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

This paper created as part of a Ph.D. thesis in mathematics at Texas A&M Univerity under the supervision of Dr. Florent Baudier. R. Malthaner was partially supported by National Science Foundation Grant Number DMS-2055604.

On Matoušek-like Embedding Obstructions of Countably Branching Graphs

Ryan Malthaner
Abstract.

In this paper we present new proofs of the non-embeddability of countably branching trees into Banach spaces satisfying property (βp)(\beta_{p}) and of countably branching diamonds into Banach spaces which are pp-AMUC for p>1p>1. These proofs are entirely metric in nature and are inspired by previous work of Jiří Matoušek. In addition, using this metric method, we succeed in extending these results to metric spaces satisfying certain curvature-like inequalities. Finally, we extend an embedding result of Tessera to give lower bounds on the compression for a class of Lipschitz embeddings of the countably branching trees into Banach spaces containing p\ell_{p}-asymptotic models for p1p\geq 1.

Key words and phrases:
Rolewicz’s property (β\beta), coarse embedding, non-embeddability, compression rate, trees, diamonds, umbels, metric spaces, distortion, bi-Lipschitz embeddings
1991 Mathematics Subject Classification:
05C63, 46B06, 46B20, 46B85, 51F30.

1. Introduction

In the world of Banach spaces, it is meaningful to pursue methods of differentiating the zoo of spaces available at a researcher’s fingertips. Given a Banach space XX satisfying property PP, one reasonable direction to pursue in that regard is to look at which spaces or families of spaces do not embed “well” into XX as a result of PP. Some of the simplest examples to consider are families of graphs {Gn}n=1\{G_{n}\}_{n=1}^{\infty} equipped with the shortest path metric where the diameter of the graphs (largest distance between two points) gets larger as nn increases. While normally we might consider linear maps, it is natural in this metric setting to instead consider families of bi-Lipschitz embeddings where if (Gn,dGn)(G_{n},d_{G_{n}}) is a connected graph with metric dGnd_{G_{n}} on the vertices and fn:GnXf_{n}:G_{n}\to X, we have some K1K\geq 1 and λ>0\lambda>0 such that

λdGn(x,y)fn(x)fn(y)KλdGn(x,y)\lambda d_{G_{n}}(x,y)\leq\|f_{n}(x)-f_{n}(y)\|\leq K\lambda d_{G_{n}}(x,y)

for every x,yGnx,y\in G_{n}. (We can assume λ=1\lambda=1 by rescaling in the Banach space, but in a metric space, we may not necessarily be able to do that.) In this case, the quantitative measure of how “well” GnG_{n} embeds into XX under this map fnf_{n} is given by the smallest KK which satisfies the above inequality, called the distortion of fnf_{n}, and is denoted by dist(fn)\textnormal{dist}(f_{n}). This, however, only gives us information for one particular embedding. If we wish to show that these graphs do not embed well as a family, we must consider supncX(Gn)=supninf{dist(f)|f:GnX and f bi-Lipschitz}\sup_{n}c_{X}(G_{n})=\sup_{n}\inf\{\textnormal{dist}(f)\ |\ f:G_{n}\to X\textnormal{ and }f\textnormal{ bi-Lipschitz}\}. If this quantity is bounded, then we say that {Gn}n=1\{G_{n}\}_{n=1}^{\infty} embeds equi-bi-Lipschitzly, meaning there exist maps fn:GnXf_{n}:G_{n}\to X such that dist(fn)supncX(Gn)\textnormal{dist}(f_{n})\leq\sup_{n}c_{X}(G_{n}) for all nn\in\mathbb{N}.

Some results in this vein come from the work of Bourgain [Bou86] and Matoušek [Mat99] who each proved that the complete binary trees (Bh)h=1(B_{h})_{h=1}^{\infty} do not embed equi-bi-Lipschitzly into any Banach space whose norm satisfies the pp-uniform convexity inequality. In essence, their result states that if the unit ball of a Banach space is sufficiently round, then any bi-Lipschitz map f:BhXf:B_{h}\to X must incur a penalty of at least O(log2(h)1/p)O(\log_{2}(h)^{1/p}). Much later, in [JS09], it was shown that the binary diamond graphs {Dn2}n=1\{D_{n}^{2}\}_{n=1}^{\infty} and Laakso graphs {Ln2}n=1\{L_{n}^{2}\}_{n=1}^{\infty} also do not embed equi-bi-Lipschitzly into pp-uniformly convex Banach spaces.

Each of these proofs, however, differed greatly from the others. Bourgain’s proof relied heavily on the linear structure of XX, utilizing Pisier’s martingale inequality to accomplish his proof, while the Johnson and Schechtman argument utilized a self-improvement argument and the recursive nature of the diamond and Laakso graphs. (This self-improvement argument was also later utitilized in a new proof of the non-embeddability of the binary trees by Kloeckner in [Klo14].) Finally, Matoušek’s argument was based on coloring and partitioning, only using the linear structure of XX in a single lemma. This would seem to indicate that the linear structure is unnecessary, and indeed, this was shown in [MN13]. In that spirit, the primary contribution of this paper is showing that these techniques can be extended to countably branching trees, and more interestingly, countably branching diamonds.

In the first part of this paper, we will show that Matousek’s arguments can be adapted and generalized to study the non-embeddability of countably branching trees into metric spaces that satisfy the infrasup pp-umbel inequality.

In the second part of the paper, we will extend Matousek’s technique to diamonds by first showing that Banach spaces satisfying the asymptotic midpoint uniform convexity property do not contain countably branching diamonds. From here, we will show a diamond analogue for the infrasup pp-umbel inequality. We will then extend Matousek’s arguments to show that the countably branching diamonds do not embed equi-bi-Lipschitzly into a metric space satisfying this inequality for p1p\geq 1.

In the last third, we will also provide an embedding of countably branching trees into certain Banach spaces, weaker than bi-Lipschitz, generalizing a result of [BG21] and [Tes11].

2. Countably Branching Trees

In this section, we will prove that for any metric space (X,dX)(X,d_{X}) satisfying the infrasup pp-umbel inequality for p1p\geq 1, the countably branching trees of finite but arbitrary height do not equi-bi-Lipschitzly embed into XX. To begin, we present these definitions and the extension from the linear to the metric setting.

2.1. Definitions and Lemmas

2.1.1. Rolewicz’s Property (βp)(\beta_{p})

Originally, in Matoušek’s proof, the obstruction to the embedding came from the pp-uniform convexity of the Banach space. There are several different asymptotic inequalities that one could consider, however, in the case of countably branching trees, the (βp)(\beta_{p}) property seems to be the most natural. As a reminder, pp-uniform convexity of a Banach space for p2p\geq 2 and some c>0c>0 is usually defined as for every t>0t>0, there exists δ(t)tpc\delta(t)\geq\frac{t^{p}}{c} such that for every x,yBXx,y\in B_{X}, the unit ball of a Banach space XX, with xyt\|x-y\|\geq t, we have x+y21δ(t)\|\frac{x+y}{2}\|\leq 1-\delta(t). This version of uniform convexity heavily focuses on the midpoint between xx and yy, an inconvenient formulation for our purposes, as would its asymptotic generalization. An equivalent version is given in [BG21] in Lemma 10 which they called a “tripod” variant of uniform convexity. Here, we will use the convention pp-tripod uniform convexity.

Definition 2.1.1 (pp-Tripod Uniform Convexity).

A Banach space XX is pp-tripod uniformly convex for some p2p\geq 2 and c>0c>0 if for every t>0t>0 and x1,x2,zBXx_{1},x_{2},z\in B_{X} with x1x2t\|x_{1}-x_{2}\|\geq t, there exists i0{1,2}i_{0}\in\{1,2\} such that zxi02(1tpc)\|z-x_{i_{0}}\|\leq 2(1-\frac{t^{p}}{c}).

This version very naturally extends to property (βp)(\beta_{p}) given below.

Definition 2.1.2 (Property (βp)(\beta_{p}) or Rolewicz’s property).

A Banach space (X,)(X,\|\cdot\|) has Rolewicz’s property (β)(\beta) if for all t>0t>0 there exists β(t)>0\beta(t)>0 such that for all zBXz\in B_{X} and {xn}nBX\{x_{n}\}_{n\in\mathbb{N}}\subset B_{X} with infijxixjXt\inf_{i\neq j}||x_{i}-x_{j}||_{X}\geq t, there exists i0i_{0}\in\mathbb{N} such that

||zxi02||X1β(t).\Bigg{\lvert}\Bigg{\lvert}\frac{z-x_{i_{0}}}{2}\Bigg{\rvert}\Bigg{\rvert}_{X}\leq 1-\beta(t).

Moreover, XX is said to have property (βp)(\beta_{p}) for p>1p>1 and constant Cβ>0C_{\beta}>0 if β(t)tpCβ\beta(t)\geq\frac{t^{p}}{C_{\beta}}. [Kut91]

This is the first of the asymptotic inequalities that we will be utilizing in this paper. Its various properties and characterizations are summarized in [Dil+16]. For our purposes, however, we will be using this as a starting point for moving to the metric inequality, the infrasup pp-umbel inequality.

Definition 2.1.3 (Infrasup pp-Umbel Inequality).

An infinite metric space (X,d)(X,d) satisfies the infrasup pp-umbel inequality for some p>0p>0 if there exists CU>0C_{U}>0 such that for any infinite collection of points w,z,x1,x2,x3,w,z,x_{1},x_{2},x_{3},..., we have

12pinfnd(w,xn)p+1CUpinfijd(xi,xj)pmax{d(w,z)p,supnd(xn,z)p}.\frac{1}{2^{p}}\inf_{n\in\mathbb{N}}d(w,x_{n})^{p}+\frac{1}{C_{U}^{p}}\inf_{i\neq j\in\mathbb{N}}d(x_{i},x_{j})^{p}\leq\max\{d(w,z)^{p},\sup_{n\in\mathbb{N}}d(x_{n},z)^{p}\}.

Importantly, we show that this inequality is satisfied by any Banach space with Rolewicz’s property, an idea implicitly found in [BG21].

Lemma 2.1 (Rolewicz implies Umbel).

Let XX be a Banach space satisfying (βp)(\beta_{p}) for p>1p>1, then there exists CU>0C_{U}>0 such that XX satisfies the infrasup pp-umbel inequality.

Proof.

Let w,z,x1,x2,Xw,z,x_{1},x_{2},...\in X. Without loss of generality, by translation of the norm, we may assume that z=0z=0. Suppose now that supnxn=\sup_{n}\|x_{n}\|=\infty. If this is the case, then the infrasup pp-umbel inequality holds trivially. As a result, we assume that supnxn<\sup_{n}\|x_{n}\|<\infty.

By the invariance of the infrasup pp-umbel inequality under scalings of the metric, we may assume that w,x1,x2,BXw,x_{1},x_{2},...\in B_{X}, so we need to show that

12pinfnwxnp+1CUpinfijxixjp1.\frac{1}{2^{p}}\inf_{n\in\mathbb{N}}\|w-x_{n}\|^{p}+\frac{1}{C^{p}_{U}}\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|^{p}\leq 1.

There are now two cases to consider, infijxixjp>0\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|^{p}>0 and infijxixjp=0\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|^{p}=0. In the case where it is equal to zero, then we can observe that since w,xnBXw,x_{n}\in B_{X}, we trivially satisfy the inequality and have

infnwxnp2p2p2p=1.\inf_{n\in\mathbb{N}}\frac{\|w-x_{n}\|^{p}}{2^{p}}\leq\frac{2^{p}}{2^{p}}=1.

In the case where infijxixjp>0\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|^{p}>0, in the definition of the (βp)(\beta_{p}) property, we can let t=infijxixjt=\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|. In addition, we observe that since infnwxn2\inf_{n\in\mathbb{N}}\|w-x_{n}\|\leq 2 and p>1p>1 we have by utilizing the (βp)(\beta_{p}) property

infnwxnp2pinfnwxn211Cβinfijxixjp.\inf_{n\in\mathbb{N}}\frac{\|w-x_{n}\|^{p}}{2^{p}}\leq\inf_{n\in\mathbb{N}}\frac{\|w-x_{n}\|}{2}\leq 1-\frac{1}{C_{\beta}}\inf_{i\neq j\in\mathbb{N}}\|x_{i}-x_{j}\|^{p}.

If we now let CU=Cβ1/pC_{U}=C_{\beta}^{1/p}, we see that XX satisfies the infrasup pp-umbel inequality for this choice of CUC_{U}. ∎

2.1.2. Tree Definitions

Here, we introduce the relevant tree definitions we will be using throughout this paper. We denote the complete, countably branching tree of height hh by ThωT_{h}^{\omega} where ω\omega is the first countable ordinal. Rather than working with the graph theoretic version of the tree, we will instead use the standard encoding of the tree given by ([]h,dT)([\mathbb{\mathbb{N}}]^{\leq h},d_{T}) where []h={A| 0|A|h}[\mathbb{\mathbb{N}}]^{\leq h}=\{A\subset\mathbb{N}\ |\ 0\leq|A|\leq h\}. By convention, we will write non-root elements of []h[\mathbb{\mathbb{N}}]^{\leq h} as n¯=(n1,n2,,nl)\bar{n}=(n_{1},n_{2},...,n_{l}) for 1lh1\leq l\leq h with n1<n2<<nln_{1}<n_{2}<...<n_{l}, and the root, rr, as the empty set. In order to reference individual pieces of a given element n¯=(n1,,nl)\bar{n}=(n_{1},...,n_{l}) of []h[\mathbb{\mathbb{N}}]^{\leq h}, we will use bracket notation, meaning n¯[i:j]=(ni,ni+1,,nj1,nj)\bar{n}[i:j]=(n_{i},n_{i+1},...,n_{j-1},n_{j}) for 1ijh1\leq i\leq j\leq h. In the case that i=ji=j, we write n¯[i]=(ni)\bar{n}[i]=(n_{i}), and in the case that j=0j=0, this is the root rr. We may also, by an abuse of notation, use concatenation to refer to pieces of a given element, i.e., if u¯=(u1,,us)\bar{u}=(u_{1},...,u_{s}), then we could construct m¯=(u¯,m1,,ml)\bar{m}=(\bar{u},m_{1},...,m_{l}) where s+lhs+l\leq h. We can equip the tree with a natural partial order by defining m¯<n¯\bar{m}<\bar{n} if n¯=(m¯,n1,,nl)\bar{n}=(\bar{m},n_{1},...,n_{l}) for some l1l\geq 1. Finally, the metric on []h[\mathbb{\mathbb{N}}]^{\leq h}, given by dTd_{T}, is defined in terms of the first common ancestor of m¯\bar{m} and n¯\bar{n}. If u¯=(u1,,us)\bar{u}=(u_{1},...,u_{s}) is the first common ancestor of m¯=(u¯,m1,,mk)\bar{m}=(\bar{u},m_{1},...,m_{k}) and n¯=(u¯,n1,,nl)\bar{n}=(\bar{u},n_{1},...,n_{l}), then dT(m¯,n¯)=l+kd_{T}(\bar{m},\bar{n})=l+k.

2.1.3. Umbel Lemma

Matoušek’s original argument showing that the complete binary trees {Bh}h=1\{B_{h}\}_{h=1}^{\infty} do not embed into pp-uniformly convex Banach spaces required three main lemmas: a fork lemma, a coloring lemma, and a path embedding lemma. The fork lemma uses the convexity of the target space to prove that fork-like structures in the target space must have fork tips that are close together. The coloring lemma provides a way for refining a coloring on pairs of vertices in a tree to something which can be more effectively used to extract a subtree with particular properties. Finally, the path lemma provides the means by which a contradiction and bound on the distortion of an arbitrary bi-Lipschitz map from the trees into our target space. In essence, it shows that embedded paths of sufficient length must have at least one subpath which is embedded arbitrarily well. Combining these three lemmas, Matoušek was able to prove the result in the binary trees case. In the extension of this theorem from the local theory to the asymptotic metric theory, we will derive three similar results for the countably branching trees, then connect them together to get the final result.

To this end, let us define, in the spirit of Matoušek, a vertical δ\delta-umbel.

Definition 2.1.4 (Vertical δ\delta-umbel).

