_{i}-q_{j}\right\|=\rho_{ij} for all 1i<jn1\leq i<j\leq n.

To show the claim, consider the vector subspace UU of n×n\mathbb{R}^{n\times n} defined by the upper-triangular matrices, i.e., the elements of UU are elements of the form (aij)ij(a_{ij})_{ij} with 1i<jn1\leq i<j\leq n and aija_{ij}\in\mathbb{R}. Let F:n×nUF:\mathbb{R}^{n\times n}\to U be given by

F(q1,,qn):=(qiqj2)ij for q1,,qnn.F(q_{1},\ldots,q_{n}):=(\left\|q_{i}-q_{j}\right\|^{2})_{ij}\quad\text{ for }\,q_{1},\ldots,q_{n}\in\mathbb{R}^{n}.

Denote by 𝐩{\bf p} the vector of vectors 𝐩:=(12e1,,12en){\bf p}:=(\tfrac{1}{\sqrt{2}}e_{1},\ldots,\tfrac{1}{\sqrt{2}}e_{n}). Notice that F(𝐩)=(1)ijF({\bf p})=(1)_{ij}.

Claim A holds if and only if F(𝐩)F(\bf p) lies in the interior of the image of FF. Thus it suffices to show that the differential (dF)𝐩(dF)_{\bf p} has maximal rank and apply the inverse function theorem.

Denote by qlkq_{l}^{k} the kk-th component of an nn-vector qlq_{l}. Notice that elk=δlke_{l}^{k}=\delta_{l}^{k}, where δlk\delta_{l}^{k} is the Kronecker symbol. Set Elk=(δilδjk)ijE_{lk}=(\delta_{i}^{l}\delta_{j}^{k})_{ij} Notice that {Elk,1i<jn}\{E_{lk},1\leq i<j\leq n\} span the set UU. Let us show that any ElkE_{lk} is in the image of (dF)𝐩(dF)_{\bf p}. Fix ll and kk so that 1l<kn1\leq l<k\leq n. By direct calculation,

F(𝐪)qlk=(2qiqj,δilekδjlek)ij=(2(qikqjk)(δilδjl))ij.\displaystyle\dfrac{\partial F({\bf q})}{\partial q_{l}^{k}}=\left(2\langle q_{i}-q_{j},\delta_{i}^{l}e_{k}-\delta_{j}^{l}e_{k}\rangle\right)_{ij}=\left(2(q_{i}^{k}-q_{j}^{k})(\delta_{i}^{l}-\delta_{j}^{l})\right)_{ij}.

Evaluating at 𝐩{\bf p}, we get

F(𝐪)qlk|𝐪=𝐩=(212(δikδjk)(δilδjl))ij.\left.\dfrac{\partial F({\bf q})}{\partial q_{l}^{k}}\right|_{{\bf q}={\bf p}}=\left(2\frac{1}{\sqrt{2}}(\delta_{i}^{k}-\delta_{j}^{k})(\delta_{i}^{l}-\delta_{j}^{l})\right)_{ij}.

Since lkl\neq k, we have δikδil0\delta_{i}^{k}\delta_{i}^{l}\equiv 0 and δjkδjl0\delta_{j}^{k}\delta_{j}^{l}\equiv 0. Since l<kl<k, and i<ji<j, we have δikδjl0\delta_{i}^{k}\delta_{j}^{l}\equiv 0. Hence

F(𝐪)qlk|𝐪=𝐩=(2(δjk)δil)ij=2Elk.\left.\dfrac{\partial F({\bf q})}{\partial q_{l}^{k}}\right|_{{\bf q}={\bf p}}=\left(\sqrt{2}(-\delta_{j}^{k})\delta_{i}^{l}\right)_{ij}=-\sqrt{2}E_{lk}.

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 n\mathbb{R}^{n} 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 hh is a degenerate snowflaking function then there exists an infinite metric space whose hh-snowflake isometrically embeds into 22-dimensional Euclidean space.

