We define harmonic maps between sub-Riemannian manifolds by generalizing known definitions for Riemannian manifolds. We establish conditions for when a horizontal map into a Lie group with a left-invariant metric structure is a harmonic map. We show that sub-Riemannian harmonic maps can be abnormal or normal, just as sub-Riemannian geodesics. We illustrate our study by presenting the equations for harmonic maps into the Heisenberg group.
Citation: Erlend Grong, Irina Markina. Harmonic maps into sub-Riemannian Lie groups[J]. Communications in Analysis and Mechanics, 2023, 15(3): 515-532. doi: 10.3934/cam.2023025
[1] | Xiaotian Hao, Lingzhong Zeng . Eigenvalues of the bi-Xin-Laplacian on complete Riemannian manifolds. Communications in Analysis and Mechanics, 2023, 15(2): 162-176. doi: 10.3934/cam.2023009 |
[2] | Vladimir Rovenski . Generalized Ricci solitons and Einstein metrics on weak $ K $-contact manifolds. Communications in Analysis and Mechanics, 2023, 15(2): 177-188. doi: 10.3934/cam.2023010 |
[3] | Velimir Jurdjevic . Time optimal problems on Lie groups and applications to quantum control. Communications in Analysis and Mechanics, 2024, 16(2): 345-387. doi: 10.3934/cam.2024017 |
[4] | Floyd L. Williams . From a magnetoacoustic system to a J-T black hole: A little trip down memory lane. Communications in Analysis and Mechanics, 2023, 15(3): 342-361. doi: 10.3934/cam.2023017 |
[5] | Jinguo Zhang, Shuhai Zhu . On criticality coupled sub-Laplacian systems with Hardy type potentials on Stratified Lie groups. Communications in Analysis and Mechanics, 2023, 15(2): 70-90. doi: 10.3934/cam.2023005 |
[6] | Jonas Schnitzer . No-go theorems for $ r $-matrices in symplectic geometry. Communications in Analysis and Mechanics, 2024, 16(3): 448-456. doi: 10.3934/cam.2024021 |
[7] | Ying Chu, Bo Wen, Libo Cheng . Existence and blow up for viscoelastic hyperbolic equations with variable exponents. Communications in Analysis and Mechanics, 2024, 16(4): 717-737. doi: 10.3934/cam.2024032 |
[8] | Richard Cushman . Normalization and reduction of the Stark Hamiltonian. Communications in Analysis and Mechanics, 2023, 15(3): 457-469. doi: 10.3934/cam.2023022 |
[9] | Zhiyong Wang, Kai Zhao, Pengtao Li, Yu Liu . Boundedness of square functions related with fractional Schrödinger semigroups on stratified Lie groups. Communications in Analysis and Mechanics, 2023, 15(3): 410-435. doi: 10.3934/cam.2023020 |
[10] | Jinli Yang, Jiajing Miao . Algebraic Schouten solitons of Lorentzian Lie groups with Yano connections. Communications in Analysis and Mechanics, 2023, 15(4): 763-791. doi: 10.3934/cam.2023037 |
We define harmonic maps between sub-Riemannian manifolds by generalizing known definitions for Riemannian manifolds. We establish conditions for when a horizontal map into a Lie group with a left-invariant metric structure is a harmonic map. We show that sub-Riemannian harmonic maps can be abnormal or normal, just as sub-Riemannian geodesics. We illustrate our study by presenting the equations for harmonic maps into the Heisenberg group.
A harmonic map between Riemannian manifolds (M,g), dim (M)=m, and (N,h), dim (N)=n, are smooth maps giving the minimum to the energy functional
E(f)=∫Me(f)dμ,e(f)(x)=n∑j=1m∑i=1h(wj,df(vi))2,f:M→N, | (1.1) |
where dμ is the Riemannian volume density on M, {vi}mi=1 is an orthonormal basis in TxM, and {wj}nj=1 is an orthonormal basis in Tf(x)N. Particular examples are maps f:[a,b]→N, describing the Riemannian geodesics in N and harmonic functions f:M→R. Other examples are minimal surfaces. For instance, a minimal surface in R3 can be seen as a harmonic map f:[a,b]×[a,b]→R3; see, e.g., [1,2,3] or more recent survey for minimal submanifolds [4]. The Euler-Lagrange equations of (1.1) correspond to the solution of τ(f)=0, where τ(f)=trg∇∇×df(×) denotes the tension field of f, defined by using an induced connection on T∗M⊗f∗TN from the Levi-Civita connections on respectively M and N. The celebrated result of [5] states that any smooth map f∈C∞(M,N) from a compact Riemannian manifold M to a manifold N of non-positive scalar curvature can be deformed to a harmonic map.
A generalization of this terminology has been suggested for sub-Riemannian manifolds. A sub-Riemannian manifold is a triplet (M,D,g) consisting of a smooth, connected manifold M, a subbundle D of the tangent bundle TM, and a sub-Riemannian metric g defined only on vectors in D. We assume that D is bracket-generating, meaning that sections of D and a sufficient number of their Lie brackets span TxM at each point x∈M. Studies of harmonic maps f:M→N from a sub-Riemannian (M,D,g) into a Riemannian manifold (N,h) of non-positive curvature was made, for instance, in [6,7,8,9]. Here, the energy functional (1.1) is modified by letting v1,…,vm be an orthonormal basis of Dx, and the corresponding equation τ(f)=0 turns to be of a hypoelliptic type. The existence and regularity of the solution to τ(f)=0 was established in [7] under some convexity condition on N, and uniqueness has been shown in [10].
In the present paper we consider harmonic maps allowing the target space to be a sub-Riemannian manifold. Already the study of curves in sub-Riemannian manifolds shows that it is not sufficient to deal exclusively with the Euler-Lagrange equation when it comes to minimizers of (1.1). More precisely, there are examples of curves that are energy minimizers, and hence the length minimizers, which are not solutions to the Euler-Lagrange equation. Such curves are necessarily singular points in the space of curves of finite sub-Riemannian length, also called horizontal curves, fixing two given points. Minimizers that are solutions of the Euler-Lagrange equation are called normal, and they are smooth [11,12]. There are several open questions related to the regularity of minimizers which are singular curves [13]; see also [14]. To simplify the exposition we choose the target sub-Riemannian manifold (N,E,h) to be a Lie group with a left-invariant sub-Riemannian structure (E,h); see also [15], where the target space is a Carnot group. The restriction of the target space to a Lie group allows one to avoid some of the complications of L2 and Sobolev maps between general manifolds; see, e.g., [16, Section 4]. Furthermore, applying the Maurer-Cartan form on a Lie group simplifies calculations and prevents the need of to choose an explicit connection for the target manifold as well. The map f:M→N is required to be horizontal, that is df(D)⊂E. We consider the harmonic maps to be analogous of "normal" and singular geodesics, based on the study of the maps that are regular or singular points of an analogue of the end-point-map. We finally produce equations for both types of horizontal maps: the singular (or abnormal) maps and the normal, latter being solutions of the Euler-Lagrange equation. We will not address conditions for existence or non-existence of such harmonic maps, rather leaving such questions for future research.
