Abstract
In this paper, we introduce the notion of horizontally affine, h-affine in short, function and give a complete description of such functions on step-2 Carnot algebras. We show that the vector space of h-affine functions on the free step-2 rank-n Carnot algebra is isomorphic to the exterior algebra of . Using that every Carnot algebra can be written as a quotient of a free Carnot algebra, we shall deduce from the free case a description of h-affine functions on arbitrary step-2 Carnot algebras, together with several characterizations of those step-2 Carnot algebras where h-affine functions are affine in the usual sense of vector spaces. Our interest for h-affine functions stems from their relationship with a class of sets called precisely monotone, recently introduced in the literature, as well as from their relationship with minimal hypersurfaces.
Keywords: Step-2 Carnot groups, Step-2 Carnot algebras, Horizontally affine functions
Introduction
In this paper, we introduce the notion of horizontally affine function and give a complete description of such functions on step-2 Carnot algebras, or equivalently on step-2 Carnot groups. In the free step-2 rank-n case, we shall see that the vector space of horizontally affine functions is isomorphic to the exterior algebra of . Using the known fact that every step-2 Carnot algebra can be written as a quotient of a free step-2 Carnot algebra, we shall next deduce from the free case a description of horizontally affine functions on arbitrary step-2 Carnot algebras, together with several characterizations of those step-2 Carnot algebras where h-affine functions are affine in the usual sense of vector spaces.
To introduce the discussion, let us recall some definitions. We refer to Sect. 2 for more details. Let be a step-2 Carnot algebra, which means that is a finite dimensional real1 nilpotent Lie algebra of step 2, denotes the derived algebra, and denotes a linear subspace of that is in direct sum with . Such a Lie algebra is naturally endowed with the group law given by
for that makes it a step-2 Carnot group. Actually every step-2 Carnot group can be realized in this way. We shall therefore view a step-2 Carnot algebra both as a Lie algebra and as a Lie group. We also adopt the notation for all and . A function is said to be horizontally affine, and for brevity, we say that f is h-affine and write , if for all and the function is affine. Note that this definition is purely algebraic - it has in particular no connection with the choice of a subRiemannian metric structure on – and can be equivalently restated in geometrical terms as follows. A function is h-affine if and only if its restriction to each integral curve of every left-invariant horizontal vector field is affine when seen as a function from to , where a left-invariant vector field is said to be horizontal whenever it belongs to .
Horizontally affine functions appear naturally in relation with monotone sets, an important class of sets introduced by Cheeger and Kleiner [5], see also [4, 8, 12, 14] and the discussion below. However, h-affine functions are studied systematically for the first time here. See also [1] for a further study of a related notion in more general settings.
Our purposes in the present paper are twofold. We first give a description of h-affine functions on step-2 Carnot algebras, starting with the free case from which the general case will follow. We shall next deduce from this description several characterizations of those step-2 Carnot algebras where h-affine functions are affine. We shall keep the standard terminology saying that a function is affine, writing , to mean that f is affine in the usual sense considering the vector space structure on . Note indeed that by elementary properties of step-2 Carnot algebras, each affine function is h-affine. In other words, in every step-2 Carnot algebra , the vector space is a linear subspace of , see the discussion in Sect. 2. This inclusion may, however, be strict, as we shall see.
More explicitly, h-affine functions can be described with the help of the Carnot dilations given by for , , and . Given a non-negative integer i we define the vector space of i-homogeneous h-affine functions as
We shall prove that every can be written in a unique way as a finite sum of i-homogeneous h-affine functions for some i’s in where if is a free step-2 Carnot algebra and if is a nonfree step-2 Carnot algebra. Recall that the rank of is defined as . Furthermore, denoting by the space of alternating k-multilinear forms over (see Sect. 6 for our conventions about exterior algebra), we shall also prove that for every the vector space is isomorphic to a linear subspace of . See Theorems 1.1, 1.2, and 1.3 for detailed statements.
Let us first consider the free case. Throughout this paper, given an integer , we shall use the model for the free step-2 rank-n Carnot algebra given by
equipped with the Lie bracket where the only nontrivial relations are given by
The induced group law takes the form
for , . For notational convenience, we shall frequently identify with writing elements in as where , .
Given integers , , and , we define as
1.1 |
The description of h-affine functions on can then be given in terms of the functions ’s and reads as follows.
Theorem 1.1
For , we have
-
(i)
.
Furthermore, given and , we have
-
(ii)
if and only if there is , which is unique, such that .
Therefore, for , the spaces and are isomorphic as vector spaces, and hence, so are and . In particular, is a finite dimensional vector space with dimension .
Let us briefly explain our strategy to prove Theorem 1.1. Let be fixed. It is rather easy to verify that if a function is such that for some with then , see Lemma 3.1. The injectivity of the map is also not hard to verify and follows from general facts about exterior algebra, see Corollary 6.5. The main difficulties are thus to get the decomposition given in Theorem 1.1 (i), see Theorem 3.2 and Proposition 3.4 (i), as well as the fact that every function can be written as for some , see Proposition 3.5. This will occupy most of Sect. 3 to which we refer for more details. For the sake of completeness, let us mention the geometric interpretation behind the decomposition in Theorem 1.1 when passing from for to any Lie subalgebra of that is isomorphic to . It can be proved that the zero level set of non-zero n-homogeneous h-affine functions on , namely, the set if n is even, if n is odd, coincides with the union of all Lie subalgebras of that are isomorphic to . Therefore, if with , one gets that its restriction to any Lie subalgebra isomorphic to coincides with the restriction to this subalgebra of the sum of the i-homogeneous terms for that show up in the decomposition of f.
Let us now turn to the general case of arbitrary step-2 Carnot algebras. Our starting point is the known fact that every step-2 Carnot algebra can be written as a quotient of free step-2 Carnot algebras. Namely, by the universal property of free step-2 Carnot algebras, for every , there is a surjective Carnot morphism , see the discussion in Sect. 2. It turns out that there is a one-to-one correspondence between h-affine functions on and h-affine functions on that factor through , see Lemma 2.3 and Corollary 2.4. The description of h-affine functions on can therefore be deduced from the characterization of those functions that factor through . Namely, we shall verify that for the function factors through if and only if annihilates , which means that where
see Lemma 4.1. In the genuinely nonfree setting, such a characterization implies the following decomposition of .