3.1. Geometric lemmas

Throughout this section VV is an nn-dimensional normed linear space with norm \|\cdot\|. For the remainder of this section we fix an inner product ,\langle\cdot,\cdot\rangle on VV for which John ellipsoid property holds. Namely, we have

(3.1) BBVnB,B\subset B_{V}\subset\sqrt{n}B,

where BVB_{V} is the \|\cdot\|-unit ball and BB is the the unit ball in the l2l_{2} metric associated to the inner product ,\langle\cdot,\cdot\rangle. For u,vVu,v\in V we denote the length given by the inner product by uv:=uv,uvuv:=\sqrt{\langle u-v,u-v\rangle}.

Given two vectors u,vVu,v\in V let (u,v)\angle(u,v) be their angle with respect to the above inner product ,\langle\cdot,\cdot\rangle:

(u,v):=arccosu,vu,uv,v.\angle(u,v):=\arccos\dfrac{\langle u,v\rangle}{\langle u,u\rangle\langle v,v\rangle}.

Given three points x,y,zVx,y,z\in V we set y(x,z):=(xy,zy)\angle_{y}(x,z):=\angle(x-y,z-y).

Let x,y,zVx,y,z\in V and let z:=x+zx,yx(yx)z^{\prime}:=x+\langle z-x,y-x\rangle(y-x) be the projection of zz onto the line through xx and yy.

Lemma 3.2.

There exist constants ε,C(0,)\varepsilon,C\in(0,\infty) depending only on nn such that if x(y,z)<ε\angle_{x}(y,z)<\varepsilon then

|xzxz|Cxzx(y,z).\big{|}\|x-z\|-\|x-z^{\prime}\|\big{|}\leq C\|x-z^{\prime}\|\angle_{x}(y,z).
Proof.

By the John Ellipsoid Theorem, (3.1), there is an angle θ\theta and a height hh only depending on nn such that for any point pp on the sphere with respect to \left\|\cdot\right\|. The cone C(p,θ,,0)C^{-}(p,\theta,\ell,0) at pp with opening θ\theta and height \ell in the direction of 0 is inside the ball, whereas the cone C+(p,θ,,p)C^{+}(p,\theta,\ell,p) at pp with opening θ\theta and height \ell in the direction of pp is outside the ball.

The inequality that we need to prove is translation and dilation invariant. So we assume that x=0x=0 and that zz^{\prime} lies on the unit sphere. Let β:=x(y,z)\beta:=\angle_{x}(y,z). We shall require β<ε<π2\beta<\varepsilon<\frac{\pi}{2} for a small enough ϵ\epsilon. Let t,s,t1,t2,t3t,s,t_{1},t_{2},t_{3} be as in the picture.

xxβ\betat2t_{2}zz^{\prime}θ\thetat3t_{3}ttt1t_{1}zzss
Figure 1. The point ww is between tt and ss.

If ϵ\epsilon is small enough, then sinθcosβ2cosθsinβ>12sinθ.\sin\theta\cos\beta-2\cos\theta\sin\beta>{1\over 2}\sin\theta. Setting r:=tzr:=tz^{\prime} we have

t2y=t3yt3t2=rsinθsinβcosβ2rcosθ>rsinθ2sinβ.t_{2}y=t_{3}y-t_{3}t_{2}=r\dfrac{\sin\theta}{\sin\beta}\cos\beta-2r\cos\theta>\dfrac{r\sin\theta}{2\sin\beta}.

Also, if ϵ\epsilon is small enough, then the point tt is in C+(p,θ,,p)C^{+}(p,\theta,\ell,p) and so sC(p,θ,,0)s\in C^{-}(p,\theta,\ell,0). In particular, between ss and tt there is a point ww such that xw=xz\|x-w\|=\|x-z^{\prime}\|.

We conclude using, in order, the definition of ww, the fact that on a line any two norms are a multiple of each other, the triangle inequality, the properties of ss and tt, and the previous bound:

1\displaystyle 1 \displaystyle\leq xzxz=xzxw=xzxwxw+wzxw=1+wzxw1+tssy\displaystyle\dfrac{\|x-z\|}{\|x-z^{\prime}\|}=\dfrac{\|x-z\|}{\|x-w\|}=\dfrac{xz}{xw}\leq\dfrac{xw+wz}{xw}=1+\dfrac{wz}{xw}\leq 1+\dfrac{ts}{sy}
\displaystyle\leq 1+tt1+tzt2y1+2rcosθ+rr2sinθ/sinβ1+Cθsinβ,\displaystyle 1+\dfrac{tt_{1}+tz^{\prime}}{t_{2}y}\leq 1+\dfrac{2r\cos\theta+r}{\frac{r}{2}\sin\theta/\sin\beta}\leq 1+C_{\theta}\sin\beta,

where Cθ=(4cosθ+2)/sinθC_{\theta}=(4\cos\theta+2)/\sin\theta. ∎

The following bound is another easy consequence of the John Ellipsoid theorem.

