Notes on Alexandros Eskenazis Rennes lectures january 26th 2023

When does {L_p} embed into {L_q}?

The question goes back to Banach. This will be an excuse to review some modern tools. Banach space theory has evolved into metric geometry. The leitmotiv will be: how does one prove non-embeddability?

1. Linear embeddings

1.1. Banach’s question

Notation. {L_p=L_p(0,1)=} measurable functions on {(0,1)}. {\ell_p =} {p}-summable sequences.

{\ell_p^n={\mathbb R}^n} in its {\ell_p} norm.

Definition 1 A linear operator {T:X\rightarrow Y} between normed spaces is a {D}-isomorphic embedding, where {D\ge 1}, if there exists {s>0} such that {\forall x\in X},

\displaystyle  s\|x\|_X \le \|Tx\|_Y\le DS\|x\|_X.

Example 1 {\ell_p} embeds in {L_p} isometrically.

Remark. {\ell_p} is not isomorphic to {L_p} when {p\not=2}. This is a nontrivial fact, probably going back to Banach.

Question (Banach). For which {p\not=q\in[1,\infty)} does {L_p} embed isomorphically into {L_q}?

Theorem 2 (Banach 1932, Paley 1936) {L_p} does not embed into {L_q} unless {p=2} or {1\le q<p<2}.

Theorem 3 (Kadec 1958) {L_p} embeds isometrically into {L_q} if {p=2} or {1\le q<p<2}.

Theorem 2 is much harder than Theorem 3 (although older).

1.2. Proof of Kadec’ theorem

Let {g_1,g_2,\ldots} be a sequence of iid standard gaussian random variables on a {\sigma}-finite probability space.

Gaussians have the following property: if standard gaussian r.v. {g} and {g'} are independent, the random variable {\lambda g+\mu g'} has the same distribution as {\sqrt{\lambda^2+\mu^2}g}. Indeed, both have the same characteristic function.

Define the embedding {T:\ell_2\rightarrow L_q} as follows,

\displaystyle  \forall a=(a_n)_n \in\ell_2,\quad Ta=\sum a_i g_i.

Then {Ta} has the same distribution as {\|a\|_2 g}. Therefore

\displaystyle  \|Ta\|_p=\|a\|_2\|g\|_p .

Question. Why doesn’t this work for other values of {p}? In other words, do there exist iid symmetric random variables {X,X'} such that {\lambda g+\mu g'} has the same distribution as {(\lambda^p+\mu^p)^{1/p}g} ?

The answer arises in Paul Lévy’s book.

Theorem 4 (Levy 1951) Such random variables exist iff {0<p\le 2}. They are called standard {p}-stable random variables, and for {0<p<2}, they satisfy

\displaystyle  \lim_{t\rightarrow\infty}t^p\mathop{\mathbb P}\{|x|\ge t\}=\sigma_p\in(0,\infty).

Therefore,

\displaystyle  \mathop{\mathbb E}(|X|^q) =\mathop{\mathbb E}(\int_0^\infty qt^{q-1}\mathbf{1}_{\{t\le|x|\}}dt) =q\int_0^\infty t^{q-1}\mathop{\mathbb P}\{|x|\ge t\}\,dt

is finite iff {q<p}.

This shows that {\ell_p} embeds isometrically into {L_q} for {1\le q<p<2}.

Remark 1 To prove that {L_p} embeds into {L_q} in this range, one needs two more nontrivial results:

  • {L_p} is finitely representable into {\ell_p}.
  • If {X} is finitely representable in {\ell_q} and is separable, then {X} embeds into {L_q}.

This ends our excursion in the construction of embeddings. From now on, we switch to nonembeddability results.

1.3. Linear distorsion

Definition 5 The linear distorsion of {X} into {Y} is the smallest {D\ge 1} such that there exists a {D}-isomorphic embedding of {X} into {Y}. It is denoted by

\displaystyle  c_Y^{lin}(X),

and if {Y=L_q}, by

\displaystyle  c_q^{lin}(X).