We emphasize that we consider a sub-Riemannian analogue of (1.1) which is only defined for horizontal maps and a map f is harmonic if it is a critical value under horizontal variations. See (4.1) for the definition of the sub-Riemannian energy functional. Such an approach can be considered as the limiting case when the length of vectors outside of E in TN approach infinity. This is in contrast to work in [17, Proposition 5.1] on CR manifolds, which uses an orthogonal projection to define an energy functional for all maps, and where maps are considered harmonic if it is a critical value relative to all variations. The latter can be considered as a limiting case where the length of vectors orthogonal to E in TN approach zero. However, we note that if a map f is horizontal and harmonic in the sense of the definition in [17], then f will also be harmonic according to our definition, as being critical under all variations implies that f is also critical with respect to horizontal variations.
The structure of the paper is as follows. In Section 2 we introduce sub-Riemannian manifolds, sub-Riemannian measure spaces, and connections compatible with such structures. In Section 3, we define horizontal maps from a compact sub-Riemannian measure space into a Lie group with a left-invariant sub-Riemannian structure, and we show the Hilbert manifold structure of the space of these maps. For the rest of the paper, we use the the convention that M is compact, which ensures that the functional in (1.1) is finite. Similar to what is done for Riemannian harmonic maps (see, e.g., [18, Section 2]) the case of M non-compact can be considered by calling f harmonic if it is a critical value of the energy functional when restricted to any (relatively) compact subdomain. For simplicity, we will also assume that M is simply connected. See Remark 3 where we suggest modifications for a non-simply connected manifold. We introduce the idea of regular and singular maps, as well as some conditions for these maps. Finally, in Section 4, we find equations for both the normal and abnormal harmonic maps. We show that these equations are a natural generalization of above-mentioned cases of maps into Riemannian manifolds, as well as abnormal and normal sub-Riemannian geodesics. We also give an explicit differential equation for harmonic maps into the Heisenberg group.
A sub-Riemannian manifold is a triple (M,D,g) where M is a connected manifold, D is a subbundle of TM and g=⟨⋅,⋅⟩g is a metric tensor defined on sections of D. Throughout the paper, unless otherwise stated, the subbundle D is assumed to be bracket-generating, meaning that the sections of D and their iterated brackets span the tangent space at each point of M. This condition is sufficient to ensure that any pair of points x0 and x1 in M can be connected by a horizontal curve γ, i.e., an absolutely continuous curve such that ˙γ(t)∈Dγ(t) for almost every t; see [19,20]. Thus, the distance
dg(x0,x1)=inf{∫10|˙γ(t)|gdt:γ is horizontalγ(0)=x0,γ(1)=x1} | (2.1) |
is well defined. Furthermore, the metric topology with respect to dg coincides with the manifold topology on M. We do not exclude the possibility D=TM.
Associated with the sub-Riemannian metric g, there is a vector bundle morphism
♯g:T∗M→D, | (2.2) |
defined by
α(v)=⟨♯gα,v⟩g |
for any x∈M, α∈T∗xM, and v∈Dx. Define a cometric g∗=⟨⋅,⋅⟩g∗ on T∗M by
⟨α,β⟩g∗=α(♯gβ)=⟨♯gα,♯gβ⟩g,α,β∈Γ(T∗M). |
This cometric is exactly degenerated along the subbundle Ann(D)⊆T∗M of covectors vanishing on D. Conversely, given a cometric g∗ on T∗M that is degenerated along a subbundle of T∗M, we can define the subbundle D of TM as the image of the map ♯g:α↦⟨α,⋅⟩g∗ in (2.2), and a metric g on D by the relation
⟨♯gα,♯gβ⟩g=⟨α,β⟩g∗. |
Hence, a sub-Riemannian manifold can equivalently be defined as a connected manifold with a symmetric positive semi-definite tensor g∗ on Γ(TM⊗2) degenerating on a subbundle of T∗M. In what follows, we will speak about a sub-Riemannian structure interchangeably as (D,g) or g∗, assuming that the subbundle D is bracket-generating. For more on sub-Riemannian manifolds, see, e.g., [14,21].
Definition 2.1. A sub-Riemannian measure space (M,D,g,dμ) is a sub-Riemannian manifold (M,D,g) with a choice of smooth volume density dμ on M. If D=TM, then dμ is the volume density of the Riemannian metric g.
On a sub-Riemannian measure space (M,D,g,dμ) there is a unique choice of second order operator
Δg,dμϕ=dvdμ♯gdϕ,ϕ∈C∞(M). | (2.3) |
We call the operator in (2.3) the sub-Laplacian of the sub-Riemannian measure space. Since D is bracket-generating, the classical result of Hörmander [22] states that Δg,dμ is a hypoelliptic operator. If the measure dμ is clear from the context, we simply write Δg. We also denote the sub-Riemannian measure space as (M,g∗,dμ).
We say that a Riemannian metric ˉg=⟨⋅,⋅⟩ˉg is a taming metric of (M,D,g,dμ) if g is the restriction ˉg|D of ˉg to D and the volume density of ˉg equals dμ.
Lemma 2.2. Any sub-Riemannian measure space has a taming Riemannian metric.
Proof. Let (M,D,g,dμ) be any sub-Riemannian measure space with dim M=m and rank D=k. If k=m, then by convention ˉg=g is a taming Riemannian metric. For k<m, we take an arbitrary Riemannian metric ˉg0 on M and let D⊥ denote the orthogonal complement of D with respect to ˉg0. The rank of D⊥ equals m−k. Define a Riemannian metric ˉg1 such that D and D⊥ are still orthogonal with respect to ˉg1, and ˉg1|D=g, ˉg1|D⊥=ˉg0|D⊥. Let dˉμ be the Riemannian volume density with respect to ˉg1, and write dˉμ=ρdμ.
Finally, we define the metric ˉg to be such that D and D⊥ are orthogonal with respect to ˉg and
ˉg|D=g,ˉg|D⊥=ρ−1/(m−k)ˉg1|D⊥, |
which gives us the desired Riemannian metric.
