Skip to main content
Log in

An application of a global lifting method for homogeneous Hörmander vector fields to the Gibbons conjecture

  • Published:
Nonlinear Differential Equations and Applications NoDEA Aims and scope Submit manuscript

Abstract

In this paper we exploit a global lifting method for homogeneous Hörmander vector fields in order to extend the Gibbons conjecture to any second-order differential operator \(\mathcal {L}_X = \sum _{j = 1}^mX_j^2\), where the \(X_j\)’s are linearly independent smooth vector fields on \(\mathbb {R}^n\) satisfying Hörmander’s rank condition and fulfilling a suitable homogeneity property with respect to a family of non-isotropic dilations. The class of these operators comprehends the sub-Laplacians on Carnot groups, the smooth Grushin-type operators and the smooth \(\Delta _\lambda \)-Laplacians studied by Franchi, Lanconelli and Kogoj. We also establish a comparison result for the solutions of the semi-linear equation \(\mathcal {L}_Xu+f(u) = 0\) under suitable assumptions on the function f.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. This means that, if there exist \(c_1,\ldots ,c_m\in \mathbb {R}\) such that

    $$\begin{aligned} (c_1X_1+\cdots +c_mX_m)(f) \equiv 0 \quad \text {for every } f\in C^\infty (\mathbb {R}^n,\mathbb {R}), \end{aligned}$$

    then \(c_1 = \cdots = c_m = 0\).

  2. We remind that a smooth vector field Y on \(\mathbb {R}^n\) is complete if, for every \(x\in \mathbb {R}^n\), the integral curve of Y starting at x is defined on the whole of \(\mathbb {R}\).

References

  1. Ambrosio, L., Cabré, X.: Entire solutions of semilinear elliptic equations in \({\mathbb{R}}^3\) and a conjecture of De Giorgi. J. Am. Math. Soc. 13, 725–739 (2000)

    Article  Google Scholar 

  2. Barlow, M.T., Bass, R.F., Gui, C.: The Liouville property and a conjecture of De Giorgi. Commun. Pure Appl. Math. 8, 1007–1038 (2000)

    Article  MathSciNet  Google Scholar 

  3. Battaglia, E., Biagi, S., Bonfiglioli, A.: The strong maximum principle and the Harnack inequality for a class of hypoelliptic non-Hörmander operators. Ann. Inst. Fourier (Grenoble) 66, 589–631 (2016)

    Article  MathSciNet  Google Scholar 

  4. Berestycki, H., Hamel, F., Monneau, R.: One-dimensional symmetry of bounded entire solutions of some elliptic equations. Duke Math. J. 103, 375–396 (2000)

    Article  MathSciNet  Google Scholar 

  5. Biagi, S., Bonfiglioli, A.: The existence of a global fundamental solution for homogeneous Hörmander operators via a global lifting method. Proc. Lond. Math. Soc. (3) 114, 855–889 (2017)

    Article  MathSciNet  Google Scholar 

  6. Biagi, S., Bonfiglioli, A.: Global Heat Kernels for Parabolic Homogeneous Hörmander Operators arXiv:1910.09907 (2019)

  7. Biagi, S., Bonfiglioli, A.: Introduction to the Geometrical Analysis of Vector Fields. With Applications To Maximum Principles and Lie Groups. World Scientific Publishing Company, Singapore (2018)

    Book  Google Scholar 

  8. Biagi, S., Bonfiglioli, A., Bramanti, M.: Global Estimates for the Fundamental Solution of Homogeneous Hörmander Sums of Squares. ArXiv: 1906.07836 (2019)

  9. Biagi, S., Bonfiglioli, A., Bramanti, M.: Global estimates in Sobolev spaces for homogeneous Hörmander sums of squares. ArXiv: 1906.07835 (2019)

  10. Biagi, S., Lanconelli, E.: Large sets at infinity and maximum principle on unbounded domains for a class of sub-elliptic operators arXiv:1908.10257 (2018)

  11. Birindelli, I., Lanconelli, E.: A note on one dimensional symmetry in Carnot groups. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Nat. Rend. Lincei (9) Mat. Appl. 13, 17–22 (2002)

    MathSciNet  MATH  Google Scholar 

  12. Birindelli, I., Prajapat, J.: One dimensional symmetry in the Heisenberg group. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 30, 269–284 (2001)

    MathSciNet  MATH  Google Scholar 

  13. Bonfiglioli, A., Lanconelli, E.: On left-invariant Hörmander operators in \(\mathbb{R}^N\): applications to the Kolmogorov–Fokker–Planck equations. J. Math. Sci. 171, 22–33 (2010)

