Notions of Convexity in Carnot Groups
Notions of Convexity in Carnot Groups
Notions of Convexity in Carnot Groups
CONTENTS
1. Introduction
2. Preliminaries
3. Strongly convex functions: A notion to abandon
4. Mildly convex functions: The appearance of horizontal planes
5. Weakly convex functions: The horizontal Hessian
6. Weak convexity of the gauge in groups of Heisenberg type
7. Convexity of sets
8. Further properties of second derivatives of weakly convex functions
9. Basic properties of first derivatives of weakly convex functions: L∞ − L1
estimates
10. The horizontal Monge-Ampère equation
11. Yet another notion of convexity
References
1. Introduction.
263
264 D. Danielli, N. Garofalo, and D.-M. Nhieu
r
for any ξ = ξ1 + ξ2 + ... + ξr ∈ g. Let Q = j dim Vj be the homogeneous
j=1
dimension of G associated with such dilations.
Conjecture: Given a connected, bounded open set Ω ⊂ G, let f ∈ LQ (Ω).
Suppose that u ∈ L2,Q
loc (Ω) ∩ C(Ω) satisfy
def 1
m
Lu = aij XiXj u + Xj Xi u ≥ F
2
i,j=1
Here, L2,Q Q
loc (Ω) indicates the Sobolev space of functions u ∈ Lloc (Ω) having
weak derivatives Xi Xj u ∈ LQloc (Ω). We note explicitly that
2
Lu = tr A X u ,
where we have denoted by X 2 u the symmetrized horizontal Hessian of u,
see (1.3), or Section 5. Concerning the optimality of the LQ norm in the
estimate (1.1), we refer the reader to [DGN4]. In the abelian case, when
G = Rn with the standard 2 homogeneous dilations, one has g = V1 = Rn ,
2
so that Q = n, and X u = D u, the standard Hessian matrix of u.
The above conjecture, in this situation, is in fact the celebrated geometric
maximum principle of Alexandrov-Bakelman-Pucci (except that the matrix
(aij ) need not be uniformly elliptic), see [A], [Ba1], [Pu], and also [Ba3].
Given the pervasive role that convexity plays in such principle, in order
to advance our comprehension of the above conjecture it seems natural to
examine appropriate notions of convexity in the non-abelian setting. This is
the goal of the present paper.
We next give a description of our results. The notion of weakly H-convex
function set forth in Definition 5.5 is central to our work. Sections 3 and 4
are devoted to justifying how we arrived to such notion starting from that of
Pansu differential. A function on a Carnot group u : G → R is called weakly
H-convex if for any g ∈ G and every λ ∈ [0, 1] one has
(1.2) u g δλ (g −1 g ) ≤ (1 − λ) u(g) + λ u(g ) for every g ∈ Hg .
represents the twisted convex combination of g and g , with base point g. The
notation Hg indicates the horizontal plane passing through g, see Definition
4.1. The collection of all such planes determines the horizontal subbundle
generated by a given orthonormal basis of the Lie algebra g. The trajectories
of the dilations gδλ (g −1 g ) are in general quite twisted and they badly fail
to be sub-unitary. They possess however the important redeeming feature
that when the end-point g belongs to the horizontal plane through g, then
such trajectories are just straight line segments, at least at the level of the
Lie algebra, see Proposition 4.4.
This leads us to our first main result, Theorem 5.12, which states that a
function u in the Folland-Stein class Γ2 (G) is weakly H-convex if and only
if its symmetrized horizontal Hessian
1
(1.3) X 2 u(g) = X 2 u(g) + X 2 u(g)T ,
2
is positive semi-definite at every g ∈ G. Here,
and
C 1
(1.6) ess sup |Xu| ≤ |u| dg .
B(g,r) r |B(g, 15r)| B(g,15r)
The main trust of the inequalities (1.5), (1.6) is that their right-hand
sides involve a set of full measure. The way we accomplish this result is to
Notions of convexity in Carnot groups 267
2. Preliminaries.
N = m1 + ... + mr
Since Carnot groups of step 2 will play a special role it will be convenient
to have a separate notation for the first two layers V1 and V2 . For simplicity,
we set m = m1 , k = m2 , and indicate
where (Lg )∗ denotes the differential of Lg . The system X defines a basis for
the so-called horizontal sub-bundle HG of the tangent bundle T G. For a
given function f : G → R, the action of Xj on f is specified by the equation
By the assumption on the Lie algebra one immediately sees that the
system X satisfies the finite rank condition in Hörmander’s theorem [H1],
therefore the operator L is hypoelliptic. However, when r > 1 it fails to be
elliptic, and the loss of regularity is measured by the step r of the stratifica-
tion of g.
For a function u on G we let
Xu = X1 u X1 + ... + Xm u Xm ,
m
and set |Xu| = ( j=1 (Xj u)2 )1/2 . We denote by dg the push-forward of
Lebesgue measure on g via the exponential map. Such dg defines a bi-
invariant Haar measure on G. One has d(g ◦ δλ) = λQdg, so that the number
Q = m1 + 2 m2 + ... + r mr
plays the role of a dimension with respect to the group dilations. For this
reason Q is called the homogeneous dimension of G. Such number is larger
than the topological dimension N of G defined above.
The Euclidean distance to the origin | · | on g induces a homogeneous
pseudo-norm | · |g on g and (via the exponential map) one on the group G in
the following way (see [F2], [FS]). For ξ ∈ g, with ξ = ξ1 + ... + ξr , ξi ∈ Vi ,
we let
2r!
r
2r!/i
(2.10) |ξ|g = |ξi| ,
i=1
where ω = |B(e, 1)|, with e indicating the group identity. A similar transla-
tion invariance holds for the metric balls Bρ (g, R).
The Heisenberg group. A model of Carnot group of paramount impor-
tance is the Heisenberg group Hn , see [S], [G]. The underlying manifold of
this Lie group is simply R2n+1 , with the non-commutative group law
(2.15)
g g = (x, y, t) (x , y , t ) = (x + x , y + y , t + t + 2(< x , y > − < x, y >)) ,
where we have let x, x , y, y ∈ Rn , t, t ∈ R. Let (Lg )∗ be the differential of
the left-translation (2.15). A simple computation shows that
∂ def ∂ ∂
(2.16) (Lg )∗ = Xj = + 2 yj , j = 1, ..., n ,
∂xj ∂xj ∂t
∂ def ∂ ∂
(Lg )∗ = Xn+j = − 2 xj , j = 1, ..., n ,
∂yj ∂yj ∂t
∂ def ∂
(Lg )∗ = X2n+1 =
∂t ∂t
We note that the only non-trivial commutator is
[Xj , Xn+k ] = − 4 X2n+1 δjk , j, k = 1, ..., n ,
therefore the vector fields {X1 , ..., X2n} constitute a basis of the Lie algebra
hn = R2n+1 = V1 ⊕ V2 , where V1 = R2n × {0}t, V2 = {0}(x,y) × R. The
non-isotropic group dilations are therefore
(2.17) δλ (g) = (λx, λy, λ2t) .
