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

Haar Measures

Stephan Tornier
Abstract.

This article provides a concise introduction to the theory of Haar measures on locally compact Hausdorff groups. We cover the necessary preliminaries on topological groups and measure theory, the Haar correspondence, unimodularity and Haar measures on coset spaces.

1. Preliminaries

References appear throughout the article. Apart from the classics by Haar [Haa33], Weil [Wei65] and Bourbaki [Bou04], the neat introduction by Knightly-Li [KL06, Sec. 7] deserves highlighting.

1.1. Locally Compact Hausdorff Groups

The natural class of groups for which to consider Haar measures is that of locally compact Hausdorff groups which we review presently.

A topological group is a group GG with a topology such that the multiplication map G×GGG\times G\to G and the inversion map GGG\to G are continuous. As a consequence, left and right multiplication by elements of GG as well as inversion are homeomorphisms of GG. Therefore, the neighbourhood system of the identity eGe\in G determines the topology on GG. A topological space is locally compact if every point has a compact neighbourhood; and it is Hausdorff if any two distinct points have disjoint neighbourhoods. In the Hausdorff case, local compactness is equivalent to every point admitting a relatively compact open neighbourhood, i.e. an open neighbourhood with compact closure.

The class of locally compact Hausdorff groups is stable under taking closed subgroups as the following proposition shows. Recall that if XX is a topological space and AA is a subset of XX, we may equip AA with the subspace topology, for which UAU\subseteq A is open if and only if there is an open set VXV\subseteq X, such that U=AVU=A\cap V.

Proposition 1.1.

Let XX be a locally compact Hausdorff space and let AA be a closed subset. Then AA is locally compact Hausdorff.

Proof.

Recalling that compact subsets of Hausdorff spaces are closed and that closed subsets of compact sets are compact, this is immediate following the definitions. ∎

As to coset spaces, we record the following lemma on a property of neighbourhoods that comes with the group structure.

Lemma 1.2.

Let GG be a topological group. Then for every xGx\in G and every neighbourhood UU of eGe\in G, there exists an open neighbourhood VV of xx such that V1VUV^{-1}V\subseteq U.

Proof.

The map φ:G×GG,(g,h)g1h\varphi:G\times G\to G,\ (g,h)\mapsto g^{-1}h is continuous. Hence there are open sets V1,V2GV_{1},V_{2}\subseteq G such that V11V2=φ(V1×V2)UV_{1}^{-1}V_{2}=\varphi(V_{1}\times V_{2})\subseteq U. Then V=V1V2V=V_{1}\cap V_{2} serves. ∎

When GG is a topological group and HGH\leq G is a subgroup of GG, we equip the set of cosets G/HG/H with the quotient topology, i.e. UG/HU\subseteq G/H is open if and only if π1(U)G\pi^{-1}(U)\subseteq G is open, where π:GG/H,ggH\pi:G\to G/H,\ g\mapsto gH. Then π\pi is continuous and open, and left multiplication by gGg\in G is a homeomorphism of G/HG/H.

Proposition 1.3.

Let GG be a topological group and let HGH\leq G be closed. Then G/HG/H is Hausdorff.

Proof.

Let xH,yHG/HxH,yH\in G/H be distinct. Then yHx1GyHx^{-1}\subseteq G is closed and does not contain eGe\in G. Hence, by Lemma 1.2, there is an open neighbourhood VGV\subseteq G of eGe\in G with V1VG\yHx1V^{-1}V\subseteq G\backslash yHx^{-1}. Then VxHVxH and VyHVyH are disjoint neighbourhoods of xHG/HxH\in G/H and yHG/HyH\in G/H respectively. ∎

Proposition 1.4.

Let GG be a locally compact group and let HGH\leq G. Then G/HG/H is locally compact.

Proof.

It suffices to show that HG/HH\in G/H has a compact neighbourhood. Since GG is locally compact, there is a compact neighbourhood UU of eGe\in G. Let VV be as in Lemma 1.2. Then π(V)\pi(V) is an open neighbourhood of HG/HH\in G/H since π\pi is open. We show that π(V)¯\smash{\overline{\pi(V)}} is compact. If gHπ(V)¯\smash{gH\in\overline{\pi(V)}} then VgHVHVgH\cap VH\neq\emptyset and hence gH=v11v2HgH=v_{1}^{-1}v_{2}H for some v1,v2Vv_{1},v_{2}\in V. Thus π(V)¯π(U)\smash{\overline{\pi(V)}}\subseteq\pi(U). The latter set is compact since π\pi is continuous and hence so is π(V)¯π(U)\smash{\overline{\pi(V)}}\subseteq\pi(U). ∎

We now state a version of Urysohn’s Lemma which guarantees the existence of certain compactly supported functions on locally compact Hausdorff spaces. Recall that when XX is a topological space, Cc(X)C_{c}(X) denotes the set of continuous, complex-valued functions ff on XX with compact support supp(f):={xXf(x)0}¯\smash{\mathrm{supp}(f):=\overline{\{x\in X\mid f(x)\neq 0\}}}. When fCc(X)f\in C_{c}(X) is such that 0f(x)10\leq f(x)\leq 1 for all xXx\in X, UXU\subseteq X is open and KXK\subseteq X is compact, write fUf\prec U if supp(f)U\operatorname{supp}(f)\subseteq U and KfK\prec f if f(k)=1f(k)=1 for all kKk\in K.

Lemma 1.5.

Let XX be a locally compact Hausdorff space. When KXK\subseteq X is compact and UXU\subseteq X is open with KUK\subseteq U, there exists an open set VXV\subseteq X with compact closure such that KVV¯XK\subseteq V\subseteq\overline{V}\subseteq X.

Proof.

By compactness of KK and local compactness of XX, there is a relatively compact open set WW containing KK. Using once more that KK is compact, and that XX is Hausdorff, there is for every pUcp\in U^{c} an open set VpV_{p} containing KK such that pV¯pp\not\in\overline{V}_{p}. Then (UcW¯V¯p)pC\smash{(U^{c}\cap\overline{W}\cap\overline{V}_{p})_{p\in C}} is a family of compact sets with empty intersection. Hence there are p1,,pnUcp_{1},\ldots,p_{n}\in U^{c} such that i=1nUcW¯V¯pi\smash{\bigcap_{i=1}^{n}U^{c}\cap\overline{W}\cap\overline{V}_{p_{i}}} is empty as well. Set V:=Wi=1nVpV:=W\cap\bigcap_{i=1}^{n}V_{p}. ∎

Lemma 1.6 (Urysohn).

Let XX be a locally compact Hausdorff space. When KXK\subseteq X is compact and UXU\subseteq X is open such that KUK\subseteq U, then there exists fCc(G)f\in C_{c}(G) satisfying KfUK\prec f\prec U.

Proof.

Let r:0[0,1]r:\operatorname{\mathbb{N}}_{0}\to\operatorname{\mathbb{Q}}\cap[0,1] be a bijection with r(0)=0r(0)=0 and r(1)=1r(1)=1. Using Lemma 1.5, pick open sets Ur(1)U_{r(1)} and Ur(0)U_{r(0)} with compact closure such that KUr(1)U¯r(1)Ur(0)U¯r(0)U\smash{K\subseteq U_{r(1)}\subseteq\overline{U}_{r(1)}\subseteq U_{r(0)}\subseteq\overline{U}_{r(0)}\subseteq U}. Then, by induction on n0n\in\operatorname{\mathbb{N}}_{0} and using Lemma 1.5, construct open sets Ur(n)U_{r(n)} with compact closure such that for all s,t[0,1]s,t\in\operatorname{\mathbb{Q}}\cap[0,1] with s>r(n)>ts>r(n)>t we have V¯sVr(n)V¯r(n)Vt\smash{\overline{V}_{s}\subseteq V_{r(n)}\subseteq\overline{V}_{r(n)}\subseteq V_{t}}. Given α[0,1]\alpha\in[0,1], set Uα:=t[α,1]UtU_{\alpha}:=\bigcup_{t\in\operatorname{\mathbb{Q}}\cap[\alpha,1]}U_{t} and define

 f:X,x{1xU1inf{α[0,1]xUα}xU1.\hbox to0.0pt{\hfil}f:X\to\operatorname{\mathbb{R}},\ x\mapsto\begin{cases}1&x\in U_{1}\\ \inf\{\alpha\in[0,1]\mid x\in U_{\alpha}\}&x\not\in U_{1}\end{cases}.

For continuity, let xXx\in X and 0<δ<ε0<\delta<\varepsilon. Then xUf(x)εδ\U¯f(x)+εf1((f(x)ε,f(x)+ε))x\in U_{f(x)-\varepsilon-\delta}\backslash\overline{U}_{f(x)+\varepsilon}\subseteq f^{-1}((f(x)-\varepsilon,f(x)+\varepsilon)), where Uα:=XU_{\alpha}:=X for α<0\alpha<0 and Uα:=U_{\alpha}:=\emptyset for α>1\alpha>1. Overall, fCc(G)f\in C_{c}(G) and KfUK\prec f\prec U. ∎

We also need the notion of uniform continuity for functions on topological groups (which comes from giving the group the structure of a uniform space). Let GG be a topological group. A function f:Gf:G\to\operatorname{\mathbb{C}} is uniformly continuous on the left (right) if for all ε>0\varepsilon>0 there is an open neighbourhood UU of eGe\in G such that for all xGx\in G and gUg\in U we have |f(gx)f(x)|<ε|f(gx)-f(x)|<\varepsilon (|f(xg)f(x)|<ε)(|f(xg)-f(x)|<\varepsilon).

Proposition 1.7.

Let GG be a locally compact Hausdorff group. Then any fCc(G)f\in C_{c}(G) is uniformly continuous on the left and right.

Proof.