Thus we have stated and partly proved that

\displaystyle  c_q(L_p)=\begin{cases} 1 & \text{ if }p=2 \text{ or }1\le q<p<2, \\ \infty & \text{otherwise}. \end{cases}

Goal. Understand the asymptotics of {c_q^{lin}(\ell_p^n)} when {p,q} are in the second range.

Theorem 6 Let {1\le p\not=q<\infty}. Then

\displaystyle  c_q^{lin}(\ell_p^n)=_{p,q}~1 \text{ if }1\le q<p\le 2,\quad \text{ (Kadec)}

\displaystyle  n^{1/p - 1/q} \text{ if }1\le p<q\le 2, \quad \text{ (type range)}

\displaystyle  n^{1/p - 1/2} \text{ if }1\le p\le 2\le q,

\displaystyle  n^{1/q - 1/p} \text{ if }1\le p<q\le 2,\quad \text{ (cotype range)}

\displaystyle  n^{1/2 - 1/p} \text{ if }1\le p<q\le 2,

\displaystyle  n^{\frac{(q-p)(p-2)}{p^2(q-2)}} \text{ if }1\le p<q\le 2,\quad (X_p \text{ range)}.

Easy. The upper bounds are easy, the embedding is identity or identity to {\ell_2^n} followed with Kadec’s isometric embedding.

Our goal is to prove lower bounds, by designing invariants.

2. Smoothness and convexity in {L_p} spaces

2.1. Smoothness and convexity

Definition 7 (Ball – Carlen – Lieb 1993) Fix {1\le p\le 2}. Say a Banach space {X} is {p}-uniformly smooth with constant {s} if

\displaystyle  \forall x,y\in X,\quad \frac{\|x\|^p+\|y\|^p}{2}\le \|\frac{x+y}{2}\|^p+s^p\|\frac{x-y}{2}\|.

Fix {2\le q\le \infty}. Say a Banach space {X} is {q}-uniformly convex with constant {K} if

\displaystyle  \forall x,y\in X,\quad \|\frac{x+y}{2}\|^p+\frac{1}{K^{p}}\|\frac{x-y}{2}\|\le \frac{\|x\|^q+\|y\|^q}{2}.

The best constants are denoted by {s_p(X)} and {K_q(X)}.

Exercise. (Lindenstrauss’ duality formula à la Ball-Carlen-Lieb). For any normed space {X} and {\frac{1}{p}+\frac{1}{p'}=1},

\displaystyle  s_p(X)=K_{p'}(X^*)\quad\text{ and }\quad s_p(X^*)=K_{p'}(X).

2.2. The case of {L_p} spaces

Theorem 8 (Clarkson’s inequality) For {1\le p\le 2}, {s_p(L_p)=1}.



For {2\le q <\infty}, {K_q(L_q)=1}.

Proof. Since the inequality involves only {\|.\|_q^q}, is suffices to prove that for all {a,b\in{\mathbb R}},

\displaystyle  |\frac{a+b}{2}|^q+|\frac{a-b}{2}|^q\le\frac{|a|^q+|b|^q}{2}.

By monotonicity of {\ell^q} norms and the parallelogram identity,

\displaystyle  (|\frac{a+b}{2}|^q+|\frac{a-b}{2}|^q)^{1/q}\le(|\frac{a+b}{2}|^2+|\frac{a-b}{2}|^2)^{1/2}=(\frac{a^2+b^2}{2})^{1/2}\le(\frac{|a|^q+|b|^q}{2})^{1/q}.

Theorem 9 For {1<p\le 2}, {K_2(L_p)\le\frac{1}{\sqrt{p-1}}}.



For {2\le q<\infty}, {s_2(L_q)\le\sqrt{q-1}}.

2.3. Two lemmata

The proof of Theorem 10 requires two lemmata, the first one is Bonami’s hypercontractivity inequality on the {1}-cube.

