On the multiplicity formula for discrete automorphic representations of disconnected tori
Abstract
Kaletha extended the local Langlands conjectures to a certain class of disconnected groups and proved them for disconnected tori. Our first main result is a reinterpretation of the local Langlands correspondence for disconnected tori. Our second, and the central objective of this paper, is to establish an automorphic multiplicity formula for disconnected tori.
1 Introduction
Let be a number field and be the adele ring of . Let be a connected reductive group defined over with centre . Let be its adelic quotient, where is the identity component of -points of the largest -split subtorus of . Then there is a right regular representation of on the Hilbert space .
In the theory of automorphic representations, it is a crucial question to understand the decomposition of the discrete part and the multiplicities of its irreducible constituents. The work of Labesse and Langlands ([labesse1979indistinguishability],[langlands1982debuts]) and the work of Kottwitz ([kottwitz1984stable]) suggest a conjectural answer for tempered automorphic representations. They expect that an admissible tempered discrete homomorphism gives rise to an adelic -packet of tempered representations of and conversely, each tempered discrete automorphic representation belongs to some -packet. The topological group above is usually called the (hypothetical) Langlands group of and it should have the Weil group as a quotient. Then it is expected that the multiplicity of (which can be ) in is the sum of with running over the equivalence classes of such parameters, where we have denoted by the -contribution towards the total multiplicity of . Moreover, it is conjectured that is given by
(1.1) |
where is a finite group related to the centraliser of in , and is a complex-valued pairing relying on the local Langlands conjectures. In [kaletha2018global], after introducing the global Galois gerbes, Kaletha makes the definitions of and the pairing precise for general connected reductive groups .
In [kaletha2022local], Kaletha initiated an extension of the (refined) local Langlands conjectures to disconnected groups. To be precise, he treats the inner forms of “quasi-split” disconnected reductive groups, which are of the form , where is a connected quasi-split reductive group and acts on by preserving an -pinning. Kaletha extends the (refined) local Langlands conjectures to this setting and proves the conjectures in the case of disconnected tori.
1.1 Main results
Based on Kaletha’s pioneering work in the local aspects, the principal goal of this thesis is to explore the global aspects for the disconnected tori and establish a multiplicity formula. It turns out that the formula we obtain in this setting takes a similar form as (1.1).
First, we introduce the disconnected tori we treat, which are the main players of this thesis, and state the multiplicity formula on the automorphic side. Let be a quasi-split disconnected torus defined over with identity component a torus , that is is a semidirect product of with some finite group defined over . Given , we twist the rational structure of via through inner automorphisms, and obtain a pure inner form . Regarding its rational points, there is a short exact sequence
(1.2) |
where is the stabiliser of in . Moreover, we see that acts on the set of Hecke characters of by conjugation. When is a Hecke character of , we denote the stabiliser of in by .
To ease the exposition, in this introduction, we assume that is anisotropic. Under this assumption, is compact. Let be an irreducible constituent of . By an argument involving the Clifford Theory, the multiplicity of in can be determined by eventually passing to a finite group and calculating its the character. To be precise, we have (Theorem LABEL:automult)
where runs over the -orbits (or equivalently, -orbits) of Hecke characters of .
Before entering the dual side, we need to review the local aspects first. We briefly recall the local Langlands correspondence for disconnected torus in terms of pure inner forms. Let be a place of , then is a quasi-split disconnected torus over under base change. We can consider the local pure inner form . It is noteworthy that the rational points of can be related to certain -hypercocycles in an obvious manner. Let be a local -parameter for . By suitably enlarging the -group, one can extend the group of self-equivalences of from to , and define as the set of irreducible representations of whose restrictions to contains (which is a character of in Kottwitz’s sense). On the other hand, by the local Langlands correspondence for tori, gives rises to a character of . One can define as the set of irreducible representations of whose restrictions to contain .
