On uniqueness of submaximally symmetric
parabolic geometries
Abstract.
Among (regular, normal) parabolic geometries of type , there is a locally unique maximally symmetric structure and it has symmetry dimension . The symmetry gap problem concerns the determination of the next realizable (submaximal) symmetry dimension. When is a complex or split-real simple Lie group of rank at least three or when , we establish a local uniqueness result for submaximally symmetric structures of type .
Key words and phrases:
Submaximal symmetry, parabolic geometry, harmonic curvature, Tanaka theory2010 Mathematics Subject Classification:
Primary 58J70, Secondary 53B99, 22E46, 17B701. Introduction
For a given (local) differential geometric structure, our interest here will be on the dimension of its Lie algebra of infinitesimal symmetries. Many types of structures (e.g. Riemannian metrics on manifolds of fixed dimension) admit a finite maximal symmetry dimension , and there is broad interest to (locally) classify all such maximally symmetric structures. Letting denote the next possible realizable (submaximal) symmetry dimension, there is often a significant gap arising between and . The symmetry gap problem refers to the determination of and in doing so the task of exhibiting (local) models realizing this submaximal symmetry dimension. With this goal in mind, one can make a detailed case-by-case study of the PDE determining the symmetry vector fields for a given structure, but in many situations such a direct investigation using analytic tools becomes cumbersome. Our approach here is to draw upon strong algebraic tools that are present for an important broad class of structures that admit an equivalent reformulation as Cartan geometries.
Parabolic geometries [6] admit such a reformulation – they are a diverse and interesting class of geometries whose underlying structures include conformal, projective, CR, 2nd order ODE systems, and many classes of generic distributions, e.g. -distributions. Their description as parabolic geometries (see §2.1) gives a solution to the equivalence problem for such structures in the sense of Élie Cartan. Briefly, such a geometric structure on (henceforth, always assumed connected) admits a categorically equivalent description as a (regular, normal) Cartan geometry of type , where is a semisimple Lie group and is a parabolic subgroup. (For more details on the passage from to the “upstairs” Cartan perspective, we refer the reader to [6, 4].) The Cartan connection provides a canonical coframing on and its symmetry algebra is isomorphic to the symmetry algebra of the underlying structure on . We have for such structures, and there is a (locally) unique maximally symmetric model, namely the flat model of type , where is the Maurer–Cartan form of . Any Cartan geometry of type can be viewed as a curved version of this flat model, and our starting point is to take the (normalized) Cartan geometry as the basic input to the problem.
Substantial progress was made on the symmetry gap problem for parabolic geometries in [16]. In that joint work with Kruglikov, we proved that for any in terms of a universal (algebraically-defined) upper bound . Moreover, when is complex or split-real simple:
-
(i)
can be efficiently calculated via Dynkin diagram combinatorics, and
-
(ii)
almost always, with some exceptions when .
We uniformly proved by exhibiting a particular homogeneous structure, encoded “Cartan-theoretically” by what we refer to here as an algebraic model (see §2.4). We remark that for more general real forms, the determination of and sharpness of is still largely open, although numerous interesting cases have been resolved – see for example [9, 13, 14, 15].
Not addressed in [16] was the broader classification problem for submaximally symmetric structures, and our goal in this article is to resolve this. In order to formulate our main result, we briefly recall some notions here. (Precise definitions will be given later.) For any (regular, normal) parabolic geometry, there is a fundamental quantity called harmonic curvature , which completely obstructs local equivalence to the flat model. The codomain of is a filtrand of a certain Lie algebra homology group, which is a completely reducible -representation, so only the action on it by the (reductive) Levi factor is relevant. Consider a -irrep . We say that is of type if it is of type and , and let be the maximal symmetry dimension among regular, normal parabolic geometries of type with . We can now formulate our main result111See §3.1 for our subscript notation for parabolics in the complex or split-real setting.:
Theorem 1.1.
Let be a complex or split-real simple Lie group, a parabolic subgroup, and its Levi factor. Let be a regular, normal parabolic geometry of type , where is a -irrep. Suppose that , and or . Then the geometry is locally homogeneous about any with . The corresponding algebraic model with is (up to -equivalences , ):
-
(1)
complex case: unique.
-
(2)
split-real case: one of at most two possibilities. Uniqueness holds if and only if there exists such that , where is a lowest weight vector.
Our result is constructive (see §3.4): over , the distinguished algebraic model encoding the corresponding submaximally symmetric geometry is what we refer to here as the canonical curved model of type , which has curvature (interpreted as a harmonic 2-cochain). The Lie algebra arises as a filtered deformation of a graded subalgebra (see §2.2), namely as vector subspaces, but with bracket , where is the bracket on (restricted to ). This is the same abstract model used in [16]. In the split-real setting, the second possibility is with .
For fixed , Theorem 1.1 can be used to deduce the analogous classification of all submaximally symmetric structures, i.e. is not constrained to a specific . See §4 for some examples.
We now give numerous examples illustrating that one cannot in general weaken the hypotheses of Theorem 1.1 and expect such a uniform conclusion.
