Notes of Will Sawin first Hadamard lecture, 15-05-2023

Number theory over function fields

1. The classical theory

Let us start with classical stuff.

Theorem 1 (Dirichlet) {N} a positive integer, {a} an integer. There exist infinitely many prime numbers {p} such that {p=a \mod N} if and only if {gcd(a,N)=1}.

A more precise statement is the prime number theorem on arithmetic progressions:

\displaystyle  \#\{p\in \mathcal{P}\,;\,p=a\mod N,\, p<x\}=\frac{1}{\phi(N)}\int_{2}^{x}\frac{dy}{\log y}+O_N (xe^{-C\sqrt{\log x}})

for some explicit {C}.

In other words, remainders mod {N} of primes are evenly distributed.

The proofs of both theorems rely on properties of Dirichlet characters.

Definition 2 A function {\chi:{\mathbb Z}^{>0}\rightarrow{\mathbb C}} is a Dirichlet character mod {N} if

  1. {\chi(mn)=\chi(m)\chi(n)} for all {m,n},
  2. {\chi(d+N)=\chi(d)} for all {d},
  3. {\chi(d)=0\iff gcd(d,N)\not=1}.

From a Dirichlet character, one constructs a Dirichlet {L}-function

\displaystyle  L(x,\chi)=\sum_{n=1}^{\infty}\chi(n)n^{-s}=\prod_{p \in \mathcal{P}}\frac{1}{1-\chi(p)p^{-s}}.

1.1. Properties

If {\chi\not=\chi_0}, the trivial character, then {L(s,\chi)} is entire. Dirichlet’s theorem follows from {L(1,\chi)\not=0}. The prime number theorem on arithmetic progressions follows from {L(1+t,\chi)\not=0} for all {t\in{\mathbb R}} and on a neighborhood of that line. Improving the remainder to be polynomial of degree {\alpha} is {x} amounts to nonvanishing of {L(s,\chi)} on {\Re(s)>\alpha}. This is hard. The special case {\alpha=\frac{1}{2}} is known as Generalized Riemann Hypothesis.

If one thinks of primes as random, i.e. an integer {x} has a probability to be prime which is {\frac{1}{\log x}}, the expected error from this random model is {O_\epsilon((x/N)^{\frac{1}{2}+\epsilon})}.

Other interesting questions about Dirichlet {L}-functions are statistical: what happens on the average over {\chi} ?

1.2. Example: moments

We are interested in

\displaystyle  \sum_{\chi\mod N}L(\frac{1}{2},\chi)^a\overline{L(\frac{1}{2},\chi)}^b.

An exact expression is known only for {a\le 2}, {b\le 2} or {a=3}, {b=0}.

2. Function fields

These problems being too hard, I will study similar questions in a different, hopefully easier, setting. Let {F_q} denote a field with {q} elements, {F_q[t]} denotes the ring of polynomials with coefficients in {F_q}. I intend to replace integers with {F_q[t]}. For instance, the Euclidean algorithm works on polynomials.

I denote by {F_q[t]^+} the set of monic polynomials (thought of as an analogue of positive integers).

For {f\in F_q[t]}, its absolute value is {|f|=q^{\mathrm{deg(f)}}}, which is equal to the cardinality of the quotient ring {F_q[t]/fF_q[t]}.

The advantage of the this shift of setting is that new connections with other fields of mathematics appear. Today, I will give one instance of that.

2.1. Dirichlet characters

Say that a function {\chi:F_q[t]^+\rightarrow{\mathbb C}} is a Dirichlet character mod {g\in F_q[t]^+} if

  1. {\chi(fh)=\chi(f)\chi(h)} for all {f,h\in F_q[t]^+},
  2. {\chi(f+gh)=\chi(f)} for all {f,g\in F_q[t]^+},
  3. {\chi(f)=0\iff gcd(f,g)\not=1}.

The Dirichlet {L}-function is

\displaystyle  L(s,\chi)=\sum_{f\in F_q[t]^+}\chi(f)|f|^{-s}.

It is a power series in {q^{-s}}. {L(s,\chi)} is entire for {\chi\not=\chi_0}, because it is a polynomial in {q^{-s}}. Indeed,

\displaystyle  L(s,\chi)=\sum_{j=0}^{\infty}(\sum_{f\in F_q[t]^+,\,\mathrm{deg(f)=d}}\chi(f))q^{-ds}

and the sum {\displaystyle\sum_{f\in F_q[t]^+,\,\mathrm{deg(f)=d}}\chi(f)} vanishes for {d\ge \mathrm{deg}(g)}, since each residue class mod {g} occurs {q^{d-\mathrm{deg}(g)}} times and, by orthogonality of characters, {\sum_{a\in F_q[t]/g}\chi(a)=0}.

The Euler product formula holds,

\displaystyle  L(s,\chi)=\prod_{\pi\in F_q[t]^+,\,\pi\text{ irreducible}}\frac{1}{1-\chi(\pi)q^{-s\mathrm{deg}(\pi)}},

showing that {L(s,\chi)\not=0} for {\Re(s)>1}. If we can improve nonvanishing we get information on the number of irreducible {\pi\in F_q[t]^+} such that {\pi=a\mod g} and {\mathrm{deg}(\pi)=n}.

2.2. Generalized Riemann Hypothesis

In the new setting, the analogue of the Generalized Riemann Hypothesis is known, this is

Theorem 3 (Weil) {L(s,\chi)\not=0} for {\Re(s)>\frac{1}{2}}.

The proof is geometric.

Weil’s theorem implies that

\displaystyle  \#\{\pi\in F_q[t]^+\,\pi \text{ irreducible},\,\pi=a\mod g\,\mathrm{deg}(\pi)=n\}=\frac{q^n}{\phi(q)n}+O(\mathrm{deg}(g)q^{n/2}).

2.3. Statistical issues

The theorem leaves open statistical questions about {L(s,\chi)}. Katz, Katz-Sarnak, Deligne answered statistical questions in the limit where {q} tends to infinity.

Definition 4 Say a character {\chi} is primitive if there is no character {\chi'} of smaller modulus {g'} such that {\chi(t)=\chi'(t)} for all {f} with {gcd(f,g)=1}.



Say {\chi} is odd if {\chi(g+\alpha)\not=1} for some {\alpha\in F_q}.

Theorem 5 Assume {\chi} is primitive and odd. Then {L(s,\chi)} is a polynomial in {q^{-s}} of degree exactly {\mathrm{deg}(g)-1} with all roots on {\frac{1}{2}+it\in{\mathbb C}}. It follows that {L(s,\chi)} is the characteristic polynomial of {q^{\frac{1}{2}-s}\theta_\chi} where {\theta_\chi} is a unitary matrix.

Theorem 6 (Katz) If {g} is squarefree and {\mathrm{deg}(g)=m}, there a exists a map

\displaystyle  F:U(m-1)\rightarrow{\mathbb C}

which is continuous, conjugacy invariant, such that

\displaystyle  \lim_{q\rightarrow\infty}\frac{1}{q^m}\sum_{\chi\text{ primitive, odd}}f(\theta_\chi)=\int_{U(m-1)}F(\theta)\,d\theta.

In other words, the {\theta_\chi} equidistribute in the unitary group as {q} tends to infinity. It kills hope to control its nonvanishing simultaneously for all {q}.

2.4. A sample result

Here is a theorem of mine.

Theorem 7 Let {g} be squarefree of degree {m}. Then

\displaystyle  \frac{q-1}{(q^m-1)(q-2)}\sum_{\chi\text{ primitive, odd}}L(s,\chi)^k = 1+O(k^m 2^{mk-k-m}q^{\frac{1-m}{2}}).

In other words, the average over primitive, odd characters is 1 up to fluctuations which become small if {q>k^2 2^{2(k-1)}}.

Whereas, in the classical setting, only a few moments are understood, in the new setting one can study arbitrarily high moments, provided {q} is large enough.

The question has a geometric nature because it involves counting solutions to polynomial equations over {F_q}. Since the middle of XXth century, one knows that such a counting can follow from the same topological techniques used to describe solutions over the complex numbers.

2.5. Scheme of proof

Let me give a scheme of the proof of Theorem 7. By the Euler product formula, the lefthand sum can be rewritten

\displaystyle  \sum_\chi\sum_{f_1,\ldots,f_k\in F_q[t]^+}\chi(f_1\cdots f_k)q^{s(\sum\mathrm{deg}(f_i))}.

For a fixed {h\in F_q[t]^+},

\displaystyle  \sum_{\chi\text{ odd}} \chi(h)=\begin{cases} 1 & \text{ if }h=1\mod g, \\ -\frac{1}{q-2} & \text{ if }h=\alpha\mod g,\,\alpha\in F_q[t]^\times, \\ 0 &\text{otherwise}. \end{cases}

Therefore the sum becomes a sum over {\alpha\in F_q[t]^\times} involving the number of {f_1,\ldots,f_k\in F_q[t]^+} such that {f_1\cdots f_k=\alpha\mod g} and {\sum\mathrm{deg}(f_i)=n}.

This set is the union over {k}-tuples of natural numbers summing to {n} of sets, each of which is the set of solutions to {m} equations in {n} variables, over {F_q} (the unknowns are the coefficients of the monic polynomials {f_i}).

The answer, obtained by Grothendieck’s school, can be found on the windows of Orsay’s math building.

Theorem 8 (Grothendieck-Lefschetz formula) Let {X} be a scheme of finite type over {F_q}. Then the number of points of {X} over {F_q} is

\displaystyle  \# X(F_q)=\sum_{j=0}^{2\mathrm{dim}(X)}(-1)^{i}\mathrm{trace}(Frob_q H^i(X(\overline{F_q}),{\mathbb Q}_\ell)).

We also need

Theorem 9 (Deligne’s Riemann hypothesis) The eigenvalues of {Frob_q} on {H^i(X(\overline{F_q}),{\mathbb Q}_\ell)} are algebraic integers of absolue value {\le q^{1/2}}.

This implies that the trace is bounded above by {q^{1/2}} times the dimension of cohomology. This allows to conclude

Theorem 10 The number of solutions to the counting problem is

\displaystyle  \frac{1}{q^m-1}\sum_{j=0}^k (-1)^j\begin{pmatrix}k\\ j \end{pmatrix}\begin{pmatrix}n-jm+k-1\\ k-1 \end{pmatrix}q^{n-jk}+O(...).

Thus we see that geometry has made it possible to go beyond the classical results in the function field setting.

Here is the key geometric statement. Let {g} be a monic squarefree polynomial of degree {m} over a field {F}. Let {a\in F[t]} be prime to {g}. Consider the union of the schemes parametrizing the set of {f_1,\ldots,f_k} whose product equals {a\mod g}. Then the cohomology of the points over {\bar F} splits as the sum of a boring piece (independent on {a}) and an interesting piece which is {0}-dimensional if {i\not=n-m,n+1-m}, and otherwise its dimension is at most

\displaystyle  \begin{cases} k^{m-1} \begin{pmatrix}mk-k-m+1\\ n-m+1 \end{pmatrix} & \text{ if }i=n+1-m, \\ (k^m-\begin{pmatrix}m+k-1\\ m-1 \end{pmatrix})\begin{pmatrix}mk-m-k\\ n-m \end{pmatrix} & \text{ if }i=n-m. \end{cases}

Posted in Uncategorized | Leave a comment

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.

Posted in Course | Leave a comment

Notes of Maryna Viazovska’s Orsay lecture, january 2023

Sphere packings

This lecture has been designed for an audience of 150 master’s students in Orsay, Universite Paris-Saclay.

1. The sphere packing problem

1.1. Density

In Euclidean space, one tries to pack equal balls as densely as possible. Pack means that interiors of balls are disjoint.

The density of a sphere packing in a box is the ratio

\displaystyle  \frac{\text{volume occupied by intersections of balls and box}}{\text{volume of the box}}.

The density of the packing is defined as a {\limsup}, over larger and larger boxes, of densities of packings in boxes.

The supremum of densities of ball packings in {{\mathbb R}^d} is denoted by {\Delta_d}, this is the sphere packing constant in {{\mathbb R}^d}.

Problem. Determine {\Delta_d}.

This is a hard problem. It is an instance of a family of similar problems, in other metric spaces.

1.2. Codes

Let {M} be a metric space, and {r>0}. An {r}-code in {M} is a subset of {M} such that all pairwise distances are {\ge r}.

Example. The centers of a packing of {{\mathbb R}^d} by balls of radius {\frac{r}{2}} is an {r}-code.

Example. In a graph, equipped with the graph metric (which is integer valued), an independent set is a {2}-code.

1.3. What are {r}-codes good for

Example. Let {M} be the Hamming cube, i.e. {M=\{ 0,1 \}^d} and the distance between two points (strings) is the number of coordinates where they differ.

Imagine being given the task of transmitting strings of {0}s and {1}s of length {\ell} to your friend. You are provided with a device which converts such strings into strings of length {d}, which belong to an {r}-code {X} of {M=\{ 0,1 \}^d}. Your friend owns a device that converts back each element of {X} into the {\ell}-string it arose from. Now you send him converted strings by radio. As soon as the radio transmission introduces no more than {\frac{r}{2}} errors, your friend can unambiguously recover the {d}-string you broadcast from the noisy string: it is the only element of {X} that sits at distance {<\frac{r}{2}} from it. Thus he can correct the errors. Using his own converter, he can recover the original {\ell}-string.

Thus {r}-codes in the Hamming cube of high density are desirable, since they allow to transmit long strings without error (after correction).

2. Results in low dimensions

{\Delta_2} has been guessed for centuries: it is achieved by the hexagonal packing in the plane. The first published proof is due to the norwegian mathematician Axel Thue in 1910, but considered as insufficiently rigorous. In 1943, the hungarian mathematician Laszlo Fejes Toth published a complete proof. A good reference for this proof is Brass, Moser and Pach’s book on Research problems in discrete geometry. Other, simpler, proofs have been found since.

Tom Hales’s determination of {\Delta_3} requires a lot of 3-dimensional geometry, plus help from a computer.

3. Convex programming

Today, I explain the method of convex programming and how it is used for the packing problem.

How can one replace the problem, which is not convex, by a convex one? By duality, points will be replaced with functions.

3.1. Copositive and positive definite functions

Let {(M,\rho)} be a metric space. Let {supp(\rho)} denote the set of values of {\rho}.

Definition 1 A function {f:supp(\rho)\rightarrow\mathbb{C}} is copositive if

\displaystyle  \sum_{x,y\in X}f(\rho(x,y))\ge 0

for all finite subsets {X} of {M}.

Copositive function form a convex cone. If we can find a good description of it, we would be in good position to solve the packing problem. Unfortunately, this is rarely the case. Therefore we switch to an easier class of functions.

Definition 2 A function {f:supp(\rho)\rightarrow\mathbb{C}} is positive definite if, for all finite subsets {X\subset M} and all complex coefficients {w:X\rightarrow\mathbb{C}},

\displaystyle  \sum_{x,y\in X}w_x \overline{w_y}f(\rho(x,y))\ge 0.

Clearly,

\displaystyle  \text{positive definite} \Rightarrow \text{copositive}.

Usually, the cone of positive definite functions is much easier to describe.

3.2. A toy theorem

Here is a toy theorem that illustrates the power of the method, although it does not suffice to determine maximal densities.

Theorem 3 Let {(M,\rho)} be a metric space. Let {N} be a integer and {r_0>0}. Suppose there exists a copositive function {f:supp(\rho)\rightarrow{\mathbb R}} such that {f(0)=1} and

\displaystyle  \forall r\ge r_0,\quad f(r)\le -\frac{1}{N-1}.

Then an {r_0}-code in {(M,\rho)} contains at most {N} points.

Proof. Let {X\subset M} be an {r_0}-code. Then

\displaystyle  \begin{array}{rcl}  0 &\le&\sum_{x,y\in X}f(\rho(x,y))\\ &=&\sum_{x\in X}f(\rho(x,x))+\sum_{x\not=y\in X}f(\rho(x,y))\\ &\le& |X| +|X|(|X|-1)(-\frac{1}{N-1}), \end{array}

hence {|X|\le N}. q.e.d.

So a hard task is to guess a clever function {f} and show that it is copositive. The procedure usually works only in very symmetric situations, where {f} turns out to be positive definite.

4. Success stories

Now I describe a few instances where the procedure works.

4.1. The Hamming code {(8,4)}

We view the Hamming cube {M} as a {d}-dimensional vectorspace {\mathbb{F}_2^d} over the field {\mathbb{F}_2} with {2} elements.

The Hamming code {\mathcal{H}_8} is a {4}-dimensional subvectorspace of {\mathbb{F}_2^8}. It is an {r}-code for {r=4}. This is optimal, as can be proved using a positive definite function on {{\mathbb N}} (which turns out to be a polynomial).

4.2. The binary Golay code

{\mathcal{G}_{24}} is a {12}-dimensional subvectorspace of {\mathbb{F}_2^{24}}. It is an {r}-code for {r=8}.

5. Back to Euclidean space

Here, the sought for function on {{\mathbb R}_+} will be a Schwartz function, i.e. it is infinitely differentiable, and it decays, as well as all its derivatives, faster that any power of {\frac{1}{r}}. To such a function {\tilde f}, we associate the radial function {f(x)=\tilde f(|x|)} on {{\mathbb R}^n}.

Theorem 4 (Cohn-Elkies 2003) Suppose that there exists a Schwartz function {\tilde f:{\mathbb R}_+\rightarrow{\mathbb R}} such that

  1. {\forall r\ge r_0}, {f(r)\le 0}.
  2. the Fourier transform {\hat f} is nonnegative on {{\mathbb R}^n}.
  3. {f(0)=\hat f(0)=1}.

Then each periodic configuration {X\subset{\mathbb R}^n} which is an {r_0}-code must have density at most {1}.

Note that the assumption {\hat f\ge 0} is related to being positive definite.

Thus Theorem 4 allows to bound from above the densities of periodic packings. We have little hope to obtain a sharp bound. However, Theorem 4 is merely a generic statement, it can be refined, yielding sharp bounds.

In the same paper, Cohn-Elkies found numerically very good solutions {\tilde f} in dimensions 8 and 24 :

\displaystyle  \begin{array}{rcl}  \Delta_8&\le& 1.00016 \,\Delta_{E_8},\\ \Delta_{24}&\le& 1.019 \,\Delta_{\text{Leech}},\end{array}

where {E_8} and Leech denote very regular periodic configurations which exist only in those dimensions.

6. Sharp results

My contribution is to replace the numerical constant by {1}.

Theorem 5 (Viazovska 2016) There exists a Schwartz function {\tilde f_{E_8}} that satisfies Cohn-Elkies’ assumptions in dimension {8} for {r_0=\sqrt{2}}. This implies that the {E_8} lattice has maximal density among all configurations (periodic or not) in {{\mathbb R}^8}.

Theorem 6 (Cohn-Kumar-Miller-Radchenko-Viazovska 2016) There exists a Schwartz function {\tilde f_{\text{Leech}}} that satisfies Cohn-Elkies’ assumptions in dimension {24} for {r_0=2}. This implies that the Leech lattice has maximal density in {{\mathbb R}^{24}}.

6.1. Further remarks

The method provides upper bounds on {\Delta_d} in other dimensions, but it does not seem to provide sharp bounds in most dimensions (this statement can be made precise).

In 2 dimensions, numerical evidence indicates that there could exist a Schwartz function yielding the sharp bound, but one is unable to prove it yet.

Posted in Uncategorized | Leave a comment

Notes of Corinna Ulcigrai’s Hadamard Lectures, june 2022

Parabolic dynamics

1. Survey

1.1. What does hyperbolic, elliptic mean in dynamics?


Let {\phi_{\mathbb R}=(\phi_t)_{t\in{\mathbb R}}} be a dynamical system. Even if deterministic, it can exhibit a chaotic behaviour. This has several characteristics. One of them is the Butterfly effect: sensitive dependence on initial conditions (SDIC).

Definition 1 A flow {\phi_{\mathbb R}} has SDIC if there exists {K>0} such that {\forall x\in X}, {\forall \epsilon>0}, {\exists y\in X} such that {d(x,y)<\epsilon} and {\exists t\in{\mathbb R}} such that {d(\phi_t(x),\phi_t(y))\ge K}.


A quantitative measurement of SDIC is provided by the dependence of {t} on {\frac{1}{\epsilon}}.

Definition 2 Let {S:{\mathbb R}_{>0}\rightarrow{\mathbb R}_{>0}} be a nondecreasing function. A flow {\phi_{\mathbb R}} has SDIC of order {S} if in the above definition, one can take {t} such that {S(t)\le \frac{1}{\epsilon}}.


This leads us to a rough division of dynamical systems:

  1. Elliptic: no SDIC, or if any, subpolynomial.
  2. Hyperbolic: SDIC is fast, exponential.
  3. Parabolic: SDIC is slow, subexponential.


This trichotomy is advertised in Katok-Hasselblatt’s book.


For instance, entropy is a measure of chaos. Elliptic or parabolic dynamical systems have zero entropy. Hyperbolic dynamical systems have positive entropy.

1.2. Examples of elliptic dynamical systems

  1. Circle diffeomorphisms.
  2. Linear flows on the torus (these two examples are related, one is the suspension of the other).
  3. Billiards in convex domains. Usually, there are many periodic orbits, trapping regions, caustics.

This the realm of Hamiltonian dynamics and KAM theory.

1.3. Examples of hyperbolic dynamical systems

  1. Automorphisms of a torus which are Anosov, i.e. all eigenvalues have absolute values {\not=1}. Then orbits diverge exponentially: if {Av=\lambda v}, {|\lambda|>1}, set {y=x+\epsilon v}. Then {T^n y=T^n x+\epsilon\lambda^n v} mod {{\mathbb Z}^2} and {|\epsilon\lambda^n|} reaches {K=\frac{1}{2}} in time {n} such that {S(n)\le \frac{1}{\epsilon}} for {S(t)=2|\lambda|^t}, an exponential.
  2. Geodesic flows on constant curvature surfaces.
  3. Sinai’s billiard: a rectangle with a circular obstacle. Somewhat equivalent to the motion of two hard spheres on a torus. Scattering occurs after hitting the obstacle, due to its strict convexity.

This is the realm of Anosov-Sinai and others’ dynamics. Structural stability occurs: hyperbolic systems form an open set.

1.4. Examples of parabolic dynamical systems

  1. Horocycle flows on constant curvature surfaces. They were introduced by Hedlund, followed by Dani, Furstenberg, Marcus, Ratner.
  2. Nilflows on nilmanifolds. If {\Gamma<N} is a cocompact lattive in a nilpotent Lie group, let {\phi_{\mathbb R}} be a {1}-parameter subgroup of {N} acting by right translations on {N}. It descends to {X=\Gamma\setminus N}. Both examples are algebraic, this provides us with tools to study them. They are a bit too special to illustrate parabolic dynamics.
  3. Smooth area-preserving flows on higher genus surfaces.
  4. Ehrenfest’s billiard: rectangular, with a rectangular obstacle. Here, SDIC is only caused by discontinuities due to corners. More generally, billiards in rational polygons (angles belong to {\pi{\mathbb Q}}), or equivalently linear flows on translation surfaces. This field is known as Teichmüller dynamics. Forni considers them as elliptic systems with singularities, I prefer to stress their parabolic character.

1.5. Uniformity


Within hyperbolic dynamics, there is a subdivision in uniformly hyperbolic, nonuniformly hyperbolic and partially hyperbolic.


In the same manner, we see horocycle flows as uniformly parabolic, and nilflows as partially parabolic, with both elliptic and parabolic directions. Area-preserving flows on surfaces have fixed points which introduce partially parabolic behaviour: shearing is uniform or not. However, there is no formal definition.

1.6. More examples of parabolic behaviours


Parabolicity is not stable. However, Ravotti has discovered a {1}-parameter perturbation of unipotent flows in {\Gamma\setminus Sl(3,{\mathbb R})}.


A flow {\tilde h_{\mathbb R}} is a time-change of a given flow {h_{\mathbb R}} if there exists a function {\tau:x\times{\mathbb R}\rightarrow{\mathbb R}} such that

\displaystyle \forall x\in X,~\forall t\in{\mathbb R},\quad \tilde h_t(x)=h_{\tau(x,t)}(x).

For {\tilde h_{\mathbb R}} to be a flow, it is necessary that {\tau} be a cocycle.


Both flows have the same trajectories. A feature of parabolic dynamics is that a typical time-change {\tilde h_{\mathbb R}} is not isomorphic to {h_{\mathbb R}} and has new chaotic features. Indeed, an isomorphism would solve the cohomology equation, and there are obstructions.

1.7. Program


Study smooth time-changes of algebraic flows.


Goes back to Marcus in the 1970’s. Algebraic tools break down, softer methods are required: geometric mechanisms. Also, we expect the features exhibited by time changes to be more typical.