The associated homogeneous dimension is Q = 2n + 2, and the gauge is
given by
1/4
N (g) = (|x|2 + |y|2 )2 + t2 .
We notice that the sub-Laplacian associated with the basis {X1 , ..., X2n}
is
2n
n
∂2 ∂ ∂ ∂
(2.18) L = Xj2 2 2
= ∆x,y +4 (|x| +|y| ) 2 +4 yj −xj ,
∂t ∂t ∂xj ∂yj
j=1 j=1
which coincides with the real part of the complex Kohn Laplacian, see [S].
Returning to general Carnot groups, we recall that for a given open set
Ω ⊂ G, the classes Γ1 (Ω) (respectively, Γ2 (Ω)) introduced by Folland and
Stein [FS], [F2], represent the collection of all functions u : Ω → R such that
the derivatives Xα u (respectively, XαXα u), α, α = 1, ..., m, exist and are
continuous functions in Ω. We also need the definition of Lipschitz function.
Notions of convexity in Carnot groups 273
Theorem 2.2. Let G, G be two Carnot groups, and consider open sets
Ω ⊂ G, Ω ⊂ G . If u : Ω → Ω is a Lipschitz map with respect to the
relative intrinsic metrics, i.e., there exists a constant L > 0 such that
We note explicitly that when the target group is G = (R, +), then the
definition (2.20) of Pansu differential of a function u : G → R becomes
m
Du(g)(h) = xj (h) Xj u(g) = < ξ1 (h), Xu(g) >, g, h ∈ G.
j=1
When u ∈ C 1 (G) the proof of the above result can be found in [DGS],
or also in [MSC]. The passage from C 1 (G) to Γ1 (G) is easily justified by
mollification with group convolution, see equation (3.5) for the relevant def-
inition.
We will need the following quantitative version of Theorem 2.2.
Theorem 2.4. Let Ω ⊂ G be a bounded open set. If for u ∈ Γ0,1 (Ω) and
(2.19) holds, then the distributional derivatives of u along the vector fields
X1 , ..., Xm belong to L∞ (Ω), and moreover there exists C = C(G) > 0 such
that
||Xαu||L∞ (Ω) ≤ C L α = 1, ..., m .
Theorem 2.4 is just a special case of a result which in [GN2] was proved
for a general family of Lipschitz vector fields in Rn , see also [FSS], where
the same result was independently obtained. We will also need the following
converse to Theorem 2.4 which is also a special case of a general result in
[GN2]. Given a measurable set E ⊂ G, we say that a function u ∈ Lp (E)
belongs to L1,p (E), 1 ≤ p ≤ ∞, if the distributional derivatives Xαu ∈
Lp (E), for α = 1, ...m. The space L1,p (E) becomes a Banach space with the
norm
m
||u||L1,p(E) = ||u||Lp(E) + ||Xαu||Lp(E) .
α=1
Notions of convexity in Carnot groups 275
Theorem 2.5. Given a Carnot group G, let u ∈ L1,∞ (B(go , 3R)), with
go ∈ G and R > 0, then u can be modified on a set of dg-measure zero in
B̄ = B̄(go, R), so as to satisfy for every g, g ∈ B̄(go , R)
|u(g) − u(g )| ≤ C d(g, g ) u L1,∞ (B(go,3R)) ,
for some C = C(G) > 0. If, furthermore, u ∈ C ∞ (B(go , 3R)), then in
the right hand side of the previous inequality one can replace the term
||u||L1,∞ (B(go,3R)) with ||Xu||L∞(B(go,3R)).
Combining Theorems 2.5 and 2.4 we obtain a characterization of the
Lipschitz class Γ0,1 1,∞
loc (G) as the space Lloc (G).
The notion of Pansu differential expressed by equation (2.21), and its connec-
tion with the horizontal gradient established by Proposition 2.3, suggests a
first notion of convexity in a Carnot group, which we call strong H-convexity.
At the beginning of Section 4 we will see that such concept is too restrictive
for the development of a satisfactory theory, and has to be modified. An-
other basic reason to renounce the notion of strong H-convexity lies in the
fact that the gauge balls in the Heisenberg group are not strongly H-convex,
see Remark 7.5 in Section 7. But...we are getting ahead of our story. It is
time to introduce a basic definition.
Definition
3.6. A function A : G → R is called H-affine if there exists
p= m α=1 α Xα ∈ V1 , and β ∈ R such that
p
m
(3.3) A(g) = pj xj (g) + β = < p, ξ1 (g) > + β .
j=1
Notions of convexity in Carnot groups 277
therefore
Proof. For 0 ≤ λ ≤ 1 one has from (3.5) and from Definition 3.2
P (x, y, t) = α x2 + 2 β xy + γ y 2 + δ t ,
αβ
where the matrix is assumed ≥ 0. Since, thanks to Proposition 3.8 we
βγ
know that Q(x, y, t) = αx2 +2βxy +γy 2 is strongly H-convex, it is natural to
ask whether such is also the function u(x, y, t) = t. This question is further
motivated by the fact that such “quadratic function” is also L-harmonic on
Hn , i.e., it is an entire solution of the Kohn sub-Laplacian (2.18). To verify
the strong H-convexity of u, let g = (x, y, t), g = (x , y , t ) be two points in
H1 , and consider the point defined by (3.1)
(4.1) gλ = x + λ(x − x) , y + λ(y − y) ,
t + 2λ(x y − xy ) + λ2 (t − t + 2xy − 2x y) .
so that gg ∈ He only when the two vectors (x , y ), (y, −x) ∈ R2n are orthog-
onal. However, it is true that g ∈ He if and only if g −1 ∈ He , and δλ g ∈ He
for λ > 0. From Definition 4.1 we immediately obtain the following elemen-
tary, yet important, result.
Proof. Only the second and third parts require a proof. If g ∈ Hg , then
g −1 g ∈ He . If ξ(g) = ξ1 (g) + ... + ξr (g) denote the exponential coordinates
of g, we see that the condition g −1 g ∈ He translates into
(4.3) ξ2 (g −1 g ) = 0 , ... , ξr (g −1 g ) = 0 .
These are the general equations of the horizontal plane Hg . Since by
(2.7) we have ξ(g −1 ) = −ξ(g), we conclude that g −1 g ∈ He if and only if
(g )−1 g ∈ He , which is equivalent to g ∈ Hg .
For the third part we observe that for every λ ∈ [0, 1] one has gλ ∈ Hg
if and only if g −1 gλ = δλ (g −1 g ) ∈ He . Since for any point h ∈ G we have
δλ h ∈ He if and only if h ∈ He , we conclude that gλ ∈ Hg is equivalent to
g −1 g ∈ He , i.e., g ∈ Hg . This completes the proof.