Remark 2.3 (Hausdorff and Popp's measure). A manifold M carries a measure dx which is the pushforward of the Lebesgue measure by the chart map. The distance dg in (2.1) generated by the sub-Riemannian metric tensor g produces the Hausdorff measure dH. Relative to any coordinate system defined sufficiently close to a regular point, dH=q(x)dx is absolutely continuous with respect to dx. It is not clear whether q is a smooth function. Another construction of a measure near regular point has been provided by O. Popp (see [21, Chapter 10]) which gives a measure dμ with a smooth Radon-Nikodym derivative with respect to dx. The latter allows one to define the sub-Laplacian by making use of the integration by parts with respect to the smooth measure dμ, which leads to the sub-Laplacian introduced in [23]. For the case of the Carnot groups, both the Hausdorff and the Haar measures are equal up to a constant, and are hence all smooth.
Consider a sub-Riemannian structure g∗ on M. For a two-tensor field ξ∈Γ(T∗M⊗2) we write
trgξ(×,×)=ξ(g∗),i.e., trgξ(×,×)(x)=k∑i=1ξ(vi,vi) |
for an arbitrary orthonormal basis v1,…,vk of Dx with k=rank D. We want to consider connections on sub-Riemannian manifolds and sub-Riemannian measure spaces. We begin with the following definition of a connection on tensor fields; see, for instance, [24, Chapter 4].
Definition 2.4. Let ∇ be an affine connection on TM.
(a) We say that ∇ is compatible with (D,g) (equiv. g∗) if it satisfies the following equivalent conditions:
(i) ∇g∗=0,
(ii) ∇♯g=♯g∇,
(iii) For any X1,X2∈Γ(D), Z∈Γ(TM), we have that ∇ZX1∈Γ(D) and
Z⟨X1,X2⟩g=⟨∇ZX1,X2⟩g+⟨X1,∇ZX2⟩g. |
(b) We say that ∇ is compatible with (D,g,dμ) (equiv. (g∗,dμ)) if ∇ is compatible with (D,g) (equiv. g∗) and for any ϕ∈C∞(M)
trg∇×dϕ(×)=k∑i=1∇vidϕ(vi)=Δg,dμϕ |
for an orthonormal basis v1,…,vk of Dx.
The following is known on sub-Riemannian manifolds.
Proposition 2.5. [25] Let g∗ be a sub-Riemannian structure and dμ a volume density on M. Then (g∗,dμ) has a compatible connection.
We also prove the following result.
Lemma 2.6. (a) A connection ∇ is compatible with g∗ if and only if for every point x∈M there exists a local orthonormal frame X1,…,Xk of D around x such that ∇Xj(x)=0.
(b) A connection ∇ is compatible with (g∗,dμ) if and only if for every point x∈M there exists a local orthonormal frame X1,…,Xk of D around x such that ∇Xj(x)=0 and dvdμXi(x)=0.
Proof. If ∇ preserves D, then ∇|D is a connection on D preserving the inner product g. Hence there is a local orthonormal frame of D that is parallel with respect to ∇ at a given point x; see, e.g., [26, Theorem 2.1 and Remark 2.2] for details. Conversely, let α be an arbitrary one-form and x∈M an arbitrary point. Assume that there exists an orthonormal frame X1,…,Xk of D around x such that it is ∇-parallel at x. Completing calculations at x, we obtain
X|α|2g∗(x)=k∑i=1X(α(Xi))2(x)=2k∑i=1α(Xi)(x)(∇Xα)(Xi)(x)=2⟨α,∇Xα⟩g∗(x). |
If we can find such a basis for every point in M, it follows that ∇ is compatible with g∗. This proves (a).
The result in (b) follows from the identity
Δg,dμf=k∑i=1X2if+k∑i=1(Xif)dvdμXi, |
that holds for any local orthonormal basis of M.
Corollary 2.7. Let ∇ be a connection compatible with (g∗,μ) and let X be a horizontal vector field. Then
dvdμX=trg⟨∇×X,×⟩g. |
Proof. For a given point x∈M, choose an orthonormal frame X1,…,Xk of D around x with ∇Xi(x)=0 and dvdμXi(x)=0. Write X=∑ki=1fiXi. Then
dvdμX(x)=k∑i=1(Xifi(x)+fi(x)dvdμXi(x))=k∑i=1Xifi(x)=k∑i=1⟨∇XiX,Xi⟩g(x). |
Since x∈M was arbitrary, the result follows.
Let G be a Lie group with Lie algebra g. Let (E,h) be a sub-Riemannian structure on G. We say that the sub-Riemannian structure is left-invariant if E is a left-invariant distribution and if
⟨v,w⟩h=⟨a⋅v,a⋅w⟩h,for any a∈G, v,w∈E1=e⊆g, |
where we denote by a⋅v the action on v∈g by the differential of the left translation by a∈G, and 1∈G is the identity element. Equivalently, let ω be the left Maurer-Cartan form, given by ω(v)=a−1⋅v∈g for any v∈TaG. Then v∈E if and only if ω(v)∈e=E1⊆g. We then say that (E,h) is obtained by left translation of (e,⟨⋅,⋅⟩).
Example 2.8 (The Heisenberg group). We consider the space Hn=R2n+1 with coordinates (a,b,c)=(a1,…,an,b1,…,bn,c). We give this space a global frame
Aj=∂aj−12bj∂c,Bj=∂bj+12aj∂c,C=∂c. | (2.4) |
The corresponding coframe is given by daj, dbj and θ=dc+12∑nj=1(bjdaj−ajdbj). Note the bracket relations
[C,Ai]=[C,Bj]=[Ai,Aj]=[Bi,Bj]=0,[Ai,Bj]=δijC. | (2.5) |
Hence, these vector fields form a Lie algebra which we will write as hn. We can give R2n+1 a group structure such that the vector fields in (2.4) become left-invariant. The group multiplication is given by
(a,b,c)⋅(˜a,˜b,˜c)=(a+˜a,b+˜b,c+˜c+12(⟨a,˜b⟩Rn−⟨˜a,b⟩Rn)). |
We will define a sub-Riemannian structure (E,h) on Hn by letting A1,…,An, B1,…,Bn be an orthonormal basis.
In what follows, we will let Ωp(M,g) be the space of g-valued differential p-forms on a manifold M. We recall the definition and properties of the Darboux derivative; referring to [27] for more details. Let G be a Lie group with Lie algebra g. Let ω∈Ω1(G,g) be the left Maurer-Cartan form as defined in Section 2.3. This form satisfies the left Maurer-Cartan equation
dω+12[ω,ω]=0, | (3.1) |
with [ω,ω] being the two-form (v,w)↦2[ω(v),ω(w)]. See Appendix A for more details. If M is a given manifold and f:M→G is a smooth map, we say that αf:=f∗ω is the left Darboux derivative of f. It follow from definition that αf satisfies (3.1). Conversely, if β∈Ω1(M,g) satisfies dβ+12[β,β]=0, then locally β is the Darboux derivative of some function. If the monodromy representation of β is trivial (see [27, Chapter 3, Theorem 7.14]) then the structural equation implies that β=αf for some map f:M→G. Particularly, for a connected, simply connected manifold M the monodromy representation of any g-valued one-form is trivial, meaning that any form satisfying the left Maurer-Cartan equation can be represented as a Darboux derivative. Through the rest of the paper, we assume that M is connected and simply connected.
