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

Exceptional families of measures on Carnot groups

Bruno Franchi and Irina Markina Bruno Franchi. Department of Mathematics, University of Bologna, Piazza di Porta S. Donato, 5 40126 Bologna, Italy [email protected] Irina Markina. Department of Mathematics, University of Bergen, P.O. Box 7803, Bergen N-5020, Norway [email protected]
Abstract.

We study the families of measures on Carnot groups that have vanishing pp-module, which we call pp-exceptional families. We found necessary and sufficient condition for the family of intrinsic Lipschitz surfaces passing through a common point to be pp-exceptional for p1p\geq 1. We described a wide class of pp-exceptional intrinsic Lipschitz surfaces for p(0,)p\in(0,\infty).

Key words and phrases:
Nilpotent Lie group, module of families of measures, Hausdorff measure, intrinsic Lipschitz graph
2010 Mathematics Subject Classification:
Primary 28A78, 53C17; Secondary 31B15, 22E30
The work of the second author was partially supported by the project Pure Mathematics in Norway, funded by Trond Mohn Foundation and Tromsø Research Foundation.

1. Introduction and motivation

“Negligible” sets appear customarily in measure theory and stochastic theory, as the sets of measure zero, as well as the sets of vanishing pp-capacity when dealing with regularity issues for solutions of PDE, and as thin and polar sets in potential theory. Sets of a family of measures having the so-called pp-module zero belong to this category of negligible or exceptional subsets of families of measures, see definitions in Section 3.

The notion of a module of a family of curves or in another terminology extremal length originated in the theory of complex analytic functions as a conformal invariant [AB50], and later was widely used for the quasiconformal analysis and extremal problems of functional spaces [AO99, BFP11, Oht03, Ric93, Str76, Š60]. B. Fuglede in his seminal paper [Fug57] proposed to extend the notion of the module from families of curves to families of measures. He characterized the completion, with respect to LpL^{p} norm, of some functional classes by using a family of surfaces having vanishing pp-module. He also described some classes of systems of measures with vanishing module and related it to the potential theory. In spite of the fact that the definition of the module of a family of measures is given for an arbitrary measure space, most of the applications in [Fug57] were done for n\mathbb{R}^{n}.

The development of the analysis on metric measure spaces inspired us to look for examples of interesting systems of measures in a more general setting then the Euclidean space. We are not the first ones, just to name [BFP13, Bjo02, Mar04, Sha00, Sha01]. However, most of the preceding works were dealing with families of curves. Our main interest focuses on families of (suitably defined) intrinsic surfaces on Carnot groups [FSSC03b, FS16, Vit12].

Carnot groups are connected, simply connected, nilpotent Lie groups and are one of the most popular examples of metric measure spaces. Being endowed with a rich structure of translations and dilations, makes the Carnot groups akin to Euclidean spaces. Euclidean spaces are commutative Carnot groups, and, more precisely, the only commutative Carnot groups. The simplest but, at the same time, non-trivial instance of non-abelian Carnot groups is provided by Heisenberg groups HnH^{n}.

Carnot groups possess an intrinsic metric, the so-called Carnot-Carathéodory metric (cccc-distance), see for instance, [BLU07, FSSC03a, Gro96]. It is also well known that non commutative Carnot groups, endowed with the cccc-distance, are not Riemannian manifolds because the cccc-distance makes them not locally Lipschitz equivalent to Riemannian at any scale [Sem96]. The Carnot groups are particular instances of the so-called sub-Riemannian manifolds.

Though Carnot groups are analytic manifolds, the study of measures supported on submanifolds (for instance the Hausdorff measures associated with their cccc-distance) cannot be reduced to the well established theory for submanifolds of Euclidean spaces, since it has been clear for a long time that considering Euclidean regular submanifolds, even in Heisenberg groups, may be both too general and too restrictive, see [KSC04] for a striking example related to the second instance. Through this paper, we shall rely on the theory of intrinsic submanifolds in Carnot groups that has been recently developed by making use of the notion of intrinsic graphs, see e.g. [FSSC07, FS16, FSSC03b]. A discussion of different alternatives leading to this notion can be found e.g. in [FS16], together with the main properties of the most relevant instances, the so-called intrinsic Lipschitz graphs. Let us sketch this construction, restricting ourselves to stress the difficulties arising when we want to extend the theory of pp-modules from the Euclidean setting to Carnot groups. For deep algebraic reasons, due to the non-commutativity of the group, the most flexible notion of submanifold of a Carnot group is the counterpart of the Euclidean notion of graph. However, the notion of intrinsic graph is not a straightforward translation of the corresponding Euclidean notion, since Carnot groups not always can be expressed as a direct product of subgroups. Because of that, we argue as follows: an intrinsic graph inside 𝔾\mathbb{G} is associated with a decomposition of 𝔾\mathbb{G} as a product 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} of two homogeneous complementary subgroups 𝕄\mathbb{M}, \mathbb{H}, see Section 3.4.1. Then the intrinsic (left) graph of f:Ωf\colon\Omega\to\mathbb{H}, where Ω\Omega is an open subset of 𝕄\mathbb{M}, is the set

graph(f)={gf(g):gΩ}.\mathrm{graph}\,(f)=\{g\cdot f(g):g\in\Omega\}.

Another deep peculiarity of Carnot groups is the poor structure of the isometry group preserving the grading structure of the their Lie algebras.

The main results of the present work are formulated in Theorem 7, Section 3.5, where we show that quite a wide class of families of intrinsic Lipschitz surfaces, (sets which are locally intrinsic Lipschitz graphs of the same “metric dimension”) has vanishing pp-module for p(0,1)p\in(0,1). We did not reach the full generality as in the Euclidean space due to the lack of knowledge about decompositions of an arbitrary Carnot groups into the product of two homogeneous subgroups. Another result contained in Theorem 8, Section 3.5.2, which is the sufficient condition for a family of surfaces passing through a common point to be pp-exceptional. In order to find a necessary condition we construct a family of intrinsic Lipschitz graphs passing through one point by making use of the orthogonal Grassmannians on some specific 2-step Carnot groups, see Section 4. The construction of the orthogonal Grassmannians and the study of measures on them have an independent interest and, as to the knowledge of the authors, were not presented in the literature. Examples of exceptional families of measures that are not related to intrinsic Lipschitz graphs are contained in Examples 1 and 2 in Section 3.3.

We are trying to keep the paper as accessible as possible for a wider audience. The structure of the paper is visible from the Contents.

2. Carnot groups

In the present section we establish notation and collect the basic notions concerning Carnot groups and their Lie algebras.

2.1. General definition of Carnot groups

A Carnot group 𝔾\mathbb{G} is a connected, simply connected Lie group whose Lie algebra 𝔤\mathfrak{g} of the left-invariant vector fields is a graded stratified nilpotent Lie algebra of step ll, i.e. the Lie algebra 𝔤\mathfrak{g} satisfies:

𝔤=k=1l𝔤k,[𝔤1,𝔤k]=𝔤k+1,𝔤l+1={0}.\mathfrak{g}=\oplus_{k=1}^{l}\mathfrak{g}_{k},\quad[\mathfrak{g}_{1},\mathfrak{g}_{k}]=\mathfrak{g}_{k+1},\quad\mathfrak{g}_{l+1}=\{0\}.

We denote by N=k=1ldim(𝔤k)N=\sum_{k=1}^{l}\,\dim(\mathfrak{g}_{k}) the topological dimension of 𝔾\mathbb{G}. The number Q=k=1lkdim(𝔤k)Q=\sum_{k=1}^{l}\,k\dim(\mathfrak{g}_{k}) is called the homogeneous dimension of the group 𝔾\mathbb{G}. Since 𝔤\mathfrak{g} is nilpotent, the exponential map exp:𝔤𝔾\exp\colon\mathfrak{g}\to\mathbb{G} is a global diffeomorphism.

One can identify the group 𝔾\mathbb{G} with N𝔤\mathbb{R}^{N}\cong\mathfrak{g} by making use of exponential coordinates of the first kind by the following procedure. We fix a basis

(1) X11,,X1𝐝1,X21,,X2𝐝2,,Xl1,,Xl𝐝l,𝐝k=dim(𝔤k),X_{11},\ldots,X_{1\mathbf{d}_{1}},X_{21},\ldots,X_{2\mathbf{d}_{2}},\ldots,X_{l1},\ldots,X_{l\mathbf{d}_{l}},\qquad\mathbf{d}_{k}=\dim(\mathfrak{g}_{k}),

of the Lie algebra 𝔤\mathfrak{g} which is adapted to the stratification. If g𝔾g\in\mathbb{G} and V𝔤V\in\mathfrak{g} are such that

g=exp(V)=exp(k=1lj=1𝐝kxkjXkj),g=\exp(V)=\exp\Big{(}\sum_{k=1}^{l}\sum_{j=1}^{\mathbf{d}_{k}}x_{kj}X_{kj}\Big{)},

then (with a slight abuse of notations) we associate with the point g𝔾g\in\mathbb{G} a point xNx\in\mathbb{R}^{N} having the following coordinates

(2) g=(x11,,x1𝐝1,x21,,x2𝐝2,,xl1,,xl𝐝l)=x.g=(x_{11},\ldots,x_{1\mathbf{d}_{1}},x_{21},\ldots,x_{2\mathbf{d}_{2}},\ldots,x_{l1},\ldots,x_{l\mathbf{d}_{l}})=x.

Thus, the identity e𝔾e\in\mathbb{G} is identified with the origin in N\mathbb{R}^{N} and the inverse g1g^{-1} with x-x.

The stratification 𝔤=k=1k=l𝔤k\mathfrak{g}=\oplus_{k=1}^{k=l}\mathfrak{g}_{k} of 𝔤\mathfrak{g} induces the one-parameter family {δλ}λ>0\{\delta_{\lambda}\}_{\lambda>0} of automorphisms of 𝔤\mathfrak{g}, where each δλ:𝔤𝔤\delta_{\lambda}\colon\mathfrak{g}\to\mathfrak{g} is defined as

δλ(X):=λkX,for allX𝔤kandλ>0.\delta_{\lambda}(X):=\lambda^{k}X,\quad\text{for all}\quad X\in\mathfrak{g}_{k}\quad\text{and}\quad\lambda>0.

The exponential map allows to transfer these automorphisms of 𝔤\mathfrak{g} to a family of automorphisms of the Lie group 𝔾\mathbb{G}: δλ𝔾:𝔾𝔾\delta_{\lambda}^{\mathbb{G}}:\mathbb{G}\to\mathbb{G}, so-called intrinsic dilations, defined as

δλ𝔾:=expδλexp1,for allλ>0.\delta_{\lambda}^{\mathbb{G}}:=\mathrm{exp}\circ\delta_{\lambda}\circ\mathrm{exp}^{-1},\quad\text{for all}\quad\lambda>0.

We keep denoting by δλ:𝔾𝔾\delta_{\lambda}:\mathbb{G}\to\mathbb{G} the intrinsic dilations, if no confusion arises.

In exponential coordinates, the group automorphism δλ:𝔾𝔾\delta_{\lambda}\colon\mathbb{G}\to\mathbb{G} for λ>0\lambda>0 is written as

(3) δλg\displaystyle\delta_{\lambda}g =\displaystyle= δλ(x11,,x1𝐝1,x21,,x2𝐝2,,xl1,,xl𝐝l)\displaystyle\delta_{\lambda}(x_{11},\ldots,x_{1\mathbf{d}_{1}},x_{21},\ldots,x_{2\mathbf{d}_{2}},\ldots,x_{l1},\ldots,x_{l\mathbf{d}_{l}})
=\displaystyle= (λx11,,λx1𝐝1,λ2x21,,λ2x2𝐝2,,λlxl1,,λlxl𝐝l).\displaystyle(\lambda x_{11},\ldots,\lambda x_{1\mathbf{d}_{1}},\lambda^{2}x_{21},\ldots,\lambda^{2}x_{2\mathbf{d}_{2}},\ldots,\lambda^{l}x_{l1},\ldots,\lambda^{l}x_{l\mathbf{d}_{l}}).

The group product on 𝔾\mathbb{G} written in coordinates (2) has the form

(4) xy=x+y+𝒬(x,y),for allx,yN,x\cdot y=x+y+\mathcal{Q}(x,y),\quad\text{for all}\quad x,y\in\mathbb{R}^{N},

where 𝒬=(𝒬1,,𝒬N):N×NN\mathcal{Q}=(\mathcal{Q}_{1},\dots,\mathcal{Q}_{N}):\mathbb{R}^{N}\times\mathbb{R}^{N}\to\mathbb{R}^{N} and each 𝒬k\mathcal{Q}_{k} is a homogeneous polynomial with respect to group dilations, see, for instance [FSSC03a, Propositions 2.1 and 2.2]. When the grading structure is not important, we will use the one-index notation

X1,,XN,x1,,xNX_{1},\ldots,X_{N},\qquad x_{1},\ldots,x_{N}

for the basis (1) of the Lie algebra 𝔤\mathfrak{g} and for the coordinates on 𝔾\mathbb{G}. The basis vectors of the Lie algebra 𝔤\mathfrak{g} viewed as left invariant vector fields XjX_{j}, j=1,,Nj=1,\ldots,N on 𝔾\mathbb{G} have polynomial coefficients and take the form in the coordinate frame:

(5) Xj=j+i>jNqi,j(x)i,forj=1,,N,X_{j}=\partial_{j}+\sum_{i>j}^{N}q_{i,j}(x)\partial_{i},\quad\text{for}\;j=1,\dots,N,

where qi,j(x)=𝒬iyj(x,y)|y=0q_{i,j}(x)=\frac{\partial\mathcal{Q}_{i}}{\partial y_{j}}(x,y){|_{y=0}}. The vector fields XjX_{j}, j=1,,𝐝1j=1,\ldots,\mathbf{d}_{1} are called horizontal and they are homogeneous of degree 11 with respect to the group dilation. Their span at q𝔾q\in\mathbb{G} is called the horizontal vector space Hq𝔾Tq𝔾H_{q}\mathbb{G}\subset T_{q}\mathbb{G}.

2.1.1. Distance functions

In the present paper we will use the following distance functions on a Carnot group 𝔾\mathbb{G} identified with N\mathbb{R}^{N} through exponential coordinates.

  • (D1D_{1})

    The standard Euclidean distance dEd_{E} associated with the Euclidean norm |x|E|x|_{E}: dE(x,y)=i=1N(xiyi)2d_{E}(x,y)=\sqrt{\sum_{i=1}^{N}(x_{i}-y_{i})^{2}}.

However, such a distance is neither left-invariant under group translations, nor 1-homogeneous with respect to group dilations. Thus, let us introduce further (not Lipschitz equivalent to dEd_{E}) distances enjoying these properties.

Definition 1.

Let 𝔾\mathbb{G} be a Carnot group. A homogeneous norm \|\cdot\| is a continuous function :𝔾[0,+)\|\cdot\|:\,\mathbb{G}\to[0,+\infty) such that

(6) p=0if and only ifp=0;p1=p,δλ(p)=λpfor allp𝔾andλ>0;pqp+qfor allp,q𝔾.\begin{split}&\|p\|=0\quad\text{if and only if}\quad p=0\,;\\ &\|p^{-1}\|=\|p\|,\quad\|\delta_{\lambda}(p)\|=\lambda\|p\|\quad\text{for all}\quad p\in\mathbb{G}\quad\text{and}\quad\lambda>0;\\ &\|{p\cdot q}\|\leq\,\|p\|+\,\|q\|\qquad\text{for all}\quad p,q\in\mathbb{G}.\end{split}
Remark 1.

A homogeneous norm \|\cdot\| induces a homogeneous left invariant distance in 𝔾\mathbb{G} as follows:

(7) d(p,q):=d(q1p,0):=q1pfor allp,q𝔾.d(p,q):=d(q^{-1}\cdot p,0):=\|q^{-1}\cdot p\|\quad\text{for all}\quad p,q\in\mathbb{G}.
  • (D2D_{2})

    On any Carnot group 𝔾\mathbb{G} there exists a homogeneous norm 𝔾\|\cdot\|_{\mathbb{G}} that is smooth away of the origin and induces a distance d𝔾(x,y):=y1x𝔾d_{\mathbb{G}}(x,y):=\|y^{-1}\cdot x\|_{\mathbb{G}} on 𝔾\mathbb{G}, see [Ste93, Page 638].

  • (D3D_{3})

    We also use the homogeneous norm H\|\cdot\|_{H}:

    xH=max{ϵ1𝐱1E,ϵ2𝐱2E1/2,,ϵl𝐱lE1/l},\|x\|_{H}=\max\{\epsilon_{1}\|\mathbf{x}_{1}\|_{E},\epsilon_{2}\|\mathbf{x}_{2}\|^{1/2}_{E},\ldots,\epsilon_{l}\|\mathbf{x}_{l}\|^{1/l}_{E}\},

    where 𝐱k=(xk1,,xk𝐝k)𝔤k\mathbf{x}_{k}=(x_{k1},\ldots,x_{k\mathbf{d}_{k}})\in\mathfrak{g}_{k}, E\|\mathbf{\cdot}\|_{E} is the Euclidean norm, making the adapted basis (1) orthonormal. The suitable constants ϵ1,,ϵl\epsilon_{1},\dots,\epsilon_{l} are positive, see [FSSC03a, Theorem 5.1]. The induced distance is dH(x,y):=y1xHd_{H}(x,y):=\|y^{-1}\cdot x\|_{H};

  • (D4D_{4})

    The Carnot-Carathéodory distance dcc(x,y)d_{cc}(x,y) which is induced by the Euclidean scalar product .,.𝔤1\langle.\,,.\rangle_{\mathfrak{g}_{1}} on 𝔤1\mathfrak{g}_{1}, making the horizontal vector fields XjX_{j}, j=1,,𝐝1j=1,\ldots,\mathbf{d}_{1} orthonormal [ABB20, Gro96].

The distances defined in (D2)(D4)(D_{2})-(D_{4}) are Lipschitz equivalent, since they are invariant under the left translation on 𝔾\mathbb{G} and are homogeneous functions of degree 1 with respect to dilation (3). We denote by dρd_{\rho} any of the distances mentioned in (D2)(D4)(D_{2})-(D_{4}). Then the above observation and [BLU07, Corollaries 5.15.1 and 5.15.2] imply:

Proposition 1.

Let 𝔾\mathbb{G} be a Carnot group of step ll. Then

  • (i)

    a set A𝔾A\subset\mathbb{G} is dρd_{\rho}-bounded if and only if it is dEd_{E}-bounded;

  • (ii)

    for any bounded set A𝔾A\subset\mathbb{G} there is CA>0C_{A}>0 such that

    CA1dE(x,y)dρ(x,y)CAdE(x,y)1/lC_{A}^{-1}\,d_{E}(x,y)\leq\,d_{\rho}(x,y)\leq\,C_{A}\,d_{E}(x,y)^{1/l}

    for all x,yAx,y\in A;

  • (iii)

    the topologies induced by dρd_{\rho} and dEd_{E} coincide.

2.1.2. Measures on the Carnot groups

The pushforward of NN-dimensional Lebesgue measure N\mathcal{L}^{N} on n\mathbb{R}^{n} to the group 𝔾\mathbb{G} under the exponential map is the Haar measure 𝐠𝔾\mathbf{g}_{\mathbb{G}} on the group 𝔾\mathbb{G}. Hence if EN𝔾E\subset\mathbb{R}^{N}\cong\mathbb{G} is measurable, then N(xE)=N(Ex)=N(E)\mathcal{L}^{N}(x\cdot E)=\mathcal{L}^{N}(E\cdot x)=\mathcal{L}^{N}(E) for all x𝔾x\in\mathbb{G}. Moreover, if λ>0\lambda>0 then N(δλ(E))=λQN(E)\mathcal{L}^{N}(\delta_{\lambda}(E))=\lambda^{Q}\mathcal{L}^{N}(E).

Recall the definition of the Hausdorff measure in a metric space (X,ρ)(X,\rho). For all sets E,EiXE,E_{i}\subseteq X, closed balls Bρ(xi,ri)B_{\rho}(x_{i},r_{i}), real numbers m[0,)m\in[0,\infty), and δ>0\delta>0 one writes

ρ,δm(E)\displaystyle\mathcal{H}_{\rho,\delta}^{m}(E) :=inf{idiam(Ei)m:EiEi,diam(Ei)δ},\displaystyle:=\inf\Big{\{}\sum_{i}\text{\rm diam}(E_{i})^{m}:\ E\subset\cup_{i}E_{i},\,\text{\rm diam}(E_{i})\leq\delta\Big{\}},
𝒮ρ,δm(E)\displaystyle\mathcal{S}_{\rho,\delta}^{m}(E) :=inf{idiam(Ei)m:EiBρ(xi,ri),diam(Bρ(xi,ri))δ},\displaystyle:=\inf\Big{\{}\sum_{i}\text{\rm diam}(E_{i})^{m}:\ E\subset\cup_{i}B_{\rho}(x_{i},r_{i}),\text{\rm diam}(B_{\rho}(x_{i},r_{i}))\leq\delta\Big{\}},

where we assume diam(Ei)0=1\text{\rm diam}(E_{i})^{0}=1 for EiE_{i}\neq\emptyset, and ρ,δm()=𝒮ρ,δm()=0\mathcal{H}_{\rho,\delta}^{m}(\emptyset)=\mathcal{S}_{\rho,\delta}^{m}(\emptyset)=0. For all EXE\subseteq X and m[0,)m\in[0,\infty) the mm-Hausdorff measure ρm(E)\mathcal{H}_{\rho}^{m}(E) and the spherical mm-Hausdorff measure 𝒮ρm(E)\mathcal{S}_{\rho}^{m}(E) are defined respectively as

ρm(E)=limδ0ρ,δm(E),𝒮ρm(E)=limδ0𝒮ρ,δm(E).\mathcal{H}_{\rho}^{m}(E)=\lim_{\delta\to 0}\mathcal{H}_{\rho,\delta}^{m}(E),\qquad\mathcal{S}_{\rho}^{m}(E)=\lim_{\delta\to 0}\mathcal{S}_{\rho,\delta}^{m}(E).

Both ρm\mathcal{H}_{\rho}^{m} and 𝒮ρm\mathcal{S}_{\rho}^{m} are Borel regular measures.

When 𝔾\mathbb{G} is a Carnot group considered as a metric space (𝔾,dρ)(\mathbb{G},d_{\rho}), where dρd_{\rho} is one of the distances (D1)(D4)(D_{1})-(D_{4}), we denote by dρ𝐝𝐦\mathcal{H}_{d_{\rho}}^{\bf d_{m}} and 𝒮dρ𝐝𝐦\mathcal{S}_{d_{\rho}}^{\bf d_{m}} the 𝐝𝐦{\bf d_{m}}-dimensional Hausdorff and spherical Hausdorff measures associated with the distance dρd_{\rho}, respectively. The measures dρ𝐝𝐦\mathcal{H}_{d_{\rho}}^{\bf d_{m}} and 𝒮dρ𝐝𝐦\mathcal{S}_{d_{\rho^{\prime}}}^{\bf d_{m}}, where dρd_{\rho} and dρd_{\rho^{\prime}} are the distance functions of types (D2)(D4)(D_{2})-(D_{4}), satisfy

cdρ𝐝𝐦(E)dρ𝐝𝐦(E)Cdρ𝐝𝐦(E),kdρ𝐝𝐦(E)𝒮dρ𝐝𝐦(E)Kdρ𝐝𝐦(E),c\mathcal{H}_{d_{\rho}}^{\bf d_{m}}(E)\leq\mathcal{H}_{d_{\rho^{\prime}}}^{\bf d_{m}}(E)\leq C\mathcal{H}_{d_{\rho}}^{\bf d_{m}}(E),\qquad k\mathcal{H}_{d_{\rho}}^{\bf d_{m}}(E)\leq\mathcal{S}_{d_{\rho^{\prime}}}^{\bf d_{m}}(E)\leq K\mathcal{H}_{d_{\rho}}^{\bf d_{m}}(E),

for some positive constants c,k,C,Kc,k,C,K and a set E𝔾E\subset\mathbb{G}. The same is true if we interchange dρ𝐝𝐦\mathcal{H}_{d_{\rho}}^{\bf d_{m}} and 𝒮dρ𝐝𝐦\mathcal{S}_{d_{\rho}}^{\bf d_{m}}. Finally

(8) c~dρQ(E)N(E)C~dρQ(E),E𝔾,\tilde{c}\mathcal{H}_{d_{\rho}}^{Q}(E)\leq\mathcal{L}^{N}(E)\leq\tilde{C}\mathcal{H}_{d_{\rho}}^{Q}(E),\quad E\subset\mathbb{G},

for some positive constants c~,C~\tilde{c},\tilde{C}, the homogeneous dimension QQ, and the topological dimension NN of the Carnot group.

2.1.3. HH-type Lie groups

One of the core examples for the present paper will be HH-type Lie groups, that are particular examples of 2-step Carnot groups. Consider a real Lie algebra (𝔤,[.,.],.,.)(\mathfrak{g},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}) with the underlying vector space 𝔤=𝔤1𝔤2\mathfrak{g}=\mathfrak{g}_{1}\oplus\mathfrak{g}_{2}, where the decomposition is orthogonal with respect to the inner product .,.\langle.\,,.\rangle_{\mathbb{R}}, and 𝔤2\mathfrak{g}_{2} is the center of the Lie algebra 𝔤\mathfrak{g}. The inner product space (𝔤2,.,.)(\mathfrak{g}_{2},\langle.\,,.\rangle_{\mathbb{R}}), where .,.\langle.\,,.\rangle_{\mathbb{R}} is the restriction of the scalar product to the subspace 𝔤2𝔤\mathfrak{g}_{2}\subset\mathfrak{g}, generates the Clifford algebra Cl(𝔤2,.,.)\text{\rm Cl}(\mathfrak{g}_{2},\langle.\,,.\rangle_{\mathbb{R}}). The Clifford algebra Cl(𝔤2,.,.)\text{\rm Cl}(\mathfrak{g}_{2},\langle.\,,.\rangle_{\mathbb{R}}) admits a representation on the vector space 𝔤1\mathfrak{g}_{1}:

J:Cl(𝔤2,.,.)End(𝔤1).J\colon\text{\rm Cl}(\mathfrak{g}_{2},\langle.\,,.\rangle_{\mathbb{R}})\to\text{\rm End}(\mathfrak{g}_{1}).

We use the notation JzJ_{z}, z𝔤2z\in\mathfrak{g}_{2}, for the value of the map JJ restricted to the vector space 𝔤2Cl(𝔤2,.,.)\mathfrak{g}_{2}\subset\text{\rm Cl}(\mathfrak{g}_{2},\langle.\,,.\rangle_{\mathbb{R}}). From the definition of the Clifford algebra we have

(9) Jz2=z,zId𝔤1,z𝔤2.J_{z}^{2}=-\langle z,z\rangle_{\mathbb{R}}\text{\rm Id}_{\mathfrak{g}_{1}},\quad z\in\mathfrak{g}_{2}.

The Lie algebra (𝔤,[.,.],.,.)(\mathfrak{g},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}) is called of HH-type if

(10) Jzu,v=z,[u,v],z𝔤2,u,v𝔤1.\langle J_{z}u,v\rangle_{\mathbb{R}}=\langle z,[u,v]\rangle_{\mathbb{R}},\quad z\in\mathfrak{g}_{2},\ \ u,v\in\mathfrak{g}_{1}.

An HH-type Lie group is a connected simply connected Lie group whose Lie algebra is an HH-type Lie algebra (𝔤,[.,.],.,.)(\mathfrak{g},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}).

2.1.4. The Heisenberg group

The nn-th Heisenberg group HnH^{n} is diffeomorphic to 2n+1\mathbb{R}^{2n+1} and is the simplest example of HH-type Lie group. Its (2n+1)(2n+1)-dimensional Lie algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{R}} has one dimensional center 𝔥2\mathfrak{h}_{2}. Let 𝔥1\mathfrak{h}_{1} be the 2n2n-dimensional orthogonal complement to 𝔥2\mathfrak{h}_{2} with respect to an inner product .,.\langle.\,,.\rangle on 𝔥n\mathfrak{h}^{n}_{\mathbb{R}}. We choose an orthonormal basis

(11) X1,,Xn,Y1,,Ynfor𝔥1andϵfor𝔥2,X_{1},\ldots,X_{n},Y_{1},\ldots,Y_{n}\ \ \text{for}\ \ \mathfrak{h}_{1}\ \ \text{and}\ \ \epsilon\ \ \text{for}\ \ \mathfrak{h}_{2},

satisfying the commutation relations

(12) [Xj,Yi]=δjiϵ,[Xj,Xi]=[Yj,Yi]=0.[X_{j},Y_{i}]=\delta_{ji}\epsilon,\quad[X_{j},X_{i}]=[Y_{j},Y_{i}]=0.

Then the map JϵJ_{\epsilon} defined in (10) satisfies

Jϵ2=Id𝔥1,Jϵ(Xi)=Yi,Jϵ(Yi)=Xi.J_{\epsilon}^{2}=-\text{\rm Id}_{\mathfrak{h}_{1}},\quad J_{\epsilon}(X_{i})=Y_{i},\quad J_{\epsilon}(Y_{i})=-X_{i}.

3. Module of a family of measures

We start from the explaining the notion of a pp-module of a system of measures, that B. Fuglede introduced in his celebrated paper [Fug57]. Let (X,𝔐,m)(X,\mathfrak{M},m) be an abstract measure space with a fixed basic measure m:𝔐[0,+]m\colon\mathfrak{M}\to[0,+\infty] defined on a σ\sigma-algebra 𝔐\mathfrak{M} of subsets of XX. We denote by 𝐌\mathbf{M} the system of all measures on XX, whose domains of definition contain 𝔐\mathfrak{M}.

