Colliot-Thelene J. Quadratic Forms, Linear Algebraic Groups, and Cohomology PDF
Colliot-Thelene J. Quadratic Forms, Linear Algebraic Groups, and Cohomology PDF
VOLUME 18
Series Editor:
Krishnaswami Alladi, University of Florida, U.S.A.
QUADRATIC FORMS,
LINEAR ALGEBRAIC GROUPS,
AND COHOMOLOGY
Edited by
JEAN-LOUIS COLLIOT-THÉLÈNE
CNRS, Université Paris-Sud, Orsay, France
SKIP GARIBALDI
Emory University, Atlanta, GA, USA
R. SUJATHA
Tata Institute of Fundamental Research, Mumbai, India
VENAPALLY SURESH
University of Hyderabad, Hyderabad, India
123
Editors
Prof. Jean-Louis Colliot-Thélène Prof. Skip Garibaldi
CNRS Emory University
Université Paris-Sud Dept. Mathematics &
Labo. Mathematiques Computer Science
91405 Orsay CX Dowman Dr. 400
France 30322 Atlanta Georgia
[email protected] MSC W421
USA
[email protected]
Prof. R. Sujatha
Tata Institute of Fundamental Research Venapally Suresh
School of Mathematics University of Hyderabad
Homi Bhabha Rd. Dept. Mathematics & Statistics
400005 Mumbai 500046 Hyderabad
Colaba P.O. Central University
India India
[email protected] [email protected]
ISSN 1389-2177
ISBN 978-1-4419-6210-2 e-ISBN 978-1-4419-6211-9
DOI 10.1007/978-1-4419-6211-9
Springer New York Dordrecht Heidelberg London
We dedicate this volume to Professor Parimala on the occasion of her 60th birthday.
It contains a variety of papers related to the themes of her research. Parimala’s first
striking result was a counterexample to a quadratic analogue of Serre’s conjecture
(Bulletin of the American Mathematical Society, 1976). Her influence has contin-
ued through her tenure at the Tata Institute of Fundamental Research in Mumbai
(1976–2006), and now her time at Emory University in Atlanta (2005–present).
A conference was held from 30 December 2008 to 4 January 2009, at the Uni-
versity of Hyderabad, India, to celebrate Parimala’s 60th birthday (see the confer-
ence’s Web site at http://mathstat.uohyd.ernet.in/conf/quadforms2008). The orga-
nizing committee consisted of J.-L. Colliot-Thélène, Skip Garibaldi, R. Sujatha,
and V. Suresh. The present volume is an outcome of this event.
We would like to thank all the participants of the conference, the authors who
have contributed to this volume, and the referees who carefully examined the sub-
mitted papers. We would also like to thank Springer-Verlag for readily accepting to
publish the volume. In addition, the other three editors of the volume would like to
place on record their deep appreciation of Skip Garibaldi’s untiring efforts toward
the final publication.
We are grateful for the support and the hospitality of the University of Hyder-
abad, especially the members of the Department of Mathematics and Statistics. We
would like to thank the office staff of the Department of Mathematics and Statistics
and the other staff of the University responsible for providing administrative and
logistical support.
We are also extremely grateful to the University Grants Commission, India and
the National Board for Higher Mathematics for financial support.
v
vi Preface
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Part I Surveys
Multiples of forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Eva Bayer-Fluckiger
ix
x Contents
Eva Bayer-Fluckiger
To my friend Parimala
Summary The aim of this paper is to survey and extend some results concerning
multiples of (quadratic, hermitian, bilinear...) forms.
Introduction
E. Bayer-Fluckiger
Ecole Polytechnique Fédérale de Lausanne, Mathématiques, EPFL–FSB–IMB–CSAG,
Station 8, 1015 Lausanne, Switzerland, e-mail: [email protected]
2010 Mathematics subject classification. Primary: 12G05. Secondary: 12F99.
Theorem. Suppose that cd2 (Γk ) ≤ d, and let G be a finite group that has no quotient
of order 2. Let q and q be two G-quadratic forms defined on the same k[G]-module,
and let φ ∈ I d−1 . Then
φ ⊗ q φ ⊗ q .
Let ks be a separable closure of k, and set Γk = Gal(ks /k). For any discrete Γk -module
C, set H i (k,C) = H i (Γk ,C). We say that the 2-cohomological dimension of Γk is at
most d, denoted by cd2 (Γk ) ≤ d, if H i (k,C) = 0 for all i > d and for every finite
2-primary Γk -module C.
Set H i (k) = H i (k, Z/2Z), and recall that H 1 (k) k∗ /k∗2 . For all a ∈ k∗ , let us
denote by (a) ∈ H 1 (k) the corresponding cohomology class. We use the additive
notation for H 1 (k). If a1 , . . . , an ∈ k∗ , we denote by (a1 ) ∪ · · · ∪ (an ) ∈ H n (k) their
cup product.
If U is a linear algebraic group defined over k, let H 1 (k,U) be the pointed set
H (Γk ,U(ks )) (cf. [11], [12, Chap. 10]).
1
All quadratic forms are supposed to be nondegenerate. We denote by W (k) the Witt
ring of k, and by I = I(k) the fundamental ideal of W (k). For all a1 , . . . , an ∈ k∗ , let
us denote by
a1 , . . . , an =
1, −a1 ⊗ · · · ⊗
1, −an the associated n-fold Pfister
form. It is well known that I n is generated by the n-fold Pfister forms. The following
has been conjectured by Milnor
en : I n /I n+1 → H n (k)
such that
en (
a1 , . . . , an ) = (a1 ) ∪ · · · ∪ (an )
for all a1 , . . . , an ∈ k∗ .
It is easy to see that the above theorem has the following consequences (cf. [3]):
Corollary 1.2.2. Suppose that cd2 (Γk ) ≤ d. Let q and q be two quadratic forms with
dim(q) = dim(q ), and let φ ∈ I d . Then φ ⊗ q φ ⊗ q .
Multiples of forms 5
For every quadratic form q, let us denote by disc(q) ∈ H 1 (k) its discriminant.
Recall that if n = dim(q), then disc(q) = (−1)n(n−1)/2 det(q).
Corollary 1.2.3. Suppose that cd2 (Γk ) ≤ d. Let q and q be two quadratic forms with
dim(q) = dim(q ), and let φ ∈ I d−1 . Then
For any quadratic form q, let us denote by w2 (q) ∈ Br2 (k) the Hasse–Witt invari-
ant of q. Recall that if q
a1 , . . . , an , then w2 (q) = ∏i< j (ai , a j ), where (ai , a j ) is
the quaternion algebra over k determined by ai and a j . We can extend the previous
results as follows:
Corollary 1.2.4. Suppose that cd2 (Γk ) ≤ d. Let q and q be two quadratic forms.
Suppose that dim(q) = dim(q ) and det(q) = det(q ). Let φ ∈ I d−2 . Then
Theorem 2.1.1.
(a) We have I d J = 0.
(b) If σ is of the second kind, then I d−1 J = 0.
(c) If σ is of the first kind and of the symplectic type, then I d−2 J = 0.
Theorem 2.1.2. Suppose that D is a quaternion algebra, and that σ is of the first
kind and of the orthogonal type. Suppose that cd2 (Γk ) ≤ d. Let h ∈ J, and let φ ∈
I d−1 . Then φ ⊗ h is hyperbolic if and only if ed−1 (φ ) ∪ (disc(h)) = 0.
Corollary 2.1.3. Suppose that D is a quaternion algebra, and that σ is of the first
kind and of the orthogonal type. Suppose that cd2 (Γk ) ≤ d. Let h and h be two
hermitian forms over (D, σ ) with dim(h) = dim(h ), and let φ ∈ I d−1 . Then φ ⊗ h
φ ⊗ h if and only if ed−1 (φ ) ∪ (disc(h)) = ed−1 (φ ) ∪ (disc(h )).
Corollary 2.1.4. Suppose that D is a quaternion algebra, and that σ is of the first
kind and of the orthogonal type. Suppose that cd2 (Γk ) ≤ d. Then I d−1 J2 = 0.
Theorem 2.1.5. Suppose that D is a quaternion algebra, and that σ is of the first
kind and of the orthogonal type. Suppose that cd2 (Γk ) ≤ d. Let h ∈ J2 , and let φ ∈
I d−2 . Then φ ⊗ h is hyperbolic if and only if ed−2 (φ ) ∪ C (h) = 0.
Multiples of forms 7
Proof. By Berhuy [6, Th. 13], it suffices to show that ed−2 (φ ) ∪ C (h) = 0 if and
only if en,D (φ ⊗ h) = 0 for all n ≥ 0 (see [6, 2.2] for the definition of the invariant
en,D ). As cd2 (Γk ) ≤ d, we have en,D (φ ⊗ h) = 0 for n > d, so it suffices to check that
ed−2 (φ ) ∪ C (h) = 0 is equivalent with en,D (φ ⊗ h) = 0 for all n = 0, . . . , d. Let k(D)
be the function field of the quadric associated to D. Then D ⊗ k(D) M2 (k(D)), and
hk(D) corresponds via Morita equivalence to a quadratic form qh over k(D). Note that
e2 (qh ) = C (h). Similarly, the hermitian form (φ ⊗ h)k(D) corresponds to a quadratic
form qφ h over k(D), and we have qφ h φ ⊗ qh.
For all n = 0, . . . , d, we have by construction that en,D (φ ⊗ h) = 0 if and only if
en (qφ h ) = 0 (cf. [6, 2.2]). But qφ h φ ⊗ qh , hence
If n < d, then en−2 (φ ) = 0 as φ ∈ I d−2 . We have ed (qφ h ) = ed−2 (φ )∪(C (h)). Hence
en (qφ h ) = 0 for all n ≥ 0 if and only if ed−2 (φ )∪C (h) = 0. This concludes the proof.
Corollary 2.1.6. Suppose that D is a quaternion algebra, and that σ is of the first
kind and of the orthogonal type. Suppose that cd2 (Γk ) ≤ d. Then I d−2 J3 = 0.
for all i = 1, . . . , s. The image of the isomorphism class of the hermitian form hi is
the hermitian form φ ⊗ hi .
If φ ∈ I d , then by 2.1.1 (a) φ ⊗ hi φ ⊗ hi for every hermitian form hi with
dim(hi ) = dim(hi ). This implies that fi is trivial for all i = 1, . . . , s, hence f is trivial.
Therefore f is trivial, which proves (i).
Let us prove (ii). Let us denote by f the composition H 1 (k,UA ) → H 1 (k,UA ) →
H 1 (k,Uφ ). Let Ui be the connected component of the identity in Ui . If σi is unitary or
symplectic, then Ui = Ui . If σi is orthogonal, then by hypothesis ai = 1, so H 1 (Fi ,Ui )
is in bijection with the isomorphism classes of the hermitian (actually, quadratic)
forms over Di = Fi of the same dimension and same discriminant as hi .
Let us denote by fi the composition H 1 (Fi ,Ui ) → H 1 (Fi ,Ui ) → H 1 (k,UMr (Ai ) )
for all i = 1, . . . , s. Suppose that φ ∈ I d−1 . Then 2.1.1 (b) and (c) imply that fi
has trivial image if σi is unitary or symplectic. If σi is orthogonal, then we have
di = 1, therefore by 1.2.3 the map fi is trivial. This implies that f is trivial, and this
completes the proof.
Multiples of forms 9
5 G-Quadratic Forms
Let G be a finite group, and let us denote by k[G] the associated group ring.
A G-quadratic form is a pair (M, q), where M is a k[G]-module that is a finite di-
mensional k-vector space, and q : M × M → k is a nondegenerate symmetric bilinear
form such that
q(gx, gy) = q(x, y) for all x, y ∈ M and all g ∈ G.
10 Eva Bayer-Fluckiger
We say that two G-quadratic forms (M, q) and (M , q ) are isomorphic if there exists
an isomorphism of k[G]-modules f : M → M such that q( f (x), f (y)) = q (x, y) for
all x, y ∈ M. If this is the case, we write (M, q) G (M , q ), or q G q .
If φ is a quadratic form over k, and q a G-quadratic form, then the tensor product
φ ⊗ q is a G-quadratic form.
For any G-quadratic form (M, q), let
q(ex, y) = q(x, f y) for all
A = A(M, q) = (e, f ) ∈ Endk[G] (M) × Endk[G] (M)
x, y ∈ M
be the associated algebra (cf. [1]). Then the unitary group UA can be identified with
the group of automorphisms of (M, q), and the set of isomorphism classes of G-
quadratic forms (M, q ) is in bijection with the set H 1 (k,UA ).
Theorem 5.1.1. Suppose cd2 (Γk ) ≤ d. Let (M, q) and (M, q ) be two G-quadratic
forms.
(i) Let φ ∈ I d . Then φ ⊗ q φ ⊗ q .
(ii) Suppose moreover that G has no quotient of order 2, and let φ ∈ I d−1 . Then
φ ⊗ q φ ⊗ q .
Proof. Part (i) follows immediately from 3.1.1(i). In order to prove part (ii), note
that as G has no quotient of order 2, the group G has no nontrivial orthogonal char-
acters, hence UA = UA . Therefore, 3.1.1(ii) implies part (ii) of the theorem.
Acknowledgements The author was partially supported by the Swiss National Science Fundation,
grant 200020-109174/1.
References
10. W. Scharlau, Quadratic and Hermitian Forms, Grundlehren der Math. Wiss, Springer–Verlag,
Berlin (1985).
11. J-P. Serre, Cohomologie galoisienne, Lecture Notes in Mathematics, Springer–Verlag, Berlin
(1964 and 1994).
12. , Corps locaux, Hermann (1968).
On Saltman’s p-Adic Curves Papers
Eric Brussel
These notes are based on Saltman’s seminal work on the Brauer group Br(K) of
the function field K of a p-adic curve XQ p . In [S1, S2] Saltman showed the index
of an element in the prime-to-p part of Br(K) divides the square of its period, and
that all division algebras of prime degree q = p are cyclic. He also gave a geo-
metric criterion for a class of prime period q = p to have index q. We reprove these
theorems, modulo some of [S2, Sect. 1], expanding some proofs, consolidating some
results, and providing some additional background along the way.
Eric Brussel
Department of Mathematics & Computer Science, Emory University, Atlanta, GA 30322, USA
e-mail: [email protected]
2010 Mathematics subject classification. Primary: 16K20. Secondary: 11S15, 13A20, 16H05,
16K50.
Saltman based his analysis on four observations: First, that while K is the func-
tion field of the curve XQ p over Q p , it is also the function field of a regular model
X of that curve, over Z p ; second, that associated to every α ∈ H2 (K) is a divi-
sor Dα ⊂ X , which may be assumed to have normal crossings; third, that purity
holds for all such X ; and fourth, that the unramified Brauer group H2nr (K) is trivial.
The goal in splitting, then, is to construct a field extension L/K over which α is
unramified:
∂w (αL ) = 0 for all w ∈ V (L).
In particular, we avoid the construction of an explicit model for XL , over which α
would have zero ramification divisor. Constructing L turns out to be tricky, mainly
due to the prospect of divisors in the ramification locus of αL , on any (hypothetical)
model of XL that are centered on closed points of X . The general strategy is to use
local equations for Dα to construct an element f ∈ K, set L = K( f 1/n ), compute the
residue of αL at all w using a local structure theorem (this uses purity), and adjust
f accordingly, if necessary. The methods are extremely valuation-theoretic. If we
can manage αL = 0 and ind(α ) = n is prime, we have cyclicity, by a theorem of
Albert.
It is possible to see why the index should divide the square of the period; in light
of the triviality of H2nr (K), it is traceable to the dimension of X (see also the excel-
lent summary of [S1] in the review [C-T2], which includes additional citations and
context). The existence of a radical splitting field in prime degree is not apparent,
even locally, and the proof is quite technical.
If w’s center on X has codimension one, then it is easy to show w(π1 ) = 0 (mod n).
However, if the center is a closed point, in general w(π1 ) = 0, and since we cannot
expect (−θ0 + θ1 )κ (w) to be zero, the strategy fails in general. We try to salvage it as
follows. Let w0 be a discrete valuation on L extending the valuation v0 on K defined
On Saltman’s p-Adic Curves Papers 15
by π0 , and let α (w0 ) = αKv (θ0 ) be the value of α at w0 . Then α (w0 ) is defined over
0
κ (v0 ), and we can prove
where v̄1 is the valuation induced by the image of π1 in κ (v0 ). If we could zero out
α (w0 ), we could solve the problem for all w. This strategy turns out to be viable.
If α has prime degree, then α (w0 ) is cyclic, and the parameter part of α (w0 ) varies
with f = π0 π1 . Finding an “anti-parameter” u ∈ K, we zero out α (w0 ) by replacing
f with u f . But now, div u f may no longer match Dα and if (div u) ∩ Dα is not
empty, there is a problem.
In general, there are other, more serious problems. Dα may not be principal in
the first place, forcing us to contend with the extra “rogue” components in any div f .
If Dα has more than one nodal point, there is a globalization problem. In the end we
produce f by a type of successive approximation, and at each stage we must check
that the earlier stages have not come undone.
1 Discrete Valuations
Let X be an integral noetherian scheme, and let K be a field containing the func-
tion field F(X) of X . If v is a discrete valuation on K, let Ov ⊂ K denote the cor-
responding discrete valuation ring, pv ⊂ Ov the prime ideal, and κ (v) the residue
field. Spec K is an open subset of Spec Ov , so Spec K → X induces a rational map
Spec Ov X. We say v has a center on X if this is a morphism, and then denote
by xv the image of the closed point Spec κ (v). Thus, v has a center on X if there is a
point xv ∈ X such that OX,xv ⊂ Ov and mxv = pv ∩ OX,xv (see [Liu, Def. 8.3.17]). We
then say v is an X-valuation of K. Denote the set of all normalized X -valuations of
K by
VX (K).
If S is a fixed base, or by default if S = SpecZ, write V (K) instead of VS (K). Sim-
ilarly, we write Vx (K) for those centered on x, and V (X ) for the discrete valuations
on F(X ) that have codimension-one center on X. Substitute κ (x) and A for x and X
when x = Spec κ (x) and X = Spec A. If xv belongs to the set X (1) of codimension-
one points on X, we have a prime divisor
df
Dv = {xv }.
Remark 1.1. If v ∈ Vx (K), then κ (x) ⊂ κ (v), and v ∈ VOX,z (K) for all z ∈ {x}. If
φ : Y → X is a morphism of integral schemes, then Vy (F(Y )) ⊂ Vφ (y) (F(Y )) for all
y ∈ Y , and conversely, if φ is proper then VY (F(Y )) = VX (F(Y )).
Lemma 1.2. Suppose A = (A, m, k) is a two-dimensional regular local domain with
fraction field K, and m = (π0 , π1 ). Suppose f ∈ A, L = K( f 1/n ), and v ∈ V (K)
16 E. Brussel
extends to w ∈ V (L). Then w|K = e(w/v) v, and e(w/v) = n/ gcd(v( f ), n). In partic-
ular, if v( f ) ∈ (Z/n)∗ , then v is totally ramified in L.
e(w/v)
w= v ◦ NLw /Kv
[Lw : Kv ]
2 Residue Map
The material in this section is standard; see [C-T,GMS] or [GS] for additional back-
ground. Let X be a scheme, and let n be invertible on X , meaning that n is prime to
the residue characteristic of κ (x) for all x ∈ X. We write
df ⊗(q−1)
Hq (X ) = Hq (Xét , μn )
resX|Y : Hq (X ) → Hq (Y ).
The map ∂v is called the residue map, the images are called residues, and α ∈ Hq (K)
is said to be unramified at v if ∂v (α ) = 0.
Diagrams 2.4. If L/K is a finite separable extension and w ∈ V (L), then for v = w|K
we have a commutative diagram
∂
0 −−−−→ Hq (Ow ) −−−−→ Hq (L) −−−w−→ Hq−1 (κ (w)) −−−−→ 0
⏐ ⏐ ⏐e(w/v) res
⏐ ⏐ ⏐ κ (v)|κ (w)
∂
0 −−−−→ Hq (Ov ) −−−−→ Hq (K) −−−v−→ Hq−1 (κ (v)) −−−−→ 0.
Let Kv denote the completion of K with respect to v, and let O v be the valuation
q
ring of v on Kv . Then the inflation map H (κ (v)) → H (Ov ) is an isomorphism, and
q
α = α ◦ + (θv , πv )
v via infla-
where α ◦ ∈ Hq (κ (v)) and θv = ∂v (α ) ∈ Hq−1 (κ (v)) are defined over O
tion. The element α ◦ , of course, depends on the choice of πv . If α ∈ H p (Kv ) and
β ∈ Hq (Kv ), then by (2.2) we have the formula
interpreted over Kv .
where Lw (θw )/Lw is the cyclic extension defined by the inflation of the character θw
from κ (w) to Lw . Note in any case α (w) is defined over κ (w)(θw ).
18 E. Brussel
3 Surfaces
We state some well-known facts and definitions. See [Liu, Lbm, Si] for additional
background and proofs.
Remarks 3.2. If X → S is an arithmetic surface, then for each closed point z ∈ X (2) ,
A = OX,z is a two-dimensional regular noetherian local domain, which is factorial by
Auslander-Buchbaum’s theorem ([Mat, 20.3]). If S is excellent, then X is excellent,
and A is excellent. Since X → S is proper, V (K) = VX (K) by the valuative criterion
for properness.
If X is a normal projective flat S-curve, and S is a one-dimensional Dedekind
scheme, then the generic fiber XK is a nonsingular integral curve over K, and each
closed fiber Xs , s ∈ S, is a projective curve over κ (s) ([Liu, 8.3.3]). If X → S is
arithmetic, then each Xs → κ (s) is a local complete intersection ([Liu, 8.3.6]), hence
Xs has no embedded points. Xs is generally not reduced. Every effective irreducible
divisor D ⊂ X is either a component of some Xs (D is vertical), or the closure of a
closed point x ∈ XK of the generic fiber (D is horizontal).
div f = ∑ v( f )Dv ,
V (X)
the sum being finite since X is noetherian. If D and E are two effective divisors with
no common irreducible components, the intersection multiplicity (D·E)z at a closed
point z is defined to be
df
(D · E)z = lengthOX,z OX,z /(ID,z + IE,z )
where ID and IE are the ideal sheaves, and the subscript z means “localized at z”. If
D is prime and g ∈ K is a local equation for E at z, then (D · E)z = vz̄ (g), where vz̄
is the valuation on D induced by z.
X if it has normal crossings at every point. If D has normal crossings, we will also
say that the set {vi } of discrete valuations centered on the generic points of D has
normal crossings.
This definition, from [Liu, 9.1.6], is different from the relative version in [SGA1,
XIII, 2.1.0]. We obtain that version by replacing “normal crossings” by “strictly
normal crossings” and “smooth over” by “regular”. A normal crossings divisor need
not be reduced, but its irreducible components are nonsingular, and meet each other
transversally, meaning their local equations form part of a system of parameters at
each intersection point.
Existence of a Model 3.5. Let S be an excellent one-dimensional affine Dedekind
scheme, with function field F, and let K be a field extension of F of transcendence
degree one. Then there exists a normal connected projective curve XF with function
field K, and a normal projective flat S-curve X with generic fiber XF . If F is perfect,
XF is smooth over F.
In fact, there is a category equivalence between normal connected projective
curves over F and function fields of transcendence degree one over F, where the
morphisms between curves are dominant morphisms ([Liu, 7.3.13]). For curves,
normal and regular are the same, and if F is perfect, regular and smooth are the
same ([Liu, 4.3.33]), so XF is smooth if F is perfect. The rest is stated in [Liu,
10.1.6].
Strong Desingularization 3.6. Let X be a two-dimensional, excellent, reduced,
noetherian scheme. Then X admits a desingularization in the strong sense, i.e., a
proper birational morphism f : X → X with X regular, and f an isomorphism
above every regular point of X . In particular, this holds for X a projective flat
S-curve over an excellent Dedekind scheme. If S is affine, X → S is then an arith-
metic surface.
This is a theorem of Lipman ([Liu, 8.3.44]). To prove the last statement, note that
since f is birational, X is integral, and X → S is flat, the induced map X → S is
flat, since X dominates S ([Liu, 4.3.9]). Since S is affine, f is projective by [Liu,
8.3.50], and since the composition of projective morphisms is projective, X → S is
an arithmetic surface.
Embedded Resolution 3.7. Let X → S be an arithmetic surface over an excellent
Dedekind scheme, and let D be a divisor on X. Then there exists a projective bi-
rational morphism f : X → X with X an arithmetic surface, such that f ∗ D has
normal crossings.
This is [Liu, 9.2.26]. We will need the following lemma from the proof.
Blowup Lemma 3.8. Suppose X → S and D are as in the statement of Embedded
Resolution (3.7), f : X → X blows up a closed point z ∈ X, and D has normal cross-
ings at z. Then X → S is an arithmetic surface, and f ∗ D has normal crossings at
every point of the exceptional fiber E. Moreover, E meets the irreducible compo-
nents of the strict transform D in at most two points, and these points are rational
over κ (z).
20 E. Brussel
Proof. By Existence of a Model 3.5 there exists a normal projective flat S-curve
X → S with function field K. Therefore by Strong Desingularization 3.6, there ex-
ists an arithmetic surface X → S with function field K. Suppose α ∈ H2 (K). Then
by Embedded Resolution (3.7), there exists an arithmetic surface X → S and a pro-
jective birational morphism f : X → X such that f ∗ (Dα ∪ Xs ) has normal crossings,
where Dα is the divisor of α on X . Since the irreducible components of f ∗ (Dα ∪Xs )
equal those of Dα ∪ Xs , this proves the result.
Purity for curves 4.1. Let C be a smooth integral projective curve over a field k,
with function field F. Then
This is proved in the affine case for the Brauer group in [Ho, Th.], and extended
using the fact that Br is a Zariski sheaf, as in [S1, Th. 1.4].
Let X be a two-dimensional regular noetherian scheme. Then by [GB, II, 2.7],
Br(X) = H2 (X , Gm ). Let
2 df
H̄ (X ) = n H2 (X , Gm )
denote the n-torsion subgroup, where n is as in Sect. 2. For any regular quasi-
compact integral scheme X with function field K, the natural map H2 (X , Gm ) →
2
H2 (K, Gm ) is injective [M, III.2.22]. Thus we may view H̄ (X ) as the image of
H (X ) in H (K). Since H (X ) → H (K) factors through H (Ov ) for any v ∈ VX (K),
2 2 2 2 2
we have
2
H̄ (X ) ⊂ H2 (Ov ).
VX (K)
On Saltman’s p-Adic Curves Papers 21
Using Auslander–Goldman’s theorem [AG, Prop. 7.4] and the fact that Br is a
Zariski sheaf, it is not hard to prove the following ([GB, II, Prop. 2.3], or [S1,
Th. 1.4]).
H̄ (X ) = H2 (Ov ).
V (X)
2
If X is projective over S = SpecR, then H̄ (X ) = H2nr (K/R).
We will use purity for surfaces in two ways, first to show that when S = SpecZ p ,
to split α ∈ H2 (K) it is sufficient to construct a finite separable extension L/K over
which ∂w (α ) = 0 for all w ∈ V (L) (Theorem 4.5); and second, to prove a local
structure theorem for H2 (K), which provides the computational foothold we need
in order to construct such an L locally.
Hq (Y ) = Hq (Y0 ).
This is a corollary to the proper base change theorem ([GB, III, 3] or [M, VI.2.7]).
This is [S1, Lemma 3.2], and it is also in [GB, III]. The following fundamental
result is [S2, Th. 0.9].
Theorem 4.5. If L is the function field of a p-adic curve, and n is prime-to-p, then
H2nr (L) = 0. Thus if αL ∈ H2 (L), and ∂w (αL ) = 0 for all w ∈ V (L), then αL = 0.
H2 (C) = H2 (Ov )
V (F)
22 E. Brussel
(see [M, III.2.22(g)] and [GB, III, Remark 2.5b]). We conclude H2 (C) = 0 by (2.1),
hence H2 (Y0 ) = 0, as desired.
We briefly summarize [S2, Sect. 1], which shows how, on an arithmetic surface over
a complete discrete valuation ring, we can represent a divisor class whose restriction
to the closed fiber is an n-th power by a divisor that is itself an n-th power, while
avoiding a fixed finite set of closed points. The results of this section will not be
applied until the very end.
Let X be a projective scheme with no embedded points over an affine scheme.
This hypothesis applies to an arithmetic surface over an affine Dedekind scheme, or
one of its closed fibers. Let z = {zi }I ⊂ X be a finite
set of closed points, and let
ι : z → X be the closed immersion so that ι∗ O∗z = I ιi∗ κ (zi )∗ . Let z O∗X denote the
sheaf of units with value 1 at each zi , defined by the exact sequence
Set
Div X = H0 (X , KX∗ /O∗X ) Pic X = H1 (X , O∗X )
z Div X = H0 (X , KX∗ /z O∗X ) z Pic X = H1 (X , z O∗X )
z Div X = H0 (X, KX∗ /z O∗X ) is the set of {(U j , f j )} such that not only is f j / fk a unit
on U j ∩Uk , but ( f j / fk )(z) = 1 for each z ∈ z ∩ (U j ∩Uk ). By definition, we have an
exact sequence
Since ι∗ O∗z and KX∗ are flasque, by (5.1) we have an exact diagram
zK
∗ −−−−→ z Div X −−−−→ z Pic X −−−−→ 1
⏐ ⏐
⏐ ⏐
zK
∗ −−−−→ Div X −−−−→ Pic X −−−−→ 1 (5.3)
⏐ ⏐
⏐ ⏐
1 1
K∗
where z ≤ H 0
is the inverse image of the subgroup z Div X ≤ z Div X
(X , KX∗ )
∗
in H (X , KX ). These are the meromorphic functions whose values at z ∈ z are
0
invertible.
The long exact sequence associated to (5.2) induces a split exact sequence
1 −→
z
κ (zi )∗ −→ z Div X −→ Div X −→ 1.
The injection splits via the evaluation map, which is well defined since ( f j / fk )(z) =
1 for all z ∈ z ∩ (U j ∩Uk ), whenever {(U j , f j )} ∈ z Div X . We obtain
z Pic X κ (zi )∗ ⊕ Div X /z K ∗
z
(5.4)
where f ∈ z K ∗ maps to the pair (( f (zi ))z , {(X , f )}). This is [S2, Prop. 1.6]. Thus,
z Pic X comprises the classes of divisors on X whose support avoids z, paired with a
tuple of nonzero values at each z ∈ z.
Proposition 5.5. Suppose S = Spec R for a complete discrete valuation ring R, and
X → S is an arithmetic surface such that X0 has normal crossings. Let n be a number
invertible on X. Then the canonical map induces an isomorphism
∼
z Pic X/n −
→ z̄ Pic((X0 )red )/n.
This is [S2, Prop. 1.7]. The proof uses standard algebraic geometry facts and an
elementary version of Grothendieck’s Existence Theorem [Liu, 10.Exercise 4.4]. If
z = ∅, we obtain Pic X/n Pic(X0 )red /n, which is [S2, Th. 1.2].
24 E. Brussel
Standing setup 6.1. We need to consider splitting at the three types of closed points
that may occur when D has normal crossings on an arithmetic surface. We say z is
a distant point of D if it does not belong to D, a curve point of D if it belongs to
exactly one component of D, and a nodal point of D if it belongs to two. We will
use the following notation for the rest of the paper.
• S = Spec R is an excellent one-dimensional affine Dedekind scheme.
• X → S is an arithmetic surface with function field K = F(X ).
• X0 is a closed fiber of X → S.
• z is a closed point on X.
• A = (A, m, k) is the local ring OX,z , a two-dimensional excellent regular noethe-
rian local ring with fraction field K.
• π0 , π1 are two (regular) primes of A generating m.
• v0 , v1 ∈ V (A) are the discrete valuations corresponding to π0 , π1 .
• n is a number relatively prime to char κ (x) for all x ∈ X .
• α ∈ H2 (K) = H2 (K, μn ).
• θv = ∂v (α ) ∈ H1 (κ (v)) for each v ∈ V (K).
• Dα is the ramification divisor of α on X.
• D+α = Dα ∪ (X0 )red .
• If Dα has normal crossings and α ramifies on A, then divA (α ) = {v0 } if z ∈ Dα
is a curve point, and divA (α ) = {v0 , v1 } if z ∈ Dα is a nodal point.
The excellent hypothesis on R induces excellence on A and X , and this is needed
for Lipman’s embedded resolution theorem, to produce an arithmetic surface on
which α ramifies at at most two normal crossings divisors.
Exactness at H2 (K) is Purity for Surfaces 4.2, and at H2 (A) it is [AG, Th. 7.2].
6.3. Assume Setup 6.1, and suppose v ∈ V (A), p = pv ∩ A, and Rv = A/p ⊂ κ (v). Let
z be the closed point of SpecA, and z̄ its image on SpecRv . Let R v denote the nor-
malization of Rv in κ (v), a one-dimensional semi-local normal noetherian ring. Let
n1 , . . . , nd denote the primes of R v extending the image m̄ of m in Rv , let w1 , . . . , wd
denote the corresponding discrete valuations on κ (v), let R i be the localization of
R v at ni , and let ki = R i /ni . Define
d
∑ corki |k ◦∂wi : Hq (κ (v)) → Hq−1(k)
df
∂v,z̄ =
i=1
∑ ∂v,z̄ :
df
∂z = Hq (κ (v)) → Hq−1 (k)
V (A) V (A)
On Saltman’s p-Adic Curves Papers 25
Let mi = |μn (ki )| and fix a primitive n-th root of unity ζn . Then
Exactness at the direct sum is [S3, Th. 5.2, Th. 6.12], and the rest is (6.2).
7 Computations
Hypothesis 7.1. Assume Setup 6.1, and Dα has normal crossings at z. Then (a) μr ⊂
A, where r = |∂z (θv0 )|, and (b) the map H1 (A) → H1 (A/(πi )) is onto, for i = 0, 1.
Proposition 7.2. In Setup 6.1, assume Dα has normal crossings at z. Then Hypoth-
esis 7.1(b) holds if μn ⊂ K, or if n is prime. In particular, if μn ⊂ K, or if n is prime
and ∂z (θv0 ) = 0, then Hypothesis 7.1 holds.
is onto by base change. We claim H1 (A ) → H1 (Ri ) is onto. Since A and Ri are
factorial, and each contains n-th roots of unity, we have compatible isomorphisms
H1 (A ) (A )∗ /n and H1 (Ri ) (Ri )∗ /n, and it suffices to show (A )∗ → (Ri )∗
is onto. It then suffices to show every nonunit of A maps to a nonunit Ri . But
A → A ⊗A A/p is surjective by base change, and pA ⊂ mA is contained in ev-
ery maximal ideal of A , since A → A is étale. Thus each nonunit of A maps to a
nonunit of A /pA , so (A )∗ → (Ri )∗ is onto, which proves the claim.
If θi ∈ H1 (Ri ), then cor◦ res(θi ) = (n − 1)θi . Let θ ∈ H1 (A ) map to θi = res(θi ).
Then cor(θ ) ∈ H1 (A) maps to (n − 1)θi by the compatibility of corestriction and
base change, and since n − 1 is invertible (mod n), we conclude θi has a preimage
in H1 (A). Therefore H1 (A) → H1 (Ri ) is onto. Therefore (7.1(b)) holds.
The next result gives us our main computational foothold, the second main ap-
plication of Purity for Surfaces (4.2).
Local Structure Theorem 7.3. Assume Hypothesis 7.1. Let ζm∗ be as in (6.3), with
m = |μn (K)|. Then there exist elements α ◦ ∈ H2 (A) and θi ∈ H1 (A) such that
where sζm∗ = ∂v0 ,z̄ (θv0 ) = −∂v1 ,z̄ (θv1 ), for θvi = ∂vi (α ). If z is a curve point of Dα ,
θ1 = 0 and s = 0.
Proof. Since {v0 , v1 } has normal crossings, we have ∂v0 ,z̄ (θv0 ) + ∂v1 ,z̄ (θv1 ) = 0 by
Theorem 6.4, hence ∂v0 ,z̄ (θv0 ) = −∂v1 ,z̄ (θv1 ) = sζm∗ , for some 0 ≤ s < m. Of course,
if z is a curve point for α then θv1 = 0 and s = 0.
Since {v0 , v1 } has normal crossings, v1 defines a discrete valuation on κ (v0 )
with residue field k. Let Ri = A/(πi ). By Hypothesis 7.1, μr ⊂ A, where r = |sζm∗ |.
The choice of π1 then determines a splitting H0 (k, μr−1 ) → H1 (κ (v0 ), Z/r) ≤
H1 (κ (v0 ), Z/n) of (2.1), so that θv0 = θv◦0 + sζm∗ · (π̄1 ) for some θv◦0 ∈ H1 (R0 ). Since
H1 (A) → H1 (Ri ) is onto by (7.1), θv◦0 lifts to some θ0 ∈ H1 (A). Choose θ1 similarly,
and set
β = (θ0 , π0 ) + (θ1 , π1 ) + (sζm∗ · (π1 ), π0 ).
Then ∂v (α ) = ∂v (β ) at all v ∈ V (K) by (2.2), so α ◦ = α − β is in H2 (A) by Purity
for Surfaces (4.2).
Lemma 7.4. Assume Hypothesis 7.1. Then in the notation of Theorem 7.3, for each
v ∈ V (K),
∂v (α ) = a(θ0 + sζm∗ · (π1 )) + b(θ1 − sζm∗ · (π1 )) + absζm∗ · (−1) κ (v)
where the terms are interpreted in Kv , a = v(π0 ), and b = v(π1 ). If v ∈ Vz (K), then
∂v factors through k.
Proof. Since we assume (7.1) we can apply Theorem 7.3. Using (2.2) and formula
(2.5), we compute directly over Kv
On Saltman’s p-Adic Curves Papers 27
If v ∈ Vz (K) then k ⊂ κ (v) by (1.1), and since the terms in ∂v (α ) are originally
defined over A, over κ (v) they are defined over k. Thus ∂v factors through k.
The following simple computation gives us the main tool we need for splitting
α ’s residues.
Lemma 7.5. Assume Hypothesis 7.1, and
not divide u0 and v ∈ V (A), we compute similarly w(u0 ) = 0( mod n). In any case
c0 ∂w (α ) = 0, and since c0 ∈ (Z/n)∗ , we conclude ∂w (α ) = 0 by (b).
In (d), if z is a curve point of div f then Dα = div π0 , θ1 = 0, s = 0, and u0 is a
unit of A. Since c0 ∈ (Z/n)∗ this implies ∂w (α ) = 0 by (b). Part (e) is immediate by
Lemma 7.5 and (2.2).
In (f), each wi /vi is totally ramified of degree n by Lemma 1.2, hence κ (wi ) =
κ (vi ), and αLwi = α (wi ) by definition, since ∂wi (α ) = 0 by (2.4). Using the compu-
tation of c0 αL and −c1 αL in (7.5), we find the computation of ∂vi ,z̄ (α (wi )) is now
simple, and we obtain c0 ∂v0 ,z̄ (α (w0 )) = −c1 ∂v1 ,z̄ (α (w1 )). Note (−1) has order 2,
so if n is odd, then sζm∗ · (−1) = 0, and if n is even, then both c0 and c1 are odd, so
(c0 + c1 )sζm∗ · (−1) = 0 in any case.
The following elementary lemma, which appears as [S1 , Lemma], was supplied
by Colliot-Thélène.
Global Splitting Lemma 7.8. Assume Setup 6.1 and μn ⊂ K. Then there exists a
field extension L/K of degree n2 such that ∂w (αL ) = 0 for all w ∈ V (L).
Proof. Since μn ⊂ K, (7.1) holds, by Proposition 7.2. Since every w ∈ V (L) has
a center on X , by Purity for Surfaces 4.2 it is enough to construct an L such that
∂w (α ) = 0 for all w ∈ VA (L), for all A = (A, mz , κ (z)). Apply Lemma 7.7 to choose
f ∈ K such that
div f = Dα + Dε
where Dε contains none of the (finitely many) self-intersection points of Dα . Let L
be an extension containing K( f 1/n ), w ∈ VA (L). By Corollary 7.6(a), (c), and (d),
we may assume z lies on Dα , on more than one irreducible component of div f , and
that w ∈ Vz (L). By Lemma 7.5, αK( f 1/n ) has the form
αK( f 1/n ) = α0 + (λ , π0 )
On Saltman’s p-Adic Curves Papers 29
where α0 and λ are defined over A. Since A contains the n-th roots of unity, λ =
ζm∗ · (uz ) for some uz ∈ A∗ by Kummer theory. By Lemma 7.7, there exists g ∈ K ∗
such that g ∈ A and g = uz (mod mz ), for each z. Let
L = K( f 1/n , g1/n ).
We compute ∂w (α ) = w(π0 )λκ (w) . Since w ∈ Vz (L), κ (z) ⊂ κ (w) by (1.1). Since λ
is defined over A, its restriction to κ (w) factors through κ (z), and by construction
λκ (w) = 0, since (uz )κ (w) = 0. Therefore, ∂w (α ) = 0 for all w ∈ V (L).
Proof. Recall ind(α ) is the smallest degree of any separable splitting field, and
per(α ) is the order of α in Br(K). Suppose α ∈ H2 (K) has period n. If μn ⊂ K and
L is as in Lemma 7.8, then αL = 0 by Theorem 4.5. If n is prime, then K(μn )/K has
prime-to-n degree, and we obtain the result by the standard restriction-corestriction
argument. In the general case, we induct on the prime degree case. For if q is a prime
dividing n, and qα , which has period n/q, has index dividing (n/q)2 , then there is a
separable extension F/K of degree (n/q)2 such that qαF = 0, hence αF has period
q, so by the base case there is an extension L/F of degree q2 splitting α , and now
L/F has degree n2 .
Remark 8.3. In Saltman’s terminology ([S2, p. 821]), α (w) is split by the ramifica-
tion whenever L/K is v-totally ramified and α (w) ∈ H2 (κ (v)) is split by κ (v)(θv ).
Since we are calling the images of ∂v “residues” instead of “ramification”, we would
say instead that α (w) is split by the residue. When α is semiramified at v, the values
of α are split by the residue at all v-totally ramified L/K. This happens, for example,
whenever α has prime index and Dα = ∅.
Nodal points 8.4. Assume Hypothesis 7.1. In the notation of Theorem 7.3,
If L = K( f 1/n ) for f = uπ0 π1c and u ∈ A∗ , with c ∈ (Z/n)∗ , then by Corollary 7.6(b),
where if θvi = ∂vi (α ), we have sζm∗ = ∂v0 ,z̄ (θv0 ) = −∂v1 ,z̄ (θv1 ). If s = 0, then θvi (z̄) =
θi (z) ∈ H1 (k). This guides the following classification scheme:
Definition 8.5. Suppose that z is a nodal point for α , and divA (α ) = {v0 , v1 } has
normal crossings. In the notation of (8.4), we say that, with respect to Dα , z is a
1. Cold point if s = 0.
2. Cool point if s = 0 and θ0 (z) = θ1 (z) = 0.
3. Chilly point if s = 0 and θ0 (z) = θ1 (z) = 0.
4. Hot point if s = 0 and θ0 (z) = θ1 (z).
If z is a chilly point, the coefficient of z with respect to v0 is the number c such that
θ1 (z) = c θ0 (z).
The next lemma shows that when α is semiramified we can ignore hot points.
Lemma 8.6. Assume Hypothesis 7.1 and the notation of (8.4), and suppose s = 0.
Let vz̄ denote a discrete valuation extending vz̄ from κ (v0 ) to κ (v0 )(θv0 ). Then
Proof. Note κ (v0 ) has a discrete valuation vz̄ induced by v1 since {v0 , v1 } has nor-
mal crossings. Since θv0 is vz̄ -unramified, any extension vz̄ from κ (v0 ) to κ (v0 )(θv0 )
measures vz̄ (π1 ) = 1, and by definition, the residue field of κ (v0 )(θv0 ) with respect
On Saltman’s p-Adic Curves Papers 31
to vz̄ is k(θ0 (z)). Since we assume (7.1), we can invoke Theorem 7.3, which implies
α (v0 ) = (α ◦ + (θ1 , π1 ))κ (v0 )(θv ) , where θ1 = 0 if z is a curve point for Dα . We then
0
compute directly ∂vz̄ (α (v0 )) = θ1 (z)k(θ0 (z)) , as desired. If α is semiramified at v0
and v1 , then α (v0 ) = α (v1 ) = 0, hence θ1 (z) ∈ θ0 (z) and θ0 (z) ∈ θ1 (z), hence
θ0 (z) = θ1 (z).
Chilly loops present the following obstacle. Suppose D = ∑m i=1 Di is a chilly loop,
where each Di is prime. Let ci j be the coefficient of zi j = Di ∩ D j with respect
to vi so that θi (zi j ) = ci j θ j (zi j ). If L = K( f 1/n ) where vi ( f ) = ci ∈ (Z/n)∗ and
c j /ci = ci j , then we compute ∂w (α ) = 0 for all w ∈ Vzi j (L) by Corollary 7.6(b). But
then proceeding around the loop we find ∏i (mod n) ci j = 1. There is no reason to
expect such a relation from a general set of residues {θi }, so such an f does not
exist in general. We then resort to the strategy of zeroing out the values α (wi ), or at
least showing they are unramified, cf. Proposition 8.11(C) below. However, we find
this strategy is also blocked, by the same compatibility problem! The solution is to
eliminate chilly loops, as follows.
Lemma 8.9. Assume α ∈ H2 (K) has prime period, and is semiramified at every
v ∈ divK (α ). This holds, for example, if α has prime index. Then there exists an
arithmetic surface X with function field K on which D+
α is well-conditioned.
Proof. By (3.9), there exists an arithmetic surface X on which D+ α has normal cross-
ings. Since α is semiramified at each v ∈ divA (α ), Dα has no hot points by Lemma
8.6. Since D+ α has finitely many irreducible components, there can only be finitely
many cool points or chilly loops.
Let f : X̃ → X be the blow-up centered at z, and let v ∈ Vz (K) be the valuation
defined by the exceptional divisor. Then by Lemma 3.8, X̃ is an arithmetic surface,
isomorphic to X away from z, and D̃+ α ⊂ X̃ has normal crossings.
Suppose z = D0 ∩ D1 is a cool point with respect to Dα . Since per(α ) is prime,
Hypothesis 7.1 holds, by Proposition 7.2. Since v is centered on z, by Lemma 7.4,
we compute ∂v (α ) = v(π0 )θ0 (z) + v(π1 )θ1 (z) = 0, by the definition of cool point.
Since D0 and D1 intersect transversally, they do not intersect on X̃ , so the cool point
z is transformed into two curve points on X̃. In this way, we eliminate all cool points
with finitely many blow-ups.
Suppose z = D0 ∩ D1 is a chilly point in a chilly loop, with coefficient c with
respect to v0 . Again Hypothesis 7.1 holds. Since v is the valuation of the exceptional
divisor, v(π0 ) = v(π1 ) = 1, and ∂v (α ) = θ0 (z) + θ1 (z) = (c + 1)θ0 (z) by Lemma
7.4. If c + 1 is a multiple of n, then ∂v (α ) = 0, and we have replaced z with two
curve points on X̃. Otherwise we have replaced z with two chilly points, and one
32 E. Brussel
Since q is prime, the field extensions have prime-to-q degree, hence the vertical
restriction arrows are injective by the standard restriction-corestriction argument.
Moreover, since A → A is flat, each v ∈ V (A ) lies over some v ∈ V (A), so that
the residues of α = resK|K (α ) at the v ∈ V (A ) are exactly those restricted from
residues of α at v ∈ V (A). Since m = mA = (π0 , π1 ), Dα has normal crossings at
z = Spec(k ) on Spec A , and Hypothesis 7.1 holds for α on A by Proposition 7.2.
Thus the classification of z in Definition 8.5 is identical to that of z, should z be a
nodal point of Dα .
Note the splitting of residues at z can now be computed at z . For if L/K is a finite
separable field extension of degree q, L = L ⊗K K , and ∂w αL = 0 for w ∈ V (L )
extending w ∈ V (L), then ∂w (α ) = 0 by (2.4), the injectivity of the restriction, and
the fact that L /L is unramified. If L/K is v0 -totally ramified, w0 ∈ V (L ) extends
v0 to L and restricts to w0 and v0 on L and K , then L /K is v0 -totally ramified,
α (w0 ) and α (w0 ) are defined over κ (v0 ) and κ (v0 ), respectively, and α (w0 ) =
resκ (v0 )|κ (v ) (α (w0 )) by the compatibility of restriction maps. Then, since restric-
0
tion commutes with the residue map (2.4), ∂v ,z̄ (α (w0 )) = resk|k (∂v0 ,z̄ (α (w0 )).
0
Proposition 8.11. In Setup 6.1, assume α ∈ H 2 (K) has prime period n, D+α is well-
conditioned, and Dα = ∑I Di , where I indexes the prime divisors Di . Suppose L =
K( f 1/n ), where f ∈ K satisfies
A. div f = cI Dα + Dε , where cI = (ci )I , for ci ∈ (Z/n)∗ , and Dε contains no com-
ponents or nodal points of Dα .
On Saltman’s p-Adic Curves Papers 33
Proof. Assume f satisfies (A), (B), and (C), and set A = OX,z . If z is a cold point and
A → A is as in (8.10), then the hypotheses transfer to A by local parallel transport,
and the conclusion, which is verified locally, descends back to K. Therefore we may
assume μn ⊂ K, so that Hypothesis 7.1 holds, by Proposition 7.2. Let w ∈ VA (L), and
let v be the normalized valuation corresponding to w|K . By Purity for Surfaces (4.2),
it is enough to show ∂w (α ) = 0, and invoking (A), we may assume by Corollary
7.6(a), (c), and (d) that z ∈ Dα , w ∈ Vz (L), and z lies on more than one component
of div f .
Suppose z is a curve point of Dα , so divA (α ) = {v0 }. By (A) we have f = u0 π0 0 ,
c
For any α ∈ H2 (K) of prime index relatively prime to all char k, there exists an
arithmetic surface X → S on which α ’s ramification locus is well-conditioned, by
Lemma 8.9. Proposition 8.11 provides us with a blueprint for splitting the residues
of such a class, under Hypothesis 7.1. We want to produce an element f ∈ K ∗ satis-
fying
A. div f = cI Dα + Dε , where cI = (ci )I , for ci ∈ (Z/n)∗ , and Dε contains no com-
ponents or nodal points of Dα .
B. If (Di · Dε )z = 0 (mod n) for some i ∈ I, then θvi (z̄) = 0.
C. ∂vi ,z̄ (α (wi )) = 0 whenever z is a nodal point of Dα on Di , and wi ∈ V (L)
extends vi .
We now carry out the program when K is the function field of a p-adic curve. The
following lemma is the technical heart of the computation (see [S2, Th. 4.6]).
α (w0 ) = (−c−1
0 θv0 ,t0 ) ∈ H (κ (v0 ))
2
θv0 is z̄-unramified and vz̄ (t0 ) ∈ (Z/q)∗ . Therefore we are finished at cold points,
since for them θv0 is z̄-totally ramified. At chilly points, we compute directly
∂v0 ,z̄ (α (w0 )) = −c−10 vz̄ (t0 )θv0 (z̄). But α (w0 ) = αLw0 by Corollary 7.6(f), and since
s = 0 and −c1 θ0 (z) + c0 θ1 (z) = 0 by the choice of the ci , by Corollary 7.6(b) we
have ∂v0 ,z̄ (α (w0 )) = 0. Since θv0 (z̄) is nontrivial (definition of chilly point), we con-
clude vz̄ (t0 ) = 0 (mod q). Therefore there is no problem at chilly points, and we
have uz for all z ∈ N .
Suppose z ∈ D0 is in Q. By the choice of f , Dε avoids z, so f has only one
irreducible component going through z, and since by definition f = t0 πvc00 , again
vz̄ (t0 ) = 0 (mod q), and there is no problem. Thus each ti is a unit at all z ∈ N ∪ Q,
and we have (a).
Suppose z ∈ N is on D0 ∩ D1 . Since t0 and t1 are units at z, each has an
image t¯i in κ (z). We show that we may assume t¯0 = t¯1 . If z is a chilly point,
then c1 θ0 (z) = c0 θ1 (z), so α (w0 )(z̄) = α (w1 )(z̄) by Corollary 7.6(e). That is,
(−c−1 −1 −1 −1
0 θ0 (z), t0 ) = (−c1 θ1 (z), t1 ). Since c1 θ1 (z) = c0 θ0 (z), t0 and t1 differ by
¯ ¯ ¯ ¯
a norm from κ (z)(θ0 (z)) to κ (z). If z is a cold point, then by Corollary 7.6(f) (ap-
plying (8.10) if necessary),
We have c0 α (w0 ) = (−θv0 ,t0 ). Writing θv0 = θv◦0 + sζm∗ · (π1 ) over κ (z)(v0 )z̄ , and
noting that vz̄ (t0 ) = 0 by (a), we compute
and similarly −c1 · ∂v1 ,z̄ (α (w1 )) = sζm∗ · (t1 )κ (z) . Since sζm∗ has order q, we conclude
t¯0 = t¯1 (mod κ (z)q ). In any case, whether z is a chilly or cold point, using (standard)
weak approximation we may adjust t0 by a norm from κ (v0 )(θv0 ) to κ (v0 ) so that
t¯0 = t¯1 simultaneously at each z ∈ N on D0 . We may apply the same procedure on
D1 without affecting the result on D0 , and by induction we may arrange so that the
t¯i are all equal, at all z ∈ N . This proves (b), and completes Step 1.
Let {pi }I denote the set of prime ideals of T corresponding to the irreducible
components of Dα , and let {m j }J denote the set of maximal ideals corresponding
36 E. Brussel
α (wi ) = (−c−1 −1
i θvi ,ti /t) = (−ci θvi , 1) = 0
where wi is an extension of vi to L = K(( f /t)1/q ). Thus the claim will give
us (C).