Denote by A⊆Ω1(M,g) the collection of forms α satisfying dα+12[α,α]=0, and define
TαA={˙β(0):β:(−ε,ε)→Ω1(M,g) is smooth,β(0)=α,β(t)∈A for any t∈(−ε,ε).}. |
Lemma 3.1. We have
TαA={dF+[α,F]:F∈C∞(M,g)}. |
Proof. Let β(t) be a differentiable curve in Ω1(M,g) and assume that dβ(t)+12[β(t),β(t)]=0. If we differentiate this relation and assume that β(0)=α and ˙β(0)=η, then
dη+[α,η]=0. |
If α=f∗ω and we write η=Ad(f−1)˜η, then
Ad(f−1)d˜η=0. | (3.2) |
We remark that here we are abusing notation to write η|x=Ad(f(x)−1)˜η|x, where f(x)−1 is the inverse of f(x) with respect to the group operation in G. To see that (3.2) holds, recall first that for any curve A(t) in a Lie group G, we have that ddtAd(A(t)−1)=−ad(ω(˙A(t)))Ad(A(t)−1), where ω is the Maurer-Cartan form for the group G. Considering the special case where A(t)=f(γ(t)) for an arbitrary smooth curve γ(t) in M with f∗ω=α, we get the formula for the differential
dAd(f−1)=−ad(f∗ω)Ad(f−1)=−ad(α)Ad(f−1). |
Using the definition of η=Ad(f−1)˜η, this leads to
0=dη+[α,η]=(dAd(f−1))∧˜η+Ad(f−1)d˜η+[α,Ad(f−1)˜η]=Ad(f−1)d˜η. |
In summary, the form ˜η∈Ω1(M,g) is closed and we can find a function ˜F:M→g such that ˜η=d˜F due to the vanishing de Rham cohomology; see [28, Theorem 11.14]. Furthermore, if we define F=Ad(f−1)˜F, then
β=Ad(f−1)d˜F=dF+[α,F]. |
Conversely, for any F∈C∞(M,g), we can define a curve g(t)=f⋅exp(tF) in the space of smooth maps M→G and β(t)=g(t)∗ω. Here exp:g→G is the group exponential. Let v∈TxM, x∈M, be arbitrary and define γ(s):(−ϵ,ϵ)→M, ϵ>0, as a curve with γ(0)=x and ∂sγ(0)=v. If we set Γ(s,t)=g(t)(γ(s)), then we compute
˙β(0)(v)=∂tω(∂sΓ(s,t))|(s,t)=(0,0)=∂tΓ∗ω(∂s)|(s,t)=(0,0)=(∂sΓ∗ω(∂t)−d(Γ∗ω)(∂s,∂t))|(s,t)=(0,0)=(∂s(ω(∂tΓ))−d(Γ∗ω)(∂s,∂t))|(s,t)=(0,0)=∂sF(γ(s))|s=0+[ω(∂sΓ),ω(∂tΓ)]|(s,t)=(0,0)=dF(v)+[α(v),F(x)]. |
Recall that α=f∗ω. The result follows.
We want to close our space of Darboux derivatives into a Hilbert space. Let (M,D,g,dμ) be a sub-Riemannian measure space and let ˉg be a taming Riemannian metric. Extend the inner product on e to a full inner product on g. These choices give us an induced inner product on ∧kT∗M⊗g, which allows us to define an L2-inner product ⟨β,β⟩=∫M⟨β(x),β(x)⟩dμ(x) for any β∈Ωk(M,g). With this definition, we consider L2Ω(M,g)=⊕dim Mk=0L2Ωk(M,g) as the space of L2-forms with values in g. We remark that since M is compact and g is finite dimensional, any other choice of taming Riemannian metric ˉg and inner product of g will give us an equivalent L2-inner product, meaning in particular that L2-forms are independent of these choices. More about the theory of Lp forms can be found, for instance in [29].
Corollary 3.2. Assume that M is simply connected and compact. Then the closure ¯A of A in L2Ω1(M,g) is a Hilbert submanifold of L2Ω1(M,g) with tangent space ¯TαA⊆L2Ω1(M,g).
Proof. Let α∈A be the Darboux derivative α=αf=f∗ω of a map f. We consider an arbitrary curve β∈A that can be written as β=dF+[α,F] for some F∈C∞(M,g). From the proof of Lemma 3.1, we note that if F=Ad(f−1)˜F, then Ad(f−1)β=d˜F. We denote by ˜Fβ a unique solution to this equation satisfying ∫M˜Fβdμ=0, and we define Fβ=Ad(f−1)˜Fβ. Then,
‖Fβ‖L2≤‖Ad(f−1)‖L∞‖˜Fβ‖L2Poincaré≤C‖Ad(f−1)‖L∞‖d˜Fβ‖L2≤C‖Ad(f−1)‖L∞‖Ad(f)‖L∞‖β‖L2 |
for some constant C>0. Note that the linear map β↦˜Fβ is bounded and invertible with respect to the L2 metric, which, in particular, is smooth. Here we have used the Poincaré inequality for compact Riemannian manifolds found in, e.g., [30, Theorem 2.10]. It follows that this map can be extended by limits to be well defined as a map from ¯TαA to {F∈L2(M,g):∫MAd(f)Fdμ=0}.
Continuing, we introduce a map Φ:TαA↦A as
Φ(β)=(f⋅eFβ)∗ω,forβ∈TαA. |
We observe then that for any v∈TxM, we can apply the formula for the differential of the Lie group exponential to obtain
Φ(β)(v)=e−Fβ(x)f(x)−1df(v)eFβ(x)+1−e−ad(Fβ(x))ad(Fβ(x))dFβ(v)=e−ad(Fβ(x))α+1−e−ad(Fβ(x))ad(Fβ(x))(β(v)+ad(Fβ(x))α(v))=α(v)+1−e−ad(Fβ(x))ad(Fβ(x))β(v), |
meaning that
Φ(β)=α+1−e−ad(Fβ)ad(Fβ)β=α+∞∑n=0(−1)nad(Fβ)n(n+1)!β. |
This map is well defined for any β∈¯TαA, giving a smooth map Φ:¯TαA→¯A⊆L2Ω1(M,g). Furthermore, we see that its Fréchet differential at β=0 is given by
DΦ|0(β)=β. |
Thus, Φ is locally injective, so it can be used as a chart close to 0∈¯TαA. Since A is dense in ¯A the result follows.