We prove that ff is uniformly continuous on the left. Uniform continuity on the right can be handled analogously. Let ε>0\varepsilon>0. By continuity of ff, there is for each xsuppfx\in\operatorname{supp}f an open neighbourhood UxU_{x} of eGe\in G such that |f(gx)f(x)|<ε/2|f(gx)-f(x)|<\varepsilon/2 for all gUxg\in U_{x}. For every UxU_{x} (xG)(x\in G), pick a symmetric open neighbourhood VxV_{x} of eGe\in G such that Vx2UxV_{x}^{2}\subseteq U_{x} using Lemma 1.2. Since suppf\operatorname{supp}f is compact, finitely many of the sets VxxV_{x}x (xsuppf)(x\in\operatorname{supp}f) cover suppf\operatorname{supp}f, say (Vxkxk)k=1n(V_{x_{k}}x_{k})_{k=1}^{n}. Define V=k=1nVkV=\bigcap_{k=1}^{n}V_{k}. Then for all xsuppfx\in\operatorname{supp}f and for all gVg\in V we have

|f(gx)f(x)||f(gx)f(xk)|+|f(xk)f(x)|<ε2+ε2=ε|f(gx)-f(x)|\leq|f(gx)-f(x_{k})|+|f(x_{k})-f(x)|<\frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon

where k{1,,n}k\in\{1,\ldots,n\} is chosen such that xVxkxkx\in V_{x_{k}}x_{k}. If xsuppfx\notin\operatorname{supp}f then for every gVg\in V either gxsuppfgx\notin\operatorname{supp}f, in which case the above inequality is trivial, or gxsuppfgx\in\operatorname{supp}f and we set y:=gxy:=gx. Then |f(gx)f(x)|=|f(g1y)f(y)||f(gx)-f(x)|=|f(g^{-1}y)-f(y)| where ysuppfy\in\operatorname{supp}f and g1Vg^{-1}\in V; we may then argue as before. ∎

Finally, the following facts are useful in various places.

Proposition 1.8.

Let GG be a topological group and A,BGA,B\subseteq G. If AA and BB are compact, then ABAB is compact. If either AA or BB is open, then ABAB is open.

Proof.

If AA and BB are compact, then so is ABAB as the image of the compact set (A,B)(A,B) under the continuous multiplication map G×GGG\times G\to G. If either AA or BB is open, then ABAB is open as a union of open sets since aAaB=AB=bBAb\bigcup_{a\in A}aB=AB=\bigcup_{b\in B}Ab. ∎

Proposition 1.9.

Let GG be a locally compact Hausdorff group and let HH be a subgroup of GG. Further, let CG/HC\subseteq G/H be compact. Then there exists a compact set KGK\subseteq G such that π(K)C\pi(K)\supseteq C.

Proof.

We may cover GG by relatively compact open sets UiU_{i} (iI)(i\in I). Since π\pi is open and CG/HC\subseteq G/H is compact, finitely many of the π(Ui)\pi(U_{i}) (iI)(i\in I) cover CC, say (π(Uk))k=1n(\pi(U_{k}))_{k=1}^{n}. Then K=k=1nUk¯K=\bigcup_{k=1}^{n}\overline{U_{k}} serves. ∎

1.2. Measure Theory

We now review some basic measure theory in order to give the definition of a Haar measure and some first properties.

Let XX be a non-empty set. A σ\sigma-algebra on XX is a set 𝒫(X)\operatorname{\mathcal{M}}\subseteq\operatorname{\mathcal{P}}(X) of subsets of XX which contains the empty set and is closed under taking both complements and countable unions. A pair (X,)(X,\operatorname{\mathcal{M}}) where XX is a set and \operatorname{\mathcal{M}} is a σ\sigma-algebra on XX is a measurable space; the sets EE\in\operatorname{\mathcal{M}} are measurable. Given two measurable spaces (X,)(X,\operatorname{\mathcal{M}}) and (Y,𝒩)(Y,\operatorname{\mathcal{N}}), a map f:XYf:X\to Y is measurable if f1(F)f^{-1}(F)\in\operatorname{\mathcal{M}} for all F𝒩F\in\operatorname{\mathcal{N}}. For example, let XX and YY be topological spaces equipped with their Borel σ\sigma-algebras (X)\operatorname{\mathcal{B}}(X) and (Y)\operatorname{\mathcal{B}}(Y) respectively, i.e. the σ\sigma-algebra generated by the open sets. Then any continuous map from XX to YY is measurable. We shall always equip topological spaces with their Borel σ\sigma-algebra.

A measure on a measurable space (X,)(X,\operatorname{\mathcal{M}}) is a map μ:0{}\mu:\operatorname{\mathcal{M}}\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\} which satisfies μ()=0\mu(\emptyset)=0 and is countably additive: whenever (En)n(E_{n})_{n\in\operatorname{\mathbb{N}}} is a sequence of pairwise disjoint measurable sets then μ(nEn)=n=1μ(En)\mu(\bigcup_{n\in\operatorname{\mathbb{N}}}E_{n})=\sum_{n=1}^{\infty}\mu(E_{n}). A triple (X,,μ)(X,\operatorname{\mathcal{M}},\mu) where (X,)(X,\operatorname{\mathcal{M}}) is a measurable space and μ\mu is a measure on (X,)(X,\operatorname{\mathcal{M}}) is a measure space. A set of measure zero is a null set and its complement conull.

If (X,,μ)(X,\operatorname{\mathcal{M}},\mu) is a measure space, (Y,𝒩)(Y,\operatorname{\mathcal{N}}) a measurable space and φ:XY\varphi:X\to Y a measurable map, then φμ:𝒩0{},Fμ(φ1(F))\varphi_{\ast}\mu:\operatorname{\mathcal{N}}\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\},\ F\mapsto\mu(\varphi^{-1}(F)) is the push-forward measure on (Y,𝒩)(Y,\operatorname{\mathcal{N}}) under φ\varphi.

The category of measure spaces is designed to allow for the following notion of integration of certain measurable, complex-valued functions on (X,,μ)(X,\operatorname{\mathcal{M}},\mu).

  1. (1)

    When χE\chi_{E} is the characteristic function of a measurable set EE\in\operatorname{\mathcal{M}}, define

    XχE(x)μ(x)=μ(E).\int_{X}\chi_{E}(x)\ \mu(x)=\mu(E).
  2. (2)

    When f=i=1nλiχEif=\sum_{i=1}^{n}\lambda_{i}\chi_{E_{i}} is a positive, real linear combination of characteristic functions of measurable sets, a simple function, define

    Xf(x)μ(x)=i=1nλiXχEi(x)μ(x).\int_{X}f(x)\ \mu(x)=\sum_{i=1}^{n}\lambda_{i}\int_{X}\chi_{E_{i}}(x)\ \mu(x).
  3. (3)

    When f:Xf:X\to\operatorname{\mathbb{R}} is measurable and non-negative, define

    Xf(x)μ(x)=supφXφ(x)μ(x)\int_{X}f(x)\ \mu(x)=\sup_{\varphi}\int_{X}\varphi(x)\ \mu(x)

    where φ\varphi ranges over all real-valued simple functions on XX with 0φf0\leq\varphi\leq f.

  4. (4)

    When f:Xf:X\to\operatorname{\mathbb{R}} is measurable, decompose

    f=f+fwheref±(x)=max(±f(x),0).f=f_{+}-f_{-}\quad\text{where}\quad f_{\pm}(x)=\max(\pm f(x),0).

    When X|f(x)|μ(x)<\int_{X}|f(x)|\ \mu(x)<\infty, define

    Xf(x)μ(x)=Xf+(x)μ(x)Xf(x)μ(x).\int_{X}f(x)\ \mu(x)=\int_{X}f_{+}(x)\ \mu(x)-\int_{X}f_{-}(x)\ \mu(x).
  5. (5)

    When f:Xf:X\to\operatorname{\mathbb{C}} is measurable and integrable, i.e. X|f(x)|μ(x)<\int_{X}|f(x)|\ \mu(x)<\infty, define

    Xf(x)μ(x)=XRe(f(x))μ(x)+iXIm(f(x))μ(x).\int_{X}f(x)\ \mu(x)=\int_{X}\text{Re}(f(x))\ \mu(x)+i\int_{X}\text{Im}(f(x))\ \mu(x).

The vector space of equivalence classes of measurable, integrable complex-valued functions on XX modulo equality on a conull set is denoted by L1(X,μ)L^{1}(X,\mu). Integration constitutes a linear map from L1(X,μ)L^{1}(X,\mu) to \operatorname{\mathbb{C}}. There is the following change of variables formula.

Proposition 1.10 (Change of variables).

Let (X,,μ)(X,\operatorname{\mathcal{M}},\mu) be a measure space, (Y,𝒩)(Y,\operatorname{\mathcal{N}}) a measurable space and φ:XY\varphi:X\to Y a measurable. For every measurable function f:Yf:Y\to\operatorname{\mathbb{C}} and F𝒩F\in\operatorname{\mathcal{N}} we have

Ff(y)φμ(y)=φ1(F)f(φ(x))μ(x).\int_{F}f(y)\ \varphi_{\ast}\mu(y)=\int_{\varphi^{-1}(F)}f(\varphi(x))\ \mu(x).

whenever either of the two expressions is defined.

Next, we recall Fubini’s Theorem which reduces integrating over a product space to integrating over the factors. Let (X,,μ)(X,\operatorname{\mathcal{M}},\mu) and (Y,𝒩,ν)(Y,\operatorname{\mathcal{N}},\nu) be measure spaces. Then so is (X×Y,×𝒩,μ×ν)(X\times Y,\operatorname{\mathcal{M}}\times\operatorname{\mathcal{N}},\mu\times\nu) where (μ×ν)(\mu\times\nu) is defined by (μ×ν)(E,F):=μ(E)ν(F)(\mu\times\nu)(E,F):=\mu(E)\nu(F) for all (E,F)×𝒩(E,F)\in\operatorname{\mathcal{M}}\times\operatorname{\mathcal{N}}. Recall that (X,,μ)(X,\operatorname{\mathcal{M}},\mu) is σ\sigma-finite if XX is a countable union of sets of finite measure.

Theorem 1.11 (Fubini).

Let (X,,μ)(X,\operatorname{\mathcal{M}},\mu) and (Y,𝒩,ν)(Y,\operatorname{\mathcal{N}},\nu) be σ\sigma-finite measure spaces. Further, let f:X×Yf:X\times Y\to\operatorname{\mathbb{C}} be measurable with XY|f(x,y)|ν(y)μ(x)<\int_{X}\int_{Y}|f(x,y)|\ \nu(y)\ \mu(x)<\infty. Then fL1(X×Y,μ×ν)f\in L^{1}(X\times Y,\mu\times\nu) and

XYf(x,y)ν(y)μ(x)=X×Yf(x,y)(μ×ν)(x,y)=YXf(x,y)μ(x)ν(y).\int_{X}\int_{Y}f(x,y)\ \nu(y)\ \mu(x)=\int_{X\times Y}f(x,y)\ (\mu\times\nu)(x,y)=\int_{Y}\int_{X}f(x,y)\ \mu(x)\ \nu(y).

Measures on topological spaces which appear in practice often satisfy the following additional regularity properties.

Definition 1.12 (Radon measure).