Now the local Langlands correspondence for disconnected tori asserts there is a natural bijection between and , which satisfies the character identities. Kaletha constructed the bijection and verified the character identities. One of the main ingredients in his construction is the Tate-Nakayama pairing (or duality) for hypercohomology introduced in [kottwitz1999foundations], which organically combines the (classical) Tate-Nakayama duality and the local Langlands correspondence for (connected) tori. After passing both the group side and the dual side to certain extensions of by , Kaletha shows that there is a canonical isomorphism between them, which induces the desired bijection. The slight drawback of this approach is that the isomorphism is given in a less transparent manner, due to the need of choosing sections for the extensions.
The first main result of this work is an intrinsic reinterpretation of the Kaletha’s LLC for disconnected tori. To be precise, we are able to define a simple relation (LABEL:reinterrelation) to characterise the 1-1 correspondence between and intrinsically, in the sense that there is no choice of “coordinate systems” involved.
Now we return to the global aspect. Let be a global -parameter and let be the associated Hecke character of under the global Langlands correspondence. One can define the adelic -packet associated to . Essentially, it consists of irreducible representations of that are unramified almost everywhere and whose local component at any place contains . Then we define a complex-valued pairing (LABEL:thepairing) given by
where and are arbitrarily chosen. Despite the fact that the choices of local data appearing in the above pairing have great flexibility, we can show the well-definedness of the pairing and its independence on the choices of the local data.
At this point, a key input is our reinterpretation of the LLC for disconnected tori, which enables us to find that this pairing has its counterpart on the automorphic side. In fact, we have
(1.3) |
After a comparison with the (actual) multiplicity obtained from the automorphic side, we reach the desired multiplicity formula
where runs over the near -equivalence classes of global -parameters, which are in 1-1 correspondence to the -orbits (or equivalently -orbits) of Hecke characters of .
1.2 Structure of the paper
In Chapter 2, we review the (local and global) Tate-Nakayama duality for hypercohomology introduced in [kottwitz1999foundations, Appendix A.3]. We aim to explicitly elucidate the isomorphisms on the chain level.
In Chapter 3, we recall the conventions (following [kaletha2022local]) on the disconnected groups considered in Kaletha’s framework. In particular, we give some examples of disconnected tori.
Chapter 4 and Chapter 5 are devoted to the local aspects. In Chapter 4, we closely follow [kaletha2022local] and review the statement of the local Langlands correspondence for disconnected tori in terms of pure inner forms. In Chapter 5, we give a new construction of the LLC, which can be eventually proved to coincide with Kaletha’s construction.
The rest of the chapters are focused on establishing the multiplicity formula. In Chapter 6, we lay the necessary foundation for the discussion of global aspects. In Chapter 7, we aim to calculate the multiplicity on the automorphic side (which means the actual multiplicity of an automorphic representation). For this purpose, we first extract a dense smooth subspace from the -space. After a preliminary decomposition, we are able to break the calculation of multiplicity into a sum of -contribution with running over the -orbits of Hecke characters. Eventually, the multiplicity can be calculated through some arguments in Clifford theory. In Chapter 8, we enter the dual side. After defining the pairing , we apply our reinterpretation of the LLC so as to conclude that the pairing has an incarnation on the automorphic side. Finally, the multiplicity formula is established after a straightforward comparison.
Acknowledgements.
The author is grateful to Wee Teck Gan for his invaluable insights, as well as for his guidance and proof-reading. The author also appreciates Tasho Kaletha for his interest in this project and helpful suggestions. The author gratefully acknowledges the support of NUS Research Scholarship.
2 The Tate-Nakayama Duality for hypercohomology
In this section, we review the local and global duality between the hypercohomology of the Galois group and that of the Weil group. These results are due to Kottwitz-Shelstad [kottwitz1999foundations]. We will see that the duality is able to encompass the LLC for (connected) tori and Tate-Nakayama duality simultaneously. We refer the reader to Appendix LABEL:appA for the notion and basic properties of group hypercohomology, Appendix LABEL:appLanglands for the (local and global) Langlands correspondences for connected tori, and Appendix LABEL:appTN for the (local and global) Tate-Nakayama dualities.