For the rest of this section, (M,D,g,dμ) will be a simply connected, compact sub-Riemannian measure space while G will be a Lie group with Lie algebra g and left Maurer-Cartan form ω. The structure (E,h) on G will be defined by left translation of (e,⟨⋅,⋅⟩). We introduce the following concept.
Definition 3.3. Let (M,D,g) and (N,E,h) be two sub-Riemannian manifolds. We say that a smooth map f:M→N is horizontal if df(D)⊆E.
To simplify the discussion in this paper, we only consider the case when N=G is a Lie group G with a left-invariant sub-Riemannian structure (E,g) that is the left translation of a vector space e⊂g and a scalar product ⟨⋅,⋅⟩ on e. Then, f:M→G is horizontal if and only if αf=f∗ω sends D into e. We write AD,E for the collection of such forms αf.
Consider Ω1(D,V)=Γ(D∗⊗V) as partial one-forms only defined on D with values in a vector space V. Write L2Ω1(D,V) for its L2-closure. Consider ¯AD,E⊆¯A. Define a linear map
P:L2Ω1(M,g)→L2Ω1(D,g/e)byP(α)=α|Dmode. | (3.3) |
Then ¯AD,E=kerP∩¯A.
Definition 3.4. We say that α∈¯AD,E is regular (respectively, singular) if α is a regular (respectively, singular) point of P|A; that is the differential of the map P is surjective (not surjective) at α∈A. We say that a sub-Riemannian horizontal map f:M→G is regular (respectively, singular), if its Darboux derivative αf∈AD is regular (respectively, singular).
Since ¯AD,E=kerP∩¯A, the implicit function theorem implies that ¯AD,E has the local structure of a manifold around any regular α.
We list the conditions for distributions on M and G, which guarantee the absence of singular morphisms.
Definition 3.5. We say that e⊂g is a strongly q-bracket generating subspace of g if for any 1≤l≤q and any set of linearly independent vectors A1,…,Al∈e and any Z1,…,Zl∈g, there exists an element B∈e such that
Zj−[Aj,B]∈e,j=1,…,l. |
Example 3.6. Let θ be a left-invariant one-form on a (2n+1)-dimensional Lie group G and define kerθ|1=e⊆g. Assume that dθ|(∧2e) is non-degenerate, i.e., θ is a contact form on G. We can find a basis A1,…,An,B1,…,Bn of e such that dθ(Ai,Aj)=dθ(Bi,Bj)=0 and dθ(Bi,Aj)=δij. Let Z∈g be the unique element satisfying θ(Z)=1 and dθ(Z,⋅)=0. To find B∈e, we need to solve the equations
[pjAi,B]=˜pjZ,[qjBi,B]=˜qjZ,pj≠0,qj≠0. |
One can easily check that
B=n∑i=1(˜pipiBi−˜qiqiAi). |
is a solution. This shows that such structures are strongly 2n-bracket generating. In particular, we note that the Heisenberg group Hn is has a strong 2n-bracket generating distribution.
Proposition 3.7. Let (M,D,g) be a sub-Riemannian manifold, where M is simply connected and D has rank k≥2. Let g be a Lie algebra with a generating subspace e⊆g of positive codimension. Let α∈Ω1(M,g) be a one-form satisfying α(D)⊆e.
(a) Assume that there exists a non-intersecting horizontal loop γ:[0,1]→M such that α(˙γ(t))=0 for almost every t∈[0,1]. Then α is singular.
(b) Assume that e⊆g is strongly k-bracket generating. If α|D is injective at every point, then α is regular.
Proof. Choose a complement f to e in g. Let F∈C∞(M,g) be a function and write F=Fe+Ff according to the decomposition g=e⊕f. Recall that the regularity of α is equivalent to the assumption that for any one-form ψ∈Ω1(M,g), one can choose Fe and Ff such that
dFf|D+[α|D,Fe]+[α|D,Ff]=ψ|Dmode. | (3.4) |
(a) If γ:[0,1]→M is a non-intersecting horizontal loop, then x1=γ(1/2)≠x0=γ(0)=γ(1) by assumption. Define γ1,γ2:[0,1]→M by γ1(t)=γ(t/2) and γ2(t)=γ(1−t/2), which are non-intersecting horizontal curves from x0 to x1. Let U be an open set that does not intersect γ2, but intersects with a subset of γ1 of positive length. Let ψ=ψ0⊗Z, where Z∈f, Z≠0, and ψ0 denotes a real valued one-form with support in U such that C=∫10ψ0(˙γ(t))dt>0. If we find a function F=Fe+Ff solving (3.4) then
Ff(x1)−Ff(x0)=∫γ1ψ=CZ≠0. |
However, in order for (3.4) to hold, we would also need to Ff(x1)−Ff(x0) tp equal ∫γ2ψ which is clearly 0 by the definition of ψ, giving us a contradiction.
(b) If α|D is injective, then we can choose Ff=0. To show the regularity of α we need to solve the equation [α|D,Fe]=ψ|Dmode. The assumption of D being strongly k-bracket generating implies that the equation [α|D,Fe]=ψ|Dmode has a solution for any one-form ψ∈Ω1(M,g). To be more precise, let X1, …, Xk be a local basis of D and ψ∈Ω1(M,g). We respectively define Aj∈C∞(M,e) and Zj∈C∞(M,g) by Aj=α(Xj) and Zj=ψ(Xj), j=1,…,k. Then we can then define Fe such that [Aj,Fe]=Zj by the strongly k-bracket generating condition on D.
Let (M,D,g,dμ) be a given sub-Riemannian measure space and let (G,E,h) be a Lie group with a left-invariant sub-Riemannian structure. For a horizontal map f:M→G with Darboux derivative αf, we define its energy as
E(f)=12∫M|df|2g∗⊗f∗hdμ=12∫Mtrg(f∗h)(×,×)dμ=12∫M|αf|2g∗dμ=:ˆE(αf). | (4.1) |
We note that if v1,…,vk∈Dx and w1,…,wn∈Ef(x) are respective orthonormal bases then
|df|2g∗⊗f∗h(x)=k∑i=1|df(vi)|2h=n∑j=1k∑i=1⟨wj,df(vi)⟩2h. |
We would generalize the definition of harmonic maps from [5] to the sub-Riemannian setting, saying that f is harmonic if it is a critical value of E. Instead, we use the Darboux derivative to make this definition precise. For α∈¯AD,E, we define a variation αs of α as a differentiable curve (−ε,ε)→¯AD,E, s↦αs, such that α0=α.