With an arbitrary subset 𝐄\bf E of the system of measures 𝐌\mathbf{M} we associate a class of functions that we call admissible for 𝐄\bf E and denote by Adm(𝐄)\text{\rm Adm}(\bf E). Namely,

Adm(𝐄)\displaystyle\text{\rm Adm}({\bf E}) =\displaystyle= {f:X:fismmeasurable,f0,and\displaystyle\Big{\{}f\colon X\to\mathbb{R}:\ f\ \text{is}\ m-\text{measurable},\ \ f\geq 0,\ \text{and}
Xfdμ1,for allμ𝐄}.\displaystyle\int_{X}f\,d\mu\geq 1,\ \text{for all}\ \mu\in{\bf E}\Big{\}}.
Definition 2.

For 0<p<0<p<\infty, the module Mp(𝐄)M_{p}({\bf E}) of a system of measures 𝐄{\bf E} is defined as

Mp(𝐄)=inffAdm(𝐄)Xfp𝑑m,M_{p}({\bf E})=\inf\limits_{f\in\text{\rm Adm}({\bf E})}\int_{X}f^{p}\,dm,

interpreted as ++\infty if Adm(𝐄)=\text{\rm Adm}({\bf E})=\emptyset.

The reader can find the fundamental properties of the pp-module of measures in [Fug57, Chapter 1].

A system of measures 𝐄𝐌{\bf E}\subset\mathbf{M} can be associated with the set where the measures are supported [Oht03, V7̈1].

(I) Consider, for instance, a family of rectifiable curves Γ={γ:[aγ,bγ]n}\Gamma=\{\gamma\colon[a_{\gamma},b_{\gamma}]\to\mathbb{R}^{n}\} and the associated system of measures

𝐄={Var(dγdt)=dE1(|γ|):γΓ}.{\mathbf{E}}=\{{\rm Var}\Big{(}\frac{d\gamma}{dt}\Big{)}=\mathcal{H}^{1}_{d_{E}}(|\gamma|):\ \ \gamma\in\Gamma\}.

Here Var(dγdt){\rm Var}\Big{(}\frac{d\gamma}{dt}\Big{)} is the total variation of the vector valued Radon measure dγdt\frac{d\gamma}{dt}, which coincides with the Hausdorff measure dE1(|γ|)\mathcal{H}^{1}_{d_{E}}(|\gamma|) of the locus |γ||\gamma| of the curve γΓ\gamma\in\Gamma.

If we consider a subfamily Γ~={γ~:[a~γ,b~γ]n}Γ\mathaccent 869{\Gamma}=\{\tilde{\gamma}\colon[\tilde{a}_{\gamma},\tilde{b}_{\gamma}]\to\mathbb{R}^{n}\}\subset\Gamma of absolutely continuous curves, then the corresponding measures can be calculated by

Var(dγ~dt)=aγ~bγ~|dγ~(t)dt|𝑑t=|γ~|𝑑dE1,γ~Γ~.{\rm Var}\Big{(}\frac{d\tilde{\gamma}}{dt}\Big{)}=\int_{a_{\tilde{\gamma}}}^{b_{\tilde{\gamma}}}\Big{|}\frac{d\tilde{\gamma}(t)}{dt}\Big{|}\,dt=\int_{|\tilde{\gamma}|}d\mathcal{H}^{1}_{d_{E}},\quad\tilde{\gamma}\in\mathaccent 869{\Gamma}.

(II) A family of locally Lipschitz kk-dimensional surfaces in n\mathbb{R}^{n}. Each surface locally is the image of an open set of k\mathbb{R}^{k} under a Lipschitz map ff. In this case a surface measure locally coincides with dσ=|J(f,t)|dtd\sigma=|J(f,t)|dt. Here J(f,t)J(f,t) is the Jacobian of ff for the points tkt\in\mathbb{R}^{k}, where the Jacobian J(f,t)J(f,t) is defined. The corresponding family of measures is supported on these Lipschitz kk-dimensional surfaces.

(III) A family of countable dEk\mathcal{H}^{k}_{d_{E}}-rectifiable subsets in n\mathbb{R}^{n}, where we understand the rectifiability in the sense of [Fed69]. The kk-dimensional Hausdorff measures dEk\mathcal{H}^{k}_{d_{E}}, is the system of measures, associated with the family of countable dEk\mathcal{H}^{k}_{d_{E}}-rectifiable sets.

3.1. Exceptional families of measures

A system 𝐄0𝐌\mathbf{E}_{0}\subset\mathbf{M} is called pp-exceptional, if Mp(𝐄0)=0M_{p}(\mathbf{E}_{0})=0. A statement concerning measures μ𝐌\mu\in\mathbf{M} is said to hold MpM_{p}-almost everywhere if it fails to hold for a pp-exceptional system 𝐄0\mathbf{E}_{0}. The question that we are interested in is to study pp-exceptional sets of measures on Carnot groups. Let us remind that a point-set 𝐄0X\mathbf{E}_{0}\subset X in a measure space (X,m)(X,m) has vanishing measure: m(𝐄0)=0m(\mathbf{E}_{0})=0 if and only if there is a function fLp(X,m)f\in L^{p}(X,m) such that f(x)=+f(x)=+\infty for all x𝐄0x\in\mathbf{E}_{0}. A generalisation of this fact to a system of measures 𝐄0𝐌\mathbf{E}_{0}\subset\mathbf{M} is given by B. Fuglede.

Theorem 1.

[Fug57, Theorem 2] A system of measures 𝐄0𝐌{\bf E}_{0}\subset\mathbf{M} is pp-excep-tional if and only if there exists a non-negative function fLp(X,m)f\in L^{p}(X,m) such that

Xf𝑑μ=+for everyμ𝐄0.\int_{X}f\,d\mu=+\infty\quad\text{for every}\quad\mu\in{\bf E}_{0}.
Remark 2.

By Theorem 1, it is easy to see that a family of curves in n\mathbb{R}^{n} that are not locally rectifiable is a pp-exceptional set for p1p\geq 1.

3.2. Exceptional family of curves on Carnot groups

Recall the definition of horizontal subbundle H𝔾T𝔾H\mathbb{G}\subset T\mathbb{G} from Section 2.1. We say that a function f:I𝔾f\colon I\to\mathbb{G}, II\subset\mathbb{R} is dρd_{\rho}-Lipschitz continuous if it is Lipschitz continuous between metric spaces (I,dE)(I,d_{E}) and (𝔾,dρ)(\mathbb{G},d_{\rho}).

Definition 3.

A continuous curve γ:I𝔾\gamma\colon I\to\mathbb{G}, II\subset\mathbb{R}, is called horizontal if it is dEd_{E}-Lipschitz continuous and the tangent vector γ˙(t)\dot{\gamma}(t) belongs to Hγ(t)𝔾H_{\gamma(t)}\mathbb{G} for almost all tIt\in I.

We mention a well know example of a pp-exceptional family of curves on a Carnot group.

Example 1.

We define 𝐌:={dρ1  γ,γΓ}{\bf M}:=\{\mathcal{H}^{1}_{d_{\rho}}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits\gamma,\ \gamma\in\Gamma\}, where

Γ:={γ:[0,1]𝔾is a dE-Lipschitz continuous curve},\Gamma:=\{\gamma\colon[0,1]\to\mathbb{G}\ \mbox{is a $d_{E}$-Lipschitz continuous curve}\},

and

ΓH:={γ:[0,1]𝔾is a horizontal curve},𝐌H:={dρ1  γ,γΓH}.\Gamma_{H}:=\{\gamma:[0,1]\to\mathbb{G}\ \mbox{is a horizontal curve}\},\quad{\bf M}_{H}:=\{\mathcal{H}^{1}_{d_{\rho}}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits\gamma,\ \gamma\in\Gamma_{H}\}.

We claim that 𝐌𝐌His p-exceptional for all p>0.{\bf M}\setminus{\bf M}_{H}\quad\mbox{is $p$-exceptional for all $p>0$.} By Theorem 1 it is enough to find a nonnegative function fLp(𝔾,𝐠𝔾)f\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) such that

γf𝑑dρ1=\int_{\gamma}f\,d\mathcal{H}^{1}_{d_{\rho}}=\infty

for all γΓΓH\gamma\in\Gamma\setminus\Gamma_{H}. Without loss of generality we can assume that the loci of curves γΓ\gamma\in\Gamma are contained in a compact set K𝔾K\subset\mathbb{G}. Then the function

f(x)={1,ifxK,0,ifxK,f(x)=\begin{cases}1,\quad&\text{if}\quad x\in K,\\ 0,\quad&\text{if}\quad x\notin K,\end{cases}

belongs to Lp(𝔾,N)L^{p}(\mathbb{G},\mathcal{L}^{N}). Assume now, by contradiction, that there is γΓΓH\gamma\in\Gamma\setminus\Gamma_{H} such that γf𝑑dρ1=dρ1(γ)<\int_{\gamma}f\,d\mathcal{H}^{1}_{d_{\rho}}=\mathcal{H}^{1}_{d_{\rho}}(\gamma)<\infty. We will show that γΓH\gamma\in\Gamma_{H}, yielding a contradiction. The proof is more or less standard, but we prefer to give complete arguments. By the property of module of a minorized family of curves, see [V7̈1, Theorem 6.4] we can assume that the curve γ\gamma is injective. Since γ([0,1])\gamma([0,1]) is a closed and connected set, by [AT04, Theorem 4.4.8] we can write

(13) γ([0,1])=γ0(k=1γk([0,1])).\gamma([0,1])=\gamma_{0}\bigcup\Big{(}\cup_{k=1}^{\infty}\gamma_{k}([0,1])\Big{)}.

Here γ0\gamma_{0} is a Borel set such that dρ1(γ0)=0\mathcal{H}_{d_{\rho}}^{1}(\gamma_{0})=0 and γk:[0,1]γ([0,1])\gamma_{k}\colon[0,1]\to\gamma([0,1]) are dρd_{\rho}-Lipschitz continuous functions. Note also that

(14) dE(x,y)Cdρ(x,y)0dE1(γ)dρ1(γ)d_{E}(x,y)\leq Cd_{\rho}(x,y)\quad\Longrightarrow\quad 0\leq\mathcal{H}^{1}_{d_{E}}(\gamma)\leq\mathcal{H}^{1}_{d_{\rho}}(\gamma)

and that γ1(γ0)\gamma^{-1}(\gamma_{0}), has zero Lebesgue measure in [0,1][0,1]. Indeed (14) implies that dE1(γ0)=0.\mathcal{H}^{1}_{d_{E}}(\gamma_{0})=0. Thus we can apply the area formula of [AT04, Theorem 3.3.1] for A0:=γ1(γ0)[0,1]A_{0}:=\gamma^{-1}(\gamma_{0})\subset[0,1]:

(15) 1(A0)=A0|dγ(s)ds|𝑑1(s)=γ0card(γ1(y))𝑑dE1(y)=0.\begin{split}\mathcal{L}^{1}(A_{0})=\int_{A_{0}}\Big{|}\frac{d\gamma(s)}{ds}\Big{|}\,d\mathcal{L}^{1}(s)=\int_{\gamma_{0}}\mathrm{card}\,(\gamma^{-1}(y))d\mathcal{H}^{1}_{d_{E}}(y)=0.\end{split}

Thus, by Rademacher’s theorem, there exists A[0,1]A\subset[0,1] with 1(A)=0\mathcal{L}^{1}(A)=0 such that for all tAt\in A we have that γ(t)kγk([0,1])\gamma(t)\in\cup_{k}\gamma_{k}([0,1]) and γ\gamma is differentiable at tt.

Since both \mathbb{R} and 𝔾\mathbb{G} are Carnot groups, by Pansu-Rademacher theorem, all γk\gamma_{k}’s are Pansu differentiable in a set [0,1]A1[0,1]\setminus A_{1} with 1(A1)=0\mathcal{L}^{1}(A_{1})=0. For any kk\in\mathbb{N}, let us denote by dPγk(t)d_{P}\gamma_{k}(t) the Pansu differential at a point t[0,1]A1t\in[0,1]\setminus A_{1}. Set now A:=A0A1A:=A_{0}\cup A_{1}. Arguing as in the proof of [FSSC07, Theorem 3.5 (2)] the Euclidean tangent space to γ\gamma at a point γ(t)\gamma(t), t[0,1]At\in[0,1]\setminus A, coincides with dPγk(τ)()d_{P}\gamma_{k}(\tau)(\mathbb{R}) if k,τk,\tau are such that γ(t)=γk(τ)\gamma(t)=\gamma_{k}(\tau). Since dPγkd_{P}\gamma_{k} is a group homomorphism between \mathbb{R} and 𝔾\mathbb{G}, it maps \mathbb{R} into the first (horizontal) layer of 𝔾\mathbb{G}, so that γ˙(t)Hγ(t)𝔾\dot{\gamma}(t)\in H_{\gamma(t)}\mathbb{G}, yielding a contradiction.

Remark 3.

By [AT04, Theorem 4.2.1] and [AT04, Remark 4.1.3], we can always assume that if t[0,1]At\in[0,1]\setminus A then |γ˙(t)|=1|\dot{\gamma}(t)|=1. Thus, if γ(t)=γk(τ)\gamma(t)=\gamma_{k}(\tau), and

γk(τ)=j=1𝐝1uj(τ)Xj1(γk(τ)),\gamma_{k}(\tau)=\sum_{j=1}^{{\bf d}_{1}}u_{j}(\tau)X_{j1}(\gamma_{k}(\tau)),

then uLC\|u\|_{L^{\infty}}\leq C, where CC is independent of τ\tau. Thus, in the definition of ΓH\Gamma_{H} we can replace “horizontal” by “admissible”, see [ABB20].

3.3. Exceptional families of Radon measures on Carnot groups

Definition 4.

If μ\mu is a measure on a metric space (X,ρ)(X,\rho), and h>0h>0, then the values

Θh(μ,x)=lim infr0μ(Bρ(x,r))rh,andΘh,(μ,x)=lim supr0μ(Bρ(x,r))rh,\Theta^{h}_{*}(\mu,x)=\liminf\limits_{r\to 0}\frac{\mu(B_{\rho}(x,r))}{r^{h}},\quad\text{and}\quad\Theta^{h,*}(\mu,x)=\limsup\limits_{r\to 0}\frac{\mu(B_{\rho}(x,r))}{r^{h}},

are called the lower and upper hh-density of μ\mu at the point xXx\in X, respectively. We say that measure μ\mu has hh-density Θh(μ,x)\Theta^{h}(\mu,x) if

0<Θh(μ,x)=Θh(μ,x)=Θh,(μ,x)<.0<\Theta^{h}_{*}(\mu,x)=\Theta^{h}(\mu,x)=\Theta^{h,*}(\mu,x)<\infty.
Lemma 1.

Let a Carnot group 𝔾\mathbb{G} of topological dimension NN and homogeneous dimension QQ be endowed with a distance function dρd_{\rho} of type (D2)(D4)(D_{2})-(D_{4}). Let μ\mu be a Radon measure on 𝔾\mathbb{G}. If Θh(μ,x)>0\Theta^{h}(\mu,x)>0 for 1<h<Q1<h<Q and μ\mu-a.e. x𝔾x\in\mathbb{G}, then Θh(μ,)Lp(𝔾,𝐠𝔾)\Theta^{h}(\mu,\cdot)\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) for any p>0p>0.

Proof.

Recall the relation 𝐠𝔾N\mathbf{g}_{\mathbb{G}}\sim\mathcal{L}^{N}. We notice that the map xΘh(μ,x)x\to\Theta^{h}(\mu,x) is Borel measurable, see e.g. [Sim83, Remark 3.1]. Let us show that

(16) 𝐠𝔾({x𝔾;Θh(μ,x)>0})=0.\mathbf{g}_{\mathbb{G}}(\{x\in\mathbb{G};\ \Theta^{h}(\mu,x)>0\})=0.

Fix R>0R>0 and assume by contradiction that

𝐠𝔾(Bdρ(e,R){x𝔾;Θh(μ,x)>0})>0.\mathbf{g}_{\mathbb{G}}\big{(}B_{d_{\rho}}(e,R)\cap\{x\in\mathbb{G}\,;\,\Theta^{h}(\mu,x)>0\}\big{)}>0.

We have

𝐠𝔾(Bdρ(e,R){x𝔾;Θh(μ,x)>0})=limk𝐠𝔾(Bdρ(e,R){x𝔾;Θd(μ,x)>1k}).\begin{split}\mathbf{g}_{\mathbb{G}}&\big{(}B_{d_{\rho}}(e,R)\cap\{x\in\mathbb{G}\,;\,\Theta^{h}(\mu,x)>0\}\big{)}\\ &=\lim_{k\to\infty}\mathbf{g}_{\mathbb{G}}\big{(}B_{d_{\rho}}(e,R)\cap\{x\in\mathbb{G}\,;\,\Theta^{d}(\mu,x)>\frac{1}{k}\}\big{)}.\end{split}

Denote Ek=Bdρ(e,R){x𝔾;Θh(μ,x)>1k}E_{k}=B_{d_{\rho}}(e,R)\cap\{x\in\mathbb{G}\,;\,\Theta^{h}(\mu,x)>\frac{1}{k}\}. Then there exists kk\in\mathbb{N} such that 0<𝐠𝔾(Ek)<0<\mathbf{g}_{\mathbb{G}}\big{(}E_{k}\big{)}<\infty. Thus, by [AT04, Theorem 2.4.3],

(17) dρh(Ek)kωμ(Ek)kωμ(Bdρ(e,R))<,\mathcal{H}_{d_{\rho}}^{h}(E_{k})\leq k\omega\,\mu(E_{k})\leq k\omega\mu\,(B_{d_{\rho}}(e,R))<\infty,

where ω\omega is a normalisation constant. On the other hand, the equivalence (8) implies that dρh(Ek)=\mathcal{H}_{d_{\rho}}^{h}(E_{k})=\infty contradicting (17). Letting RR\to\infty we obtain (16). This accomplishes the proof. ∎

Example 2.

Consider the Heisenberg group H1H^{1} endowed with a distance function dρd_{\rho} of type (D2)(D4)(D_{2})-(D_{4}). Let 𝐌\mathbf{M} be the set of all Radon measures μ\mu on H1H^{1} satisfying

(18) Θ1(μ,x):=limr0μ(Bdρ(x,r))r>0for μ-a.e. xH1.\Theta^{1}(\mu,x):=\lim_{r\to 0}\,\dfrac{\mu(B_{d_{\rho}}(x,r))}{r}>0\qquad\mbox{for $\mu$-a.e. $x\in H^{1}$}.

We let 𝐌0𝐌\mathbf{M}_{0}\subset\mathbf{M} to be the measures for which there exists a countable family of Lipschitz maps Φi:AiH1\Phi_{i}\colon A_{i}\to H^{1}, AiA_{i}\subset\mathbb{R}, such that

μ(H1iΦi(Ai))=0.\mu\big{(}H^{1}\setminus\bigcup_{i}\Phi_{i}(A_{i})\big{)}=0.

We want to show that 𝐌𝐌0\mathbf{M}\setminus\mathbf{M}_{0} is pp-exceptional for p>0p>0. By Theorem 1 and Lemma 1 it is enough to show that

H1Θ1(μ,x)𝑑μ(x)=if μ𝐌𝐌0.\int_{H^{1}}\Theta^{1}(\mu,x)\,d\mu(x)=\infty\qquad\mbox{if $\mu\in\mathbf{M}\setminus\mathbf{M}_{0}$.}

Suppose by contradiction that the above integral is finite for a given measure μ𝐌\mu\in\mathbf{M}. Then Θ1(μ,x)<\Theta^{1}(\mu,x)<\infty for μ\mu-a.e. xH1x\in H^{1}. Then applying [AM21, Theorem 1.4], which is an analog of one-dimensional Preiss’ theorem for H1H^{1}, we obtain that μ𝐌0\mu\in\mathbf{M}_{0}. That is a contradiction.

Example 3.

In this example we refer to the definition of a tangent measure Tanh(μ,x)\mathrm{Tan}_{h}(\mu,x) in [Mat95, Chapter 14],[AM21]. Consider a Carnot group 𝔾\mathbb{G} as a metric space (𝔾,dρ)(\mathbb{G},d_{\rho}) where dρd_{\rho} is one of the distance functions (D2)(D4)(D_{2})-(D_{4}). For 1<h<Q1<h<Q denote by 𝐌\mathbf{M} the set of all Radon measures μ\mu on 𝔾\mathbb{G} such that

  • i)

    Θh(μ,x)>0\Theta_{*}^{h}(\mu,x)>0 for μ\mu-a.e. x𝔾x\in\mathbb{G};

  • ii)

    Tanh(μ,x){λ𝒮dρh  𝕍(x)}\mathrm{Tan}_{h}(\mu,x)\subset\{\lambda\mathcal{S}^{h}_{d_{\rho}}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits\mathbb{V}(x)\}, where 𝕍(x)\mathbb{V}(x) is a complemented homogeneous subgroup in 𝔾\mathbb{G}.

We let 𝐌0𝐌\mathbf{M}_{0}\subset\mathbf{M} to be a family of Radon measures such that there exists a countable family {Γi:=graph(ϕi);i=1,2,}\{\Gamma_{i}:=\mathrm{graph}\,(\phi_{i})\,;\,i=1,2,\dots\} of compact intrinsic Lipschitz graphs of metric dimension hh, see Section 3.4.1, which are intrinsically differentiable almost everywhere, see [FS16], and such that

μ(𝔾i=1Γi)=0.\mu\Big{(}\mathbb{G}\setminus\bigcup_{i=1}^{\infty}\Gamma_{i}\Big{)}=0.

Then we claim that 𝐌𝐌0\mathbf{M}\setminus\mathbf{M}_{0} is pp-exceptional family of measures. By the methods of Lemma 1 one can prove that Θh,(μ,)Lp(𝔾,𝐠𝔾)\Theta^{h,*}(\mu,\cdot)\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) for any p>0p>0. Thus it is enough to show that

𝔾Θh,(μ,x)𝑑μ(x)=if μ𝐌𝐌0\int_{\mathbb{G}}\Theta^{h,*}(\mu,x)\,d\mu(x)=\infty\qquad\mbox{if $\mu\in\mathbf{M}\setminus\mathbf{M}_{0}$}

by Theorem 1. Suppose by contradiction that the above integral is finite for a measure μ𝐌𝐌0\mu\in\mathbf{M}\setminus\mathbf{M}_{0}. Then Θh,(μ,x)<\Theta^{h,*}(\mu,x)<\infty for μ\mu-a.e. x𝔾x\in\mathbb{G} and therefore [AM21, Theorem 1.8] implies that μ𝐌0\mu\in\mathbf{M}_{0}, which is a contradiction.

3.4. Families of surfaces on Carnot groups

In this section we aim to define families of surfaces that will be of our interest.

3.4.1. Intrinsic Lipschitz surfaces on the Carnot groups

In the present section we recall the definition of intrinsic Lipschitz graphs, i.e. graphs of intrinsic Lipschitz functions, see [FSSC06, FS16]. Then we give a definition of an intrinsic Lipschitz surface. A subgroup 𝕄\mathbb{M} of a Carnot group 𝔾\mathbb{G} is called a homogeneous subgroup if 𝕄\mathbb{M} is a homogeneous group with respect to the dilation δλ\delta_{\lambda} defined in (3). Let us assume that 𝔾\mathbb{G} is decomposed into complementary homogeneous subgroups: 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}, 𝕄=e\mathbb{M}\cap\mathbb{H}=e, and let 𝐏𝕄\mathbf{P}_{\mathbb{M}} and 𝐏\mathbf{P}_{\mathbb{H}} be the canonical projections: 𝐏𝕄:𝔾𝕄\mathbf{P}_{\mathbb{M}}\colon\mathbb{G}\to\mathbb{M} and 𝐏:𝔾\mathbf{P}_{\mathbb{H}}\colon\mathbb{G}\to\mathbb{H} defined by the identity 𝐏𝕄q𝐏qq\mathbf{P}_{\mathbb{M}}q\cdot\mathbf{P}_{\mathbb{H}}q\equiv q for q𝔾q\in\mathbb{G}. The projections define intrinsic cones:

C𝕄,(e,β)={p𝔾𝐏𝕄pβ𝐏p},C𝕄,(q,β)=qC𝕄,(e,β),C_{\mathbb{M},\mathbb{H}}(e,\beta)=\{p\in\mathbb{G}\mid\ \|\mathbf{P}_{\mathbb{M}}p\|\leq\beta\|\mathbf{P}_{\mathbb{H}}p\|\},\qquad C_{\mathbb{M},\mathbb{H}}(q,\beta)=q\cdot C_{\mathbb{M},\mathbb{H}}(e,\beta),

where β>0\beta>0 is called the opening of the cone C𝕄,(q,β)C_{\mathbb{M},\mathbb{H}}(q,\beta) and qq is the vertex.

Definition 5.

The graph of a function f:Ωf\colon\Omega\to\mathbb{H}, where Ω\Omega is an open set of 𝕄\mathbb{M}, is the set

graph(f)={qf(q)𝔾=𝕄qΩ𝕄}.\mathrm{graph}\,(f)\ =\{\ q\cdot f(q)\in\mathbb{G}=\mathbb{M}\cdot\mathbb{H}\mid\ q\in\Omega\subset\mathbb{M}\}.

A function f:Ωf\colon\Omega\to\mathbb{H}, Ω𝕄\Omega\subset\mathbb{M}, is an intrinsic Lipschitz function in Ω\Omega with the Lipschitz constant L>0L>0 if

C𝕄,(p,1/L)graph(f)={p}for allpgraph(f).C_{\mathbb{M},\mathbb{H}}(p,1/L)\cap\mathrm{graph}\,(f)=\{p\}\quad\text{for all}\quad p\in\mathrm{graph}\,(f).

An intrinsic Lipschitz graph is the graph of an intrinsic Lipschitz function.

Left translation of intrinsic Lipschitz graphs are still intrinsic Lipschitz graphs. Following [FS16, Lemma 2.12], we set

(19) c0(𝕄,):=inf{mh:m+h=1}.c_{0}(\mathbb{M},\mathbb{H}):=\inf\{\left\|{mh}\right\|\,:\,\left\|{m}\right\|+\left\|{h}\right\|=1\}.
Remark 4.

We emphasise that there is a subtle difference in the notions of a Lipschitz function between metric spaces and that of intrinsic Lipschitz function within a Carnot group. We refer the reader to [FS16].

Definition 6.

The topological dimension 𝐝𝐭{\bf d_{t}} of a (sub)group is the dimension of its Lie algebra. The metric dimension 𝐝𝐦{\bf d_{m}} of a Borel set U𝔾U\subset\mathbb{G} is its Hausdorff dimension, with respect to the Hausdorff measure dρ\mathcal{H}_{d_{\rho}} (or 𝒮dρ\mathcal{S}_{d_{\rho}}) for a distance function dρd_{\rho} of type (D2)(D4)(D_{2})-(D_{4}). We say that 𝕄\mathbb{M} is a (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-subgroup of 𝔾\mathbb{G} if 𝕄\mathbb{M} is a homogeneous subgroup of 𝔾\mathbb{G} with topological dimension 𝐝𝐭{\bf d_{t}} and metric dimension 𝐝𝐦{\bf d_{m}}. We say that graph(f)\mathrm{graph}\,(f) is intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz grapf if f:Ωf\colon\Omega\to\mathbb{H}, Ω𝕄\Omega\subset\mathbb{M}, is an intrinsic Lipschitz function and 𝕄\mathbb{M} is a (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-subgroup of 𝔾\mathbb{G}.

The metric dimension 𝐝𝐦{\bf d_{m}} of a homogeneous subgroup is an integer usually larger than its topological dimension 𝐝𝐭{\bf d_{t}}, see [Mit85] and coincides with the homogeneous dimension, defined in Section 2.1.

Definition 7.

Suppose 1𝐝𝐭N11\leq{\bf d_{t}}\leq N-1 and 1𝐝𝐦Q11\leq{\bf d_{m}}\leq Q-1. A non-empty subset S𝔾S\subset\mathbb{G} is called an intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surface (or manifold in the graph representation) in 𝔾\mathbb{G} if to every point xSx\in S there correspond an open neighbourhood U(x,r)𝔾U(x,r)\subset\mathbb{G}, a decomposition 𝔾=𝕄UU\mathbb{G}=\mathbb{M}_{U}\cdot\mathbb{H}_{U} and an open set Ω𝕄U\Omega\subset\mathbb{M}_{U} such that

  • xUx\in U;

  • 𝕄U\mathbb{M}_{U} is a (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-subgroup of 𝔾\mathbb{G};

  • there exists an intrinsic Lipschitz map fU:ΩUf_{U}\colon\Omega\to\mathbb{H}_{U} such that SU=graph(fU)U(x,r)S\cap U=\mathrm{graph}\,(f_{U})\cap U(x,r).

3.4.2. Measures on the intrinsic Lipschitz graphs and surfaces

We suppose that (𝔾,,N,dρ)(\mathbb{G},\mathcal{B},\mathcal{L}^{N},d_{\rho}) is a Carnot group with the Borel σ\sigma-algebra \mathcal{B}, the Lebesgue measure N\mathcal{L}^{N} which is identified with the Haar measure 𝐠𝔾\mathbf{g}_{\mathbb{G}}, and a distance function dρd_{\rho} of types (D2D4)(D_{2}-D_{4}). We assume that 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}. The following result provides the construction of a Borel measure on an intrinsic Lipschitz graph.

Theorem 2.