A Radon measure on a topological space XX is a measure on (X,(X))(X,\operatorname{\mathcal{B}}(X)) which satisfies the following properties:

  1. (LF)

    If KXK\subseteq X is compact, then μ(K)<\mu(K)<\infty. (locally finite)

  2. (OR)

    If EXE\subseteq X is measurable, then μ(E)=inf{μ(U)UE,U open}\mu(E)=\inf\{\mu(U)\mid U\supseteq E,U\text{ open}\}. (outer regular)

  3. (IR)

    If UXU\subseteq X is open, then μ(U)=sup{μ(K)KU,K compact}\mu(U)=\sup\{\mu(K)\mid K\subseteq U,K\text{ compact}\}. (inner regular)

The importance of Radon measures is also due to the following result of Riesz which is often employed to define a measure on a given space in the first place.

Theorem 1.13 (Riesz).

Let XX be a locally compact Hausdorff space. Further, let λ:Cc(X)\lambda:C_{c}(X)\to\operatorname{\mathbb{C}} be a positive, i.e. λ(f)[0,)\lambda(f)\in[0,\infty) whenever f(x)[0,)f(x)\in[0,\infty) for all xXx\in X, linear functional. Then there exists a unique Radon measure μ\mu on XX such that

λ(f)=Xf(x)μ(x)for allfCc(X).\lambda(f)=\int_{X}f(x)\ \mu(x)\quad\text{for all}\quad f\in C_{c}(X).

Furthermore, μ\mu satisfies μ(U)=sup{λ(f)fU}\mu(U)=\sup\{\lambda(f)\mid f\prec U\} and μ(K)=inf{T(f)Kf}\mu(K)=\inf\{T(f)\mid K\prec f\} for every open set UXU\subseteq X and every compact set KXK\subseteq X respectively.

2. Definition

In the context of topological groups it is natural to look for measures which are invariant under translation. Such measures always exist for locally compact Hausdorff groups.

Definition 2.1 (Haar measure).

Let GG be a locally compact Hausdorff group. A left (right) Haar measure on GG is a Radon measure μ\mu on (G,(G))(G,\operatorname{\mathcal{B}}(G)) which is non-zero on non-empty open sets and invariant under left-translation (right-translation):

  1. (NT)

    If UXU\subseteq X is open and non-empty, then μ(U)0\mu(U)\gneq 0. (non-trivial)

  2. (TI)

    For all E(G)E\in\operatorname{\mathcal{B}}(G) and gGg\in G: μ(gE)=μ(E)\mu(gE)=\mu(E) (μ(Eg)=μ(E)\mu(Eg)=\mu(E)). (translation-invariant)

Theorem 2.2 (Haar, Weil).

Let GG be a locally compact Hausdorff group. Then there exists a left (right) Haar measure on GG which is unique up to strictly positive scalar multiples.

We do not prove this theorem here but make the following remark.

Remark 2.3.

Whereas the uniqueness statement of Theorem 2.2 is not too hard to establish, the existence proof is more involved and not particularly fruitful. For both, see e.g. [Wei65]. However, there are several classes of locally compact Hausdorff groups for which the existence of a Haar measure may be established by more classical means, see Remark 2.8.

Example 2.4.

Let GG be a discrete group. Then (G)=𝒫(G)\operatorname{\mathcal{B}}(G)=\operatorname{\mathcal{P}}(G) and the counting measure on GG, defined by μ:𝒫(G)0{}\mu:\operatorname{\mathcal{P}}(G)\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\}, E|E|E\mapsto|E| is a left and right Haar measure.

More examples of Haar measures are given in Example 2.7. For now, consider the following alternative description of Haar measures: Due to Riesz’ Theorem 1.13, there is a one-to-one correspondence between Haar measures and Haar functionals, to be defined shortly, on a given group which is often used to define a Haar measure. Recall that a topological group GG acts on Cc(G)C_{c}(G) via the left-regular and the right-regular representations λG(g)f(x)=f(g1x)\lambda_{G}(g)f(x)=f(g^{-1}x) and ϱG(g)f(x)=f(xg)\varrho_{G}(g)f(x)=f(xg) respectively, where g,xGg,x\in G and fCc(G)f\in C_{c}(G).

Definition 2.5.

Let GG be a locally compact Hausdorff group. A left (right) Haar functional on GG is a non-trivial positive linear functional on Cc(G)C_{c}(G) which is invariant under λG\lambda_{G} (ϱG)(\varrho_{G}).

Proposition 2.6.

Let GG be a locally compact Hausdorff group. Then there are the following mutually inverse maps.

Φ:{Haar measures on G}\textstyle{\Phi:\{\text{Haar measures on $G$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Integration{Haar functionals on G}:Ψ\textstyle{\{\text{Haar functionals on $G$}\}:\Psi\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Riesz
Proof.

The map Φ\Phi is readily checked to range in the positive linear functionals on Cc(G)C_{c}(G). For λG\lambda_{G}-invariance (ϱG\varrho_{G}-invariance), use the change of variables formula given by Proposition 1.10. As to non-triviality, let μ\mu be a left (right) Haar measure on GG and let KK be a compact neighbourhood of some point in GG. Then μ(K)(0,)\mu(K)\in(0,\infty) by (LF) and (NT), and by Urysohn’s Lemma 1.6 there is fCc(G)f\in C_{c}(G) such that KfGK\prec f\prec G and therefore Φμ(f)=Gf(g)μ(g)μ(K)0\Phi\mu(f)=\int_{G}f(g)\ \mu(g)\geq\mu(K)\gneq 0.

Conversely, if λ\lambda is a left (right) Haar functional on GG, its non-triviality translates to (NT) for μ:=Ψλ\mu:=\Psi\lambda and its invariance under λG\lambda_{G} (ϱG)(\varrho_{G}) translates to (TI) for μ\mu:

Suppose UU is a non-empty open set of measure zero with respect to μ\mu. Then any compact set admits a finite cover by left (right) translates of UU and hence has measure zero as well. Thus λ(f)=Gf(g)μ(g)=suppff(g)μ(g)=0\lambda(f)=\!\int_{G}f(g)\ \mu(g)=\!\int_{\operatorname{supp}f}f(g)\ \mu(g)=0 for all fCc(G)f\in C_{c}(G), contradicting the non-triviality of λ\lambda.

As for invariance, suppose that λ\lambda is λG\lambda_{G}-invariant (ϱG\varrho_{G}-invariance being handled analogously) and let E(G)E\in\operatorname{\mathcal{B}}(G) and gGg\in G. Then (OR) implies

μ(gE)=inf{μ(U)UgE,U open}=inf{μ(gU)UE,U open}.\mu(gE)=\inf\{\mu(U)\mid U\supseteq gE,\ U\text{ open}\}=\inf\{\mu(gU)\mid U\supseteq E,\ U\text{ open}\}.

Furthermore, by Theorem 1.13 and the λG\lambda_{G}-invariance of λ\lambda we have

μ(gU)=sup{λ(f)fgU}=sup{λ(λG(g)f)fU}=μ(U).\mu(gU)=\sup\{\lambda(f)\mid f\prec gU\}=\sup\{\lambda(\lambda_{G}(g)f)\mid f\prec U\}=\mu(U).

Hence μ\mu is left-invariant. The assertions ΨΦ=id\Psi\circ\Phi=\operatorname{id} and ΦΨ=id\Phi\circ\Psi=\operatorname{id} are immediate. ∎

Example 2.7.

Using Proposition 2.6 we now provide further examples of Haar measures.

  1. (i)

    On G=(,+)G=(\operatorname{\mathbb{R}},+), a left- and right Haar measure is given by the Lebesgue measure λ\lambda which can be defined as the Radon measure associated to the Riemann integral :Cc()\int_{\operatorname{\mathbb{R}}}:C_{c}(\operatorname{\mathbb{R}})\to\operatorname{\mathbb{C}}.

  2. (ii)

    On G=(n,+)G=(\operatorname{\mathbb{R}}^{n},+), n1n\geq 1, a left- and right Haar measure is given by the nn-th power of the Lebesgue measure λ\lambda.

  3. (iii)

    On G=(,)G=(\operatorname{\mathbb{R}}^{\ast},\cdot), the Lebesgue measure is not translation-invariant. However, the map

     μ:Cc(G),ff(x)λ(x)|x|\hbox to0.0pt{\hfil}\mu:C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{\operatorname{\mathbb{R}}}f(x)\ \frac{\lambda(x)}{|x|}

    can be checked to be a left- and right Haar functional using the classical substitution rule. Note that the above integral is always finite as the integrand has compact support. Hence μ\mu defines a left- and right Haar measure on GG.

  4. (iv)

    On G=GL(n,)G=\operatorname{GL}(n,\operatorname{\mathbb{R}}), n1n\geq 1, the map

     μ:Cc(G),fGf(X)λ(X)|detX|n\hbox to0.0pt{\hfil}\mu:C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G}f(X)\ \frac{\lambda(X)}{|\det X|^{n}}

    defines a left- and right Haar functional. Here, λ(X):=i,j=1nλ(xij)\lambda(X):=\prod_{i,j=1}^{n}\lambda(x_{ij}), where X=(xij)i,jX=(x_{ij})_{i,j}, is the Lebesgue measure on nn\operatorname{\mathbb{R}}^{n\cdot n} of which GL(n,)\operatorname{GL}(n,\operatorname{\mathbb{R}}) is an open subset. Again, the integral is finite by compactness of the support of the integrand and invariance is checked by changing variables. Note that the case G=(,)G=(\operatorname{\mathbb{R}}^{\ast},\cdot) is contained via n=1n=1 in this example.

    The fact that GL(n,)\operatorname{GL}(n,\operatorname{\mathbb{R}}) is an open subset of nn\operatorname{\mathbb{R}}^{n\cdot n} is key: The above construction does not work for e.g. SL(n,)\operatorname{SL}(n,\operatorname{\mathbb{R}}) which is a submanifold of nn\operatorname{\mathbb{R}}^{n\cdot n} of strictly smaller dimension. A left- and right Haar measure for SL(2,)\operatorname{SL}(2,\operatorname{\mathbb{R}}) will be constructed in Example 4.5.

Remark 2.8.

With the correspondence between Haar functionals and Haar measures at hand, we now outline existence proofs of Theorem 2.2 for compact Hausdorff groups, Lie groups and totally disconnected locally compact separable Hausdorff groups.