Definition 4.1. We say that α∈¯AD,E is harmonic if it is a critical point of ˆE, i.e., for every variation αs of α, we have ddsˆE(αs)|s=0=0. We say that f is harmonic if αf is harmonic.
We have the following result.
Theorem 4.2. Let M be a simply connected, compact manifold, and ∇ a connection compatible with the sub-Riemannian measure space (M,g∗,dμ). Let the map ♯=♯h1:g∗→e correspond to the sub-Riemannian metric h at the identity. Assume that α∈AD,E is harmonic. Then at least one of the following statements holds.
(a) (Abnormal case) There exists form a η∈Ω1(M,g∗) with η|D non-zero, satisfying ♯η|D=0 and
δDη−trgad∗(α(×))η(×)=0. |
with δDη=−trg∇×η(×).
(a) (Normal case) There exists a form λ∈Ω1(M,g∗) satisfying ♯λ|D=α|D and
δDλ−trgad∗(α(×))λ(×)=0. |
with δDλ=−trg∇×λ(×).
Recall that ad∗ denotes the adjoint representation of g on g∗ given by (ad∗(A)β)(B)=−β([A,B]) for any A,B∈g, β∈g∗.
Remark 4.3. We remark the following about the result of Theorem 4.2.
● Case (a), which we call abnormal, occurs when α=αf∈AD,E is singular. It is a property that holds for all singular elements, and it is not related to optimality. The proof of the result in (a) does not use the property that α is a harmonic form.
● Case (b), which is called normal, occurs when α=αf∈AD,E is both regular and is a critical value of ˆE. However, there are also singular forms α∈AD,E that are critical values of ˆE and also have a corresponding λ∈Ω1(M,g∗) satisfying the equations as in (b), but such an extremal form α is not called normal. Thus, Cases (a) and (b) are not completely disjoint.
● We remark also that the results of Theorem 4.2 only depend on restrictions α|D and λ|D of forms to D and do not depend on their extension to the entire tangent bundle. Hence, we could have stated Theorem 4.2 by using only the restrictions α|D and λ|D.
Before proceeding to the proof, we observe how the result of Theorem 4.2 satisfies known examples in literature.
Example 4.4. We note that if E=TG, so that G is a Riemannian Lie group, then there cannot exist any abnormal harmonic maps. Indeed, since ♯ is now is a bijective map, we cannot have that η|D≠0, while still having that ♯η|D=0.
For the normal case, we have ♯λ=α and if α=αf, then
δDα+trgad(α(×))†α(×)=0, | (4.2) |
where ad(α(×))† is the transpose map with respect to left-invariant metric h on G. This is just the classical tension field equation for harmonic maps. In order to explain this, we write (4.2) as
τ(f)=trg∇∇×df(×)=0,∇∇=∇⊗f∗∇h. | (4.3) |
Here, ∇h is the Levi-Civita connection on G, which for left-invariant vector fields, can be written as,
2∇hAB=ad(A)B−ad(A)†B−ad(B)†A. |
Furthermore, ∇∇ is the induced connection on T∗M⊗f∗TG, which can be described by
(∇∇Xdf)(Y)=∇hdf(X)df(Y)−df(∇XY). |
Equation (4.3) coincides with the tension field τ(f) for maps between the Riemannian manifolds in [5] or from sub-Riemannian manifolds to Riemannian manifolds in [7,9]. For the special case G=R, we have that τ(f)=Δg,dμf.
Example 4.5. Consider M=[0,1]. Although this is not within the scope of the theorem, as M is a Riemannian manifold with boundary, the theorem is still valid under the assumption that any variation is constant on ∂M={0,1}. Define ∇ℓ to be the left-invariant connection on G, i.e., the connection such that ∇ℓA=0 for any left-invariant vector field A. This connection is compatible with the sub-Riemannian structure (E,h). Let Tℓ be the torsion of ∇ℓ, given for left-invariant vector fields by
Tℓ(A,B)=−ad(A)B. |
We say that the adjoint connection to ∇ℓ is given by ˆ∇ℓAB=∇ℓAB−Tℓ(A,B). For the special case of ∇ℓ, its adjoint will be the right invariant connection. If f:[0,1]→G, then the equation in Theorem 4.2 (a) is written as
ˆ∇˙fη=0,♯hη=0, |
where η is a one-form along f(t). The equation in Theorem 4.2 (b) becomes
ˆ∇˙fλ=0,♯hλ=˙f. |
These are the respective equations for abnormal curves and normal geodesics see [31,32] for details.
Proof of Theorem 4.2. Recall that we have defined L2-forms with respect to a taming metric ˉg and an inner product on g. Assume that α∈AD,E is harmonic. Write
Q={η∈Ω1(M,g∗):η(D)⊆Ann(e)},Λα={λ∈Ω1(M,g∗):♯λ|D=α|D}. |
Note that Q is a vector space, while Λα is an affine space with λ1−λ2∈Q for λ1,λ2∈Λα. We first observe the following. Consider the operator LαF:=dF+[α,F]. Then for any η∈L2Ω1(M,g∗), we note that
∫Mtrgη(×)dF(×)dμ=∫Mtrgη(×)∇×Fdμ=∫M(δDη)Fdμ; |
hence
∫Mtrgη(×)(LαF)(×)dμ=∫Mtrgη(×)(dF(×)+[α(×),F])=∫M(δDη−trgad∗(α(×))η(×))Fdμ=:∫M(L∗αη)Fdμ. |
If α∈¯AD,E is singular, then there is a non-zero form ˇη∈L2Ω1(D,g/e) orthogonal to the image of DαP(TαA) where P is given in (3.3). Define η=⟨ˇη,DαP⋅⟩L2∈ˉQ. Then for any element LαF in TαA, F∈C∞(M,g), we have
0=⟨η,DαPLαF⟩=∫Mη(LαF)dμ=∫M(L∗αη)Fdμ. |
As this holds for any F∈C∞(M,g), the result in (a) follows.