Lemma 3.3.

There exists a constant K>0K>0 depending only on nn such that when x(y,z),y(x,z)(0,π4)\angle_{x}(y,z),\angle_{y}(x,z)\in(0,\frac{\pi}{4}) we have

(3.4) K1zyzxx(y,z)y(x,z)Kzyzx.K^{-1}\frac{\|z^{\prime}-y\|}{\|z^{\prime}-x\|}\leq\frac{\angle_{x}(y,z)}{\angle_{y}(x,z)}\leq K\frac{\|z^{\prime}-y\|}{\|z^{\prime}-x\|}.
Proof.

Set α=x(y,z)\alpha={\angle_{x}(y,z)}, and β:=y(x,z)\beta:={\angle_{y}(x,z)}. By (3.1)

αβ\displaystyle\frac{\alpha}{\beta} \displaystyle\leq tanα12tanβ=2zz/xzzz/yz=2yzxz2nzyzx.\displaystyle\frac{\tan\alpha}{{1\over 2}\tan\beta}=2\frac{zz^{\prime}/xz^{\prime}}{zz^{\prime}/yz^{\prime}}=2\frac{yz^{\prime}}{xz^{\prime}}\leq 2\frac{\sqrt{n}\|z^{\prime}-y\|}{\|z^{\prime}-x\|}.

By symmetry, we get the other inequality. ∎

Our arguments rely on the following Ramsey-theoretic result. Explicit bounds on the number of points that one can have in n\mathbb{R}^{n} 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 nn\in\mathbb{N} and 0<β<π0<\beta<\pi there is an NN\in\mathbb{N} such that if SnS\subseteq\mathbb{R}^{n} has cardinality at least NN then there are distinct x,y,zSx,y,z\in S such that β(xyz)π\beta\leq\angle(xyz)\leq\pi.

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.

Fix nn and α\alpha and let C,ϵ,KC,\epsilon,K be given by Lemma 3.2 and Lemma 3.3. Now take θ<min{ϵ,π4,22α3C(2K+2α)}\theta<\min\{\epsilon,\frac{\pi}{4},\frac{2-2^{\alpha}}{3C(2K+2^{\alpha})}\}.

We let NN\in\mathbb{N} be the constant in Lemma 3.5 with β=πθ\beta=\pi-\theta. We suppose towards a contradiction that (X,d)(X,d) has cardinality NN and that there is an isometric embedding ι\iota of (X,dα)(X,d^{\alpha}) into an nn-dimensional normed vector space (V,)(V,\|\cdot\|). By Lemma 3.5 there exist three isometrically embedded points x,y,zι(X)x,y,z\in\iota(X) such that πθ<z(x,y)π\pi-\theta<\angle_{z}(x,y)\leq\pi. We declare δz:=z(x,y).\delta_{z}:=\angle_{z}(x,y). We may assume that δzπ\delta_{z}\neq\pi. We also declare δx:=x(y,z)\delta_{x}:=\angle_{x}(y,z) and δy:=y(x,z)\delta_{y}:=\angle_{y}(x,z). Note that 0<δx,δy<θ0<\delta_{x},\delta_{y}<\theta.

Let znz^{\prime}\in\mathbb{R}^{n} be the orthogonal projection of zz on the line passing through xx and yy, i.e. z:=x+zx,yx(yx)z^{\prime}:=x+\langle z-x,y-x\rangle(y-x). Lemma 3.2 yields

(3.6) |xzxz|Cxzδx\left|\|x-z\|-\|x-z^{\prime}\|\right|\leq C\|x-z^{\prime}\|\delta_{x}

and

(3.7) |yzyz|Cyzδy.\left|\|y-z\|-\|y-z^{\prime}\|\right|\leq C\|y-z^{\prime}\|\delta_{y}.