2.1 Convention
Whenever we say hypercohomology of profinite groups in a (complex of) discrete module(s), we always refer to the continuous version.
Whenever we consider (hyper-)cochains/cocycles/cohomology of Weil group or relative Weil groups , we always assume continuity, unless otherwise stated or indicated as the abstract version by the subscript “abs”. In contrast, in this work, all the (hyper-)chains/cycles/homology of (relative) Weil groups are always in the abstract sense.
2.2 Local Tate-Nakayama duality for hypercohomology
In this part, we review the duality for the hypercohomology of the Galois group and the Weil group. These results were established by Kottwitz-Shelstad [kottwitz1999foundations]. We will see that the duality is able to encompass the LLC for tori and Tate-Nakayama duality (the case in the previous section) simultaneously. We refer the reader to Appendix LABEL:appA for the notion and basic properties of group hypercohomology.
Let be a local field of characteristic zero. Let and be -tori with cocharacter groups and . Let be a morphism defined over . Let and be the maps induced by . We aim to define a pairing between and . For convenience, we fix a finite Galois extension of such that both and split over .
Throughout this chapter, once and for all, we fix a section with , so that we have maps and on the chain level with explicit formulae (LABEL:themapphi_formula) and (LABEL:themappsi), respectively. We will eventually show that the choice of section has no effect on the (co)homology level (see Proposition 2.3).
2.2.1 Step 1
We quickly remind the reader that, whenever talking about (hyper)homology of (relative) Weil groups, we ignore the topology and regard them as abstract groups. First, we note that the image of the differential
lies in , the subgroup of norm- elements in . Indeed, given , we have
Now for each fixed , the inner sum vanishes and it follows that has norm .
Next, we define a modified (norm-) hyperhomology group as a subgroup of . Given , we consider the following complex used in defining group hyperhomology:
(2.1) |
where and . We recall from Appendix LABEL:appA that is defined as the quotient . We define to be the subgroup of pairs in with . According to the observation made above, we have . Now we define the modified hyperhomology group
(2.2) |
Analogous to Fact LABEL:Exactforhyperho, we have an exact sequence
The main goal of Step 1 is to define a natural isomorphism
For this purpose, we recall from (LABEL:themapphi) and (LABEL:themappsi) that we have defined
and | ||||
The key observation is that and sit in the following diagram to make it commute:
(2.3) |
For brevity, we dropped the (Galois/Weil) groups concerned in each term in the above diagram, but we remind the reader that the chains are of , while the cochains and cocycles are of .
Lemma 2.1.
The diagram (2.3) commutes.
Proof.
We recall that, by definition, . The restriction is a chain map (see (LABEL:reschain)), thus we have . Now we notice that the map (on the chain level)
is trivial on -boundaries, which implies . Thus, the commutativity of the first square follows.
Now we check the commutativity of the second square. Let . On the one hand, we find that is an element in sending to
We let and alter the domain over which the sum is taken accordingly. Then the above expression becomes
(2.4) |
On the other hand, through an elementary computation involving certain changes of variables, we find that sends to
(2.5) |
where we have used the -cocyle relation in the second to last equality.
We consider the following diagram with the first row a modification of (2.1) (th-chains modified to norm- subgroups), and the second row mixed with cochains and cocycles of :
(2.6) |
where and , hence the cohomology of the second row at is nothing but . And the vertical maps are indicated in the diagram. Using the commutativity of the diagram (2.3), one can immediately check:
Fact 2.2.
The diagram (2.6) commutes.
At this point, the commutativity of diagram (2.6) immediately suggests that the middle vertical map actually passes to (co)homology:
(2.7) |
We quickly check the well-definedness of :
Proposition 2.3.
does not depend on the choice of section .
Proof.