    Article  MathSciNet  Google Scholar 

  14. Bonfiglioli, A., Lanconelli, E., Uguzzoni, F.: Stratified Lie Groups and Potential Theory for Their Sub-Laplacians. Springer Monographs in Mathematics. Springer, Berlin (2007)

    MATH  Google Scholar 

  15. Bony, J.-M.: Principe du maximum, inégalite de Harnack et unicité du problème de Cauchy pour les opérateurs elliptiques dégénérés. Ann. Inst. Fourier (Grenoble) 19, 277–304 (1969)

    Article  MathSciNet  Google Scholar 

  16. Carbou, G.: Unicité et minimalité des solutions d’une équation de Ginzburg–Landau. Ann. Inst. Henri Poincaré Anal. Non Linéaire 12, 305–318 (1995)

    Article  MathSciNet  Google Scholar 

  17. Cesaroni, A., Novaga, M., Valdinoci, E.: A symmetry result for the Ornstein–Uhlenbeck operator. Discrete Contin. Dyn. Syst. 34, 2451–2467 (2014)

    Article  MathSciNet  Google Scholar 

  18. del Pino, M., Kowalczyk, M., Wei, J.: On De Giorgi’s conjecture in dimension \(N\ge 9\). Ann. Math. (2) 174, 1485–1569 (2011)

    Article  MathSciNet  Google Scholar 

  19. Farina, A.: Symmetry for solutions of semilinear elliptic equations in \({\mathbb{R}}^{N}\) and related conjectures. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl. 10, 255–265 (1999)

    MathSciNet  MATH  Google Scholar 

  20. Farina, A., Valdinoci, E.: Rigidity results for elliptic PDEs with uniform limits: an abstract framework with applications. Indiana Univ. Math. J. 60, 121–141 (2011)

    Article  MathSciNet  Google Scholar 

  21. Farina, A., Valdinoci, E.: The State of the Art for a Conjecture of De Giorgi and Related Problems. Recent Progress on Reaction–Diffusion Systems and Viscosity Solutions. World Sci. Publ., Hackensack (2009)

    MATH  Google Scholar 

  22. Franchi, B., Lanconelli, E.: Une métrique associée à une classe d’opérateurs elliptiques dégénérés. Rend. Sem. Mat. Univ. Politec. Torino 105, 114 (1983)

    MATH  Google Scholar 

  23. Franchi, B., Lanconelli, E.: Hölder regularity theorem for a class of linear nonuniformly elliptic operators with measurable coefficients. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 10, 523–541 (1983)

    MathSciNet  MATH  Google Scholar 

  24. Folland, G.B.: On the Rothschild–Stein lifting theorem. Commun. Partial Differ. Eq. 2, 161–207 (1977)

    Article  MathSciNet  Google Scholar 

  25. Ghoussoub, N., Gui, C.: On a conjecture of De Giorgi and some related problems. Math. Ann. 311, 481–491 (1998)

    Article  MathSciNet  Google Scholar 

  26. Gibbons, G.W., Townsend, P.K.: Bogomol‘nyi equation for intersecting domain walls. Phys. Rev. Lett. 83, 1727–1730 (1999)

    Article  Google Scholar 

  27. Kogoj, A.E., Lanconelli, E.: On semilinear \({\Delta }_{\lambda }\)-Laplace equation. Nonlinear Anal. 75, 4637–4649 (2012)

    Article  MathSciNet  Google Scholar 

  28. Kogoj, A.E., Lanconelli, E.: Linear and semilinear problems involving \(\Delta _\lambda \)-Laplacians. In: Proceedings of the International Conference “Two Nonlinear Days in Urbino 2017”, Electronic Journal of Differential Equations Conference, vol. 25, pp. 167–178 (2018)

  29. Rothschild, L.P., Stein, E.M.: Hypoelliptic differential operators and nilpotent groups. Acta Math. 137, 247–320 (1976)

    Article  MathSciNet  Google Scholar 

  30. Savin, O.: Regularity of flat level sets in phase transitions. Ann. Math. (2) 169, 41–78 (2009)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

We would like to thank Professor E. Lanconelli for pointing to our attention the paper [11], which has been the starting point for this work, and for many useful conversations and suggestions. We also express our gratitude to the anonymous referees for the careful reading of the paper and for their precious comments, leading to an improved version of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stefano Biagi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: A brief review on Carnot groups

Appendix: A brief review on Carnot groups

In this appendix we review the basic definitions and facts on homogeneous Carnot groups which have been used in the paper. We follow the exposition in [14], to which we refer for all the omitted proofs and the details.

