centertableaux
Intersections of Deligne–Lusztig varieties and Springer fibres
Abstract.
In this paper we prove a direct geometric relation between Deligne–Lusztig varieties and Springer fibres in type : For any rational unipotent element, the Springer fibre cuts out a unique component of a specific Deligne–Lusztig variety; moreover, this component forms an open dense subset of a component of the Springer fibre. This boils down to a map from the unipotent variety to the Weyl group, and combines several constructions with a combinatorial flavour (like Weyr normal forms, Robinson–Schensted correspondence, and Spaltenstein’s and Steinberg’s labellings); it also provides a geometric interpretation of a classical dimension formula of unipotent centralisers.
1. Introduction
Let be a connected reductive group over a finite field , and let be the geometric Frobenius endomorphism on . Fix an -stable Borel subgroup and an -stable maximal torus . There are two important classes of varieties lying in the flag variety , namely, Deligne–Lusztig varieties and Springer fibres.
Deligne–Lusztig varieties (see Definition 2.1) are parametrised by , and their -adic cohomology (with coefficients in suitable local systems) affords all the irreducible representations of . Meanwhile, Springer fibres (see Definition 2.4) are parametrised by the unipotent elements , and their -adic cohomology affords all the irreducible representations of . Since their births in the seminal works [DL76] and [Spr76], respectively, these two families of varieties play crucial roles in the study of representations of Lie type groups and Weyl groups. In this paper we give a study of their relations in the case of type .
In the remaining part of this paper we take to be , and let be the standard upper Borel subgroup and the diagonal maximal torus.
Indeed recently we found that, for a specific unipotent , the intersection appears in the study of smooth representations of the profinite group , where is a divisor of (see [Che20b]); this serves the initial motivation for our attention on the relations between these two varieties. On the other hand, on the level of representations there are already striking relations found between these two constructions (see for example [Lus90] and [BV21, 1.3.5]). In this paper, instead of representations we focus on the geometric relations. We remark that a similar theme has been considered over in [Tym06], in which case Hessenberg varieties and Schubert cells took the roles of Springer fibres and Deligne–Lusztig varieties. In any case, providing the facts that both and are vital in geometric representation theory, and that they share the same ambient space , it is very interesting and natural to seek their geometric interactions.
We found the following surprising simple relation between the components: (See Theorem 4.1 for the formal statement.)
-
(a)
Any Springer fibre at a rational unipotent element of has a component containing a component of a specific Deligne–Lusztig variety as a dense open subset; moreover, this Deligne–Lusztig component forms the whole intersection.
-
(b)
Conversely, every component of a Deligne–Lusztig variety at an involution of a specific shape is a dense open subset of a component of some Springer fibre.
As we will see in Theorem 4.1, this component relation boils down to a map from the unipotent variety to the Weyl group.
In Section 2 we make some preparations on Deligne–Lusztig varieties and Springer fibres.
In Section 3 we give a brief recall of some ingredients used in the proof of Theorem 4.1, like Weyr normal forms, Robinson–Schensted correspondence, and Spaltenstein’s and Steinberg’s descriptions of Springer components.
In Section 4 we present the proof of Theorem 4.1, which is a careful combination of the above constructions. We also derive a geometric proof of a classical dimension formula of unipotent centralisers (see Corollary 4.2).
In Section 5 we give a few remarks, including three illustrating examples (one concerning the boundary of the theorem, one concerning a uniqueness property, and one concerning an opposite phenomenon), and a short discussion on the representations associated with .
Throughout this paper: We use the convention notation for elements in an algebraic group; all varieties are assumed to be reduced; by a component we always mean an irreducible component.
Acknowledgement. The author thanks George Lusztig and Alexander Stasinski for helpful comments and suggestions, and thanks Guangyi Yue for a helpful communication. During the preparation of this work the author is partially supported by the NSFC funding no.12001351.
2. Deligne–Lusztig varieties and Springer fibres
In this section we recall some basics of Deligne–Lusztig varieties and Springer fibres. The details can be found in [DL76], [Sho88], [Ste88], [Car93].
Definition 2.1.