Non-uniqueness over can occur if we do not require to be -irreducible:
Example 1.2.
A Legendrian contact geometry (over or ) is a contact manifold with contact distribution endowed with a splitting into Legendrian subbundles. (Second order ODE is the case.) Such a structure underlies a parabolic geometry of type , , and for we have an -irreducible decomposition . From [16, Table 11], we have . The corresponding canonical curved models are inequivalent.
If one weakens the complex / split-real assumptions, varying phenomena can occur:
Example 1.3.
Real hypersurfaces in having positive-definite Levi form yield 5D (integrable) CR geometries, which are specific real forms of complex Legendrian contact geometries (Example 1.2) when . They underlie regular, normal parabolic geometries of type , where (not split-real), and the complexification of would take values only in . We have , while it is known that , with infinitely many inequivalent submaximally symmetric models; see [8, Table 8 (D.7)]. In the Levi-indefinite case, (again, not split-real), , and there is a unique local model realizing ; see [8, Table 7 (N.8)].
Now suppose . In contrast to local uniqueness in the case (both over and , see §3.2), there is a 1-parameter family of submaximally symmetric models in the case:
Example 1.4.
A -geometry is a 5-manifold equipped with a rank 2 distribution having generic growth under the Lie bracket, i.e. and . Locally, any such admits a Monge normal form: there exist local coordinates and a function with such that is spanned by the vector fields
Such a structure underlies a parabolic geometry of type , so , with realizing maximal symmetry. Here, in either the complex or real case. Over , a well-known list of submaximally symmetric models is given by (for ) and .
Other rank two cases include 3-dimensional conformal geometry, i.e. type , and the contact geometry of scalar 3rd order ODE, i.e. type . Submaximally symmetric models are non-unique for both – in the former case see the classification in [12], while in the latter case they are given by , where is constant. The rank two case of 2nd order ODE exhibits several exceptional phenomena:
Example 1.5.
Scalar 2nd order ODE (up to point transformations) underlie geometries, for which and . Locally, one has a 3-manifold with coordinates and split contact distribution on with
(1.1) |
We have , and corresponds to arbitrary rescalings along and . We have , with each being a 1-dim -irrep. The components of along and correspond to the well-known Tresse relative invariants and . For , we refer to [16, eqn (5.8)] and replace there with . Two submaximally symmetric models are:
-
(i)
: symmetries are . We have and both nonvanishing. Thus, is not concentrated in a single irreducible component.
-
(ii)
: symmetries are . The evaluation map is surjective except along the singular set , so neighbourhoods of and (endowed with restricted geometric structures) are not locally equivalent. We have and , so vanishes along .
A priori, we cannot exclude the possibility of similar limiting singular behavior as in Example 1.5(ii) for submaximally symmetric structures occurring in geometries with , so we always work near a point where is nonvanishing. Constraining ourselves to the hypotheses of Theorem 1.1 ultimately leads to a classification problem for homogeneous structures.
We note that Cartan reduction is a general method for classifying (homogeneous) geometric structures. (See for example [7] for a recent application.) While this is a powerful, systematic method, it is typically applied on a case-by-case basis, and for any given structure it takes a substantial amount of effort to set up the correct structure equations (via the Cartan equivalence method, for instance). Moreover, its implementation can be extremely cumbersome to do by-hand (often being done in a symbolic algebra system such as Maple or Mathematica), and normalizations generally proceed in an ad-hoc manner. In principle, it can be used to analyze submaximally symmetric structures, but in practice it is not a feasible method to arrive at the claimed generality of Theorem 1.1. Our approach will be to proceed in a uniform manner by taking the Cartan-geometric viewpoint as the basic input, and make efficient use of representation theory.
Let us briefly outline our article. In §2, we recall relevant background from parabolic geometries and our earlier work on symmetry gaps, and formulate the notion of an algebraic model encoding any homogeneous parabolic geometry. In §3, we recall Kostant’s theorem, define the canonical curved model, and formulate the algebraic model classification problem (Problem 3.3). We then solve it, first for geometries (§3.2), and then the general case (§3.4). We conclude in §4 with concrete examples of submaximally symmetric structures, which are asserted to be unique (over ) from Theorem 1.1.
Conventions: The base manifold is always assumed to be connected. We work in the smooth and holomorphic categories when referring to real and complex geometries, respectively. For simple roots, we use the same ordering as in LiE [19].
2. Parabolic geometries and algebraic models
We begin by reviewing background from parabolic geometries and symmetry gaps – see [6, 16] for more details.
2.1. Parabolic geometries
Let be a real or complex semisimple Lie group, a parabolic subgroup, and be their Lie algebras. Then admits a natural -invariant (decreasing) filtration (we put for , for ), is the nilradical of , and for all . There always exists grading element whose -eigenvalues (degrees) and eigenspaces () endow with the structure of a graded Lie algebra compatible with the filtration, i.e. and . The associated-graded Lie algebra is defined by . Given as above, we identify as -modules, and if , we denote by the projection to its leading part. We have (centre of ), , and the Killing form on identifies as -modules. Finally, letting be the Levi subgroup (with Lie algebra ), and , we have .