2. Chaotic properties

2.1. Definitions

Definition 3 Let {(X,\mathcal{A},\mu)} be a measure space with finite measure. Let {\phi_{\mathbb R}} be a measure preserving flow. The trajectory of a point {x\in X} is equidistributed with respect to {\mu} if for every smooth observable {f:X\rightarrow{\mathbb R}},

\displaystyle \frac{1}{T}I_T(f,x):=\frac{1}{T}\int_{0}^{T}f(\phi_t(x))\,dt tends to {0} as {T} tends to {\infty}.

{\phi_{\mathbb R}} is ergodic if {\mu} almost every {x\in X} has equidistributed orbit with respect to {\mu}.

This is Boltzmann hypothesis.

Definition 4 Say {\phi_{\mathbb R}} is mixing if for all {f,g\in L^2(X,\mu)}, the correlation

\displaystyle \mathcal{C}_{f,g}(t):=\int f\circ\phi_t \,g\,d\mu-(\int f\,d\mu)(\int g\,d\mu) tends to {0} as {t} tends to {\infty}.

This means decorrelation of functions. This implies ergodicity.


The speed at which decorrelation occurs is a significative feature too.

Definition 5 The speed of mixing is a function {S:{\mathbb R}_+\rightarrow{\mathbb R}_+} such that for all smooth observables {f,g:X\rightarrow{\mathbb R}}, the correlation decays at speed {S}, i.e.

\displaystyle \mathcal{C}_{f,g}(t)=O(S(t)) as {t} tends to {\infty}.

2.2. Relation to the trichotomy


Elliptic systems often are not ergodic, but even when they are, they are not mixing.


Hyperbolic and parabolic systems can be mixing, but at different speeds:

  • hyperbolic {\Rightarrow} exponential decay of correlations.
  • in parabolic systems, we expect that, if mixing occurs, the decay of correlations is slower: polynomial or subpolynomial.


A related concept is that of polynomial deviations of ergodic averages: if {f} is smooth and has vanishing integral, {\phi_{\mathbb R}} s ergodic and {x} is an equidistribution point,

\displaystyle |I_T(f,x)|=O(T^\alpha) \quad \text{for some}\quad 0<\alpha<1,

and no faster. This phenomenon was first discovered on horocycle flows, then Teichmüller flows (Zorich, experimentally, Kontsevitch-Zorich for a proof).

2.3. Other features


Spectral properties. Let {U_t} be the operator {f\mapsto f\circ\phi_t} on {L^2}. Then {U_t} is unitary. What is its spectrum?


Disjointness of rescalings. Rescaling means linear time change {\tilde h_t=h_{kt}}.

3. Results


Horocycle flow is mixing (Ratner). The spectrum is Lebesgue absolutely continuous.


Time changes of horocycle flows are mixing (Marcus, by shearing). This can be made quantitative (Forni-Ulcigrai).


Disjointness of rescalings fails for horocycle flows, but hold for nontrivial time changes (Kanigowski-Ulcigrai and Flaminio-Forni). The spectrum is Lebesgue absolutely continuous as well.


Nilflows themselves are not mixing, but typical time changes of nilflows are mixing (Avila-Forni-Ravotti-Ulcigrai).


We shall see geometric mechanisms at work:

  • Mixing via shearing.
  • Ratner property of shearing.
  • Renormizable parabolic flows.
  • Deviations of ergodic integrals.

4. Horocycle flows


Today’s goal is to explain the technique of mixing by shearing.

4.1. Algebraic viewpoint


Consider {G=PSl(2,{\mathbb R})} and its subgroups

\displaystyle N=\{h_s:=\begin{pmatrix} 1 & s \\ 0 & 1 \end{pmatrix}\,,s\in{\mathbb R}\}, \quad N=\{h^-_s:=\begin{pmatrix} 1 & 0 \\ s & 1 \end{pmatrix}\,,s\in{\mathbb R}\},

\displaystyle A=\{g_t:=\begin{pmatrix} e^{t/2} & 0 \\ 0 & e^{-t/2} \end{pmatrix}\,;\,t\in{\mathbb R}\},\quad K=\{r_\theta :=\begin{pmatrix} \cos\theta & -\sin\theta \\ \sin\theta & \cos\theta \end{pmatrix}\,;\,\theta\in {\mathbb R}/2\pi{\mathbb Z}\}.

Every matrix can be uniquely written {g=h_s g_t r_\theta}, hence {G=NAK}. The factors do not commute, because of the key relation

\displaystyle g_t h_s = h_{e^{t}s} g_t.

The key relation can be interpreted as a selfsimilarity property: {h_\mathbb{R}} is a fixed point of renormalization by {g_{\mathbb R}}.


Take a discrete and cocompact subgroup {\Gamma<G}. Then {A} and {N} act on {X=G/\Gamma} by left multiplication. The key relation implies that the rescaled flow $latex {h_{\mathbb R}^k=(h_{ks})_{s\in{\mathbb R}}}&fg=000000$ is conjugated to $latex {h_{\mathbb R}}&fg=000000$. This fails for other (nonrescaling) time changes, as I proved recently with Fraczek and Kanigowski.

4.2. Geometric viewpoint


Let {\mathbb{H}} denote the upper half plane, with metric {ds^2=\frac{dx^2+dy^2}{y^2}}.


Fact. {G} acts isometrically and transitively on {\mathbb{H}}, and this yields a diffeomorphism of {G} with the unit tangent bundle {T^1 \mathbb{H}}.


Indeed, the action is by Möbius transformations

\displaystyle A=\begin{pmatrix} a & b \\ c & d \end{pmatrix}\mapsto (z\mapsto \frac{az+b}{cz+d}),

on {\mathbb{H}}, and by their derivatives on {T^1 \mathbb{H}}.


With this identification, orbits of {A} are curves which project to geodesics of {\mathbb{H}}, and coincide with their lifts by their unit speed vector. On the other hand, orbits of {N} are curves which projects to horocycles of {\mathbb{H}}, and coincide with their lifts by their unit normal outward pointing vectors. Lifts by inward pointing normal vectors are orbits of {N^-}.


{A=g_{\mathbb R}} is a hyperbolic flow: it contracts in the direction of {N^-}-orbits, it dilates in the direction of {N}-orbits,

4.3. Classical results


Let {\mu} denote Haar measure on {G}. It maps via {G\rightarrow T^1\mathbb{H}\rightarrow\mathbb{H}} to hyperbolic volume. Consider the induced measure on {X=G/\Gamma} (still denoted by {\mu}). Then {\mu(X)<+\infty} and {A} acts on {X} by measure preserving transformations.


Then

  1. {A} is ergodic (Hopf). It is far from being unique ergodic (plenty of periodic orbits).
  2. {A} is mixing.
  3. {N} is uniquely ergodic (Furstenberg). This means that every orbit is equidistributed with respect to {\mu}.

5. Shearing


This is an alternate way to prove mixing. The idea goes back to Marcus (Annals of Math. 1977). Marcus covered a more general situation, and proved mixing of all orders (i.e. for multicorrelations, integrals involving an arbitrarily large number of functions).


Let us shift viewpoint on the key relation. Let {\gamma} be a piece of {A}-orbit of length {\sigma}. Let

\displaystyle \gamma^s:=h_s\circ\gamma=\{h_s g_t x\,;\,0\le t \le \sigma\}.

Then {\gamma^s} is sheared or tilted in the direction of the geodesic flow.


Key idea in parabolic dynamics: In several parabolic systems, the Butterfly effect happens in a special way, e.g. shearing. Points nearby move parallel, but with different speeds. This implies that transverse arcs shear.

5.1. Recipe for mixing in parabolic dynamics


Here are the ingredients:

  • A uniquely ergodic flow {\phi_{\mathbb R}}.
  • A transverse direction which is sheared in the direction of the flow.


By assumption, for every {x\in X}, the trajectory {\phi_{{\mathbb R}_+}(x)} equidistributes with respect to {\mu}. We want to upgrade it to mixing, which is a property of sets: indeed

\displaystyle \int f\circ\phi_t \,g\,d\mu \rightarrow (\int f\,d\mu)(\int g\,d\mu)

is equivalent to

\displaystyle \forall A,B\in\mathcal{A},\quad\mu(\phi_t(A)\cap B)\rightarrow \mu(A)\mu(B)

i.e {\phi_t(A)} equidistributes as {t\rightarrow+\infty}.


The idea is to cover {A} by short arcs in the transverse direction. We prove that each such arc equidistributes, and apply Fubini. I.e. if {A=\bigcup \gamma_\alpha}, apply {\phi_t}. Then {\phi_t(A)=\bigcup\phi_t(\gamma_\alpha)}.


Each {\phi_t(\gamma_\alpha)} becomes close to a long piece of orbit of {\phi_{\mathbb R}}. By unique ergodicity, that piece equidistributes, and this implies equidistribution for {\gamma_\alpha}.

5.2. Time change


Let {\tau:X\times{\mathbb R}\rightarrow {\mathbb R}} be a smooth time change, which is a cocycle with respect to a given smooth flow {h_{\mathbb R}}, i.e.

\displaystyle \tau(x,t+s)=\tau(x,t)+\tau(h_t(x),s).

We are interested in the flow {\tilde h} defined by

\displaystyle h_t(x)=\tilde h_{\tau(x,t)}(x).

The generator of {\tau} is

\displaystyle \alpha(x):=\frac{\partial \tau(x,t)}{\partial t}_{|t=0}.

It is a smooth nonnegative function on {X}. We assume that {\int \alpha\,d\mu=1}. We denote {\tilde h} by {h^\alpha_{\mathbb R}}.


Remark. If {h_{\mathbb R}} is generated by a smooth vectorfield {U}, then {\tilde h_{\mathbb R}} is generated by the vectorfield {\frac{1}{\alpha}U}.

Theorem 6 (Forni-Ulcigrai) Let {h_{\mathbb R}} be the horocyclic flow of a compact constant curvature surface. For any smooth function {\alpha}, the flow {h^\alpha_{\mathbb R}} is mixing, with quantitative estimates which imply that the spectrum is absolutely continuous with respect to Lebesgue measure.

Lemma 7 Let {X} denote the generator of {g_{\mathbb R}} and {U} the generator of {h_{\mathbb R}}. Take a segment of {g_{\mathbb R}}-orbit

\displaystyle \gamma=\{g_t x\,;\,0\le t \le \sigma\}. Let \displaystyle \gamma_s:=h^\alpha_s\circ\gamma. Then \displaystyle \frac{d\gamma_s}{dt}=v_s(x,t)U_\alpha +X, where \displaystyle v_s(x,t)=\int_{0}^{s}(\frac{X\alpha}{\alpha}-1)h^\alpha_\tau\circ g_t(x)\,d\tau.


Remark. {\int \frac{X\alpha}{\alpha}\,d\mu=0}, hence {\int(\frac{X\alpha}{\alpha}-1)\,d\mu=-1}. Thus, as {s} tends to {+\infty}, {\frac{1}{s}v_s(x,t)} tends to a finite limit, the shear rate.


Proof of Lemma. It relies on {[U,X]=U}, which implies that {[U_\alpha,X]=(\frac{X\alpha}{\alpha}-1)U_\alpha}.


We see that we need compute integrals over sheared arcs {\gamma_s}. Let {f:X\rightarrow{\mathbb R}} be a smooth function. Then

\displaystyle \int_{0}^{s}f(h^\alpha_\tau\circ g_t(x)\,dt =\int_{0}^{s}f(h^\alpha_\tau\circ g_t(x)\frac{v_s(x,t)}{s}\,dt+\int_{0}^{s}f(h^\alpha_\tau\circ g_t(x))(\frac{v_s(x,t)}{s}+1)\,dt.

The second term is an ergodic integral which is easy to handle. The main term is the first term, which can be rewritten

\displaystyle -\frac{1}{s}\int_{\gamma_s}f\hat U_\alpha,

where {i_{U_\alpha}\hat U_\alpha=1}, {i_X \hat U_\alpha=0}.


To deduce mixing, one must estimate {L^2} inner products {\langle f\circ h^\alpha_s,a\rangle}. We integrate by parts

\displaystyle \langle f\circ h^\alpha_s,g\rangle =\frac{1}{\sigma}\int_{0}^{\sigma}\langle f\circ h^\alpha_s\circ g_t,g\circ g_t\rangle\,dt

\displaystyle =\frac{1}{\sigma}\int_{0}^{\sigma}\langle f(h^\alpha_s\circ g_t\,dt,g\circ g_\sigma\rangle\,dt

\displaystyle -\frac{1}{\sigma}\int_{0}^{\sigma}\langle\int_{0}^{s}f\circ h^\alpha_s\circ g_t(x)\,ds,(L_Xg)\circ g_t\rangle\,dt.

The second term is again an ergodic integral that tends to {0}.

5.3. Quantitative equidistribution estimates


We are interested in ergodic integrals of the form

\displaystyle I_T(f,x)=\int_{0}^{T}f\circ h^\alpha_s(x)\,ds.

Flaminio-Forni treat the un-time-changed case {\alpha=1} and show that

\displaystyle \|I_T(f,x)\|_{L^\infty}\le C\, T^{(1+\nu_0)/2}

for some {0<\nu_0\le 1}. One can adapt their arguments, using estimates by Bufetov-Forni, to the time-changed case, and get similar estimates.

5.4. Additional references


Marcus original technique already proved mixing. His setting was Anosov flows, with their stable and unstable foliations. From these, a flow {h_{\mathbb R}} can be defined, which satisfies

\displaystyle g_t \circ h_s =h_{ss^*(t,s,x)}\circ g_t,

where $latex {s^*}&fg=000000$ has a continuous mixed partial second derivative $latex {\frac{\partial s^*}{\partial t \partial s}}&fg=000000$. So we see that time-changes were already in the picture.


Kushnirenko was able to prove mixing for smooth time-changes, assuming

\displaystyle (KC)\quad\quad \|\frac{X\alpha}{\alpha}\|_{\infty}<1.

Thus small time-changes are mixing. What about larger ones? This is still open.


Tiedra de Aldecoa uses a different method to prove absolute continuity of the spectrum for time changes satisfying (KC).


Generalizations. The setting is algebraic dynamics: a unipotent {1}-parameter subgroup acting on {G/\Gamma}, {G} semisimple Lie group.


Lucia Simonelli (Forni’s student) could prove absolute continuity of the spectrum for time changes satisfying (KC).


Davide Ravotti (my student) could prove quantitative mixing.


Kanigowski and Ravotti could prove quantitative {3}-mixing.

6. Heisenberg nilfows


Let {H} denote the Lie group of unipotent {3\times 3} matrices, with {X,Y,Z} as standard generators of its Lie algebra, {Z=[X,Y]}. Let {\Gamma<H} be a discrete cocompact lattice (for instance, unipotent matrices with integer entries). We call {X=\Gamma\setminus H} the Heisenberg nilmanifold. {H} acts on {X} by right multiplication. The action of a {1}-parameter subgroup is called a nilflow.

6.1. Classical results


Auslander-Green-Hahn (1963) studied unique ergodicity of nilflows. They showed that is {W\in Lie(H)} can be written {W=aX+bY+cZ}, for the nilflow defined by {W},


unique ergodicity {\Leftrightarrow} ergodicity {\Leftrightarrow} minimality {\Leftrightarrow} {a,b} and {1} are rationally independent.


Rational independence means that no linear relation with nonzero integral coefficients {ka+\ell b+ n=0} can hold.


In other words, if we project the situation to the {2}-torus {\bar X=\bar\Gamma\setminus\bar H} where {\bar H=H/[H,H]}, {\bar\Gamma=}, then the flow {\phi^W_{\mathbb R}} projects to a flow {\bar \phi_{\mathbb R}} on {\bar X}, and


unique ergodicity for {\phi_{\mathbb R}} {\Leftrightarrow} unique ergodicity for {\bar\phi_{\mathbb R}}.


Here, we have used a theorem of Furstenberg on skew-products of rotations of the circle.

Definition 8 For real numbers {\alpha,\beta\in{\mathbb R}}, let {f_{\alpha,\beta}} be the diffeomorphism of the {2}-torus defined by

\displaystyle f_{\alpha,\beta}(x,y)=(x+\alpha,y+x+\beta) \mod {\mathbb Z}^2.


In general, a skew-product over a map {T:X\rightarrow X} is a map {f:X\times Y\rightarrow X\times Y} which is fiber-preserving (with respect to the projection {X\times Y\rightarrow X}) and the permutation of fibers is given by {T}. In the example at hand, {f} is isometric on fibers.

6.2. First return map

Lemma 9 Assume that the Heisenberg nilflow {\phi^W_{\mathbb R}} is uniquely ergodic. There is a transverse submanifold {\Sigma\subset X}, diffeomorphic to a torus, such that the Poincaré return map {P:\Sigma\rightarrow\Sigma}, given by {P(g)=\Phi^W_{r(g)}(g)}, {r} the first return time to {\Sigma}, is one of the Furstenberg skew-products {f_{\alpha,\beta}}.


Proof of the Lemma. {\Sigma} lifts to a vertical plane {{\exp(xX+zZ)\,;\,x,z\in {\mathbb R}}} in {H}. Since {X} and {Z} commute, {\Sigma} is diffeomorphic to a torus. If {\phi^W_{\mathbb R}} is uniquely ergodic, {b\not=0}, so {\Sigma} is transverse to the flow.


We show that {\frac{1}{b}} is a return time. We use the fact that {\exp(-Y)\in\Gamma}. So using the Campbell-Hausdorff-Dynkin formula, we compute

\displaystyle \exp(-Y)\exp(xX+zZ)\exp(\frac{W}{b}) =\exp((x+\frac{a}{b})X+(z+x+\frac{c}{b}+\frac{a}{2b})Z).

This point belongs to {\Sigma}, so {P(x,z)=f_{\frac{a}{b},\frac{c}{b}+\frac{a}{2b}}(x,z)}.

6.3. Lack of mixing


The above Lemma shows th at we can now focus on Furstenberg skew-products. We shall see that the parameter {\beta} plays no role, so we focus on {f_{\alpha,0}}.

Definition 10 Given a map {f:X\rightarrow X} and a function {\Phi:X\rightarrow{\mathbb R}} (called the roof function), we define the special flow {f^\Phi_{\mathbb R}} over {f} under {\Phi} as follows: it is a flow on an {X}-bundle {Y} over the circle, the quotient of the vertical unit speed flow on {X\times{\mathbb R}} under the identification

\displaystyle (x,t)\sim(f(x),t+1).


Fact. If a flow {f_{\mathbb R}} admits a global Poincaré section {\Sigma} with first return time {r}, then {f_{\mathbb R}} is isomorphic to the special flow of the first return map with roof function {r}.


In the case at hand, the roof function is constant. Therefore, the special flow is not mixing: if {\bar A\subset X\times I}, {I} a short interval, so do its images by the vertical unit speed flow, and so do their projections to {Y}, which are the images of a set {A\subset Y} and its images by the special flow.


This shows that Heisenberg nilflows are never mixing.

7. Mixing time-changes


The following contents can be found in Avila-Forni-Ulcigrai. A recent generalization to all step 2 nilflows can be found in Avila-Forni-Ravotti-Ulcigrai. Ravotti has treated filiform nilflows.


Let {\phi_{\mathbb R}} be a Heisenberg nilflow and {\alpha} a smooth positive function on {X}. Let

\displaystyle \tau(g,t)=\int_0^t \alpha(\phi_s(g))\,ds.

Then the time-change is given by

\displaystyle \phi^\alpha_t=\phi_{\tau(g,t)}.

7.1. Time-changes versus special flows


We have seen that Heisenberg nilflows {\phi_{\mathbb R}} are special flows with constant roof function. The time-change {\phi_{\mathbb R}^\alpha} is again a special flow, over the same skew-product, but with roof function

\displaystyle \Phi(g)=\tau(g,r(g)).

7.2. Trivial time-changes


Beware that there exist smooth time-changes which are trivial, i.e. smoothly conjugate to the original nilflow.


In general, adding a coboundary {u\circ f-u} to the roof function of a special flow produces an isomorphic flow. This leads us to the following problem: understand cohomology of nilflows. Here are our ultimate results.

Theorem 11 Let {f} be a There exists a dense set {\mathcal{R}} in {C^{\infty}(T^2)} of roof functions, and a vectorspace {\mathcal{T}_f\subset\mathcal{R}} of countable dimension and codimension, such that if a roof function {\Phi} is chosen in {\mathcal{M}_f:=\mathcal{R}\setminus\mathcal{T}_f}, the corresponding special flow {f^\Phi} is mixing.

Moreover, for {\Phi\in\mathcal{R}},

\displaystyle \Phi\in \mathcal{M}_f \Leftrightarrow f^\Phi \text{ is not smoothly trivial}.

In fact, Katok has found a nice characterization of which {f^\Phi} are smoothly trivial. I will come back to this next week. Today, I merely give one example.


Example. {\Phi(x,y)=\sin(2\pi y)+2} is a smoothly trivial roof function.


Under the assumption that {\alpha} has bounded type, Kanigowski and Forni have proved quantitative mixing.

7.3. Idea of proof


We start from a Furstenberg skew-product {f_{\alpha,0}} and play with roof functions. We use again mixing by shearing. We consider intervals in the fiber (i.e. in the {y} direction). We shall see that many of them shear in the flow ({z}) direction. But there are intervals which do not shear or shear in the other direction.

7.4. Special flow dynamics


Given a point {(x,y)} in the torus and {t>0}, we compute {f^\Phi_t(x,y)}. When {t} is large, we join bottom to roof several times. Let

\displaystyle \Phi_n(x,y)=\sum_{k=0}^{n-1}\Phi(f^k(x,y))

and

\displaystyle n_t(x,y):=\max\{n\,;\,\Phi_x(x,y)<t\}.

Then

\displaystyle f^\Phi_t(x,y)=(f^{n_t(x,y)}(x,y),t-\Phi_{n_t(x,y)}(x,y)).

In order to exhibit shearing, we want to see how this changes in {y}. Since {f} is an isometry in the {y}-direction, the {y}-derivative of the sum {\Phi_n} is the sum of {y}-derivatives, i.e.

\displaystyle \frac{\partial\Phi_n}{\partial y}=(\frac{\partial\Phi}{\partial y})_n.


Take {\mathcal{R}=} trigonometric polynomials on the torus, which are positive.


Given {\Phi\in\mathcal{R}}, let {\phi=\Phi} minus its average on the {y}-fiber. Define {\mathcal{M}_f} as the set of {\Phi} such that {\phi} is not a measurable coboundary.


Step 1. Since {\phi} is not a measurable coboundary, the sums {\phi_n} must grow,

\displaystyle \forall C>0,\quad \mathrm{Lebesgue}(\{(x,y)\,;\,|\phi_n|(x,y)>C\})\rightarrow 1.

This relies on a result by Gottschalk-Hedlund, plus decoupling.


Step 2. The sums {\phi_n} are trigonometric polynomials of bounded degree. It follows that

\displaystyle \forall C>0,\quad \mathrm{Lebesgue}(\{(x,y)\,;\,|\frac{\partial\phi_n}{\partial y}|(x,y)>C\})\rightarrow 1.


There can be flat intervals where no stretch occur, so one must throw them away. But the larger {t}, the shorter these intervals are.


Step 3. We use a polynomial bound on level sets of trigonometric polynomials.


The next class will deal with renormalization, deviations of ergodic averages. Only later shall we get back to Katok’s characterization of smoothly trivial special flows and to nilflows.

8. Short recap


Parabolicity is (a bit heuristically) defined by slow butterfly effect. This can be formalized for smooth flows, in terms of growth of derivatives under iteration. It is not that easy to build examples.


Presently, parabolic flows is the following list of examples,

  • Horocycle flows of compact constant curvature surfaces.
  • Unipotent flows (a generalization of the above).
  • Nilflows and their time-changes.
  • Smooth area-preserving flows.
  • Linear flows on flat surfaces with conical singularites.

The two first are uniformly parabolic. The next is partially parabolic. The fourth is nonuniformly parabolic, the last is elliptic with singularities.


We have proven mixing via the technique of shearing.


Here is a further example where this technique works, due to B. Fayad, of an elliptic flavour. Start with a linear flow on the {n}-torus. If {n\geq 3}, Fayad has been able to construct anaytic time-changes which are mixing.


Such examples are very rare (they rely on parameters being very Liouville numbers).

9. Renormalization


The word here is taken in a meaning which differs from its use in holomorphic dynamics (not to speak of quantum mechanics).


The idea is to analyze systems which are approximately self-similar, and exhibit several time scales.


We introduce the renormalization flow {\mathcal{R}_t} which rescales: long trajectories become short. Given a map {T}, one way to zoom in is to restrict {T} to a subspace {Y\subset X} and replace {T} with the first return map to {Y}. Eventually rescale space afterwards. But there are other means.

9.1. A series of examples


Example. Start with the horocycle flow {h_{\mathbb R}}. The key relation is

\displaystyle g_t h_s = h_{e^{t}s} g_t.

Applying the geodesic flow to a length {e^t} trajectory {\gamma_{e^t}}, we get a trajectory of length {1}. So the geodesic flow achieves renormalization, on the same space, with no effort.