Proposition 4.3. Let G be a Carnot group of step 2, then for any given
go ∈ G the horizontal plane passing through go is the collection of all points
g ∈ G whose exponential coordinates (x(g), y(g)) verify the k linear equa-
tions
1 β
m
yβ (g) = yβ (go ) + bαα xα (go) xα (g) , β = 1, ..., k ,
2 α,α =1
Proof. We only present the proof for groups of step 2, the details are
similar but more (tediously) complicated in the case of higher step. We have
already observed in Proposition 4.2 that g ∈ Hg if and only if gλ ∈ Hg for
every 0 ≤ λ ≤ 1. Proposition 4.3 gives
1
m
yβ (gλ ) = yβ (g) + bβαα xα (g) xα (gλ ) , β = 1, ..., k ,
2
α,α=1
Since
bβαα = < [Xα, Xα ], Yβ > = < J(Yβ )Xα, Xα > ,
we easily conclude that
m
(4.9) bβαα xα(g) xα (g)
α,α=1
m
= < ξ1 (g), Xα >< ξ1 (g), Xα >< J(Yβ )Xα, Xα >
α,α=1
m
= < J(Yβ )Xα, ξ1 (g) >< ξ1 (g), Xα >
α=1
m
= < J(Yβ )T ξ1 (g), Xα >< ξ1 (g), Xα >
α=1
m
= − < J(Yβ )ξ1 (g), Xα >< ξ1 (g), Xα >
α=1
= − < J(Yβ )ξ1 (g), ξ1(g) > = 0 ,
where in the latter equality we have used (4.5). Substitution of (4.9) in (4.8)
gives
λ
m
(4.10) yβ (gλ ) = yβ (g) + bβαα xα (g) xα (g ) .
2
α,α=1
Notions of convexity in Carnot groups 283
1
m
bβαα xα (g) xα (g ) = yβ (g ) − yβ (g) ,
2
α,α=1
ξ2 (gλ ) = (1 − λ) ξ2 (g) + λ ξ2 (g ) .
Since ξ(g) = ξ1 (g) + ξ2 (g), the latter equation and (4.7) imply the con-
clusion.
We denote by CH
m
(Hn ) the collection of all mildly H-convex functions on Hn .
284 D. Danielli, N. Garofalo, and D.-M. Nhieu
now reads
2 λ (1 − λ) |x y − xy | ≤ 0 ,
which is clearly impossible, except in the trivial cases λ = 0, 1. We observe,
however, that if we impose to g = (x , y , 0) to be on the horizontal plane
through g = (x, y, 0), then |x y − xy | = 0, and therefore the inequality is
satisfied for all λ ∈ [0, 1]. This negative example brings us to the subject of
the next section.
(5.2) φ(0) = u(g) , φ(1) = u(g ) , φ (0) = < Xu(g), x(g) − x(g) > .
(5.5) φ (λ) = X12 u(gλ )(x − x)2 + X1 X2 u(gλ )(x − x)(y − y)
+ X2 X1 u(gλ )(x − x)(y − y) + X22 u(gλ )(y − y)2
+ 2(t − t + 2(xy − x y)) ut (gλ )
+ 2λ X1 (ut)(gλ )(x − x) + X2 (ut)(gλ )(y − y)
2
+ 2λ (t − t + 2(xy − x y)) utt(gλ ) .
(5.6) φ (λ) = X12 u(gλ )(x − x)2 + X1 X2 u(gλ )(x − x)(y − y)
+ X2 X1 u(gλ )(x − x)(y − y) + X22 u(gλ )(y − y)2
= < (X 2 u(gλ )) ζ , ζ > ,
where we have let ζ T = (x −x, y −y). We have thus established the following
result.
Proposition 5.2 continues to be valid for any Carnot group. We will need
the following useful lemma from [GV1].
1
k
∂ ∂
(5.8) Xα = + < [ξ1 , Xα], Yβ >
∂xα 2 ∂yβ
β=1
1 β
k m
∂ ∂
= + bαα xα (g) .
∂xα 2
∂yβ
β=1 α =1
1
k
∂2 ∂
(5.9) Xα X α = + < [Xα, Xα ], Yβ >
∂xα∂xα 2 ∂yβ
β=1
1
k
∂2
+ < [ξ1 , Xα ], Yβ >
2 ∂xα∂yβ
β=1
1
k
∂2
+ < [ξ1 , Xα], Yβ >
2 ∂xα ∂yβ
β=1
1
k
∂2
+ < [ξ1 , Xα], Yβ >< [ξ1 , Xα ], Yβ > .
4 ∂yβ ∂yβ
β,β =1
k
∂
(5.10) [Xα , Xα ] u = bβα α .
∂yβ
β=1
Notions of convexity in Carnot groups 287
Proof. Equation (5.8) is formula (3.3) in [GV1]. The formula (5.9) follows
from (5.8), once we observe that
∂
< [ξ1 , Xα], Yβ > = 0 ,
∂yβ
and that
∂ ∂
m
< [ξ1 , Xα ], Yβ >= xγ < [Xγ , Xα ], Yβ >=< [Xα, Xα ], Yβ > .
∂xα γ=1
∂xα
In the case of arbitrary step one has to replace formulas (5.8), (5.9) with
more complicated ones. Let ms = dim Vs , s = 1, ..., r, and indicate with
X1 , ..., Xm1 an orthonormal basis of V1 . We denote by (x1 (g), ..., xr(g)) the
exponential coordinates of g = exp (ξ1 (g) + ... + ξr (g)), where for each s =
1, ..., r we have presently let xs (g) = (xs,1 (g), ..., xs,ms (g)) for the components
of ξs (g) with respect to a fixed orthonormal basis of Vs . The following
formula holds
(5.11)
∂ r ms
∂
Xα = + bβs,α (x1 (g), ..., xs−1(g)) , α = 1, ..., m1 ,
∂x1,α ∂xs,β
s=2 β=1
Proposition 5.4. Let G be a Carnot group and let u ∈ Γ2 (G). For every
fixed g ∈ G, and for any g ∈ Hg one has
m
(5.12) φ (λ) = Xαu(gλ ) (xα(g ) − xα (g)) ,
α=1
288 D. Danielli, N. Garofalo, and D.-M. Nhieu
where ζ = ξ1 (g ) − ξ1 (g) ∈ V1 .
λ
m
φ(λ) = u xα(g) + λ(xα(g ) − xα(g)) , yβ (g) + bβαα xα (g) xα (g ) ,
2
α,α =1
m
∂u 1 β
k m
∂u
φ (λ) = (gλ )(xα(g ) − xα (g)) + bαα xα (g)xα (g ) (gλ )
∂xα 2
∂yβ
α=1 β=1 α,α =1
(5.14)
m
φ (λ) = Xαu(gλ ) (xα (g ) − xα(g))
α=1
1
k m
∂u
+ bβαα xα(g) xα (g ) (gλ )
2 ∂yβ
β=1 α,α=1
1 β
k m
∂u
− bα α xα (g) + λ(xα (g ) − xα (g)) (xα (g ) − xα(g)) (gλ )
2
∂yβ
β=1 α,α =1
Notions of convexity in Carnot groups 289
1 β
k m
∂u
=− bαα xα (g)xα(g ) (gλ )
2
∂yβ
β=1 α,α =1
λ β
k m
∂u
+ bα α xα (g )xα(g) (gλ )
2
∂yβ
β=1 α,α =1
λ β
k m
∂u
+ bαα xα (g)xα(g ) (gλ )
2
∂yβ
β=1 α,α =1
1 β
k m
∂u
= − bαα xα(g)xα (g ) (gλ ) ,
2
∂yβ
β=1 α,α =1
where we have used the equality bβαα = −bβα α . Substituting (5.15) in (5.14)
we conclude that (5.12) holds.