Let \(*\) be a (binary) group operation on Euclidean space \(\mathbb {R}^N\). We say that \(\mathbb {G}= (\mathbb {R}^N,*)\) is a Lie group (on \(\mathbb {R}^N\)) if the map

$$\begin{aligned} *:\mathbb {R}^N\times \mathbb {R}^N\longrightarrow \mathbb {R}^N, \qquad (z,\zeta )\mapsto z*\zeta , \end{aligned}$$

is smooth. For every fixed \(z\in \mathbb {G}\equiv \mathbb {R}^N\) we indicate, respectively, by \(\tau _\alpha \) and \(\rho _\alpha \) the left-translation and the right-translation by z, on \(\mathbb {G}\), that is,

  1. (a)

    \(\tau _z:\mathbb {R}^N\rightarrow \mathbb {R}^N, \qquad \tau _z(\zeta ) := z*\zeta \);

  2. (b)

    \(\rho _z:\mathbb {R}^N\rightarrow \mathbb {R}^N, \qquad \rho _z(\zeta ) := \zeta *z\).

Given a Lie group \(\mathbb {G}= (\mathbb {R}^N,*)\) and a smooth vector field \(X = \sum _j a_j(z)\,\partial _{z_j}\) on \(\mathbb {R}^N\), we say that X is left-invariant on \(\mathbb {G}\) if

$$\begin{aligned} X\big (u\circ \tau _\alpha ) = (Xu)\circ \tau _\alpha , \end{aligned}$$
(A.1)

for every \(u\in C^\infty (\mathbb {R}^N,\mathbb {R})\) and every \(\alpha \in \mathbb {R}^N\). We denote by \(\mathrm {Lie}(\mathbb {G})\) the set of all left-invariant vector fields on \(\mathbb {G}\) and we call it the Lie algebra of \(\mathbb {G}\).

We then have the following very standard result.

Proposition A.1

Let \(\mathbb {G}= (\mathbb {R}^N,*)\) be a Lie group with neutral element e, and let \(\mathrm {Lie}(\mathbb {G})\) be the Lie algebra of \(\mathbb {G}\). Then the following facts hold.

  1. 1.

    \(\mathrm {Lie}(\mathbb {G})\) is a Lie algebra and \(\dim _{\mathbb {R}}(\mathrm {Lie}(\mathbb {G})) = N\);

  2. 2.

    for every \(i \in \{1,\ldots ,N\}\) there exists precisely one \(J_i\in \mathrm {Lie}(\mathbb {G})\) such that

    $$\begin{aligned} J_iu(e) = \frac{\partial u}{\partial z_i}(e) \qquad \big (\text {for every } u\in C^\infty (\mathbb {R}^N,\mathbb {R})\big ); \end{aligned}$$

    more precisely, we have

    $$\begin{aligned} J_i = \sum _{j = 1}^N a_{ij}(z)\,\frac{\partial }{\partial z_j}, \quad \text {where } \begin{pmatrix} a_{i1}(z) \\ \vdots \\ a_{iN}(z) \end{pmatrix} = \mathcal {J}_{\tau _z}(e)\cdot e_i \end{aligned}$$

    (and \(\mathcal {J}_{\tau _x}\) denotes the Jacobian matrix of \(\tau _z\)).

  3. 3.

    If \(J_1,\ldots ,J_N\) are as above, then \(\mathcal {J} = \{J_1,\ldots ,J_N\}\) is a (linear) basis of \(\mathrm {Lie}(\mathbb {G})\), which is usually called the Jacobian basis of \(\mathrm {Lie}(\mathbb {G})\).

Let now \(\mathbb {G}= (\mathbb {R}^N,*)\) be a Lie group on \(\mathbb {R}^N\), and let us assume that there exists a N-tuple \((\upsilon _1,\ldots , \upsilon _N)\) of positive real numbers such that

$$\begin{aligned} D_\lambda :\mathbb {R}^N\longrightarrow \mathbb {R}^N,\qquad D_\lambda (z) := (\lambda ^{\upsilon _1}z_1,\ldots ,\lambda ^{\upsilon _N}z_N), \end{aligned}$$
(A.2)

is an automorphism of \(\mathbb {G}\) for every fixed \(\lambda > 0\), that is,

$$\begin{aligned} D_\lambda (z*\zeta ) = D_\lambda (z)*D_\lambda (\zeta ), \qquad \text {for every }z,\zeta \in \mathbb {R}^N. \end{aligned}$$

Then, the triple \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) is called a homogeneous (Lie) group (on \(\mathbb {R}^N\)) and \(\{D_\lambda \}_{\lambda > 0}\) is usually referred to as the family of dilations of \(\mathbb {G}\).

