\( \newcommand{\R}{\mathbb{R}} \newcommand{\Z}{\mathbb{Z}} \newcommand{\N}{\mathbb{N}} \newcommand{\from}{\colon} \DeclareMathOperator{\vol}{vol} \newcommand{\HH}{\mathbb{H}}\)

Notes on nilpotent groups and subriemannian manifolds

Table of Contents

Lecture 3: Subriemannian geometry

Examples of nilpotent groups and Lie algebras

Some examples of nilpotent groups and Lie algebras:

The Heisenberg group
\[\mathfrak{h} = \langle X, Y, Z \mid [X,Y] = Z, \text{all other pairs commute}\rangle\]
The Engel group
\[\mathfrak{e} = \langle X, Y, Z, W \mid [X,Y] = Z, [X,Z] = W, \text{all other pairs commute}\rangle\] This is a step–3 nilpotent group.
Filiform nilpotent Lie algebras
A nilpotent Lie algebra is filiform if the dimension is equal to the step plus \(1\); the Heisenberg group and the Engel group are both filiform. For \(k>1\), one filiform Lie algebra of dimension \(k\) is given by: \[\mathfrak{g}_k = \langle e_1,\dots, e_k\mid [e_i,e_j] = (j-i) e_{i+j} \text{ for all } i+j\le k\rangle;\] (Exercise: Check that this satsifies the Jacobi identity.)
Free nilpotent groups and Lie algebras

Let \(F_r\) be the free group on \(r\) generators, \(F_r=\langle g_1,\dots, g_r\rangle\) and let \(k\ge 1\). The free step--\(k\) nilpotent group of rank \(r\) is the group \(F_{r,k} := F_r/(F_r)_{k+1}\); any step--\(k\) nilpotent group with at most \(r\) generators is a quotient of \(F_{r,k}\).

Notably, the group \(F_{2,2}\) is isomorphic to the Heisenberg group.

The easiest way to describe this group is to embed it as a lattice in a nilpotent Lie group. Let \(\mathfrak{f}_r\) be the free Lie algebra of rank \(r\); then \(F_{r,k}\) is a lattice in the Lie group corresponding to \(\mathfrak{f}_{r,k} = \mathfrak{f}_r/(\mathfrak{f}_r)_{k+1}\).

One can describe \(\mathfrak{f}_r\) combinatorially; if \(e_1,\dots, e_r\) are the generators of \(\mathfrak{r}\), then each quotient \((\mathfrak{f}_r)_n/(\mathfrak{f}_r)_{n+1}\) is an abelian Lie algebra spanned by iterated brackets of the \(e_i\)'s. Because of the Jacobi identity, these iterated brackets are not linearly independent; Witt showed that \(\mathfrak{f}_{r,k}/\mathfrak{f}_{r,k+1}\) has a basis consisting of \[M_r(k) = \frac{1}{k} \sum_{d\mid k} \mu(d) r^{k/d}\] iterated brackets, where \(\mu\) is the Möbius function.

The Riemannian geometry of \(\mathbb{H}\)

The geometry of \(\mathbb{H}\) is a good starting point for understanding the geometry of nilpotent groups in general.

We represent \(\HH\) as \(\R^3\) equipped with the multiplication \[(x,y,z)\cdot(x',y',z') = \left(x+x', y+y', z + z' + \frac{xy'-yx'}{2}\right).\] (This is a little different from the coordinates we used before – this is the multiplication on \(\mathfrak{h}\) coming from the Baker–Campbell–Hausdorff formula.)

We define a left-invariant metric on \(\HH\): the standard basis of \(\R^3\) extends to three left-invariant vector fields

\begin{align*} X & = (1,0,-\tfrac{y}{2}) \\ Y & = (0,1,\tfrac{x}{2}) \\ Z & = (0,0,1). \end{align*}

We define a Riemannian metric so that these fields are orthogonal: \[dg^2 = dx^2 + dy^2 + \left(dz + \frac{y}{2} dx - \frac{x}{2} dy\right);\] this metric is left-invariant.

Since this is a Riemannian metric, we can calculate its geodesics: it turns out that the geodesics are based on helices. For example, suppose that \(\gamma=(\gamma_x,\gamma_y,\gamma_z)\) is a path from \(0\) to \((0,0,C)\) for \(C>0\). We will find conditions such that \(\gamma\) is length-minimizing.