Let (X,d)(X,d) be an infinite metric space. A vertical δ\delta-umbel is a subspace U={w,z,x1,x2,}U=\{w,z,x_{1},x_{2},...\} of XX such that

θd(w,z)\displaystyle\theta\leq d(w,z) (1+δ)θ\displaystyle\leq(1+\delta)\theta
θd(z,xi)\displaystyle\theta\leq d(z,x_{i}) (1+δ)θ\displaystyle\leq(1+\delta)\theta
2θd(w,xi)\displaystyle 2\theta\leq d(w,x_{i}) 2(1+δ)θ\displaystyle\leq 2(1+\delta)\theta

for some θ>0\theta>0, or, in other words, for every ii\in\mathbb{N} we have that {w,z,xi}\{w,z,x_{i}\} is (1+δ)(1+\delta)-isomorphic (in the sense of distortion) to the metric space P2={0,1,2}P_{2}=\{0,1,2\} with the absolute value metric on it. (See Fig. 1)

Refer to caption
Figure 1. Umbel Graph

From here, we can now present the vertical δ\delta-umbel lemma, demonstrating that in a metric space satisfying the infrasup pp-umbel inequality, any vertical δ\delta-umbel must have at least two tips which are close together.

Lemma 2.2 (Infrasup Umbel Inequality Lemma).

Let (X,d)(X,d) be a metric space satisfying the infrasup pp-umbel inequality for some CU1C_{U}\geq 1 and p1p\geq 1 and let U={w,z,x1,x2,}U=\{w,z,x_{1},x_{2},...\} be a vertical δ\delta-umbel in XX with constant θ>0\theta>0 and δ(0,1)\delta\in(0,1). Then,

infijd(xi,xj)6CUθδ1/p.\inf_{i\neq j\in\mathbb{N}}d(x_{i},x_{j})\leq 6C_{U}\theta\delta^{1/p}.
Proof.

Since UU is a vertical δ\delta-umbel with θ>0\theta>0, we have that

θ\displaystyle\theta d(w,z)(1+δ)θ\displaystyle\leq d(w,z)\leq(1+\delta)\theta
θ\displaystyle\theta d(z,xi)(1+δ)θ\displaystyle\leq d(z,x_{i})\leq(1+\delta)\theta
2θ\displaystyle 2\theta d(w,xi)2(1+δ)θ.\displaystyle\leq d(w,x_{i})\leq 2(1+\delta)\theta.

We will now use these inequalities in three of the four pieces of the infrasup pp-umbel inequality to bound the distance between the branches of the umbel. Observe that

θp\displaystyle\theta^{p} 12pinfnd(w,xi)p\displaystyle\leq\frac{1}{2^{p}}\inf_{n\in\mathbb{N}}d(w,x_{i})^{p}
d(z,w)p\displaystyle d(z,w)^{p} (1+δ)pθp\displaystyle\leq(1+\delta)^{p}\theta^{p}
supnd(xi,z)p\displaystyle\sup_{n\in\mathbb{N}}d(x_{i},z)^{p} (1+δ)pθp\displaystyle\leq(1+\delta)^{p}\theta^{p}

which, by taking the pp-th root of both sides, gives us that

1CUinfijd(xi,xj)[(1+δ)p1]1/pθ.\frac{1}{C_{U}}\inf_{i\neq j\in\mathbb{N}}d(x_{i},x_{j})\leq[(1+\delta)^{p}-1]^{1/p}\theta.

By applying Taylor’s Theorem to (1+δ)p(1+\delta)^{p} and utilizing δ<1\delta<1, we see for some ξ(0,δ)\xi\in(0,\delta)

[(1+δ)p1]1/p\displaystyle[(1+\delta)^{p}-1]^{1/p} =[pδ+p(p1)(1+ξ)p22δ2]1/p\displaystyle=\left[p\delta+\frac{p(p-1)(1+\xi)^{p-2}}{2}\delta^{2}\right]^{1/p}
[pδ+p(p1)2p22δ2]1/p\displaystyle\leq\left[p\delta+\frac{p(p-1)2^{p-2}}{2}\delta^{2}\right]^{1/p}
p1/pδ1/p+(p(p1)2)1/p2p2pδ2/p\displaystyle\leq p^{1/p}\delta^{1/p}+\left(\frac{p(p-1)}{2}\right)^{1/p}2^{\frac{p-2}{p}}\delta^{2/p}
2δ1/p+4δ2/p6δ1/p\displaystyle\leq 2\delta^{1/p}+4\delta^{2/p}\leq 6\delta^{1/p}

Putting everything together and moving CUC_{U} to the other side, gives us the result. ∎

2.1.4. Coloring Lemma

From here, we can move on to the coloring lemma. To this end, for a tree with encoding []h[\mathbb{N}]^{\leq h}, let VP([]h)\textnormal{VP}([\mathbb{N}]^{\leq h}) be the set of vertical vertex pairs (m¯,n¯)(\bar{m},\bar{n}) such that m¯<n¯\bar{m}<\bar{n}. By coloring this set with a finite number of colors, we hope to extract a subtree whose coloring is horizontally monochromatic , i.e., (m¯,n¯)(\bar{m},\bar{n}) has the same color as (j¯,k¯)(\bar{j},\bar{k}) if d(m¯,r)=d(j¯,r)d(\bar{m},r)=d(\bar{j},r) and d(n¯,r)=d(k¯,r)d(\bar{n},r)=d(\bar{k},r). Guaranteeing the existence of such a subcoloring is the content of the next lemma.

Lemma 2.3 (Tree Vertex Pair Coloring Lemma).

Let h,ch,c\in\mathbb{N}. Let []h[\mathbb{\mathbb{N}}]^{\leq h} be the complete countably branching tree of height hh with root rr. Color each of the elements of VP([]h)\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}) by one of cc colors. Then there exists an infinite 𝕄\mathbb{M}\subset\mathbb{N} such that the coloring on VP([𝕄]h)\textnormal{VP}([\mathbb{M}]^{\leq h}) is horizontally monochromatic.

Proof.

The first step is to use the coloring on VP([]h)\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}) to induce a coloring on the leaves, []h[\mathbb{\mathbb{N}}]^{h}. From there, we apply the Infinite Ramsey Theorem to get our subset 𝕄\mathbb{M}, giving us the result.

Let χ:VP([]h){1,2,,c}\chi:\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h})\to\{1,2,...,c\} be the coloring map, and let (n1,n2,,nh)=n¯[]h(n_{1},n_{2},...,n_{h})=\bar{n}\in[\mathbb{\mathbb{N}}]^{h}. Consider for all 0i<jh0\leq i<j\leq h, vi=n¯[1:i]v_{i}=\bar{n}[1:i] and vj=n¯[1:j]v_{j}=\bar{n}[1:j] where if i=0i=0, then v0=rv_{0}=r. Observe that vi<vj<n¯v_{i}<v_{j}<\bar{n}. Thus, we have that (vi,vj)VP([]h)(v_{i},v_{j})\in\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}). Let An¯A_{\bar{n}} be the set of all such pairs. Since this set has finite cardinality r(h2)r^{\binom{h}{2}}, we can pick an arbitrary ordering of this set, but whatever ordering is chosen, it must be consistently chosen across all n¯[]h\bar{n}\in[\mathbb{\mathbb{N}}]^{h}.

Using this, the induced color for our leaf n¯\bar{n} is given by the vector of colors,

(χ(z1),χ(z2),,χ(z(h2)),(\chi(z_{1}),\chi(z_{2}),...,\chi(z_{\binom{h}{2}}),

where the ziz_{i}’s are the ordered elements of An¯A_{\bar{n}}.

Following this procedure for every element of []h[\mathbb{\mathbb{N}}]^{h} associates a color to each of the leaves of []h[\mathbb{\mathbb{N}}]^{\leq h}. By the construction of this induced coloring, the leaves are colored by at most c(h2)c^{\binom{h}{2}} colors. Since this is a finite number of colors, we can apply the Infinite Ramsey Theorem to extract 𝕄\mathbb{M}\subset\mathbb{N} such that this induced coloring is monochromatic on the leaves of [𝕄]h[\mathbb{M}]^{\leq h}. Since this induced coloring is monochromatic, we must have that the color vector for every leaf of [𝕄]h[\mathbb{M}]^{\leq h} has the same entry in every position, giving us the result. ∎

2.1.5. Path Lemma

The final necessary lemma is the path lemma. In the case of countably branching trees, Matoušek’s original lemma can be used untouched, but for more complicated graphs, such as diamonds, a more general version is necessary which we will prove in later sections. Here, we present a reformulated proof of Matoušek’s original result.

To begin, we have a small helper lemma.

Lemma 2.4 (Bound on Ratio of Bounded, Decreasing Sequences).

Let K1K\geq 1 and nn\in\mathbb{N}. Let {ki}i=0n\{k_{i}\}_{i=0}^{n} be a decreasing sequence in [1,K][1,K]. Then there exists i{0,1,,n}i\in\{0,1,...,n\} such that

1kiki+11+Knki+1.1\leq\frac{k_{i}}{k_{i+1}}\leq 1+\frac{K}{nk_{i+1}}.

In particular, if nKpn\geq K^{p} for some p1p\geq 1, then we have instead

1kiki+11+1ki+1p.1\leq\frac{k_{i}}{k_{i+1}}\leq 1+\frac{1}{k_{i+1}^{p}}.
Proof.

Divide the interval [1,K][1,K] into nn segments, each of length K1n\frac{K-1}{n}. By the pigeonhole principle, since we have n+1n+1 of the kik_{i}’s, there must exist an i{0,,n1}i\in\{0,...,n-1\} such that kik_{i} and ki+1k_{i+1} are in the same interval. This implies that

0kiki+1K1nKn.0\leq k_{i}-k_{i+1}\leq\frac{K-1}{n}\leq\frac{K}{n}.

Dividing through by ki+1k_{i+1} and adding one, we have

1kiki+11+Knki+1.1\leq\frac{k_{i}}{k_{i+1}}\leq 1+\frac{K}{nk_{i+1}}.

If we have that nKpn\geq K^{p}, then since Kp1ki+1p1K^{p-1}\geq k_{i+1}^{p-1} and p1p\geq 1, we have

1kiki+11+1ki+1p.1\leq\frac{k_{i}}{k_{i+1}}\leq 1+\frac{1}{k_{i+1}^{p}}.

Lemma 2.5 (Path Lemma).

Let n4n\geq 4, (X,d)(X,d) a metric space, and Pn={0,1,,n}P_{n}=\{0,1,...,n\} be viewed as a metric space equipped with the absolute value metric. Let f:PnXf:P_{n}\to X be a mapping such that for some λ>0\lambda>0 and K1K\geq 1, we have

λ|xy|dX(f(x),f(y))λK|xy|.\lambda|x-y|\leq d_{X}(f(x),f(y))\leq\lambda K|x-y|.

and define

Li:=sup|xy|=2id(f(x),f(y))λ2i.L_{i}:=\sup_{|x-y|=2^{i}}\frac{d(f(x),f(y))}{\lambda 2^{i}}.

Then there exists i{0,1,,log2n}i\in\{0,1,...,\lfloor\log_{2}n\rfloor\} and vPnv\in P_{n} such that for Z={v,v+2i,v+2i+1}Z=\{v,v+2^{i},v+2^{i+1}\} we have every x,yZx,y\in Z satisfies,

λLi+1(1Klog2nLi+1)|xy|d(f(x),f(y))λLi+1(1+Klog2nLi+1)|xy|.\lambda L_{i+1}\left(1-\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)|x-y|\leq d(f(x),f(y))\leq\lambda L_{i+1}\left(1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)|x-y|.

In particular, if n22CKpn\geq 2^{\lceil 2CK^{p}\rceil} for some p1p\geq 1 and C1C\geq 1, then we have

λLi+1(112CLi+1p)|xy|d(f(x),f(y))λLi+1(1+12CLi+1p)|xy|,\lambda L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)|x-y|\leq d(f(x),f(y))\leq\lambda L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)|x-y|,

and

dist(f|Z)1+2CLi+1p.\textnormal{dist}(f_{|Z})\leq 1+\frac{2}{CL_{i+1}^{p}}.
Proof.

Observe that KLiLi+11K\geq L_{i}\geq L_{i+1}\geq 1 by the triangle inequality and λK\lambda K being the Lipschitz constant of ff. With this in mind, we can consider the sequence {Li}i=0log2n\{L_{i}\}_{i=0}^{\lfloor\log_{2}n\rfloor}. Observe that this is a decreasing sequence on the interval [1,K][1,K], so we may apply the Sequence Lemma from above. This gives us that for some ii

1LiLi+11+Klog2nLi+1.1\leq\frac{L_{i}}{L_{i+1}}\leq 1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}.

Let v,v+2i+1Pnv,v+2^{i+1}\in P_{n} be such that Li+1L_{i+1} is attained on them, i.e.,

Li+1=d(f(v),f(v+2i+1))λ2i+1.L_{i+1}=\frac{d(f(v),f(v+2^{i+1}))}{\lambda 2^{i+1}}.

We now let Z={v,v+2i,v+2i+1}Z=\{v,v+2^{i},v+2^{i+1}\} and wish to get bounds on the quantity d(f(x),f(y))d(f(x),f(y)) for x,yZx,y\in Z. Let |xy|=2i|x-y|=2^{i}, then we have

d(f(x),f(y))2iλLi2iλ(1+Klog2nLi+1)Li+1.\displaystyle d(f(x),f(y))\leq 2^{i}\lambda L_{i}\leq 2^{i}\lambda\left(1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)L_{i+1}.

By using the reverse triangle inequality and the upper bound above, we have for v,v+2iv,v+2^{i}

d(f(v),f(v+2i))\displaystyle d(f(v),f(v+2^{i})) d(f(v),f(v+2i+1))d(f(v+2i+1),f(v+2i))\displaystyle\geq d(f(v),f(v+2^{i+1}))-d(f(v+2^{i+1}),f(v+2^{i}))
2i+1λLi+12iλ(1+Klog2nLi+1)Li+1\displaystyle\geq 2^{i+1}\lambda L_{i+1}-2^{i}\lambda\left(1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)L_{i+1}
=2iλLi+1(1Klog2nLi+1).\displaystyle=2^{i}\lambda L_{i+1}\left(1-\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right).

Using a similar argument for v+2i,v+2i+1v+2^{i},v+2^{i+1}, we see

d(f(v+2i),f(v+2i+1))2iλLi+1(1Klog2nLi+1).\displaystyle d(f(v+2^{i}),f(v+2^{i+1}))\geq 2^{i}\lambda L_{i+1}\left(1-\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right).

This implies that for x,yZx,y\in Z and the fact that v,v+2i+1v,v+2^{i+1} achieves Li+1L_{i+1} we have

λLi+1(1Klog2nLi+1)|xy|d(f(x),f(y))λLi+1(1+Klog2nLi+1)|xy|.\displaystyle\lambda L_{i+1}\left(1-\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)|x-y|\leq d(f(x),f(y))\leq\lambda L_{i+1}\left(1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\right)|x-y|.

Thus, since 1+x1x1+4x\frac{1+x}{1-x}\leq 1+4x for 0x1/20\leq x\leq 1/2, if Klog2nLi+11/2\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}\leq 1/2, then

dist(f|Z)1+Klog2nLi+11Klog2nLi+11+4Klog2nLi+1.\textnormal{dist}(f_{|Z})\leq\frac{1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}}{1-\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}}\leq 1+\frac{4K}{\lfloor\log_{2}n\rfloor L_{i+1}}.

This is satisfied exactly when n22CKpn\geq 2^{\lceil 2CK^{p}\rceil} for some p1p\geq 1 and C1C\geq 1, in which case, we also have the improved estimate

dist(f|Z)1+4Klog2nLi+11+4K2CKpLi+11+2CLi+1p,\textnormal{dist}(f_{|Z})\leq 1+\frac{4K}{\lfloor\log_{2}n\rfloor L_{i+1}}\leq 1+\frac{4K}{2CK^{p}L_{i+1}}\leq 1+\frac{2}{CL_{i+1}^{p}},

and we have

λLi+1(112CLi+1p)|xy|d(f(x),f(y))λLi+1(1+12CLi+1p)|xy|.\lambda L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)|x-y|\leq d(f(x),f(y))\leq\lambda L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)|x-y|.

2.2. Embedding Obstruction Theorem

Theorem 2.6 (Lower Bound for Embedding of Countably Branching Trees).