Theorem 1.2
Let be a step-2 rank-r Carnot algebra. Assume that is not isomorphic to . Then .
Note that, in contrast with the free case, one has when is a step-2 rank-r Carnot algebra that is not isomorphic to . This follows from the fact that , and hence , whenever is a surjective Carnot morphism. A description of the summands is provided by the following theorem that applies both in the free and in the nonfree cases (note that when and , one recovers Theorem 1.1) and where the space of annihilators of in is defined by
Theorem 1.3
Let be a step-2 rank-r Carnot algebra, , and be a surjective Carnot morphism. Then the following hold true. For , ,
-
(i)
for every , there is a unique such that ;
-
(ii)
for every , there is a unique such that ;
-
(iii)
via this correspondence, and are isomorphic as vector spaces.
Consequently,
-
(iv)
and are isomorphic as vector spaces.
In particular, is a finite dimensional vector space.
As a consequence of Theorem 1.3, one gets that h-affine functions on step-2 Carnot algebras are polynomials and hence smooth. Let us stress here that there is no regularity assumption in our definition of h-affine functions. Such functions are indeed only assumed to be affine when restricted to horizontal lines and were not even assumed continuous, nor measurable, beforehand. As a further consequence of their smoothness (one actually only needs local integrability), one can characterize elements in as those locally integrable functions that are harmonic in the distributional sense with respect to every subLaplacian on , see Remark 2.8. Let us mention that horizontally affine distributions have been recently studied in [1] in wider settings where they can be proved to be polynomials. Note, however, that this later notion may be different from a pointwise generalization of our present notion of h-affine functions to more general settings, as explained at the end of Remark 2.8.
Several characterizations of step-2 Carnot algebras where can easily be deduced from Theorem 1.3, see Theorem 1.4 below and Sect. 5. It turns out that one of these characterizations can be formulated using a class of Lie algebras known in literature as -null, see [11]. We recall that a step-2 Carnot algebra is -null if every bilinear form satisfying for all vanishes identically on , see Definition 2.6 and Proposition 2.7.
Theorem 1.4
Let be a step-2 Carnot algebra. Then the following are equivalent:
-
(i)
-
(ii)
-
(iii)
, equivalently,
-
(iv)
is -null
-
(v)
, equivalently, , for some, equivalently all, , surjective Carnot morphism.
Note incidentally that it follows from Theorem 1.1 that if and only if . Therefore, the equivalent conditions given in Theorem 1.4 hold true on if and only if . Theorem 1.4 can be efficiently applied in several concrete situations which will be discussed in Sect. 5.2 and to which we refer for more details.
Before closing this introduction, we briefly go back to the relationship between h-affine functions and precisely monotone sets, as defined in [5] in the Heisenberg setting. More generally, a subset of a Carnot algebra, identified with a Carnot group, is said to be precisely monotone if the restriction of its characteristic function to each integral curve of every left-invariant horizontal vector field is monotone when seen as a function from to . Equivalently, a precisely monotone set is a h-convex set with h-convex complement, see for instance [15] for more details about h-convex sets. Precisely monotone sets have been classified in the first Heisenberg algebra , in higher dimensional Heisenberg algebras, and in the direct product , see [5, 12, 14]. In the aforementioned step-2 settings, it turns out that the boundary of a nonempty precisely monotone strict subset is a hyperplane, while in step-3 Carnot algebras the same statement may be false, see [2, 3]. As a consequence of our results, we actually get plenty of examples of Carnot algebras already in the step-2 case where there are precisely monotone subsets whose boundary is not a hyperplane. Indeed, it can easily be seen that sublevel sets of h-affine functions are precisely monotone. Therefore if is a step-2 Carnot algebra that is not -null and if then every sublevel set of f is a precisely monotone set whose boundary is not a hyperplane. We refer to the recent paper [13] for a more detailed introduction to precisely monotone sets as well as for a classification of such sets in the step-2 rank-3 case in terms of sublevel sets of h-affine functions, and for further discussions about higher rank and higher step cases. To conclude these observations, let us mention that measurable precisely monotone sets, and therefore sublevel sets of h-affine functions on step-2 Carnot algebras, can be proved to be local minimizers for the intrinsic perimeter, see [18, Proposition 3.9] and [13, Proposition 2.9].
The rest of this paper is organized as follows. Section 2 contains our conventions and notations about step-2 Carnot algebras and h-affine functions. We also provide easy facts that will be useful for later arguments. In Sect. 3, we focus on the free case and prove Theorem 1.1. Theorems 1.2 and 1.3 are proved in Sect. 4. Section 5.1 is devoted to the proof of Theorem 1.4 and Sect. 5.2 to a discussion of several examples. In the final Sect. 6, we gather notations and facts in linear and exterior algebra.
Step-2 Carnot Algebras and Horizontally Affine Functions
We recall that a real2 and finite dimensional Lie algebra is said to be nilpotent of step 2 if the derived algebra is nontrivial, i.e., , and central, i.e., . Here, given , we denote by [U, V] the linear subspace of generated by elements of the form [u, v] with , . If is a linear subspace of that is in direct sum with then and the decomposition is therefore a stratification of . As a matter of fact, every stratification of a nilpotent Lie algebra of step 2 is of this form.
A step-2 Carnot algebra is a Lie algebra nilpotent of step 2 equipped with a stratification . The rank of is defined as . Such a Lie algebra is naturally endowed with the group law3 given by
for that makes it a step-2 Carnot group. It is actually well known that any step-2 Carnot group can be realized in this way. We shall therefore view a step-2 Carnot algebra both as a Lie algebra and group.
Throughout this paper, we shall always denote by a step-2 Carnot algebra. Given , , we set .
Definition 2.1
Given we say that is if for every , , the function is affine.
When , one recovers the notion of real-valued affine functions on seen as a vector space. Indeed, since is nilpotent of step 2, for , , we have and . Therefore is -affine if and only if for every , , the function is affine, i.e., f is affine, see Proposition 6.1. In particular, real-valued affine functions are A-affine for every .
In the present paper, we are interested in -affine functions, which we shall call horizontally affine, h-affine in short, namely:
Definition 2.2
(h-affine functions) We say that is horizontally affine, h-affine in short, if f is -affine. In other words, f is h-affine if for every , , the function is affine. We denote by the real vector space of h-affine functions on .