Lemma 10 (Bonami’s two-point inequality) Let {1<p\le 2}. For {a,b\in{\mathbb R}},

\displaystyle  (a^2+(p-1)b^2)^{1/2}\le(\frac{|a+b|^p+|a-b|^p}{2})^{1/p}.

Proof. One can assume that {a=1} and {b=x\le 1}. Using Taylor’s expansion, one checks that

\displaystyle  (1+(p-1)x^2)^{p/2}\le \frac{(1+x)^p+(1-x)^p}{2}.

The second lemma illustrates a general principle: an inequality for {L_p} with constant one can be nothing but a relaxation of the parallelogram identity in Euclidean space.

Lemma 11 (Hanner’s inequality) Let {1\le p\le 2}. For {f,g\in L_p},

\displaystyle  |\|f\|_p-\|g\|_p|^p +(\|f\|_p+\|g\|_p)^p\le \|f+g\|_p^p+\|f-g\|_p^p.

One first checks that for {r\in[0,1]}, the numbers

\displaystyle  \alpha(r)=(1+r)^{p-1}+(1-r)^{p-1}\quad \text{and}\quad \beta(r)=\frac{(1+r)^{p-1}-(1-r)^{p-1}}{r^{p-1}}

satisfy

\displaystyle  \forall A,B\in{\mathbb R},\quad \max_{r\in[0,1]}\{\alpha(r)|A|^p+\beta(r)B^p\}\le |A+B|^p+|A-B|^p.

(again, one can assume that {A=1} and {|B|\le 1} ; then it amounts to the monotonicity of a function).

Hanner’s inequality follows: assuming that {\|f\|_p\ge \|g\|_p}, set {A=|f(x)|}, {B=|g(x)|}, {r=\frac{\|g\|_p}{\|f\|_p}} to get

\displaystyle  |f(x)+g(x)|^p+|f(x)-g(x)|^p\ge \left( (1+\frac{\|g\|_p}{\|f\|_p})^{p-1}+(1-\frac{\|g\|_p}{\|f\|_p})^{p-1} \right)|f(x)|^p

\displaystyle  +\frac{(1+\frac{\|g\|_p}{\|f\|_p})^{p-1}-(1-\frac{\|g\|_p}{\|f\|_p})^{p-1}}{(\frac{\|g\|_p}{\|f\|_p})^{p-1}}|g(x)|^p.

Integrating with respect to {x} yields

\displaystyle  \|f+g\|_p^p+\|f-g\|_p^p \ge\left( |\|f\|_p+\|g\|_p|^{p-1} +|\|f\|_p-\|g\|_p|^{p-1} \right)\|f\|_p

\displaystyle  + \left((\|f\|_p+\|g\|_p|)^{p-1} -|\|f\|_p-\|g\|_p|^{p-1} \right)\|g\|_p

\displaystyle  =(\|f\|_p+\|g\|_p)^{p}+|\|f\|_p-\|g\|_p|^{p}.

2.4. Proof of Theorem 10

Let {f=\frac{x+y}{2}}, {g=\frac{x-y}{2}}. We show that

\displaystyle  \|f\|_p^2+(p-1)\|g\|_p^2 \le \frac{\|f+g\|^2_p+\|f-g\|^2_p}{2}.

Indeed, first apply Bonami’s inequality.

Conjecture. Does Hanner’s inequality hold in Schatten class, i.e. for matrices where the {\ell_p} norm of singular values is used?

Known for {p\ge 4} and {p\le \frac{4}{3}} (Ball-Carlen-Lieb). Heinavaara 2022 proves this for {p=3}.

Conjecture. For {1<p\le 2},

\displaystyle  \mathop{\mathbb E}|\sum\epsilon_i\|f_i\|_p|^p\le \mathop{\mathbb E}\|\sum\epsilon_i f_i\|_p^p

Theorem 12 (D. Schechtman 1995) Yes for {p\ge 3}.