  1. (i)

    Compact Hausdorff groups. Let GG be a compact Hausdorff group. Then GG acts continuously on C(G)=Cc(G)C(G)=C_{c}(G), equipped with the supremum norm, via the left-regular representation. Therefore, GG also acts on the dual space C(G)C(G)^{\ast} of C(G)C(G) via the adjoint representation λG\lambda_{G}^{\ast} of λG\lambda_{G}, which is defined by the relation

     λG(g)μ,f=μ,λG(g1)f\hbox to0.0pt{\hfil}\langle\lambda_{G}^{\ast}(g)\mu,f\rangle=\langle\mu,\lambda_{G}(g^{-1})f\rangle

    for all μC(G)\mu\in C(G)^{\ast} and fC(G)f\in C(G). Since the set P(G)P(G) of probability measures on GG is a weak\text{weak}^{\ast}-compact, convex and λG\lambda_{G}^{\ast}-invariant subset of C(G)C(G)^{\ast}, the compact version of the Kakutani-Markov Fixed Point Theorem (e.g. [Zim90, Thm. 2.23]) provides a λG\lambda_{G}^{\ast}-fixed point within P(G)P(G), i.e. a left-invariant probability measure, which turns out to be a Haar measure.

  2. (ii)

    Lie groups. Let GG be a Lie group with Lie algebra Lie(G)Γ(TG)G\text{Lie}(G)\cong\Gamma(\mathrm{T}G)^{G}, the space of left-invariant vector fields on GG, which is isomorphic to the tangent space TeG\mathrm{T}_{e}G as a vector space. Further, let X1,,XnX_{1},\ldots,X_{n} be a basis of TeG\mathrm{T}_{e}G with associated left-invariant vector fields X1G,,XnGΓ(TG)GX_{1}^{G},\ldots,X_{n}^{G}\in\Gamma(\mathrm{T}G)^{G}. Then for each pGp\in G, the tuple ((X1G)p,,(XnG)p)((X_{1}^{G})_{p},\ldots,(X_{n}^{G})_{p}) is a basis of TpG\mathrm{T}_{p}G. For each i{1,,n}i\in\{1,\ldots,n\} we may thus define a 11-form ωi\omega_{i} on GG by (ωi)p((Xj)p)=δij(\omega_{i})_{p}((X_{j})_{p})=\delta_{ij}; in other words, for every pGp\in G the tuple ((ω1)p,,(ωn)p)((\omega_{1})_{p},\ldots,(\omega_{n})_{p}) is the basis of TpG\mathrm{T}_{p}^{\ast}G dual to ((X1G)p,,(XnG)p)((X_{1}^{G})_{p},\ldots,(X_{n}^{G})_{p}). It is readily checked that the left-invariance of X1G,,XnGX_{1}^{G},\ldots,X_{n}^{G} implies left-invariance of the ωi\omega_{i} (i{1,,n})(i\in\{1,\ldots,n\}) in the sense that Lgωi=ωiL_{g}^{\ast}\omega_{i}=\omega_{i} for all gGg\in G and i{1,,n}i\in\{1,\ldots,n\}. As a consequence, the nn-form ω:=ω1ωn\omega:=\omega_{1}\wedge\cdots\wedge\omega_{n} is left-invariant as well since \wedge commutes with pullback. One checks that ω\omega is nowhere vanishing. Finally, we may orient GG so that ω\omega is positive and hence gives rise to the left Haar functional

     λω:Cc(G),fGfω\hbox to0.0pt{\hfil}\lambda_{\omega}:C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G}f\ \omega

    which in turn provides a left Haar measure on GG. See [Kna02, VIII.2] for details.

  3. (iii)

    Totally disconnected locally compact separable Hausdorff groups. Let GG be of this type. By van Dantzig’s theorem [vD31], GG contains a compact open subgroup KK. Assuming GG to be non-compact, by separability and openness of KK there are gnGg_{n}\in G (n)(n\in\operatorname{\mathbb{N}}) such that G=ngnKG=\bigsqcup_{n\in\operatorname{\mathbb{N}}}g_{n}K. Using part (i), let ν\nu be a Haar measure on KK and let νn:=gnν\nu_{n}:=g_{n\ast}\nu be the corresponding measure on gnKg_{n}K. Finally, for E(G)E\in\operatorname{\mathcal{B}}(G) define

     μ(E):=nνn(EgnK)=nν(gn1EK)\hbox to0.0pt{\hfil}\mu(E):=\sum_{n\in\operatorname{\mathbb{N}}}\nu_{n}(E\cap g_{n}K)=\sum_{n\in\operatorname{\mathbb{N}}}\nu(g_{n}^{-1}E\cap K)

    if the sum exists and infinity otherwise. Then μ\mu is a Radon measure on GG which is non-zero on non-empty open sets since ν\nu is. Also, μ\mu is left-invariant: Given gGg\in G, there is σS\sigma\in S_{\operatorname{\mathbb{N}}} such that ggnK=gσ(n)Kgg_{n}K=g_{\sigma(n)}K. Then

    μ(g1E)\displaystyle\hbox to0.0pt{\hfil}\mu(g^{-1}E) =nν(gn1g1EK)=nν(gσ(n)1ggngn1g1EK)\displaystyle=\sum_{n\in\operatorname{\mathbb{N}}}\nu(g_{n}^{-1}g^{-1}E\cap K)=\sum_{n\in\operatorname{\mathbb{N}}}\nu(g_{\sigma(n)}^{-1}gg_{n}g_{n}^{-1}g^{-1}E\cap K)
    =nν(gσ(n)1EK)=nν(gn1EK)=μ(E).\displaystyle=\sum_{n\in\operatorname{\mathbb{N}}}\nu(g_{\sigma(n)}^{-1}E\cap K)=\sum_{n\in\operatorname{\mathbb{N}}}\nu(g_{n}^{-1}E\cap K)=\mu(E).

    where the second equality uses KK-invariance of ν\nu.

By Remark 2.8i, compact Hausdorff groups have finite Haar measure. The converse also holds.

Proposition 2.9.

Let GG be a locally compact Hausdorff group and let μ\mu be a left (right) Haar measure on GG. Then μ(G)<\mu(G)<\infty if and only if GG is compact.

Proof.

If GG is compact, then μ(G)<\mu(G)<\infty by Definition (LF). Conversely, suppose that GG is not compact and let UU be a relatively compact neighbourhood of eGe\in G. Then there is an infinite sequence (gn)n(g_{n})_{n\in\operatorname{\mathbb{N}}} of elements of GG such that gnk<ngkUg_{n}\notin\bigcup_{k<n}g_{k}U; otherwise GG would be compact as a finite union of compact sets. Let VV be as in Lemma 1.2. Then the sets gnVg_{n}V (n)(n\in\operatorname{\mathbb{N}}) are pairwise disjoint by the fact that VV1UVV^{-1}\subseteq U and the definition of (gn)n(g_{n})_{n\in\operatorname{\mathbb{N}}}. Therefore, as VV has strictly positive measure, GG has infinite measure. ∎

3. Unimodularity

We now address and quantify the question whether left and right Haar measures on a given locally compact Hausdorff group coincide.

Definition 3.1.

A locally compact Hausdorff group GG is unimodular if every left Haar measure on GG is also a right Haar measure on GG and conversely.

Remark 3.2.

By Theorem 2.2, it suffices in Definition 3.1 to ask for every left Haar measure on GG to also be a right Haar measure.

Proposition 3.6 below provides several classes of unimodular groups. For now, let GG be a locally compact Hausdorff group and let μ\mu be a left Haar measure on GG. Then for every gGg\in G, the map μg:(G)0{},Eμ(Eg)\mu_{g}:\operatorname{\mathcal{B}}(G)\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\},\ E\mapsto\mu(Eg) is a left Haar measure on GG as well. Hence, by uniqueness, there exists a strictly positive real number ΔG(g)\Delta_{G}(g) such that μg=ΔG(g)μ\mu_{g}=\Delta_{G}(g)\mu, i.e.

(M) μ(Eg)=μg(E)=ΔG(g)μ(E)for allE(G).\mu(Eg)=\mu_{g}(E)=\Delta_{G}(g)\mu(E)\quad\text{for all}\quad E\in\operatorname{\mathcal{B}}(G).

The function ΔG:G>0\Delta_{G}:G\to\operatorname{\mathbb{R}}_{>0} is independent of μ\mu and called the modular function of GG.

Let λ\lambda be the left Haar functional associated to μ\mu by Proposition 2.6. Then by the change of variables formula of Proposition 1.10 applied to φ=Rg1\varphi=R_{g^{-1}}, equation (M) immediately translates to

(M’) λ(ϱG(g1)f)=ΔG(g)λ(f)for allfCc(G).\lambda(\varrho_{G}(g^{-1})f)=\Delta_{G}(g)\lambda(f)\quad\text{for all}\quad f\in C_{c}(G).
Proposition 3.3.

Let GG be a locally compact Hausdorff group. Then the modular function ΔG\Delta_{G} is a continuous homomorphism from GG to (>0,)(\operatorname{\mathbb{R}}_{>0},\cdot).

Proof.

Let μ\mu be a left Haar measure on GG. The homomorphism property is immediate from (M): For all g,hGg,h\in G we have

ΔG(gh)μ=μgh=(μg)h=ΔG(h)μg=ΔG(h)ΔG(g)μ=ΔG(g)ΔG(h)μ.\Delta_{G}(gh)\mu=\mu_{gh}=(\mu_{g})_{h}=\Delta_{G}(h)\mu_{g}=\Delta_{G}(h)\Delta_{G}(g)\mu=\Delta_{G}(g)\Delta_{G}(h)\mu.

Evaluating on a set of non-zero finite measure, e.g. a compact neighbourhood of some point, proves that indeed ΔG(gh)=ΔG(g)ΔG(h)\Delta_{G}(gh)=\Delta_{G}(g)\Delta_{G}(h).

As to continuity, note that it suffices to check continuity at eGe\in G, since ΔG\Delta_{G} is a homomorphism. Let λ\lambda be the left Haar functional associated to μ\mu by Proposition 2.6 and let KK be a compact neighbourhood of eGe\in G. Using Urysohn’s Lemma 1.6, choose φCc(G)\varphi\in C_{c}(G) such that KφGK\prec\varphi\prec G and ψCc(G)\psi\in C_{c}(G) such that KsuppφψGK\operatorname{supp}\varphi\prec\psi\prec G (see Proposition 1.8). In particular, φ\varphi is uniformly continuous on the right by Proposition 1.7. Hence, given ε>0\varepsilon>0, there is a symmetric open neighbourhood UKU\subseteq K of eGe\in G such that |φ(xg)φ(x)|<ε|\varphi(xg)-\varphi(x)|<\varepsilon for all gUg\in U. Then by (M’) we have

|ΔG(g)1|=1λ(φ)|ΔG(g)λ(φ)λ(φ)|1λ(φ)λ(|ϱG(g1)φφ|ψ)ελ(ψ)λ(φ)|\Delta_{G}(g)-1|=\frac{1}{\lambda(\varphi)}\left|\Delta_{G}(g)\lambda(\varphi)-\lambda(\varphi)\right|\leq\frac{1}{\lambda(\varphi)}\lambda(|\varrho_{G}(g^{-1})\varphi-\varphi|\psi)\leq\varepsilon\frac{\lambda(\psi)}{\lambda(\varphi)}

for all gUg\in U. Hence ΔG\Delta_{G} is continuous at eGe\in G. ∎

Remark 3.4.