We say that is a horizontal line if there are , such that . We already noticed that horizontal lines are 1-dimensional affine subspaces of and therefore h-affine functions can equivalently be defined as functions whose restriction to every horizontal line is affine.
We recall that a Carnot morphism between step-2 Carnot algebras and is a homomorphism of graded Lie algebras, which means that is a linear map such that for all and for . Note that a Carnot morphism is both a homomorphism of graded Lie algebras and a group homomorphism.
Lemma 2.3
Let be step-2 Carnot algebras and be a Carnot morphism. For , we have . If is surjective then if and only if .
Proof
Carnot morphisms map affinely horizontal lines to either horizontal lines or singletons therefore when . If the Carnot morphism is surjective then every horizontal line in is the affine image through of a horizontal line in and therefore when .
For , the Carnot dilation is defined as the linear map such that for , . The family is a one parameter group of Carnot automorphisms. Recall that, given a non-negative integer i, we denote by , the linear subspace of of i-homogeneous h-affine functions on . Since dilations commute with Carnot morphisms, we get from Lemma 2.3 the following corollary.
Corollary 2.4
Let be step-2 Carnot algebras, be a Carnot morphism, and i be a non-negative integer. For , we have . If is surjective then if and only if .
We already noticed that the set of real-valued affine functions on is a linear subspace of . More precisely, we have the following inclusion.
Lemma 2.5
is a linear subspace of .
Proof
Let . There are and linear forms , , such that for all , . Clearly, constant functions belong to and the functions and belong to and , respectively.
We say that step-2 Carnot algebras are isomorphic if there is a bijective Carnot morphism from one to the other. Note that being h-affine, respectively, affine, are intrinsic properties, in particular if and only if for isomorphic step-2 Carnot algebras . This indeed more explicitly follows from Lemma 2.3 together with the fact that Carnot morphisms are linear maps.
Let us recall that by the universal property of free step-2 Carnot algebras, see Sect. 1 for our conventions about the free step-2 rank-n Carnot algebra , given a step-2 rank-r Carnot algebra and given an integer , there is a surjective Carnot morphism , see for instance [17, p.45]. We also recall that for such a Carnot morphism, is a graded ideal in , which means that where are linear subspaces of , , such that for all , .
We now recall the definition of -null Lie algebras that will be used in one of our characterizations of those step-2 Carnot algebras where h-affine functions are affine, see Theorem 1.4.
Definition 2.6
[11] A Lie algebra is said to be -null if for every symmetric bilinear invariant form , we have . Here B is said to be invariant if for all , or equivalently, if the trilinear form is alternating on .
For step-2 Carnot algebras, the previous definition can be rephrased in the following way, of which we omit the elementary proof.
Proposition 2.7
A step-2 Carnot algebra is -null if and only if every bilinear form satisfying for all vanishes identically on .
Remark 2.8
In the present article, we focus on step-2 Carnot algebras or, equivalently, step-2 Carnot groups. Let us mention that the notion of horizontally affine function makes sense in broader generality. One may for instance consider Carnot groups of arbitrary step (see [10, 16] for a primer on the subject) or, more generally, a connected nilpotent Lie group G equipped with a vector subspace of its Lie algebra that Lie generates . Then we say that is -affine if for every the restriction of f to each integral curve of X is affine when seen as a function from to . Here an element is seen as a left-invariant vector field on G. When G is a step-2 Carnot group with stratified Lie algebra and is the first, usually called horizontal, layer of the stratification of , one recovers Definition 2.2, and this latter definition can hence be extended to Carnot groups of arbitrary step in the obvious way. Going back to the aforementioned more general setting and considering G equipped with a Haar measure, let us mention that we have the following characterizations of locally integrable -affine functions. Namely, has a representative that is -affine if and only if one of the following equivalent conditions holds true in the distributional sense:
for every
for every
for every basis of .
Indeed, if a representative of is -affine then (A.1) holds true as a consequence of the very definitions. Conversely, if satisfies (A.1) then f has a representative that is smooth by Hörmander’s hypoellipticity theorem and then it clearly follows from (A.1) that this representative is -affine. The fact that (A.1) is equivalent to (A.2) is a consequence of Hörmander’s hypoellipticity theorem together with the identity for smooth functions f. Condition (A.1) obviously implies (A.3). Conversely, if satisfies (A.3) and , one can complete X into a basis of . Then for every one has with the left-hand side converging to as and therefore . See also [1] for other generalizations of condition (A.1) for locally integrable functions.
To conclude this remark, note that in the specific setting considered in this paper, i.e., step-2 Carnot algebras , it follows from Theorem 1.3 that h-affine functions are smooth and hence locally integrable. Therefore each of the distributional sense conditions (A.1), (A.2), (A.3) with makes sense for all h-affine functions and hence characterizes such a class of functions. In the more general setting considered in the present remark, it is, however, not clear to us whether -affinity implies local integrability, and the class of locally integrable -affine functions that can be characterized through each of the equivalent conditions (A.1), (A.2), and (A.3) could therefore be a strict subset of the class of -affine functions.
Horizontally Affine Functions on Free Step-2 Carnot Algebras
This section is devoted to the proof of Theorem 1.1. The proof will proceed into 4 steps. We first verify in Lemma 3.1 that for , we have where is given by (1.1), together with the injectivity of the linear map . We shall next prove Theorem 1.1 for , see Theorem 3.2, and deduce properties of h-affine functions on for to be used in the next step, see Proposition 3.3. When , we first prove that together with preliminary information about elements in , see Proposition 3.4. We then upgrade these information in Proposition 3.5 to get the description stated in Theorem 1.1.
For notational convenience, we identify in this section with and write elements in as with , . In the next lemma, we denote by the analogue of for -valued functions. More explicitly, belongs to if and only if for every , , the function is affine, and for all .
Lemma 3.1
For , , and , we have where is given by (1.1). Furthermore, the linear map is injective.
Proof
Let . Clearly for all . If is even, we have
if is odd,
for all , . Therefore . For the injectivity of the linear map , see Corollary 6.5.
Theorem 3.2
We have .