Remark A.2

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous Lie group on \(\mathbb {R}^N\), with \(D_\lambda \) as in (A.2). We aim to show that, by eventually performing an ad-hoc change of coordinates in \(\mathbb {R}^N\), it is not restrictive to assume

$$\begin{aligned} 1 = \upsilon _1\le \cdots \le \upsilon _N. \end{aligned}$$
(A.3)

In fact, let \(\alpha \) be the unique permutation of \(\{1,\ldots ,N\}\) such that

  1. (a)

    \(\upsilon _{\alpha (1)}\le \cdots \le \upsilon _{\alpha (N)}\), whence \(\upsilon _{\alpha (1)} = \underline{\upsilon }\);

  2. (b)

    if \(\upsilon _{\alpha (i)} = \upsilon _{\alpha (j)}\) for some ij with \(i < j\), then \(\alpha (i) < \alpha (j)\);

moreover, let \(T:\mathbb {R}^N\rightarrow \mathbb {R}^N\) be the map defined as follows

$$\begin{aligned} T(z) := \big (z_{\alpha (1)},\ldots ,z_{\alpha (N)}\big ) \qquad (z\in \mathbb {R}^N). \end{aligned}$$
(A.4)

Obviously, T is a smooth diffeomorphism of \(\mathbb {R}^N\) and

$$\begin{aligned} T^{-1}(a) = \big (a_{\beta (1)},\ldots ,a_{\beta (N)}\big ),\qquad \text { where }\beta = \alpha ^{-1}; \end{aligned}$$

furthermore, each component of T is \(D_\lambda \)-homogeneous of degree \(\upsilon _{\alpha (i)}\), i.e.,

$$\begin{aligned} T_i\big (D_\lambda (z)\big ) = \lambda ^{\upsilon _{\alpha (i)}}\,T_i(z)\quad \forall \,z\in \mathbb {R}^N\,\,\text {and}\,\,\forall \,\,i = 1,\ldots ,N. \end{aligned}$$
(A.5)

By making use of this map T, we define a new group operation \(\circ \) in \(\mathbb {R}^N\) and a new family of dilations \(\{D'_\lambda \}_{\lambda > 0}\) in the following (standard) way:

$$\begin{aligned}&\qquad a \circ b := T\Big (T^{-1}(a)*T^{-1}(b)\Big ) \qquad (\text {for }a,b\in \mathbb {R}^N);&\end{aligned}$$
(A.6)
$$\begin{aligned}&\qquad \qquad D'_\lambda (a) := T\Big (D_{\lambda ^{1/\underline{\upsilon }}} \big (T^{-1}(a)\big )\Big ) \qquad (\text {for }a\in \mathbb {R}^N\text { and }\lambda > 0).&\end{aligned}$$
(A.7)

Notice that, by combining (A.5) with (A.7), we have

$$\begin{aligned} D'_{\lambda }(a) = (\lambda ^{\upsilon '_1}a_1,\ldots ,\lambda ^{\upsilon '_N}a_N), \quad \text {where }\upsilon '_i = \upsilon _{\alpha (i)}/\underline{\upsilon }; \end{aligned}$$
(A.8)

as a consequence, by the choice of \(\alpha \) in (a) one has

$$\begin{aligned} 1 = \upsilon '_1 \le \upsilon '_2\le \cdots \le \upsilon '_N. \end{aligned}$$
(A.9)

On account of (A.8), and taking into account the fact that \(\mathbb {G}\) is a homogeneous Lie group, it is not difficult to recognize that

$$\begin{aligned} \mathbb {G}' = \mathbb {G}'_\alpha := (\mathbb {R}^N,\circ ,D'_\lambda ), \end{aligned}$$

is a homogeneous Lie group on \(\mathbb {R}^N\) as well; moreover, the map T turns out to be an isomorphism of Lie groups, that is, \(T\in C^\infty (\mathbb {R}^N,\mathbb {R}^N)\) and

$$\begin{aligned} T(z*\zeta ) = T(z)\circ T(\zeta ),\qquad \text {for every }z,\zeta \in \mathbb {R}^N. \end{aligned}$$

We also point out that, by the very definition of \(D'_\lambda \), one has

$$\begin{aligned} T\big (D_\lambda (z)\big ) = D'_{\lambda ^{\overline{\upsilon }}}\big (T(z)\big ), \quad \text {for every }z\in \mathbb {R}^N\hbox { and every }\lambda > 0. \end{aligned}$$