Example. The cat map {\psi_A} associated with the matrix {A=\begin{pmatrix} 2 & 1 \ 1 & 1 \end{pmatrix}} on the {2}-torus. Let {\lambda_1>1}, {\lambda_2<1} denote the eigenvalues, {v_1,v_2} the eigenvectors. Let {\phi_{\mathbb R}} be the linear flow in direction {v_1}, with unit speed. This is an elliptic flow.


Put {T_n=\lambda_1^n}. Let

\displaystyle \gamma_{T_n}=\{\phi_t(x)\,;\,0\le t<T_n \}.

Put {\mathcal{R}=\psi_A^{-1}}. Then {\mathcal{R}} maps {\gamma_{T_n}} to a trajectory of {\phi_{\mathbb R}} of length {1}.


Express {\psi_A} as the composition of two Dehn twists,

\displaystyle A=\begin{pmatrix} 2 & 1 \\ 1 & 1 \end{pmatrix}=\begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}\begin{pmatrix} 1 & 0 \\ 1 & 1 \end{pmatrix}.

Rotate coordinates so that eigenvector {v_1} becomes vertical. Then {\mathcal{R}} acts by a diagonal matrix in {SL_2({\mathbb R})}, which was denoted by {g_t}, {e^t=\lambda_1} earlier.


Example. Higher genus toy model.


View a genus {2} surface {S} as a regular octagon in the Euclidean plane with edge identifications. Fix a direction {\theta}, consider the linear flow in direction {\theta}. It descends to a flow {\phi_{\mathbb R}} on the surface which is not well-defined at the vertex


This generalizes to translation surfaces, made of a disjoint union of polygons, with identifications given by translations. The notion of a direction {\theta} is well-defined on the surface (except at finally many singularities), whence a flow {\phi_{\mathbb R}} with singularities.


Consider matrix

\displaystyle P^+=\begin{pmatrix} 1 & 2(1+\sqrt{2}) \\ 0 & 1 \end{pmatrix}.

It applies a shear on the original regular octagon. Since {1(1+\sqrt{2})=2\cot(\frac{\pi}{8})}, This linear map induces a homeomorphism of the surface {S}. Let us repeat in different direction (rotation by 45 degrees), get matrix {P^-}. Let {A=P^+P^-}, this is a hyperbolic matrix with eigenvalues {\lambda_1>1>\lambda_2}. We get again an affine automorphism {\psi_A} of the surface. Let {\phi_{\mathbb R}} denote the linear flow in the direction of the eigenvector {v_1}. Then


B. Veech has shown that the affine group of the surface is generated by {\psi_A} and the order {8} rotation. This is as large as the affine group of a translation surface can be. Therefore, {S} is clled the Veech surface.


For almost every direction {\theta} (the condition is that {\tan(\theta)\notin{\mathbb Q}(\sqrt{2})}), there exists a sequence of hyperbolic automorphisms {\psi_n} with unstable direction {\theta_n} that converge to {\theta}. {\psi_n} can be used to renormalize {\phi^\theta_{\mathbb R}}.


Example.


For a slightly deformed octagon {O'}, the affine group of the corresponding surface {S'} is trivial. Nevertheless, almost every linear flow {\phi^\theta_{\mathbb R}} is still renormalizable: there is a sequence of surfaces {S_n} and affine hyperbolic morphisms {\psi_n:S'\rightarrow S_n} with expanding direction {\theta_n} converging to {\theta}.


Indeed, consider the space {\mathcal{C}} of linear flows on translation surfaces of genus {2}. The flow of diagonal matrices acts on this space, generating a flow {\mathcal{R}_{\mathbb R}} on {\mathcal{C}}, known as the Teichmüller flow.

Theorem 12 (Masur-Veech) {\mathcal{R}_{\mathbb R}} is recurrent.

Therefore for almost every linear flow, there exists a sequance {t_n\rightarrow\infty} such that {\mathcal{R}_{t_n}(S,\theta)} tends to {(S,\theta)}.

9.2. What is renormalization good for?


It is used to put diophantine conditions on linear flows in higher genus. See my ICM 2022 talk (watch it on line on july 11th). I do not pursue this topic further.


It is used to study deviations of ergodic averages. Assume a flow {\phi_{\mathbb R}} is uniquely ergodic. Ergodic integrals take the form

\displaystyle I_T(f,x):=\int_0^T f(\phi_t(x))\,dt.

By the ergodic theorem, {\frac{1}{T}I_T(f,x)\rightarrow\int f\,d\mu} for every {x}. Let us focus on our favourite example, the Veech surface. The invariant measure is area in the plane. Unique ergodicity is a theorem of Masur, Kerckhoff-Masur-Smillie


We show that for functions with vanishing average,

\displaystyle I_T(f,x)=O(T^\alpha),

for some {0<\alpha<1} (discovered by Zorich, conjectured by Kontsevitch-Zorich, proven by Forni).


Fix a basis {\tilde\gamma_1,\ldots;\tilde\gamma_4} of {H_1(S,{\mathbb Z})}. Fix a section {\Sigma} of the linear flow. Take trajectories {\gamma_1,\ldots,\gamma_4} from {\Sigma} and to their first return to {\Sigma}. Closing them by segments of sigma gives representatives of the basis of homology. Let

\displaystyle \gamma_i^{n}=\psi_A^n(\gamma_i).

Let {B\in \mathrm{Sp}(4,{\mathbb Z})} be the matrix expressing the homology basis {\psi_A(\tilde\gamma_1,\ldots,\tilde\gamma_4)} is the initial basis. Let {\lambda_1>\lambda_2>1>\lambda_3=1/\lambda_2 > \lambda_4=1/\lambda_1} denote the eigenvalues of {B}. By continuity of {f}, up to a small error,

\displaystyle \int_{0}^{\mathrm{length}(\gamma_{j}^{n})}f(\phi_t(x))\,dt=\int_{\gamma_{j}^{n}}f.

Let us start from a large {n_0} instead of {0},

\displaystyle \int_{\gamma_{i}^{n}}f=\sum_{j} B^{n-n_0}_{ij}\int_{\gamma_{j}^{n_0}}f:=(B^n \rho)_i.

For simplicity, let us assume that {\rho} is the eigenvector {v_1}. Then

\displaystyle \int_{\gamma_{i}^{n}}f\ge\min \{|v_1|,\|B^n\|\},

a contradiction. Therefore {\rho} must be a combination of {v_2,v_3,v_4}. Thus

\displaystyle |\int_{\gamma_{i}^{n}}f|\le\|B^n\rho\|\le \lambda_2^n.

On the other hand, {T=\mathrm{length}(\gamma_{j}^{n})} grows like {\lambda_1^n}. Let {\alpha} satisfy

\displaystyle \lambda_1^\alpha=\lambda_2.

Then {I_T(f,x)=O(T^\alpha)}, as announced.


Of course, I cheated a bit, some more regularity of {f} is needed.


In that example, the fact that {\lambda_1\not=\lambda_2} can be explicitly. Tomorrow, I will mention results on this for general translation surfaces.


And then, back to nilflows.

9.3. Area preserving flows on surfaces


{S} compact connected oriented surface of genus {g>1}. Let {\phi_{\mathbb R}} be a flow which preserves a smooth measure {\mu}. Note that {\phi_{\mathbb R}} has fixed points, which are all of saddle type. In the next theorem, the number and types of fixed points are fixed.

Theorem 13 (Zorich, Forni, Avila-Viana) There exists {g} distinct positive exponents {1=\alpha_1 >\cdots>\alpha_g>0} such that almost every choice of {\phi_{\mathbb R}} (Katok’s fundamental class, described in terms of periods) is uniquely ergodic, and for all smooth functions {f},

\displaystyle I_T(f,x)=(\int f\,d\mu)T+\mathcal{D}_2(f)O(T^{\alpha_2})+\cdots+\mathcal{D}_g(f)O(T^{\alpha_g})+ O(T^\epsilon), for all {\epsilon>0}.

In this statement, {O(T^\alpha)} means a function {u} such that

\displaystyle \limsup_{T\rightarrow+\infty}\frac{\log u}{\log T}=\alpha.

Furthermore, {\mathcal{D}_j} is a distribution on {S}, and {\mathcal{D}_2(f)\not=0}.


This type of behavior is called a power deviation spectrum. It was conjectured by Kontsevitch and Zorich, then proved by Zorich in 1997 for special functions {f}, with only one term {\alpha_2}. Then Forni obtained a proof in 2002 for functions with support away from fixed points, up to distinctness of exponents which was proven by Avila-Viana. Bufetov gave a more precise version, transforming the result into an asymptotic expansion. With Fraczek, we gave a different proof based on Marmi-Moussa-Yoccoz). Finally, Fraczek-Kim could handle generic saddles. The expansion then involves extra terms depending on saddles (and not on {f}).

9.4. Idea of proof


Choose coordinates such that {\Phi_{\mathbb R}} appears as a time-change of a linear flow on a translation surface. The time change is smooth only away from fixed points. If function {f} has support away from fixed points, one is reduced to study deviation for linear flows. For a general linear flow, a renormalization is given by a sequence of matrices {B_n\in Sp(2g,{\mathbb Z})}, the Kontsevitch-Zorich cocycle. Eigenvalues are replaced with ratios of Lyapunov exponents.

10. More on renormalization

10.1. Back to nilflows


Let {ASl(2,{\mathbb R})} denote the stabilizer of vector {(1,0,0)} is {Sl(3,{\mathbb R})}. Defined {\mathcal{R}_t} by

\displaystyle X\mapsto e^{-t}X,\quad Y\mapsto e^t Y,\quad Z\mapsto Z.

Then {\mathcal{R}_t} is recurrent, its has been used as a renormalization by Flaminio-Forni, in order to prove polynomial deviations of ergodic averages.

10.2. Renormalizable parabolic flows


Horocycle flows: yes.


Unipotent flows: unknown.


Heisenberg nilflows: yes.


Higher step nilflows: unknown in general. Special case (special flows over skew products) studied by Flaminio-Forni. In that case, the maps {\mathcal{R}_t} diverge.


Linear flows over higher genus surfaces (they are smooth and area-preserving).


More general smooth flows on surfaces (not necessarily area preserving). Then {\mathcal{R}_t} typically diverges. This is related to generalized interval exchange transformations.

11. Isomorphisms between time-changes


Recall that a time-change {\tilde\phi_{\mathbb R}} of a flow {\phi_{\mathbb R}} is

\displaystyle \tilde\phi_t(x)=\phi_{\tau(x,t)}(x).

When do such a change lead to a genuinely different, nonisomorphic flow?

11.1. Setting of special flows


Let {f:Y\rightarrow Y} be a map. Given a roof function $latex {\Phi:Y\rightarrow{\mathbb R}_{>0}}&fg=000000$, the flow of translations on {Y\times{\mathbb R}} descends to a flow $latex {\psi^{f,\Phi}_{\mathbb R}}&fg=000000$ on the quotient space

\displaystyle X=(Y\times{\mathbb R})/\sim, \quad \text{where}\quad (y,z)\sim(f(y),z+\Phi(y).

Lemma 14 Let {\psi_1,\psi_2} be special flows over the same map {f}, under roofs {\Phi^1} and {\Phi^2}. If there exists a function {u:Y\rightarrow{\mathbb R}} such that

\displaystyle \Phi^2=\Phi^1+u\circ f-u, then the two special flows are isomorphic.


Indeed, look for a conjugating homeomorphism of the form

\displaystyle (y,z)\mapsto (y,z+u(y)).

It commutes with translations and maps one equivalence relation to the other.

11.2. Cohomological equations


The operator {u\mapsto u\circ f-u} is called the coboundary operator. Hence the equation {u\circ f-u=\Phi} with unknown {u} is called a cohomological equation. It sometimes appears with a twist: {\lambda u\circ f-u=\Phi}, for some {\lambda\in{\mathbb R}}.


There are obvious obstructions.


If {x} is a periodic point, i.e. {f^n(x)=x}, then

\displaystyle \sum_{k=0}^{n-1}\Phi(f^k(x))=0.

So every periodic orbit gives an obstruction. If {f} is hyperbolic, it is essentially the only one.


More generally, if {\mu} is an invariant measure, then

\displaystyle \int \Phi\,d\mu=0.

So every invariant measure gives an obstruction.

11.3. An elliptic example


If {f} is elliptic, this is sometimes sufficient, e.g. for circle rotations {R_\alpha}: if {\Phi} is a trigonometric polynomial and {\int \Phi(x)\,dx=0}, then there exists a solution {u}. If {\Phi} is smooth, a Diophantine condition on {\alpha} is required in addition for the solution {u} to be smooth. Indeed, in Fourier, if

\displaystyle \Phi(x)=\sum_{k\in{\mathbb Z}}\Phi_k e^{2\pi ikx}, \quad \Phi_0=0,\quad u(x)=\sum_{k\in{\mathbb Z}}u_k e^{2\pi ikx},

then

\displaystyle u_k=\frac{\Phi_k}{e^{2\pi ik\alpha}-1}.

For {u_k} to decay superpolynomially, one needs that

\displaystyle |e^{2\pi ik\alpha}-1|\ge\frac{c}{k^\tau}

for some {\tau}, which amounts to {\alpha} being badly approximable by rationals.

11.4. Parabolic case


In the parabolic world, in addition to invariant measures, invariant distributions provide further obstructions.


Remember that Heisenberg nilflows are special flows with constant roof over Furstenberg skew-products of the form

\displaystyle f_{\alpha,\beta}(x,y)=(x+\alpha,y+x+\beta).

We need study the corresponding cohomological equation.

Proposition 15 In Fourier series, if

\displaystyle \Phi(x)=\sum_{n,m\in{\mathbb Z}}\Phi_k e^{2\pi i(nx+my)}, then the cohomological equation has a formal solution if and only if all \displaystyle \mathcal{D}_{m,n}(\Phi)=0 where {\mathcal{D}_{m,n}} is the distribution such that \displaystyle \mathcal{D}_{m,n}(e^{2\pi i(ax+by)}) =\begin{cases} e^{-2\pi i((\alpha n+\beta m)k+\alpha m{k\choose 2})} & \text{ if } (a,b)=(n+km,m), \\ 0 & \text{otherwise}. \end{cases}


Indeed, let {e_{a,b}(x,y):=e^{2\pi i(ax+by)}}, and

\displaystyle u=\sum_{m,\,n\in{\mathbb Z}}u_{n,m}e_{n,m}.

Then

\displaystyle u\circ f =\sum_{m,\,n\in{\mathbb Z}}e^{2\pi i(n\alpha+m\beta)}.u_{n,m}e_{n+m,m}.

The matrix {A=\begin{pmatrix} 1 & 1 \ 0 & 1 \end{pmatrix}} acts on {{\mathbb Z}^2}. It has a family of orbits {{(n,0)}} and {m} orbits

\displaystyle O_{n,m}:=(n+m{\mathbb Z})\times\{m\},

{0\le n<m}, for each {m>0}. {L^2(T^2)} splits accordingly, so the cohomological equation can be solved independently in each summand

\displaystyle H_{n,m}:=\ell^2(O_{n,m}),

and on

\displaystyle H_0:=\ell^2({\mathbb Z}\times\{0\}).

For {H_0}, {f} acts like the circle rotation {R_\alpha}, so we already understand the necessary condition, given by invariant measures.


For fixed {0\le n<m} and {m>0}, let us denote

\displaystyle e_k:=e_{n+km,m},\quad \Phi_k:=\Phi_{n+km,m},\quad u_k:=u_{n+km,m}.

The cohomological equation reads

\displaystyle \sum\Phi_k e_k=\sum(\lambda_{k-1}u_{k-1}-u_k),

where

\displaystyle \lambda_k:=e^{2\pi i(n\alpha+km\alpha+m\beta)}.

Recursively, one gets

\displaystyle u_0=-\Phi_0-\sum_{j=-\infty}^{-1}(\lambda_j\cdots\lambda_{-1})\Phi_j.

Then

\displaystyle \lambda_j\cdots\lambda_{-1}=e^{2\pi i(jn\alpha+jm\alpha+{j\choose 2}m\alpha)}.

Similarly, the equation

\displaystyle \phi\circ f^{-1}=u-u\circ f^{-1}

yields

\displaystyle u_0=\Phi_1+\sum_{j=1}^{+\infty}(\lambda_{1}\cdots \lambda_{j})\Phi_j.

Combining both leads to

\displaystyle \sum_{j\in{\mathbb Z}}e^{-2\pi i(jn\alpha+jm\alpha+{j\choose 2}m\alpha)}\Phi_j=0,

i.e. {\mathcal{D}_{n,m}(\Phi)=0}.


Conversely, one can see that this conditions are sufficient for existence of a formal solution, and for existence of smooth solutions under Diophantine conditions.

11.5. More general results on cohomological equations in parabolic dynamics

  • The case of Heisenberg nilflows, which we just treated, is due to Katok.
  • More general nilflows have been studied by Flaminio-Forni, as well as horocycle flows.
  • Linear flows on higher genus surfaces are due to Forni. In this case, one gets {g} distributions {\mathcal{D}_1,\ldots,\mathcal{D}_g}. If a smooth function {f} is killed by all of them, ergodic integrals stay bounded. This is equivalent to {f} being a coboundary, according to Gottschalk-Hedlund.

11.6. Cocycle effectiveness


Sometimes, the cohomological equation with measurable data is needed. It is much more difficult, but some miracle occurs in the parabolic setting.

Definition 16 Given a map {f:X\rightarrow X}, say a function {\Phi} on {X} is a measurable coboundary of the exists a measurable {u} such that {u\circ f-u=\Phi}.


The theorem on mixing smooth time-changes of Heisenberg nilflows (Avila-Forni-Ulcigrai) required the roof not to be a measurable coboundary. It turns out that here, this is equivalent to not being a smooth coboundary.

Proposition 17 Let us study the Furstenberg skew-product {f_{\alpha,\beta}} on the {2}-torus. Let {\Phi} be a smooth function on {T^2}. Then
{\Phi} is not a smooth coboundary {\iff} {\phi} is not a measurable coboundary.


Indeed, let

\displaystyle S_n\Phi:=\sum_{k=0}^{n-1}\Phi\circ f^k

denote the Birkhoff sums. Then Flaminio-Forni establish quadratic upper bounds: there exists a sequence {n_\ell} such that

\displaystyle (UB)\quad\quad \|S_{n_\ell}\Phi\|_\infty \le C\, n_\ell^{1/2}.

Matching lower bounds exist: if {\Phi} is not a smooth coboundary, there exists a nonvanishing {\mathcal{D}_{k,l}(\Phi)}, and

\displaystyle (LB) \quad\quad \liminf \frac{1}{n}\|S_n\Phi\|_2\geq c\,|\mathcal{D}_{k,l}(\Phi)|>0.


If {\Phi} is a measurable coboundary, {\Phi=u\circ f-u}, then

\displaystyle S_n\Phi =u\circ f^n -f

stays bounded on a set of almost full measure, and grows at most quadratically on the complement. This contradicts the quadratic lower bound.

11.7. More general nilflows

Theorem 18 (Avila-Forni-Ravotti-Ulcigrai) For general nilflows of step {k\ge 2}, there exists a dense (in {C^\infty(X)}) class {\mathcal{P}} of generators {\alpha} of time-changes, which are

  • either measurably trivial (i.e. measurably conjugate to the nilflow);
  • or mixing.

Unfortunately, the set {\mathcal{P}} is not explicitly describable like in the Heisenberg case.


The proof is an induction on central extensions. It uses mixing by shearing.

12. More examples of parabolic dynamics

12.1. Parabolic perturbations which are not time-changes


These were discovered by Ravotti during his PhD at Princeton.


Here, parabolic means that the derivative of the flow grows polynomially. It implies that smooth time-changes are still parabolic.


Start with {G=Sl(3,{\mathbb R})}, {\Gamma} a cocompact lattice, {X=\Gamma\setminus G}. Let {h_{\mathbb R}} be the flow generated by a unipotent element {U}. Let {Z} belong to the center of the minimal unipotent and {V} such that {[U,V]=-cZ}.


Let {\tilde U=U+\beta Z} where {\beta} is a function on {X} such that {|\nabla \beta|_\infty <|c|}. Let {\tilde h_{\mathbb R}} denote the corresponding flow.

Theorem 19 If {\tilde h_{\mathbb R}} preserves a smooth measure {\tilde\mu} with {C^1} density, then {\tilde h_{\mathbb R}} is

  1. parabolic: {\|\nabla\tilde h_{\mathbb R}\|=O(t^4)};
  2. ergodic;
  3. mixing.


Remark. Existence of {\tilde\mu} {\iff} there exists a time-change of the {Z}-flow which commutes with {\tilde h_{\mathbb R}}.


Remark. There exists such {\tilde h_{\mathbb R}} which are not smoothly isomorphic to {h_{\mathbb R}}. This follows from the failure of cocycle rigidity for parabolic actions, due to Wang, following many people.


Remark. Ergodicity needs be proven, it does not follow from general principles.


The proof relies on mixing by shearing, although in a setting different from what we have already met. Consider arcs of orbits of {V} and push them by {\tilde h_{\mathbb R}}. Since

\displaystyle \nabla \tilde h_{\mathbb R}(V)=V+u_t(x)Z

where {u_t(x)} is an ergodic integral for {\tilde h_{\mathbb R}}.

12.2. What else can shearing be used for?


A strong, quantitative, shearing can be used to establish spectral results. Here, I mean the spectrum of the Koopman operator

\displaystyle U_t:L^2\rightarrow L^2,\quad f\mapsto f\circ\phi_t.

To each {f\in L^2(X,\mu)}, there corresponds a spectral measure {\sigma_f} on {{\mathbb R}}. Its Fourier coefficients are given by selfcorrelations of {f}, i.e.

\displaystyle \hat\sigma_f(t)=\langle f\circ \phi_t , f \rangle_2.


The spectrum of {U_t} is absolutely continuous {\iff} for all {f\in L^2}, {\hat\sigma_f} is absolutely continuous {\iff} for all {f\in L^2}, {\int_{{\mathbb R}}\hat\sigma_f(t)^2\,dt <+\infty}.


If {\phi_{\mathbb R}} is ergodic, it is enough to study functions {f} which are smooth coboundaries.

Theorem 20 (Forni-Ulcigrai) Smooth time-changes of a horocycle flow has absolutely continuous spectrum.


The proof uses quantitative bounds

\displaystyle |\langle f\circ \phi_t , f \rangle_2|\le \frac{Cf}{t}.

Since {\frac{1}{t}\in L^2}, this implies absolute continuity.


Fayad-Forni-Kanigowski consider smooth area-preserving flows on the {2}-torus with a stopping point.

12.3. Ratner property


M. Ratner uses a quantitative form of shearing for unipotent flows.


Shearing takes some time. Ratner requires the following:


For all {\epsilon>0}, for all large enough {t_0}, there exists a set {X_\epsilon} of measure {>1-\epsilon}, for all pairs {x,y\in X_\epsilon} not in the same orbit, but such that {d(x,y)<\epsilon}, there exists {t_1>t_0} such that

\displaystyle d(\phi_{t_1}(x),\phi_{t_1+s}(y))<\epsilon

and

\displaystyle d(\phi_\tau(\phi_{t_1}(x)),\phi_\tau(\phi_{t_1+s}(y)))<\epsilon

for all {\tau\in [t_1,(1+K)t_1]}.


For a long time, this was used only in algebraic dynamics, until a more flexible variant, called switchability, was introduced. It means that one can switch past and future.


Fayad-Kanigowski and Kanigowski-Kulaga-Ulcigrai established this variant for typical smooth area-preserving flows on surfaces. This implies mixing of all orders.


Here is another application if these ideas:

Theorem 21 (Kanigowski-Lemanczyk-Ulcigrai) For all smooth time-changes {\tilde h_{\mathbb R}} of the horocycle flow, the rescaled flow {\tilde h_{\mathbb R}^K} is not isomorphic to {\tilde h_{\mathbb R}}.

They are actually disjoint in Furstenberg’s sense. Recall that the horocycle flow itself is isomorphic to its rescalings.


We use a disjointness criterion based on this switchable variant of Ratner’s property.

12.4. Summary

  • Parabolic means slow butterfly effect.
  • Typically slow mixing.
  • Slow equidistribution.
  • Disjointness of rescalings.
  • Obstructions to cohomological equation.


Tools:

  • Shearing.
  • Ratner property and switchability.
  • Renormalization.
Posted in Uncategorized | Leave a comment

Notes of Uri Bader IHES lectures october 5th, 2021

Algebraic representations of ergodic actions

Based on joint work with Alex Furman, and earlier litterature.

Thanks to Sami Douba for his help with notetaking.

Today, we start with basics on measure spaces and algebraic varieties. Later on, we shall merge both subjects together.