Suppose is another section, is the corresponding -cocycle given by (LABEL:2-cocycle), and and are the chain-level maps (LABEL:themapphi) and (LABEL:themappsi) defined in terms of and . Let . We hope to show that and differ by a -hypercoboundary. We consider the element
in . Then elementary compuations suggest that, for any , we have
We note that we have used the fact that has trivial norm in the last equality above. Since lies in , we have . Using this and some elementary manipulations, one can find that
coincides with
where the products are taken over triples
We conclude that the difference between and is a -hypercoboundary. ∎
One can further show:
Proposition 2.4.
The natural map
(2.8) |
is an isomorphism.
Proof.
We consider the following diagram with rows the long exact sequences associated to hyper(co)homology (see LABEL:LESforhyper):
(2.9) |
In the above diagram, is the key isomorphism (LABEL:keyisomodified) induced by in Deligne’s convention, is the Tate-Nakayama isomorphism (LABEL:-1TN) induced by , and is the natural map induced by .
We can see that the diagram (2.9) commutes. Indeed, on the one hand, we recall from LABEL:LESforhyper that the two arrows in the middle of each row are induced by an inclusion and a projection on the chain level. Hence the commutativity of the two squares in the middle is clear. On the other hand, the first and the last squares commute due to functoriality of and the Tate-Nakayama isomorphism.
It follows that is an isomorphism by the five lemma. ∎
2.2.2 Step 2
In this part, we will proceed as in Section LABEL:Step_2 to produce a pairing between certain hypercohomology and hyperhomology groups.
Let be the bar resolution of . Then the defining complex of :
is nothing but
(2.10) |
which we denote by .
Similar to the argument in Section LABEL:Step_2, we note that
which also holds with replaced by .
We note that is an injective abelian group, hence after applying the functor to complex (2.10) and taking cohomology, we obtain
Hence we have
The right-hand side is the abstract cohomology group, ignoring the topology on and . Similarly, for hyper(co)homology we have canonical isomorphism
Explicitly, given a 0-hypercycle in , i.e. , and given a 1-hypercocycle with and such that . The pairing between and is
(2.11) |
Remark 2.5.
We need to point out some subtlety here. On the one hand, we have regarded as a complex concentrated at degrees and . Hence the cohomology of dual complex actually gives hypercohomology of with and placed at degrees and , respectively. However, whenever we write , we are always placing and at degrees and . Hence, the inverse we add in (2.11) is due to this discrepancy.
So far, we have obtained a pairing between and . We recall from (2.2) that we defined a subgroup of . And we can restrict the pairing to subgroups
and | ||||
Here, we recall that denotes the continuous cohomology group.
Furthermore, in view of the isomorphism we established in Step 2, we have obtained a pairing
(2.12) |
Suppose is another Galois extension of . Then again, we have a pairing with replaced by above. More precisely, we have the following diagram:
(2.13) |
Proposition 2.6.
The pairing (2.12) is compatible with inflations.
Proof (Sketch)..
The desired compatibility follows from the compatibility of
(2.14) |
and the commutativity of
(2.15) |
The compatibility in (2.14) is obvious, since the deflation between homology is dual to the inflation between cohomology. The commutativity of the diagram (2.15) can be shown in the same manner as the proof of Proposition 2.3. Details can be found on pp. 136-137 of [kottwitz1999foundations]. ∎
Therefore, after passing to colimits, we have the desired functorial pairing
(2.16) |
Proposition 2.7.
The pairing (2.16) is compatible with the Langlands pairing (LABEL:Langlands-pairing) (in Langlands’ convention) and the Tate-Nakayama pairing (LABEL:TN-pairing). Precisely speaking, if we let the long exact sequences on the group side and dual side pair with each other
(2.17) |
then we have
and | ||||
for each , , and .
Proof.
Both compatibilities follow from the commutativity of Diagram (2.9) and the definition of the pairing (2.11). Compatibility with Tate-Nakayama pairing is immediate. As for compatibility with the LLC in Langlands’ convention, we first notice that is compatible with , which differs from by a sign. Then we compare the pairing (2.11) with the pairing (LABEL:co-hopairingllc), and find that they differ by a sign as well. Therefore, the two minus signs cancel and the desired compatibility follows. ∎
2.2.3 Step 3
In this final step, we take continuity into account. But first we need to endow with a topology. Recall that we have the long exact sequence
Then we topogise by stipulating that
is a continuous open map. We quickly note that induces an isomorphism of topological groups between the quotient group and the image of . Since is finite, is an open subgroup of finite index.