[FS16, Theorem 3.9] Let SS be an intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graph on a Carnot group (𝔾,dρ)(\mathbb{G},d_{\rho}). Suppose S=graph(f)S=\mathrm{graph}\,(f) is defined by a decomposition 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} and an intrinsic LL-Lipschitz function f:Ωf\colon\Omega\to\mathbb{H} in the domain Ω𝕄\Omega\subset\mathbb{M}. Then there are positive constants c0,cc_{0},c depending on the decomposition 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} such that

(20) (c0(𝕄)1+L)𝐝𝐦R𝐝𝐦𝒮dρ𝐝𝐦(SBdρ(x,R))c(𝕄)(1+L)dmR𝐝𝐦\Big{(}\dfrac{c_{0}(\mathbb{M}\cdot\mathbb{H})}{1+L}\Big{)}^{{\bf d_{m}}}R^{{\bf d_{m}}}\leq\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(S\cap B_{d_{\rho}}(x,R))\leq c(\mathbb{M}\cdot\mathbb{H})(1+L)^{d_{m}}R^{{\bf d_{m}}}

for all points xSx\in S and R>0R>0. In particular, the Hausdorff dimension of SS with respect to dρd_{\rho} equals the homogeneous dimension of the group 𝕄\mathbb{M}.

Let 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} be a decomposition of 𝔾\mathbb{G} into complementary homogeneous subgroups. If Ω𝕄\Omega\subset\mathbb{M} is an open set and f:Ωf\colon\Omega\to\mathbb{H} is an intrinsic Lipschitz function, then one can define a map Φf:Ω𝔾\Phi_{f}\colon\Omega\to\mathbb{G} by Φf(m)=mf(m)\Phi_{f}(m)=m\cdot f(m), mΩm\in\Omega, that parametrises the intrinsic Lipschitz graph of ff. We define two measures

(21) σS(A)=𝒮d𝐝𝐦  graph(f)(A)=𝒮d𝐝𝐦(graph(f)A)\sigma_{S}(A)=\mathcal{S}^{{\bf d_{m}}}_{d}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits{\mathrm{graph}\,(f)}(A)=\mathcal{S}^{{\bf d_{m}}}_{d}(\mathrm{graph}\,(f)\cap A)

and

(22) μ(A)=(((Φf))𝐠𝕄)(A)=𝐠𝕄(Φf1(A))=𝐠𝕄(Φf1(graph(f)A))\mu(A)=\Big{(}\big{(}(\Phi_{f})_{\sharp}\big{)}\mathbf{g}_{\mathbb{M}}\Big{)}(A)=\mathbf{g}_{\mathbb{M}}(\Phi_{f}^{-1}(A))=\mathbf{g}_{\mathbb{M}}(\Phi_{f}^{-1}(\mathrm{graph}\,(f)\cap A))

for any Borel measurable subset A𝔾A\subset\mathbb{G}. Both measures σS\sigma_{S} and μ\mu are concentrated on the set graph(f)𝔾\mathrm{graph}\,(f)\subset\mathbb{G}.

Theorem 3.

For the measures σS\sigma_{S} and μ\mu defined in (21) and (22) there are positive constants C1C_{1} and C2C_{2} such that

(23) C1σS(A)μ(A)C2σS(A)C_{1}\sigma_{S}(A)\leq\mu(A)\leq C_{2}\sigma_{S}(A)

for any Borel measurable set A𝔾A\subset\mathbb{G}.

Proof.

Notice that if A𝔾A\subset\mathbb{G}, then by definition

(24) μ(A)=𝐠𝕄(Φf1(graph(f)A))=𝐠𝕄(𝐏𝕄(graph(f)A)).\mu(A)=\mathbf{g}_{\mathbb{M}}(\Phi_{f}^{-1}(\mathrm{graph}\,(f)\cap A))=\mathbf{g}_{\mathbb{M}}(\mathbf{P}_{\mathbb{M}}(\mathrm{graph}\,(f)\cap A)).

Let Bdρ(x,R)𝔾B_{d_{\rho}}(x,R)\subset\mathbb{G} be a ball centred at xgraph(f)x\in\mathrm{graph}\,(f). Then, by  [FS16, Formula (44)],

𝐏𝕄(Bdρ(x,cR))𝐏𝕄(graph(f)Bdρ(x,R))𝐏𝕄(Bdρ(x,R)),\mathbf{P}_{\mathbb{M}}(B_{d_{\rho}}(x,cR))\subset\mathbf{P}_{\mathbb{M}}(\mathrm{graph}\,(f)\cap B_{d_{\rho}}(x,R))\subset\mathbf{P}_{\mathbb{M}}(B_{d_{\rho}}(x,R)),

where c=c0(𝕄)1+Lc=\dfrac{c_{0}(\mathbb{M}\cdot\mathbb{H})}{1+L}. Moreover, it was shown in [FS16, Lemma 2.20] that

𝐠𝕄(𝐏𝕄(Bdρ(x,R)))=c1R𝐝𝐦,c1=𝐠𝕄(Bdρ(e,1)).\mathbf{g}_{\mathbb{M}}\Big{(}\mathbf{P}_{\mathbb{M}}\big{(}B_{d_{\rho}}(x,R)\big{)}\Big{)}=c_{1}R^{{\bf d_{m}}},\quad c_{1}=\mathbf{g}_{\mathbb{M}}(B_{d_{\rho}}(e,1)).

It implies

(25) c1c𝐝𝐦R𝐝𝐦𝐠𝕄(𝐏𝕄(graph(f)Bdρ(x,R)))=μ(Bdρ(x,R))c1R𝐝𝐦c_{1}c^{{\bf d_{m}}}R^{{\bf d_{m}}}\leq\mathbf{g}_{\mathbb{M}}\Big{(}\mathbf{P}_{\mathbb{M}}\big{(}\mathrm{graph}\,(f)\cap B_{d_{\rho}}(x,R)\big{)}\Big{)}=\mu(B_{d_{\rho}}(x,R))\leq c_{1}R^{{\bf d_{m}}}

by (24) and the definition of the measure μ\mu. Passing to the upper and lower limits in (25) we obtain

(26) c1c𝐝𝐦lim infR0μ(Bdρ(x,R))R𝐝𝐦lim supR0μ(Bdρ(x,R))R𝐝𝐦c1.\begin{split}c_{1}c^{{\bf d_{m}}}&\leq\liminf_{R\to 0}\frac{\mu(B_{d_{\rho}}(x,R))}{R^{{\bf d_{m}}}}\leq\limsup_{R\to 0}\frac{\mu(B_{d_{\rho}}(x,R))}{R^{{\bf d_{m}}}}\leq c_{1}.\end{split}

Therefore, arguing as in [Fed69, Section 2.10.19], we can write (25) as

(27) c~1c𝐝𝐦Θ𝐝𝐦(μ,x)Θ𝐝𝐦,(μ,x)c~1\begin{split}\tilde{c}_{1}c^{{\bf d_{m}}}\leq\Theta^{{\bf d_{m}}}_{*}(\mu,x)\leq\Theta^{{\bf d_{m}},*}(\mu,x)\leq\tilde{c}_{1}\end{split}

for any xgraph(f)x\in\mathrm{graph}\,(f). Again by [Fed69, Section 2.10.19] if follows that

(28) c~1c𝐝𝐦𝒮dρ𝐝𝐦(U)μ(U)c~12𝐝𝐦𝒮dρ𝐝𝐦(U)\tilde{c}_{1}c^{{\bf d_{m}}}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(U)\leq\mu(U)\leq\tilde{c}_{1}2^{{\bf d_{m}}}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(U)

for any Borel measurable set U𝔾U\subset\mathbb{G}. As a set UU we take U:=Agraph(f)U:=A\cap\mathrm{graph}\,(f) for a Borel set A𝔾A\subset\mathbb{G}. By (28) we get

C1σS(A)=C1𝒮dρ𝐝𝐦(U)μ(U)=μ(A)C2𝒮dρ𝐝𝐦(U)=C2σS(A),C_{1}\sigma_{S}(A)=C_{1}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(U)\leq\mu(U)=\mu(A)\leq C_{2}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(U)=C_{2}\sigma_{S}(A),

where C1:=c~1c𝐝𝐦C_{1}:=\tilde{c}_{1}c^{{\bf d_{m}}} and C2:=c~12𝐝𝐦C_{2}:=\tilde{c}_{1}2^{{\bf d_{m}}}. ∎

Corollary 1.

Let S=graph(f)S=\mathrm{graph}\,(f) be an intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graph in a Carnot group (𝔾,dρ)(\mathbb{G},d_{\rho}), and let Φf:Ω𝔾\Phi_{f}\colon\Omega\to\mathbb{G} be the map defined by Φf(m)=mf(m)\Phi_{f}(m)=m\cdot f(m), that parametrizes SS. For the measure σS\sigma_{S} defined above and any Borel non-negative function hh on 𝔾\mathbb{G} one has

(29) C1𝔾h(y)𝑑σS(y)Ω(hΦf)(x)𝑑𝐠𝕄(x)C2𝔾h(y)𝑑σS(y).C_{1}\int_{\mathbb{G}}h(y)\,d\sigma_{S}(y)\leq\int_{\Omega}(h\circ\Phi_{f})(x)\,d\mathbf{g}_{\mathbb{M}}(x)\leq C_{2}\int_{\mathbb{G}}h(y)\,d\sigma_{S}(y).
Proof.

We recall a result from [Mat95, Theorem 1.19]. Let XX and YY be two separable metric spaces, Φ:XY\Phi\colon X\to Y a Borel map, ν\nu a Borel measure on XX and hh is a Borel non-negative function on YY. Then

Yh(y)d(Φν)(y)=X(hΦ)(x)𝑑ν(x).\int_{Y}h(y)\,d(\Phi_{\sharp}\nu)(y)=\int_{X}(h\circ\Phi)(x)\,d\nu(x).

We apply the result for the surface locally parametrised by an intrinsic Lipschitz graph of ff, taking X=Ω𝕄X=\Omega\subset\mathbb{M}, Y=𝔾Y=\mathbb{G}, Φ=Φf\Phi=\Phi_{f}, ν=𝐠𝕄\nu=\mathbf{g}_{\mathbb{M}}. Then (29) follows. ∎

Remark 5.

We stress that, if SS is an intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surface, then according to Definition 7, the constant c0(𝕄,)c_{0}(\mathbb{M},\mathbb{H}) in (19) depends on SS and therefore we can write c0(𝕄,)=:c0(S)c_{0}(\mathbb{M},\mathbb{H})=:c_{0}(S) when we consider this constant on a surface SS.

If in Corollary 1 we track the definition of the constants C1,C2C_{1},C_{2}, then we see that they depend (up to geometric constants) only on the decomposition 𝕄\mathbb{M}\cdot\mathbb{H} and the Lipschitz constant LL. So, we can say that C1,C2C_{1},C_{2} depend on SS and we write C1(S)C_{1}(S), C2(S)C_{2}(S).

3.4.3. Examples of families of surfaces on Carnot groups

Example 4.

(1,1)(1,1)-intrinsic Lipschitz surfaces. Let 𝕄X=exp(tX)\mathbb{M}_{X}=\exp(tX), tt\in\mathbb{R}, X𝔤1X\in\mathfrak{g}_{1} be a one dimensional commutative subgroup and X\mathbb{H}_{X} a complementary to 𝕄X\mathbb{M}_{X} subgroup. Let ϕX:𝕄XX\phi_{X}\colon\mathbb{M}_{X}\to\mathbb{H}_{X} be a Lipschitz map. The family

Φ={ϕX:𝕄XX:X𝔤1}\Phi=\{\phi_{X}\colon\mathbb{M}_{X}\to\mathbb{H}_{X}:\ \ X\in\mathfrak{g}_{1}\}

is a family of (1,1)(1,1)-intrinsic Lipschitz surfaces, that is a family of horizontal curves.

Example 5.

A family of parametrised intrinsic Lipschitz graphs. Let 𝕄\mathbb{M} and \mathbb{H} be complementary subgroups and f:Ωf\colon\Omega\to\mathbb{H}, Ω𝕄\Omega\subset\mathbb{M}, be an intrinsic Lipschitz function with S=graph(f)S=\mathrm{graph}\,(f). We define

fλ:δλΩ:fλ(m):=δλf(δ1/λm).f_{\lambda}\colon\delta_{\lambda}\Omega\to\mathbb{H}:\ f_{\lambda}(m):=\delta_{\lambda}f(\delta_{1/\lambda}m).

Then Sλ=graph(fλ)S_{\lambda}=\mathrm{graph}\,(f_{\lambda}) is a family of intrinsic Lipschitz graphs, parametrised by λ>0\lambda>0 and it coincides with δλS\delta_{\lambda}S.

We also consider

fq:Ωq,Ωq={m𝕄:𝐏𝕄(q1m)Ω},qE𝔾f_{q}\colon\Omega_{q}\to\mathbb{H},\quad\Omega_{q}=\{m\in\mathbb{M}:\ \mathbf{P}_{\mathbb{M}}(q^{-1}m)\in\Omega\},\quad q\in E\subseteq\mathbb{G}

defined by

fq(m)=(𝐏(q1m))1f(𝐏𝕄(q1m)).f_{q}(m)=\big{(}\mathbf{P}_{\mathbb{H}}(q^{-1}m)\big{)}^{-1}\cdot f(\mathbf{P}_{\mathbb{M}}(q^{-1}m)).

Then

Sq=graph(fq)={(mfq(m)):mΩq,qE𝔾,fq:Ωq}S_{q}=\mathrm{graph}\,(f_{q})=\{\big{(}m\cdot f_{q}(m)\big{)}:\ m\in\Omega_{q},\ \ q\in E\subset\mathbb{G},\ \ f_{q}\colon\Omega_{q}\to\mathbb{H}\}

is a family of intrinsic Lipschitz graphs parametrised by qE𝔾q\in E\subset\mathbb{G} and it coincides with qS=Lq(S)q\cdot S=L_{q}(S). The details about the properties of SλS_{\lambda} and SqS_{q} see in [FS16, Theorem 3.2].

Particularly, if qE=q\in E=\mathbb{H}, then 𝐏𝕄(q1m)=m\mathbf{P}_{\mathbb{M}}(q^{-1}m)=m, 𝐏(q1m)=q1\mathbf{P}_{\mathbb{H}}(q^{-1}m)=q^{-1}. It implies Ωq=Ω\Omega_{q}=\Omega, and the family fq(m)=qf(m)f_{q}(m)=q\cdot f(m) is a family of graphs shifted along the subgroup \mathbb{H}.

Example 6.

Let Aut(𝔾)\mathcal{F}\in\text{\rm Aut}(\mathbb{G}) be a grading preserving automorphism. Let f:𝕄f\colon\mathbb{M}\to\mathbb{H} be an (intrinsic) Lipschitz function with S=graph(f)S=\mathrm{graph}\,(f). We define f:(𝕄)()f_{\mathcal{F}}\colon\mathcal{F}(\mathbb{M})\to\mathcal{F}(\mathbb{H}) by f(m)=(f1)(m).f_{\mathcal{F}}(m)=\big{(}{\mathcal{F}}f{\mathcal{F}}^{-1}\big{)}(m). Then

S=graph(f)={(mf(m)):m(𝕄)}S_{\mathcal{F}}=\mathrm{graph}\,(f_{\mathcal{F}})=\{\big{(}m\cdot f_{\mathcal{F}}(m)\big{)}:\ m\in\mathcal{F}(\mathbb{M})\}

is a family of (intrinsic) Lipschitz graphs and it coincides with (S)\mathcal{F}(S). Indeed, let (mf(m))S(m\cdot f(m))\in S, then

(S)((mf(m))\displaystyle\mathcal{F}(S)\ni\mathcal{F}((m\cdot f(m)) =\displaystyle= (mf(m))=(mf(1m))\displaystyle(\mathcal{F}m\cdot\mathcal{F}f(m))=(\mathcal{F}m\cdot\mathcal{F}f(\mathcal{F}^{-1}\mathcal{F}m))
=\displaystyle= (mf(1m))=(mf(m))S.\displaystyle(m^{\prime}\cdot\mathcal{F}f(\mathcal{F}^{-1}m^{\prime}))=(m^{\prime}\cdot f_{\mathcal{F}}(m^{\prime}))\in S_{\mathcal{F}}.

If Aut(𝔾)\mathcal{F}\in\text{\rm Aut}(\mathbb{G}) preserves the homogeneous norm, then an intrinsic LL-Lipschitz graph is transformed to an intrinsic LL-Lipschitz graph.

3.5. Exceptional families of intrinsic Lipschitz surfaces

3.5.1. Exceptional families for 0<p<10<p<1

Denote by Σ(𝐝𝐭,𝐝𝐦)\Sigma^{({\bf d_{t}},{\bf d_{m}})} a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces in 𝔾\mathbb{G}. With each surface SΣ(𝐝𝐭,𝐝𝐦)S\in\Sigma^{({\bf d_{t}},{\bf d_{m}})} we associate a measure σS\sigma_{S} as in Section 3.4.2. Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a subfamily and 𝐄{\bf E} the system of measures σS\sigma_{S}, SΣS\in\Sigma, associated with Σ\Sigma. Then Mp(Σ)=Mp(𝐄)M_{p}(\Sigma)=M_{p}({\bf E}) denotes the pp-module of the family of measures 𝐄\mathbf{E} as well as the family of the surfaces Σ\Sigma.

Note that in [Fug57, page 187] it was shown that if 0<p<10<p<1, then any system of Lipschitz kk-dimensional surfaces Σ\Sigma which intersects the cube

Cubea={xn|xl|<a,l=1,,n}{\rm Cube}_{a}=\{x\in\mathbb{R}^{n}\mid\ |x_{l}|<a,\ l=1,\ldots,n\}

has vanishing pp-module for any a>0a>0. It leads to the fact that Mp(Σ)=0M_{p}(\Sigma)=0, p(0,1)p\in(0,1), by the monotonicity of pp-module. We will study an analog situation on the Carnot groups.

Suppose 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} is a decomposition of 𝔾\mathbb{G}, and denote by 𝔪\mathfrak{m} and 𝔤\mathfrak{g} the Lie algebras of the Lie groups 𝕄\mathbb{M} and 𝔾\mathbb{G}, respectively. Let us fix a weak Malcev basis {W1𝕄,,W𝐝𝐭𝕄,W𝐝𝐭+1,,WN}\{W_{1}^{\mathbb{M}},\dots,W_{{\bf d_{t}}}^{\mathbb{M}},W_{{\bf d_{t}}+1},\dots,W_{N}\} for 𝔤\mathfrak{g} through 𝔪\mathfrak{m}, see [CG90, Theorem 1.1.13]. We define the following coordinate maps

T𝕄:𝐝𝐭𝕄:ζ=(ζ1,,ζ𝐝𝐭)exp(ζ1W1𝕄)exp(ζ𝐝𝐭W𝐝𝐭𝕄),T_{\mathbb{M}}\colon\mathbb{R}^{{\bf d_{t}}}\to\mathbb{M}:\quad\zeta=(\zeta_{1},\ldots,\zeta_{{\bf d_{t}}})\mapsto\exp(\zeta_{1}W_{1}^{\mathbb{M}})\cdot\ldots\cdot\exp(\zeta_{{\bf d_{t}}}W_{{\bf d_{t}}}^{\mathbb{M}}),
T𝔾:N𝐝𝐭𝔾:s=(s𝐝𝐭+1,,sN)exp(s𝐝𝐭+1W𝐝𝐭+1)exp(sNWN),T_{\mathbb{G}}\colon\mathbb{R}^{N-{\bf d_{t}}}\to\mathbb{G}:\quad s=(s_{{\bf d_{t}}+1},\ldots,s_{N})\mapsto\exp(s_{{\bf d_{t}}+1}W_{{\bf d_{t}}+1})\cdot\ldots\cdot\exp(s_{N}W_{N}),
T:N𝐝𝐭𝕄𝔾:s=(s𝐝𝐭+1,,sN)𝕄exp(s𝐝𝐭+1W𝐝𝐭+1)exp(sNWN).T\colon\mathbb{R}^{N-{\bf d_{t}}}\to\mathbb{M}\setminus\mathbb{G}:\ s=(s_{{\bf d_{t}}+1},\ldots,s_{N})\mapsto\mathbb{M}\cdot\exp(s_{{\bf d_{t}}+1}W_{{\bf d_{t}}+1})\cdot\ldots\cdot\exp(s_{N}W_{N}).

We define the natural projection on the right coset space by

(30) π:𝔾𝕄𝔾:g𝕄g.\pi\colon\mathbb{G}\to\mathbb{M}\setminus\mathbb{G}:\quad g\mapsto\mathbb{M}\cdot g.

The group 𝔾\mathbb{G} acts on the right on the coset space 𝕄𝔾\mathbb{M}\setminus\mathbb{G} by

(𝕄𝔾)×𝔾𝕄𝔾:(𝕄g,g~)𝕄gg~.\big{(}\mathbb{M}\setminus\mathbb{G}\big{)}\times\mathbb{G}\to\mathbb{M}\setminus\mathbb{G}:\quad(\mathbb{M}\cdot g,\tilde{g})\mapsto\mathbb{M}\cdot g\tilde{g}.

Since 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}, the coset space is “parametrized” by the elements of \mathbb{H} in the following sence 𝕄g=𝕄mh=𝕄h\mathbb{M}\cdot g=\mathbb{M}\cdot mh=\mathbb{M}\cdot h. Moreover 𝕄gg~=𝕄mhm~h~=𝕄h~~\mathbb{M}\cdot g\tilde{g}=\mathbb{M}\cdot mh\tilde{m}\tilde{h}=\mathbb{M}\cdot\tilde{\tilde{h}}, where h~~\tilde{\tilde{h}} is not necessarily hh~h\tilde{h}. In the following proposition we formulate a Fubini type theorem related to the right quotient space 𝕄𝔾\mathbb{M}\setminus\mathbb{G}.

Proposition 2.

[CG90, Lemma 1.2.13] [RH63, Theorem 15.24] The map T=πT𝔾T=\pi\circ T_{\mathbb{G}} is a diffeomorphism and the push forward measure 𝐠𝕄𝔾=T#(N𝐝𝐭)\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}}=T_{\#}(\mathcal{L}^{N-{\bf d_{t}}}) is a right 𝔾\mathbb{G}-invariant measure on the coset space 𝕄𝔾\mathbb{M}\setminus\mathbb{G}[CG90, Theorem 1.2.12]. Moreover the right invariant measure 𝐠𝔾\mathbf{g}_{\mathbb{G}} on 𝔾\mathbb{G} and the measure 𝐠𝕄𝔾\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}} on 𝕄𝔾\mathbb{M}\setminus\mathbb{G} are related by

(31) 𝔾ϰ(g)𝑑𝐠𝔾=𝕄𝔾𝑑𝐠𝕄𝔾𝕄ϰ(mg)𝑑𝐠𝕄\int_{\mathbb{G}}\varkappa(g)d\mathbf{g}_{\mathbb{G}}=\int_{\mathbb{M}\setminus\mathbb{G}}d\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}}\int_{\mathbb{M}}\varkappa(mg)d\mathbf{g}_{\mathbb{M}}

for any continuous function ϰ\varkappa with a compact support.

By making use of the Vitali covering lemma, we consider a countable family of Euclidean balls {B(ξj,r)𝐝𝐭,j}\{B(\xi_{j},r)\subset\mathbb{R}^{{\bf d_{t}}},j\in\mathbb{N}\} such that

(32) {B(ξj,r),j}is an open covering of𝐝𝐭;the ballsB(ξj,r/5)are disjoint.\begin{array}[]{lll}&\bullet&\ \{B(\xi_{j},r),\ j\in\mathbb{N}\}\ \text{is an open covering of}\ \mathbb{R}^{{\bf d_{t}}};\\ &\bullet&\ \text{the balls}\ B(\xi_{j},r/5)\ \text{are disjoint}.\end{array}

In particular, if a finite family of balls 𝔅={B(ξj1,3r),,B(ξjK,3r)}\mathfrak{B}=\{B(\xi_{j_{1}},3r),\dots,B(\xi_{j_{K}},3r)\} has nonempty intersection, then #𝔅30𝐝𝐭\#\mathfrak{B}\leq 30^{{\bf d_{t}}}. Indeed, suppose for sake of simplicity that 𝔅={B(ξ1,3r),,B(ξK,3r)}\mathfrak{B}=\{B(\xi_{1},3r),\dots,B(\xi_{K},3r)\} are such that ζ0i=1KB(ξi,3r)\zeta_{0}\in\cap_{i=1}^{K}B(\xi_{i},3r). By triangle inequality i=1KB(ξi,3r)B(ξ1,6r)\cup_{i=1}^{K}B(\xi_{i},3r)\subset B(\xi_{1},6r). Since {B(ξ1,r/5),,B(ξK,r/5)}\{B(\xi_{1},r/5),\dots,B(\xi_{K},r/5)\} is a disjoint family in B(ξ1,6r)B(\xi_{1},6r) we obtain

vol(B(0,1))r𝐝𝐭5𝐝𝐭(#𝔅)=|i=1KB(ξi,r/5)||B(ξ1,6r)|=(6r)𝐝𝐭vol(B(0,1)).\begin{split}{\rm vol(B(0,1))}\frac{r^{\bf d_{t}}}{5^{{\bf d_{t}}}}\cdot(\#\mathfrak{B})&=|\cup_{i=1}^{K}B(\xi_{i},r/5)|\leq|B(\xi_{1},6r)|=(6r)^{{\bf d_{t}}}{\rm vol(B(0,1))}.\end{split}

Let now ψC0(B(0,3r))\psi\in C^{\infty}_{0}(B(0,3r)) be a cut-off function of the ball B(0,2r)B(0,2r) and set

(33) ϕ(ζ):=ϕr(ζ)=i2i|ξjζ|1ψ(ξjζ),\phi(\zeta):=\phi_{r}(\zeta)=\sum_{i}2^{-i}|\xi_{j}-\zeta|^{-1}\psi(\xi_{j}-\zeta),

where ξj\xi_{j} are centers of the balls in family (32).

Lemma 2.

For any r>0r>0 and 0<1<p0<1<p we have ϕrLp(𝐝𝐭)\phi_{r}\in L^{p}(\mathbb{R}^{{\bf d_{t}}}).

Proof.

If ζ\zeta is fixed, then ψ(ξiζ)>0\psi(\xi_{i}-\zeta)>0 if and only if ζB(ξi,3r)\zeta\in B(\xi_{i},3r), which is possible for at most 30𝐝𝐭30^{{\bf d_{t}}} values of jj. Thus, if 0<p<10<p<1, then

|ϕr(ζ)|p(i2i|ξiζ|1ψ(ξiζ))pCpi2ip|ξiζ|pψp(ξiζ),|\phi_{r}(\zeta)|^{p}\leq\big{(}\sum_{i}2^{-i}|\xi_{i}-\zeta|^{-1}\psi(\xi_{i}-\zeta)\big{)}^{p}\leq C_{p}\sum_{i}2^{-ip}|\xi_{i}-\zeta|^{-p}\psi^{p}(\xi_{i}-\zeta),

so that

𝐝𝐭ϕr(ζ)pCpi2ip𝐝𝐭|ξiζ|pψp(ξiζ)𝑑ζCpi2ipB(ξi,3r)|ξiζ|p𝑑ζ=Cpi2ipB(0,3r)|ζ|p𝑑ζ<.\begin{split}&\int_{\mathbb{R}^{{\bf d_{t}}}}\phi_{r}(\zeta)^{p}\leq C_{p}\sum_{i}2^{-ip}\int_{\mathbb{R}^{{\bf d_{t}}}}|\xi_{i}-\zeta|^{-p}\psi^{p}(\xi_{i}-\zeta)\,d\zeta\\ &\leq C_{p}\sum_{i}2^{-ip}\int_{B(\xi_{i},3r)}|\xi_{i}-\zeta|^{-p}\,d\zeta=C_{p}\sum_{i}2^{-ip}\int_{B(0,3r)}|\zeta|^{-p}\,d\zeta<\infty.\end{split}

Lemma 3.

Let us fix the Vitali covering as in (32) by balls of a radius rr. Then for any xB(ξi,r)x\in B(\xi_{i},r) the function ϕr\phi_{r} defined in (33) satisfies

(34) B(x,r)ϕr(ζ)𝑑ζ=.\int_{B(x,r)}\phi_{r}(\zeta)\,d\zeta=\infty.
Proof.

We have B(x,r)B(ξi,2r)B(x,r)\subset B(\xi_{i},2r) by the triangle inequality and xB(ξi,r)x\in B(\xi_{i},r). Then for ζB(x,r)\zeta\in B(x,r) we get

ϕr(ζ)2i|ξiζ|1ψ(ξiζ)=2i|ξiζ|1.\phi_{r}(\zeta)\geq 2^{-i}|\xi_{i}-\zeta|^{-1}\psi(\xi_{i}-\zeta)=2^{-i}|\xi_{i}-\zeta|^{-1}.

It implies

B(x,r)ϕr(ζ)𝑑ζB(x,r)2i|ξjζ|1𝑑ζB(ξi,r~)2i|ξjζ|1𝑑ζ=,\displaystyle\int_{B(x,r)}\phi_{r}(\zeta)\,d\zeta\geq\int_{B(x,r)}2^{-i}|\xi_{j}-\zeta|^{-1}\,d\zeta\geq\int_{B(\xi_{i},\tilde{r})}2^{-i}|\xi_{j}-\zeta|^{-1}\,d\zeta=\infty,

for some r~(0,r)\tilde{r}\in(0,r). ∎

A Carnot group 𝔾\mathbb{G} can admit various decompositions 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} into homogeneous subgroups of the same topological 𝐝𝐭{\bf d_{t}} and Hausdorff 𝐝𝐦{\bf d_{m}} dimensions. The general result about the structure of such kind of decompositions and their number is not known. Therefore we restrict to the following cases. The first one when there are finitely many decompositions 𝔾α=𝕄αα\mathbb{G}_{\alpha}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}, α=1,2,,l\alpha=1,2,\ldots,l into non isomorphic pairs 𝕄αα\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}. The second case when each pair 𝕄αα\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha} belongs to an orbit of the action of grading preserving isometries of 𝔾\mathbb{G}.