1. Ergodic theory of algebraic varieties

1.1. Algebraic actions

{k} is a local field. For simplicity, characteristic is {0} but most of what I will say extends to positive characteristic. Also to nonlocal complete normed fields. {\mathbb{G}} is an algebraic group acting algebraically on an algebraic variety {\mathbb{V}}. Then {G=\mathbb{G}(k)} acts on {V=\mathbb{V}(k)}. We equip {V} with the {k}-topology, it is second countable and locally compact.

Questions. What can one say of the structure of orbits?

Examples. {SL_2({\mathbb R})} acting on {{\mathbb R}^2} has {2} orbits, one open and one closed. Let {K=SO(2)}, {P=} triangular matrices, {A=} diagonal matrices, {U=} unipotent matrices. Then {K}-orbits are concentric circles, {P}-orbits are the origin, two halflines and two halfplanes. {A}-orbits are the origin, four halflines and branches of hyperbolas. {U}-orbits are points of the {s}-axis and two halplanes.

We observe that orbits are open or closed, except for halflines which are nearly closed: intersections of an open and a closed set. We call such sets locally closed

Fact: Orbits of algebraic actions are locally closed.

Consequences.

{G}-invariant open sets separate points in distinct orbit in {V}. It follows that the quotient topology on {V/G} is second countable and {T_0} (topology separates points). This is known as Chevalley Theorem (combined with a result of Borel-Serre).

The Borel structure on {V/G} is countably separated. It follows that there exists a Borel embedding of {V/G} to {[0,1]}.

A fundamental theorem of Descriptive Set Theory states that all uncountable Polish topological spaces (completely metrizable, admitting a countable dense subset) are isomorphic as Borel spaces. See Kechris’ book. Such Borel spaces are called standard Borel spaces.

Examples. Finite sets, countable sets, {[0,1]\simeq\{ 0,1 \}^{\mathbb N}\simeq} a separable Hilbert space.

A Borel space is said to be countably separated if there exists a countable collection of Borel sets that separates points (equivalently, space has a Borel embedding into {\{ 0,1 \}^{\mathbb N}}). Standard Borel sets have this property.

We take this encouraging fact as an invitation to do ergodic theory.

1.2. measures and measure classes

Say two measures on a Borel space are equivalent if they have the same sigma-ideal of null sets. A measure class could be understood as the choice of a sigma-ideal.

Warning. Not every sigma-ideal comes from a measure. For instance, the sigma-ideal of meager sets does not arise from a measure.

Examples.

{V} has a natural volume measure class.

{G} has the Haar measure class.

For every closed subgroup {S<G}, {G/S} has a unique {G}-invariant measure class, called the Haar class. Warning: very rarely does {G/S} admit a {G}-invariant measure. For instance, {SL_n(k)} acting on projective space {P^{n-1}(k)} has no invariant measure in the Haar class. More generally, when {G} is simple and {Q} is parabolic, {G/Q} has no invariant measure in the Haar class.

If {S} is a locally compact group, Haar measure is finite iff {S} is compact. If {T<S} is a closed normal subgroup, then {S/T} has a finite invariant measure iff {T} is cocompact in {S}.

Definition 1 Say that a {G}-invariant measure class on {V} is ergodic if every {G}-invariant Borel set is either null or full (complement is null).

Equivalently, every a.e. defined Borel {G}-invariant map {V\rightarrow[0,1]} is a.e. constant. Here, {[0,1]} can be replaced with any countably separated space.

Corollary 2 Every {G}-ergodic measure class on {V} is supported on a single orbit. Moreover, it coincides with the Haar class on this orbit.

Indeed, think of the action as a {G}-map {X\rightarrow V} where {X} is a {G}-ergodic space. Since {V/G} is countably separated, this map must be a.e. constant. This leads to a map {X\rightarrow G/H} to an orbit. This must be a Borel and measure class isomorphism, thanks to the uniqueness of the Haar class.

In the nonergodic case, one can use ergodic decompositions.

Definition 3 Given a {G}-measure class {[\nu]} on {V}, since {V/G} is countably separated, for every measure {\nu} in the class, there exists a family of {G}-invariant measures {t\mapsto \nu_t} on {V}, {t\in V/G}, such that {\nu_t} is a Haar measure on the orbit denoted by {t}, and

\displaystyle  \nu=\int_{V/G} \nu_t d\bar{\nu}(t).

This is the ergodic decomposition of {[\nu]}.

In particular, the measure class {[\nu]} is fully determined by the class of the pushed-forward measure {[\bar{\nu}]}.

1.3. Classification of {G}-invariant probability measures

If {N<G} is a normal {k}-algebraic subgroup, which is cocompact, then {G/N} has a finite Haar measure. By Noetherianity, there exists a minimal element {N_0} among such cocompact normal {k}-algebraic subgroups. In fact, {N_0} is a least element. Indeed, given {N} and {N'}, {G/(N\cap N')} maps to a closed subset of {(G/N)\times(G/N')} hence is compact.

Theorem 4 If {G} has no compact factors (i.e. {N_0=G}), then every {G}-invariant probability measure on {V} is supported on fixed points.
In general, every {G}-invariant probability measure on {V} is supported on the {N_0}-fixed points.

Example. For {G=SL_2({\mathbb R})} acting on {{\mathbb R}^2}, the only {G}-invariant probability measure is the Dirac mass at the origin. Same picture for {P}, {A} and {U} (with as many invariant measures as there are measure on the {x}-axis). However, {K=SO(2)} has a lot of invariant measures.

This implies Borel’s density theorem.

Theorem 5 (Borel) If has no compact factor, and {\Gamma<G} is a lattice, then {\Gamma} is Zariski-dense in {G}.

Indeed, let {Z} be the Zariski closure of {\Gamma}, then push the {G}-invariant probability measure from {G/\Gamma} to {G/Z}. It must be supported on a fixed point, i.e. {G/Z} is a point.

1.4. Generalization

Fix a locally compact sigma-compact group {\Gamma} and a representation {\rho:\Gamma\rightarrow G}. Then {\Gamma} acts on {V}. The orbit space is complicated, but still {V\rightarrow V/G} is a {\Gamma}-map. Same reasoning yields: every {\Gamma}-invariant ergodic measure is supported on a unique {G}-orbit.

Theorem 6 There exists a minimal normal {k}-subgroup {N<G} such that

\displaystyle  \Gamma\rightarrow G\rightarrow G/N

has precompact image.
Every {\Gamma}-invariant measure on a {G}-algebraic variety is supported on the {N}-fixed points.

Consider {\Gamma=Stab_G(\mu)}, for {\mu\in Prob(V)}.

Corollary 7 The stabilizer of a measure {\mu\in Prob(V)} is compact modulo the fixator of the Zariski-support of {\mu}.

Fact (Zimmer). The action of {G} on {Prob(V)} has locally closed leaves.

1.5. Some more ergodic theory

A Lebesgue space is a standard Borel space equipped with a measure class.

By a map from a Lebesgue space {X} to a Borel space {U}, we mean an equivalence class of Borel maps : {X\rightarrow U} defined almost everywhere, up to almost everywhere equality. The space of such maps is denoted by {L(X,U)}.

A morphism of Lebesgue spaces is a map which sends null sets to null sets.

Let {S} be a locally compact second countable group. Then {S} has a standard Borel space structure and a Lebesgue space structure. An {S}-Lebesgue space {X} is a Lebesgue space with a homomorphism {S\rightarrow Aut(X)} such that {S\times X\rightarrow X} is a morphism.

An action of {S} on {X} is ergodic if every {S}-invariant map {X\rightarrow U} is essentially constant, for every standard Borel space {U}.

Definition 8 Fix an action of {S} on {X}. Say it is

  1. Doubly ergodic if the diagonal action on {X\times X} is ergodic.
  2. metrically ergodic if for every isometric action of {S} on a separable metric space {U}, every {S}-equivariant map {X\rightarrow U} is (essentially) constant.
  3. Weakly mixing if for every ergodic probability measure preserving action of {S} on {Y}, the diagonal action on {X\times Y} is ergodic.
  4. Has no compact factors if for every continuous homomorphism {S\rightarrow K} to a compact group {K} and any compact subgroup {H<K}, for every map {X\rightarrow K/H} (equipped with Haar measure), {H=K}.

Easy fact. {1\implies 2\implies 3\implies 4}.

Indeed, if {X} is doubly ergodic and acts isometrically on {U}, the distance defines an invariant function on {X\times X}, hence constant. If the constant is not zero, the ilage of {X} is discrete in {U}, hence countable (since {U} is separable), contradiction.

If {X} is metrically ergodic and has a pmp action on {Y}, an invariant function {f} on {X\times Y} gives rise to an equivariant map {X\rightarrow L^\infty(Y)}. Since probability measure is invariant, {L^\infty(Y)\rightarrow L^2(Y)} is equivariant and the action on {L^2(Y)} is isometric. Now {L^2(Y)} is separable, so {f} is constant.

If {X} is weakly mixing and {K/H} is a compact factor, one can assume {X=K/H}. Take {Y=K}. The map {(x,k)\mapsto k^{-1}x} is {K}-equivariant {:X\times Y=(K/H)\times K\rightarrow K/H}, hence constant, so {K/H} is a single point.

Easy fact. If the action of {S} on {X} is probability measure preserving, then {1\iff 2\iff 3\iff 4}.

Indeed, it suffices to prove that {4\implies 2}. One can assume that metric space {U} has an invariant probability measure {\mu}, fully supported, and that {U} is complete. One easily shows that {U} is compact. Then {K=Isom(U)} is compact. It must act transitively on {U}, {U=K/H}. Under assumption 4, {U} is a point, this is {2}.

1.6. metric ergodicity

Nonexample. Let {K<G} be a compact subgroup. Let {S\rightarrow G} be a homomorphism. Then the action of {S} on {G/K} is not metrically ergodic.

Example. Let {\Gamma} be a countable group. Let {\Omega} be a probability space. Then the shift action of {\Gamma} on {\Omega^\Gamma} is metrically ergodic.

Indeed, {\Omega^\Gamma \times \Omega^\Gamma =(\Omega\times\Omega)^\Gamma}, and the shift action on {(\Omega\times\Omega)^\Gamma} is ergodic.

Claim. Let {G} be a noncompact {k}-simple algebraic group. Let {\Gamma <G} be a lattice, {H<G} a noncompact closed subgroup. Then the action of {\Gamma} on {G/H} and the action of {H} on {G/\Gamma} is metrically ergodic.

Indeed, metric ergodicity passes to lattices (pass from {U} to {Map_\Gamma(S,U)}). For the {H} action, this follows from Howe-Moore. Indeed, the action of {G} on the pmp space {G/\Gamma} is mixing. This implies decay of coefficients. Their restrictions to any closed noncompact subgroup {H} decay, hence the mixing action of {H}.

1.7. Amenability

Definition 9 (Zimmer) The action of {S} on {X} is amenable if there exists an {S}-equivariant conditional expectation

\displaystyle  L^\infty(S\times X) \rightarrow L^\infty(X).

Note that amenability implies the following weaker “baby amenability”, which is often used: for every compact convex {S}-space {C}, there exists an {S}-map {X\rightarrow C}.

Example. If {H<S} is an amenable subgroup, the action of {S} on {S/H} is amenable.

Fact. For every locally compact second countable group {S}, there exists an action of {S} on some Lebesgue space {X} which is both amenable and metrically ergodic (the Furstenberg boundary).

Example. Let {G} be a noncompact {k}-simple algebraic group. Let {\Gamma <G} be a lattice, let {H<G} be a noncompact amenable subgroup. Then the action of {\Gamma} on {G/H} is both amenable and metrically ergodic.

2. Algebraic representations of ergodic actions

Now we merge algebraic groups and ergodic actions. Fix a locally compact second countable group {S}, an action of {S} on a Lebesgue space {X}, a local field {k} and an algebraic {k}-group {\mathbb{G}}, {G=\mathbb{G}(k)}. Fix a continuous homomorphism {\rho:S\rightarrow G}.

Definition 10 An algebraic representation of the action of {S} on {X} with respect to {\rho} is a {k}{\mathbb{G}}-variety {\mathbb{V}} and an {S}-equivariant map {\phi:S\rightarrow V=\mathbb{V}(k)}.
A morphism between to such AREAs {(\mathbb{V},\phi)} and {(\mathbb{U},\psi)} is a {k}{\mathbb{G}}-morphism {\alpha:\mathbb{V}\rightarrow\mathbb{U}} such that {\alpha\circ\phi=\psi}.

Example. Let {T<S} be a closed subgroup. Consider the action of {S} on {T/S}. Every algebraic representation of this action is given by a pair {\mathbb{G},\mathbb{V})} and a point in {V} which is fixed by the Zariski closure of {\rho(T)} in {G}.

If {\rho} is Zariski dense, we get a map from {V_0=G/H_0} to {V}, where

\displaystyle  H_0=\overline{\rho(T)}^Z.

We get an AREA {\phi_0:X=S/T\rightarrow V_0}, and for every AREA {\phi:X\rightarrow V}, we have a unique morphism {V_0\rightarrow V} such that {\phi=\alpha\circ\phi_0}.
In other words, {G/H_0} is an initial object in the category of AREAs of {S/T}. This holds in general.

Theorem 11 Let {S} act ergodically on {X}. Then there exists an initial object in the category of AREAs associated with {\rho}, of the form {\phi_0:X\rightarrow G/H_0}

We think of an ergodic action of {S} as a generalization of a closed subgroup, up to conjugacy. The theorem states that the initial object is indeed a Zariski-closed subgroup.

Proof of Theorem 11. Consider the set of {k}-algebraic subgroups of {G} such that there exists a, AREA {\phi:X\rightarrow G/H}. This is nonempty. By Noetherianity, one can pick a minimal element {H_0} (it will turn out to be a minimum, up to conjugacy, but it is harder). We show that the map {\phi_0:X\rightarrow G/H_0} is an initial object.

Consider an other AREA {\phi:X\rightarrow V}. Consider the diagonal representation

\displaystyle  \phi\times\phi_0:X\rightarrow V\times G/H_0 .

By ergodicity, the image of {X} lies in one single {G}-orbit {G/H_1}. Composing with projection, we get a {G}-map {G/H_1 \rightarrow G/H_0}, hence an embedding {H_1 < H_0} up to conjugacy. By minimality, {H_1=H_0}. The other projection {G/H_1\rightarrow V} provides us with a {G}-map {G/H_0\rightarrow V}, which is unique.

Theorem 12 Assume the action of {S} on {X} is pmp, {\mathbb{G}} is {k}-simple and {\rho(S)} is unbounded (i.e. not contained in a compact subgroup). Then the initial object is trivial: any representation of {X} is constant. Indeed, an {S}-invariant probability measure on {V} exists only if {\rho(S)} is precompact.

2.1. Consequences

Theorem 13 (Bader-Furman-Gorodnik-Weiss) Let {G} be a noncompact {k}-simple algebraic group. Let {\Gamma <G} be a lattice, let {H<G} be a noncompact {k}-algebraic subgroup. Consider the {\Gamma}-action on {G/H=X}. For {\rho=} the inclusion of {\Gamma} into {G}, the initial object is the identity {X=G/H}.

Application. Every Borel map {{\mathbb R}^n\rightarrow{\mathbb R}^n} which commutes with {SL_n({\mathbb Z})} is a homothety.

Indeed, let {\phi:G/H\rightarrow G/H=V} be a {\Gamma}-map. Let {\Phi} be the composition of {\phi} with {G\rightarrow G/H}. Set

\displaystyle  \Psi:G\rightarrow V,\quad \Psi(g)=g^{-1}\Phi(g).

Then {\Psi} is right-{\Gamma}-invariant and left-{H}-invariant. Since the action of {H} on {G/\Gamma} is pmp and weakly mixing, {\Psi} is constant. Thus there exists {v\in V} such that {g^{-1}\Phi(g)=v}, {\Phi(g)=gv}, i.e. {\phi} is a {G}-map.

Theorem 14 Let {X} be an amenable and metrically ergodic {S}-space. Let {\mathbb{G}} be a {k}-simple algebraic group. Let {\rho:S\rightarrow G} be an unbounded homomorphism. Then there exists an initial object {\phi:X\rightarrow G/H} where {H<G} is a proper subgroup.

Indeed, let {P} be a parabolic subgroup of {G}. Since {S} acts on the convex space {Prob(G/P)}, by amenability, there exists an {S}-map {X\rightarrow Prob(G/P)}. Since {G}-orbits in {Prob(G/P)} are locally closed, the image of the {S}-map is contained in a single orbit {G/H_1}, where {H_1} is the stabilizer of a measure {\mu}. By the structure theorem on measure stabilizers, the fixator {H_0} of the Zariski hull of the support of {\mu} is cocompact in {H_1}. Up to conjugacy, it is contained in {P}.

Assume that {H_0} is trivial. Then {H_1} is compact, {G} acts by isometries on the separable space {G/H_1}. By metric ergodicity, the {S}-map {X\rightarrow G/H_1} is constant, which contradicts the assumption that {\rho(S)} is unbounded in {G}.

Therefore, {H_0} is not normal. {H_1} is contained in the normalizer {N} of {H_0} in {G}, which is a proper {k}-algebraic subgroup of {G}. The composition {X\rightarrow G/H_1 \rightarrow G/N} is a nontrivial AREA for {X} with respect to {\rho}.

3. Lattices in products

Today, we are aiming at rigidity results for lattices. Before entering the subject, let me sum up where we had reached last time.

3.1. AREAs continued

The tension between ergodicity and the very simple structure of algebraic actions creates an initial object in the category of AREAs. We call gate the initial object, because it is our entrance gate into the algebraic world.

Two theorems:

  1. Theorem 12. For a pmp and metrically ergodic action, the gate is trivial.
  2. Theorem 14. For amenable and metrically ergodic actions, the gate is nontrivial.

Remark: unbounded amenable subgroups of {G} are not Zariski-dense.

3.2. Leftover from last time : functoriality

Proposition 15 Fix {\rho:S\rightarrow G}. The gate defines a functor from the category of {S}-ergodic actions and the category of {k}-algebraic {G}-(coset)-varieties.

3.3. Introduction to lattices in products

Examples.

{{\mathbb Z}[\sqrt{2}]} is a lattice in {{\mathbb R}\times{\mathbb R}}.

{{\mathbb Z}[\frac{1}{p}]} is a lattice in {{\mathbb R}\times{\mathbb Q}_p}.

{SL_n({\mathbb Z}[\frac{1}{p}])} is a lattice in {SL_n({\mathbb R})\times SL_n({\mathbb Q}_p)}.

Definition 16 A lattice {\Gamma} in a product {S=S_1\times S_2} is irreducible if its projections to both factors are dense subgroups.
Equivalently, the action of {\Gamma} on each factor is ergodic.

Equivalently, the actions of {\Gamma\times S_2} and of {S_1\times \Gamma} on {S_1\times S_2} are ergodic.

Equivalently, the action of each {S_i} on {(S_1\times S_2)/\Gamma} is ergodic.


Indeed, the action of {S_i} on {L^\infty(S_i)} equipped with the weak{^*} topology is continuous.

3.4. Commensurability

Assume that {S} is totally disconnected locally compact. Then there exists a compact open subgroup {K<S}. Any two are commensurable. Say a subgroup {K<S} is commensurated if for all {s\in S}, {K^s} and {K} are commen surable.

Let {\Gamma<S_1\times S_2} be an irreducible lattice. Let {K_1<S_1} be a compact open subgroup. Then

\displaystyle  \Lambda=\gamma\cap(K_1\times S_2).

is commensurated in {\Gamma}, it is a lattice in {K_1\times S_2}. Hence {\Lambda<S_2} is a lattice which is commensurated by the dense subgroup {\Gamma<S_2}.
Conversely, assume that {\Lambda<\Gamma<T} where {\Lambda<T} is a lattice, {\Lambda<\Gamma} is commensurated and {\Gamma<T} is dense. Then one can reconstruct {S_1} from these data. There exists a totally disconnected group {T'}, a dense embedding {\Gamma\rightarrow T'}, and a precompact embedding {\Lambda\rightarrow T'} such that {\Gamma<T\times T'} is an irreducible lattice. It is called the Schlichting completion of {(\Gamma,\Lambda)}.

This indicates that lattices in products Lie{\times}tdlc are simpler that lattices in Lie groups, in the sens that we have a dual way of looking at them.

3.5. Superrigidity

Theorem 17 Let {\Gamma<S_1\times S_2} be an irreducible lattice. Let {G} be a {k}-simple group, let {\rho:\Gamma\rightarrow G} is Zariski dense and unbounded. Then superrigidity holds: {\rho} extends uniquely to a continuous homomorphism {\bar\rho:S\rightarrow G} and {\bar \rho} factors through one of the factors.

Corollary 18 If {\Lambda<\Gamma<T} are as above (i.e. {\Lambda<\Gamma} is commensurated and {\Gamma<T} is dense), {G} is {k}-simple and {\rho:\Gamma\rightarrow G} is Zariski dense and unbounded on {\Lambda}, then {\rho} extends uniquely to a continuous homomorphism {\bar\rho:T\rightarrow G}.

Indeed, the case where {\rho} extends to {T'} is excluded.

3.6. Application to arithmeticity

Apply previous theorem to a lattice in {G} and conclude that a lattice {\Gamma<G} is arithmetic iff it has a dense commensurator.

3.7. Preparation for the proof

Fix an action of {S_i} on a Lebesgue space {B_i} which is amenable and metrically ergodic.

Claim. The diagonal action of {S_1\times S_2} on {B_1\times B_2} is amenable and metrically ergodic.

Indeed, assume {C} is a nonempty {S}-compact convex space. Then {Map_{S_1}(B_1,C)} is nonempty, it is an {S_1}-compact convex space (viewed as a subset of {L^\infty(B_1,C)}). Therefore, there exists an {S_2}-map {B_2\rightarrow Map_{S_1}(B_1,C)}. I.e., there exists an {S_1\times S_2}-map {B_1\times B_2\rightarrow C}. This proves amenability.

Let {U} be an {S}-isometric metric space. Any {S}-map {B_1\times B_2\rightarrow U} is a.e. independant on the {B_1} variable, and on the {B_2} variable, therefore a.e. constant.

Corollary 19 For every irreducible lattice in {S}, the action of {\Gamma} on {B_1\times B_2} is amenable and metrically ergodic.

Of course, the action of {\Gamma} on {S} is not ergodic, it is proper, but

Claim. The action of {\Gamma} on {S_1\times B_2} is ergodic.

Before proving the claim, let us start with a general fact.

Given an action of {S} on {X}, when is the restriction to {\Gamma} ergodic? Answer is : iff the action of {S} on {(S/\Gamma)\times X} is ergodic. Indeed, one can mod out by a proper action: the space of {\Gamma}-orbits in {S\times X}, denoted by {S\times_\Gamma X}, is well defined, since the diagonal {S}-action on {S\times X} is conjugated to the action on the {S}-factor only, trivial on the {X} factor. In fact, {S\times_\Gamma X=(S/\Gamma)\times X}.

Applying this to the claim, {\Gamma} ergodic on {S_1\times B_2} {\iff} {S} ergodic on {(S/\Gamma)\times S_1\times B_2} {\iff} {S_2} ergodic on {(S/\Gamma)\times B_2)} is implies by metric ergodicity of {S_2} on {B_2}.

This proves the claim.

3.8. Proof of superrigidity theorem

Again, fix an action of {S_i} on a Lebesgue space {B_i} which is amenable and metrically ergodic. The action of {\Gamma} on {B_1\times B_2} is amenable and metrically ergodic. According to Theorem 14, the gate {G/H} is nontrivial, i.e. {H\not=G}.

Pick a generic {S_1}-orbit in {B_1\times B_2}, identify it with {S_1}. Get a map {S_1\times B_2\rightarrow G/H}. The ergodic action of {\Gamma} on {S_1\times B_2} yields a gate {\theta:\Gamma\times S_1 G\times N/H_0}, {N} the normalizer of {H_0} in {G}. Mod out by {N}, get a {\Gamma\times S_1}-map {S_1\times B_2\rightarrow G/N}. By ergodicity, we get a map {B_2\rightarrow G/N}.

3.9. Assume that {H_0\not=\{e\}}

By simplicity of {G}, {N\not=G}. So the {\Gamma}-action on {B_2} has a nontrivial representation {B_2\rightarrow G/N}, whence a {\Gamma}-representation {S_2\rightarrow G/N}. There exists a gate {S_2\rightarrow G/H_2} where {N_2<N\not=G}. By ergodicity of the {\Gamma}-action on {S_2}, we get a {\Gamma\times S_2}-equivariant map (where {\Gamma\times S_2\rightarrow G\times N_2/H_2}, where {N_2} is the normalizer of {H_2}) from {S_2} to {G/N_2}. Again by ergodicity, there is a {\Gamma}-invariant point in {G/N_2}. It also fixed by the Zariski-closure of {\Gamma}, which is {G}. Hence {N_2=G}, {H_2} is normal in {G}. By simplicity, {H_2=\{e\}}. So the above morphism to {G\times N_2/H_2} was to {G\times G}. The formula

\displaystyle  s_2 \mapsto \phi(s_2)\theta_2(s_2)^{-1}

defines an {S_2}-invariant (hence constant) map {S_2\rightarrow G}. In other words, {\phi(s_2)=g\theta(s_2)}. This implies that {\rho=\theta_2^g}. Composing with the projection {S\rightarrow S_2}, we get {\bar\rho:S\rightarrow G} whose restriction to {\Gamma} equals {\rho}. So we are done under the assumption that {H_0\not=\{e\}}. This was the easiest case.