We have \[\ell(\gamma) = \int_0^1 \sqrt{(\gamma_x')^2 + (\gamma_y')^2 + (\gamma_z' - \tfrac{\gamma_x\gamma_y'}{2} + \tfrac{\gamma_y\gamma_x'}{2})^2}\; dt.\]

We analyze this by defining two auxiliary functions. Let \(\pi\from \HH \to \R^2\) be projection to the \(xy\)–plane and let \[L(t) = \ell(\pi(\gamma([0,t]))) = \int_0^t \sqrt{(\gamma_x')^2 - (\gamma_y')^2}\; dt\] \[A(t) = \frac{1}{2} \int_0^t \gamma_y\gamma_x' - \gamma_x\gamma_y'\; dt.\] Note that \(A(t)\) is the signed area of the closed curve formed by joining the endpoints of \(\pi(\gamma([0,t]))\) by a line segment. Let \(L_\gamma =L(1)\) be the length of \(\pi \circ \gamma\) and let \(A_\gamma =A(1)\) be the signed area of \(\pi \circ \gamma\) (which is a closed curve).

Then \[\ell(\gamma) = \int_0^1 \sqrt{(L')^2 + (\gamma_z' - A')^2}\; dt.\] Consider the curve \(g(t) = (L(t), \gamma_z(t) - A(t))\). Then \(g\) connects \((0,0)\) to \((0,C-A_\gamma)\) and \(\ell(\gamma) = \ell(g)\) — in particular, we have \[\ell(\gamma) \ge \sqrt{L^2 + (C-A_\gamma)^2},\] and we can achieve \(\ell(\gamma) = \sqrt{L_\gamma^2 + (C-A_\gamma)^2}\) by reparametrizing \(\gamma_x\) and \(\gamma_y\) and adjusting \(\gamma_z\) so that \(g\) is a line segment.

So we want to minimize \(f(\gamma) = L_\gamma^2 + (C-A_\gamma)^2\) over closed curves \(\gamma\). Suppose \(A_\gamma = \delta\). By the isoperimetric inequality for the plane, \(L_\gamma \ge \sqrt{2\pi\delta}\), and \[f(\gamma) \ge 2\pi\delta + (C-\delta)^2,\] and this is sharp if \(\gamma\) is a circle. We minimize over \(\delta\) to find that if \(\gamma\) is a geodesic, then

  • if \(0\le C\le \pi\), then \(\gamma\) is a vertical line segment
  • if \(C > \pi\), then \(\pi \circ \gamma\) is a circle of area \(C - \pi\).

Scaling limits and sub-Riemannian geometry

We're interested in large-scale geometry, so what does this metric look like on the large scale? We construct the scaling limit or the asymptotic cone by letting \(d_g\) be the Riemannian distance function and considering the spaces \((\HH, R^{-1} d_g)\) – the unit ball in \((\HH, R^{-1} d_g)\) corresponds to the ball of radius \(R\) in \((\HH, d_g)\).

We first define a sequence of scalings of \(\HH\). For \(t>0\), let \(s_t \from \HH \to \HH\), \((x,y,z)\mapsto (tx,ty,t^2z)\). This is a Lie group isomorphism, and from our calculations with the discrete group, we expect that \[R^{-1} d_g(0,s_{R}(x,y,z)) = R^{-1} d_g(0,(R x,R y,R^2 z)) \approx R^{-1} \max\{|R x|, |R y|, \sqrt{|R^2 z|}\}\] stays bounded as \(R^{-1} \to 0\).

We can calculate the limit by considering the pullback

\begin{align*} dg_R^2 & = R^{-2} s_{R}^*(dg^2)\\ & = R^{-2}\left[(R dx)^2 + (R dy)^2 + \left(R^2 dz + \frac{R y}{2} R dx - \frac{R x}{2} R dy\right)^2\right] \\ & = dx^2 + dy^2 + R^2 \left(dz + \frac{y}{2} dx - \frac{x}{2} dy\right)^2. \end{align*}

This is the same left-invariant metric we saw before – except now the \(Z\) field has length \(R\). As \(R\to \infty\), the length of \(Z\) goes to \(\infty\).