We have noticed that for a locally compact Hausdorff group GG with left Haar measure μ\mu and given gGg\in G, the map μg:(G)0{},Eμ(Eg)\mu_{g}:\operatorname{\mathcal{B}}(G)\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\},\ E\mapsto\mu(Eg) is a left Haar measure on GG as well. This is an instance of the following more general observation: For every continuous automorphism αAut(G)\alpha\in\operatorname{Aut}(G), the map μα:(G)0{},Eμ(α(E))\mu_{\alpha}:\operatorname{\mathcal{B}}(G)\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\},\ E\mapsto\mu(\alpha(E)) is a left Haar measure on GG. In this setting, μg=μint(g1)\mu_{g}=\mu_{\text{int}(g^{-1})} where int(g):GG,xgxg1\text{int}(g):G\to G,\ x\mapsto gxg^{-1} denotes conjugation in GG by gg. One may then introduce the general modular function modG:Aut(G)(>0,)\text{mod}_{G}:\operatorname{Aut}(G)\to(\operatorname{\mathbb{R}}_{>0},\cdot) which remains to be a homomorphism and when Aut(G)\operatorname{Aut}(G) is equipped with the Braconnier topology, a refinement of the compact-open topology, becomes continuous. See e.g. [Pal01, 12.1.12] for details.

We obtain the following useful criterion for unimodularity.

Corollary 3.5.

A locally compact Hausdorff group GG is unimodular if and only if ΔG1\Delta_{G}\equiv 1.

Proof.

If ΔG1\Delta_{G}\equiv 1, then GG is unimodular by (M) and Remark 3.2. Conversely, if GG is unimodular, let μ\mu be a Haar measure on GG and let EE be a compact neighbourhood of some point in GG. Then μ(E)(0,)\mu(E)\in(0,\infty) and hence ΔG1\Delta_{G}\equiv 1 by (M). ∎

Corollary 3.5 provides us with the following list of classes of unimodular groups. Yet another class will be given in Proposition 4.12.

Proposition 3.6.

Let GG be a locally compact Hausdorff group. Then GG is unimodular if, in addition, it satisfies one of the following properties: being abelian, compact, discrete, topologically simple, connected semisimple Lie or connected nilpotent Lie.

Proof.

When GG is abelian then Eg=gEEg=gE for every subset EGE\subseteq G and all gGg\in G. Hence left-invariance implies right-invariance.

When GG is compact and μ\mu is a left Haar measure on GG, then μ(G)(0,)\mu(G)\in(0,\infty) by (LF) and (NT) and therefore ΔG1\Delta_{G}\equiv 1 by (M).

For a discrete group, the left Haar measures are the strictly positive scalar multiples of the counting measure which is also right-invariant.

When GG is topologically simple, then [G,G]¯\overline{[G,G]}, which is a closed normal subgroup of GG, either equals {e}\{e\} or GG. In the first case GG is abelian and hence unimodular. In the latter case, continuity of ΔG\Delta_{G} implies ΔG(G)=ΔG([G,G]¯)ΔG([G,G])¯={1}\smash{\Delta_{G}(G)=\Delta_{G}(\overline{[G,G]})\subseteq\overline{\Delta_{G}([G,G])}=\{1\}} and hence GG is unimodular.

When GG is a Lie group, the modular function ΔG:G(>0,)\Delta_{G}:G\to(\operatorname{\mathbb{R}}_{>0},\cdot) is a continuous and hence smooth ([War83, Thm. 3.39]) homomorphism of Lie groups. It is given by ΔG(g)=|detAd(g)|\Delta_{G}(g)=|\det\text{Ad}(g)|, where Ad:GAut(Lie(G))\text{Ad}:G\to\operatorname{Aut}(\text{Lie}(G)) is the adjoint representation of GG, see e.g. [Kna02, Prop. 8.27], which follows in the setting of Remark 2.8ii. In particular, the derivative DeΔG:Lie(G)D_{e}\Delta_{G}:\text{Lie}(G)\to\operatorname{\mathbb{R}} is a morphism of Lie algebras. When Lie(G)\text{Lie}(G) is semisimple we obtain

DeΔG(Lie(G))=DeΔG([Lie(G),Lie(G)])=[DeΔG(Lie(G)),DeΔG(Lie(G))]={0}D_{e}\Delta_{G}(\text{Lie}(G))=D_{e}\Delta_{G}([\text{Lie}(G),\text{Lie}(G)])=[D_{e}\Delta_{G}(\text{Lie}(G)),D_{e}\Delta_{G}(\text{Lie}(G))]=\{0\}

as (>0,)(\operatorname{\mathbb{R}}_{>0},\cdot) is abelian. Thus ΔG1\Delta_{G}\!\equiv\!1 by the Lie correspondence, passing to the universal cover of GG.

For a connected nilpotent Lie group the exponential map exp:Lie(G)G\text{exp}:\text{Lie}(G)\to G is surjective, see e.g. [Kna02, Thm. 1.127]. So for every gGg\in G there is some XLie(G)X\in\text{Lie}(G) such that g=exp(X)g=\text{exp}(X) and

ΔG(g)=|detAd(g)|=|detAd(expX)|=|deteadX|=etr(adX)=1\Delta_{G}(g)=|\det\text{Ad}(g)|=|\det\text{Ad}(\exp X)|=|\det e^{\text{ad}X}|=e^{\operatorname{tr}(\text{ad}X)}=1

where the last equality follows from the fact that adX\text{ad}X is nilpotent as Lie(G)\text{Lie}(G) is. ∎

Remark 3.7.

It can be shown that GG is unimodular if and only if G/Z(G)G/Z(G) is unimodular, see e.g. [Nac76, Proposition 25]. Hence any nilpotent locally compact Hausdorff group is unimodular. Solvable groups, however, need not be unimodular, see Example 3.9i.

The following proposition provides a class of totally disconnected locally compact Hausdorff groups that are unimodular. Recall that if TT is a locally finite tree then Aut(T)\operatorname{Aut}(T) is a totally disconnected locally compact separable Hausdorff group with the permutation topology. We adopt Serre’s graph theory conventions, see [Ser80].

Proposition 3.8.

Let T=(V,E)T=(V,E) be a locally finite tree. If GAut(T)G\leq\operatorname{Aut}(T) is closed and locally transitive then GG is unimodular.

Proof.

Let μ\mu be a left Haar measure on GG, see Remark 2.8. Since GG is locally transitive there is for every triple (x,e0,e)(x,e_{0},e) of a vertex xVx\in V and edges e0,eE(x)e_{0},e\in E(x) an element geGxg_{e}\in G_{x} such that gee0=eg_{e}e_{0}=e. Then Gx=eE(x)geGe0G_{x}=\bigsqcup_{e\in E(x)}g_{e}G_{e_{0}}. Since Ge=Ge¯G_{e}=G_{\overline{e}} for all eEe\in E we conclude that μ(Ge)=μ(Ge)\mu(G_{e})=\mu(G_{e^{\prime}}) for all e,eEe,e^{\prime}\in E. Given gGg\in G we therefore have

μ(Ge)=μ(Gge)=μ(gGeg1)=μ(Geg1)=ΔG(g1)μ(Ge).\mu(G_{e})=\mu(G_{ge})=\mu(gG_{e}g^{-1})=\mu(G_{e}g^{-1})=\Delta_{G}(g^{-1})\mu(G_{e}).

Since μ(Ge)(0,)\mu(G_{e})\in(0,\infty) as a compact open subgroup of Aut(T)\operatorname{Aut}(T) we conclude that GG is unimodular. ∎

Example 3.9.

We now provide two related examples of non-unimodular groups, cf. Remark 4.6.

  1. (i)

    Consider the group

     P:={(xyx1)|x\{0},y}SL(2,).\hbox to0.0pt{\hfil}P:=\left\{\left.\begin{pmatrix}x&y\\ &x^{-1}\end{pmatrix}\right|x\in\operatorname{\mathbb{R}}\backslash\{0\},\ y\in\operatorname{\mathbb{R}}\right\}\leq\operatorname{SL}(2,\operatorname{\mathbb{R}}).

    Then the functionals μ,ν:Cc(P)\mu,\nu:C_{c}(P)\to\operatorname{\mathbb{C}}, given by

     μ:f2f(X)λ(x)λ(y)x2andν:f2f(X)λ(x)λ(y)\hbox to0.0pt{\hfil}\mu:f\mapsto\int_{\operatorname{\mathbb{R}}^{2}}f(X)\ \frac{\lambda(x)\lambda(y)}{x^{2}}\quad\text{and}\quad\nu:f\mapsto\int_{\operatorname{\mathbb{R}}^{2}}f(X)\ \lambda(x)\lambda(y)

    are left- and right Haar functionals respectively as can be checked by changing variables. However, PP is a closed subgroup of SL(2,)\operatorname{SL}(2,\operatorname{\mathbb{R}}) which is unimodular as a connected simple Lie group by Proposition 3.6. Remark 4.6 sheds some light on the origin of this example.

  2. (ii)

    Let Td=(V,E)T_{d}=(V,E) be the dd-regular tree and let ωTd\omega\in\partial T_{d} be a boundary point of TdT_{d}. Set G:=Aut(Td)ωG:=\operatorname{Aut}(T_{d})_{\omega}, the stabiliser of ω\omega in Aut(Td)\operatorname{Aut}(T_{d}). Then GG is not unimodular: Let tGt\in G be a translation of length 11 towards ω\omega and let xVx\in V be on the translation axis of tt, then

     Δ(t)=μ(Gx)μ(Gtx)=μ(Gx)μ(Gx,tx)μ(Gx,tx)μ(Gtx)=[Gx:Gx,tx][Gtx:Gx,tx]=|Gxtx||Gtxx|=1d1.\hbox to0.0pt{\hfil}\Delta(t)=\frac{\mu(G_{x})}{\mu(G_{tx})}=\frac{\mu(G_{x})}{\mu(G_{x,tx})}\frac{\mu(G_{x,tx})}{\mu(G_{tx})}=\frac{[G_{x}:G_{x,tx}]}{[G_{tx}:G_{x,tx}]}=\frac{|G_{x}tx|}{|G_{tx}x|}=\frac{1}{d-1}.

Uilising the modular function, we can turn left Haar measures into right Haar measures as in the following Proposition. Let i:GGi:G\to G denote the inversion map of GG.