We now take a look at the Lie algebras of \(\mathbb {G}\) and \(\mathbb {G}'\). To this end, we first introduce a notation: given any smooth vector field X on \(\mathbb {R}^N\) (acting in the z variables), we denote by \(T_*(X)\) the vector field on \(\mathbb {R}^N\) (acting in the a variables) defined by

$$\begin{aligned} T_*(X)(u) := X(u\circ T)\circ T^{-1}, \qquad \text {for every }u\in C^\infty (\mathbb {R}^N,\mathbb {R}). \end{aligned}$$

Bearing in mind that T is an isomorphism of Lie groups, we have the following well-known facts (for a proof see, e.g., [7, Appendix C]):

  1. (i)

    \(T_*(X)\in \mathrm {Lie}(\mathbb {G}')\) for every \(X\in \mathrm {Lie}(\mathbb {G})\);

  2. (ii)

    the map \(\mathrm {d}T:\mathrm {Lie}(\mathbb {G})\rightarrow \mathrm {Lie}(\mathbb {G}')\) defined by

    $$\begin{aligned} \mathrm {d}T(X) := T_*(X) \end{aligned}$$
    (A.10)

    is an isomorphism of Lie algebras, that is, \(\mathrm {d}T\) is a (linear) isomorphism of vector spaces and, for every \(X,Y\in \mathrm {Lie}(\mathbb {G})\), one has the identity

    $$\begin{aligned} \mathrm {d}T\big ([X,Y]\big ) = [\mathrm {d}T(X),\mathrm {d}T(Y)]. \end{aligned}$$

In particular, if \(\mathcal {J} = \{J_1,\ldots ,J_N\}\) denotes the Jacobian basis of \(\mathrm {Lie}(\mathbb {G})\), a direct computation shows that (for every \(i = 1,\ldots ,N\))

$$\begin{aligned} \mathrm {d}T(J_i)(u)(0) = \frac{\partial u}{\partial a_{j}}(0),\qquad \text {with }j = \beta (i); \end{aligned}$$

as a consequence, from Proposition A.1-(2) we infer that

$$\begin{aligned} \mathrm {d}T(J_i) = J'_{\beta (i)}, \qquad \text {for every }i = 1,\ldots ,N, \end{aligned}$$
(A.11)

where \(\mathcal {J}' = \{J'_1,\ldots ,J'_N\}\) denotes the Jacobian basis of \(\mathrm {Lie}(\mathbb {G}')\). Summing up, the groups \(\mathbb {G}\) and \(\mathbb {G}'\) are isomorphic (via the map T) and their Lie algebras are isomorphic via \(\mathrm {d}T\); furthermore, the matrix representing \(\mathrm {d}T\) with respect to the Jacobian basis \(\mathcal {J}\) and \(\mathcal {J}'\) is the permutation matrix

$$\begin{aligned} P_\alpha = \begin{pmatrix} e_{\alpha (1)}\\ \vdots \\ e_{\alpha (N)} \end{pmatrix}. \end{aligned}$$

In view of all the computations just performed, it seems very natural to identify the group \(\mathbb {G}\) with \(\mathbb {G}'\); however, we prefer to avoid this identification since the homogeneous groups naturally associated with the homogeneous Hörmander operators considered in Sect. 2 do not satisfy (A.3).

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous group on \(\mathbb {R}^N\) [with \(D_\lambda \) as in (A.2)] and let \(\mathcal {J} = \{J_1,\ldots ,J_N\}\) be the Jacobian basis of \(\mathrm {Lie}(\mathbb {G})\). It is very easy to recognize that \(J_i\) is \(D_\lambda \)-homogeneous of degree \(\upsilon _i\) (for every fixed \(i = 1,\ldots ,N\)), that is,

$$\begin{aligned} J_i(u\circ D_\lambda ) = \lambda ^{\upsilon _i}\,(J_i u)\circ D_\lambda , \end{aligned}$$
(A.12)

for every \(u\in C^\infty (\mathbb {R}^N,\mathbb {R})\) and every \(\lambda > 0\); we then set

$$\begin{aligned} \underline{\upsilon } := \min _{1\le i\le N}\upsilon _i, \qquad \overline{\upsilon } := \max _{1\le i\le N}\upsilon _i \end{aligned}$$
(A.13)

and we define the horizontal layer \(V_1(\mathbb {G})\) of \(\mathbb {G}\) as follows:

$$\begin{aligned} V_1(\mathbb {G}) := \mathrm {span}\big \{J_i:\,\upsilon _i = \underline{\upsilon }\big \}\subseteq \mathrm {Lie}(\mathbb {G}). \end{aligned}$$
(A.14)