The only vectors whose length stays bounded are the vectors in \[V_1 = \ker(dz + \frac{y}{2} dx - \frac{x}{2} dy) = \langle X, Y\rangle.\] We call these horizontal vectors, and we call curves tangent to \(V_1\) horizontal curves.

Lecture 4: Scaling limits

Carnot groups

  • Definition
  • Examples and non-examples
  • Scaling automorphisms
  • The scale-invariant sub-riemannian metric \(d_\infty\)

The ball-box theorem

The fact that the scaling automorphisms of a Carnot group scale the subriemannian metric leads to an estimate of the subriemannian distance.

Let \(\mathfrak{g}\) be Carnot and let \(X_1,\dots, X_m \in \mathfrak{g}\) be a basis such that there are \(j_i\) satisfying \[V_i = \langle X_{j_i}, \dots, X_{j_{i+1}-1} \rangle.\] For each \(i\), let \(d_i\) be such that \(V_i\in X_{d_i}\). We parameterize \(G\) by setting \((a_1,\dots, a_m) = \exp(\sum a_i X_i)\). Then the following theorem holds:

For \(v=(v_1,\dots, v_m)\), let \(\rho(v) = \max |v_i|^{1/d_i}\). Then \(d_\infty(0,v)\approx \rho(v)\).

Consequently, \(\vol B(0,R;d_\infty) \approx R^Q\), where \(Q = \sum d_i = \sum i \dim(V_i)\).

The proof follows by scaling; there is an \(\epsilon>0\) such that if \(\rho(v)\le \epsilon\), then \(v\in B(0,1;d_\infty)\) and if \(v\in B(0,1;d_\infty)\), then \(\rho(v)<\epsilon^{-1}\). Then for any \(v\in G\), we have \(s_{\epsilon \rho(v)^{-1}}(v)\in B(0,1;d_\infty)\) and thus \(v\in B(0, \epsilon^{-1}\rho(v))\); likewise, \(v\not \in B(0, \epsilon\rho(v))\).

Asymptotic cones and the associated graded algebra

While Carnot groups have scaling automorphisms and can be given a scale-invariant metric, an arbitrary nilpotent Lie group might not have scalings. Nonetheless, we can describe the geometry using the lower central series and the associated graded algebra.

Let \(G\) be a nilpotent Lie group of step \(k\) and let \(\mathfrak{g}_i\) be the lower central series of \(\mathfrak{g}\). An adapted basis of \(\mathfrak{g}\) is a basis \(X_1,\dots, X_m\) such that for each \(i\), there is a \(j_i\) satisfying \(\mathfrak{g}_i = \langle X_{j_i},\dots, X_m\rangle\). Let \((a_1,\dots,a_m) = \exp(\sum a_i X_i)\). We can use this basis to state the ball-box theorem for \(G\).

Let \(G\) be a step-\(k\) nilpotent group and let \(X_1,\dots, X_m\) be an adapted basis. Equip \(G\) with coordinates \((a_1,\dots,a_m) = \exp(\sum a_i X_i)\) and a left-invariant Riemannian metric. Then \[d_G(0, (a_1,\dots,a_m)) \approx \sum |a_i|^{1/d_i}.\]

(We omit the proof; for full details, see one of the sources in the further reading section of the website.)

We can describe the geometry in finer detail using the associated graded algebra \(\mathfrak{g}_\infty\).

Let \(V_i=\mathfrak{g}_i/\mathfrak{g}_{i+1}\). We note that \(V_i\) is abelian and that, by the Jacobi identity, \([\mathfrak{g}_i, \mathfrak{g}_j] \subset \mathfrak{g}_{i+j}\).

Let \[\mathfrak{g}_\infty = V_1\oplus\dots\oplus V_k.\] For any \(v_i \in \mathfrak{g}_{i}\) and \(v_j\in \mathfrak{g}_{j}\), \[[v_i + \mathfrak{g}_{i+1}, v_j + \mathfrak{g}_{j+1}]\subset [v_i,v_j] + \mathfrak{g}_{i+j+1},\] so the bracket on \(\mathfrak{g}\) induces a well-defined bracket on \(\mathfrak{g}_\infty\), which makes \(\mathfrak{g}_\infty\) Carnot – we call \(\mathfrak{g}_\infty\) the associated graded algebra of \(\mathfrak{g}\).