We prove the claim, as in [S2, Prop. 4.5]. Order I ∪ J, and set a = n−1 qi , the
i=1
ideal of the scheme n−1i=1 V (q i ). The map (a, b) →
a − b defines an exact sequence
We sketch this argument. The sequence is easily seen to be a complex, and the
injectivity and surjectivity are immediate. To prove exactness in the middle, we
may assume T = Tm for some maximal ideal m, and then T is factorial, since X
is regular. If a + qn = T , the result follows by the Chinese Remainder Theorem. If
a + qn = T then all of the qi for i ≤ n are (regular) principal primes, so a = (xc ) and
qn = (y) for some c > 0 and primes x and y, and since Dα has normal crossings,
m = (x, y). If the pair (a + a, b + qn) maps into the ideal (xc , y) then we may assume
b = uxd for some d ≥ 0 and u ∈ T ∗ , and a = uxd + sye , for e > 0 and s ∈ T , and then
(a + a, b + qn) is in the image of t = uxd + sye , proving exactness in the middle.
Now induct on the cardinality of I ∪ J, the base case n = 1 being trivial. Assume
there exists a t ∈ T such that t = ti (mod qi ) for i ≤ n − 1 in I ∪ J. If qn = mn , i.e.,
V (qn ) = zn ∈ P, or if qn = pn for an isolated component Dn (of Dα ), or if V (a) = ∅,
then V (a + qn ) = ∅, and since T /(a + qn) = 0, by exactness of the above sequence
there exists a t ∈ T mapping to (t ,tn ) ∈ T /a ⊕ T /qn , and then t = ti (mod qi ) for
each i ≤ n. If qn = pn and V (a + qn ) = ∅, then since Dα has normal crossings,
V (a + qn ) is a finite subset of N . If z ∈ Di ∩ Dn is one of these points, then ti = tn
(mod mz ) by Step 1. Therefore (t ,tn ) maps into each mz containing a+qn . Again by
the normal crossings hypothesis it follows that (t ,tn ) maps to zero in T /(a + qn ),
and by exactness again we find t ∈ T mapping to (t ,tn ) so that t = ti (mod qi )
for each i ≤ n. By induction, there exists t ∈ T such that t = ti (mod qi ), for each
i ∈ I ∪ J, as claimed. This completes the proof of Lemma 9.1.
Problematic curve points 9.2. We call the z ∈ Dε ∩ Dα , which appear in (B), prob-
lematic curve points. If we choose f ∈ K as in Lemma 9.1, then we have (A) and
(C). There is no reason why (B) should follow, though we have the following.
On Saltman’s p-Adic Curves Papers 37
Proposition 9.3. Assume the hypotheses of Lemma 9.1, f ∈ K satisfies (A) and (C),
X is as in Lemma 9.1, and D0 ⊂ Dα is an irreducible component. If z ∈ D0 ∩ Dε is
a problematic curve point, then either θv0 (z̄) = 0 or (D0 · Dε )z is a multiple of q.
Proof. Write f = u0 π0c0 , where u0 ∈ O∗v0 , and set L = K( f 1/q ). Since c0 ∈ (Z/q)∗
by (A), and θv0 has order q, αLw0 = α (w0 ) and κ (w0 ) = κ (v0 ) by Corollary 7.6(f).
By Lemma 7.5, c0 α (w0 ) = c0 α ◦ − (θ0 , u0 ), hence
We still do not have (B): (D0 · Dε )z being a multiple of q only helps if Dε has
only one irreducible component passing through z; as it stands, Dε could have up to
q distinct components all intersecting at z. To resolve this problem we will invoke
Proposition 5.5, and for this we now take S = SpecR, where R = Z p .
Proof. Let z denote the set of all nodal points of D+ α , and write z̄ for the image of
z on X0 . Choose f as in Lemma 9.1, and write div f = cI Dα + Dε . We will show
that if z ∈ Dα ∩ Dε and (D0 · Dε )z is a multiple of q, then we may assume Dε is a
q-multiple at z. Then f satisfies (B) by Proposition 9.3.
By Lemma 9.1, Dε contains no irreducible component of X0 , so all of its compo-
nents are horizontal. Since S is henselian, by [Liu, 8.3.35] there exists a divisor Dgood
that is the sum of horizontal divisors, with correct multiplicities, cutting out exactly
those points of Dε ∩(X0 )red of multiplicity not divisible by q. Let Dbad = Dε − Dgood.
The restriction (Dbad )0 to the closed fiber is an effective divisor that has multiplicity
q; we want to replace Dbad with an effective divisor that has multiplicity q.
Since Dbad avoids z, by (5.3), it has a preimage z Dbad in z Div X . Let z δbad denote
the image in z Pic X , again using (5.3). Since the image of z δbad in z̄ Pic(X0 )red is a
q-th power, so is z δbad by Proposition 5.5. Thus by (5.4), there is an element g ∈ z K ∗
such that div g = q · zE − z Dbad , for some z E ∈ z Div X . By (5.3), we then have
div g = qE − Dbad
div f g = cI Dα + Dgood + qE
why α (w0 ) should be zero, but computing as in the proof of Proposition 9.3, we
set u0 = gu0 ∈ O∗v0 , c0 α (w0 ) = c0 α ◦ − (θ0 , u0 ), and c0 ∂v0 ,z̄ (α (w0 )) = −vz̄ (u0 )θ0 (z).
Therefore since g(z) is in κ (z)∗q , vz̄ (u0 ) = vz̄ (u0 ) (mod q), hence ∂v0 ,z̄ (α (w0 )) = 0,
and we have (C). In light of Proposition 9.3, we now have (B).
Theorem 9.5. Let K be the function field of a p-adic curve, and let Δ be a K-division
algebra of prime index q = p. Then Δ is cyclic.
Proof. This is [S2, Th. 5.1]. Let α ∈ H2 (K) be the class of Δ in Br(K). Since
ind(α ) = q is prime, α is semiramified at all v ∈ divK (α ) by (8.1). Let X → S =
Spec Z p be any arithmetic surface on which D+ α is well-conditioned, as in Lemma
8.9. By Proposition 9.4, there is an element f ∈ K satisfying (A), (B),
and (C). Let
L = K( f 1/q ). Since f satisfies (A), (B), and (C), αL ∈ H2nr (L) = V (L) H2 (Ov ) by
Proposition 8.11. Since L is the function field of a p-adic curve, αL = 0 by Theorem
4.5. Finally, since Δ is split by a radical extension of degree q, Δ is cyclic by a the-
orem of Albert [S2, Prop. 0.1].
Finally, we have the following geometric criterion for a Brauer class to have
prime index.
Corollary 9.6. Let K be the function field of a p-adic curve, and let α ∈ Br(K) be an
element of prime period q = p. Let X → Spec Z p be an arithmetic surface on which
Dα has normal crossings. Then ind(α ) = q if and only if Dα has no hot points.
Acknowledgements The author thanks Skip Garibaldi for encouraging him to take on this project,
and Colliot-Thélène, and especially Suresh and his staff for organizing a wonderful conference in
Hyderabad, in honor of Professor Parimala.
References
[AT] Artin, E., Tate, J.: Class Field Theory, AMS Chelsea Publishing, Am. Math. Soc.,
Rhode Island, 2009.
[AG] Auslander, M., Goldman, O.: The Brauer group of a commutative ring, Trans. Am.
Math. Soc., 97 (1960) 367–409.
[C-T] Colliot-Thélène, J.-L.: Birational invariants, purity and the Gersten conjecture, in
K-Theory and Algebraic Geometry: Connections with Quadratic Forms and Division
Algebras (B. Jacob and A. Rosenberg, eds.), Proceedings of Symposia in Pure Mathe-
matics, vol. 58.1, Am. Math. Soc., Providence, RI, 1995.
[C-T2] : Review of [S1], Zentralblatt MATH database 1931–2009, Zbl 0902.16021.
[GMS] Garibaldi, S., Serre, J.-P., Merkurjev, A.: Cohomological Invariants in Galois Cohomol-
ogy, (University Lecture Series Volume 28), American Mathematical Society, 2003.
[GS] Gille, P., Szamuely, T.: Central Simple Algebras and Galois Cohomology, Cambridge
University Press, Cambridge, 2006.
[GB] Grothendieck, A.: Le groupe de Brauer I, II, III, in Dix Exposés sur la Cohomologie des
Schémas, North-Holland, Amsterdam/Masson, Paris, 1968, 46–188.
[Ho] Hoobler, R.T.: A cohomological interpretation of Brauer groups of rings, Pact. J. Math.
86 no. 1, (1980) 89–92.
[Lbm] Lichtenbaum, S.: Curves over discrete valuation rings, Am. J. Math., 90 no. 2, (1968)
380–405.
[Liu] Liu, Q.: Algebraic Geometry and Arithmetic Curves, Oxford University Press, New
York, 2002.
[Mat] Matsumura, H.: Commutative Ring Theory, (Cambridge Studies in Advanced Mathe-
matics 8), Cambridge University Press, Cambridge, 1989.
[M] Milne, J. S.: Etale Cohomology. Princeton University Press, New Jersey, 1980.
[S1] : Division algebras over p-adic curves, J. Raman. Math. Soc., 12 (1997) 25–47.
[S1 ] : Correction to division algebras over p-adic curves, J. Raman. Math. Soc., 13
(1998) 125–129.
[S2] : Cyclic algebras over p-adic curves, J. Alg., 314 no. 2 (2007), 817–843.
[S3] : Division algebras over surfaces, J. Alg., 315, no. 4 (2008), 1543–1585.
[SGA1] Grothendieck, A., Raynaud, M.: Séminaire de géométrie algébrique du Bois Marie
1960–1961: Revêtements étales et groupe fondamental, Documents mathématiques,
vol. 3, Société Mathématique de France, 2003.
[Si] Silverman, J. H.: Advanced Topics in the Arithmetic of Elliptic Curves, Springer, New
York, 1994.
Serre’s Conjecture II: A Survey
Philippe Gille
Dedicated to Parimala
1 Introduction
Serre’s original conjecture II (1962) states that the Galois cohomology set H 1 (k, G)
vanishes for a semisimple simply connected algebraic group G defined over a perfect
field of cohomological dimension ≤ 2 [52, Sect. 4.1] [53, II.3.1]. This means that
G-torsors (or principal homogeneous spaces) over Spec(k) are trivial.
For example, if A is a central simple algebra defined over a field k and c ∈ k× ,
the subvariety
Xc := {nrd(y) = c} ⊂ GL1 (A)
of elements of reduced norm c is a torsor under the special linear group G = SL1 (A)
which is semisimple and simply connected. If cd(k) ≤ 2, we thus expect that this
/ Applying this to all c ∈ k× , we thus expect that
G-torsor is trivial, i.e., Xc (k) = 0.
× ×
the reduced norm map A → k is surjective.
For imaginary number fields, the surjectivity of reduced norms goes back to
Eichler in 1938 (see [39, Sect. 5.4]). For function fields of complex surfaces, this
follows from the Tsen-Lang theorem because the reduced norm is a homogeneous
form of degree deg(A) in deg(A)2 -indeterminates [53, II.4.5]. In the general case,
the surjectivity of reduced norms is due to Merkurjev–Suslin in 1981 [59, Th. 24.8],
Philippe Gille
DMA, UMR 8553 du CNRS, Ecole normale supérieure, 45 rue d’Ulm, F-75005 Paris, France,
e-mail: [email protected]
2010 Mathematics subject classification. 20G05.
The third example shows that central simple algebras and quadratic forms are not
in general low dimensional objects. We already mentioned the following character-
ization which uses Merkurjev–Suslin’s theorem [47].
Theorem 2.2. [59, Th. 24.8] Let l be an invertible prime in k. The following are
equivalent:
1. cdl (k) ≤ 2.
2. For any finite separable extension L/k and any l-primary central simple
L-algebra A/L, the reduced norm nrd : A× → L× is surjective.
3. For any finite extension L/k and any l-primary central simple L-algebra A/L, the
reduced norm nrd : A× → L× is surjective.
We added here the easy implication (2) =⇒ (3) which follows from the usual
transfer argument. We say that k is of cohomological dimension ≤ d if k is of
l-cohomological dimension cdl (k) ≤ d for all primes l.
If k is of positive characteristic p, we always have cd p (k) ≤ 1; this explains the
necessary change in the following analogous statement.
Theorem 2.3. [27, Th. 7] Assume that char(k) = p > 0. The following are
equivalent:
1. H p3 (L) = 0 for any finite separable extension L/k;
2. For any finite separable extension L/k and any l-primary central simple
L-algebra A/L, the reduced norm nrd : A× → L× is surjective.
Examples 2.4.
(1) The function field of a curve over a finite field is of separable dimension 2.
(2) The function field k0 (S) of a surface over an algebraically closed field k0 of
characteristic p ≥ 0 is of separable dimension 2.
(3) Given an arbitrary field F, Theorems 2.2 and 2.3 provide a way to construct a
“generic” field extension E/F of separable dimension 2, see Ducros [19].
We can now state the strong form of Serre’s conjecture II. For each simply con-
nected group G, Serre defined the set S(G) of primes in terms of the Cartan-Killing
type of G, cf. [54, Sect. 2.2]. For absolutely almost simple groups, the primes are
listed in Table 1.
1 Kato defined the p-dimension dim p (k) as follows [38]. If [k : k p ] = ∞, define dim p (k) = ∞. If
[k : k p ] = pr < ∞, dim p (k) = r if Hpr+1 (L) = 0 for any finite extension L/k, and dim p (k) = r + 1
otherwise.
44 Philippe Gille
1 → Z(G) → G → Gad → 1
ϕG
1 → Z(G)(k) → G(k) → Gad (k)−→H 1f pp f (k, Z(G)) →
δ
→ H 1f pp f (k, G) → H 1f pp f (k, Gad ) −→
G
H 2f pp f (k, Z(G)). (3.1)
Serre’s Conjecture II: A Survey 45
The homomorphism ϕG is called the characteristic map and the mapping δG is the
boundary. Since G (resp. Gad ) is smooth, the f pp f -cohomology of G (resp. Gad )
coincides with Galois cohomology [55, XXIV.8], i.e., we have a bijection
∼
H 1 (k, G)−→H 1f pp f (k, G). Following [40, 31.6], one defines the Tits class of G
by the following formula
tG = − δG νG ∈ H 2f pp f (k, Z(G)).
δG
H 1 (k, Gqad ) −−−ad
−→ H 2f pp f (k, Z(Gq )),
we see that tG = δGq ([z]) which is indeed Tits’ definition [63, Sect. 1].
Proof. (1) follows from the exact sequence (3.1). For (2), let z ∈ Z 1 (k, Gad ) be a
q
cocycle such that G ∼ = z G. We assume that tG = tG . Hence δGq ([z]) = δGq ([z ]) ∈
H f pp f (k, Z(G )). The compatibility above shows that
2 q
τz−1 ([z ]) ∈ ker H 1 (k, Gad ) → H 2f pp f (k, Z(G)) .
By 1), we have τz−1 ([z ]) = [1] ∈ H 1 (k, Gad ), hence [z] = [z ] ∈ H 1 (k, Gad ). Thus G
and G are k-isomorphic.
We would like to describe a few ways to attack the conjecture and their limitations.
This is somehow artificial because in practice we work with all tools.
Let Us Explain it within the following example due to Tits [64]. Let G/k be the split
semisimple simply connected group of type
α2
r
E6 r r r r r
α1 α3 α4 α5 α6
Assume that k is infinite. Let z ∈ Z 1 (ks /k, G) and consider the twisted group G = z G.
Since tG = 0, the 27-dimensional standard representation of G of highest weight ω 6
descends to G by [62]. We then have a representation ρ : G → GL(V ). The point is
that G has a dense orbit in the projective space X = P(V ), so there exists a k-rational
point [x] in that orbit. The connected stabilizer (Gx )0 is then semisimple of type F4
[22, 9.12]. Assuming that Conjecture 2.5 holds for groups of type F4 , it follows that
(Gx )0 is split. Hence G has relative rank ≥ 4 and a glance at Tits’ tables [61] tells
us that G is split. It is then easy to conclude that [z] = 1 ∈ H 1 (k, G).
The subgroup trick (and variants) was fully investigated by Garibaldi in his Lens
lectures [22]. The underlying topic is that of prehomogeneous spaces, namely pro-
jective G-varieties with a dense orbit.
Unfortunately, this trick works only in few cases. Tits has shown that the gen-
eral form of type E8 is “almost abelian” namely has no nontrivial other reductive
subgroups than maximal tori [64]. Together with Garibaldi, we have shown that the
general trialitarian group is almost abelian [23].
In this case, the idea is to derive Serre’s conjecture II from a more general set-
ting. The Rost invariant [25] generalizes the Arason invariant for 3-fold Pfister form
which (in characteristic = 2) attaches to a Pfister form φ = a, b, c the cup-product
e3 (φ ) = (a) ∪ (b) ∪ (c) ∈ H 3 (k, Z/2Z). We now see it as the cohomological invari-
ant H 1 (k, Spin8 ) → H 3 (k, Z/2Z(2)). More generally, for G/k simply connected and
absolutely almost simple, there is a cohomological invariant
where the p-primary part has to be understood in Kato’s setting [25]. If this invariant
has trivial kernel, then H 1 (k, G) = 1 for G/k satisfying the hypothesis of Conjecture
2.5. This is the case for Spin8 by Arason’s theorem, namely the invariant e3 (φ )
determines φ .
Serre’s Conjecture II: A Survey 47
A special case of a question raised by Serre in 1962 ([52], see also [54, Sect. 2.4]),
is the following.
Question 4.1. Let G/k be a connected linear algebraic group. Let (ki )i=1,..,r be a
family of finite field extensions of k such that g.c.d.([ki : k]) = 1. Is the kernel of the
map
H 1 (k, G) → ∏ H 1 (ki , G)
i=1,..,r
trivial ?
Remarks.
(1) The hypothesis of connectedness is necessary since there are counterexamples
with finite constant groups [34, 49].
(2) The question was generalized by Totaro [65, question 0.2], see also [24].
(3) If k is of positive characteristic p, there exists a complete DVR R with residue
field k and field of fractions K = Frac(R) of characteristic zero, and an R-group
scheme G with special fiber G. An answer for GK to Serre’s question yields an
answer for G. A fortiori and without lost of generality we can assume that the
extensions ki /k are separable.
Definition 4.2. Let l be a prime. We say that a field k is l-special if every finite
separable extension of k is of degree a power of l.
has trivial kernel for l running over all primes. If the answer to this question is yes,
then Conjecture II becomes a question for l-special fields for primes l in S(G).
Here are the few cases where a positive answer to Serre’s question is known:
unitary groups (Bayer–Lenstra [3]), groups of type G2 , quasi-split groups of type
D4 , F4 , E6 , E7 [11, 21, 26].
If we know that the Rost invariant has zero kernel, then we easily deduce that
the answer to Question 4.1 is yes. Thus we can answer Serre’s question for groups
of type G2 , and quasi-split semisimple simply connected groups of type D4 , F4 , E6 ,
and E7 .
48 Philippe Gille
Recall that a semisimple simply connected group is called classical if its factors are
of type A, B, C, or D, and there is no triality involved.
Theorem 5.1. Let G be an absolutely almost simple and simply connected classical
group over a field k as in Conjecture 2.5. Then H 1 (k, G) = 1.
If k is perfect or char(k) = 2, this is the original Serre’s conjecture II proven by
Bayer–Parimala [4]. The general case is a recent work by Berhuy–Frings–Tignol
[5]. Its proof is based on Weil’s presentation of classical groups in terms of uni-
tary groups of algebras with involutions [66]. This proof is characteristic free; in
particular, it provides a somewhat different proof of Bayer–Parimala’s theorem.
Possibly the trickiest case is that of outer groups of type A, namely unitary groups
of central simple algebras equipped with an involution of the second kind. The proof
in the number field case (which uses Landherr’s Theorem) is already difficult, see
[39, Sect. 5.5].
For such groups, the best approach is by investigating the Rost invariant.
Theorem 5.2. Let G/k be a quasi-split semisimple simply connected group of
Cartan-Killing type G2 , F4 , D4 , E6 , or E7 . Then the Rost invariant H 1 (k, G) →
H 3 (k, Q/Z(2)) has trivial kernel.
Here the field k is arbitrary, but proving the theorem boils down to the charac-
teristic zero case by a lifting argument [27]. For the cases G2 , F4 , see [4] or [54].
As pointed out by Garibaldi, the D4 case is done in The Book of Involutions but not
stated in that shape. We need to know that a trialitarian algebra whose underlying
algebra is split arises as the endomorphism of a twisted composition [40, 44.16] and
to use results on degree 3 invariants of twisted compositions (ibid., 40.16). For type
E6 and E7 , this is due independently to Chernousov [11] and Garibaldi [21].
Thus Conjecture 2.5 holds for quasi-split groups of all types except E8 ; we have
an independent proof which is quite different since it is based on Bruhat-Tits theory
[28]. For the split group of type E8 denoted by E8 , the Rost invariant in general has
a nontrivial kernel [30, appendix]. In characteristic 0, Semenov recently constructed
a higher invariant
ker H 1 (k, E8 ) → H 3 (k, Q/Z(2)) → H 5 (k, Z/2Z)
which is nontrivial since it is so already over the reals [57, Sect. 8]. Semenov’s
invariant has trivial kernel for two-special fields.
Serre’s Conjecture II: A Survey 49
Theorem 5.3. [28, Th. 6] Let G/k be a semisimple simply connected group which
satisfies the hypothesis of Conjecture 2.5. Let μ ⊂ G be a finite central subgroup of
G. Then the characteristic map
is surjective.
Flat cohomology (see [56], [5, app. B] or [33]) is the right object if the order of μ
is not invertible in k, it coincides with Galois cohomology if this order is invertible.
By continuing the exact sequence of pointed sets
we see that H 1f pp f (k, μ ) → H 1f pp f (k, G) is the trivial map. In other words, the center
of G does not contribute to H 1 (k, G). [The reason why we can avoid the type E8 is
that such groups have trivial center.]
Theorem 5.4. [11, 28] Let G/k be a semisimple group satisfying the hypothesis of
Conjecture 2.5. Then H 1 (k, G) = 1 in the following cases:
1. G is trialitarian and its Allen algebra is of index ≤ 2.
2. G is of quasi-split type 1 E6 or 2 E6 and its Tits algebra is of index ≤ 3.
3. G is of type E7 and its Tits algebra is of index ≤ 4.
α1
r α2 α4 αr3 r
r
i r ri ri r r
r α5 α6
α2 α2
r r
a) r i
r r i
r i
r i
r b) r r r r r i
r
α7 α6 α5 α4 α3 α1 α7 α6 α5 α4 α3 α1
where case (a) (resp. (b)) is that of Tits algebra of index 2 (resp. 4). One more
reason why other exceptional groups are not easy to work with is because they are
anisotropic.
50 Philippe Gille
Corollary 5.5. [15] Let G/k as in Theorem 5.4. For every separable finite field ex-
tension L/k, assume that every central simple L-algebras of period 2 (resp. 3) is of
index ≤ 2 (resp. 3). If G is trialitarian or of type E7 (resp. E6 ), then H 1 (k, G) = 1.
This is the case for function fields of surfaces as pointed out by Artin [1], thus
Corollary 5.5 holds for these fields. In the paper [15] with Colliot–Thélène and
Parimala, we exploited Serre’s conjecture II for the study of arithmetic properties
in this framework by proceeding with analogies with Sansuc’s paper [51] in the
number field case. On this topic, see also the paper by Borovoi and Kunyavskiı̌ [7].
Let K be a henselian valued field for a discrete valuation with perfect residue field κ .
A consequence of the Bruhat–Tits decomposition for Galois cohomology over com-
plete fields is the following.
Theorem 6.1. (Bruhat–Tits [8, Cor. 3.15]) Assume that κ is of cohomological dim-
ension ≤ 1. Let G/K be a simply connected semisimple group. Then H 1 (K, G) = 1.
Note that the hypotheses imply that K is of separable dimension ≤ 2. Serre asked
whether it can be generalized when assuming [κ : κ p ] ≤ p [54, 5.1]. The hypothesis
[κ : κ p ] ≤ p alone is not enough here because K = F p ((x))((y)) is of separable
dimension 3 and is complete with residue field F p ((x)).
But if κ is separably closed and [κ : κ p ] ≤ p, then K is of separable dimension 1
and enough cases of the vanishing of H 1 (κ ((x)), G) have been established in view
of the proof of Tits conjectures on unipotent subgroups [29]. The general case is
still open.
Serre’s Conjecture II: A Survey 51
Note also that the conjecture is proven for fraction fields of henselian two dimen-
sional local rings (e.g. C[[x, y]]) with algebraically closed residue field of character-
istic zero [15]. For the E8 case, a key point is that the derived group of the absolute
Galois group is of cohomological dimension 1 [17, Th. 2.2].
The number field case is due to Kneser for classical groups [39], Harder for excep-
tional groups except the type E8 [35, I, II], and Chernousov for the type E8 [9], see
[50]. The function field case is due to Harder [35, III].
He, de Jong, and Starr have proven Conjecture 2.5 for split groups over function
fields in a uniform way and in arbitrary characteristic.
Theorem 6.2. [37, Cor. 1.5] Let k be an algebraically closed field and let K be the
function field of a quasi-projective smooth surface S. Let G be a split semisimple
simply connected group over k. Then H 1 (K, G) = 1.
For cases other than E8 , the conjecture had been establised by case by case con-
siderations [15]. Hence Conjecture 2.5 is fully proven for function fields of surfaces.
The proof of Theorem 6.2 is based on the existence of sections for fibrations in ra-
tionally simply connected varieties.
Theorem 6.3. [37, Th. 1.4] Let S/k as in Theorem 6.2. Let X/S be a projective
morphism whose geometric generic fiber is a twisted flag variety. Assume that
Pic(X ) → Pic(X ×K K) is surjective. Then X → S has a rational section.
The assumption on the Picard group means that there is no “Brauer obstruction”.
By application to higher Severi-Brauer schemes, this statement yields as corollary
de Jong’s theorem “period=index” [36] for central simple algebras over such fields;
see also [14].
It is the first classification-free item in this survey.
Lemma 6.4. Let G/F be a semisimple simply connected group over a field F. Let
E/F be a G-torsor.
52 Philippe Gille
(2) Let P be an F-parabolic subgroup of G and let E/P be the variety of parabolic
subgroups of the twisted F-group E(G) of the same type as P. Then we have an
exact sequence
0
⏐
⏐
0 −−−−→ Br(F) −−−−→ Br(E/P) −−−−→ Br(E/P ×F Fs )
⏐ ⏐
⏐ ⏐
||
0 −−−−→ Br(F) −−−−→ Br(E) −−−−→ Br(E ×F Fs ).
Since the bottom sequence is exact, we get by diagram chasing that the up-
∼
per horizontal sequence is exact as well. The second isomorphism Pic(E/P) − →
Pic(E/P ×F Fs )G al(Fs /F) comes from the Hochschild-Serre spectral sequence.
For complete results on Picard and Brauer groups of twisted flag varieties, see
Merkurjev–Tignol [48, Sect. 2].
Proposition 6.5. [37, Th. 1.4] Let S/k be as in Theorem 6.2. Let G/K be a semisim-
ple simply connected K-group which is an inner form and let P be a K-parabolic
subgroup of G. Then the map H 1 (K, P) → H 1 (K, G) is bijective.
Proof. Injectivity is a general fact due to Borel–Tits ([6], théorème 4.13.a). Let E/K
be a G-torsor of class [E] ∈ H 1 (K, G). After shrinking S, we can assume that G/K
Serre’s Conjecture II: A Survey 53
Pic(V ) → Pic(V ×K Ks )
is onto. Thus the composite map Pic(V) → Pic(V ) → Pic(V ×K Ks ) is onto. Theorem
6.2 applies and shows that V (K) = 0. / Thus
the torsor E admits a reduction
to P ([53],
Sect. I.5, proposition 37), that is [E] ∈ im H 1 (K, P) → H 1 (K, G) . We conclude that
the mapping H 1 (K, P) → H 1 (K, G) is surjective.
Corollary 6.6. Let S/k be a smooth quasi-projective surface. Let G/k be a (split)
semisimple simply connected group. Let E/S be a G-torsor. Then E is locally trivial
for the Zariski topology.
Acknowledgements We thank Skip Garibaldi, Fabien Morel, and Jean-Pierre Tignol for useful
comments which improved the exposition.
References
20. R. Fossum and B. Iversen, On Picard groups of algebraic fibre spaces, J. Pure Appl. Algebra
3 (1973), 269–280.
21. R.S. Garibaldi, The Rost invariant has trivial kernel for quasi-split groups of low rank, Com-
ment. Math. Helv. 76 (2001), 684–711.
22. , Cohomological invariants: exceptional groups and spin groups, With an appendix
by Detlev W. Hoffmann. Mem. Am. Math. Soc. 200 (2009), no. 937,
23. R.S. Garibaldi and P. Gille, Algebraic groups with few subgroups, J. Lond. Math. Soc. 80
(2009), 405–430.
24. R.S. Garibaldi and D.W. Hoffmann, Totaro’s question on zero-cycles on G2 , F4 and E6 torsors,
J. Lond. Math. Soc. 73 (2006), 325–338.
25. R.S. Garibaldi, A.A. Merkurjev and J.-P. Serre, Cohomological invariants in Galois cohomol-
ogy, University Lecture Series, 28 (2003). American Mathematical Society, Providence.
26. P. Gille, La R-équivalence sur les groupes algébriques réductifs définis sur un corps global,
Inst. Hautes Études Sci. Publ. Math. 86 (1997), 199–235.
27. , Invariants cohomologiques de Rost en caractéristique positive, K-Theory 21 (2000),
57–100.
28. , Cohomologie galoisienne des groupes quasi-déployés sur des corps de dimension
cohomologique ≤ 2, Compositio Math. 125 (2001), 283–325.
29. , Unipotent subgroups of reductive groups in characteristic p > 0, Duke Math. J. 114
(2002), 307–328.
30. , An invariant of elements of finite order in semisimple simply connected algebraic
groups, J. Group Theory 5 (2002), 177–197.
31. , Le problème de Kneser-Tits, Astérisque 326 (2009), 39–82.
32. P. Gille and T. Szamuely, Central simple algebras and Galois cohomology, Cambridge Studies
in Advanced Mathematics 101 (2006), Cambridge University Press, Cambridge.
33. J. Giraud, Cohomologie non abélienne, Die Grundlehren der mathematischen Wissenschaften
179 (1971), Springer-Verlag, Berlin.
34. D. Goldstein, R.M. Guralnick, E.W. Howe, M.E. Zieve, Nonisomorphic curves that become
isomorphic over extensions of coprime degrees, J. Algebra 320 (2008), 2526–2558.
35. G. Harder, Über die Galoiskohomologie halbeinfacher Matrizengruppen I Math. Zeit. 90
(1965), 404–428, II Math. Zeit. 92 (1966), 396–415, III J. für die reine angew. Math. 274/5
(1975), 125–138.
36. A.J. de Jong, The period-index problem for the Brauer group of an algebraic surface, Duke
Math. J. 123 (2004), 71–94.
37. A.J. de Jong, X. He and J.M. Starr, Families of rationally simply connected varieties over
surfaces and torsors for semisimple groups, preprint (2008), arxiv:0809.5224.
38. K. Kato, Galois cohomology of complete discrete valuation fields, Lect. Notes in Math. 967
(1982), 215–238.
39. M. Kneser, Lectures on Galois cohomology, Tata institute (1969).
40. M.-A. Knus, A. Merkurjev, M. Rost and J.-P. Tignol, The book of involutions, avec une préface
de J. Tits, American Mathematical Society Colloquium Publications 44 (1998), American
Mathematical Society, Providence, RI.
41. M. Lieblich, Deformation theory and rational points on rationally connected varieties, in this
volume.
42. D.W. Lewis and J.-P. Tignol, Classification theorems for central simple algebras with involu-
tion, Manuscripta Math. 100 (1999), 259–276.
43. B. Margaux, Passage to the limit in non-abelian Čech cohomology, Journal of Lie Theory 17
(2007), 591–596.
44. A. S. Merkurjev, Simple algebras and quadratic forms, Izv. Akad. Nauk SSSR 55 (1991),
218–224.
45. , K-theory of simple algebras, K-theory and algebraic geometry: connections with
quadratic forms and division algebras (Santa Barbara, CA, 1992), 65–83, Proc. Sympos. Pure
Math. 58.1 (1995) American Mathematical Society, Providence, RI.
46. , K-theory and algebraic groups, European Congress of Mathematics, Vol. II
(Budapest, 1996), 43–72, Progr. Math. 169 (1998), Birkhäuser.
56 Philippe Gille
47. A.S. Merkurjev and A.A. Suslin, K -cohomology of Severi-Brauer varieties and norm residue
homomorphism, Izv. Akad. Nauk SSSR 46 (1982), 1011–1046, english translation: Math.
USSR Izv. 21 (1983), 307–340.
48. A.S. Merkurjev and J.-P. Tignol, The multipliers of similitudes and the Brauer group of homo-
geneous varieties, J. Reine Angew. Math. 461 (1995), 13–47.
49. R. Parimala, Homogeneous varieties—zero-cycles of degree one versus rational points, Asian
J. Math. 9 (2005), 251–256.
50. V. P. Platonov and A. Rapinchuk, Algebraic groups and number theory, Pure Appl. Math. 139
(1994), Academic Press, London.
51. J.-J. Sansuc, Groupe de Brauer et arithmétique des groupes algébriques linéaires, J. Reine
Angew. Math. 327 (1981), 12–80.
52. J-P. Serre, Cohomologie galoisienne des groupes algébriques linéaires, Colloque sur la théorie
des groupes algébriques linéaires, Bruxelles (1962), 53–68.
53. , Cohomologie galoisienne, cinquième édition révisée et complétée, Lecture Notes in
Mathematics 5, Springer-Verlag, Berlin.
54. , Cohomologie galoisienne: Progrès et problèmes, Séminaire Bourbaki, exposé 783
(1993–94), Astérisque 227 (1995).
55. Séminaire de Géométrie algébrique de l’I.H.E.S., 1963-1964, Schémas en groupes, dirigé par
M. Demazure et A. Grothendieck, Lecture Notes in Math. 151–153, (1970) Springer, Berlin.
56. S. Shatz, Cohomology of artinian group schemes over local fields, Ann. of Math. 79 (1964),
411–449.
57. N. Semenov, Motivic construction of cohomological invariants, preprint (2008), arxiv:0905.
4384v1.
58. J. Starr, Rational points of rationally connected and rationally simply connected varieties,
preprint (2008).
59. A. A. Suslin, Algebraic K-theory and the norm–residue homomorphism, J. Soviet. 30 (1985),
2556–2611.
60. R. Switzer, Algebraic topology—homotopy and homology, Die Grundlehren der mathematis-
chen Wissenschaften, Band 212 (1975), Springer-Verlag, Berlin.
61. J. Tits, Classification of algebraic semisimple groups, Algebraic Groups and Discontinuous
Subgroups, Proc. Sympos. Pure Math., Boulder, Colo. (1965) pp. 33–62, Amer. Math. Soc.
62. , Représentations linéaires irréductibles d’un groupe réductif sur un corps quel-
conque, J. Reine Angew. Math. 247 (1971), 196–220.
63. , Résumé des cours du Collège de France 1990–91, Annuaire du Collège de France.
64. , Sur les degrés des extensions de corps déployant les groupes algébriques simples,
C. R. Acad. Sci. Paris Sér. I Math. 315 (1992), 1131–1138.
65. B. Totaro, Splitting fields for E8 -torsors, Duke Math. J. 121 (2004), 425–455.
66. A. Weil, Algebras with involutions and the classical groups, J. Indian Math. Soc. 24 (1960),
589–623.
Field Patching, Factorization,
and Local–Global Principles
Daniel Krashen
Summary The method of field patching has proven useful in obtaining results on
Galois theory, central simple algebras, and quadratic forms. A crucial ingredient for
this was proving certain “factorization” results for connected, rational linear alge-
braic groups. In this paper, we explore other possible applications of field patching
by examining the relationship between factorization results and local–global prin-
ciples, and also by extending the known factorization results to connected, retract
rational linear algebraic groups.
1 Introduction
The goal of this paper is twofold – first to give an introduction to the method of
field patching as first presented in [HH], and later used in [HHK09], paying special
attention to the relationship between factorization and local–global principles and
second, to extend the basic factorization result in [HHK09] to the case of retract
rational groups, thereby answering a question posed to the author by J.-L. Colliot-
Thélène.
Throughout, we fix a complete discrete valuation ring T with field of fractions
K and residue field k. Let t ∈ T be a uniformizer. Let X /K be a smooth projective
curve and F its function field.
Broadly speaking, the method of field patching is a procedure for constructing
new fields Fξ which will be in certain ways simpler than F, and to reduce problems
concerning F to problems about the various Fξ . Overall, there are two ways in which
Daniel Krashen
Department of Mathematics, University of Georgia, Athens, Georgia 30602, USA,
e-mail: [email protected]
2010 Mathematics subject classification. 12J25, 20G15.
this is done. Let us suppose that we are interested in studying a particular type of
arithmetic object, such as a quadratic form, a central simple algebra, etc.
This consists in showing that under suitable hypotheses, algebraic objects defined
over the fields Fξ which are “compatible,” exactly correspond to objects defined
over F (see Theorem 3.2.3). One may then use this idea to construct new examples
and counterexamples of such objects by building them “locally.”
The fields Fξ are not canonically defined – they depend on a number of choices,
beginning with the choice of a model for X over T .
Given a model X (we will generally suppress the morphism f : X → P1T from the
notation), we let S(X) denote the set of closed points in f −1 (∞) and U (X ) denote
the set of connected (or equivalently, irreducible) components of Xkred \ S(X).
These
sets play a critical role in what follows.
Warning 2.1.2. In other sources such as [HH, HHK09, HHK], X is not given the
structure of a P1T -scheme, but rather the structure of a T scheme together with a
Field Patching, Factorization, and Local–Global Principles 59
distinguished set of closed points S. In this context, one is allowed more general
sets S. The reader must keep in mind that a model X comes with the extra structure
of a morphism to P1 throughout!
It is perhaps a bit odd to include the finite morphism to P1T as part of the def-
inition of a model – by comparison, in [HH], it is only assumed that one should
start with a projective T -curve with a set S of closed points such that there exists
a finite T -morphism to a curve with smooth reduced closed fiber and such that the
set S is the inverse image of a set of closed points under this morphism. We include
the morphism to P1T as part of our definition simply as a matter of convenience of
exposition. The following lemma shows that it is not much of an extra assumption,
however:
Lemma 2.1.3 ([HH, Prop. 6.6]). Suppose X is a projective T -curve and S ⊂ X a
finite set of closed points. Then there exists a finite morphism f : X → P1 such that
S ⊂ f −1 (∞).
For the remainder of the section, we will suppose that we are given such a model
and we let F = F(X ) be its function field. Given any nonempty subset of points
X,
we define
Z ⊂ X,
Note that there are natural maps F ⊂ FU , FP for any such P and U, as well as
inclusions FU → FV and FU → FP , whenever V ⊂ U or P ∈ U, respectively.
We may now give some examples of local–global principles. For these, we assume
that X/K is a smooth projective curve where K is a complete discretely valued field
with valuation ring T , and that we are given a model X → P1 . We let F be the
function field of X.
60 Daniel Krashen
Theorem 2.2.1 (Local–global principle for the Brauer group [HH, Th. 4.10]).
Let Br(·) denote the Brauer group. The natural homomorphism
Br(F) → ∏P∈S(X) Br(F P ) × ∏U∈U (X) Br(FU )
is injective.
We give a proof of this result on the next page. In fact, we will see later, using
patching, that this may be extended to a three term exact sequence by adding a term
on the right (see Theorem 3.3.1).
Theorem 2.2.2 (Local–global principle for isotropy [HHK09, Th. 4.2]). Sup-
pose q is a regular quadratic form of dimension at least 3, and char(F) = 2. If qFP
and U ∈ U (X ) then q is also isotropic.
and qFU are isotropic for every P ∈ S(X)
The proof is given below. See also Theorem 3.3.2 for a related result.
Both of these principles in fact, may be regarded as special cases of either of the
following results, the main new results of this paper:
The following result generalizes [HHK09, Th. 3.7] by weakening the hypothesis
of rationality to allow for retract rational groups as well:
Theorem 2.2.4 (Local–global principle for varieties with transitive actions).
Suppose G is a connected retract rational algebraic group defined over F, and H is a
variety on which G acts transitively. Then H(F) = 0/ if and only if H(FP ), H(FU ) = 0/
and U ∈ U (X).
for all P ∈ S(X)
This theorem follows quickly from Theorem 5.1.1, and its proof may be found
just after the statement of this theorem on page 69. The proof of this in the case
of retract rationality will occupy a good portion of this paper. Along the way, we
will explore the connections between these local–global principles and the notion of
“factorization” for the group G. The following corollary is particularly useful.
Proof. This follows from the fact that the action of G(F) on H(F) is transitive. This
in turn in a consequence of [Bor, Th. 20.9(iii)].
From these theorems (or even the versions assuming only rationality of G from
[HHK09]), we may prove the above local–global results concerning the Brauer
group and quadratic forms.
Field Patching, Factorization, and Local–Global Principles 61
Proof (of Theorem 2.2.1). Let α ∈ Br(F) and suppose αFP = 0, αFU = 0 for every
U ∈ U (X).
P ∈ S(X), We need to show that α = 0.
Let A be a central simple F algebra in the class of α and let H be the Severi-
Brauer variety for A. Note that this is a homogeneous variety for the group GL(A)
which is rational, connected, and reductive. Recall that for a field extension L/F,
H(L) is nonempty exactly when A ⊗F L is a split algebra – that is to say, αL = 0.
But since αFP , αFU = 0, we have H(FP ), H(FU ) = 0/ for every U, P. Consequently,
by Corollary 2.2.5, it follows that H(F) = 0/ and so α = 0 as desired.
Proof (of Theorem 2.2.2). Let q be a quadratic form over F satisfying the hypothe-
ses of the theorem. We wish to show that q is isotropic. Let H be the quadratic
hypersurface of projective space defined by the equation q = 0. Recall that this is a
homogeneous variety for the group SO(q) which under the hypotheses is a rational,
connected, reductive group (see [KMRT, p. 201, exercise 9]). As above, we immedi-
and U ∈ U (X ),
ately see that since H(FP ), H(FU ) are nonempty for each P ∈ S(X)
we have by Corollary 2.2.5, H(F) = 0/ as desired.
3 Patching
The fundamental idea of patching is that defining an algebraic object over the field
and FU for
F is equivalent to defining objects over each of the fields FP for P ∈ S(X)
together with the data of how these objects agree on overlaps. This will
U ∈ U (X),
be stated in this section in terms of an equivalence of categories. We will simply
cite the results of [HH] Sects. 6 and 7 for the most part, but we focus more on the
equivalence of tensor categories, and explore how to produce other examples of
algebraic patching.
Suppose we are given a model X for a curve X /K. Given a point P ∈ S(X),
the height 1 primes of RP which contain t correspond to the components of Xkred
incident to P. Each such component is the closure of a uniquely determined element
U ∈ U (X).
Definition 3.0.1 (Branches, and their fields). Given such a height 1 prime P of
a branch along U at P is an irreducible
RP , corresponding to an element U ∈ U (X),
component of the scheme RP /P RP . Alternately, these are in correspondence with
the height one primes of RP containing P RP . Given such a height 1 prime ℘, we
let R℘ be the t-adic completion of the localization of RP at ℘, and F℘ its field of
fractions. We let B(X) denote the set of all branches at all points in S(X).
The fields FP and FU come equipped with natural inclusions into F℘ which we
now describe. We note that the natural inclusion RP → R℘ induces an inclusion of
fields FP → F℘. Further, we note that R℘ is a 1 dimensional regular local ring, and
hence a DVR, whose valuation is determined by considering the order of vanishing
along the branch corresponding to ℘. In particular, considering the inclusion F ⊂
FP ⊂ F℘, we find that all the elements of RU , cannot have poles along any branch
62 Daniel Krashen
lying along U, and in particular, we see we have an inclusion RU ⊂ R℘. Since the
t-adic topology on R℘ is the same as the ℘-adic topology, we further find that R℘ is
t-adically complete, and we therefore have an induced inclusion FU → F℘.
f P ⊗F℘
VP ⊗FP F℘ / WP ⊗FP F℘
φ℘ ψ℘
VU ⊗FU F℘ / WU ⊗FU F℘
fU ⊗F℘
We see then that patching problems naturally form a category, which we denote
by PP(X, S). In fact, this category has a ⊗-structure as well defined by (V, φ ) ⊗
(W, ψ ) = (V ⊗ W, φ ⊗ ψ ) where (V ⊗ W )ξ = Vξ ⊗Fξ Wξ and
(φ ⊗ ψ )℘ : (V ⊗ W )P ⊗FP F℘ → (V ⊗ W )U ⊗FU F℘
is given by φ℘ ⊗F℘ ψ℘ via the above identification. One may also verify that this
monoidal structure is symmetric and closed, see [ML, VII.7].
Definition 3.1.2. If V is a vector space over F, we let (V , I) denote the patching
problem defined by V P = VFP and VU = VFU and where I℘ is induced by the natural
identifications
from the category of finite dimensional F-vector spaces to the category of patching
problems defined by sending a finite dimensional vector space V to the patching
problem (V , I). Then Ω is an equivalence of categories.
Field Patching, Factorization, and Local–Global Principles 63
A ⊗ Aop → Hom(A, A)
is an isomorphism, see [DI, Chap. 2, Th. 3.4(iii)]. In this case, the category T is
generated by a single element a, a morphism a ⊗ a → a and 1 → a (where 1 is the
unit for the monoidal structure), and such that the natural map a ⊗ a → Hom(a, a)
(where the Hom is defined by the closed structure) has an inverse.
To see quadratic forms and isometries in this way, one may simply let the
category T be generated by a single element v a morphism v ⊗ v → 1, assumed to
64 Daniel Krashen
commute with the morphism switching the order of the v’s. In the case of similarities
instead of isometries, one may add a new object , and replace v ⊗ v → 1 with a mor-
phism v ⊗ v → . To force to correspond to a 1-dimensional vector space, one may
then add to this category an inverse to the natural morphism 1 → ⊗ ∗ ∼
= Hom(, ).
defined by sending an algebraic object A to the patching problem (A, I). Then ΩT
is an equivalence of categories.
One may now check that this gives the desired equivalence.
Remark 3.2.4. It would be interesting to know if one could extend this to equiva-
lences of other kinds of objects. In particular, infinite dimensional vector spaces,
finitely generated commutative algebras, or perhaps even to (some suitably re-
stricted) categories of schemes. None of these fall under the definition of an al-
gebraic object given above, and it is therefore not at all clear if the conclusions of
Theorem 3.2.3 will still hold.
For the following results, we suppose we are given X a normal, connected, projec-
tive, finite P1T -scheme. The machinery of patching gives the exactness of various
exact sequences relating to field invariants derived from algebraic objects, such as
the Brauer group Br(F) and the Witt group W (F) of quadratic forms over F.
Field Patching, Factorization, and Local–Global Principles 65
Proof. Exactness on the left was noted in Theorem 2.2.1. To see exactness in the
middle, suppose we have classes αP , αU such that (αU )F℘ ∼= (αP )F℘ whenever ℘ is
a branch at P on U. Since there are only a finite number of points and components,
we may choose an integer n such that each of the Brauer classes αU , αP may be
represented by central simple algebras AU , AP of degree n. Now, by hypothesis, for
each branch ℘ as above, we may find an isomorphism of central simple algebras
φ℘ : (AP )F℘ → (AU )F℘ . But this gives the data of a patching problem for central
simple algebras, and therefore, we may find a central simple F-algebra A such that
AFP ∼
= AP and AFU ∼ = AU as desired.
Proof. The proof is very similar to the last one. Suppose we have Witt classes αP ,
αU such that (αU )F℘ = (αP )F℘ whenever ℘ is a branch at P on U. Since there are
only a finite number of points and components, we may choose an integer n such
that each of the Witt classes αU , αP may be represented by quadratic forms qU , qP
of the same dimension n. Now, by hypothesis and Witt’s cancellation theorems, for
each branch ℘ as above, we may find an isometry φ℘ : (qP )F℘ → (qU )Fw p . But this
gives the data of a patching problem for quadratic forms, and therefore, we may
obtain a form q over F such that the class α of q in W(F) has the property that
αFP = αP and αFU = αU .
Let us now gather together some fundamental facts which we will need in the sequel.
F(Y ) = ∏ FP , F(Y ) = ∏ FU , = ∏ F℘ ,
FP ⊗F(X) ∼ FU ⊗F(X) ∼ F℘ ⊗F(X) F(Y ) ∼
where P (resp. U , ℘ ) range over all the points (resp. components, branches) lying
over P (resp. U, ℘).
66 Daniel Krashen
Lemma 3.4.2 ([HH, Lemma 6.3]). Let X be a projective, normal, finite P1T -scheme.
Then the natural inclusions of fields yield an exact sequence of F = F(X)-vector
spaces: ⎛ ⎞ ⎛ ⎞
0→F →⎝ ∏ FP ⎠ × ⎝ ∏ FU ⎠ → ∏ F℘
P∈S(X)
U∈U (X)
℘∈B(X)
RA1 ∼
= RA1 /t RA1 , R∞ ∼
= R∞ /t R∞ , and R℘ ∼
= R℘/t R℘.
Let X → P1 be a model for X/K, and let G be an algebraic group defined over F.
Definition 4.1.1. We say that factorization holds for G, with respect to X, if for
every tuple (g℘)℘∈B(X) , there exist collections of elements g P for each P
∈ S(X)
and gU for each U ∈ U (X ) such that whenever ℘ is a branch at P on U we have
g℘ = gP gU
Definition 4.1.2. We say that the local-global principle holds for an F-scheme V ,
with respect to a model X if V (F) = 0/ holds if and only if V (FP ),V (FU ) = 0/ for
and U ∈ U (X).
every P ∈ S(X)
Definition 4.1.3. Let G be an algebraic group over F and H a scheme over F. We
say that H is a transitive G-scheme if G acts transitively on H (see Definition 2.2.3).
Proposition 4.1.4. If factorization holds for a group G, then the local–global prin-
ciple holds for all transitive schemes over G.
Proof. We essentially follow the proof of Theorem 3.7 in [HHK09]. Suppose have
a group G such that factorization holds for G, and a transitive G-scheme H. Suppose
we are given points xP ∈ H(FP ) and xU ∈ H(FU ) for all P and U. We will show that
H(F) = 0./
By transitivity of the action, whenever ℘ is a branch at P on U, we may find
an element g℘ ∈ G(F℘) such that g℘(xP )Fw p = (xU )F℘ . By hypotheses, we may
find elements gP ∈ G(FP ) and gU ∈ G(FU ) for every P and U such that g℘ = gP gU
whenever ℘ is a branch at P on U. In particular, by replacing xP by g−1 P xP and xU
by gU xU , we may assume that our points satisfy (xP )F℘ = (xU )F℘.
Now, consider these points as morphisms
Spec(FP )
6 NNNxP
llll NNN
lll x℘ NN'
Spec(F℘) /
RRR p ppp8 H
RRR p
R( pppxU
Spec(FU )
shows that the image of each of the morphisms xU , xP , x℘ are the same. But since the
closed fiber Xkred is connected, it follows that we may inductively show the image of
all the morphisms corresponding to points, components or branches must coincide.
Let κ be the residue field of this image point h ∈ H. Then we have field maps
FP fNN
ooooo NNN
N
wo
F℘ gOO κ
OOO ppp
O xppp
FU
Using Lemma 3.4.2, we find that we obtain a map κ → F which one may check
must be a homomorphism of fields. Therefore, we obtain a morphism Spec(F) → H
mapping onto the point h and so H(F) = 0/ as desired.
68 Daniel Krashen
Definition 4.2.1. We say that the local–global principle holds for an algebraic ob-
ject A (of some given type) if for any algebraic object B (of the same type), we have
A∼ = B if and only if AFP ∼ = BFP and AFU ∼ = BFU for all P and U. We say that the
local–global principle holds for a particular type of algebraic object if it holds for
all algebraic objects of this type.
Proof. Suppose that the local–global principle holds for A , and let (A, φ ), (A, ψ )
be two patching problems, such that AP ∼ = (A )FP and AU ∼ = (A )FU for each P,U.
Since we may patch algebraic objects, we may find algebraic objects B1 , B2 over
F whose patching problems are equivalent to (A, φ ), (A, ψ ), respectively. Since
(B1 )FU ∼
= AFU ∼ = (B2 )FU and similarly for FP , we find that by the local–global princi-
ple, B1 ∼
= B2 . Therefore their associated patching problems are isomorphic, implying
(A, φ ) ∼
= (A, ψ ) as desired.
Conversely, suppose that (A, φ )’s isomorphism class is independent of φ for
every patching problem. Suppose we are given A , B be algebraic objects over F
with associated patching problems (A, φ ) and (B, ψ ) respectively. Suppose further
that (A )FU ∼
= (B )FU and similarly for FP . Since AU ∼ = (A )FU ∼
= (B )FU ∼
= BU and
AP ∼= (A )FP ∼
= (B )FP ∼
= BP for all U, P by definition, we may change ψ via these
isomorphisms to find (B, ψ ) ∼ = (A, ψ ) for some ψ . But therefore by hypothesis,
(A, φ ) ∼
= (A, ψ ) ∼
= (B, ψ ). Since patching gives an equivalence of categories, we
further conclude A ∼ = B , completing the proof.
Definition 4.2.4. Let G be an algebraic group over F. We say that the local–global
principle holds for G if for α ∈ H 1 (F, G), with αFP , αFU trivial for each P, U, we
have α trivial.