3. Martingales in Banach spaces

Definition 13 Let {X} be a Banach space. A (Paley-Walsh) martingale with values in {X} is a sequence of functions {M_k:\{-1,1\}^k\rightarrow X} such that for all {\epsilon\in\{-1,1\}^k},

\displaystyle  M_k(\epsilon_1,\ldots,\epsilon_k)=\frac{M_{k+1}(\epsilon_1,\ldots,\epsilon_k,1)+M_{k+1}(\epsilon_1,\ldots,\epsilon_k,0)}{2}.

The basic example is

\displaystyle  M_k(\epsilon)=\sum\epsilon_ix_i,

for given vectors {x_1, x_2,\ldots\in X}.

Remark. If {X={\mathbb R}}, the set {\{M_k-M_{k-1}\}} is orthogonal, so

\displaystyle  \mathop{\mathbb E}|M_n-M_0|^2=\sum\mathop{\mathbb E}|M_k-M_{k-1}|^2.

Definition 14 (Pisier 1975) A Banach space {X} has martingale type {p}, {1\le p\le 2}, il for every {X}-valued martingale {\{M_k\}},

\displaystyle  \mathop{\mathbb E}\|M_n-M_0\|_X^p \le T_p(X)^p\sum\mathop{\mathbb E}\|M_k-M_{k-1}|_X^p.

{X} has martingale cotype {q}, {2\le q\le \infty}, il for every {X}-valued martingale {\{M_k\}},

\displaystyle  \mathop{\mathbb E}\|M_n-M_0\|_X^q \ge\frac{1}{c_q(X)^q}\sum\mathop{\mathbb E}\|M_k-M_{k-1}|_X^q.

These properties can be thought of as relaxations of the identity that holds in Hilbert space.

3.1. Smoothness/convexity versus type/cotype

These properties follow from the convexity properties introduced earlier.

Proposition 15 (Pisier) {p}-smoothness (resp. {q}-convexity) implies martingale type {p} (resp. cotype {q}) with the same constant.

Pisier’s renorming theorem states that the converse is true, up to changing for an equivalent norm.

The advantage of the type/cotype formulation is that the error is multiplicative, hence these properties are isomorphism invariant.

3.2. Proof of Pisier’s “smoothness implies type”

Let {\{M_k\}} be an {X}-valued martingale with {M_0=0}. Then

\displaystyle  \mathop{\mathbb E}\|M_n(\epsilon)\|^p = \mathop{\mathbb E}_{\epsilon_1,\ldots,\epsilon_{n-1}}(\frac{\|M_n(\epsilon_1,\ldots,\epsilon_{n-1},1)\|^p+\|M_n(\epsilon_1,\ldots,\epsilon_{n-1},0)\|^p}{2})

\displaystyle  \le \mathop{\mathbb E}_{\epsilon_1,\ldots,\epsilon_{n-1}}\left( \|\frac{M_n(\epsilon_1,\ldots,\epsilon_{n-1},1)+M_n(\epsilon_1,\ldots,\epsilon_{n-1},0)}{2} \|^p \right.

\displaystyle  \left.+S_p(X)^p\|\frac{M_n(\epsilon_1,\ldots,\epsilon_{n-1},1)-M_n(\epsilon_1,\ldots,\epsilon_{n-1},0)}{2}\|^p \right)

\displaystyle  =\mathop{\mathbb E}(\|M_{n-1}\|^p+s_p(X)^p\|M_{n}-M_{n-1}\|^p)

\displaystyle  \le \mathop{\mathbb E}(\|M_{n-2}\|^p+s_p(X)^p(\|M_{n}-M_{n-1}\|^p+\|M_{n-1}-M_{n-2}\|^p)

\displaystyle  \le ...

\displaystyle  \le s_p(X)^p \sum\mathop{\mathbb E}\|M_k-M_{k-1}\|^p,

thus {X} has martingale type {p} with constant {s_p(X)}.

Corollary 16 Let {x_1,\ldots,x_n\in L_q}.