Pansu proved the following theorem:

Let \((G,d)\) be a nilpotent Lie group equipped with a left-invariant Riemannian metric. Equip \(G\) with an adapted basis \(X_1,\dots, X_m\) and a left-invariant Riemannian metric such that the \(X_i\)'s are orthonormal.

Then \((G,R^{-1}d)\) converges to a limit \((G_\infty, d_\infty)\) where \(G_\infty\) is a Carnot group whose Lie algebra is the associated graded algebra of \(\mathfrak{g}\) and \(d_\infty\) is a left-invariant subriemannian metric on \(G_\infty\).

This convergence is easiest to see from an algebraic standpoint. Since \(X_1,\dots, X_m\) is an adapted basis, the subset \(X_{j_i},\dots, X_{j_{i+1}-1}\) is a basis for \(V_i=\mathfrak{g}_i/\mathfrak{g}_{i+1}\). This lets us identify (the vector spaces) \(\mathfrak{g}\) and \(\mathfrak{g}_\infty\) with (the vector space) \(\R^m\). For \(v,w\in \R^m\), we write \([v,w]\) and \([v,w]_\infty\) for the brackets on \(\R^m\) coming from the identification with \(\mathfrak{g}\) and \(\mathfrak{g}_\infty\) respectively.

Let \(d_i\) as above, so that \(X_i\in \mathfrak{g}_{d_i}\setminus \mathfrak{g}_{d_i+1}\). We can describe the bracket on \(\mathfrak{g}\) in terms of the structure coefficients \(c_{ij}^k\), i.e., the real numbers such that \[[X_i,X_j] = \sum_k c_{ij}^k X_k.\] Note that \([X_i,X_j]\in \mathfrak{g}_{d_i+d_j}\) and that \(\mathfrak{g}_{d_i+d_j}\) is generated by the \(X_k\)'s such that \(d_k \ge d_i+d_j\). Therefore, \(c_{ij}^k = 0\) unless \(d_k \ge d_i+d_j\), and \[[X_i,X_j] = \sum_{d_k\ge d_i+d_j} c_{ij}^k X_k.\]

We can describe the bracket on \(\mathfrak{g}_\infty\) similarly. Recall that \[[X_i,X_j]_\infty = [X_i,X_j]+\mathfrak{g}_{d_i+d_j+1}\in V_{d_i+d_j}.\] Since \(X_k\in \mathfrak{g}_{d_k}\), the formula above becomes \[[X_i,X_j]_\infty = \sum_{d_k=d_i+d_j} c_{ij}^k X_k.\] That is, \([,]_\infty\) ignores the parts of \([,]\) that lie in the wrong \(V_k\).

We claim that rescalings of \(\mathfrak{g}\) converge to \(\mathfrak{g}_\infty\) in the following sense:

Let \(s_t(a_1,\dots, a_m) = (t^{d_1} a_1,\dots, t^{d_m} a_m)\). For \(R>0\), let \[[v,w]_R = s_R^{-1}([s_R(v),s_R(w)]);\] note that \((\R^m,[,]_R)\cong \mathfrak{g}\). Then for any \(v,w\in \R^m\), \[\lim_{R\to \infty} s_R^{-1}([s_R(v),s_R(w)]) = [v,w]_\infty.\]

It suffices to consider \([X_i,X_j]_R\). By the formulas above, \[s_R^{-1}([s_R(X_i),s_R(X_j)])= s_R^{-1}(\sum_{d_k\ge d_i+d_j} c_{ij}^k R^{d_i + d_j} X_k) = \sum_{d_k\ge d_i+d_j} c_{ij}^k R^{d_i + d_j-d_k} X_k.\] If \(d_k > d_i+d_j\), then \(R^{d_i + d_j-d_k}\to 0\), so \[\lim_{R\to \infty} [X_i,X_j]_R = \sum_{d_k = d_i+d_j} c_{ij}^k X_k = [X_i,X_j]_\infty.\] This proves the lemma.