Note that since elements of H 1 (F, G) correspond to torsors for G, we see imme-
diately that the local–global principle will hold for G if and only if the local–global
principle holds for all G-torsors, in the sense of Definition 4.1.2. Since G-torsors are
transitive G-schemes, from Proposition 4.1.4, we immediately obtain:
Field Patching, Factorization, and Local–Global Principles 69
Proposition 4.2.5. Suppose G is a linear algebraic group defined over F, and sup-
Then the local global principle
pose that factorization holds for G with respect to X.
holds for G.
Remark 4.2.7. Theorem 4.2.6 raises the question of whether it would be possible
to show the equivalence of the local–global principle for a group G and factoriza-
tion for this group without the presence of an algebraic object with G as its au-
tomorphism group. This would give a converse to Proposition 4.2.5. In turn since
G-torsors are, in particular, transitive G-schemes, one would then also obtain a con-
verse to Proposition 4.1.4.
Theorem 5.1.1. Suppose X is a connected normal finite P1T -scheme, with function
field F and let G be a connected retract rational algebraic group over F. Then
factorization holds for G with respect to X.
Using this theorem, we may easily proceed to the proof of the local global prin-
ciple for schemes with transitive action stated earlier in Theorem 2.2.4. If G is a
70 Daniel Krashen
connected retract rational group over F, then by the theorem, factorization holds for
G with respect to X. But then by Proposition 4.1.4, the local–global principle must
hold for transitive G schemes, as desired.
The proof of this theorem will occupy the remainder of the section. Our strategy
will be to reduce this to a more abstract factorization problem, arising from the
case when X = P1T . Overall, the proof stategy is roughly parallel to that followed in
[HHK09], where retractions of open subsets of affine space take the place of open
subsets of affine space.
Note that here we are omitting from the notation the homomorphism G(Fi ) → G(F0 )
for i = 1, 2. Suppose X is a connected, normal, finite P1T -scheme. In this case, we
and we let
set F = F(X),
F1 = ∏P∈S(X)
FP , F2 = ∏U∈U (X)
FU , and F0 = ∏℘∈B(X)
F℘.
Remark 5.1.3. It follows immediately from the definitions that factorization holds
for the group G with respect to X in the sense of Definition 4.1.1 if and only if
factorization holds for G, F, F1 , F2 , F0 in the sense of Definition 5.1.2 where F, F1 ,
F2 , F0 are as above.
Back to the somewhat more abstract setting, suppose that F is some field, and
let L be a finite dimensional commutative F-algebra. Recall that if G is a linear
algebraic group scheme, we may define its Weil restriction, also referred to as its
corestriction or transfer, as the linear algebraic group with the functor of points
defined by:
RL/F G(R) = G(R ⊗F L)
where R ranges through all F-algebras, see [Gro62, Exp. 195, p. 13] for the defini-
tion and Exp. 221, p. 19 of ibid. for proof of existence. We note that the corestriction
in fact comes from a Weil restriction functor from the category of quasi-projective
L-schemes to the category of quasi-projective F-schemes, and that this functor takes
open inclusions to open inclusions, and takes affine space to affine space (of a differ-
ent dimension). In particular, it follows that the corestriction of a rational (or retract
rational) variety is itself rational (resp. retract rational).
We note the following lemma, which is a consequence of the definition of the
corestriction in terms of the functor of points given above.
Lemma 5.1.4. Let F be a field, and suppose we are given rings F ⊂ F1 , F2 ⊂ F0 , and
a finite dimensional commutative F-algebra L. Let G be a linear algebraic group
over L. Then factorization holds for G, L, L ⊗F F1 , L ⊗F F2 , L ⊗ F0 if and only if it
holds for RL/F G, F, F1 , F2 , F0 .
Field Patching, Factorization, and Local–Global Principles 71
Lemma 5.1.5. Suppose that we are given a morphism of connected projective nor-
mal finite P1T -schemes f : Y → X. Let L be the function field of Y and F the function
Then factorization holds for G, Y if and only if it holds for RL/F G, X.
field of X.
Proof. This follows immediately from the universal property of the Weil restriction,
together with Lemma 3.4.1.
Lemma 5.1.6. Let F be the function field of P1T . Suppose that for every connected
retract rational group G over F, factorization holds for G with respect to P1T (as in
Definition 4.1.1). Then for every normal finite P1T -scheme X with function field L,
and every connected retract rational group H over L, factorization holds for H with
respect to X.
Proof. This follows immediately from Lemma 5.1.5.
As a consequence of this, in order to prove Theorem 5.1.1, we may restrict to
the setting where F is the function field of P1T , and where F1 = F∞ , F2 = FA1 , and
k
where F0 = F℘ is the field associated to the unique branch ℘ along A1k at ∞. We let
R0 = R℘, V = R∞ , and W = RA1 . For convenience, in the sequel we will often refer
k
to the following hypothesis for factorization.
Hypothesis 5.1.7 (see [HHK09, Hypothesis 2.4]). We assume that the complete
discrete valuation ring R0 contains a subring T which is also a complete discrete val-
uation ring having uniformizer t, and that F1 , F2 are subfields of F0 containing T . We
further assume that V ⊂ F1 ∩ R0 , W ⊂ F2 ∩ R0 are t-adically complete T -submodules
satisfying V + W = R0 .
Lemma 5.1.8. With respect to the scheme P1T consider F = F(P1T ), F0 = F℘, F1 =
F∞ , F2 = FA1 , R0 = R℘, V = R∞ , W = RA1 . Then these rings and modules satisfy
k k
the Hypothesis 5.1.7.
Proof. The completeness of V, W is satisfied by definition. The fact that V + W =
R0 follows from Lemma 3.4.4.
YO _ _ _/ ? UO _ _ _/? YO
i p
Y
U Y
We choose Y = i−1 p−1Y ⊂ Y . We note that the hypothesis implies that p is domi-
nant, and that the image of the generic point of Y under i lies in Ũ, from which one
then concludes that Y is not empty. As pi is defined on Y , by definition we may
find Y0 ⊂ Y such that pi|Y0 = idY0 . Let U0 = p−1 (Y0 ). Then we have pi(Y0 ) ⊂ Y0 and
so i(Y0 ) ⊂ p−1 (Y0 ) = U. Since p(U0 ) ⊂ Y0 by definition, we have constructed the
desired morphisms.
Lemma 5.2.6 (Retractions shrink to closed retractions). Suppose Y is a retrac-
tion of U via morphisms i, p. Then we may find dense open subvarieties Y0 ⊂ Y and
U0 ⊂ U such that Y0 is a closed retraction of U0 via the restrictions of i, p.
Proof. Since we may identify Y with the image of i, it follows that Y is locally
constructible in U [EGA 4-1, p. 239 (Chevalley’s theorem)]. By [EGA 3-1, p. 12],
it follows that Y is the intersection of a closed and an open set in U. By setting U0
to be this open set, and Y0 = Y ∩U0 it follows that Y0 is closed in U0 . Now it is easy
to see that the restrictions of i, p exhibit Y0 as a retraction of U0 .
Corollary 5.2.7 (Rational retractions shrink to closed retractions). Suppose Y
is a rational retraction of U via rational maps i, p. Then we may find dense open
subsets Y0 ⊂ Y and U0 ⊂ U such that i, p make Y0 a closed retraction of U0 .
Proof. This follows immediately from Lemmas 5.2.5 and 5.2.6.
The following lemma gives us a first hint that retractions inherit some of the
geometry of the larger spaces.
Field Patching, Factorization, and Local–Global Principles 73
as desired.
Lemma 5.2.9 (Standard position for retractions). Suppose Y is a d-dimensional
variety which is a closed retraction of an open subscheme U ⊂ An . We also suppose
that 0 is in Y (with respect to the inclusion of Y in An ). Then we may shrink U
and choose coordinates on U so that Y smooth and is the zero locus of polynomials
f1 , . . . , fn−d with
fi = xi + Pi
where the xi ’s are the coordinate functions on An and Pi is a polynomial in the x j ’s,
each of whose terms are of degree at least 2.
Further, we may alter i and p defining the retraction so that the morphism ip :
U → Y → U is given by
(x1 , . . . , xn ) → (M1 + Q1 , . . . , Mn + Qn )
where
0 if 1 ≤ i ≤ n − d
Mi =
xi if n − d < i ≤ n
∂
and Qi is a rational function in the variables xi , regular on U, such that ∂ x j Qi 0 =0
for all i, j.
Proof. For purposes of skimmability, we have placed this proof at the end of the
section.
The basic strategy for factorization will be to produce closer and closer approxi-
mations to a particular factorization. In order to carry this out, it is necessary to
discuss notions of convergence and approximations in the adic setting, paralleling
the discussion of [HHK09, Sect. 2].
Suppose F0 is a field complete with respect to a discrete valuation v with uni-
formizer t, and let |a| = e−v(a) be a corresponding norm. Let A = F0 [x1 , . . . , xN ], m
74 Daniel Krashen
the maximal ideal at 0, Am the local ring at 0 and A = F0 [[x1 , . . . , xN ]] the complete
local ring at 0. For I = (i1 , . . . , iN ) ∈ N , we let |I| = ∑ j i j . Define for r ∈ R, r > 0
N
I
Ar = ∑ aI x lim |aI |r = 0
|I|
I |I|→∞
r , we set
and for f = ∑ aI xI ∈ A
|(a1 , . . . , aN )| = max{|ai |}
i
and we let D(a, r) be the closed disk of radius r about a ∈ An (F0 ) with respect to
the induced metric. We note that since the values of the metric are discrete, this disk
is in fact both open and closed in the t-adic topology.
We note the following elementary lemma:
r . Then:
Lemma 5.3.1. Suppose a ∈ D(0, r), and f , g ∈ A
r .
1. f + g, f g ∈ A
2. | f + g|r ≤ max{| f |r , |g|r }.
3. For every real number M > 0, the group
r | | f |r < M} ⊂ A
{f ∈ A r
r .
is complete with respect to the filtered collection of subgroups mi ∩ A
4. | f |r is finite.
5. | f g|r ≤ | f |r |g|r .
6. if r < r, then | f |r ≤ max{| f (0)|, rr | f |r }.
7. f (a) is well defined (i.e. is a convergent series).
8. | f (a)| ≤ | f |r , and if f (0) = 0 then | f (a)| ≤ | f |r |a|r−1 .
Lemma 5.3.2. Suppose f is in Am . Then for all ε ≥ | f (0)| with ε > 0, there exists
r and | f |r < ε . Further, for any δ > 0 we may choose r < δ .
r > 0 such that f ∈ A
Lemma 5.3.3. The t-adic topology on AN (F0 ) is finer than the Zariski topology.
Field Patching, Factorization, and Local–Global Principles 75
Proof. This proof is a very slight modification of Lemma 2.2 in [HHK09]. Choose
a real number s so that we may write |h| = ε |t|s . We may rearrange the quantity of
interest as:
f (a + h) − f (a) − L(h) = ∑ cν ((a + h)ν − aν ) .
|ν |≥2
Since the absolute value is nonarchimedean, it suffices to show that for every term
m = cν xν with |ν | ≥ 2 we have
Now ε ≤ |t|r2 , so ε j−1 ≤ |t| j−1 r2 j−2 . Since |t| < 1, r ≤ 1, and j ≥ 2, we have
Rearranging this gives the inequality (ε /r) j ≤ ε |t| and so (ε /r) j |t|s ≤ ε |t|s+1 .
Therefore
|m(a + h) − m(a)| ≤ r− j ε j |t|s ≤ ε |t|s+1 = |t||h|,
as desired.
Proof. By Lemma 5.3.2, since f (0) = 0, we may find 0 < r ≤ 1 such that f ∈ A r
and | f |r ≤ 1. Choose ε ≤ |t|r2 as in the statement of Proposition 5.3.4 and such that
D0 (ε ) ⊂ V (F0 ) and D0 (ε ) ⊂ U(F0 ). Let V = D0 (ε ) ⊂ V (F0 ) and U = D0 (ε ) ⊂
U(F0 ). We claim that for b ∈ U , we have | f (b)| ≤ ε and so f (b) ∈ V . To see this,
we note that Qi ∈ A r and |Qi |r ≤ 1, and hence we may apply Proposition 5.3.4 (with
0 linear and constant term) to see that |Qi (b)| ≤ |t||b| < |b| = max{|bi |}. By the
nonarchimedean property, this gives
We consider first surjectivity. Note that both U and V are both closed and open.
Since they are closed in a complete metric space, they contain all limits of their
Cauchy sequences. Let a ∈ V , and let b0 = 0. We will inductively construct el-
ements bi ∈ U such that | f (bi ) − a| ≤ ε |t|i . In particular, since |a| ≤ ε , we have
| f (b0 ) − a| = |a| ≤ ε . Assuming we have constructed bi−1 , we let h = a − f (bi−1 ),
and note |h| ≤ ε |t|i−1 by hypothesis, and |bi−1 | ≤ ε since bi−1 ∈ U . Therefore, by
Proposition 5.3.4, we have
g : U → Ad the projection onto the last d coordinates. Let U ∈ Ad (F0 ) be the disk
about the origin of radius ε . Let V be the intersection of g−1U with Y (F0 ).
Suppose a ∈ V and b ∈ U with a = b and g(a) = g(b). We claim that b ∈ V . In
particular, this would imply that g|V is injective. To see that b is not in V , first let
h = b − a. If g(a) = g(b), then by definition of g, the last d coordinates of a and b
must match. Since a = b, we therefore know that xi (h) = 0 for some i = 1, . . . , n − d
where xi is the ith coordinate function on Ad . We may therefore choose i such that
|xi (h)| has the largest possible value, and in particular, we then would have |xi (h)| =
|h|. But, estimating | fi (b) − fi (a)| = | fi (a + h) − fi (a)| using Proposition 5.3.4, we
find
| fi (a + h) − fi(a) − xi(h)| ≤ |t||h|.
We claim that | fi (a + h) − fi (a)| ≥ |h| and in particular that fi (a) = fi (b). To see
this must hold, assume by contradiction that | fi (a + h) − fi (a)| < |h|. In this case,
we have
| fi (a + h) − fi(a) − xi (h)| = |xi (h)|
since |xi (h)| = |h|. Therefore we have |h| ≤ |t||h| which is a contradiction since
|t| < 1. Therefore, fi (b) = fi (a). Since V lies in the zero locus of the functions fi ,
we have fi (a) = 0 = fi (b), and so b ∈ V as claimed. Therefore, g|V is injective.
By construction, g|V has image entirely in the ball of radius ε in Ad (F0 ) about
the origin. We claim that it in fact surjects onto this ball (possibly after shrinking
ε ). For this, let a ∈ U , and consider its image b = g(a) ∈ Ad (F0 ). Using the form
for the retraction in Lemma 5.2.9, we may apply Lemma 5.3.5 to the composition
(shrinking ε if necessary)
g|Y
/U p
/Y /U '/
Ad ∩U g Ad
78 Daniel Krashen
5.4 Factorization
Theorem 5.4.1. Under Hypothesis 5.1.7, let f : AdF0 × AdF0 AdF0 be an F0 -rational
map that is defined on a Zariski-open set U ⊆ AdF0 × AdF0 containing the origin (0, 0).
Suppose further that we may write:
f = ( f1 , . . . , fd ) and fi ∈
k[x1 , y1 . . . , xd yd ]m where fi = xi + yi + ∑ cν ,ρ ,i xν yρ .
|(ν ,ρ )|≥2
Then there is a real number ε > 0 such that for all a ∈ Ad (F0 ) with |a| ≤ ε , there
exist v ∈ Vd and w ∈ Wd such that (v, w) ∈ U(F0 ) and f (v, w) = a.
Proof. The proof of this theorem is exactly as for Theorem 2.5 in [HHK09], wherein
in the first paragraph, the problem is reduced to exactly the hypotheses which we
assume.
Theorem 5.4.2. Assume Hypothesis 5.1.7. Let m : Y × Y → Y be a rational F-
morphism defined at (0, 0), and suppose that m(y, 0) = y = m(0, y) where it is de-
fined. Suppose that Y is a closed retraction of an open subscheme of An . Then there
exists ε > 0 such that for y ∈ Y (F0 ) ⊂ An (F0 ), |y| ≤ ε , there exist yi ∈ Y (Fi ), i = 1, 2
such that y = m(y1 , y2 ).
Proof. We consider as in Corollary 5.3.7, t-adic neighborhoods of 0 U ⊂ Ad (F0 )
and V ⊂ Y (F0 ) such that we have bijections U → V and V → U defined by
algebraic rational morphisms p : Ad Y and i : Y Ad . We consider
m|V ×V
V ×O V / V
p i
U ×U / U
Field Patching, Factorization, and Local–Global Principles 79
Proof. Using Lemma 5.2.6, we may find an open subscheme Y ⊂ G that is a re-
traction of an open subscheme U of affine space. In particular, Y must contain an
F-rational point y ∈ Y (F), and after replacing Y by y−1Y if necessary, we may as-
sume Y contains the identity element of G. Using 5.4.2, where m is the multiplica-
tion map, we find that there exists ε > 0 such that factorization holds for g0 ∈ G(F0 )
provided that |g0 | < ε . Fix such an epsilon, and suppose g0 ∈ G(F0 ) is an arbitrary
element. Since G is retract rational, it follows that G(F) is Zariski dense in G(F0 ).
Therefore, we have the existence of an element g ∈ G(F) such that g −1 g0 ∈ Y .
Since Y is a retraction of affine space, it follows that Y (F2 ) is t-adically dense in
Y (F0 ). Therefore, we may find g ∈ Y (F2 ) such that |g −1 g0 g −1 | < ε . Writing
g−1 g0 g−1 = g1 g2 where gi ∈ G(Fi ), we conclude that g0 = (g g1 )(g2 g ). Since
g g1 ∈ G(F1 ) and g2 g ∈ G(F2 ), we are done.
By Lemma 5.1.5 and the comments just following, we conclude that Theorem
5.1.1 holds.
Lemma 5.5.1. Suppose f = g/h for g, h ∈ k[x1 , . . . , xn ] with h(0) = 0, g(0) = 0, and
(∂ f /∂ xi )|0 = 0 for all i. Then if R is a k-algebra with h(0) ∈ R∗ and containing an
element ε ∈ R, ε 2 = 0 then f (ε v) = 0 for v ∈ kn .
We now proceed with the proof of Lemma 5.2.9. By Lemma 5.2.8, we may as-
sume that Y is smooth. Choose f1 , . . . , fr which are regular on a neighborhood of
0 ∈ U and which cut out Y . Writing fi = gi /hi , for gi and hi with no common fac-
tors, we see that since the hi don’t vanish at 0, after shrinking U so that the hi don’t
vanish on U, we may ensure that the hi are units, and hence Y is cut out by the gi .
Therefore, we may assume (after replacing fi by gi and shrinking U) that the fi are
polynomials. Next, we write
fi = Li + Pi
where Li is a linear polynomial and Pi has degree at least 2. Note that fi has no
constant term since it must vanish at 0. Since Y is smooth of dimension d, by the
Jacobian criterion, the Li s (which we may identify with the gradient of fi at 0), span
a n − d dimensional space. After relabelling, we may assume that L1 , . . . , Ln−d give
a basis for this space. Let Y be the zero locus of f1 , . . . , fn−d . Since Y ⊂ Y we have
the codimension of Y at 0, codim0 (Y ) ≤ codim(Y ) = n − d. By construction, the
Jacobian matrix of the defining equations for Y at 0 has rank n − d, and so by [Eis,
p. 402], n − d ≤ codim(Y ). But then
n − d ≤ codim0 (Y ) ≤ codim(Y ) = n − d
so codim0 (Y ) = n − d and also by the Jacobian criterion, we conclude that Y is
smooth at 0. We may therefore, after shrinking U assume that Y is smooth, irre-
ducible, and of the same dimension as Y . But Y ⊂ Y therefore implies Y = Y , and in
particular we may assume r = n − d, and the Li are independent.
After choosing a new basis for An , it is clear that we may assume Li = xi while
preserving our hypotheses.
Finally, consider the morphism γ = ip : U → U (where i and p are as in the
definition of the retraction), and write γ (x) = (γ1 (x), . . . , γn (x)), where each γi is a
regular function on U. Since γi (0) = 0, we may write γi = Mi + Qi for the linear
function
∂
Mi = ∑ γi xi
∂ xi x=0
and have all the partial derivatives of the Qi vanishing. Let T = Speck[ε ]/(ε 2 ), and
consider a T-valued point τ : T → U given by aε = (a1 ε , . . . , an ε ) ∈ An (k[ε ]/(ε 2 )).
We note that τ maps T into Y if and only if fi (aε ) = 0 for each i. But we have (by
Lemma 5.5.1)
fi (aε ) = Li (aε ) = ε Li (a).
Field Patching, Factorization, and Local–Global Principles 81
φ : (x1 , . . . xn ) → (y1 , . . . , yn )
where
xi if 1 ≤ i ≤ n − d
yi =
xi − Mi n − d < i ≤ n
Define rational maps i = φ ◦ i : Y An and p = p ◦ φ −1 : An Y . We then
have p ◦ i = p ◦ φ −1 φ i = pi = idY as rational maps, and therefore define a rational
retraction. By Lemma 5.2.6 we may shrink U and Y to make this a closed retraction.
Note also that i p = φ ipφ −1 = φ γφ −1 .
As before, let τ : T → U be given by aε = (a1 ε , . . . , an ε ) ∈ An (k[ε ]/(ε 2 )), where
a ∈ An (k). We consider the morphism i p : U → Y → U, which we write as
with Ni linear and the first derivatives of the Pi vanishing at the origin. Computing
using Lemma 5.5.1 applied to functions Pi , we then find
and so
0 if 1 ≤ i ≤ n − d
Ni =
xi if n − d < i ≤ n
Therefore, upon replacing p, i by p , i , we obtain the desired conclusion.
Acknowledgements The author is grateful to David Harbater, Julia Hartmann, and the anonymous
referee for numerous helpful comments and corrections in the preparation of this manuscript.
References
[ASS] S.A. Amitsur, D.J. Saltman, and G.B. Seligman, editors. Algebraists’ homage: papers
in ring theory and related topics, volume 13 of Contemporary Mathematics, 1982.
American Mathematical Society, Providence, RI.
[Bor] A. Borel. Linear algebraic groups, volume 126 of Graduate Texts in Mathematics.
Springer-Verlag, New York, second edition, 1991.
[CTPS] J.-L. Colliot-Thélène, R. Parimala, and V. Suresh. Patching and local-global principles
for homogeneous spaces over function fields of p-adic curves, to appear in Commen-
tarii Math. Helv.
82 Daniel Krashen
[DI] F. DeMeyer and E. Ingraham. Separable algebras over commutative rings. Springer-
Verlag, Berlin, 1971.
[Eis] D. Eisenbud. Commutative algebra, volume 150 of Graduate Texts in Mathematics.
Springer-Verlag, New York, 1995.
[Gra] J.W. Gray. Formal category theory: adjointness for 2-categories. Lecture Notes in
Mathematics, Vol. 391. Springer-Verlag, Berlin, 1974.
[Gro62] A. Grothendieck. Fondements de la géométrie algébrique [Extraits du Séminaire
Bourbaki, 1957–1962]. Secrétariat mathématique, Paris, 1962.
[EGA] A. Grothendieck. Éléments de géométrie algébrique. 3–1: Inst. Hautes Études Sci.
Publ. Math., 11:167, 1961; 4–1: Inst. Hautes Études Sci. Publ. Math., 20:259, 1964;
4–4: Inst. Hautes Études Sci. Publ. Math., 32:321, 1967.
[HH] D. Harbater and J. Hartmann. Patching over fields. To appear in Israel J. Math., arxiv:
0710.1392.
[HHK] D. Harbater, J. Hartmann, and D. Krashen. Patching subfields of division algebras. to
appear in Trans. Am. Math. Soc., preprint available at arxiv:0904.1594.
[HHK09] . Applications of patching to quadratic forms and central simple algebras.
Invent. Math., 178:231–264, 2009.
[Hov] M. Hovey. Model categories, volume 63 of Mathematical Surveys and Monographs.
American Mathematical Society, Providence, RI, 1999.
[Ill] L. Illusie. Frobenius et dégénérescence de Hodge. In Introduction à la théorie de
Hodge, volume 3 of Panor. Synthèses, pages 113–168. Soc. Math. France, Paris, 1996.
[KMRT] M.-A. Knus, A. Merkurjev, M. Rost, and J.-P. Tignol. The book of involutions. Amer-
ican Mathematical Society, Providence, RI, 1998.
[ML] S. Mac Lane. Categories for the working mathematician. Springer-Verlag, New York,
second edition, 1998.
[Sal] D.J. Saltman. Retract rational fields and cyclic Galois extensions. Israel J. Math.,
47(2–3):165–215, 1984.
[Ser] J-P. Serre. Local fields. Springer-Verlag, New York, 1979.
Deformation Theory and Rational Points
on Rationally Connected Varieties
Max Lieblich
1 Introduction
The question we address in these notes is the following. Fix once and for all
an algebraically closed field k. In all sections but Sect. 2, we assume that k has
characteristic 0.
Lest this seem like a strange question, let us remind the reader of two theorems
along these lines when the base has lower dimension.
This is the Nullstellensatz: any maximal ideal in a finite-type k-algebra has residue
field k. Increasing the dimension of the base, we have the following more recent
result.
Max Lieblich
Padelford Hall, University of Washington, Seattle, WA 98195, USA
e-mail: [email protected]
2010 Mathematics subject classification. Primary: 14D06. Secondary: 14G05.
How can we tell if Homc →y (C,Y ) is positive dimensional? Let us start with an
easier question: how can we tell if Hom(C,Y ) is positive dimensional? One thing
we can do is to calculate the tangent space to the scheme at a point ϕ : C → Y . The
calculation is one of the gems of algebraic geometry.
Write S = Spec k[ε ]/(ε 2 ). It is a standard fact that the tangent space to a k-scheme
Q at a point q : Spec k → Q is identified with the set of maps S → Q such that
the composition with the natural closed immersion Spec k → S (given by setting ε
equal to 0) is q. For the scheme Q = Hom(C,Y ), the tangent space at a morphism
ϕ : C → Y is thus described by families of maps S ×C → Y reducing to ϕ when ε = 0.
These are basic objects of deformation theory: first-order infinitesimal deformations
of ϕ .
Proof. Suppose for a moment that all of the schemes are affine, so that C = Spec A
and Y = Spec B with A and B two k-algebras. The morphism ϕ then corresponds
to a k-algebra homomorphism γ : B → A, and an infinitesimal deformation of ϕ
corresponds to a k[ε ]-algebra homomorphism B[ε ] → A[ε ] which equals γ modulo ε .
A routine calculation identifies such extensions with B-linear derivations B → A. By
the universal property of the module of differentials, such derivations are identified
with B-linear module homomorphisms ΩB → A, and these are the same as A-linear
module homomorphisms ΩB ⊗ A → A. Dualizing, such maps are elements of the
tangent module TB ⊗ A. Covering C with affine schemes which map into affine open
subschemes of Y , we can apply the preceding argument to give local sections of
ϕ ∗ TY ; these glue together to give a global section, as desired.
Now, even if we can show that the tangent space Tϕ Hom(C,Y ) is positive di-
mensional, this does not mean that the scheme is positive dimensional. To settle the
issue, we would like to be able to compute all the jets at a point and not merely the
tangent space. In other words, we would like to describe the structure of the com-
plete local ring OHom(C,Y ),ϕ . This requires a broadening of our deformation theory.
Write Sn = Spec k[x]/(xn+1 ). Given a k-morphism ϕ : C × Sn−1 → Y (which is
the same as an Sn−1 -morphism C × Sn−1 → Y × Sn−1 ), we can try to understand
its deformations to a morphism C × Sn → Y . Write ϕ0 : C → Y for the restriction
of ϕ to the closed subscheme arising from the closed immersion Spec k → Sn−1 .
An analysis similar to that carried out above shows that for any affine SpecA ⊂ C
such that ϕ (Spec A) lands in an affine open Spec B of Y , there is an extension of ϕ
to C × Sn , and that extensions are parameterized by (are a principal homogeneous
space under) H0 (Spec A, ϕ0∗ TY ). Let {Ui }m i=1 be such an affine open covering of C.
Given i = j, choosing deformations of ϕUi yields two sections of ϕ0∗ TY over Ui ∩U j .
It is an amusing exercise to show that this is a 1-cocycle with coefficients in ϕ0∗ TY .
In fact, we have the following proposition.
In other words, the vector bundle ϕ0∗ TY governs the infinitesimal deformation
theory of the morphism ϕ0 : C → Y . A more careful study of the properties of the
deformation theory yields the following result, due to Schlessinger (implicit in the
proof of Theorem 2.11 of [13] and made explicit using obstruction theory in Propo-
sition 2.A.11 of [7]). Write di = dimk Hi (C, ϕ0∗ TY ) for i = 0, 1.
χ (C, ϕ0∗ TY (−c)) = deg ϕ0∗ TY (−c) + dimY (1 − g(C)) = − deg ϕ0∗ KY − g(C) dimY.
On the other hand, we know by Proposition 2.10(3) that dimϕ0 Homc →y (C,Y ) ≥
χ (C, ϕ0∗ TY (−c)). The result thus follows from Theorem 2.2 if we can show that
a positive-dimensional family of morphisms C → Y sending c to y containing
ϕ0 as a fiber must have image of dimension at least 2. Suppose to the con-
trary that T is a smooth curve and ϕ : C × T → Y is a morphism such that
ϕ ({c} × T ) = y and ϕ (C × T ) is a curve C ⊂ Y . Replacing C by its normal-
ization (T being itself normal), we may thus assume that Y is a smooth proper
curve of positive genus and C × T → Y is a family of surjective morphisms send-
ing c to y. But now the tangent space to Homc →y (C,Y ) at ϕ0 has dimension
dim H0 (C, ϕ0∗ TY (−c)) = 0 because deg ωY ≥ 0. In other words, the family corre-
sponds to a map T → Homc →y (C,Y ) to a discrete space and thus to a constant
family of maps, contradicting our assumption.
Deformation Theory and Rational Points on Rationally Connected Varieties 89
Putting Corollaries 2.11 and 2.13 together yields the following result. Recall that
a smooth proper variety Y is Fano if −KY is an ample divisor.
This seems to imply that Fano varieties should be covered by rational curves. We
briefly recall a basic definition.
In this section, we discuss a more systematic approach to the ways in which curves
can break when they are moving. This will lead us to stable curves and the Kontse-
vich moduli space of stable maps. We also review some classical moduli problems
92 M. Lieblich
and discuss how the method of degeneration to the boundary yields numerous re-
sults on the structure of moduli spaces. As we will see in Sect. 4, the techniques of
de Jong and Starr fit into the classical framework in which moduli problems come
with an inductive structure provided by the boundary.
In Sect. 2, we described the deformation theory of morphisms and how it can
be used to break a curve apart. In this section, we spend some time describing the
deformation theory of varieties and how this can be used to smooth a singular curve
in a variety. The easiest way to construct a degenerating family of varieties is to blow
up a trivial family. The resulting family will be surprising useful in what follows.
Proposition 3.2. Base change yields a bijection between the set homT (C ,Y × T )
and the set lim homTn (C ×T Tn ,Y × Tn ) of compatible systems of infinitesimal mor-
←−
phisms.
The second condition of Definition 3.3 is equivalent to the statement that the
morphism f has finite automorphism group. As a special case, when Y is a single
point, we recover the usual Deligne–Mumford notion of a stable curve of genus g
(and here the second condition is equivalent to finiteness of the automorphism group
of C).
The basic result on stable maps is the following, which is a special case of
Theorem 1 of Fulton and Pandharipande [4].
When we say that β is a “homology class,” the reader can choose to work with Chow
theory (modulo algebraic equivalence) or with integral or -adic cohomology. The
original work was done over C, so it is customary to fix β ∈ H2 (Y, Z).
Remark 3.5. When we say “moduli space,” we are purposely leaving things slightly
vague. The reader comfortable with stacks should read this as “Deligne–Mumford
stack” and should replace “projective” with “proper with projective coarse moduli
space.” In this case, an isomorphism between two objects C → Y and C
→ Y is
an isomorphism C → C
commuting with the morphisms to Y . Readers unfamiliar
with the theory of stacks will not lose anything essential, and should think we are
working with coarse moduli spaces as in [4]. In this case, the points of M g,o (Y, β )
parameterize isomorphism classes of stable maps C → Y .
Using modern techniques, the proof of this theorem is not especially difficult
(although projectivity is a bit more subtle than existence). Existence follows rela-
tively easily from the existence of the Deligne–Mumford moduli space M g of stable
curves of genus g; the existence of the latter is somewhat less trivial. In fact, it was
in Deligne and Mumford’s paper [3] on the irreducibility of Mg that the concept
of an algebraic stack was first codified. One especially streamlined way to see that
M g exists as an algebraic stack is to use Artin’s representability theorem, which re-
duces the algebraicity to properties of the natural deformation theory of the moduli
problem, something that is transparent for families of curves.
One especially amusing example of a Kontsevich space is the space M g,0 (C, [C]),
where C is a fixed smooth curve of genus g and [C] is the fundamental class of C
in H2 (C, Z). What does this parameterize? A stable map f : D → C with homology
class f∗ ([D]) = [C] has the following properties:
94 M. Lieblich
In other words, M g,0 (C, [C]) has only one point. (Two different isomorphisms
C
→ C differ by an automorphism of C
and thus give rise to the same point in
the moduli space.)
On the other hand, consider the space M g,0 (C, e[C]) for some integer e > 1,
and let us restrict our attention to the case C = P1 . Now, we can make interesting
elements in the space.
Remark 3.8. It is a remarkable fact that there is actually a unique irreducible compo-
nent of M g,0 (P1 , e[P1 ]) containing those points parameterizing stable maps which
are flat. (Note that even in this irreducible component, the limit points need not be
flat.) We will call this component F.
For large enough e, which is all that is needed here, one can prove the unique-
ness of F rather easily as follows. Let F ◦ denote the locus parameterizing flat stable
maps σ : C → P1 of genus g with C smooth. The forgetful functor gives a morphism
Φ : F ◦ → Mg to the stack of stable curves of genus g, which we claim is smooth for
large enough values of e. Since Φ has connected geometric fibers (Grassmannian
bundles over PicCe ), it will follow that F ◦ is itself irreducible (by Theorem 3.10). To
see that Φ is smooth, note that morphisms σ correspond to invertible sheaves L on
C of degree e along with a pair of generating sections OC⊕2 L . Given an infinites-
imal deformation of C over k[ε ], the invertible sheaf L lifts (as the obstruction to
doing so lies in H2 (C, O) = 0), and the obstruction to lifting the generating sections
lies in H1 (C, L )⊕2 , which vanishes as long as e > 2g − 2.
Now, let D → P1 be a flat stable map of genus g with D nodal. By an argument
identical to that following Proposition 3.11 (treating D as a lci P1 -scheme), one can
show that D → P1 deforms to a map D
→ P1 with D
smooth. It follows that any
flat stable map is in the closure of F ◦ , showing that there is thus a unique irreducible
component containing F ◦ , as desired.
The geometry of the Kontsevich spaces (in particular, the uniqueness of the com-
ponent F in Remark 3.8) leads to the existence of sections for rationally connected
fibrations over curves.
Deformation Theory and Rational Points on Rationally Connected Varieties 95
DK K
KKK
KKρK
KKK
K%
∼
=
s9 Y
sss
ss
ss
ss Φ |C ×T SpecK
C ×T Spec K
and χ restricted to the closed point of R is one of the curves given by Construc-
tion 3.7. Let η ∈ D be the generic point of one of the irreducible components of
Dκ . Consider the proper morphism g : Y ×P1 SpecOD ,η → Spec OD ,η . By construc-
tion there is a section of g over the generic point of SpecOD ,η , so by the valuative
criterion of properness there is a section of g. Evaluating g at η (and extending
to a section again using the valuative criterion of properness) yields a morphism
P1κ → Yκ giving a section of Yκ → P1κ . Since the base field is algebraically closed,
it follows from standard limiting and specialization arguments (write κ as a direct
limit of finite-type k-algebras, etc.) that the existence of a section over κ implies the
existence of a section over k.
The strategy employed by Graber, Harris, and Starr is thus to find a multisection
C ⊂ Y whose deformations inside Y dominate F ⊂ Mg,0 (P1 , e[P1 ]). Now, a random
curve C will not have enough deformations in Y to dominate F. But we can make
curves of genus g that admit larger deformation spaces when we know that the
general fibers of f are rationally connected. Here is how: given C ⊂ Y , choose points
c1 , . . . , cn which lie in general (smooth rationally connected) fibers Y1 , . . . ,Yn . Since
96 M. Lieblich
the Yi are rationally connected, there are many rational curves Ri on Yi through ci
(in fact, the rational curves generate the tangent space to Yi at ci ) such that the
tangent bundle TYi is ample when restricted to Ri . Given such curves R1 , . . . , Rn , we
can form the nodal union C
= C ∪ R1 ∪ · · · ∪ Rn , which gives a stable map to Y .
Since each Ri is rational, C
still has genus g. Moreover, one can show that given m,
for sufficiently large n we have that the normal sheaf N f
to the map f
: C
→ Y
is generated by global sections taking any prescribed values on m smooth points
q1 , . . . , qm of C
, and that H1 (C
, N f
(−q1 − · · · − qm )) = 0. The latter group is the
group of obstructions to deforming C
as a Y -scheme; combining both calculations,
we see that C
→ Y deforms to a stable map C
→ Y with C
together at each point), then this high degree of deformability shows that the defor-
mations of C
in Y are dominant over the component F ⊂ Mg,0 (P1 , e[P1 ]), allowing
us to conclude as in the second paragraph above. Producing such a curve in Y that
is only mildly ramified over P1 is beyond the scope of this survey and the reader is
referred to Sect. 3 of Graber et al. [5] for the details.
This proves Theorem 3.9 when the base curve is P1 . As observed by de Jong,
generalizing the result to other base curves B is a triviality using the technique of
Weil restriction: choose a finite generically étale morphism π : B → P1 and consider
the variety π∗Y whose functor of points sends T → P1 to HomB (T ×P1 B,Y ). One
checks that π∗Y is indeed a variety (e.g., it is a proper algebraic space with a finite
cover by a projective scheme) with a proper morphism π∗Y → P1 . The description of
the functor of points shows that for a general point p ∈ P1 with preimages {q ∈ B},
the fiber of π∗Y over p is isomorphic to ∏ Yq . Thus, π∗Y → P1 has rationally
connected geometric generic fiber. Since π∗Y (P1 ) = Y (B), the result follows.
The short form of the proof. Attach vertical rational curves to a multisection so that
it deforms enough to break into a union of sections (and possibly a few vertical com-
ponents). This combines the techniques of deforming off of the boundary (smooth-
ing the singular curve) and degenerating to the boundary (following the moving
curve to the nice part of the boundary of M g,0 (P1 , e[P1 ]) which consists of singular
curves containing sections). Continuing along these lines, one can in fact see that
there are many sections of the morphism, with any nonempty open subscheme of Y
containing infinitely many sections [8, Theorem IV.6.10].
As we have seen, “degenerating to the boundary” is a way of carrying complex
information into a simpler situation (and therefore achieving greater understanding
using the surplus). As another illustration of this principle, we conclude this section
by describing the mapping space to a point: the moduli space Mg of curves of
genus g. In particular, we will sketch the basic idea of Deligne and Mumford’s proof
that Mg is geometrically irreducible. The proof is an archetype: one compactifies
the moduli space by enlarging the moduli problem, and the boundary then endows
the moduli problem with an inductive structure in which the interiors of “lower”
problems appear as boundary strata of “higher” problems, allowing a transfer of
Deformation Theory and Rational Points on Rationally Connected Varieties 97
information up the chain and (in many cases) resulting in an irreducibility proof.
This kind of inductive structure is ubiquitous in the study of moduli and will appear
in the next section when we discuss de Jong and Starr’s theory of porcupines.
Note that as stated this is not a theorem about the space of stable curves, only
about the space of smooth curves. In fact, the space of stable curves was introduced
to compactify the space of smooth curves in such a way that the argument we sketch
here (using the inductive structure of the compactification) would be possible. Here
is how it works. Assume that g > 1:
Step 1. Embed Mg as an open subspace of the space M g of stable curves of
genus g.
Step 2. Show that the embedding from step 1 is dense, and never an isomorphism.
We can accomplish both of these things by studying the deformation theory of
singular stable curves. For example, a nodal union E of a curve of genus g − 1
with a curve of genus 1 will be a stable curve of genus g. We can smooth such
a nodal union by using infinitesimal deformation theory: the formal smooth-
ing of a node works just as in the blowup picture in Degeneration 3.1. On the
other hand, the infinitesimal deformation theory of the curve E maps to the in-
finitesimal deformation theory of a node. One checks using abstract nonsense
that this map is formally smooth. This says that a formal smoothing of the node
can be extended to a formal smoothing of E. More generally, one can make
this argument for any stable curve. Thus, there are singular stable curves, and
any singular stable curve can be deformed to a smooth one. This shows that the
inclusion is dense and never an isomorphism, as desired.
It is worth making this more precise, as it gives another illustration of basic de-
formation theory. For the sake of simplicity, we will study deformations over the
bases Sn used in Sect. 2.
Proposition 3.11. Let X → Sn−1 be a flat morphism of finite type with X0 a curve
with at worst nodal singularities. Then
1. There is no obstruction to deforming X to a flat scheme over Sn .
2. The space of deformations is a principal homogeneous space under the group
Ext1 (ΩX0 , OX0 ). Moreover, this description is functorial in local immersions
on X.
Given a stable curve C with nodal locus N, applying the proposition to C and the
localization of C at N, and noting (via the spectral sequence connecting local to
global Ext) that Ext1 (ΩC , OC ) is concentrated on N, gives the statement on surjec-
tivity and formal smoothness of the restriction map on infinitesimal deformation
theories. More precisely, writing FC,N for the restriction of a coherent sheaf F
98 M. Lieblich
on C to the semilocal ring OC,N , the spectral sequence and vanishing of H2 give a
surjection (as an edge map)
which one checks corresponds to the natural restriction map, which in turn corre-
sponds to the restriction of infinitesimal deformation theories. The rest (smoothing
the node and following this by a smoothing of the curve C) uses the second pillar:
results similar to those of Proposition 3.2 for the integration of infinitesimal defor-
mations of curves (instead of morphisms).
Step 3. Show that the space M g is proper (in fact, projective). This is the famous
stable reduction theorem.
Step 4. Show that the space M g is normal (in fact, has finite quotient similarities).
This is a formal consequence of the fact that the infinitesimal deformation the-
ory of a curve is formally smooth; a proper understanding of this point uses the
theory of stacks.
Step 5. To show that a normal variety is irreducible, it suffices to show that it is
connected. In other words, to show that Mg is irreducible, it is enough to show
that given two points a and b of Mg , there is a connected curve C in M g which
contains a and b.
Step 6. To produce such a path, we proceed as follows. The point a parameterizes a
smooth curve D of genus g. Using a stable reduction theorem, we can degener-
ate D to a nodal union D0 of curves of lower genera; this degeneration connects
D to D0 by a curve in M g . We can do a similar thing for the curve D
parameter-
ized by b, yielding another singular curve D
0 . On the other hand, by induction
we know that the moduli spaces Mh with h < g are irreducible. Taking enough
care with the degenerations, we can use this inductive fact to prove that there is
a curve in M g connecting D0 to D
0 .
Step 7. We are done! (Obviously, there are many technical details that are left out
of this sketch. In particular, we did not explain how to do the degenerations
properly to ensure that the limits are connected by a curve in a simple way.)
We follow the same basic pattern as in the Graber–Harris–Starr theorem: we
combined degenerating to the boundary with deformation off of the boundary
to connect together points in the open moduli problem that was originally of
interest to us.
In Sect. 4, we will see another technique for putting an inductive structure on a
moduli problem. While the problem in question is a moduli problem of stable maps,
the inductive structure on the problem is actually more reminiscent of what hap-
pens for the moduli of vector bundles on algebraic surfaces: the structural results
described (irreducibility, rational connectivity) only hold asymptotically in a dis-
crete parameter (in contrast to the irreducibility of Mg for any genus g). For moduli
problems of this type, the inductive structure involves an increasing understanding
of the problem as one ascends the chain of moduli spaces, and not merely a transfer
of perfect information from lower problems to higher ones.
Deformation Theory and Rational Points on Rationally Connected Varieties 99
In this section, we fix a smooth projective surface S over k. Our goal is to under-
stand when there exists a rational section to a morphism f : X → S whose fibers are
sufficiently rationally connected. As explained in Sect. 1, there are many rationally
connected fibrations over S which lack rational sections. If we suppose for a moment
that X and S are CW complexes with S two dimensional, then a consideration of the
long exact sequence in homotopy associated to a fibration shows that if the fibers of
f are simply connected, then there is no topological obstruction to lifting the largest
cells of S into X. This is not meant to give any kind of an argument in the classical
complex topology; for example, the dimensions are wrong. Instead, this is meant to
inspire us to think of what it should mean for the fibers of f to be simply connected
in some algebraic sense. Ultimately, de Jong and Starr came up with a careful list
of hypotheses that seem to express simple connectivity in the context of rational
curves.
To motivate the definition, let us consider an approach to finding sections to f .
Let C ⊂ S be a general smooth curve. (Here “general” can be taken to mean that
the fiber of f over C still has smooth total space and rationally connected geometric
generic fiber.) By the Graber–Harris–Starr theorem, the restriction fC has many sec-
tions. In fact, there is a moduli space Σ (XC /C) → Speck parameterizing sections.
Moreover, this moduli space varies nicely as C varies. (Of course, Σ (XC /C) can be
complicated – it can have many connected components, etc. – but we still get a rea-
sonable family of such moduli spaces as the curve C moves in S.) The question is:
how can we properly formulate what it would mean to make a family of sections of
f over a variable curve C?
There is a well-known resolution of this type of question which illustrates the
power of Grothendieck’s theory of generic points. Suppose the curve C is a member
of a very ample linear system on S. We can choose two members C1 and C2 of the
linear system |C| which are smooth and intersect transversely in the locus of S over
which f is smooth. The total space S̃ of the pencil spanned by C1 and C2 will be
the blowup of S in the intersection C1 ∩ C2 . Since S̃ and S are birational, to find a
rational section of f it is enough to find a rational section of the pullback of X to
S̃. On the other hand, there is a natural proper flat generically smooth morphism
S̃ → P1 . Thus, we may assume that our original base surface S fibers are a family of
curves over P1 ; this expresses our surface as the total space of a moving family of
curves.
Now, what does it mean to paste together sections of the fibration over these
curves? Such a compatible family of sections should be precisely a section of the
fibration X → S over the generic member of the family parameterized by P1 . In
other words, it should be a section of the restriction of f to the generic fiber of
S̃ → P1 . This generic fiber is a proper smooth geometrically connected curve C over
the field k(t).
100 M. Lieblich
To summarize: we can solve our original problem if we can solve the analogous
problem for fibrations f : X → C with proper smooth rationally simply connected
fibers over curves, but now over the base field k(t). Associated to the fibration f is a
moduli space Σ → Speck(t) parameterizing sections. If we knew that suitable com-
ponents of Σ were rationally connected, then we could apply the Graber–Harris–
Starr theorem to conclude a section exists. This is precisely what de Jong and Starr
do. What is perhaps more interesting, they deduce the rational connectivity of certain
spaces of sections from a condition on the generic fiber of f which is analogous (by
a loose topological analogy) to the path connectedness of spaces of paths between
two points (where now a space of paths is to be interpreted as a space of rational
curves, and path connectedness is to be interpreted as rational connectedness).
With this introduction, we are now ready to put some conditions on f which will
allow us to make all of this (more) rigorous. We will suppose the following. (de Jong
and Starr allow more general hypotheses, but nothing essential about their argument
is changed by using the stricter hypotheses we have chosen here.)
Hypothesis 1. The morphism f : X → C is a proper flat morphism of smooth ge-
ometrically connected varieties over k(t). Following Jason Starr’s notes [14],
we will write Y for the geometric generic fiber of f (so Y is a variety over
K = k(C)).
Hypothesis 2. There is an invertible sheaf L on X which is ample and globally
generated.
Hypothesis 3. The restriction L |Y is very ample.
Hypotheses 2 and 3 together say that L defines a morphism to a projective space
and that this morphism restricted to a general fiber is finite. There may be “horizon-
tal” collapsing induced by the map, but no curve in a general fiber is collapsed.
To state the rest of the hypotheses, we first need to develop a small amount of
theory. For the moment, let us work with the geometric generic fiber; thus, we re-
strict our attention to the variety Y over the field K. Since L is very ample on Y ,
we can imagine Y embedded in some PN in such a way that L is the restriction
of O(1). Any curve D ⊂ Y has a degree (via the induced embedding in PN ). The
curves of degree 1 in Y are precisely the lines in PN which lie in Y . For historical
reasons (and to harmonize our notation with that of Starr), we will denote the space
of pairs (L, p) with L a line in Y and p a point of L by M 0,1 (Y /K, 1). Since lines
cannot degenerate, the space is actually projective. There is a natural morphism
ev : M 0,1 (Y /K, 1) → Y which is defined by sending a pair (L, p) to the point p ∈ Y .
Recall the following definition.
Definition 4.1. A closed subscheme Z ⊂ W of a smooth variety is free if the restric-
tion TW |Z is globally generated.
Following de Jong and Starr, given a natural number n, a chain of n-free lines is a
sequence of triples (Li , pi , qi ) with each Li a free line in Y and pi and qi two points
of Li , such that for each i we have that qi−1 = pi , with this intersection point being
a node of the total curve given by the union of the Li . It is easy to see that there is
a quasiprojective K-scheme F2 (Y /K, n) parameterizing chains of n-free lines, with
Deformation Theory and Rational Points on Rationally Connected Varieties 101
p1 and qn marked on the total curve. This again comes with an evaluation morphism
ev : F2 (Y /K, n) → Y × Y given by evaluating maps at p1 and qn .
Using these two evaluation morphisms, we can state the next two hypotheses:
Hypothesis 4. The evaluation morphism M 0,1 (Y /K, 1) → Y is surjective, the geo-
metric generic fiber is irreducible and rationally connected, and the fibers over
codimension one points of Y are only allowed to be mildly singular (in a fashion
which de Jong and Starr make precise).
Hypothesis 5. For some natural number n, the evaluation morphism F2 (Y /K, n) →
Y × Y is dominant, and the geometric generic fiber is isomorphic to an open
subscheme of a projective rationally connected scheme.
To state the final hypothesis, I need to remind you of one more geometric object. A
scroll in a projective space PN is a smooth surface R ⊂ PN together with a smooth
morphism R → P1 whose fibers are lines.
Hypothesis 6. There is a very twisting scroll in Y . This is a special type of scroll
R ⊂ Y which comes with a section, the whole package required to satisfy some
cohomological conditions. In particular, R should be free in P1 ×Y (in the sense
of Definition 4.1). (There is more, and the meaning is very clearly explained in
the lecture notes of Starr [14].)
As Starr points out, all of these hypotheses are necessary for the veracity of the the-
orem which we will ultimately state. The reader should know that these hypotheses
are algebraic analogues of natural topological conditions: Hypothesis 4 says that the
space of paths through point is path connected (a trivial hypothesis in usual topo-
logical settings), while Hypothesis 5 says that the space of paths connecting two
points is itself path connected (which is equivalent to being simply connected in the
usual topological settings). The fact that Hypothesis 6 has no reasonable (nontauto-
logical) topological analogue gives an indication of the complexity of the algebraic
situation.
Before I can state the theorem, I need to introduce one more object: the Abel
map on the space of stable sections. Using the invertible sheaf L , we can define a
degree for any stable map D → X : we simply compute the degree of the preimage
of L on D.
It is relatively easy to see that stable sections of degree e form a projective moduli
space, which we will denote Σ e (X /C/k(t)). Colloquially, we get D → X by taking
a section of f and attaching some vertical components in the fibers of f . The hori-
zontal component is referred to as the “handle” of the stable map. Among the stable
sections of f of degree e are those which are actually sections C → X whose degree
is e. The sections form an open subscheme which we will denote Se (X /C/k(t)).
Note that Se (X /C/k(t)) is nonempty by Theorem 3.9.
102 M. Lieblich
has lower degree, the vertical components make up the difference.) This morphism
α : Σ e (X /C/k(t)) → PicC/k(t)
e
As we will see below, there is a fairly explicit description of the irreducible com-
ponent Ze using porcupines. But before we give the (mildest) indication of how one
proves such a theorem, let me show you how to deduce the existence of rational
sections from this theorem.
α : Me ⊗ k(t) → PicC/k(t)
e
⊗k(t)
has the property that the restriction of α to each irreducible component of Me ⊗ k(t)
is dominant with geometrically irreducible generic fiber.
Idea of proof. There is a beautiful trick here: fix a component Ze of Me ⊗ k(t). Since
the porcupines are smooth points of Ze , to show that α |Ze is dominant with geomet-
rically irreducible generic fiber, it suffices to prove the statement for the locus of
porcupines W ⊂ Ze . It is now relatively easy to analyze the Abel map: taking the
body and quill attachment points gives a map W → Ze0 × Syme−e0 (C), where Ze0 is
now a smooth scheme parameterizing free sections (the image of Ze in Me0 ⊗ k(t)
under the natural rational map). This is smooth and dominant with geometrically
irreducible fibers by Hypothesis 4 and the definition of peaceful.
The Abel map factors through the product
Pic
e0 e−e
× Pic 0 / Pice ,
C⊗k(t)/k(t) C⊗k(t)/k(t) C⊗k(t)/k(t)
Deformation Theory and Rational Points on Rationally Connected Varieties 107
where the first map is the divisor class map and the last is the tensor product of in-
vertible sheaves. For large enough e, the divisor class map is smooth and surjective.
e
Fix s ∈ Ze0 with Abel image L ∈ Pic 0 (k(t)). Translating by elements
C⊗k(t)/k(t)
of Pice−e0 covers Pice , so the composition of the latter two maps is surjective.