Let . Then the Deligne–Lusztig variety at is
where is the Lang isogeny given by .
Viewing as the variety of complete flags, one can describe in the following way.
Definition 2.2.
Let
be two complete flags of , with . We say that and are in relative position , if
for any .
-
(*)
Let be viewed as the automorphism group of . Then is the variety consisting of the complete flags such that and are in relative position , namely
for any .
Proposition 2.3 (Deligne–Lusztig).
The variety is a smooth locally closed subvariety of of pure dimension , where denotes the length of .
Proof.
See [DL76, Page 107]. ∎
Definition 2.4.
Let be a unipotent element. Then the Springer fibre at is
Using the term of flags, can be viewed as the closed subvariety consisting of complete flags stabilised by (namely, for all ).
Unlike , usually is singular, but one still has:
Proposition 2.5 (Spaltenstein, Steinberg).
The variety is of pure dimension , where denotes the number of positive roots and denotes the conjugacy class of .
Proof.
See e.g. [Sho88, 1.2]. ∎
In this paper we will very often take the viewpoint that elements in and are flags. We put .
3. Some preliminaries
In this section we recall some ingredients needed in the proof of Theorem 4.1. We first fix the notation that will be used throughout this paper:
Notation 3.1.
Let be a unipotent element. Then
-
•
is the standard Jordan normal form of , where the ’s are the Jordan blocks (with non-increasing sizes).
-
•
is the size of the Jordan block (that is, is an -matrix); in particular .
-
•
is the Young diagram associated with , that is, a Young diagram whose -th row has boxes.
-
•
is the number of boxes in the -th column of . For convenience, we also put . Note that there are totally columns and rows.
One ingredient we would need is the so-called Weyr normal form , which is a “dual” of the Jordan normal form :
Definition 3.2.
For a unipotent , the matrix is blocked upper triangular, and is characterised by the following rules:
-
(i)
The -th diagonal block is the identity matrix.
-
(ii)
The block just right to the -th diagonal block is of the form , where denotes the -identity matrix, and denotes the zero matrix of a suitable size.
-
(iii)
All other blocks are zero.
So is a blocked matrix of the shape
where the ’s denote some identity matrices of possibly different sizes and the ’s denote some zero matrices of possibly different sizes.
Although Weyr normal forms and Jordan normal forms were both discovered in the second half of the 19th century, the Weyr form appears to be much lesser known; a comprehensive reference on these normal forms is [OCV11]. Note that recently there has been a (very different) application of Weyr normal forms in the representation theory of Lie type groups over local rings; see [Sta21].
Remark 3.3.
One of the main features we need from Weyr form (instead of Jordan form) is that the block sizes are with respect to the columns of Young diagrams, which allows one to combine the other elements in the proof of Theorem 4.1 in a natural way.
Proposition 3.4.
The unipotent elements and are in the same conjugacy class.
Proof.
See e.g. [OCV11, 2.2.2]. ∎
Another ingredient we need is the Robinson–Schensted correspondence. Recall that a very basic property of finite group representation theory is the identity , where is a finite group and suns over the irreducible representations. If we take , then this identity can be re-written as
where runs over the Young diagrams of boxes, and denotes the set of standard -tableaux (we use the convention that the numbers in a standard tableaux go increasingly from left to right and from up to down). The Robinson–Schensted correspondence, which was later generalised by Knuth to a more general situation, gives a combinatorial explanation of this identity.
Proposition 3.5 (Robinson–Schensted correspondence).
There is an explicit computable bijection between the sets
satisfying the property , given by the following algorithm:
-
(i)
Take . If is in the -th box of , then we remove this box from , and denote the new tableau by .
-
(ii)
Suppose the number in the -th box of is . We remove this box from and move up by one row to replace the largest number smaller than .
-
(iii)
Suppose the number replaced by is , then we move up by one row to replace the largest number smaller than , and so on, until we replaced a number in the first row.
-
(iv)
Denote the number been replaced from the first row by , and denote the resulting tableau by .
-
(v)
Repeat the above process for with the tableaux pair , and so on, until we find all . Then
One often use the word notation .