Let (X,dX)(X,d_{X}) be a metric space which satisfies the infrasup pp-umbel inequality (Definition 2.1.3) for some CU1C_{U}\geq 1 and p1p\geq 1. Then the minimum distortion necessary for embedding []h[\mathbb{\mathbb{N}}]^{\leq h} into XX is at least C(log2h)1/pC(\log_{2}h)^{1/p} for some C>0C>0, dependent only on pp and CUC_{U}.

Proof.

Let hh\in\mathbb{N} and suppose that the tree []h[\mathbb{\mathbb{N}}]^{\leq h} embeds into XX with Lipschitz map f:[]hXf:[\mathbb{\mathbb{N}}]^{\leq h}\to X for some λ>0\lambda>0 and K1K\geq 1 such that for every m¯,n¯[]h\bar{m},\bar{n}\in[\mathbb{\mathbb{N}}]^{\leq h},

λdT(m¯,n¯)dX(f(m¯),f(n¯))KλdT(m¯,n¯).\lambda d_{T}(\bar{m},\bar{n})\leq d_{X}(f(\bar{m}),f(\bar{n}))\leq K\lambda d_{T}(\bar{m},\bar{n}).

Observe, however, that without loss of generality we can assume that λ=1\lambda=1 by rescaling the metric dXd_{X}, which preserves the infrasup pp-umbel inequality and its constant, CUC_{U}. We will show that this distortion must grow at a particular rate with respect to hh, i.e., K>O((log2h)1/p)K>O((\log_{2}h)^{1/p}) for sufficiently large hh.

This proof will have four main parts. First, using the coloring lemma (Lemma 2.3) we will show that there exists a subtree with a horizontally monochromatic coloring created from the log distortion of pairs of vertices in VP([]h)\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}). Second, we will use the path lemma (Lemma 2.5) to find a subset of an arbitrary root leaf path in []h[\mathbb{\mathbb{N}}]^{\leq h} that is well embedded into XX. Third, we will leverage our horizontally monochromatic coloring and the well embedded subset to construct a vertical δ\delta-umbel in XX. Finally, we can then apply the umbel lemma (Lemma 2.2) to realize a contradiction to the conditions of the path lemma, giving us the desired bound.

Let γ>0\gamma>0 be a parameter to be chosen later, and set r=log1+γKr=\lceil\log_{1+\gamma}K\rceil. Color the pairs in VP([]h)\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}) according to the log distortion of their embedding by ff, namely a pair (m¯,n¯)VP([]h)(\bar{m},\bar{n})\in\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}) gets the color

log1+γ(dX(f(m¯),f(n¯))dT(m¯,n¯)){0,,r}\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(\bar{m}),f(\bar{n}))}{d_{T}(\bar{m},\bar{n})}\right)\right\rfloor\in\{0,...,r\}

We use the log distortion rather than the distortion in order to facilitate later computations and give credit to [MN13] for this technical argument. By the coloring lemma (Lemma 2.3), there exists [𝕄]h[]h[\mathbb{M}]^{\leq h}\subset[\mathbb{\mathbb{N}}]^{\leq h} such that the color of pairs (m¯,n¯)VP([𝕄]h)(\bar{m},\bar{n})\in\textnormal{VP}([\mathbb{M}]^{\leq h}) only depends on the level of m¯\bar{m} and n¯\bar{n}.

With this in mind, let m¯\bar{m} be a leaf in [𝕄]h[\mathbb{M}]^{\leq h}. This creates a path of length hh from the root rr to m¯\bar{m} given by rm¯[1]m¯[1:2]m¯r\to\bar{m}[1]\to\bar{m}[1:2]\to...\to\bar{m}. This is isometric to PhP_{h}, so by the path lemma, there exists i{0,,log2h}i\in\{0,...,\lfloor\log_{2}h\rfloor\} and Z={w,z,x1}Z=\{w,z,x_{1}\} such that dT(w,z)=dT(z,x1)=dT(w,x1)2d_{T}(w,z)=d_{T}(z,x_{1})=\frac{d_{T}(w,x_{1})}{2} and we have for x,yZx,y\in Z

Li+1(1Klog2hLi+1)dT(x,y)dX(f(x),f(y))Li+1(1+Klog2hLi+1)dT(x,y),L_{i+1}\left(1-\frac{K}{\lfloor\log_{2}h\rfloor L_{i+1}}\right)d_{T}(x,y)\leq d_{X}(f(x),f(y))\leq L_{i+1}\left(1+\frac{K}{\lfloor\log_{2}h\rfloor L_{i+1}}\right)d_{T}(x,y),

where Li+1=dX(f(w),f(x1))dT(w,x1)L_{i+1}=\frac{d_{X}(f(w),f(x_{1}))}{d_{T}(w,x_{1})}. We now wish to extend this to a vertical umbel in the tree by choosing good parameters for γ\gamma. To this end, we have the following claim:

Claim.

If we have

γ=17(6CU)pKp17(6CU)pLi+1p,θ=Li+1(1114(6CU)pLi+1p)2i1+17(6CU)pKp\displaystyle\gamma=\frac{1}{7(6C_{U})^{p}K^{p}}\leq\frac{1}{7(6C_{U})^{p}L_{i+1}^{p}},\quad\quad\theta=\frac{L_{i+1}\left(1-\frac{1}{14(6C_{U})^{p}L_{i+1}^{p}}\right)2^{i}}{1+\frac{1}{7(6C_{U})^{p}K^{p}}}

with h256(6CU)pKp228(6CU)pKph\geq 2^{56(6C_{U})^{p}K^{p}}\geq 2^{\lceil 28(6C_{U})^{p}K^{p}\rceil}, then for δ=1(6CU)pLi+1p\delta=\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}, there exists x2,x3,Xx_{2},x_{3},...\in X such that U={f(w),f(z),f(x1),}XU=\{f(w),f(z),f(x_{1}),...\}\subset X forms a δ\delta-umbel.

Assuming the claim, we show how to proceed. If K((561/p)6CU)1(log2h)1/pK\geq((56^{1/p})6C_{U})^{-1}(\log_{2}h)^{1/p}, then this is the result. Suppose on the other hand that K<((561/p)6CU)1(log2h)1/pK<((56^{1/p})6C_{U})^{-1}(\log_{2}h)^{1/p}. By solving for hh, this implies that h>256(6CU)pKph>2^{56(6C_{U})^{p}K^{p}}. Thus, by utilizing the claim, we have that U={f(w),f(z),f(x1),f(x2),}XU=\{f(w),f(z),f(x_{1}),f(x_{2}),...\}\subset X is a vertical δ\delta-umbel in XX with δ=1(6CU)pLi+1p\delta=\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}. Observe that θ2iLi+1\theta\leq 2^{i}L_{i+1}. Thus, when we apply the Umbel Lemma to our δ\delta-umbel, we have

2i+1infijdX(f(xi),f(xj))6CUθδ1/p6CU(2iLi+1)6CULi+1=2i,\displaystyle 2^{i+1}\leq\inf_{i\neq j}d_{X}(f(x_{i}),f(x_{j}))\leq 6C_{U}\theta\delta^{1/p}\leq\frac{6C_{U}(2^{i}L_{i+1})}{6C_{U}L_{i+1}}=2^{i},

which is a contradiction, thus proving the result. ∎

Proof of claim.

For now, to better illustrate where exactly these values come from, we will give the exact values of the parameters only when it becomes necessary in the course of this proof. To this end, if h22CKph\geq 2^{\lceil 2CK^{p}\rceil} for some C1C\geq 1 to be determined later and p1p\geq 1, then we have by the path lemma for some ii

Li+1(112CLi+1p)2i\displaystyle L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i} dX(f(w),f(z))Li+1(1+12CLi+1p)2i\displaystyle\leq d_{X}(f(w),f(z))\leq L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)2^{i}
Li+1(112CLi+1p)2i\displaystyle L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i} dX(f(z),f(x1))Li+1(1+12CLi+1p)2i\displaystyle\leq d_{X}(f(z),f(x_{1}))\leq L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)2^{i}
Li+1(112CLi+1p)2i+1\displaystyle L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i+1} dX(f(w),f(x1))Li+1(1+12CLi+1p)2i+1\displaystyle\leq d_{X}(f(w),f(x_{1}))\leq L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)2^{i+1}

where w,z,x1[𝕄]hw,z,x_{1}\in[\mathbb{M}]^{\leq h} were the elements generated by the application of the path lemma.

Now observe that by the coloring lemma, we have for every n¯[𝕄]h\bar{n}\in[\mathbb{M}]^{\leq h} such that z<n¯z<\bar{n} and dT(z,n¯)=2id_{T}(z,\bar{n})=2^{i}, that the color of (z,n¯)(z,\bar{n}) is the same as the color of (z,x1)(z,x_{1}). Thus we can enumerate all of these n¯[𝕄]h\bar{n}\in[\mathbb{M}]^{\leq h} as {x2,x3,x4,.}\{x_{2},x_{3},x_{4},....\} where for every ii\in\mathbb{N}, we have dT(z,xi)=dT(w,z)=dT(w,z)2d_{T}(z,x_{i})=d_{T}(w,z)=\frac{d_{T}(w,z)}{2}. This gives us that the colors of (z,xi)(z,x_{i}) and (z,xj)(z,x_{j}) for every iji\neq j\in\mathbb{N} are the same. Armed with this information, we see that for every ii\in\mathbb{N},

log1+γ(dX(f(z),f(x1))dT(z,x1))=log1+γ(dX(f(z),f(xi))dT(z,xi)).\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{1}))}{d_{T}(z,x_{1})}\right)\right\rfloor=\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{i}))}{d_{T}(z,x_{i})}\right)\right\rfloor.

Observe now by the definition of the floor function, we have

log1+γ(dX(f(z),f(x1))dT(z,x1))1log1+γ(dX(f(z),f(xi))dT(z,xi))log1+γ(dX(f(z),f(x1))dT(z,x1))+1.\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{1}))}{d_{T}(z,x_{1})}\right)-1\leq\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{i}))}{d_{T}(z,x_{i})}\right)\leq\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{1}))}{d_{T}(z,x_{1})}\right)+1.

Similarly, since we know that log1+a(1+a)=1\log_{1+a}(1+a)=1 for all a>0a>0, we have

log1+γ(dX(f(z),f(x1))(1+γ)dT(z,x1))log1+γ(dX(f(z),f(xi))dT(z,xi))log1+γ((1+γ)dX(f(z),f(x1))dT(z,x1)).\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{1}))}{(1+\gamma)d_{T}(z,x_{1})}\right)\leq\log_{1+\gamma}\left(\frac{d_{X}(f(z),f(x_{i}))}{d_{T}(z,x_{i})}\right)\leq\log_{1+\gamma}\left(\frac{(1+\gamma)d_{X}(f(z),f(x_{1}))}{d_{T}(z,x_{1})}\right).

Now, applying the upper and lower bounds for dX(f(z),f(x1))d_{X}(f(z),f(x_{1})) given by the path lemma, this gives us

Li+1(112CLi+1p)2i(1+γ)dT(z,x1)\displaystyle\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{(1+\gamma)d_{T}(z,x_{1})} dX(f(z),f(xi))dT(z,xi)(1+γ)Li+1(1+12CLi+1p)2idT(z,xi).\displaystyle\leq\frac{d_{X}(f(z),f(x_{i}))}{d_{T}(z,x_{i})}\leq\frac{(1+\gamma)L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{d_{T}(z,x_{i})}.

Notice here that this line implies that moving from one tip of the umbel to another, only incurs a 1+γ1+\gamma penalty for some γ\gamma, and the distortion gains a factor of (1+γ)2(1+\gamma)^{2}. In addition, by following the same steps for ww and xix_{i} for ii\in\mathbb{N}, we have the same inequality chain with a factor of two appearing on the left and the right. This gives us then that for all i1i\geq 1, since 1+γ11+\gamma\geq 1

Li+1(112CLi+1p)2i1+γ\displaystyle\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{1+\gamma} dX(f(w),f(z))Li+1(1+12CLi+1p)(1+γ)2i\displaystyle\leq d_{X}(f(w),f(z))\leq L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)(1+\gamma)2^{i}
Li+1(112CLi+1p)2i1+γ\displaystyle\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{1+\gamma} dX(f(z),f(xi))Li+1(1+12CLi+1p)(1+γ)2i\displaystyle\leq d_{X}(f(z),f(x_{i}))\leq L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)(1+\gamma)2^{i}
Li+1(112CLi+1p)2i+11+γ\displaystyle\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i+1}}{1+\gamma} dX(f(w),f(xi))Li+1(1+12CLp)(1+γ)2i+1.\displaystyle\leq d_{X}(f(w),f(x_{i}))\leq L_{i+1}\left(1+\frac{1}{2CL^{p}}\right)(1+\gamma)2^{i+1}.

From here, we are ultimately trying to find a way to rewrite this as a δ\delta-umbel for some δ\delta, but this requires choosing θ\theta. In this case, the most logical option would be

θ=Li+1(112CLi+1p)2i1+γ.\theta=\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{1+\gamma}.

This change of variable trick, along with the estimate 1+x1x1+4x\frac{1+x}{1-x}\leq 1+4x for 0x120\leq x\leq\frac{1}{2}, enables us to rewrite these inequalities with the distortion easily calculable, i.e.,

θ\displaystyle\theta dX(f(w),f(z))θ(1+γ)2(1+2CLi+1p)\displaystyle\leq d_{X}(f(w),f(z))\leq\theta(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right)
θ\displaystyle\theta dX(f(z),f(xi))θ(1+γ)2(1+2CLi+1p)\displaystyle\leq d_{X}(f(z),f(x_{i}))\leq\theta(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right)
2θ\displaystyle 2\theta dX(f(w),f(xi))2θ(1+γ)2(1+2CLi+1p).\displaystyle\leq d_{X}(f(w),f(x_{i}))\leq 2\theta(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right).

This gives us a distortion of (1+γ)2(1+2CLi+1p)(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right) when ff is restricted to any triple {w,z,xi}\{w,z,x_{i}\}. We would like, however, for this to be of the form 1+δ1+\delta. This is where our choice of γ\gamma and CC comes into play. It is important to notice that γ\gamma, due to how early in the proof it was chosen, can only depend on K,p,K,p, and CUC_{U}. As a result, we will choose γ=2CKp2CLi+1p\gamma=\frac{2}{CK^{p}}\leq\frac{2}{CL_{i+1}^{p}}. Observe that this inequality respects the upper and lower portions of the chain of inequalities we have established thus far, i.e., replacing γ\gamma with 2CLi+1p\frac{2}{CL_{i+1}^{p}} instead of 2CKp\frac{2}{CK^{p}} inside the (1+γ)3(1+\gamma)^{3} term will force the rightmost inequalities to be larger, preserving the inequalities. Thus, we now have for γ=2CKp2CLi+1p\gamma=\frac{2}{CK^{p}}\leq\frac{2}{CL_{i+1}^{p}}

θ\displaystyle\theta dX(f(w),f(z))θ(1+2CLi+1p)3\displaystyle\leq d_{X}(f(w),f(z))\leq\theta\left(1+\frac{2}{CL_{i+1}^{p}}\right)^{3}
θ\displaystyle\theta dX(f(z),f(xi))θ(1+2CLi+1p)3\displaystyle\leq d_{X}(f(z),f(x_{i}))\leq\theta\left(1+\frac{2}{CL_{i+1}^{p}}\right)^{3}
2θ\displaystyle 2\theta dX(f(w),f(xi))2θ(1+2CLi+1p)3\displaystyle\leq d_{X}(f(w),f(x_{i}))\leq 2\theta\left(1+\frac{2}{CL_{i+1}^{p}}\right)^{3}

In addition, by recognizing that (1+x)31+7x(1+x)^{3}\leq 1+7x so long as 0x10\leq x\leq 1, our choice for CC should be one which guarantees that 2CLi+1p1\frac{2}{CL_{i+1}^{p}}\leq 1 which happens for any C2C\geq 2 since Li+1p1L_{i+1}^{p}\geq 1. Looking to when we apply the umbel lemma, however, we would like to cancel the term 6CU6C_{U} by our choice of CC and force δ<1\delta<1. As a result, we let C=14(6CU)p2C=14(6C_{U})^{p}\geq 2 for p1p\geq 1, though this may be improved with a more careful choice of CC. In addition, notice that this necessitates CUp17(6)pC_{U}^{p}\geq\frac{1}{7(6)^{p}}, but if CUpC_{U}^{p} is smaller than this quantity, we simply omit it from our constant CC. This final choice now forces the distortion on w,z,w,z, and xix_{i} to be 1+1(6CU)pLi+1p1+\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}, giving us for

γ=17(6CU)pKp,θ=Li+1(112CLi+1p)2i1+17(6CU)pKp,δ=1(6CU)pLi+1p,\displaystyle\gamma=\frac{1}{7(6C_{U})^{p}K^{p}},\quad\theta=\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)2^{i}}{1+\frac{1}{7(6C_{U})^{p}K^{p}}},\quad\delta=\frac{1}{(6C_{U})^{p}L_{i+1}^{p}},

that the following estimate holds

θ\displaystyle\theta dX(f(w),f(z))θ(1+δ)=θ(1+1(6CU)pLi+1p)\displaystyle\leq d_{X}(f(w),f(z))\leq\theta(1+\delta)=\theta\left(1+\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}\right)
θ\displaystyle\theta dX(f(z),f(xi))θ(1+δ)=θ(1+1(6CU)pLi+1p)\displaystyle\leq d_{X}(f(z),f(x_{i}))\leq\theta(1+\delta)=\theta\left(1+\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}\right)
2θ\displaystyle 2\theta dX(f(w),f(xi))2θ(1+δ)=2θ(1+1(6CU)pLi+1p).\displaystyle\leq d_{X}(f(w),f(x_{i}))\leq 2\theta(1+\delta)=2\theta\left(1+\frac{1}{(6C_{U})^{p}L_{i+1}^{p}}\right).