3.10. From no on, assume that {H_0=\{e\}}

Then {N=G}, so {\theta:S_1\rightarrow G}. We have an {\Gamma\times S_1}-equivariant map {S_1\times B_2\rightarrow G} (equivariant w.r.t. {\rho\times\theta:\Gamma\times S_1\rightarrow G\times G}).

Compose this map with {G\times G\rightarrow G\times G/\Delta}, the diagonal. This is an AREA of {\Gamma\times S_1} to {G\times G}. By ergodicity of {\Gamma\times S_1} on {S_1\times S_2}, there is a gate {S_1\times S_2\rightarrow G\times G/M} where {M<\Delta}. Let us show that we are done if {M\not=\Delta}.

Assume that {M\not=\Delta}. Let us mod out the left {G}-action. This mods out the {S_1}-action in the gate, thus I get an {S_2}-map from {S_2} to a nontrivial quotient of {G}. As before, we get an extension of {\rho} to {S}.

From now on, assume that {M=\Delta}. {S_2} acts on {S_1\times S_2}. Since {\Delta} is equal ti its own normalizer in {G\times G}, the gate factors via a map {S_1\rightarrow S\times G/\Delta=G}, so {\rho} extends as before.

4. Margulis superrigidity

Let {S} be a real semisimple group of higher rank, let {\Gamma<G} be an irreducible lattice. Margulis superrigidity states that every Zariski-dense unbounded homomorphism {\rho:\Gamma\rightarrow G} uniquely extends to {S}.

The case when {S} is a product has been treated. Next time, I will prove the case when {S} is simple. I explain now that we are not too far from it.

Consider {S=SL_3(\ell)}, {T=\{\begin{pmatrix} \lambda & 0 & 0 \\ 0 & \lambda & 0 \\ 0 & 0 & \lambda^{-2} \end{pmatrix} \,;\, \lambda\in\ell\}}. It is amenable, noncompact, and its centralizer is the upper diagonal {GL_2(\ell)}. The {\Gamma}-action on {S/T} is amenable and metrically ergodic. {\Gamma\times PGL_2(\ell)} acts on {S/T} and we get a gate {S/T\rightarrow G/H}.

I would like to indicate a related geometric context without groups acting, that of {\tilde A_2}-buildings, covered by Caprace-Lecureux. At infinity, such buildings have an exotic projective plane. It has a large group of projectivities (in fact a pseudogroup of maps from lines to lines). This yields a large group acting on the boundary of a tree. It plays the role of {PGL_2(\ell)}. It is not too hard to show that this group is linear iff the building is classical. It follows that {\Gamma} is linear iff the building is classical.

4.1. Remarks on lattices in products

  1. The superrigidity theorem holds for lattices in products of {n} factors for any {n}.
  2. Nonarithmetic examples are known only for {n=2} yet.
  3. The examples are Kac-Moody groups acting on twin buildings and Burger-Mozes and Wise examples acting on products of trees.

4.2. A new category of representations

For the proof of Margulis superrigidity, we need to modify the concept of AREA. Up to now, we represented an {S} action on Lebesgue space {X}. Now we need to represent pairs of closed subgroups of a locally compact second countable group {S}.

The objects in our category are now the following data:

  • two closed subgroups {\Gamma} and {T} of {S},
  • an algebraic {k}-group {G},
  • a {k}-algebraic subgroup {L<Aut_k(V)} that commutes with {G},
  • a homomorphism {\rho:\Gamma\rightarrow G} and a continuous and Zariski-dense homomorphism {\theta:T\rightarrow L},
  • a representation, i.e. a measurable map {\phi:S\rightarrow V} which is {\rho\times\theta}-equivariant.

Theorem 20 Let {\Gamma<S} be a lattice. Assume that the action of {T} on {S/\Gamma} is ME. Then there exists an initial object (a gate), i.e. a {\Gamma\times T}-equivariant map

\displaystyle  S\rightarrow G/H_0,

where {\rho\times\theta:\Gamma\times T\rightarrow (G\times N)/H_0} and {N=Norm_G(H_0)}.

The proof follows similar lines as Theorem 11. Pick a minimal {H_0}.

4.3. The nontriviality theorem

Theorem 21 Let {\Gamma<S} be a lattice. Assume that the action of {T} on {S/\Gamma} is ME. Assume further that {T<S} is amenable, {G} is {k}-simple and {\rho} is unbounded, then the gate is not trivial.

Indeed, consider the {\Gamma} action on {X=S/T}. By amenability and ME, there exists a proper subgroup {H<G} and a representation {S\rightarrow S/T\rightarrow G/H}, under homomorphism {\rho\times e:\Gamma\times T\rightarrow G\times\{e\}}. The gate will be a deeper object, but this suffices to prove that the gate is nontrivial.

4.4. Functoriality

Assume that {S} is a noncompact simple {\ell}-algebraic group. Let {\Gamma<S} be a lattice. Recall that for every closed noncompact {T<S}, the actions of {\Gamma} on {S/T} and of {T} on {S/\Gamma} are ME (this is Howe-Moore’s theorem). We now show that this gives right to extra invariants.

Fix {S}, {\Gamma}. Consider the category of {(\Gamma,S)}-actions, i.e.

  • the objects are closed, noncompact subgroups {T<S},
  • the morphisms are elements {s} of {S} acting on the right and conjugating the right action of {T_1} into the right action of {T_2} on {S}.

Consider next the category of {G}-spaces:

  • the objects are {L}‘s acting on {G}-space {V}, commuting with {G},
  • the morphisms are {k}{G}-morphisms of varieties {V\rightarrow U}.

The gate functor assigns to a {(\Gamma,S)}-action of {\Gamma\times T_1} on {S} an orbit {G/H_1} with action of {G\times N/H_1}. To a morphism {S\rightarrow S} given by element {s\in S}, the gate associates a unique morphism {\alpha(s)=gate(s)} of algebraic varieties {G/H_1\rightarrow G/H_2}.
In particular, we get a map {\alpha:Aut(\Gamma,S,T)\rightarrow Aut(G/H_0)}. The group {Aut(G/H_0)=N_G(H_0)/H_0}. On the other hand, {Aut(\Gamma,S,T)=N_S(T)}. This homomorphism {\alpha} is a nontrivial datum: it gives extra invariance to the representation {S\rightarrow G/H_0}.

Corollary 22 If {T_1,T_2<S} normalize each other. Then they have the same gate {S\rightarrow G/H_0}.

I really mean, the same map serves as a gate for both.

4.5. Proof of Margulis superrigidity

Recall our goal. Let {S} be a real semisimple group of higher rank, let {\Gamma<G} be an irreducible lattice. Margulis superrigidity states that every Zariski-dense unbounded homomorphism {\rho:\Gamma\rightarrow G} uniquely extends to {S}.

Fact. There exist closed noncompact abelian subgroups {T_1,\ldots,T_n<S} generating {S} such that each {T_i} commutes with {T_{i+1}}.

In fact, this is equivalent to higher rank.

Example. {S=SL_3({\mathbb R})}. The first three {T_i}‘s are {1}-parameter subgroups of the Heisenberg group of upper unipotent matrices, the three next are {1}-parameter subgroups of the opposite Heisenberg group of lower unipotent matrices.

Geometrically speaking, it means that one can move from any geodesic to any other by travelling within finitely many maximal flats.

Now we embark in the proof. By Corollary 22, all the {T_i}‘s have the same gate {\phi:S\rightarrow G/H}, equivariant with respect to {\rho\times\theta_i:\Gamma\times T_i \rightarrow G\times N/H}. By ergodicity, the {\Gamma}-invariant point is invariant under the Zariski closure of {\rho(\Gamma)}, i.e; b y {G}, so {N=G}, {H} is normal in {G}. Since {T_i} is amenable, {H\not=G} so {H=\{e\}}.

So the gate is {\phi:S\rightarrow G}, equivariant with respect to {\rho\times\theta_i:\Gamma\times T_i \rightarrow G\times G}.

{\phi} defines a pull-back map from the algebra of polynomial functions on {G}, {k[G]}, to {k}-valued functions on {S}, {L(S,k)}. It is injective, since its Zariski support is {G}. This embeds {k[G]} as a {T_i}-invariant subalgebra of {L(S,k)} for all {i}, hence an {S}-invariant subalgebra. This gives an {S}-action of {S} on {k[G]} extending the actions of the {T_i}. Thus {S} acts on {G} on the right, when a homomorphism {\theta:S\rightarrow G}. The gate {\phi:S\rightarrow G} is {\rho\times\theta}-equivariant, this implies that {\rho} extends to {S}. End of proof.

4.6. Rank one

Let {S} be a real semisimple group of rank one, let {\Gamma<G} be an irreducible lattice. Let {\rho:\Gamma\rightarrow G} be a Zariski-dense unbounded homomorphism. What extra assumptions should we make to show that {\rho} extends?

For {T<S}, let {\phi_T}, {\theta_T}, {N_T}, {H_T},… denote the {T}-gate.

Lemma 23 Let {T=P=MAU} be a parabolic. If {H_P=\{e\}}, then {\rho} extends.

Indeed, consider {\phi_A}, {H_A<H_P} hence {H_A=\{e\}}, {\phi_A=\phi_{N_S(A)}}.

Since {P} and {N_S(A)} generate {S}, {\rho} extends as before.

Lemma 24 Assume that there exists a simple, noncompact subgroup {W<S} and a {\Gamma}-representation {S/W\rightarrow G/H}, equivariant with respect to {\rho}, with {H\not=G}. Then {U} is not in the kernel of {\theta_P}.

Indeed, {U'=W\cap U} is noncompact. By a finite number of up and down steps of taking normalizers we reach {P} (two steps should be sufficient). All these intermediate normalizers have the same gate. Assume by contradiction that {\theta_P(U)=\{e\}}. Then the {U'}-gate equals the {P}-gate, which thus factors by {U}. This map is both {U} and {W}-invariant, but {U} and {W} generate {S}. Thus {H=G}, contradiction.

Note that the assumption of Lemma 24 never holds if {U<\mathrm{ker}(\theta_P)} and {K} is nonarchimedean.

4.7. Relation to arithmeticity

Definition 25 Say that {G} is compatible (with {S}) if for all proposer subgroups {H} of {G}, {U<\mathrm{ker}(\theta_P)}, where {\theta_P:P\rightarrow N_G(H)/H}.

Let {S=\mathbb{S}({\mathbb R})} where {S\not=Sl(2,{\mathbb R})}. Le {\Gamma<S} ne an orreducible lattice.

Fact. There exists a unique minimal number field {i_0:\ell\rightarrow{\mathbb R}} such that {S} is defined over {\ell} and {\Gamma<S(\ell)} up to conjugation and finite index.

Theorem 26 (Margulis arthmeticity criterion) {\Gamma} is arithmetic if and only iff its image is precompact in any place other than {i_0}. I.e. for every embedding {j:\ell\rightarrow k} to a local field, either {j(\Gamma)\subset\mathbb{S}(k)} is precompact or {k={\mathbb R}} and {j=i_0}.

Supperrigidity implies arithmeticity. Indeed, take {\mathbb{G}=\mathbb{S}}. Can assume that it is adjoint. For every embedding {j:\ell\rightarrow k}, the image of the composition {\Gamma \rightarrow\mathbb{S}(\ell)\rightarrow\mathbb{S}(k)} is precompact unless {S=\mathbb{S}({\mathbb R})}. By an argument of Borel-Tits, one gets {j=i_0}.

Observe that one need only a few targets {\mathbb{G}} to get arithmeticity.

Theorem 27 (Bader-Fisher-Miller-Stover) If there exists in {\Gamma\setminus S/K} infinitely many immersed maximal totally geodesic subspaces of dimension {\geq 2}, then the assumption of Lemma 24 holds for all {\mathbb{G}}‘s relevant to arithmeticity.

Corollary 28 {\Gamma} is defined over the ring of integers of a number field.

Exercise. For {n\geq 4}, {S=SO(n,1)}, the group {\mathbb{G}=\mathbb{S}({\mathbb C})} is compatible. In particular, if there exist infinitely many immersed maximal totally geodesic subspaces of dimension {\geq 2}, then {\Gamma} is arithmetic. (Special care is needed for {n=3}).

5. Apafic Gregs

Talk given in Fanny Kassel’s seminar, on Oct. 11th, 2021, dedicated to Margulis, Perlman and Baldi.

Joint work with Alex Furman.

How to produce a random element of a group? In fact, a sequence. We will study linear representations of such random elements.

5.1. GREGs

Definition 29 Let {X} be a probability space. Let {T} be an ergodic invertible pmp transformation of {X}. Let {\Gamma} be a locally compact second countable group. Let {\phi:X\rightarrow\Gamma} be a map. The Greg is the data {(X,B,m,T,\Gamma,\phi)}.

Examples.

Compact. Let {X=\Gamma=S^1}, {\phi=id}, {T} an irrational rotation.

Random walk. Fix {\Gamma} and a probability measure {\mu} on {\Gamma}. Let {X=\Gamma^{\mathbb Z}}, {T=} shift, {\phi=} projection to the {0}-th coordinate. The resulting Greg is a random walk.

Markov chain. Fix a graph with probability transitions (i.e. for each vertex {v}, a probability measure on the set of edges emanating from {v}), {X=} set of paths in the graph, {m=} some Gibbs measure associate. Decorate edges with elements of {\Gamma}. This defines a map {\phi:X\rightarrow \Gamma}.

Geodesic flow. Let {M} be a compact negatively curved Riemannian manifold. Let {X=T^1 M} denote the unit tangent bundle of {M}. {T_t=} geodesic flow. Let {\Gamma=\pi_1(M)}. Fix a fundamental domain. For each time {t}, there is a map {\phi_t:X\rightarrow\Gamma} which tells which group element drags {T_t(v)} back to the fundamental domain. One can decorate the picture with a flat bundle.

5.2. Associated constructions

Let {B} be the minimal {\sigma}-algebra such that {\phi} is measurable. Given {m<n}, let {F_m^n=\bigvee_{k=m}^n T^k F}. Assume that {B=F_{-\infty}^{+\infty}}.

Let {X_+=F} equipped with the future {\sigma}-algebra {F_0^{+\infty}}. Let {X_-=F} equipped with the past {\sigma}-algebra {F_{-\infty}^0} (I view these as factors of {X}). Then {T} maps {X_+\rightarrow X_+}, {T^{-1}:X_-\rightarrow X_-}. It turns out that {X} can be reconstructed from the pair {(X_+,T)}, using the natural extension construction.

Definition 30 Let {S:Y\rightarrow Y} be a pmp map which is not invertible. One constructs another system, with an invertible transformation, by taking the inverse limit {\tilde Y} of

\displaystyle  \cdots\rightarrow Y\rightarrow Y\rightarrow Y\rightarrow\cdots

Example. {X=\Gamma^{\mathbb N}}, {T=} shift, then the natural extension is {\Gamma^{\mathbb Z}}.

Theorem 31 If {T} is ergodic on {Y}, then its natural extension {\tilde T} on {\tilde Y} is metrically ergodic.

Here, {\tilde T} metrically ergodic means that for every metric extension {U\rightarrow \tilde Y} (i.e. a family of complete separable metric spaces on which a lift of {\tilde T} acts isometrically, there exists an equivariant section.

Fact. The map {\phi:X\rightarrow \Gamma} defines a cocycle as follows. For {n=0}, {\phi_0=e}, and for {n>0},

\displaystyle  \phi_n(x)=\phi(T^{n-1}x)\phi(T^{n-2}x)\cdots \phi(Tx)\phi(x),\quad \phi_{-n}(x)=\phi_n(T^{-n}x)^{-1}.

The cocycle identity

\displaystyle  \phi_{n+m}(x)=\phi_n(T^m x)\phi_m(x)

holds. Then a map {X\times\Gamma\rightarrow\Gamma^{\mathbb Z}} is defined by

\displaystyle  (x,\gamma)\mapsto (\phi_n(x)\gamma^{-1}).

Whence an action of {{\mathbb Z}\times\Gamma} on {\Gamma^{\mathbb Z}}.
The projections {\Gamma^{\mathbb Z}\rightarrow\Gamma^{{\mathbb Z}_{\ge 0}}} and {\Gamma^{\mathbb Z}\rightarrow\Gamma^{{\mathbb Z}_{\le 0}}}, when modded out by {\Gamma}, yield the factors {X\rightarrow X_+} and {X\rightarrow X_-}. If instead one mods out by {{\mathbb Z}}, one gets the space {E} of {{\mathbb Z}}-ergodic components of {\Gamma^{\mathbb Z}}, and factors {E\rightarrow E_+} and {E\rightarrow E_-}. I think of these as ideal futures and pasts.

Theorem 32 {E}, {E_+} and {E_-} are amenable {\Gamma}-spaces, the maps {E\rightarrow E_+} and {E\rightarrow E_-} are {\Gamma}-metrically ergodic.

5.3. Asymptotic past and future independence condition

Definition 33 Say that the asymptotic past and future independence condition (apafic) is satisfied if the map {E\rightarrow E_+\times E_-} is measure class preserving.

In other words, if you were born poor, you still may become rich. It holds for all examples above but the compact example.

If the Greg is Apafic, then the pair {(E_-,E_+)} is called a Boundary pair. The spaces are amenable and the maps {E_+\times E_-\rightarrow E_+} and {E_-} are metrically ergodic.

Theorem 34 Let {G} be a simple {k}-algebraic group. Let {\rho:\Gamma\rightarrow G} be Zariski-dense. Then there exist opposite parabolic subgroups {Q_+,Q_-<G} and measurable equivariant maps {E_+\rightarrow G/Q_+}, {E_-\rightarrow G/Q_-} and {E_+\times E_-\rightarrow G/(Q_+\cap Q_-)}.
Moreover, if {k={\mathbb R}}, {Q_+=Q_-} is the minimal parabolic.

View the Cartan projection as a leftinvariant {\mathfrak{a}^+}-valued metric on {G}. Let {F_n} be the composition of {\phi_n:X\rightarrow \Gamma} with {\rho} and Cartan projection,

\displaystyle  F_n:X\rightarrow \Gamma\rightarrow G\rightarrow \mathfrak{a}^+ .

Subadditivity shows that {\frac{1}{n}F_n} converges to an interior point of the Weyl chamber {\mathfrak{a}^+}. This is called “simplicity of the spectrum”.
I view apafic gregs as a useful generalization of random walks, and we see that the simplicity of the spectrum theorem holds for them.

Tholozan, Ledrappier: See Avila-Viana for the Markov chain version.

> Let {G} be a simple {k}-algebraic group. Let {\rho:\Gamma\rightarrow G} be Zariski-dense. Then there exist opposite parabolic subgroups {Q_+,Q_-<G} and measurable equivariant maps {E_+\rightarrow G/Q_+}, {E_-\rightarrow G/Q_-} and {E_+\times E_-\rightarrow G/(Q_+\cap Q_-)}.
Moreover, if {k={\mathbb R}}, {Q_+=Q_-} is the minimal parabolic.

View the Cartan projection as a leftinvariant {\mathfrak{a}^+}-valued metric on {G}. Let {F_n} be the composition of {\phi_n:X\rightarrow \Gamma} with {\rho} and Cartan projection,

\displaystyle F_n:X\rightarrow \Gamma\rightarrow G\rightarrow \mathfrak{a}^+ .

Subadditivity shows that {\frac{1}{n}F_n} converges to an interior point of the Weyl chamber {\mathfrak{a}^+}. This is called “simplicity of the spectrum”.

I view apafic gregs as a useful generalization of random walks, and we see that the simplicity of the spectrum theorem holds for them.

Tholozan, Ledrappier: See Avila-Viana for the Markov chain version.

Posted in Uncategorized | Leave a comment

Notes of Karim Adiprasito’s Hadamard Lectures 2021-04-19

Karim Adiprasito

I want to show how combinatorial tools can be used to prove results which previously came from algebraic geometry.


Today, I give an

1. Overview of applications

1.1. Matroids and characteristic polynomials


Example. The chromatic polynomial {\chi_G(t)} of a graph {G} counts the number of colorings with {t} colors. To compute it, one can use inclusion/exclusion, removing an edge and putting it back.


A matroid is an abstraction for the concept of linear independence in vectorspaces.


Example. Given a configuration of vectors {V} in a vectorspace, let {f_i(V)} be the number of independent sets of size {i}.


Often, the sequence of coefficients has a log concavity property:

\displaystyle a_i^2 \ge a_{i-d}a_{i+d}

(in the chromatic case, the signs are alternating, but this does not affect the inequality).

1.2. Face vectors in embeddings


A seemingly simple problem. Let {\Delta} be a simplicial complex. I want to embed it PL into a vectorspace {{\mathbb R}^{2k}}. Subdivision is allowed. How many faces are needed? Let {f_i(\Delta)} be the number of faces of dimension {i} after subdivision.

Theorem 1 A necessary condition is

\displaystyle \forall k,\quad f_n(\Delta)\le (k+2)f_{k-1}(\Delta).

This is asymptotically tight.


Example: for an embedding of a graph in the plane, {E\le 3V}. For a {2}-complex in {{\mathbb R}^4}, the numbers of triangles and edges must satisfy

\displaystyle T\le 4E.


Here is why it is easy for graphs. Adding edges, one can modify the embedding in order that every complementary region is a triangle. Then one uses Euler’s formula {V-E+T=2} and {3T=2E}. The point is that the modification changed only the left-hand side of the inequality.


The 2D case is much harder.

1.3. Triangulations of manifolds


Let {M} be a closed manifold, fix a triangulation, the {f}-vector counts simplices of each dimension starting from {-1}.

Theorem 2 The {f}-vectors of simplicial spheres are realized by {f}-vectors of simplicial polytopes, and are characterized by an algebraic condition.

To make a polytope into a simplicial polytope, one merely needs put its vertices in general position.


Later I will explain this algebraic characterization.

2. Basics

2.1. The algebra associated to a simplicial complex


Let {\mathbf{k}} be any infinite field. Let {\Delta} be a finite simplicial complex (by convention, the empty set is a face of {\Delta}). Consider the polynomial ring with one indeterminate for each vertex of {\Delta}. Let {I_\Delta} be the ideal generated by all monomials which are not supported on {\Delta}. Let

\displaystyle \mathbf{k}[\Delta]=\mathbf{k}[X]/I_\Delta.

Let

\displaystyle H(\mathbf{k}[\Delta],t)=\frac{1}{(1-t)^d }\sum f_{i-1}(\Delta)t^i(1-t)^{d-i}.

denote the Hilbert polynomial of {\Delta}. If {\Delta} has dimension {d-1},

\displaystyle H(\mathbf{k}[\Delta],t)=\sum_{k,i}{k\choose i}f_{i-1}(\Delta)t^k.


In order to remove a useless infinite dimensional part in {\mathbf{k}[\Delta]}, we pick a matrix {\Theta} whose rows are linear combinations of vertices. The quotient

\displaystyle A(\Delta,\Theta)=\mathbf{k}[\Delta]/<\Theta>

is of Krull dimension {0} if and only if the rank of the restriction of {\Theta} to {\sigma} is equal to {|\sigma|}. This holds generically.

2.2. Cohen-Macaulay


An interesting case. Assume {\mathbf{k}[\Delta]} is Cohen-Macaulay. I.e. if I define the length of {\Theta} to be the Krull dimension of {\mathbf{k}[\Delta]}, then {\Theta} is a regular system for {\mathbf{k}[\Delta]}, i.e. if multiplication by {\Theta} is injective on {\mathbf{k}[\Delta]}. If

\displaystyle H(\mathbf{k}[\Delta],t)=\frac{1}{(1-t)^d }h(t),

then

\displaystyle H(A(\Delta,\Theta),t)=h(t).


In this case, {h} can be computed as follows. Write a Pascal triangle with {f}-numbers on the right-hand side, ones on the left-hand side and the skew-Pascal rule (every entry is the difference of both entries above it). Then the coefficients of {h} appear on the bottom line. Conversely, one can recover {f}-numbers from the coefficients of {h} (the {h}-vector).

Theorem 3 (Hochster) {\mathbf{k}[\Delta]} is Cohen-Macaulay if and only if for all nonempty simplices {\sigma} in {\Delta}, the link of {\sigma} in {\Delta} has homology concentrated in dimension {d-1-|\sigma|}.

MK: Gorenstein corresponds to links being homology spheres? Yes, see below.