Proof.
See e.g. [Knu98, 5.1.4]. ∎
Besides the above two ingredients, we also need the Young tableaux labelling of components of the Springer fibre given in [Spa76] and [Ste76]. We follow Steinberg’s description in [Ste88].
Proposition 3.6 (Tableaux labelling of components).
There is a bijection
More explicitly, for a given tableau , the corresponding component is characterised as the closure of an open subset , where consists of the flags
constructed via the following steps:
-
(I)
Let , the nilpotent element associated with ;
-
(II)
if is in the position of , then is any hyperplane satisfying that
-
(III)
once such a is chosen, we repeat the above process to construct a (by replacing by ), and so on.
Proof.
See [Ste88, Section 2]. ∎
Moreover, they proved the following property of generic relative position for the components of :
Proposition 3.7 (Generic relative position).
Let and be two standard -tableaux. Then there is an open dense subsvariety such that any closed point has the relative position .
4. Distribution of components
Our main theorem is:
Theorem 4.1.
Let be the variety of unipotent elements of . Then we have:
-
(a)
There is a canonical map
such that, for , the Springer fibre intersects exactly one component of the Deligne–Lusztig variety , and this component is an open dense subset of an irreducible component of .
-
(b)
Conversely, each component of , where is an involution of the shape (the block means the reversion along the word), is a dense open subset of a component of some Springer fibre.
Proof.
Proof of (a).
We first prove (a) for the Weyr form of , in a constructible manner.
Let us construct a special -tableau which will be critical for us: This is done by filling into in the way that, first fill the 1st column of from up to down, and then the 2nd column of from up to down, and so on. So we have
(1) |
(Recall that we have made the convention .)
By the Robinson–Schensted correspondence (Proposition 3.5) this gives an involution
(2) |
where is the reversion along the -th column of :
Now consider the corresponding Deligne–Lusztig variety ; we want to compute it using flags via (*). Let us fix a set as a standard basis of (over ), and view (resp. ) as the standard upper triangular subgroup (resp. diagonal maximal torus) with respect to this basis. Then the Weyl group is identified as the symmetric group permuting the subscripts of the basis and the simple reflections are identified as the transpositions . Then by (*), a complete flag
lies in if and only if the following two conditions hold:
-
(i)
For each , one has
(i.e. the spaces are -stable);
-
(ii)
for any in the interval , one has
For each , let be the corresponding reversion “modulo ”, namely
Then the condition (ii) can be re-written as:
-
(ii)’
For any in the interval , one has
where and .
Note that, considering in the quotient space , the condition (ii)’ can be viewed as the flag condition for the Deligne–Lusztig variety of
at . Therefore, the conditions (i) and (ii)’ tell us that, to construct a flag in is the same as to do the following two steps:
-
(iii)
Construct an -stable partial flag
of spaces of dimensions , which is the same as to choose a point in , where is the standard parabolic subgroup fixing the partial flag
-
(iv)
for each quotient space , take a point in the Deligne–Lusztig variety .
Therefore we get a decomposition
(3) |
into closed subvarieties. (To fix a partial part of a complete flag in the form given in the step (iii) is the same as to require a Borel subgroup to be inside a fixed parabolic subgroup, hence constitute a closed condition.) Note that corresponds to the longest element of the Weyl group of , so is an open subvariety of the flag variety of , hence irreducible. Thus the disjoint union (3) is actually the component decomposition of , and acts transitively on the components by translating the partial flag corresponding to .
Denote by the component of corresponding to . By construction consists of the flags satisfying that
-
(v)
for all ;
-
(vi)
(ii).
We shall show that , where the latter is an open subset of the component of corresponding to (see Proposition 3.6). Consider the nilpotent element , which is a blocked matrix of shape
where the ’s are the identity matrices of sizes and the ’s are the zero matrices of possibly different suitable sizes. Note that
for any flag in . Thus by Proposition 3.6, to show that , it is sufficient to show that every satisfies:
-
(vii)
If is in the interval , then
-
(viii)
if (with ), then
Actually this is clear, because a direct blocked-matrix computation gives that
for . So .