A parabolic geometry is a Cartan geometry of type , i.e. a (right) principal -bundle with a Cartan connection :
-
(i)
is a linear isomorphism ;
-
(ii)
is -equivariant: , ;
-
(iii)
, , where is the fundamental vertical vector field corresponding to .
The curvature of is (which is -equivariant and horizontal, i.e. ), or equivalently we have the curvature function given by . The geometry is flat if , which characterizes local equivalence to the flat model , where is the (left-invariant) Maurer–Cartan form on . Via the Killing form, the codomain of identifies (as a -module) with . These are 2-chains in the complex with the Lie algebra homology differential. We say that is normal if and it is regular if for any . Equivalently, if we naturally extend the filtration on to a filtration on , then we have . This is the subspace of on which acts with positive eigenvalues (degrees). There is a well-known equivalence of categories between regular, normal parabolic geometries and underlying geometric structures on (see [6] for details).
For any regular, normal parabolic geometry, a key invariant is its harmonic curvature , given by , and this -equivariant function is a complete obstruction to flatness. Moreover, is a completely reducible -representation, i.e. -acts trivially. As -modules, , and yields a complex with respect to the standard Lie algebra cohomology differential , for which we have the (-invariant) algebraic Hodge decomposition:
(2.1) |
where is the (-equivariant) algebraic Laplacian, with . Then as -modules, which may be efficiently computed via Kostant’s theorem (§3.1). By regularity, has image in the subspace on which acts with positive eigenvalues. This corresponds to some -submodule under the above identification.
Finally, by [6, Thm.3.1.12], if has lowest non-trivial degree , then its leading part is harmonic and coincides with the degree component of . In particular, being a complete obstruction to flatness follows from this.
2.2. Symmetry and Tanaka prolongation
Two (regular, normal) parabolic geometries of type are equivalent if there is a principal bundle isomorphism that pulls back one Cartan connection to the other, and an automorphism is a self-equivalence. A Cartan geometry is (locally) homogeneous if there is a Lie group acting by (local) automorphisms whose projection to yields a (locally) transitive action on . Infinitesimally, the symmetry algebra is
(2.2) |
where are the -invariant vector fields on .
Let us now summarize how to equivalently view in a more algebraic manner [2, 5, 16]. Fix any . Then restricts to a linear injection on . Letting , the Lie bracket on transfers to the bracket on given by
(2.3) |
The -invariant filtration on induces a filtration on via . By regularity, , so , and becomes a filtered Lie algebra (generally not a Lie subalgebra of ). By regularity, the associated-graded , defined by is identified as a graded subalgebra of (via ). The filtrand satisfies the important algebraic condition , which implies . Since acts trivially on , then always, so , i.e. is contained in the annihilator .
Now define the following (extrinsic) Tanaka prolongation algebra as in [16]:
Definition 2.1 (Extrinsic Tanaka prolongation).
Let be a Lie subalgebra. Extend this to a -graded Lie subalgebra by defining and for . Denote by . When lies in some -representation, we write .
The constraint propagates via Tanaka prolongation to the higher levels. More precisely, the following important inclusion holds:
(2.4) |
Otherwise put, the symmetry algebra is a constrained filtered sub-deformation of , i.e.
-
(i)
is a filtered deformation of the graded subspace , and
-
(ii)
is constrained: e.g. it is a filtered subspace of and satisfies (2.3) for some .
The inclusion (2.4) was established in [16, Thm.2.4.6] on the open dense set of so-called regular points, i.e. those on which are locally constant functions , and was generalized to all points in [17, Thm.3.3]. If the given geometry is not flat, then at some , so by the regularity assumption on , and hence . Thus, the flat model is locally the unique maximally symmetric geometry. Defining
(2.5) | ||||
(2.6) |
equation (2.4) immediately implies
(2.7) |
A (regular, normal) geometry with is submaximally symmetric. A priori, it should not be assumed that these are locally homogeneous, particularly if .
Definition 2.2.
Let be a -invariant subset. Let
Lemma 2.3.
For regular, normal parabolic geometries of type , and a -invariant subset, suppose that . Then any with and is locally homogeneous near any with .
2.3. Homogeneous parabolic geometries
Let be homogeneous with respect to the Lie group . Fix , and let be the stabilizer of . Given any , we have for some Lie group homomorphism . This defines a right -action on via and we let be the collection of all -orbits . Letting and be the Lie algebras of and respectively, we have [6, Prop.1.5.15]:
-
(1)
is equivalent to the associated bundle .
-
(2)
Any -invariant Cartan connection of type is completely determined by the following:
Definition 2.4.
An algebraic Cartan connection of type on is a linear map with:
-
(C1)
, where is the differential of .
-
(C2)
, . Infinitesimally:
(C2’) where and are the Lie brackets on and respectively. If is connected, then and (C2’) are equivalent.
-
(C3)
induces a vector space isomorphism .
Indeed, given as above, we obtain by factoring given by
The basepoint change leaves unchanged, but a fibrewise change induces .