On the other hand, fixing L ∈ PicC⊗k(t)/k(t)
e
(k(t)) and s ∈ Ze0 , there is a unique
L
∈ Pic 0 (k(t)) such that L
+ α (s) = L, namely L − α (s)! This shows
e−e
C⊗k(t)/k(t)
that the fibers of the composition are in bijection with Ze0 , hence geometrically
irreducible.
Combining the statements shows that the Abel map on W is dominant with geo-
metrically irreducible generic fiber, as desired.
Corollary 4.17. For sufficiently large e, a general point of any irreducible compo-
nent of Me ⊗ k(t) maps to a general point of PicC/k(t)
e
⊗k(t).
Remark 4.19. The reader will note that the chain of curves need not stay in the locus
of porcupines. This is why the results mentioned here do not show that the locus of
porcupines has rationally connected Abel fibers.
Combining this with the lemma, we see that given two general porcupines
D and D
, up to adding quills to each, we can connect them by a chain of ratio-
nal curves lying in a single irreducible component of Σ e (X /C/k(t)) with nodes at
smooth points. This completes the proof of Theorem 4.14.
To complete the proof of Theorem 4.3(2) (that general fibers of the Abel map
are rationally connected), we thus need to see how to connect a general point of Me
to a general porcupine with the same Abel image. It is here that one must use the
108 M. Lieblich
existence of a very twisting scroll in X. In fact, this part of the argument is quite
difficult, as one sees must be the case: it requires understanding a general point of
the mysterious component Me , when all we know about is a part of its boundary –
the locus of porcupines.
The work of de Jong and Starr is a tour de force of algebraic geometry, and
yet it barely scratches the surface of the concept of rational simple connectiv-
ity. What is the most general form of the condition? Is there a theorem on the
existence of rational sections for fibrations with higher Picard numbers? (What
about higher-dimensional bases? Will we soon be reading about connecting “gen-
eral Dimetrodons” by rational curves?) The recent theorems on rational sections of
fibrations are a sign that we have hit a rich seam at the intersection of arithmetic,
geometry, topology, and deformation theory. Much digging remains to be done!
Acknowledgements I am grateful to Aise Johan de Jong and Jason Starr for discussing their ideas
with me, and to the referee and Jean-Louis Colliot-Thélène for helpful comments and corrections.
References
1. F. Campana. Une version géométrique généralisée du théorème du produit de Nadel. Bull. Soc.
Math. France, 119(4):479–493, 1991
2. O. Debarre. Higher-dimensional algebraic geometry. Universitext. Springer, New York, 2001
3. P. Deligne, D. Mumford. The irreducibility of the space of curves of given genus. Inst. Hautes
Études Sci. Publ. Math., 36:75–109, 1969
4. W. Fulton, R. Pandharipande. Notes on stable maps and quantum cohomology. In Algebraic
geometry – Santa Cruz 1995, vol. 62: Proceedings of the Symposium on Pure Mathematics,
pp. 45–96. American Mathematical Society, Providence, RI, 1997
5. T. Graber, J. Harris, J. Starr. Families of rationally connected varieties. J. Am. Math. Soc.,
16(1):57–67 (electronic), 2003
6. A. Grothendieck. Techniques de construction et théorèmes d’existence en géométrie
algébrique. IV. Les schémas de Hilbert. In Séminaire Bourbaki, vol. 6, pages Exp. No. 221,
249–276. Socit Mathmatique de France, Paris, 1995
7. D. Huybrechts, M. Lehn. The geometry of moduli spaces of sheaves. Aspects of Mathematics,
E31. Friedrick Vieweg & Sohn, Braunschweig, 1997
8. J. Kollár. Rational curves on algebraic varieties, Ergebnisse der Mathematik und ihrer Gren-
zgebiete 3, vol. 32. Springer, Berlin, 1996
9. J. Kollár, Y. Miyaoka, S. Mori. Rational curves on Fano varieties. In Classification of irregu-
lar varieties (Trento, 1990), vol. 1515: Lecture Notes in Mathematics, pp. 100–105. Springer,
Berlin, 1992
10. J. Kollár, S. Mori. Birational geometry of algebraic varieties, vol. 134: Cambridge Tracts
in Mathematics. Cambridge University Press, Cambridge, 1998 (translated from the 1998
Japanese original)
11. S. Mori. Projective manifolds with ample tangent bundles. Ann. Math. (2), 110(3):593–606,
1979
12. D. Mumford, J. Fogarty, F. Kirwan. Geometric invariant theory. Ergebnisse der Mathematik
und ihrer Grenzgebiete (2), 3rd ed., vol. 34. Springer, Berlin, 1994
13. M. Schlessinger. Functors of Artin rings. Trans. Am. Math. Soc., 130:208–222, 1968
14. J. Starr. Rational points of rationally simply connected varieties, 2009 (preprint)
15. J. Starr, A.J. de Jong, X. He. Families of rationally simply connected varieties over surfaces
and torsors for semisimple groups, 2009. arXiv:0809.5224
Recent Progress on the Kato Conjecture
Shuji Saito
Summary This is a survey paper on recent works in progress by Jannsen, Kerz, and
the author on the Kato conjecture on the cohomological Hasse principle.
1 Statements of the Kato Conjectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
2 Known Results and Announcement of New Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3 Outline of Proof of Theorem 2.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
This is a survey paper on recent works [17, 19] and a work in progress [19]
by Jannsen, Kerz, and the author on the Kato conjecture on the cohomological
Hasse principle. In [17, 19], general approaches are proposed to solve the conjec-
ture for schemes over a finite field assuming resolution of singularities. Based on
the idea in [19], a new approach is proposed in [24] to solve the conjecture for
schemes over a finite field or the ring of integers in a local field, restricted to the
prime-to-characteristic part. A key ingredient in [24], which replaces resolution of
singularities, is a recently announced result on refined alterations due to Gabber
(see [16]). We will give an outline of the proof. As an application, it implies a finite-
ness result on higher Chow groups of arithmetic schemes using the Bloch–Kato
conjecture whose proof has been announced by Rost and Voevodsky ([31, 35], see
also [14, 36–38]).
We start with a review on the following fundamental fact in number theory. Let k
be a global field, namely either a finite extension of Q or a function field in one
Shuji Saito
Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1 Komaba,
Tokyo 153-8914, Japan
e-mail: [email protected]
2010 Mathematics subject classification. 14F20, 14F42, 14G17, 14G26, 11R34.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 109
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 6,
c Springer Science+Business Media, LLC 2010
110 S. Saito
variable over a finite field. For simplicity, we assume that char(k) > 0 or k is totally
imaginary. Let P be the set of all finite places of k, and denote by kv the completion
of k at v ∈ P. For a field L, let Br(L) be its Brauer group, and identify the Galois
cohomology group H 1 (L, Q/Z) with the group of continuous characters on the ab-
solute Galois group of L with values in Q/Z:
1. For v ∈ P, there is a natural isomorphism
Br(kv ) −−−−→ H 1 (Fv , Q/Z) −−−−→ Q/Z,
αv βv
where Fv be the residue field of v and αv is the residue map and βv is the evalua-
tion of characters at the Frobenius element.
2. There is an exact sequence
0 → Br(k) → H 1 (Fv , Q/Z) → Q/Z → 0,
v∈P
where the first map is the composite of the restriction maps and αv and the second
map is the sum of βv . The injectivity of the first map is the so-called Hasse prin-
ciple for central simples algebras of k, which is a celebrated theorem of Hasse–
Brauer–Noether.
Kato [21] proposed a fascinating framework of conjectures that generalizes the
above facts to higher-dimensional arithmetic schemes. To review these conjectures,
we introduce some notation. Let L be a field with p = char(L). Let n be an integer
n > 0 and write n = mpr with (p, m) = 1. We define the following Galois cohomol-
ogy groups:
where μm is the Galois module of nth roots of unity and Wr ΩL,logi is the logarithmic
part of the de Rham-Witt sheaf Wr ΩL [15, I, (5.7)]. Note that there is a canonical
i
identification H 2 (L, Z/nZ(1)) = Br(L)[n], where [n] denotes the n-torsion part.
Now, let X be a scheme of finite type over F p or the integer ring of a number
field or a (p-adic) local field. Kato introduced the following complex KC• (X , Z/nZ)
which we call the Kato complex:
··· → H a+1 (x, Z/nZ(a)) → H a (x, Z/nZ(a − 1)) → · · ·
x∈X(a) x∈X(a−1)
··· → H 2 (x, Z/nZ(1)) → H 1 (x, Z/nZ), (2)
x∈X(1) x∈X(0)
where X(a) = {x ∈ X | dim {x} = a} and the term ⊕x∈X(a) is put in degree a. We will
also use the complexes:
KC• (X, Q/Z) = lim KC• (X , Z/nZ) and KC• (X , Q /Z ) = lim KC• (X , Z/n Z),
−→ −→
n n
Recent Progress on the Kato Conjecture 111
are called the Kato homology of X with coefficient Λ . (It is indeed a homology
theory in the sense of Definition 3.1.)
Now let X be a projective smooth connected curve over a finite field Fq with
the function field k = Fq (X ), or X = Spec(Ok ) for the integer ring Ok of a number
field or a local field. Then the Kato complex KC• (X , Q/Z) is identified with the
following complex: 1
Br(k) −−−−→ H (x, Q/Z).
x∈X(0)
Hence the above facts (1) and (2) are equivalent to the following:
0 if k is local,
KH1 (X , Q/Z) = 0 and KH0 (X , Q/Z) =
Q/Z if k is global.
Then
∼
= 0 if a = 0,
KHa (X , Z/nZ) −→
Z/nZ if a = 0.
We note that the assumption (∗) may be removed by modifying KHa (X , Q/Z)
(see Conjecture C on page 482 of Jannsen and Saito [18]).
Conjecture 1.3. Let X be a regular scheme proper and flat over Spec(Ok ) where Ok
is the integer ring of a local field k. Then
As was already noticed, the Kato conjectures in case dim(X ) = 1 rephrase the clas-
sical fundamental facts on the Brauer group of a global field and a local field.
112 S. Saito
Theorem 2.1 (Saito [30]). Let X be a smooth projective 3-fold over a finite field F.
Then KH3 (X, Q /Z ) = 0 for any prime = char(F).
Theorem 2.2 (Colliot-Thélène [7], Suwa [33]). Let X be a smooth projective vari-
ety over a finite field F. Then
Theorem 2.3 (Jannsen–Saito [18]). Let X be a regular projective flat scheme over
S = Spec(Ok ) where k is a number field or a p-adic field. Assume that k is totally
imaginary if k is a number field. Fix a prime p. Assume that for any closed point
v ∈ S, the reduced part of Xv = X ×S v is a simple normal crossing divisor on X and
that Xv is reduced if v|p. Then we have
(RS)d For any X integral and proper of dimension ≤ d over F, there exists a
proper birational morphism π : X → X such that X is smooth over F. For any U
smooth of dimension ≤ d over F, there is an open immersion U → X such that
X is projective smooth over F with X −U a simple normal crossing divisor on X .
(RES)t For any smooth projective variety X over F, any simple normal crossing
divisor Y on X with U = X − Y , and any integral closed subscheme W ⊂ X of
dimension ≤ t such that W ∩ U is regular, there exists a projective smooth X
Recent Progress on the Kato Conjecture 113
over F and a birational proper map π : X → X such that π −1 (U) U, and that
Y = X − π −1(U) is a simple normal crossing divisor on X , and that the proper
transform of W in X is regular and intersects transversally with Y .
We note that a proof of (RES)2 is given in [9] based on an idea of Hironaka,
which enables us to obtain the unconditional vanishing of the Kato homology with
Q/Z-coefficients in degree a ≤ 4.
Finally, the above approach has been improved to remove the assumptions (RS)d
and (RES)t on resolution of singularities, at least if we are restricted to the prime-
to-char(F) part.
Theorem 2.5 (Kerz–Saito [24]). Let X be a projective smooth variety over a finite
field F. For a prime = char(F), we have KHa (X , Q /Z ) = 0 for a > 0.
A key to the proof is the following refinement of de Jong’s alteration theorem
due to Gabber (see [16]).
Theorem 2.6 (Gabber). Let F be a perfect field and X be a variety over F. Let
W ⊂ X be a proper closed subscheme. Let be a prime different from char(F). Then
there exists a projective morphism π : X → X such that:
• X is smooth over F and the reduced part of π −1 (W ) is a simple normal crossing
divisor on X.
• π is generically finite of degree prime to .
The same technique proves the following arithmetic version as well.
Theorem 2.7 (Kerz–Saito [24]). Let X be a regular projective flat scheme over a
henselian discrete valuation ring with finite residue field F. Then, for a prime =
char(F), we have KHa (X , Q /Z ) = 0 for a ≥ 0.
Finally we remark that one can prove the above results with Z/n Z-coefficients
instead of Q /Z -coefficients by using the Bloch–Kato conjecture: For a prime
and a field L, we have the symbol map
where KtM (L) denotes the Milnor K-group of L. It is conjectured that htL, is surjec-
tive. The conjecture is called the Bloch–Kato conjecture in case l = char(L). For a
scheme X, we introduce the following condition:
(BK)tX, For every field L finitely generated over a residue field of X , htL, is sur-
jective.
The surjectivity of htL, is known if t = 1 (Kummer theory) or t = 2 (Merkurjev–
Suslin [26]) or = char(L) (Bloch–Gabber–Kato [5]) or = 2 (Voevodsky [34]). It
is conjectured (the Bloch–Kato conjecture) that htL, is always an isomorphism for
every field L. Recently, a complete proof of the conjecture has been announced by
Rost and Voevodsky ([31, 35], see also [14, 36–38]).
Theorem 2.8 (see [19, Sect. 5]). Let X and be as in either of Theorem 2.5 or
of Theorem 2.7. Assume (BK)tX, holds. Then we have KHa (X , Z/n Z) = 0 for
0 < a ≤ t.
114 S. Saito
In this section, we give an outline of the proof of Theorem 2.5. We fix a finite field
F with p = char(F) and work in the category C of schemes separated of finite type
over F. We first recall the following.
Definition 3.1. Let C∗ be the category with the same objects as C , but morphisms
are just the proper maps in C . Let Mod be the category of modules. A homology
theory H = {Ha }a∈Z on C is a sequence of covariant functors:
Ha (−) : C∗ → Mod
∂ i
∗ j∗ ∂
· · · −→ Ha (Y ) −→ Ha (X ) −→ Ha (V ) −→ Ha−1 (Y ) −→ · · · .
(The maps ∂ are called the connecting morphisms.) This sequence is functorial
with respect to proper maps or open immersions, in an obvious way.
It is an easy exercise to check that the Kato homology (3)
Here, the limit is over all open nonempty subschemes V ⊆ {x}. This spectral se-
quence is covariant with respect to proper morphisms in C and contravariant with
respect to open immersions. The functoriality of the spectral sequence is a direct
consequence of that of the homology theory H.
In what follows, we assume Λ = Z/nZ with n prime to p or Λ = Q /Z with
= p.
Here R f ! is the right adjoint of R f! defined in [2, XVIII, (3.1.4)]. This is a homology
theory in the sense of Definition 3.1. For X smooth of pure dimension d over F, we
have (cf. [6] and [18, Theorem 2.14])
where, for an integer r > 0, Λ (r) is the Tate twist by the étale sheaf of roots of unity.
We then look at the spectral sequence (4) arising from this homology theory. The
first step of the proof is the following lemma.
isomorphism of complexes
KC• (X , Λ ) E•,−1
1
(X ),
d1 d1 d1 d1
· · · → Ea,−1
1
(X ) −→ Ea−1,−1
1
(X ) −→ · · · −→ E1,−1
1
(X ) −→ E0,−1
1
(X ).
KHa (X , Λ ) Ea,−1
2
(X ).
The first assertion follows easily from the fact cd(F) = 1 and the second
from [20, Theorem 1.1.1].
By the above lemma, we get the edge homomorphism
where the second isomorphism is due to (5) and the third is the trace map and the
last is the natural isomorphism in (1).
Let Φ = (X,Y ) be a log pair and let Y1 , . . . ,YN be the irreducible components of
Y . For an integer a ≥ 1 write
Y (a) = Yi1 ,...,ia (Yi1 ,...,ia = Yi1 ∩ · · · ∩Yia ). (8)
1≤ii <···<ia ≤N
which we call the graphication of the Kato homology. To control γΦa , we use
which is defined as the composite of γΦa with εUa (cf. (6)). We note that the right-hand
side of (12) is nonzero only if 0 ≤ a ≤ d = dim(X ) while the left-hand side could
be nonzero for any a with 0 ≤ a ≤ 2d + 1.
Now the following theorem [19, Lemmas 3.3 and 3.4] is crucial.
Theorem 3.6. Take Λ = Q /Z . Let Φ = (X ,Y ;U) be an ample log pair, which
means by definition that one of the irreducible components of Y is an ample divisor
on X. Then Φ is clean in degree q for all q ≤ dim(X ).
The proof of the above theorem hinges on the affine Lefschetz theorem and the
Weil conjecture proved by Deligne [10].
The above theorem implies that for an ample log pair Φ and for any integer a
with 0 ≤ a ≤ dim(X ), γΦa is surjective, and an isomorphism if εUa is surjective. This
is already a big step in the proof of Theorem 2.5.
We will need a variant of the above construction.
Definition 3.7. The reduced graph complex of a log pair Φ = (X ,Y ) is the complex:
(a) )
where Λ π0 (Y is put in degree a. We have evident maps of complexes
ι
Λ π0 (X) [0] → G• (Φ , Λ ) −→ G• (Φ̂ , Λ )
ι
0 → GH1 (Φ , Λ )) −→ GH1 (Φ̂ , Λ ) → Λ → 0. (14)
In the same way as above, one may also define the natural homomorphism [24]
∂
KHa (U, Λ ) −−−−→ KHa−1 (Y, Λ )
⏐ ⏐
⏐γ a ⏐γ a
Φ Φ̂
ι
GHa (Φ , Λ ) −−−−→ GHa (Φ̂ , Λ )
Another key ingredient to the proof of Theorem 2.5 is the construction of the pull-
back map for Kato homology, which is stated in the following form.
118 S. Saito
εq
Hq−1 (Y, Λ ) −−−Y−→ KHq (Y, Λ )
⏐ ⏐
⏐ ∗ ⏐ ∗
f f
εq
Hq−1 (X , Λ ) −−−X−→ KHq (X , Λ )
is the multiplication by the degree [F(X) : F(Y )] of the extension of the function
fields.
In [24] the above lemma is shown based on the intersection theory on cycle mod-
ules due to Rost [28]. (The theory was originally developed over a field but it may
be extended to a more general base.)
Let q ≥ 1 be an integer. For a log pair Φ = (X,Y ;U) consider the condition:
(LG)q The composite map
εa ∂
∂ εΦa : Ha−1 (U, Λ ) −→
U
KHa (U, Λ ) −→ KHa−1 (Y, Λ )
Lemma 3.9. Let q ≥ 1 be an integer. Let Φ = (X ,Y ;U) be a log pair that satisfies the
condition (LG)q . Let j∗ : KHq (X , Λ ) → KHq (U, Λ ) be the pullback via j : U → X
and εUq : Hq−1 (U, Λ ) → KHq (U, Λ ) be as in (6). Then the map j∗ is injective and
Image( j∗ ) ∩ Image(εU ) = 0.
q
Recent Progress on the Kato Conjecture 119
Proof. First we claim that j∗ is injective. Indeed we have the exact sequence
∂ j∗
KHq+1 (U, Λ ) −→ KHq (Y, Λ ) → KHq (X , Λ ) −→ KHq (U).
Since ∂ εUq+1 is surjective by the assumption, ∂ is surjective and the claim follows.
By the above claim it suffices to show Image( j∗ ) ∩ Image(εUq ) = 0. We have the
exact sequence
j∗ ∂
KHq (X, Λ ) −−−−→ KHq (U, Λ ) −−−−→ KHq−1 (Y, Λ ).
Let β ∈ Hq−1 (U, Λ ) and assume α = εU (β ) ∈ KHq (U, Λ ) lies in Image( j∗ ). It im-
q
Lemma 3.10. Take Λ = Q /Z . Fix integers d and q with d ≥ q ≥ 1, and assume
KC(q, d − 1). For any log pair Φ = (X ,Y ) with d = dim(X ), there exists a log pair
Φ = (X,Y ) with Y ⊂ Y such that Φ satisfies the condition (LG)q .
Proof. It follows from Bertini’s theorem [1, 27] that for any log pair Φ = (X ,Y )
with dim(X) = d, one can take Z ⊂ X , a smooth section of a sufficiently ample
line bundle on X, such that Φ = (X,Y ∪ Z) is an ample log pair so that it is clean
in degree q for all q ≤ dim(X) by Theorem 3.6. Hence, it suffices to show that if
Φ = (X,Y ;U) is clean in degree q, then it satisfies the condition (LG)q . We consider
the commutative diagram
q+1
ε ∂
Hq (U, Λ ) −−−
U
−→ KHq+1 (U, Λ ) −−−−→ KHq (Y, Λ )
⏐ ⏐
⏐ q+1 ⏐γ q+1
γΦ Φ̂
GHq+1 (Φ , Λ ) −−−−→ GHq+1 (Φ̂ , Λ )
ι
where ι is an isomorphism by (13), and γεΦq+1 = γΦq+1 ◦ εUq+1 is surjective by the as-
sumption. Moreover, γΦ̂q+1 is an isomorphism. Indeed we have the spectral sequence
(Mayer–Vietoris for homology of closed coverings):
1
Es,t = KHt (Y (s) , Λ ) ⇒ KHs+t−1 (Y, Λ ) (cf. (8))
q
ε ∂
Hq−1 (U, Λ ) −−−U−→ KHq (U, Λ ) −−−−→ KHq−1 (Y, Λ )
⏐ ⏐
⏐γ q ⏐γ q−1
Φ Φ̂
→
GHq (Φ , Λ ) −−−−→ GHq (Φ̂ , Λ )
ι
where ι is injective by (13) and (14), and γεΦq = γΦq ◦ εUq is injective by the assump-
tion. As before one can show by using KC(q, d − 1) that γΦ̂ is an isomorphism.
q−1
We fix a prime = char(F) and take Λ = Q /Z . We finish the proof of Theorem 2.5
by induction on d = dim(X) ≥ 0. The case d = 0 is trivial. Assume d ≥ 1 and
that KC(q, d − 1) holds for 1 ≤ q ≤ d. Let X ∈ Ob(S ) with d = dim(X ). Let α ∈
KHq (X, Λ ). By recalling that
⏐ε q ⏐ε q ⏐ε q
⏐U ⏐ U ⏐V
π∗ j∗
Hq−1 (U, Λ ) −−−−→ Hq−1 (U , Λ ) −−−−→ Hq−1 (V, Λ )
4 Applications
ρX,Z/nZ
r,q 2r−q
: CHr (X , q; Z/nZ) → Hét (X , Z/nZ(r)). (16)
Here,
CHr (X , q; Z/nZ) = Hq (zr (X , •) ⊗L Z/nZ)
is the higher Chow group with finite coefficients which fits into an exact sequence
where n = mpr and (p, m) = 1 with p = char(F) (cf. (1)). The same construction
has been carried out for a regular scheme X of finite type over a Dedekind domain
by Levine [25] (see also [11]), assuming that n is invertible on X .
We recall the following result due to Suslin–Voevodsky [32] and Geisser-
Levine [12].
Theorem 4.1. Let the assumption be as above (the case over a Dedekind domain
is included). Assume (BK)tX, for all t ≥ 0 (cf. Sect. 1). Then ρX,Z/n Z is an isomor-
r,q
is a regular scheme over either a finite field F or a henselian discrete valuation ring
with finite residue field F. In case r > d it is easily shown (see [19, Lemma 6.2])
that ρX,Z/n Z is an isomorphism assuming (BK)X, . An interesting phenomenon
r,q q+1
d,q−1
ρX,Z/ nZ
→ KHq+1 (X , Z/n Z) → CHd (X , q − 1; Z/nZ) −−−−−→ · · ·
Theorem 4.2 (Kerz and Saito [24]). Let X be a regular projective scheme over
either a finite field F or a henselian discrete valuation ring with finite residue field
F. Let q ≥ 0 be an integer and n > 0 be an integer prime to char(F) and assume
q+2
(BK)X, for all primes dividing n. Let d = dim(X ). Then
∼
=
ρX,Z/nZ
d,q 2d−q
: CHd (X , q; Z/nZ) −→ Hét (X , Z/nZ(d)).
The above theorem implies the following affirmative result on the finiteness con-
jecture on motivic cohomology.
Proof. For simplicity, we only treat the case over a finite field F. We may assume
n = m for a prime = char(F). We proceed by the induction on dim(X ). First we
remark that the localization sequence for higher Chow groups implies that for a
dense open subscheme U ⊂ X, the finiteness of CHr (X , q; Z/nZ) for all r ≥ dim(X )
and q is equivalent to that of CHr (U, q; Z/nZ). Thus it suffices to show the assertion
for any smooth variety U over F. If U is an open subscheme of a smooth projective
variety X over F, the assertion holds for X by Theorem 4.2 and hence for U by
the above remark. In general Gabbers’s theorem 2.6 implies that there exist an open
subscheme V of a smooth projective variety X over F, an open subscheme W of
U, and a finite étale morphism π : V → W of degree prime to . We know that the
assertion holds for V so that it holds for W by a standard norm argument. This
completes the proof by the above remark.
Finally, we note that the above corollary implies the following affirmative result
on the Bass conjecture. Let Ki (X , Z/nZ) be Quillen’s higher K-groups with finite
Recent Progress on the Kato Conjecture 123
Corollary 4.4. Under the assumption of Corollary 4.3, Ki (X , Z/nZ) is finite for
i ≥ dim(X) − 2.
Proof. Theorem 4.1 implies that CHr (X , q; Z/nZ) is finite for r ≤ q + 1. Hence the
assertion follows from the Atiyah–Hirzebruch spectral sequence (see [25] for its
construction in the most general case):
References
1. A. Altman and S. Kleiman, Bertini theorems for hypersurface sections containing a sub-
scheme, Commun. Algebra 7 (1979), 775–790
2. M. Artin, A. Grothendieck and J.-L. Verdier, eds., Séminaire de Géomt́rie Algébrique du Bois
Marie (1963–64) – SGA4, Lecture Notes in Mathematics, vols. 269, 270, and 305, Springer,
Berlin, 1972
3. S. Bloch, Higher algebraic K-theory and class field theory for arithmetic surfaces, Ann. Math.
114 (1981), 229–265
4. , Algebraic cycles and higher algebraic K-theory, Adv. Math. 61 (1986), 267–304
5. S. Bloch and K. Kato, p-adic étale cohomology, Publ. Math. IHES 63 (1986), 107–152
6. S. Bloch and A. Ogus, Gersten’s conjecture and the homology of schemes, Ann. Sci. Ecole.
Norm. Sup. 4 série 7 (1974), 181–202
7. J.-L. Colliot-Thélène, On the reciprocity sequence in the higher class field theory of function
fields, in: Algebraic K-Theory and Algebraic Topology (Lake Louise, AB, 1991), (J.F. Jardine
and V.P. Snaith, eds), Kluwer, Dordrecht, 1993, pp. 35–55
8. J.-L. Colliot-Thélène, J.-J. Sansuc and C. Soulé, Torsion dans le groupe de Chow de codimen-
sion deux, Duke Math. J. 50 (1983), 763–801
9. V. Cossart, U. Jannsen and S. Saito, Canonical embedded and non-embedded resolution of
singularities for excellent two-dimensional schemes, arxiv:0905.2191 (2009)
10. P. Deligne, La conjecture de Weil II, Publ. Math. IHES 52 (1981), 313–428
11. T. Geisser, Motivic cohomology over Dedekind rings, Math. Z. 248 (2004), 773–794
12. T. Geisser and M. Levine, The Bloch–Kato conjecture and a theorem of Suslin–Voevodsky,
J. Reine Angew. 530 (2001), 55–103
13. M. Gros, Sur la partie p-primaire du groupe de Chow de codimension deux, Commun. Algebra
13 (1985) 2407–2420
14. C. Haesemeyer and C. Weibel, Norm varieties and the chain lemma (after Markus Rost), in:
Algebraic Topology, Abel Symposia, vol. 4, Springer, Berlin, 2009, pp. 95–130
15. L. Illusie, Complexe de de Rham-Witt et cohomologie cristalline, Ann. Sci. Ecole. Norm. Sup.
4 série 12 (1979), 501–661
16. , On Gabber’s refined uniformization, preprint available at http://www.math.u-psud.
fr/illusie/ (2008)
17. U. Jannsen, Hasse principles for higher-dimensional fields, arxiv:0910.2803 (2009)
18. U. Jannsen and S. Saito, Kato homology of arithmetic schemes and higher class field theory,
Doc. Math. Extra Volume: Kazuya Kato’s Fiftieth Birthday (2003), 479–538
124 S. Saito
19. , Kato conjecture and motivic cohomology over finite fields, arxiv:0910.2815 (2009)
20. U. Jannsen, S. Saito and K. Sato, Étale duality for constructible sheaves on arithmetic
schemes, arxiv:0910.3759 (2009)
21. K. Kato, A Hasse principle for two dimensional global fields, J. Reine Angew. Math. 366
(1986), 142–183
22. K. Kato and S. Saito, Unramified class field theory of arithmetic surfaces, Ann. Math. 118
(1985), 241–275
23. , Global class field theory of arithmetic schemes, Am. J. Math. 108 (1986), 297–360
24. M. Kerz and S. Saito, Kato conjecture and motivic cohomology for arithmetic schemes (in
preparation)
25. M. Levine, K-theory and motivic cohomology of schemes, preprint (1999)
26. A.S. Merkurjev and A.A. Suslin, K-cohomology of Severi-Brauer varieties and the norm
residue homomorphism, Math. USSR Izvestiya 21 (1983), 307–340
27. B. Poonen, Bertini theorems over finite fields, Ann. Math. 160 (2004), 1099–1127
28. M. Rost, Chow groups with coefficients, Doc. Math. J. 1 (1996), 319–393
29. S. Saito, Class field theory for curves over local fields, J. Number Theory 21 (1985), 44–80
30. , Cohomological Hasse principle for a threefold over a finite field, in: Algebraic
K-Theory and Algebraic Topology, NATO ASI Series, vol. 407, Kluwer, Dordrecht, 1994,
pp. 229–241
31. A. Suslin and S. Joukhovitski, Norm varieties, J. Pure Appl. Algebra 206 (2006), 245–276
32. A. Suslin and V. Voevodsky, Bloch–Kato conjecture and motivic cohomology with finite coeffi-
cients, in: Cycles, Transfer, and Motivic Homology Theories, Annals of Mathematics Studies,
vol. 143, Princeton University Press, Princeton, 2000
33. N. Suwa, A note on Gersten’s conjecture for logarithmic Hodge-Witt sheaves, K-Theory 9
(1995), 245–271
34. V. Voevodsky, Motivic cohomology with Z/2-coefficients, Publ. IHES 98 (2003), 1–57
35. , On motivic cohomology with Z/l-coefficients, K-Theory Preprint Archives, http://
www.math.uiuc.edu/K-theory/0639/ (2003)
36. , Motivic Eilenberg–MacLane spaces, K-Theory Preprint Archives, http://www.math.
uiuc.edu/K-theory/0864/ (2007)
37. C. Weibel, Axioms for the norm residue isomorphism, in: K-Theory and Noncommutative
Geometry, European Mathematical Society, Zürich, 2008, pp. 427–435
38. , The norm residue isomorphism theorem, J. Topol. 2 (2009), 346–372
Elliptic Curves and Iwasawa’s μ = 0 Conjecture
R. Sujatha
1 Introduction
R. Sujatha
School of Mathematics, Tata Institute of Fundamental Research, Homi Bhabha Road,
Mumbai 400005, India
e-mail: [email protected]
2010 Mathematics subject classification. 11R23, 11G05.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 125
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 7,
c Springer Science+Business Media, LLC 2010
126 R. Sujatha
explain how Iwasawa’s μ = 0 conjecture proves the freeness of certain pro-p Galois
groups of infinite p-adic Lie extensions of Q. We then proceed to describe the ana-
logue of this conjecture for the motive of an elliptic curve, whose formulation was
carried out in a joint work with J. Coates a few years ago (“Conjecture A” in [3]).
Finally, we discuss some recent results that prove Conjecture A in special cases.
Throughout, p will denote an odd prime and Z p (resp., Q p ), will denote the ring
of p-adic integers (resp., the field of p-adic numbers). Given any number field F, we
shall denote by F cyc the cyclotomic Z p -extension of F. Recall that an extension L
of F is said to be a p-adic Lie extension if the Galois group Gal(L /Q) is a p-adic
Lie group [12]. The cyclotomic extension F cyc is a basic example of a p-adic Lie
extension of F, and has been studied extensively for over a century, culminating in
Iwasawa’s seminal works [8,9]. For any finite set S of primes in a number field F, FS
denotes the maximal extension of F unramified outside S. We shall always assume
that S contains the set S p of primes in F that lie above p, and the archimedean
primes. Given a field L, the p-cohomological dimension of the field will be denoted
by cd p (L). If G is a profinite group, the Iwasawa algebra of G is denoted by Λ (G),
and we recall that it is defined as
where U runs over all open normal subgroups of G, with the inverse limit being
taken with respect to the natural surjections. For any number field F, we shall de-
note the Galois group Gal(F cyc /F) of the cyclotomic extension by Γ . Given a dis-
crete module M over the Iwasawa algebra Λ (G), we denote its compact Pontryagin
dual by M ∨ . For any module M over the Iwasawa algebra Λ (G), M(p) denotes the
p-primary torsion submodule of M.
2 Iwasawa’s μ = 0 Conjecture
where the pi s are prime ideals of height 1. Further, the set of prime ideals {pi } and
the set of integers {ni } are unique up to a bijection of the indexing set.
The integer r is called the rank of M. The importance of this theorem lies in
the fact that it enables us to define two key invariants for finitely generated torsion
Λ (G)-modules. Let M be a finitely generated torsion Λ (G)-module so that it has
rank zero. Write pi j (with 1 ≤ j ≤ k) for the set of prime ideals occurring in the
structure theorem such that pi j = p. Then the μ -invariant of M is defined as
μ (M) = ∑ ni j ,
j
is a principal ideal. Any generator of chG (M) is called a characteristic power series
of M, which can be assumed to be a distinguished polynomial in Z p [[T ]], thanks to
Weierstrass’ preparation theorem. The degree of the characteristic polynomial of M
is clearly the λ -invariant. If M is an arbitrary finitely generated Λ (G)-module, then
we set μ (M) := μ (Mtors ) where Mtors is the Λ (G)-torsion submodule of M.
Let F be a number field and F cyc be the cyclotomic Z p -extension of F, with
Γ = Gal(F cyc /F) Z p . We denote by F∞ the maximal abelian p-extension of F cyc
which is unramified everywhere. In other words, F∞ is the p-Hilbert class field of
F cyc . Let X∞ denote the Galois group Gal(F∞ /F cyc ). As X∞ is abelian, it has a natural
structure of a Γ -module. Indeed, given an element γ in Γ and an element x in X∞ ,
the action is defined by
γ .x = γ̃ x γ̃ −1 ,
where γ̃ is any lift of γ to Gal(F∞ /F). It is easily checked that the action is inde-
pendent of the lift since X∞ is abelian. Further, it is well known that any Z p -module
with a continuous Γ -action has a natural structure as a compact Z p [[Γ ]]-module and
thus X∞ is a compact module over the Iwasawa algebra. The following theorem is
due to Iwasawa [9].
class group of Q is zero, we thus have that the group (X∞ )Γ of Γ -coinvariants is
trivial, and hence X∞ is itself zero. Sinnott [16] later gave a different proof of the
Ferrero-Washington theorem. It should be noted that Iwasawa gave examples of
noncyclotomic Z p -extensions of Q whose p-Hilbert class fields have positive μ -
invariant [10].
We shall next consider other infinite extensions of F cyc . Given a finite set S of
primes in F that contains S p and the archimedean primes, we recall that FS denotes
the maximal extension of F that is unramified outside S. As the primes in S p are the
only nonarchimedean primes that ramify in F cyc , we have FS ⊇ F cyc .
Note that the profinite degree of FS /FS (p) is not necessarily prime to p. Also,
FS (p) ⊇ F cyc and FS (p) has no Galois p-extensions of F in FS , that is, the group
H 1 (Gal(FS /FS (p)), Z/p) is zero. For a group H, let H ab denote the abelianization
of H.
Definition 2.4. FS (p)ab is the Galois extension of F cyc such that Gal(FS (p)ab /F cyc )
equals (Gal(FS (p)/F cyc ))ab . We denote its Galois group by XS .
Recall that a free pro-p group G over a set T is a pro-p group G, along with a
map i : T → G satisfying the following properties: (1) every open subgroup of G
contains all but a finite number of elements of i(T ), and (2) if j : T → G̃ is any
other map with the property (1) into a pro-p group G̃, then there exists a unique
homomorphism f : G → G̃ such that j = f ◦ i.
For any group G, let cd p (G) denote the p-cohomological dimension of G. Recall
that cd p (G) = n if the cohomology groups H k (G, M) are trivial for all k ≥ n + 1, and
all p-primary torsion modules M. In particular, if G is a pro-p group, then cd p (G) =
n if and only if H k (G, M) = 0 for all k > n, and M any discrete p-primary torsion
G-module. The following theorem (see [15]) characterizes free pro-p groups.
We return to the number field F. The Weak Leopoldt Conjecture for F cyc is the
assertion
H 2 (Gal(FS /F cyc ), Q p /Z p ) = 0. (3.2)
Elliptic Curves and Iwasawa’s μ = 0 Conjecture 129
Theorem 3.3. The Galois group GS,p (F cyc ) := Gal(FS (p)/F cyc ) is a free pro-p
group if and only if μ (XS ) = 0 (equivalently, if μ (X∞ ) = 0).
0 → Z/p → Q p /Z p → Q p /Z p → 0.
Suppose that GS,p (F cyc ) is a free pro-p group. Then by Theorem 3.1, we have
cd p (GS,p (F cyc )) is 1 and hence H 2 (GS,p (F cyc ), Z/p) = 0. Taking Pontryagin du-
als of the long exact sequence, along with the vanishing result of the Weak Leopoldt
Conjecture (3.2), we see that
But this implies that GS,p (F cyc )ab has no nontrivial p-torsion, and hence clearly
μ (XS ) = 0. For the converse, we need to use one additional fact from Iwasawa
theory, namely that the Pontryagin dual (H 1 (GS,p (F cyc ), Q p /Z p ))∨ has no fi-
nite nonzero Λ (Γ )-submodule [9]. Suppose now that μ (XS ) = 0. By the structure
theorem for finitely generated Λ (Γ )-modules, it follows that the p-torsion of the
Pontryagin dual is finite, and further, by the above remark, it is in fact trivial.
We therefore have H 2 (GS,p (F cyc ), Z/p) = 0. But the cohomological dimension of
GS,p (F cyc ) is at most 2, as it is a closed subgroup of Gal(FS /F) which has cohomo-
logical dimension 2 [15]. The conclusion now follows from Theorem 3.1.
The next simplest motive after the trivial Tate motive is the motive associated with
an elliptic curve. Guided by the philosophy that results for the Tate motive have
elliptic curve analogues, one seeks the corresponding results for elliptic curves of
the material in the previous section. Let E be an elliptic curve over the number field
F and as before, let p be an odd prime. We take S to be the finite set consisting
of primes S p in F that lie above p, the archimedean primes, and the primes of bad
reduction for the elliptic curve. Put GF = Gal(F/F) for the absolute Galois group
of F. Let
E p∞ := E pn ,
n≥0
130 R. Sujatha
Tp (E) = lim E pn ,
←−
and Vp (E) = Tp (E) ⊗ Q p , is the corresponding two-dimensional Q p -vector space
with a continuous action of GF . For the purposes of this chapter, we shall choose
to be simple minded and consider the GF -module Vp(E) as the “dual of the Tate
motive of the elliptic curve E.” We first discuss the analogous Λ (Γ )-modules in this
context.
As before, let F∞ denote the p-Hilbert class field of F cyc and X∞ the Galois group
Gal(F∞ /F cyc ). At a first glance, classical Iwasawa theory for elliptic curves suggests
that the corresponding Λ (Γ )-module in this context is the Pontryagin dual of the
Selmer group over the cyclotomic extension. Recall that for a finite Galois extension
L of F, the Selmer group S(E/L) of E over L is a discrete Gal(L/F)-module and is
defined as
S(E/L) = Ker H 1 (Gal(FS /L), E p∞ )) → ⊕v∈S Jv (E/L) , (4.1)
where
Jv (E/L) = ⊕w|v H 1 (Gal(L̄w /Lw ), E)(p).
The Selmer group of E over F cyc is defined as the direct limit of S(E/L) as L varies
over finite extensions of F in F cyc . Its Pontryagin dual is denoted by X(E/F cyc )
and is well known to be a finitely generated module over Λ (Γ ). It is a deep con-
jecture due to Mazur that if E has good ordinary reduction at the primes above p,
X(E/F cyc ) is a torsion Λ (Γ )-module and there are plenty of numerical examples
known where this conjecture holds. With this knowledge, one is naturally led to
wonder if X(E/F cyc ) is a finitely generated Z p -module, which would be the ana-
logue of Iwasawa’s μ = 0 conjecture. However, Mazur already gave examples where
this is not true, and the fact that the μ -invariant of X(E/F cyc ) is not zero causes end-
less technical difficulties in Iwasawa theory! An explicit example is the elliptic curve
E/Q of conductor 11, defined by
E : y2 + y = x3 − x2 − 10x − 20.
Definition 4.2. The fine Selmer group for E over a finite extension L of F, denoted
R(E/L) is defined by
R(E/L) = Ker H 1 (Gal(FS /L), E p∞ ) → ⊕v∈S Kv1 (L) ,
where
Kv1 (E/L) = ⊕w|v H 1 (Gal(L̄w /Lw ), E p∞ ).
Elliptic Curves and Iwasawa’s μ = 0 Conjecture 131
The fine Selmer group of E over F cyc , denoted R(E/F cyc ) is defined to be the direct
limit of R(E/L) as L varies over finite extensions of F in F cyc .
The fine Selmer group R(E/F cyc ) is a discrete finitely generated module over the
Iwasawa algebra Λ (Γ ), and its compact Pontryagin dual is denoted by Y(E/F cyc ).
This latter module is clearly a quotient of X(E/F cyc ). The Weak Leopoldt Conjec-
ture for elliptic curves in this context is the assertion that
This conjecture is still open and is known to be true, for instance, if F = Q and
E(Q) has Mordell–Weil rank at most 1. A deep result of Kato [11] proves (4.3) for
all but a finite number of primes p for every elliptic curve E over Q. If (4.3) holds,
then one can show (see [3, Lemma 3.1]) that Y(E/F cyc ) is a torsion Λ (Γ )-module.
We formulated the following Conjecture jointly with Coates [3].
We believe that Y(E/F cyc ) is the analogue of X∞ in this context, and that the
above conjecture is the right analogue of Iwasawa’s μ = 0 conjecture for the motive
of an elliptic curve. In fact, the Galois group X∞ has a natural quotient X∞ , which
is the Galois group of the maximal abelian p-extension of F cyc that is unramified
everywhere, and in which all primes above p split completely. Note that all other
primes which do not lie above p automatically split in any unramified p-extension.
Iwasawa showed that μ (X∞ ) = μ (X∞ ). Let F(E p∞ ) denote the Galois extension of F
obtained by attaching to F the coordinates of all p-primary torsion points of E(F).
The following result is proved in [3, Theorem 3.4].
Theorem 4.5. Let p be an odd prime number such that the extension F(E p∞ )/F is
pro-p. Then Conjecture 4.4 holds for E over F cyc if and only if the Iwasawa μ = 0
conjecture holds for F cyc .
Thus, if Iwasawa’s μ = 0 conjecture were known to be true for all number fields,
then Conjecture 4.4 would follow. In general however, Conjecture 4.4 turns out to
be rather delicate to prove. This is to be expected, given its close relationship to
Iwasawa’s μ = 0 conjecture. Analogous to the case of the Tate motive, one can
establish the following result, which was independently noticed by Greenberg [7].
Proposition 4.6. Assume (4.3) holds. Then Conjecture 4.4 is equivalent to the as-
sertion that H 2 (Gal(FS /F cyc ), E p ) = 0.
and
Z 2 (E p /F cyc ) := lim H 2 (Gal(FS /F ), E p ),
←−
F
132 R. Sujatha
where the inverse limit is taken with respect to the natural corestriction maps over
all finite extensions F of F contained in F cyc . It can be shown (see [3]) that the
modules Z 2 (E p /F cyc ) and (H 2 (Gal(FS /F cyc ), E p ))∨ have the same μ -invariant.
Hence, the vanishing of H 2 (Gal(FS /F cyc ), E p ) is equivalent to the assertion that
Z 2 (E p /F cyc ) is finite. On the other hand, if (4.3) holds, then it can be shown [3]
that Z 2 (Tp E/F cyc ) is Λ (Γ )-torsion. Combining this with the fact that
and that the latter has μ -invariant zero if it is finite, it is easily seen that the finiteness
of Z 2 (E p /F cyc ) is equivalent to the assertion that Z 2 (Tp E/F cyc ) is a finitely gener-
ated Z p -module. To complete the proof, we only need to remark that Z 2 (Tp E/F cyc )
and Y(E/F cyc ) differ by finitely generated Z p -modules, a fact that can be deduced
from the Poitou–Tate exact sequence.
The module E p is trivialized over FS (p) as the field F1 is contained in FS (p) by our
hypothesis. Thus H 1 (Gal(QS /FS (p)), E p ) H 1 (Gal(QS /FS (p)), Z/p) = 0, by the
definition of FS (p). Further, H 2 (Gal(QS /FS (p)), E p ) H 2 (Gal(QS /FS (p)), Z/p ⊕
Z/p) = 0 (see [15]). Finally, since cd p (Gal(FS (p)/F cyc ) = 2, we conclude from the
Elliptic Curves and Iwasawa’s μ = 0 Conjecture 133
Then K(Ep∞ ) is Z p -extension of K(Ep ) and we write G = Gal(K(Ep∞ )/K) for the
corresponding Galois group and Λ (G) for the associated Iwasawa algebra. Let X∞
be the maximal abelian p-extension of K∞ that is unramified outside of the set of
primes above p. It is known [13] that X∞ is a finitely generated torsion Λ (G)-module.
Let Sp (E/K∞ ) (resp., Rp (E/K∞ )) denote the p-Selmer group (resp., p-fine Selmer
group) of E over K∞ . To be precise, these modules are defined by taking p instead
of p in (4.1) and Definition 4.2, respectively. Let Xp (E/K∞ ) and Yp (E/K∞ ) be the
respective compact duals of the Selmer group and the fine Selmer group. We then
have [13]
Sp (E/K∞ ) = Hom(X∞ , Ep∞ ),
and hence it follows that Xp (E/K∞ ) and Yp (E/K∞ ) are both finitely generated
Λ (G)-torsion modules. It was shown by Gillard [6] and Schneps [14] independently
that X∞ has μ -invariant zero. We denote by KS the maximal extension of K that is
unramified outside the primes of S. Let GS,p (K∞ ) denote the maximal pro-p quotient
of the Galois group Gal(KS /K∞ ). By a result of Perrin-Riou [13], it is known that
the Weak Leopoldt Conjecture holds for K∞ and the Galois modules Ep∞ and E p∞ ;
in other words that
Arguing as in Sect. 3 (cf. Theorem 3.3), it then follows that GS,p (K∞ ) is free and
hence has p-cohomological dimension at most 1. Further, using the Hochschild–
Serre spectral sequence as in the proof of Theorem 4.7, one then sees that
It seems difficult to directly deduce Conjecture A for K cyc or F cyc , where F = K(Ep )
by these methods. In [2], the following stronger conjecture was put forth (see
also [4]).
Conjecture 4.9. Let K be an imaginary quadratic field and let E be an elliptic
curve defined over K such that EndK (E) ⊗ Q is isomorphic to K. Let p be an odd
prime which splits in K, and such that E has good reduction at both primes of K
above p. Then the dual Selmer group of E over K cyc is a finitely generated Z p -
module. In particular, if E is defined over Q, the dual Selmer group of E over Qcyc
is a finitely generated Z p -module.
In this case, let F = K(E p ) and F∞ = K(E p∞ ), and put H = Gal(F∞ /F cyc ). All
we know at present is that the dual Selmer group of E over F∞ is a finitely gener-
ated Λ (H)-module if and only if the dual Selmer group of E over F cyc is a finitely
generated Z p -module.
References
16. W. S INNOTT, On the μ -invariant of the Γ -transform of a rational function, Invent. Math. 75
(1984), 273–282
17. C. W UTHRICH , Extending Kato’s result to curves with p-isogenies, Math. Res. Lett. 13
(2006), 713–718
Cohomological Invariants of Central Simple
Algebras with Involution
Jean-Pierre Tignol
Summary This survey reviews the various invariants with values in Galois
cohomology groups that have been defined for involutions on central simple al-
gebras following the model of the discriminant, Clifford invariant, and Arason
invariant of quadratic forms. From the orthogonal case to the unitary case to the
symplectic case, the degree of the invariants increases but their properties are
similar.
This text is based on two lectures delivered in January 2009 at the conference
“Quadratic forms, linear algebraic groups, and Galois cohomology” held at the
Jean-Pierre Tignol
Département de mathématique, Université catholique de Louvain, Chemin du cyclotron, 2, B-1348
Louvain-la-Neuve, Belgium
e-mail: [email protected]
2010 Mathematics subject classification. 11E72.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 137
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 8,
c Springer Science+Business Media, LLC 2010
138 J.-P. Tignol
1 → μ2 → Fs× −
→ Fs× → 1
2
the notation · for the cup product in the cohomology ring H ∗ (F), so for a, b ∈ F × the
Brauer class of the quaternion algebra (a, b)F is (a) · (b).
Various invariants are classically defined to determine whether quadratic forms over
an arbitrary field F are isometric. The first invariant is of course the dimension. To
obtain an invariant that vanishes on hyperbolic forms, one considers the dimension
modulo 2:
e0 (q) = dim q (mod 2) ∈ Z/2Z.
The next invariant is the discriminant: for q a quadratic form of dimension n, we set
They were vastly generalized during the last quarter of the twentieth century: see
Pfister’s survey [42] for an account of the historical development of the subject.
After Voevodsky’s proof of the Milnor conjecture and additional work of Orlov–
Vishik–Voevodsky, we have surjective homomorphisms
On the other hand, the Arason–Pfister Hauptsatz [13, Theorem 23.7] states that the
dimension of any anisotropic quadratic
form q representing a nonzero element in
I n F satisfies dim q ≥ 2n , hence n I n F = {0}. Therefore, the problem of deciding
whether two quadratic forms q1 and q2 over F are isometric can in principle be
solved by computing cohomology classes: indeed, q1 and q2 are isometric if and
only if dim q1 = dim q2 and q1 − q2 is hyperbolic, and the latter condition can be
checked by computing successively2 e0 (q1 − q2 ), e1 (q1 − q2 ), e2 (q1 − q2 ), . . ., which
should all vanish. In view of the Arason–Pfister Hauptsatz, it actually suffices to
check that ed (q1 − q2) = 0 for all d ≥ 0 with 2d ≤ dim q1 + dim q2 .
This remarkably complete classification result is a model that we want to emu-
late for other algebraic objects. The problem can be formalized as follows: we are
given a base field F and a functor A : Fields F → Sets from the category of fields
containing F, where the morphisms are F-algebra homomorphisms, to the category
of sets. Typically, A(K) is the set of K-isomorphism classes of some objects (like
quadratic forms) that allow scalar extension. An invariant of A with values in a
functor H : Fields F → Sets is a natural transformation of functors:
e : A → H.
cohomological invariants of quadratic forms are determined in [17, Sect. 17]: they
are generated by the so-called Stiefel–Whitney invariants. (The invariants e1 and e2
above can be computed in terms of Stiefel–Whitney classes, but not ei for i ≥ 3,
see [32, pp. 135, 367].)
In the present notes, we consider the case where A(F) is the set of isomorphism
classes of central simple F-algebras with involution, as explained in the next sec-
tion. Our aim is not to describe the collection of all cohomological invariants of A,
but rather to define a sequence of invariants e1 , e2 , . . ., each of which is defined on
the subfunctor of A on which the preceding invariant vanishes, just as in the case of
quadratic forms. Invariants of degree 2 are essentially Brauer classes of the Tits al-
gebras that arise in the representation theory of linear algebraic groups [53]. Indeed,
central simple algebras with involution define torsors under adjoint classical groups
(see Sect. 2), and our approach via central isogenies owes much to Tits’s discussion
of what he calls the β -invariant in [53, 54].
From the Bayer-Fluckiger–Parimala proof of Serre’s conjecture II for classical
groups [2], it follows that the invariants of degree 1 and 2 are sufficient to classify
involutions over a field of cohomological dimension 1 or 2. We shall not address
this topic here, and refer to [8, 35] for this classification. Below, we aim at classifi-
cation results of a different kind, restricting the dimension of the algebra rather than
the cohomological dimension of the center (see Theorem 3.1 for a paradigmatic
case).
Note that the involution is really adjoint to the bilinear form, not to the quadratic
form, hence we also use the notation adb for adq . Since every F-automorphism of
EndF V is inner, it is easy to check that every F-linear involution on EndF V is
adjoint to a nonsingular bilinear form that is either symmetric or skew-symmetric.