If α is regular, then ¯AD,E is locally a manifold with Tα¯AD,E being the closure of elements LαF such that LαF(D)⊆e. In other words, elements in Tα¯AD,E are in the closure of elements LαF, F∈C∞(M,g), that are orthogonal to Q, which can be written as
Tα¯AD,E=¯{LαF:⟨F,ϕ⟩L2=0 for any ϕ∈L∗αQ}. |
Let F be an arbitrary such element in C∞(M,g) that is orthogonal to L∗αQ. For such a tangent vector in Tα¯AD,E, let αs be the corresponding variation with α0=α and ddsαs|s=0=LαF. We observe that for any smooth ˜λ∈Λα,
ddsˆE(αs)|s=0=∫Mtrg⟨α(×),dF(×)+[α(×),F]⟩dμ=∫M(L∗α˜λ)Fdμ. |
If this vanishes for all such variations, then L∗α˜λ∈(C∞(M,g)∩(L∗αQ)⊥)⊥. We remark that since the elements of L2(M,g) can be considered as equivalence classes of sequences (ϕn)∞n=1 of smooth functions convergent in L2, we have
(C∞(M,g)∩(L∗αQ)⊥)⊥=¯L∗αQ, and ¯L∗αQ∩C∞(M,g)=L∗αQ. |
Furthermore, since ˜λ is smooth, then so is L∗α˜λ; hence we can write L∗α˜λ=L∗αη∈L∗αQ. By defining λ=˜λ−η∈Λα, we find that λ satisfies L∗αλ=0.
Remark 4.6. The results in Theorem 4.2 can be generalized to a non simply connected manifold. If M is not simply connected, we consider its universal cover Π:˜M→M. We note that ˜M might not be compact, but, as mentioned in our introduction, we can consider compact subdomains. We can then lift functions from M to ˜M as f↦f∘Π. By using a partition of unity, we decompose an integral over ˜M or one of its compact subdomains as integrals over open sets that are mapped bijectively to an open set in M. It leads to the conclusion that if f is a harmonic map, then so is f∘Π. Looking at the equations in Theorem 4.2, we see that they are all local and can hence be projected to M.
We consider the case of harmonic maps f:M→Hn. Let (a,b,c) be the coordinates on Hn as described in Example 2.8. We then have the following corollary.
Proposition 4.7. Let f:M→Hn be a horizontal map from (M,D,g,dμ) into the Heisenberg group (Hn,E,h) with its standard sub-Riemannian structure. Write
(u,v,w)=(a,b,c)∘f,ζ=u+iv. |
Then f is a normal harmonic if and only if for some horizontal vector field Y∈Γ(M) satisfying dvdμY=0 we have
(Δg,dμ−iY)ζ=0. |
We note that the operator Δg,dμ−iY is hypoelliptic by [33]. However, recall that we are also assuming that f is horizontal, meaning that
(dw+12n∑j=1(vjduj−ujdvj))|D=0. |
Note that if rank D≤2n, then all harmonic maps are normal by Example 3.6. and Proposition 3.7.
Proof. Since f is horizontal, then
α|D=αf|D=n∑j=1duj|D⊗Aj+n∑j=1dvj|D⊗Bj. |
From the requirement that ♯λ|D=α|D, we have that
λ|D=n∑j=1duj|D⊗daj+n∑j=1dvj|D⊗dbj+λ0|D⊗θ, |
where λ0 is a one-form on M. Write Y=♯gλ0 as a vector field. The harmonic equation is given by
0=δDλ−trgad∗(α(×))λ(×)=−n∑j=1trg(∇×duj)(×)⊗daj−n∑j=1trg(∇×dvj)(×)⊗dbj+(∇×λ0)(×)⊗θ+trgλ0(×)(duj(×)⊗dbj)−trgλ0(×)(dvj(×)⊗daj)=−n∑j=1(ΔDuj+Yvj)⊗daj−n∑j=1(ΔDvj−Yuj)⊗dbj+(dvμY)⊗θ. |
It follows that dvμY=0 and
Δg,μζ=(Δg,μuj+iΔg,μvj)j=(−Yvj+iYuj)j=iYζ, |
completing the proof.
Example 4.8. If we choose M=[0,1] in Proposition 4.7, allowing a manifold with boundary, we obtain the normal sub-Riemannian geodesics on the Heisenberg group. More precisely, the horizontality requirement for f can be written as
˙w=−12n∑j=1(vj˙uj−uj˙vj), |
while ζ=u+iv now has to satisfy
¨ζ+iy˙ζ=0,for some constant y. |
In other words ˙ζ=eiyt˙ζ(0), which is exactly the equation for normal geodesics on the Heisenberg group with solutions being the horizontal lifts of circular arcs on the (u,v)-plain.
We recall here the definition of brackets of forms with values in a Lie algebra g. Let α∈Ωk(M,g) be a g valued k-form, that is a section of the vector bundle ∧kT∗M⊗g. Note that all such elements can be written as a finite sum of elements ˇα⊗A where ˇα∈Ωk(M) is a real valued k-form, and A∈g. For ˇα∈Ωk(M), ˇβ∈Ωl(M) and A,B∈g, we define
[ˇα⊗A,ˇβ⊗B]=(ˇα∧ˇβ)⊗[A,B]. |
We can extend this definition by linearity to arbitrary forms α∈Ωk(M,g) and β∈Ωl(M,g), to obtain a form [α,β]∈Ωk+l(M,g). We note that [α,β]=(−1)kl+1[β,α]. We look at the particular case when k=l=1. For α,β∈Ω1(M,g) and a basis A1,…,An of g, we write α=∑nj=1ˇαj⊗Aj and β=∑nj=1ˇβj⊗Aj. We then observe that for any v,w∈TM,
[α,β](v,w)=[β,α](v,w)=n∑i,j=1(ˇαi(v)ˇβj(w)−ˇβj(v)ˇαi(w))[Ai,Aj]=[α(v),β(w)]−[α(w),β(v)]=[α(v),β(w)]+[β(v),α(w)]. |
In particular, [α,α](v,w)=2[α(v),α(w)]. If α is a one-form and F is a zero-form, i.e., a function, then [α,F](v)=[α(v),F(x)] for every v∈TxM.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The first author was by Trond Mohn Foundation-Grant TMS2021STG02 (GeoProCo). The work of both authors was partially supported by the project Pure Mathematics in Norway, funded by the Trond Mohn Foundation and Tromsø Research Foundation. We thank Pierre Pansu and Mauricio Godoy Molina for helpful discussions.
The authors declare there is no conflict of interest.