In particular, a character of is continuous if and only if it is continuous on the image . According to Fact 2.7, the pairing (2.16) is continuous on indeed, giving us a map
In virtue of Proposition 2.7, the following diagram commutes:
(2.18) |
In the above diagram, the first row consists of (hyper)cohomology groups of applied by . We denote by for brevity. The second row consists of continuous (hyper)cohomology groups of . The two vertical maps to the right are the LLC map for tori, and the two vertical maps to the left are induced by the Tate-Nakayama pairing (LABEL:TN-pairing). The isomorphism (LABEL:TN-Kottwitz) implies that the two vertical maps to the left are surjective, with kernels and , respectively.
We define the quotient
and modify the above commutative diagram to
(2.19) |
By the five lemma, the vertical map in the middle must be an isomorphism, because all the other four are so. We have thus shown
Proposition 2.8.
The pairing (2.16) induces a functorial isomorphism
2.3 A cohomological lemma
For later convenience, we record an interesting result. This will play an important role in our reinterpretation of the local Langlands correspondence for disconnected tori (Theorem LABEL:reinterthm) and its proof (especially the proof of Lemma LABEL:lem:homo).
Theorem 2.9.
Let , and be -tori, and consider -morphisms and :
Let and be the morphisms between the dual tori induced by :
We consider the natural map induced by
sending the class of to that of , and the natural map induced by
sending the class of to that of . Then the Tate-Nakayama pairing between the image of and the image of vanishes:
Proof.
We go back to the definition of Tate-Nakayama pairing. First we fix a finite Galois extension such that , and split over . Under the isomorphism, we let correspond to . Then corresponds to . Now the Tate-Nakayama pairing reads as
(2.20) |
Since and , we have
(2.21) |
and
(2.22) |
for each . Now we can plug (2.21) and (2.22) into (2.20):
∎
2.4 Global Tate-Nakayama Duality for hypercohomology
In this section, we turn to the global Tate-Nakayama duality for hypercohomology. As its local analogue, this combines the global Langlands correspondence and the global Tate-Nakayama duality. Since the constructions are carried out in the way as the local case, details are omitted.
Let be a number field, and be its idele class group. Idele class groups will play the same role as multiplicative groups have played in the local setting. Let be the global Weil group of . Let and be tori defined over with cocharacter groups and , respectively, and be an -morphism. We fix a finite Galois extension of over which both and split.
First, we introduce some Galois hypercohomology groups:
It is the last group that will appear in the duality and it can be topologized in the same manner as Step 3 of Section 2.2 (see also Kottwitz-Shelstad [kottwitz1999foundations] for details). We note an elementary long exact sequence involving all the three groups above:
(2.23) |
On the dual side, we can define the continuous hypercohomology and its reduced version as in the local setting. Now we state the global Tate-Nakayama duality for hypercohomology:
Theorem 2.10.
There is a natural functorial isomorphism
that is compatible with the global Langlands correspondence (or more precisely, the isomorphism (LABEL:glcisom)) and the global Tate-Nakayama pairing (LABEL:TN-pairing-global). Moreover, this isomorphism is compatible with the local Tate-Nakayama duality for hypercohomogy, in the sense that the following diagram commutes:
The construction is the same as in the local case, and so is the compatibility with the GLC and the Tate-Nakayama pairing. Meanwhile, the local-global compatibility can be readily checked from the construction.
3 Disconnected reductive groups
In this chapter, we closely follow [kaletha2022local] and introduce a certain class of disconnected groups, for which we reserve the term “disconnected reductive groups” in the rest of this work. Due to our emphasis on disconnected tori, we will present some rank- examples.