If {q\le 2}, then

\displaystyle  \mathop{\mathbb E}\|\sum\epsilon_i x_i\|_q^q \le \sum\|_i\|_q^q, \quad (\text{Rademacher type }q)

\displaystyle  \mathop{\mathbb E}\|\sum\epsilon_i x_i\|_q^2 \ge(q-1) \sum\|_i\|_2^q,\quad (\text{Rademacher cotype }2).

If {q\ge 2}, then

\displaystyle  \mathop{\mathbb E}\|\sum\epsilon_i x_i\|_q^2 \le (q-1)\sum\|_i\|_q^2, \quad (\text{Rademacher type }2)

\displaystyle  \mathop{\mathbb E}\|\sum\epsilon_i x_i\|_q^q \ge \sum\|_i\|_q^q,\quad (\text{Rademacher cotype }q).

Beware that {L_1} has Rademacher cotype {2} with constant {\sqrt{2}} but no nontrivial martingale cotype.

3.3. Proof of Theorem 7, type and cotype range

Let {T:\ell_p^n\rightarrow L_q} be an embedding of distorsion {D}. Apply type {q} of {L_q} to {x_i=Te_i}. Then

\displaystyle  \sum\|Te_i\|_q^q\le nD^q.

On the other hand,

\displaystyle  \mathop{\mathbb E}\|T(\sum\epsilon_i e_i)\|_q^q \ge \mathop{\mathbb E}\|\sim\epsilon_ie_i\|_p^p=n^{q/p}.

This yields {D\ge n^{1/p - 1/q}}.

The proofs of the three other lower bounds on {c_q^{lin}(\ell_p^n)} are similar. The results are sharp.

Note that the argument used linearity very strongly.

4. Nonlinear embeddings

Given metric spaces {M} and {N}, one can speak of the biLipschitz distorsion of {M} into {N}, denoted by {c_N(M)}, as the least {D} such that there exists {s>0} and {f:M\rightarrow N} satisfying

\displaystyle  \forall x,x'\in M,\quad s\,d(x,x')\le d(f(x),f(x'))\le Ds\,d(x,x').

Now we discretize spaces. Let {[m]_p^n=(\{1,2,\ldots,n\},\|.\|_p)}, viewed as an approximation of {\ell_p^n}. What can one say about {c_q([m]_p^n)}?

Theorem 17 Let {m,n\ge 2}. Then {c_q([m]_p^n)} is of the order of

\displaystyle  1 \text{ if }1\le q<p\le 2,\quad \text{ (Kadec)}

\displaystyle  n^{1/p - 1/q} \text{ if }1\le p<q\le 2, \quad \text{ (metric type range)}

\displaystyle  n^{1/p - 1/2} \text{ if }1\le p\le 2\le q,

\displaystyle  ?? \text{ if }2\le q\le p,

\displaystyle  n^{1/q - 1/p} \text{ if }1\le p<q\le 2,\quad \text{ (metric cotype range)}

\displaystyle  n^{1/2 - 1/p} \text{ if }1\le p<q\le 2,

\displaystyle  n^{\frac{(q-p)(p-2)}{p^2(q-2)}} \text{ if }1\le 2\le p<q,\quad (\text{metric }X_p \text{ range)}.

So we see that a phase transition occurs when {q\le 2\le p}.

In certain cases, the upper bounds are smart, but I will not focus on them.

In the unknown range {2\le q\le p}, the best we know now is

\displaystyle  \min\{n^{1/q - 1/p},m^{1- q/p}\}\le_{p,q}c_q([m]_p^n)\le_{p,q}\min\{n^{1/q - 1/p},m^{1- 2/p}\}.

The left-hand side would be sharp if the following was true: for every {r>2} and every {0<\theta<1}, the metric space {(L_r,\|x-y\|_r^{\theta})} has a biLipschitz embedding into {L_r}.

Theorem 18 (Bretagnolle – Dacunha-Castelle – Krivine 1965) This is true for {r\le 2}.