We now estimate

(yz1α+zx1α)2α(\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}})^{2\alpha}

from above. Subadditivity of ttαt\mapsto t^{\alpha} yields:

(yz1α+zx1α)2α\displaystyle(\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}})^{2\alpha} =((yz1α+zx1α)2)α\displaystyle=((\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}})^{2})^{\alpha}
=(yz2α+zx2α+2yz1αzx1α)α\displaystyle=(\|y-z\|^{\frac{2}{\alpha}}+\|z-x\|^{\frac{2}{\alpha}}+2\|y-z\|^{\frac{1}{\alpha}}\|z-x\|^{\frac{1}{\alpha}})^{\alpha}
yz2+zx2+2αyzzx.\displaystyle\leq\|y-z\|^{2}+\|z-x\|^{2}+2^{\alpha}\|y-z\|\|z-x\|.

Estimating the obtained terms from above using (3.7) and (3.6) we have

(yz1α+zx1α)2α\displaystyle(\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}})^{2\alpha} yz2(1+Cδy)2+zx2(1+Cδx)2\displaystyle\leq\|y-z^{\prime}\|^{2}(1+C\delta_{y})^{2}+\|z^{\prime}-x\|^{2}(1+C\delta_{x})^{2}
+2αyzzx(1+Cδx)(1+Cδy).\displaystyle\quad\,+2^{\alpha}\|y-z^{\prime}\|\|z^{\prime}-x\|(1+C\delta_{x})(1+C\delta_{y}).

Now we use the fact that Cδx<1C\delta_{x}<1 and Lemma 3.3 to obtain

yz2(1+Cδy)2\displaystyle\|y-z^{\prime}\|^{2}(1+C\delta_{y})^{2} yz2(1+3Cδy)yz2(1+3CKδxzxzy)\displaystyle\leq\|y-z^{\prime}\|^{2}(1+3C\delta_{y})\leq\|y-z^{\prime}\|^{2}\left(1+3CK\delta_{x}\frac{\|z^{\prime}-x\|}{\|z^{\prime}-y\|}\right)
yz2+3CKθyzxz\displaystyle\leq\|y-z^{\prime}\|^{2}+3CK\theta\|y-z^{\prime}\|\|x-z^{\prime}\|

and similarly

xz2(1+Cδx)2xz2+3CKθyzxz.\|x-z^{\prime}\|^{2}(1+C\delta_{x})^{2}\leq\|x-z^{\prime}\|^{2}+3CK\theta\|y-z^{\prime}\|\|x-z^{\prime}\|.

Again by Cδx<1C\delta_{x}<1 we get (1+Cδx)(1+Cδy)<1+3Cθ(1+C\delta_{x})(1+C\delta_{y})<1+3C\theta. Collecting the estimates together and using the fact that 6CKθ+2α(1+3Cθ)<26CK\theta+2^{\alpha}(1+3C\theta)<2 we have:

(yz1α+zx1α)2α\displaystyle(\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}})^{2\alpha} yz2+3CKθyzxz\displaystyle\leq\|y-z^{\prime}\|^{2}+3CK\theta\|y-z^{\prime}\|\|x-z^{\prime}\|
+xz2+3CKθyzxz\displaystyle\quad\,+\|x-z^{\prime}\|^{2}+3CK\theta\|y-z^{\prime}\|\|x-z^{\prime}\|
+2α(1+3Cθ)yzzx\displaystyle\quad\,+2^{\alpha}(1+3C\theta)\|y-z^{\prime}\|\|z^{\prime}-x\|
=yz2+xz2\displaystyle=\|y-z^{\prime}\|^{2}+\|x-z^{\prime}\|^{2}
+(6CKθ+2α(1+3Cθ))yzzx\displaystyle\quad\,+(6CK\theta+2^{\alpha}(1+3C\theta))\|y-z^{\prime}\|\|z^{\prime}-x\|
<xz2+zy2+2xzzy\displaystyle<\|x-z^{\prime}\|^{2}+\|z^{\prime}-y\|^{2}+2\|x-z^{\prime}\|\|z^{\prime}-y\|
=(xz+zy)2=xy2.\displaystyle=(\|x-z^{\prime}\|+\|z^{\prime}-y\|)^{2}=\|x-y\|^{2}.