[1] |
S. Bochner, Harmonic surfaces in Riemann metric, Trans. Amer. Math. Soc., 47 (1940), 146–154. https://doi.org/10.1090/S0002-9947-1940-0001632-1 doi: 10.1090/S0002-9947-1940-0001632-1
![]() |
[2] | R. Courant, Dirichlet's Principle, Conformal Mapping, and Minimal Surfaces, Interscience Publishers, Inc., New York, N.Y., 1950. |
[3] | R. Osserman, A survey of minimal surfaces. Dover Publications, Inc., New York, 1986. |
[4] | H. B Lawson. Jr, Lectures on minimal submanifolds, Vol. I, volume 9 of Mathematics Lecture Series. Publish or Perish, Inc., Wilmington, Del., second edition, 1980. |
[5] |
J. Eells, J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math., 86 (1964), 109–160. https://doi.org/10.2307/2373037 doi: 10.2307/2373037
![]() |
[6] |
E. Barletta, S. Dragomir, H. Urakawa, Pseudoharmonic maps from nondegenerate CR manifolds to Riemannian manifolds, Indiana Univ. Math. J., 50 (2001), 719–746. https://doi.org/10.1512/iumj.2001.50.1931 doi: 10.1512/iumj.2001.50.1931
![]() |
[7] |
Y. Dong, Eells-Sampson type theorems for subelliptic harmonic maps from sub-Riemannian manifolds, J. Geom. Anal., 31 (2021), 3608–3655. https://doi.org/10.1007/s12220-020-00408-z doi: 10.1007/s12220-020-00408-z
![]() |
[8] |
S. Dragomir, D. Perrone, Levi harmonic maps of contact Riemannian manifolds, J. Geom. Anal., 24 (2014), 1233–1275. https://doi.org/10.1007/s12220-012-9371-8 doi: 10.1007/s12220-012-9371-8
![]() |
[9] |
J. Jost, C. J. Xu, Subelliptic harmonic maps, Trans. Amer. Math. Soc., 350 (1998), 4633–4649. https://doi.org/10.1090/S0002-9947-98-01992-8 doi: 10.1090/S0002-9947-98-01992-8
![]() |
[10] | Z. R. Zhou, Uniqueness of subelliptic harmonic maps, Ann. Global Anal. Geom., 17 (1999), 581–594. |
[11] |
U. Hamenstädt, Some regularity theorems for Carnot-Carathéodory metrics, J. Differential Geom., 32 (1990), 819–850. https://doi.org/10.4310/jdg/1214445536 doi: 10.4310/jdg/1214445536
![]() |
[12] |
R. S. Strichartz, Sub-Riemannian geometry, J. Differential Geom., 24 (1986), 221–263. https://doi.org/10.4310/jdg/1214440436 doi: 10.4310/jdg/1214440436
![]() |
[13] | R. Monti, The regularity problem for geodesics of the control distance, In Bruno Pini Mathematical Analysis Seminar 2018, volume 9 of Bruno Pini Math. Anal. Semin., Univ. Bologna, Alma Mater Stud., Bologna, 2018. |
[14] | A. Agrachev, D. Barilari, U. Boscain, A comprehensive introduction to sub-Riemannian geometry, volume 181 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2020. https://doi.org/10.1017/9781108677325 |
[15] |
G. He, J. Li, P. Zhao, Liouville-type theorems for CC-F-harmonic maps into a Carnot group, J. Geom. Anal., 31 (2021), 4024–4050. https://doi.org/10.1007/s12220-020-00424-z doi: 10.1007/s12220-020-00424-z
![]() |
[16] |
R. Schoen, K. Uhlenbeck, Boundary regularity and the Dirichlet problem for harmonic maps, J. Differential Geom., 18 (1983), 253–268. https://doi.org/10.4310/jdg/1214437663 doi: 10.4310/jdg/1214437663
![]() |
[17] |
R. Petit, Mok-Siu-Yeung type formulas on contact locally sub-symmetric spaces, Ann. Global Anal. Geom., 35 (2009), 1–37. https://doi.org/10.1007/s10455-008-9120-1 doi: 10.1007/s10455-008-9120-1
![]() |
[18] |
J. Eells and L. Lemaire. Another report on harmonic maps, Bull. London Math. Soc., 20 (1988), 385–524. https://doi.org/10.1112/blms/20.5.385 doi: 10.1112/blms/20.5.385
![]() |
[19] |
W. L. Chow, Über Systeme von linearen partiellen Differentialgleichungen erster Ordnung, Math. Ann., 117 (1939), 98–105. https://doi.org/10.1007/BF01450011 doi: 10.1007/BF01450011
![]() |
[20] | P. K. Rashevskii, On joining any two points of a completely nonholonomic space by an admissible line, Math. Ann., 3 (1938), 83–94. |
[21] | R. Montgomery, A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002. |
[22] |
L. Hörmander, Hypoelliptic second order differential equations, Acta Math., 119 (1967), 147–171. https://doi.org/10.1007/BF02392081 doi: 10.1007/BF02392081
![]() |
[23] |
A. Agrachev, U. Boscain, J. P. Gauthier, F. Rossi, The intrinsic hypoelliptic Laplacian and its heat kernel on unimodular Lie groups, J. Funct. Anal., 256 (2009), 2621–2655. https://doi.org/10.1016/j.jfa.2009.01.006 doi: 10.1016/j.jfa.2009.01.006
![]() |
[24] | J. M. Lee, Riemannian manifolds, volume 176 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1997. An introduction to curvature. |
[25] |
E. Grong, A. Thalmaier, Stochastic completeness and gradient representations for sub-Riemannian manifolds, Potential Anal., 51 (2019), 219–254. https://doi.org/10.1007/s11118-018-9710-x doi: 10.1007/s11118-018-9710-x
![]() |
[26] |
E. Grong, Curvature and the equivalence problem in sub-Riemannian geometry, Arch. Math. (Brno), 58 (2022), 295–327. https://doi.org/10.5817/AM2022-5-295 doi: 10.5817/AM2022-5-295
![]() |
[27] | R. W. Sharpe, Differential geometry, volume 166 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1997. Cartan's generalization of Klein's Erlangen program, With a foreword by S. S. Chern. |
[28] | J. M. Lee, Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2003. |
[29] |
C. Scott, Lp theory of differential forms on manifolds, Trans. Amer. Math. Soc., 347(1995), 2075–2096. https://doi.org/10.1090/S0002-9947-1995-1297538-7 doi: 10.1090/S0002-9947-1995-1297538-7
![]() |
[30] | E. Hebey, Nonlinear analysis on manifolds: Sobolev spaces and inequalities, volume 5 of Courant Lecture Notes in Mathematics. New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 1999. https://doi.org/10.1090/cln/005 |
[31] |
M. Godoy Molina, E. Grong, Riemannian and sub-Riemannian geodesic flows, J. Geom. Anal., 27 (2017), 1260–1273. https://doi.org/10.1007/s12220-016-9717-8 doi: 10.1007/s12220-016-9717-8
![]() |
[32] | E. Grong, Affine connections and curvature in sub-riemannian geometry, arXiv preprint arXiv: 2001.03817, 2020. |
[33] | B. Helffer, J. Nourrigat, Hypoellipticité maximale pour des opérateurs polynômes de champs de vecteurs, volume 58 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1985. |