Theorem 4 (Macaulay) ({f}-vectors and) {h}-vectors of Cohen-Macaulay simplicial complexes are the same as Hilbert polynomials of commutative graded algebras generated in degree {1}.

They are sometimes called {M}-vectors.

2.3. Poincaré duality


Now we go to MK’s spoiler.


Boundaries of simplicial polytopes are a source of simplicial spheres. Think of {\Sigma=\partial P} as a fan. Consider the algebra {P(\Sigma)} of cone-wise polynomial functions. If {\Sigma} is rational, {P(\Sigma)} is isomorphic to the equivariant cohomology of the toric variety over {\Sigma}. Each vertex {v} of {\Sigma} defines an element {\chi_v} of {P(\Sigma)}, which is linear (equal to {1} at {v}) on $latex {{\mathbb R}_+ v}&fg=000000$ and {0} on all other vertex rays. These generate {P(\Sigma)}. Multiplying vertex functions gives simplex functions $latex {\chi_\sigma}&fg=000000$. Let us mod out by globally linear functions. Claim: the result is {A(\Sigma,\Theta)}, where {\Theta} is given by vertex coordinates in ambient {{\mathbb R}^d}.

Theorem 5 (Poincaré duality. Probably due to Hochster) Let {\Sigma=\partial P} be the boundary of a homology sphere of dimension {d-1}. Then the degree {d} component of {A(\Sigma)} is {1}-dimensional, and for all {k\le d}, the pairing

\displaystyle A^k(\Sigma)\times A^{d-k}(\Sigma) \rightarrow A^d(\Sigma) is perfect.

Here, a homology sphere means that {\mathbf{k}[\Delta]} is Cohen-Macaulay and the {k}-homology of every link is concentrated in top dimension and has dimension {1}.


I will explain where this comes from just below.

2.4. Shellability


Say a simplicial complex of pure dimension {d-1} is shellable if {d-1=0} or {\Delta} is a {d-1}-simplex or there exists a {d-1}-simplex {F} in {\Delta} such that the simplicial complex {\Delta\setminus F} (induced by other {d-1}-facets) is shellable and {(\Delta-F)\cap F} is shellable of codimension {1}.


For instance, two simplices touching at a vertex do not form a shellable complex, since {(\Delta-F)\cap F} has codimension {\geq 2}. On the other hand, two boundaries of simplices touching at a vertex is shellable. In general, boundaries of simplicial polytopes are shellable. Proof by picture: start from a facet in {{\mathbb R}^d} and launch a rocket. You travel in the rocket and look back at the earth. Remove facets in the order in which they appear in view. Go to infinity and come back the opposite side.


Shellable implies Cohen-Macaulay over all fields. Indeed, it implies that {\Delta} is homotopically equivalent to a wedge of {d-1}-spheres.


For {\Delta} shellable, let {F} the facet removed in shelling step. Then the kernel of the (surjective) restriction map

\displaystyle A(\Delta)\rightarrow A(\Delta -F)

is generated by {\chi_{\sigma(F)}}, where {\sigma(F)} is the minimal simplex of {F} not in {\Delta-F} (everything here once modded out by {\Theta}).


If {\Sigma=\partial P}, given an element {\alpha\in A^k(\Sigma)}, I first express it in shelling steps. The shelling turned around (in reverse order), is a shelling too. Therefore I can construct an element {\beta} step by step that pairs nontrivially with {\alpha}.

3. Hard Lefschetz theorem


Today, I want to go to the hard Lefschetz theorem and its combinatorial applications, “beyond positivity”.


Let {\Sigma} be the boundary of a {d}-dimensional polytope {P}. Remember the algebra {A:=A(\Sigma,\Theta)} is expressible in terms of cone-wise polynomials:

\displaystyle A=P(\Sigma)/<globally \,linear \,functions>.


Let {\ell\in A^1} be a cone-wise linear function. Assume that {\ell} is strictly convex (i.e. the domains of linearity are exactly the cones of the fan). Think of {\ell} as defining an ample line-bundle.


OG: should one assume {\ell} is rational? No, this is not necessary. It allows to interpret the algebra in terms of a toric orbifold, but we shall not use it.

Theorem 6 (Hard Lefschetz theorem) For all {k\le d/2}, multiplication with {\ell^{d-2k}} yields an isomorphism {A^k \rightarrow A^{d-k}}.


Remember that when {\Delta} is Cohen-Macaulay over {\mathbf{k}}, the {h}-vector, given by dimensions of homogeneous components {A^i}, is an M-vector (i.e. realizable as the dimensions in a polynomial ring). Define, for {i\le d/2},

\displaystyle g_i=h_i-h_{i+1}=\mathrm{dim}(A^i/\ell A^{i-1}).

Then Lefschetz theorem implies that the {g}-vector is again an M-vector. This observation is due to Stanley. In fact, the converse is true: all M-vectors arise in this way (Billera-Lee).

3.1. Hodge-Riemann relations


Define a quadratic form on {A^k} as follows.

\displaystyle Q_{k,\ell}:A^k\times A^k \rightarrow A^d,\quad (a,b) \mapsto \mathrm{deg}(ab\ell^{d-2k}).

The primitive part {P_\ell A^k} is the kernel of the multiplication by {\ell^{d-2k+1}}

\displaystyle A^k \rightarrow A^{d-k+1},\quad

Theorem 7 (Hodge-Riemann relations) The quadratic form {Q_{k,\ell}} is definite of sign {(-1)^k} on {P_\ell A^k}.


Let {\alpha,\beta} be two strictly convex elements of {A^1}. The restriction of {Q_{1,\ell}} to the span of {\alpha} and {\beta} cannot be definite, because it has a plus from degree 0 and a minus from degree {1}. Therefore

\displaystyle \mathrm{deg}(\alpha^2\ell^{d-2})\mathrm{deg}(\beta^2\ell^{d-2})-\mathrm{deg}(\alpha\beta\ell^{d-2})^2\leq 0.

This is Alexander-Fenchel’s inequality, that states that for convex sets in {{\mathbb R}^d}, the function

\displaystyle t \mapsto \mathrm{vol}(t_A A+t_B B+...)

is log concave.

3.2. Application to Grunbaum’s conjecture


Question 1. We know that Poincaré duality holds in a broader generality. Does Stanley’s observation (“{g}-theorem”) extend to simplicial {\mathbf{k}}-homology spheres?


Assuming this, I prove

Theorem 8 Let {\Delta} be a simplicial complex embedded PL into {{\mathbb R}^{2k}}. The number of faces {f_k} satisfies

\displaystyle f_k \le (k+2)f_{k-1}.


Think of {\Delta} as a subcomplex of a triangulated {2k}-sphere {\Sigma} (see Bing’s notes on the Geometric Topology of {3}-manifolds, 1983). There is a restriction map : {A(\Sigma)\rightarrow A(\Delta)}. Faces provide a generating system for {A^k(\Delta)} (because {A} is generated, as a vectorspace, by square-free elements), therefore

\displaystyle \mathrm{dim}(A^k(\Delta))\le f_{k-1}(\Delta).

(Remember that the length of {\Theta} is the Krull dimension of {\mathbf{k}(\Sigma)}, which is larger than the Krull dimension of {\mathbf{k}(\Delta)}. Later on we shall see a model for {A} where this inequality will become obvious).


Relations come from codimension {1} faces of {k+1}-faces, therefore

\displaystyle \mathrm{dim}(A^{k+1}(\Delta))\ge f_{k}(\Delta)-(k+1)f_{k-1}(\Delta).


Claim. {\mathrm{dim}(A^{k+1}(\Delta))\leq \mathrm{dim}(A^{k}(\Delta))}. Indeed, by Poincaré duality and if Lefschetz property applies to {\Sigma}, then a commutatiove square holds, and multiplication by {\ell} is surjective

\displaystyle A^k(\Delta)\rightarrow A^{k+1}(\Delta).

Therefore

\displaystyle \mathrm{dim}(A^{k+1}(\Delta))\leq \mathrm{dim}(A^{k}(\Delta)).

This completes the proof of Theorem 8.


Our next goal is the following theorem.

Theorem 9 (Adiprasito 2018, Adiprasito-Papadakis-Petrotou 2021) Let {\Sigma} be a {\mathbf{k}}-homology sphere (i.e. {A} is Gorenstein*). For a generic choice of {\Theta} and a generic choice of {\ell\in A^1}, the Hard Lefschetz theorem holds for {A(\Sigma,\Theta)}. Furthermore (2018 version), the restriction of quadratic form {Q_{k,\ell}} to any square-free monomial ideal is perfect.

Note that I do not claim any positivity, merely nondegeneracy. OA: sign main depend on the choice of {\Theta}.


Continuation of theorem (2021 version). In characteristic {2}, there exists a field extension {\tilde{\mathbf{k}}} of {\mathbf{k}} such that {Q_{k,\ell}(\alpha,\alpha)\not=0} for all nonzero {\alpha\in \tilde{A}}.

4. The partition complex and Poincaré duality


Now we go to cohomological tools.

4.1. The works


How do I prove Poincaré duality? I show that the socle of {A} has dimension {1}. For this, I need two ingredients.


a) For all {k<d}, the map to stars of vertices

\displaystyle A^k \rightarrow \bigoplus_{v\,vertex\,of\,\Sigma} A^k(st_v(\Sigma))

is injective.


b) For every vertex {v}, the multiplication by the indeterminate {X_v}

\displaystyle A^k(st_v \Sigma)\rightarrow A^{k+1}(\Sigma)

is injective.


Let us prove b) first. I ignore modding out by {\Theta}. By Cohen-Macaulay property, the exact sequence

\displaystyle 0\rightarrow \mathbf{k}[st_v\Sigma]\rightarrow \mathbf{k}[\Sigma]\rightarrow\mathbf{k}[\Sigma - v]\rightarrow 0

(where the central map is multiplication by {X_v}) stays exact after modding out, for all {\Theta}.


Let us prove a) now. The complex {\tilde P}

\displaystyle A^k \rightarrow \bigoplus_{v\,vertex\,of\,\Sigma} A^k(st_v(\Sigma))\rightarrow \bigoplus_{e\,edge\,of\,\Sigma} A^k(st_e(\Sigma))\rightarrow\cdots

with suitable choices of signs, is exact is positive homogeneity degrees (in degree {0}, this is the \v Cech complex).


Consider the Koszul complex {K(\Theta)}. The tensor product

\displaystyle \tilde P\otimes K(\Theta)

is a bi-complex. It is exact in the Koszul direction, and exact in the {\tilde P} in positive homogeneity degrees. Thus the diagonal complex is acyclic in positive homogeneity degrees.


If {\Sigma} is Buchsbaum (i.e., it is of pure dimension {d-1} and for every vertex, the star is Cohen-Macaulay), the kernel of the map {A^k \rightarrow \bigoplus_{v\,vertex\,of\,\Sigma} A^k(st_v(\Sigma))} is isomorphic to the cohomology

\displaystyle H^{k-1}(\Sigma,\mathbf{k})^{{d\choose k}}.


In general, if {M} is a closed orientable simplicial manifold of dimension {d-1}, {A(M)} is not Poincaré duality, but the quotient

\displaystyle B(M)=A(M)/H^{k-1}(\Sigma,\mathbf{k})^{{d\choose k}}

is.

4.2. The cheats


Let me state the most general version of the Hard Lefschtez theorem we have now.


Let {\Delta} be a simplicial complex of dimension {d-1}. Let {\mu} be any {1}-dimensional quotient of {A^d(\mu)}. The kernel is always {H^{d-1}(\Delta,\mathbf{k})}. Then $latex {B^(\mu)}&fg=000000$ is a minimal quotient of $latex {A^(\Delta)}&fg=000000$ such that {B^d(\mu)} is isomorphic to {\mu}.


Let {\check{\mu}} be an element in {H_{d-1}(\Delta,\mathbf{k})} which is dual to {\mu}. Let {T} be a simplex of cardinality {d}. Let

\displaystyle X_T=\prod_{v\,vertex\,of\,T}X_v

be the corresponding monomial. Then

\displaystyle \mathrm{deg}(X_T)=\frac{\check{\mu}_T}{\mathrm{det}(\Theta_{|T})}

where {\mathrm{det}(\Theta_{|T})} is the minor of {T} in the matrix {\Theta}.


Fix {\mu}. For a generic choice of {\ell\in B^1(\mu,\Theta)} (and generic {\Theta}), {B(\mu,\Theta)} satisfies the Hard Lefschetz theorem.

5. Proof of the classical case of Hard Lefschetz Theorem


I first state McMullen’s version (“On simple polytopes”, following Fleming-Karu 2009.

Theorem 10 (McMullen) Let {\Sigma} be the boundary of a simplicial {d}-polytope. Let {\ell\in A^1(\Sigma)} be a strictly convex cone-wise linear function.

HL:

\displaystyle \times \ell^{d-2k}:A^k\rightarrow A^{d-k} is an isomorphism.

HR: \displaystyle Q_{k,\ell}: A^l\times A^k \rightarrow A^d={\mathbb R},\quad (a,b)\mapsto \mathrm{deg}(ab\ell^{d-2k}) is definite of sign {(-1)^k} o n primitive elements {P_\ell A^k=\mathrm{ker}(\times\ell^{d-2k+1}:A^k\rightarrow A^{d-k+1}}.


Next is the nonclassical version (due to Adiprasito 2018, Adiprasito-Papadakis-Petrotou 2021).

Theorem 11 Let {\Sigma} be a triangulated {d-2}-sphere. Thee exists {\Theta} and there exists {\ell\in A^1(\Sigma,\Theta)} such that

HL:

\displaystyle \times \ell^{d-2k}:A^k\rightarrow A^{d-k} is an isomorphism.

HR: \displaystyle Q_{k,\ell}: A^l\times A^k \rightarrow A^d={\mathbb R},\quad (a,b)\mapsto \mathrm{deg}(ab\ell^{d-2k}) is nondegenerate on any squarefree monomial ideal.


OG: does this extend to other spaces than spheres?


Good spoiler! No. For {\mathop{\mathbb P}^1\times \mathop{\mathbb P}^1}, {Q_{1,\ell}(a,b)=\mathrm{deg}(ab)}, consider the monomial ideal generated by {\chi_v}. Then elements of this ideal are cone-wise linear functions that vanish on a half-plane, they must be restrictions of globally linear functions on the opposite halfplane, hence the square vanishes.

5.1. Classical proof


Today, I give the classical proof. It is a double induction.

5.2. The climb (raise in dimension)

Lemma 12 Let {\Sigma=\partial P}. Let {\ell\in A^1} be strictly convex. Assume that the theorem (HL and HR) holds in codimension {1}. Then HL holds for {\Sigma}.


Proof is by contradiction. Assume there exists {\alpha\not=0} in {A^k(\Sigma)} such that {\ell^{d-2k}\alpha=0}. For every vertex {v} of {\Sigma}, there is a restriction map to the link {A(lk_v\Sigma)}. The restriction of {\ell} to this fan is still strictly convex. Thus the image of {\alpha} in the sum of links {\bigoplus_v A(lk_v\Sigma)} is nonzero. Since

\displaystyle (\alpha_{|lk_v})\ell_{|lk_v}^{d-2k}=0,

\displaystyle \alpha_{|v}\in P_{\ell_{lk_v}}A^k(lk_v).

We compute $latex \displaystyle (-1)^k Q^\Sigma(\alpha,\alpha)=(-1)^k \mathrm{deg}_\Sigma(\ell^{d-2k}\alpha^2)=(-1)^k \sum_i \mathrm{deg}_\Sigma(\ell_i x_i \alpha^2 \ell^{d-2k-1})&fg=000000$

$latex \displaystyle =(-1)^k \sum_i \ell_i\mathrm{deg}_{lk_v}(\alpha_{|lk_v}^2 \ell_{|lk_v}^{d-2k-1}).&fg=000000$ In the last term, the sign of the degree is {(-1)^k} provided {\alpha_{lk_v}\not=0}. The sum is strictly positive. This proves that {\alpha^2\not=0}, and HL follows.

5.3. The hike (stay in the same dimension)


Next we prove HR. First assume that {\Sigma_0} is the boundary of the {d}-simplex. Then {A(\Sigma_0)={\mathbb R}[x]/x^{d+1}}, and HR is proven by direct calculation.


Next we deform a general simplicial polytope to the simplex continuously. Start with the simplex, add vertices in general position in order to have as many as in {\Sigma} and let them move. Assuming general position, for all but finitely many times {t}, the obtained convex hull is simplicial. In each interval, HL is true hence HR is preserved. So we need only understand what happens when crossing special times.


Assume that {d=3}. What we observe are Pachner moves: – flip of diagonals in a quadrilateral face, – upraisal of a vertex in the middle of a face. In the first case, there is a common refinement (blow up by adding a vertex at the intersection of diagonals). I denote by {m} the cardinality of the minimal simplex introduced ({m\le \frac{d+1}{2}}). I call {\Sigma_+} the side where this minimal cardinality is smallest. I denote by {\Delta_+} the flip locus after the flip (it is the star of the minimal simplex introduced). Then

\displaystyle \Gamma=\Sigma_+ -\Delta_+ = \Sigma_- -\Delta_-

is the part which is unaffected by the flip. My goal is to construct a map {A(\Sigma_-) \rightarrow A(\Sigma_+)}.


I introduce relative algebras: in the can of boundaries of polytopes, {P(A,B)} is the ring of cone-wise polynomials on {A} that vanish on {B}. In the general case,

\displaystyle \mathbf{k}[A,B]=I_B/I_A,

where {I_A} is the ideal generated by monomials whose support is not in {A}.


The following sequences are exact:

\displaystyle 0\rightarrow A(\Delta_\pm,\partial\Delta_\pm)\rightarrow A(\Sigma_\pm)\rightarrow A(\Gamma)\rightarrow 0.

\displaystyle 0\rightarrow A(\Gamma,\partial\Gamma)\rightarrow A(\Sigma_\pm)\rightarrow A(\Delta_\pm)\rightarrow 0.

The generator for {A(\Delta_+,\partial\Delta_+)} is in homogeneity degree {m}. The generator for {A(\Delta_-,\partial\Delta_-)} is in homogeneity degree {d+1-m}. Thus, for {k<m}, the maps

\displaystyle A^k(\Sigma_\pm)\rightarrow A^k(\Gamma)

are isomorphisms. For {k\ge m}, the maps

\displaystyle A^k(\Gamma,\partial\Gamma)\rightarrow A^k(\Sigma_\pm)

are isomorphisms. Therefore, grosso modo,

\displaystyle A(\Sigma_+)=A(\Sigma_-)\oplus K

where {K=A(\Delta_+,\partial\Delta_+)} truncated after {d-m}. This decomposition is orthogonal with respect to the Poincaré pairing. Indeed, if {\beta\in K} has degree {k\geq m} and {\alpha\in A(\Sigma_-)} has degree {d-k\leq d-m}, then {\alpha\in A^{d-k}(\Gamma,\partial\Gamma)} which injects into {A^{d-k}(\Sigma_-)}, therefore {\alpha\beta=0}. If instead {\beta} is the generator of {A(\Delta_+,\partial\Delta_+)} of degree {m}, i.e. {\beta=\chi_\sigma} of minimal face {\sigma} introduced, then {\beta} is a product of {m} concave cone-wise linear functions, so the quadratic form is

\displaystyle (concave\, linear)^m \cdot \ell^{d-2m}

its sign is {(-1)^m}.

This prove the HR relations.

5.4. Beyond polytopes (but still convex)


I explain the application to matroids. A matroid {M} is a pair {(E,L)} where {E} is a finite set and {L\subset 2^E} (ordered by inclusion) is a set of subsets of {E} satisfying

  • {\emptyset}, {E\in L}.
  • {S,T\in L\implies S\cap T\in L}.
  • If {S\in L}, the set \displaystyle C_S=\{T-S\,;\, T\in L,\, T>S\} partitions {E-S}.


The fan {B_M} associated to {M} in {{\mathbb R}^n}, {n=|E|-1}, is defined as follows. Let $latex {(e_i)_{i\in E}}&fg=000000$ be a generating set in {{\mathbb R}^n} such that {\sum_i e_i=0}. For every subset {F\subset E}, set $latex {e_F=\sum_{i\in F} e_i}&fg=000000$. To a flag {F_1\subset F_2\subset\cdots} of {E}, we associate the cone on the vectors {e_{F_1}, e_{F_2},\ldots}. Then the fan {B_M} is the collection of all theses cones.


Example. The set of singletons in a set {E}. The fan consists of only one cone, the whole plane.


Example. The full power set {2^E}. The fan consists of 6 cones in the plane.

Theorem 13 (Adiprasito, Huh, Katz) Let {M} be a matroid. The algebra

\displaystyle A(B_M)=cone-wise\,polynomials/<global\,linear\,functions>

satisfies Poincaré duality (the socle degree is the longest chain in {\check{L}=L-\{\emptyset,E\}}). Hard Lefschetz. Hodge-Riemann.


Next time, I will explain the non classical Lefschetz theorem. Before, I will explain a bit of the context of matroids.


OG: worries about the notion of ampleness.


JK: does the ring you define correspond to the Chow ring in algebraic geometry? Yes.


JK: does the proof follow the same inductive scheme as McMullen’s ? Yes.


MK: do secondary polytopes arise in matroids? yes.

6. Matroids, continued


Today, I quickly finish with matroids and then proceed to Lefschetz without positivity.

Theorem 14 (Adiprasito-Huh-Katz) Let {M} be a matroid, let {B_M} denote its Bergman fan. Then the algebra {A(B_M)} satisfies

  • Poincaré duality (the fundamental class belongs to the degree corresponding to the longest chain).
  • If {\ell\in A^1} is ample, then Hard Lefschetz and Hodge-Riemann hold.

6.1. Characteristic polynomial of a matroid


Given a vector {e}, a matroid can be reduced in two ways:

  1. By removing {e}, get {M\setminus e}.
  2. By projecting mod {e} (and removing {0}), get {M/e}.

Recursively, let us define the polynomial (in indeterminate {\lambda})

\displaystyle \chi(M,\lambda)=\chi(M\setminus e,\lambda)-\chi(M/e,\lambda),

starting from the following chromatic polynomials:

\displaystyle \chi(\text{loop},\lambda)=0,\quad \chi(\text{edge},\lambda)=\lambda-1.

(this gives rather the characteristic polynomial divided by {\lambda}).


Question. Can one give a formula for {\chi(M,\lambda)}?


Empirically, one sees that coefficients {a_i} alternate in signs and their absolute values are unimodal, i.e. increase and then decrease.


From the above theorem, it follows that

\displaystyle a_i^2\geq a_{i-1}a_{i+1}.


One can access to {a_i} from certain intersection numbers. Let {d} denote the degree of the fundamental class. The relevant intersection numbers are

\displaystyle \mathrm{deg}(\cdot\alpha^i \ell^{d-i}),

where

\displaystyle \alpha_j=\sum_{j\in F} x_F

where {F} is a subset of {E} and {j\in E}, and

\displaystyle \beta_j =\sum_{j\notin F} x_F .

Then all {[\alpha_j]\in A(B _n)} (resp. all {[\beta_j]}) are the same (independent of {j}). This is because one mods out by global linear functions. Denote them by {\alpha} and {\beta}. These classes are nef. They can be approximated by ample classes.


I will not prove that {\mathrm{deg}(\cdot\alpha^i \ell^{d-i})} is a coefficient; merely explain its combinatorial content.


The Hodge-Riemann relations imply the log-concavity. Indeed,

\displaystyle a_i^2-a_{i-1}a_{i+1}

can be interpreted as the determinant of a {2\times 2} matrix which is indefinite, so the determinant is negative.


The proof, in the case with positivity (ample classes) follows the same lines as McMullen’s argument for polytopes: induction.

7. Hard Lefschetz without positivity


I will focus on the following special case. {\mathbf{k}} is an arbitrary infinite field.

Theorem 15 Let {\Sigma} be a triangulated {d-1}-sphere (meaning a homology manifold, which is a homology sphere). Then for a generic Artinian reduction {A(\Sigma,\Theta)} and a generic element {\ell\in A^1},

  • the Hard Lefschetz theorem holds
  • Hall-Lamac relations hold, i.e. \displaystyle Q_{k,\ell}:A^k\rightarrow\mathbf{k},\quad (a,b)\rightarrow \mathrm{deg}(ab\ell^{d-2k}) does not degenerate on squarefree monomial ideals.


Assume that {d=2k+1}. The first nontrivial isomorphism is multiplication by {\ell} : {A^k\rightarrow A^{k+1}}.


Let {v} be a vertex. The multiplication with {x_v} is restriction to the star of {v}. Its kernel is the relative space {A^k(\Sigma,st_v\Sigma)}. Its image is {A^k(st_v \Sigma)x_v}. Let {w} be an other vertex. One expects that the kernel of multiplication with a generic linear combination of {x_v} and {x_w} be as small as possible, i.e. {\mathrm{ker}(x_v)\cap\mathrm{ker}(x_w)}. Dually, the image of this combination should be {\mathrm{im}(x_v)+\mathrm{im}(x_w)}. Now {\mathrm{ker}(x_v)\cap\mathrm{ker}(x_w)} is expressible in terms of the homology of the intersection of two stars, i.e a link, which vanishes.