Remark A.3

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous group and let \(V_1(\mathbb {G})\) be the horizontal layer of \(\mathbb {G}\). Since \(V_1(\mathbb {G})\) is spanned by vector fields which are \(D_\lambda \)-homogeneous of degree \(\underline{\upsilon }\), it is straightforward to recognize that

$$\begin{aligned} X\in V_1(\mathbb {G})\,\,\Longrightarrow \,\, X\text { is }D_\lambda \text {-homogeneous of degree }\underline{\upsilon }. \end{aligned}$$

If \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) is a homogeneous Lie group on \(\mathbb {R}^N\) and if

$$\begin{aligned} \mathrm {Lie}\big (V_1(\mathbb {G})\big ) = \mathrm {Lie}(\mathbb {G}), \end{aligned}$$

we say that \(\mathbb {G}\) is a (homogeneous) Carnot group (here, by \(\mathrm {Lie}(V_1(\mathbb {G}))\) we mean the smallest Lie sub-algebra of \(\mathrm {Lie}(\mathbb {G})\) containing \(V_1(\mathbb {G})\)).

Remark A.4

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous Lie group on \(\mathbb {R}^n\) and let \(\mathbb {G}' = (\mathbb {R}^N,\circ ,D'_\lambda )\) be the group (isomorphic to \(\mathbb {G}\)) constructed in the previous Remark A.2.

Bearing in mind that the Lie algebras of \(\mathbb {G}\) and of \(\mathbb {G}'\) are isomorphic via the map \(\mathrm {d}T\) defined in (A.10), one has

$$\begin{aligned} \mathrm {d}T\big (V_1(\mathbb {G})\big )&\,\,=\,\, \mathrm {span}\big \{\mathrm {d}T(J_i):\,\upsilon _i = \underline{\upsilon }\big \} {\mathop {=}\limits ^{(A.11)}} \mathrm {span}\big \{J'_{\beta (i)}:\,\upsilon _i = \underline{\upsilon }\big \}\\&\,\,=\,\, \mathrm {span}\big \{J'_p:\,\upsilon '_p = 1\big \} \qquad \big (\text {see (A.8) and (A.9)}\big )\\&\,\,=\,\,V_1(\mathbb {G}') \qquad \big (\text {since }\min _{j}\upsilon '_j = 1,\text { see (A.9)}\big ) \end{aligned}$$

(here, as usual, \(\mathcal {J} = \{J_1,\ldots ,J_N\}\) and \(\mathcal {J}' = \{J'_1,\ldots ,J'_N\}\) denote the Jacobian basis of \(\mathrm {Lie}(\mathbb {G})\) and \(\mathrm {Lie}(\mathbb {G}')\), respectively). As a consequence, one gets

$$\begin{aligned} \mathrm {d}T\Big (\mathrm {Lie}\big (V_1(\mathbb {G})\big )\Big ) = \mathrm {Lie}\Big (\mathrm {d}T\big (V_1(\mathbb {G})\big )\Big ) = \mathrm {Lie}\big (V_1(\mathbb {G}')\big ), \end{aligned}$$

and thus \(\mathbb {G}\) is a Carnot group if and only if \(\mathbb {G}'\) is.

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous Carnot group [with \(D_\lambda \) as in (A.2)] and let \(V_1(\mathbb {G})\) be the horizontal layer of \(\mathbb {G}\). The (integer) number

$$\begin{aligned} m = m(\mathbb {G}) := \dim _{\mathbb {R}}\big (V_1(\mathbb {G})\big ) = \mathrm {card}\big \{i\in \{1,\ldots ,N\}:\,\upsilon _i = \underline{\upsilon } \big \}, \end{aligned}$$

is usually referred to as the number of generators of \(\mathbb {G}\). If \(\mathcal {Z} = \{Z_1,\ldots ,Z_m\}\) is any linear basis of \(V_1(\mathbb {G})\), the second-order differential operator

$$\begin{aligned} \Delta _\mathcal {Z} := \sum _{j = 1}^mZ_j^2, \end{aligned}$$

is called a sub-Laplacian on \(\mathbb {G}\). In the particular case when

$$\begin{aligned} \mathcal {Z} = \{J_i:\,\upsilon _i = \underline{\upsilon }\} \end{aligned}$$

(where \(\{J_1,\ldots ,J_N\}\) is the Jacobian basis of \(\mathrm {Lie}(\mathbb {G})\)), we write \(\Delta _\mathbb {G}\) in place of \(\Delta _\mathcal {Z}\) and we call \(\Delta _\mathbb {G}\) the canonical sub-Laplacian on \(\mathbb {G}\).