Therefore

d(ι1(x),ι1(y))\displaystyle d(\iota^{-1}(x),\iota^{-1}(y)) =xy1α>yz1α+zx1α\displaystyle=\|x-y\|^{\frac{1}{\alpha}}>\|y-z\|^{\frac{1}{\alpha}}+\|z-x\|^{\frac{1}{\alpha}}
=d(ι1(z),ι1(y))+d(ι1(x),ι1(z)).\displaystyle=d(\iota^{-1}(z),\iota^{-1}(y))+d(\iota^{-1}(x),\iota^{-1}(z)).

This contradicts the triangle inequality in (X,d)(X,d). ∎

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 hh, but also on the first element of the sequence. Therefore no upper bound (depending on hh and nn) 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 XX is infinite and that there exists an isometric embedding ι:(X,hd)V\iota\colon(X,h\circ d)\to V where VV is an nn-dimensional normed vector space. We divide our proof into two cases. An infinite bounded subset of n\mathbb{R}^{n} is not discrete, so one of the following holds:

  1. (i)

    ι(X)\iota(X) is unbounded;

  2. (ii)

    ι(X)\iota(X) 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 ι(X)\iota(X) is unbounded
Observe that (S4) implies that the existence of a function T:T\colon\mathbb{R}_{\geq}\to\mathbb{R}_{\geq} such that for any t>0t>0 and ST(t)S\geq T(t) we have

(3.8) th(t)h(S)S12.\frac{t}{h(t)}\frac{h(S)}{S}\leq\frac{1}{2}.

Combining (3.8) with (S1) and (S2) we get

(3.9) h(S+t)h(S)+th(S)S=h(S)+th(t)h(S)Sh(t)h(S)+12h(t).h(S+t)\leq h(S)+t\frac{h(S)}{S}=h(S)+\frac{t}{h(t)}\frac{h(S)}{S}h(t)\leq h(S)+\frac{1}{2}h(t).

Now fix x0,x1Xx_{0},x_{1}\in X, x0x1x_{0}\neq x_{1}. Since (X,d)(X,d) is unbounded, there exists a point x2Xx_{2}\in X with ι(x2)(ι(x0),ι(x1))π/4\angle_{\iota(x_{2})}(\iota(x_{0}),\iota(x_{1}))\leq\pi/4 and d(x2,xi)>T(d(x0,x1))d(x_{2},x_{i})>T(d(x_{0},x_{1})) for i=0,1i=0,1. We continue inductively. Suppose (xi)i=0N1X(x_{i})_{i=0}^{N-1}\subset X have been chosen. Now we select xNXx_{N}\in X satisfying

(3.10) ι(xN)(ι(xi),ι(xj))π/4 and d(xi,xN)>T(d(xi,xj)),i,j<N.\angle_{\iota(x_{N})}(\iota(x_{i}),\iota(x_{j}))\leq\pi/4\text{ and }d(x_{i},x_{N})>T(d(x_{i},x_{j})),\quad\forall i,j<N.

Let ϵ,C,K\epsilon,C,K be the constants from Lemma 3.2 and Lemma 3.3. Set

(3.11) δ=min{ϵ,π/4,12C(1+K)}.\delta=\min\{\epsilon,\pi/4,\frac{1}{2C(1+K)}\}.

By Lemma 3.5 there exist x,y,zx,y,z in {ι(x)}\{\iota(x_{\ell})\}_{\ell\in\mathbb{N}} such that z(x,y)>πδ\angle_{z}(x,y)>\pi-\delta. By the condition (3.10), there exist i,j,ki,j,k\in\mathbb{N} with k>max{i,j}k>\max\{i,j\} such that x=ι(xj)x=\iota(x_{j}), y=ι(xk)y=\iota(x_{k}), z=ι(xi)z=\iota(x_{i}). Let zz^{\prime} be the orthogonal projection of zz to the line passing through xx and yy. On the one hand, we have that