Next we show that does not intersect any other component of , namely,
By the decomposition (3), it is sufficient to show that, if a flag is stabilised by (or equivalently, , ) and satisfies the condition (ii), then it automatically satisfies the condition (v), i.e. for all . We first prove this for . Note that, for any , the condition (ii) implies that there is an internal direct sum decomposition
(4) |
which gives an internal direct sum decomposition of
where is some non-zero vector in . So by we get
(5) |
Since is -stable, (5) implies that
however, by the internal decomposition (4), this happens if and only if . Thus from (5) we see that for any . Therefore
which gives that . Now, suppose for some , then by applying the above argument to the quotient space we also get . So by induction the condition (v) holds, and we conclude that intersects at exactly the component , or more precisely,
(6) |
It remains to discuss the openness of in the component . First note that, since our is -stable, the step (II) in Proposition 3.6 implies that
So there is a graph embedding
It is well-known that (see e.g. [Car93, 7.7]) the subsets of pairs at various relative positions give a finite stratification of into locally closed subvarieties, so their intersections with also give such a stratification of ; let us denote the unique dense open strata of by . Then Proposition 3.7 implies that every pair in is in the relative position . In particular we see that
and that this is an open subvariety of (and hence of ). So, as is pure dimensional (see Proposition 2.3), by (6) we conclude that is a dense open subvariety of as desired.
The above proves (a) for . For a general unipotent , by Proposition 3.4 we know that
for some . Meanwhile, as both and are in , by the fact that two matrices similar over a field extension are similar over the original field, we can take . (Another way to see this is to apply the Lang–Steinberg theorem, by noting that centralisers in are always connected.) Thus is still a component of and is also an open dense subset of the component
The uniqueness of the component is also clear by taking a conjugation. This completes the proof of (a).
Proof of (b).
For such an involution , by Proposition 3.5 we see that for some tableau of the form specified in (1). Let be any unipotent element making a -tableau. Then, as we did in the proof of (a), there is a component of lying as an open dense subvariety in the component of .
Now, from the discussion of given in (a) we know that the translation action of on is transitive on the components; in particular, each component of is of the form for some , which is then an open dense subset of some component of . This completes the proof. ∎
Recall the dimension formula:
Corollary 4.2.
Let be a unipotent element, then
where is the number of boxes in the -th column of the Young diagram of .
This can be proved by a matrix manipulation together with a combinatorial consideration (see [SS70, IV.1] and [Hum95, 1.2 and 1.3]); here we derive a geometric proof from the argument of Theorem 4.1.
Proof.
By Proposition 3.4 we can assume that . Then from the proof of Theorem 4.1 we see that
By Proposition 2.3 and Proposition 2.5 this equality can be written as
where denotes the conjugacy class of . Note that, by the component decomposition in the argument of Theorem 4.1 (see (3)), we have
So
as desired. ∎
5. Examples and remarks
In this section we give some examples and remarks related to our main theorem, and give a short discussion on the representations associated with for rectangular .
1.
One may wonder that, in Theorem 4.1, does (a) applies to an arbitrary component of , or does (b) applies to an arbitrary involution of ? As illustrated below, without weakening the assertions usually the answer is no:
Example 5.1.
Let and let
By Proposition 2.5 and Proposition 3.6, the variety has two irreducible components, both of dimension , indexed by the Young tableaux
respectively. The Robinson–Schensted correspondence gives two involutions,
and
(The second equalities are for writing the elements as reduced products of simple reflections.) First consider the component . Explicitly, by computing the flag condition (*) for the Deligne–Lusztig variety , and by computing the step (II) of Proposition 3.6 for , one can see that
and
which suggests that contains a component of as a dense open subset, and from the argument of Theorem 4.1 we know that this is exactly the case, and this component is given as .