Proof
We recall that a set is said to be a horizontal line if for some , . Define the h-affine hull of a set as the smallest set C containing A with the property that if a horizontal line meets C in more than one point then . It follows from [5, Lemma 4.10] that there are 4 points in whose h-affine hull is . Indeed, given linearly independent , the h-affine hull C of contains a pair of parallel lines with distinct projection in the sense of [5], namely, the horizontal line through (0, 0) and and the horizontal line through and , therefore by [5, Lemma 4.10]. This implies that is a vector space with dimension 4. Since is a 4-dimensional linear subspace of , we get that , as claimed.
Note that Theorem 1.1 for follows from Lemma 2.5, Lemma 3.1 and Theorem 3.2. For , we set where denotes the Lie subalgebra of generated by . We refer to Definition 2.1 for the definition of -affine functions.
Proposition 3.3
For , , the following hold true:
3.1 |
3.2 |
where is given by and denotes the space of real-valued affine functions on .
Proof
Clearly, composing h-affine functions with left-translations yields h-affine functions. Therefore, to prove that every is -affine, we only need to verify that for every , , , , the function is affine. Set and denote by the restriction of f to . On the one hand, the structure of step-2 Carnot algebra of induces on a structure of step-2 Carnot algebra that makes it isomorphic to . Therefore by Theorem 3.2. On the other hand, . Thus , which implies that for all , the function is affine and concludes the proof of (3.1). To prove (3.2), note that for , , , we have . Since , it follows from Proposition 6.1 that for every -affine function f, and hence, in particular for by (3.1).
In addition to the notations given in the appendix, see Sect. 6, we shall use the following ones in the rest of this section. Recall that denotes a basis of . For , we set for , see (6.4) for the definition of , with the convention .
In the following, for , we write to denote the multi-index with 2 indices. We set
and we equip with the lexicographic order, i.e., we write to mean either that and or that . We set ,
for , and . We write and for . Given , we denote by the unique element in such that and we write to mean that .
We recall from Sect. 6 that for a multi-index . For , we set and for .
We write and for . We set and
for . Note that , . When , we have for , and for .
Proposition 3.4
For , the following holds true:
-
(i)
,
-
(ii)for , every can be written as
for constants , -
(iii)for , every can be written as
for linear forms .
Proof
For , the are linear subspaces of that are in direct sum. Therefore . Conversely, let be given.
We first prove that there are functions , , such that
3.3 |
Let be given. We know from (3.1) that, for every , , the function is affine. Since , it follows from elementary properties of multiaffine maps, see Proposition 6.2 applied to , that there are , , such that (3.3) holds true.
Next, we prove that
3.4 |
We have and we prove by induction on k that for all . For , we have with and we apply (3.2) with to get that . Given , assume that for all . For , we apply (3.2) with to get that
By induction hypothesis, we get that , which concludes the proof of (3.4).
It follows from (3.4) that there are constants and linear forms , , such that for every . For , we set
so that . We claim that for all . Indeed, let , , be given. Since the dilations are Carnot automorphisms, we know from Lemma 2.3 that for all . Therefore
for all , which implies that for all . Since this holds true for all , , , we get that . Clearly, we also have for all . Therefore for all , as claimed.
For , we now claim that
3.5 |
For , we have and there is nothing to prove. Let be given. First, note that if and only if there are integers such that either , or , or . Now, let be given. On the one hand, since , we know that, for every , the function is affine. On the other hand, this function is a polynomial for which the coefficient of , namely,
must therefore vanish. Since this holds true for every , it follows that for every such that . Considering the function , respectively, , and arguing in a similar way, we get that for every such that , respectively, such that , which concludes the proof of (3.5).
For , we claim that
3.6 |
Indeed, let , be given. Consider the function given by for . We have for all , . Since , it follows from (3.1) that . We then argue as for the proof of (3.5) to get that for every , which concludes the proof of (3.6).
For , we have and it follows from (3.5) and (3.6) that .
We now prove that whenever n is even. Assume that with . Let be given. Note that . Therefore, to show that , we need to verify that for every . Let be given. Set . On the one hand, since , the function is affine. On the other hand,
Therefore the coefficient of vanishes, i.e., . To prove that , we argue in a similar way considering the function .
All together, we have shown that with that can be written as in (ii) when is even, respectively, as in (iii) when is odd, which concludes the proof of the proposition.
Proposition 3.5
For , , , every can be written as for some .
Proof
Assume with no loss of generality that . We first prove the proposition when is even. Let and let , , be given by Proposition 3.4 (ii) so that
For we get that f is constant and the required conclusion clearly holds true. Next, let us consider the case with and . For , we have
where, given , denotes the signature of the permutation of given by . Therefore it suffices to prove that
3.7 |
On the one hand, since , we know that for all , , the function
is affine. On the other hand, this function is a polynomial for which the coefficient of is given by
and must therefore vanish. Since this holds true for all , it follows that for all , ,
Now let be given and let be such that . Then the previous equality reads as
for all . Looking at the coefficient of we get that . Looking at the coefficient of we get that . Therefore, we have proved that for all such that for some . Since one can pass from any to any by a finite number of such steps, (3.7) follows.
Let us now consider the case and is even with . For set and define by so that
Set and . We have for all , . Since , it follows that the restriction of to belongs to . Since for all and all , we get that the restriction of to belongs to and it follows from the previous case that there is such that for all . Since does not depend on , this equality holds actually true for all , . For , we have where denotes the projection map given by . Therefore, for , we have
where is such that . Now, note that , therefore , and the previous equality becomes
where , which concludes the proof of the proposition when is even.
We now consider the case where is odd. Let and let , , be linear forms given by Proposition 3.4 (iii) so that
For we get that and the required conclusion clearly holds true. Thus assume that and let be given. As in the proof of (3.6), consider the function given by for . We have , see the proof of (3.6), and since , it follows that . Then we know from the previous cases and Corollary 6.5 that there is a unique such that for all . Next, it follows from the linearity of the that the map is linear for every . Since , see (6.1), we get that is linear. We now claim that for all . Indeed, on the one hand, we know that the function is affine for all , . On the other hand, we have
![]() |
and hence the coefficient of vanishes, i.e., for every , . Since , see (6.2), we get that for all , as claimed. Then the required conclusion follows from Proposition 6.6, and this concludes the proof of the proposition.