h(d(xj,xk))\displaystyle h(d(x_{j},x_{k})) =xy\displaystyle=\|x-y\|
=xz+zy\displaystyle=\|x-z^{\prime}\|+\|z^{\prime}-y\|
(Using Lemma 3.2)\displaystyle(\text{Using Lemma~\ref{lma:closetoEuclidean}})\quad xzCxzx(z,y)\displaystyle\geq\|x-z\|-C\|x-z^{\prime}\|\angle_{x}(z,y)
+zyCyzy(z,x)\displaystyle\hskip 56.9055pt+\|z-y\|-C\|y-z^{\prime}\|\angle_{y}(z,x)
(Using Lemma 3.3)\displaystyle(\text{Using Lemma~\ref{lemma234124}})\quad xzCxzx(z,y)\displaystyle\geq\|x-z\|-C\|x-z^{\prime}\|\angle_{x}(z,y)
+zyCKxzx(z,y)\displaystyle\hskip 56.9055pt+\|z-y\|-CK\|x-z^{\prime}\|\angle_{x}(z,y)
=xz+zyC(1+K)xzx(z,y)\displaystyle=\|x-z\|+\|z-y\|-C(1+K)\|x-z^{\prime}\|\angle_{x}(z,y)
xz+zyC(1+K)xzx(z,y)\displaystyle\geq\|x-z\|+\|z-y\|-C(1+K)\|x-z\|\angle_{x}(z,y)
=h(d(xi,xk))+(1C(1+K)x(z,y))h(d(xi,xj))\displaystyle=h(d(x_{i},x_{k}))+(1-C(1+K)\angle_{x}(z,y))h(d(x_{i},x_{j}))
(Using that x(z,y)<δ)\displaystyle(\text{Using that }\angle_{x}(z,y)<\delta)\quad >h(d(xi,xk))+12h(d(xi,xj)).\displaystyle>h(d(x_{i},x_{k}))+\frac{1}{2}h(d(x_{i},x_{j})).

On the other hand, first notice that since hh is concave and positive, then hh has to be weakly increasing. Secondly, by the definition of the function TT we then have that for any three i,j,ki,j,k\in\mathbb{N} with k>max{i,j}k>\max\{i,j\}

(3.12) h(d(xj,xk))h(d(xi,xk)+d(xi,xj))h(d(xi,xk))+12h(d(xi,xj)).h(d(x_{j},x_{k}))\leq h(d(x_{i},x_{k})+d(x_{i},x_{j}))\leq h(d(x_{i},x_{k}))+\frac{1}{2}h(d(x_{i},x_{j})).

Therefore we have a contradiction.

Case (ii): Suppose ι(X)\iota(X) is not discrete
This time we observe that (S3) implies the existence of a function T~:\tilde{T}\colon\mathbb{R}_{\geq}\to\mathbb{R}_{\geq} such that T~(r)r\tilde{T}(r)\leq r for all rr and for any S>0S>0 and 0<tT~(S)0<t\leq\tilde{T}(S) we have (3.8), and hence (3.9), using (S1) and (S2).

Let yy be an accumulation point of XX. First we select x0X{y}x_{0}\in X\setminus\{y\}. Next we take a radius r0>0r_{0}>0 so that for all y1,y2B(y,r0)y_{1},y_{2}\in B(y,r_{0}) we have both ι(x0)(ι(y1),ι(y2))π/4\angle_{\iota(x_{0})}(\iota(y_{1}),\iota(y_{2}))\leq\pi/4 and d(y1,y2)<T~(d(x0,yi))d(y_{1},y_{2})<\tilde{T}(d(x_{0},y_{i})) for i=0,1i=0,1. Now we select a point x1B(y,r1){y}x_{1}\in B(y,r_{1})\setminus\{y\}. We continue inductively. Suppose (xi)i=0N1X(x_{i})_{i=0}^{N-1}\subset X have been chosen. Now we take a radius rN1<rN2r_{N-1}<r_{N-2} such that for all y1,y2B(y,r0)y_{1},y_{2}\in B(y,r_{0}) we have

(3.13) ι(xi)(ι(y1),ι(y2))π/4 and d(y1,y2)<T~(d(xi,yj)),i<N,j=1,2.\angle_{\iota(x_{i})}(\iota(y_{1}),\iota(y_{2}))\leq\pi/4\text{ and }d(y_{1},y_{2})<\tilde{T}(d(x_{i},y_{j})),\quad\forall i<N,j=1,2.