We conclude this section by stating a theorem by Birindelli and Lanconelli which proves the Gibbons conjecture on any Carnot group. To this end, we first introduce another useful notation which shall be adopted throughout the sequel.

Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous Carnot group, with \(D_\lambda \) as in (A.2); given any \(i\in \{1,\ldots ,N\}\), we say that \(z_i\) is a \(\mathbb {G}\)-horizontal variable if

$$\begin{aligned} \upsilon _i = \underline{\upsilon } = \min _{1\le i\le N}\upsilon _i. \end{aligned}$$

Theorem A.5

([11, Theorems 1.3 and 1.4]) Let \(\mathbb {G}= (\mathbb {R}^N,*,D_\lambda )\) be a homogeneous Carnot group with \(m = m_\mathbb {G}\ge 1\) generators and let \(\mathcal {Z} = \{Z_1,\ldots ,Z_m\}\) be a fixed linear basis of the first layer \(V_1(\mathbb {G})\) of \(\mathbb {G}\).

Moreover, let \(f:[-1,1]\rightarrow \mathbb {R}\) be a Lipschitz-continuous function satisfying the following two properties:

  1. (f1)

    \(f(-1) = f(1) = 0\);

  2. (f2)

    there exists a real \(\varepsilon < 1\) such that

    $$\begin{aligned} f\text { is non-increasing on }[-1,-1+\varepsilon ] \text { and on }[1-\varepsilon ,1]. \end{aligned}$$

Finally, let \(v\in C^2(\mathbb {R}^N,\mathbb {R})\) be a solution of

$$\begin{aligned} \Delta _\mathcal {Z}v + f(v) = \sum _{j = 1}^mZ_j^2v + f(v) = 0\qquad \text {on }\mathbb {G}\equiv \mathbb {R}^N, \end{aligned}$$
(A.15)

which satisfies the next two conditions:

  1. (a)

    \(|v(x)|\le 1\) for every \(x\in \mathbb {R}^N\);

  2. (b)

    there exists \(i\in \{1,\ldots ,N\}\) such that

    $$\begin{aligned} \lim _{z_i\rightarrow \pm \infty }v(z) = \pm 1, \end{aligned}$$

    uniformly for \(\hat{z} = (z_1,\ldots ,z_{i-1},z_{i+1},\ldots ,z_N)\in \mathbb {R}^{N-1}\).

Then the following facts hold true.

  1. (1)

    If \(z_i\) is a \(\mathbb {G}\)-horizontal variable, then there exists \(V\in C^2(\mathbb {R},\mathbb {R})\) such that

    $$\begin{aligned} v(z) = V(z_i)\,\,\text {for every }z\in \mathbb {R}^N\quad \text {and}\quad V' > 0\,\,\text {on }\mathbb {R}. \end{aligned}$$
    (A.16)
  2. (2)

    If \(z_i\) is not a \(\mathbb {G}\)-horizontal variable, then one can say that

    $$\begin{aligned} \frac{\partial v}{\partial z_i}(z)\ge 0\quad \text {for every }z\in \mathbb {R}^N, \end{aligned}$$
    (A.17)

    and the inequality is strict if \(\partial _{z_i}\) commutes with \(Z_1,\ldots ,Z_m\).

As a matter of fact, Theorem A.5 is proved in [11] under the additional assumption that the exponents of \(D_\lambda \) fulfill (A.3); for the sake of completeness, we show here how one can pass from a general Carnot group (whose family of dilations does not fulfill (A.3)) to the case considered in [11].

Proof

Let \(\mathbb {G},\,f,\,v\) be as in the statement of the theorem and let

$$\begin{aligned} \mathbb {G}' = (\mathbb {R}^N,\circ ,D'_\lambda ) \end{aligned}$$

be the homogeneous group (isomorphic to \(\mathbb {G}\)) constructed in Remark A.2.