In my thesis, we have the following result:

Theorem 19 (Eskenazis – Naor 2016) For {r>2}, {0<\theta<1}, {0<\alpha<\theta}, the space

\displaystyle  (L_r,\frac{\|x-y\|_r^{\theta}}{1+\log^\alpha(1+\|x-y\|_r)})

does not embed in {L_r}.

5. Metric type

5.1. Enflo type

Definition 20 (Enflo 1969, modern terminology) A metric space {M} has Enflo type {p} with constant {T>0} if for all {n\in{\mathbb N}}, for all {f:\{-1,1\}^n\rightarrow M},

\displaystyle  \mathop{\mathbb E} d(f(\epsilon),f(-\epsilon))^p\le T^p\sum_i \mathop{\mathbb E} d(f(\epsilon),f(\epsilon_1,\ldots,\epsilon_{i-1},-\epsilon_i,\epsilon_{i+1},\ldots,\epsilon_n))^p.

Remark. If {X} is a normed space and {f(\epsilon)=\sum\epsilon_i x_i}, then the lefthand side is

\displaystyle  2^p\mathop{\mathbb E}\|\sum\epsilon_i x_i\|^p

and the righthand side is

\displaystyle  2^p\sum\|x_i\|^p,

so this is really a metric analogue of Rademacher type: for linear spaces,

\displaystyle  \text{Enflo type }p \Rightarrow \text{Rademacher type }p.

Theorem 21 (Khot – Naor 2006) For normed linear spaces,

\displaystyle  \text{Martingale type }p \Rightarrow \text{Enflo type }p.

Proof. Let {f:\{-1,1\}^n\rightarrow X} which has martingale type {p}. Define a martingale

\displaystyle  M_k(\epsilon)=\mathop{\mathbb E}_\epsilon f(\epsilon_1,\ldots,\epsilon_k,\delta_{k+1},\ldots,\delta_n).

From martingale type, we know that {\mathop{\mathbb E}\|f-\mathop{\mathbb E} f\|p\le T^p\sum_k\mathop{\mathbb E}\|M_k-M_{k+1}\|^p}. But

\displaystyle  M_k-M_{k+1}=\frac{1}{2}\mathop{\mathbb E}_\delta(f(\epsilon,\delta)-f(\epsilon',\delta)),

where {\epsilon'} has one sign changed. So

\displaystyle  \mathop{\mathbb E}\|M_k-M_{k+1}\|^p\le \frac{1}{2^p}\mathop{\mathbb E}_{\epsilon}\|f(\epsilon)-f(\epsilon')\|^p

where {\epsilon} has length {n} and {\epsilon'} one sign flipped at position {k}. So summing over {k} yields the expected righthand side. Now add and subtract the expectation,

\displaystyle  \mathop{\mathbb E}\|f(\epsilon)-f(-\epsilon)\|^p\le 2^{p-1}(\mathop{\mathbb E}\|f(\epsilon)-\mathop{\mathbb E} f\|^p +\mathop{\mathbb E}\|\mathop{\mathbb E} f-f(-\epsilon)\|^p)

\displaystyle  \le 2^p\mathop{\mathbb E}\|f(\epsilon)-\mathop{\mathbb E} f\|^p\le 2^p\sum_k\mathop{\mathbb E}\|M_k-M_{k+1}\|^p

\displaystyle  \le \sum_k\mathop{\mathbb E}_{\epsilon}\|f(\epsilon,\epsilon')\|^p,

which completes the proof.

The converse is a recent breakthrough, still in the linear case:

Theorem 22 (Ivanisvili – von Handel – Volberg 2020) For linear spaces, Rademacher type implies Enflo type.

5.2. Proof of distorsion lower bounds for {p\le 2}

The upper bound is easy (use global embedding of {L_p}).

Here is another easy lower bound:

\displaystyle  c_q([m]_p^n)\ge c_q([2]_p^n)=c_q(\{-1,1\}^n,\|.\|_p).