From (5.12) we now easily obtain
m
φ (λ) = Xα Xα u(gλ ) xα(g ) − xα (g) xα (g ) − xα (g) ,
α,α =1
Proposition (5.4) paves the way to the following definition, which plays
a central role in this paper.
for every g ∈ Hg . When the strict inequality holds we say that u is strictly
weakly H-convex. The effective domain of u is the set
We denote by CH
w
(G) the class of all weakly H-convex functions on G.
290 D. Danielli, N. Garofalo, and D.-M. Nhieu
Remark 5.6. It is worth noting at this point that Definition 5.5 involves
a one-sided information on u on lower-dimensional manifolds, namely the
horizontal planes. This new aspect, which is of course not present in the
classical theory of convex functions, is a source of difficulties when one wants
to pass from such information to conclusions on a set of full measure.
u(gλ ) − u(g)
≤ u(g ) − u(g) .
λ
We now pass to the limit for λ → 0 obtaining the conclusion from Propo-
sition 2.3.
(5.17) < DUg (ξ1 ), ξ1 − ξ1 > + Ug (ξ1 ) ≤ Ug (ξ1 ) for every ξ1 ∈ V1 .
Notions of convexity in Carnot groups 291
(5.18) CH
s
CH
m
CH
w
.
Remark 5.9. It is important to notice that Propositions 3.4, 3.5 and 3.9
continue to be valid if in their statements we replace “strongly H-convex”
with “weakly H-convex”.
The following two theorems provide the main motivation for Definition
5.5 : the weak H-convexity of a function in the class Γ2 is characterized by
the positivity of its horizontal Hessian. To facilitate the presentation we will
first analyze with full details the Heisenberg group model in Theorem 5.11.
292 D. Danielli, N. Garofalo, and D.-M. Nhieu
Noting that gδλ (δτ ζ̃) → g, as τ → 0, and using the continuity of the
second horizontal derivatives of u, we conclude from (5.21)
The principal curvatures at (xo , yo ) of the graph of Ugo are given by the
equation
κ = H ± H2 − K ,
where H and K respectively denote the mean curvature and the Gaussian
curvature of the graph at (xo , yo ). Since
2
Uxx Uyy − Uxy DU
K = , 2 H = div ,
(1 + Ux2 + Uy2 )2 1 + |DU |2
We stress that when go = (0, 0, to) then the plane Hgo has a characteristic
point at go (i.e., a point where the vector fields X1 and X2 become tangent
to the plane), and we have κ1 = κ2 = 0. We will come back to the equation
det (X 2 u) = F in Section 10.
(6.2) < J(η )ξ, J(η )ξ > = < η , η > |ξ|2 , η , η ∈ V2 , ξ ∈ V1 .
2
Given the sub-Laplacian L = m j=1 Xj associated with an orthonormal
basis of V1 , Kaplan [K] proved that there exists a constant C = C(G) > 0
such that the locally integrable function
C
(6.4) Γ(g, g ) = ,
N (g −1 g )Q−2
satisfies the equation LΓ(g, ·) = −δg in D (G). This result generalized Fol-
land’s fundamental solution for the Heisenberg group Hn [F1]. In [CDG] it
was discovered that, remarkably, the gauge (6.3) enters also in the expres-
sion of the fundamental solution of the following one-parameter family of
non-linear operators
m
(6.5) Lp u = Xj (|Xu|p−2 Xj u) , 1<p<∞.
j=1
296 D. Danielli, N. Garofalo, and D.-M. Nhieu
It was proved there that for every 1 < p < ∞, p = Q, there exists
Cp = C(G, p) > 0 such that the fundamental solution of −Lp is given by
Cp
(6.6) Γp (g, g ) = .
N (g −1 g )(Q−p)/(p−1)
def
m
(6.7) L∞ u = XαXα u Xαu Xα u = < (X 2 u) Xu, Xu > .
α,α=1
def
Lemma 6.2. In a Carnot group G of step 2, consider the function ψ(g) =
|x(g)|2. For any s = 1, ..., k, one has for every g ∈ G
As a consequence,
and
|x(g)|2
(6.11) |XN (g)|2 = .
N (g)2
1
(6.12) |X(yl )(g)|2 = |x(g)|2 , l = 1, ..., k .
4
L∞ N = 0 , in G \ {e} .
Finally, we obtain
1 −3
XN = N X(ψ 2 + 16|y|2 ) = N −3 |x|2 ξ1 + 4X|y|2 ,
4
which gives
where in the last equality we have used (6.10) in Lemma 6.2. Replacing the
latter equation in (6.15) we reach the conclusion.
298 D. Danielli, N. Garofalo, and D.-M. Nhieu
Consider the function u(g) = N (g)4 = |x(g)|4 + 16|y(g)|2. From the re-
sults in Section 3, we easily conclude that g → |x(g)|4 is strongly H−convex,
and therefore it is weakly H-convex, see (5.18). Next, we prove that also the
function g → |y(g)|2 is weakly-convex, and therefore the fourth power of the
gauge is weakly H-convex.
def
Proposition 6.5. In a Carnot group of step 2 the function χ(g) = |y(g)|2
w (G), therefore so does u(g) = N (g)4 = |x(g)|4 + 16|y(g)|2.
belongs to CH
k
Xα Xα χ = < [Xα, Xα ], Yβ > yβ
β=1
1
k
+ < [ξ1 , Xα], Yβ >< [ξ1 , Xα ], Yβ > δββ
2
β,β =1
k
= < [Xα, Xα ], Yβ > yβ
β=1
1
k
+ < [ξ1 , Xα], Yβ >< [ξ1 , Xα ], Yβ > .
2
β=1
(6.16)
1 1
k
Xα Xα χ + Xα Xα χ = < [ξ1 , Xα], Yβ >< [ξ1 , Xα ], Yβ > .
2 2
β=1
1
k
= < J(Yβ )ξ1 (g), ζ >2 ≥ 0 .
2
β=1
for every ζ ∈ V1 .
Proof. Let χ(g) = |y(g)|2, and consider the function ψ(g) = |x(g)|2. We
thus have u(g) = N (g)4 = ψ(g)2 + 16χ(g). We first compute the quadratic
form associated with the symmetrized horizontal Hessian of u. Since
we need to compute < (X 2 ψ 2 (g))ζ, ζ > and < (X 2χ(g)) ζ, ζ >, where
ζ ∈ V1 . The latter of these quantities has been obtained in (6.17), so we are
left with computing the former. From (5.8), (5.9) in Lemma 5.3 we find
(6.24)
k
Xα χ = yβ < [ξ1 , Xα], Yβ >
β=1
(6.25)
k
= < ξ2 , Yβ >< [ξ1 , Xα], Yβ >=< [ξ1 , Xα], ξ2 >=< J(ξ2 )ξ1 , Xα > .