It is enough to consider surfaces belonging to an open bounded set 𝒰𝔾\mathcal{U}\subset\mathbb{G}, for instance a ball. Let Σ\Sigma be a system of Lipschitz surfaces with some specific property, that will be specified in theorems below, and let

𝐄:={σS=𝒮dρ𝐝𝐦  S;SΣ}\mathbf{E}:=\{\sigma_{S}=\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits S\,;\,S\in\Sigma\}

be the system of the associated measures in (𝔾,dρ)(\mathbb{G},d_{\rho}). Then we denote by Σ𝒰\Sigma_{\mathcal{U}} the system of the Lipschitz surfaces S𝒰S\cap\mathcal{U}, SΣS\in\Sigma, and by 𝐄𝒰{\bf E}_{\mathcal{U}} the family of associated measures. If we show that Mp(𝐄𝒰)=0M_{p}(\mathbf{E}_{\mathcal{U}})=0, then it will imply that the system of measures 𝐄\mathbf{E} is exceptional by [Fug57, Theorem 3 (b)].

We start from the family of a most simple nature, that is a family of graphs parametrised over a single decomposition 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}. Then we consider multiple decompositions and more complicate families of surfaces.

Theorem 4.

Let 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} be a decomposition of 𝔾\mathbb{G}, and let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs over 𝕄\mathbb{M} and Σ𝒰\Sigma_{\mathcal{U}} the family of (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs in a bounded open set 𝒰𝔾\mathcal{U}\subset\mathbb{G}. Then the system 𝐄𝒰\mathbf{E}_{\mathcal{U}} is pp-exceptional for p(0,1)p\in(0,1).

Proof.

By definition, for any SΣ𝒰S\in\Sigma_{\mathcal{U}} there exists an intrinsic Lipschitz function fS:ΩSf_{S}\colon\Omega_{S}\to\mathbb{H}, , such that

  • ΩS\Omega_{S} is an open subset of 𝕄\mathbb{M};

  • S=graph(fS)S=\mathrm{graph}\,(f_{S}), i.e. S=ΦS(ΩS)S=\Phi_{S}(\Omega_{S}), where ΦS(m)=mfS(m)\Phi_{S}(m)=m\cdot f_{S}(m), mΩSm\in\Omega_{S};

If R,N>0R,N>0 we denote by Σ𝒰(R,N)Σ𝒰\Sigma_{\mathcal{U}}(R,N)\subset\Sigma_{\mathcal{U}} the family of graphs such that

  • the open set T𝕄1ΦS1(𝒰)T_{\mathbb{M}}^{-1}\Phi_{S}^{-1}(\mathcal{U}) contains an Euclidean ball B(ζS,R)𝐝𝐭B(\zeta_{S},R)\subset\mathbb{R}^{{\bf d_{t}}};

  • the associated measures σS=𝒮dρ𝐝𝐦  S\sigma_{S}=\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}\mathop{\hbox{\vrule height=7.0pt,width=0.5pt,depth=0.0pt\vrule height=0.5pt,width=6.0pt,depth=0.0pt}}\nolimits S satisfy

    ΦS1(𝒰)(hΦS)(x)𝑑𝐠𝕄(x)NS𝒰h(y)𝑑σS(y).\int_{\Phi_{S}^{-1}(\mathcal{U})}(h\circ\Phi_{S})(x)\,d\mathbf{g}_{\mathbb{M}}(x)\leq N\int_{S\cap\mathcal{U}}h(y)\,d\sigma_{S}(y).

We fix RR and NN and denote 𝐄𝒰(R,N){\bf E}_{\mathcal{U}}(R,N) the family of associated measures to Σ𝒰(R,N)\Sigma_{\mathcal{U}}(R,N). To show that Mp(𝐄𝒰(R,N))=0M_{p}\big{(}{\bf E}_{\mathcal{U}}(R,N)\big{)}=0 it is enough to find FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) such that 𝔾F𝑑σS=\int_{\mathbb{G}}F\,d\sigma_{S}=\infty for all σS𝐄𝒰(R,N)\sigma_{S}\in\mathbf{E}_{\mathcal{U}}(R,N), see Theorem 1. We set

(35) F=ϕRT𝕄1Π𝕄:𝔾[0,],F=\phi_{R}\circ T^{-1}_{\mathbb{M}}\circ\Pi_{\mathbb{M}}\colon\mathbb{G}\to[0,\infty],

where Π𝕄\Pi_{\mathbb{M}} is the projection over 𝕄\mathbb{M} associated with 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}, as in [FS16, Formula (28)], and ϕR\phi_{R} is the function defined in (33) for r=Rr=R.

Step 1: we claim that FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}). Since 𝒰K𝔾\mathcal{U}\subset K\subset\mathbb{G} for some compact set KK, it is enough to show that FK=χKFLp(𝔾,𝐠𝔾)F_{K}=\chi_{K}F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}). If we put K~:=π(K)\tilde{K}:=\pi(K), then K~𝕄/𝔾\tilde{K}\subset\mathbb{M}/\mathbb{G} is compact by the continuity of π\pi and χK(g)χK~(π(g))\chi_{K}(g)\leq\chi_{\tilde{K}}\big{(}\pi(g)\big{)}, since, if gKg\in K, then π(g)K~\pi(g)\in\tilde{K}.

We apply (31) with ϰ=FKp\varkappa=F^{p}_{K} and obtain

𝔾|FK|p(g)𝑑𝐠𝔾(g)=𝕄𝔾𝑑𝐠𝕄𝔾𝕄|FK|p(mg)𝑑𝐠𝕄(m)=𝕄𝔾𝑑𝐠𝕄𝔾𝕄χK(mg)|ϕRT𝕄1Π𝕄(m)|p𝑑𝐠𝕄(m)𝕄𝔾χK~(π(g))𝑑𝐠𝕄𝔾𝕄|ϕRT𝕄1(m)|p𝑑𝐠𝕄(m)=c(K~)𝕄|ϕRT𝕄1(m)|p𝑑𝐠𝕄(m)=c(K~)𝐝𝐭|ϕR|p(ζ)𝑑𝐝𝐭(ζ)<.\begin{split}&\int_{\mathbb{G}}|F_{K}|^{p}(g)\,d\mathbf{g}_{\mathbb{G}}(g)=\int_{\mathbb{M}\setminus\mathbb{G}}\,d\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}}\int_{\mathbb{M}}|F_{K}|^{p}(mg)\,d\mathbf{g}_{\mathbb{M}}(m)\\ &=\int_{\mathbb{M}\setminus\mathbb{G}}\,d\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}}\int_{\mathbb{M}}\chi_{K}(mg)|\phi_{R}\circ T^{-1}_{\mathbb{M}}\circ\Pi_{\mathbb{M}}(m)|^{p}\,d\mathbf{g}_{\mathbb{M}}(m)\\ &\leq\int_{\mathbb{M}\setminus\mathbb{G}}\,\chi_{\tilde{K}}(\pi(g))d\mathbf{g}_{\mathbb{M}\setminus\mathbb{G}}\int_{\mathbb{M}}|\phi_{R}\circ T^{-1}_{\mathbb{M}}(m)|^{p}\,d\mathbf{g}_{\mathbb{M}}(m)\\ &=c(\tilde{K})\int_{\mathbb{M}}|\phi_{R}\circ T^{-1}_{\mathbb{M}}(m)|^{p}\,d\mathbf{g}_{\mathbb{M}}(m)=c(\tilde{K})\int_{\mathbb{R}^{{\bf d_{t}}}}|\phi_{R}|^{p}(\zeta)\,d\mathcal{L}^{{\bf d_{t}}}(\zeta)<\infty.\end{split}

Step 2: we claim that SF𝑑σS=\int_{S}F\,d\sigma_{S}=\infty for any σS𝐄𝒰(R,N)\sigma_{S}\in\mathbf{E}_{\mathcal{U}}(R,N). Assume that SΣ𝒰(R,N)S\in\Sigma_{\mathcal{U}}(R,N) is the graph of a Lipschitz function fS:ΩSf_{S}\colon\Omega_{S}\to\mathbb{H}. We denote by ΦS(m):=mfS(m)\Phi_{S}(m):=m\cdot f_{S}(m), mΩSm\in\Omega_{S}, the parametrization of SS associated with fSf_{S}. We stress that (Π𝕄ΦS)(m)=m\big{(}\Pi_{\mathbb{M}}\circ\Phi_{S}\big{)}(m)=m. By Corollary 1 with N=C2N=C_{2}, and (35) we have:

NS𝒰F𝑑σSΦS1(𝒰)FΦS(m)𝑑𝐠𝕄=ΦS1(𝒰)ϕRT𝕄1Π𝕄ΦS(m)𝑑𝐠𝕄=ΦS1(𝒰)ϕRT𝕄1(m)𝑑𝐠𝕄=T𝕄1ΦS1(𝒰)ϕR(ζ)𝑑ζB(ζS,R)ϕR(ζ)𝑑ζ.\begin{split}&N\int_{S\cap\mathcal{U}}F\,d\sigma_{S}\geq\int_{\Phi_{S}^{-1}(\mathcal{U})}F\circ\Phi_{S}(m)\,d\mathbf{g}_{\mathbb{M}}=\int_{\Phi_{S}^{-1}(\mathcal{U})}\phi_{R}\circ T^{-1}_{\mathbb{M}}\circ\Pi_{\mathbb{M}}\circ\Phi_{S}(m)\,d\mathbf{g}_{\mathbb{M}}\\ &=\int_{\Phi_{S}^{-1}(\mathcal{U})}\phi_{R}\circ T^{-1}_{\mathbb{M}}(m)\,d\mathbf{g}_{\mathbb{M}}=\int_{T_{\mathbb{M}}^{-1}\Phi_{S}^{-1}(\mathcal{U})}\phi_{R}(\zeta)\,d\zeta\geq\int_{B(\zeta_{S},R)}\phi_{R}(\zeta)\,d\zeta.\end{split}

By the Vitali covering lemma, there exists ξi\xi_{i} such that ζSB(ξi,R)\zeta_{S}\in B(\xi_{i},R). Thus we apply Lemma 3 to show that the integral on the right-hand diverges.

Step 3: we claim that Mp(Σ𝒰)=0M_{p}(\Sigma_{\mathcal{U}})=0. From Steps 1 and 2 we conclude that Mp(Σ𝒰(R,N))=0M_{p}(\Sigma_{\mathcal{U}}(R,N))=0 for any R,N>0R,N>0. We set R=1MR=\frac{1}{M} for MM\in\mathbb{N}. Then

Σ𝒰=NMΣ𝒰(1M,N).\Sigma_{\mathcal{U}}=\bigcup_{N\in\mathbb{N}}\bigcup_{M\in\mathbb{N}}\Sigma_{\mathcal{U}}(\frac{1}{M},N).

We conclude by [Fug57, Theorem 3 (b)] that Mp(Σ𝒰)=0M_{p}(\Sigma_{\mathcal{U}})=0. ∎

Corollary 2.

Let 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} be a decomposition of 𝔾\mathbb{G} and let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces, such that locally each surface is represented by an intrinsic Lipschitz graph over 𝕄\mathbb{M}. Let Σ𝒰\Sigma_{\mathcal{U}} be the family of Lipschitz surfaces in a bounded open set 𝒰𝔾\mathcal{U}\subset\mathbb{G}. Then the system 𝐄𝒰\mathbf{E}_{\mathcal{U}} is pp-exceptional for p(0,1)p\in(0,1).

In the next step we assume that the family of graphs is parametrised over a decomposition 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} that belongs to the orbit of a subgroup KIso(𝔾)K\subset{\rm Iso}(\mathbb{G}) preserving the decomposition. Here we denote by Iso(𝔾){\rm Iso}(\mathbb{G}) the group of grading preserving isometries of 𝔾\mathbb{G}.

Theorem 5.

Let 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H}, and KK a subgroup of the group of isometries Iso(𝔾){\rm Iso}(\mathbb{G}) preserving the decomposition. Let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs over the orbit K(𝕄)K(\mathbb{M}) and Σ𝒰={S𝒰:SΣ}\Sigma_{\mathcal{U}}=\{S\cap\mathcal{U}:\ S\in\Sigma\}. Then the system of measures 𝐄𝒰\mathbf{E}_{\mathcal{U}} is pp-exceptional for p(0,1)p\in(0,1).

Proof.

By definition, for any SΣ𝒰S\in\Sigma_{\mathcal{U}} there exists an intrinsic Lipschitz function f^S:Ω^S^\hat{f}_{S}\colon\hat{\Omega}_{S}\to\hat{\mathbb{H}}, where Ω^S\hat{\Omega}_{S} is an open set in the group 𝕄^\hat{\mathbb{M}} such that 𝕄K(𝕄^^)\mathbb{M}\cdot\mathbb{H}\in K(\hat{\mathbb{M}}\cdot\hat{\mathbb{H}}). Then there is an isometric diffeomorphism kKk\in K such that 𝕄=k(𝕄^^)\mathbb{M}\cdot\mathbb{H}=k(\hat{\mathbb{M}}\cdot\hat{\mathbb{H}}). We write

fS=kf^Sk1:ΩS,f_{S}=k\circ\hat{f}_{S}\circ k^{-1}\colon\Omega_{S}\to\mathbb{H},

where ΩS\Omega_{S} is an open subset in 𝕄\mathbb{M}, such that k(Ω^S)=ΩSk(\hat{\Omega}_{S})=\Omega_{S}. Thus for any SΣ𝒰S\in\Sigma_{\mathcal{U}} there exists an intrinsic Lipschitz function fS:ΩSf_{S}\colon\Omega_{S}\to\mathbb{H}, such that

  • ΩS\Omega_{S} is an open subset of 𝕄\mathbb{M};

  • S=graph(fS)S=\mathrm{graph}\,(f_{S}), i.e. S=ΦS(ΩS)S=\Phi_{S}(\Omega_{S}), where

    ΦS(m)=k(m^f^S(m^))=k(m^)(kf^Sk1(k(m^))),mΩS.\Phi_{S}(m)=k\big{(}\hat{m}\cdot\hat{f}_{S}(\hat{m})\big{)}=k(\hat{m})\cdot\Big{(}k\circ\hat{f}_{S}\circ k^{-1}\big{(}k(\hat{m})\big{)}\Big{)},\quad m\in\Omega_{S}.

We define FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) as in (35) and argue as in Theorem 4. ∎

Corollary 3.

Let 𝔾=𝕄\mathbb{G}=\mathbb{M}\cdot\mathbb{H} and KIso(𝔾)K\subset{\rm Iso}(\mathbb{G}). Let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces, such that locally each surface is represented by an intrinsic Lipschitz graph over an element of the orbit K(𝕄)K(\mathbb{M}) as in Theorem 5. Then 𝐄𝒰\mathbf{E}_{\mathcal{U}} is pp-exceptional for p(0,1)p\in(0,1).

Now we assume that the group 𝔾\mathbb{G} can be written as 𝔾=𝕄αα\mathbb{G}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha} for finitely many α=1,2,,l\alpha=1,2,\ldots,l and 𝕄α\mathbb{M}_{\alpha} being (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-homogeneous non isomorphic subgroups for all α\alpha. Under this assumption we state the following result.

Theorem 6.

Let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs where each graph is parametrised over one of the decompositions 𝔾=𝕄αα\mathbb{G}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}, α=1,2,,l\alpha=1,2,\ldots,l. Let Σ𝒰=S𝒰\Sigma_{\mathcal{U}}=S\cap\,\mathcal{U}, SΣS\in\Sigma, and 𝒰𝔾\mathcal{U}\subset\mathbb{G} be an open set. Then Mp(𝐄𝒰)=0M_{p}(\mathbf{E}_{\mathcal{U}})=0 for p(0,1)p\in(0,1).

Proof.

We use the notation 𝔾α=𝕄αα\mathbb{G}_{\alpha}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}, α=1,2,,l\alpha=1,2,\ldots,l and 𝔾l=𝔾1××𝔾l\mathbb{G}^{l}=\mathbb{G}_{1}\times\ldots\times\mathbb{G}_{l}. We define the selection map χα(g)=mαhα\chi_{\alpha}(g)=m_{\alpha}h_{\alpha}, mα𝕄αm_{\alpha}\in\mathbb{M}_{\alpha}, hααh_{\alpha}\in\mathbb{H}_{\alpha}. It can be considered as a composition of the map

P:𝔾𝔾l=𝔾1××𝔾lg(m1h1,,mlhl);\begin{array}[]{cccccc}P\colon&\mathbb{G}&\to&\mathbb{G}^{l}=&\mathbb{G}_{1}\times\ldots\times\mathbb{G}_{l}\\ &g&\mapsto&&(m_{1}h_{1},\ldots,m_{l}h_{l});\end{array}

followed by the projection on α\alpha-slot. Then we define

(36) F(g)=α=1lϕrαT𝕄α1Π𝕄αχα(g).F(g)=\sum_{\alpha=1}^{l}\phi_{r_{\alpha}}\circ T^{-1}_{\mathbb{M}_{\alpha}}\circ\Pi_{\mathbb{M}_{\alpha}}\circ\chi_{\alpha}(g).

Then as in Theorem 4 we show that FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}). Moreover if SΣ𝒰(R,N)S\in\Sigma_{\mathcal{U}}(R,N) is the graph of a Lipschitz function fS:(Ωα)Sαf_{S}\colon(\Omega_{\alpha})_{S}\to\mathbb{H}_{\alpha}, for some α=1,,l\alpha=1,\ldots,l, where (Ωα)S(\Omega_{\alpha})_{S} is an open set in 𝕄α\mathbb{M}_{\alpha}, then ΦS(mα):=mαfS(mα)\Phi_{S}(m_{\alpha}):=m_{\alpha}\cdot f_{S}(m_{\alpha}) the parametrization of SS associated with fSf_{S}. Then

NS𝒰F𝑑σSΦS1(𝒰)FΦS(mα)𝑑𝐠𝕄αB((ζ¯α)S,rα)ϕrα(ζ¯α)𝑑ζ¯α=,\displaystyle N\int_{S\cap\mathcal{U}}F\,d\sigma_{S}\geq\int_{\Phi_{S}^{-1}(\mathcal{U})}F\circ\Phi_{S}(m_{\alpha})\,d\mathbf{g}_{\mathbb{M}_{\alpha}}\geq\int_{B((\overline{\zeta}_{\alpha})_{S},r_{\alpha})}\phi_{r_{\alpha}}(\overline{\zeta}_{\alpha})\,d\overline{\zeta}_{\alpha}=\infty,

as in the proof of Theorem 4. ∎

Corollary 4.

Let assume that 𝔾\mathbb{G} can be decomposed in a finitely many ways 𝔾=𝕄αα\mathbb{G}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}, α=1,2,,l\alpha=1,2,\ldots,l, with non isomorphic subgroups 𝕄α\mathbb{M}_{\alpha}.

Let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces, such that locally each surface is represented by an intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graph over one of the groups 𝕄α\mathbb{M}_{\alpha} as in Theorem 6. Then Mp(𝐄𝒰)=0M_{p}(\mathbf{E}_{\mathcal{U}})=0 for p(0,1)p\in(0,1).

The last result in this section is a combination of Theorem 5 and Theorem 6.

Theorem 7.

Let Σ\Sigma be a family of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces where each graph is parametrised over one of the decompositions 𝔾=𝕄αα\mathbb{G}=\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}, α=1,2,,l\alpha=1,2,\ldots,l. Moreover any term 𝕄αα\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha} is an element in the orbit of a subgroup KαIso(𝔾)K_{\alpha}\subset{\rm Iso}(\mathbb{G}) preserving the decomposition 𝕄αα\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}.

Let Σ𝒰=S𝒰\Sigma_{\mathcal{U}}=S\cap\,\mathcal{U}, SΣS\in\Sigma, and 𝒰𝔾\mathcal{U}\subset\mathbb{G} be an open set. Then Mp(𝐄𝒰)=0M_{p}(\mathbf{E}_{\mathcal{U}})=0 for p(0,1)p\in(0,1).

Proof.

We define function FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) as in (36). Moreover for any point qS𝒰q\in S\cap\mathcal{U} there is a neighbourhood VV, such that for S𝒰VS\cap\mathcal{U}\cap V there exists an intrinsic Lipschitz function f^S:(Ω^S)α^α\hat{f}_{S}\colon(\hat{\Omega}_{S})_{\alpha}\to\hat{\mathbb{H}}_{\alpha}, where (Ω^S)α𝕄^α(\hat{\Omega}_{S})_{\alpha}\subset\hat{\mathbb{M}}_{\alpha} such that 𝕄ααKα(𝕄^α^α)\mathbb{M}_{\alpha}\cdot\mathbb{H}_{\alpha}\in K_{\alpha}(\hat{\mathbb{M}}_{\alpha}\cdot\hat{\mathbb{H}}_{\alpha}). By choosing an isometry kαKαk_{\alpha}\in K_{\alpha} we find an intrinsic Lipschitz function fS:(ΩS)ααf_{S}\colon(\Omega_{S})_{\alpha}\to\mathbb{H}_{\alpha}, such that

  • (ΩS)α(\Omega_{S})_{\alpha} is an open subset of 𝕄α\mathbb{M}_{\alpha}, kα((Ω^S)α)=(ΩS)αk_{\alpha}\big{(}(\hat{\Omega}_{S})_{\alpha}\big{)}=(\Omega_{S})_{\alpha};

  • S𝒰V=graph(fS)S\cap\mathcal{U}\cap V=\mathrm{graph}\,(f_{S}), i.e. S=ΦS((ΩS)α)S=\Phi_{S}\big{(}(\Omega_{S})_{\alpha}\big{)}, where

    ΦS(m)=kα(m^)(kαf^Skα1(kα(m^))),m(ΩS)α,m^(Ω^S)α.\Phi_{S}(m)=k_{\alpha}(\hat{m})\cdot\Big{(}k_{\alpha}\circ\hat{f}_{S}\circ k^{-1}_{\alpha}\big{(}k_{\alpha}(\hat{m})\big{)}\Big{)},\quad m\in(\Omega_{S})_{\alpha},\quad\hat{m}\in(\hat{\Omega}_{S})_{\alpha}.

Then we proceed in the same way as in Theorem 4 and show that

NS𝒰VF𝑑σSB(ζS,R)ϕR(ζ)𝑑ζ=.N\int_{S\cap\mathcal{U}\cap V}F\,d\sigma_{S}\geq\int_{B(\zeta_{S},R)}\phi_{R}(\zeta)\,d\zeta=\infty.

3.5.2. Exceptional families for p1p\geq 1

B. Fuglede proved that the system of kk-dimensional Lipschitz surfaces in n\mathbb{R}^{n} which pass through a given point is pp-exceptional if and only if kpnkp\leq n [Fug57]. Here we show the sufficient part of the analogous statement for Carnot groups.

Theorem 8.

Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a collection of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs. Suppose that all the graphs SΣS\in\Sigma contain a common point g0𝔾g_{0}\in\mathbb{G}. Then for 𝐝𝐦pQ{\bf d_{m}}p\leq Q we have Mp(Σ)=0M_{p}(\Sigma)=0.

Proof.

Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a collection of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs containing a common point g0𝔾g_{0}\in\mathbb{G}. We can assume that g0=e𝔾g_{0}=e\in\mathbb{G} by the translation invariance of measures and the fact that the translation of an intrinsic Lipschitz graph is still an intrinsic Lipschitz graph, see [FS16, Theorem 3.2]. We need to find a non-negative measurable function F:𝔾F\colon\mathbb{G}\to\mathbb{R} such that 𝔾Fp𝑑𝐠𝔾<\int_{\mathbb{G}}F^{p}\,d\mathbf{g}_{\mathbb{G}}<\infty, but SF𝑑σS=\int_{S}F\,d\sigma_{S}=\infty for any SΣS\in\Sigma.

Let \|\cdot\| be a homogeneous norm on 𝔾\mathbb{G}, for instance one of the types (D2)(D4)(D_{2})-(D_{4}) and let dρd_{\rho} be a metric associated with the norm \|\cdot\|. We set

(37) F(g)={g𝐝𝐦,ifg<1,0,ifg1,𝐝𝐦p<Q.F(g)=\begin{cases}\|g\|^{-{\bf d_{m}}},\quad&\text{if}\quad\|g\|<1,\\ 0,\quad&\text{if}\quad\|g\|\geq 1,\end{cases}\qquad{\bf d_{m}}p<Q.

Then FLp(𝔾,𝐠𝔾)F\in L_{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) since

Bdρ(e,1)|F|p𝑑N=ω01rp𝐝𝐦+Q1𝑑r<\int_{B_{d_{\rho}}(e,1)}|F|^{p}\,d\mathcal{L}^{N}=\omega\int_{0}^{1}r^{-p{\bf d_{m}}+Q-1}dr<\infty

for p𝐝𝐦+Q>0-p{\bf d_{m}}+Q>0, where ω\omega is a suitable constant depending only on \|\cdot\|, see, e.g. [FS82, Proposition 1.15].

Consider intersections SBdρ(e,12j)S\cap B_{d_{\rho}}(e,\frac{1}{2^{j}}), jj\in\mathbb{N}. We divide the ball Bdρ(e,1)B_{d_{\rho}}(e,1) into the spherical rings Rj=Bdρ(e,12j)B¯dρ(e,12j+1)R_{j}=B_{d_{\rho}}(e,\frac{1}{2^{j}})\setminus\overline{B}_{d_{\rho}}(e,\frac{1}{2^{j+1}}). In each ring RjR_{j} we choose a point pjSBdρ(e,32j+2)p_{j}\in S\cap\partial B_{d_{\rho}}(e,\frac{3}{2^{j+2}}), Then Bdρ(pj,12j+3)RjB_{d_{\rho}}(p_{j},\frac{1}{2^{j+3}})\subset R_{j}. We observe that 2(j+1)𝐝𝐦g𝐝𝐦>2j𝐝𝐦2^{(j+1){\bf d_{m}}}\geq\|g\|^{-{\bf d_{m}}}>2^{j{\bf d_{m}}} for gRjg\in R_{j}. Then

SF𝑑σS\displaystyle\int_{S}F\,d\sigma_{S} \displaystyle\geq Bdρ(e,1)Sg𝐝𝐦𝑑σS=jRjSg𝐝𝐦𝑑σS\displaystyle\int_{B_{d_{\rho}}(e,1)\cap S}\|g\|^{-{\bf d_{m}}}\,d\sigma_{S}=\sum_{j}\int_{R_{j}\cap S}\|g\|^{-{\bf d_{m}}}\,d\sigma_{S}
>\displaystyle> j2j𝐝𝐦𝒮dρ𝐝𝐦(RjS)>j2j𝐝𝐦𝒮dρ𝐝𝐦(Bdρ(pj,2j3)S)\displaystyle\sum_{j}2^{j{\bf d_{m}}}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(R_{j}\cap S)>\sum_{j}2^{j{\bf d_{m}}}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(B_{d_{\rho}}(p_{j},2^{-j-3})\cap S)
\displaystyle\geq c123𝐝𝐦j2j𝐝𝐦2j𝐝𝐦=.\displaystyle c_{1}2^{-3{\bf d_{m}}}\sum_{j}2^{j{\bf d_{m}}}2^{-j{\bf d_{m}}}=\infty.

If 𝐝𝐦p=Q{\bf d_{m}}p=Q, then we need to change the function FF to

(38) F(g)={g𝐝𝐦(ln2g)α,ifg<1,0,ifg1.F(g)=\begin{cases}\|g\|^{-{\bf d_{m}}}\big{(}\ln\frac{2}{\|g\|}\big{)}^{-\alpha},\quad&\text{if}\quad\|g\|<1,\\ 0,\quad&\text{if}\quad\|g\|\geq 1.\end{cases}

If we choose α[𝐝𝐦Q,1]\alpha\in[\frac{{\bf d_{m}}}{Q},1], then FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) and SF(g)𝑑σS=\int_{S}F(g)\,d\sigma_{S}=\infty. Indeed,

Bdρ(0,1)|F|p𝑑𝐠𝔾=ω01rp𝐝𝐦+Q1(ln2r)αp𝑑r=ωln2tαp𝑑t<\int_{B_{d_{\rho}}(0,1)}|F|^{p}\,d\mathbf{g}_{\mathbb{G}}=\omega\int_{0}^{1}r^{-p{\bf d_{m}}+Q-1}\big{(}\ln\frac{2}{r}\big{)}^{-\alpha p}dr=\omega\int_{\ln 2}^{\infty}t^{-\alpha p}dt<\infty

if αp>1\alpha p>1 or equivalently α>𝐝𝐦Q\alpha>\frac{{\bf d_{m}}}{Q}. From the other side

SF𝑑σS\displaystyle\int_{S}F\,d\sigma_{S} \displaystyle\geq RjSg𝐝𝐦(ln2g)α𝑑σS\displaystyle\int_{R_{j}\cap S}\|g\|^{-{\bf d_{m}}}\Big{(}\ln\frac{2}{\|g\|}\Big{)}^{-\alpha}\,d\sigma_{S}
>\displaystyle> j2j𝐝𝐦(ln2j+2)α𝒮dρ𝐝𝐦(Bdρ(pj,2j3)S)\displaystyle\sum_{j}2^{j{\bf d_{m}}}(\ln 2^{j+2})^{-\alpha}\mathcal{S}^{{\bf d_{m}}}_{d_{\rho}}(B_{d_{\rho}}(p_{j},2^{-j-3})\cap S)
\displaystyle\geq c123𝐝𝐦(ln2)αj(j+2)α=\displaystyle c_{1}2^{-3{\bf d_{m}}}(\ln 2)^{-\alpha}\sum_{j}(j+2)^{-\alpha}=\infty

for α<1\alpha<1. ∎

Corollary 5.

Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a collection of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz surfaces. Suppose that all the surfaces S^Σ\mathaccent 866{S}\in\Sigma contain a common point g0𝔾g_{0}\in\mathbb{G}. Then for 𝐝𝐦pQ{\bf d_{m}}p\leq Q we have Mp(Σ)=0M_{p}(\Sigma)=0.

Proof.

We assume that g0=eg_{0}=e and, as in Theorem 8, consider the function FLp(𝔾,𝐠𝔾)F\in L^{p}(\mathbb{G},\mathbf{g}_{\mathbb{G}}) for 𝐝𝐦pQ{\bf d_{m}}p\leq Q. To show that S^F𝑑σS^=\int_{\mathaccent 866{S}}F\,d\sigma_{\mathaccent 866{S}}=\infty, we fix a surface S^Σ\mathaccent 866{S}\in\Sigma and a neighbourhood UU of ee, such that S=S^US=\mathaccent 866{S}\cap U is an intrinsic Lipschitz graph. Then we argue as in Theorem 8. ∎

In the following sections we will study pp-exeptional sets of intrinsic Lipschitz surfaces passing through a common point when 𝐝𝐦p>Q{\bf d_{m}}p>Q. We will make special constructions of intrinsic Lipschitz surfaces that will reveal the situation of being pp-exceptional. The first step to the construction is the study of an analog of a Grassmann manifold on special type of Carnot groups.

4. Orthogonal Grassmannians

In this section we construct orthogonal Grassmannians of Lie subalgebras on specially chosen H-type Lie algebras. We start from the overview of Grassmannians of kk-plains of nn-dimensional Euclidean space, reminding that they are orbits under the action (modulo the isotropy subgroup) of the isometry group O(n)O(n) of the Euclidean space, see Section 4.1. In order to make a proper construction of the Grassmannian of subalgebras we first remind in Section 4.2 the structure of the isometry group Iso of H-type algebras. Then we make construction of orthogonal Grassmannians of subalgebras and study their properties in Section 4.3

4.1. Overview over the Grassmannians in 𝕂n\mathbb{K}^{n}

4.1.1. Definition of the groups O(n)\text{\rm O}(n), U(n)\text{\rm U}(n), and Sp(n)\text{\rm Sp}(n)

The orthogonal group O(n)\text{\rm O}(n) in n\mathbb{R}^{n}, endowed with the standard Euclidean inner product .,.\langle.\,,.\rangle_{\mathbb{R}}, is

O(n)={A:nn:Av,Av=v,v}.\text{\rm O}(n)=\{A\colon\mathbb{R}^{n}\to\mathbb{R}^{n}:\ \langle Av,Av\rangle_{\mathbb{R}}=\langle v,v\rangle_{\mathbb{R}}\}.

The unitary group U(n)\text{\rm U}(n) acting in n\mathbb{C}^{n}, endowed with the standard Hermitian inner product .,.\langle.\,,.\rangle_{\mathbb{C}}, is defined analogously by

U(n)={A:nn:Av,Av=v,v}.\text{\rm U}(n)=\{A\colon\mathbb{C}^{n}\to\mathbb{C}^{n}:\ \langle Av,Av\rangle_{\mathbb{C}}=\langle v,v\rangle_{\mathbb{C}}\}.

Finally, the quaternion unitary group (or compact symplectic group) Sp(n)\text{\rm Sp}(n) acting in right quaternion space n\mathbb{Q}^{n}, is defined by

Sp(n)={A:nn:Av,Av=v,v}.\text{\rm Sp}(n)=\{A\colon\mathbb{Q}^{n}\to\mathbb{Q}^{n}:\ \langle Av,Av\rangle_{\mathbb{Q}}=\langle v,v\rangle_{\mathbb{Q}}\}.

where .,.\langle.\,,.\rangle_{\mathbb{Q}} is a quaternion Hermitian product in n\mathbb{Q}^{n}, see, for instance [Ros06].

Let us write 𝕂\mathbb{K} for the division algebras \mathbb{R}, \mathbb{C}, or \mathbb{Q} and U(n,𝕂)\text{\rm U}(n,\mathbb{K}) for the groups O(n)\text{\rm O}(n), U(n)\text{\rm U}(n), and Sp(n)\text{\rm Sp}(n), respectively. We let kk be an integer satisfying 0<kn0<k\leq n. The Grassmann manifold or Grassmannian Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}) is defined as the set of kk-dimensional vector subspaces in 𝕂n\mathbb{K}^{n}:

Grk(𝕂n)={Vis akdimensional vector subspace of𝕂n}.\text{\rm Gr}_{k}(\mathbb{K}^{n})=\{V\ \text{is a}\ k-\text{dimensional vector subspace of}\ \ \mathbb{K}^{n}\}.

Note that the vector space n\mathbb{Q}^{n} is defined as the right vector space with respect to the right multiplication by scalars from \mathbb{Q}. The same agreement is done for the subspaces VnV\subset\mathbb{Q}^{n}.

The group U(n,𝕂)\text{\rm U}(n,\mathbb{K}) acts transitively on the set Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}) via

A.V={Av𝕂n:vV,AU(n,𝕂)}.A.V=\{Av\in\mathbb{K}^{n}:\ v\in V,\ A\in\text{\rm U}(n,\mathbb{K})\}.

Fix a plain V^Grk(𝕂n)\hat{V}\in\text{\rm Gr}_{k}(\mathbb{K}^{n}). Let KV^={AU(n,𝕂):A.V^=V^}K_{\hat{V}}=\{A\in\text{\rm U}(n,\mathbb{K}):\ A.\hat{V}=\hat{V}\} be the isotropy group of V^\hat{V}. It follows that the Grassmannian Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}) admits the structure of a compact manifold [War83] given by

Grk(𝕂n)=U(n,𝕂)/KV^,\text{\rm Gr}_{k}(\mathbb{K}^{n})=\text{\rm U}(n,\mathbb{K})/K_{\hat{V}},

where KV^K_{\hat{V}} is isomorphic to U(k,𝕂)×U(nk,𝕂)\text{\rm U}(k,\mathbb{K})\times\text{\rm U}(n-k,\mathbb{K}). Note that there is a diffeomorphism Grnk(𝕂n)Grk(𝕂n)\text{\rm Gr}_{n-k}(\mathbb{K}^{n})\cong\text{\rm Gr}_{k}(\mathbb{K}^{n}) mapping any VGrnk(𝕂n)V\in\text{\rm Gr}_{n-k}(\mathbb{K}^{n}) to its orthogonal complements VGrk(𝕂n)V^{\bot}\in\text{\rm Gr}_{k}(\mathbb{K}^{n}).

4.1.2. Measure on the Grassmannians

Let us remind the definition of a measure on the Grassmannian Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}). The continuous map

ψ:U(n,𝕂)Grk(𝕂n),\psi\colon\text{\rm U}(n,\mathbb{K})\to\text{\rm Gr}_{k}(\mathbb{K}^{n}),

which is the composition of the projection map to the quotient and the diffeomorpism giving the manifold structure to Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}) is used to push-forward a measure from U(n,𝕂)\text{\rm U}(n,\mathbb{K}) to Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}). We take for granted that the group U(n,𝕂)\text{\rm U}(n,\mathbb{K}) carries bi-invariant normalised measure λ\lambda:

λ(AU)=λ(UA)=λ(U),λ(U(n,𝕂))=1,\lambda(AU)=\lambda(UA)=\lambda(U),\qquad\lambda(\text{\rm U}(n,\mathbb{K}))=1,

for any Borel set UU(n,𝕂)U\subset\text{\rm U}(n,\mathbb{K}) and any isometry AU(n,𝕂)A\in\text{\rm U}(n,\mathbb{K}). The measure μ\mu on the Borel sets ΩGrk(𝕂n)\Omega\subset\text{\rm Gr}_{k}(\mathbb{K}^{n}) is defined as a pushforward of λ\lambda:

μ(Ω)=(ψλ)(Ω)=λ(ψ1(Ω))=λ{AU(n,𝕂):V=A.V^Ω}.\mu(\Omega)=(\psi_{\sharp}\lambda)(\Omega)=\lambda(\psi^{-1}(\Omega))=\lambda\{A\in\text{\rm U}(n,\mathbb{K}):\ V=A.\hat{V}\in\Omega\}.

It can be verified that μ\mu is normalised and it is U(n,𝕂)\text{\rm U}(n,\mathbb{K})-invariant. The converse is also true: a normalised U(n,𝕂)\text{\rm U}(n,\mathbb{K})-invariant measure on the homogeneous space Grk(𝕂n)\text{\rm Gr}_{k}(\mathbb{K}^{n}) is a push forward of the normalised Haar measure from U(n,𝕂)\text{\rm U}(n,\mathbb{K}), see for instance [Mat95].

Note that an (n1)(n-1)-dimensional sphere S(0,R)S(0,R) in n\mathbb{R}^{n} can be considered as a particular case of the Grassmanniann Grn1(n)\text{\rm Gr}_{n-1}(\mathbb{R}^{n}). If we denote by Kx()K_{x}(\mathbb{R}) an isotropy group of a point xS(0,R)x\in S(0,R) under the action of O(n)\text{\rm O}(n), then the following manifolds are diffeomorphic

S(0,R)Grn1(n)O(n)/Kx().S(0,R)\sim\text{\rm Gr}_{n-1}(\mathbb{R}^{n})\sim\text{\rm O}(n)/K_{x}(\mathbb{R}).

The push-forward of the normalised measure λ\lambda on O(n)\text{\rm O}(n) to S(0,R)S(0,R) coincides with the normalised surface measure on S(0,R)S(0,R), see [Mat95, Theorem 3.7].

4.2. Isometry groups of special HH-type Lie algebras

Before we make the construction of orthogonal Grassmannians on some special HH-type Lie algebras, we describe the group of isometries of these Lie algebras.

Recall the definition of an HH-type Lie algebra 𝔥=(𝔥1𝔥2,[.,.],.,.)\mathfrak{h}=(\mathfrak{h}_{1}\otimes\mathfrak{h}_{2},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}) from Section 2.1.3. The group of isometries Iso(𝔥)\text{\rm Iso}(\mathfrak{h}) of HH-type Lie algebras were studied in [Rie82, Rie84]. It was shown that

Iso(𝔥)={(A,C)O(𝔥1)×O(𝔥2):AtJzA=JCt(z)for anyz𝔥2}.\text{\rm Iso}(\mathfrak{h})=\{(A,C)\in\text{\rm O}(\mathfrak{h}_{1})\times\text{\rm O}(\mathfrak{h}_{2}):\ A^{t}J_{z}A=J_{C^{t}(z)}\quad\text{for any}\quad z\in\mathfrak{h}_{2}\}.

The group Iso(𝔥)\text{\rm Iso}(\mathfrak{h}) is isogenous to the product of the Pin group Pin(𝔥2,.,.)\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}}) of the Clifford algebra Cl(𝔥2,.,.)\text{\rm Cl}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}}) and a classical group 𝔸\mathbb{A}. The latter means that there is a surjective morphism

(39) ϕ:Pin(𝔥2,.,.)×𝔸Iso(𝔥)O(𝔥1)×O(𝔥2)θ=(α,A)ϕ(θ)=(JαA,κ(α)).\begin{array}[]{ccc}\phi\colon\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}})\times\mathbb{A}&\to&\text{\rm Iso}(\mathfrak{h})\subset\text{\rm O}(\mathfrak{h}_{1})\times\text{\rm O}(\mathfrak{h}_{2})\\ \theta=(\alpha,A)&\mapsto&\phi(\theta)=\big{(}J_{\alpha}\circ A,\kappa(\alpha)\big{)}.\end{array}

with a finite kernel of order 2 or 4. Here

(40) κ:Pin(𝔥2,.,.)O(𝔥2)\kappa\colon\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}})\to\text{\rm O}(\mathfrak{h}_{2})

is the double cover of the orthogonal group defined by

κ(α)v=αvα1,v𝔥2,αPin(𝔥2,.,.).\kappa(\alpha)v=\alpha v\alpha^{-1},\quad v\in\mathfrak{h}_{2},\ \ \alpha\in\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}}).

We will not give the full description of the group of isometries, but rather concentrate on the cases when the restriction Iso(𝔥)|𝔥1\text{\rm Iso}(\mathfrak{h})|_{\mathfrak{h}_{1}} on 𝔥1\mathfrak{h}_{1} acts transitively on the spheres S(0,r)𝔥1S(0,r)\in\mathfrak{h}_{1} and the vector space 𝔥1\mathfrak{h}_{1}, considered as a Clifford module, is not irreducible. Only in these cases the construction of the Grassmannian on the Lie algebra 𝔥\mathfrak{h} is not trivial. Thus we will consider the following HH-type Lie algebras:

  • RH:

    The Heisenberg algebra 𝔥n=(2n+1,[.,.],.,.)\mathfrak{h}_{\mathbb{R}}^{n}=(\mathbb{R}^{2n+1},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}), n>1n>1 with 𝔥=𝔥1𝔥22n\mathfrak{h}_{\mathbb{R}}=\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}\cong\mathbb{R}^{2n}\otimes\mathbb{R};

  • CH:

    The complex Heisenberg algebra 𝔥n=(4n+2,[.,.],.,.)\mathfrak{h}_{\mathbb{C}}^{n}=(\mathbb{R}^{4n+2},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}), n>1n>1 with 𝔥=𝔥1𝔥24n2\mathfrak{h}_{\mathbb{C}}=\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}\cong\mathbb{R}^{4n}\otimes\mathbb{R}^{2};

  • QH:

    The quaternion Heisenberg algebra 𝔥n=(4n+3,[.,.],.,.)\mathfrak{h}_{\mathbb{Q}}^{n}=(\mathbb{R}^{4n+3},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}), n>1n>1 with 𝔥=𝔥1𝔥24n3\mathfrak{h}_{\mathbb{Q}}=\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}\cong\mathbb{R}^{4n}\otimes\mathbb{R}^{3}.

The commutators in all the cases are defined by (10) and the scalar products are the standard Euclidean products making the direct sums orthogonal. The names real, complex, and quaternion are attached to the names of the algebras by the following reasons. For the Lie algebra 𝔥n\mathfrak{h}_{\mathbb{R}}^{n}, the group of real symplectic transformations Sp(n,)\text{\rm Sp}(n,\mathbb{R}) is a subgroup of automorphisms, leaving the center 𝔥2\mathfrak{h}_{2} of the Lie algebra 𝔥n\mathfrak{h}_{\mathbb{R}}^{n} invariant. For the Lie algebra 𝔥n\mathfrak{h}_{\mathbb{C}}^{n}, the group of complex symplectic transformations Sp(2n,)\text{\rm Sp}(2n,\mathbb{C}) is a subgroup of the Lie algebra automorphisms, leaving the center invariant and also 𝔥n\mathfrak{h}_{\mathbb{C}}^{n} is isomorphic to the complexification of 𝔥n\mathfrak{h}_{\mathbb{R}}^{n}. For 𝔥n\mathfrak{h}_{\mathbb{Q}}^{n} the group of quaternion unitary transformations Sp(n)\text{\rm Sp}(n) is a subgroup of the Lie algebra automorphisms, leaving the center invariant. The center 𝔥2\mathfrak{h}_{2} of 𝔥n\mathfrak{h}_{\mathbb{Q}}^{n} is isomorphic to the space of pure imaginary quaternions.

Now we describe the isometry groups in each case.

4.2.1. The isometry group Iso(𝔥)\text{\rm Iso}(\mathfrak{h}_{\mathbb{R}})

Recall the definition of the Heisenberg algebra in Section 2.1.4. We write 𝔥=𝔥1𝔥2\mathfrak{h}_{\mathbb{R}}=\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}. Let ϵ𝔥2\epsilon\in\mathfrak{h}_{2}\cong\mathbb{R} be a vector satisfying ϵ,ϵ=1\langle\epsilon,\epsilon\rangle_{\mathbb{R}}=1. The vector space 𝔥12n\mathfrak{h}_{1}\cong\mathbb{R}^{2n} carries natural almost complex structure given by the action of JϵJ_{\epsilon}, ϵ𝔥2\epsilon\in\mathfrak{h}_{2}. Therefore, the space 𝔥12n\mathfrak{h}_{1}\cong\mathbb{R}^{2n} can be considered as a complex vector space 𝔥1n\mathfrak{h}_{1}\cong\mathbb{C}^{n}, where the multiplication by a complex number α\alpha\in\mathbb{C} is defined by αv=(a+ib)v=av+bJϵv\alpha v=(a+ib)v=av+bJ_{\epsilon}v, v𝔥1v\in\mathfrak{h}_{1}. The maximal compact subgroup 𝔸=U(n)Sp(2n,)\mathbb{A}=\text{\rm U}(n)\subset\text{\rm Sp}(2n,\mathbb{R}) is an isometry on 𝔥12n\mathfrak{h}_{1}\cong\mathbb{R}^{2n}. The corresponding Hermitian form .|.\langle.|.\rangle_{\mathbb{C}} on 𝔥1n\mathfrak{h}_{1}\cong\mathbb{C}^{n} is defined by making use of the real scalar product .,.\langle.\,,.\rangle_{\mathbb{R}} by the following

(41) u|v=u,viJϵu,v.\langle u|v\rangle_{\mathbb{C}}=\langle u,v\rangle_{\mathbb{R}}-i\langle J_{\epsilon}u,v\rangle_{\mathbb{R}}.

The imaginary part of the Hermitian product (41) defines the symplectic form ω:𝔥1×𝔥1\omega\colon\mathfrak{h}_{1}\times\mathfrak{h}_{1}\to\mathbb{R}, that is compatible with the real inner product on 𝔥1\mathfrak{h}_{1}. Namely

(42) ω(u,v)=Jϵu,v=[u,v],ϵ.\omega(u,v)=\langle J_{\epsilon}u,v\rangle_{\mathbb{R}}=\langle[u,v],\epsilon\rangle_{\mathbb{R}}.

which implies ω(Jϵu,Jϵv)=ω(u,v),ω(u,Jϵv)=u,v.\omega(J_{\epsilon}u,J_{\epsilon}v)=\omega(u,v),\quad\omega(u,J_{\epsilon}v)=\langle u,v\rangle_{\mathbb{R}}.

We use the basis described in Section 2.1.4 and give the coordinates to 𝔥n\mathfrak{h}^{n}_{\mathbb{R}}. Consider the product of two spheres on 𝔥n2n×\mathfrak{h}^{n}_{\mathbb{R}}\cong\mathbb{R}^{2n}\times\mathbb{R} centered at the origin

(43) 𝒮(0,r1,r2)\displaystyle\mathcal{S}_{\mathbb{R}}(0,r_{1},r_{2}) =\displaystyle= S2n1(0,r1)×S0(0,r2)\displaystyle S^{2n-1}(0,r_{1})\times S^{0}(0,r_{2})
=\displaystyle= {v𝔥1:v,v=r12}×{z𝔥2,z,z=r22}\displaystyle\{v\in\mathfrak{h}_{1}:\ \langle v,v\rangle_{\mathbb{R}}=r_{1}^{2}\}\times\{z\in\mathfrak{h}_{2},\langle z,z\rangle_{\mathbb{R}}=r_{2}^{2}\}
Lemma 4.

The group Iso(𝔥n)\text{\rm Iso}(\mathfrak{h}_{\mathbb{R}}^{n}) acts transitively on 𝒮(0,r1,r2)\mathcal{S}_{\mathbb{R}}(0,r_{1},r_{2}).

Proof.

The map κ(ϵ)O(𝔥2)\kappa(\epsilon)\in\text{\rm O}(\mathfrak{h}_{2}), defined in (40) acts either as a reflection with respect to the origin of 𝔥2\mathfrak{h}_{2} or as the identity, since 𝔥2\mathfrak{h}_{2}\cong\mathbb{R}. Let (x1,z),(x2,z)𝒮(0,r1,r2)(x_{1},z),(x_{2},-z)\in\mathcal{S}_{\mathbb{R}}(0,r_{1},r_{2}). Then we find a map (JϵA,κ(ϵ))Iso(𝔥n)(J_{\epsilon}\circ A,\kappa(\epsilon))\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{R}}) that maps the point (x1,z)(x_{1},z) to (x3,z)=(JϵAx1,κ(ϵ)z)(x_{3},-z)=(J_{\epsilon}\circ Ax_{1},\kappa(\epsilon)z), AU(n)A\in\text{\rm U}(n). Since the subgroup (U(n)×Id𝔥2)Iso(𝔥n)(\text{\rm U}(n)\times\text{\rm Id}_{\mathfrak{h}_{2}})\subset\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{R}}) acts transitively on S2n1(0,r1)S^{2n-1}(0,r_{1}) we can find the transformation mapping (x3,z)(x_{3},-z) to (x2,z)(x_{2},-z). ∎

At the end we recall that the basis of left invariant vector fields on the Heisenberg group HnH_{\mathbb{R}}^{n} are

(44) X1k=x1kx2k2ϵ1,X2k=x2k+x1k2ϵ2,X_{1k}=\frac{\partial}{\partial x_{1k}}-\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{1}},\quad X_{2k}=\frac{\partial}{\partial x_{2k}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{2}},

for k=1,,nk=1,\ldots,n.

4.2.2. Isometry groups of H-type Lie algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{C}}

Let 𝔥14n𝔥n\mathfrak{h}_{1}\cong\mathbb{R}^{4n}\subset\mathfrak{h}^{n}_{\mathbb{C}} be the horizontal subspace and 𝔥22𝔥n\mathfrak{h}_{2}\cong\mathbb{R}^{2}\subset\mathfrak{h}^{n}_{\mathbb{C}} be the vertical subspace. Then Cl(𝔥2,.,.)Cl(2)\text{\rm Cl}(\mathfrak{h}_{2},\langle.\,,.\rangle_{\mathbb{R}})\cong\text{\rm Cl}(\mathbb{R}^{2}) contains two elements ϵ1,ϵ2\epsilon_{1},\epsilon_{2} such that

ϵ12=ϵ22=1,ϵ1,ϵ1=ϵ2,ϵ2=1,ϵ1,ϵ2=0,\epsilon_{1}^{2}=\epsilon_{2}^{2}=-1,\quad\langle\epsilon_{1},\epsilon_{1}\rangle_{\mathbb{R}}=\langle\epsilon_{2},\epsilon_{2}\rangle_{\mathbb{R}}=1,\quad\langle\epsilon_{1},\epsilon_{2}\rangle_{\mathbb{R}}=0,

and moreover ϵ3=ϵ1ϵ2\epsilon_{3}=\epsilon_{1}\epsilon_{2} satisfies ϵ32=1\epsilon_{3}^{2}=-1.

We first consider the case n=1n=1, that is 𝔥14𝔥1\mathfrak{h}_{1}\cong\mathbb{R}^{4}\subset\mathfrak{h}^{1}_{\mathbb{C}}. The maps Jϵi:𝔥1𝔥1J_{\epsilon_{i}}\colon\mathfrak{h}_{1}\to\mathfrak{h}_{1}, i=1,2i=1,2, are orthogonal transformations. A convenient orthonormal basis for 𝔥1\mathfrak{h}_{1} can be constructed by the following. We choose a vector v𝔥1v\in\mathfrak{h}_{1}, such that v,v=1\langle v\,,v\rangle_{\mathbb{R}}=1 and define the orthonormal vectors

(45) X1=v,X2=Jϵ1v,X3=Jϵ2v,X4=Jϵ2Jϵ1v.X_{1}=v,\quad X_{2}=J_{\epsilon_{1}}v,\quad X_{3}=J_{\epsilon_{2}}v,\quad X_{4}=J_{\epsilon_{2}}J_{\epsilon_{1}}v.

The commutation relations according to (10) are

(46) [X1,X2]=[X3,X4]=ϵ1,[X1,X3]=[X2,X4]=ϵ2.[X_{1},X_{2}]=-[X_{3},X_{4}]=\epsilon_{1},\quad[X_{1},X_{3}]=[X_{2},X_{4}]=\epsilon_{2}.

We show now that 𝔥1\mathfrak{h}^{1}_{\mathbb{C}} is the complexified Lie algebra of 𝔥1\mathfrak{h}^{1}_{\mathbb{R}}. Let X1,X2=JϵX1X_{1},X_{2}=J_{\epsilon}X_{1} be an orthonormal basis of 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{R}}^{1} and 𝔥2=span{ϵ}\mathfrak{h}_{2}=\text{\rm span}\,_{\mathbb{R}}\{\epsilon\} is the center of the Lie algebra 𝔥1\mathfrak{h}^{1}_{\mathbb{R}}. Then the complexification 𝔥1=𝔥1\mathfrak{h}_{1}^{\mathbb{C}}=\mathbb{C}\otimes\mathfrak{h}_{1} of the vector space 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}^{1}_{\mathbb{R}} can be described as any of the following direct sums

𝔥1\displaystyle\mathfrak{h}_{1}^{\mathbb{C}} =\displaystyle= span{Z=X1+iX2}span{Z¯=X1iX2}\displaystyle\text{\rm span}\,_{\mathbb{R}}\{Z=X_{1}+iX_{2}\}\oplus\text{\rm span}\,_{\mathbb{R}}\{\bar{Z}=X_{1}-iX_{2}\}
=\displaystyle= span{X1,X2}span{iX1,iX2}\displaystyle\text{\rm span}\,_{\mathbb{R}}\{X_{1},X_{2}\}\oplus\text{\rm span}\,_{\mathbb{R}}\{iX_{1},iX_{2}\}
=\displaystyle= span{X1,X2}.\displaystyle\text{\rm span}\,_{\mathbb{C}}\{X_{1},X_{2}\}.

The Lie bracket on 𝔥1\mathfrak{h}_{1}^{\mathbb{C}} is a complex linear Lie bracket on 𝔥1\mathfrak{h}_{1}:

(48) [X1,X2]=[iX1,iX2]=ϵ,[X1,iX2]=[iX1,X2]=iϵ.[X_{1},X_{2}]=-[iX_{1},iX_{2}]=\epsilon,\quad[X_{1},iX_{2}]=[iX_{1},X_{2}]=i\epsilon.

Thus the center 𝔥2\mathfrak{h}_{2}^{\mathbb{C}} of the complexified Heisenberg algebra is given by

(49) 𝔥2=span{ϵ}=span{ϵ,iϵ}.\displaystyle\mathfrak{h}_{2}^{\mathbb{C}}=\text{\rm span}\,_{\mathbb{C}}\{\epsilon\}=\text{\rm span}\,_{\mathbb{R}}\{\epsilon,i\epsilon\}.

Recall that the real Heisenberg algebra has the complex structure Jϵ:𝔥1𝔥1J_{\epsilon}\colon\mathfrak{h}_{1}\to\mathfrak{h}_{1} defined by JϵX1=X2J_{\epsilon}X_{1}=X_{2}, JϵX2=X1J_{\epsilon}X_{2}=-X_{1}. We extend JϵJ_{\epsilon} to 𝔥1\mathfrak{h}_{1}^{\mathbb{C}} by linearity, meaning that Jϵ(iu)=iJϵ(u)J_{\epsilon}(iu)=iJ_{\epsilon}(u) for uspan{X1,X2}u\in\text{\rm span}\,_{\mathbb{R}}\{X_{1},X_{2}\}. We define another complex structure

Jiϵ:𝔥1𝔥1:JiϵX1=iX2,JiϵX2=iX1.J_{i\epsilon}\colon\mathfrak{h}_{1}^{\mathbb{C}}\to\mathfrak{h}_{1}^{\mathbb{C}}:\quad J_{i\epsilon}X_{1}=iX_{2},\quad J_{i\epsilon}X_{2}=iX_{1}.

It is easy to check that JϵJiϵ=JiϵJϵJ_{\epsilon}J_{i\epsilon}=-J_{i\epsilon}J_{\epsilon}. Note that if we denote

(50) ϵ1=ϵ,ϵ2=iϵ,\displaystyle\epsilon_{1}=\epsilon,\quad\epsilon_{2}=i\epsilon,
X1=X1,X2=Jϵ1X1,X3=iX2=Jϵ2X1,X4=iX1=Jϵ2Jϵ1X1,\displaystyle X_{1}=X_{1},\quad X_{2}=J_{\epsilon_{1}}X_{1},\quad X_{3}=iX_{2}=J_{\epsilon_{2}}X_{1},\quad X_{4}=iX_{1}=J_{\epsilon_{2}}J_{\epsilon_{1}}X_{1},

then we recover basis (45) and commutation relations (46) of 𝔥1\mathfrak{h}_{\mathbb{C}}^{1}.

We show now that the horizontal space 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{C}}^{1} of the complexified Heisenberg algebra is a complex symplectic space. The real symplectic form ω\omega defined in (42) can be extended to the complex symplectic form ω:𝔥1×𝔥1\omega^{\mathbb{C}}\colon\mathfrak{h}_{1}\times{\mathfrak{h}_{1}}\to\mathbb{C}, 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{C}}^{1}, by

(51) ω(u+iv,x+iy)=ω(u,x)+ω(v,y)+i(ω(v,x)+ω(y,u)).\omega^{\mathbb{C}}(u+iv,x+iy)=\omega(u,x)+\omega(v,y)+i\big{(}\omega(v,x)+\omega(y,u)\big{)}.