Assume that {1\le p\le q\le 2} (the other case is similar). Let {f:\{-1,1\}^n\rightarrow L_q} have distorsion {D} with rescaling {s}. Since {L_q} has Enflo type {q} with constant {1},

\displaystyle  \mathop{\mathbb E}\|f(\epsilon)-f(-\epsilon)\|_q^q \le \sum\|f(\epsilon)-f(\epsilon')\|_q^q,

(one sign change in {\epsilon'}), so the righthand side is bounded above by {(sD)^q n 2^q}, and the lefthand side is bounded below by {s^q\mathop{\mathbb E}\|\epsilon-(-\epsilon)\|_p^q=2^q s^q n^{q/p}}, so {D\ge n^{1/p -1/q}}.

6. Metric cotype

Formally, Rademacher cotype is the reverse of Rademacher type, but a naive delinearization does not work. The cube does not suffice, one needs an extra scaling factor {m}.

Definition 23 (Mendel – Naor 2008) A metric space {M} has metric cotype {q>0} with constant {c>0} if for any {n}, there exists {m=m(n,M)} such that any function {f:{\mathbb Z}_{2m}^n\rightarrow M} satisfies

\displaystyle  \sum_{t=1}^n \mathop{\mathbb E}_{x\in{\mathbb Z}_{2m}^n} d(f(x+me_i),f(x))^q \le C^q m^q \mathop{\mathbb E}_x\mathop{\mathbb E}_\epsilon d(f(x+\epsilon),f(x)).

Remark. As soon as {M} has at least 2 points, {m(n,M)\ge n^{1/q}}.

Theorem 24 (Mendel – Naor 2008, Giladi – Mendel – Naor 2011) For normed spaces, Rademacher cotype is equivalent to metric cotype, and one can always take {m\le n^{1+ 1/q}}.

The major open question is wether one can take {m=n^{1/q}}. Indeed, this would have many geometric applications.

Theorem 25 (Eskenazis – Mendel – Naor 2019) For normed spaces, martingale cotype implies metric cotype, with {m\le n^{1/q}}.

6.1. Proof of the remaining distorsion lower bounds in Theorem 18

I will cheat and identify {{\mathbb Z}_{2m}^{n}} with {[m]_p^n}. This is not a serious matter, since

\displaystyle  [m]_p^n \subset {\mathbb Z}_{m}^{n}\subset [m+1]_p^{2n}.

We must show that

\displaystyle  c_q([m]_p^n)\ge_{p,q} m^{1-\frac{q}{p}} \quad \text{ if }m\le n^{1/q},

\displaystyle  c_q([m]_p^n)\ge_{p,q} n^{\frac{1}{q}-\frac{1}{p}} \quad \text{ if }m\ge n^{1/q}.

Since decreasing {m} decreases the {L^p}-distorsion, one can assume that {m=n^{1/q}}. Let {f:\rightarrow L_q} have distorsion {D}. Then the lefthand side of metric cotype assumption is {\ge s^qm^qn}, whereas the right hand side is {\le m^q(sD)^q n^{q/p}}, so {D\ge n^{\frac{1}{q}-\frac{1}{p}}}.

We see that we really need {m=n^{1/q}}.

6.2. Proof of Theorem 26

The strategy is inspired from hypercontractivity: smoothing the space using an averaging operator puts functions closer to linear functions, i.e. leads us closer to the linear setting.

Given {h:\{-1,1\}^n \rightarrow X}, we define a martingale

\displaystyle  E_i h(\epsilon)=\mathop{\mathbb E}_{\delta}h(\epsilon,\delta),

where {\epsilon} has length {i}. We view {E_i} as a smoothing operator.