Since \(\mathbb {G}\) is a Carnot group, we know from Remark A.4 that the same is true of \(\mathbb {G}'\); moreover, since \(\mathcal {Z}\) is a basis of \(V_1(\mathbb {G})\) and since the Lie algebras of \(\mathbb {G}\) and of \(\mathbb {G}'\) are isomorphic via the map \(\mathrm {d}T\) in (A.10), the set

$$\begin{aligned} \mathcal {Z}' := \{Y_1,\ldots ,Y_N\}, \qquad \text {where } Y_i := \mathrm {d}T(Z_i) \end{aligned}$$

is a basis of \(V_1(\mathbb {G}') = \mathrm {d}T(V_1(\mathbb {G}))\). As a consequence, the differential operator

$$\begin{aligned} \Delta _{\mathcal {Z}'} := \sum _{j = 1}^N Y_j^2, \end{aligned}$$

is a sub-Laplacian on \(\mathbb {G}'\). Now, if T is the map defined in (A.4) (which is an isomorphism between \(\mathbb {G}\) and \(\mathbb {G}'\)), we consider the following function

$$\begin{aligned} w:\mathbb {R}^N\longrightarrow \mathbb {R}^N, \qquad w(a) := v\big (T^{-1}(a)\big ). \end{aligned}$$

Obviously, \(w\in C^2(\mathbb {R}^N,\mathbb {R})\); moreover, from the very definition of \(\mathrm {d}T\) (and since v solves (A.15) and satisfies assumptions (a)-(b)), we easily infer that:

  • \(\Delta _{\mathcal {Z}'}w+f(w) = 0\) on \(\mathbb {G}'\equiv \mathbb {R}^N\);

  • \(|w(a)|\le 1\) for every \(a\in \mathbb {R}^N\);

  • setting \(j := \beta (i)\) (where \(\beta = \alpha ^{-1}\), see Remark A.2), we have

    $$\begin{aligned} \lim _{a_j\rightarrow \pm \infty }w(a) = \pm 1, \end{aligned}$$

    uniformly for \(\hat{a} = (a_1,\ldots ,a_{j-1},a_{j+1},\ldots ,a_N)\in \mathbb {R}^{N-1}\).

Since Theorem A.5 holds in the group \(\mathbb {G}'\) [see [11] and remember that the exponents \(\upsilon '_1,\ldots ,\upsilon '_N\) of \(D'_\lambda \) satisfy (A.9)], we have the following two cases:

  1. (1)

    \(z_i\) is a \(\mathbb {G}\)-horizontal variable. In this case, since \(\upsilon _i = \underline{\upsilon }\) we have

    $$\begin{aligned} \upsilon '_{j}\,\,=\,\,\upsilon '_{\beta (i)} {\mathop {=}\limits ^{(A.8)}} 1, \end{aligned}$$
    (A.18)

    and thus \(a_j = a_{\beta (i)}\) is a \(\mathbb {G}'\)-horizontal variable. By Theorem A.5 (applied to the group \(\mathbb {G}'\)) we then infer the existence of \(V\in C^2(\mathbb {R},\mathbb {R})\) such that

    1. (i)

      \(V' > 0\) throughout \(\mathbb {R}\);

    2. (ii)

      \(v\big (a_{\beta (1)},\ldots ,a_{\beta (N)}\big ) = w(a) = V(a_j) = V(a_{\beta (i)})\) for every \(a\in \mathbb {R}^N\);

    as a consequence, \(v(z) = V(z_i)\) for every \(z\in \mathbb {R}^N\), as desired.

  2. (2)

    \(z_i\) is not a \(\mathbb {G}\)-horizontal variable. In this case, \(a_j = a_{\beta (i)}\) is not a \(\mathbb {G}'\)-horizontal variable [see (A.18)] and thus, by applying Theorem A.5 to the group \(\mathbb {G}'\), we obtain

    $$\begin{aligned} \frac{\partial w}{\partial a_j}(a) = \frac{\partial v}{\partial z_i}\big (T^{-1}(a)\big ) \ge 0, \quad \text {for every }a\in \mathbb {R}^N; \end{aligned}$$

    since T is a diffeomorphism, this proves that \(\partial _{z_i}v\ge 0\) on \(\mathbb {R}^N\). Finally, if \(\partial _{z_i}\) commutes with \(Z_1,\ldots ,Z_N\), a direct computation shows that

    $$\begin{aligned}{}[\partial _{a_j},Y_k] = [\partial _{a_j}, \mathrm {d}T(Z_k)] = 0, \quad \text {for all }k = 1,\ldots ,N, \end{aligned}$$

    and thus, again by Theorem A.5, we get \(\partial _{a_j}w = \partial _{z_i}u\circ T^{-1} > 0\) on \(\mathbb {R}^N\).

This ends the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Biagi, S. An application of a global lifting method for homogeneous Hörmander vector fields to the Gibbons conjecture. Nonlinear Differ. Equ. Appl. 26, 49 (2019). https://doi.org/10.1007/s00030-019-0594-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00030-019-0594-2

Keywords

Mathematics Subject Classification

Navigation