3.1 Convention
Let be a field of characteristic with absolute Galois group . In this work, we say an affine algebraic group is a disconnected reductive group, if satisfies the following conditions:
-
•
There is an -isomorphism
where is a connected reductive group and is a (nontrivial) finite group.
-
•
The action of on preserves some fixed -pinning of .
Non-example 3.1.
The normaliser of the diagonal torus in does not split as a semidirect product (even over ). Indeed, there does not exist any element of order two in the nonidentity component of -points.
3.2 Classification
In this part, we will recall from [kaletha2022local] the classification of disconnected reductive groups. It is a well-known fact that each connected reductive group has a unique split form, and moreover has a unique quasi-split inner form. The same notions can be extended to the disconnected setting, although there turns out to be an additional type, “translation form”, in the term of [kaletha2022local]. We start from extending the notions of “quasi-split” and “split” groups:
Definition 3.2.
We call a split disconnected reductive group, if there is an -isomorphism
where is a split connected reductive group, and is a (nontrivial) constant group scheme acting on by preserving some -pinning of it.
Definition 3.3.
We call a quasi-split disconnected reductive group, if there is an -isomorphism
where is a quasi-split connected reductive group, and is a (not necessarily constant) finite group scheme acting on (the action is defined over ) and preserving some -pinning of it.
We note that split disconnected reductive groups can be classified by the root datum of and the action of on the root datum. It can be immediately seen from the definitions that every disconnected reductive group is a unique form of a split disconnected reductive group. Now, to obtain a classification of all disconnected reductive groups, it suffices to classify all the forms of a given split disconnected reductive group . Equivalently, it suffices to understand the Galois cohomology . This leads us to a close examination on the automorphism group first. From now on, we omit the subscript for brevity.
We highlight three subgroups of :
-
•
, the group of inner automorphisms.
-
•
, the group of translation automorphisms. Each -cocycle induces an automorphism .
-
•
, the group of pinned automorphisms over , as which we refer to automorphisms preserving the pinning of and the subgroup .
We note that intersects nontrivially with . The intersection is (as a subgroup of ) or (as a subgroup of ). Now, one can characterise the structure of (see [kaletha2022local, Section 3.1]):
Fact 3.4.
There is a semidirect product decomposition:
In view of this result, we resume discussing the classification of disconnected reductive groups. Let be a split disconnected reductive group. One observes that twisting the rational structure of by a cocycle in yields a quasi-split disconnected reductive group . It is also clear that each quasi-split disconnected reductive group arises in this way.
Then, one can twist the rational structure on the quasi-split group by a -cocycle to obtain , an inner form of the quasi-split group. Furthermore, twisting by some yields a translation form of the inner form . Now Fact 3.4 implies that each disconnected reductive group can be obtained in this manner. To put it in a concise way, each disconnected reductive group is a translation form of an inner form of some quasi-split disconnected reductive group.
3.3 Examples of some simple disconnected tori
In this work, we focus on are disconnected tori. As the name suggests, a disconneted torus is a group that has a torus as its identity component and satisfies the Convention 3.1.
In order to provide the reader with a glimpse into disconnected tori and their rational points, we present instances of the most elementary (nontrivial) case, when the identity component is and the component group is . In this case, there are two split forms: the direct product , and the semi-direct product with acting by inverting. We classify their forms and also exhibit the -rational points of all the forms.
3.3.1 Forms of
Let . We write . The automorphism group of can be worked out as
We write the nontrivial element in the first (resp. second) as (resp. ). As an automorphism of , fixes the identity component pointwise and sends any from the non-identity component to . And is the automorphism sending to . And clearly, the Galois action on is trivial. Thus, the forms of are classified by the Galois cohomology
Any nontrivial cocycle falls into one of the four categories to be described below, and we denote by the group obtained from twisting by .