Define , so by (C2’). The curvature of corresponds to given by . The notions of regularity and normality of are immediately specialized to this algebraic setting, as is the quotient object .
2.4. Algebraic models
Note that and (C2’) forces to be an ideal in . The -action on can always assumed to be infinitesimally effective, i.e. does not contain any non-trivial ideals of (hence, ). (Otherwise, we may without loss of generality quotient both and by the corresponding normal subgroup.) Consequently, we assume that is injective and identify with its image in . This motivates the following definition.
Definition 2.5.
An algebraic model is a Lie algebra such that:
-
(M1)
is a vector subspace with inherited filtration such that satisfies .
-
(M2)
inserts trivially into , so identify with .
-
(M3)
is regular and normal, i.e. .
The result below immediately follows from the general theory recalled in §2.2, but it is instructive to give proofs directly following from Definition 2.5 above.
Proposition 2.6.
Let be an algebraic model. Then
-
(1)
is a filtered Lie algebra. (In general, is not a Lie subalgebra of .)
-
(2)
, i.e. , and .
-
(3)
, where .
Proof.
For (1), if and , then by regularity of , i.e. (M3), so (1) implies . But is a Lie algebra, so and . For (2), we use the Jacobi identity and the fact that vanishes under -insertions by (M2). Namely, let and . Then
Finally, we prove (3). Since is -equivariant and , then , so (2) implies , which factors to by complete reducibility of . Letting , this means . By regularity, is a graded Lie subalgebra, so for any , . Inductively, we have for all . ∎
Importantly, we note that the set of algebraic models of type :
-
•
admits a -action via for any . All algebraic models in the same -orbit are to be regarded as equivalent, so we must always account for this redundancy.
-
•
is partially ordered: declare that if and only if is a Lie subalgebra of . We will focus on maximal elements . (We view non-maximal elements as non-optimal descriptions of the same geometric structure.)
Remark 2.7.
Conversely, by [16, Lemma 4.1.4], to each algebraic model , there exists a locally homogeneous geometry of type with containing a subalgebra isomorphic to . If is maximal with respect to the partial order defined above, then .
Since , we may view as a graph over . Namely, choosing some graded subspace so that (in fact, since by hypothesis), we can write
(2.8) |
for some unique linear map of positive degree, i.e. if , then . We refer to as the deformation map and as the tail of .
Lemma 2.8.
Let with and being -invariant. Then , i.e. .
Proof.
Given , we have and . By (M2), . But and since has positive degree, so by uniqueness of , we have . ∎
3. The canonical curved model and local uniqueness
We focus on proving Theorem 1.1 in the complex case. The arguments in the split-real case are almost exactly the same, and potentially differ only in the final step of §3.4. This is described at the end of §3.
3.1. Kostant’s theorem and the canonical curved model
Let be a complex semisimple Lie group, a Cartan subalgebra with , the associated root system, the root space for , the positive roots relative to a choice of simple root system , and its dual basis, i.e. . If is an -invariant subspace, we write , and . A parabolic subgroup with Lie algebra is encoded by a subset , with associated grading element , and relative to it. On the Dynkin diagram of , we put crosses at nodes corresponding to , and refer to using subscripts, e.g. if , then , etc. (In our convention, the Borel subalgebra has crosses on all Dynkin diagram nodes.)
The Killing form of induces a non-degenerate pairing on . Letting be the coroot of , we have the Cartan matrix with . The fundamental weights are defined by , and these satisfy . Corresponding to is the simple reflection defined by , and the Weyl group is the group generated by all simple reflections.
Kostant’s theorem [11] yields an efficient -module description of : it is the direct sum of -irreps , each of multiplicity one, and having lowest weight , where:
-
(1)
is the highest weight of a simple ideal of ;
-
(2)
is a length 2 word of the Hasse diagram . Concretely for our purposes here, this is equivalent to: and either: (i) , or (ii) .
-
(3)
refers to the affine action of on weights: letting , we have
(3.1)
Via the -module isomorphism from (2.1), a representative lowest weight vector is given in terms of root vectors by:
(3.2) |
To interpret as a 2-cochain, we identify and as the dual elements and via the Killing form. (Here, we fix root vectors yielding a basis on .)
For , define and . By [16, Prop.3.1.1], we have
(3.3) |
with equality precisely when the projectivizations and lie in the same -orbit in . Hence, . Concerning realizability of this upper bound, we have:
Definition 3.1.
Use notations as above with simple. Suppose satisfies and . The canonical curved model of type is the algebraic model given by defining as a vector subspace of , equipped with the filtration inherited from , and deformed bracket .
By [16, Lemma 4.1.1], is indeed a Lie algebra, and clearly . The filtration on is inherited: . Since , then is clearly normal, and guarantees regularity. Thus, is an algebraic model, which is clearly maximal with respect to the aforementioned partial order.
Proposition 3.2.
Use notations as above with simple. Suppose that , and exclude and . Then for any . For with , the canonical curved model of type exists, and so . Moreover, any regular, normal parabolic geometry of type with is locally homogeneous about any with .