Notice, now, that as a direct consequence of this theorem and Lemma 2.1, we have the following theorem, proven previously in [BZ16].

Theorem 2.7 ((βp)(\beta_{p}) Embedding Obstruction).

Let XX be a Banach space which satisfies the (βp)(\beta_{p}) property for some p>1p>1. Then the minimum distortion necessary for embedding TnωT_{n}^{\omega} into XX is at least C(logn)1/pC(\log n)^{1/p} for some C>0C>0, dependent only on pp and CβC_{\beta}.

3. Countably Branching Diamonds

Abstracting from the proof of Theorem 2.6, there are three steps needed to check the embeddability of certain classes of graphs into various Banach spaces satisfying asymptotic inequalities. First, an appropriate subgraph must be identified which is in some way intrinsic to the family of graphs considered and can appropriately utilize the given asymptotic inequality. In the case of trees, we considered the δ\delta-umbel and property (βp)(\beta_{p}), but in the case of the countably branching diamond graphs DnωD_{n}^{\omega}, we will be looking for sub-diamonds D1ωD_{1}^{\omega} and asymptotic midpoint uniform convexity (AMUC). These diamonds, as we will show, form the base case for constructing the diamond graphs and have very natural midpoints to consider.

Secondly, we need to be able to extract subgraphs which have “good” colorings. These graphs should also be, in the asymptotic setting, graph isomorphic to a copy of the original graph. For trees, we extracted a subgraph on which every element of VP([]h)\textnormal{VP}([\mathbb{\mathbb{N}}]^{\leq h}) at the same level was monochromatic, but for diamonds, the situation is a bit more precarious. With trees, every vertex can be uniquely identified with a subpath of a root-leaf path, but diamonds, by virtue of having cycles, do not have this property. As a result, we will present a weaker formulation of a coloring lemma which will yield a subgraph whose coloring is horizontally monochromatic when restricted to scaled D1ωD_{1}^{\omega} sub-diamonds of DnωD_{n}^{\omega}.

Finally, we need some way of finding well embedded copies of these important subgraphs. A more generalized version of the path lemma can be used just as in the case of trees with one of the key differences being the importance of the progression of points yielded by the lemma. In the case of trees, any progression of three equally spaced vertices on an arbitrary root-leaf path could be used to generate a δ\delta-umbel. In the case of diamonds, however, as we will see, we will need to consider specific subsets of paths connecting the ss and tt vertices of the diamond.

3.1. Definitions

To begin, we shall give a graph theoretic definition of stst-graphs, slash products, and the diamond graphs.

For the reader’s convenience, we reproduce here the definitions of stst-graphs and slash products from [Bau+17]. All graphs in this paper are assumed to be connected, simple, unweighted, and equipped with the shortest path metric dd.

Definition 3.1.1 (stst-Graphs and Slash Products).

A directed graph G=(V,E)G=(V,E) is called an stst-graph if it has two distinguished points ss and tt, which we will denote s(G)s(G) and t(G)t(G). In this paper, the orientation of an edge, if not explicitly stated, will be viewed as “flowing” from s(G)s(G) to t(G)t(G), i.e., if e=(u,v)E(G)e=(u,v)\in E(G), then d(u,s(G))<d(v,s(G))d(u,s(G))<d(v,s(G)). Given two stst-graphs, HH and GG, there is a natural operation which composes them. We call this composition the slash product, given by HGH\varoslash G. This composition can be viewed as replacing every edge of HH with a copy of GG. In some cases, we may take e=(u,v)E(H)e=(u,v)\in E(H) and use eGe\varoslash G to refer to the copy of GG between uu and vv, or evie\varoslash v_{i} to refer to the vertex viv_{i} of GG which comes from replacing the edge eE(H)e\in E(H) with GG. We also take V(eG)V(e\varoslash G) to be the union of V(G)/{s(G),t(G)}V(G)/\{s(G),t(G)\} and {u,v}\{u,v\}. In these cases, we view the directed edge ee as an stst-graph with s(e)=us(e)=u and t(e)=vt(e)=v and treat this as a way of indexing into specific pieces of the graph HGH\varoslash G. The following three steps rigorously define this process:

  1. (1)

    V(HG):=V(H)[E(H)×(V(G)/{s(G),t(G)})]V(H\varoslash G):=V(H)\cup[E(H)\times(V(G)/\{s(G),t(G)\})].

  2. (2)

    Given an oriented edge e=(u,v)E(H)e=(u,v)\in E(H), there are |E(G)||E(G)| edges given by the union of the following sets which account for edges contained within GG that do not connect to s(G)s(G) or t(G)t(G) :

    1. (a)

      {((e,v1),(e,v2))|(v1,v2)E(G) and v1,v2{s(G),t(G)}}\left\{\ \left((e,v_{1}),(e,v_{2})\right)\ |\ (v_{1},v_{2})\in E(G)\textnormal{ and }v_{1},v_{2}\notin\{s(G),t(G)\}\right\}

    2. (b)

      {(u,(e,w))|(s(G),w)E(G)}{((e,w),u)|(w,s(G))E(G)}\left\{\ (u,(e,w))\ |\ (s(G),w)\in E(G)\right\}\cup\left\{\ ((e,w),u)\ |\ (w,s(G))\in E(G)\right\}

    3. (c)

      {((e,w),v)|(w,t(G))E(G)}{(v,(e,w))|(t(G),w)E(G)}\left\{\ ((e,w),v)\ |\ (w,t(G))\in E(G)\right\}\cup\left\{\ (v,(e,w))\ |\ (t(G),w)\in E(G)\right\}.

  3. (3)

    s(HG)=s(H)s(H\varoslash G)=s(H) and t(HG)=t(H)t(H\varoslash G)=t(H), and similarly, for e=(u,v)E(H)e=(u,v)\in E(H), s(eG)=us(e\varoslash G)=u and t(eG)=vt(e\varoslash G)=v.

Since the slash product is associative and also a directed stst-graph, one can define the slash power of a directed stst-graph, GnG^{\varoslash^{n}}, for all nn\in\mathbb{N} by setting G1:=GG^{\varoslash^{1}}:=G and Gn+1:=GnGG^{\varoslash^{n+1}}:=G^{\varoslash^{n}}\varoslash G. As a convention, we let G0G^{\varoslash^{0}} be the graph consisting of a single edge connecting s(G)s(G) and t(G)t(G). As also noted in [Bau+17], symmetric graphs (viewed as embedded in the plane) do not depend on the orientation of the edges of GG.

Following the example of [Bau+17], we can now set forth a definition of D1kD_{1}^{k}, the kk-branching diamond graph for k{ω}k\in\mathbb{N}\cup\{\omega\}. Let K2,kK_{2,k} be the complete bipartite graph with two vertices on the left and kk many vertices on the right, where every vertex on the left is connected to every vertex on the right. The vertices on the left can be labeled ss and tt, and the vertices on the right are labeled in an arbitrary way by {xi}i=1k\{x_{i}\}_{i=1}^{k}. The set of edges here is given by {(s,xi)}i=1k{(xi,t)}i=1k\{(s,x_{i})\}_{i=1}^{k}\cup\{(x_{i},t)\}_{i=1}^{k}, giving a natural orientation for edges flowing from ss to tt. This gives us D1kD_{1}^{k}. (We can think of moving the tt vertex to the other side of the xix_{i} vertices for a more natural shape.) Now we can define the nn-th kk-branching diamond as Dnk=(D1k)nD_{n}^{k}=(D_{1}^{k})^{\varoslash^{n}}. There are also non-recursive methods of defining the diamond graphs, and we refer the reader to [Bau+17] for a treatment of this topic.

Refer to caption
(a) D1ωD_{1}^{\omega}
Refer to caption
(b) D2ωD_{2}^{\omega}
Figure 2. Examples of D1ωD_{1}^{\omega} and D2ωD_{2}^{\omega} along with their vertex labeling

Next, we must define the convexity condition on our Banach space which forces the embedding obstruction. The (βp)(\beta_{p}) property was natural for studying embedding obstructions of trees, as it dealt with countably branching tripods, an unavoidable structure in the countably branching trees. Similarly, the midpoints of the countably branching diamonds are a natural target for asymptotic inequalities. To that end, we first define δ\delta-approximate midpoint sets.

Definition 3.1.2 (δ\delta-Approximate Midpoint Sets).

Given a metric space (X,d)(X,d), x,yXx,y\in X, and δ(0,1)\delta\in(0,1), we define the δ\delta-approximate midpoint set for xx and yy, Mid(x,y,δ)\textnormal{Mid}(x,y,\delta), as

Mid(x,y,δ):={zX|max{d(x,z),d(z,y)}(1+δ)2d(x,y)}.\textnormal{Mid}(x,y,\delta):=\left\{z\in X\ |\ \max\left\{d(x,z),d(z,y)\right\}\leq\frac{(1+\delta)}{2}d(x,y)\right\}.

From here, we can begin to define asymptotic midpoint uniform convexity, and the characterization which will be most useful to us in this context.

Definition 3.1.3 (Asymptotic Midpoint Uniform Convexity (AMUC)).

A Banach space XX is said to be asymptotically midpoint uniformly convex if for every ϵ(0,1)\epsilon\in(0,1), the quantity

δ~X(ϵ)=infxSXsupYcof(X)infySYmax{x+ϵy,xϵy}1\tilde{\delta}_{X}(\epsilon)=\inf_{x\in S_{X}}\sup_{Y\in\textnormal{cof}(X)}\inf_{y\in S_{Y}}\max\{\|x+\epsilon y\|,\|x-\epsilon y\|\}-1

is greater than zero where δ~X(ϵ)\tilde{\delta}_{X}(\epsilon) is called the modulus of asymptotic midpoint uniform convexity. We say that XX has power type pp modulus of asymptotic midpoint uniform convexity, or is pp-AMUC, if there exists some p(1,)p\in(1,\infty) and CM>0C_{M}>0 such that for every ϵ(0,1)\epsilon\in(0,1), δ~X(ϵ)CMϵp\tilde{\delta}_{X}(\epsilon)\geq C_{M}\epsilon^{p}.

It was shown in [Dil+16] that this is equivalent to the condition that

limδ0supxSxα(BX(x,1+δ)BX(x,1+δ))=0,\lim_{\delta\to 0}\sup_{x\in S_{x}}\alpha(B_{X}(x,1+\delta)\cap B_{X}(-x,1+\delta))=0,

where α(S)\alpha(S), the Kuratowski measure of non-compactness, is the infimum of d>0d>0 such that SS can be covered by finitely many sets of diameter dd. Notice that if x=1\|x\|=1 , then BX(x,1+δ)BX(x,1+δ)=Mid(x,x,δ)B_{X}(x,1+\delta)\cap B_{X}(-x,1+\delta)=\textnormal{Mid}(x,-x,\delta). In [Bau+17], Lemma 4.2 gives a quantitative estimate of this value, stating that for every ϵ(0,1)\epsilon\in(0,1) and xXx\in X in an AMUC Banach space, we have

α(Mid(x,x,δ~X(ϵ)/4)<4ϵ.\alpha(\textnormal{Mid}(-x,x,\tilde{\delta}_{X}(\epsilon)/4)<4\epsilon.

In the case where XX is pp-AMUC for some p>1p>1 and CM>0C_{M}>0, then we have a more refined estimate for every ϵ(0,1)\epsilon\in(0,1) (via a change in variables)

α(Mid(x,x,ϵ))<4(4ϵCM)1/p.\alpha(\textnormal{Mid}(-x,x,\epsilon))<4\left(\frac{4\epsilon}{C_{M}}\right)^{1/p}.

The last result we will need from [Bau+17] (Lemma 4.3) states that the δ\delta-approximate midpoint set is a subset of a “small” set. Quantitatively, we repeat the lemma here, adding the specific case where XX has a power type pp modulus of asymptotic midpoint uniform convexity.

Lemma 3.1 (Small Midpoint Set Lemma).

Let XX be a Banach space which is asymptotically midpoint uniformly convex, then for every ϵ(0,1)\epsilon\in(0,1) and s,tXs,t\in X, there exists a finite set SXS\subset X such that

Mid(s,t,δ~(ϵ)/4)S+2ϵstBX.\textnormal{Mid}(s,t,\tilde{\delta}(\epsilon)/4)\subset S+2\epsilon\|s-t\|B_{X}.

In the particular case where XX is pp-AMUC for p>1p>1, we have

Mid(s,t,ϵ)S+8(4ϵCM)1/pstBX.\textnormal{Mid}(s,t,\epsilon)\subset S+8\left(\frac{4\epsilon}{C_{M}}\right)^{1/p}\|s-t\|B_{X}.

Already with this lemma, we have the beginnings of a statement regarding well-embedded diamonds into an asymptotically midpoint uniformly convex Banach space XX. If each path from ss to tt in D1ωD_{1}^{\omega} is embedded into XX with distortion at most 1+δ1+\delta, this lemma tells us that each of the midpoints of our graph must be of the very specific format given by Lemma 3.1, allowing us to find two elements of the midpoint set which must be close together. With this in mind, we can now define a vertical ϵ\epsilon-diamond, creating our vertical δ\delta-umbel lemma equivalent.

3.1.1. ϵ\epsilon-Diamond Lemma

As mentioned previously, we will be looking this time for vertical ϵ\epsilon-diamonds which are defined similarly to vertical δ\delta-umbels.

Definition 3.1.4 (Vertical ϵ\epsilon-diamond).

Let (X,d)(X,d) be an infinite metric space. An ϵ\epsilon-diamond is a subspace D={s,t,x1,x2,x3,}D=\{s,t,x_{1},x_{2},x_{3},...\} of XX such that for some θ>0\theta>0 and for every ii\in\mathbb{N}

2θd(s,t)\displaystyle 2\theta\leq d(s,t) 2(1+ϵ)θ\displaystyle\leq 2(1+\epsilon)\theta
θd(s,xi)\displaystyle\theta\leq d(s,x_{i}) (1+ϵ)θ\displaystyle\leq(1+\epsilon)\theta
θd(xi,t)\displaystyle\theta\leq d(x_{i},t) (1+ϵ)θ\displaystyle\leq(1+\epsilon)\theta

or in other words, for every ii\in\mathbb{N} we have that {s,xi,t}\{s,x_{i},t\} is (1+ϵ)(1+\epsilon)-isomorphic (in the sense of distortion) to the metric space P2={0,1,2}P_{2}=\{0,1,2\} with the absolute value metric on it.

With a definition of a vertical ϵ\epsilon-diamond, we are now in a position to define a similar “good” graph embedding lemma for vertical ϵ\epsilon-diamonds.

Lemma 3.2 (ϵ\epsilon-Diamond Lemma).

Let D={s,t,x1,x2,}D=\{s,t,x_{1},x_{2},...\} be a vertical ϵ\epsilon-diamond for ϵ(0,1)\epsilon\in(0,1), and let XX be a pp-AMUC space for some p>1p>1 and CM>0C_{M}>0. This implies that for some i,ji,j\in\mathbb{N} and C>0C>0 which depends only on pp and CMC_{M},

xixjCϵ1/pθ.\|x_{i}-x_{j}\|\leq C\epsilon^{1/p}\theta.
Proof.