Horizontally Affine Functions on Arbitrary Step-2 Carnot Algebras
In this section, we prove Theorem 1.2, that will be deduced from Theorem 1.1 writing a step-2 rank-r Carnot algebra that is not isomorphic to as a proper quotient of , and Theorem 1.3. The main argument for proving both theorems is given in Lemma 4.1 where we characterize those functions in that factor through where a graded ideal of . We refer to (6.5) and (6.6) for the notions of annihilators. For notational convenience, we shall again identify with throughout this section.
Lemma 4.1
Let and be a graded ideal of . For , , the map given by (1.1) factors through if and only if .
Proof
Write the graded ideal of as where , are linear subspaces of , such that for all , . Recall that factors through if and only if for all , . If then is constant and therefore clearly factors through . Since , see (6.7), this proves the lemma for . If then . Therefore factors through if and only if for all , i.e., , where the last equality comes from (6.8), which proves the lemma for . Now let . For , it easily follows from (1.1) that factors through . Conversely, let and assume that factors through . Then, for all , , , we have , i.e.,
Identifying the coefficient of degree 1 in t, we get that for all , , ,
It then follows from (6.1) when j is even, (6.2) when j is odd, and Lemma 6.4 that for all , i.e., , where the last equality comes from (6.9), and this concludes the proof of the lemma.
We first prove Theorem 1.2.
Proof of Theorem 1.2
Let be a step-2 rank-r Carnot algebra that is not isomorphic to . Clearly, . To prove the converse inclusion, let be a surjective Carnot morphism and be fixed. Recall for further use that is a nontrivial graded ideal of . Let . Then by Lemma 2.3 and it follows from Theorem 1.1 that there are , , such that . We claim that each factors through . Indeed, since commutes with dilations, we have for all , , ,
This implies that for all , for all , , i.e., factors through , as claimed. Since is surjective, it follows that for each there is such that . Since , we get from Corollary 2.4 that (note indeed that the analogue of Corollary 2.4 holds true for -valued functions). Let us now verify that . Since factors through , we know from Lemma 4.1 that . Since is not isomorphic to , we have and hence . Therefore and hence . All together we get that and this concludes the proof of Theorem 1.2.
We now prove Theorem 1.3
Proof of Theorem 1.3
Let be a step-2 rank-r Carnot algebra, , and be a surjective Carnot morphism. Let , , . Since is a graded ideal of , we know from Lemma 4.1 that factors through and since is surjective we get the existence of a unique function such that . Furthermore, since , we get from Corollary 2.4 that , which concludes the proof of Theorem 1.3 (i). Conversely, let . Then by Corollary 2.4 and it follows from Theorem 1.1 (ii) that there is a unique such that . This equality shows in turn that factors through and hence by Lemma 4.1, which concludes the proof Theorem 1.3 (ii). By linearity of the map , we get that the bijective map where f is given by Theorem 1.3 (i) is linear and therefore is an isomorphism of vector spaces, which concludes the proof of Theorem 1.3 (iii). By Theorem 1.1 (i) and Theorem 1.2, it follows that and are isomorphic as vector spaces. Finally, since is a graded ideal of , we get from Corollary 6.9 that . Therefore and are isomorphic as vector spaces, which concludes the proof of Theorem 1.3 (iv).
Remark 4.2
If is a step-2 rank-r Carnot algebra that is not isomorphic to then . Indeed consider a surjective Carnot morphism . Then is a nontrivial graded ideal of that is contained in . Therefore by (6.7), where the first equality follows from (6.8), and for . This latter claim indeed follows from the inclusion together with the fact that for , see Lemma 6.4. By Theorem 1.3 (iii) we get that and are isomorphic for and for . Therefore by Theorem 1.2.
Remark 4.3
It follows from Theorems 1.1, 1.2, 1.3 (iii), and Lemma 6.11 that if is a step-2 Carnot algebra then for some non-negative integer i if and only if .
Remark 4.4
Note that it follows from Theorems 1.1 and 1.3 (iii) that if and is a surjective Carnot morphism then for . Similarly, it follows from Theorems 1.2 and 1.3 (iii) that if is a step-2 rank-r Carnot algebra that is not isomorphic to , , and is a surjective Carnot morphism then for .
Step-2 Carnot Algebras Where Horizontally Affine Functions are Affine
Characterization
This section is devoted to the proof of Theorem 1.4 that characterizes step-2 Carnot algebras where h-affine functions are affine. We begin with an easy consequence of Theorem 1.3.
Lemma 5.1
Let be a step-2 Carnot algebra. Then .
Proof
We already know from Lemma 2.5 that . Conversely, let . Let , be a surjective Carnot morphism, and . By Theorem 1.3 (ii) there are , , such that . Then it clearly follows from the form of , see (1.1), together with the fact that is a surjective Carnot morphism that .
We now turn to the proof of Theorem 1.4.
Proof of Theorem 1.4
The equivalence between Theorem 1.4 (i) and (ii) follows from Lemmas 2.5 and 5.1. Next, Theorem 1.4 (ii) clearly implies Theorem 1.4 (iii) since are linear subspaces of that are in direct sum, recalling also that if and only if , see Remark 4.3.
Now, assume that . Let be a bilinear form such that for all . Identifying with , we have for , , ,
and . Therefore and hence , which proves that is -null.
Next, assume that is -null. Let and be a surjective Carnot morphism and let us verify that . Let be given. By Theorem 1.3 (i), for there is such that . Identifying with , we have . Therefore is bilinear. Since is a surjective Carnot morphism, it follows that is bilinear as well. Furthermore, for , let be such that , . Then and hence, identifying with , we have . Therefore . Since is -null, it follows that and hence by Corollary 6.5. Therefore , as wanted. Recall that this is in turn equivalent to by Lemma 6.11.
To conclude the proof of Theorem 1.4, assume that there are and a surjective Carnot morphism such that . By Theorem 1.3 (iii) we get that and Theorems 1.1 (i) and 1.2 imply in turn Theorem 1.4 (ii).