Then we select a point xNB(y,rN1){y}x_{N}\in B(y,r_{N-1})\setminus\{y\}.

With the points {xi}\{x_{i}\} we now arrive at a contradiction with the same argument as in the case (i). Let δ\delta be as in (3.11). Again by Lemma 3.5 there exist x,y,zx,y,z in {ι(x)}\{\iota(x_{\ell})\}_{\ell\in\mathbb{N}} such that z(x,y)>πδ\angle_{z}(x,y)>\pi-\delta, but this time by the condition (3.13), there exist i,j,ki,j,k\in\mathbb{N} with k<min{i,j}k<\min\{i,j\} such that x=ι(xj)x=\iota(x_{j}), y=ι(xk)y=\iota(x_{k}), z=ι(xi)z=\iota(x_{i}). Now we continue verbatim the proof in the case (i). ∎

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 h:h\colon\mathbb{R}_{\geq}\to\mathbb{R}_{\geq} satisfies (S1) and (S2) but fails to satisfy (S3) or (S4). Then there is an infinite metric space (X,d)(X,d) such that (X,hd)(X,h\circ d) admits an isometric embeddeding into 22-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 {xi}i=0\{x_{i}\}_{i=0}^{\infty} in 2\mathbb{R}^{2} such that (X,d):=({xi}i=1,h1dE)(X,d):=(\{x_{i}\}_{i=1}^{\infty},h^{-1}\circ d_{E}) is a metric space. For this purpose we fix a sequence {αi}i=1\{\alpha_{i}\}_{i=1}^{\infty} of positive angles such that i=1αi<π2\sum_{i=1}^{\infty}\alpha_{i}<\frac{\pi}{2}. Depending on the sequence {αi}i=1\{\alpha_{i}\}_{i=1}^{\infty} and the function hh we construct an increasing sequence {ti}i=1\{t_{i}\}_{i=1}^{\infty} of positive real numbers which determine the Euclidean distance between xi1x_{i-1} and xix_{i}. For notational convenience we let c(t)=h(t)tc(t)=\frac{h(t)}{t}. Notice that by assumption c(t)c>0c(t)\searrow c>0 as tt\to\infty. This allows us to select for every ii\in\mathbb{N} a real number ti>0t_{i}>0 such that for all s,ttis,t\geq t_{i} we have

(3.15) c(s)(c(s+t)c)+c(t)(c(s+t)c)2(c(s)c(t)cos(παi)+c(s+t)2).c(s)(c(s+t)-c)+c(t)(c(s+t)-c)\leq 2(c(s)c(t)\cos(\pi-\alpha_{i})+c(s+t)^{2}).

Now, using the sequences {αi}i=1\{\alpha_{i}\}_{i=1}^{\infty} and {ti}i=1\{t_{i}\}_{i=1}^{\infty} we define the sequence {xi}i\{x_{i}\}_{i\in\mathbb{N}} as follows. We set x0:=(0,0)x_{0}:=(0,0), x1:=(t1,0)x_{1}:=(t_{1},0), and inductively for n2n\geq 2 declare

xn:=xn1+(tnsin(j=1n1aj),tncos(j=1n1aj)).x_{n}:=x_{n-1}+\left(t_{n}\sin(\sum_{j=1}^{n-1}a_{j}),t_{n}\cos(\sum_{j=1}^{n-1}a_{j})\right).

In order to see that (X,d)(X,d) is a metric space we need to check that the triangle inequality holds. For this purpose let 0i<j<k0\leq i<j<k be three integers. Let dEd_{E} be the Euclidean metric. The only nontrivial inequality that we have to verify is

h1(dE(xi,xk))h1(dE(xj,xk))+h1(dE(xj,xk)).h^{-1}(d_{E}(x_{i},x_{k}))\leq h^{-1}(d_{E}(x_{j},x_{k}))+h^{-1}(d_{E}(x_{j},x_{k})).

Denoting s:=h1(dE(xi,xj))s:=h^{-1}(d_{E}(x_{i},x_{j})) and t:=h1(dE(xj,xk))t:=h^{-1}(d_{E}(x_{j},x_{k})) the above inequality is equivalent to

(3.16) dE(xi,xk)h(s+t).d_{E}(x_{i},x_{k})\leq h(s+t).