Under the hypotheses of Proposition 3.2, our problem of classifying submaximally symmetric models relative to becomes that of classifying algebraic models with and . From Proposition 2.6, . The equality forces with for some . Using the induced -action on , we may henceforth assume that , and so satisfies
(3.4) |
In summary, to establish Theorem 1.1, it suffices to answer the following question:
Problem 3.3.
When or , classify algebraic models with satisfying (3.4), up to the action by the residual subgroup .)
Over , we will show that the canonical curved model is the unique solution to Problem 3.3.
3.2. The case
We first answer Problem 3.3 in the case. For , recall:
(3.5) |
Let be the dual basis to the simple roots . The highest weight is . The root diagram is given in Figure 1. Let be a root vector for the root .
We have , which induces the grading . Moreover,
(3.15) |
Let us now classify algebraic models with . Let with . Since for all , we use the -action to normalize to .
Defining , let be the associated deformation map (of positive degree). The decomposition is -invariant, so by Lemma 2.8, i.e. is a sum of weight vectors for weights that are multiples of . The weights of and agree, and they are both:
(3.16) |
Since has coefficients with respect to of opposite sign, there is no sum of two weights in (3.16) that is: (i) a multiple of , and (ii) has positive degree. Thus, , and as filtered subspaces of .
Now consider curvature . Since by Proposition 2.6, then we are interested in weights (of 2-cochains) that are multiples of :
(3.17) |
(We have since regularity and the final statement in §2.1 imply .) Recall that and note that . Then , and so . However, and has integer coefficients in the simple root basis, so the only possibility is . Since is a -irrep, then must be a nonzero multiple of . Use to rescale over so that , so we obtain the canonical curved model.
Over , we may rescale to . Let us study the action by more concretely. Let be the standard basis of , and the dual basis. Then as -modules and we can identify
We regard as a multiple of . Hence, acts as , so taking induces . Again, we obtain the canonical curved model.
The underlying structures for regular, normal parabolic geometries of type are called -contact geometries. See [18] for and a coordinate realization of a submaximally symmetric structure given in [18, Table 8]. By uniqueness proved above, this corresponds to the canonical curved model. We have shown:
Proposition 3.4.
There is a locally unique (complex or real) -contact geometry that is submaximally symmetric () about any point where harmonic curvature is nonvanishing.
3.3. Preparation for the general case and the twistor simplification
The case for Problem 3.3 is treated in a similar spirit to the case, but will require some further preparations. We will need more details about and . First, observe that for , we have by (3.1) and (3.2). If , then (3.1) becomes:
(3.18) |
Second, from [16, Thm. 3.3.3], is the Tanaka prolongation of:
(3.19) |
This is the direct sum of and root spaces for the roots
(3.20) |
where is a secondary grading with respect to the set
(3.21) |
In [16], the weight was encoded on a Dynkin diagram by inscribing over corresponding nodes the coefficients of with respect to . The set corresponds to uncrossed nodes with a nonzero coefficient.
Example 3.5.
Consider . Here, , , , and . Applying Kostant’s theorem, we find that
with , and so . (Also, .) See [16, §3.3] for more examples.
According to [3], any parabolic geometry can be lifted to a correspondence space, and conversely a parabolic geometry may be descended to a twistor space if a suitable curvature condition is satisfied. (The latter amounts to viewing the given geometry of type as a geometry of type , where .) These are categorical constructions, so symmetries are naturally mapped to symmetries. We will not recall here the general theory developed in [3], but only summarize various results from [16, §3.5] in order to emphasize a “twistor simplification” (3.22) relevant for our purposes. The main reason for doing so is to assert (3.28), which facilitates the classification of filtered sub-deformations of in §3.4.
Under the assumption , one may always descend to a minimal twistor space. Concretely, if , where , then [16, Prop.3.5.1 & Cor.3.5.2] indicates that we may instead view a given geometry as a geometry, where
(3.22) |
In [16, Thm.3.5.4], we showed that the Lie algebra structure of is unchanged with respect to the above change of parabolics, in spite of the grading change, cf. [16, Example 3.5.5].
Normality of the geometry is preserved in passing to a correspondence or twistor space, but a priori regularity is not.
Example 3.6.
For regular, normal geometries of type , we have and . Then yields , so . Viewed on the correspondence space as a geometry, which has grading element , the corresponding harmonic curvature would take values in a module with degree , i.e. regularity is not preserved.
Despite regularity not being preserved when passing upwards to a correspondence space, let us consider the passage downwards to the minimal twistor space. In the simple setting, preservation of regularity upon such descent can be observed a posteriori through the tables compiled in [16, Appendix C]. We now give a uniform proof of this:
Lemma 3.7.
Let be a complex simple Lie algebra of with highest weight , a parabolic subalgebra. Fix and . Define , where the parabolic subalgebra is defined by Let be the grading elements corresponding to respectively. If we have , then .
Proof.
From (3.18), we have for any . (Recall that all coefficients of in the simple root basis are positive.) We have . Suppose that (i) (hence, ), or (ii) and . In either case, , where , so since .