Observe that since DD is a vertical ϵ\epsilon-diamond, for every ii\in\mathbb{N}, xiMid(s,t,ϵ)x_{i}\in\textnormal{Mid}(s,t,\epsilon). Next, since XX is pp-AMUC, we can leverage Lemma 3.1 to see that

Mid(s,t,ϵ)S+8(4ϵCM)1/pstBX\textnormal{Mid}(s,t,\epsilon)\subset S+8\left(\frac{4\epsilon}{C_{M}}\right)^{1/p}\|s-t\|B_{X}

for some finite set S={s1,s2,,sn}XS=\{s_{1},s_{2},...,s_{n}\}\subset X. Observe now that since we have infinitely many xiMid(s,t,ϵ)x_{i}\in\textnormal{Mid}(s,t,\epsilon), there must be at least two xi,xjx_{i},x_{j} with iji\neq j such that xi=sk+zxix_{i}=s_{k}+z_{x_{i}} and xj=sk+zxjx_{j}=s_{k}+z_{x_{j}} for zxi,zxj8(4ϵCM)1/pstBXz_{x_{i}},z_{x_{j}}\in 8\left(\frac{4\epsilon}{C_{M}}\right)^{1/p}\|s-t\|B_{X}. Notice, however, that for these two elements, we must have

xixj=zxizxj16(4ϵCM)1/pst32(4)1/pCM1/pϵ1/p(1+ϵ)θ64(4)1/pCM1/pϵ1/pθ,\|x_{i}-x_{j}\|=\|z_{x_{i}}-z_{x_{j}}\|\leq 16\left(\frac{4\epsilon}{C_{M}}\right)^{1/p}\|s-t\|\leq\frac{32(4)^{1/p}}{C_{M}^{1/p}}\epsilon^{1/p}(1+\epsilon)\theta\leq\frac{64(4)^{1/p}}{C_{M}^{1/p}}\epsilon^{1/p}\theta,

where the final inequality relies on the fact that st2θ(1+ϵ)\|s-t\|\leq 2\theta(1+\epsilon) and ϵ(0,1)\epsilon\in(0,1). ∎

3.1.2. Diamond Coloring Lemma

We now turn our attention to the coloring lemma.

First, we make some observations about the proof of the non-embeddability of the countably branching trees. In that proof, we made very little usage of the level-respecting nature of the coloring on the subgraph. We never used the fact that all pairs of vertices at the same level were monochromatic. What we really used was that for every umbel in the countably branching tree, the pairs of vertices at the same level within that umbel were monochromatic. This is a substantially weaker assumption, allowing for umbels at the same level to be different colors. This observation will guide our coloring lemma for diamonds.

To begin, we need to define the vertical pairs for DnωD_{n}^{\omega}. Just as in the tree case, our definition will rely on the simple paths which begin at the “root” (in this case the vertex ss) and end at a “leaf” (the vertex tt). With this in mind, we let a pair of vertices (x,y)VP(Dnω)(x,y)\in\textnormal{VP}(D_{n}^{\omega}) if d(x,s)<d(y,s)d(x,s)<d(y,s) and xx and yy are on the same simple stst-path. The level of a vertex xx is given by d(x,s)d(x,s).

Now, we must define what substructures of DnωD_{n}^{\omega} our coloring should respect. Observe that for every 0in10\leq i\leq n-1, by the associativity of \varoslash, we have that Dnω=Dni+1ω(D1ωDiω)D_{n}^{\omega}=D_{n-i+1}^{\omega}\varoslash(D_{1}^{\omega}\varoslash D_{i}^{\omega}). This implies that for every eDni1ωe\in D_{n-i-1}^{\omega}, eD1ωDiωe\varoslash D_{1}^{\omega}\varoslash D_{i}^{\omega} is graph isomorphic to D1ωDiωD_{1}^{\omega}\varoslash D_{i}^{\omega}. Notice that the copy of D1ωD_{1}^{\omega} here is seen as having its edges replaced with copies of DiωD_{i}^{\omega}, meaning that for every kk\in\mathbb{N}, d(s(D1ω),xD1ωk)=d(xD1ωk,t(D1ω))=2id(s(D_{1}^{\omega}),x^{k}_{D_{1}^{\omega}})=d(x^{k}_{D_{1}^{\omega}},t(D_{1}^{\omega}))=2^{i}, where xD1ωkx^{k}_{D_{1}^{\omega}} refers to the xkx_{k} vertex of our D1ωD_{1}^{\omega} copy. This is what we will refer to as a 2i2^{i}-scaled diamond. For every choice of edges, this gives us a new and distinct 2i2^{i}-scaled diamond, except for the corresponding ss and tt vertices. Once we have colored VP(Dnω)\textnormal{VP}(D_{n}^{\omega}), we will be asking for a subgraph 𝔇\mathfrak{D} that is graph isomorphic to DnωD_{n}^{\omega} and one for which every 2i2^{i}-scaled diamond is horizontally monochromatic, i.e., for every 2i2^{i}-scaled diamond, D1ωD_{1}^{\omega}, and for every (a,b),(x,y)VP(D1ω)VP(Dnω)(a,b),(x,y)\in\textnormal{VP}(D_{1}^{\omega})\subset\textnormal{VP}(D_{n}^{\omega}) such that d(a,s(Dnω))=d(x,s(Dnω))d(a,s(D_{n}^{\omega}))=d(x,s(D_{n}^{\omega})) and d(b,s(Dnω))=d(y,s(Dnω))d(b,s(D_{n}^{\omega}))=d(y,s(D_{n}^{\omega})), (a,b)(a,b) and (x,y)(x,y) have the same color.

In the particular case of n=1n=1, this is a very straightforward task. Here, the only diamond is the entire graph. As a result, if we color VP(D1ω)\textnormal{VP}(D_{1}^{\omega}) with cc\in\mathbb{N} colors and color map χ:VP(D1ω){1,,c}\chi:\textnormal{VP}(D_{1}^{\omega})\to\{1,...,c\}, then for every stst-path, just as in the tree case, we can collect and order the pairs into a vector (χ((s,xi)),χ((xi,t)),χ((s,t)))(\chi((s,x_{i})),\chi((x_{i},t)),\chi((s,t))). There are at most c3c^{3} possible vectors, inducing a new coloring on the stst-paths of D1ωD_{1}^{\omega}. Since there are countably many stst-paths and only c3c^{3} colors on those paths, we can extract 𝔇D1ω\mathfrak{D}\subset D_{1}^{\omega} which is graph isomorphic to D1ωD_{1}^{\omega}, such that the stst-paths of 𝔇\mathfrak{D} are monochromatic, giving us the result. The general case for DnωD_{n}^{\omega} will follow in a very similar way, but we will run this argument for every 2i2^{i}-scaled diamond, rather than all of DnωD_{n}^{\omega} at once. In doing so, we will not just be removing vertices, but instead the entire copy of DiωD_{i}^{\omega} that is attached to the edges of that scaled diamond.

Lemma 3.3 (DnωD_{n}^{\omega} Coloring Lemma).

Let cc\in\mathbb{N} and let DnωD_{n}^{\omega} be the nn-th countably branching diamond graph. Color the elements of VP(Dnω)\textnormal{VP}(D_{n}^{\omega}) by cc colors. Then there exists a 𝔇Dnω\mathfrak{D}\subset D_{n}^{\omega} that is graph isomorphic to DnωD_{n}^{\omega} such that for every 2j2^{j}-scaled D1ω𝔇D_{1}^{\omega}\subset\mathfrak{D} with 0jn10\leq j\leq n-1, D1ωD_{1}^{\omega} is horizontally monochromatic.

Proof.

We will start from the largest scale and work down to the smallest scale. Thus, we first consider D1ωDn1ωD_{1}^{\omega}\varoslash D_{n-1}^{\omega}. This copy of D1ωD_{1}^{\omega} is a 2n12^{n-1}-scaled diamond, so we can use the remarks prior to this lemma to extract a subdiamond 𝔇1D1ω\mathfrak{D}_{1}\subset D_{1}^{\omega} which is horizontally monochromatic. Since this is a countably branching diamond, the ss and tt vertices are not removed, and 𝔇1\mathfrak{D}_{1} is graph isomorphic to D1ωD_{1}^{\omega}, we have that 𝔇1Dn1ω\mathfrak{D}_{1}\varoslash D_{n-1}^{\omega} is graph isomorphic to DnωD_{n}^{\omega}. This handles the largest scale.

From here, enumerate the edges of 𝔇1\mathfrak{D}_{1} and let eiE(𝔇1)e_{i}\in E(\mathfrak{D}_{1}). Consider eiD1ωDn2ωe_{i}\varoslash D_{1}^{\omega}\varoslash D_{n-2}^{\omega}. This is a 2n22^{n-2}-scaled diamond, and following the steps from before, we extract a horizontally monochromatic subdiamond, call it 𝔇2i\mathfrak{D}_{2}^{i}, which is graph isomorphic to D1ωD_{1}^{\omega}. For the same reasons as before, this extraction process preserves the structure of DnωD_{n}^{\omega} and gives us that 𝔇1𝔇2Dn2ω\mathfrak{D}_{1}\varoslash\mathfrak{D}_{2}\varoslash D_{n-2}^{\omega} is graph isomorphic to DnωD_{n}^{\omega}, where 𝔇1𝔇2\mathfrak{D}_{1}\varoslash\mathfrak{D}_{2} is understood to be replacement of every edge in 𝔇1\mathfrak{D}_{1} with its corresponding 𝔇2i\mathfrak{D}_{2}^{i}. Continuing in this way, since there are only nn many scales, this process terminates and leaves us with the desired 𝔇=𝔇1𝔇2𝔇n\mathfrak{D}=\mathfrak{D}_{1}\varoslash\mathfrak{D}_{2}\varoslash...\varoslash\mathfrak{D}_{n}. ∎

3.1.3. Generalized Path Lemma

As we move to the path lemma for the diamonds, there is a need to generalize Matoušek’s original path lemma. Matoušek’s original proof for the path lemma only allowed for subpaths of length 2 and allowed for no control over how the subpath was chosen. The coloring lemma for diamonds only gives horizontally monochromatic 2i2^{i}-scaled diamonds, not the entire graph. As a result, when we look to apply the path lemma, we need to ensure that the subpath selected by the lemma lies along one of these 2i2^{i}-scaled diamonds. Observe that for any given stst-path in DnωD_{n}^{\omega}, s=x0x1x2n=ts=x_{0}\to x_{1}\to...\to x_{2^{n}}=t, if we consider the vertices on this path which are a multiple of 2i2^{i}, this would give us x0,x2i,x22i,,x2ni2ix_{0},x_{2^{i}},x_{2\cdot 2^{i}},...,x_{2^{n-i}\cdot 2^{i}}. Each of these vertices are elements of V(Dni1D1ω)V(D_{n-i-1}\varoslash D_{1}^{\omega}) viewed as a subset of V(Dni1ωD1ωDiω)V(D_{n-i-1}^{\omega}\varoslash D_{1}^{\omega}\varoslash D_{i}^{\omega}). Notice that this is true only because these paths start at ss and end at tt, otherwise, we would not have this guarantee. Thus, for every 1k2ni11\leq k\leq 2^{n-i}-1, we have that the edges (x(k1)2i,xk2i),(xk2i,x(k+1)2i)E(Dni1ωD1ω)(x_{(k-1)2^{i}},x_{k2^{i}}),(x_{k2^{i}},x_{(k+1)2^{i}})\in E(D_{n-i-1}^{\omega}\varoslash D_{1}^{\omega}) are edges of ejD1ωe_{j}\varoslash D_{1}^{\omega} for some ejE(Dni1ω)e_{j}\in E(D_{n-i-1}^{\omega}). Therefore, in the larger context of Dni1D1ωDiωD_{n-i-1}\varoslash D_{1}^{\omega}\varoslash D_{i}^{\omega}, we have that the D1ωD_{1}^{\omega} here is a 2i2^{i}-scaled diamond with s(D1ω)=x(k1)2is(D_{1}^{\omega})=x_{(k-1)2^{i}} and t(D1ω)=x(k+1)2it(D_{1}^{\omega})=x_{(k+1)2^{i}}. The rest of the vertices for this 2i2^{i}-scaled diamond are simply the remaining vertices from ejD1ωe_{j}\varoslash D_{1}^{\omega}.

This gives us a candidate for how the generalized path lemma should go for \varoslash powers of graphs. We will need to be able to extract subpaths whose initial vertex is the ss vertex of some scaled subgraph and whose final vertex is the tt vertex of some scaled subgraph, while the distances between consecutive vertices in the subpath are the same. In the lemma, the Ai+1A_{i+1} subsets perform the role of picking the s,ts,t vertices, while the AiA_{i} subsets are uniformly spaced subsets which collect the other vertices for our scaled subgraph copy. Here, we will only show the result for subpaths of length two, but the methods we use here readily generalize to arbitrary subpath lengths.

Lemma 3.4 (Generalized Path Lemma).

Let p1p\geq 1 and C1C\geq 1. Let ff be a mapping from the path of length n4n\geq 4, Pn={0,1,2,,n}P_{n}=\{0,1,2,...,n\}, with the metric dPn(x,y)=|xy|d_{P_{n}}(x,y)=|x-y| into a metric space (X,dX)(X,d_{X}) such that for some K1K\geq 1 and λ>0\lambda>0, if x,yPnx,y\in P_{n},

λdPn(x,y)dX(f(x),f(y))λKdPn(x,y).\lambda d_{P_{n}}(x,y)\leq d_{X}(f(x),f(y))\leq\lambda Kd_{P_{n}}(x,y).

For i{0,1,,log2n}i\in\{0,1,...,\lfloor\log_{2}n\rfloor\}, let Ai:={0,2i,22i,32i,n2i2i}A_{i}:=\{0,2^{i},2\cdot 2^{i},3\cdot 2^{i}...,\lfloor\frac{n}{2^{i}}\rfloor 2^{i}\}. Define LiL_{i} as a type of scaled Lipschitz norm of ff given by

Li=supx,yAi,dPn(x,y)=2idX(f(x),f(y))λdPn(x,y).L_{i}=\sup_{x,y\in A_{i},d_{P_{n}}(x,y)=2^{i}}\frac{d_{X}(f(x),f(y))}{\lambda d_{P_{n}}(x,y)}.

Then, there exists i{0,1,,log2n}i\in\{0,1,...,\lfloor\log_{2}n\rfloor\} and a subset Z={z0,z1,z2}Z=\{z_{0},z_{1},z_{2}\} of AiA_{i} with
dPn(zj,zj+1)=2id_{P_{n}}(z_{j},z_{j+1})=2^{i} and z0,z2Ai+1z_{0},z_{2}\in A_{i+1} such that for B=Klog2nLi+1B=\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}

λ(1B)Li+1dPn(zj,zk)dX(f(zj),f(zk))λ(1+B)Li+1dPn(zj,zk).\displaystyle\lambda\left(1-B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k})\leq d_{X}(f(z_{j}),f(z_{k}))\leq\lambda\left(1+B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}).

In particular, if n22CKpn\geq 2^{\lceil 2CK^{p}\rceil}, then there is the refined estimate

λ(112CLi+1p)Li+1dPn(zj,zk)dX(f(zj),f(zk))λ(1+12CLi+1p)Li+1dPn(zj,zk),\displaystyle\lambda\left(1-\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}d_{P_{n}}(z_{j},z_{k})\leq d_{X}(f(z_{j}),f(z_{k}))\leq\lambda\left(1+\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}),

with dist(f|Z)1+2CLi+1p\textnormal{dist}(f_{|Z})\leq 1+\frac{2}{CL_{i+1}^{p}}.

Proof.