β=1
(6.26)
m
2 1 2 3 α,α =1 Xα uXα uζα ζα
< (X N ) ζ, ζ > =
< (X u) ζ, ζ > − .
4N 3 4 N4
k
1 2 2 2
= |ξ 1 | |ζ| + 2 < ξ 1 , ζ > + 2 < J(Yβ )ξ1 , ζ >2
N3
β=1
3
− |ξ1 |4 < ξ1 , ζ >2 + 8|ξ1 |2 < ξ1 , ζ >< J(ξ2 )ξ1 , ζ >
N4
+ 16 < J(ξ2 )ξ1 , ζ >2
k
1
= |ξ1 |2 |ζ|2 + 2 < ξ1 , ζ >2 + 2 < J(Yβ )ξ1 , ζ >2
N3
β=1
2
3
− |ξ1 |2 < ξ1 , ζ > + 4 < J(ξ2 )ξ1 , ζ > .
N4
that
2
3 2
(6.27) |ξ1 | < ξ1 , ζ > + 4 < J(ξ2 )ξ1 , ζ >
N4
k
≤ |ξ1 |2 |ζ|2 + 2 < ξ1 , ζ >2 + 2 < J(Yβ )ξ1 , ζ >2
β=1
k
Recalling that ξ2 = β=1 yβ Yβ , we now have from Cauchy-Schwarz in-
equality
2
2
(6.28) |ξ1 | < ξ1 , ζ > + 4 < J(ξ2 )ξ1 , ζ >
k 2
2
= |ξ1 | < ξ1 , ζ > + 4 yβ < J(Yβ )ξ1 , ζ >
β=1
k
≤ |ξ1 |4 + 16 |ξ2 |2 < ξ1 , ζ >2 + < J(Yβ )ξ1 , ζ >2
β=1
k
= N 4 < ξ1 , ζ >2 + < J(Yβ )ξ1 , ζ >2
β=1
k
2 2 2
≤ |ξ1 | |ζ| + 2 < ξ1 , ζ > +2 < J(Yβ )ξ1 , ζ >2 .
β=1
k
(6.29) < ξ1 , ζ >2 + < J(Yβ )ξ1 , ζ >2 ≤ |ξ1 |2 |ζ|2 .
β=1
Remark 6.7. For the first Heisenberg group H1 the proof of Theorem 6.6
can be simplified to some extent due to the fact that in R1 one has < x, y >=
xy. This causes some special cancellations in the expression of the quadratic
form < (X 2N ) ζ, ζ > which, however, fail to occur in the higher-dimensional
case.
Proof. We need to prove that given any g ∈ G, for any g ∈ Hg one has
for 0 ≤ λ ≤ 1
and (6.30) holds. We are thus left with case (ii). If ξ(g ) is not on the
straight line segment joining ξ(g) to the origin in g, then we argue as in case
(a) using Theorem 6.6. The final case is when ξ(g ) belongs to the straight
line segment joining ξ(g) to the origin in g, i.e.,
which gives
7. Convexity of sets.
In this section we introduce a notion of convexity for sets which turns out
to be equivalent, at level of the epigraph, to that of weakly H-convex func-
tion. Such characterization is very desirable since, quoting from p.25 of [Ro],
“...it emphasizes the geometry which is fundamental to the theory of convex
functions”.
Remark 7.2. We remark explicitly that, in view of Proposition 4.4, the ge-
ometric interpretation of Definition 7.1 is that for any g ∈ A the intersection
of A with the horizontal plane Hg is starlike in the Euclidean sense with
respect to g at the level of the Lie algebra.
Ωa = {g ∈ G | u(g) ≤ a} ,
u(gλ ) ≤ (1 − λ) a + λ a = a .
The second part now follows from the first and from Theorem 6.8.
u(g ) ≤ s . We want to prove that for every λ ∈ [0, 1] one has g̃λ ∈ epi u.
From Proposition 4.4 we obtain
˜ λ ) = (1 − λ)ξ(g) + λξ(g ), (1 − λ)s + λs .
ξ(g̃
This shows that the condition g̃λ ∈ epi u is equivalent to
u(gλ ) ≤ (1 − λ)s + λs ,
with gλ given by (4.1). On the other hand, the weak H-convexity of u implies
u(gλ ) ≤ (1 − λ) u(g) + λ u(g ) ≤ (1 − λ)s + λs .
This proves the necessity. The sufficiency is established by simply re-
versing the above implications, and choosing s = u(g), s = u(g ) in the end.
φ ∈ Co∞ (G) .
ζ
Tζ (φ) = φ(g) d νH (g) ,
G
(8.1)
ζ 1 αα
φ(g) d νHαα (g) = φ(g) d νH (g)
G 2 G
1 1
+ u(g) {XαXα φ(g) + Xα Xαφ(g)} dg + αα
φ(g) d νH (g) .
2 G 2 G
The notation BVH,loc (Ω) indicates the collection of functions u ∈ L1loc (Ω),
such that u ∈ BVH (ω), for every ω ⊂⊂ Ω.
Returning to the proof of the corollary, let then ω ⊂⊂ G, and consider
ζ ∈ FH (ω). For any α = 1, ..., m we have
m
m
Xα u Xα ζα dg = − u Xα Xα ζα dg .
ω α =1 α =1 ω
m
m
m
Xα u Xα ζα dg = − 2 αα
ζα (g) dνH (g) + (XαXα u, ζα ) ,
ω α=1 α=1 G α =1
where we have denoted by (·, ·) the duality between D (G) and D(G). By
Proposition 8.3 we know that also XαXα u are Radon measures, therefore
we conclude
m
m
m
αα
Xα u Xα ζα dg ≤ 2 νH (ω) + XαXα u(ω) < ∞ .
ω α =1 α =1 α =1
for g ∈ G \ {e}
k
∂N
(8.2) [Xα, X ] N (g) =
α bβαα
∂yβ
β=1
8
k
= bβαα yβ (g) .
N (g)3
β=1
The trivial observation that |4yβ (g)| ≤ N (g)2 allows to conclude from
(8.2)
Cαα
|[Xα, Xα ] N (g)| ≤ ,
N (g)
where we have let Cαα = 2 kβ=1 |bβαα |. Since N −1 ∈ Lqloc (G) for every
1 ≤ q < Q, where Q represents the homogeneous dimension of G, the first
part of the proposition follows.