-
•
Case 1: factors through a quadratic extension and maps onto the first -factor: . Let . Then after passing to , the Galois action twisted by (which we denote by adding a subscript ) is given by and . Then the group of rational points is
-
•
Case 2: factors through a quadratic extension and maps onto the second -factor: . Let . Then after passing to , the Galois action twisted by is given by . Thus we have
-
•
Case 3: factors through a quadratic extension and maps into diagonally: . Let . Then after passing to , the Galois action twisted by is given by and . So we have
-
•
Case 4: factors through a biquadratic extension and maps isomorphically to : . We write and assume and . After passing to , the twisted Galois action thus obtained is , , and . Then we have
3.3.2 Forms of
Let . The semidirect product is given by the inverting action of on . An elementary calculation suggests that the inner automorphisms and translation automorphisms coincide over . For convenience, we consider them as translation automorphisms, and one can check
in which acts on by inverting. For , we denote the automorphism that fixes and sends by . And we denote the nontrivial element in by . Again, is the automorphism sending .
One can further check that the Galois group acts on by . The forms of are classified by the Galois cohomology
To compute the Galois cohomology , we consider the following long exact sequence
where the surjectivity of the last map follows from the existence of a splitting due to the semidirect product. In view of Hilbert 90, we have , and hence we conclude that the fibre over the trivial element in is a singleton, which is nothing but the split form. Now, it suffices to understand the fibre of any nontrivial element (which corresponds to a quadratic extension ). To this end, we consider the cocycle sending to . Clearly, we have .
It remains to investigate whether the fibre contains any other element than . From now on, we fix the quadratic extension . If we twist the Galois action on the short exact sequence by (abbreviated as below) and take the long exact sequence, then we obtain
We note that there are identifications
Remark 3.5.
More generally speaking, given a short exact sequence of (not necessarily abelian) -modules , the kernel of the map (in the associated long exact sequence) is in 1-1 correspondence to the orbit space , where the action of on is given by: ( is any lift of ). When the -modules are abelian, this recovers the usual quotient. See [serre1979galois, Chapter I] for more details on nonabelian cohomology.
Claim 3.6.
The action of on is trivial.
Proof.
Due to the exactness at , it suffices to show the surjectivity of the projection . It is clear that lies in according to the definition of . So the surjectivity follows. ∎
Therefore, is in bijection with . It is clear from the construction that (with the Galois action twisted by ) coincides with -points of the norm torus determined by the quadratic extension . We have . Explicitly, the isomorphism is given by sending to , where is the nontrivial element.
Based on discussion above, each coset represents a class in , and we can define by setting after passing to the quotient . After composing with , we are able to obtain a -cocycle in the original (untwisted) sense. Let defined by (after passing to ) . We conclude that each form in the fibre arises as a thus obtained. The -rational points can be easily computed:
-
•
After passing to , the Galois action twisted by is given by and . The -points are given by
We note that, whenever , there are no rational points on the non-identity component.
3.4 Rational points on inner forms of quasi-split disconnected groups
In this work, we focus on inner forms of quasi-split groups and do not treat translation forms (that do not fall into the former category). Let be a quasi-split disconnected reductive group. Although the semidirect product is defined over and is not necessarily constant as a finite group scheme, we will still abbreviate as .
Let . We obtain by twisting the rational structure of via . After this twisting process, there is still a short exact sequence of -modules:
where the twisted Galois action on is given by
for , while the Galois action on is unchanged.
After taking -fixed points, we have
The last projection is not always surjective. In fact, one can see that if and only if and
(3.1) |
for any . We define as the subgroup of , comprising elements for which there exists some such that (3.1) is satisfied for any . Then there is a short exact sequence
4 The LLC for disconnected tori
Let be a local field of characteristic zero with absolute Galois group . The goal of this chapter is to state the LLC for disconnected tori in terms of pure inner forms.
4.1 Pure inner forms
We start with a quasi-split disconnected torus defined over . To be precise, is a (not necessarily split) torus over , is a (not necessarily constant) finite group scheme defined over acting on , and the action of on is also defined over .
Let . Under the natural map , is brought to a -cocycle taking values in the group of inner automorphisms . We twist the rational structure of by (or more precisely, by ) and obtain an inner form , which we call a pure inner form.