There are therefore two types of involutions of the first kind on EndF V , which can be
distinguished by the dimension of their space of symmetric elements: the orthogonal
involutions are adjoint to quadratic forms (or symmetric bilinear forms) and the sym-
plectic involutions are adjoint to skew-symmetric bilinear forms. The involutions
Cohomological Invariants of Central Simple Algebras with Involution 141
actually determine up to a scalar factor the form to which they are adjoint, so
mapping each nonsingular symmetric or skew-symmetric form b to the adjoint in-
volution adb defines bijections:
nonsingular symmetric
orthogonal involutions
bilinear forms on V up ←→
on EndF V
to a scalar factor
and
nonsingular skew-symmetric
symplectic involutions
bilinear forms on V up to a ←→
on EndF V
scalar factor
Central simple F-algebras are twisted forms (in the sense of Galois cohomology)
of endomorphism algebras; therefore, one may also distinguish two types of invo-
lutions of the first kind on a central simple F-algebra A: an involution σ : A → A is
called orthogonal (resp., symplectic) if for any splitting field K and under any K-
isomorphism A ⊗F K EndK V , the involution σ ⊗ IdK obtained by scalar extension
is adjoint to a nonsingular symmetric (resp., skew-symmetric) bilinear form. Note
that symplectic involutions exist only on central simple algebras of even degree,
since every skew-symmetric bilinear form on an odd-dimensional vector space is
singular.
Involutions on a central simple algebra can also be obtained by adjunction, as in
the split case: if we fix a representation A = EndD V , where D is a central division
F-algebra Brauer equivalent to A and V is a (right) D-vector space, and choose an
involution of the first kind θ on D, then every involution of the first kind on A is
adjoint to a nonsingular hermitian or skew-hermitian form on V with respect to θ ,
and this hermitian form is uniquely determined up to a factor in F × .
The correspondence between involutions of the first kind and bilinear forms can
also be described in terms of the corresponding automorphism groups. Recall that
the orthogonal group On is the group of isometries of the standard form 1, . . . , 1
of dimension n:
On (F) = Aut(1, . . . , 1).
Let H 1 (F, On ) denote the nonabelian Galois cohomology set:
H 1 (F, On ) = H 1 Gal(Fs /F), On (Fs ) .
The usual technique of nonabelian Galois cohomology (see [49, Chap. III, Sect. 1]
or [29, (29.28)]) yields a canonical bijection:
This bijection maps the isometry class of the standard form to the distinguished
element in H 1 (F, On ).
142 J.-P. Tignol
Int(g) ◦ t = t ◦ Int(g).
The automorphisms of the transpose involution are the Int(g) with g ∈ GOn (F),
hence
Aut(t) = PGOn (F) := GOn (F)/F × .
Again, Galois cohomology yields a bijection (see [29, Sect. 29.F]):
Under this bijection, the isomorphism class of the transpose involution corre-
sponds to the distinguished element of H 1 (F, PGOn ). The canonical map On (Fs ) →
PGOn (Fs ) yields a map H 1 (F, On ) → H 1 (F, PGOn ). Under the bijections (2.1) and
(2.2), this map carries the isometry class of any quadratic form q of dimension n to
the isomorphism class of the adjoint involution adq on the split algebra of degree n.
In view of the close relationship between quadratic forms and (orthogonal) invo-
lutions, it may be expected that invariants for quadratic forms have counterparts for
involutions. There is, however, a significant difference to keep in mind: in contrast
with quadratic forms, there is no3 direct sum of involutions. Therefore, the problem
of deciding whether two involutions are isomorphic cannot be readily reduced to the
problem of deciding whether an involution is hyperbolic (i.e., adjoint to a hyperbolic
hermitian or skew-hermitian form).
In the next sections, we successively discuss orthogonal involutions, involutions
of the second kind (also called unitary involutions), and symplectic involutions. In
each case, some invariants with striking common features are defined.
3 Orthogonal Involutions
Any involution of the first kind on a central simple algebra A defines an isomorphism
between A and its opposite algebra, hence 2[A] = 0 in the Brauer group. Therefore, A
is split if its degree is odd, and in this case every involution is orthogonal and adjoint
to a quadratic form q. Upon scaling, q can be assumed to have trivial discriminant;
it is then uniquely determined. (These observations also follow from the equality
3 More exactly, direct sums of involutions are defined only with some additional data (see [11]).
Cohomological Invariants of Central Simple Algebras with Involution 143
PGOn = On for n odd.) Thus, the case of involutions on central simple algebras of
odd degree immediately reduces to the case of quadratic forms; we shall not discuss
it further.
In this section, we consider orthogonal involutions on central simple algebras
of even degree. We first review in Sects. 3.1 and 3.2 the cases where the algebra
has Schur index 1 or 2. A collection of invariants ei can then be defined for all i.
In the following sections, we successively discuss invariants e1 , e2 , and e3 without
restriction on the index of the algebra.
Just as for involutions on central simple algebras of odd degree, the case of orthog-
onal involutions on split central simple algebras of even degree reduces to the case
of quadratic forms. Note that if q ∈ I d F for some d ≥ 1, then for any λ ∈ F × we
have
q ≡ λ q mod I d+1 F,
hence ed (q) = ed (λ q). Since ed only depends on the similarity class of q, we may
consider it as attached to the isomorphism class of the adjoint involution and set
The right-hand side is a tensor product of (split) quaternion algebras with (orthogo-
nal) involution. Conversely, if (A, σ ) is a tensor product of quaternion algebras with
involution, the solution of the Pfister factor conjecture by Becher [5] shows that q is
a multiple of some (d + 1)-fold Pfister form, hence ei (q) = 0 for i = 1, . . . , d.
Now, suppose deg A = 2d and σ = adq , σ = adq are orthogonal involutions on
A with ei (σ ) = ei (σ ) for i = 1, . . . , d − 1, that is, q, q ∈ I d F. Let λ ∈ F × (resp.,
λ ∈ F × ) be a represented value of q (resp., q ). As observed at the beginning of this
section, we have
In this section, we assume that the Schur index ind A is 2, that is, A is Brauer equiv-
alent to some quaternion division F-algebra Q. Then A can be represented as
A = EndQ V
for some right Q-vector space V of dimension dimQ V = (1/2) degA, and every
orthogonal involution σ on A is adjoint to some skew-hermitian form h on V with
respect to the conjugation involution on Q. The form h is uniquely determined up
Cohomological Invariants of Central Simple Algebras with Involution 145
such that
ei (σ )F(X) = ei (σF(X) ).
Note that for i ≥ 2, the group [A] · H i−2(F) can be identified with the kernel of the
scalar extension map H i (F, μ4⊗i−1 ) → H i (F(X), μ4⊗i−1 ), as shown in the proof of
Berhuy [6, Proposition 9].
I do not know whether the analogue of the classification result from Theorem 3.1
holds for algebras of index 2 (except for the cases of low degree discussed in
Theorems 3.6 and 3.10). The issue is whether skew-hermitian forms over Q are
similar when they are similar after scalar extension to F(X ).
146 J.-P. Tignol
3.3 Discriminant
PGO2m = PGO(mH).
It follows that the e1 -invariant defined earlier coincides with the e1 -invariant defined
in Sects. 3.1 and 3.2 when ind A ≤ 2.
Knus–Parimala–Sridharan [30] give a nice direct definition of the determinant:
if σ is an orthogonal involution on a central simple algebra A of even degree, they
show that all the skew-symmetric units have the same reduced norm up to squares
and that
In a slightly different form, this formula already appears in Tits’s seminal work [52,
Sect. 2.6].
This first invariant is of course rather weak. Yet, we have the following result.
Part (b) was first observed by Knus–Parimala–Sridharan [30]. Note that in any
decomposition of the form (3.7), the involutions σ1 and σ2 are of the same type
(orthogonal or symplectic) since σ is orthogonal. However, when there is any de-
composition of the type (3.7), there is one where σ1 and σ2 are the quaternion con-
jugations (which are the unique symplectic involutions on Q1 and Q2 ). The subal-
gebras Q1 and Q2 are then uniquely determined by σ (see [29, p. 215]) (or (3.11)).
An analogue of the even Clifford algebra of quadratic forms has been defined
for orthogonal involutions by Jacobson [23] (by Galois descent) and by Tits [52]
(rationally) (see also [29, Sect. 8]). For σ an orthogonal involution on a central
simple F-algebra A of degree n = 2m, the Clifford √ algebra C(A, σ ) has dimen-
sion 2n−1 and its center Z(A, σ ) is isomorphic to F( disc σ ) if disc σ = 1, to F × F
if disc σ = 1. It is simple if Z(A, σ ) is a field and is a direct product of two central
simple F-algebras C+ (A, σ ) and C− (A, σ ) of degree 2m−1 if Z(A, σ ) F × F. For
any quadratic form q on an F-vector space V of even dimension, we have
If disc q = 1, then C+ (EndF V, adq ) and C− (EndF V, adq ) are isomorphic and Brauer
equivalent to the full Clifford algebra C(q).
The construction of the Clifford algebra C(A, σ ) is functorial and may be used
to obtain a second invariant of orthogonal involutions of even degree and trivial dis-
criminant in the same spirit as the Witt invariant of quadratic forms, as we proceed
to show.
The set of isomorphism classes of orthogonal involutions of trivial discriminant
on central simple algebras of even degree n = 2m corresponds under the bijec-
tion (2.2) to the kernel of the map δ 1 of (3.5), which is the image of H 1 (F, PGO+ n)
in H 1 (F, PGOn ). These involutions can therefore be classified by H 1 (F, PGO+n ), but
148 J.-P. Tignol
the map H 1 (F, PGO+n ) → H (F, PGOn ) is generally not injective: a given central
1
∼
If ϕ , ϕ : Z(A, σ ) −
→ F × F are the two different isomorphisms, the triples (A, σ , ϕ )
and (A, σ , ϕ ) are isomorphic if and only if (A, σ ) has an automorphism whose
induced action on C(A, σ ) swaps the two components or, equivalently, if A contains
an improper similitude, that is, an element g such that σ (g)g ∈ F × and NrdA (g) =
−(σ (g)g)n/2 . This readily follows from the exact sequence below, which is a part
of the cohomology sequence derived from a twisted version of (3.4):
δ δ1
→ μ2 → H 1 (F, PGO+ (A, σ )) → H 1 (F, PGO(A, σ )) −→ H 1 (F).
PGO(A, σ ) −
In particular, if A is split, then the triples (A, σ , ϕ ) and (A, σ , ϕ ) are isomorphic.
The group PGO+ n has a simply connected cover Spinn and we have an exact
sequence of algebraic groups:
1 → μ → Spinn → PGO+
n → 1, (3.8)
∂ : H 1 (F, PGO+
n ) → H (F, μ ),
2
and, if n ≡ 0 mod 4,
[C+ (A, σ )], [C− (A, σ )] and [C− (A, σ )], [C+ (A, σ )] ∈ H 2 (F) × H 2 (F).
2[C+ (A, σ )] = 2[C− (A, σ )] = [A] and [C+ (A, σ )] + [C− (A, σ )] = 0. (3.9)
Cohomological Invariants of Central Simple Algebras with Involution 149
2[C+ (A, σ )] = 2[C− (A, σ )] = 0 and [C+ (A, σ )] + [C− (A, σ )] = [A].
Therefore, letting B2A denote the subgroup of Br(F) generated by [A], that is, B2A =
{0, [A]} = [A] · H 0(F), we may set
Sketch of proof. If degA = 4, the two components C± (A, σ ) of the Clifford algebra
are quaternion algebras. Letting σ± denote the quaternion conjugation on C± (A, σ ),
we have a decomposition
λ 2C+ (A, σ ) (resp., λ 2C− (A, σ )) of degree 6 endowed with a canonical involution
γ+ (resp., γ− ). As a consequence of the coincidence of Dynkin diagrams D3 = A3
(see [29, (15.32)]), we have isomorphisms
Since e2 (σ ) determines the pair {[C+ (A, σ )], [C− (A, σ )]}, it follows that orthogonal
involutions of trivial discriminant on A are classified by their e2 -invariant.
Now, suppose deg A ≡ 2 mod 4. If C+ (A, σ ) is split, it follows from (3.9) that
[A] = 0. Similarly, if [C+ (A, σ )] = [A], then since (3.9) shows that 2[C+ (A, σ )] =
[A] it follows that [C+ (A, σ )] = [A] = 0. Therefore, in each case we have (A, σ )
(EndF V, adq ) for some quadratic form q, and
e2 (q) = e2 (σ ) = 0,
If C+ (A, σ ) is split, then (A, σ ) is isomorphic to one of the components of the even
Clifford algebra of a quadratic form of dimension 8, from which the existence of a
decomposition easily follows.
In view of part (d) of Theorem 3.10, the case of orthogonal involutions with
e1 = e2 = 0 on central simple algebras of degree 2 (mod 4) is reduced to the split
case. We thus have invariants ei as in Sect. 3.1, and Theorem 3.1 applies. Therefore,
in the rest of this section, we only consider central simple algebras of degree 0
(mod 4).
The hope to define further invariants of orthogonal involutions on the model of the
Arason e3 -invariant of quadratic forms was dashed by an example of Quéguiner-
Mathieu [45] or [4, Sect. 3.4]. The following variation on Quéguiner-Mathieu’s ex-
ample was suggested by Becher: consider quaternion algebras with orthogonal in-
volutions (Q1 , σ1 ), (Q2 , σ2 ), and (Q3 , σ3 ) over an arbitrary field F, and let
and
[C− (A, σ ) ⊗F C+ (A, σ0 )], [C+ (A, σ ) ⊗F C− (A, σ0 )] .
Since e2 (σ ) = 0, one of the components C± (A, σ ) is split and the other is Brauer
equivalent to A. The same holds for σ0 since e2 (σ0 ) = 0; therefore, we have
∂ (η ) = 0 or ∂ (η ) = 0, and at least one of η , η lifts to some ξ ∈ H 1 F, Spin(A, σ0 ) .
The analysis of the exact sequence (3.12) given in [15] shows how the Rost invari-
ant of ξ varies when a different lift ξ is chosen: the difference of Rost invariants
R(ξ ) − R(ξ ) lies in the subgroup B3A ⊂ H 3 (F, μ4⊗2 ), which is defined to be the
image under the injection H 3 (F) → H 3 (F, μ4⊗2 ) of the subgroup
In the particular case where the Schur index ind A divides (1/2) degA, that is,
when A M2 (A ) for some central simple algebra A , we may choose for σ0 a
hyperbolic involution and thus consider e3 (σ /σ0 ) as an absolute invariant e3 (σ ).
When (A, σ ) (EndF V, adq ), we have B3A = {0} and
e3 (adq ) = e3 (q)
in view of the relation between the Rost and the Arason invariants. Therefore, the
e3 -invariant defined earlier coincides with the e3 -invariant of Sects. 3.1 and 3.2 when
ind A ≤ 2.
In the case where deg A = 8 (and ind A divides 4), an explicit computation of the
e3 -invariant was recently obtained by Quéguiner-Mathieu and Tignol: if e1 (σ ) =
e2 (σ ) = 0, we may find a decomposition
for some λ ∈ F × , some central simple F-algebra D of degree 4, and some orthogo-
nal involution θ on D such that the quaternion F-algebra (disc θ , λ )F is split: see [5]
if ind A = 1 or 2 and [38] if ind A = 4. The Clifford algebra C(D, θ ) is a√quaternion
algebra over a quadratic étale F-algebra Z, which is isomorphic to F( disc θ ) if
disc θ = 1 and to F × F if disc θ = 1. Therefore, we may find an element ∈ Z such
that NZ/F () = λ . The e3 -invariant is
e3 (σ ) = corZ/F () · [C(D, θ )] + B3A ∈ H 3 (F)/B3A ,
Proof. The “only if” part is clear. We may thus assume ei (σ ) = 0 for i = 1, 2, 3,
and we have to show that σ is hyperbolic. If deg A < 8, the theorem follows from
Theorem 3.10(a). If A is split (which, by Theorem 3.10(d), occurs in particular when
deg A ≡ 2 (mod 4)), the theorem follows from Theorem 3.1. Likewise, the theorem
follows from Theorem 3.3 if ind A = 2. Therefore, it suffices to consider the case
where deg A = 8 and ind A = 4. Let FA be the function field of the Severi-Brauer
variety of A. Since A splits over FA , and since the theorem holds in the split case,
(A, σ )FA is hyperbolic. It follows that σ is hyperbolic by a theorem of Sivatski [50,
Proposition 3] (based on a result of Laghribi [31, Théorème 4]).
Theorem 3.13 yields the expected hyperbolicity criterion for degA < 16, under
the hypothesis that ind A | (1/2) degA (which is necessary for σ to be hyperbolic).
Whether the decomposition criterion holds when deg A = 16 is an open question,
which is addressed by Garibaldi [15]. By contrast, Quéguiner-Mathieu and Tignol
have used an example of Hoffmann [22] and a result of Sivatski [50, Proposition 4]
to construct nonisomorphic orthogonal involutions σ , σ on a central simple algebra
A of degree 8 and index 4 such that
e1 (σ ) = e1 (σ ) = e2 (σ ) = e2 (σ ) = 0 and e3 (σ ) = e3 (σ ),
hence the classification criterion does not hold in degree 8. If σ0 is a fixed orthogonal
involution with e1 (σ0 ) = e2 (σ0 ) = 0 on a central simple algebra A of degree 8 and
index 4, orthogonal involutions σ on A with e1 (σ ) = e2 (σ ) = 0 and e3 (σ ) = e3 (σ0 )
are classified by a relative invariant:
e4 (σ /σ0 ) ∈ H 4 (F)/B4σ0 ,
where
B4σ0 = {(a) · e3(σ0 ) | a ∈ F × , (a) · [A] = 0}.
Garibaldi [15] shows how to use the Rost invariant of groups of type E8 to obtain
an absolute e3 -invariant for orthogonal involutions on central simple algebras of
degree 16.
4 Unitary Involutions
√
Let K = F( a) be a quadratic field extension of F with nontrivial automorphism ι .
For each integer n ≥ 1, let hn be the n-dimensional hermitian form over K with
maximal Witt index, defined for x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) ∈ K n by
More precisely, the set on the left consists of the isomorphism classes of triples
(B, τ , ϕ ), where B is a central simple algebra of degree n over a quadratic extension
Z of F, τ is an involution on B that restricts to the nontrivial F-automorphism of Z,
∼
and ϕ : Z − → K is an F-algebra isomorphism.
When B is split, the study of unitary involutions reduces to the quadratic form case
by an observation due to Jacobson: every hermitian form h : V × V → K on a K-
vector space V yields a quadratic form qh : V → F on V (viewed as an F-vector
space) by
qh (x) = h(x, x) for x ∈ V .
1
h(x, y) = qh (x + y) − qh (x) − qh (y) + qh (x + yα )α −1 − qh (x)α −1 − qh (yα )α −1 .
2
Cohomological Invariants of Central Simple Algebras with Involution 155
Therefore, the invariants of h and qh are the same and the invariants of adh are
the invariants of the similarity class of qh (which is always even dimensional since
dimF V = 2 dimK V ). If ei (qh ) is defined for some i, we set
(B, τ ) → (A, σ ).
h = a1 , . . . , an K ,
Therefore,
(a) if n is odd,
e1 (adh ) =
0 if n is even,
and, when n is even,
Comparing with Theorem 3.1, note the shift in the degree of the algebra.
Proof. Let τ = adh , so ei (τ ) = ei (qh ) when defined. Note that dim qh = 2 degB, so
in case (a) qh is hyperbolic if and only if qh ∈ I d+1 F, and this condition is equivalent
to ei (qh ) = 0 for i = 2, . . . , d. If degB = 2d and ei (τ ) = 0 for i = 2, . . . , d, then qh is
a (d + 1)-Pfister form of the type
qh = a, a1 , . . . , ad for some a1 , . . . , ad ∈ F × ,
hence
h = a1 , . . . , ad K .
Therefore,
Conversely, suppose there are quaternion F-algebras Q1 , . . . , Qd with conjugation
involutions γ1 , . . . , γd such that
Since the group PGUn,K is connected, the procedure that yields the discriminant of
orthogonal involutions in Sect. 3.3 does not apply here. The group PGUn,K is not
simply connected however, so we may obtain cohomological invariants of degree 2
as in the orthogonal case. The simply connected cover of PGUn,K is the special
unitary group:
SUn,K := {u ∈ Un,K | det(u) = 1}.
Its center is a twisted version of the group of nth roots of unity:
Henceforth, we assume that the characteristic of F does not divide n, so that μn,K is
a smooth algebraic group. Its group of rational points over Fs is isomorphic to the
group of nth roots of unity in Fs , with a twisted Galois action that disappears after
scalar extension to K. From the exact sequence
The features of this map are significantly different according to whether n is odd
or even.
158 J.-P. Tignol
is injective, and identifies H 2 (F, μn,K ) with the kernel of the corestriction map
see [29, (30.12)]. Comparing the exact sequence (4.3) with the corresponding exact
sequence after scalar extension to K, which is
1 → μn → SLn → PGLn → 1,
Let n = 2m. Besides the restriction map resK/F , we may also consider the map
defined by raising to the mth power:
Let (B, τ ) be a central simple K-algebra of degree n with unitary involution, and
let Δ (B, τ ) ∈ H 2 (F, μn,K ) be the image under ∂ of the corresponding cohomology
class in H 1 (F, PGUn,K ). As in the case where n is odd, we have
the discriminant algebra of (B, τ ) in [29, Sect. 10], and the e2 -invariant defined ear-
lier coincides with the e2 -invariant of unitary involutions on split algebras defined
in Sect. 4.1.
The same method as in Sect. 3.5 allows one to derive from Rost’s invariant for SUn,K
a relative
√ invariant of unitary involutions on a given central simple algebra B over
K = F( a). Suppose deg B = n is even and not divisible by the characteristic, and
let τ0 be a unitary involution
on B with e2 (τ0 ) = 0. By Proposition 4.4, the coho-
mology class in H 1 F, PGU(B, τ0 ) corresponding to a unitary involution τ on B
with e2 (τ ) = 0 lifts to some ξ ∈ H 1 F, SU(B, τ0 ) . The Rost invariant R(ξ ) lies in
H 3 (F, μd⊗2 ), where d is the Dynkin index of SU(B, τ0 ), determined in [17, (12.6),
p. 142]:
exp B if n is a 2-power and exp B = n or n/2,
d=
2 exp B otherwise.
In view of the description of the Rost invariant on the center of SU(B, τ0 ) given by
Merkurjev–Parimala–Tignol [40], one should be able to define
where
B3B = corK/F [B] · H 1 (K, μd ) .
As far as I know, this invariant has not been investigated yet.
Although the case where the degree n is odd does not pertain to the line of investi-
gation developed so far in this text, it is worth mentioning that a similar construction
based on the Rost invariant is better documented in this case: consider the F-vector
space Sym(τ ) of τ -symmetric elements in B and the quadratic form
so that f3 (τ /τ0 ) = f3 (τ ) − f3 (τ0 ). (The form 1 + n−1 2 2, 2a is Witt equivalent
to the form Qτ for τ the adjoint involution of the n-dimensional hermitian form of
maximal Witt index.) The absolute invariant f3 was first investigated in the particu-
lar case where deg B = 3 by Haile–Knus–Rost–Tignol [20]. It classifies the unitary
involutions on a given central simple algebra of degree 3 up to isomorphism (see [29,
Sect. 19.B and (30.21)]).
There√is also an absolute invariant of degree 4 defined just for degB = 4 and
K = F( −1) by Rost–Serre–Tignol [46]: letting nD denote the norm form of the
quaternion F-algebra Brauer equivalent to the discriminant algebra D(B, τ ), the in-
variant is
f4 (τ ) = e4 (Qτ − nD ) ∈ H 4 (F).
It vanishes if and only if B is generated
√ by two elements x, y ∈ Sym(τ ) such that x4 ,
×
y ∈ F and yx = ixy, where i = −1 ∈ K. If B is split and τ = adh with
4
h = a1 , a2 , a3 , a4 K ,
we have
f4 (τ ) = (−1) · (−a1a2 ) · (−a1 a3 ) · (−a1a4 ).
Note that the fourfold Pfister form −1, −a1a2 , −a1 a3 , −a1 a4 is indeed an in-
variant of the hermitian form h: in the notation of Garibaldi et al. [17, p. 67], we
have
−1, −a1 a2 , −a1 a3 , −a1 a4 = 2 + λ22(h) + λ24(h).
Cohomological Invariants of Central Simple Algebras with Involution 161
5 Symplectic Involutions
Let n be an even integer, n = 2m, and let sn be the following skew-symmetric matrix
of order n: 0 1 0 1
sn = diag −1 0 , . . . , −1 0 .
m
σ (m) = s−1
n · m · sn for m ∈ Mn (F).
The symplectic group is the automorphism group of the bilinear form with Gram
matrix sn :
Let A be a central simple F-algebra of degree n and index 2, that is, A is Brauer
equivalent to a quaternion F-algebra Q. As in Sect. 3.2, we may represent A as
A = EndQ V
for some right Q-vector space V with dimQ V = n/2. Every symplectic involution on
A is adjoint to some hermitian form h on V with respect to the conjugation involution
162 J.-P. Tignol
The correspondence between the symplectic involution adh and the orthogonal
involution adjoint to qh was used by MacDonald [36, Theorem 4.7] to determine all
the invariants of symplectic involutions on central simple algebras A with ind A = 2
and deg A ≡ 2 mod 4. This correspondence can also be described directly, as the
next lemma shows.
ei (σ ) = ei (σ ⊗ γ ).
Let m = n/2 and let (vα )mα =1 be an orthogonal Q-base of V with respect to h, in
which h has the diagonalization
h = λ1 , . . . , λm Q .
Note that the involution adh always decomposes: if V0 ⊂ V denotes the F-span of
(vα )m
α =1 , then V = V0 ⊗F Q and
Proof. Using the same notation as in Lemma 5.1, part (a) readily follows from
Theorem 3.1 since ei (σ ) = ei (σ ⊗ γ ) and σ is hyperbolic if and only if σ ⊗ γ is
hyperbolic.
If degA = 2d and ei (σ ) = 0 for i = 2, . . . , d, then qh is a (d + 1)-Pfister form
Part (a) is clear. Part (b) was first observed by Rowen [47, Theorem B]. For a
proof based on the coincidence of Dynkin diagrams B2 = C2 , see [54, p. 131] or
[29, (16.16)].
Nrp : Sym(A, σ0 ) → F
Every symplectic involution σ on A has the form σ = Int(s) ◦ σ0 for some unit
s ∈ Sym(σ0 ) uniquely determined up to a scalar factor.
Since Nrp(1λ s) = λ Nrp(s)
m
for λ ∈ F, and since m is even, the square class Nrp(s) ∈ H (F) is uniquely
determined by σ . We may thus set
e3 (σ /σ0 ) = Nrp(s) · [A] ∈ H 3 (F).
4 This condition determines q uniquely up to a scalar factor (see [29, Sect. 15.C]).
166 J.-P. Tignol
hence
e3 (σ /σ0 ) = e3 (σ /σ0 ).
Therefore, e3 (σ /σ0 ) depends only on the isomorphism class of σ : it is an invariant
of symplectic involutions on A.
Alternatively, as pointed out by Garibaldi, the invariant e3 (σ /σ0 ) may be ob-
tained by mimicking the argument in Sect. 3.5, using the following twisted version
of (5.5):
1 → μ2 → Sp(A, σ0 ) → PGSp(A, σ0 ) → 1.
The involution σ defines a cohomology class ξ ∈ H 1 F, PGSp(A, σ0 ) that can be
lifted to H 1 F, Sp(A, σ0 ) . The Rost invariant of the lift is the invariant e3 (σ /σ0 ).
Note that in this case the Rost invariant does not depend on the choice of the lift of
ξ since it vanishes on cocycles that come from the center of Sp(A, σ0 ) (see [40]).
Therefore, we do not have to factor out H 3 (F) by a subgroup depending on A to
obtain a well-defined relative invariant.
Theorem 5.7. If degA = 4, then the symplectic involutions σ and σ0 are conjugate
if and only if e3 (σ /σ0 ) = 0.
Then Q1 ⊗ Q2 is division, and remains division over the generic splitting field FQ3
of Q3 . Let X be the projective quadric defined by the vanishing of an Albert form of
Q1 ⊗ Q2 . Over the function field F(X ), the product Q1 ⊗ Q2 is not division, hence we
may find e3 (σF(X) ) ∈ H 3 (F(X)). Similarly, since Q3 splits over FQ3 , we may find
e3 (σFQ3 ) ∈ H 3 (FQ3 ). One may check that e3 (σF(X) ) is unramified over F, that is,
it is in the kernel of all the residue maps corresponding to points of codimension 1
on X. Arason [1, (5.6)] proved that the scalar extension map H 3 (F) → Hnr 3 (F(X ))
is injective, and Kahn [25] showed that its cokernel is Z/2Z, the nontrivial el-
ement being represented by the relative invariant e3 (ρF(X) /ρ0 ), where ρ is any
symplectic involution on Q1 ⊗ Q2 and ρ0 is a hyperbolic symplectic involution on
(Q1 ⊗ Q2 )F(X) . Since Q1 ⊗ Q2 remains division after scalar extension to FQ3 , the
form ϕ is anisotropic over FQ3 , and we have a commutative diagram where the ver-
tical maps are given by scalar extension:
hence a diagram chase shows that e3 (σF(X) ) is the image of a unique element
e3 (σ ) ∈ H 3 (F). Thus, e3 is well defined and unique on the set of isomorphism
classes of symplectic involutions on tensor products of three quaternion algebras.
The proof in the general case relies on the same arguments, using induction on the
minimal number of terms in a decomposition of [A] into a sum of Brauer classes of
quaternion algebras. See [18, Sect. 2].
Besides those that are explicitly defined in terms of the trace form (at the end of
Sect. 4.3), several invariants defined earlier5 have an alternative description in terms
of the trace form. Throughout this appendix, we use the following notation: for σ
an involution of arbitrary type on a central simple algebra A, we let Qσ denote the
quadratic form
Qσ : Sym(σ ) → F, x → TrdA (x2 ),
where F is the subfield fixed under σ in the center of A.
Suppose first σ is orthogonal and deg A = n = 2m. Lewis [33, Theorem 1] and
Quéguiner [44, Sect. 2.2] computed
det Qσ = 2m det σ .
det Qτ = (−a)n(n−1)/2 · F ×2 ∈ F × /F ×2 .
5 Essentially the first nontrivial invariant in each of the orthogonal, unitary, and symplectic case.
Cohomological Invariants of Central Simple Algebras with Involution 169
References
18. Garibaldi, S., Parimala, R., Tignol, J.P.: Discriminant of symplectic involutions. Pure Appl.
Math. Q. 5(1), 349–374 (2009)
19. Garibaldi, S., Quéguiner-Mathieu, A.: Pfister’s theorem for orthogonal involutions of degree
12. Proc. Am. Math. Soc. 137(4), 1215–1222 (2009)
20. Haile, D.E., Knus, M.A., Rost, M., Tignol, J.P.: Algebras of odd degree with involution, trace
forms and dihedral extensions. Israel J. Math. B 96, 299–340 (1996)
21. Hoffmann, D.W.: Isotropy of 5-dimensional quadratic forms over the function field of a
quadric. In: K-theory and algebraic geometry: connections with quadratic forms and division
algebras (Santa Barbara, CA, 1992). Proceedings of the Symposium on Pure Mathematics,
vol. 58, pp. 217–225. American Mathematical Society, Providence, RI (1995)
22. : Similarity of quadratic forms and half-neighbors. J. Algebra 204(1), 255–280 (1998)
23. Jacobson, N.: Clifford algebras for algebras with involution of type D. J. Algebra 1, 288–300
(1964)
24. : Some applications of Jordan norms to involutorial simple associative algebras. Adv.
Math. 48(2), 149–165 (1983)
25. Kahn, B.: Lower H -cohomology of higher-dimensional quadrics. Arch. Math. (Basel) 65(3),
244–250 (1995)
26. Karpenko, N., Quéguiner, A.: A criterion of decomposability for degree 4 algebras with uni-
tary involution. J. Pure Appl. Algebra 147(3), 303–309 (2000)
27. Karpenko, N.A.: On anisotropy of orthogonal involutions. J. Ramanujan Math. Soc. 15(1),
1–22 (2000)
28. Knus, M.A., Lam, T.Y., Shapiro, D.B., Tignol, J.P.: Discriminants of involutions on biquater-
nion algebras. In: K-theory and algebraic geometry: connections with quadratic forms and
division algebras (Santa Barbara, CA, 1992). Proceedings of the Symposium on Pure Mathe-
matics, vol. 58, pp. 279–303. American Mathematical Society, Providence, RI (1995)
29. Knus, M.A., Merkurjev, A., Rost, M., Tignol, J.P.: The book of involutions. American Math-
ematical Society Colloquium Publications, vol. 44. American Mathematical Society, Provi-
dence, RI (1998); with a preface in French by J. Tits
30. Knus, M.A., Parimala, R., Sridharan, R.: Pfaffians, central simple algebras and similitudes.
Math. Z. 206(4), 589–604 (1991)
31. Laghribi, A.: Isotropie de certaines formes quadratiques de dimensions 7 et 8 sur le corps des
fonctions d’une quadrique. Duke Math. J. 85(2), 397–410 (1996)
32. Lam, T.Y.: Introduction to quadratic forms over fields. Graduate Studies in Mathematics,
vol. 67. American Mathematical Society, Providence, RI (2005)
33. Lewis, D.W.: A note on trace forms and involutions of the first kind. Exposition. Math. 15(3),
265–272 (1997)
34. Lewis, D.W., Tignol, J.P.: On the signature of an involution. Arch. Math. (Basel) 60(2),
128–135 (1993)
35. : Classification theorems for central simple algebras with involution. Manuscripta
Math. 100(3), 259–276 (1999); with an appendix by R. Parimala
36. MacDonald, M.L.: Cohomological invariants of odd degree Jordan algebras. Math. Proc.
Camb. Philos. Soc. 145(2), 295–303 (2008)
37. Mammone, P., Shapiro, D.B.: The Albert quadratic form for an algebra of degree four. Proc.
Am. Math. Soc. 105(3), 525–530 (1989)
38. Masquelein, A., Quéguiner-Mathieu, A., Tignol, J.P.: Quadratic forms of dimension 8 with
trivial discriminant and Clifford algebra of index 4. Arch. Math. (Basel) 93(2), 129–138 (2009)
39. Merkurjev, A.S.: Simple algebras and quadratic forms. Izv. Akad. Nauk SSSR Ser. Mat. 55(1),
218–224 (1991)
40. Merkurjev, A.S., Parimala, R., Tignol, J.P.: Invariants of quasitrivial tori and the Rost invariant.
Algebra i Analiz 14(5), 110–151 (2002)
41. Parimala, R., Sridharan, R., Suresh, V.: Hermitian analogue of a theorem of Springer. J. Alge-
bra 243(2), 780–789 (2001)
42. Pfister, A.: On the Milnor conjectures: history, influence, applications. Jahresber. Deutsch.
Math.-Verein. 102(1), 15–41 (2000)
Cohomological Invariants of Central Simple Algebras with Involution 171
43. Quéguiner, A.: Signature des involutions de deuxième espèce. Arch. Math. (Basel) 65(5),
408–412 (1995)
44. : Cohomological invariants of algebras with involution. J. Algebra 194(1), 299–330
(1997)
45. Quéguiner-Mathieu, A.: Invariants cohomologiques: des formes quadratiques aux algèbres à
involution. In: Théorie des nombres (Besançon, 2002). Publications mathmatiques de l’UFR
Sciences et techniques de Besançon, p. 12. University of Franche-Comté, Besançon (2002)
46. Rost, M., Serre, J.P., Tignol, J.P.: La forme trace d’une algèbre simple centrale de degré 4.
C.R. Math. Acad. Sci. Paris 342(2), 83–87 (2006)
47. Rowen, L.H.: Central simple algebras. Israel J. Math. 29(2–3), 285–301 (1978)
48. Scharlau, W.: Quadratic and Hermitian forms. Grundlehren der mathematischen
Wissenschaften, vol. 270. Springer, Berlin (1985)
49. Serre, J.P.: Cohomologie galoisienne. Lecture Notes in Mathematics, vol. 5, 5th edn. Springer,
Berlin (1994)
50. Sivatski, A.S.: Applications of Clifford algebras to involutions and quadratic forms. Commun.
Algebra 33(3), 937–951 (2005)
51. Tamagawa, T.: Representation theory and the notion of the discriminant. In: Algebraic num-
ber theory. Kyoto International Symposium on Research Institute for Mathematical Sciences,
University of Kyoto, Kyoto, 1976, pp. 219–227. Japan Society for the Promotion of Science,
Tokyo (1977)
52. Tits, J.: Formes quadratiques, groupes orthogonaux et algèbres de Clifford. Invent. Math. 5,
19–41 (1968)
53. : Représentations linéaires irréductibles d’un groupe réductif sur un corps quelconque.
J. Reine Angew. Math. 247, 196–220 (1971)
54. : Théorie des groupes. Ann. Collège France 91, 125–138 (1992) (1990/1991)
Witt Groups of Varieties and the Purity Problem
Kirill Zainoulline
Summary We provide a general algorithm used to prove purity for functors with
transfers. As a basic example we consider the Witt group of an algebraic variety.
Kirill Zainoulline
Department of Mathematics and Statistics, University of Ottawa, 585 King Edward, Ottawa, ON,
Canada
e-mail: [email protected]
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 173
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 9,
c Springer Science+Business Media, LLC 2010
174 K. Zainoulline
ϕ
V / V
b b
ϕ#
V# o V #
commutes. For the respective matrices Mb and Mb , it means that
Mb = Cϕt · Mb ·Cϕ ,
where Cϕ is the transformation matrix corresponding to ϕ .
The Witt ring. We define the orthogonal sum and the tensor product of symmetric
bilinear spaces as
(V, b) ⊥ (V , b ) := (V ⊕ V , b ⊕ b),
(V, b) ⊗ (V , b ) := (V ⊗ V , b ⊗ b).
The orthogonal sum and the tensor product respect isometries.
Let KO(k) denote the Grothendieck group of isometry classes of symmetric bi-
linear spaces with respect to the orthogonal sum. Let H be the subgroup of KO(k)
generated by the classes of hyperbolic spaces. The Witt group of k is defined to be
the quotient
W (k) := KO(k)/H.
The tensor product of spaces turns W (k) into a commutative ring with a unit 1 =
(k, id), where k is a vector space of rank 1.
Properties:
• Let l/k be a field extension. The base change induces a ring homomorphism
i∗ : W (k) → W (l). Hence, the assignment l/k → W (l) is a covariant functor
from the category of field extensions to the category of commutative rings.
• Let l/k be a finite field extension and s : l → k be a nontrivial k-linear map. Then
the composite s ◦ B, where B is a bilinear form over l, defines a bilinear form over
k and, moreover, induces a well-defined group homomorphism s∗ : W (l) → W (k)
called a Scharlau transfer.
• There is the projection formula:
q(x) = ∑ a i j xi x j with ai j ∈ k.
i, j=1,...,n
Assume we are given a symmetric bilinear space (V, b). Let Mb be the symmetric
matrix corresponding to the space (V, b) and the chosen basis {e1 , e2 , . . . , en }. We
can associate to Mb a quadratic form
q(x) = xt · Mb · x.
In the opposite direction, let q(x) be a quadratic form over k. Then we may define a
symmetric matrix Mq as
1
Mq := (ai j + a ji )
2 ij
isometry classes
isometry classes of
of nonsingular ↔
symmetric bilinear spaces
quadratic forms
which is compatible with the orthogonal sum and the tensor product of spaces.
Hence, to compute the Witt group we can use the language of quadratic forms.
The following two properties turn out to be very important for computations:
q = qan ⊥ H.
Examples.
• W (C) = Z/2 or, more generally, W (k) = Z/2 whenever k is quadratically closed,
that is, any element of k is a square. Indeed, in this case any form can be diago-
nalized to
1, 1, . . . , 1
• W (R) = Z (use the signature)
176 K. Zainoulline
Z/2[k∗ /k∗ 2 ] if p ≡ 1 mod 4
• W (F p ) =
Z/4 if p ≡ 3 mod 4
The affine case. Let R be a commutative ring with unit, 1/2 ∈ R. The previous
definition of the Witt ring perfectly works if one replaces
a field k by a ring R
a k-vector space V by a finitely generated projective R-module P
an isomorphism b by an isomorphism of R-modules b : P → P#
In this way, we obtain the definition of the Witt ring of a commutative ring R.
Let X = Spec(R) be the associated affine scheme. Then we define
W (X) := W (R).
The Euler trace. We extend the notion of the Scharlau transfer to the affine case
as follows.
Let T /S be a finite flat extension of commutative rings. Let s : T → S be an
S-linear map such that the induced map
λ : T → HomS (T, S)
which induces an S-linear map ε : T → S via ε (x) := λ (1)(x), called an Euler trace.
Observe that from ε we get back λ as λ (x)(y) := ε (xy).
The general case. Let X be a scheme with structure sheaf OX , 1/2 ∈ Γ (OX ).
A symmetric bilinear space over X is a pair (E , b), consisting of a vector bundle
E and an isomorphism b : E → E # , where E # = HomOX (E , OX ) such that b# = b
(we identify E with its double dual E ## ).
Witt Groups of Varieties and the Purity Problem 177
W (X ) := KO(X)/M.
As before, the tensor product turns W (X ) into a ring. Note that in the affine case
this definition gives the previously defined Witt ring.
Examples.
• W (AnX ) = W (X), where X is affine – Karoubi [10]
• W (Pnk ) = W (k) for n = 1 follows from the description of vector bundles over the
projective line and for n ≥ 2 – Arason [1]
• For an irreducible smooth quasiprojective complex curve C
W (C) = (Z/2)1+2g .
For real projective surfaces, see Sujatha [21]. For affine threefolds, see
Parimala [19].
where X is an integral regular scheme over k, k(X ) is its quotient field, k(x) is the
residue field at a point x of codimension 1, and ∂x is a second residue homomor-
phism which depends on the choice of a local parameter. The exactness of this
sequence at the W (k(X ))-term is called purity and is the main subject of these lec-
tures.
Purity for the Witt groups is known for the following cases:
• X is a regular integral noetherian scheme of dimension at most 2 by Colliot-
Thélène and Sansuc [8].
• X is a regular integral noetherian affine scheme of dimension 3 by Ojanguren,
Parimala, Sridharan, Suresh [15].
• X = Spec(R), where R is a local regular ring containing a field by Ojanguren and
Panin [14].
Triangular Witt groups. A vast generalization of the notion of the Witt group of
a scheme was provided by Balmer [2]. He introduced and studied the notion of the
Witt group W of a triangulated category with duality. For triangular Witt groups
the following computations were obtained:
• W (PX (E )), W (quadric) by Nenashev [12], Walter
• W (Grassmannian) by Balmer and Calmès [4]
3 Purity
Let A be a smooth algebra over an infinite field k and let K be its quotient field. Let
R = Aq be the localization of A at a prime ideal q ∈ Spec(A). Such a ring R is called
a local regular ring of geometric type.
Let F : A-Alg → Ab be a covariant functor from the category of A-algebras to the
category of Abelian groups. From now on
• the local regular ring of geometric type R
• the covariant functor F
will be the main objects of our discussion.
Witt Groups of Varieties and the Purity Problem 179
Definition 1. (cf. [6, 2.1.1]) For every prime ideal p ∈ Spec(R), consider the group
homomorphism F(Rp ) → F(K) induced by the canonical inclusion of Rp into its
quotient field K.
We say that an element α ∈ F(K) is unramified at a prime ideal p if it is in the
image of F(Rp ). We say that an element α ∈ F(K) is unramified over R if it is
unramified at every prime ideal p ∈ Spec(R) of height 1.
Definition 2. (cf. [6, 2.1.4.(b)]) We say that purity holds for the functor F over R if
every unramified element of F(K) belongs to the image of F(R) in F(K). In other
words, the following equality holds between subgroups of F(K):
im{F(Rp ) → F(K)} = im{F(R) → F(K)}.
ht p=1
The subgroup of unramified elements (the left-hand side of the equality) will be
denoted by Fnr (K).
Remark. Purity is a particular case of the following more general problem, called
the Gersten conjecture: Given a cohomology theory F = H ∗ over R, show that the
Gersten complex
0 → H ∗ (R) → H ∗ (K) → H ∗−1 (k(x)) → H ∗−2 (k(x)) → · · ·
x∈U (1) x∈U (2)
is exact.
Here H ∗ is a functor to the category of graded abelian groups which satisfies
certain axioms (homotopy invariance, excision, localization, etc.) as in [7, Sect. 5]
or [16, Sect. 2]. Observe that exactness at the H ∗ (K)-term gives purity.
For the Witt groups the Gersten conjecture was proven
• for a local regular ring of geometric type over an infinite field of characteristic
not 2 – Balmer [3] and Schmid [20]
• for a local regular ring R containing a field of characteristic not 2 – Balmer, Gille,
Panin, and Walter [5]
• The proof of Theorem 3 uses the same techniques as the original proof by
Ojanguren–Panin.
• The Witt group formally does not satisfy the condition of being a functor with
transfers. Nevertheless, since for all varieties appearing in the proof the canoni-
cal sheaves ω turn out to be trivial, the proof works after replacing transfers by
Euler traces.
Functors with transfers. Let R be a local regular ring of geometric type obtained
by localizing a smooth k-algebra A. Let F : A-Alg → Ab be a covariant functor. We
say the F is a functor with transfers if it satisfies the following axioms:
(C) (Continuity) Roughly speaking, it says that F commutes with filtered direct
limits of localizations. More precisely, for any A-algebra S essentially smooth
over k and for any multiplicative system M in S the canonical map
(a) TrRR = idR and for any finite étale R-algebras T1 and T2 the following
relation holds
(b1) Let R[t] denote a polynomial ring over R. For a finite R[t]-algebra
S such that S/(t) and S/(t − 1) are finite étale over R the following
diagram commutes:
FR (S) / FR (S/(t))
Tr
FR (S/(t − 1)) / FR (R)
Tr
Witt Groups of Varieties and the Purity Problem 181
(b2) The transfer map TrKT ⊗R K satisfies conditions (a) and (b1) above and
the following diagram induced by extension of scalars via the canoni-
cal inclusion R → K commutes:
FR (T ) / FR (T ⊗R K)
Tr Tr
FR (R) / FR (K)
is exact.
• F = coker(μ ), where μ : G → Gm is a surjective morphism of group schemes and
G is a linear algebraic group over k, char(k) = 0, which is rational as a k-variety.
• F = Het∗ (−, C ), where C is a locally constant sheaf with finite stalks of Z/n-
modules over R, (n, char(k)) = 1.
• F = coker(Nrd), where Nrd : GL1,A → Gm is the reduced norm of an Azumaya
algebra A over R.
• F = coker(Sn), where Sn : SOq → Het1 (−, μ2 ) is the spinor norm of a nonsingular
quadratic form q defined over R.
• F = coker(Nrd), where Nrd : U(A , σ ) → U(Z, σ |Z ) is the reduced norm from
the unitary group of an Azumaya algebra A with involution of the second kind
σ over R to the unitary group of its center Z.
We also have the torsion versions of the previous cases. Here d > 1 and S is an
R-algebra.
• F : S → S∗ /Nrd(AS∗ ) · (S∗)d
• F: S→ U(Z, σ )/Nrd(U(A , σ )) ·U(Z, σ )d
182 K. Zainoulline
First, we discuss two main geometric ingredients of the proof: the Geometric Pre-
sentation Lemma and the Quillen trick.
Lemma 5 (Geometric Presentation Lemma). Let R be a local essentially smooth
k-algebra with maximal ideal m over an infinite field k, and S an essentially smooth
integral k-algebra, finite over the polynomial algebra R[t]. Suppose that ε : S → R is
an R-augmentation and let I = ker ε . Assume that S/mS is smooth over the residue
field R/m at the maximal ideal ε −1 (m)/mS. Then, given a regular function f ∈ S
such that S/( f ) is finite over R, we can find a t ∈ I such that:
• S is finite over R[t ].
• There is an ideal J comaximal with I such that I ∩ J = (t ).
• The ideals J and (t − 1) are comaximal with the ideal ( f ).
• S/(t ) and S/(t − 1) are étale over R.
Proof. We will only show how to construct this t and establish the first property
and the part of the last.
Replacing t by t − ε (t) we may assume that t ∈ I. We denote by “bar” the re-
duction modulo m. By the assumptions made on S the quotient S̄ = S/mS is smooth
over the residue field R̄ = R/m at its maximal ideal I¯ = ε −1 (m)/mS.
Choose an α ∈ S such that ᾱ is a local parameter of the localization of S̄ at I.
¯ By
the Chinese Remainder theorem we may assume that ᾱ does not vanish at the zeros
of f¯ different from I.
¯
Without changing ᾱ we may replace α by α − ε (α ) and assume that α ∈ I. Since
S is integral over R[t], there exists a relation of integral dependence
For any r ∈ k× and any N larger than the degree of each pi (t) we put
t = α − rt N .
By the equation (∗), t is integral over R[t ]. Hence S, which is integral over R[t], is
integral over R[t ] and the first property is proven.
By Bertini’s theorem we may choose α such that the algebra S̄/(ᾱ ) is étale over
R̄. Consider the fiber product diagram
u →r /k
k[u]
Since the fiber of π at u = 0 is étale, the fibers of π at almost all rational points
u ∈ k× are étale. In other words, S̄/(t¯ ) is étale over k and, hence, over R̄ for almost
all r ∈ k× .
Witt Groups of Varieties and the Purity Problem 183
By assumption S and R[t ] are smooth. Since S is finite over R[t ], S is finitely
generated projective as an R[t ]-module and, hence, S/(t ) is free as an R-module.
In particular, S/(t ) is flat over R. Finally, the fact that S̄/(t¯ ) is étale over R̄ implies
that S/(t ) is étale over R.
a →a⊗1 / A ⊗W R
AO O
r →1⊗r
i /R
W
ϕ : F(A f ) → F((A ⊗W R) f )
ψ : F(S f ) → F(R)
be the map defined by ψ = Tr1 ◦ p∗1 − TrJ ◦ p∗J , where p1 : S f → S/(t − 1), pJ : S f →
S/J are the quotient maps and Tr1 , TrJ are the respective transfers.
This is a key point of the proof: There is no map F(S f ) → F(R) which comes
from a regular one, however, by the Geometric Presentation Lemma one can define
its substitute ψ using transfers.
Step 3. Consider the commutative diagram
ϕ i∗K
F(A f ) / F(S f ) / F((S ⊗R K) f )
ψ ψK
i∗K
F(R) / F(K)
where the square is obtained by the base change and the maps ϕ and ψ were defined
before. By commutativity, the image of α = ψ (ϕ (α f )) in F(K) coincides with
ψK (β ), where β = i∗K (ϕ (α f )). By the homotopy invariance (b2) we obtain that
and by the additivity (a) the latter coincides with the image of the augmentation map
εK∗ (β ). Here εK : (S ⊗R K) f → K is given by the usual multiplication.
On the other hand, one can show that εK (β ) = εK (i∗K (ϕ (α f ))) = α . Therefore, α
is the image of α by means of the canonical map i∗K , and the proof is finished.
References
1. Arason, J.Kr. Der Wittring projektiver Räume. Math. Ann. 253 (1980), 205–212
2. Balmer, P. Derived Witt groups of a scheme. J. Pure Appl. Algebra 141(2) (1999), 101–129
3. . Witt cohomology, Mayer–Vietoris, homotopy invariance and the Gersten conjecture.
K-Theory 23(1) (2001), 15–30
4. Balmer, P., Calmès, B. Witt groups of Grassmann varieties. Preprint (2008), 28pp
5. Balmer, P., Gille, S., Panin, I., Walter, C. The Gersten conjecture for Witt groups in the
equicharacteristic case. Doc. Math. 7 (2002), 203–217
6. Colliot-Thélène, J.-L. Birational invariants, purity and the Gersten conjecture, in K-Theory
and Algebraic Geometry: Connections with Quadratic Forms and Division Algebras, AMS
Summer Research Institute, Santa Barbara 1992, eds. W. Jacob and A. Rosenberg, Proceedings
of Symposia in Pure Mathematics 58, Part I (1995) 1–64
7. Colliot-Thélène, J.-L., Hoobler, R., Kahn, B. The Bloch–Ogus–Gabber theorem, in Algebraic
K-Theory, 31-94, Fields Institute Communications 16, American Mathematics Society, Prov-
idence, RI, 1997
Witt Groups of Varieties and the Purity Problem 185
24. Baeza, R. Quadratic forms over semi-local rings, Lecture Notes in Mathematics, Vol. 655,
Springer, Berlin, 1978
25. Knus, M.-A. Quadratic and Hermitian forms over rings. Grundlehren der Mathematischen
Wissenschaften, Vol. 294, Springer, Berlin, 1991
26. Knus, M.-A., Merkurjev, A., Rost, M., Tignol, J.-P. The Book of Involutions. AMS Collo-
quium Publications, Vol. 44, AMS, Providence, RI, 1998
27. Lam, T.Y. The algebraic theory of quadratic forms. Lecture Notes Series, W.A. Benjamin,
Reading, MA, 1980
28. Scharlau, W. Quadratic and Hermitian forms. Grundlehren der Mathematischen Wis-
senschaften, Vol. 270, Springer, Berlin, 1985
Part II
Invited Articles
Some Extensions and Applications
of the Eisenstein Irreducibility Criterion
Anuj Bishnoi
Department of Mathematics, Panjab University, Chandigarh-160014, India,
e-mail: [email protected]
Sudesh K. Khanduja (corresponding author)
Department of Mathematics, Panjab University, Chandigarh-160014, India,
e-mail: [email protected]
2010 Mathematics subject classification. 12E05, 12J10, 12J25.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 189
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 10,
c Springer Science+Business Media, LLC 2010
190 Anuj Bishnoi and Sudesh K. Khanduja
where G is a totally ordered (additively written with infinity greater than each ele-
ment of G) abelian group and v satisfies properties (i), (ii) and (iii) of a real valua-
tion. The pair (K, v) is called a valued field and G is called the value group of v.