The following classical fact supports this.

Lemma 16 (Kronecker) Let {X} and {Y} be two vectorspaces over {\mathbf{k}}. Let {A,B:X\rightarrow Y} be linear maps such that {B(ker A)} and {im A} intersect trivially. Then the kernel of a generic linear combination of {A} and {B} is {\mathrm{ker}(A)\cap\mathrm{ker}(B)}.

Proof. It suffices to show that {B(\mathrm{ker}(A))\cap \mathrm{im}(A)=0} and {B^{-1}(\mathrm{im}(A))+ \mathrm{ker}(A)=X}.


Here, {A^{k+1}} and {A^k} are dual to each other, so one can speak of orthogonal complements. Also, {\mathrm{ker}(x_v)} is the orthogonal complement of {\mathrm{im}(x_v)}.


So now we see that constructing an isomorphism is related to nondegeneracy of the pairing on certain subspaces.

7.1. Biaised pairing theory


Again, we want to prove inductively the “transversal primes property” for all subsets {W} of the vertex set of {\Sigma} (by adding vertices one by one): for a generic linear combination of {x_w}‘s, {w\in W}, the kernel of the multiplication is the intersection of kernels. If we can do it for {W=} all vertices, we shall be done for Hard Lefschetz.

7.2. The genericity of {\Theta} is needed


Let me construct a counterexample. The simplest sphere is the tetrahedron, boundary of the simplex on {{0,1,2,3}}. {\mathbf{k}(\Sigma)} has Krull dimension {3}, so {\Theta} consists of {3} linear forms. Let us pick generic vectors for the two first and {0} for the third. This is not a linear system of parameters, so we must modify it. Subdivide each facet, this produces {4} more vertices {0',1',2',3'}, hence we continue the linear forms with {3} generic vectors. We must check that minor corresponding to do not vanish. This now works. No choice can satisfy Hard Lefschetz.

7.3. Biased pairing property

Say that and ideal {I} of {A(\Sigma)} satisfies the biased pairing property (bpp) in degree {k} if

\displaystyle I^k\times I^{d-k}\rightarrow\mathbf{k}

is nondegenerate in the first factor.


Say that and {A(\Sigma)} satisfies the biased pairing property in degree {k} if all its squarefree monomials ideals do.


We use a descent lemma.

Lemma 17 Let {\Sigma} be a {d-1}-sphere. Let {k<d/2}. Then {A(\Sigma)} satisfies the biased pairing property in degree {k} if all links of vertices do.

So we only need to consider the middle case, in degree {k}, when the sphere has dimension {2k-1}.


From now on, this is the case. I will explain a special case, when the ideal {I} is the nonface ideal of a subcomplex {\Delta}, i.e. the kernel of

\displaystyle 0\rightarrow I \rightarrow A(\Sigma)\rightarrow A(\Delta)\rightarrow 0,

where {\Delta} is a codimension {1} sphere in {\Sigma}. Denote by {D} and {\bar D} the two hemispheres determined by {\Sigma}. Then

\displaystyle I(\Sigma,\Delta)=I(\Sigma,D)+I(\Sigma,\bar D)

is an orthogonal splitting. So we merely need prove the bpp for {I(\Sigma,\bar D)}.


Observation 1.

\displaystyle I(\Sigma,\bar D)=A(\Sigma, \bar D)=I_{\bar D}/I_\Sigma+\Theta,

since

\displaystyle 0\rightarrow A^\times (\Sigma,\bar D)\rightarrow A^\times(\Sigma)\rightarrow A^\times(\bar D)\rightarrow 0.


Observation 2.

Lemma 18 Say we want to prove bpp for an ideal {J} of {A(\Sigma} in degree {k}. This is equivalent to an injection

\displaystyle J^k\rightarrow (A(\Sigma)/Annihilator(J))^k in degree {k}.


Let us combine these two observations. When {J=I(\Sigma,\bar D)}, the annihilator of {J} is {I(\Sigma,D)}, hence

\displaystyle A(\Sigma)/Annihilator(I(\Sigma,\bar D))=A(D).

So we want an injection

\displaystyle A^k(D,\partial D)\rightarrow A^k(D).


Write {\Theta=(\tilde\Theta,\ell)}. Before Artinian reduction, the map

\displaystyle 0\rightarrow \mathbf{k}[D,\partial D]\rightarrow \mathbf{k}[D]\rightarrow\mathbf{k}[\partial D]\rightarrow 0

is an injection. Since all terms in the sequence are Cohen-Macaulay, a generic choice of {\tilde\Theta} is a regular system of parameters for {\partial D}, and the sequence remains exact when modding out by {\tilde\Theta}.


Since {\partial D} is a {2K-2}-sphere, by Poincaré duality, multiplication by {\ell},

\displaystyle \mathbf{k-1}[\partial D]/\tilde\Theta \rightarrow \mathbf{k}[\partial D]/\tilde\Theta

is injective, and thus bpp for {I(\Sigma,\bar D)} in degree {k} is equivalent to Lefschetz property for

\displaystyle A^k(D,\partial D)\rightarrow A^k(D).

Since {\partial D=\Delta} is a lower dimension sphere, this allows for induction.

8. Hard Lefschetz in codimension {1} implies bpp


Last time, I explained how to iteratively construct a Lefschetz element using Kronecker’s Lemma. Specifically, bpp (the nondegeneracy of the Poincare pairing at ideals) in codimension {1} inmplies the existence of Lefschetz elements.


To close the loop, I must show that Hard Lefschetz in codimension {1} implies bpp.


Later on, I will give another proof, based on transcendental arguments (no induction).


Last time, we dealt with ideals of the form {I(\Sigma,E)} where {E} was a codimension {1} sphere. Today, I cover the case where {E} is a codimension {1} manifold which is {k-2}-acyclic over {\mathbf{k}}. {k=1} is a trivial case.


As before,

\displaystyle I(\Sigma,E)=I(\Sigma,M)+I(\Sigma,\bar M),

where {E} separates {\Sigma} into two pieces {M} and {\bar M}.


Remember that the partition complex yields a map

\displaystyle \mathbf{k})^{{d\choose k}}\rightarrow A^k(E).

Theorem 19 {I(\Sigma,\bar M)} satisfies bpp if and only if the composition

\displaystyle K^{k-1}(M,\mathbf{k})^{{d\choose k}}\rightarrow H^{k-1}(E,\mathbf{k})^{{d\choose k}}\rightarrow A^k(E) is an isomorphism.


This is mainly diagram chasing. Under the assumption of the theorem,

\displaystyle I^k(M,E)\rightarrow B^k(M)

is an isomorphism, where

\displaystyle B(M)=A(M)/\bigoplus_{v\text{ vertex of }M} A(st_v M)=A(M)/\text{annihilator}(I(\Sigma,\bar M)).

This is exactly what we want.

8.1. How to prove that {I(\Sigma,\Delta)} satisfies bpp in general?

Assume again that {\Sigma} has dimension {2k-1} and {\Delta} has dimension {k-1}. The idea is to construct a hypersurface {E} such that {E} contains {\Delta} and {A(E)} is isomorphic to {A(\Delta)}.


Assume that {\Sigma} is a PL sphere.


Observation. Subdivision of {\Sigma} does not affect the fact that {\Delta} has bpp (we shall only use stellar subdivisions). Indeed, {A(\Sigma)} injects into {A(\Sigma')}, the subdivision, and there is an orthogonal decomposition

\displaystyle A(\Sigma')=A(\Sigma)\oplus G

Since {I(\Sigma',\Delta)} consists of all monomials not in {\Delta}, the same orthogonal decomposition holds for

\displaystyle I(\Sigma',\Delta)=A(\Sigma,\Delta)\oplus G,

so the required nondegeneracy holds after subdivision if and only if in holds before.


Since {\Delta} can be embedded in the boundary of its regular neighborhood {N} (radial projection), there exists a refinement {\tilde\Sigma} of {\Sigma}, not affecting {\Delta}, such that

\displaystyle \Delta\subset \partial N\subset \tilde\Sigma.

This provides us with a sujective map

\displaystyle A^k(\partial N)\rightarrow A^k(\Delta).

{\partial N} is not exactly the hypersurface we need. Every element of the kernel of that map can be killed by introducing holes. This produces a hypersurface with boundary {E}. It is orientable. Double it and compactify it. By this, I mean I use scissors to open {\tilde\Sigma} along {E}. Then {E} becomes {D(E)}, a closed hypersurface, in a new {\hat\Sigma} which is still a homology sphere. By diagram chasing, one shows that {I^k(\tilde\Sigma,E)} satisfies bpp if and only if {I^k(\hat \Sigma,D(E))} does,

8.2. Summary of what we have done

  1. Kronecker’s Lemma allows to construct Lefschetz elements in {\Sigma} provided we can prove nondegeneracy of pairing

\displaystyle \mathrm{ker}(\sum_{v\in W}) \text{ in }A(lk_w \Sigma)_k

holds.

  1. Nondegeneracy of pairing at {I(\Sigma,\Delta)} is implied by Hard Lefschetz for hypersurfaces. Using a hypersurface {E} containing {\Delta}, with {A^k(E)\simeq A^k(\Delta)} covers the general case.

    Each step wins a dimension.

    There is an issue: the above kernel, intersection of two monomial ideals, is not monomial in general. The above kernel is the orthogonal complement to a monomial ideal, provided

\displaystyle \bigcup_{v\in W} st_v\Sigma

and

\displaystyle (\bigcup_{v\in W} st_v\Sigma)\cup st_w\Sigma

are submanifolds. This fact follows from shellability, but it is too strong an assumption.


Recall that we want Hard Lefschetz for {\Sigma} a {2k}dimensional sphere. Order vertices arbitrarily. Embed {\Sigma} into a {2k-1}-manifold {\bar\Sigma}. Then Hard Lefschetz holds for {\Sigma} in degree {k} if and only if {I(\bar\Sigma,\Sigma)} satisfies bpp in degree {k+1}, i.e. if and only if {I(\bar\Sigma,\Sigma^{(\le k)})} satisfies bpp. We manage to find a refinement {\hat\Sigma} of {\bar\Sigma} such that there exists a hypersurface {E} with boundary such that {I(\Sigma,E)} is unchanged, and…

9. Transcendentality


Let {\mu} be a simplicial {\mathbf{k}}-cycle of dimension {d-1}. I mean a pair {(|\mu,\mu)} where {|\mu|} is a simplicial complex and {\mu} is a {d-1}-cycle in it, hence represents a class in {H_{d-1}(|\mu|,\mathbf{k})}.


On defines

\displaystyle A(|\mu|,\Theta)=\mathbf{k}[X]/I(\mu).

The top-degree chain {\mu} can also be viewed as a cochain {\check{\mu}}. It defines a pairing of degree {d} to {\mathbf{k}}. I view {\check{\mu}} as a quotient of {H^{d-1}(|\mu|)}, hence of {A^d(|\mu|)}. Let {B(\mu)} be the smallest quotient of {A(|\mu|)} such that {B^d(\mu)=\check{\mu}}. Then

\displaystyle B(\mu):=B(\mu,\Theta)

is a Poincare duality algebra, with fundamental class in degree {d}. Does {B(\mu)} satisfy Hard Lefschetz?


Consider the field extension {\tilde{\mathbf{k}}} of {\mathbf{k}} generated by extra algebraically independent elements, the rows of {\Theta} and {\ell}.

Theorem 20 Let {\mu} be a {d-1}-cycle over {\mathbf{k}}. Then, over {\tilde{\mathbf{k}}},

  1. {B(\mu)} satisfies Lefschetz property.
  2. {B(\mu)} satisfies the Hall-Lamac relations.
  3. If {char(\mathbf{k}=2}, then {Q_{k,\ell}(u,u)\not=0} for all nonzero elements {u\in B^k(\mu)}.


Open question: what happens in other characteristics?

9.1. Corollaries


Observe that if {\Sigma} is a sphere, then

\displaystyle A(\Sigma)=B(\text{fundamental class}).

In general, if {M} is an orientable, closed manifold, we define

\displaystyle B(\text{fundamental class})=A(M)/\mathrm{ker}(A(M)\rightarrow\bigoplus_{v}A(st_v M)).

Corollary 21 If {P} is an orientable pseudomanifold, then {B(\text{fundamental class}).} satisfies Hard Lefschetz.


We are now very far from any geometric interpretation.

9.2. The degree function


Define {\mathrm{deg}:B^d(\mu)\rightarrow\mathbf{k}} as follows. For {F} a facet of {|\mu|} of dimension {d-1-}, set

\displaystyle \mathrm{deg}(x_F)=\frac{\mu_F}{\mathrm{det}(|\Theta_F|)},

where {|\Theta_F|} denotes the minor in the matrix {\Theta} defined by {F}.


AB: Morozov’s Lemma.

10. Transcendentality (continued)


Recall that {\tilde{\mathbf{k}}} is the field of rational functions in {\ell} and the rows of {\Theta}. Today, we want to prove the following

Theorem 22 Let {\mu} be a {d-1}-cycle over {\mathbf{k}}. Then, over {\tilde{\mathbf{k}}},

  1. {B(\mu)} satisfies the Hard Lefschetz property.
  2. {B(\mu)} satisfies the Hall-Lamac relations.
  3. If {char(\mathbf{k})=2}, then {Q_{k,\ell}(u,u)\not=0} for all nonzero elements {u\in B^k(\mu)}.

11. Basic reductions and lemmas

11.1. Reduction to the middle dimension


Let {\mu} be a {d-1}-cycle. Let {I} be a graded ideal in {B(\mu)}, {\ell\in B^1(\mu)}. Then {I} satisfies Hall-Lamac relations with respect to {\ell} in degree {k} if

\displaystyle Q_{k,\ell}:B^k(\mu)\times B^k(\mu)\rightarrow\mathbf{k}

does not degenerate on {I^k}. We want to establish this by only considering {d=2k}.


The suspension of {\mu} is the free join of {\mu} with a {0}-cycle, i.e. {[-1,1]\times\mu/\sim} where {1\times\mu} and {-1\times \mu} are collapsed to distinct points {n} and {s}. Then {\mathrm{susp}(\mu)} is a {d}-cycle. Let {\Theta} be a linear system of parameters (for an Artinian reduction) for {A(|\mathrm{susp}(\mu)|)}. Projection along {n} yields a linear system of parameters {\bar\Theta} for {A(|\mu|)}.

Theorem 23 Assume that {k<d/2}. Let {J} be any ideal in {B(\mu,\bar\Theta)}. The following are equivalent:

  1. The Hall-Lamac relations for {J} in {B(\mu,\bar\Theta)} in degree {k} and with respect to {\ell}.
  2. The Hall-Lamac relations for {x_n J} in {B(\mathrm{susp}(\mu),\bar\Theta)} in degree {k+1} and with respect to the coordinate {x_n}.


Let {\Gamma\subset |\mu|} be a subcomplex. If

\displaystyle J=I(\mu,\Gamma):=\mathrm{ker}(B(\mu)\rightarrow B(\Gamma\subset \mu))

where {B(\mu\subset\Gamma)} is the quotient of {A(\Gamma)} obtained like {B(\mu)} is obtained from {A(|\mu|)}.


Observe that

\displaystyle x_n J=I(\mathrm{susp}(\mu),\mathrm{susp}(\Gamma\cup|\mathrm{susp}(\mu))).


Let us prove Theorem 23. Consider the element {x_n+x_s+\ell_{|\mu|^0}} of {\bar\Theta}. Then there is a commuting square between multiplication by {\ell^{d-2k}} at the level of {\mu} and multiplication by {x_n^{d-2k-1}} at the level of the suspension. Restricting to {J} proves the theorem.

11.2. Squarefree monomial generation


Let {\Delta} be a simplicial complex. Then, for any linear system of parameters {\Theta}, {A(\Delta,\Theta)} is generated by squarefree monomials in each graded component.


I explain this for a generic choice of {\Theta}. I first construct a larger complex {K} where it is true. Let {K} be the {d-1}-skeleton of the simplex on {n} vertices. Let {\Theta} be a linear system of parameters for {\mathbf{k}[K]}. Since {K} is shellable, we know how to give a basis for {A} when we attach a new face {F}: we add a generator {x_\sigma} where {\sigma} is the smallest face of {F} which is not contained in the complement of {F}. Now generation by squarefree monomials for {K} implies the same result for {\Delta}.

11.3. Continuation of the proof


Recall: if {F} is a facet of {|\mu|}, then

\displaystyle \mathrm{deg}(x_F)=\frac{\mu_F}{\mathrm{det}(|\Theta_F|)},

where {|\Theta_F|} denotes the minor in the matrix {\Theta} defined by {F}.

Lemma 24 (Lee) Let {\tau} be a simplex of cardinality {k}, then

\displaystyle \mathrm{deg}(x_\tau^2)=\sum_{F\text{ facet of }|\mu|\text{ containing }\tau}\mathrm{deg}(x_F)\frac{\prod_{i\subset\tau}[\Theta_{F\setminus i}]}{\prod_{j\in F\setminus\tau}[\Theta_{F\setminus j}]}.


I want to evaluate {\mathrm{deg}(u^2)} for {u\in B^k(\mu)}. Let {\sigma} and {\tau} be disjoint simplices of cardinality {k} in {|\mu|}. Let me view them as differential operators. Let {\rho\in\mathbf{k}(\Theta)} denote the element used to define the degree. Then {\partial_\sigma^\tau} is

\displaystyle \frac{d \sigma_1+t\tau_1}{dt}_{|t=0}\rho,

i.e. a derivative after performing a column operation on {\Theta}.

Lemma 25

\displaystyle \partial_\sigma^\tau (\mathrm{deg}(x_\sigma^2))= (\mathrm{deg}(x_\tau x_\sigma))^2. In particular, in characteristic {2}, \displaystyle \partial_\sigma^\tau\mathrm{deg}(u^2)=(\mathrm{deg}(x_\sigma u))^2.


Thanks to Lee’s Lemma, if {v} is any vertex not in {st_\tau(|\mu|)}, then {\mathrm{deg}(x_\tau^2)} is independent of {\Theta_v}.

11.4. Let me first treat characteristic {2}


Let {u\in B^k(\mu)}. We must show that {\mathrm{deg}(u^2)\not=0}. By Poincare duality, because {B(\mu)} is generated by squarefree monomials, there exists {\sigma} of cardinality {k} such that {x_\sigma\cdot u\not=0}. The ideal of {x_\sigma} of {B(\mu)} in degree {d} is isomorphic to {B^k(lk_\sigma(\mu))}, and {lk_\sigma(\mu)} is again a cycle. {B(lk_\sigma(\mu))} is generated by {x_\tau}, {\tau} of cardinality {k}. Therefore

\displaystyle u=\lambda_\tau x_\tau+\sum_{\omega\notin st_v(\mu|)} \lambda_\omega x_\omega .

Compute, using the lemma,

\displaystyle \partial_\sigma^\tau \mathrm{deg}(u^2)=\partial_\sigma^\tau (\mathrm{deg}(\lambda_\tau^2 x_\tau^2)+\sum_\omega \mathrm{deg}(\lambda_\omega^2 x_\omega^2)=\lambda_\tau^2 (\mathrm{deg}(x_\sigma x_\tau) )^2,

since all other terms

\displaystyle \partial_\sigma^\tau \mathrm{deg}(\lambda_\omega^2 x_\omega^2)

vanish. Hence

\displaystyle \partial_\sigma^\tau \mathrm{deg}(u^2)=(\mathrm{deg}(\lambda_\tau x_\sigma x_\tau))^2=(\mathrm{deg}(x_\sigma u))^2\not=0.

11.5. General characteristic


Let {u=\sum \lambda_a x_a}.

Definition 26 Say {u} is compatible with {\sigma} if

  1. The support {|u|} of {u} intersects {st_\sigma(|\mu|)} is a single face {\tau} and {u_{|\tau}=1}.
  2. If {\tau'} is any other face in {|u|}, then {st_\sigma(|\mu|)\cap\sigma} is a simplex {\sigma'}, face of {\sigma}, or empty, and {u_{|\tau'}} is independent of one of the variables of {\sigma-\sigma'}.


Let us extend again the ground field into {\mathbf{k}(\Theta,\Theta')} using a second copy of {\Theta}. Let us associate to {u} an element {u'\in\mathbf{k}(\Theta')} obtained by replacing {\Theta_{ij}} with {\Theta'_{ij}}.

Lemma 27 For {u} compatible with {\sigma},

\displaystyle \partial_\sigma^\tau \mathrm{deg}(uu')=\mathrm{deg}((\partial_\sigma^\tau u)u')+(\mathrm{deg}(x_\sigma u))^2.


If one substitues {u=u'}, one gets

\displaystyle \partial_\sigma^\tau \mathrm{deg}(u^2)=\mathrm{deg}((\partial_\sigma^\tau u)u)+(\mathrm{deg}(x_\sigma u))^2.

By Poincare duality, we may assume that {x_\sigma u\not=0}. If {u} is in a monomial ideal of {\mathbf{k}(|\mu|)}, then so are {u'}and {\partial_\sigma^\tau u}.


There remains to show how to find a compatible element. For us, the ideal is {I=I(\mu,\Delta)} for some subcomplex {\Delta} of {|\mu|}, {\mu} has dimension {2k-1}. Consider the pairing on {I^k}, with values in {\mathbf{k}(\Theta,\Theta')}. Let {u} be represented by squarefree monomials. We know that there exists {x_\sigma} such that {u\cdot x_\sigma\not=0} if {\sigma\in|\mu|-\Delta}, hence {\sigma\subset\Delta}.


Intuitively, {B^k(lk_\sigma(|\mu|))} has dimension {1}. We can assume that the class of {u} in {I(\mu,\Delta)} is represented by an element {u_0} such that the support of {u_0} intersect {st_\tau(|\mu|)} in a single face {\tau}. Without loss of generality, we can assume that the coefficient is {1}. Pick a face {\tau'} is the support of {u_0}. I can assume that {u_0 \cdot x_{\tau'\cup v}\not=0}. Furthermore, {B(lk_{\sigma\cup v})} is generated by a single element in top degree. This combines into compatibility.

11.6. Application

Definition 28 Say a simplicial complex is {2}-Cohen-Macaulay if for every vertex {v\in\Delta}, {A(\Delta-v)} is Cohen-Macaulay.


Example. Triangulations of disks are not Cohen-Macaulay, but {2}-Cohen-Macaulay.


Algebraically, {2}-Cohen-Macaulay complexes are level, that is, for all {\alpha\in A^k(\Delta)}, there exists {\beta\in A^{d-k}(\Delta)} such that {\alpha\cdot \beta\not=0} (see Stanley’s book).


Let {M} be a quotient of {A^d(\Delta)}. Let us define

\displaystyle B(M)=A(\Delta)/\mathrm{ker}(A(\Delta)\rightarrow\bigoplus_{\mu\text{ generator for }M}B(\mu))

The level property implies that

\displaystyle A(\Delta)=B(A^d(\Delta)).

Hard Lefschetz implies that multiplication by {\ell^{d-2k}} is an injection on {B^k(M)}. Thus we get

Corollary 29 (of Hard Lefschetz for cycles) The dimensions of {B^i(M)} are increasing while {i\le d/2}.

Beware that Poincare duality need not hold, so this does not imply unimodality of these dimensions (in fact, counterexamples exist). Nevertheless, such a corollary is valuable, since the class of {2}-Cohen-Macaulay complexes is much wider that homology spheres.

Posted in Course, Uncategorized | Leave a comment

Notes of Mikolaj Fraczyk’s YGGT IX lecture, february 27th, 2020

Growth of mod-{p} homology of higher rank lattices and mapping class groups

1. Homology growth

Let {\Gamma} be a finitely generated group, and {\Gamma_i} a sequence of finite index subgroups. Let {F} be some field. I am interested in the growth of Betti numbers. When

\displaystyle  \lim_{i\rightarrow\infty}\frac{b_k(\Gamma_i,F)}{[\Gamma:\Gamma_i]}

exists, what is it?

Note that if {\Gamma} is of class {\mathcal{F}_k}, i.e. acts freely on a {k}-connected complex {X}, {b_k(\Gamma_i,F)} grows at most like the number of cells of {\Gamma_i\setminus X}, i.e. linearly with the index. Also, the Euler characteristic is proportional to the index, therefore linear growth is to be expected.

The answer is known in characteristic zero.

Theorem 1 (Lück 1994) If {\Gamma} is of class {\mathcal{F}_k}, {F={\mathbb R}} and {\Gamma_j} are nested normal subgroups, then the limit exists and for all {j< k},

\displaystyle  \lim_{i\rightarrow\infty}\frac{b_j(\Gamma_i,{\mathbb R})}{[\Gamma:\Gamma_i]}=b_j^{(2)}(\Gamma)

is equal to an {\ell^2}-Betti number.