If {f:{\mathbb Z}_{4m}^n\rightarrow X} and {x\in {\mathbb Z}_{4m}^n}, let {f_x(\epsilon)=f(x+\epsilon)}. Then

\displaystyle  \sum_i \mathop{\mathbb E}\|f(x+me_i)-f(x)\|^p = \sum_i \mathop{\mathbb E}_{x,\epsilon}\|f_{x+2me_i}(\epsilon)-f(\epsilon)\|^p

\displaystyle  =\sum_i \mathop{\mathbb E}_{x,\epsilon}(\|f_{x+2me_i}(\epsilon)-E_i f_{x+2me_i}(\epsilon)\|+\|E_i f_{x+2me_i}(\epsilon)-E_i f_x(\epsilon)\|+\|E_i f_{x}(\epsilon)-f_x(\epsilon)\|)^p

\displaystyle  \le (1)+(2),

where (1) is an approximation term,

\displaystyle  (1)=\sum_i \|E_i f_{x}(\epsilon)-f_x(\epsilon)\|^p,

and (2) is a smoothing term,

\displaystyle  (2)=\sum_i \|E_i f_{x+2me_i}(\epsilon)-E_if_x(\epsilon)\|^p

We first estimate (1) with Hölder,

\displaystyle  (1)\le \sum_i (\mathop{\mathbb E}_{x,\epsilon}\|E_if_x(\epsilon)-f(\epsilon)\|^p +\|f_x(\epsilon)-f(x)\|^p).

Jensen’s inequality gives

\displaystyle  \mathop{\mathbb E}_\epsilon\|E_i h(\epsilon)\|^p\le \mathop{\mathbb E}\|h(\epsilon)\|^p.

In other words, {E_i} contracts {L_p}, so

\displaystyle  (1)\le \sum_i \mathop{\mathbb E}_{x,\epsilon}\|f_x(\epsilon)-f(x)\|^p=n\mathop{\mathbb E}_{x,\epsilon}\|f_x(\epsilon)-f(x)\|^p,

which is ok since {n=m^p}, this what we want to see on the righthand side.

Next we bound (2). We can replace {2m} with {-2m} mod {4m}. For fixed {i}, write the summand in (2) as a telescopic sum,

\displaystyle  \mathop{\mathbb E}_{x,\epsilon}\|E_i f_{x-2me_i}(\epsilon)-E_if_x(\epsilon)\|^p \le \mathop{\mathbb E}_{x,\epsilon}\|\sum_\ell(E_i f_{x-2\ell e_i}(\epsilon)-E_if_{x-2(\ell-1)e_i}(\epsilon)\|^p

\displaystyle  \le m^{p-1}\sum_\ell\mathop{\mathbb E}_{x,\epsilon}\|E_i f_{x+2\ell e_i}(\epsilon)-E_i f_{x-2(\ell-1)e_i}(\epsilon)\|^p

\displaystyle  =m^p\mathop{\mathbb E}_{x,\epsilon}\|E_i f_{x+2 e_i}(\epsilon)-E_i f_{x}(\epsilon)\|^p

But

\displaystyle  \|E_i f_{x+2e_i}(\epsilon)-E_i f_{x}(\epsilon)\|=2\|E_i f_x(\epsilon)-E_{i-1}f_x(\epsilon)\|^p.

By martingale cotype,

\displaystyle  (2)\le m^p \mathop{\mathbb E}_{x,\epsilon}\|E_n f_x (\epsilon)-E_0 f_x(\epsilon)\|^p

\displaystyle  =m^p\mathop{\mathbb E}_{x,\epsilon}\|f(x+\epsilon)-\mathop{\mathbb E}_\delta f(x+\delta)\|^p\le m^p\mathop{\mathbb E}_{x,\epsilon}\|f(x+\epsilon)-f(x)\|^p.

Adding {(1)+(2)} gives the expected bound {m^p\mathop{\mathbb E}_{x,\epsilon}\|f(x+\epsilon)-f(x)\|^p}.

6.3. Final comment

What {m=n^{1/q}} does for you is that it provides invariance under coarse embeddings, and not merely biLipschitz embeddings. So the story is far from being finished.

About metric2011

metric2011 is a program of Centre Emile Borel, an activity of Institut Henri Poincaré, 11 rue Pierre et Marie Curie, 75005 Paris, France. See http://www.math.ens.fr/metric2011/
This entry was posted in Course. Bookmark the permalink.

Leave a comment