In 1997, Khanduja and Saha [13] generalized the Eisenstein–Dumas–Kürschák
criterion to polynomials with coefficients from any valued field (K, v).
Theorem 1. Let v be a Krull valuation of any rank of a field K with value group G
and F(x) = a0 xs + a1 xs−1 + · · · + as be a polynomial over K. If v(a0 ) = 0, v(ai )/i ≥
v(as )/s for 1 ≤ i ≤ s and there does not exit any integer d > 1 dividing s such that
v(as )/d ∈ G, then F(x) is irreducible over K.
The following criterion which is more general than the Eisenstein Irreducibility
Criterion was proved by Schönemann [19] in 1846.
Some Extensions and Applications of the Eisenstein Irreducibility Criterion 191
Theorem 2. Let v be a valuation of a field K with value group the set of integers.
Let g(x) = xm + a1 xm−1 + · · · + am be a polynomial with coefficients in K such that
v(ai )/i > v(am )/m for 1 ≤ i ≤ m − 1. Let r denote gcd(v(am ), m) and b be an el-
ement of K with v(b) = v(am )/r. Suppose that the polynomial zr + (am /br ) in the
indeterminate z is irreducible over the residue field of v. Then g(x) is irreducible
over K.
deg Fi (y) d
< for 1 ≤ i ≤ n − 1.
i n
If r denotes gcd(d, n) and if the polynomial zr + c0 /c is irreducible over K, then so
is F(x, y).
Proof. Let
P(x, y) = a0 F1 (x, y) · · · Fs (x, y) (1)
be a factorization of P(x, y) as a product of irreducible factors over K. Let x be given
the weight n and y the weight m. Let Hi (x, y) denote the sum of those monomials
occurring in Fi (x, y) which have the highest weight, then Hi (x, y) is a nonconstant
polynomial. Denote b0 /a0 by −b. Keeping in mind the definition of a generalized
difference polynomial and comparing the terms of highest weight on both sides of
(1), it can be easily seen that:
Let ξ be a primitive r1 -th root of unity, where r = pt r1 and p does not divide r1 .
The integers m/r and n/r will be denoted by m1 and n1 , respectively. Choose c
belonging to the algebraic closure K of K such that cr = b. Rewrite (2) as:
r
∏(xm1 − cξ iyn1 ) = H1 (x, y) · · · Hs (x, y). (3)
i=1
The desired result follows from (3) because a polynomial of the type x j − ayk is
when j and k are co-prime.
irreducible over K
For polynomials with rational coefficients, the following theorem gives a better
estimate than the one given by Theorem 6.
Proof. Let b = −b0 /a0 be as in the proof of Theorem 6. Fix a complex number
c satisfying cr = b and a primitive d-th root of unity to be denoted by ζd for any
number d dividing r. Define a polynomial Pd (x) with complex coefficients by:
Pd (x) = ∏ (x − cζdi ).
1≤i≤d,(i,d)=1
Then
xr − b = ∏ Pd (x). (4)
d|r
Corollary. Let f (x) and h(y) be polynomials having coefficients in Q. Assume that
deg f (x) is a prime number. Then f (x) − f (h(y)) is a product of two irreducible
factors over Q. In particular the polynomial f (x) − f (y)/x − y is irreducible over
Q.
The above corollary sheds some light on a problem posed by Cassels [6] which
asks for what polynomials f is the polynomial f (x) − f (y)/x − y reducible over Q.
Definition. A polynomial with coefficients from a valued field (K, v) which satis-
fies the hypothesis of Theorem 1 is called an Eisenstein–Dumas polynomial with
respect to v. A polynomial satisfying the hypothesis of Generalized Schönemann Ir-
reducibility Criterion (stated before Theorem 2) will be referred to as a Generalized
Schönemann polynomial with respect to v and f (x).
Some Extensions and Applications of the Eisenstein Irreducibility Criterion 195
Definition. Let v be a henselian valuation of a field K and let ṽ be the unique prolon-
of K. A pair (θ , α ) of elements of K
gation of v to the algebraic closure K is called a
distinguished pair (more precisely a (K, v)-distinguished pair) if the following three
conditions are satisfied:
(i) [K(θ ) : K] > [K(α ) : K],
with [K(β ) : K] < [K(θ ) : K], and
(ii) ṽ(θ − β ) ṽ(θ − α ) for every β in K
(iii) whenever β belonging to K is such that [K(β ) : K] < [K(α ) : K], then
ṽ(θ − β ) < ṽ(θ − α ).
Properties of distinguished pairs have been studied extensively in [17] and [1].
The following result, which generalizes a result of Juras [12] proved in 2006, has
been quickly deduced from the earlier theorem.
Notation. For α separable over K of degree > 1, ωK (α ) will stand for the Krasner’s
constant defined by:
ωK (α ) = max ṽ(α − α ) | α = α runs over K-conjugates of α .
196 Anuj Bishnoi and Sudesh K. Khanduja
Example. Let K be the field of 2-adic numbers with the usual valuation v2 given
by v2 (2) = 1. The prolongation of v2 to the algebraic closure of K will be denoted
by v2 again. Consider θ = 2 + 2(2−1/2) + 22 (2−1/2 ) and θ1 = 2 + 2(2−1/2). It will
2
be shown that K(θ ) = K(21/4 ) and (θ , θ1 ) is a distinguished pair. Note that the
Krasner’s constant ωK (θ1 ) = 3/2 and v2 (θ − θ1 ) = 7/4 > ωK (θ1 ). Therefore, by
Krasner’s Lemma [10, Theorem 4.1.7], K(θ1 ) ⊆ K(θ ) and hence 22 (2−1/4 ) = θ − θ1
belongs to K(θ ) as asserted. To show that (θ , θ1 ) is a distinguished pair, we first
verify that whenever γ belonging to K satisfies v2 (θ − γ ) > v2 (θ − θ1 ) = 7/4, then
deg γ 4. If γ is as above, we have by the strong triangle law
Also for any b ∈ K with deg b < deg θ1 , we have b ∈ K and clearly v2 (θ1 − b)
1/2 < v2 (θ − θ1 ). So (θ , θ1 ) is a distinguished pair. As can be easily checked, θ
is a root of g(x) = x4 − 8x3 + 20x2 − 80x + 4 which must be irreducible over K.
By Theorem 8, no translate of g(x) can be an Eisenstein–Dumas polynomial with
respect to v2 because (θ , θ1 ) is a distinguished pair with θ1 ∈
/ K and consequently
(θ , a) cannot be a distinguished pair for any a in K. Moreover, if p is a prime
number different from 2, then no translate of g(x) can be an Eisenstein–Dumas poly-
nomial with respect to the p-adic valuation v p , for otherwise in view of Theorem 9,
g(x + 2) = x4 − 4x2 − 64x − 124 would be an Eisenstein–Dumas polynomial with
respect to v p , which is not so.
Acknowledgements The first author gratefully acknowledges the financial support of CSIR (grant
no. 09/135(0539)/2008-EMR-I).
References
1. K. Aghigh and S. K. Khanduja, On chains associated with elements algebraic over a henselian
valued field, Algebra Colloq. 12:4 (2005) 607–616.
2. V. Alexandru, N. Popescu and A. Zaharescu, A theorem of characterizaton of residual tran-
scendental extension of a valuation, J. Math. Kyoto Univ. 28 (1988) 579–592.
3. S. Bhatia and S. K. Khanduja, Difference polynomials and their generalizations, Mathematika
48 (2001) 293–299.
4. A. Bishnoi and S. K. Khanduja, On Eisenstein-Dumas and generalized Schönemann polyno-
mials, Comm. Algebra (2010).
5. R. Brown, Schönemann Polynomials in Henselian extension fields, Indian J. Pure Appl. Math.
39(5) (2008) 403–410.
Some Extensions and Applications of the Eisenstein Irreducibility Criterion 197
Vladimir Chernousov
Summary We show that the kernel of the Rost invariant for a split group of type E8
is trivial modulo 3.
1 Introduction
Let F be a field and let G be a simple simply connected algebraic group over F.
With these data one can associate two natural functors f , g : Fields → Sets, where
Fields denotes the category of field extensions of F and Sets denotes the category
of sets, given by f (K/F) = H 1 (K, G) and g (K/F) = H 3 (K, Q/Z(2)).
In the 1990s, Rost constructed a natural transformation between f and g which
nowadays is called the Rost invariant of G and it is denoted by RG . Thus for every
field extension K/F we have a natural (functorial) mapping of pointed sets RG,K :
H 1 (K, G) → H 3 (K, Q/Z(2)). By abuse of language, instead of RG,K we will use
the abbreviated notation RG whenever there is no danger of confusion.
For applications it is important to describe the kernel of RG . If G is split or quasi-
split over F and has small rank then the kernel is trivial by [Gar] (see also [Ch03]).
In particular, Ker RG = 1 for all groups G of exceptional types but E8 , i.e., for
every field extension K/F we have Ker RG,K = 1. For type E8 the situation is much
more difficult and not much information is available. For instance, we know that
the kernel of the Rost invariant is not trivial. Indeed, if F = Q and K = R then
one checks that the class [ξ ] ∈ H 1 (K, G) corresponding to the compact form of E8
belongs to the kernel of the Rost invariant.
Vladimir Chernousov
Department of Mathematics, University of Alberta, Edmonton, Alberta T6G 2G1, Canada,
e-mail: [email protected]
2010 Mathematics subject classification. 20G07, 20G10, 20G15, 20G41.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 199
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 11,
c Springer Science+Business Media, LLC 2010
200 Vladimir Chernousov
In this paper, we address the (slightly simpler) problem of describing the kernel
of RG modulo a prime number p where G is a split group over F of type E8 . We say
that a class [ξ ] ∈ H 1 (K, G) is in the kernel of RG modulo p if there is a finite field
extension L/K of degree prime to p such that RG,L ([ξL ]) = 1. We also say that the
kernel of the Rost invariant of G is trivial modulo p if every class [ξ ] ∈ H 1 (K, G)
which is in the kernel of RG modulo p can be killed by a finite field extension L/K
of degree prime to p.
Clearly, for all p = 2, 3, 5 the kernel of the Rost invariant of a split group of
type E8 is trivial modulo p because every torsor over K of type E8 can be killed
by a finite extension of K of degree dividing 26 32 5 by [To] (for a simpler proof see
also [Ch06]). If p = 5 then the main result in [Ch94] says that the kernel of the Rost
invariant for E8 modulo 5 is trivial. The aim of the paper is to consider the next case
p = 3.
Theorem 1.1. Let F be a field of characteristic = 3 and let G0 be the split group of
type E8 over F. Then the kernel of the Rost invariant for G0 is trivial modulo 3.
Assume, additionally, that G is a simple simply connected group split over F and
that T ⊂ G is a split maximal torus. Let g be the Lie algebra of G. A choice of
Borel subgroup B ⊂ G containing T determines an ordering of Σ , hence the system
On the Kernel of the Rost Invariant for E8 Modulo 3 201
{Hα1 , . . . , Hαn , Xα , α ∈ Σ }
in g corresponding to the pair (T, B). This basis is unique up to signs and automor-
phisms of g which preserve B and T (see [St67], Sect. 1, Remark 1).
Let T and T∗ denote the corresponding group of characters and cocharacters of T .
There exists a natural pairing T∗ × T → Z. We may view the root system Σ as
a subset in T. For each root α ∈ Σ Steinberg in his book [St67] associates the
cocharacter in T∗ denoted by hα . Thus, hα is a homomorphism hα : Gm → T . For
instance, if G = SLn and α = εi − ε j then hα (t) is a diagonal matrix with t in the
ith column and t −1 in the jth. The image of hα coincides with Tα = T ∩ Gα . If all
roots in Σ have the same length then for every sum α = β + γ , where α , β , and γ
are roots we have hα = hβ hγ .
As G is a simply connected group, the following relations hold in G (cf. [St67],
Lemma 28 b), Lemma 20 c) ):
(A) T = Tα1 × · · · × Tαn ; hence T∗ is generated by hα1 , . . . , hαn .
(B) For any two roots α , β ∈ Σ we have
hα (t) Xβ hα (t)−1 = t β ,α Xβ ,
where β , α = 2 (β , α )/(α , α ).
If Δ ⊂ Π , we let GΔ denote the subgroup generated by U±α , α ∈ Δ .
We state now two theorems which are due to Steinberg [St65]. Although they
are not formulated explicitly in [St65], their proofs can be easily obtained from
the arguments contained in [St65], Sect. 10 (see also [PR, Propositions 6.18, 6.19,
p. 338–339] and [Ch03, Theorems 3.1, 3.2]).
Let G0 be a simple (not necessary simply connected) linear algebraic group split
or quasi-split over F. Let ξ ∈ Z 1 (F, G0 ) be a cocycle and let G = ξ G0 be the
corresponding twisted group.
Theorem 2.1. For any maximal torus S ⊂ G over F there is an F-embedding
S → G0 such that the class [ξ ] lies in the image of H 1 (F, S) → H 1 (F, G0 ).
Theorem 2.2. In the notation earlier assume that G0 is a simple simply connected
group split over F and that G is isotropic over F. Then ξ is equivalent to a cocycle
with coefficients in a proper semisimple simply connected F-split subgroup H0 in
G0 which is standard and isomorphic over F s to a semisimple anisotropic kernel
of G.
202 Vladimir Chernousov
as required.
In this section for later use we state (without proofs) two properties of the Rost
invariant. For proofs we refer to [KMRT] (Sect. 31), [GMS]. We note that an explicit
formula for RG is known in a few cases only.
Let A be a central simple algebra over F of degree p, where p is prime and let
G = SL(1, A). One has:
Assume that char(F) = p. Then the formula for the Rost invariant RG is given (up
to a sign) by
Let G, H be almost simple simply connected linear algebraic groups over F and let
ρ : H → G be an F-embedding. The restriction RG at ρ (H) gives a cohomological
invariant of H, hence is equal to nρ RH for a positive integer nρ (see [KMRT],
Sect. 31). The smallest such integer is called the Rost multiplier of the embedding ρ .
Proof. Let Σ1 ⊂ Σ (G, T ) be a subsystem such that ρ (H) = GΣ1 . Let S be the con-
nected component of T ∩ ρ (H). Clearly S is a maximal torus in ρ (H) and there is a
natural embedding of the root systems
Σ (ρ (H), S) → Σ (G, T ).
Hence the coroots of ρ (H) are also the coroots of G (we used the fact that all roots
of the root system Σ = Σ (G, T ) have the same length). The rest of the proof follows
from [GMS, Proposition 7.9, p. 122].
In what follows we assume char (F) = 3 and G0 is the split group of type E8 over F.
Let T0 ⊂ G0 be a maximal split torus over F. For the proof of Theorem 1.1 we may
assume without loss of generality that F has no finite extension of degree prime
to 3. In particular, this implies that F is perfect (because char F = 3) and F contains
a primitive cube root of unity.
Now consider a cocycle ξ ∈ Z 1 (F, G0 ) that is in the kernel of the Rost invariant
RG0 ,F . Take a finite extension L/F of minimal possible degree killing ξ . As F has
no finite extensions of degree prime to 3 we have [L : F] = 3n . Consider a chain
F = L0 ⊂ L1 ⊂ · · · ⊂ Ln = L
Proof. Let H0 be the subgroup in G0 given by Theorem 2.2. We may assume that
ξ takes values in H0 . Since H0 is proper, rank of H0 is at most 7. Note that H0
is simply connected since so is G0 . By Proposition 3.3, we know that RH0 ([ξ ]) =
RG0 ([ξ ]) = 1. Furthermore for all split or quasi-split F-groups of rank ≤7 the Rost
invariant has trivial kernel by [Gar], hence [ξ ] = 1.
Thus, for the proof of Theorem 1.1 we may assume that ξ is F-anisotropic, i.e.,
the twisted group G = ξ G0 is F-anisotropic. To complete the reduction to special
cocycles we need the following
Proposition 4.2. Let θ ∈ Z 1 (F, G0 ) be a cocycle such that the twisted group θ G0 is
F-anisotropic and split over a cubic Galois extension L/F. Then G = θ G0 contains
a maximal F-torus S ⊂ G which is split over L.
Proof. We apply the same argument as in [Ch03], Proposition 5.2. Namely, let
P ⊂ G be a maximal parabolic subgroup over L corresponding to the root α8 . It
has dimension 191. Now consider the connected component C of the intersection
P ∩ σ (P) ∩ σ 2 (P). As F is perfect and G is F-anisotropic, C is a reductive group
over F of dimension at least 77. Let C(1) = [C,C] = C1 · · ·Cs be the decomposition
of the semisimple part of C into an almost direct product of the simple components
over F s . It easily follows from dimC ≥ 77 that C(1) contains a simple component,
say C1 , of type not An . As in [Ch03] we may additionally assume that C1 is defined
over F. Because F has no quadratic extensions, C1 is not of type Bn , Cn , G2 . By a
dimension argument, it also cannot be of type Dn or F4 . So only one of the following
cases can occur.
Case: C1 is of type E7 . From C1 ⊂ C ⊂ P we conclude that C1 is the semisimple
part of a Levi subgroup of P. In particular, CG (C1 ) contains a 1-dimensional torus
which is split over L (the center of the Levi subgroup). Thus, by dimension con-
sideration CG (C1 ) is an F-defined reductive group of rank 1. But any such group is
split over F. This contradicts our assumption that G is F-anisotropic.
Case: C1 is a group of type E6 . As earlier we find that CG (C1 ) is a reductive
group of rank 2 which is isotropic over L and hence split over L. By classification
results, every such group contains a maximal torus, say S1 , defined over F and
splitting over L. Now consider the reductive group CG (S1 ). Its semisimple part has
rank at most 6 and contains C1 . Hence its semisimple part is precisely C1 . As S1 is
split over L so is C1 . By [Ch03], C1 contains a maximal F-defined torus S2 splitting
over L. Then the torus S = S1 · S2 is maximal and splits over L.
On the Kernel of the Rost Invariant for E8 Modulo 3 205
We first recall an explicit description of a 3-Sylow subgroup of the Weyl group WE6 .
For the proof we refer to [Ch03]. Let ΣE6 be a root system of type E6 . Fix a basis
Π = { α1 , . . . , α6 } of ΣE6 and denote by β the highest root. The extended Dynkin
diagram of type E6 is of the form
r −β
r α2
r r r r r
α1 α3 α4 α5 α6
It follows from the picture that ΣE6 contains a root subsystem Σ1 of type A2 ×A2 ×A2
generated by roots {α1 , α3 }, {α5 , α6 }, and {α2 , −β }. Since the Weyl group WA2 of
type A2 is isomorphic to S3 it contains a unique 3-Sylow subgroup which is cyclic
of order 3. Let v1 , v2 , v3 be elements in WΣ1 S3 × S3 × S3 given by:
v v v v v v
α3 −→
1
α1 −→
1
−(α1 + α3 ) −→
1
α3 , α5 −→
2
α6 −→
2
−(α5 + α6 ) −→
1
α5 ,
v v v
and − β −→
3
α2 −→
3
β − α2 −→
3
−β . (5.1)
It is a central subgroup of order 3 of the following shape. It is easy to see that the
images of the embeddings
and
μ3 → G{α5 ,α6 } , x → hα5 (x) hα6 (x2 )
are the centers of the groups in question.
Lemma 5.3. The image of ϕ : μ3 → G{−β ,α2 } × G{α5 ,α6 } given by x → (x, x2 ), i.e.,
ϕ (x) = h−β (x)hα2 (x2 )hα5 (x2 )hα6 (x) = hα1 (x2 )hα3 (x),
r α2
r r r r r r r r
α1 α3 α4 α5 α6 α7 α8 −α
Here α is the highest root. From the picture we conclude that ΣE8 contains a root
subsystem Σ2 of type E6 × A2 generated by roots {α1 , . . . , α6 } and {α8 , −α }. Recall
that the order of a 3-Sylow subgroup WE8 (3) is 35 . As WΣ2 (3) = WE6 (3) × WA2 (3)
has the same order we obtain
Lemma 5.4. One has
WE8 (3) WE6 (3) × Z/3 ((Z/3 × Z/3 × Z/3) Z/3) × Z/3.
Finally, we describe a 3-Sylow subgroup of WE8 . Consider the Weyl group WA2
corresponding to the roots α8 , −α . Let v4 ∈ WA2 S3 be the element given by:
v v v
α8 −→
4
−α −→
4
α − α8 −→
4
α8 , (5.5)
From the earlier argument we know that the subgroup in WE8 generated by v1 , v2 , v3 ,
v4 and v is a 3-Sylow subgroup of WE8 . We will call it a standard 3-Sylow subgroup.
On the Kernel of the Rost Invariant for E8 Modulo 3 207
in gL = g ⊗F L. Recall that
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
010 000 000
Xα1 = ⎝ 0 0 0 ⎠ , Xα2 = ⎝ 0 0 1 ⎠ , and X−(α1 +α2 ) = ⎝ 0 0 0 ⎠ .
000 000 100
7 Construction of ξ1
where ∼ is an equivalence relation given by conjugation. The map NG0 (T0 )(L) →
WE8 = WE8 (L) is surjective. So passing to an equivalent cocycle, if necessary, we
may assume that the image ξ of ξ corresponds to a mapping ξ : Gal (L/F) → WE8
On the Kernel of the Rost Invariant for E8 Modulo 3 209
Consider a cocycle η
= (aσ ) ∈ Z 1 (L/F, N
(L)) where aσ = u1 u2 u3 u4 and its
image η in Z 1 (L/F, NG0 (T0 )). As η
is trivial so is η . Twisting G0 by η gives rise
to a canonical bijection λ : H 1 (F, G0 ) → H 1 (F, η G0 ) (see [Se]). As by construction
ξ and η have the same image under H 1 (L/F, NG0 (T0 )(L)) → H 1 (L/F,WE8 ), the
image ξ1 = λ (ξ ) takes values in the twisted torus S1 = η T0 . Note that because η
is trivial we may identify η G0 = G0 and this allows us to view S1 as a maximal
F-anisotropic torus in G0 splitting by L.
We have thus proved that there is a maximal F-defined torus S1 ⊂ G0 splitting
over L such that ξ is equivalent to a cocycle ξ1 with coefficients in S1 . By our
We keep the above notation. In particular, we deal with the maximal torus S1 ⊂ G0
and the cocycle ξ1 ∈ Z 1 (L/F, S1 (L)) constructed in the previous section. Let Σ1 =
Σ (G0 , S1 ). Abusing notation we still denote its basis by α1 , . . . , α8 . As we saw
210 Vladimir Chernousov
earlier S1 is F-anisotropic and hence the lattice S1 contains no F-defined characters.
This implies in turn that for every root α ∈ Σ1 we have α + σ (α ) + σ 2 (α ) = 0.
Thus the σ -orbit
o(α ) = { α , σ (α ), σ 2 (α ) } ⊂ Σ1
of α generates a root subsystem Σo(α ) ⊂ Σ1 of rank at most 2. Because σ is an
automorphism of Σo(α ) of order 3 it immediately follows that Σo(α ) is a root system
of type A2 with basis α , σ (α ).
As usual we denote by:
{ Hα1 , . . . , Hα8 , Xα , α ∈ Σ1 }
and
σ (α7 ) = α1 + 2α3 + 2α2 + 3α4 + 3α5 + 2α6 + α7 + α8 .
Proof. We know that σ (β ) = −α2 and that β = α1 + 2α2 + 2α3 + 3α4 + 2α5 + α6 .
It follows that
Substituting
is an F-isomorphism.
h α 4 , h σ (α 4 ) , h α 7 + α 6 , h σ (α 7 + α 6 ) , h α 2 , h σ (α 2 ) , h α 5 , h σ (α 5 ) ,
9 Construction of ξ2
We start with the following general formalism. Let G be an algebraic group (not
necessary split) over F and let S ⊂ G be a maximal F-torus splitting over a finite
Galois extension L/F. Assume that we are given two cocycles η1 = (aσ ) and η2 =
(bσ ) with coefficients in S(L) such that η2 viewed in Z 1 (L/F, G(L)) is trivial, say
bσ = g1−σ for some element g ∈ G(L).
Let now η = η1 η2 . As η1 , η2 take values in a commutative group S(L), η is a co-
cycle. Consider an inner conjugation int(g): G → G given by x → g−1 xg. We claim
that the torus S1 = g−1 Sg is an F-defined maximal torus in G and the restriction of
int(g) to S is an F-defined isomorphism S → S1 . Indeed, if s ∈ S(L) then we have
(We used the fact that g1−σ ∈ S(L) implies that int(g) is L-defined on S.)
Now take the cocycle
η1
= (g−1 aσ bσ gσ ) = g−1 aσ g,
which is equivalent to η and takes values in S1 . We say that the passage from η to
η1
is an untwisting of η , and we call the torus S1 and the cocycle η1
the syndrome
of S and η with respect to g.
Lastly we note that if η1 takes values in a subtorus S
⊂ S which is a maximal
torus in a standard subgroup S
⊂ H ⊂ G over F then the syndrome of η can be
viewed as a cocycle with coefficients in an F-defined H1 = g−1 Hg subgroup in G
which will be called a syndrome of H.
212 Vladimir Chernousov
Remark 9.1. (1) The restriction of int(g) at H is not necessarily F-defined and the
syndrome H1 of H is not necessary isomorphic to H over F.
(2) Even if both η1 , η2 are trivial when viewed as cocycles in Z 1 (F, G), the cocycle
η is not necessarily trivial in G.
respectively.
Using Lemma 8.3 one easily checks that the roots in o(α7 + α6 ) and o(α4 ) are or-
thogonal to each other, in particular Go(α7 +α6 ) and Go(α4 ) are commuting subgroups
in G0 each of type A2 and their intersection is trivial. Hence the cocycle η2 = θ3 θ4
can be viewed as a cocycle in Z 1 (L/F, G0,o(α7+α6 ) × G0,o(α4) ).
By Corollary 6.3, θ3 and θ4 can be written in the form θ3 = (aσ ), θ4 = (bσ )
where aσ = hα7 +α6 (u3 ) and bσ = hα4 (u4 ) for some parameters u3 , u4 ∈ F × . By
Lemma 8.1, Go(α7 +α6 ) × Go(α4 ) is isomorphic to SL3 × SL3 . Hence there is g ∈
(G0,o(α7 +α6 ) × G0,o(α4 ) )(L) such that θ3 θ4 = (g1−σ ). We now denote by ξ2 the
syndrome of ξ1 with respect to g.
α , H
{H α , ±Xα , ±Xα , ±Xα +α },
5 6 5 6 5 6
where
α = g−1 Hα g,
H α = g−1 Hα g,
H
5 5 6 6
and
Xα5 = g−1 Xα5 g, Xα6 = g−1 Xα6 g, Xα5 +α6 = g−1 Hα5 +α6 g.
On the Kernel of the Rost Invariant for E8 Modulo 3 213
It follows that the action of σ ∈ Gal (L/F) on Xα5 , Xα6 is the composition of the
old action given by (8.2) and the twisting action by g1−σ = hα7 +α6 (u3 )hα4 (u4 ). For
instance,
σ (Xα5 ) = g−1 (g1−σ ε Xα6 gσ −1 )g = ε u3 Xα6
where ε = ±1. Computing the second structure constant of T1 we see that T1 =
(a, u3 u24 ). The argument for T2 is similar.
Proof of (b) from §4: By the above lemma the cocycle ξ2 lives in SL (1, T1 ) ×
SL (1, T1 ), hence it can be presented by a pair of elements
Since ξ2 is in the kernel of the Rost invariant, by (3.1) and (3.2) we have
RG0 ([ξ2 ]) = a1 ∪ T1 + a2 ∪ T1 = 0.
Acknowledgements The author was partially supported by the Canada Research Chairs Program
and an NSERC research grant.
References
[Ch94] V. Chernousov, Remark on (mod 5)–Serre’s invariant for groups of type E8 , Math.
Zametki, 56 (1994), no. 1, 116–121.
[Ch03] , The kernel of the Rost invariant, Serre’s Conjecture II and the Hasse principle
for quasi-split groups 3,6 D4 , E6 , E7 , Math. Annalen 326 (2003), 297– 330.
[Ch06] , Another proof of Totaro’s theorem on E8 -torsors, Can. Math. Bull 49 (2006),
196–202.
[Gar] R. S. Garibaldi, The Rost invariant has trivial kernel for quasi-split groups of low rank,
Comment. Math. Helvetici 76 (2001), 684–711.
[GMS] S. Garibaldi, A. Merkurjev, and J.-P. Serre, Cohomological invariants in Galois coho-
mology, University Lecture Series, vol. 28, Amer. Math. Soc., 2003.
[GS] P. Gille, T. Szamuely, Central simple algebras and Galois cohomology, Cambridge stud-
ies in advanced mathematics, vol. 101, 2006.
[KMRT] M. Knus, A. Merkurjev, M. Rost, J–P. Tignol, The Book of Involutions, 44 Colloquium
publications: American Mathematical Society 1998.
[MS] A. Merkurjev, A. Suslin, K-cohomology of Severi-Brauer varieties and the norm residue
homomorphism, Izv. Akad. Nauk SSSR 46 (1982), 1011–1046.
214 Vladimir Chernousov
[PR] V. Platonov, A. Rapinchuk, Algebraic Groups and Number Theory, Academic Press,
London 1994.
[Se] J.-P. Serre, Galois Cohomology, Springer, Berlin 1997.
[St65] R. Steinberg, Regular elements of semisimple groups, Publ. Math. I.H.E.S. 25 (1965),
281–312.
[St67] , Lectures on Chevalley Groups, Yale University 1967.
[To] B. Totaro, Splitting fields for E8 -torsors, Duke Math. J 121 (2004), 425–455.
Une version du théorème d’Amer et Brumer
pour les zéro-cycles
À Parimala
Résumé (anglais). M. Amer and A. Brumer have shown that, for two homogeneous
quadratic forms f and g over a field k, the locus f = g = 0 has a non-trivial solution
over k if and only if, for a variable t, the equation f +tg = 0 has a non-trivial solution
over k(t). We consider a modified version of this result. We show that the projective
variety over k defined by f0 = · · · = fr = 0, where the fi are homogeneous forms
over k of the same degree d ≥ 2 in n + 1 variables (with n + 1 ≥ r + 2), has a 0-cycle
of degree 1 over k if and only if the generic hypersurface f0 + t1 f1 + · · · + tr fr = 0
has a 0-cycle of degree 1 over k(t1 , . . . ,tr ).
1 Introduction
Jean-Louis Colliot-Thélène
C.N.R.S., Mathématiques, Bâtiment 425, Université Paris-Sud, F-91405 Orsay, France
e-mail: [email protected]
Marc Levine
Department of Mathematics, Northeastern University, 360 Huntington Avenue, Boston,
MA 02115, USA and Universität Duisburg-Essen Fakultät Mathematik, Campus Essen,
D-45117 Essen, Germany; e-mail: [email protected]
Classification thématique AMS. 14C25
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 215
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 12,
c Springer Science+Business Media, LLC 2010
216 Jean-Louis Colliot-Thélène et Marc Levine
de deux formes de même degré d (Théorème 1), et des versions pour un système
quelconque de formes de même degré d (Théorèmes 2 et 3).
Pour K/k une extension de corps et X un k-schéma, on note XK = X ×k K. Une
variété algébrique sur un corps k est un k-schéma séparé de type fini.
Preuve. Seul le point (c) requiert une explication. Soit K = k(t). Si L/k est une
extension finie de corps avec X (L) = 0/ alors Xk(t) (L(t)) = 0./ Ainsi I(XK /K) divise
I(X/k). Soit P un point fermé de XK de degré n. On a donc une extension finie de
corps L/K et une K-immersion fermée SpecL → XK . Le corps L est le corps des
fonctions d’une k-courbe normale C, finie sur Spec k[t]. Il existe un ouvert non vide
U ⊂ Spec k[t] tel que la restriction CU /U soit finie de degré n, et qu’il existe une
U-immersion fermée CU → X ×k U. Si le corps k est infini, on choisit un k-point
R ∈ U(k). La fibre de CU /U au-dessus de ce k-point est le spectre d’une k-algèbre
finie de degré n qui admet une k-immersion dans X . Une telle situation définit un
zéro-cycle effectif de degré n sur la k-variété X (voir [5, Appendices A1, A2, A3]).
Si le corps k est fini, il existe un zéro-cycle ∑ ni Ri tel que tous les points fermés Ri
soient dans U ⊂ Spec k[t] et que ∑i ni [k(Ri ) : k] = 1. A chaque Ri on associe par la
méthode ci-dessus un zéro-cycle effectif zi de degré n[k(Ri ) : k] sur la k-variété X .
Le zéro-cycle ∑i ni .zi est alors de degré n sur la k-variété X . Ainsi I(X /k) divise
I(XK /K). L’énoncé sur les indices réduits résulte immédiatement de celui sur les
indices.
Une version du théorème d’Amer et Brumer pour les zéro-cycles 217
λ f + μ g = 0.
Notons
∼
=
i : CH1 (Z) → CH1 (P1 × Pn ) → CH1 (P1 ) ⊕ CH1 (Pn ) = Z ⊕ Z
Ired (W /K) = pgcd(Ired (W1 /K), Ired (X2,K /K)) = pgcd(Ired (W1 /K), Ired (X2 /k))
d’après le lemme 1 et
pgcd(Ired (W1 /K), Ired (X2 /k)) = pgcd(Ired (X1 /k), Ired (X2 /k)) = Ired (X /k),
la première égalité résultant de Ired (W1 /K) = Ired (X1 /k) établi ci-dessus, la seconde
égalité provenant du lemme 1. Ceci achève la démonstration.
Preuve. Par l’argument donné au lemme 1 (c), on peut supposer le corps k infini.
Cette hypothèse sera utilisée de façon constante dans ce qui suit. Soient g0 , g1 , . . . , gr
des formes de degré d, à coefficients dans k, dont l’annulation définit une sous-k-
variété intersection complète et lisse Y ⊂ Pn . Soit X ⊂ Pn × A1 le sous-schéma
fermé défini par l’idéal (homogène en les xi )
est constante sur A1 . En particulier, le 0-cycle associé au sous-schéma fermé p−1 (0)
de X a degré d r+1 sur k.
Preuve. Pour les k-variétés, on omet l’indice k. Si L/k est une extension finie de
corps avec X(L) = 0, / alors W (L(t)) = 0.
/ Ainsi l’indice I(W /K) divise I(X /k),
et donc l’indice réduit Ired (W /K) divise l’indice réduit Ired (X /k). D’après le lemme
1 (c), pour établir l’énoncé on peut remplacer le corps k par une extension transcen-
dante pure. On supposera donc le corps k infini.
Supposons d’abord que les formes fi n’ont pas de facteur commun non constant.
Soient V = Pn \ X et I = I(X/k). Comme au théorème 1, on a CH0 (V ) = Z/I.
Soit Z ⊂ Pr × Pn la k-variété définie par la proportionalité de (λ0 , . . . , λr ) et de
220 Jean-Louis Colliot-Thélène et Marc Levine
U → U1 → V
Notons
∼
=
i : CH1 (Z) → CH1 (Pr × Pn ) → CH1 (Pr ) ⊕ CH1 (Pn ) = Z ⊕ Z
Supposons maintenant que fi = h.gi pour tout i avec les gi homogènes de même
degré sans facteur commun non constant et h homogène non constant.
Soit X ⊂ Pnk , resp. X1 ⊂ Pnk , resp. X2 ⊂ Pnk la k-variété définie par l’annulation
des fi , resp. par l’annulation des gi , resp. par h = 0. Soit W ⊂ PnK , resp. W1 /K la
variété définie par l’annulation des fi − ti f0 (i = 1, . . . , r), resp. par l’annulation des
gi − ti g0 (i = 1, . . . , r). On a
Ired (W /K) = pgcd(Ired (W1 /K), Ired (X2,K /K)) = pgcd(Ired (W1 /K), Ired (X2 /k))
d’après le lemme 1 et
pgcd(Ired (W1 /K), Ired (X2 /k)) = pgcd(Ired (X1 /k), Ired (X2 /k)) = Ired (X /k),
la première égalité résultant de Ired (W1 /K) = Ired (X1 /k) établi ci-dessus, la seconde
égalité provenant du lemme 1. Ceci achève la démonstration.
Preuve. Soit Er l’énoncé de ce théorème pour r fixé et tout corps k. L’énoncé E1 est
le théorème 1. Supposons établi Er−1 .
Supposons r ≥ 2. Soient t1 , . . . ,tr des variables indépendantes et K = k(t1 , . . . ,tr ).
D’après le théorème 2, on a Ired (X /k) = Ired (W /K), où la K-variété W ⊂ PnK est
définie par
f1 − t1 f0 = · · · = fr − tr f0 = 0.
222 Jean-Louis Colliot-Thélène et Marc Levine
( f1 − t1 f0 ) + s2 ( f2 − t2 f0 ) + · · · + sr ( fr − tr f0 ) = 0.
Ceci se réécrit
f1 − (t1 + s2t2 + · · · + sr tr ) f0 + s2 f2 + · · · + sr fr = 0.
f0 + w1 f1 + · · · + wr fr = 0.
L’inclusion
est une égalité. L’extension F = L(t2 , . . . ,tr ) est transcendante pure. D’après le
lemme 1, l’indice réduit sur L de l’hypersurface définie par
f0 + w1 f1 + · · · + wr fr = 0
dans PnL est égal à l’indice réduit de cette hypersurface sur F. Ceci achève la
démonstration.
Remarque. A. Pfister, J.W.S. Cassels et D. F. Coray (voir les références dans [3]) ont
donné des exemples d’intersections complètes de trois quadriques f0 = f1 = f2 = 0
dans Pnk (sur un corps k de caractéristique différente de 2) qui possèdent un zéro-
cycle de degré 1 sans posséder de point rationnel. La quadrique f0 + t1 f1 + t2 f2 =
0 sur le corps k(t1 ,t2 ) possède alors un zéro-cycle de degré 1. Comme c’est une
quadrique, un théorème de Springer [6] assure que cette quadrique admet un point
k(t1 ,t2 )-rationnel. On voit ainsi que le théorème 3 ne vaut pas lorsque l’on remplace
les zéro-cycles de degré 1 par des points rationnels : le théorème d’Amer et Brumer
ne s’étend pas à un système de 3 formes.
Remerciements. Les auteurs savent gré au rapporteur de ses lectures attentives. Marc Levine
remercie la NSF (grant number DMS-0801220) et la fondation Alexander von Humboldt.
Littérature
3. D. F. Coray, On a problem of Pfister about intersections of three quadrics. Arch. Math. (Basel)
34 (1980), no. 5, 403–411.
4. R. Elman, N. Karpenko, and A. Merkurjev, The algebraic and geometric theory of quadratic
forms, AMS Colloquium Publications, vol. 56, American Mathematical Society, Providence,
RI, 2008.
5. W. Fulton, Intersection theory, Ergebnisse der Mathematik und ihrer Grenzgebiete, 3. Folge,
Band 2, Springer-Verlag 1984.
6. T. A. Springer, Sur les formes quadratiques d’indice zéro. C. R. Acad. Sci. Paris 234 (1952)
1517–1519.
Quaternion Algebras with the Same Subfields
Summary Prasad and Rapinchuk asked if two quaternion division F-algebras that
have the same subfields are necessarily isomorphic. The answer is known to be “no”
for some very large fields. We prove that the answer is “yes” if F is an extension
of a global field K so that F/K is unirational and has zero unramified Brauer group.
We also prove a similar result for Pfister forms and give an application to tractable
fields.
1 Introduction
Gopal Prasad and Andrei Rapinchuk asked the following question in Remark 5.4 of
their paper [PR]:1
If two quaternion division algebras over a field F have the same maximal
(1.1)
subfields, are the algebras necessarily isomorphic?
The answer is “no” for some fields F, see Sect. 2 below. The answer is “yes” if F
is a global field by the Albert-Brauer-Hasse-Minkowski Theorem [NSW, 8.1.17].
Prasad and Rapinchuk note that the answer is unknown even for fields like Q(x).
We prove that the answer is “yes” for this field.
Skip Garibaldi
Department of Mathematics and Computer Science, Emory University, Atlanta, GA 30322, USA,
e-mail: [email protected]
David J. Saltman
Center for Communications Research-Princeton, 805 Bunn Drive, Princeton, NJ 008540, USA,
e-mail: [email protected]
2010 Mathematics subject classification. Primary: 16K20. Secondary: 11E04, 11E72.
1 The reference [40] in [PR] should point to this paper.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 225
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 13,
c Springer Science+Business Media, LLC 2010
226 Skip Garibaldi and David J. Saltman
The term “transparent” is defined in Sect. 6 below. Every retract rational exten-
sion of a local, global, real-closed, or algebraically closed field is transparent; in
particular K(x1 , . . . , xn ) is transparent for every global field K of characteristic = 2
and every n. There are many other examples.
The theorem can be viewed as a statement about symbols in the Galois cohomol-
ogy group H 2 (F, Z/2Z). We also state and prove an analogue for symbols in Galois
cohomology of H d (F, Z/2Z) – i.e., d-Pfister quadratic forms – in Theorem 9.1 later.
Notation. A global field is a finite extension of Q or of F p (t) for some prime p.
A local field is a completion of a global field with respect to a discrete valuation.
We follow [Lam 05] for notation on quadratic forms, except that we define the
symbol a1 , . . . , ad to be the d-Pfister form ⊗di=1 1, −ai . For a Pfister form ϕ , we
put ϕ for the unique form such that ϕ ∼ = 1 ⊕ ϕ .
2 An Example
Example 2.1. Several people (in no particular order: Markus Rost, Kelly McKinnie,
Adrian Wadsworth, Murray Schacher, Daniel Goldstein, etc.) noted that some hy-
pothesis on the field F is necessary for the conclusion of the theorem to hold. Here
is an example to illustrate this.
Let F0 be a field of characteristic = 2 that has quaternion division algebras Q1 ,
Q2 that are not isomorphic. Replacing F0 if necessary by the function field of a
generalized Severi-Brauer variety, we may assume that √ Q1 ⊗ Q2 has index 2 and
so that Q1 and Q2 both contain a maximal subfield F0 ( b). Write ϕi := ai , b
for the norm form of Qi . The Albert form ϕ1 − ϕ2 has anisotropic part similar to
γ = a1 a2 , b. √
Fix an extension E/F0 and a proper quadratic extension E( c) contained in Qi
but not Qi+1 , with subscripts taken modulo 2. Equivalently, −c is represented by
ϕi but not by ϕi+1 . Put q := c ⊕ ϕ ; it is anisotropic. Further, its determinant
i,c i+1
c is a nonsquare, so qi,c is not similar to ϕi , ϕi+1 , nor γ . All three of those forms
remain anisotropic over the function field E(qi,c ) of qi,c by [Lam 05,√X.4.10(3)].
That is, Q1 ⊗ E(qi,c ) and Q2 ⊗ E(qi,c ) are division and contain E(qi,c )( c), and the
two algebras are still distinct.
We build a field F from F0 inductively. Suppose we have constructed the field Fj
for some j ≥ 0. Let Fj+1 be the composita of the extensions Fj (qi,c ) for c as in the
previous paragraph with E = Fj and i = 1, 2. Let F be the colimit of the Fj .
By construction, Q1 ⊗ F and Q2 ⊗ F are both division, are not isomorphic to each
other, and have the same quadratic subfields. This shows that some hypothesis on
the field F is necessary for the conclusion of the theorem to hold.
Quaternion Algebras with the Same Subfields 227
One cannot simply omit the hypothesis “division” in the theorem, by Exam-
ple 3.6.
Furthermore, Theorem 1.2 does not generalize to division algebras of prime de-
gree > 2, even over number fields. One can easily construct examples using the
local–global principle for division algebras over a number field, see [PR, Example
6.5] for details.
Fix a field F with a discrete valuation v. We write F̄ for the residue field and F̂
for the completion. (Throughout this note, for fields with a discrete valuation and
elements of such a field, we use bars and hats to indicate the corresponding residue
field/completion and residues/image over the completion.)
For a central simple F-algebra A, we write  for A ⊗ F̂. It is Brauer-equivalent
to a central division algebra B over F̂. As v is complete on F̂, it extends uniquely to
a valuation on B by setting v(b) := deg1 B v(NrdB (b)) for b ∈ B× , see, e.g., [R, 12.6].
We say that A is unramified if v(B× ) = v(F × ) – e.g., if F̂ splits A – and ramified if
v(B× ) is strictly larger. That is, A is ramified or unramified if v is so on B. Note that
these meanings for ramified and unramified are quite a bit weaker than the usual
definitions (which usually require that the residue division algebra of B be separable
over F̄).
The valuation ring in F̂ is contained in a noncommutative discrete valuation ring
S in B, and the residue algebra of S is a (possibly commutative) division algebra B̄
whose center contains F̄. One has the formula
(see [W, p. 393] for references). That is, A is unramified if and only if dimF̄ B̄ =
dimF̂ B.
We use a few warm-up lemmas.
Lemma 3.1. Let A1 , A2 be central simple algebras over a field F that has a discrete
valuation. If there is a separable maximal commutative subalgebra that embeds in
one of the Âi and not the other, then the same is true for the Ai .
Lemma 3.2. Let A1 , A2 be central division algebras of prime degree over a field F
with a discrete valuation. If
1. Â1 is split and Â2 is not; or
2. Both Â1 and Â2 are division, but only one is ramified,
then there is a separable maximal subfield of A1 or A2 that does not embed in the
other.
Proof. The statement is obvious from Lemma 3.1. In case (1), we take the diagonal
subalgebra in Â1 . In case (2), we take a maximal separable subfield that is ramified.
For the rest of this section, we fix a prime p and let F be a field with a primitive
p-th root of unity ζ ; we suppose that F has a discrete valuation v with residue field
of characteristic different from p.
A degree p symbol algebra is an (associative) F-algebra generated by elements
α , β subject to the relations α p = a, β p = b, and αβ = ζ β α . It is central simple
over F [D, p. 78] and we denote it by (a, b)F or simply (a, b).
Lemma 3.3. If A is a degree p symbol algebra over F, then A is isomorphic to
(a, b)F for some a, b ∈ F × with v(a) = 0.
Proof. Write A as (a, b) where a = xπ na and b = yπ nb where x, y have value 0 and π
is a uniformizer for F. We may assume that p divides neither na nor nb , so na = snb
modulo p. But A is isomorphic to (a(−b)−s , b) because (−b, b) is split [D, p. 82,
Cor. 5], and v(a(−b)−s ) is divisible by p.
3.4. Let (a, b) be a symbol algebra as in Lemma 3.3 and write F̂ for the completion
of F at v. If (a, b) ⊗ F̂ is not division, certainly (a, b) is unramified. Otherwise, there
are two cases:
• If v(b) is divisible by p, then we can assume v(b) = 0, (a, b) is unramified, and
(a, b) is the symbol algebra (ā, b̄)F̄ . (The residue algebra – which is division –
contains (ā, b̄)F̄ , hence they must be equal.)
• If v(b)
√ is not divisible by p, then (a, b) is ramified and (a, b) is the field extension
F̄( p ā). (By hypothesis, a is not a p-th power in F̂, so by Hensel’s Lemma ā is
not a p-th power in F̄.)
Proposition 3.5. Let A1 and A2 be degree p symbol algebras over F that are division
algebras. If
1. A1 ramifies and A2 does not; or
2. A1 and A2 both ramify but their ramification defines different extensions of the
residue field F̄,
then there is a maximal subfield L of one of the Ai that does not split the other.
Proof. Write Ai as (ai , bi ). We can assume v(a1 ) = v(a2 ) = 0. If both the Ai ramify,
then the v(bi ) √
must be prime to p and the field extensions of the ramification of the
√
Ai must be F̄( p āi ). If these are different fields, then the field L2 := F( p a2 ) cannot
split A1 because it does not split the ramification of A1 .
Quaternion Algebras with the Same Subfields 229
Example 3.6. In the proposition, the hypothesis that A1 and A2 are both division
is necessary. Indeed, for = 2, the field Q of -adic numbers has two quaternion
algebras: one split and one division. Only one is ramified, but both are split by every
quadratic extension of Q .
In this section, we consider quaternion algebras over a field with a dyadic discrete
valuation. The purpose of this section is to prove:
What really helps nail down the structure of A1 ⊗ A2 is the hypothesis that the
residue algebra is a separable quadratic extension. In particular, the division algebra
underlying A1 ⊗ A2 is “inertially split” in the language of [JW] or “tame” in the
language of [GP].
Proof (of Proposition 4.1). Put B for the quaternion division F-algebra underlying
A1 ⊗ A2 . By the hypothesis on its residue algebra, B remains division when we
230 Skip Garibaldi and David J. Saltman
extend scalars to the completion F̂. Hence at least one of Â1 , Â2 is division; by
Lemma 3.2 (using that A1 and A2 are division) we may assume that both Â1 and Â2
are division and that the valuation has the same ramification index on both algebras.
By Lemma 3.1, we may assume that F is complete.
For sake of contradiction, suppose that A1 and A2 have the same subfields. The
residue division algebra Ā1 is distinct from F̄ and we can find a quadratic subfield
L of A1 so that the residue field L̄ is a quadratic extension of F̄. As L is a maximal
subfield of A1 , it is also one for A2 and hence B. Certainly, L̄ is a subfield of the
residue algebra B̄; as both have dimension 2 over F̄ they are equal. That is, L̄/F̄ is
separable. We write Ai = (L, ai ) and B = (L, b) for some ai , b ∈ F × . As the valuation
does not ramify on L we may assume that v(b) = 1. As Āi contains the separable
extension L̄/F̄, we have v(a1 ) = v(a2 ) as in Example 4.2. Then A1 ⊗ A2 ⊗ B – which
is split – is Brauer-equivalent to (L, a1 a2 b), which is ramified (because v(a1 a2 b) =
1+2v(ai ) is odd) and in particular not split. This is a contradiction, which completes
the proof of the proposition.
5 Unramified Cohomology
For a field F of characteristic not 2, we write H d (F) for the Galois cohomology
group H d (F, Z/2Z). We remind the reader that for d = 1, this group is F × /F ×2 and
for d = 2 it is identified with the 2-torsion in the Brauer group of F.
Definition 5.1. Fix an integer d ≥ 2. We define Hud (F) to be the subgroup of H d (F)
consisting of classes that are unramified at every nondyadic discrete valuation of F
(in the usual sense, as in [GMS, p. 19]) and are killed by H d (F) → H d (R) for every
real closure R of F.
Example 5.2. 1. A local field F (of characteristic = 2) has Hu2 (F) = 0 if and only if
F is nondyadic. Indeed, F has no orderings, so it suffices to consider the unique
discrete valuation on F.
2. A global field F (of characteristic = 2) has Hu2 (F) = 0 if and only if F has at most
1 dyadic place. This is Hasse’s local–global theorem for central simple algebras
[NSW, 8.1.17].
3. A real closed field F has Hud (F) = 0 for all d ≥ 2; this is trivial.
A local field F (of characteristic = 2) has Hud (F) = 0 for d ≥ 3. This is trivial
because such a field has no orderings and cohomological 2-dimension 2.
√
Lemma 5.3. If F( −1) has cohomological 2-dimension < d, then Hud (F) is zero.
Proof. If an element x ∈ H d (F) is zero at every real closure, then by [A, Satz 3]
x · (−1)r is zero for some r ≥ 0. On the other hand, there is an exact sequence for
every n [KMRT, 30.12(1)]:
√ cor ·(−1) res √
H n (F( −1)) −−−−→ H n (F) −−−−→ H n+1 (F) −−−−→ H n+1 (F( −1)
Quaternion Algebras with the Same Subfields 231
For n ≥ d, the two end terms are zero, so the cup product with (−1)r defines an
∼
isomorphism H d (F) −
→ H d+r (F) for all r ≥ 0.
We immediately obtain:
Corollary 5.4. A global field F of characteristic = 2 has Hud (F) = 0 for all d ≥ 3.
Proposition 5.6. For every extension F/F0 , the homomorphism res : H d (F0 ) →
H d (F) restricts to a homomorphism Hud (F0 ) → Hud (F). If, additionally, F/F0 is
unirational and
every class in H d (F) that is unramified at every discrete valuation
(5.7)
of F/F0 comes from H d (F0 )
Lemma 5.8. If R1 ⊆ R2 are real-closed fields, then the natural map H d (R1 ) →
H d (R2 ) is an isomorphism for every d ≥ 0.
Proof. Trivially, H 0 (Ri ) = Z/2Z. Also, H 1 (Ri ) = Z/2Z with nonzero element
(−1) ∈ R× ×2
i /Ri . For all d, H (Ri ) (Galois cohomology) equals H (Z/2Z, Z/2Z)
d d
(group cohomology), and two-periodicity for the cohomology of finite cyclic groups
shows that H d (Ri ) = Z/2Z for all d ≥ 0, with nonzero element (−1)d .
6 Transparent Fields
The unramified Brauer group of an extension F/F0 is the subgroup of the Brauer
group of F consisting of elements that are unramified at every discrete valuation of
F/F0.
Proof. Suppose there is a nonzero x ∈ Hu2 (F). By Proposition 5.6, x is the image
of some nonzero x0 ∈ Hu2 (F0 ) and there is a dyadic valuation v of F0 such that con-
ditions (1) and (2) of Definition 6.1 holds for the division algebra D0 represented
by x0 . Extend v to a valuation on F as in the proof of Proposition 5.6. The maxi-
mal unramified extension of F̂ (with respect to v) contains the maximal unramified
extension of F̂0 , so it kills x0 and hence also x; this verifies condition (6.1.1). For
condition (6.1.2), it suffices to note that the map H 1 (F̄0 , Z/2Z) → H 1 (F̄, Z/2Z) is
injective as in the proof of Proposition 5.6.
We now prove Theorem 1.2. We assume that D1 and D2 are not isomorphic and that
there is some quadratic extension of F that is contained in both of them, hence that
D1 ⊗ D2 has index 2. We will produce a quadratic extension of F that is contained
in one and not the other.