From §3.1, it remains to consider the case and . Then
(3.23) |
From (3.18), note that
(3.24) |
If , then . As above, and we are done. So let us suppose that . We can examine all such possibilities from knowledge of the well-known highest roots of simple Lie algebras:
(3.27) |
If is not type A or C, then from (3.27), we have , and it is well-known that yields a contact grading on . So , from (3.24), and follows from (3.23). For the type A and C cases, we show that independent of the hypothesis on :
-
(1)
Type C: We have , , and . Since , then from (3.18), we have .
-
(2)
Type A: We have and using a Dynkin diagram symmetry, we may assume , so . Since , we have .
∎
3.4. Proof of the main theorem
Let us turn now to the proof of Theorem 1.1.
Lemma 3.8.
Let be a complex simple Lie algebra with and its highest root. Let such that satisfies . Then:
-
(MU1)
has coefficients of opposite sign. More precisely, , , and either or .
-
(MU2)
with for all with .
Proof.
From (3.18), . Since is simple, then all coefficients of with respect to the basis of simple roots are strictly positive. Since , then for all . At least one of or must be positive, since by hypothesis.
Fix any , and , so by the first claim, is not a multiple of . Thus, and are distinct hyperplanes in . Their sum must be , while is a hyperplane in . Since is finite, the finite union has non-empty complement in (being the finite intersection of open sets ). Picking in this (open) complement completes the proof. ∎
Assume the hypotheses of Lemma 3.8. From the previous subsections, we have reduced our submaximal symmetry classification problem to studying algebraic models with
(3.29) |
where is given by (3.2). We will classify these up to the action of . Moreover, we may assume the twistor simplification, which implies that , where we have moved to the grading associated with the larger parabolic subgroup indicated in (3.22).
Step 1: Using the -action, normalize so that .
As in Lemma 3.8, fix . Let with leading part , so , where . If , let for some minimal . Let us normalize via the -action. Letting , we have:
(3.30) |
where the dots indicate terms of degree . Fixing root vectors , we have . By (MU2) in Lemma 3.8, , so defining , we have . Redefining as and as , the latter has with leading part of degree . Inductively, we may normalize , and so . Since for all , the -part of the structure group has been completely reduced.
Step 2: Observe that
Fix any , so . Write with . By (M2) from Definition 2.5, we have:
(3.31) |
where the twistor simplification was invoked for the last equality. Since (MU2) implies for all , then necessarily . Thus, .
Step 3: Show that as subspaces of .
Recall from (3.21) and the secondary grading . From (3.19), we have where
(3.32) |
and is a 1-dimensional complement to inside . Both and are -invariant, so in particular they are invariant under . Defining the associated deformation map , Lemma 2.8 implies that , , so lies in the direct sum of weight spaces of for weights that are multiples of .
Note , so let and examine . From (3.19), we have , and the weights of are of the form , where . These all have non-negative coefficients in the simple root basis. By (MU1), these weights cannot be multiples of . Hence, , i.e. . (This argument is very similar to the case from §3.2.)
For our Step 3 claim, it suffices to consider and show that , . First recall that as in Lemma 3.8. We claim that we may assume
(3.33) |
Via the twistor simplification, we have either: (a) , hence ; or (b) with , hence , so either or . Since , then swap if necessary to assume that . Since , (3.18) implies the rest of (3.33).
Since has positive degree, then , so let us consider a weight for corresponding to a possible term in . Using , we have two cases:
-
(1)
: Since , then , while for any , we have . By (3.33), cannot be a multiple of .
-
(2)
: Since , fix any and note that (by standard properties of Cartan matrices). By definition of , we have . Recalling that for all , (3.18) implies:
(3.34) Hence, , , i.e. every is not connected in the Dynkin diagram to either or . Since is simple (with ), its Dynkin diagram is connected, so this is a contradiction, i.e. this case is vacuous.
We conclude that , , and hence . Thus, as subspaces of .
Step 4: Study curvature
By (M3) and Proposition 2.6, we have and . Since , then is valued in the direct sum of weight spaces of for weights with and . For the same reasons there (regularity and the final statement in §2.1), we again have . Let us show that . Write the highest weight of as , where for all since is simple. Since is the lowest root of , then . Thus, for any ,
(3.35) |
where the last equality follows from (3.18). Since , then follows. Thus, and so has weight . The multiplicity of (lowest weight) is the same as that occurring in the -irrep , i.e. multiplicity one, by Kostant’s theorem. Under the identification with harmonic 2-cochains, must be a nonzero multiple of . Using , we may do a complex rescaling to arrange . Thus, we have obtained the canonical curved model.
Working with split-real geometries, we similarly arrive at being a nonzero multiple of using almost exactly the same arguments as in the complex case. The only part that differs concerns the use of [16, Prop.3.1.1] to assert (3.3) and the subsequent statement characterizing equality there. A key ingredient for that Proposition is that is the unique closed -orbit in , and this orbit is of minimal dimension. This is a well-known result in the complex setting, and the result remains true in the split-real setting – see [20, Cor.1]. All other arguments in [16, Prop.3.1.1] and this section are exactly the same to arrive to being a nonzero multiple of .