Here we consider 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{C}}^{1} as a complex space (4.2.2). Then ω\omega^{\mathbb{C}} satisfies

ω(z1,z2)=ω(z2,z1)¯,\omega^{\mathbb{C}}(z_{1},z_{2})=-\overline{\omega^{\mathbb{C}}(z_{2},z_{1})},
ω(αz1,z2)=αω(z1,z2),ω(z1,αz2)=α¯ω(z1,z2).\omega^{\mathbb{C}}(\alpha z_{1},z_{2})=\alpha\omega^{\mathbb{C}}(z_{1},z_{2}),\quad\omega^{\mathbb{C}}(z_{1},\alpha z_{2})=\bar{\alpha}\omega^{\mathbb{C}}(z_{1},z_{2}).

The symplectic basis X1,X2=Jϵ1X1X_{1},X_{2}=J_{\epsilon_{1}}X_{1} over \mathbb{R} for the real symplectic vector space 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{R}}^{1} is the symplectic basis over \mathbb{C} for the complex symplectic vector space 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}^{1}_{\mathbb{C}}.

The multidimensional algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} as a vector space is the Cartesian product

(52) 𝔥n=(𝔥1)1××(𝔥1)n×𝔥2,\mathfrak{h}^{n}_{\mathbb{C}}=\big{(}\mathfrak{h}_{1}^{\mathbb{C}}\big{)}_{1}\times\ldots\times\big{(}\mathfrak{h}_{1}^{\mathbb{C}}\big{)}_{n}\times\mathfrak{h}_{2}^{\mathbb{C}},

where 𝔥1\mathfrak{h}_{1}^{\mathbb{C}} is defined in (4.2.2) and 𝔥2\mathfrak{h}_{2}^{\mathbb{C}} is defined in (49). The commutation relations are given by (48), if the vectors belong to the same slott (𝔥1)k\big{(}\mathfrak{h}_{1}^{\mathbb{C}}\big{)}_{k} in the Cartesian product (59) and zero otherwise. The real scalar product .,.\langle.\,,.\rangle_{\mathbb{R}} is extended to the Cartesian product (59) by making the different slots orthogonal. The multidimensional algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} has real topological dimension 4n+24n+2 and the Hausdorff dimension 4n+44n+4. The corresponding left invariant vector fields on the complexified Heisenberg group HnH_{\mathbb{C}}^{n} are

(53) X1k=x1kx2k2ϵ1x3k2ϵ2,X2k=x2k+x1k2ϵ1x4k2ϵ2,X3k=x3k+x4k2ϵ1+x1k2ϵ2,X4k=x4kx3k2ϵ1+x2k2ϵ2,k=1,,n.\begin{array}[]{ccccccc}X_{1k}&=&\frac{\partial}{\partial x_{1k}}-\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{1}}-\frac{x_{3k}}{2}\frac{\partial}{\partial\epsilon_{2}},\\ X_{2k}&=&\frac{\partial}{\partial x_{2k}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{1}}-\frac{x_{4k}}{2}\frac{\partial}{\partial\epsilon_{2}},\\ X_{3k}&=&\frac{\partial}{\partial x_{3k}}+\frac{x_{4k}}{2}\frac{\partial}{\partial\epsilon_{1}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{2}},\\ X_{4k}&=&\frac{\partial}{\partial x_{4k}}-\frac{x_{3k}}{2}\frac{\partial}{\partial\epsilon_{1}}+\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{2}},\end{array}\qquad k=1,\ldots,n.

The subgroup 𝔸Sp(n)\mathbb{A}\cong\text{\rm Sp}(n) of the isometry group Iso(𝔥n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{C}}) preserves the quaternion Hermitian product .|.\langle.|.\rangle_{\mathbb{Q}} on 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{C}}, defined by

(54) z|w=z,wiJϵ1z,wjJϵ2z,wkJϵ2Jϵ1z,w.\langle z|w\rangle_{\mathbb{Q}}=\langle z,w\rangle_{\mathbb{R}}-i\langle J_{\epsilon_{1}}z,w\rangle_{\mathbb{R}}-j\langle J_{\epsilon_{2}}z,w\rangle_{\mathbb{R}}-k\langle J_{\epsilon_{2}}J_{\epsilon_{1}}z,w\rangle_{\mathbb{R}}.

The group 𝔸Sp(n)\mathbb{A}\cong\text{\rm Sp}(n) acts transitively on the unit sphere S4n1𝔥12nS^{4n-1}\subset\mathfrak{h}_{1}\cong\mathbb{C}^{2n}, where 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{C}}.

4.2.3. Isometry groups of H-type Lie algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}}

Since 3𝔥2𝔥1\mathbb{R}^{3}\cong\mathfrak{h}_{2}\subset\mathfrak{h}^{1}_{\mathbb{Q}}, there is three linearly independent length one elements ϵiCl(3,.,.)\epsilon_{i}\in\text{\rm Cl}(\mathbb{R}^{3},\langle.\,,.\rangle_{\mathbb{R}}), i=1,2,3i=1,2,3, satisfying the quaternion relations

(55) ϵ12=ϵ32=ϵ32=ϵ1ϵ2ϵ3=1.\epsilon_{1}^{2}=\epsilon_{3}^{2}=\epsilon_{3}^{2}=\epsilon_{1}\epsilon_{2}\epsilon_{3}=-1.

We introduce a quaternion structure on 4𝔥1𝔥1\mathbb{R}^{4}\cong\mathfrak{h}_{1}\subset\mathfrak{h}^{1}_{\mathbb{Q}} by defining the multiplication by a quaternion number q=a+ib+jc+kdq=a+ib+jc+kd\in\mathbb{Q} as follows

qv=(a+ib+jc+kd)v=av+bJϵ1v+cJϵ2v+dJϵ3v,v𝔥1.qv=(a+ib+jc+kd)v=av+bJ_{\epsilon_{1}}v+cJ_{\epsilon_{2}}v+dJ_{\epsilon_{3}}v,\quad v\in\mathfrak{h}_{1}.

The quaternion Hermitian product is

(56) u|v=u,viJϵ1u,vjJϵ2u,vkJϵ3u,v.\langle u|v\rangle_{\mathbb{Q}}=\langle u,v\rangle_{\mathbb{R}}-i\langle J_{\epsilon_{1}}u,v\rangle_{\mathbb{R}}-j\langle J_{\epsilon_{2}}u,v\rangle_{\mathbb{R}}-k\langle J_{\epsilon_{3}}u,v\rangle_{\mathbb{R}}.

To construct an orthonormal basis we choose v𝔥1v\in\mathfrak{h}_{1} with v,v=1\langle v,v\rangle_{\mathbb{R}}=1 and set

(57) X1=v,X2=Jϵ1v,X3=Jϵ2v,X4=Jϵ3v.X_{1}=v,\quad X_{2}=J_{\epsilon_{1}}v,\quad X_{3}=J_{\epsilon_{2}}v,\quad X_{4}=J_{\epsilon_{3}}v.

The commutation relations are

(58) [X1,X2]=[X3,X4]=ϵ1,[X1,X3]=[X2,X4]=ϵ2,\displaystyle[X_{1},X_{2}]=-[X_{3},X_{4}]=\epsilon_{1},\quad[X_{1},X_{3}]=[X_{2},X_{4}]=\epsilon_{2},
[X1,X4]=[X2,X3]=ϵ3.\displaystyle[X_{1},X_{4}]=-[X_{2},X_{3}]=\epsilon_{3}.

The space 𝔥1=span{X1,X2,X3,X4}𝔥1\mathfrak{h}_{1}=\text{\rm span}\,_{\mathbb{R}}\{X_{1},X_{2},X_{3},X_{4}\}\subset\mathfrak{h}^{1}_{\mathbb{C}} is isomorphic and isometric to the 1-dimensional quaternion space endowed by the quaternion Hermitian product (56).

We notice the relation between 𝔥1\mathfrak{h}^{1}_{\mathbb{C}} and 𝔥1\mathfrak{h}^{1}_{\mathbb{Q}}. Due to (55), the action of Jϵ3J_{\epsilon_{3}} can be expressed as Jϵ3=Jϵ1Jϵ2J_{\epsilon_{3}}=J_{\epsilon_{1}}J_{\epsilon_{2}} and therefore the space 𝔥1𝔥1\mathfrak{h}_{1}\subset\mathfrak{h}^{1}_{\mathbb{Q}} can be considered as a complex symplectic space with respect to ω\omega^{\mathbb{C}} in (51). In addition to that we add a real symplectic form ω3(u,v)=Jϵ3u,v.\omega_{3}(u,v)=\langle J_{\epsilon_{3}}u,v\rangle_{\mathbb{R}}.

The multidimensional algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} as a vector space is the Cartesian product

(59) 𝔥n=(𝔥1)1××(𝔥1)n×𝔥2.\mathfrak{h}^{n}_{\mathbb{Q}}=\big{(}\mathfrak{h}_{1}\big{)}_{1}\times\ldots\times\big{(}\mathfrak{h}_{1}\big{)}_{n}\times\mathfrak{h}_{2}.

The commutation relations are given by (58), if the vectors belong to the same slott (𝔥1)k\big{(}\mathfrak{h}_{1}\big{)}_{k} in the Cartesian product (59) and zero otherwise. The multidimensional algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} has real topological dimension 4n+34n+3 and the Hausdorff dimension 4n+64n+6. We note that the Clifford module (𝔥1)1××(𝔥1)n\big{(}\mathfrak{h}_{1}\big{)}_{1}\times\ldots\times\big{(}\mathfrak{h}_{1}\big{)}_{n} in this case is assumed to be isotypic, that corresponds the fact that the product Jϵ1Jϵ2Jϵ3J_{\epsilon_{1}}J_{\epsilon_{2}}J_{\epsilon_{3}} acts as minus identity on every slot (𝔥1)k\big{(}\mathfrak{h}_{1}\big{)}_{k}. The subgroup 𝔸Sp(n)\mathbb{A}\cong\text{\rm Sp}(n) of the isometry group Iso(𝔥n)\text{\rm Iso}(\mathfrak{h}_{\mathbb{Q}}^{n}) preserves the quaternion Hermitian product on 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{Q}}, defined in (56). The corresponding basis of left invariant vector fields on the group HnH_{\mathbb{Q}}^{n} is

(60) X1k=x1kx2k2ϵ1x3k2ϵ2x4k2ϵ3X2k=x2k+x1k2ϵ1x4k2ϵ2+x3k2ϵ3,X3k=x3k+x4k2ϵ1+x1k2ϵ2x2k2ϵ3,X4k=x4kx3k2ϵ1+x2k2ϵ2+x1k2ϵ3,k=1,,n.\begin{array}[]{ccccccc}X_{1k}&=&\frac{\partial}{\partial x_{1k}}-\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{1}}-\frac{x_{3k}}{2}\frac{\partial}{\partial\epsilon_{2}}-\frac{x_{4k}}{2}\frac{\partial}{\partial\epsilon_{3}}\\ X_{2k}&=&\frac{\partial}{\partial x_{2k}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{1}}-\frac{x_{4k}}{2}\frac{\partial}{\partial\epsilon_{2}}+\frac{x_{3k}}{2}\frac{\partial}{\partial\epsilon_{3}},\\ X_{3k}&=&\frac{\partial}{\partial x_{3k}}+\frac{x_{4k}}{2}\frac{\partial}{\partial\epsilon_{1}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{2}}-\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{3}},\\ X_{4k}&=&\frac{\partial}{\partial x_{4k}}-\frac{x_{3k}}{2}\frac{\partial}{\partial\epsilon_{1}}+\frac{x_{2k}}{2}\frac{\partial}{\partial\epsilon_{2}}+\frac{x_{1k}}{2}\frac{\partial}{\partial\epsilon_{3}},\end{array}\qquad k=1,\ldots,n.

We use the bases (45) and (57) to identify 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} with 4n×2\mathbb{R}^{4n}\times\mathbb{R}^{2} and 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} with 4n×3\mathbb{R}^{4n}\times\mathbb{R}^{3}. Define the product of spheres

𝒮(0,r1,r2)\displaystyle\mathcal{S}_{\mathbb{C}}(0,r_{1},r_{2}) =\displaystyle= {(x,z)𝔥n:xS4n1(0,r1)𝔥1,zS1(0,r2)𝔥2}\displaystyle\{(x,z)\in\mathfrak{h}^{n}_{\mathbb{C}}:\ x\in S^{4n-1}(0,r_{1})\subset\mathfrak{h}_{1},z\in S^{1}(0,r_{2})\subset\mathfrak{h}_{2}\}
\displaystyle\cong S4n1(0,r1)×S1(0,r2),𝒮(0,r1,r2)𝔥n\displaystyle S^{4n-1}(0,r_{1})\times S^{1}(0,r_{2}),\quad\mathcal{S}_{\mathbb{C}}(0,r_{1},r_{2})\subset\mathfrak{h}^{n}_{\mathbb{C}}

and

𝒮(0,r1,r2)\displaystyle\mathcal{S}_{\mathbb{Q}}(0,r_{1},r_{2}) =\displaystyle= {(x,z)𝔥n:xS4n1(0,r1)𝔥1,zS2(0,r2)𝔥2}\displaystyle\{(x,z)\in\mathfrak{h}^{n}_{\mathbb{Q}}:\ x\in S^{4n-1}(0,r_{1})\subset\mathfrak{h}_{1},z\in S^{2}(0,r_{2})\subset\mathfrak{h}_{2}\}
\displaystyle\cong S4n1(0,r1)×S2(0,r2),𝒮(0,r1,r2)𝔥n.\displaystyle S^{4n-1}(0,r_{1})\times S^{2}(0,r_{2}),\quad\mathcal{S}_{\mathbb{Q}}(0,r_{1},r_{2})\subset\mathfrak{h}^{n}_{\mathbb{Q}}.
Lemma 5.

The groups of isometries Iso(𝔥n)\text{\rm Iso}(\mathfrak{h}_{\mathbb{C}}^{n}) and Iso(𝔥n)\text{\rm Iso}(\mathfrak{h}_{\mathbb{Q}}^{n}) act transitively on the respective products of spheres 𝒮(0,r1,r2)\mathcal{S}_{\mathbb{C}}(0,r_{1},r_{2}) and 𝒮(0,r1,r2)\mathcal{S}_{\mathbb{Q}}(0,r_{1},r_{2}).

4.3. Grassmannians on special HH-type Lie algebras

We will construct orthogonal Grassmannians on the H-type Lie algebras mentioned in Section 4.2. In these cases the isometry groups act transitively on the spheres 𝒮𝕂(0,r1,r2)\mathcal{S}_{\mathbb{K}}(0,r_{1},r_{2}), 𝕂=,\mathbb{K}=\mathbb{R},\mathbb{C}, or \mathbb{Q}. This allows to define the measure on the Grassmannians. Moreover, the transitive action permits to realise the Grassmannians as orbit spaces under the action of the isometry groups. Note that the spaces n\mathbb{R}^{n}, n\mathbb{C}^{n}, and n\mathbb{Q}^{n} are Abelian algebras with respect to the summation. Any kk dimensional vector subspace VV is a subalgebra that has nkn-k dimensional orthogonal complement that is also a subalgebra. This property is not trivial for the non-commutative subalgebras and therefore we restrict ourself to the consideration of orthogonally complemented Grassmannians.

4.3.1. Orthogonally complemented Grassmannians

A subalgebra V𝔥𝕂nV\subset\mathfrak{h}^{n}_{\mathbb{K}} is called homogeneous if it is invariant under the action of dilation (3). In the following definition we use the inner product from the definition of HH-type Lie algebra 𝔥=(𝔥1𝔥2,[.,.],.,.)\mathfrak{h}=(\mathfrak{h}_{1}\oplus\mathfrak{h}_{2},[.\,,.],\langle.\,,.\rangle_{\mathbb{R}}).

Definition 8.

We say that a homogeneous subalgebra V𝔥𝕂nV\subset\mathfrak{h}^{n}_{\mathbb{K}} is an orthogonally complemented homogeneous subalgebra of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}} if the orthogonal complement VV^{\perp} is a homogeneous subalgebra of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}.

Lemma 6.

Let V𝔥𝕂nV\subset\mathfrak{h}^{n}_{\mathbb{K}} be an orthogonally complemented homogeneous subalgebra. Then:

  • i)

    In the case 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}} for 𝕂=,\mathbb{K}=\mathbb{R},\mathbb{H} we have

    • i-1)

      if dimVn\dim_{\mathbb{R}}V\leq n, then V𝔥1V\subset\mathfrak{h}_{1} (and hence VV is commutative);

    • i-2)

      if dimV>n\dim_{\mathbb{R}}V>n, then 𝔥2V\mathfrak{h}_{2}\subset V.

  • ii)

    In the case 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} we have

    • ii-1)

      if dimV2n\dim_{\mathbb{R}}V\leq 2n, then V𝔥1V\subset\mathfrak{h}_{1} (and hence VV is commutative);

    • ii-2)

      if dimV>2n\dim_{\mathbb{R}}V>2n, then 𝔥2V\mathfrak{h}_{2}\subset V.

Proof.

We start from the case 𝔥n\mathfrak{h}^{n}_{\mathbb{R}}. We sketch the construction of the complementary homogeneous subalgebras. The horizontal vector space 𝔥1\mathfrak{h}_{1} is a symplectic vector space with the symplectic form (42), see Section 4.2.1. Let Wh1W\subset h_{1} be a vector space of dimension d=dim(W)12dim(𝔥1)=nd=\dim_{\mathbb{R}}(W)\geq\frac{1}{2}\dim_{\mathbb{R}}(\mathfrak{h}_{1})=n. It was shown in [FSSC07, Lemma 3.26], that there is a vector space W𝔥1W^{\prime}\subset\mathfrak{h}_{1} such that WW=𝔥1W^{\prime}\oplus W=\mathfrak{h}_{1} and ω(v,w)=0\omega(v,w)=0 for all v,wWv,w\in W^{\prime}. The relation between the symplectic form and the commutation relation shows that WW^{\prime} is a commutative subalgebra of 𝔥1\mathfrak{h}_{1}. We choose an orthonormal basis {e1,,ed}\{e_{1},\ldots,e_{d^{\prime}}\}, d=dim(W)nd^{\prime}=\dim(W^{\prime})\leq n for WW^{\prime} and extend it to an orthonormal basis

{e1,,ed,ed+1,,en,f1=Jϵe1,,fn=Jϵen}.\{e_{1},\ldots,e_{d^{\prime}},e_{d^{\prime}+1},\ldots,e_{n},f_{1}=J_{\epsilon}e_{1},\ldots,f_{n}=J_{\epsilon}e_{n}\}.

We denote

V=W,V=span{ed+1,,en,f1=Jϵe1,,fn=Jϵen}span{ϵ}.V=W^{\prime},\quad V^{\prime}=\text{\rm span}\,\{e_{d^{\prime}+1},\ldots,e_{n},f_{1}=J_{\epsilon}e_{1},\ldots,f_{n}=J_{\epsilon}e_{n}\}\oplus\text{\rm span}\,\{\epsilon\}.

Then it is easy to see that VVV\oplus V^{\prime} are orthogonally complemented subalgebras, satisfying hypothesis of Lemma 6. Note that the construction above gives all possible complementary subalgebras.

We turn to 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} and will consider 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{C}} as a complex symplectic space with the complex symplectic form ω\omega^{\mathbb{C}}. We have dim(𝔥1)=2n\dim_{\mathbb{C}}(\mathfrak{h}_{1})=2n and nn is even. We can show that for any WW such that dim(W)n\dim_{\mathbb{C}}(W)\geq n there is a vector space W𝔥1W^{\prime}\subset\mathfrak{h}_{1} satisfying WW=𝔥1W^{\prime}\oplus W=\mathfrak{h}_{1} and ω(v,w)=0\omega^{\mathbb{C}}(v,w)=0 for all v,wWv,w\in W^{\prime}. The arguments are the same as at the beginning of the proof, since the arguments do not depend on the choice of the fields \mathbb{R} or \mathbb{C}, but only on the construction of the basis.

Thus, by the construction, the vector space WW^{\prime} is a complex isotropic vector space with respect to the complex symplectic form ω\omega^{\mathbb{C}} with dimWn\dim_{\mathbb{C}}W^{\prime}\leq n. It contains the isotropic subspace W~\tilde{W}^{\prime} with respect to the real symplectic form ω\omega and the complex span of W~\tilde{W}^{\prime} coincides with WW^{\prime}. The vector space W~\tilde{W}^{\prime} is a commutative real subalgebra and its complexification WW^{\prime} will be a commutative complex subalgebra of the complex Heisenberg algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{C}}. By making use notation (50) we conclude that dimW2n\dim_{\mathbb{R}}W^{\prime}\leq 2n, and WW^{\prime} is a commutative subalgebra of 𝔥1\mathfrak{h}^{1}_{\mathbb{C}} considered as a real Lie algebra.

We consider two cases: dimW=2kn\dim_{\mathbb{R}}W^{\prime}=2k\leq n and dimW=2k>n\dim_{\mathbb{R}}W^{\prime}=2k>n. Let dimW=2kn\dim_{\mathbb{R}}W^{\prime}=2k\leq n. We find a real commutative orthonormal basis {e1,,e2k}\{e_{1},\ldots,e_{2k}\} for WW^{\prime} and extend it to an orthonormal basis {e1,,en}\{e_{1},\ldots,e_{n}\}. Then we denote V=WV=W^{\prime}, and set

V=span{e2k+1,,en,Jϵ1ei,Jϵ2ei,Jϵ2Jϵ1ei;i=1,,n}span{ϵ1,ϵ2}.V^{\prime}=\text{\rm span}\,_{\mathbb{R}}\{e_{2k+1},\ldots,e_{n},J_{\epsilon_{1}}e_{i},J_{\epsilon_{2}}e_{i},J_{\epsilon_{2}}J_{\epsilon_{1}}e_{i};\ i=1,\ldots,n\}\oplus\text{\rm span}\,_{\mathbb{R}}\{\epsilon_{1},\epsilon_{2}\}.

In the case dimW=2k>n\dim_{\mathbb{R}}W^{\prime}=2k>n, we choose orthonormal vectors {e1,,en}\{e_{1},\ldots,e_{n}\} in WW^{\prime} and extend them to the orthonormal basis

{e1,,en,Jϵ2Jϵ1e1,,Jϵ2Jϵ1en}\{e_{1},\ldots,e_{n},J_{\epsilon_{2}}J_{\epsilon_{1}}e_{1},\ldots,J_{\epsilon_{2}}J_{\epsilon_{1}}e_{n}\}

of the maximal commutative subalgebra of 𝔥1\mathfrak{h}^{1}_{\mathbb{C}}. Without loss of generality we can assume that

{e1,,en,Jϵ2Jϵ1e1,,Jϵ2Jϵ1ep},n+p=2k\{e_{1},\ldots,e_{n},J_{\epsilon_{2}}J_{\epsilon_{1}}e_{1},\ldots,J_{\epsilon_{2}}J_{\epsilon_{1}}e_{p}\},\quad n+p=2k

is an orthonormal basis for WW^{\prime}. Now we denote V=WV=W^{\prime} and set

V=span{Jϵ1ei,Jϵ2ei;i=1,,n,Jϵ2Jϵ1ep+1,,Jϵ2Jϵ1en}span{ϵ1,ϵ2}.V^{\prime}=\text{\rm span}\,_{\mathbb{R}}\{J_{\epsilon_{1}}e_{i},J_{\epsilon_{2}}e_{i};\ i=1,\ldots,n,\ J_{\epsilon_{2}}J_{\epsilon_{1}}e_{p+1},\ldots,J_{\epsilon_{2}}J_{\epsilon_{1}}e_{n}\}\oplus\text{\rm span}\,_{\mathbb{R}}\{\epsilon_{1},\epsilon_{2}\}.

We recall that

ei,Jϵ1Jϵ2ei=Jϵ1ei,Jϵ2ei=ϵ1,ϵ2ei,ei=0\langle e_{i},J_{\epsilon_{1}}J_{\epsilon_{2}}e_{i}\rangle_{\mathbb{R}}=-\langle J_{\epsilon_{1}}e_{i},J_{\epsilon_{2}}e_{i}\rangle_{\mathbb{R}}=\langle\epsilon_{1},\epsilon_{2}\rangle_{\mathbb{R}}\langle e_{i},e_{i}\rangle_{\mathbb{R}}=0

because of the orthogonality of vectors ϵ1,ϵ2\epsilon_{1},\epsilon_{2}. Also ei,Jϵ1Jϵ2ej=0\langle e_{i},J_{\epsilon_{1}}J_{\epsilon_{2}}e_{j}\rangle_{\mathbb{R}}=0 for iji\neq j because of the orthogonal decomposition (59).

The last case concerns with 𝔥n\mathfrak{h}^{n}_{\mathbb{H}}. We start from lower dimensional subalgebra 𝔥1\mathfrak{h}^{1}_{\mathbb{H}}. We choose a vector v𝔥1𝔥1v\in\mathfrak{h}_{1}\subset\mathfrak{h}^{1}_{\mathbb{H}} with v,v=1\langle v,v\rangle_{\mathbb{R}}=1 and define an orthonormal basis of the real Heisenberg algebra

X1=v,X2=Jϵ1v,,ϵ1X_{1}=v,\quad X_{2}=J_{\epsilon_{1}}v,\quad,\epsilon_{1}

We use ϵ2\epsilon_{2} and construct the symplectic complex space 𝔥1\mathfrak{h}_{1}^{\mathbb{C}} as in the previous case. Then the constructed multidimensional complexified Heisenberg algebra satisfies the commutation relations of the first line in (58). As in the previous case we find a space WW^{\prime} such that ω(u,v)=0\omega^{\mathbb{C}}(u,v)=0 for all u,vWu,v\in W^{\prime}. Since WW^{\prime} is a complex vector space it has even real dimension dim(W)=2k2n\dim_{\mathbb{R}}(W^{\prime})=2k\leq 2n and therefore we can define one more real symplectic form on WW^{\prime} (considered as a real vector space) by

ω3(u1,u2)=Jϵ3u1,u2,u1,u2W.\omega_{3}(u_{1},u_{2})=\langle J_{\epsilon_{3}}u_{1},u_{2}\rangle_{\mathbb{R}},\quad u_{1},u_{2}\in W^{\prime}.

Then we set V=WL(W)V^{\prime}=W^{\prime}\cap L(W^{\prime}), where L(W)L(W^{\prime}) is the Lagrangian subspace of the real symplectic space (W,ω3)(W^{\prime},\omega_{3}). We obtain dim(V)=pkn\dim_{\mathbb{R}}(V)=p\leq k\leq n and it is by construction a commutative subspace of 𝔥1\mathfrak{h}_{\mathbb{H}}^{1}. Now we choose an orthonormal basis {e1,,ep}\{e_{1},\ldots,e_{p}\} for VV and complement it to an orthonormal basis

(61) {e1,,ep,ep+1,,en}.\{e_{1},\ldots,e_{p},e_{p+1},\ldots,e_{n}\}.

In the last step we extend (61) to an orthonormal basis of 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{H}}^{n} by

{e1,,en,Jϵle1,,Jϵlen;ł=1,2,3}.\{e_{1},\ldots,e_{n},J_{\epsilon_{l}}e_{1},\ldots,J_{\epsilon_{l}}e_{n};\ \l=1,2,3\}.

We have obtained the orthogonally complemented subalgebras

V,V=span{ep+1,,en,Jϵle1,,Jϵlen;ł=1,2,3}𝔥2V,\quad V^{\prime}=\text{\rm span}\,_{\mathbb{R}}\{e_{p+1},\ldots,e_{n},J_{\epsilon_{l}}e_{1},\ldots,J_{\epsilon_{l}}e_{n};\ \l=1,2,3\}\oplus\mathfrak{h}_{2}

satisfying the statement of Lemma 6. ∎

Theorem 9.

The group Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) acts transitively on the family of orthogonally complemented subalgebras of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}, i.e., if VV and VV^{\prime} are subalgebras of the same dimension that have orthogonal subalgebras VV^{\perp} and VV^{\prime\perp}, then there exists 𝒜Iso(𝔥𝕂n)\mathcal{A}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) such that

V=𝒜(V),V=𝒜(V).V^{\prime}=\mathcal{A}(V),\qquad V^{\prime\perp}=\mathcal{A}(V^{\perp}).
Proof.

We start from 𝔥n\mathfrak{h}^{n}_{\mathbb{R}}. Suppose first that k=dimV=dimVnk=\dim_{\mathbb{R}}V=\dim_{\mathbb{R}}V^{\prime}\leq n. Then both VV and VV^{\prime} are commutative by Lemma 6. Let {z1,,zk}\{z_{1},\dots,z_{k}\} and {z1,,zk}\{z^{\prime}_{1},\dots,z^{\prime}_{k}\} be orthonormal bases of the corresponding VV and VV^{\prime} with respect to the scalar product .,.\langle.\,,.\rangle_{\mathbb{R}}. Since each of the bases belong to the isotropic space of the real symplectic form ω\omega in (42), then the bases are orthonormal with respect to the Hermitian scalar profuct (41); that is

(62) zi|zj=zi,zjiω(zi,zj)=0.\langle z_{i}|z_{j}\rangle_{\mathbb{C}}=\langle z_{i},z_{j}\rangle_{\mathbb{R}}-i\omega(z_{i},z_{j})=0.

The same holds for {z1,,zk}\{z^{\prime}_{1},\dots,z^{\prime}_{k}\}.