Suppose that first there is a real closure R of F that does not split D1 ⊗ D2 . Then
one of the Di is split by R and the other is not; say D1 is split. We can write
√
D1 = (a1 , b1 ), where a1 is positive in R. The field F( a1 ) is contained in D1 , but
not contained in D2 because D2 ⊗ R is division.
If there is a nondyadic discrete valuation of F where D1 ⊗ D2 is ramified then
[D1 ] and [D2 ] have different images under the residue map H 2 (F) → H 1 (F̄) and
Proposition 3.5 provides the desired quadratic extension.
Finally, suppose that D1 ⊗ D2 is split by every real closure and is unramified at
every nondyadic place, so D1 ⊗ D2 belongs to Hu2 (F). By the original hypotheses
on D1 and D2 , the class [D1 ⊗ D2 ] is represented by a quaternion division algebra
B. But F is transparent, so Proposition 4.1 provides a quadratic extension of F that
embeds in D1 or D2 but not the other.
234 Skip Garibaldi and David J. Saltman
With the theorem from the introduction proved, we now set to proving a version of
it for Pfister forms. The goal of this section is the following proposition, which is
an analogue of Proposition 3.5.
The proof given later for case (1) of the proposition was suggested to us by
Adrian Wadsworth, and is much simpler than our original proof for that case. It
makes use of the following lemma:
Proof (of Proposition 8.1). Obviously, xπ , yπ ∼ = −xy, yπ , so we may assume
that
ϕi = ai2 , ai3 , . . . , aid , bi ,
where ai j has value 0 and bi has value 0 or 1.
In case (2), b1 and b2 have value 1 and the residue forms
are not isomorphic. We take γ to be a12 , a13 , . . . , a1d ; it divides ϕ1 . The projective
quadric X defined by γ = 0 is defined over the discrete valuation ring, and F̄(X ) does
not split r(ϕ2 ) by [Lam 05, X.4.10]. Hence, F(X ) does not split ϕ2 and γ does not
divide ϕ2 .
Now suppose we are in case (1); obviously v(b1 ) = 1. If ϕ2 ⊗ F̂ is anisotropic,
then as ϕ2 does not ramify, v(b2 ) = 0 and the second residue form of ϕ2 is zero.
Certainly, γ := a13 , . . . , a1d , b1 divides ϕ1 . On the other hand, every anisotropic
multiple of γ ⊗ F̂ over F̂ will have a nonzero second residue form, so ϕ2 cannot
contain γ .
Otherwise, ϕ2 ⊗ F̂ is hyperbolic, and there exists some 2-dimensional subform
q of ϕ2 that becomes hyperbolic over F̂ by Lemma 8.2. It follows that there is a
(d − 1)-Pfister ψ that is contained in ϕ2 and that becomes hyperbolic over F̂. But
ϕ1 ⊗ F̂ is ramified, hence anisotropic, so ϕ1 cannot contain ψ .
Theorem 9.1. Let F be a field of characteristic = 2 such that Hud (F) = 0 for some
d ≥ 2. Let ϕ1 and ϕ2 be anisotropic d-Pfister forms over F such that, for every
(d − 1)-Pfister form γ , we have: γ divides ϕ1 if and only if γ divides ϕ2 . Then ϕ1 is
isomorphic to ϕ2 .
We remark that for d = 2, Theorem 9.1 does not include the case where F is a
retract rational extension of a global field with more than one dyadic place, a case
included in Theorem 1.2.
Proof (of Theorem 9.1). Suppose that ϕ1 and ϕ2 are d-Pfister forms over F for
some d ≥ 2, Hud (F) is zero, and ϕ1 is not isomorphic to ϕ2 . Then ϕ1 + ϕ2 is in
H d (F) \ Hud (F).
If there is a nondyadic discrete valuation where ϕ1 + ϕ2 ramifies, then Proposi-
tion 8.1 gives a (d − 1)-Pfister form that divides one of the ϕi and not the other.
Otherwise, there is an ordering v of F such that ϕ1 and ϕ2 are not isomorphic
over the real closure Fv , i.e., one is locally hyperbolic and the other is not. Write
ϕi = ai1 , ai2 , . . . , aid and suppose that ϕ1 is locally hyperbolic, so some a1 j is
positive; renumbering if necessary, we may assume that it is a1d . Then the form
a12 , a13 , . . . , a1d divides ϕ1 , but it is hyperbolic over Fv and so does not divide ϕ2 .
We now use the notion of unramified cohomology from Sect. 5 to prove some new
results and give new proofs of some known results regarding tractable fields. Recall
from [CTW] that a field F of characteristic = 2 is called tractable if for every a1 ,
236 Skip Garibaldi and David J. Saltman
a2 , a3 , b1 , b2 , b3 ∈ F × such that the six quaternion algebras (ai , b j ) are split for all
i = j and the three quaternion algebras (ai , bi ) are pairwise isomorphic, the algebras
(ai , bi ) are necessarily split. (This condition was motivated by studying conditions
for decomposability of central simple algebras of degree 8 and exponent 2.) We
prove:
Proof. For the sake of contradiction, suppose we are given ai , bi as in the definition
of tractable where the quaternion algebras (ai , bi ) are not split. If the (ai , bi ) do
not split at some real closure of F, then the ai and bi are all negative there. Hence
(ai , b j ) for i = j is also not split, a contradiction.
So the (ai , bi ) ramify at some nondyadic discrete valuation v of F and in particu-
lar are division over the completion of F at v. We replace F with its completion and
derive a contradiction. Modifying each ai or bi by a square if necessary, we may
assume that each one has value 0 or 1. Further, as each (ai , bi ) is ramified, at most
one of the two slots has value 0.
Suppose first that one of the ai or bi – say, a1 – has value 0. Then b1 has value 1
and a1 is not a square in F. As (a1 , b j ) is split for j = 1, we deduce that v(b j ) = 0.
Hence v(a2 ) = v(a3 ) = 1. As (a3 , b3 ) is division, b3 is not a square in F. Since
v(a2 ) = 1 and v(b3 ) = 0, it follows that (a2 , b3 ) is division, a contradiction.
It remains to consider the case where the ai ’s and bi ’s all have value 1. We can
write ai = παi and bi = πβi , where π is a prime element – say, a1 so α1 = 1 – and
αi , βi have value 0. In the Brauer group of F, we have for i = j:
a contradiction.
Example 10.2. In [CTW], the authors proved that a local field is tractable if and
only if it is nondyadic (ibid., Cor. 2.3) and a global field is tractable if and only if it
has at most 1 dyadic place (ibid., Th. 2.10). Combining the preceding proposition
and Example 5.2 gives a proof of the “if” direction for both of these statements.
Corollary 10.3 ([CTW, Th. 3.17]). Every field of transcendence degree 1 over a
real-closed is tractable.
Example 10.4. The converse of Proposition 10.1 is false. For every odd prime
p, Q p ((x)) is tractable by [CTW, Prop. 2.5]. On the other hand, the usual dis-
crete valuation on the formal power series ring is the unique one [EP, 4.4.1], so
Hu2 (Q p ((x))) = H 2 (Q p ) = Z/2Z.
Quaternion Algebras with the Same Subfields 237
Proof. We note that F0 has Hu2 (F0 ) = 0. In case F0 is global or local, then F0 has
one dyadic place or is nondyadic, respectively, by [CTW], and the claim follows,
see Example 5.2. Then Hu2 (F) is also zero by Proposition 5.6, hence F is tractable.
Corollary 10.6. Let F be a global field with at most one dyadic place. If G is an
isotropic, simply connected, absolutely almost simple linear algebraic group over
F, then the function field F(G) is tractable.
Acknowledgements The first author’s research was partially supported by National Science
Foundation grant no. DMS-0653502. Both authors thank Adrian Wadsworth and Kelly McKin-
nie for their comments on an earlier version of this paper.
References
[JW] B. Jacob and A. Wadsworth, Division algebras over Henselian fields, J. Algebra 128
(1990), 126–179.
[KMRT] M.-A. Knus, A.S. Merkurjev, M. Rost, and J.-P. Tignol, The book of involutions, Col-
loquium Publications, vol. 44, Am. Math. Soc., 1998.
[Lam 80] T.Y. Lam, The theory of ordered fields, Ring theory and algebra, III (Proc. Third Conf.,
Univ. Oklahoma, Norman, Okla., 1979), Lecture Notes in Pure and Appl. Math.,
vol. 55, Dekker, New York, 1980, pp. 1–152.
[Lam 05] , Introduction to quadratic forms over fields, Graduate Studies in Mathematics,
vol. 67, Am. Math. Soc., Providence, RI, 2005.
[M] A. Merkurjev, Unramified elements in cycle modules, J. Lond. Math. Soc. 78 (2008),
no. 2, 51–64.
[NSW] J. Neukirch, A. Schmidt, and K. Wingberg, Cohomology of number fields, 2nd ed.,
Grundlehren der math. Wissenschaften, vol. 323, Springer, Berlin, 2008.
[PR] G. Prasad and A.S. Rapinchuk, Weakly commensurable arithmetic groups and isospec-
tral locally symmetric spaces, Pub. Math. IHES 109 (2009), 113–184.
[R] I. Reiner, Maximal orders, LMS Monographs, vol. 5, Academic Press, London, 1975.
[Sa] D.J. Saltman, Lectures on division algebras, CBMS Regional Conference Series in
Mathematics, vol. 94, American Mathematical Society, Providence, RI, 1999.
[Se] J-P. Serre, Galois cohomology, Springer-Verlag, Berlin, 2002, originally published as
Cohomologie galoisienne (1965).
[T] J.-P. Tignol, Classification of wild cyclic field extensions and division algebras of
prime degree over a Henselian field, Proceedings of the International Conference on
Algebra, Part 2 (Novosibirsk, 1989), Contemp. Math., vol. 131, Am. Math. Soc., 1992,
pp. 491–508.
[W] A.R. Wadsworth, Valuation theory on finite dimensional division algebras, Valua-
tion theory and its applications, Vol. I (Saskatoon, SK, 1999), Fields Inst. Commun.,
vol. 32, Am. Math. Soc., Providence, RI, 2002, pp. 385–449.
Lifting of Coefficients for Chow Motives
of Quadrics
Olivier Haution
Summary We prove that the natural functor from the category of Chow motives
of smooth projective quadrics with integral coefficients to the category with coeffi-
cients modulo 2 induces a bijection on the isomorphism classes of objects.
1 Introduction
Alexander Vishik has given a description of the Chow motives of quadrics with
integral coefficients in [6]. It uses much subtler methods than the ones used to
give a similar description with coefficients in Z/2, found for example in [2], but the
description obtained is the same ([2, Theorems 93.1 and 94.1]). The result presented
here allows one to recover Vishik’s results from the modulo 2 description.
In order to state the main result, we first define the categories involved. Let Λ be
a commutative ring. We write QF for the class of smooth projective quadrics over a
field F. We consider the additive category C (QF , Λ ), where objects are (coproducts
of) quadrics in QF and if X ,Y are two such quadrics, Hom(X ,Y ) is the group of
correspondences of degree 0, namely CHdim X (X × Y, Λ ). We write C M (QF , Λ )
for the idempotent completion of C (QF , Λ ). This is the category of graded Chow
motives of smooth projective quadrics with coefficients in Λ . If (X , ρ ), (Y, σ ) are
two such motives then we have:
Olivier Haution
Institut de mathématiques de Jussieu, 175 rue du Chevaleret 75013 Paris, France,
e-mail: [email protected]
2010 Mathematics subject classification. 11E04, 14C25.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 239
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 14,
c Springer Science+Business Media, LLC 2010
240 Olivier Haution
We first recall some facts and fix the notations that we will use.
If L/F is a field extension, and S is a scheme over F, we write SL for the scheme
S ×Spec(F) Spec(L). Similarly, for an F-vector space U, we write UL for U ⊗F L, and
for a cycle x ∈ CH(S), the element xL ∈ CH(SL ) is the pull-back of x along the flat
morphism SL → S.
We say that a cycle in CH(SL ) is F-rational (or simply rational when no confu-
sion seems possible) if it can be written as xL for some cycle x ∈ CH(S), i.e., if it
belongs to the image of the pull-back homomorphism CH(S) → CH(SL ).
Let F be a field and ϕ be a nondegenerate quadratic form on an F-vector space
V of dimension D + 2. The associated projective quadric X is smooth of dimension
D = 2d or 2d + 1. Let L/F be a splitting extension for X , i.e., a field extension
such that VL has a totally isotropic subspace of dimension d + 1. We write hi , li
for the usual basis of CH(XL ), where 0 ≤ i ≤ d. The class h is the pull-back of
the hyperplane class of the projective space of VL , the class li is the class of the
projectivisation of a totally isotropic subspace of VL of dimension i + 1.
If D is even, then CHd (XL ) is freely generated by hd and ld . In this case, there are
exactly two classes of maximal totally isotropic spaces, ld and ld . They correspond
to spaces exchanged by a reflection and verify the relation ld + ld = hd .
The group Aut(L/F) acts on CH(XL ). It acts trivially on the ith homogeneous
component of CH(XL ), as long as 2i = D.
See [2] for proofs of all these facts.
In the next proposition, X is a smooth projective quadric of dimension D = 2d
associated with a quadratic space (V, ϕ ) over a field F, L/F is a splitting extension
for X , and disc X is the discriminant algebra of ϕ .
Proposition 2. Under the natural Aut(L/F)-actions, the pair {ld , ld } can be
identified with the connected components of Spec(disc X ⊗ L).
Proof. We consider the scheme G(ϕ ) of maximal totally isotropic subspaces of V ,
i.e., the grassmannian variety of isotropic (d + 1)-dimensional subspaces of V . The
Lifting of Coefficients for Chow Motives of Quadrics 241
scheme G(ϕ )L has two connected components exchanged by any reflection of the
quadratic space (V, ϕ ). There is a faithfully flat morphism G(ϕ ) → Spec(disc X )
(see [2, §85, p. 357]), hence the connected components of G(ϕ )L are in correspon-
dence with those of Spec(disc X ⊗ L), in a way respecting the natural Aut(L/F)-
actions.
Now two maximal totally isotropic subspaces lie in the same connected compo-
nent of G(ϕ )L if and only if the corresponding d-dimensional closed subvarieties of
the quadric XL are rationally equivalent (see [2, §86, p. 358]). Therefore, the pair
{ld , ld } is Aut(L/F)-isomorphic to the pair of connected components of G(ϕ )L .
The statement follows.
3 Lifting of Coefficients
We now give a useful characterization of rational cycles (Proposition 7). The proof
will rely on the following theorem from [5, Proposition 9]:
Theorem 3 (Rost’s nilpotence for quadrics). Let X be a smooth projective quadric
over a field F, and let α ∈ EndC M (QF ,Λ ) (X ). If αL ∈ CH(XL2 ) vanishes for some
field extension L/F then α is nilpotent.
We will use the following classical corollaries:
Corollary 4 ([2, Corollary 92.5]). Let X be a smooth projective quadric over a
field F and L/F a field extension. Let π a projector in EndC M (QL ,Λ ) (XL ) that is the
restriction of some element in EndC M (QF ,Λ ) (X ). Then there exist a projector ϕ in
EndC M (QF ,Λ ) (X ) such that ϕL = π .
Corollary 5 ([2, Corollary 92.7]). Let f : (X , ρ ) → (Y, σ ) be a morphism in the
category C M (QF , Λ ). If fL is an isomorphism for some field extension L/F then
f is an isomorphism.
Proposition 6. For any n ≥ 1, the functor C M (QF , Z/2n ) → C M (QF , Z/2) is
bijective on the isomorphism classes of objects.
Proof. We are clearly in the situation () of [7, Sect. 2, p. 587] for the obvious
functor C (QF , Z/2n ) → C (QF , Z/2). The statement then follows from [7, Propo-
sitions 2.5 and 2.2] (see also [4, Corollary 2.7]).
Any smooth projective quadric admits a (noncanonical) finite Galois splitting
extension, of degree a power of 2. This will be used together with the following
proposition when we will need to prove that a given cycle with integral coefficients
(and defined over some extension of the base field) is rational.
Proposition 7. Let X ,Y ∈ QF and L/F be a splitting Galois extension of degree
m for X and Y . A correspondence in CH((X × Y )L , Z) is rational if and only if
it is invariant under the group Gal(L/F) and its image in CH((X × Y )L , Z/m) is
rational.
242 Olivier Haution
ZL o ZL⊗F̄
Zo ZF̄
CH(ZL ) / CH(ZL⊗F̄ )
CH(Z) / CH(ZF̄ )
x → ∑ t ∗ (γ x)
γ ∈Gal(L/F)
where t : ZF̄ → ZL is the map induced by τ . Using the commutativity of the diagram
and the injectivity of t ∗ , we see that the composite CH(ZL ) → CH(Z) → CH(ZL )
maps x to ∑ γ x, where γ runs in Gal(L/F). The claim follows.
Now suppose that u is a cycle in CH(ZL , Z) invariant under Gal(L/F), and that
its image in CH(ZL , Z/m) is rational. We can find a rational cycle v in CH(ZL , Z)
and a cycle δ in CH(ZL , Z) such that mδ = v − u. As CH(ZL , Z) is torsion-free, δ is
invariant under Gal(L/F) . The first claim ensures that v − u is rational, hence u is
rational, and we have proven the proposition.
Let us remark that if X ∈ QF , L/F is a splitting extension, and 2i < dim X then
2li = hdim X−i ∈ CH(XL ) is always rational. It follows that 2 CHi (XL ) consists of
rational cycles when 2i = dim X.
Proof. Let (X, π ) ∈ C M (QF , Z/2) and L/F a finite splitting Galois extension for
X of degree 2n . By Proposition 6, we can lift the isomorphism class of (X , π ) to the
isomorphism class of some (X, τ ) ∈ C M (QF , Z/2n ).
Lifting of Coefficients for Chow Motives of Quadrics 243
Lemma 9 ([4, Lemma 2.14]). For any positive integers k and p, the reduction
homomorphism SLk (Z) → SLk (Z/p) is surjective.
Proof. As Z/p is a semilocal commutative ring, it follows from [1, Corollary 9.3,
Chap. V, p. 267], applied with A = Z and q = pZ , that any matrix in SLk (Z/p) is
the image modulo p of a product of elementary matrices with integral coefficients.
Such a product in particular belongs to SLk (Z), as required.
In order to prove injectivity in Theorem 1, we may assume that we are given two
motives (X, ρ ), (Y, σ ) in C M (QF , Z) and an isomorphism between their images in
C M (QF , Z/2). We will build an isomorphism with integral coefficients between
the two motives (which will not, in general, be an integral lifting of the original
isomorphism with finite coefficients).
244 Olivier Haution
We fix a finite Galois splitting extension L/F for X and Y of degree 2n . Using
Proposition 6 we may assume that there exists an isomorphism α between (X , ρ )
and (Y, σ ) in C M (QF , Z/2n ). By Proposition 7 and Corollary 5, it is enough to
build an isomorphism (XL , ρL ) → (YL , σL ) which reduces to a rational correspon-
dence modulo 2n and which is equivariant under Gal(L/F).
Let dX be such that dim X = 2dX or 2dX + 1 and dY defined similarly for Y . Let
r(X, ρ ) be the rank of CHdX (XL ) ∩ im (ρL )∗ if dim X is even and r(X , ρ ) = 0 if dim X
is odd. We define r(Y, σ ) in a similar fashion. We will distinguish cases using these
integers.
A basis of CH(XL ) ∩ im (ρL )∗ gives an isomorphism of (XL , ρL ) with twists of
Tate motives, thus choosing bases for the groups CH(XL ) ∩ im (ρL )∗ and CH(YL ) ∩
im (σL )∗ , we can see morphisms between the two motives as matrices.
We fix a basis (ei ) of CH(XL ) ∩ im (ρL )∗ as follows: we choose ei ∈ CHi (XL )
among the cycles hdim X−i , li for 2i = dim X . We are done when r(X , ρ ) = 0.
If r(X, ρ ) = 2 we complete the basis with edX = ldX , edX = ld X ∈ CHdX (XL ).
If r(X, ρ ) = 1 we choose a generator edX of CHdX (XL ) ∩ im (ρL )∗ to complete the
basis.
We choose a basis ( fi ) for CH(YL ) ∩ im (σL )∗ in a similar way.
If we write ρ̃ and σ̃ for the reduction modulo 2n of ρ and σ , these bases reduce
to bases (e˜i ) of CH(XL , Z/2n ) ∩ im(ρ̃L )∗ and ( f̃i ) of CH(YL , Z/2n ) ∩ im(σ̃L )∗ . In
these homogeneous bases the matrix of a correspondence of degree 0 is diagonal by
blocks. The sizes of the blocks are the ranks of the homogeneous components of
im (ρL )∗ .
Lemma 10. If r(X , ρ ) = 1 then disc X is trivial.
Proof. Assume disc X is not trivial. The correspondence ρ induces a projection of
CHdX (XL ) which is equivariant under the action of Gal(L/F). But CHdX (XL ) is
indecomposable as a Gal(L/F)-module. It follows that (ρL )∗ is either the identity
or 0 when restricted to CHdX (XL ) , hence r(X , ρ ) = 1.
Corollary 11. If r(X, ρ ) = 2 then Gal(L/F) acts trivially on im (ρL )∗ .
Lemma 12. If r(X, ρ ) = 2 then r(Y, ρ ) = 2, dimY = dim X and discY = disc X.
Proof. As the isomorphism (αL )∗ is graded, the dX -th homogeneous component of
im(αL )∗ has rank 2. This image is a subgroup of the Chow group with coefficients
in Z/2n of a split quadric, thus the only possibility is that dimY is even, dX = dY
and r(Y, σ ) = 2.
The isomorphism (αL )∗ is equivariant under the action of Gal(L/F). It follows
that an element of the group Gal(L/F) acts trivially on CH(XL , Z/2n ) if and only if it
acts trivially on CH(YL , Z/2n ). But such an element acts trivially on CH(XL , Z/2n )
(respectively, CH(YL , Z/2n )) if and only if it acts trivially on the pair of integral
cycles {ldX , ldX } ⊂ CH(XL ) (respectively, {ldY , ldY } ⊂ CH(YL )). Therefore, the
pair of integral cycles {ldX , ldX } is Gal(L/F)-isomorphic to the pair {ldY , ldY }. By
proposition 2, we have a Gal(L/F)-isomorphism between the split étale algebras
disc X ⊗ L and discY ⊗ L. Hence, disc X and discY correspond to the same cocycle
in H 1 (Gal(L/F), Z/2), thus are isomorphic.
Lifting of Coefficients for Chow Motives of Quadrics 245
γ = Δ + 2 ∑ ki ẽi × ẽdimX−i ,
where the sum is taken over all i such that CHi (XL ) ⊂ im (ρL )∗ (which implies in
case r(X, ρ ) = 1 that we do not take i = dX ). The composite β ◦ γ is rational,
and its matrix in our bases is the identity matrix. This correspondence lifts to an
isomorphism with integral coefficients (XL , ρL ) → (YL , σL ).
Next assume that r(X , ρ ) = 2. Then we have dim X = dimY , r(Y, σ ) = 2, and
disc X = discY by Lemma 12. The matrix of (αL )∗ is diagonal by blocks:
⎛ ⎞
νi1
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ νir ⎟
⎜ ⎟
⎜ B ⎟,
⎜ ⎟
⎜ νir+1 ⎟
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
νi p
the sum being taken over all i such that CHi (XL ) ∩ im (ρL )∗ = 0/ and i =
dX .
Therefore, we may assume that νi = 1 for all i and that we have a matrix of the
shape (Ir being the identity block of size r):
⎛ ⎞
Is 0
⎜ a+1 a ⎟
⎜ ⎟
⎝ a a+1 ⎠ .
0 It
The matrix of the rational cycle hdX × hdY ∈ CH((X × Y )L , Z/2n ) is:
⎛ ⎞
0 0
⎜ 11 ⎟
⎜ ⎟
⎝ 11 ⎠
0 0
Now αL − a(hdX × hdY ) is rational and its matrix is the identity. This cycle is invari-
ant under Gal(L/F) and lifts to an isomorphism (XL , ρL ) → (YL , σL ).
It remains to treat the case when disc X is trivial. In this case the group Gal(L/F)
acts trivially on CH(XL ) (and on CH(YL ) as discY is also trivial). As before,
composing with a rational cycle, we may assume that νi = 1 for all i. We write
det B−1 = 2k + 1.
The cycle Δ +k(hdX ×hdX ) ∈ End(XL , ρ̃L ) is rational and its matrix in our basis is:
⎛ ⎞
Ip 0
⎜ 1+k k ⎟
⎜ ⎟.
⎝ k 1+k ⎠
0 Ir
We see that the determinant of this matrix is 1 + 2k. Therefore, the composite
αL ◦ (Δ + k(hdX × hdX )) has determinant 1. We use Lemma 9 to conclude, which
completes the proof of Theorem 1.
Acknowledgements This work is part of my PhD thesis at the University of Paris 6 under the
direction of Nikita Karpenko. I am very grateful to him for his useful suggestions.
References
5. Rost M., The Motive of a Pfister Form, preprint (1998), available at http://www.math.
uni-bielefeld.de/∼ rost/motive.html
6. Vishik A., Motives of Quadrics with Applications to the Theory of Quadratic Forms (2004)
in Geometric methods in the algebraic theory of quadratic forms, Lect. Notes in Math. 1835,
Springer, Berlin, 25–101.
7. Vishik A., Yagita N., Algebraic Cobordisms of a Pfister Quadric (2007), J. Lond. Math. Soc.
(2) 76 (2007) 586–604.
Upper Motives of Outer Algebraic Groups
Nikita A. Karpenko
1 Introduction
We fix an arbitrary base field F. Besides that, we fix a finite field F and we consider
the Grothendieck Chow motives over F with coefficients in F. These are the objects
of the category CM(F, F), defined as in [4].
Let G be a semisimple affine algebraic group over F. According to [3, Corollary
35(4)] (see also Corollary 2.2 here), the motive of any projective G-homogeneous
variety decomposes (in a unique way) into a finite direct sum of indecomposable
motives. One would like to describe the indecomposable motives that appear this
way. In this paper we do it under certain assumption on G formulated in terms of the
(unique up to F-isomorphism) minimal field extension E/F such that the group GE
is of inner type: the degree of E/F is assumed to be a power of p, where p = char F.
Nikita A. Karpenko
UPMC Univ Paris 06, Institut de Mathématiques de Jussieu, Paris, France,
e-mail: [email protected]
2010 Mathematics subject classification. 14L17, 14C25.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 249
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 15,
c Springer Science+Business Media, LLC 2010
250 Nikita A. Karpenko
Note that this has already been done in [5] in the case when E = F, that is, when
G itself is of inner type. Therefore, though the inner case is formally included in
the present paper, we concentrate here on the special effects of the outer case. This
remark explains the choice of the title.
Note that the extension E/F is galois. Actually, we do not use the minimality
condition on the extension E/F in the paper. Therefore, E/F could be any finite
p-primary galois field extension with GE of inner type. However, it is reasonable to
keep the minimality condition at least for the sake of the definition of the set of the
upper motives of G which we give now.
For any intermediate field L of the extension E/F and any projective GL -
homogeneous variety Y , we consider the upper (see [5, Definition 2.10]) inde-
composable summand MY of the motive M(Y ) ∈ CM(F, F) of Y (considered as
an F-variety at this point). By definition, this is the (unique up to isomorphism)
indecomposable summand of M(Y ) with nonzero 0-codimensional Chow group.
The set of the isomorphism classes of the motives MY for all L and Y is called the
set of upper motives of the algebraic group G.
The summand MY is definitely the “easiest” indecomposable summand of M(Y )
over which we have the best control. For instance, the motive MY is isomorphic to
the motive MY for another projective homogeneous (not necessarily under an action
of the same group G) variety Y if and only if there exist multiplicity 1 correspon-
dences Y Y and Y Y , [5, Corollary 2.15]. Here a correspondence Y Y is
an element of the (dimY )-dimensional Chow group of Y ×F Y with coefficients in
F; its multiplicity is its image under the push-forward to the (dimY )-dimensional
Chow group of Y identified with F.
One more nice property of MY (which will be used in the proof of Theorem 1.1)
is an easy control over the condition that MY is a summand of an arbitrary motive
M: by [5, Lemma 2.14], this condition holds if and only if there exist morphisms
α : M(Y ) → M and β : M → M(Y ) such that the correspondence β ◦ α is of multi-
plicity 1.
We are going to claim that the complete motivic decomposition of any projective
G-homogeneous variety X consists of shifts of upper motives of G. In fact, the
information we have is a bit more precise:
Remark 1.2. Theorem 1.1 fails if the degree of the extension E/F is divisible by a
prime different from p (see Example 3.3).
The proof of Theorem 1.1 is given in Sect. 4. Before this, we get some prepara-
tory results which are also of independent interest. In Sect. 2, we prove the nilpo-
tence principle for the quasi-homogeneous varieties. In Sect. 3, we establish some
properties of a motivic corestriction functor.
Upper Motives of Outer Algebraic Groups 251
Let us consider the category CM(F, Λ ) of Grothendieck Chow motives over a field
F with coefficients in an arbitrary associative commutative unital ring Λ .
We say that a smooth complete F-variety X satisfies the nilpotence principle if for
any Λ and any field extension K/F, the kernel of the change of field homomorphism
Proof. By [4, Theorem 92.4] it suffices to show that the quasi-homogeneous vari-
eties form a tractable class. We first recall the definition of a tractable class C (over
F). This is a disjoint union of classes CK of smooth complete K-varieties, where K
runs over all field extensions of F, having the following properties:
(1) If Y1 and Y2 are in CK for some K, then the disjoint union of Y1 and Y2 is also in
CK ;
(2) If Y is in CK for some K, then each component of Y is also in CK ;
(3) If Y is in CK for some K, then for every field extension K /K the K -variety YK
is in CK ;
(4) If Y is in CK for some K, Y is irreducible, dimY > 0, and Y (K) = 0, / then CK
contains a (not necessarily connected) variety Y0 such that dimY0 < dimY and
M(Y ) M(Y0 ) in CM(K, Λ ) (for any Λ or, equivalently, for Λ = Z).
Let us define a class C as follows. For every field extension K/F, CK is the class
of all quasi-homogeneous K-varieties.
We claim that the class C is tractable. Indeed, the properties (1)–(3) are trivial
and the property (4) is [2, Theorem 7.2].
252 Nikita A. Karpenko
We turn back to the case where the coefficient ring Λ is a finite field F.
as follows: on the objects corL/F (X , i) = (X, i), where on the right-hand side X
is considered as an F-variety via the composition X → SpecL → SpecF; on the
morphisms, the map
HomC(L,Λ ) ((X, i), (Y, j)) → HomC(F,Λ ) ((X, i), (Y, j))
Chi (M) and Chi (corL/F M) are canonically isomorphic as well as the Chow groups
Chi (M) and Chi (corL/F M) are. Indeed, since resL/F Λ = Λ ∈ CM(L, Λ ), we have
Chi (M) := Hom (M, Λ (i)) = Hom corL/F M, Λ (i) =: Chi (corL/F M) and
Chi (M) := Hom (Λ (i), M) = Hom Λ (i), corL/F M =: Chi (corL/F M).
Proposition 3.1. The following three conditions on a finite galois field extension
E/F are equivalent:
(1) For any intermediate field F ⊂ K ⊂ E, the K-motive of Spec E is indecompos-
able;
(2) For any intermediate fields F ⊂ K ⊂ L ⊂ E and any indecomposable L-motive
M, the K-motive corL/K (M) is indecomposable;
(3) The degree of E/F is a power of p (where p is the characteristic of the coeffi-
cient field F).
Proof. We start by showing that (3) ⇒ (2). So, we assume that [E : F] is a power
of p and we prove (2). The extension L/K decomposes in a finite chain of galois
degree p extensions. Therefore, we may assume that L/K itself is a galois degree
p extension. Let R = End(M). This is an associative, unital, but not necessarily
commutative F-algebra. Moreover, since M is indecomposable,
the ring R has no
nontrivial idempotents. We have End corL/K (M) = R ⊗F End (MK (Spec L)) where
MK (SpecL) ∈ CM(K, F) is the motive of the K-variety Spec L. According to [3,
Sect. 7], the ring End (MK (SpecL)) is isomorphic to the group ring FΓ , where Γ
is the Galois group of L/K. As the group Γ is (cyclic) of order p, we have FΓ
F[t]/(t p − 1). Because p = char F, F[t]/(t − 1) F[t]/(t
p p
). It follows that the
ring End corL/K (M) is isomorphic to the ring R[t]/(t ). We prove (2) by showing
p
that the latter ring does not contain nontrivial idempotents. An arbitrary element of
R[t]/(t p) can be written in the form a + b in a unique way, where a ∈ R and b is
an element of R[t]/(t p) divisible by the class of t. Note that b is nilpotent. Let us
take an idempotent of R[t]/(t p) and write it in the earlier form a + b. Then a is an
idempotent of R. Therefore a = 1 or a = 0. If a = 1 then a + b is invertible and
therefore a + b = 1. If a = 0 then a + b is nilpotent and therefore a + b = 0.
We have proved the implication (3) ⇒ (2). The implication (2) ⇒ (1) is trivial.
We finish by proving that (1) ⇒ (3).
We assume that [E : F] is divisible by a different from p prime q and we show that
(1) does not hold. Indeed, the galois group of E/F contains an element σ of order q.
Let K be the subfield of E consisting of the elements of E fixed by σ . We have F ⊂
K ⊂ E and E/K is galois of degree q. The endomorphisms ring of MK (Spec E) is
isomorphic to F[t]/(t q − 1). As q = char F, the factors of the decomposition t q − 1 =
(t − 1) · (t q−1 + t q−2 + · · · + 1) ∈ F[t] are coprime. Therefore, the ring F[t]/(t q − 1)
254 Nikita A. Karpenko
is the direct product of the rings F[t]/(t − 1) = F and F[t]/(t q−1 + · · · + 1), and it
follows that the motive MK (Spec E) is not indecomposable.
Corollary 3.2. Let E/F be a finite p-primary galois field extension and let L be an
intermediate field: F ⊂ L ⊂ E. Let M ∈ CM(L, F) be an upper indecomposable
motivic summand of an irreducible smooth complete L-variety X. Then corL/F M is
an upper indecomposable summand of the F-variety X.
Before starting the proof of Theorem 1.1, let us recall some classical facts and in-
troduce some notation.
We write D (or DG ) for the set of vertices of the Dynkin diagram of G. The
absolute galois group ΓF of the field F acts on D. The subgroup ΓE ⊂ ΓF is the
kernel of the action.
Let L be a field extension of F. The set DGL is identified with D = DG . The
action of ΓL on D is trivial if and only if the group GL is of inner type. Any ΓL -
stable subset τ in D determines a projective GL -homogeneous variety Xτ ,GL in the
way described in [5, Sect. 3]. This is the variety corresponding to the set D \ τ in
the sense of [6]. For instance, XD ,GL is the variety of the Borel subgroups of GL , and
/ L = SpecL. Any projective GL -homogeneous variety is GL -isomorphic to Xτ ,GL
X0,G
for some ΓL -stable τ ⊂ D.
If the extension L/F is finite separable, we write Mτ ,GL for the upper indecom-
posable motivic summand of the F-variety Xτ ,GL (where τ is a ΓL -stable subset in
D). If L ⊂ E, the isomorphism class of Mτ ,GL is an upper motive of G.
For any field extension L/F, the set DG , attached to the semisimple anisotropic
kernel G of GL , is identified with a (ΓL -invariant) subset in D. We write τL (or τL,G )
for its complement. The subset τL ⊂ D is (the set of circled vertices of) the Tits
index of GL defined in [6]. For any ΓL -stable subset τ ⊂ D, the variety Xτ ,GL has a
rational point if and only if τ ⊂ τL .
Upper Motives of Outer Algebraic Groups 255
Proof (of Theorem 1.1). This is a recast of [5, proof of Theorem 3.5].
We prove Theorem 1.1 simultaneously for all F, G, X using an induction on
n = dim X. The base of the induction is n = 0 where X = Spec F and the statement
is trivial.
From now on we assume that n ≥ 1 and that Theorem 1.1 is already proven for
all varieties of dimension < n.
For any field extension L/F, we write L̃ for the function field L(X ).
Let M be an indecomposable summand of M(X ). We have to show that M is
isomorphic to a shift of Mτ ,GL for some intermediate field L of E/F and some
Gal(E/L)-stable subset τ ⊂ DG containing τF̃ .
Let G /F̃ be the semisimple anisotropic kernel of the group GF̃ . The set DG is
identified with DG \ τF̃,G .
We note that the group GẼ is of inner type. The field extension Ẽ/F̃ is galois
with the galois group Gal(E/F). In particular, its degree is a power of p and every
intermediate field is of the form L̃ for some intermediate field L of the extension
E/F.
According to [1, Theorem 4.2], the motive of XF̃ decomposes into a sum of
shifts of motives of projective GL̃ -homogeneous (where L runs over intermediate
fields of the extension E/F) varieties Y , satisfying dimY < dim X = n. (We are
using the assumption that n > 0 here.) It follows by the induction hypothesis and
Corollary 3.2 that each summand of the complete motivic decomposition of XF̃ is
a shift of Mτ ,G for some L and some τ ⊂ DG . By Corollary 2.2, the complete
L̃
decomposition of MF̃ also consists of shifts of Mτ ,G .
L̃
Let us choose a summand Mτ ,G (i) with minimal i in the complete decomposi-
L̃
tion of MF̃ . We set τ = τ ∪ τF̃ ⊂ DG . We shall show that M Mτ ,GL (i) for these L,
τ , and i.
We write Y for the F-variety Xτ ,GL and we write Y for the F̃-variety Xτ ,G . We
L̃
write N for the F-motive Mτ ,GL and we write N for the F̃-motive Mτ ,G .
L̃
By [5, Lemma 2.14] (also formulated in Sect. 1 here) and since M is indecom-
posable, it suffices to construct morphisms
where the first map is the push-forward with respect to the closed imbedding
Since τF̃ ⊂ τ , the variety X has an F(Y )-point and, therefore, the field extension
F̃(Y )/F(Y ) is purely transcendental. Consequently, the element α2 is F(Y )-rational
and lifts to an element α3 ∈ ChdimY +i (Y × X ). We mean here a lifting with respect
to the composition
resF̃(Y )/F(Y )
ChdimY +i (Y × X ) →→ Chi (XF(Y ) ) −−−−−−→ Chi (XF̃(Y ) ),
where the first map is the epimorphism given by the pull-back with respect to the
morphism XF(Y ) → Y × X induced by the generic point of the variety Y .
We define the morphism α as the composition
α
M(Y )(i) −−−3−→ M(X) −−−−→ M,
where the second map is the projection of M(X ) onto its summand M.
We proceed by constructing β . Let β1 ∈ ChdimY (Y ×F̃ YF̃ ) be the class of the
closure of the graph of a rational map of L̃-varieties Y YF̃ (which exists because
τ ⊂ τF̃ ∪ τ ). Note that this graph is a subset of Y ×L̃ YF̃ , which we consider as a
subset of Y ×F̃ YF̃ via the closed imbedding Y ×L̃ YF̃ → Y ×F̃ YF̃ . Let β2 be the
image of β1 under the composition
ChdimY (Y ×F̃ YF̃ ) = ChdimY M(Y ) ⊗ M(YF̃ ) → ChdimY N ⊗ M(YF̃ ) →
ChdimY +i (M(XF̃ ) ⊗ M(YF̃ )) = ChdimY +i ((X × Y )F̃ ) ,
where the first arrow is induced by the projection M(Y ) → N and the second arrow
is induced by the imbedding N (i) → M(XF̃ ). The element β2 lifts to an element
β3 ∈ ChdimY +i (X × X × Y ).
β4 ∈ ChdimY +i (X × X × Y )
as the image of β4 under the pull-back with respect to the diagonal of X . Finally, we
define the morphism β as the composition
β
M −−−−→ M(X ) −−−5−→ M(Y )(i).
Remark 4.1. Theorem 1.1 can be also proven using a weaker result in place of
[1, Theorem 4.2], namely, [2, Theorem 7.5].
Acknowledgements The author was supported by the Collaborative Research Centre 701 of the
Bielefeld University.
References
1. Brosnan, P.: On motivic decompositions arising from the method of Białynicki-Birula. Invent.
Math. 161(1), 91–111 (2005)
2. Chernousov, V., Gille, S., Merkurjev, A.: Motivic decomposition of isotropic projective homo-
geneous varieties. Duke Math. J. 126(1), 137–159 (2005)
3. Chernousov, V., Merkurjev, A.: Motivic decomposition of projective homogeneous varieties
and the Krull-Schmidt theorem. Transform. Groups 11(3), 371–386 (2006)
4. Elman, R., Karpenko, N., Merkurjev, A.: The algebraic and geometric theory of quadratic
forms, American Mathematical Society Colloquium Publications, vol. 56. American Mathe-
matical Society, Providence, RI (2008)
5. Karpenko, N.A.: Upper motives of algebraic groups and incompressibility of Severi-Brauer
varieties. Linear Algebraic Groups and Related Structures (preprint server) 333 (2009, Apr 3,
revised: 2009, Apr 24), 18 pages
6. Tits, J.: Classification of algebraic semisimple groups. In: Algebraic Groups and Discontinuous
Subgroups (Proc. Sympos. Pure Math., Boulder, Colo., 1965), pp. 33–62. American Mathemat-
ical Society, Providence, R.I. (1966)
Triality and étale algebras
1 Introduction
All Dynkin diagrams but one admit at most automorphisms of order two, which are
related to duality in algebra and geometry. The Dynkin diagram of D4
c α3
c c
α1 α2@
@ c α4
Max-Albert Knus
Departement Mathematik, ETH Zentrum, CH-8092 Zürich, Switzerland,
e-mail: [email protected]
Jean-Pierre Tignol
Institut de Mathématique Pure et Appliquée, Université catholique de Louvain,
B-1348 Louvain-la-Neuve, Belgium, e-mail: [email protected]
2010 Mathematics subject classification. 13B40, 12F10, 12G0.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 259
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 16,
c Springer Science+Business Media, LLC 2010
260 Max-Albert Knus and Jean-Pierre Tignol
Set Γ ≡ É t F (2.1)
(see [7, Proposition (4.3), p. 25] or [12, (18.4)]). Under this antiequivalence, the
cardinality of Γ -sets corresponds to the dimension of étale F-algebras, the disjoint
union in Set Γ corresponds to the product × in É t F , and the product × in Set Γ
n
to the tensor product ⊗ in É t F . For any integer n ≥ 1, we let É t F denote the
groupoid1 whose objects are n-dimensional étale F-algebras and whose morphisms
are F-algebra isomorphisms, and Set Γn the groupoid of Γ -sets with n elements.
n
The antiequivalence (2.1) restricts to an antiequivalence Set Γn ≡ É t F . The split
étale algebra F corresponds to the Γ -set n of n elements with trivial Γ -action.
n
We recall from [12, Sect. 18] and [13, Sect. 2.1] the construction of the discriminant
Δ (X) of a Γ -set X with |X| = n ≥ 2. Consider the set of n-tuples of elements in X :
This Γ -set carries an obvious transitive (right) action of the symmetric group Sn .
The discriminant Δ (X ) is the set of orbits under the alternating group An :
Δ (X ) = Σn (X )/An .
Δ : Set Γn → Set Γ2 .
π
For any covering Z0 ← − Z of degree 2 with |Z0 | = n, (hence |Z| = 2n), we consider
the set of (not necessarily Γ -equivariant) sections of π :
C(Z/Z0 ) = {z1 , . . . , zn } ⊂ Z | {zπ1 , . . . , zπn } = Z0 .
Triality and étale algebras 263
called the Clifford functor (see [14]). The Γ -set C(Z/Z0 ) is equipped with a canon-
ical surjective morphism
δ
Δ (Z) ←
− C(Z/Z0 ), (2.3)
which is defined in [13, Sect. 2.2] as follows: let σ : Z → Z be the involution canon-
π
ically associated to the double covering Z0 ← − Z, so the fiber of zπ is {z, zσ } for
each z ∈ Z; then δ maps each section {z1 , . . . , zn } to the A2n -orbit of the 2n-tuple
(z1 , . . . , zn , zσ1 , . . . , zσn ),
and
Oriented Γ -sets
κ : (Cov Γ )+ → (Cov Γ )+ .
d/n d/n
2/n 2/2n−1
Proposition 2.6. If n is even the functor C : Cov Γ → Cov Γ of (2.4) restricts
to a pair of functors
2/2n−2
C1 , C2 : (Cov Γ )+ → Cov Γ
2/n
.
Moreover, two sections ω and ω of the oriented Γ -covering (Z/Z0 , ∂Z ) lie in the
same set C1 (Z/Z0 , ∂Z ) or C2 (Z/Z0 , ∂Z ) if and only if |ω ∩ ω | ≡ 0 mod 2.
Proof. Let Z/Z0 be a 2/n-covering. Proposition 2.5 implies that the covering
δ
Δ (Z) ←
− C(Z/Z0 ) factors through C(Z/Z0 )/σ , where σ is the canonical involution
of C(Z/Z0 ):
Δ (Z) ← C(Z/Z0 )/σ ← C(Z/Z0 ).
∂
Thus, if Z/Z0 is oriented, we may use the given isomorphism 2 ←−
Z
Δ (Z) to define
the Γ -sets
C1 (Z/Z0 , ∂Z ) = {ω ∈ C(Z/Z0 ) | ω δ ∂Z = 1}
and
We call the two functors C1 and C2 the spinor functors. Note that when n is even
an orientation ∂Z on Z/Z0 can also be defined by specifying whether a given section
ω ∈ C(Z/Z0 ) lies in C1 (Z/Z0 , ∂Z ) or C2 (Z/Z0 , ∂Z ). Indeed, ω ∈ C1 (Z/Z0 , ∂Z ) if and
only if ω δ ∈ Δ (Z) is mapped to 1, which determines ∂Z uniquely. We shall avail
2/4
ourselves of this possibility to define orientations on coverings in Cov Γ in Sect. 4.
Triality and étale algebras 265
We now consider analogues of the functors Δ and C for étale algebras and étale
extensions. For S an étale F-algebra of dimension n ≥ 2, the discriminant Δ (S) is a
quadratic étale F-algebra such that
X Δ (S) = Δ X(S) .
We thus have a functor
n 2
Δ : É t F → É t F for n ≥ 2.
We call the 2n -dimensional algebra C(T /S) the Clifford algebra of T /S. It admits a
canonical involution σ . If dimF S is even σ is the identity on Δ (T ). The canonical
morphism δ of (2.3)
δ
Δ X(T ) ← − C X(T )/X(S)
yields a canonical F-algebra homomorphism which we again denote by δ ,
δ
Δ (T ) −
→ C(T /S),
so that C(T /S) is an étale extension of degree 2n−1 of a quadratic étale F-algebra.
As for oriented Γ -sets we define oriented étale algebras as pairs (S, ∂S ) where S
∼
is an étale algebra and ∂S : Δ (S) −
→ F × F is an isomorphism of F-algebras. Ori-
ented extensions of étale algebras are pairs (S/S0, ∂S ) where S/S0 is an extension of
∼
étale algebras and ∂S : Δ (S) −
→ F × F is an isomorphism of F-algebras. We have
n d/n
corresponding groupoids (É t F )+ , (É tex F )+ and antiequivalences
n d/n
(Set Γn )+ ≡ (É t F )+ and (Cov Γ )+ ≡ (É tex F )+ .
d/n
if n is even.
Remark 2.9. The terminology used earlier owes its origin to the fact that the Clifford
functor is related to the theory of Clifford algebras in the framework of quadratic
forms and central simple algebras with involution. We refer to [14] for details and
more properties of the Clifford construction.
3 Cohomology
For any integer n ≥ 1, we consider the Γ -set n = {1, . . . , n} with the trivial Γ -action
and let Sn denote the symmetric group on n, i.e., the automorphism group of n,
Sn = Aut(n).
Recall from [12, Sect. 28.A] that the cohomology set H 1 (Γ , Sn ) (for the trivial ac-
tion of Γ on Sn ) is the set of continuous group homomorphisms Γ → Sn (“cocy-
cles”) up to conjugation.
Letting Iso(Set Γn ) denote the set of isomorphism classes in Set Γn , we have a
canonical bijection of pointed setsv
∼
Iso(Set Γn ) −
→ H 1 (Γ , Sn ). (3.1)
Aut(d/n) = Sd Sn (= Snd Sn ).
where the Γ -action on Sd Sn is trivial; see [13, Sect. 4.2]. The automorphism
group of the oriented Γ -covering (d/n, ∂n×d ) is the group
so that d/n ∼
Iso (Cov Γ )+ −→ H 1 Γ , (Sd Sn )+ . (3.3)
We now assume that Γ is the absolute Galois group Γ = Gal(Fs /F) of a field F and
n
use the notation H 1 (F, Sn ) for H 1 (Γ , Sn ). The antiequivalence Set Γn ≡ É t F and
the bijection (3.1) induce canonical bijections
Iso(É t F ) ∼
= Iso(Set Γn ) ∼
n
= H 1 (F, Sn )
AutF-alg (F n ) ∼
= Sn ,
see [12, (29.9)]. Similarly, it follows from (3.2), (3.3), and the antiequivalence
d/n d/n
of groupoids É tex F ≡ Cov Γ , (É tex F )+ ≡ (Cov Γ )+ , that we have canonical
d/n d/n
Iso(É tex Γ ) ∼
= Iso(Cov Γ ) ∼
d/n d/n
= H 1 (F, Sd Sn ) (3.4)
and
d/n d/n
Iso (É tex Γ )+ ∼= Iso (Cov Γ )+ ∼= H 1 F, (Sd Sn )+ . (3.5)
Recall the functor C, which associates to any double covering its set of sections. For
oriented 2/4-coverings of Γ -sets, it leads to two functors:
C1 , C2 : (Cov Γ )+ → Cov Γ ,
2/4 2/4
see Proposition 2.6. The functors C1 and C2 together with the functor κ , which
changes the orientation, give an explicit description of an action of the group S3
2/4
on the pointed set Iso (Cov Γ )+ .
268 Max-Albert Knus and Jean-Pierre Tignol
Theorem 4.1. The functors C1 , C2 : (Cov Γ )+ → Cov Γ factor through the forget-
2/4 2/4
Proof. Let (Z/Z0 , ∂ ) be an object in (Cov Γ )+ and let σ denote the involution of
2/4
Z/Z0 . Consider a real vector space V with basis (e1 , e2 , e3 , e4 ). Fixing a bijection ϕ
between a section ω ∈ C1 (Z/Z0 , ∂ ) and {e1 , . . . , e4 }, we identify Z with a subset of
V by:
zϕ if z ∈ ω ,
z → σ ϕ
−z if z ∈ / ω.
Thus, Z = {±e1 , ±e2 , ±e3 , ±e4 } and σ acts on Z by mapping each element to its
opposite. The action of Γ on Z extends to a linear action on V because it commutes
with σ . We also identify C(Z/Z0 ) with a subset of V by the map
ω → 1
2 ∑ z.
z∈ω
The set C(Z/Z0 ) then consists of the following vectors and their opposite:
C1 (Z/Z0 , ∂ ) = {± f1 , ± f2 , ± f3 , ± f4 }
and
The canonical involutions on C1 (Z/Z0 , ∂ ) and C2 (Z/Z0 , ∂ ) map each vector to its
opposite. Note that these identifications are independent of the choice of the sec-
tion ω in C1 (Z/Z0 , ∂ ) and of the bijection ϕ .
Let ∂1 be the orientation of C1 (Z/Z0 , ∂ ) such that C1 C1 (Z/Z0 , ∂ ), ∂1 contains
the section { f1 , f2 , f3 , f4 }. Thus, by Proposition 2.6, C1 C1 (Z/Z0 , ∂ ), ∂1 consists
of the following sections:
±{ f1 , f2 , f3 , f4 }, ±{ f1 , f2 , − f3 , − f4 }, ±{ f1 , − f2 , f3 , − f4 }, ±{− f1 , f2 , f3 , − f4 }.
Triality and étale algebras 269
They are characterized by the property that for each i = 1, . . . , 4 they contain a sec-
tion ± f j containing ei and a section ± fk containing −ei . Identifying these sections
to vectors in V as earlier, we obtain
C1 C1 (Z/Z0 , ∂ ), ∂1 = {± 12 ( f1 + f2 + f3 + f4 ), ± 12 ( f1 + f2 − f3 − f4 ),
± 12 ( f1 − f2 + f3 − f4 ), ± 12 (− f1 + f2 + f3 − f4 )}.
C1+ (Z/Z0 , ∂ ) = C1 (Z/Z0 ), ∂1 and C2+ (Z/Z0 , ∂ ) = C2 (Z/Z0 ), ∂2 .