Examples
In this section, we deduce from Theorem 1.4 sufficient conditions implying that h-affine functions are affine and necessary conditions that must be satisfied when this is the case. These conditions may be easier to verify on concrete examples than those given in the characterization obtained in Theorem 1.4. We, however, illustrate with explicit examples to what extent some of these easier conditions cannot be turned into characterizations of step-2 Carnot algebras where h-affine functions are affine. We shall also see from some of these examples that, unlike affine functions, a h-affine function defined on a Lie subalgebra of a step-2 Carnot algebra may not admit a h-affine extension to the whole algebra.
Proposition 5.2
Let be a step-2 Carnot algebra and assume there is such that is surjective. Then .
Proof
If is surjective for some then so is for for some open neighborhood of x. If is a bilinear form such that for all then, by bilinearity of b and surjectivity of , we get for all , , and finally , using once again the bilinearity of b together with the fact that U is a nonempty open subset of . Therefore is -null and hence by Theorem 1.4 (i)–(iv).
Proposition 5.2 applies in particular to step-2 Carnot algebras of Métivier’s type, i.e., step-2 Carnot algebras where is surjective for all .
The condition given in Proposition 5.2 about the surjectivity of for some implying that is -null should not be confused with the surjectivity of the Lie bracket . Indeed, step-2 Carnot algebras of Métivier’s type are examples of -null Lie algebras where the Lie bracket is surjective, whereas Example 5.3 below gives an example of a -null step-2 Carnot algebra where the Lie bracket is not surjective. On the other hand, free step-2 Carnot algebras of rank 3 or higher are not -null whereas the Lie bracket is surjective if and only if .
Example 5.3
Let be the graded ideal of given by and let . Elementary computations show that . Therefore and is -null by Theorem 1.4 (i)–(iv)–(v). To see that the Lie bracket is not surjective we identify with where . The only nontrivial bracket relations are given by
for and we claim that
Indeed, let with . Set for . We have for . If or then and we obviously have . Assume now that , , and . Then and , respectively, and , are colinear. Therefore there are such that and . Since , we get that , as claimed.
Given a step-2 Carnot algebra , we say that is a quotient of if is a step-2 Carnot algebra and there is a surjective Carnot morphism .
Proposition 5.4
Let be a step-2 Carnot algebra such that . Then for every quotient of .
Proof
By [11, Lemma 2.3] every quotient of a -null Lie algebra is -null and hence the proposition follows from Theorem 1.4 (i)–(iv).
Note that it may happen that while for every proper quotient of , i.e., for every quotient of that is not isomorphic to . A simple example is given by the free step-2 rank-3 Carnot algebra . Indeed, we know from Theorem 1.1 that , whereas every proper quotient of is either isomorphic to or has rank 3 and is not isomorphic to , therefore by Theorems 3.2, 1.2, and 1.4 (i)–(ii). See also Example 5.6 for another example that is not isomorphic to .
Note also that since , Proposition 5.4 has the following immediate corollary.
Corollary 5.5
Let be a step-2 Carnot algebra that has as one of its quotients. Then .
It may happen that while does not have as one of its quotients, as shows Example 5.6 where is a step-2 rank-5 Carnot algebra such that and for every proper quotient of .
Example 5.6
Let be the graded ideal of given by where and let so that is in particular a step-2 rank-5 Carnot algebra and therefore is not isomorphic to . We have and hence by Theorem 1.4 (i)–(v). We now claim that for every proper quotient of . Indeed, let be a proper quotient of and let be a surjective Carnot morphism. Let denote the quotient map so that is a surjective Carnot morphism and let us verify that . Let . We have . Since , we have , implying , and for , , implying for , . Therefore . If then which implies in turn and therefore . This contradicts the fact that is a proper quotient of and hence . Therefore and it follows from Theorem 1.4 (i)–(v) that , as claimed.
We recall that the direct product of a step-2 Carnot algebra with an abelian Lie algebra and the direct product of step-2 Carnot algebras inherit naturally of a structure of step-2 Carnot algebra from those of and .
Proposition 5.7
Let be a step-2 Carnot algebras and be an integer. Then the following hold true:
-
(i)
if and only if and
-
(ii)
if and only if .
Proof
By [11, Lemma 2.3], any finite direct product of -null Lie algebras is -null. Therefore it follows from Theorem 1.4 (i)–(iv) that whenever and . Similarly, since abelian Lie algebras are -null (see Definition 2.6), we have whenever . The converse implications in (i) and (ii) follow from Proposition 5.4 noting that the projection maps from onto either or and from onto are surjective Carnot morphisms.
The next proposition is another simple consequence of Theorem 1.4 that gives a sufficient condition ensuring that h-affine maps are affine.
Proposition 5.8
Let be a step-2 Carnot algebra such that . Then there is a Lie subalgebra of isomorphic to .
Proof
By Theorem 1.4 (i)–(iv) we know that is not -null. Then let be a non-zero bilinear form such that for all . Since , there are , such that . By bilinearity together with the fact that , it follows that there are such that . We claim that are linearly independent and therefore the Lie subalgebra of generated by is isomorphic to . Indeed, note that since for all , the trilinear form is alternating and therefore . Now let be such that . For , we have where and . Since for such indices i, k, l, it follows that , which concludes the proof of the lemma.
We stress that -null step-2 Carnot algebras may have Lie subalgebras isomorphic to , as shown in the following two examples.
Example 5.9
[The quaternionic Heisenberg algebra.] Let i, j, k denote the quaternion units satisfying and denote by the set of quaternions. Given , denote by its imaginary part and its conjugate. Equip with the Lie bracket for which the only nontrivial relations are given by for which makes a step-2 Carnot algebra that is well known to be of Heisenberg type, and therefore of Métivier’s type (see for instance [9]). Therefore, we have and is -null by Proposition 5.2. We now claim that the Lie subalgebra of generated by any three linearly independent elements in is isomorphic to . Indeed, for , we have if and only if q and are colinear. This indeed follows from the fact that for , the linear map is surjective with , together with the fact that and . Then let be linearly independent and assume by contradiction that . Exchanging the role of if necessary, there are such that . Then which implies that and are colinear and contradicts the fact that are linearly independent. Therefore and the Lie subalgebra generated by is isomorphic to , as claimed.
Example 5.10
Let where be the -null step-2 Carnot algebra given by Example 5.3. We claim that the Lie subalgebra of generated by any three linearly independent elements in is isomorphic to . Indeed let denote the quotient map. Let be linearly independent and let be such that , i.e., . We have while with . It follows that and hence . This proves that and therefore the Lie subalgebra of generated by is isomorphic to , as claimed.