By (S1) and (S2) we can estimate

(3.17) c(t)=h(t)th(t+s)cst=c(t+s)+(c(t+s)c)st.c(t)=\frac{h(t)}{t}\leq\frac{h(t+s)-cs}{t}=c(t+s)+(c(t+s)-c)\frac{s}{t}.

Since s,ttjs,t\geq t_{j}, by applying the law of cosines, (3.17) and (3.15) we obtain

dE(xi,xk)2h(s+t)2\displaystyle d_{E}(x_{i},x_{k})^{2}-h(s+t)^{2} =dE(xi,xj)2+dE(xj,xk)2\displaystyle=d_{E}(x_{i},x_{j})^{2}+d_{E}(x_{j},x_{k})^{2}
2dE(xi,xj)dE(xj,xk)cos(xj(xi,xk))h(s+t)2\displaystyle\qquad-2d_{E}(x_{i},x_{j})d_{E}(x_{j},x_{k})\cos(\angle_{x_{j}}(x_{i},x_{k}))-h(s+t)^{2}
h(s)2+h(t)22h(s)h(t)cos(παj)h(s+t)2\displaystyle\leq h(s)^{2}+h(t)^{2}-2h(s)h(t)\cos(\pi-\alpha_{j})-h(s+t)^{2}
=s2c(s)2+t2c(t)22stc(s)c(t)cos(παj)(s+t)2c(s+t)2\displaystyle=s^{2}c(s)^{2}+t^{2}c(t)^{2}-2stc(s)c(t)\cos(\pi-\alpha_{j})-(s+t)^{2}c(s+t)^{2}
=s2(c(s)2c(s+t)2)+t2(c(t)2c(s+t)2)\displaystyle=s^{2}(c(s)^{2}-c(s+t)^{2})+t^{2}(c(t)^{2}-c(s+t)^{2})
2st(c(s)c(t)cos(παj)+c(s+t)2)\displaystyle\qquad-2st(c(s)c(t)\cos(\pi-\alpha_{j})+c(s+t)^{2})
(Using (3.17)) st(c(s)(c(s+t)c)+c(t)(c(s+t)c)\displaystyle\leq st\big{(}c(s)(c(s+t)-c)+c(t)(c(s+t)-c)
2(c(s)c(t)cos(παj)+c(s+t)2))\displaystyle\qquad-2(c(s)c(t)\cos(\pi-\alpha_{j})+c(s+t)^{2})\big{)}
(Using (3.15)) 0\displaystyle\leq 0

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 hh (satisfying all the conditions (S1)–(S4)) such that for every nn\in\mathbb{N} there exists a metric space (Xn,dn)(X_{n},d_{n}) with cardinality nn so that (Xn,hdn)(X_{n},h\circ d_{n}) embeds isometrically into 22-dimensional Euclidean space.

Indeed, suppose we are given nn\in\mathbb{N} and we have already defined hh on [0,Tn1][0,T_{n-1}] and that the slope of hh at Tn1T_{n-1} is cn1>0c_{n-1}>0. Then we can define hh on an arbitrary long interval [Tn1,Tn][T_{n-1},T_{n}] as h(t)=H(Tn1)+cnth(t)=H(T_{n-1})+c_{n}t, where 0<cn<cn10<c_{n}<c_{n-1}. By the proof of Proposition 3.14, by taking SS large enough, there exist nn points in 2\mathbb{R}^{2} with the distance between any two of them between h(T)h(T) and h(S)h(S) so that they are an hh-snowflake of some metric space. We can define hh near 0 so that it satisfies (S3) and if we take cn0c_{n}\searrow 0 then (S4) is also satisfied. Therefore hh will have all the required properties.

References

  • [Ass83] Patrice Assouad, Plongements lipschitziens dans 𝐑n{\bf R}^{n}, 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 L1L^{1}, 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 nn points in the dd-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 kk-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 AA_{\infty}-weighted metric not bi-Lipschitz embeddable to N{\mathbb{R}}^{N}, Bull. London Math. Soc. 34 (2002), no. 6, 667–676.
  • [LN14] Vincent Lafforgue and Assaf Naor, A doubling subset of LpL_{p} for p>2p>2 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 𝐑n{\bf R}^{n}, 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 AA_{\infty}-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.