Now consider the other component . As mentioned in the argument of Theorem 4.1, Proposition 3.7 implies that the variety of pairs of flags at relative position cut out an open dense subset of , and so the preimage of along the Frobenius graph embedding
is an open subset of ; indeed, this open subset is non-empty: By fixing a standard basis of (over ), one easily checks that it contains the flag
for any . This means that some component of cuts out an open dense subset of . However, since
and since the Deligne–Lusztig varieties partition the flag variety, the component cannot contain any component of a Deligne–Lusztig variety as an open dense subset. On the other hand, by Proposition 2.5 and Corollary 4.2 we see that the possible dimensions of Springer fibres for are , so none of the components of can be an open subset of a Springer fibre.
2.
It is a natural desire that, the map is uniquely characterised by the property given in Theorem 4.1. To make this true one shall add further conditions:
Example 5.2.
Let and let
Then the components of are labelled by the two tableaux of hook shape
corresponding to the simple reflections and , via the Robinson–Schensted correspondence, respectively. The three varieties , , and , are all of dimension . From the argument of Theorem 4.1 we know that contains a component of as a dense open subset and does not intersect other components of . Let us consider . Given a flag , by direct computations with the flag condition (*) and the condition that takes into , one sees that intersects only at the component (of ) consisting of the flags
where runs over . Meanwhile, it follows from the step (II) of Proposition 3.6 that this component is contained in . So, at this , there are two choices of the value of fulfilling the requirement in Theorem 4.1.
Thus we hope to state here a question:
Question 5.3.
3.
Let be the longest element of , then contributes a generic (i.e. open dense) part of the flag variety. Quite opposite to the component containment relation in Theorem 4.1, Springer fibres missed this “largest” Deligne–Lusztig variety at all:
Example 5.4.
In this example we show that is always empty unless . Suppose that is non-empty; let be a point in the component for some -tableau . Consider the Frobenius graph embedding
(Here preserves because preserves and is a homeomorphism.) By Proposition 3.7 and by considering Bruhat order we get
which in turn implies that . So, via the Robinson–Schensted correspondence we see that must be a single column diagram, that is, , in which case is the whole flag variety.
4.
Recall that (see Notation 3.1) we have let be the Jordan normal form of , the number of Jordan blocks in , the sizes of each Jordan block, and the associated Young diagram.
Definition 5.5.
A unipotent element is called rectangular, if the Young diagram is rectangular.
Note that if is rectangular, then () is a divisor of and for all ; when this is the case we denote by .
As mentioned in the introduction, our original focus on the relations between Deligne–Lusztig varieties and Springer fibres comes from the smooth representation theory of the profinite group . Indeed, if is rectangular, then the finite quotient group acts on , because is naturally isomorphic to the -centraliser of
where denotes the identity matrix, by the ring injection from to :
(7) |
where . (See [Che20a, 5.3] and [Che20b, 4.6]; note that the notation used there is different up to a blocked transpose.) Thus we get a virtual representation
of . Here denotes the -th compactly supported -adic cohomology with a prime not equal to . Note that if , then this construction gives the unipotent representations of in the sense of [DL76, 7.8].
Definition 5.6.
A representation of , where , is called primitive, if it does not factor through .
For example, when is rectangular with , from [Che20b, 4.8] we known that is a primitive representation of if is a cycle with .
Question 5.7.
For a given rectangular (resp. Weyl element ), is there a reasonable characterisation of the Weyl element (resp. rectangular ) making primitive?
We note that there is a uniform description of the representations for various rectangular , using a complex on : First consider the character-sheaf type diagram (see [Lus85])
where
and are the natural projections; this diagram is -equivariant, where acts on by , on by left conjugation, and on by left multiplication. Then the character-sheaf type complex
encodes all :
Proposition 5.8.
If is rectangular, then .
Proof.
Note that , so the assertion formally follows from the proper base change along . ∎
We end with a (non-)smoothness property of the above diagram.
Proposition 5.9.
The variety and the morphism are smooth. However, the morphism is smooth only for .
Proof.
We follow the idea of [Lus85, 2.5.2] to use a faithful flat descent. Let be the base change of along the morphism
This is faithfully flat since is solvable (hence the quotient admits local sections). Then by [GD67, 17.7.7] it suffices to prove that the variety and the morphism (extending ) are smooth. Applying the variable change we get
and reads as . As the Lang morphism is étale, the assertion on and follows.