Observe that KLiLi+11K\geq L_{i}\geq L_{i+1}\geq 1 for all i{0,1,2,,log2n}i\in\{0,1,2,...,\lfloor\log_{2}n\rfloor\}. This can be seen by applying the triangle inequality to Li+1L_{i+1} as in the basic path lemma and utilizing the fact that AiAi+1A_{i}\supset A_{i+1}. Thus, we can now apply the Sequence Lemma to {Li}i=0log2n\{L_{i}\}_{i=0}^{\lfloor\log_{2}n\rfloor} to see that there exists i{0,1,2,,log2n1}i\in\{0,1,2,...,\lfloor\log_{2}n\rfloor-1\} such that

1LiLi+11+Klog2nLi+11\leq\frac{L_{i}}{L_{i+1}}\leq 1+\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}}

Once again, for the sake of space and notation, let B=Klog2nLi+1B=\frac{K}{\lfloor\log_{2}n\rfloor L_{i+1}} for the remainder of the proof.

Now fix z0,z2Ai+1z_{0},z_{2}\in A_{i+1} such that Li+1L_{i+1} is attained on them, i.e., dX(f(z0),f(z2))=λ2i+1Li+1d_{X}(f(z_{0}),f(z_{2}))=\lambda 2^{i+1}L_{i+1}. By the definition of Ai+1A_{i+1}, we also have Z={z0,z1,z2}AiZ=\{z_{0},z_{1},z_{2}\}\subset A_{i} with the property that dPn(zj,zj+1)=2id_{P_{n}}(z_{j},z_{j+1})=2^{i}. Notice as well that dPn(zk,zj)=|kj|2id_{P_{n}}(z_{k},z_{j})=|k-j|2^{i}. From here, we need to demonstrate upper and lower bounds on dX(f(zj),f(zk))d_{X}(f(z_{j}),f(z_{k})) for 0j<k20\leq j<k\leq 2 and use these bounds to achieve a distortion bound on the entire subpath. We begin by getting the upper bound which will be used to show the lower bound.

First, consider for j<kj<k by using the triangle inequality and bounds given by the Sequence Lemma

dX(f(zj),f(zk))\displaystyle d_{X}(f(z_{j}),f(z_{k})) m=jk1dX(f(zm),f(zm+1))\displaystyle\leq\sum_{m=j}^{k-1}d_{X}(f(z_{m}),f(z_{m+1}))
m=jk1λ2iLi\displaystyle\leq\sum_{m=j}^{k-1}\lambda 2^{i}L_{i}
(1) λ(1+B)Li+1(kj)2i\displaystyle\leq\lambda\left(1+B\right)L_{i+1}(k-j)2^{i}
=λ(1+B)Li+1dPn(zj,zk).\displaystyle=\lambda\left(1+B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}).

Once again with j<kj<k, for the lower bound, we can see by the reverse triangle inequality and Section 3.1.3 in the proof of the upper bound

dX(f(zj),f(zk))\displaystyle d_{X}(f(z_{j}),f(z_{k})) dX(f(z0),f(x2))dX(f(z0),f(zj))dX(f(zk),f(z2))\displaystyle\geq d_{X}(f(z_{0}),f(x_{2}))-d_{X}(f(z_{0}),f(z_{j}))-d_{X}(f(z_{k}),f(z_{2}))
=λ2i+1Li+1dX(f(z0),f(zj))dX(f(zk),f(z2))\displaystyle=\lambda 2^{i+1}L_{i+1}-d_{X}(f(z_{0}),f(z_{j}))-d_{X}(f(z_{k}),f(z_{2}))
λ2i+1Li+1λj2i(1+B)Li+1λ(2k)2i(1+B)Li+1\displaystyle\geq\lambda 2^{i+1}L_{i+1}-\lambda j2^{i}(1+B)L_{i+1}-\lambda(2-k)2^{i}(1+B)L_{i+1}
=λ2iLi+1[2j(1+B)(2k)(1+B)]\displaystyle=\lambda 2^{i}L_{i+1}[2-j(1+B)-(2-k)(1+B)]
=λ2iLi+1[(kj)B(2(kj))]\displaystyle=\lambda 2^{i}L_{i+1}[(k-j)-B(2-(k-j))]
=λ(kj)2iLi+1(1B2(kj)kj).\displaystyle=\lambda(k-j)2^{i}L_{i+1}\left(1-B\frac{2-(k-j)}{k-j}\right).

Now, in order to ensure that there is no dependency on jj or kk in 1B2(kj)kj1-B\frac{2-(k-j)}{k-j}, we utilize the fact that for x1x\geq 1, 2xx1\frac{2-x}{x}\leq 1 to get the bound

dX(f(zj),f(zk))\displaystyle d_{X}(f(z_{j}),f(z_{k})) λ(1B)Li+1(kj)2i\displaystyle\geq\lambda\left(1-B\right)L_{i+1}(k-j)2^{i}
=λ(1B)Li+1dPn(zj,zk).\displaystyle=\lambda\left(1-B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}).

This now gives us the first statement of the lemma

λ(1B)Li+1dPn(zj,zk)dX(f(zj),f(zk))λ(1+B)Li+1dPn(zj,zk).\lambda\left(1-B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k})\leq d_{X}(f(z_{j}),f(z_{k}))\leq\lambda\left(1+B\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}).

In addition, if we wish to get a distortion bound, then we must consider

supzj,zkAidX(f(zj),f(zk))λdPn(zj,zk)supzj,zkAiλdPn(zj,zk)dX(f(zj),f(zk)).\sup_{z_{j},z_{k}\in A_{i}}\frac{d_{X}(f(z_{j}),f(z_{k}))}{\lambda d_{P_{n}}(z_{j},z_{k})}\sup_{z_{j},z_{k}\in A_{i}}\frac{\lambda d_{P_{n}}(z_{j},z_{k})}{d_{X}(f(z_{j}),f(z_{k}))}.

By the upper bound, however, we can already achieve a bound for the first supremum,

supzj,zkAidX(f(zj),f(zk))λdPn(zj,zk)(1+B)Li+1.\sup_{z_{j},z_{k}\in A_{i}}\frac{d_{X}(f(z_{j}),f(z_{k}))}{\lambda d_{P_{n}}(z_{j},z_{k})}\leq\left(1+B\right)L_{i+1}.

The bound for the second supremum, however, requires a bit more precision. Notice that the quantity [1B][1-B] is not necessarily positive. It is entirely possible that BB is sufficiently large to cause a contradiction with the supremum being strictly positive by definition of ff, so we must apply constraints to BB in order to proceed. For pedagogical purposes, however, it will be more clear to see these constraints by continuing the proof and assuming a BB that works.

In a similar way to the first supremum, the bound for the second supremum is achieved by using the lower bound we proved earlier,

supzj,zkAiλdPn(zj,zk)dX(f(zj),f(zk))1(1B)Li+1.\sup_{z_{j},z_{k}\in A_{i}}\frac{\lambda d_{P_{n}}(z_{j},z_{k})}{d_{X}(f(z_{j}),f(z_{k}))}\leq\frac{1}{\left(1-B\right)L_{i+1}}.

Putting these bounds together gives us

supzj,zkAidX(f(zj),f(zk))λdPn(zj,zk)supzj,zkAiλdPn(zj,zk)dX(f(zj),f(zk))1+B1B.\sup_{z_{j},z_{k}\in A_{i}}\frac{d_{X}(f(z_{j}),f(z_{k}))}{\lambda d_{P_{n}}(z_{j},z_{k})}\sup_{z_{j},z_{k}\in A_{i}}\frac{\lambda d_{P_{n}}(z_{j},z_{k})}{d_{X}(f(z_{j}),f(z_{k}))}\leq\frac{1+B}{1-B}.

Just as with the original Path Lemma, we want the distortion to be of the form 1+δ1+\delta for some δ>0\delta>0. We can achieve this by recognizing that for 0x120\leq x\leq\frac{1}{2}, 1+x1x1+4x\frac{1+x}{1-x}\leq 1+4x, so we see

1+B1B1+4B.\frac{1+B}{1-B}\leq 1+4B.

Since the terms of BB are all positive if n2n\geq 2, the only interesting constraint is forcing B12B\leq\frac{1}{2}. In this case, we can exercise control over BB by adding a requirement to the length nn of our path PnP_{n}, and this constraint is achieved exactly when n22CKpn\geq 2^{\lceil 2CK^{p}\rceil} for any C,p1C,p\geq 1. In the case where we have this lower bound for nn, then our upper and lower bounds from earlier are improved to

λ(112CLi+1p)Li+1dPn(zj,zk)dX(f(zj),f(zk))λ(1+12CLi+1p)Li+1dPn(zj,zk).\displaystyle\lambda\left(1-\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}d_{P_{n}}(z_{j},z_{k})\leq d_{X}(f(z_{j}),f(z_{k}))\leq\lambda\left(1+\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}d_{P_{n}}(z_{j},z_{k}).

Using our distortion estimates from above in terms of BB, we now have the distortion estimate stated in the lemma,

dist(f|Z)1+4Klog2nLi+11+2CLi+1p1+2CLi+1p\textnormal{dist}(f_{|Z})\leq 1+\frac{4K}{\lfloor\log_{2}n\rfloor L_{i+1}}\leq 1+\frac{2}{CL_{i+1}^{p}}\leq 1+\frac{2}{CL_{i+1}^{p}}

3.2. Embedding Obstruction Theorem

Theorem 3.5 (Lower Bound for Embedding of Countably Branching Diamonds).

Let XX be a Banach space which is pp-AMUC for some p>1p>1 and CM>0C_{M}>0. Then the minimum distortion necessary for embedding DnωD_{n}^{\omega} for sufficiently large nn is at least Cn1/pCn^{1/p} for some C>0C>0, dependent only on CMC_{M} and pp.

Proof.

Let nn be sufficiently large and suppose that DnωD_{n}^{\omega} embeds into XX with Lipschitz map f:DnωXf:D_{n}^{\omega}\to X such that for some λ>0\lambda>0 and for every x,yDnωx,y\in D_{n}^{\omega},

λdDnω(x,y)dX(f(x),f(y))λKdDnω(x,y),\lambda d_{D_{n}^{\omega}}(x,y)\leq d_{X}(f(x),f(y))\leq\lambda Kd_{D_{n}^{\omega}}(x,y),

where dX(x,y)=xyd_{X}(x,y)=\|x-y\|. We choose this notation instead of the norm to highlight the lack of dependency on the linear structure of our space and to facilitate the move to metric spaces in later parts. Observe that by rescaling the metric, we may assume that λ=1\lambda=1.

Just as before, this proof will have four main parts. First, using the diamond coloring lemma (Lemma 3.3), we will show that there exists a 𝔇Dnω\mathfrak{D}\subset D_{n}^{\omega} which is graphically isomorphic to DnωD_{n}^{\omega} such that every 2i2^{i}-scaled D1ωD_{1}^{\omega} in 𝔇\mathfrak{D} is horizontally monochromatic under coloring by the log distortion of pairs of vertices in VP(Dnω)\textnormal{VP}(D_{n}^{\omega}). Second, we will utilize the generalized path lemma (Lemma 3.4) to extract a well-embedded subpath. Third, we will leverage our coloring on 𝔇\mathfrak{D} and the well embedded subpath to construct a vertical ϵ\epsilon-diamond in XX. Finally, we will then apply our vertical ϵ\epsilon-diamond lemma (Lemma 3.2) to realize a contradiction to the conditions necessary for the construction of the vertical ϵ\epsilon-diamond to get our result.

Let γ>0\gamma>0 be a parameter to be chosen later. Set r=log1+γKr=\lceil\log_{1+\gamma}K\rceil and give {x,y}VP(Dnω)\{x,y\}\in\textnormal{VP}(D_{n}^{\omega}) the color

log1+γ(dX(f(x),f(y))dDnω(x,y)){0,1,,r}.\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(x),f(y))}{d_{D_{n}^{\omega}}(x,y)}\right)\right\rfloor\in\{0,1,...,r\}.

By the diamond coloring lemma (Lemma 3.3), there exists 𝔇Dnω\mathfrak{D}\subset D_{n}^{\omega} whose vertically faithful 2i2^{i}-scaled copies of D1ωD_{1}^{\omega} are monochromatic.

Now, we pick an arbitrary stst-path in 𝔇\mathfrak{D} and apply the generalized path lemma with the sets Ai={0,2i,22i,32i,,2n}A_{i}=\{0,2^{i},2\cdot 2^{i},3\cdot 2^{i},...,2^{n}\} for 0in0\leq i\leq n. This gives us, for some ii, a set of vertices Z={sD1ω,x1,tD1ω}Z=\{s_{D_{1}^{\omega}},x_{1},t_{D_{1}^{\omega}}\} that correspond to the ss, tt, and midpoint vertex of a 2i2^{i}-scaled copy of D1ωD_{1}^{\omega} in 𝔇\mathfrak{D} with the property that for x,yZx,y\in Z

Li+1(1KnLi+1)dDnω(x,y)\displaystyle L_{i+1}\left(1-\frac{K}{nL_{i+1}}\right)d_{D_{n}^{\omega}}(x,y) dX(f(x),f(y)))Li+1(1+KnLi+1)dDnω(x,y).\displaystyle\leq d_{X}(f(x),f(y)))\leq L_{i+1}\left(1+\frac{K}{nL_{i+1}}\right)d_{D_{n}^{\omega}}(x,y).

(Notice that since the length of the stst-path is 2n2^{n}, the log\log term simplifies to just an nn.)

From here, we wish to create a vertical ϵ\epsilon-diamond in XX for choices of ϵ\epsilon, γ\gamma, and θ\theta. To this end, we have, as before, the following claim.

Claim.

If

γ=2CKp,θ=Li+1(112CLi+1p)2i1+2CKp,ϵ=14CLi+1p,\displaystyle\gamma=\frac{2}{CK^{p}},\quad\theta=\frac{L_{i+1}(1-\frac{1}{2CL_{i+1}^{p}})2^{i}}{1+\frac{2}{CK^{p}}},\quad\epsilon=\frac{14}{CL_{i+1}^{p}},

and 2n24CKp22CKp2^{n}\geq 2^{4CK^{p}}\geq 2^{\lceil 2CK^{p}\rceil} where C=(56)(64)pCMC=\frac{(56)(64)^{p}}{C_{M}} if CM28(64)pC_{M}\leq 28(64)^{p} or C=(56)(64)pC=(56)(64)^{p} otherwise, then there exists x2,x3,𝔇x_{2},x_{3},...\in\mathfrak{D} such that {s,t,x1,x2,}\{s,t,x_{1},x_{2},...\} is a 2i2^{i}-scaled copy of D1ωD_{1}^{\omega} whose image in XX is a vertical ϵ\epsilon-diamond.

Assuming the claim, we show how to proceed. If we have that 2n<24CKp2^{n}<2^{4CK^{p}}, then K>(n4C)1/pK>\left(\frac{n}{4C}\right)^{1/p} and we are done. Otherwise, 2n24CKp2^{n}\geq 2^{4CK^{p}}, so we can use the claim. This gives us a vertical ϵ\epsilon-diamond, so we can apply the ϵ\epsilon-diamond lemma while utilizing that dDnω(xi,xj)=2i+1d_{D_{n}^{\omega}}(x_{i},x_{j})=2^{i+1} because the xi,xjx_{i},x_{j} are in a 2i2^{i}-scaled copy of D1ωD_{1}^{\omega} to get for CM28(64)pC_{M}\leq 28(64)^{p}

2i+1infijdX(f(xi),f(xj))64(4)1/pCM1/pϵ1/pθ=64(4)1/pCM1/pCM1/p64(4)1/pLi+1θ=θLi+12i.\displaystyle 2^{i+1}\leq\inf_{i\neq j}d_{X}(f(x_{i}),f(x_{j}))\leq\frac{64(4)^{1/p}}{C_{M}^{1/p}}\epsilon^{1/p}\theta=\frac{64(4)^{1/p}}{C_{M}^{1/p}}\frac{C_{M}^{1/p}}{64(4)^{1/p}L_{i+1}}\theta=\frac{\theta}{L_{i+1}}\leq 2^{i}.

Similarly, if CM>28(64)pC_{M}>28(64)^{p}, then we disregard it in the constant CC and we remove it from the chain of inequalities above, making everything bigger, but maintaining the contradiction. ∎

Proof of claim.