Suppose now that G be of Heisenberg type, then by Theorem 6.8 we
know that N is weakly H-convex. Furthermore, by (6.11) in Lemma 6.3 we
have
|x(g)|2
|XN (g)|2 = ≤ 1,
N (g)2
and therefore N ∈ L1,∞ loc (G). All the hypothesis of Proposition 8.3, and
therefore of Corollary 8.1, are thus fulfilled. The conclusion follows.
our Theorem 9.2 in the next section, and of (8.3). We plan to come back to
these questions in a forthcoming article.
for any go ∈ G and every R > 0. In particular, (9.1) shows that there exist
no bounded functions in CH
w
(G) ∩ C(G), other than the constant functions.
Proof.
Given a kernel K ∈ Co∞ (G), K ≥ 0, supp K ⊂ B(e, 1),
G K(g)dg = 1, consider the corresponding approximation to the identity
{K }>0 associated with it. Let u = K u, then u ∈ C ∞ (G), and by the
hypothesis u ∈ C(G) we have u → u uniformly on compact subsets of G.
We now fix a gauge ball B(go , R) ⊂ B(go , 3R) ⊂ G. By Remark 5.9
the functions u are weakly H-convex. Proposition 5.7 implies for any fixed
g ∈ B(go , R) and every g ∈ Hg
the latter equality being justified by the fact ξ2 (h) = 0, ... , ξr (h) = 0. On
the other hand, (2.7) gives x(h) = x(g −1 g ) = x(g ) − x(g), and (9.3) follows.
If we now take g ∈ ∂B(g, R) ∩ Hg in (9.2), for 0 < < R/2 we obtain
2
|Xu(g)| ≤ ||u||L∞(B(go,3R)) .
R
By the arbitrariness of g ∈ B(go , R) we conclude
2
(9.5) ||Xu||L∞ (B(go,R)) ≤ ||u||L∞(B(go,3R)) .
R
From Theorem 2.5, and from (9.5) we obtain
C
|u (g) − u (g )| ≤ ||u||L∞(B(go,3R)) d(g, g ) , g, g ∈ B(go, R) ,
R
where C > 0 is an absolute constant. Passing to the limit as → 0 we find
C
(9.6) |u(g) − u(g )| ≤ ||u||L∞(B(go,3R)) d(g, g ) , g, g ∈ B(go , R) .
R
Theorem 9.2. Let u ∈ CH w (G) ∩ C(G), then there exists C = C(G) > 0
1
(9.7) sup |u| ≤ C |u| dg ,
B(g,r) |B(g, 5r)| B(g,5r)
and
C 1
(9.8) ess sup |Xu| ≤ |u| dg .
B(g,r) r |B(g, 15r)| B(g,15r)
now conclude that φ (s) ≥ 0 for every s ∈ [−1, 1], therefore φ is a convex
function on the interval [−1, 1]. Jensen’s inequality then implies
1
1
φ(0) ≤ φ(s) ds ,
2 −1
which we rewrite
1
1
(9.9) u(e) ≤ u(γ(s)) ds .
2 −1
For any s ∈ [−1, 1], we apply Proposition 5.7 obtaining for every g ∈
Hγ(s)
u(γ(s)) ≤ u(g ) − < Xu(γ(s)), x(g) − x(γ(s)) > .
We integrate the latter inequality with respect to g and to 2-dimensional
Lebesgue measure dσ(g ) on the set B(γ(s), 1) ∩ Hγ(s) , obtaining
(9.10)
u(γ(s)) dσ(g ) ≤ u(g ) dσ(g )
B(γ(s),1)∩Hγ(s) B(γ(s),1)∩Hγ(s)
whereas
the vanishing of the latter integral being justified by the fact that the inte-
grand is an odd function of the variable of integration (x, y). Inserting these
two equations in (9.10) we conclude
1
(9.11) u(γ(s)) ≤ |u(g )| dσ(g ) .
π B(γ(s),1)∩Hγ(s)
Notions of convexity in Carnot groups 315
We now observe that the integral in the right-hand side of the latter in-
equality is finite. To see this we notice that as s varies in [−1, 1] the regions
B(γ(s), 1) ∩ Hγ(s) overlap only along the segment of the x-axis (which has
2-dimensional measure zero) joining the points (−1, 0, 0), (1, 0, 0). Further-
more, one easily sees that
(9.14) B(γ(s), 1) ∩ Hγ(s) ⊂ Ws = {(x, y, t) | |x| < 2, |y| < 1, t = −2sy} ,
and
(9.15) Ws ⊂ B(e, 4) .
s∈[−1,1]
≥ ζ(g )u(g ) dσ(g ) − < Xu(g ), x(g ) > ζ(g ) dσ(g ) .
B(e,1)∩He B(e,1)∩He
316 D. Danielli, N. Garofalo, and D.-M. Nhieu
= x ux (g ) + y uy (g ) ζ(g ) dσ(g )
B(e,1)∩He
and from (9.18), we conclude the existence of a constant C2 > 0 such that
We now assert that an estimate similar to (9.19) holds at any point γ(s),
s ∈ [−1, 1], with an absolute constant C2 > 0 independent of s, that is
The details of the verification of (9.20) are similar to those in the proof
of (9.19) and are left to the reader. If we now let, as in the proof of (9.16),
φ(s) = u(γ(s)), then we can rewrite (9.20) as follows
Integrating (9.21) on the interval [0, 1], and using (9.14), (9.15), we obtain
1
(9.22) φ(s) ds ≥ − C2 |u(g )| dg ≥ − C2 |u(g )| dg .
0 B(e,4) B(g,5)
Notions of convexity in Carnot groups 317
where we have let go = (1, 0, 0) ∈ B(e, 1). Invoking (9.16), in which we take
g = e, we obtain
This establishes the sought for estimate from below, which, together with
(9.16), proves (9.7) when G = H1 . In the case of a general Carnot group G
we define γ : [−1, 1] → G as follows
One of the most important aspects of the theory of convex functions is its
connection with geometry through the Monge-Ampère equation. Beautiful
318 D. Danielli, N. Garofalo, and D.-M. Nhieu
n
det D 2 u
(10.1) K = κj = ,
j=1
(1 + |Du|2 )(n+2)/2
Definition
10.2. The equation (10.2) is called H-elliptic if the matrix
∂F
∂Mij is positive definite, i.e., if for every ζ ∈ Rm \ {0} one has
i,j=1,...,m
m
∂F
ζi ζj > 0 .
∂Mij
i,j=1
Notions of convexity in Carnot groups 319
def
m−1
KH (S; go) = κi (xo) .
i=1
and we cannot apply the above considerations. In this case the situation
is more complex, as one needs to identify geometric conditions on S near
go which allow to define the horizontal principal curvatures by appropriate
limiting relations. For these aspects we refer the reader to the forthcoming
article [DGN3].
We close this section by introducing some geometric notions of convexity
which will prove important in the study of the equation (10.3), and of its
connections to the problem stated in the introduction. In the process, we
will also propose an appropriate notion of viscosity solution. We begin by
developing a few preliminary properties. We recall Definition 3.6 of an H-
affine function A : G → R
Γ−
H (u; Ω) = {go ∈ Ω | ∂H u(go) = ∅} .
It is clear from (10.5) and (10.6) that ∂H u(go) = ∅ if and only if u admits
a support at go .