By the Gram-Schmit procedure for Hermitian scalar products the orthonormal families {z1,,zk}\{z_{1},\dots,z_{k}\} and {z1,,zk}\{z^{\prime}_{1},\dots,z^{\prime}_{k}\} can be extended to orthonormal bases 𝒵={z1,,zk,zk+1,,zn}\mathcal{Z}=\{z_{1},\dots,z_{k},z_{k+1},\dots,z_{n}\} and 𝒵={z1,,zk,zk+1,,zn}\mathcal{Z}^{\prime}=\{z^{\prime}_{1},\dots,z^{\prime}_{k},z^{\prime}_{k+1},\dots,z^{\prime}_{n}\} of 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{R}}. Then 𝔥1\mathfrak{h}_{1} spanned over \mathbb{C} by 𝒵\mathcal{Z} and also by 𝒵\mathcal{Z}^{\prime} is an nn-dimensional complex space. We can find AU(n)A\in\text{\rm U}(n) such that A(zj)=zjA(z_{j})=z^{\prime}_{j}. Then 𝒜=A×Id𝔥2Iso(𝔥n)\mathcal{A}=A\times\text{\rm Id}_{\mathfrak{h}_{2}}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{R}}) and A(V)=VA(V)=V^{\prime}. Since

U(n)=O(2n)GL(n,)Sp(2n,)\text{\rm U}(n)=\text{\rm O}(2n)\cap\text{\rm GL}(n,\mathbb{C})\cap\text{\rm Sp}(2n,\mathbb{R})

we conclude that AO(2n)A\in\text{\rm O}(2n) and 𝒜O(2n+1)\mathcal{A}\in\text{\rm O}(2n+1), and therefore 𝒜(V)=V\mathcal{A}(V^{\perp})=V^{\prime\perp}. This completes the proof of the assertion when knk\leq n.

Suppose now k>nk>n and let VV and VV^{\prime} be two orthogonally complemented subalgebras of 𝔥n\mathfrak{h}^{n}_{\mathbb{R}}. The assertion follows by the previous arguments applied to the orthogonal complements VV^{\perp} and VV^{\prime\perp}.

Consider the Lie algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{C}}. The complex dimension of 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{C}} is equal to 2n2n. Let VV and VV^{\prime} be orthogonally complemented subalgebras of complex dimension knk\leq n. By the construction in Lemma 6 the bases of VV or VV^{\prime} will be also orthonormal bases with respect to the quaternion Hermitian product (54). Then we extend the bases of VV and VV^{\prime} to bases of 𝔥1𝔥n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{C}} and apply the Gram-Schmidt procedure to make the bases orthonormal with respect to the quaternion Hermitian product. Noticing that Sp(n)Sp(2n,)U(2n),\text{\rm Sp}(n)\cong\text{\rm Sp}(2n,\mathbb{C})\cap\text{\rm U}(2n), we obtain that the bases will be orthogonal with respect to the original real scalar product and therefore will preserve the orthogonally complemented subalgebras. We finish the proof as in the previous case.

The last case is the Lie algebra 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}}. In this case we use the similar arguments noting that an orthonormal basis (with respect to .,.\langle.\,,.\rangle_{\mathbb{R}}) for a commutative subalgebra VV will be orthogonal with respect to the quaternion Hermitian form (56). Therefore, the basis can be extended to an orthonormal basis for 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} with respect to the quaternion Hermitian form (56). The group Sp(n)\text{\rm Sp}(n) acts transitively on a set of such kind of extended bases for 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}}, preserving the orthogonality. ∎

Definition 9.

The set Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}), 1kdim𝔥𝕂n1\leq k\leq\dim_{\mathbb{R}}\mathfrak{h}^{n}_{\mathbb{K}} of orthogonally complemented homogeneous subalgebras of the same topological dimension kk is called the Grassmannian of the Heisenberg algebra 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}.

According to (39) we can write 𝒜Iso(𝔥𝕂n)\mathcal{A}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) as 𝒜=(𝒰,𝒱)O(𝔥1)×O(𝔥2)\mathcal{A}=(\mathcal{U},\mathcal{V})\subset\text{\rm O}(\mathfrak{h}_{1})\times\text{\rm O}(\mathfrak{h}_{2}). If V=HTV=H\otimes T, H𝔥1H\subset\mathfrak{h}_{1}, T𝔥2T\subset\mathfrak{h}_{2}, is an orthogonally complemented subalgebra then the action of Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) on Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is given by

(63) 𝒜.V=𝒰.H×𝒱.T.\mathcal{A}.V=\mathcal{U}.H\times\mathcal{V}.T.

Theorem 9 immediately implies the corollary.

Corollary 6.

The Grassmannian Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is the orbit of the action of the isometry group Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) issued from an orthogonally complemented homogeneous kk-dimensional subalgebra V^𝔥𝕂n\hat{V}\subset\mathfrak{h}^{n}_{\mathbb{K}}.

Example 7.

Consider the Heisenberg algebra 𝔥2\mathfrak{h}^{2}_{\mathbb{R}} with the basis (11) and the commutation relations (12). We write V^=span{X1,X2}.\hat{V}=\text{\rm span}\,_{\mathbb{R}}\{X_{1},X_{2}\}. The space V^\hat{V} is a commutative subalgebra of 𝔥2\mathfrak{h}^{2}_{\mathbb{R}} orthogonally complemented by the commutative subalgebra V=span{Y1,Y2,ϵ}V=\text{\rm span}\,_{\mathbb{R}}\{Y_{1},Y_{2},\epsilon\}. The isometry group Iso(𝔥2)\text{\rm Iso}(\mathfrak{h}^{2}_{\mathbb{R}}) is given by (39) with 𝔸=U(2)\mathbb{A}=\text{\rm U}(2). The orbit of the isometry group Iso(𝔥2)\text{\rm Iso}(\mathfrak{h}^{2}_{\mathbb{R}}) acting on V^\hat{V} is the Grassmannian Gr(2,𝔥2)\text{\rm Gr}(2,\mathfrak{h}^{2}_{\mathbb{R}}). The planes in this Grassmannian intersect in one point and it is analogue of the Grassmann manifold of 2 dimensional planes in 4𝔥1𝔥2\mathbb{R}^{4}\cong\mathfrak{h}_{1}\subset\mathfrak{h}^{2}_{\mathbb{R}}. The orbit of the isometry group Iso(𝔥2)\text{\rm Iso}(\mathfrak{h}^{2}_{\mathbb{R}}) acting on the complementary subalgebra VV is the Grassmannian Gr(3,𝔥2)\text{\rm Gr}(3,\mathfrak{h}^{2}_{\mathbb{R}}). The planes in Gr(3,𝔥2)\text{\rm Gr}(3,\mathfrak{h}^{2}_{\mathbb{R}}) intersect in the stright line coinciding with the center of 𝔥2\mathfrak{h}^{2}_{\mathbb{R}}.

In the following example we show that the complementary subalgebras can both contain the elements of the center and be both commutative.

Example 8.

Let us consider the Lie algebra 𝔥1\mathfrak{h}^{1}_{\mathbb{C}} with the basis (45) and the commutation relations (46). We set

V^=span{X1,X4,ϵ1}=span{X1}span{ϵ1}.\hat{V}=\text{\rm span}\,_{\mathbb{R}}\{X_{1},X_{4},\epsilon_{1}\}=\text{\rm span}\,_{\mathbb{C}}\{X_{1}\}\oplus\text{\rm span}\,_{\mathbb{R}}\{\epsilon_{1}\}.

It is a commutative subalgebra of 𝔥\mathfrak{h}_{\mathbb{C}} that is orthogonally complemented by the commutative subalgebra V=span{X2,X3,ϵ2}V=\text{\rm span}\,_{\mathbb{R}}\{X_{2},X_{3},\epsilon_{2}\}. The isometry group Iso(𝔥1)\text{\rm Iso}(\mathfrak{h}^{1}_{\mathbb{C}}) is given by (39) with 𝔸=Sp(1)=Sp(2,)U(2)\mathbb{A}=\text{\rm Sp}(1)=\text{\rm Sp}(2,\mathbb{C})\cap\text{\rm U}(2). The orbit of the isometry group Iso(𝔥1)\text{\rm Iso}(\mathfrak{h}^{1}_{\mathbb{C}}) acting on V^\hat{V} is the Grassmannian Gr(3,𝔥1)\text{\rm Gr}(3,\mathfrak{h}^{1}_{\mathbb{C}}). The planes in Gr(3,𝔥1)\text{\rm Gr}(3,\mathfrak{h}^{1}_{\mathbb{C}}) intersect in one point.

4.3.2. Grassmannians Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) as quotient spaces

Let us fix an orthogonally complemented subalgebra V^𝔥𝕂n\hat{V}\subset\mathfrak{h}^{n}_{\mathbb{K}} and write

K(V^)={𝒜Iso(𝔥𝕂n):𝒜.V^=V^}K(\hat{V})=\Big{\{}\mathcal{A}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}):\ \mathcal{A}.\hat{V}=\hat{V}\Big{\}}

for the isotropy group of V^\hat{V}. The canonical projection

(64) Π:Iso(𝔥𝕂n)Iso(𝔥𝕂n)/K(V^)\Pi\colon\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})\to\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})/K(\hat{V})

is a continuous map. The action of Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) on Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is transitive, by Lemma 9. We identify the left cosets from Iso(𝔥𝕂n)/K(V^)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})/K(\hat{V}) with elements in Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) by

(65) Γ:Iso(𝔥𝕂n)/K(V^)Gr(k,𝔥𝕂n)𝒜.K(V^)V=𝒜.V^.\begin{array}[]{ccccccc}\Gamma\colon&\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})/K(\hat{V})&\to&\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})\\ &\mathcal{A}.\,K(\hat{V})&\mapsto&V=\mathcal{A}.\hat{V}.\end{array}

The map Γ\Gamma is a diffeomorphism, see [War83, Theorem 3.62].

4.3.3. Measure on the Grassmannians Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})

The groups Pin(𝔥2,.,.)\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle), 𝔥2𝔥𝕂n\mathfrak{h}_{2}\subset\mathfrak{h}^{n}_{\mathbb{K}} and 𝔸\mathbb{A} from Section 4.2 are compact Lie groups and therefore they carry normalised Haar measures that we will denote by λPin\lambda_{\text{\rm Pin}} and λ𝔸\lambda_{\mathbb{A}}, respectively. We also will denote λ=λPin×λ𝔸\lambda=\lambda_{\text{\rm Pin}}\times\lambda_{\mathbb{A}} the normalised product measure on the space =Pin(𝔥2,.,.)×𝔸\mathbb{P}=\text{\rm Pin}(\mathfrak{h}_{2},\langle.\,,.\rangle)\times\mathbb{A}. The map ϕ:Iso(𝔥𝕂n)\phi\colon\mathbb{P}\to\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) from (39) mapping θϕ(θ)=𝒜=(𝒰,𝒱)Iso(𝔥𝕂n)\theta\in\mathbb{P}\to\phi(\theta)=\mathcal{A}=(\mathcal{U},\mathcal{V})\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) is continuous surjective map that makes possible to push forward the measure λ\lambda to Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}). Then the map Ψ=ΓΠϕ\Psi=\Gamma\circ\Pi\circ\phi, where Π\Pi and Γ\Gamma are defined in (64) and (65) respectively, allows us to push forward the normalised Haar measure λ\lambda from \mathbb{P} to Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}). We say that a set ΩGr(k,𝔥𝕂n)\Omega\subset\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is measurable if Ψ1(Ω)\Psi^{-1}(\Omega)\subset\mathbb{P} is measurable with respect to the measure λ\lambda. The measure μ\mu on Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is defined by

μ(Ω)=(Ψλ)(Ω)=λ(Ψ1(Ω))=λ{θ:V=ϕ(θ).V^=𝒜.V^Ω},\mu(\Omega)=(\Psi_{\sharp}\lambda)(\Omega)=\lambda(\Psi^{-1}(\Omega)\big{)}=\lambda\Big{\{}\theta\in\mathbb{P}:\ V=\phi(\theta).\hat{V}=\mathcal{A}.\hat{V}\in\Omega\Big{\}},

for any measurable ΩGr(k,𝔥𝕂n)\Omega\subset\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}). We express the push forward in the integral form

(66) Gr(k,𝔥𝕂n)f(V)dμ(V)=f(ϕ(θ).V^)dλ(θ)\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}f(V)\,d\mu(V)=\int_{\mathbb{P}}f(\phi(\theta).\hat{V})\,d\lambda(\theta)

for any measurable function ff on the Grassmannian.

4.3.4. The groups Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) and the product of spheres

According to Lemmas 4 and 5 the groups Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) act transitively on the product of two spheres

𝒮𝕂(0,r1,r2)=Sh(0,r1)×Sv(0,r2)𝔥1𝔥2𝔥𝕂n.\mathcal{S}_{\mathbb{K}}(0,r_{1},r_{2})=S^{h}(0,r_{1})\times S^{v}(0,r_{2})\subset\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}\subset\mathfrak{h}^{n}_{\mathbb{K}}.

with

Sh(0,r1)={g=(x,0)𝔥𝕂n:xE=r1},S^{h}(0,r_{1})=\{g=(x,0)\in\mathfrak{h}^{n}_{\mathbb{K}}:\ \|x\|_{E}=r_{1}\},\quad
Sv(0,r2)={g=(0,t)𝔥𝕂n:tE=r2}.S^{v}(0,r_{2})=\{g=(0,t)\in\mathfrak{h}^{n}_{\mathbb{K}}:\ \|t\|_{E}=r_{2}\}.

The group Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) acts on 𝒮𝕂(0,r1,r2)\mathcal{S}_{\mathbb{K}}(0,r_{1},r_{2}) by the following

𝒜.(y,w)=(𝒰y,𝒱w)for any𝒜=(𝒰,𝒱)Iso(𝔥𝕂n),(y,w)𝒮𝕂(0,r1,r2).\mathcal{A}.(y,w)=(\mathcal{U}y,\mathcal{V}w)\quad\text{for any}\quad\mathcal{A}=(\mathcal{U},\mathcal{V})\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}),\ \ (y,w)\in\mathcal{S}_{\mathbb{K}}(0,r_{1},r_{2}).

We fix (x,t)𝒮𝕂(0,r1,r2)(x,t)\in\mathcal{S}_{\mathbb{K}}(0,r_{1},r_{2}) and define the isotropy subgroups

(67) K(x,t)h={𝒜Iso(𝔥𝕂n):𝒜.(x,t)=(𝒰x,𝒱t)=(x,𝒱t)},K(x,t)v={𝒜Iso(𝔥𝕂n):𝒜.(x,t)=(𝒰x,𝒱t)=(𝒰x,t)}.\begin{split}K^{h}_{(x,t)}=\{\mathcal{A}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}):\ \mathcal{A}.(x,t)=(\mathcal{U}x,\mathcal{V}t)=(x,\mathcal{V}t)\},\\ K^{v}_{(x,t)}=\{\mathcal{A}\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}):\ \mathcal{A}.(x,t)=(\mathcal{U}x,\mathcal{V}t)=(\mathcal{U}x,t)\}.\end{split}

We can realise both spheres as homogeneous spaces under the action of the respective groups, see [War83, Theorem 3.62]. Namely, we write

Π:Iso(𝔥𝕂n)Iso(𝔥𝕂n)/(K(x,t)h×K(x,t)v)Sh(0,r1)×Sv(0,r2).\Pi\colon\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})\to\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})/\Big{(}K^{h}_{(x,t)}\times K^{v}_{(x,t)}\Big{)}\cong S^{h}(0,r_{1})\times S^{v}(0,r_{2}).

We will use the projections

Πh:Iso(𝔥𝕂n)Sh(0,r1),Πv:Iso(𝔥𝕂n)Sv(0,r2).\Pi^{h}\colon\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})\to S^{h}(0,r_{1}),\quad\Pi^{v}\colon\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}})\to S^{v}(0,r_{2}).

The map ϕ\phi from (39) is continuous and surjective. It allows us to define the push forward measures

μh=(Πhϕ)λPinandμv=(Πvϕ)λ𝔸.\mu^{h}=(\Pi^{h}\circ\phi)_{\sharp}\lambda_{\text{\rm Pin}}\quad\text{and}\quad\mu^{v}=(\Pi^{v}\circ\phi)_{\sharp}\lambda_{\mathbb{A}}.
Lemma 7.

The measures μh\mu^{h} and μv\mu^{v} are normalised measures on the spheres Sh(0,r1)S^{h}(0,r_{1}) and Sv(0,r2)S^{v}(0,r_{2}), respectively. Moreover,

Sv(0,r2)dμv(w)Sh(0,r1)f(y,w)dμh(y)=f(ϕ(θ).(x,t))dλ(θ)\int_{S^{v}(0,r_{2})}d\mu^{v}(w)\int_{S^{h}(0,r_{1})}f(y,w)d\mu^{h}(y)=\int_{\mathbb{P}}f\big{(}\phi(\theta).(x,t)\big{)}\,d\lambda(\theta)

for any measurable function ff on S𝕂(0,r1,r2)S_{\mathbb{K}}(0,r_{1},r_{2}) and the isotropy point (x,t)(x,t) from (67).

Proof.

The transitive action of the isometry group on Sh(0,r1)S^{h}(0,r_{1}) and Sv(0,r2)S^{v}(0,r_{2}) ensures that the measures μh\mu^{h} and μv\mu^{v} are uniformly distributed on the respective spheres. Therefore, they are the spherical measures up to constants, see [Mat95].

Let ChSh(0,r1)C^{h}\subset S^{h}(0,r_{1}), CvSv(0,r2)C^{v}\subset S^{v}(0,r_{2}) be measurable sets and let CC\subset\mathbb{P} be its preimage under the map Πϕ=(Πhϕ,Πvϕ)\Pi\circ\phi=(\Pi^{h}\circ\phi,\Pi^{v}\circ\phi) from (39). We write θ=(α,A)Pin(𝔥2,.,.)×𝔸\theta=(\alpha,A)\in\text{\rm Pin}(\mathfrak{h}_{2},\langle.,\,.\rangle)\times\mathbb{A} and ϕ(θ)=ϕ(α,A)=(𝒰,𝒱)Iso(𝔥𝕂n)\phi(\theta)=\phi(\alpha,A)=(\mathcal{U},\mathcal{V})\in\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}), where 𝒰=JαA\mathcal{U}=J_{\alpha}\circ A and 𝒱=κ(α)\mathcal{V}=\kappa(\alpha) then

Cf(ϕ(θ).(x,t))dλ(θ)\displaystyle\int_{C}f\big{(}\phi(\theta).(x,t)\big{)}\,d\lambda(\theta)
=\displaystyle= {αPin:(κ(α).t)Cv}dλPin(α){A𝔸:(JαA.x)Ch}f(𝒰.x,𝒱.t)dλ𝔸(A)\displaystyle\int\limits_{\{\alpha\in\text{\rm Pin}:\ (\kappa(\alpha).t)\in C^{v}\}}d\lambda_{\text{\rm Pin}}(\alpha)\int\limits_{\{A\in\mathbb{A}:\ (J_{\alpha}\circ A.x)\in C^{h}\}}f(\mathcal{U}.x,\mathcal{V}.t)\,d\lambda_{\mathbb{A}}(A)
=\displaystyle= {αPin:(κ(α).tCv}dλPin(α){A𝔸:(A.x)Ch}f(𝒰.x,𝒱.t)dλ𝔸(A)\displaystyle\int\limits_{\{\alpha\in\text{\rm Pin}:\ (\kappa(\alpha).t\in C^{v}\}}d\lambda_{\text{\rm Pin}(\alpha)}\int\limits_{\{A\in\mathbb{A}:\ (A.x)\in C^{h}\}}f(\mathcal{U}.x,\mathcal{V}.t)\,d\lambda_{\mathbb{A}}(A)
=\displaystyle= cCv𝑑μv(w)Chf(y,w)𝑑μh(y).\displaystyle c\int\limits_{C^{v}}d\mu^{v}(w)\int\limits_{C^{h}}f(y,w)\,d\mu^{h}(y).

In the third line we used the fact that the group 𝔸\mathbb{A} is already acts transitively on the spheres Sh(0,r1)S^{h}(0,r_{1}) and therefore the push forward measure μh=(Πhϕ)λPin\mu^{h}=(\Pi^{h}\circ\phi)_{\sharp}\lambda_{\text{\rm Pin}} is up to a constant the Hausdorff spherical measure of Sh(0,r1)S^{h}(0,r_{1}) due to the fact that both measures will be uniformly distributed on Sh(0,r1)S^{h}(0,r_{1}). ∎

Remark 6.

Consider the subgroup 𝔸×Id𝔥2Iso(𝔥𝕂n)\mathbb{A}\times\text{\rm Id}_{\mathfrak{h}_{2}}\subset\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}). Since 𝔸×Id𝔥2\mathbb{A}\times\text{\rm Id}_{\mathfrak{h}_{2}} leaves the center 𝔥2\mathfrak{h}_{2} invariant we can define the action of 𝔸\mathbb{A} only on the horizontal slot of coordinates. Since the group 𝔸\mathbb{A} acts transitively on the spheres Sh(0,r)S^{h}(0,r), the sphere Sh(0,r)S^{h}(0,r) passing through the point x𝔥1x\in\mathfrak{h}_{1} with xE=r\|x\|_{E}=r is a homogeneous manifold realised as a quotient of 𝔸\mathbb{A} by a subgroup fixing xx. Then the integral form of the push forward of a measure λ~\tilde{\lambda} from 𝔸\mathbb{A} is the following

(68) Sh(0,r)f(y)𝑑μh(y)=𝔸f(𝒰x)𝑑λ~(𝒰),xSh(0,r),\int_{S^{h}(0,r)}f(y)\,d\mu^{h}(y)=\int_{\mathbb{A}}f(\mathcal{U}x)\,d\tilde{\lambda}(\mathcal{U}),\quad x\in S^{h}(0,r),

for any measurable function ff on Sh(0,r)S^{h}(0,r).

5. Integral formula on “special” HH-type algebras

5.1. Overview of the formula in n\mathbb{R}^{n}

For the Grassmann manifolds in the Euclidean space the following formula is known [Fug58]. Let VGrk(n)V\in\text{\rm Gr}_{k}(\mathbb{R}^{n}) and

F(V)=Vf(x)𝑑σ(x),F(V)=\int_{V}f(x)d\sigma(x),

where ff is a non-negative measurable function in n\mathbb{R}^{n} and σ\sigma is the kk-dimensional Lebesgue measure on the plane VV. Then

(69) Grk(n)F(V)𝑑μ(V)=m(Sk1(0,1))m(Sn1(0,1))nxEknf(x)𝑑x,\int_{\text{\rm Gr}_{k}(\mathbb{R}^{n})}F(V)\,d\mu(V)=\frac{m(S^{k-1}(0,1))}{m(S^{n-1}(0,1))}\int_{\mathbb{R}^{n}}\|x\|^{k-n}_{E}f(x)\,dx,

where μ\mu is a normalised invariant under the rotational group measure on the Grassmann manifold, m(Sk1(0,1))m(S^{k-1}(0,1)) is the measure of the unit sphere Sk1(0,1)kS^{k-1}(0,1)\subset\mathbb{R}^{k}, dxdx is the Lebesgue measure on n\mathbb{R}^{n}, and xE\|x\|_{E} is the Euclidean norm of xx. Our aim is finding an analogous expression for the three types of the Heisenberg algebras mentioned in Section 4.2.

5.2. Formula for special HH-type Lie algebras

We start from the case of orthogonally complemented Grassmannians Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) which elements consist of the commutative subalgebras and do not include elements of the center. We call them shortly ”horizontal” Grassmannians. In this case we recover formula (69).

5.2.1. Formula for the ”horizontal” Grassmannians

We consider both manifolds Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) and Sh(0,x)𝔥𝕂nS^{h}(0,\|x\|)\subset\mathfrak{h}^{n}_{\mathbb{K}} as homogeneous subspaces under the action of the subgroup 𝔸×IdIso(𝔥𝕂n)\mathbb{A}\times\text{\rm Id}\subset\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}). Here x\|x\| is the Euclidean, or Hermitian norm on 𝔥1𝔥𝕂n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{K}}, accordingly to 𝕂\mathbb{K}.

Let Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) be the Grassmannian consisting of orthogonally complemented commutative subalgebras that do not contain elements of the center. In this case the topological and homogeneous dimensions of VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) coincide and we denote them by 𝐝𝐦=𝐝𝐭=k{\bf d_{m}}={\bf d_{t}}=k. Let also k\mathcal{L}^{k} denote kk-dimensional Lebesgue measure on a generic plain VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) with V𝔥1V\subset\mathfrak{h}_{1}, 𝔥1𝔥𝕂n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{K}}. Set

(70) F(V)=Vf(y)𝑑k(y),yV𝔥1.F(V)=\int_{V}f(y)\,d\mathcal{L}^{k}(y),\quad y\in V\subset\mathfrak{h}_{1}.

Here f:𝔥1f\colon\mathfrak{h}_{1}\to\mathbb{R} is a non-negative measurable function.

Theorem 10.

The formula

Gr(k,𝔥𝕂n)F(V)𝑑μ(V)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}F(V)\,d\mu(V) =\displaystyle= Gr(k,𝔥𝕂n)𝑑μ(V)Vf(y)𝑑k(y)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}\,d\mu(V)\int_{V}f(y)\,d\mathcal{L}^{k}(y)
=\displaystyle= C𝔥1zkm1f(z)𝑑m1(z),\displaystyle C\int_{\mathfrak{h}_{1}}\|z\|^{k-m_{1}}f(z)d\mathcal{L}^{m_{1}}(z),

holds for any measurable non-negative function f:𝔥1f\colon\mathfrak{h}_{1}\to\mathbb{R} and an orthogonally complemented Grassmannian Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) of commutative subalgebras that do not contain elements of the center of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}. Here m1\mathcal{L}^{m_{1}}, m1=dim(𝔥1)m_{1}=\dim_{\mathbb{R}}(\mathfrak{h}_{1}) is the Lebesgue measure on 𝔥1𝔥𝕂n\mathfrak{h}_{1}\subset\mathfrak{h}^{n}_{\mathbb{K}} and C>0C>0 is a constant.

Proof.

Note first that V=A.V^V=A.\hat{V} for any VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) and some A𝔸A\in\mathbb{A}. Therefore

F(V)=F(A.V^)=A.V^f(y)dk(y)=V^f(Ax)dk(x)F(V)=F(A.\hat{V})=\int_{A.\hat{V}}f(y)\,d\mathcal{L}^{k}(y)=\int_{\hat{V}}f(Ax)\,d\mathcal{L}^{k}(x)
(71) Gr(k,𝔥𝕂n)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})} F(V)dμ(V)=Gr(k,𝔥𝕂n)F(A.V^)dμ(A.V^)=𝔸F(A.V^)dλ(A)\displaystyle F(V)\,d\mu(V)=\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}F(A.\hat{V})\,d\mu(A.\hat{V})=\int_{\mathbb{A}}F(A.\hat{V})\,d\lambda(A)
=\displaystyle= 𝔸𝑑λ(A)V^f(Ax)𝑑k(x)=V^𝑑k(x)𝔸f(Ax)𝑑λ(A)\displaystyle\int_{\mathbb{A}}\,d\lambda(A)\int_{\hat{V}}f(Ax)\,d\mathcal{L}^{k}(x)=\int_{\hat{V}}\,d\mathcal{L}^{k}(x)\int_{\mathbb{A}}f(Ax)d\lambda(A)

by (66). ∎

Let us consider the last integral, where xV^x\in\hat{V} will be also considered as a point on the sphere Sh(0,x)V^𝔥1S^{h}(0,\|x\|)\subset\hat{V}\subset\mathfrak{h}_{1}. For that xSh(0,x)x\in S^{h}(0,\|x\|) we can consider Sh(0,x)S^{h}(0,\|x\|) as a homogeneous manifold under the action of 𝔸\mathbb{A} with the isotropy group that fixes xx. We denote that sphere by Sh,x(0,x)S^{h,x}(0,\|x\|), emphasising the fixed point on the sphere. Then we use the push forward μh\mu^{h} of the normalized measure λ\lambda from 𝔸\mathbb{A} to Sh,x(0,x)S^{h,x}(0,\|x\|) and obtain

(72) 𝔸f(Ax)𝑑λ(A)=Sh,x(0,x)f(z)𝑑μh(z)=C~Sh(0,1)f(xξ)𝑑𝒮m11(ξ),\int_{\mathbb{A}}f(Ax)d\lambda(A)=\int_{S^{h,x}(0,\|x\|)}f(z)d\mu^{h}(z)=\tilde{C}\int_{S^{h}(0,1)}f(\|x\|\xi)d\mathcal{S}^{m_{1}-1}(\xi),

where d𝒮m11(ξ)d\mathcal{S}^{m_{1}-1}(\xi) is the surface measure on the unit sphere Sh(0,1)S^{h}(0,1). In the last step we used the following calculations

Sh,x(0,x)f(z)𝑑μh(z)=cSh,x(0,x)x1m1f(xξ)𝑑S(xξ)\displaystyle\int_{S^{h,x}(0,\|x\|)}f(z)d\mu^{h}(z)=c\int_{S^{h,x}(0,\|x\|)}\|x\|^{1-m_{1}}f(\|x\|\xi)dS(\|x\|\xi)
=\displaystyle= CSh(0,1)x1m1xm11f(xξ)𝑑𝒮m11(ξ)=CSh(0,1)f(xξ)𝑑𝒮m11(ξ),\displaystyle C\int_{S^{h}(0,1)}\|x\|^{1-m_{1}}\|x\|^{m_{1}-1}f(\|x\|\xi)d\mathcal{S}^{m_{1}-1}(\xi)=C\int_{S^{h}(0,1)}f(\|x\|\xi)d\mathcal{S}^{m_{1}-1}(\xi),

where dS(xξ)dS(\|x\|\xi) is the surface measure on the sphere Sh,x(0,x)S^{h,x}(0,\|x\|), xV^x\in\hat{V}. Substituting integral (72) into (71), we obtain (for ρ=x\rho=\|x\|)

V^𝑑k(x)Sh(0,1)f(xξ)𝑑𝒮m11(ξ)\displaystyle\int_{\hat{V}}d\mathcal{L}^{k}(x)\int_{S^{h}(0,1)}f(\|x\|\xi)d\mathcal{S}^{m_{1}-1}(\xi)
=\displaystyle= 0ρk1𝑑ρSk1(0,1)𝑑𝒮k1(ζ)constantSh(0,1)f(xξ)𝑑𝒮m11(ξ)\displaystyle\int_{0}^{\infty}\rho^{k-1}d\rho\underbrace{\int_{S^{k-1}(0,1)}d\mathcal{S}^{k-1}(\zeta)}_{\text{constant}}\int_{S^{h}(0,1)}f(\|x\|\xi)d\mathcal{S}^{m_{1}-1}(\xi)
=\displaystyle= C^0ρk1(m11)𝑑ρSh(0,1)f(ρξ)ρm11𝑑𝒮m11(ξ)\displaystyle\hat{C}\int_{0}^{\infty}\rho^{k-1-(m_{1}-1)}d\rho\int_{S^{h}(0,1)}f(\rho\xi)\rho^{m_{1}-1}d\mathcal{S}^{m_{1}-1}(\xi)
=\displaystyle= C^𝔥1zkm1f(z)𝑑m1(z).\displaystyle\hat{C}\int_{\mathfrak{h}_{1}}\|z\|^{k-m_{1}}f(z)d\mathcal{L}^{m_{1}}(z).