Theorem 1.4 (i)–(iv) together with Proposition 5.8 has the following immediate consequence.
Proposition 5.11
Let be a step-2 Carnot algebra with . Then .
If then is isomorphic to and therefore . This fact generalizes to higher rank step-2 Carnot algebras with in the following way.
Proposition 5.12
Let be a step-2 Carnot algebra with . Then if and only if is not isomorphic to for some non-negative integer d.
Proof
If is isomorphic to the direct product for some non-negative integer d then has as one of its quotients and we know by Corollary 5.5 that . Conversely, assume that . First, if , since , then is isomorphic to . Next, assume that . By Proposition 5.8, there are generating a Lie subalgebra of isomorphic to and there is a bilinear form such that for all and . Set . We claim that for every , there is such that lies in the center of , i.e., for all . To prove this claim, let be given. Since , there are , , such that for . Set and let us verify that . We first verify that for . Since for all , the bilinear form is skew-symmetric. For and , we have . By skew-symmetry, we get for and it follows that for . Now let and write with . Since the trilinear form is alternating, we have . On the other hand, . Hence for and therefore , as wanted. It now clearly follows from this claim that one can complete into a basis of in such a way that , i.e., is isomorphic to .
To conclude this section, let us remark that a h-affine function defined on a Lie subalgebra of a step-2 Carnot algebra may not admit a h-affine extension to the whole algebra. Indeed, let be a step-2 Carnot algebra such that and such that there is a Lie subalgebra of isomorphic to , see Examples 5.9 and 5.10. Then and therefore there is . Assume there is whose restriction to is h. By assumption on , we have . Since is a linear subspace of , it follows that the restriction of to is affine, i.e., , which gives a contradiction. Recall that on the contrary an affine function defined on an affine subspace of a vector space can always be extended to an affine function on the whole space.
Appendix About Linear and Exterior Algebra
We gather in this appendix some basic facts about linear and exterior algebra not pertaining to h-affine functions that have been used in the previous sections.
We start with a characterization of affine maps between real vector spaces whose elementary proof is left to the reader.
Proposition 6.1
Let E, F be real vector spaces. A map is affine if and only if for every , the map is affine.
For the sake of completeness, we state below an elementary property of multiaffine functions that has been used in the proof of Proposition 3.4. The proof can easily be done by induction on the dimension and is left to the reader.
Proposition 6.2
Let be an integer and E be a p-dimensional real vector space. Let and assume that there is a basis of E such that the map is affine for every and . Then f is a linear combination of the , where J ranges over the subsets of and for with the convention .
The rest of this appendix is devoted to (basic) facts about exterior algebra that have been used throughout this paper. Although some of them look quite elementary, we were, however, unable to find references in the literature and thus provide proofs for the reader’s convenience.
Given integers and , we denote by the set of alternating k-multilinear forms over . For , we set . We denote by the exterior algebra equipped with exterior product . We recall that if . For , we set and for .
Lemma 6.3
For , we have in
6.1 |
6.2 |
for all .
Proof
We only need to consider the nontrivial cases where and . Then (6.1) and (6.2) follow from the identity
6.3 |
for all .
Given , we set ,
6.4 |
for , and . We write and for . Given , we denote by the unique element in such that and we set . We fix a basis of and denote by its dual basis. For , we set with the convention .
The space of exterior annihilators of an element in of some given order k has been introduced in [7], see also [6, Section 2.2]. More generally, given , , and , we define the annihilator of A in , respectively, , as
6.5 |
6.6 |
We also set .
Lemma 6.4
For , , we have .
Proof
For , we clearly have . Let and be given. Clearly . Since is graded, i.e., (see Lemma 6.8), if equality fails then for some . Let and be such that . Then . However, we can write for some , , and since , we get , which gives a contradiction.
Writing to denote the set of maps we deduce from Lemmas 6.3 and 6.4 the following corollary.
Corollary 6.5
For , the linear map given by is injective. For , the linear map given by is injective.
Proof
By linearity we only need to verify that these maps have trivial kernel. Let and be such that for all . On the one hand, by (6.1) we have . On the other hand, by Lemma 6.4, we have . Therefore . Similarly, for and such that for all , we have by (6.2) and Lemma 6.4 that .
The next proposition played a key role at the end of the proof of Proposition 3.5.
Proposition 6.6
For the following holds. Let and be linear. Assume that for all . Then there is such that for all .
Proof
When every map satisfies the assumption for all . If is in addition assumed to be linear and is such that , then for all where and is such that .
Let us now argue by induction on n. First, if , the conclusion follows from the previous remark. Next, let . By the previous remark, we only need to consider the case where . By linearity of , we only need to prove that there is such that for . Set for .
For , since , we can write for some . Next, write for some and (when , ). Define and to be linear and such that and for all . For , we have . Thus for all and this implies in turn that and for all . By induction, there are (again when ) and such that and for all . It follows that for all . We set . To conclude the proof of the proposition, it remains to verify that . Since , this is equivalent to showing that for all , see Lemma 6.4. By assumption, we have for all and therefore for all . It follows that for . Since , we finally get , and this concludes the proof of the proposition.
Recall that a graded ideal of can be seen as a linear subspace of of the form where are linear subspaces of, respectively, such that for all , . The structure of annihilators of such subsets of , and in particular Corollary 6.9 and Lemma 6.10, played a major role in Sect. 4. Before proving Corollary 6.9 and Lemma 6.10, we first state two elementary lemmas. The easy proof of Lemma 6.7 is left to the reader.
Lemma 6.7
Let V, W be linear subspaces of that are in direct sum. Then and for all .
Lemma 6.8
Let , , . Then .
Proof
Clearly, . Conversely, let where . For , we have with therefore for all . Since this holds true for all , we get that for all , and therefore .
Corollary 6.9
Let be linear subspaces of, respectively, . Then .