Finally, a real scaling using normalizes . The algebraic models are -equivalent if and only if there exists such that . The proof of Theorem 1.1 is complete.
4. Examples
In this final section, we apply Theorem 1.1 and give concrete examples of submaximally symmetric parabolic geometries, expressed as underlying geometric structures. Implicit here are known equivalences of categories, in particular the parabolic geometry types associated to given structures. We do not provide details here, but instead refer the reader to [6].
We will use the following notation. Let denote the standard square matrix (of size to be specified) with a 1 in the -position and elsewhere. We continue to use for the highest weight of , and for a lowest weight vector of a -irreducible submodule of , obtained via Kostant’s theorem (§3.1).
4.1. Projective structures
On a manifold , two torsion-free affine connections are equivalent if and only if they admit the same unparametrized geodesics, and an equivalence class is called a projective structure. These well-known structures underlie geometries of type , for which , and harmonic curvature corresponds to the projective Weyl curvature. Here, (for or ) realized as matrices of the form , where . In [16], we found that for , realized in particular by the Egorov projective structure [10], [16, (5.11)]. We can now assert:
Corollary 4.1.
Let , and a submaximally symmetric projective structure with non-vanishing projective Weyl curvature at . Then about , is locally equivalent to the Egorov projective structure (in either the real or complex settings).
Proof.
Remark 4.2.
Over , some attention should be given to the choice of Lie group . Choosing with as above, the induced -action on is , where , so , which is always positive when is odd. In these cases, one is in fact working with oriented manifolds. In the unoriented setting, one could work with (i.e. modulo its centre , used as in [6, Prop.4.1.5]) or use when is odd.
4.2. 2nd order ODE systems
Any system , of 2nd order ODE in dependent variables (viewed up to point transformations) admits an equivalent description as a (regular, normal) parabolic geometry of type . (In [6, §4.4.3], these are formulated as generalized path geometries. When (or ), these can all be locally realized as 2nd order ODE systems, while for , we additionally have the constraint that vanishes in degree .) We have , locally uniquely realized by the trivial ODE . Here,
(4.1) |
and harmonic curvature decomposes into two components: Fels curvature (degree +3) and Fels torsion (degree +2). Referring to [17, §5.3 and §5.4], we have (using and and notation , from §3.1):
-
•
(“Segré branch”, i.e. vanishing Fels torsion): has , and , realized in the Segré branch by:
(4.2) -
•
(“projective branch”, i.e. vanishing Fels curvature): has and , realized in the projective branch by the geodesic equations of the Egorov projective structure:
(4.3) Using the point transformation given in [1], a simpler alternative model to (4.3) is
(4.4) (All ODE in the projective branch are geodesic equations for some projective structure, and Theorem 4.1 asserts the classification of submaximal symmetry models in this branch.)
Uniqueness of the submaximally symmetric ODE (4.2) and (4.4) was recently asserted in [1, Theorems 2 & 3] without proof. Applying our Theorem 1.1, we obtain:
Corollary 4.3.
Let . Over or , suppose that a given 2nd order ODE system , is submaximally symmetric, i.e. it has point symmetry algebra of dimension . Then the system has vanishing Fels torsion everywhere, and about any point where Fels curvature is non-vanishing, the system is locally point equivalent to (4.2).
Within the projective branch (i.e. vanishing Fels curvature), about any point where Fels torsion is non-vanishing, any submaximally symmetric system (realizing ) is locally point equivalent to (4.4).
Proof.
Note that we have and . Since , then local homogeneity follows from Lemma 2.3. Write for and , where are the -irreducible submodules of corresponding to Fels curvature and Fels torsion respectively. We have , which has maximal dimension when . By (2.4) and the symmetry dimension being , the Fels torsion must vanish everywhere.
We remark that in [16], we used instead of . This small change does not affect and , but the notion of point equivalence is slightly restricted with the former, as we now explain. Consider with . The ODE structure is modelled on , which is split into the direct sum of (corresponding to the line field spanned by the total derivative ) and (corresponding to ). On , induces:
(4.5) |
But then . When is even, the signs of and are aligned, and the point transformation would not be an admissible equivalence.
If we consider , then for , setting for except yields . When , we have , and no exists with . In this case, , would be inequivalent submaximally symmetric models.
4.3. Conformal structures
Given a smooth manifold with and a metric of signature , we let , and refer to as a conformal structure. This admits an equivalent description as a parabolic geometry of type , where is the stabilizer of a null line in , so . Restrict now to . See [9] for in the Riemannian / Lorentzian cases, which are exceptional. In non-Riemannian / non-Lorentzian signatures, [16, §5.1] indicates , realized by , with given by the direct product of a flat Euclidean metric of signature and the -metric
(4.6) |
Restrict now to the split-real form, so . We view as matrices that are skew with respect to the anti-diagonal, and let consist of block diagonal matrices with blocks with and , and so the -action on is given by . In particular, any scalar product on is only positively rescaled, so any conformal structure and its “negative” are inequivalent. Together with Theorem 1.1, we deduce:
Corollary 4.4.