We write \(G\) and \(G_\infty\) as two different group operations on \(\R^m\): for \(v,w\in \R^m\), let \[v\cdot w = v + w + \frac{1}{2}[v,w]+\dots\] \[v\ast w = v + w + \frac{1}{2}[v,w]_\infty+\dots\] By the Baker-Campbell-Hausdorff formula, the lemma implies \[\lim_{R\to \infty} s_R^{-1}(s_R(v)s_R(w)) = v\ast w.\] That is, the group operation on \(G\) can be approximated by the group operation on \(G_\infty\) on large balls.

Lecture 5: Maps of subRiemannian manifolds

Limits of nilpotent groups and subRiemannian manifolds

  • If \(G\) is nilpotent and \(V\) is a left-invariant subbundle of the tangent bundle such that \(V(0)\) is isomorphic to \(\mathfrak{g}/[\mathfrak{g}, \mathfrak{g}]\), then there's a scaling map \(s_t\) so that rescalings \((s_R^{-1})_*(V)\) converges to the horizontal bundle of \(G_\infty\).
  • Thus, if \(G\) is given a Riemannian metric and \(\gamma_R\from [0,1]\to G\) is constant-speed with \(\ell(\gamma_R)=R\), then \(s_{R^{-1}}(\gamma_R)\) is close to a horizontal curve with subunit speed in \(G_\infty\).
  • Therefore, if \(\gamma\) is a geodesic from \(s_R(v)\) to \(s_R(w)\), then there's a horizontal curve of length \(\ell(\gamma)/R\) that connects \(v\) to \(w'\), where \(w'\) has coordinates close to \(w\).
  • By the Ball-Box theorem, \(w'\) is metrically close to \(w\). Taking these arguments further, we get \[\lim_{R\to \infty} R^{-1}d_G(s_R(v),s_R(w)) = d_{G_\infty}(v,w).\]
  • Similar results apply to equiregular subRiemannian manifolds, in which case, by a theorem of Mitchell, scalings of the balls \(B_\epsilon(p)\) converge to a Carnot group as \(\epsilon\to 0\). In order to prove this, one proves that there are privileged coordinates for which the ball-box theorem holds. Using these coordinates, one can define a scaling so that scalings of the horizontal distribution approach a limit. Finally, one proves that this limit is the horizontal distribution of a Carnot group.

Maps between manifolds

The scaling limits of nilpotent groups are crucial to understanding quasi-isometries between nilpotent groups. If \(G\) and \(H\) are nilpotent Lie groups with Riemannian metrics and \(f\from G\to H\) is a quasi-isometry, then scalings of \(f\) converge to a bilipschitz map \(f_\infty\from G_\infty \to H_\infty\).

We can then apply the following theorem of Pansu:

Let \(G, H\) be Carnot groups (graded groups equipped with a left-invariant subriemannian metric whose horizontal distribution is \(V_1\)). Let \(f\from G \to H\) be a Lipschitz map. Then \(f\) is Pansu differentiable almost everywhere. That is, for almost every \(p\in G\), there is a homeomorphism \(\alpha\from G\to H\) such that \[\lim_{h\to 0} \frac{d(f(ph),f(p)\alpha(h))}{|h|}=0.\]

Before proving the theorem, we first prove a special case.

Let \(f\from \R\to G\) be Lipschitz. Then \(f\) is Pansu differentiable almost everywhere.

First, by the Ball-Box theorem, \(f'(t)\) is horizontal almost everywhere. In general, this is weaker than Pansu differentiability; the curve \(t\mapsto (t,0,t^2)\) in the Heisenberg group has horizontal velocity vector at \(0\), but it is not Pansu differentiable.

We use this condition by defining horizontal lifts. Let \(G\) be a Carnot group, let \(V=G/[G,G]\), and let \(\pi\from G\to V\) be the quotient map. If \(\gamma\from \R\to V\) is Lipschitz and \(\tilde{\gamma}\from \R\to G\) is a Lipschitz curve such that \(\tilde{\gamma}'\) is horizontal almost everywhere and \(\pi\circ \tilde{\gamma}=\gamma\), then we say that \(\tilde{\gamma}\) is a horizontal lift of \(\gamma\). The uniqueness of solutions to ODEs implies that for any Lipschitz curve \(\gamma\from \R\to V\) and any point \(p\in \pi^{-1}(\gamma(0))\), there is a unique horizontal lift \(\tilde{\gamma}\) such that \(\tilde{\gamma}(0)=p\). In particular, \(f\) is a horizontal lift of \(\pi\circ f\).