Γ−
H (u; Ω) = Ω ,
then u ∈ CH
w
(Ω).
for any g ∈ Ω ∩ Hgo , and every λ ∈ [0, 1]. From Proposition 4.2 we know
gλ ∈ Ω ∩ Hgo . Let A be a support for u at gλ . From (10.6) we have
where we have denoted by Be 0, Ra the m-dimensional Euclidean ball in V1
centered at the origin and having radius a/R.
ψ (0) 2
ψ(λ) = ψ(0) + ψ (0) λ + λ + o(λ2 ) .
2
Notions of convexity in Carnot groups 325
From the latter equation, and from (5.12) and (5.13) in Proposition 5.4,
we conclude
(10.12)
φ(gλ ) = φ(go) + < Xφ(go), ξ1 (g) − ξ1 (go ) > λ
1
+ < (X 2 φ(go)) (ξ1 (g) − ξ1 (go)), (ξ1(g) − ξ1 (go)) > λ2 + o(λ2 ) .
2
From Proposition 4.2 and from the weak H-convexity of Ω we know that
gλ ∈ Ω ∩ Hgo for every λ ∈ [0, 1]. Since the restriction of u − φ to Hgo has a
local maximum in go , we conclude that there exists λo > 0 sufficiently small
such that
(10.13)
u(gλ ) ≤ u(go ) + < Xφ(go), ξ1(g) − ξ1 (go) > λ
1
+ < (X 2φ(go )) (ξ1 (g) − ξ1 (go )), (ξ1(g) − ξ1 (go )) > λ2 + o(λ2 ) .
2
From the assumption (10.10) we infer the existence of p(go ) ∈ V1 such
that
(10.14)
< p(go), ξ1 (gλ ) − ξ1 (go ) > ≤ < Xφ(go), ξ1(g) − ξ1 (go ) > λ
1
+ < (X 2 φ(go)) (ξ1 (g) − ξ1 (go)), (ξ1(g) − ξ1 (go)) > λ2 + o(λ2 ) .
2
By (2.7) one has ξ1 (gλ ) − ξ1 (go) = λ(ξ1(g) − ξ1 (go)), therefore we obtain
from (10.14)
(10.15) < p(go), ξ1 (g1 ) − ξ1 (go) > ≤ < Xφ(go), ξ1 (g) − ξ1 (go ) >
1
+ < (X 2 φ(go )) (ξ1 (g) − ξ1 (go)), (ξ1(g) − ξ1 (go)) > λ + o(λ) .
2
326 D. Danielli, N. Garofalo, and D.-M. Nhieu
det (X 2 N (g)) = 0 if g = e .
Proof. To prove the theorem we show that the horizontal Hessian matrix
(X 2 N (g)) is singular for g = e by exhibiting a non-zero vector p = p(g) ∈
Rm V1 in the kernel of the corresponding linear map associated to this
matrix. In fact, we take p = Xu = (X1 u, ..., Xmu) where u(g) = N (g)4.
Since by (6.23),
2 1 2 3
(X N ) =
(X u) − (Xu ⊗ Xu) ,
4 N3 4 N4
to prove the theorem it will suffice to show
3
(10.17) (X 2u) − (Xu ⊗ Xu) (Xu) = 0 in G \ {e} .
4 N4
|Xu|2 = 16 N 6 |XN |2 = 16 N 4 ψ .
(10.21) (X 2ψ 2 ) (Xψ 2) = 48 ψ 2 ξ1 .
328 D. Danielli, N. Garofalo, and D.-M. Nhieu
¿From (6.21) and from (6.24) we obtain for the α−th component of
16(X 2ψ 2 ) (Xχ)
m
16 4δαα ψ + 8xαxα < J(ξ2 )ξ1 , Xα >
α =1
= 16 4ψ < J(ξ2 )ξ1 , Xα > + 8xα < J(ξ2 )ξ1 , ξ1 > .
Next, from (6.16) and (6.20) we obtain for the α−th component of
16(X 2χ) (Xψ 2)
k
m
32 ψ < [ξ1 , Xα], Yβ >< [ξ1 , Xα ], Yβ > xα
β=1 α =1
k
= 32 ψ < [ξ1 , Xα], Yβ >< J(Yβ )ξ1 , ξ1 > = 0 .
β=1
This gives
Finally, from (6.16) and (6.24) we find for the α−th component of
16 (X 2 χ) (Xχ)
2
162
k m
< [ξ1 , Xα], Yβ > < [ξ1 , Xα ], Yβ >< J(ξ2 )ξ1 , Xα >
2
β=1 α =1
162
k
= < [ξ1 , Xα], Yβ >< J(Yβ )ξ1 , J(ξ2 )ξ1 >
2
β=1
162
k
= < [ξ1 , Xα], Yβ >< Yβ , ξ2 > |ξ1 |2
2
β=1
2
16 162
= ψ < [ξ1 , Xα], ξ2 > = ψ < J(ξ2 )ξ1 , Xα > .
2 2
This gives
162
(10.24) 162 (X 2 χ) (Xχ) = ψ J(ξ2 )ξ1 .
2
Notions of convexity in Carnot groups 329
The most elementary properties of classical convex functions are their be-
ing locally bounded and continuous on their effective domain. Correspond-
ing properties for weakly H-convex functions are far from obvious and are
presently being investigated. The purpose of this section is to prove that
by strengthening Definition 5.5, but of course without falling back into the
notion of strong H-convexity, one obtains an “intermediate” class of convex
functions which, remarkably, share with the classical ones the above men-
tioned properties. This implies in particular that the results in Sections 8
and 9 are valid generically for such functions, without any additional as-
sumption.
To set the stage, consider in the Heisenberg group a function u : Hn → R.
For every go = (xo, yo , to ) ∈ Hn denote by Ugo the restriction of u to the
horizontal plane Hgo , i.e.,
def
Ugo (x, y) = u(x, y, to + 2 (< yo , x > − < xo , y >)) , (x, y) ∈ R2n .
Proof. The inclusion CH (Hn ) ⊂ CHw (Hn ) is clear since for any given g ∈ Hn ,
Definition 11.1 gives for any g ∈ Hg and 0 ≤ λ ≤ 1
(1 − λ) g + λ g = gλ ,
point go = (−1/2, 0, 0). The horizontal plane Hgo has equation t = y, and
therefore the restriction of N (g) to Hgo is given by
1/4
Ugo (x, y) = (x2 + y 2 )2 + y 2 , (x, y) ∈ R2 .
If Ugo (x, y) were an ordinary convex function, then its restriction to the
y-axis
1/4
Ugo (0, y) = y 4 + y 2
would be a convex function of y, and this is clearly false since Ugo (0, y) ≈
|y|1/2 for |y| ≈ 0.
Proof. We present the proof in the case of the first Heisenberg group H1 .
The case n ≥ 2 only requires minor modifications, which are left to the
reader. Consider the four points
e+
1 = (1, 0, 0) , e−
1 = (−1, 0, 0) , e+
2 = (0, 1, 0) , e+
2 = (0, 1, 0) .