Let and . Suppose that a conformal structure of signature is submaximally symmetric, i.e. its conformal symmetry algebra has dimension . Then about any point where the Weyl curvature is non-vanishing, the structure is locally conformally equivalent to one of the two models or described above.
The split-signature assumption may likely be relaxed so that the same conclusion would hold in general non-Riemannian / non-Lorentzian signatures, but this would require a more careful investigation into related real forms, which is beyond our scope here. For the more subtle conformal Riemannian and Lorentzian cases, finding the complete local classification of submaximally symmetric models is an open problem. (See [9] for known models.)
4.4. Parabolic contact structures
Generalizing §3.2, parabolic contact structures of type (or “-contact structures”) are underlying structures for (regular, normal) geometries of types:
(4.7) |
As shown in [18], these structures all admit descriptions (possibly passing to a correspondence space) in terms of differential equations. The cases and are classical, and correspond to scalar 2nd order ODE (up to point transformations) and scalar 3rd order ODE (up to contact transformations). These are exceptions: they admit non-unique submaximally symmetric structures with and symmetries respectively. For all other cases, explicit submaximally symmetric structures (with respect to a given -irrep of ) were given in [18, §4.2]. Over , these are locally unique by Theorem 1.1.
Acknowledgements
We thank Boris Kruglikov and Henrik Winther for helpful discussions. The research leading to these results has received funding from the Norwegian Financial Mechanism 2014-2021 (project registration number 2019/34/H/ST1/00636), the Tromsø Research Foundation (project “Pure Mathematics in Norway”), the UiT Aurora project MASCOT, and this article/publication is based upon work from COST Action CaLISTA CA21109 supported by COST (European Cooperation in Science and Technology), https://www.cost.eu.
References
- [1] V.M. Boyko, O.V. Lokaziuk, R.O. Popovych, Admissible transformations and Lie symmetries of linear systems of second-order ordinary differential equations, arXiv:2105.05139 (2021).
- [2] A. Čap, Automorphism groups of parabolic geometries, Rend. Circ. Mat. Palermo (2) Suppl. 75 (2005), 233–239.
- [3] A. Čap, Correspondence spaces and twistor spaces for parabolic geometries, J. Reine Angew. Math 582 (2005), 143–172.
- [4] A. Čap, On canonical Cartan connections associated to filtered G-structures, arXiv:1707.05627 (2017).
- [5] A. Čap, K. Neusser, On automorphism groups of some types of generic distributions, Diff. Geom. Appl. 27 (2009), no. 6, 769–779.
- [6] A. Čap, J. Slovák, Parabolic Geometries I: Background and General Theory, Mathematical Surveys and Monographs, vol. 154, American Mathematical Society, 2009.
- [7] B. Doubrov, A. Medvedev, and D. The, Homogeneous integrable Legendrian contact structures in dimension five, J. Geom. Anal. 30 (2020), 3806–3858.
- [8] B. Doubrov, A. Medvedev, and D. The, Homogeneous Levi non-degenerate hypersurfaces in , Math. Z. 297 (2021), 669–709.
- [9] B. Doubrov and D. The, Maximally degenerate Weyl tensors in Riemannian and Lorentzian signatures, Diff. Geom. Appl. 34 (2014), 25–44.
- [10] P. Egorov, Collineations of projectively connected spaces (Russian), Dokl. Akad. Nauk SSSR 80 (1951), 709–712.
- [11] B. Kostant, Lie algebra cohomology and the generalized Borel–Weil theorem, Ann. of Math. (2) 74 (1961), 329–387.
- [12] G. I. Kruchkovich, Classification of three-dimensional Riemannian spaces according to groups of motions, Usp. Mat. Nauk (N.S.) 9, no.1(59) (1954), 3-40; correction: Usp. Mat. Nauk (N.S.) 9, no.3 (61) (1954), 285. (Russian)
- [13] B. Kruglikov, Submaximally symmetric CR-structures, J. Geom. Anal. 26 (2016), 3090–3097.
- [14] B. Kruglikov, V. Matveev, D. The, Submaximally symmetric c-projective structures, Internat. J. Math. 27 (2016), no. 3, 1650022.
- [15] B. Kruglikov, H. Winther, L. Zalabova, Submaximally symmetric almost quaternionic structures, Transform. Groups 23 (2018), 723–741.
- [16] B. Kruglikov, D. The, The gap phenomenon in parabolic geometries, J. Reine Angew. Math. 723 (2017), 153–215.
- [17] B. Kruglikov, D. The, Jet-determination of symmetries of parabolic geometries, Math. Ann. 371 (2018), 1575–1613.
- [18] D. The, Exceptionally simple PDE, Diff. Geom. Appl. 56 (2018), 13–41.
- [19] M.A.A. van Leeuwen, A. M. Cohen, B. Lisser, LiE, A package for Lie group computations (1992), http://wwwmathlabo.univ-poitiers.fr/maavl/LiE/.
- [20] H. Winther, Minimal Projective Orbits of Semi-simple Lie Groups, arXiv:2302.12138 (2023).