Suppose that \(f\) is differentiable at \(t\). Without loss of generality, we may suppose that \(t=0\) and \(f(t)=0\). Let \(\alpha\from \R\to V\), \(\alpha(t) = (\pi\circ f)'(0) t\). Note that the horizontal lift \(\tilde{\alpha}\) is a homomorphism from \(\R\) to \(G\).

Let \(f_\epsilon\) be the rescaling \(f_\epsilon(t) = s_{\epsilon^{-1}}(f(\epsilon t))\). Then \(\pi\circ f_\epsilon\) converges to \(\alpha\) as \(\epsilon\to 0\). With some work, one can show that the lifts \(\widetilde{\pi\circ f_\epsilon}\) converge to \(\tilde{\alpha}\), but \(\widetilde{\pi\circ f_\epsilon} = f_\epsilon\), so \(f_\epsilon\) converges to a homomorphism as \(\epsilon \to 0\), which implies that \(f\) is Pansu differentiable at \(t\).

Lecture 6-7: Quasi-isometries of nilpotent groups

One of the main questions about the geometric group theory of nilpotent groups is the following: Suppose \(G\) and \(H\) are quasi-isometric nilpotent Lie groups. Are \(G\) and \(H\) isomorphic?

Pansu's theorem implies that if \(G\) and \(H\) are nilpotent Lie groups and \(f\) is a quasi-isometry from \(G\) to \(H\), then the associated graded groups \(G_\infty\) and \(H_\infty\) are isomorphic. So far, however, it has proven difficult to distinguish pairs of groups with the same associated graded algebra.

Results:

  • (Shalom) If \(\Gamma\subset G\) and \(\Lambda \subset H\) are quasi-isometric lattices, then \(\dim H^i(\Gamma;\R)=\dim H^i(\Lambda;\R)\) for all \(i\).
  • (Sauer) If \(G\) and \(H\) are quasi-isometric, then \(H^*(\Gamma;\R)\) and \(H^*(\Lambda;\R)\) are isomorphic rings.
  • (Llosa Isenrich–Pallier–Tessera) The Dehn function can be used to distinguish certain nilpotent groups with isomorphic associated graded algebras.

We will sketch the proof of Shalom's theorem. We'll need some background on group cohomology and amenability.

Group cohomology

Recall some facts from algebraic topology:

  • Let \(\Gamma\) be a discrete group. An Eilenberg–MacLane space of type \(K(\Gamma,1)\) is a CW complex \(X\) with \(\pi_1(X)=\Gamma\) and \(\pi_n(X)=0\) for \(n > 1\). Equivalently, a \(K(\Gamma,1)\) is a CW complex with \(\pi_1(X)=\Gamma\) such that the universal cover is contractible. Any two \(K(\Gamma,1)\)'s are homotopy equivalent.
  • Given a group \(\Gamma\), we define \(H^n(\Gamma;\R) = H^n(K(\Gamma,1);\R)\) for any \(K(\Gamma,1)\). Since any two \(K(\Gamma,1)\)'s are homotopy equivalent, this doesn't depend on the choice of space.

When \(\Gamma\) is a lattice in a nilpotent group, we can describe the cohomology of \(\Gamma\) in terms of the Lie algebra.

If \(G\) is a nilpotent Lie group and \(\Gamma\) is a lattice in \(G\), then \(H^*(\Gamma;\R)\cong H^*(\mathfrak{g})\).

Here, \(H^*(\mathfrak{g})\) is the cohomology of the complex \(C^*(\mathfrak{g})\) of left-invariant differential forms on \(G\). Any such form is determined by its value at \(0\), so we can write this complex as: \[\R \to \mathfrak{g}^* \to \wedge^2 \mathfrak{g}^* \to \dots \to \wedge^n \mathfrak{g}^*.\] If \(\omega\in \wedge^d \mathfrak{g}^*\) and \(\Sigma\) is an \(d\)–dimensional surface in \(G\), we define \[\omega(\Sigma) = \int_{\Sigma} \omega.\]