Proposition 3.10.

Let GG be a locally compact Hausdorff group with left Haar measure μ\mu. Then μ¯=iμ:(G)0{},Eμ(E1)\overline{\mu}=i_{\ast}\mu:\operatorname{\mathcal{B}}(G)\to\operatorname{\mathbb{R}}_{\geq 0}\cup\{\infty\},\ E\mapsto\mu(E^{-1}) is a right Haar measure on GG with associated right Haar functional ϱ:Cc(G),fGf(x)ΔG(x1)μ(x)\varrho:C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G}f(x)\Delta_{G}(x^{-1})\ \mu(x). If GG is unimodular, then μ¯=μ\overline{\mu}=\mu.

Proof.

The map μ¯\overline{\mu} is readily checked to be a right Haar measure on GG. The map ϱ\varrho is clearly positive and linear. Its non-triviality follows as in the proof of Proposition 2.6 using that ΔG(g)0\Delta_{G}(g)\gneq 0 for all gGg\in G. As to ϱG\varrho_{G}-invariance, changing variables and using Rgμ=μg1R_{g\ast}\mu=\mu_{g^{-1}} yields

ϱ(ϱG\displaystyle\varrho(\varrho_{G} (g)f)=Gf(xg)ΔG(x1)μ(x)=Gf(x)ΔG(gx1)μg1(x)=\displaystyle(g)f)=\int_{G}f(xg)\Delta_{G}(x^{-1})\ \mu(x)=\int_{G}f(x)\Delta_{G}(gx^{-1})\ \mu_{g^{-1}}(x)=
=Gf(x)ΔG(g)ΔG(x1)ΔG(g1)μ(x)=Gf(x)ΔG(x1)μ(x)=ϱ(f).\displaystyle=\int_{G}f(x)\Delta_{G}(g)\Delta_{G}(x^{-1})\Delta_{G}(g^{-1})\ \mu(x)=\int_{G}f(x)\Delta_{G}(x^{-1})\ \mu(x)=\varrho(f).

for every fCc(G)f\in C_{c}(G) and gGg\in G. Overall, ϱ\varrho is a right Haar functional on GG.

Now, let Φμ¯\Phi\overline{\mu} denote the right Haar functional associated to μ¯\overline{\mu} by Proposition 2.6. Then there is a strictly positive real number cc such that Φμ¯=cϱ\Phi\overline{\mu}=c\varrho. Applying the change of variables formula given by Proposition 1.10, we obtain for all fCc(G)f\in C_{c}(G):

Gf(x)μ¯(x)\displaystyle\int_{G}f(x)\ \overline{\mu}(x) =cGf(x)ΔG(x1)μ(x)=cGf(x1)ΔG(x)μ¯(x)\displaystyle=c\int_{G}f(x)\Delta_{G}(x^{-1})\ \mu(x)=c\int_{G}f(x^{-1})\Delta_{G}(x)\ \overline{\mu}(x)
=c2Gf(x1)ΔG(x)ΔG(x1)μ(x)=c2Gf(x)μ¯(x).\displaystyle=c^{2}\int_{G}f(x^{-1})\Delta_{G}(x)\Delta_{G}(x^{-1})\ \mu(x)=c^{2}\int_{G}f(x)\ \overline{\mu}(x).

Let KK be a compact symmetric neighbourhood of a point in GG and fCc(G)f\in C_{c}(G) with KfGK\prec f\prec G. Then Gf(x1)μ(x)(0,)\int_{G}f(x^{-1})\ \mu(x)\in(0,\infty) and hence c=1c=1. Henceu unimodularity of GG implies μ=μ¯\mu=\overline{\mu}. ∎

4. Coset spaces

Let GG be a locally compact Hausdorff group and let HH be a closed subgroup of GG. When HH is normal in GG, there exists a left (right) Haar measure on G/HG/H by Theorem 2.2. We now address the question under which circumstances there exists a GG-invariant Radon measure on G/HG/H that is non-zero on non-empty open sets when HH is not normal in GG. We shall refer to such a measure as a Haar measure on G/HG/H by abuse of notation. The following example shows that a Haar measure on G/HG/H may or may not exist.

Example 4.1.

Let G=SL(2,)G=\operatorname{SL}(2,\operatorname{\mathbb{R}}).

  1. (i)

    Consider the natural, transitive action of GG on X=2\{0}X=\operatorname{\mathbb{R}}^{2}\backslash\{0\} and the stabiliser

     H:=stabG((1,0)T)={(1x1)|x}.\hbox to0.0pt{\hfil}H:=\operatorname{stab}_{G}((1,0)^{T})=\left\{\left.\begin{pmatrix}1&x\\ &1\end{pmatrix}\right|x\in\operatorname{\mathbb{R}}\right\}.

    Then G/HXG/H\cong X on which the restricted 22-dimensional Lebesgue measure is a Haar measure.

  2. (ii)

    Consider the natural, transitive action of GG on X=1={V2dimV=1}X=\operatorname{\mathbb{P}}^{1}\operatorname{\mathbb{R}}=\{V\leq\operatorname{\mathbb{R}}^{2}\mid\dim V=1\} and

     H:=stabG(e1)={(xyx1)|x\{0},y}.\hbox to0.0pt{\hfil}H:=\operatorname{stab}_{G}(\langle e_{1}\rangle)=\left\{\left.\begin{pmatrix}x&y\\ &x^{-1}\end{pmatrix}\right|x\in\operatorname{\mathbb{R}}\backslash\{0\},y\in\operatorname{\mathbb{R}}\right\}.

    Here, G/HXG/H\cong X does not admit a Haar measure: Indeed, consider the compact subsets E1:={(1,t)Tt[0,1]}E_{1}:=\{\langle(1,t)^{T}\rangle\mid t\in[0,1]\} and E2:={(t,1)Tt[0,1]}E_{2}:=\{\langle(t,1)^{T}\rangle\mid t\in[0,1]\} of 1\operatorname{\mathbb{P}}^{1}\operatorname{\mathbb{R}}. Then

     (111)E1=E1E2and(111)E1=E2.\hbox to0.0pt{\hfil}\begin{pmatrix}1&-1\\ &1\end{pmatrix}E_{1}=E_{1}\cup E_{2}\quad\text{and}\quad\begin{pmatrix}1&-1\\ 1&\end{pmatrix}E_{1}=E_{2}.

    A Haar measure on G/HG/H would assign finite non-zero measure to the compact sets E1E_{1} and E2E_{2}. Combined with GG-invariance contradicts the above two equalities. Note that HH is the non-unimodular group of Example 3.9.

Theorem 4.2.

Let GG be a locally compact Hausdorff group with left Haar measure μ\mu and let HH be a closed subgroup of GG with left Haar measure ν\nu. Then there exists a Haar measure ξ\xi on G/HG/H if and only if ΔG|HΔH\Delta_{G}|_{H}\equiv\Delta_{H}. In this case, ξ\xi is unique up to strictly positive scalar multiples and suitably normalized satisfies for all fCc(G)f\in C_{c}(G):

(W) Gf(g)μ(g)=G/HHf(gh)ν(h)ξ(gH).\int_{G}f(g)\ \mu(g)=\int_{G/H}\int_{H}f(gh)\ \nu(h)\ \xi(gH).

In the context of Theorem 4.2, formula (W) can be extended to hold for fL1(G)f\in L^{1}(G), see e.g. [KL06, Theorem 7.12] and the explanations around it.

Proof.

(Theorem 4.2, “\Rightarrow”). If ξ\xi exists as above, then the map

λ:Cc(G),fG/HHf(gh)ν(h)ξ(gH)\lambda:C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G/H}\int_{H}f(gh)\ \nu(h)\ \xi(gH)

is a left Haar functional on GG and therefore defines a left Haar measure μ\mu on GG. In particular, λ(ϱG(t1)f)=ΔG(t)λ(f)\lambda(\varrho_{G}(t^{-1})f)=\Delta_{G}(t)\lambda(f) for all tGt\in G and fCc(G)f\in C_{c}(G) by (M’). On the other hand, we have for all tHt\in H and fCc(G)f\in C_{c}(G):

λ(ϱG(t1)f)\displaystyle\lambda(\varrho_{G}(t^{-1})f) =G/HH(ϱG(t1)f)(gh)ν(h)ξ(gH)\displaystyle=\int_{G/H}\int_{H}(\varrho_{G}(t^{-1})f)(gh)\ \nu(h)\ \xi(gH)
=G/HHΔH(t)f(gh)ν(h)ξ(gH)=ΔH(t)λ(f).\displaystyle=\int_{G/H}\int_{H}\Delta_{H}(t)f(gh)\ \nu(h)\ \xi(gH)=\Delta_{H}(t)\lambda(f).

Using Urysohn’s Lemma 1.6, choose fCc(G)f\in C_{c}(G) to satisfy KfGK\prec f\prec G where KK is a compact neighbourhood of some point in GG. Then Gf(g)μ(g)=λ(f)(0,)\int_{G}f(g)\ \mu(g)=\lambda(f)\in(0,\infty) and hence ΔG|HΔH\Delta_{G}|_{H}\equiv\Delta_{H}. ∎

The proof of the converse assertion of Theorem 4.2 relies on the following description of compactly supported functions on G/HG/H. Once more, Riesz’ Theorem 1.13 is used to produce a measure.

Lemma 4.3.

Let GG be a locally compact Hausdorff group and HH a closed subgroup of GG with left Haar measure ν\nu. Then the following map is surjective:

Cc(G)Cc(G/H),f(fH:gHHf(gh)ν(h)).C_{c}(G)\to C_{c}(G/H),\ f\mapsto\left(f_{H}:gH\mapsto\int_{H}f(gh)\ \nu(h)\right).
Proof.