5.2.2. Formula for the ”vertical” Grassmannians

Let Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) be a Grassmannian, where a typical orthogonally complemented subalgebra VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) contains a non-trivial part of the center of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}. Let V^=V^hV^v\hat{V}=\hat{V}_{h}\oplus\hat{V}_{v}, V^v{0}\hat{V}_{v}\neq\{0\} be an orthogonally complemented subalgebra, such that V^h𝔥1𝔥𝕂n\hat{V}_{h}\subset\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{K}}^{n} and V^v𝔥2𝔥𝕂n\hat{V}_{v}\subset\mathfrak{h}_{2}\subset\mathfrak{h}_{\mathbb{K}}^{n}. We write kv=dim(V^h)k_{v}=\dim(\hat{V}_{h}), kv=dim(V^v)k_{v}=\dim(\hat{V}_{v}) for the topological dimensions of the vector spaces V^h\hat{V}_{h} and V^v\hat{V}_{v}. Thus k=dimV^=kh+kvk=\dim\hat{V}=k_{h}+k_{v} is the topological dimension of orthogonally complemented subalgebra V^\hat{V}. A generic element VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is the image of V^\hat{V} under the action of Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}). We write y=(x,t)VGr(k,𝔥𝕂n)y=(x,t)\in V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}). Let

F(V)=Vf(x,t)𝑑k(x,t),F(V)=\int_{V}f(x,t)\,d\mathcal{L}^{k}(x,t),

where f:𝔥𝕂nf\colon\mathfrak{h}^{n}_{\mathbb{K}}\to\mathbb{R} is a measurable non-negative function and k\mathcal{L}^{k}, k=dim(V)k=\dim_{\mathbb{R}}(V) is the Lebesgue measure on VV. We denote m1=dim(𝔥1)m_{1}=\dim_{\mathbb{R}}(\mathfrak{h}_{1}), m2=dim(𝔥2)m_{2}=\dim_{\mathbb{R}}(\mathfrak{h}_{2}), the topological dimensions of the horizontal 𝔥1\mathfrak{h}_{1} and vertical 𝔥2\mathfrak{h}_{2} layers of 𝔥𝕂n=𝔥1𝔥2\mathfrak{h}^{n}_{\mathbb{K}}=\mathfrak{h}_{1}\oplus\mathfrak{h}_{2}. Thus N=m1+m2N=m_{1}+m_{2} is the topological dimension of the Lie algebra 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}. Moreover, m1\mathcal{L}^{m_{1}} and m2\mathcal{L}^{m_{2}} are the respective Lebesgue measures on the vector spaces 𝔥1\mathfrak{h}_{1} and 𝔥2\mathfrak{h}_{2}.

Theorem 11.

The formula

Gr(k,𝔥𝕂n)F(V)𝑑μ(V)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}F(V)\,d\mu(V) =\displaystyle= Gr(k,𝔥𝕂n)𝑑μ(V)Vf(x,t)𝑑k(x,t)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}\,d\mu(V)\int_{V}f(x,t)\,d\mathcal{L}^{k}(x,t)
=\displaystyle= Cm1×m2xkhm1tkvm2f(x,t)𝑑m1(x)𝑑m2(t).\displaystyle C\int\limits_{\mathbb{R}^{m_{1}}\times\mathbb{R}^{m_{2}}}\|x\|^{k_{h}-m_{1}}\|t\|^{k_{v}-m_{2}}f(x,t)\,d\mathcal{L}^{m_{1}}(x)d\mathcal{L}^{m_{2}}(t).

holds for any measurable non-negative function f:𝔥𝕂nf\colon\mathfrak{h}^{n}_{\mathbb{K}}\to\mathbb{R} and an orthogonally complemented Grassmannian Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) of subalgebras that contain a nontrivial element of the center of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}. Here C>0C>0 is a constant.

Proof.

The isometry group Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) does not act transitively on the spheres with respect to any of the metrics (D2)(D4)(D_{2})-(D_{4}). This fact does not allow to obtain a uniformly distributed measure on a sub-Riemannian sphere by pushing forward the measure from the isometry group Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}). Nevertheless, the transitive action of Iso(𝔥𝕂n)\text{\rm Iso}(\mathfrak{h}^{n}_{\mathbb{K}}) on the product of spheres allows to prove Lemma 7. In the product of the integrals

(74) f(ϕ(θ).(x,t))dλ(θ)=Sv(0,r2)dμv(w)Sh(0,r1)f(y,w)dμh(y)\int_{\mathbb{P}}f\big{(}\phi(\theta).(x,t)\big{)}\,d\lambda(\theta)=\int_{S^{v}(0,r_{2})}d\mu^{v}(w)\int_{S^{h}(0,r_{1})}f(y,w)d\mu^{h}(y)

the measures dμvd\mu^{v} and dμhd\mu^{h} are the normalised measures on the spheres Sv(0,r2)S^{v}(0,r_{2}) and Sh(0,r1)S^{h}(0,r_{1}), respectively. We can use successive independent dilations in the vertical and horizontal variables and write right-hand side of (74) as a product of integrals with respect to the Hausdorff measures on the unit spheres

(75) CSv(0,1)𝑑𝒮m21(η)Sh(0,1)f(r2ξ,r1η)𝑑𝒮m11(ξ).C\int_{S^{v}(0,1)}d\mathcal{S}^{m_{2}-1}(\eta)\int_{S^{h}(0,1)}f(r_{2}\xi,r_{1}\eta)\,d\mathcal{S}^{m_{1}-1}(\xi).

Futhermore, by making use of the polar coordinates in each of the vector spaces V^h\hat{V}_{h} and V^v\hat{V}_{v} we obtain

V^𝑑𝐝𝐭(x,t)=V^h𝑑kh(x)V^v𝑑kv(t)\displaystyle\int_{\hat{V}}d\mathcal{L}^{{\bf d_{t}}}(x,t)=\int_{\hat{V}_{h}}d\mathcal{L}^{k_{h}}(x)\int_{\hat{V}_{v}}d\mathcal{L}^{k_{v}}(t)
=\displaystyle= 0ρkh1𝑑ρSkh1(0,1)𝑑𝒮kh1(ϕ)0rkv1𝑑rSkv1(0,1)𝑑𝒮kv1(ψ)\displaystyle\int_{0}^{\infty}\rho^{k_{h}-1}d\rho\int\limits_{S^{k_{h}-1}(0,1)}d\mathcal{S}^{k_{h}-1}(\phi)\int_{0}^{\infty}r^{k_{v}-1}dr\int\limits_{S^{k_{v}-1}(0,1)}d\mathcal{S}^{k_{v}-1}(\psi)
=\displaystyle= C~0ρkh1𝑑ρ0rkv1𝑑r.\displaystyle\tilde{C}\int_{0}^{\infty}\rho^{k_{h}-1}d\rho\int_{0}^{\infty}r^{k_{v}-1}dr.

We recall that the measure μ(V)\mu(V) on Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is the pushforward of the measure λ(θ)\lambda(\theta) from the group \mathbb{P}. It allows us to write

Gr(k,𝔥𝕂n)F(V)𝑑μ(V)=Gr(k,𝔥𝕂n)𝑑μ(V)Vf(x,t)𝑑k(x,t)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}F(V)\,d\mu(V)=\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}\,d\mu(V)\int_{V}f(x,t)\,d\mathcal{L}^{k}(x,t)
=\displaystyle= V^dk(x,t)f(ϕ(θ).(x,t))dλ(θ)\displaystyle\int_{\hat{V}}\,d\mathcal{L}^{k}(x,t)\int_{\mathbb{P}}f\big{(}\phi(\theta).(x,t)\big{)}\,d\lambda(\theta)
=\displaystyle= C0ρkh1𝑑ρ0rkv1𝑑rSv(0,1)𝑑𝒮m21(η)Sh(0,1)f(ρξ,rη)𝑑𝒮m11(ξ)\displaystyle C\int_{0}^{\infty}\rho^{k_{h}-1}d\rho\int_{0}^{\infty}r^{k_{v}-1}dr\int_{S^{v}(0,1)}d\mathcal{S}^{m_{2}-1}(\eta)\int_{S^{h}(0,1)}f(\rho\xi,r\eta)d\mathcal{S}^{m_{1}-1}(\xi)
=\displaystyle= C0ρkhm1𝑑ρ0rkvm2𝑑r\displaystyle C\int_{0}^{\infty}\rho^{k_{h}-m_{1}}d\rho\int_{0}^{\infty}r^{k_{v}-m_{2}}dr
×\displaystyle\times Sv(0,1)𝑑𝒮m21(η)Sh(0,1)ρm11rm21f(ρξ,rη)𝑑𝒮m11(ξ)\displaystyle\int_{S^{v}(0,1)}d\mathcal{S}^{m_{2}-1}(\eta)\int_{S^{h}(0,1)}\rho^{m_{1}-1}r^{m_{2}-1}f(\rho\xi,r\eta)d\mathcal{S}^{m_{1}-1}(\xi)
=\displaystyle= m2×m1xkhm1tkvm2f(x,t)𝑑m1(x)𝑑m2(t).\displaystyle\int\limits_{\mathbb{R}^{m_{2}}\times\mathbb{R}^{m_{1}}}\|x\|^{k_{h}-m_{1}}\|t\|^{k_{v}-m_{2}}f(x,t)d\mathcal{L}^{m_{1}}(x)d\mathcal{L}^{m_{2}}(t).

It finishes the proof. ∎

5.3. Application of the integral formula

Let 𝔾\mathbb{G} be one of the special Heisenberg-type Lie groups of the topological dimension NN and the homogeneous dimension QQ with the Lie algebra 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}, such that dim(𝔥1)=m1\dim(\mathfrak{h}_{1})=m_{1}.

Corollary 7.

Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a collection of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs on 𝔾\mathbb{G}. Suppose that all the graphs SΣS\in\Sigma contain a common point g0𝔾g_{0}\in\mathbb{G}. Then for 𝐝𝐦pQ{\bf d_{m}}p\leq Q, p>1p>1, we have Mp(Σ)=0M_{p}(\Sigma)=0. In the case 𝐝𝐭=𝐝𝐦{\bf d_{t}}={\bf d_{m}} if p𝐝𝐭>m1p{\bf d_{t}}>m_{1}, then there is a family Σ\Sigma of intrinsic Lipschitz graphs such that Mp(Σ)0M_{p}(\Sigma)\neq 0.

Proof.

The proof of the statement that 𝐝𝐦pQ{\bf d_{m}}p\leq Q implies Mp(Σ)=0M_{p}(\Sigma)=0 is the proof of Theorem 8. To show the statement in the opposite direction, we consider the Grassmannian Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) consisting of orthogonally complemented commutative subalgebras that do not contain elements of the center. In this case the topological and homogeneous dimensions of VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) coincide; that is 𝐝𝐦=𝐝𝐭=k{\bf d_{m}}={\bf d_{t}}=k.

Consider S=VBh(0,1)S=V\cap B^{h}(0,1), where VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) and Bh(0,1)𝔥1B^{h}(0,1)\in\mathfrak{h}_{1} is the Euclidean unit ball and assume by contrary, that Mp(Σ)=0M_{p}(\Sigma)=0, where

Σ={S=VBh(0,1),VGr(k,𝔥𝕂n)}.\Sigma=\{S=V\cap B^{h}(0,1),\ V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})\}.

Let fLp(𝔥1,m1)f\in L^{p}(\mathfrak{h}_{1},\mathcal{L}^{m_{1}}), m1=dim(𝔥1)m_{1}=\dim_{\mathbb{R}}(\mathfrak{h}_{1}). Then from Theorem 10 and the Hölder inequality for 1p+1q=1\frac{1}{p}+\frac{1}{q}=1 we obtain

Gr(k,𝔥𝕂n)𝑑μ(V)VBh(0,1)f(y)𝑑𝐝𝐭(y)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}\,d\mu(V)\int_{V\cap B^{h}(0,1)}f(y)\,d\mathcal{L}^{{\bf d_{t}}}(y) =\displaystyle= C𝔥1Bh(0,1)x𝐝𝐭m1f(x)𝑑m1(x)\displaystyle C\int_{\mathfrak{h}_{1}\cap B^{h}(0,1)}\|x\|^{{\bf d_{t}}-m_{1}}f(x)d\mathcal{L}^{m_{1}}(x)
\displaystyle\leq CfLp(𝔥1)(01rp𝐝𝐭m1p11)p1p.\displaystyle C\|f\|_{L^{p}(\mathfrak{h}_{1})}\Big{(}\int_{0}^{1}r^{\frac{p{\bf d_{t}}-m_{1}}{p-1}-1}\Big{)}^{\frac{p-1}{p}}.

Then since p𝐝𝐭>m1p{\bf d_{t}}>m_{1} the last integral is finite. It implies

(76) VBh(0,1)f(y)𝑑𝐝𝐭(y)<\int_{V\cap B^{h}(0,1)}f(y)\,d\mathcal{L}^{{\bf d_{t}}}(y)<\infty

for μ\mu-almost all S=VBh(0,1)ΣS=V\cap B^{h}(0,1)\in\Sigma, that contradicts the assumption Mp(Σ)=0M_{p}(\Sigma)=0. ∎

For the following corollary we assume that Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) is a Grassmannian, where a typical orthogonally complemented subalgebra VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) contains a non-trivial part of the center of 𝔥𝕂n\mathfrak{h}^{n}_{\mathbb{K}}. In this case necessarily 𝐝𝐭<𝐝𝐦{\bf d_{t}}<{\bf d_{m}}. Let V=VhVvV=V_{h}\oplus V_{v} where Vv{0}V_{v}\neq\{0\}, be an orthogonally complemented subalgebra, such that Vh𝔥1𝔥𝕂nV_{h}\subset\mathfrak{h}_{1}\subset\mathfrak{h}_{\mathbb{K}}^{n} and Vv𝔥2𝔥𝕂nV_{v}\subset\mathfrak{h}_{2}\subset\mathfrak{h}_{\mathbb{K}}^{n}. We write kh=dim(Vh)k_{h}=\dim(V_{h}), kv=dim(Vv)k_{v}=\dim(V_{v}), 𝐝𝐭=k=kh+kv{\bf d_{t}}=k=k_{h}+k_{v}, for the topological dimensions of the vector spaces VhV_{h} and VvV_{v}, and m1=dim(𝔥1)m_{1}=\dim(\mathfrak{h}_{1}), m2=dim(𝔥2)m_{2}=\dim(\mathfrak{h}_{2}), the topological dimensions of the horizontal and the vertical layers of 𝔥𝕂n\mathfrak{h}_{\mathbb{K}}^{n}.

Corollary 8.

Let ΣΣ(𝐝𝐭,𝐝𝐦)\Sigma\subset\Sigma^{({\bf d_{t}},{\bf d_{m}})} be a collection of intrinsic (𝐝𝐭,𝐝𝐦)({\bf d_{t}},{\bf d_{m}})-Lipschitz graphs. Suppose that all the graphs SΣS\in\Sigma contain a common point g0𝔾g_{0}\in\mathbb{G}. Then for 𝐝𝐦pQ{\bf d_{m}}p\leq Q, p>1p>1, we have Mp(Σ)=0M_{p}(\Sigma)=0. In the case 𝐝𝐭<𝐝𝐦{\bf d_{t}}<{\bf d_{m}}, if pkh>m1pk_{h}>m_{1}, pkv>m2pk_{v}>m_{2} then there is a family Σ\Sigma of intrinsic Lipschitz graphs such that Mp(Σ)0M_{p}(\Sigma)\neq 0.

Proof.

We argue as in the proof of Corollary 8. Consider S=V(Bh(0,1)×Bv(0,1))S=V\cap\Big{(}B^{h}(0,1)\times B^{v}(0,1)\Big{)}, where VGr(k,𝔥𝕂n)V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) and Bh(0,1)𝔥1B^{h}(0,1)\in\mathfrak{h}_{1}, Bv(0,1)𝔥2B^{v}(0,1)\in\mathfrak{h}_{2} are the Euclidean unit balls. Asssume that Mp(Σ)=0M_{p}(\Sigma)=0, where

Σ={S=V(Bh(0,1)×Bv(0,1)),VGr(k,𝔥𝕂n)}.\Sigma=\{S=V\cap\Big{(}B^{h}(0,1)\times B^{v}(0,1)\Big{)},\ V\in\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})\}.

Let fLp(𝔥𝕂n,N)f\in L^{p}(\mathfrak{h}^{n}_{\mathbb{K}},\mathcal{L}^{N}), N=dim(𝔥𝕂n)N=\dim_{\mathbb{R}}(\mathfrak{h}^{n}_{\mathbb{K}}). Then from Theorem 11 and the Hölder inequality for 1p+1q=1\frac{1}{p}+\frac{1}{q}=1 we obtain

Gr(k,𝔥𝕂n)𝑑μ(V)Vf(x,t)𝑑k(x,t)\displaystyle\int_{\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}})}\,d\mu(V)\int_{V}f(x,t)\,d\mathcal{L}^{k}(x,t)
=\displaystyle= Cm2Bv(0,1)×m1Bh(0,1)xkhm1tkvm2f(x,t)𝑑m1(x)𝑑m2(t)\displaystyle C\int\limits_{\mathbb{R}^{m_{2}}\cap B^{v}(0,1)\times\mathbb{R}^{m_{1}}\cap B^{h}(0,1)}\|x\|^{k_{h}-m_{1}}\|t\|^{k_{v}-m_{2}}f(x,t)d\mathcal{L}^{m_{1}}(x)d\mathcal{L}^{m_{2}}(t)
\displaystyle\leq CfLp(𝔥𝕂n)m2Bv(0,1)×m1Bh(0,1)x(khm1)pp1t(kvm2)pp1𝑑m1(x)𝑑m2(t)\displaystyle C\|f\|_{L^{p}(\mathfrak{h}^{n}_{\mathbb{K}})}\int\limits_{\mathbb{R}^{m_{2}}\cap B^{v}(0,1)\times\mathbb{R}^{m_{1}}\cap B^{h}(0,1)}\|x\|^{(k_{h}-m_{1})\frac{p}{p-1}}\|t\|^{(k_{v}-m_{2})\frac{p}{p-1}}d\mathcal{L}^{m_{1}}(x)d\mathcal{L}^{m_{2}}(t)
=\displaystyle= C~fLp(𝔥𝕂n)01rpkhm1p11𝑑r01ρpkvm2p11𝑑ρ<\displaystyle\tilde{C}\|f\|_{L^{p}(\mathfrak{h}^{n}_{\mathbb{K}})}\int\limits_{0}^{1}r^{\frac{pk_{h}-m_{1}}{p-1}-1}dr\int\limits_{0}^{1}\rho^{\frac{pk_{v}-m_{2}}{p-1}-1}d\rho<\infty

Since pkh>m1pk_{h}>m_{1} and pkv>m2pk_{v}>m_{2}, then Vf(x,t)𝑑k(x,t)<\int_{V}f(x,t)d\mathcal{L}^{k}(x,t)<\infty for μ\mu-almost all plains SΣS\in\Sigma, that contrudicts to the assumption Mp(Σ)=0M_{p}(\Sigma)=0. ∎

Remark 7.

Two conditions

pkh>m1andpkv>m2pk_{h}>m_{1}\quad\text{and}\quad pk_{v}>m_{2}

imply

p(kh+kv)=p𝐝𝐭>m1+m2=N,p(kh+2kv)=p𝐝𝐦>m1+2m2=Qp(k_{h}+k_{v})=p{\bf d_{t}}>m_{1}+m_{2}=N,\quad p(k_{h}+2k_{v})=p{\bf d_{m}}>m_{1}+2m_{2}=Q

In general p𝐝𝐦>Qp{\bf d_{m}}>Q, p>1p>1, does not imply both conditions pkh>m1pk_{h}>m_{1} and pkv>m2pk_{v}>m_{2} in spite that the second one for 𝔥n\mathfrak{h}^{n}_{\mathbb{R}} and 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} is always fulfilled for the Grassmannians Gr(k,𝔥𝕂n)\text{\rm Gr}(k,\mathfrak{h}^{n}_{\mathbb{K}}) where a typical orthogonally complemented subalgebra VV contains a non-trivial part of the center. In both cases 𝔥n\mathfrak{h}^{n}_{\mathbb{R}} and 𝔥n\mathfrak{h}^{n}_{\mathbb{Q}} the subalgebra VV necessarily contains the entire center, and therefore kv=m2k_{v}=m_{2}. From the other side we note that the Lipschitz surfaces meet each other not only at one point but at the entire center.

In the case of 𝔥n\mathfrak{h}^{n}_{\mathbb{C}} and V=VhVvV=V_{h}\oplus V_{v} with kh=dim(Vh)k_{h}=\dim(V_{h}) and kv=dim(Vv)=1k_{v}=\dim(V_{v})=1 we obtain that

p𝐝𝐦=p(kh+2)>Q=m1+4pkh>m1+42pm1ifp2p{\bf d_{m}}=p(k_{h}+2)>Q=m_{1}+4\quad\Longrightarrow\quad pk_{h}>m_{1}+4-2p\geq m_{1}\quad\text{if}\quad p\leq 2

but

pkv=p>m2=2ifp>2pk_{v}=p>m_{2}=2\quad\text{if}\quad p>2

Thus even if the Lipschitz surfaces intersects in one point, our example does not give the answer to the question: is there an example on a Carnot group that is not n\mathbb{R}^{n} where the condition p𝐝𝐦Qp{\bf d_{m}}\leq Q with 𝐝𝐭<𝐝𝐦{\bf d_{t}}<{\bf d_{m}} is necessary for the system of Lipschitz surfaces intersecting in one point to be pp-exceptional?

References

  • [AB50] Lars Ahlfors and Arne Beurling. Conformal invariants and function-theoretic null-sets. Acta Math., 83:101–129, 1950.
  • [ABB20] Andrei Agrachev, Davide Barilari, and Ugo Boscain. A comprehensive introduction to sub-Riemannian geometry, volume 181 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2020. From the Hamiltonian viewpoint, With an appendix by Igor Zelenko.
  • [AM21] Gioacchino Antonelli and Andrea Merlo. On rectifiable measures in carnot groups: representation, 2021.
  • [AO99] Hiroaki Aikawa and Makoto Ohtsuka. Extremal length of vector measures. Ann. Acad. Sci. Fenn. Math., 24(1):61–88, 1999.
  • [AT04] Luigi Ambrosio and Paolo Tilli. Topics on analysis in metric spaces, volume 25 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford, 2004.
  • [BFP11] Zoltán M. Balogh, Katrin Fässler, and Ioannis D. Platis. Modulus of curve families and extremality of spiral-stretch maps. J. Anal. Math., 113:265–291, 2011.
  • [BFP13] Zoltán M. Balogh, Katrin Fässler, and Ioannis D. Platis. Modulus method and radial stretch map in the Heisenberg group. Ann. Acad. Sci. Fenn. Math., 38(1):149–180, 2013.
  • [Bjo02] Jana Bjorn. Boundary continuity for quasiminimizers on metric spaces. Illinois J. Math., 46(2):383–403, 2002.
  • [BLU07] Andrea Bonfiglioli, Ermanno Lanconelli, and Francesco Uguzzoni. Stratified Lie groups and potential theory for their sub-Laplacians. Springer Monographs in Mathematics. Springer, Berlin, 2007.
  • [CG90] Laurence Corwin and Frederick P Greenleaf. Representations of nilpotent Lie groups and their applications: Volume 1, Part 1, Basic theory and examples, volume 18. Cambridge university press, 1990.
  • [Fed69] Herbert Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.
  • [FS82] Gerald B. Folland and Elias M. Stein. Hardy spaces on homogeneous groups, volume 28 of Mathematical Notes. Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1982.
  • [FS16] Bruno Franchi and Raul Paolo Serapioni. Intrinsic Lipschitz graphs within Carnot groups. J. Geom. Anal., 26(3):1946–1994, 2016.
  • [FSSC03a] Bruno Franchi, Raul Serapioni, and Francesco Serra Cassano. On the structure of finite perimeter sets in step 2 Carnot groups. J. Geom. Anal., 13(3):421–466, 2003.
  • [FSSC03b] Bruno Franchi, Raul Serapioni, and Francesco Serra Cassano. Regular hypersurfaces, intrinsic perimeter and implicit function theorem in Carnot groups. Comm. Anal. Geom., 11(5):909–944, 2003.
  • [FSSC06] Bruno Franchi, Raul Serapioni, and Francesco Serra Cassano. Intrinsic Lipschitz graphs in Heisenberg groups. J. Nonlinear Convex Anal., 7(3):423–441, 2006.
  • [FSSC07] Bruno Franchi, Raul Serapioni, and Francesco Serra Cassano. Regular submanifolds, graphs and area formula in Heisenberg groups. Adv. Math., 211(1):152–203, 2007.
  • [Fug57] Bent Fuglede. Extremal length and functional completion. Acta Math., 98:171–219, 1957.
  • [Fug58] Bent Fuglede. An integral formula. Math. Scand., 6:207–212, 1958.
  • [Gro96] Mikhael Gromov. Carnot-Carathéodory spaces seen from within. In Sub-Riemannian geometry, volume 144 of Progr. Math., pages 79–323. Birkhäuser, Basel, 1996.
  • [KSC04] Bernd Kirchheim and Francesco Serra Cassano. Rectifiability and parameterization of intrinsic regular surfaces in the Heisenberg group. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 3(4):871–896, 2004.
  • [Mar04] Irina Markina. pp-module of vector measures in domains with intrinsic metric on Carnot groups. Tohoku Math. J. (2), 56(4):553–569, 2004.
  • [Mat95] Pertti Mattila. Geometry of sets and measures in Euclidean spaces, volume 44 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. Fractals and rectifiability.
  • [Mit85] John Mitchell. On Carnot-Carathéodory metrics. J. Differential Geom., 21(1):35–45, 1985.
  • [Oht03] Makoto Ohtsuka. Extremal length and precise functions, volume 19 of GAKUTO International Series. Mathematical Sciences and Applications. Gakkōtosho Co., Ltd., Tokyo, 2003. With a preface by Fumi-Yuki Maeda.
  • [RH63] Kenneth A Ross and Edwin Hewitt. Abstract Harmonic Analysis: Structure of Topological Groups: Integration Theory. Group Representations. 1963.
  • [Ric93] Seppo Rickman. Quasiregular mappings, volume 26 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1993.
  • [Rie82] Carl R. Riehm. The automorphism group of a composition of quadratic forms. Trans. Amer. Math. Soc., 269(2):403–414, 1982.
  • [Rie84] Carl R. Riehm. Explicit spin representations and Lie algebras of Heisenberg type. J. London Math. Soc. (2), 29(1):49–62, 1984.
  • [Ros06] Wulf Rossmann. Lie groups: an introduction through linear groups, volume 5. Oxford University Press on Demand, 2006.
  • [Sem96] Stephen Semmes. On the nonexistence of bi-lipschitz parameterizations and geometric problems about aa_{\infty}-weights. Rev. Mat. Iberoamericana, 12(2):337–410, 1996.
  • [Sha00] Nageswari Shanmugalingam. Newtonian spaces: an extension of Sobolev spaces to metric measure spaces. Rev. Mat. Iberoamericana, 16(2):243–279, 2000.
  • [Sha01] Nageswari Shanmugalingam. Harmonic functions on metric spaces. Illinois J. Math., 45(3):1021–1050, 2001.
  • [Sim83] Leon Simon. Lectures on geometric measure theory, volume 3 of Proceedings of the Centre for Mathematical Analysis, Australian National University. Australian National University, Centre for Mathematical Analysis, Canberra, 1983.
  • [Ste93] Elias M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
  • [Str76] Kurt Strebel. On the existence of extremal Teichmueller mappings. J. Analyse Math., 30:464–480, 1976.
  • [V7̈1] Jussi Väisälä. Lectures on nn-dimensional quasiconformal mappings. Lecture Notes in Mathematics, Vol. 229. Springer-Verlag, Berlin-New York, 1971.
  • [Vit12] Davide Vittone. Lipschitz surfaces, perimeter and trace theorems for BV functions in Carnot-Carathéodory spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 11(4):939–998, 2012.
  • [Š60] Boris V. Šabat. The modulus method in space. Soviet Math. Dokl., 1:165–168, 1960.
  • [War83] Frank W. Warner. Foundations of differentiable manifolds and Lie groups, volume 94 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1983. Corrected reprint of the 1971 edition.