Proof
By Lemmas 6.7 and 6.8, we have
Lemma 6.10
Let , be linear subspaces of, respectively, such that for all , . Then
6.7 |
6.8 |
6.9 |
Proof
Clearly, for all and (6.7) follows. By Lemma 6.7, we have . For , we have therefore and (6.8) follows. For , clearly . Conversely, let and . By assumption, for all . Therefore for all , i.e., , where the last equality follows from Lemma 6.4. Since , we have , therefore and hence . Since this holds true for all , we get that , where the last equality comes from Lemma 6.7 and concludes the proof of (6.9).
We end this section with an observation that has been useful for our purposes in Remark 4.3 and Sect. 5.
Lemma 6.11
Let , , be such that . Then for all .
Proof
Let , . For all , , we have , i.e., . Since , it follows that , where the last equality follows from Lemma 6.4. Since , we have , therefore and hence .
Acknowledgements
The authors are grateful to an anonymous reader of a previous version of this paper for valuable comments and remarks that helped them to improve and simplify the exposition.
Funding
Open Access funding provided by University of Jyväskylä (JYU).
Footnotes
It is worth to stress that our arguments and results can be verbatim extended to finite dimensional nilpotent Lie algebras of step 2 over an arbitrary field of characteristic zero.
As already mentioned in the introduction, our arguments and results can be verbatim extended to finite dimensional nilpotent Lie algebras of step 2 over an arbitrary field of characteristic zero.
Our convention for the group law is nothing but a technical convenience. Any other choice where [x, y] is replaced by for some independent of x and y leads to the same results.
E.L.D. was partially supported by the Academy of Finland (Grant 288501 ‘Geometry of subRiemannian groups’ and by Grant 322898 ‘Sub-Riemannian Geometry via Metric-geometry and Lie- group Theory’) and by the European Research Council (ERC Starting Grant 713998 GeoMeG ‘Geometry of Metric Groups’). S.R. is partially supported by ANR Project SRGI (Sub-Riemannian Geometry and Interactions) ANR-15-CE40-0018.
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Enrico Le Donne, Email: enrico.ledonne@unifr.ch.
Daniele Morbidelli, Email: daniele.morbidelli@unibo.it.
Séverine Rigot, Email: severine.rigot@univ-cotedazur.fr.
References
- 1.Antonelli G, Le Donne E. Polynomial and horizontally polynomial functions on Lie groups. Ann. Mat. Pura Appl. (4) 2022;201(5):2063–2100. doi: 10.1007/s10231-022-01192-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bellettini C, Le Donne E. Regularity of sets with constant horizontal normal in the Engel group. Commun. Anal. Geom. 2013;21(3):469–507. doi: 10.4310/CAG.2013.v21.n3.a1. [DOI] [Google Scholar]
- 3.Bellettini C, Le Donne E. Sets with constant normal in Carnot groups: properties and examples. Comment. Math. Helv. 2021;96(1):149–198. doi: 10.4171/CMH/510. [DOI] [Google Scholar]
- 4.Cheeger J, Kleiner B, Naor A. Compression bounds for Lipschitz maps from the Heisenberg group to . Acta Math. 2011;207(2):291–373. doi: 10.1007/s11511-012-0071-9. [DOI] [Google Scholar]
- 5.Cheeger J, Kleiner B. Metric differentiation, monotonicity and maps to . Invent. Math. 2010;182(2):335–370. doi: 10.1007/s00222-010-0264-9. [DOI] [Google Scholar]
- 6.Csató G , Dacorogna B, Kneuss O. The Pullback Equation for Differential Forms, Progress in Nonlinear Differential Equations and their Applications. New York: Birkhäuser/Springer; 2012. [Google Scholar]
- 7.Dacorogna B, Kneuss O. Divisibility in Grassmann algebra. Linear Multilinear Algebra. 2011;59(11):1201–1220. doi: 10.1080/03081087.2010.495389. [DOI] [Google Scholar]
- 8.Fässler K, Orponen T, Rigot S. Semmes surfaces and intrinsic Lipschitz graphs in the Heisenberg group. Trans. Am. Math. Soc. 2020;373(8):5957–5996. doi: 10.1090/tran/8146. [DOI] [Google Scholar]
- 9.Kaplan A. Fundamental solutions for a class of hypoelliptic PDE generated by composition of quadratic forms. Trans. Am. Math. Soc. 1980;258(1):147–153. doi: 10.1090/S0002-9947-1980-0554324-X. [DOI] [Google Scholar]
- 10.Le Donne E. A primer on Carnot groups: homogenous groups, Carnot-Carathéodory spaces, and regularity of their isometries. Anal. Geom. Metr. Spaces. 2017;5(1):116–137. doi: 10.1515/agms-2017-0007. [DOI] [Google Scholar]
- 11.Magnin L. On -null Lie algebras. Rev. Un. Mat. Argentina. 2012;53(2):37–58. [Google Scholar]
- 12.Morbidelli D. On the inner cone property for convex sets in two-step Carnot groups, with applications to monotone sets. Publ. Mat. 2020;64(2):391–421. doi: 10.5565/PUBLMAT6422002. [DOI] [Google Scholar]
- 13.Morbidelli, D., Rigot, S.: Precisely monotone sets in step-2 rank-3 Carnot algebras. Math. Z. 303(3) (2023), 54
- 14.Naor A, Young R. Vertical perimeter versus horizontal perimeter. Ann. Math. (2) 2018;188(1):171–279. doi: 10.4007/annals.2018.188.1.4. [DOI] [Google Scholar]
- 15.Rickly M. First-order regularity of convex functions on Carnot groups. J. Geom. Anal. 2006;16(4):679–702. doi: 10.1007/BF02922136. [DOI] [Google Scholar]
- 16.Serra Cassano, F.: Some Topics of Geometric Measure Theory in Carnot Groups, Geometry, Analysis and Dynamics on Sub-Riemannian Manifolds, vol. 1, pp. 1–121. EMS Ser. Lect. Math., Eur. Math. Soc., Zürich (2016)
- 17.Varopoulos NT, Saloff-Coste L, Coulhon T. Analysis and Geometry on Groups, Cambridge Tracts in Mathematics. Cambridge: Cambridge University Press; 1992. [Google Scholar]
- 18.Young R. Area-minimizing ruled graphs and the Bernstein problem in the Heisenberg group. Calc. Var. Partial Differ. Equ. 2022;61(4):142,32. doi: 10.1007/s00526-022-02264-x. [DOI] [Google Scholar]