Just as before, we will only give values for constants at the very end. For now, if 2n22CKp2^{n}\geq 2^{\lceil 2CK^{p}\rceil} for some C1C\geq 1 to be determined later and p>1p>1, then the generalized path lemma improves to give us

(112CLi+1p)Li+12i\displaystyle\left(1-\frac{1}{2CL^{p}_{i+1}}\right)L_{i+1}2^{i} dX(f(sD1ω),f(x1))(1+12CLi+1p)Li+12i\displaystyle\leq d_{X}(f(s_{D_{1}^{\omega}}),f(x_{1}))\leq\left(1+\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}2^{i}
(112CLi+1p)Li+12i\displaystyle\left(1-\frac{1}{2CL^{p}_{i+1}}\right)L_{i+1}2^{i} dX(f(x1),f(tD1ω))(1+12CLi+1p)Li+12i\displaystyle\leq d_{X}(f(x_{1}),f(t_{D_{1}^{\omega}}))\leq\left(1+\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}2^{i}
(112CLi+1p)Li+12i+1\displaystyle\left(1-\frac{1}{2CL^{p}_{i+1}}\right)L_{i+1}2^{i+1} dX(f(sD1ω),f(tD1ω))(1+12CLi+1p)Li+12i+1.\displaystyle\leq d_{X}(f(s_{D_{1}^{\omega}}),f(t_{D_{1}^{\omega}}))\leq\left(1+\frac{1}{2CL_{i+1}^{p}}\right)L_{i+1}2^{i+1}.

We will only work with the vertices sD1ωs_{D_{1}^{\omega}} and x1x_{1}, but the argument holds for the other vertices as well. By the horizontally monochromatic property of our 2i2^{i}-scaled copy of D1ωD_{1}^{\omega}, we have for every j2j\geq 2, the existence of xjx_{j} vertices such that

log1+γ(dX(f(sD1ω),f(x1))2i)=log1+γ(dX(f(sD1ω),f(xj))2i).\displaystyle\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(s_{D_{1}^{\omega}}),f(x_{1}))}{2^{i}}\right)\right\rfloor=\left\lfloor\log_{1+\gamma}\left(\frac{d_{X}(f(s_{D_{1}^{\omega}}),f(x_{j}))}{2^{i}}\right)\right\rfloor.

Applying the same logic as the tree case, this eventually gives us by utilizing the bound given by the generalized path lemma and the properties of the logarithm

Li+1(112CLi+1p)1+γ2idX(f(sD1ω),f(xj))(1+γ)Li+1(1+12CLi+1p)2i.\displaystyle\frac{L_{i+1}\left(1-\frac{1}{2CL_{i+1}^{p}}\right)}{1+\gamma}2^{i}\leq d_{X}(f(s_{D_{1}^{\omega}}),f(x_{j}))\leq(1+\gamma)L_{i+1}\left(1+\frac{1}{2CL_{i+1}^{p}}\right)2^{i}.

Thus, moving from one midpoint of the diamond to another only incurs a 1+γ1+\gamma penalty. We now wish to rewrite this inequality in a form that is conducive to creating a vertical ϵ\epsilon-diamond. Here, we choose

θ=Li+1(112CLi+1p)2i1+γ.\theta=\frac{L_{i+1}(1-\frac{1}{2CL_{i+1}^{p}})2^{i}}{1+\gamma}.

This allows us, after utilizing that 1+x1x1+4x\frac{1+x}{1-x}\leq 1+4x for 0x120\leq x\leq\frac{1}{2}, to rewrite this as

θdX(f(sD1ω),f(xj))θ(1+γ)2(1+2CLi+1p).\theta\leq d_{X}(f(s_{D_{1}^{\omega}}),f(x_{j}))\leq\theta(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right).

Now, our distortion is (1+γ)2(1+2CLi+1p)(1+\gamma)^{2}\left(1+\frac{2}{CL_{i+1}^{p}}\right), but we need something of the form 1+ϵ1+\epsilon. To this end, we begin by choosing γ=2CKp2CLi+1p\gamma=\frac{2}{CK^{p}}\leq\frac{2}{CL_{i+1}^{p}} and want to utilize the fact that (1+x)31+7x(1+x)^{3}\leq 1+7x for 0x10\leq x\leq 1. This motivates our choice of CC which we will choose so that 2CLi+1p1\frac{2}{CL_{i+1}^{p}}\leq 1. Notice that because Li+1p1L_{i+1}^{p}\geq 1, our CC is independent of all choices made thus far. As a result, any C2C\geq 2 will suffice, however, we also want to cancel with the constants in the ϵ\epsilon-diamond lemma, so we will choose C=56(64)pCMC=\frac{56(64)^{p}}{C_{M}} if CM28(64)pC_{M}\leq 28(64)^{p}. If not, then we simply omit CM1/pC_{M}^{1/p}. With this in mind, we then have our final estimate for every ii\in\mathbb{N}

θ\displaystyle\theta dX(f(sD1ω),f(xi))θ(1+14CLi+1p)\displaystyle\leq d_{X}(f(s_{D_{1}^{\omega}}),f(x_{i}))\leq\theta\left(1+\frac{14}{CL_{i+1}^{p}}\right)
θ\displaystyle\theta dX(f(xi),f(tD1ω))θ(1+14CLi+1p)\displaystyle\leq d_{X}(f(x_{i}),f(t_{D_{1}^{\omega}}))\leq\theta\left(1+\frac{14}{CL_{i+1}^{p}}\right)
2θ\displaystyle 2\theta dX(f(sD1ω),f(tD1ω))2θ(1+14CLi+1p).\displaystyle\leq d_{X}(f(s_{D_{1}^{\omega}}),f(t_{D_{1}^{\omega}}))\leq 2\theta\left(1+\frac{14}{CL_{i+1}^{p}}\right).

Notice that we once again do not utilize any Banach structure in the final proof of the embedding obstruction theorem. As a result, we can redo the vertical ϵ\epsilon-diamond lemma only assuming a metric inequality.

Lemma 3.6 (Metric Inequality ϵ\epsilon-diamond).

Let (X,d)(X,d) be a metric space such that for every s,t,x1,x2,Xs,t,x_{1},x_{2},...\in X we have

(2) d(s,t)p2p+infijd(xi,xj)pCDpmax(supid(s,xi)p,supid(t,xi)p).\frac{d(s,t)^{p}}{2^{p}}+\frac{\inf_{i\neq j}d(x_{i},x_{j})^{p}}{C_{D}^{p}}\leq\max\left(\sup_{i}d(s,x_{i})^{p},\sup_{i}d(t,x_{i})^{p}\right).

for some p1p\geq 1 and CD>0C_{D}>0. Let F={s,t,x1,}F=\{s,t,x_{1},...\} be a vertical ϵ\epsilon-diamond for some ϵ(0,1)\epsilon\in(0,1) and θ>0\theta>0, then

infijd(xi,xj)p6CDθϵ1/p.\inf_{i\neq j}d(x_{i},x_{j})^{p}\leq 6C_{D}\theta\epsilon^{1/p}.
Proof.

By leveraging the definition of a ϵ\epsilon-diamond, we have

infijd(xi,xj)pCDp\displaystyle\frac{\inf_{i\neq j}d(x_{i},x_{j})^{p}}{C_{D}^{p}} (1+ϵ)pθpd(s,t)p2p(1+ϵ)pθpθp,\displaystyle\leq(1+\epsilon)^{p}\theta^{p}-\frac{d(s,t)^{p}}{2^{p}}\leq(1+\epsilon)^{p}\theta^{p}-\theta^{p},

which implies

infijd(xi,xj)CDθ((1+ϵ)p1)1/p.\displaystyle\inf_{i\neq j}d(x_{i},x_{j})\leq C_{D}\theta\left((1+\epsilon)^{p}-1\right)^{1/p}.

Now utilizing the same approximation argument as in the proof of the δ\delta-umbel lemma (Lemma 2.2), we have the result. ∎

Inequalities such as (2) are studied in [Bau] where it is shown in particular that a reflexive and pp-AUC Banach space satisfies (2).

With this final piece completed, we can replace the pp-AMUC version of the vertical ϵ\epsilon-diamond lemma with the metric version in the proof of the pp-AMUC embeddability obstruction and observe that (2) is preserved by rescalings of the metric. This, then, gives us the following theorem.

Theorem 3.7 (Metric Diamond Inequality Embedding Obstruction).

Let (X,d)(X,d) be a metric space satisfying (2) for some p1p\geq 1 and CD>0C_{D}>0. Then, we have for sufficiently large nn, DnωD_{n}^{\omega} embeds into XX with distortion at least Cn1/pCn^{1/p} where CC is only dependent on CDC_{D} and pp.

4. Coarse Embeddings of Countably Branching Trees

In this section, we will work with the other side of tree embeddings, finding examples of embeddings into spaces. Previously, we worked with bi-Lipschitz embeddings, but here, we will be dealing with coarse embeddings. This category of embeddings glosses over much of the local geometry of spaces and instead focuses on large scale geometric trends. Rigorously, to define a coarse embedding, we must consider for two metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) and a map F:XYF:X\to Y the compression, ρF(t)\rho_{F}(t), of FF given by

ρF(t)=infdX(x,y)tdY(F(x),F(y)),\rho_{F}(t)=\inf_{d_{X}(x,y)\geq t}d_{Y}(F(x),F(y)),

and the expansion, ωF(t)\omega_{F}(t), given by

ωF(t)=supdX(x,y)tdY(F(x),F(y)).\omega_{F}(t)=\sup_{d_{X}(x,y)\leq t}d_{Y}(F(x),F(y)).

We say FF is a coarse embedding if ωF(t)\omega_{F}(t) is finite for all t>0t>0 and limtρF(t)=\lim_{t\to\infty}\rho_{F}(t)=\infty. Similar to the way we defined equi-bi-Lipschitz embeddings, we can also define equi-coarse embeddings for a family of metric spaces {Xi}i=1\{X_{i}\}_{i=1}^{\infty} if there exists ρ,ω:[0,)[0,)\rho,\omega:[0,\infty)\to[0,\infty) and maps fi:XiYf_{i}:X_{i}\to Y such that ρρfi\rho\leq\rho_{f_{i}} and ωfiω\omega_{f_{i}}\leq\omega with limtρ(t)=\lim_{t\to\infty}\rho(t)=\infty and ω(t)<\omega(t)<\infty for all t>0t>0. Notice that in both definitions we have no indication about the Lipschitz-ness or continuity of FF; coarse embeddings are more general than bi-Lipschitz embeddings. These types of embeddings were first introduced by Gromov [Gro93].

In [Tes11], it was shown that any tree TT is coarsely embeddable into p(T)\ell_{p}(T) for 1p<1\leq p<\infty by a map FF with compression bounded below by any function f:[0,)(0,)f:[0,\infty)\to(0,\infty) such that 1f(t)tp1t𝑑t<\int_{1}^{\infty}\frac{f(t)}{t}^{p}\frac{1}{t}dt<\infty. This result relied heavily on the properties of unit basis in p\ell_{p}, namely that the unit basis is 1-suppression unconditional. This condition can be weakened, however, at the expense of moving from coarse embeddings of any tree to equi-coarse embeddings of the countably branching trees of depth kk.

Theorem 4.1 (Compression Theorem).

Let {[]k}k=1\{[\mathbb{\mathbb{N}}]^{\leq k}\}_{k=1}^{\infty} be the countably branching trees of depth kk\in\mathbb{N} equipped with the standard tree metric. Let XX be a Banach space with an p\ell^{p}-asymptotic model for 1p<1\leq p<\infty generated by a weakly null array. Then, for every increasing function f:[0,)[0,)f:[0,\infty)\to[0,\infty) satisfying 1(f(t)t)p1t𝑑t<\int_{1}^{\infty}(\frac{f(t)}{t})^{p}\frac{1}{t}dt<\infty and for every kk\in\mathbb{N}, there exist Fk:[]kXF_{k}:[\mathbb{\mathbb{N}}]^{\leq k}\to X with uniformly bounded Lipschitz norms such that ρFk(t)f(t/8)\rho_{F_{k}}(t)\succeq f(t/8) for all t3t\geq 3.

Here, the condition that we are embedding into p(T)\ell_{p}(T) has been replaced with a Banach space XX containing an p\ell_{p}-asymptotic model, a generalization of p\ell_{p}-spreading models that was first defined in [HO04]. Various properties about p\ell_{p}-asymptotic models were proven in [Bau+21], but here we will only give their definition and restate the relevant lemma from [Bau+21, Lemma 3.8] needed for the theorem.

Definition 4.0.1 (p\ell_{p}-Asymptotic Model).

Given a Banach space XX, a basic sequence (ei)i=1(e_{i})_{i=1}^{\infty} is an p\ell_{p}-asymptotic model of XX for some p1p\geq 1 if there exists (xj(i):i,j)SX\left(x_{j}^{(i)}:i,j\in\mathbb{N}\right)\subset S_{X}, the unit sphere of XX, and a null-sequence of scalars (ϵn)n=1(0,1)(\epsilon_{n})_{n=1}^{\infty}\subset(0,1) such that for all nn\in\mathbb{N} and all {ai}i=1n[1,1]\{a_{i}\}_{i=1}^{n}\subset[-1,1] and nk1<<knn\leq k_{1}<...<k_{n}, then

|||i=1naixki(i)||(i=1n|ai|p)1/p|<ϵn.\Bigg{\lvert}\Bigg{\lvert}\Bigg{\lvert}\sum_{i=1}^{n}a_{i}x_{k_{i}}^{(i)}\Bigg{\rvert}\Bigg{\rvert}-\left(\sum_{i=1}^{n}|a_{i}|^{p}\right)^{1/p}\Bigg{\rvert}<\epsilon_{n}.
Lemma 4.2 ((1+ϵ)(1+\epsilon)-suppression unconditional finite subsequences in p\ell_{p}-asymptotic model).

Let XX be a Banach space and (xj(i):1ik,j)(x_{j}^{(i)}:1\leq i\leq k,j\in\mathbb{N}) be a normalized weakly null array of height kk. Then for every ϵ>0\epsilon>0 and mm\in\mathbb{N}, there exists 𝕄[]ω\mathbb{M}\in[\mathbb{N}]^{\omega} so that for every i1,,im{1,,k}i_{1},...,i_{m}\in\{1,...,k\} (not necessarily distinct) and pairwise distinct l1,,lm𝕄l_{1},...,l_{m}\in\mathbb{M}, (xlj(ij))j=1m(x_{l_{j}}^{(i_{j})})_{j=1}^{m} is (1+ϵ)(1+\epsilon)-suppression unconditional.

This lemma is of vital importance as it allows us to extract exactly the subsequence of the infinite array which has the suppression unconditionality that is necessary for the lower bound of our given embedding. Below we present the proof of Theorem 4.1.

Proof.

Following the example in [BG21], let kk\in\mathbb{N} and fix a bijection Φ:[]k{2k,2k+1,}\Phi:[\mathbb{\mathbb{N}}]^{\leq k}\to\{2k,2k+1,...\} such that Φ((n1,n2,,nl))Φ((n1,n2,,nl+1))\Phi((n_{1},n_{2},...,n_{l}))\leq\Phi((n_{1},n_{2},...,n_{l+1})) for lk1l\leq k-1 and Φ()=r\Phi(\emptyset)=r where rr is the root of the tree. Let S=maxi{[(f(i)i)p1i],1}S=\max_{i\in\mathbb{N}}\left\{\left[\left(\frac{f(i)}{i}\right)^{p}\frac{1}{i}\right],1\right\} which is finite by definition of ff. Fix ϵ>0\epsilon>0. By Lemma 4.2, for every i1,i2,,i2ki_{1},i_{2},...,i_{2k} in 1,,k{1,...,k}, not necessarily distinct, and any pairwise different l1,,l2kl_{1},...,l_{2k}\in\mathbb{N}, we have that the sequence (xljij)j=12k(x_{l_{j}}^{i_{j}})_{j=1}^{2k} is (1+ϵ)(1+\epsilon)-suppression unconditional. This enables us to assume, in addition, by taking a subarray if necessary, that for every (aj)12k[S,S](a_{j})_{1}^{2k}\in[-S,S], we have

|||j=12kajxljij||(j=12kajp)1/p|ϵ.\displaystyle\Bigg{\lvert}\Bigg{\lvert}\Bigg{\lvert}\sum_{j=1}^{2k}a_{j}x_{l_{j}}^{i_{j}}\Bigg{\rvert}\Bigg{\rvert}-\left(\sum_{j=1}^{2k}a_{j}^{p}\right)^{1/p}\Bigg{\rvert}\leq\epsilon.