What about other fields as coefficients?

1.1. Locally symmetric spaces

I especially like the following examples. If {G} is a semisimple Lie group and {K<G} a maximal compact subgroup, then the symmetric space {X=G/K} has nonpositively curved left-invariant metrics, hence is contractible. Lattices exist, thery need not be cocompact. Nevertheless, they have finite covolume {Vol(\Gamma\setminus X)}.

2. Benjamini-Schramm convergence

Lück restricted to normal subgroups, this restriction can be weakened. The replacement is a geometric condition: Benjamini-Schramm convergence. This means that for every {R>0}, the relative volume of the {R}-thin part tends to zero. {R}-thin means the locus of points where injectivity radius is {\le R}.

Theorem 2 (Abert, Bergeron, Biringer, Gelander) If the sequence {\Gamma_i\setminus X} converges to {X} in Benjamini-Schramm’s sense, then

\displaystyle  \lim_{i\rightarrow\infty}\frac{b_j(\Gamma_i,{\mathbb R})}{Vol(\Gamma_i\setminus X)}=b_j^{(2)}(X).

Again, this is over the reals, generalization to other fields is unclear.

Over the reals, one uses harmonic representatives, and spectral properties of the Laplacian. This is not available for finite fields.

3. Results

Theorem 3 (Fraczyk 2018) If {X} is a higher rank symmetric space and {\Gamma_i\setminus X} converges to {X} in Benjamini-Schramm’s sense, then

\displaystyle  \lim_{i\rightarrow\infty}\frac{b_1(\Gamma_i,\mathbb{F}_2)}{Vol(\Gamma_i\setminus X)}=0.

Ongoing work: growth of torsion.

Theorem 4 (Abert-Bergeron-Fraczyk-Gaboriau) For congruence subgroups {\Gamma(N)} of {Sl_n({\mathbb Z})}, and all {k\leq n-2},

\displaystyle  \lim_{i\rightarrow\infty}\log|H_k(\Gamma(N),{\mathbb Z})_{tor}|=o([\Gamma:\Gamma(N)]^{n/(n+1)}).

4. Sketch of proof of Theorem 4

One uses a less naive simplicial complex on which {\Gamma_i} acts. The requirements are

  1. {\tilde Y_i} is {k+1}-connected,
  2. {|Y_i|=o([\Gamma:\Gamma_i])}.

Lemma 5 Assume that {\Gamma} acts on a complex {\Delta} satisfying

  • (A1) {\Gamma\setminus\Delta} is finite.
  • (A2) {\Delta} is {k}-connected.
  • (A3) Each simplex stabilizer {\Gamma_\sigma} is of class {\mathcal{F}_k} and contains a normal subgroup isomorphic to {{\mathbb Z}^{n_\sigma}}.

Then we can build the needed spaces {Y_i} for every normal chain {(\Gamma_i)}.

Indeed, let {\Gamma_{\sigma,i}=\Gamma_\sigma\cap\Gamma_i}. Using standard homotopy theory (see Geoghegan’s book, the keyword in stack of CW complexes), one glues together spaces {Y_{\sigma,i}\sim K(\Gamma_{\sigma,i},1)} into a space {Y_i} fibering over {\Gamma_i\setminus \Delta} with fibers {Y_{\sigma,i}}.

One counts cells,

\displaystyle  |Y_i|=\sum_{\sigma\in \Gamma_i\setminus\Delta}|Y_{\sigma,i}|,

\displaystyle  [\Gamma_\sigma\cap\Gamma_i]=\sum_{\sigma\in \Gamma_i\setminus\Delta}[\Gamma:\Gamma_{\sigma,i}]

Claim. The exist classifying spaces {Y_{\sigma,i}} such that

\displaystyle  |Y_{\sigma,i}|=o(\frac{[\Gamma:\Gamma_{\sigma,i}]}{[{\mathbb Z}^{n_\sigma}:{\mathbb Z}^{n_\sigma}\cap\Gamma_{i}]}).

This implies that {|Y_i|=o([\Gamma:\Gamma_{\sigma,i}])} and proves a weak form of Theorem 4. To win an exponent, we need control on norms of differentials.

Proof of the claim. We take for {\Delta} the Solomon-Tits complex, whose cells are index by {{\mathbb Q}}-parabolics, and cells are incidents when parabolics are contained in each other. Vertices are maximal parabolics et top dimensional celles correspond to Borel subgroups. This complex is homotopic to a wedge of {n-2}-spheres. The cell stabilizers are intersections of {\Gamma} with parabolic subgroups (stabilizers of flags in case of {SL_n}).

5. Proof of Theorem 3

Take {G=Sl_3({\mathbb R})}. We want to estimate {b_1(\Gamma_i\setminus G,\mathbb{F}_2)}, assuming that injectivity radius tends to infinity.

We show that every class is represented by a cycle of length which is {o(Volume)}. Start with any representative, typically of length {\sim Volume}. If two pieces of the cycle are nearby, I can remove these pieces. Mod {2}, this is possible. Then realize by geodesic. Due to Euclidean character, it can be wobbled inside maximal flats without increasing length too much. This puts again pieces close to each other. Repeating, one reduced length substantially.

5.1. Generalization

Our methods applies to Mapping Class Groups, with {G/K} replaced with the curve complex.

Posted in Workshop lecture | Tagged | Leave a comment

Short presentations, YGGT IX, february 25th, 2020

5 minutes presentations

1. Anschel Schaffer-Cohen

I am interested in big mapping class groups. Let such a group {G} act continuously on a graph {\Gamma}, with unbounded orbits. Assume the graph is hyperbolic, plus some kind of other property. Then {\Gamma} is quasiisometric to {G}. This should lead to examples of hyperbolic mapping class groups.

2. Vladimir Vankov

I study cube complexes {X} with groups {G} acting. If one knows some finite quotient of {G}, can one say something on {X}. To each edge, there is a matrix associated matrix, containing the quotient. Take determinants, yielding characters. This has led sufficient conditions for cube complexes to be special.

An interesting situation is provided by the short exact sequence where a RAAG appears like an extension of {{\mathbb Z}} by a Bestvina-Brady group.

3. Josh Faber

Let {H,G} be finitely generated subgroups of a free group {F}. The Hanna Neumann conjecture states that

\displaystyle  \bar r(H\cap G)\leq \bar r(H)\bar r(G),

where {\bar r(H)}

Joel Friedman’s solution, using sheaves, suggests a new conjecture:

\displaystyle  \bar r(H\cap G)\leq \frac{1}{6}\bar r(H)\bar r(G),

I proved a slightly stronger result, with a finer constant. I am also interested in generalizations to cube complexes.

4. Sam Hughes

I am interested in irreducible lattices {\Gamma<H_1\times H_2} in a product of groups which are either Lie or tree automorphism groups. Example: {SL_2({\mathbb Z}[\sqrt{2}])<SL_2({\mathbb R})^2}.

There exist nonuniform examples in the Lie group case. They are linear, just infinite. In the tree case, there exist irreducible lattices, sometimes simple, sometimes just infinite. I provide examples illustrating these properties in mixed products.

5. Nick Bell

Mirzakhani 2016: Fix a closed geodesic {\gamma_0} in a surface {S}. The number of curves in {Mod(S)\gamma_0} of length {\leq L} is {\sim C_s D_p E_{\gamma_0}\,L^{6g-6+2n}}, where {C_s, D_p, E_{\gamma_0}} are known.

I have shown that this works for arcs as well.

6. Sam Shepherd

Leighton’s theorem states that if two finite graphs have a common universal cover, then they have a common finite cover.

Question: does Leighton’s theorem hold for other spaces?

I proved that the answer is yes in certain cases: graphs with fins, certain {2}-complexes. I showed that answer is no for NPC cube complexes, but the question remains open for special NPC cube complexes.

Gardan-Woodhouse did it independently.

7. Lucas De Souza

I work on group compactifications. Gerassimov’s Attractor-Sum Theorem states that if {G} acts properly and cocompactly on {X} and has a convergence action on {Y}, a convergence action of {G} can be defined on the join of {X} and {Y}.

I introduce a property of a compactification of {X}, called perspectivity, that garantees that such sums go through…

8. Shivam Arora

I study subgroups of hyperbolic groups. When are they hyperbolic? Note that hyperbolicity implies finite presentation. The converse fails (example by Bestvina-Brady).

Gersten proved that if {G} is hyperbolic with cohomological dimension {2} over {{\mathbb Z}}, then every finitely presented subgroup is hyperbolic.

I have been able to replace {{\mathbb Z}} with {{\mathbb Q}} coefficients.

9. Sahana H Balasubramanya

I study the class of acylindrically hyperbolic groups. It contains relatively hyperbolic groups, most RAAGs, {MCG}s, {Out(F_n)},…

Acylindrically hyperbolic groups come with a generalization of peripheral groups, hyperbolically embedded subgroups. Are they QI invariant? Seems hard.

I prove that if {H<G} is normal of finite index and AH-accessible, then {G} is also AH-accessible (hence acylindrically hyperbolic). AH-accessible means that {G} admits a largest acylindrical action on a hyperbolic space.

10. Feng Zhu

A really inefficient way of proving that a group is hyperbolic.

Let {\Gamma} be finitely generated and torsion free. Say that a reprrsentation {\rho:\Gamma\rightarrow Sl_n({\mathbb R})} is {P_1}-Anosov if for every element {\gamma\in\Gamma},

\displaystyle  \log\frac{\sigma_1}{\sigma_2}(\rho(\gamma))\geq \mu|\gamma|-C.

Here, {\sigma_i} are singular values.

Kapovich-Leeb-Porti, BPS proved that if {\Gamma} admits a {P_1}-Anosov representation, then {\Gamma} is hyperbolic.

The point is that is shows that simple Lie groups contain plenty of nonfree hyperbolic subgroups.

I proved a relative version of this.

11. David Sheard

Flag manifolds contain Schubert varieties, usually not smooth, sometimes rationnally smooth, this has led me to Coxeter groups. Indeed, there is a Weyl group {W} associated with a flag variety. Rationnal smoothness is related with a Poincaré polynomial associated to {W}.

There is an order on a Weyl group {W}, the corresponding Poincaré polynomial is

\displaystyle  P_W(q)=\sum_{u\in [e,w]}q^{\ell(u)}

Rationnal smoothness is equivalent to {P_W} being palindromic. There is a criterion for that.

12. Nora Szoke

I would like to tell you what sofic groups are. An {(n,\epsilon)}-approximation of a Cayley graph {Cay(G,S)} is a finite graph {F} such that for most vertices of {F}, the {n}-ball at that vertex is isomorphic to the {n}-ball in {Cay(G,S)}.

Say {G} is sofic if {\forall\epsilon} and {n}, there exists an {(n,\epsilon)}-approximation. Obviously, residually finite groups are sofic. Amenable groups are sofic, because Folner sets do the job. The class of sofic groups is stable under direct products, subgroups, limits, free products. We do not know wether there exist nonsofic groups.

13. Dounnu Sasaki

I want to introduce the notion of subset current, due to Bowditch. Let {G} be a hyperbolic group. Subset currents {SC} form a completion of the semigroup of positive linear combinations of quasiconvex subgroups. Reduced rank gives rise to a morphism to {{\mathbb R}_+}.

I showed that the inequality of the Strengthened Hanna Neumann’s conjecture for subgroups of a free group,

\displaystyle  \sum_{k\in }\bar r(H\cap k^{-1}Gk)\leq \bar r(H)\bar r(G),

extends to subset currents. I think this is a useful tool.

14. Jonathan Fruchter

I want to study the first order theory of acylindrically hyperbolic groups. First order theory is the collection of sentences with only quantifyers, constants, variables and monomials.

Lubotzky proved that a finitely generated profinite group is uniquely determined by its first order theory. Free groups are not (Sela), they all have the same first order theory. Say a group has property (*) if {Th(G)=Th(G*Z)}.

With Simon André, we proved that if {G} is acylindrically hyperbolic, then {Th^+(G)=Th^+(F_2)}. Such groups also slightly a slightly weakened form of property (*).

15. Hannah Hoganson

According to Gromov, {G} is nonamenable {\iff} {G} admits a Ponzi scheme. Thompson’s group {F} acts on a rooted binary tree, with two generators. I have a good understanding of its Cayley graph. I have tried to construct a Ponzi scheme, without success.

Instead, I studied {L^p} metrics on Teichmüller space. They turn out to be incomplete. The completion includes pinched curves, forming a stratification by lower genus Teichmüller spaces.

16. Xenia Flamm

I begin studying higher Teichmüller theory. This deals with homomorphisms of a surface group into a Lie group (up to conjugacy). A higher Teichmüller is a union of connected components of this space, consists of discrete and faithful homomorphisms. This exists for {Sl_n({\mathbb R})}, {Sp(2n,{\mathbb R})}, {SO(p,q)}, for instance.

I want to understand compactifications of these higher Teichmüller spaces. A possible approach uses real spectra, following Brumfiel.

17. George Domat

Projection complexes, constructed by Bestvina-Bromberg-Fujiwara, deal with the projections to one of them of the lifts in the universal cover of a geodesic. These are quasitrees on which the fundamental group {G} acts. This gives rise to quasimorphisms. I use this to show that the abelianizations of certain big mapping class groups contain infinite direct sums of copies of {{\mathbb Q}}.

18. Yuri Santos Rego

To a conjugacy class in the free group, I associate a combinatorial object I call a “circle”. Artin did the same with braids and produced knots. Sapir and Guba did the same for diagram groups, Thompson groups is one of them. Inside the big MCG of the complement of a Cantor set, there is a “braided Thompson group” to which a similar procedure can be applied, the objects associated to conjugacy classes are Haken manifolds.

19. Thomas Ng

I like to decide wether two elements in a finitely presented group generate a free goup or not. Hard. Replace them by some high powers. Still hard. If group has some nonpositive curvature, there is some hope. Free semigroup would do.

With Radhika Gupta ans Junkiewicz, we prove that if {G} acts on a {2}-dimensional {CAT(0)} cube complex without global fixed point, for any generating set, either {G} is virtually abelian or there exist distinct elements of length {<30000} that generate a free subgroup. This applies to Higman’s group (which acts of a {CAT(0)} cube complex with stabilizers Baumslag-Solitar groups.

20. Macarena Arenas

In hyperbolic groups, two loops which bound an annulus bound an annulus with area linear in their lengths. With Wise, we generalize this to any number of loops, first in genus {0} and then in higher genus.

21. Laurent Hayez

There are two distinct 6 vertex 7 edges graphs which have the same spectrum. I am interested in spectral properties of infinite graphs, and especially Cayley graphs. For those, there is a spectral measure {\mu}.

Question. Does {\mu} determine the group up to isomorphism?

22. Nima Hoda

Bisimplices are a trick to add higher dimensional cells to a bipartite graph, for instance to make it contractible. Cubes just won’t do (for instance, one cannot glue a cube to {K_{3,3}}).

Are there bipartite analogues to simplices? Answer: inductively define cells filling complete bipartite graphs. Use discrete Morse theory

23. Sami Douba

In a group {\Gamma}, let {\gamma,\gamma_1,\ldots,\gamma_n} be elements such that {\gamma=\prod[\gamma_i,\gamma_j]} commutes with the others. Then in any finite dimensional unitary representation of {\Gamma}, {\gamma} has finite order.

This implies for instance that Heisenberg group, fundamental groups of unit tangent bundles of surfaces, most mapping class groups cannot embed in a unitary group.

24. Matthieu Joseph

Let me share with you what I call the Infinite Noodle Lemma. I used it to build {L^p} orbit equivalences between groups.

Theorem 1 Let {\phi} a fixedpoint free involution on the set {{\mathbb Z}}. I draw an arc in the upper half plane between {x} and {\phi(x)} for every {x}. Assume that the arcs do not intersect. Reflect and translate the picture by {1}. Then one gets a curve.

25. Matthew Conder

Question. Given matrices {A,B\in Sl_2({\mathbb R})}, when is the subgroup {\langle A,B\rangle} they generate free?

For instance, there is no complete answer for

\displaystyle  A=\begin{pmatrix} 1 & x \\ 0 & 1 \end{pmatrix},\quad B=\begin{pmatrix} 1 & 0 \\ x & 1 \end{pmatrix}

The ping-pong lemma gives a sufficient condition. If you ask simultaneously that {\langle A,B\rangle} be also discrete, i becomes easier: there is an algorithm that decides it (Eick-Kirschner-Leedham-Green).

I have solved the corresponding problem in {SL_2({\mathbb Q}_p)}, where it is easier, because ping-pong is easier there.

26. Holli Chopra

I plan to work on Cannon’s conjecture: if {G} is a Gromov-hyperbolic group whose boundary is homeomorphic to a {2}-sphere, does {G} act geometrically on hyperbolic {3}-sphere?

My approach is to modify Perelman’s methods.

27. Michal Buran

Let me tell you of nightmare towns, with randomly chosen railway system.

Two random permutations of {\mathfrak{S}_n} generate {\mathfrak{S}_n} with high probability, even if a number of constraints, encoded in a graph with {n} vertices, are imposed.

28. Francisco Nicolas

Kähler groups are fundamental groups of compact Kähler manifolds. These are complex manifolds equipped with hermitian metrics whose imaginary part os a closed {2}-form. Example: complex tori, Riemann surfaces, complex projective spaces and their complex submanifolds…

Which finitely presented groups are Kähler? {{\mathbb Z}^{2n+1}} is not, since the first Betti number of a Kähler manifold is always even (Hodge theory).

I plan to study finiteness properties of Kähler groups, by constructing examples of Kähler manifolds as level sets of functions, in the spirit of Bestvina-Brady.

Posted in Workshop lecture | Tagged | Leave a comment

Notes of Radhika Gupta’s YGGT IX lecture, february 25th, 2020

Non-uniquely ergodic arational trees in the boundary of Outer space

With M. Bestvina and J. Tao.

1. Surface case

1.1. Laminations

Let {S} be a hyperbolic surface. A geodesic lamination is filling if the complement is made of discs or once-punctured discs. It is minimal if every leaf is dense in it.

Question. For measured laminations, does filling, minimal imply uniquely ergodic?

Minimal, filling laminations occur as early as 1969 in Veech’s work on interval exchange transformations.

Nonuniquely ergodic filling laminations appear in Gabai’s work.

Leminger-Lenzhen-Rafi have constructed nonuniquely ergodic minimal filling laminations.

1.2. Trees

The associated tree is the set of leaves in the universal cover, it inherits a metric from the transverse measure. If the lamination is nonuniquely ergodic, the tree supports two projectively distinct metrics: it is non ergometric.

2. Free group

Let {F_n} denote the free group on {n} generators. We collect all the simplicial actions of {F_n} on trees which are free and minimal, up to homothety. Collapsing edges equivariantly allows to connect combinatorially distinct actions, it provides a topology, this is Outerspace {CV_n}. The outer automorphism group {Out(F_n)} acts on it. We think that Outerspace plays for free groups the role played by Teichmüller space for surface groups.

{CV_n} has a boundary, consisting of isometric actions on real trees.

The analogue of the curve complex, a hyperbolic space with ideal boundary the space of minimal filling laminations, is the free factor complex, with boundary the space of arational trees.

Question. Do there exist arational trees which support non uniquely ergodic measures?

Theorem 1 (Bestvina-Gupta-Tao) There exist non-uniquely ergometric non-geometric arational trees.

2.1. Main construction

We start with two automorphisms of {F_7}, {\phi} which is fully irreducible on the first {3} generators and identity on the others, and {\rho} which cyclically permutes all generators

Set {\phi_r:=\rho \phi^r}. One picks a sequence {r_j} and let act successively {\phi_{r_1},\phi_{r_2},\ldots}. One orbit in {CV_n} is a zig-zag whose closure in compactified Outerspace contains an arc. On the other hand, the corresponding orbit in the free factor complex converges to some arational tree. This produces two different metrics on that tree.

Posted in Workshop lecture | Tagged | Leave a comment

Notes of Yair Hartman’s YGGT IX lecture, february 25th, 2020

Which groups have bounded harmonic functions?

Joint with Frisch, Tamuz and Vahidi-Ferdowsi.

Familiar groups act on boundaries

1. Dirichlet’s problem

Impose the temperature along the boundary {B} of a domain {D}, wait until equilibrium. What is the temperature inside? Temperature is a bounded function {f}, harmonic in the interior. In case of a round disk, the solution is provided by Poisson’s transform. For general domains, it can be obtained as follows. Given {x} inside, launch a random walk and record where it hits the boundary. This produces a probability measure {\nu_x} on the boundary. The temperature satisfies

\displaystyle  h(x)=\int_{B}f(b)d\nu_x(b).

This provides a bijection between {L^\infty} functions on {B} and bounded harmonic functions on {D}.

Note that all {\nu_x} define the same measure class, so {L^\infty} is unambiguous.

Liouville’s theorem states that all bounded harmonic functions on the plane are constant. I think of it in these terms: the boundary of the plane has only one point.

2. Measured groups

In order to define harmonicity, one merely needs a notion of average over neighbours.

Definition 1 A measured group is a countable discrete group {\Gamma} equipped with a symmetric probability measure whose support contains a generating set. Symmetric means that {\mu(g^{-1})=\mu(g)}.

A function {h:\Gamma\rightarrow{\mathbb R}} is {\mu}-harmonic if for all {g\in\Gamma},

\displaystyle  h(g)=\sum_{\gamma\in\Gamma}h(g\gamma)\mu(\gamma).

{H^\infty(\Gamma,\mu)} denotes the space of bounded harmonic functions.

A notion of boundary for a measured group has been developped since the 1960’s.

Theorem 2 (Furstenberg) Given {(\Gamma,\mu)}, there exists a standard probability {\Gamma}-space {\Pi(\Gamma,\mu)=(B,\nu)} such that

\displaystyle  L^\infty(B,\nu)=H^\infty(\Gamma,\mu).

This is called the (Furstenberg-)Poisson boundary of {(\Gamma,\mu)}.

Note that {B} has no topology on it, only a family of measures.

The random walk interpretation is still valid. Given {(\Gamma,\mu)}, pick independently elements {g_1,g_2,\ldots} according to {\mu} and consider the sequence {e\rightarrow g_1\rightarrow g_1g_2\rightarrow g_1g_2g_3\ldots}. Morally, {B} is the set of trajectories and {\nu_e} the hitting measure. This is litterally true when {(\Gamma,\mu)} is a free group with {\mu} the uniform measure on generators and their inverses.

Definition 3 Say {(\Gamma,\mu)} is Liouville if {\Pi(\Gamma,\mu)} is one point and {H^\infty(\Gamma,\mu)={\mathbb R}} has only constant functions in it.

Example. When {(\Gamma,\mu)} is {{\mathbb Z}} with {\frac{1}{2}} mass on {1} and {-1}, harmonic functions are arithmetic progressions. Only constant arithmetic progressions can be bounded.

Poisson boundaries behave well under taking quotients.

3. Which groups are Liouville

Question. Determine the classes of groups

\displaystyle  NA:=\{\Gamma\,;\,\forall\mu,\,\Pi(\Gamma,\mu)\not=\{*\}\} \quad\text{and}\quad H:=\{\Gamma\,;\,\forall\mu,\,\Pi(\Gamma,\mu)=\{*\}\}.

Theorem 4 (Furstenberg, Rosenblatt, Kaimanovich-Vershik) {NA} is the set of nonamenable groups.

Theorem 5 (Blackwell, Choquet-Deny, Dynkin-Malyotov, Margulis) Virtually nilpotent groups belong to {H}.

Theorem 6 (Frisch, Hartman,Tamuz,Vahidi-Ferdowsi) Among finitely generated groups, {H} is the class of virtually nilpotent groups.

3.1. Proof

The key concept in our proof is

Definition 7 {\Gamma} is ICC if all nonidentity conjugacy classes are infinite.

Starting from a finitely generated group, kill all FC central elements. In the quotient, do it again. Do it again and again (transfinite induction). Ultimately, one gets an ICC quotient. Since Poisson boundary behaves well under taking quotient, the problem is reduced to the ICC case.

Proposition 8 If {\Gamma} is finitely generated and ICC, it admits a probability measure with nontrivial Poisson boundary.

The ICC property is exploited as follows.

Lemma 9 If {\Gamma} is ICC, for every finite subset {B\subset\Gamma}, there exists {g\in G} such that {g^{-1}Bg\cap B\subset\{e\}}.

It means that starting with a ball, applying a suitable right translation blows it out into a sparse set of points.

We choose a probability measure {\mu} concentrated at a sequence of such snipers leaving to infinity very fast. Two different trajectories will frequently encounter this blowing out phenomenon and stay far apart. They must exit by different points, showing that {\Pi(\Gamma,\mu)} is more than one point.

The measures we construct are not finitely supported. Both variants of the above problems where one restricts to finitely supported measures are open.

Posted in Workshop lecture | Tagged | Leave a comment