The differential on \(C^*(\mathfrak{g})\) can be written in terms of the Lie bracket. For \(\omega\in \mathfrak{g}^*\), \[d\omega(v,w) = \omega([v,w]).\]

Example: The Heisenberg groups

Let \[\mathfrak{h}_n = \langle X_1,\dots, X_n, Y_1,\dots, Y_n, Z \mid [X_i, Y_i] = Z, \text{ all others commute}\rangle.\] Then

\begin{align*} dX_i^* &= 0 & dY_i^* &= 0 & dZ^* &= \sum_i X_i^*\wedge Y_i^*. \end{align*}

Let \(\omega=\sum_i X_i^*\wedge Y_i^*\). One can calculate \(H^1(\mathfrak{h}_n) = \langle X_1,\dots, X_n, Y_1,\dots, Y_n\rangle\). To calculate \(H^2(\mathfrak{h}_n)\), we note that \[d(C^1(\mathfrak{h}_n)) = \langle \omega \rangle.\] If \(\mathfrak{g}=V_1\oplus V_k\) is Carnot, then \(d(\wedge^j V_1^*) = 0\), so any horizontal form is a coboundary, i.e., \[\langle X_i\wedge X_j, X_i\wedge Y_j,Y_i\wedge Y_j\rangle/\omega \subset H^2(\mathfrak{h}_n);\] this subspace has dimension \(\binom{2n}{2} - 1\).

The only \(2\)–forms with nonzero differential are of the form \(X_i^*\wedge Z^*\) or \(Y_i^*\wedge Z^*\), and we have \[d(X_j^* \wedge Z^*) = - X_j^* \wedge \omega = - \sum_{i\ne j} X_j \wedge X_i \wedge Y_i.\] \[d(Y_j^* \wedge Z^*) = - Y_j^* \wedge \omega = - \sum_{i\ne j} Y_j \wedge X_i \wedge Y_i.\]

If \(n>1\), then these vectors are linearly independent, so \(d\) is injective on \(Z^*\wedge V_1^*\) and \[\dim H^2(\mathfrak{h}_n) = \binom{2n}{2} - 1.\]

If \(n=1\), both of these are zero, so \(X^*\wedge Z^*\) and \(Y^*\wedge Z^*\) are cocycles and \[\dim H^2(\mathfrak{h}_1) = \binom{2}{2} - 1 + 2 = 2.\]

Last year, we saw how these cohomology classes affect the Dehn function of the Heisenberg groups (the amount of area that can be enclosed by a curve of length \(L\)). See the course notes for Lecture 6 for details.

Amenability

A locally compact group \(G\) is amenable if there is a finitely-additive left-invariant probability measure on \(G\). That is, there is a function \(\mu\from 2^G\to [0,1]\) such that:

  • \(\mu(G)=1\)
  • for \(g\in G\), \(\mu(gS)=\mu(S)\)
  • if \(S,T\subset G\) and \(S\cap T=\emptyset\), then \(\mu(S\cup T) = \mu(S) + \mu(T)\)

Roughly, \(\mu\) is a way to measure how big a subset of \(G\) is. For example, when \(G=\Z\) and \(a>0\), this implies \(\mu(a\Z) = \frac{1}{a}\), because \(\mu(a\Z)=\mu(a\Z+b)\) for any \(b\) by left-invariance, and \[\mu(\Z)=\mu(a\Z) + \mu(a\Z+1) + \dots + \mu(a\Z+a-1).\] Equivalently, an invariant mean can be viewed as a left-invariant functional \(\mu\from L_\infty(G)\to \R\) such that:

  • if \(f\ge 0\), then \(\mu(f)\ge 0\)
  • \(\mu\) has norm \(1\).

It's not obvious how to construct a function like this, and in fact, the construction uses the axiom of choice, in the form of the Hahn–Banach theorem. We first construct an ultralimit. Let \(V\subset \ell_\infty(\N)\) be the set of convergent subsequences. Then \(\lim\from V\to \R\) is a linear map with norm \(1\). By Hahn–Banach, it has a linear extension \(\underline{\lim}\from \ell_\infty(\N)\to \R\), also with norm \(1\); we think of \(\underline{\lim}\) as a way to assign a limit to any bounded sequence.