Several things need to be checked. First of all, for all fCc(G)f\in C_{c}(G) and for all gHG/HgH\in G/H, the integral Hf(gh)ν(h)\int_{H}f(gh)\ \nu(h) is independent of the representative of gHgH and finite. Next, for all fCc(G)f\in C_{c}(G), the function fHf_{H} is continuous as a parametrized integral as in the proof of the continuity of the modular function. Clearly, suppfHπ(supp(f))\operatorname{supp}f_{H}\subseteq\pi(\operatorname{supp}(f)) and hence fHCc(G/H)f_{H}\in C_{c}(G/H). It remains to prove surjectivity. To this end, let FCc(G/H)F\in C_{c}(G/H). Pick KGK\subseteq G such that π(K)suppF\pi(K)\supseteq\operatorname{supp}F (Proposition 1.9) and let ηCc(G)\eta\in C_{c}(G) satisfying KηK\prec\eta (Urysohn’s Lemma 1.6). Now define fCc(G)f\in C_{c}(G) by

f:G,g{F(gH)η(g)ηH(gH)ηH(gH)00ηH(gH)=0f:G\to\operatorname{\mathbb{C}},\ g\mapsto\begin{cases}\frac{F(gH)\eta(g)}{\eta_{H}(gH)}&\eta_{H}(gH)\neq 0\\ 0&\eta_{H}(gH)=0\end{cases}

Again, we need to show that this function is continuous and has compact support. As for compact support, clearly suppfsuppη\operatorname{supp}f\subseteq\operatorname{supp}\eta. In fact, if GG was compact, we could choose η1\eta\equiv 1. To show that ff is continuous, we show that it is continuous at every point of two open sets U1GU_{1}\subseteq G and U2GU_{2}\subseteq G satisfying U1U2=GU_{1}\cup U_{2}=G. On the set U1:={gGηH(gH)0}U_{1}:=\{g\in G\mid\eta_{H}(gH)\neq 0\} it is continuous as a quotient of continuous functions; and on the set U2:=G\KHU_{2}:=G\backslash KH it is continuous because it vanishes. Further, if gU1g\not\in U_{1}, then 0=ηH(gH)=Hη(gh)ν(h)0=\eta_{H}(gH)=\int_{H}\eta(gh)\ \nu(h). Since η\eta is a non-negative continuous function, this implies η(gh)=0\eta(gh)=0 for all hHh\in H, hence gKHg\not\in KH, i.e. gU2g\in U_{2}. With continuity and compact support established, it remains to show that fHFf_{H}\equiv F. To this end, we compute

fH(gH)=HF(ghH)η(gh)ηH(ghHν(h)=F(gH)Hη(gh)ν(h)ηH(gH)=F(gH).f_{H}(gH)=\int_{H}\frac{F(ghH)\eta(gh)}{\eta_{H}(ghH}\ \nu(h)=F(gH)\frac{\int_{H}\eta(gh)\ \nu(h)}{\eta_{H}(gH)}=F(gH).

Hence the map ()H:Cc(G)Cc(G/H)(-)_{H}:C_{c}(G)\to C_{c}(G/H) is surjective. ∎

Proof.

(Theorem 4.2, “\Leftarrow”). Lemma 4.3 allows us to pick a be a right-inverse s:Cc(G/H)Cc(G)s:C_{c}(G/H)\to C_{c}(G) for the map ()H:Cc(G)Cc(G/H),ffH(-)_{H}:C_{c}(G)\to C_{c}(G/H),\ f\mapsto f_{H} of the same lemma. Now consider the map

λ:Cc(G/H),fG(sf)(g)μ(g).\lambda:C_{c}(G/H)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G}(sf)(g)\ \mu(g).

Once λ\lambda is independent of ss, it is a positive linear functional. To prove that it is independent of ss, it suffices to show that Gf(g)μ(g)=0\int_{G}f(g)\ \mu(g)=0 whenever fH0f_{H}\equiv 0. By Lemma 4.3 and Urysohn’s Lemma 1.6 there is a function ηCc(G)\eta\in C_{c}(G) such that (suppf)HηHG/H(\operatorname{supp}f)H\!\prec\eta_{H}\!\prec G/H. Then by Proposition 3.10 we have

Gf(g)μ(g)\displaystyle\int_{G}f(g)\ \mu(g) =GηH(gH)f(g)μ(g)=GHη(gh)f(g)ν(h)μ(g)\displaystyle=\int_{G}\eta_{H}(gH)f(g)\ \mu(g)=\int_{G}\int_{H}\eta(gh)f(g)\ \nu(h)\ \mu(g)
=GHη(gh1)f(g)ΔH(h1)ν(h)μ(g).\displaystyle=\int_{G}\int_{H}\eta(gh^{-1})f(g)\Delta_{H}(h^{-1})\ \nu(h)\ \mu(g).

We may as well integrate over the compact spaces suppfG\operatorname{supp}f\subseteq G and (suppη)1suppfHH(\operatorname{supp}\eta)^{-1}\operatorname{supp}f\cap H\subseteq H (Proposition 1.8). Fubini’s Theorem 1.11 then allows us to continue the above computation by

=HGη(gh1)f(g)ΔH(h1)μ(g)ν(h)=HGη(g)f(gh)ΔH(h1)ΔG(h)μ(g)ν(h).=\int_{H}\int_{G}\eta(gh^{-1})f(g)\Delta_{H}(h^{-1})\ \mu(g)\ \nu(h)=\int_{H}\int_{G}\eta(g)f(gh)\Delta_{H}(h^{-1})\Delta_{G}(h)\ \mu(g)\ \nu(h).

Applying Fubini’s Theorem 1.11 again, we deduce using that ΔG|HΔH\Delta_{G}|_{H}\equiv\Delta_{H} and fH0f_{H}\equiv 0:

=Gη(g)Hf(gh)ν(h)μ(g)=Gη(g)fH(gH)=0=\int_{G}\eta(g)\int_{H}f(gh)\ \nu(h)\ \mu(g)=\int_{G}\eta(g)f_{H}(gH)=0

which completes the proof that λ\lambda is a positive linear functional. Hence, by Riesz’ Theorem 1.13, there exists a unique Radon measure ξ\xi on G/HG/H such that

G(sf)(g)μ(g)=λ(f)\displaystyle\int_{G}(sf)(g)\ \mu(g)=\lambda(f) =G/Hf(gH)ξ(gH)\displaystyle=\int_{G/H}f(gH)\ \xi(gH)
=G/H(sf)H(gH)ξ(gH)=G/HH(sf)(gh)ν(h)ξ(gH)\displaystyle=\int_{G/H}(sf)_{H}(gH)\ \xi(gH)=\int_{G/H}\int_{H}(sf)(gh)\ \nu(h)\ \xi(gH)

for all fCc(G/H)f\in C_{c}(G/H). The measure ξ\xi is checked to be non-zero on non-empty open sets as well as GG-invariant, i.e. ξ\xi is a Haar measure on G/HG/H. Since the above equation is independent of ss, we may as well start with a function fCc(G)f\in C_{c}(G), thus proving the existence of a unique Haar measure ξ\xi on G/HG/H satisfying (W). To complete the proof, we need to show that any Haar measure on G/HG/H (not necessarily satisfying (W)) is a strictly positive scalar multiple of ξ\xi: Let ξ1,ξ2\xi_{1},\xi_{2} be Haar measures on G/HG/H. Then there are left Haar measures μ1,μ2\mu_{1},\mu_{2} on GG satisfying (W) for ξ1\xi_{1} and ξ2\xi_{2} respectively (see the converse direction of the proof). By uniqueness, μ2=cμ1\mu_{2}=c\mu_{1} for some strictly positive real number cc. Then ξ2\xi_{2} and cξ1c\xi_{1} both satisfy (W) for μ2\mu_{2}. By the uniqueness proven above, ξ2=cξ1\xi_{2}=c\xi_{1}. ∎

Remark 4.4.

Retain the notation of Theorem 4.2. When GG is compact, we may choose η1\eta\equiv 1 in the proof of Lemma 4.3. The constructed left Haar functional on G/HG/H is then given by

λ:Cc(G/H),fGf(gH)1H(gH)μ(g)=1ν(H)Gf(gH)μ(g).\lambda:C_{c}(G/H)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G}\frac{f(gH)}{1_{H}(gH)}\ \mu(g)=\frac{1}{\nu(H)}\int_{G}f(gH)\ \mu(g).

Notice that ν(H)\nu(H) is finite by Proposition 2.9 given that HH is compact as a closed subset of a compact space. Now, it is a fact (see [KL06, Thm. 7.12]) that the Haar measure ξ\xi on G/HG/H associated to λ\lambda can be computed by evaluating λ\lambda on characteristic functions. Thus, when EG/HE\subseteq G/H is measurable,

ξ(E)=μ(π1(E))ν(H),in particularξ(G/H)=μ(G)ν(H).\xi(E)=\frac{\mu(\pi^{-1}(E))}{\nu(H)},\quad\text{in particular}\quad\xi(G/H)=\frac{\mu(G)}{\nu(H)}.

The auxiliary function η\eta merely mends the issues that arise when GG is not compact.

Example 4.5.

To illustrate the usefulness of Theorem 4.2, we now provide a Haar functional for G:=SL(2,)G:=\operatorname{SL}(2,\operatorname{\mathbb{R}}). Recall that GG acts transitively on the upper half plane :={zIm(z)>0}\operatorname{\mathbb{H}}:=\{z\in\operatorname{\mathbb{C}}\mid\text{Im}(z)>0\} via fractional linear transformations:

(abcd)z:=az+bcz+dand(yxy1y1)i=x+iy\begin{pmatrix}a&b\\ c&d\end{pmatrix}z:=\frac{az+b}{cz+d}\quad\text{and}\quad\begin{pmatrix}\sqrt{y}&x\sqrt{y}^{-1}\\ &\sqrt{y}^{-1}\end{pmatrix}i=x+iy

for xx\in\operatorname{\mathbb{R}} and y>0y\in\operatorname{\mathbb{R}}_{>0}. Also, one readily verifies that H:=stabG(i)=SO(2,)H:=\text{stab}_{G}(i)=\operatorname{SO}(2,\operatorname{\mathbb{R}}). Hence the maps

G/H,gHgiandG/H,x+iy(yxy1y1)G/H\to\operatorname{\mathbb{H}},\ gH\mapsto gi\quad\text{and}\quad\operatorname{\mathbb{H}}\to G/H,\ x+iy\mapsto\begin{pmatrix}\sqrt{y}&x\sqrt{y}^{-1}\\ &\sqrt{y}^{-1}\end{pmatrix}

are mutually inverse GG-isomorphisms. In fact they are homeomorphisms. Since GG is unimodular as a connected semisimple Lie group and HH is unimodular as a compact group by Proposition 3.6, we obtain a Haar measure ξ\xi on G/HG/H\cong\operatorname{\mathbb{H}} by Theorem 4.2. Let ν\nu be a left Haar measure on HH. Then

Cc(G),fG/HHf(gH)ν(h)ξ(gH)C_{c}(G)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{G/H}\int_{H}f(gH)\ \nu(h)\ \xi(gH)

is a left Haar functional on GG. To make this computable, we use the homeomorphisms HS1H\cong S^{1} and G/HG/H\cong\operatorname{\mathbb{H}} to change variables via Proposition 1.10, and the SL(2,)\operatorname{SL}(2,\operatorname{\mathbb{R}})-invariant Radon measure on \operatorname{\mathbb{H}} that stems from hyperbolic geometry. All together, the Haar functional on G=SL(2,)G=\operatorname{SL}(2,\operatorname{\mathbb{R}}) reads

f002πf((yxy1y1)(cosθsinθsinθcosθ))𝑑θdλ(y)dλ(x)y2.f\mapsto\int_{-\infty}^{\infty}\int_{0}^{\infty}\int_{0}^{2\pi}f\left(\begin{pmatrix}\sqrt{y}&x\sqrt{y}^{-1}\\ &\sqrt{y}^{-1}\end{pmatrix}\begin{pmatrix}\cos\theta&\sin\theta\\ -\sin\theta&\cos\theta\end{pmatrix}\right)\ d\theta\ \frac{d\lambda(y)\ d\lambda(x)}{y^{2}}.
Remark 4.6.