With this in mind, we are now in position to define our map Fk:[]kXF_{k}:[\mathbb{\mathbb{N}}]^{\leq k}\to X which we will show is Lipschitz and bounded below. Let (ξi)i=0(\xi_{i})_{i=0}^{\infty} be a non-negative sequence such that i=1|ξiξi1|p<\sum_{i=1}^{\infty}|\xi_{i}-\xi_{i-1}|^{p}<\infty whose exact definition in terms of ff will be given later. Let Fk((n1,,nl))=i=0lξlixΦ((n1,,ni))iF_{k}((n_{1},...,n_{l}))=\sum_{i=0}^{l}\xi_{l-i}x_{\Phi((n_{1},...,n_{i}))}^{i} where if i=0i=0, we have Φ()=r\Phi(\emptyset)=r. Notice that since []k[\mathbb{\mathbb{N}}]^{\leq k} is quasi-geodesic, we need only check that FkF_{k} is Lipschitz on neighbor elements of the tree. Let x=(n1,,nl)x=(n_{1},...,n_{l}) and y=(n1,,nl,nl+1)y=(n_{1},...,n_{l},n_{l+1}). Then we see that

Fk(x)Fk(y)\displaystyle\|F_{k}(x)-F_{k}(y)\| =||i=0lξlixΦ((n1,,ni))ii=0l+1ξl+1ixΦ((n1,,ni))i||\displaystyle=\Bigg{\lvert}\Bigg{\lvert}\sum_{i=0}^{l}\xi_{l-i}x^{i}_{\Phi((n_{1},...,n_{i}))}-\sum_{i=0}^{l+1}\xi_{l+1-i}x^{i}_{\Phi((n_{1},...,n_{i}))}\Bigg{\rvert}\Bigg{\rvert}
=||i=0l(ξl+1iξli)xΦ((n1,,ni))i+ξ0xΦ((n1,..,nl+1))l+1||\displaystyle=\Bigg{\lvert}\Bigg{\lvert}\sum_{i=0}^{l}(\xi_{l+1-i}-\xi_{l-i})x^{i}_{\Phi((n_{1},...,n_{i}))}+\xi_{0}x_{\Phi((n_{1},..,n_{l+1}))}^{l+1}\Bigg{\rvert}\Bigg{\rvert}
(|ξ0|p+i=0l|ξl+1iξli|p)1/p+ϵ\displaystyle\leq\left(|\xi_{0}|^{p}+\sum_{i=0}^{l}|\xi_{l+1-i}-\xi_{l-i}|^{p}\right)^{1/p}+\epsilon
=(|ξ0|p+i=1l+1|ξiξi1|p)1/p+ϵ\displaystyle=\left(|\xi_{0}|^{p}+\sum_{i=1}^{l+1}|\xi_{i}-\xi_{i-1}|^{p}\right)^{1/p}+\epsilon

where the second to last step uses the p\ell^{p}-asymptotic property of the sequence. Taking the limit as ll goes to infinity, we see that the boundedness property guarantees that this map is Lipschitz regardless of the depth of the tree chosen, i.e., there exists an upper bound for the Lipschitz constant which is completely independent of kk.

From here, we need to prove lower boundedness. Let x=(u¯,x1,,xn)x=(\bar{u},x_{1},...,x_{n}) and let y=(u¯,y1,y2,,ym)y=(\bar{u},y_{1},y_{2},...,y_{m}) be elements of []k[\mathbb{\mathbb{N}}]^{\leq k} such that u¯\bar{u} is the first element both xx and yy have in common on their paths back to the root rr. This element may be the root itself, but we only need to have that if u¯=(u1,,us)\bar{u}=(u_{1},...,u_{s}), then max(s+n,s+m)k\max(s+n,s+m)\leq k so as to stay in []k[\mathbb{\mathbb{N}}]^{\leq k}. With this in mind, we observe

Fk(x)Fk(y)=||\displaystyle\|F_{k}(x)-F_{k}(y)\|=\Bigg{\lvert}\Bigg{\lvert} i=0s(ξs+niξs+mi)xΦ((u1,,ui))i+i=1nξnixΦ((u¯,x1,,xi))s+i\displaystyle\sum_{i=0}^{s}(\xi_{s+n-i}-\xi_{s+m-i})x_{\Phi((u_{1},...,u_{i}))}^{i}+\sum_{i=1}^{n}\xi_{n-i}x_{\Phi((\bar{u},x_{1},...,x_{i}))}^{s+i}-
i=1mξmixΦ((u¯,y1,,yi))s+i||\displaystyle\sum_{i=1}^{m}\xi_{m-i}x_{\Phi((\bar{u},y_{1},...,y_{i}))}^{s+i}\Bigg{\rvert}\Bigg{\rvert}
11+ϵ||i=1nξnixΦ((u¯,x1,,xi))s+i||\displaystyle\geq\frac{1}{1+\epsilon}\Bigg{\lvert}\Bigg{\lvert}\sum_{i=1}^{n}\xi_{n-i}x_{\Phi((\bar{u},x_{1},...,x_{i}))}^{s+i}\Bigg{\rvert}\Bigg{\rvert}
11+ϵ(i=1n(ξni)p)1/pϵ1+ϵ\displaystyle\geq\frac{1}{1+\epsilon}\left(\sum_{i=1}^{n}(\xi_{n-i})^{p}\right)^{1/p}-\frac{\epsilon}{1+\epsilon}
=11+ϵ(i=0n1(ξi)p)1/pϵ1+ϵ\displaystyle=\frac{1}{1+\epsilon}\left(\sum_{i=0}^{n-1}(\xi_{i})^{p}\right)^{1/p}-\frac{\epsilon}{1+\epsilon}

Similarly, we can apply this suppression and asymptotic model argument to the last third of the sum and for ϵ1/2\epsilon\leq 1/2, get the combined lower bound given by

Fk(x)Fk(y)13[(i=0n1(ξi)p)1/p+(i=0m1(ξi)p)1/p]ϵ1+ϵ.\|F_{k}(x)-F_{k}(y)\|\geq\frac{1}{3}\left[\left(\sum_{i=0}^{n-1}(\xi_{i})^{p}\right)^{1/p}+\left(\sum_{i=0}^{m-1}(\xi_{i})^{p}\right)^{1/p}\right]-\frac{\epsilon}{1+\epsilon}.

Notice that this is a bound for the compression in terms of the p\ell^{p}-sums of the (ξi)(\xi_{i}) sequence, i.e., ρFk(m+n)Ωϵ((i=0M1ξip)1/p)\rho_{F_{k}}(m+n)\geq\Omega_{\epsilon}((\sum_{i=0}^{M-1}\xi_{i}^{p})^{1/p}) where M=max(m,n)M=\max(m,n). Ultimately, we will desire a lower bound in terms of d(x,y)d(x,y), but converting from MM to d(x,y)d(x,y) can be done by observing that max(m,n)d(x,y)/2\max(m,n)\geq d(x,y)/2.

From here, we can now demonstrate the dependency on ff. Notice that up until this point, the only requirement for the ξi\xi_{i}’s has been that the p\ell^{p}-sum of their sequential differences is finite. If we choose ξ0=f(0)\xi_{0}=f(0) and ξiξi1=1i1/pf(i)i\xi_{i}-\xi_{i-1}=\frac{1}{i^{1/p}}\frac{f(i)}{i} for i1i\geq 1, then we can first observe that the properties of ff guarantee that i=0|ξiξi1|p<\sum_{i=0}^{\infty}|\xi_{i}-\xi_{i-1}|^{p}<\infty, if ξ1\xi_{-1} is defined as zero. Next, for the technical portion of this proof, we will need some kind of rounding function to ensure integer indices in our summations, since we will be frequently dividing these integers in half. As a result, if A={m2|m}A=\{\frac{m}{2}\ |\ m\in\mathbb{Z}\}, then our rounding function, R:AR:\mathbb{Z}\cup A\to\mathbb{Z}, is defined as R(x)=xR(x)=x if xx\in\mathbb{Z} and R(x)=x+1/2R(x)=x+1/2 if xAx\in A, essentially rounding up whenever xx is a fraction. Finally, since we are only searching for a bound on the distances greater than or equal to three, we have M2M\geq 2. This is important as it prevents the quantity MR(M/2)M-R(M/2) from zeroing out and trivializing our lower bound.

We now follow the example of [Tes11] for these calculations,

Fk(x)Fk(y)\displaystyle\|F_{k}(x)-F_{k}(y)\| 13(i=0M1ξip)1/pϵ1+ϵ\displaystyle\geq\frac{1}{3}\left(\sum_{i=0}^{M-1}\xi_{i}^{p}\right)^{1/p}-\frac{\epsilon}{1+\epsilon}
13[i=R(M/2)M1(j=0iξjξj1)p]1/pϵ1+ϵ\displaystyle\geq\frac{1}{3}\left[\sum_{i=R(M/2)}^{M-1}\left(\sum_{j=0}^{i}\xi_{j}-\xi_{j-1}\right)^{p}\right]^{1/p}-\frac{\epsilon}{1+\epsilon}
13[M1R(M/2)+12(j=0R(M/2)ξjξj1)p]1/pϵ1+ϵ\displaystyle\geq\frac{1}{3}\left[\frac{M-1-R(M/2)+1}{2}\left(\sum_{j=0}^{R(M/2)}\xi_{j}-\xi_{j-1}\right)^{p}\right]^{1/p}-\frac{\epsilon}{1+\epsilon}
13(MR(M/2)2)1/p(j=0R(M/2)ξjξj1)ϵ1+ϵ\displaystyle\geq\frac{1}{3}\left(\frac{M-R(M/2)}{2}\right)^{1/p}\left(\sum_{j=0}^{R(M/2)}\xi_{j}-\xi_{j-1}\right)-\frac{\epsilon}{1+\epsilon}
13(4)1/p(M1)1/p(j=R(R(M/2)/2)R(M/2)1j1/pf(j)j)ϵ1+ϵ\displaystyle\geq\frac{1}{3(4)^{1/p}}\left(M-1\right)^{1/p}\left(\sum_{j=R(R(M/2)/2)}^{R(M/2)}\frac{1}{j^{1/p}}\frac{f(j)}{j}\right)-\frac{\epsilon}{1+\epsilon}
f(M/4)3(4)1/p(M1)1/p(j=R(R(M/2)/2)R(M/2)1j1p+1)ϵ1+ϵ.\displaystyle\geq\frac{f(M/4)}{3(4)^{1/p}}\left(M-1\right)^{1/p}\left(\sum_{j=R(R(M/2)/2)}^{R(M/2)}\frac{1}{j^{\frac{1}{p}+1}}\right)-\frac{\epsilon}{1+\epsilon}.

From here, let us process the sum separately from the rest of the chain of inequalities, then substitute it back in later. This gives us

j=R(R(M/2)/2)R(M/2)1j1p+1\displaystyle\sum_{j=R(R(M/2)/2)}^{R(M/2)}\frac{1}{j^{\frac{1}{p}+1}} (R(M/2)R(R(M/2)/2)+12)(1(R(M/2))1p+1)\displaystyle\geq\left(\frac{R(M/2)-R(R(M/2)/2)+1}{2}\right)\left(\frac{1}{(R(M/2))^{\frac{1}{p}+1}}\right)
(R(M/2)+14)1(R(M/2))1p+1\displaystyle\geq\left(\frac{R(M/2)+1}{4}\right)\frac{1}{\left(R(M/2)\right)^{\frac{1}{p}+1}}
14(R(M/2))1p\displaystyle\geq\frac{1}{4(R(M/2))^{\frac{1}{p}}}
14(M+12)1p\displaystyle\geq\frac{1}{4(\frac{M+1}{2})^{\frac{1}{p}}}
=21p4(M+1)1p\displaystyle=\frac{2^{\frac{1}{p}}}{4(M+1)^{\frac{1}{p}}}

Finally, combining these and applying that M2M\geq 2 to see that M1M+113\frac{M-1}{M+1}\geq\frac{1}{3}, we have

f(M/4)12(2)1p(M1M+1)1/pϵ1+ϵ\displaystyle\frac{f(M/4)}{12(2)^{\frac{1}{p}}}\left(\frac{M-1}{M+1}\right)^{1/p}-\frac{\epsilon}{1+\epsilon} 112(6)1/pf(M/4)ϵ1+ϵ\displaystyle\geq\frac{1}{12(6)^{1/p}}f(M/4)-\frac{\epsilon}{1+\epsilon}
112(6)1/pf(d(x,y)/8)ϵ1+ϵ.\displaystyle\geq\frac{1}{12(6)^{1/p}}f(d(x,y)/8)-\frac{\epsilon}{1+\epsilon}.

This enables us to bound ρFk\rho_{F_{k}} below by 112(6)1/pf(d(x,y)8)ϵ1+ϵ\frac{1}{12(6)^{1/p}}f(\frac{d(x,y)}{8})-\frac{\epsilon}{1+\epsilon} where ϵ\epsilon has been chosen sufficiently small. ∎

References

  • [Bau] Florent Baudier “Bicone Convexity and the Geometry of the Diamond Graphs, in preparation”
  • [Bau+17] Florent Baudier, Ryan Causey, Stephen Dilworth, Denka Kutzarova, Nirina L. Randrianarivony, Thomas Schlumprecht and Sheng Zhang “On the geometry of the countably branching diamond graphs” In Journal of Functional Analysis 273.10, 2017, pp. 3150–3199 DOI: https://doi.org/10.1016/j.jfa.2017.05.013
  • [BG21] Florent Baudier and Christopher Gartland “Umbel convexity and the geometry of trees”, 2021 arXiv: https://arxiv.org/pdf/2103.16011.pdf
  • [Bau+21] Florent Baudier, Gilles Lancien, Pavlos Motakis and Thomas Schlumprecht “The geometry of Hamming-type metrics and their embeddings into Banach spaces” In Israel Journal of Mathematics 244.2, 2021, pp. 681–725 DOI: 10.1007/s11856-021-2187-0
  • [BZ16] Florent Baudier and Sheng Zhang “(β\beta)-distortion of some infinite graphs” In Journal of the London Mathematical Society 93.2, 2016, pp. 481–501 DOI: 10.1112/jlms/jdv074
  • [Bou86] J. Bourgain “The metrical interpretation of superreflexivity in Banach spaces” In Israel Journal of Mathematics 56.2, 1986, pp. 222–230 DOI: 10.1007/BF02766125
  • [Dil+16] S.J. Dilworth, Denka Kutzarova, N. Randrianarivony, J.P. Revalski and N.V. Zhivkov “Lenses and asymptotic midpoint uniform convexity” In Journal of Mathematical Analysis and Applications 436.2, 2016, pp. 810–821 DOI: https://doi.org/10.1016/j.jmaa.2015.11.061
  • [Gro93] Mikhael Gromov “Filling Invariants” In Geometric Group Theory 2, London Mathematical Society Lecture Note Series Cambridge University Press, 1993, pp. 81–118 DOI: 10.1017/CBO9780511629273.007
  • [HO04] Lorenz Halbeisen and Edward Odell “On asymptotic models in Banach Spaces” In Israel Journal of Mathematics 139.1, 2004, pp. 253–291 DOI: 10.1007/BF02787552
  • [JS09] William B. Johnson and Gideon Schechtman “Diamond Graphs and Super-reflexivity” In Journal of Topology and Analysis 01, 2009, pp. 177–189
  • [Klo14] Benoît R. Kloeckner “Yet another short proof of Bourgain’s distortion estimate for embedding of trees into uniformly convex Banach spaces” In Israel Journal of Mathematics 200.1, 2014, pp. 419–422 DOI: 10.1007/s11856-014-0024-4
  • [Kut91] Denka Kutzarova kk-β\beta and kk-nearly uniformly convex Banach spaces” In Journal of Mathematical Analysis and Applications 162.2, 1991, pp. 322–338 DOI: https://doi.org/10.1016/0022-247X(91)90153-Q
  • [Mat99] Jiří Matoušek “On embedding trees into uniformly convex Banach spaces” In Israel Journal of Mathematics 114.1, 1999, pp. 221–237 DOI: 10.1007/BF02785579
  • [MN13] Manor Mendel and Assaf Naor “Markov convexity and local rigidty of distorted metrics” In Journal of the European Mathematical Society 15.1, 2013, pp. 287–337 DOI: https://doi.org/10.4171/jems/362
  • [Tes11] Romain Tessera “Asymptotic isoperimetry on groups and uniform embeddings into Banach spaces” In Commentarii Mathematiici Helvetici 86.3, 2011, pp. 499–535 DOI: https://doi.org/10.4171/cmh/232