L+
1 = (1, s, −2s) , L−
1 = (−1, s, 2s) ,
L+
2 = (s, 1, 2s) , L−
2 = (s, −1, −2s) ,
t t
P1+ (t) = 1, − , t , P1− (t) = − 1, , t ,
2 2
+
t −
t
P2 (t) = , 1, t , P2 (t) = − , −1, t .
2 2
We now consider the four additional points
1 1
ẽ±
1 = (± , 0, 0) ẽ±
2 = (0, ± , 0) ,
2 2
and construct corresponding straight lines segments
1 1
L̃±
1 = (± , s, ∓s) , L̃±
2 = (s, ± , ±s) , −1 ≤ s ≤ 1 ,
2 2
as above. Such segments intersect the plane Hgt in the four points
1 1
(11.4) P̃1± (t) = ± , ∓t, t , P̃2± (t) = ± t, ± , t .
2 2
Notions of convexity in Carnot groups 333
Denote by S̃(t) the simplex in the plane Hgt generated by these four
points. A moment’s thought and elementary geometric considerations will
convince the reader that
U = S̃(t)
−1≤t≤1
where P̃o (t) is an arbitrarily fixed point in S̃(t), for instance P̃1+ (t). However,
the two estimates (11.3) and (11.5) guarantee that
Remark 11.5. It should be clear to the reader that by suitably varying the
choice of the points e±
i in the proof of Theorem 11.4 we can make the radius
δ in the estimate (11.7) arbitrarily large.
Proof. Again, we present the proof only for H1 , and leave the minor
adjustements for n ≥ 2 to the reader. By left-translation it will suffice
to prove that u is continuous at the group identity e = (0, 0, 0). From
Theorem 11.4, and Remark 11.5 after it, we know that u ∈ L∞ (B(e, 1)).
Set M = ||u||L∞ ((B(e,1)). We will prove that there exist ηo = ηo > 0, and
C = C(M ) > 0 such that for every go = (xo , yo , to) ∈ B(e, ηo) we have
We first consider the case when go belongs to the center of the group, i.e.,
go = (0, 0, to), and without loss of generality we assume to > 0. Furthermore,
we impose that to ≤ ηo2√, where ηo > 0 will be suitably chosen in the sequel.
Consider the point (− 2to , 0, 0) ∈ He , and the straight-line segment L1 (to )
lying on He which joins e to such point. The horizontal plane H(− √to ,0,0)
√ 2
has equation t = to y. On such√plane we consider √
the straight-line segment
√
L2 (to ) which joins the points (− 2to , 0, 0) and (− 2to , to , to ). Finally, denote
by L3 (to ) the straight-line segment lying on the horizontal plane H(0,0,to)
√ √
joining (− 2to , to , to ) to (0, 0, to). We have
√
to √
(11.8) |u(go) − u(e)| ≤ |u(0, 0, to) − u(− , to , to )|
√ 2 √
to √ to
+ |u(− , to , to ) − u(− , 0, 0)|
√2 2
to
+ |u(− , 0, 0) − u(0, 0, 0)|
2 √
to √
= |U(0,0,to)(0, 0) − U(0,0,to) (− , to )|
√ 2 √
to √ to
+ |U(− to ,0,0)(−
√
, to ) − U(− to ,0,0)(−
√
, 0)|
2 2 2 2
√
to
+ |U(0,0,0)(− , 0) − U(0,0,0)(0, 0)| ,
2
where we have denoted by Ugo (x, y) the restriction of the function u(x, y, t)
to the horizontal plane Hgo . By the assumption, each of the functions
U(0,0,to) (x, y), U(− √to ,0,0)(x, y), U(0,0,0)(x, y) is convex, and therefore locally
2
lipschitz. As to varies in the interval [0, ηo2], the segments L1 (to ), L2 (to ) and
L3 (to ) respectively vary in compact subsets K1 (ηo), K2 (ηo ), K3 (ηo ) of the
horizontal planes H(0,0,0), H(− √to ,√t ,t ) , H(0,0,to) . We now choose η in such
2 o o
We next consider the case in which go = (xo , yo, to ) does not belong to
the center, i.e., x2o + yo2 = 0. We now assume that x2o + yo2 ≤ ηo2 , 0 ≤ to ≤ ηo2 ,
where as before we are going to choose ηo sufficiently small. In the present
case we first move along the same polygonal path L1 (to ) ∪ L2 (to ) ∪ L3 (to )
as before, joining the identity to the point (0, 0, to). Once we reach (0, 0, to),
we move from there to the point (xo , yo , to ) along a straight-line segment
L4 (xo , yo, to ) lying on the horizontal plane H(0,0,to) . Similarly to before we
obtain
(11.9) |u(go) − u(e)| ≤ |u(xo, yo , to) − u(0, 0, to)|
√
to √
+ |u(0, 0, to) − u(− , to , to)|
√ 2 √
to √ to
+ |u(− , to , to ) − u(− , 0, 0)|
√2 2
to
+ |u(− , 0, 0) − u(0, 0, 0)|
2
= |U(0,0,to) (xo , yo) − |U(0,0,to) (0, 0)|
√
to √
+ |U(0,0,to) (0, 0) − U(0,0,to) (− , to )|
√ 2 √
to √ to
+ |U(− √to ,0,0)(− , to ) − U(− √to ,0,0)(− , 0)|
2 2 2 2
√
to
+ |U(0,0,0)(− , 0) − U(0,0,0)(0, 0)| .
2
At this point, the same argument given in the case in which go is on the
center allows to infer
|U(0,0,to) (xo , yo) − |U(0,0,to) (0, 0)| ≤ M x2o + yo2 ≤ M d(go, e) .
The remaining three terms in the right-hand side of (11.9) are estimated
as before. We conclude
|u(go) − u(e)| ≤ C(M ) d(go, e) .
336 D. Danielli, N. Garofalo, and D.-M. Nhieu
where for a function f (x, y) we have let τ(−ξ,−η) f (x, y) = f (−ξ + x, −η + y).
By the hypothesis on u, for any fixed h ∈ Hn the function Uh−1 go is a convex
function of the variables (x, y) ∈ R2n , therefore such is also the function
τ(−ξ,−η) Uh−1 go . This implies
1
(11.12) sup |u| ≤ C |u| dg ,
B(g,r) |B(g, 5r)| B(g,5r)
and
C 1
(11.13) ess sup |Xu| ≤ |u| dg .
B(g,r) r |B(g, 15r)| B(g,15r)
References.
[FS] G. B. Folland & E. M. Stein, Estimates for the ∂¯b Complex and
Analysis on the Heisenberg Group, Comm. Pure Appl. Math.,
27 (1974), 429-522.
Department of Mathematics
Purdue University
West Lafayette, IN 47907
E-mail : [email protected]
Department of Mathematics
Purdue University
Notions of convexity in Carnot groups 341
Department of Mathematics
Georgetown University
Washington DC 20057-1233
E-mail : [email protected]