For the assertion on , again by the faithful flat descent it suffices to show that the morphism given by
is not smooth. Actually this morphism is not flat: Note that the fibre of at any central element is . So, if is flat then we have (see e.g. [Liu06, 4.3.12])
which is impossible unless . ∎
In the above argument, note that for any , the elements and have the same image under . So, the fibre of at a closed point is either empty or has dimension at least , which implies that (via [Liu06, 4.3.12])
if is flat at . Thus actually cannot be flat at any point unless is the longest element .
References
- [BV21] Roman Bezrukavnikov and Yakov Varshavsky. Affine Springer fibers and depth zero L-packets. arXiv preprint arXiv:2104.13123, 2021.
- [Car93] Roger W. Carter. Finite groups of Lie type. Wiley Classics Library. John Wiley & Sons Ltd., Chichester, 1993. Conjugacy classes and complex characters, Reprint of the 1985 original, A Wiley-Interscience Publication.
- [Che20a] Zhe Chen. Flags and orbits of connected reductive groups over local rings. Math. Ann., 376(3-4):1449–1466, 2020.
- [Che20b] Zhe Chen. Twisting operators and centralisers of lie type groups over local rings. arXiv preprint arXiv:2006.02145, 2020.
- [DL76] Pierre Deligne and George Lusztig. Representations of reductive groups over finite fields. Ann. of Math. (2), 103(1):103–161, 1976.
- [GD67] Alexander Grothendieck and Jean Dieudonné. Éléments de géométrie algébrique IV Étude locale des schémas et des morphismes de schémas. Inst. Hautes Études Sci. Publ. Math., 1964-1967.
- [Hum95] James E. Humphreys. Conjugacy classes in semisimple algebraic groups, volume 43 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1995.
- [Knu98] Donald E. Knuth. The art of computer programming. Vol. 3. Addison-Wesley, Reading, MA, 1998. Sorting and searching, Second edition.
- [Liu06] Qing Liu. Algebraic geometry and arithmetic curves, volume 6 of Oxford Graduate Texts in Mathematics. Oxford University Press, 2006.
- [Lus85] George Lusztig. Character sheaves I. Adv. in Math., 56(3):193 – 237, 1985.
- [Lus90] George Lusztig. Green functions and character sheaves. Ann. of Math. (2), 131(2):355–408, 1990.
- [OCV11] Kevin C. O’Meara, John Clark, and Charles I. Vinsonhaler. Advanced topics in linear algebra. Oxford University Press, Oxford, 2011. Weaving matrix problems through the Weyr form.
- [Sho88] Toshiaki Shoji. Geometry of orbits and Springer correspondence. Astérisque, (168):9, 61–140, 1988. Orbites unipotentes et représentations, I.
- [Spa76] Nicolas Spaltenstein. The fixed point set of a unipotent transformation on the flag manifold. Nederl. Akad. Wetensch. Proc. Ser. A, 38(5):452–456, 1976.
- [Spa82] Nicolas Spaltenstein. Classes unipotentes et sous-groupes de Borel, volume 946 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1982.
- [Spr76] Tonny A. Springer. Trigonometric sums, Green functions of finite groups and representations of Weyl groups. Invent. Math., 36:173–207, 1976.
- [SS70] Tonny A. Springer and Robert Steinberg. Conjugacy classes. In Seminar on Algebraic Groups and Related Finite Groups (The Institute for Advanced Study, Princeton, N.J., 1968/69), Lecture Notes in Mathematics, Vol. 131, pages 167–266. Springer, Berlin, 1970.
- [Sta21] Alexander Stasinski. Representations of over finite local rings of length two. J. Algebra, 566:119–135, 2021.
- [Ste76] Robert Steinberg. On the desingularization of the unipotent variety. Invent. Math., 36:209–224, 1976.
- [Ste88] Robert Steinberg. An occurrence of the Robinson-Schensted correspondence. J. Algebra, 113(2):523–528, 1988.
- [Tym06] Julianna S. Tymoczko. Linear conditions imposed on flag varieties. Amer. J. Math., 128(6):1587–1604, 2006.