In the setting of Example 4.5, the group PP of Example 3.9 is the stabiliser in SL(2,)\operatorname{SL}(2,\operatorname{\mathbb{R}}) of the boundary point of \operatorname{\mathbb{H}} associated to the (unit-speed) geodesic γ:[0,),ti+ieit\gamma:[0,\infty)\to\operatorname{\mathbb{H}},\ t\mapsto i+ie^{it}. Basically, PP translates γ\gamma to asymptotic geodesics. More general, when MM is a symmetric space of non-compact type such as SL(n,)/SO(n)\operatorname{SL}(n,\operatorname{\mathbb{R}})/\operatorname{SO}(n), let G:=Iso(M)G:=\mathrm{Iso}(M)^{\circ}, pMp\!\in\!M and xMx\in\partial M a boundary point. Then there is a strong dichotomy between stabG(p)\mathrm{stab}_{G}(p) and stabG(x)\mathrm{stab}_{G}(x) that pertains to compactness, connnectedness, transitivity, conjugacy and unimodularity. See [Ebe96, 2.17] for details.

4.1. Discrete Subgroups

When, in the discussion above, Γ:=H\Gamma:=H is a discrete subgroup of GG and GG is second-countable, then integration over G/ΓG/\Gamma can be realized by integrating over a fundamental domain for G/ΓG/\Gamma in GG. In the following, we pick the counting measure ν\nu as the Haar measure on Γ\Gamma.

Definition 4.7.

Let GG be a locally compact Hausdorff group and let Γ\Gamma be a discrete subgroup of GG. A strict fundamental domain for G/ΓG/\Gamma in GG is a set F(G)F\in\operatorname{\mathcal{B}}(G) such that π:FG/Γ\pi:F\to G/\Gamma is a bijection. A fundamental domain for G/ΓG/\Gamma in GG is a set F(G)F\in\operatorname{\mathcal{B}}(G) which differs from a strict fundamental domain by a set of measure zero with respect to any left Haar measure on GG.

Proposition 4.8.

Let GG be a locally compact Hausdorff, second-countable group with a discrete subgroup Γ\Gamma. Then there exists a fundamental domain for G/ΓG/\Gamma in GG.

Remark 4.9.

Note that, in Proposition 4.8, second-countability of GG implies that Γ\Gamma is countable.

Proof.

(Proposition 4.8). The canonical projection π:GG/Γ\pi:G\to G/\Gamma is a local homeomorphism. In view of second-countability, this implies the existence of an open cover (Un)n(U_{n})_{n\in\operatorname{\mathbb{N}}} of GG such that π:Unπ(Un)\pi:U_{n}\to\pi(U_{n}) is a homeomorphism for every nn\in\operatorname{\mathbb{N}}. Let F1=U1F_{1}=U_{1} and define inductively Fn=Un\(Unπ1π(k<nUk))F_{n}=U_{n}\backslash(U_{n}\cap\pi^{-1}\pi(\bigcup_{k<n}U_{k})). Then F:=nFnF:=\bigcup_{n\in\operatorname{\mathbb{N}}}F_{n} is a fundamental domain for G/ΓG/\Gamma in GG. ∎

Integration over G/ΓG/\Gamma now reduces to integration over GG as follows.

Proposition 4.10.

Let GG be a locally compact Hausdorff, second-countable group with left Haar measure μ\mu and let Γ\Gamma be a discrete subgroup of GG. Assume that ΔG|ΓΔΓ\Delta_{G}|_{\Gamma}\equiv\Delta_{\Gamma}. Further, let FF be a fundamental domain for G/ΓG/\Gamma in GG. Then a Haar measure ξ\xi on G/ΓG/\Gamma satisfying (W) exists and is associated to the following functional: λ:Cc(G/Γ),fFf(gΓ)μ(g)\lambda:C_{c}(G/\Gamma)\to\operatorname{\mathbb{C}},\ f\mapsto\int_{F}f(g\Gamma)\ \mu(g), i.e.

G/Γf(gΓ)ξ(gΓ)=Ff(gΓ)μ(g)for allfCc(G/Γ).\int_{G/\Gamma}f(g\Gamma)\ \xi(g\Gamma)=\int_{F}f(g\Gamma)\ \mu(g)\quad\text{for all}\quad f\in C_{c}(G/\Gamma).
Proof.

The functional λ\lambda is positive and linear. The associated Radon measure ξ\xi on G/ΓG/\Gamma is checked to be non-zero on non-empty open sets and GG-invariant. Hence ξ\xi is a Haar measure on G/ΓG/\Gamma. To prove that it satisfies (W), note that changing FF by a set of measure zero, we may assume that FF is a strict fundamental domain. Then GG is a countable disjoint union G=γΓFγG=\bigsqcup_{\gamma\in\Gamma}F\gamma and hence

G\displaystyle\int_{G} f(g)μ(g)=γΓFγf(g)μ(g)=γΓFf(gγ)μ(g)=ΓFf(gγ)μ(g)ν(γ)\displaystyle f(g)\ \mu(g)=\sum_{\gamma\in\Gamma}\int_{F\gamma}f(g)\ \mu(g)=\sum_{\gamma\in\Gamma}\int_{F}f(g\gamma)\ \mu(g)=\int_{\Gamma}\int_{F}f(g\gamma)\ \mu(g)\ \nu(\gamma)
=FΓf(gγ)ν(γ)μ(g)=FfΓ(gΓ)μ(g)=G/ΓfΓ(gΓ)ξ(gΓ)=G/ΓΓf(gγ)ν(γ)ξ(gΓ)\displaystyle=\int_{F}\int_{\Gamma}f(g\gamma)\ \nu(\gamma)\ \mu(g)=\int_{F}f_{\Gamma}(g\Gamma)\ \mu(g)=\int_{G/\Gamma}f_{\Gamma}(g\Gamma)\ \xi(g\Gamma)=\int_{G/\Gamma}\int_{\Gamma}f(g\gamma)\ \nu(\gamma)\ \xi(g\Gamma)

for all fCc(G)f\in C_{c}(G), where the second equality follows from the assumption that ΔG|ΓΔΓ1\Delta_{G}|_{\Gamma}\equiv\Delta_{\Gamma}\equiv 1, and the application of Fubini’s Theorem 1.11 is valid since GG is σ\sigma-finite as a locally compact, second-countable space and Γ\Gamma is σ\sigma-finite because it is countable. ∎

Remark 4.11.

Retain the notation of Proposition 4.10. The assumption that ΔG|ΓΔΓ\Delta_{G}|_{\Gamma}\equiv\Delta_{\Gamma} is not automatic. For instance, the subgroup

Γ:={(etet)|t}\Gamma:=\left\{\left.\begin{pmatrix}e^{t}&\\ &e^{-t}\end{pmatrix}\right|t\in\operatorname{\mathbb{Z}}\right\}

of the group PP of Example 3.9 is isomorphic to \operatorname{\mathbb{Z}} and discrete in PP. However, for γ:=diag(et,et)\gamma:=\mathrm{diag}(e^{t},e^{-t}) we have ΔP(γ)=e2t1ΔΓ\Delta_{P}(\gamma)=e^{-2t}\neq 1\equiv\Delta_{\Gamma} by Example 3.9 whenever t0t\neq 0.

We end with a result about groups containing lattices. Recall that a lattice Γ\Gamma in a locally compact Hausdorff group GG is discrete subgroup such that G/ΓG/\Gamma supports a finite Haar measure.

Proposition 4.12.

A locally compact Hausdorff group GG containing a lattice Γ\Gamma is unimodular.

Proof.

Suppose Γ\Gamma is a lattice in GG. Since G/ΓG/\Gamma supports a finite Haar measure ξ\xi, Theorem 4.2 implies that ΔG|ΓΔΓ1\Delta_{G}|_{\Gamma}\equiv\Delta_{\Gamma}\equiv 1 and hence kerΔGΓ\ker\Delta_{G}\supseteq\Gamma. Therefore, ΔG\Delta_{G} factors through GG/ΓG\to G/\Gamma via Δ~G:G/Γ(0,)\smash{\widetilde{\Delta}_{G}:G/\Gamma\to(\operatorname{\mathbb{R}}_{\geq 0}^{\ast},\cdot)}. Then (Δ~G)ξ\smash{(\widetilde{\Delta}_{G})_{\ast}\xi} is a non-zero, finite measure on 0\operatorname{\mathbb{R}}_{\geq 0}^{\ast} which is invariant under the image of ΔG\Delta_{G}. This forces ΔG1\Delta_{G}\equiv 1. ∎

References

  • [Bou04] N. Bourbaki, Integration II: Chapters 7-9, vol. 2, Springer, 2004.
  • [Ebe96] P. Eberlein, Geometry of Nonpositively Curved Manifolds, University of Chicago Press, 1996.
  • [Haa33] A. Haar, Der Massbegriff in der Theorie der kontinuierlichen Gruppen, Annals of Mathematics (1933), 147–169.
  • [KL06] A. Knightly and C. Li, Traces of Hecke operators, vol. 133, American Mathematical Society, 2006.
  • [Kna02] A. W. Knapp, Lie Groups Beyond an Introduction, vol. 140, Birkhäuser, 2002.
  • [Nac76] L. Nachbin, The Haar integral, RE Krieger Pub. Co., 1976.
  • [Pal01] T. W. Palmer, Banach Algebras and the General Theory of *-Algebras, vol. 2, Cambridge University Press, 2001.
  • [Ser80] J. P. Serre, Trees, Springer, 1980.
  • [vD31] D. van Dantzig, Studien over topologische algebra, University of Groningen, 1931.
  • [War83] F. W. Warner, Foundations of Differentiable Manifolds and Lie Groups, vol. 94, Springer, 1983.
  • [Wei65] A. Weil, L’intégration dans les groupes topologiques et ses applications, vol. 1145, Hermann, 1965.
  • [Zim90] R. J. Zimmer, Essential Results of Functional Analysis, University of Chicago Press, 1990.