Next, we define a Følner sequence for a group. This is a sequence of subsets of the group with small boundary. Let \(G\) be a group and let \(K\) be a finite generating set. For \(S\subset G\), let \(\partial S = \{g\in S\mid gK\not\subset S\}\). A sequence of nonempty finite sets \(S_1,\dots\) is a Følner sequence if \(\frac{\# \partial S_i}{\#S_i}\to 0\). If a group has polynomial growth, then the metric balls \(B_n\) form a Følner sequence, but many groups with nonpolynomial growth also have Følner sequences, including all solvable groups.

We can use a Følner sequence for \(\Gamma\) to construct an invariant mean. Let \(S_1,\dots\subset \Gamma\) be a Følner sequence. For \(T\subset \Gamma\), let \[\mu(T) = \underline{\lim} \frac{|T\cap S_n^{-1}|}{|S_n|}.\] Since \(\frac{|T\cap S_n^{-1}|}{|S_n|}\in [0,1]\), this limit is between \(0\) and \(1\). It satisfies \(\mu(\Gamma)=1\) and, by the linearity of \(\underline{\lim}\), \(\mu\) is finitely additive.

To show left-invariance, let \(k\) be a generator of \(\Gamma\). Then \[|kT\cap S_n^{-1}| = |T\cap k^{-1} S_n^{-1}| = |T\cap (k S_n)^{-1}|;\] it follows that \[\big||kT\cap S_n^{-1}| - |T\cap S_n^{-1}|\big| \le |k S_n\setminus S_n| \le |\partial S_n|.\] Therefore, \[\underline{\lim} \frac{|T\cap S_n^{-1}|}{|S_n|} - \frac{|kT\cap S_n^{-1}|}{|S_n|} = 0\] and \(\mu(T) = \mu(kT)\). Since this holds for every generator of \(\Gamma\), we have \(\mu(gT)=\mu(T)\) for all \(g\in \Gamma\).

The map induced by a quasi-isometry

Let \(\Lambda\) and \(\Gamma\) be torsion-free and let \(f\from \Lambda \to \Gamma\) be a quasi-isometry. Let \(\Delta(\Lambda)\) be the complete simplex with vertex set \(\Lambda\). Then we can write \(H^*(\Lambda;\R)\) as the cohomology of the complex of invariant cocycles, \[C^n(\Lambda\backslash \Delta(\Lambda);\R) = \{\omega\from \Lambda^{n+1}\to \R \mid \omega \text{ is alternating and left-invariant}\}.\]

For \(\omega\in C^n(\Lambda\backslash \Delta(\Lambda);\R)\) and \(D>0\), let \(f^*\omega\in C^n(\Gamma;\R)\), \[f^*\omega(\gamma_0,\dots,\gamma_n) = \omega(f(\gamma_0),\dots, f(\gamma_n)).\] Note that \(f^*\omega\) is generally not invariant, but we can average over translates to construct an invariant cocycle. \[M_\omega(D) = \max_{\diam\{\lambda_0,\dots, \lambda_n\}\le D} |\omega(\lambda_0,\dots, \lambda_n)|;\] since \(\omega\) is left-invariant, this maximum is finite. Since \(f\) is a quasi-isometry, there is a \(C>0\) such that for any \(\omega\in C^n(\Lambda\backslash \Delta(\Lambda);\R)\) and any \(\gamma_0, \dots, \gamma_n, H\in \Gamma\), \[|f^*\omega(h\gamma_0,\dots,g\gamma_n)| = |\omega(f(h \gamma_0), \dots, f(h\gamma_n))| \le M_\omega(C\diam \{\lambda_0, \dots, \lambda_n\}+C).\] In particular, if \(\mu\) is an invariant mean on \(\Lambda\), we can define an invariant cocycle \[\overline{f^*}\omega(\gamma_0,\dots, \gamma_n) = \int_\Gamma f^*\omega(h\gamma_0,\dots,h\gamma_n)\ud \mu.\] One can check that \(\overline{f^*}(d\omega) = d(\overline{f^*}\omega)\), so this induces a map \(\overline{f^*}\from H^n(\Gamma;\R)\to H^n(\Lambda;\R)\) – what can we say about this map?

Author: Robert Young

Created: 2023-03-07 Tue 18:09