C(Z/Z0 ) = { 12 (±e1 ± e2 ± e3 ± e4 )}
8 a5
a
B
aH
1
B
C a4
HH6 B a a
Ha a 2
3
7
A B
B A D
a4 a 1
5 aa HHH
8
a 7 B H
a6
D C B
a
Ba
2 3
Fig. 1 Hypercube
We identify
Z = {A, Ā, B, B̄,C, C̄, D, D̄},
where A, . . . , C̄ are as in Fig. 1, D is the big cell and D̄ the small cell inside. The invo-
lution permutes a cell with its opposite cell and the set Z0 is obtained by identifying
pairs of opposite cells
opposite. Four such cells intersect in exactly one vertex and conversely each vertex
lies in four cells. With the notation in Fig. 1 we have the following identification:
1 = {A, B̄,C, D} 1̄ = {Ā, B, C̄, D̄} 2 = {A, B̄, C̄, D} 2̄ = {Ā, B,C, D̄}
3 = {A, B, C̄, D} 3̄ = {Ā, B̄,C, D̄} 4 = {A, B,C, D} 4̄ = {Ā, B̄, C̄, D̄}
5 = {Ā, B,C, D} 5̄ = {A, B̄, C̄, D̄} 6 = {Ā, B, C̄, D} 6̄ = {A, B̄,C, D̄}
7 = {Ā, B̄, C̄, D} 7̄ = {A, B,C, D̄} 8 = {Ā, B̄,C, D} 8̄ = {A, B, C̄, D̄}
This set of vertices decomposes into two classes, two vertices being in the same
class if the number of edges in any path connecting them is even. One class is
A = {1, 3, 5̄, 7̄} = {2, 4, 6̄, 8̄} Ā = {1̄, 3̄, 5, 7} = {2̄, 4̄, 6, 8}
B = {1̄, 3, 5, 7̄} = {2̄, 4, 6, 8̄} B̄ = {1, 3̄, 5̄, 7} = {2, 4̄, 6̄, 8}
(4.3)
C = {1, 3̄, 5, 7̄} = {2̄, 4, 6̄, 8} C̄ = {1̄, 3, 5̄, 7} = {2, 4̄, 6, 8̄}
D = {1, 3, 5, 7} = {2, 4, 6, 8} D̄ = {1̄, 3̄, 5̄, 7̄} = {2̄, 4̄, 6̄, 8̄}
and
1 = {A, B̄,C, D} = {2, 4, 6̄, 8} 1̄ = {Ā, B, C̄, D̄} = {2̄, 4̄, 6, 8̄}
3 = {A, B, C̄, D} = {2, 4, 6, 8̄} 3̄ = {Ā, B̄,C, D̄} = {2̄, 4̄, 6̄, 8}
5 = {Ā, B,C, D} = {2̄, 4, 6, 8} 5̄ = {A, B̄, C̄, D̄} = {2, 4̄, 6̄, 8̄}
7 = {Ā, B̄, C̄, D} = {2, 4̄, 6, 8} 7̄ = {A, B,C, D̄} = {2̄, 4, 6̄, 8̄}
(4.4)
2 = {A, B̄, C̄, D} = {1, 3, 5̄, 7} 2̄ = {Ā, B,C, D̄} = {1̄, 3̄, 5, 7̄}
4 = {A, B,C, D} = {1, 3, 5, 7̄} 4̄ = {Ā, B̄, C̄, D̄} = {1̄, 3̄, 5̄, 7}
6 = {Ā, B, C̄, D} = {1̄, 3, 5, 7} 6̄ = {A, B̄,C, D̄} = {1, 3̄, 5̄, 7̄}
8 = {Ā, B̄,C, D} = {1, 3̄, 5, 7} 8̄ = {A, B, C̄, D̄} = {1̄, 3, 5̄, 7̄},
hence the existence of decompositions C(X/X0 ) = Y /Y0 Z/Z0 and C(Y /Y0 ) =
Z/Z0 X/X0 which, in fact, are decompositions as Γ -sets.
Remark 4.5. In the proof of Theorem 4.1, μ is not the only linear map that can be
used to describe the C1+ and the C2+ construction. An alternative description uses
272 Max-Albert Knus and Jean-Pierre Tignol
Hurwitz’ quaternions. Choosing for V the skew field of real quaternions H and for
(e1 , e2 , e3 , e4 ) the standard basis (1, i, j, k), we have
H 1 = Q 8 C3 ,
where the cyclic group of three elements C3 operates on Q8 via conjugation with ρ .
+
If ∂ is the orientation of Z/Z0 such that ρ ∈ C1 (Z/Z0 , ∂ ), we have C+1 (Z/Z0 , ∂ ) =
+
ρ · Z with the orientation such that ρ ∈ C1 C1 (Z/Z0 , ∂ ) , and C2 (Z/Z0 , ∂ ) =
2
(C1+ )2 (Z/Z0 , ∂ ) = ρ 2 · Z with the orientation such that 1 ∈ C1 C2+ (Z/Z0 , ∂ ) . Note
that, with respect to the standard basis, multiplication by ρ is given by the matrix
⎛ ⎞
−1 1 1 1
1 ⎜−1 −1 1 −1⎟
ρ= ⎜ ⎟ (4.6)
2 ⎝−1 −1 −1 1 ⎠
−1 1 −1 −1
has the permutation group S3 as a group of automorphisms. The vertices of the dia-
gram are labeled by the simple roots of the Lie algebra of type D4 . Let (e1 , e2 , e3 , e4 )
be the standard orthonormal basis of the Euclidean space R4 . The simple roots are
α1 = e1 − e2 , α2 = e2 − e3 , α3 = e3 − e4 , and α4 = e3 + e4
Triality and étale algebras 273
is the Weyl group of the split adjoint algebraic group PGO+ 8 , which is of type D4 .
As a subgroup of the orthogonal group O4 , the group W (D4 ) is generated by the
reflections with respect to the roots of the Lie algebra of PGO+ 8 . Elements of S2 S4
can be written as matrix products
w = D · P(π ), (5.3)
where D is the diagonal matrix diag(ε1 , ε2 , ε3 , ε4 ), εi = ±1, and P(π ) is the permu-
tation matrix of π ∈ S4 . The group S2 S4 fits into the exact sequence
β
1 → S42 → S2 S4 → S4 → 1, (5.4)
Automorphisms of W (D4 )
Thus, W (D4 )/w0 acts on W (D4 ) as the group of inner automorphisms. Let ψ be
the automorphism of order 2 of W (D4 ) given by:
Corollary 5.7.
Aut W (D4 ) / Int W (D4 ) S3 × ψ .
Proposition
5.8. Any trialitarian automorphism of W (D4 ) is conjugate in the group
Aut W (D4 ) to either μ̃ or ρ̃ .
Proof. Explicit computation (with the help of the algebra computational system
Magma [2]) shows that the conjugation class of ρ̃ contains 16 elements and the
conjugation class of μ̃ contains 32 elements. In view of Proposition 5.6 any tri-
alitarian automorphism of W (D4 ) is the restriction to W (D4 ) of conjugation by an
element u ∈ O4 of the form u = μ · w or u = μ 2 · w, with w ∈ W (D4 ) and u3 = 1.
There are 48 elements u of this form, hence the claim.
Corollary 5.9. There are up to isomorphism two types of subgroups of fixed points
of trialitarian automorphisms of W (D4 ), those isomorphic to Fix(μ̃ ) and those iso-
morphic to Fix(ρ̃ ) .
Proof. Trialitarian automorphisms which are conjugate in Aut W (D4 ) have iso-
morphic groups of fixed points.
Triality and étale algebras 275
Proposition 5.11. The group Fix(ρ̃ ) is isomorphic to the group of order 24 of Hur-
witz quaternions H1 = Q8 C3 . This group is isomorphic to the double covering
Ã4 of A4 .
α∗ : [ϕ ] → [α ◦ ϕ ],
4 The corresponding Γ -set λ 2 X is obtained by removing the diagonal from X × X and dividing by
the involution (x, y) → (y, x).
Triality and étale algebras 277
Observe that fixed étale algebras in class N = 61 are of the form K × (E0 ⊗ K),
where K is quadratic and E0 is cubic. Hence they are not fields over F, in contrast
to algebras in class N = 85.
278 Max-Albert Knus and Jean-Pierre Tignol
Table 1 List of isomorphism classes of étale algebras corresponding to the conjugacy classes of
subgroups of W (D4 ), together with a description of the triality action
N S0 S G1 → G → G0 |G| MS T
1 F4 F8 1→1→1 1 1 1 1, 1
2 F4 K4 C → S2 → 1 2 6 1 2, 2
3 K2 K4 1 → S 2 → S 2 ⊂ V4 2 6 1 3, 5
4 K2 K4 1 → S 2 → S 2 ⊂ V4 2 6 1 4, 3
5 F4 F 2 × K2 w1 → S2 → 1 2 6 1 5, 4
6 F2 × K F 2 × K2 1 → S2 → S2 ⊂ V4 2 12 1 6, 6
7 F2 × K F 2 × K2 1 → S2 → S2 ⊂ V4 2 12 1 7, 7
8 F × E0 F 2 × S02 1 → C3 → C3 3 16 1 8, 8
9 F4 K12 × K22 C, w1 → S22 → 1 4 3 25 11, 10
10 K2 K ⊗ K12 C → S22 → S2 ⊂ V4 4 3 23 9, 11
11 K2 K ⊗ K12 C → S22 → S2 ⊂ V4 4 3 24 10, 9
12 K1 ⊗ K2 K1 ⊗ K22 1 → S22 → V4 4 4 4 14, 13
13 F4 K12 × K22 w1 , w2 → S22 → 1 4 4 5 12, 14
14 K1 ⊗ K2 K1 ⊗ K22 1 → S22 → V4 4 4 3 13, 12
15 F2 × K F 4 × K1 ⊗ K w1 → S22 → S2 ⊂ V4 4 6 56 18, 17
16 K1 × K2 K1 ⊗ K22 1 → S22 → S22 4 6 47 21, 19
17 K1 × K2 K12 × K22 1 → S22 → S22 4 6 36 15, 18
18 K1 × K2 K12 × K22 1 → S22 → S22 4 6 46 17, 15
19 F2 × K K12 × K1 × K w1 → S22 → S2 ⊂ V4 4 6 57 16, 21
20 K2 E2 C → C 4 → S 2 ⊂ V4 4 2 2 20, 20
21 K1 × K2 K1 ⊗ K22 1 → S22 → S22 4 6 37 19, 16
22 K1 × K2 K12 × K1 ⊗ K2 1 → S22 → S22 4 12 467 23, 27
23 K1 × K2 K12 × K1 ⊗ K2 1 → S22 → S22 4 12 367 27, 22
24 K2 K 2 × K ⊗ K1 w1 → S22 → S2 ⊂ V4 4 12 345 24, 24
25 F2 × K F4 × E w1 → C4 → S2 ⊂ V4 4 12 5 26, 28
26 E0 E02 1 → C4 → C4 4 12 4 28, 25
27 F2 × K K12 × K1 ⊗ K −w1 → S22 → S2 ⊂ V4 4 12 567 22, 23
28 E0 E0 × E0 1 → C4 → C4 4 12 3 25, 26
29 F2 × K K 2 × K ⊗ K1 C → S22 → S2 ⊂ V4 4 12 267 29, 29
30 F × E0 F 2 × E0 ⊗ Δ (E0 ) 1 → S3 → S3 6 16 78 30, 30
31 F × E0 F2 × E C → C6 → C3 6 16 28 31, 31
32 F × E0 F 2 × E02 1 → S3 → S3 6 16 68 32, 32
33 E0 E C → S32 → V4 8 1 11 12 34, 35
34 E0 E C → S32 → V4 8 1 10 14 35, 33
35 F4 K1 × K2 × K3 × K123 S32 → S32 → 1 8 1 9 13 33, 34
36 E0 E C → Q 8 → V4 8 2 20 36, 36
37 K2 E1 × E2 C, w1 → S2 × C4 → S2 ⊂ V4 8 3 3 20 38, 40
38 E0 E0 ⊗ K C → S2 × C4 → C4 8 3 11 20 40, 37
39 K2 K ⊗ K1 × K ⊗ K2 C, w1 → S32 → S2 ⊂ V4 8 3 9 10 11 24 39, 39
40 E0 E0 ⊗ K C → S2 × C4 → C4 8 3 10 20 37, 38
41 K2K E2 C, w2 → D4 → S2 ⊂ V4 8 6 9 11 20 49, 45
42 F2 × K K1 × E C, w1 → S2 × C4 → S2 ⊂ V4 8 6 9 25 44, 47
43 K1 × K2 K ⊗ K1 × K ⊗ K2 C → S32 → S22 8 6 10 17 21 23 29 48, 46
44 E0 E C → S2 × C4 → C4 8 6 11 26 47, 42
45 K2 E2 C, w2 → D4 → S2 ⊂ V4 8 6 9 10 20 41, 49
46 K1 × K2 K ⊗ K1 × K ⊗ K2 C → S32 → S22 8 6 11 16 18 22 29 43, 48
47 E0 E0 ⊗ K C → S2 × C4 → C4 8 6 10 28 42, 44
(continued)
Triality and étale algebras 279
Table 1 (continued)
N S0 S G1 → G → G0 |G| MS T
48 F2 × K K12 × K ⊗ K2 C, w1 → S32 → S2 ⊂ V4 8 6 9 15 19 27 29 46, 43
49 E0 E C → D4 → V 4 8 6 10 11 20 45, 41
50 F2 × K K2 × E w1 , w2 → D4 → S2 ⊂ V4 8 6 13 19 25 55, 51
51 E0 E 1 → D4 → D 4 8 12 14 21 28 50, 55
52 E0 E02 1 → D4 → D 4 8 12 14 17 28 54, 57
53 K1 × K2 K ⊗ K1 × K ⊗ K2 w1 → S32 → S22 8 12 16 19 21 22 23 24 27 53, 53
54 F2 × K F2 × K × E w1 , w2 → D4 → S2 ⊂ V4 8 12 13 15 25 57, 52
55 E0 Ē0 1 → D4 → D 4 8 12 12 16 26 51, 50
56 K1 × K2 K12 × E w1 → S32 → S22 8 12 15 17 18 22 23 24 27 56, 56
57 E0 E02 1 → D4 → D 4 8 12 12 18 26 52, 54
58 E0 E02 1 → A4 → A4 12 4 8 14 59, 60
59 F × E0 F 2 × E02 w1 , w2 → S22 C3 → C3 12 4 8 13 60, 58
60 E0 E02 1 → A4 → A4 12 4 8 12 58, 59
61 F × E0 K × K ⊗ E0 C → S2 × S3 → S3 12 4 29, 30, 31, 32 61, 61
62 K1 × K2 E2 C, w1 → [22 ]4 → S22 16 3 39 42 66, 63
63 E0 E C, w1 → [22 ]4 → C4 16 3 39 47 62, 66
64 K2 E1 × E2 S32 → S2 × D4 → S2 ⊂ V4 16 3 35 37 39 41 45 68, 67
65 K1 × K2 K1 ⊗ K3 × K2 ⊗ K4 C, w1 → S42 → S22 16 3 39 43 46 48 53 56 65, 65
66 E0 E C, w1 → [22 ]4 → C4 16 3 39 44 63, 62
67 E0 E C, w1 → S2 × D4 → V4 16 3 34 39 40 45 49 64, 68
68 E0 E C, w2 → S2 × D4 → V4 16 3 33 38 39 41 49 67, 64
69 K1 × K2 E2 C, w1 → [22 ]4 → S22 16 6 37 42 48 73, 72
70 E0 E C, w1 → Q8 : 2 → V4 16 6 36 37 38 40 41 45 49 70, 70
71 F2 × K K12 × E S32 → S2 × D4 → S2 ⊂ V4 16 6 35 42 48 50 54 75, 74
72 E0 E C, w1 → [22 ]4 → V4 16 6 40 43 47 69, 73
73 E0 E C → [22 ]4 → D4 16 6 38 44 46 72, 69
74 E0 E0 ⊗ K C → S 2 × D4 → D 4 16 6 34 43 47 51 52 71, 75
75 E0 E0 ⊗ K C → S 2 × D4 → D 4 16 6 33 44 46 55 57 74, 71
76 E0 E02 1 → S4 → S4 24 4 32 52 58 79, 81
77 F × R(E) Δ (E) × λ 2 E w1 , w3 → S22 S3 → S3 24 4 30 50 59 84, 82
78 F × E0 Δ (E) × λ 2 E S32 → S32 C3 → C3 24 4 31 35 59 83, 80
79 F × R(E) F 2 × λ 2E w1 , w3 → S22 S3 → S3 24 4 32 54 59 76, 81
80 E0 E0 ⊗ K C → S2 × A4 → A4 24 4 31 34 58 78, 83
81 E0 E02 1 → S4 → S4 24 4 32 57 60 76, 79
82 E0 E0 ⊗ Δ (E0 ) 1 → S4 → S4 24 4 30 51 58 77, 84
83 E0 E C → S2 × A4 → A4 24 4 31 33 60 78, 80
84 E0 E0 ⊗ Δ (E0 ) 1 → S4 → S4 24 4 30 55 66 82, 77
85 E0 E S2 → Ã4 → A4 24 4 31 36 85, 85
86 E0 E S32 → S32 V4 → V4 32 1 64 67 68 70 86, 86
87 E0 E C, w1 → S22 D4 → D4 32 3 63 65 67 72 74 92, 90
88 E0 E S32 → S32 C4 → C4 32 3 63 64 66 91, 89
89 E0 E C, w1 → S32 C4 → D4 32 3 62 66 67 88, 91
90 E0 E C, w1 → S22 D4 → D4 32 3 65 66 68 73 75 92, 87
91 E0 E C, w1 → S32 C4 → D4 32 3 62 63 68 89, 88
92 K1 × K2 E1 × E2 S32 → S22 D4 → S22 32 3 62 64 65 69 71 87, 90
93 F × R(E) Δ (E) × K ∗ λ 2 E S32 → S2 × S4 → S3 48 4 61 71 77 78 79 94, 95
94 E0 E0 ⊗ K C → S2 × S4 → S4 48 4 61 75 81 83 84 95, 93
95 E0 E0 ⊗ K C → S2 × S4 → S4 48 4 61 74 76 80 82 93, 94
96 E0 E S32 → S32 D4 → D4 64 3 86 87 88 89 90 91 92 96, 96
97 E0 E S32 → S32 A4 → A4 96 1 78 80 83 85 86 97, 97
98 E0 E S32 → S32 S4 → S4 192 1 93 94 95 96 97 98, 98
280 Max-Albert Knus and Jean-Pierre Tignol
7 Trialitarian Resolvents
Trialitarian triples of étale algebras can be viewed as one étale algebra with two
attached resolvents (see Remark 3.6). For example, let E be a quartic separable
field with Galois group S4 . The field E ⊗ Δ (E) is octic with the same Galois group
S4 and the extension E ⊗ Δ (E)/E corresponds to Class N = 82 in Table 1. Class
N = 77 in the same trialitarian triple corresponds to the extension
Δ (E) × λ 2E / F × R(E) ,
where Δ (E) is the discriminant, R(E) is the cubic resolvent of E and λ 2 E is the
second lambda power of E, as defined in [13].
In this section we consider the situation where one étale algebra in the triple is
given by a separable polynomial and compute polynomials for the two other étale
algebras. We assume that the base field F is infinite and has characteristic different
from 2.
Proposition 7.1. Let S/S0 be an étale algebra with involution of dimension 2n
over F.
1) There exists an invertible element x ∈ S such that x generates S and x2 gener-
ates S0 .
2) There exists a polynomial
Proof. To prove (1) we are looking for invertible elements x of S such that
TrS/S0 (x) = 0 and such that the discriminant of the characteristic polynomial of
x2 is not zero. Any such element generates S and x2 ∈ S0 generates S0 . These ele-
ments form a Zariski open subset of the space of trace zero elements. One checks
that this open subspace is not empty by going to an algebraic closure of F.
(2) follows from (1) and (3) follows from a discriminant formula (see [4, p. 51]) for
the discriminant D( f2n ) of f2n :
2
D( f2n ) = (−1)n a0 · 2n D( fn )
∼
(recall that Δ (S) −
→ F[x]/ x2 − D( f2n ) ).
Theorem 7.2. Let S/S0, dimF S = 8, with trivial discriminant, be given as in Propo-
sition 7.1 by a polynomial
f4 (x) = x4 + ax3 + bx2 + cx + e2. (7.3)
Triality and étale algebras 281
The polynomials
define extensions of étale algebras S /S0 , S /S0 such that the isomorphism classes
of S/S0, S /S0 , and S /S0 are in triality.
Proof. Let {y1 , y2 , y3 , y4 } be the set of zeroes of f4 in a separable closure Fs of F.
√
The set {±xi = ± yi , i = 1, . . . 4} is the set of zeroes of f8 . Let ξ be the column
vector [x1 , x2 , x3 , x4 ]T . If ϕ : Γ → W (D4 ) ⊂ O4 is the cocycle corresponding to S/S0,
the group ϕ (Γ ) permutes the elements ±xi through left matrix multiplication on ξ .
The cocycle corresponding to S /S0 is given by:
ϕ : γ → μϕ (γ )μ −1 , γ ∈ Γ ,
f8 (x) = f4 (x2 ) = ∏(x − xi )(x + xi ) = ∏(x2 − xi ).
2
i i
The xi are the components of ξ = μξ . Thus, the coefficients of f8 (x) can be ex-
pressed as functions of the xi . Using that the symmetric functions in the xi can be
expressed as functions of the coefficients of f8 one gets (for example with Magma
[2]) the expression given in Proposition 7.2 for fi . Similar computations with μ 2
instead of μ lead to the formula for f4 .
Remark 7.4. Observe that we move from f8 to f8 by replacing e by −e, as it
should be.
The results of this section were communicated to us by Serre [16]. They are based
on results of [15] and [10]. Similar results can be obtained for cohomological in-
variants of étale algebras instead of Witt invariants. Let k be a fixed base field
of characteristic not 2 and F/k be a field extension. Let W Gr(F ) be the Witt-
Grothendieck ring and W (F) the Witt ring of F, viewed as functors of F. We recall
that elements of W Gr(F) are formal differences q − q of isomorphism classes of
nonsingular quadratic forms over F and that the sum and product are those induced
by the orthogonal sum and the tensor product of quadratic forms. The Witt ring
W (F) is the quotient of W Gr(F) by the ideal consisting of integral multiples of the
2-dimensional diagonal form 1, −1.
282 Max-Albert Knus and Jean-Pierre Tignol
Some of the following considerations hold for oriented quadratic extensions S/S0
of étale algebras of arbitrary dimension. To simplify notation we assume from now
on that dimF S = 8.
2/4 2/4
Let (É tex )+ be the functor which associates to F the set (É tex F )+ of iso-
morphism classes of oriented quadratic extensions S/S0 of étale algebras over F
2/4
such that dimF S = 8. A Witt invariant on (É tex )+ , more precisely on W (D4 ), is
a map
H 1 F,W (D4 ) → W (F)
for each F/k, subject to compatibility and specialization conditions (see [10]). The
set of Witt invariants
2/4
Inv W (D4 ),W = Inv (É tex )+ ,W
is a module over W (k). The aim of this section is to describe this module and how
triality acts on it. A main tool is the following splitting principle, which is a special
case of a variant of the splitting principle for étale algebras (see [10, Theorem 24.9])
and which can be proved following the same lines.
2/4
Theorem 8.1. If a ∈ Inv (É tex )+ ,W satisfies a(S/S0) = 0 for every product of
two biquadratic algebras
√ √ √ √ √ √
S = F x, y × F z, t , S0 = F xy × F zt .
over every extension F of k, then a = 0.
is injective.
The decomposition
S = Sym(S, σ ) ⊕ Skew(S, σ )
Triality and étale algebras 283
Q = Q + ⊥ Q− , Q = Q+ ⊥ −Q− ,
hence the forms Q+ and Q− define two Witt invariants attached to S/S0. The étale
algebras associated to S/S0 by triality lead to corresponding invariants. We intro-
duce the following notation: S/S0 = S1 /S0,1 and Si /S0,i , i = 2, 3, for the associ-
ated étale algebras. We denote the corresponding Witt invariants by Q+ +
i = QSi and
− −
Qi = QSi , i = 1, 2, and 3.
Another construction of Witt invariants is through orthogonal representations.
Let On be the orthogonal group of the n-dimensional form 1, . . . , 1. Quadratic
forms over F of dimension n are classified by the cohomology set H 1 (F, On ). Thus
any group homomorphism W (D4 ) → On gives rise to a Witt invariant. In particular
we get a Witt invariant q associated with the orthogonal representation W (D4 ) → O4
described in (5.3). Moreover the group W (D4 ) has three normal subgroups Hi of
type (2, 2, 2) (i.e., isomorphic to S32 ), corresponding to the classes N = 33, 34, 35
of Table 1. As the factor groups are isomorphic to S4 , the canonical representation
S4 → O4 through permutation matrices leads to three Witt invariants q1 , q2 , q3 .
Proposition 8.3. 1) The Witt invariant q is invariant under triality and coincides
with Q− i , i = 1, 2, 3.
2) We have qi = Q+ i , i = 1, 2, 3, and the three invariants q1 , q2 , q3 are permuted by
triality.
Proof. The fact that q is invariant under triality follows from the fact that triality
acts on W (D4 ) by an inner automorphism of O4 . Moreover the trialitarian action
on W (D4 ) permutes the normal subgroups Hi , hence the invariants qi . For the other
claims we may assume by the splitting principle that S1 is a product of two bi-
quadratic algebras
√ √ √ √ √ √
S1 = F x, y × F z, t , S0,1 = F xy × F zt . (8.4)
An explicit computation, using for example the description of triality given in the
proof of Theorem 4.1 (see [1] and [20]) shows that one can make the following
identifications
√ √ √ √ √ √
S2 = F x, z × F y, t , S0,2 = F xz × F yt
and √ √ √ √ √ √
S3 = F x, t × F y, z , S0,3 = F xt × F yz .
Observe that, with this identification, the three-cycle (y, z,t) permutes cyclically the
algebras Si /S0,i . We get
Q−
i = x, y, z,t
for i = 1, 2, 3, and
Q+
1 = 1, 1, xy, zt, Q+
2 = 1, 1, xz, yt, Q+
3 = 1, 1, xt, yz.
284 Max-Albert Knus and Jean-Pierre Tignol
The equalities q = Q− +
i and qi = Qi follow from the fact that the corresponding
cocycles are conjugate in O4 .
Further basic invariants are the constant invariant 1 and the discriminant
d = Disc(q) = Disc(qi ), i = 1, 2, 3,
which corresponds to the 1-dimensional representation det : W (D4 ) → O1 = ±1.
Since Q+ +
i (1) = 1, the quadratic forms qi = Qi represent 1 and one can replace them
⊥
by 3-dimensional invariants i = (1) ⊂ qi , i = 1, 2, 3.
In the following result λ 2 q denotes the second exterior power of the quadratic
form q (see [10]). If q = α1 , . . . , αn is diagonal, then λ 2 q is the n(n − 1)/2-
dimensional form λ 2 q = α1 α2 , . . . , αn−1 αn .
2/4
Theorem 8.5. 1) The W (k)-module Inv W (D4 ),W = Inv (É tex )+ ,W is free
over W (k) with basis
2) The elements 1, d, q, d · q are fixed under triality and the elements 1 , 2 ,
and 3 are permuted.
Proof. 1) The proof follows the pattern of the proof of [10, Theorem 29.2]. Let G be
an elementary subgroup of W (D4 ) of type (2, 2, 2, 2), i.e., isomorphic to S42 . An ar-
4
bitrary element of H 1 (F, G) is given by a four-tuple (α1 , α2 , α3 , α4 ) ∈ F × /F ×2 .
For I a subset of 4 = {1, 2, 3, 4}, we write αI for the product of the αi for i ∈ I. By
[10, Theorem 27.15] the set Inv(G,W ) is a free W (k)-module with basis (αI )I⊂4 . It
then follows from Theorem 8.2 that the family of elements given in (8.6) is linearly
2/4
independent over W (k). Let a be an element of Inv (É tex )+ ,W and let Sα be
the algebra (8.4) for α1 = x, α2 = y, α3 = z, α4 = t. The map α → a(Sα ) is a Witt
invariant of G, hence by [10, Theorem 27.15] can be uniquely written as a linear
combination
∑ cI · αI with cI ∈ W (k). (8.7)
The claim will follow if we show that the invariant α is in fact a linear combi-
nation of the elements
given in (8.6). By [10, Prop. 13.2], the image of the re-
striction map Inv W (D4 ),W → Inv(G,W ) is contained in the W (k)-submodule of
Inv(G,W ) fixed by the normalizer H = S32 D4 of G in W (D4 ) (conjugacy class
N = 96 in Table 1). The group H acts on the set of √isomorphisms classes
√ of algebras
√ √
Si by acting in the obvious way on the symbols ± x, ± y, ± z, ± t. It follows
that H acts trivially on this set of isomorphisms classes. This shows that only linear
combinations of elements in the family
Triality and étale algebras 285
B = {1, x + y + z + t, xy + zt, xz + yt, xt + yz,
xyz + xyt + xzt + yzt, xyzt}
can occur in the sum (8.7). The family B and the family given in (8.6) are equiv-
alent bases. This implies the first claim of Theorem 8.5. Claim 2) follows from
Proposition 8.3 and 3) is easy to check for a product of biquadratic extensions.
Acknowledgements We are grateful to Parimala for her unshakable interest in triality, in particu-
lar for many discussions at earlier stages of this work and we specially thank Serre for communi-
cating to us his results on Witt and cohomological invariants of the group W (D4 ). We also thank
Emmanuel Kowalski who introduced us to Magma [2] with much patience, Jean Barge for his help
with Galois cohomology and Humphreys and Mühlherr for the reference to the paper [9]. The
paper [11] on octic fields was a very useful source of inspiration. Finally, we are highly thankful
to the referee for many improvements.
References
1. L. Beltrametti Über quadratische Erweiterungen étaler Algebren der Dimension vier. Diplo-
marbeit, Mathematikdepartement, ETH Zürich (2006). http://www.math.ethz.ch/∼knus
2. W. Bosma, J. Cannon, C. Playoust. The Magma algebra system. I. The user language. J. Sym-
bolic Comput. 24 (3–4), 235–265 (1997)
3. N. Bourbaki. Éléments de mathématique. Groupes and algèbres de Lie. Chapitres 4, 5 et 6.
(Masson, Paris, 1981)
4. J. Brillhart. On the Euler and Bernoulli polynomials. J. Reine Angew. Math. 234, 45–64 (1969)
5. E. Cartan. Le principe de dualité et la théorie des groupes simples et semi-simples. Bull. Sci.
Math. 49, 361–374 (1925)
6. J. H. Conway, A. Hulpke, J. McKay. On transitive permutation groups. LMS J. Comput. Math.
1 (1998), 1–8
7. P. Deligne. Séminaire de géométrie algébrique du Bois-Marie, SGA 4 1/2, Cohomologie étale,
avec la collaboration de J. F. Boutot, A. Grothendieck, L. Illusie and J. L. Verdier, Lecture
Notes in Mathematics vol. 569. (Springer-Verlag, Berlin, 1977)
8. W. N. Franzsen. Automorphisms of Coxeter Groups. PhD thesis, School of Mathematics and
Statistics, University of Sydney (2001). http://www.maths.usyd.edu.au/u/PG/theses.html
9. W. N. Franzsen, R. B. Howlett. Automorphisms of nearly finite Coxeter groups. Adv. Geom.
3 (3), 301–338 (2003)
10. S. Garibaldi, A. A. Merkurjev, J-P. Serre. Cohomological invariants in Galois cohomology.
University Lecture Series vol. 28. (American Mathematical Society, Providence, RI, 2003)
11. J. W. Jones, D. P. Roberts. Octic 2-adic fields. J. Number Theory, 128 1410–1429 (2008)
12. M.-A. Knus, A. A. Merkurjev, M. Rost, J.-P. Tignol. The Book of Involutions. American Math-
ematical Society Colloquium Publications vol. 44. (American Mathematical Society, Provi-
dence, R.I., 1998)
13. M.-A. Knus, J.-P. Tignol. Quartic exercises. Inter. J. Math. Math. Sci. 2003 (68), 4263–4323
(2003)
14. . Thin Severi-Brauer varieties. preprint, 2009, arxiv:0912.3359
15. J.-P. Serre. Witt invariants and trace forms. Minicourse, Workshop “From quadratic forms
to algebraic groups”, Ascona, organized by Paul Balmer, Eva Bayer and Max-Albert Knus,
February 18–23 (2007)
16. . Les invariants de W (D4 ). Emails, June 5, June 6, June 18, 2009
286 Max-Albert Knus and Jean-Pierre Tignol
17. J. Tits. Sur les analogues algébriques des groupes semi-simples complexes. In Colloque
d’algèbre supérieure, tenu à Bruxelles du 19 au 22 décembre 1956, 261–289. (Centre Belge
de Recherches Mathématiques, Établissements Ceuterick, Louvain; Librairie Gauthier-Villars,
Paris, 1957)
18. . Sur la trialité et certains groupes qui s’en déduisent. Publ. Math. IHES 2, 14–60
(1959)
19. . Buildings of Spherical Type and Finite BN-Pairs. Lecture Notes in Mathematics
vol. 386. (Springer-Verlag, Berlin, 1974)
20. S. Zweifel. Etale Algebren und Trialität. Diplomarbeit, Mathematikdepartement, ETH Zürich.
(2006). http://www.math.ethz.ch/∼knus
Remarks on Unimodular Rows
To Parimala Raman
1 Introduction
N. Mohan Kumar
Department of Mathematics, Washington University in St. Louis, St. Louis, MO 63130, USA,
e-mail: [email protected]
M. Pavaman Murthy
Department of Mathematics, University of Chicago, 5734 University Avenue, Chicago, IL 60637
e-mail: [email protected]
2010 Mathematics subject classification. 13C10.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 287
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 17,
c Springer Science+Business Media, LLC 2010
288 N. Mohan Kumar and M. Pavaman Murthy
The first nontrivial condition of this kind was obtained by Swan and Towber [4]
and was later generalized by Suslin [3].
Theorem 1 (Suslin). Let (a1 , a2 , . . . , an ) be a unimodular row over a ring A. Assume
that ai = bri i where the ri ’s are non-negative integers and bi ∈ A. If (n − 1)! divides
∏ ri , then (a1 , a2 , . . . , an ) is completable.
In an attempt to generalize the above theorem, Nori conjectured the following.
Conjecture 2 (Nori). Let ϕ : R = k[x1 , x2 , . . . , xn ] → A be a homomorphism of
k-algebras, where k is a field and the xi ’s are indeterminates. Assume that the row
(ϕ (x1 ), . . . , ϕ (xn )) is unimodular. Let fi ∈ R with 1 ≤ i ≤ n be such that the radical
of the ideal generated by the fi ’s is (x1 , . . . , xn ) and the length of R/( f1 , . . . , fn ) is a
multiple of (n − 1)!. Then the unimodular row (a1 , a2 , . . . , an ) where ai = ϕ ( fi ) is
completable.
This conjecture is still open, but the first author proved a partial result in this
direction [2].
Theorem 3 (Mohan Kumar). Assume in the above that k is algebraically closed
and the fi ’s are homogeneous. Then the conjecture is true.
In the present article we prove yet another sufficient condition for a unimodular
row of length three to be completable, which does not seem to follow from the
earlier results.
Theorem 4. Let (a, b, c) ∈ A3 be unimodular. Suppose that the polynomial z2 + bz +
ac has a root in A. Then (a, b, c) is completable.
As an immediate corollary we have
Corollary 5. Suppose 1/2 ∈ A and (a, b, c) ∈ A3 unimodular. If b2 − 4ac is a square
in A, then (a, b, c) is completable.
For larger length unimodular rows we have a somewhat weaker result which can
be found in Sect. 3.
2 Proof of Theorem 4
Let π : P1A → SpecA be the structure morphism and let OP1 (1) be the tautological
A
line bundle. Then π∗ OP1 (n) = H0 (P1A , OP1 (n)) is a free A-module of rank n + 1
A A
for all n ≥ 0. Let x, y be the homogeneous coordinates of P1A . Given unimodular
(a, b, c) ∈ A3 , we may define a subscheme X of P1A by the vanishing of s = ax2 +
bxy + cy2. Then we have an exact sequence
s
0 → OP1 (−2) −
→ OP1 → OX → 0. (1)
A A
As H (OP1 ) ∼
0
=A∼ = H1 (OP1 (−2)), we see that B ∼ = A ⊕ A as A-modules.
A A
We first identify the ring B. It is generated by one element over A. Since the
generator is identified as a generator of H1 (P1A , OP1 (−2)), we use Čech complex to
A
identify this element. Let Ux (respectively Uy ) be the open set x = 0 (respectively
y = 0) of P1A . With respect to this open cover, the Čech complexes for the exact
sequence (1) fit into a diagram as follows.
0 0
⏐ ⏐
⏐ ⏐
ϕ1
H0 (Ux , OP1 (−2)) ⊕ H0(Uy , OP1 (−2)) −−−−→ H0 (Ux ∩Uy , OP1 (−2))
A
⏐ A
⏐ A
⏐
ψ1
⏐ψ
2
ϕ
H0 (Ux , OP1 ) ⊕ H0(Uy , OP1 ) −−−2−→ H0 (Ux ∩Uy , OP1 )
A
⏐ A
⏐ A
⏐ ⏐
ϕ
H0 (Ux , OX ) ⊕ H0(Uy , OX ) −−−3−→ H0 (Ux ∩Uy , OX )
0 0
⏐ ⏐
⏐ ⏐
ϕ
A[t] ⊕ A[t −1] −−−1−→ A[t,t −1 ]
⏐ ⏐
⏐
ψ1
⏐ψ
2
ϕ
A[t] ⊕ A[t −1] −−−2−→ A[t,t −1 ]
⏐ ⏐
⏐ ⏐
ϕ
A[t]/(h) ⊕ A[t −1]/(ht −2 ) −−−3−→ A[t,t −1 ]/(h)
ϕ1 (α , β ) = α − t −2 β , ϕ2 (α , β ) = α − β , ϕ3 (α , β ) = α − β ,
−2
ψ1 (α , β ) = (hα , ht β ), and ψ2 (α ) = hα .
The following are easy to check. The kernel of ϕ1 is zero (since it equals
H0 (OP1 (−2))) and similarly, the kernel of ϕ2 is A and the kernel of ϕ3 is B. The
A
cokernel of ϕ1 is naturally identified with At −1 and the cokernels of ϕ2 and ϕ3 are
both zero. The element z = (at, −(b + ct −1 )) ∈ A[t]/(h) ⊕ A[t −1 ]/(ht −2) goes to
zero under ϕ3 and hence defines an element in B. We claim that this element gen-
erates B as an A-algebra. To check this, it suffices to check that this element goes
to t −1 ∈ At −1 = H1 (OP1 (−2)). We do this by a simple diagram chase. Clearly z
A
can be lifted to (at, −(b + ct −1)) ∈ A[t] ⊕ A[t −1]. When we apply ϕ2 to this we get
ht −1 ∈ A[t,t −1 ], and thus it is the image of t −1 under ψ2 proving the claim. Next
we claim that z satisfies the equation z2 + bz + ac = 0. It suffices to check that
at ∈ A[t]/(h) and −(b + ct −1) ∈ A[t −1 ]/(ht −2) satisfy this equation:
(a,b,c)
0 −→ A −−−→ A −→ P −→ 0,
3
(2)
where P = π∗ OX (2).
We have a surjection π ∗ (P) → OX (2) and thus we see that π ∗ (P) = OX (2) ⊕
OX (−2). But we have seen that π∗ OX (1) = π∗ OP1 (1) = A2 and thus OX (1) is two
A
generated. So, by Lemma 6, we see that π ∗ P is free. If z2 + bz + ac has a root in
A, we have a retraction B → A. Since B ⊗A P is free, we see that P must be free as
well, proving Theorem 4.
Remarks on Unimodular Rows 291
Remark 8 (Nori). Let (a, b, c) ∈ A3 be unimodular and let P be the associated pro-
jective module. Then for any n ≥ 2, there exists an A-algebra B, which is A-free of
rank n and such that B ⊗A P is B-free.
We now justify this claim. Let s = axn + bxn−1 y + cyn , where as before, x and y
are the homogeneous coordinates of P1A . The zeros of s define a subscheme X of P1A .
We have an exact sequence
s
0 → OP1 (−n) −
→ OP1 → OX → 0.
A A
Then as before, H0 (X, OX ) = B is a free A-module of rank n, since H1 (P1A , OP1 (−n))
A
is a free A-module of rank n − 1 and the map X → Spec A is a finite map of degree
n since (a, b, c) is unimodular. We have as before a surjection B2 → OX (n) given
by sending the basis elements to xn , yn . Thus, we get a surjection B3 → OX (n) by
sending the basis elements to xn , xn−1 y, yn . As axn + bxn−1 y + cyn = 0 on X, we see
that this surjection factors through B3 /B(a, b, c) = Q. Thus, the B-projective module
Q is isomorphic to OX (n) ⊕ OX (−n) and since OX (n) is two generated, we see that
Q is free by Lemma 6. But,
Q = B3 /B(a, b, c) = A3 /A(a, b, c) ⊗A B = P ⊗A B,
as desired.
Remark 10. From the proof earlier, we see that with the condition (∗) on A, having
a section of π : X → SpecA implies that s has a linear factor. Conversely, if s has a
linear factor, say bx − ay, then the coefficients of s are contained in the ideal (a, b)
and thus (a, b) is unimodular, since the coefficients of s generate the unit ideal.
Then it is clear that the map π : X → Spec A has a section. Thus, Theorem 9 can
be restated as follows: The unimodular row (a1 , a2 , . . . , an ) is completable if the
associated polynomial s has a linear factor.
Acknowledgements The first author was partially supported by a grant from the NSA.
Remarks on Unimodular Rows 293
References
Summary In this short note we study vector bundles generated by sections and the
associated morphism to the Grassmannian.
1 Introduction
f : Pm → Gr(2, Cn+2 )
F. Laytimi
Mathématiques - bât. M2, Université Lille 1, F-59655 Villeneuve d’Ascq Cedex, France,
e-mail: [email protected]
D.S. Nagaraj
Institute of Mathematical Sciences C.I.T. campus, Taramani, Chennai 600113, India,
e-mail: [email protected]
2010 Mathematics subject classification. 14F05.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 295
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 18,
c Springer Science+Business Media, LLC 2010
296 F. Laytimi and D.S. Nagaraj
However, our paper contains some further new observations regarding morphisms
to Grassmannians and our proof of the above theorem is straightforward.
Definition 2.1. Let X be a variety over the field C and s be a positive integer. A
vector bundle E on X is said to be generated by s sections if there is a surjection of
vector bundles
Cs ⊗ OX → E.
Note that in this definition we do not assume that the induced map Cs →
H 0 (X , E)
is injective.
We recall the following well-known result. For the sake of completeness we will
sketch its proof.
Lemma 2.3. Let X be a projective variety, r and k be two positive integers. Consider
the following two sets:
1) The set of morphisms f : X → Gr(r, Cr+k ).
2) The set of equivalence classes of vector bundle surjections
ϕ : Cr+k ⊗ OX → E,
ϕ0 : Cr+k ⊗ OGr(r,Cr+k ) → Q,
on Gr(r, Cr+k ) gives a natural map from the set in (1) to the set in (2). This map is
a bijection between the two sets.
Vector Bundles Generated by Sections and Morphisms to Grassmannians 297
This itself defines a map from the set in (2) to the set in (1). Moreover, it is easy to
see that
f ∗ (ϕ0 ) : f ∗ (Cr+k ⊗ OGr(r,Cr+k ) ) → f ∗ (Q)
is equivalent to ϕ : Cr+k ⊗ OX → E.
These two maps are clearly inverse to each other, hence the natural map from set
in (1) to set in (2) is a bijection.
Remark 2.4. Lemma 2.3, in the case r = 1, gives a correspondence between finite
morphisms
f : X −→ Gr(1, C1+k ) Pk
and ample line bundles generated by k + 1 sections.
Theorem 2.5. Let X be a projective variety. The bijective map of Lemma 2.3 induces
a natural bijection between the following two sets:
1) The set of morphisms f : X −→ Gr(r, Cr+k ) with f finite (onto its image).
2) The set of equivalence classes of vector bundle surjections
Cr+k ⊗ OX −→ E
with det(E) ample, where E is a vector bundle of rank r and det(E) is the deter-
minant line bundle of E.
Proof. Given an element of the set in (2), i.e., a surjection, Cr+k ⊗ OX −→ E with
rk(E) = r and det(E) ample, then as in Lemma 2.3 it determines a morphism
f : X −→ Gr(r, Cr+k ) and f ∗ (Q) E, where Q is the universal quotient bundle
of rank r on Gr(r, Cr+k ). Hence f ∗ (det(Q)) det(E). Since det(E)| f −1 ( f (x))
f ∗ (det(Q))| f −1 ( f (x)) is trivial, the ampleness assumption on det(E) implies that
dim( f −1 ( f (x))) ≤ 0. Thus, if det(E) is ample, then f is finite onto its image.
In the other direction, if f : X −→ Gr(r, Cr+k ) is a finite morphism onto its image,
then
det( f ∗ (Q)) f ∗ (det(Q))
is ample, since the pull-back of an ample bundle under a finite morphism is ample.
298 F. Laytimi and D.S. Nagaraj
Corollary 2.6. Let X be a projective variety of dimension n > r.k, where r and k are
positive integers. Then there is no vector bundle E on X of rank r generated by r + k
sections and with det(E) ample.
C1+k ⊗ OX → L
Cr+k ⊗ OX −→ E ?
Or, equivalently, does there exist a finite morphism (onto its image)
f : X −→ Gr(r, Cr+k ) ?
The answer to this question is in general negative. To see this, we recall the
following:
Theorem 2.8. (Lazarsfeld [3, p. 55, Theorem 4.1]) Let X be a smooth projective
variety of positive dimension and let f : Pn −→ X be a surjective map. Then X Pn .
Corollary 2.9. There is no rank two vector bundle E on P4 with det(E) ample and
generated by four sections.
Remark 2.10. If dim X ≤ 2, then it is easy to see that there is always a finite (onto its
image) morphism f : X −→ Gr(2, C4 ). Hence, if dim(X) ≤ 2 there always exists a
rank 2 vector bundle E on X and a surjection
C4 ⊗ OX −→ E.
In the next section we consider among other things the question of existence of
nonconstant morphisms from P3 → Gr(2, C4 ).
Vector Bundles Generated by Sections and Morphisms to Grassmannians 299
The simplest projective variety of dimension three is P3 . With Remark 2.10 in mind
we consider the following:
Question. Does there exist a rank 2 vector bundle E on P3 with det(E) ample and
with a surjection
C4 ⊗ OP3 −→ E?
Remark 3.1. A result of Fulton and Lazarsfeld [2] implies the following: if E is an
ample vector bundle on P3 then there is no surjective morphism of vector bundles
C4 ⊗ OP3 −→ E. Thus, if there is a rank 2 vector bundle E on P3 generated by four
sections and with det(E) ample, then E has to be nontrivial and nonample. It is also
easy to see that the bundles of the form:
The following examples give an affirmative answer to the above question. These
examples were pointed out to us by Gruson.
where IC denotes the ideal sheaf of C in P3 and c1 (E) denotes the first Chern class
of E. It is easy to see that c1 (E) = 2 and h0 (IC (2)) = 4, where h0 stands for the
dimension of the vector space of global sections. Thus we conclude that h0 (E) =
5. Note that C can be expressed as the intersection of three independent quadrics
vanishing on C. (For example, if
C = {X = 0 = Y } ∪ {Z = 0 = W },
C4 ⊗ OP3 → E.
It is easy to see that c1 (E) = 4 and h0 (IC (2)) = 3. Thus we conclude that h0 (E) = 4.
Since C has no 5-secant, C is scheme theoretically defined by the three linearly
independent quartics containing it [1, p. 579]. This shows that there is a vector
bundle surjection
C4 ⊗ OP3 → E.
By Theorem 2.5, this surjection defines a finite morphism f : P3 → Gr(2, C4 ).
Remark 3.4. In Example 3.3, we do not know what the image variety is nor whether
the morphism is generically one–one. It would be interesting to investigate these
questions.
Lemma 3.5. Let E be a nontrivial rank two vector bundle on P3 together with a
surjection
C4 ⊗ OP3 → E.
Then c1 (E) = 2m and c2 (E) = 2m2 for some integer m ≥ 1.
Proof. The surjection
C4 ⊗ OP3 → E
gives rise to an exact sequence of vector bundles
0 → S → C4 ⊗ OP3 → E → 0.
Let c1 (E) = a and c2 (E) = b. Since E is generated by sections and nontrivial we see
by Lemma 3.9 that a > 0. Now, since the rank of S is two, we obtain 0 = c3 (S) =
2b.a − a3. But a
= 0 implies a2 − 2b = 0. This gives the required result.
Remark 3.6. Examples 3.2 and 3.3 give some bundles corresponding to the cases
m = 1 and m = 2 of Lemma 3.5. It will be an interesting problem to study the bun-
dles for the other values of m and to see whether they are generated by 4-sections
and if they are generated by 4-section, what are the properties of the induced fi-
nite morphisms from P3 to Gr(2, C4 ). For example, for each n ≥ 1, consider the
morphism
f n : P3 → P3
defined by (x0 , x1 , x2 , x3 ) → (xn0 , xn1 , xn2 , xn3 ). Since fn∗ (OP3 (1)) = OP3 (n), we see that
the rank two vector bundle En = fn∗ (E) is generated by 4 sections with c1 (En ) = 2n
and c2 (En ) = 2n2 , where E is the bundle in Example 3.2 above. But in this case we
do not get any new morphism from P3 to Gr(2, C4 ).
The following result is a generalization of Corollary 2.9 to higher-dimensional
projective spaces.
Vector Bundles Generated by Sections and Morphisms to Grassmannians 301
Theorem 3.7. Let n ≥ 1 and m ≥ n+2 be two integers. Then there is no nonconstant
morphism f : Pm → Gr(2, Cn+2 ).
Proof. First note that every nontrivial line bundle on Pm which is generated by
sections is necessarily ample. Hence, if X is a projective variety and f : Pm → X is
a nonconstant morphism then the pull back by f of an ample bundle on X is ample
on Pm . This shows that any nonconstant morphism from Pm is necessarily finite onto
its image.
The earlier observation shows that, if m ≥ n + 2 and f : Pm → Gr(2, Cn+2 ) is a
nonconstant morphism, then f |L is also a nonconstant morphism, where L is a linear
subspace of Pm of dimension n + 2. Thus, if we show that there is no nonconstant
morphism from Pn+2 → Gr(2, Cn+2 ), then it follows that for any m ≥ n + 2 there are
no nonconstant morphism from Pm → Gr(2, Cn+2 ). From these remarks, to prove
the theorem, it is enough to show that there is no nonconstant morphism f : Pn+2 →
Gr(2, Cn+2 ).
Let f : Pn+2 → Gr(2, Cn+2 ) be a morphism. Then by pulling back the universal
exact sequence
0 → S → Cn+2 ⊗ OGr(2,Cn+2 ) → Q → 0
on Pn+2 , where Q is the universal rank 2 quotient bundle and S is the kernel of the
natural surjection Cn+2 ⊗ OGr(2,Cn+2 ) → Q. Let c1 ( f ∗ (Q)) = a.H and c2 ( f ∗ (Q)) =
b.H 2 be the Chern class of f ∗ (Q), where H ∈ H2 (Pn+2 , Z) is the generator, corre-
sponding to the ample line bundle OPn+2 (1), of the cohomology algebra of Pn+2 .
Then the total Chern class of f ∗ (S) is
Since the rank of f ∗ (S) is n, the Chern classes cn+1 ( f ∗ (S)) and cn+2 ( f ∗ (S)) are
zero. Hence we see from (1) that
[ n+1
2 ]
i n+1−i
cn+1 ( f (S)) = ∑ (−1)
∗
c1 ( f ∗ (Q))n+1−2i c2 ( f ∗ (Q))i
i=0 i
⎛ n+1 ⎞
[ 2 ]
n + 1 − i
= ⎝ ∑ (−1)i an+1−2ibi ⎠ H n+1 = 0,
i=0 i
2 ]
[ n+1
i n+1−i
∑ (−1) i
an+1−2ibi = 0. (2)
i=0
302 F. Laytimi and D.S. Nagaraj
Similarly, the vanishing of cn+2 ( f ∗ (S)) implies that the integers a, b satisfy the
equation
[ n+2
2 ]
i n+2−i
∑ (−1) i
an+2−2ibi = 0. (3)
i=0
By Lemma 3.8 below we see that the only simultaneous integral solution of
equations (2) and (3) is (a, b) = (0, 0). Hence c1 ( f ∗ (Q)) = 0. Since E is generated
by sections, Lemma 3.9 below implies f ∗ (Q) OPn+2 ⊕ OPn+2 and hence f is a
constant map. This completes the proof of the theorem.
Lemma 3.8. Let n ≥ 1 be a fixed integer. Then the only simultaneous integral solu-
tion of the two equations
[ n+1
2 ]
i n+1−i
∑ (−1) i
X n+1−2iY i = 0 (4)
i=0
2 ]
[ n+2
i n+2−i
∑ (−1) i
X n+2−2iY i = 0 (5)
i=0
is (0, 0).
Proof. Proof by induction on n. For n = 1 the equations (4) and (5) are
X 2 − Y = 0 and X 3 − 2XY = 0.
In this case the result is obvious. Assume n > 1 and that the result holds for n − 1.
Note that if (a, b) is a simultaneous solution of (4) and (5) and if one of a or b is
zero then clearly the other is also zero. Assume both a and b are nonzero and (a, b)
is a simultaneous solution of (4) and (5). Then multiplying (4) by X and subtracting
by (5) we see that (a, b) is a solution of
[ n+2
2 ]
i n+1−i
∑ (−1)
i − 1
X n+2−2iY i = 0. (6)
i=1
[n/2]
n − i n−2i i
∑ (−1)i i
X Y = 0. (7)
i=0
This implies that (a, b)
= (0, 0) is a simultaneous solution of the equations (7)
and (4), which leads to a contradiction to the induction assumption. This com-
pletes the proof of the lemma.
Proof. Choose n linearly independent sections which generate the stalk at one point,
where n is the rank of E. Such sections exist because of our assumption that E is
generated by sections. Then the set S of points where these sections fail to generate
is the zero set of a nonzero section of det(E). But det(E) is generated by sections
and c1 (E) = 0, hence det(E) = OX . Since X is projective any nonzero section of OX
is a nonzero constant. Hence we see that E is trivial.
Acknowledgements We thank L. Gruson for his encouragement and P. Sankaran for going
through an earlier version and suggesting some improvements. The second author would like
to thank the University of Lille, France, for its hospitality and IFIM for the financial grants.
References
1. J. D’Almeida: Une involution sur un espace de modules de fibrés instantons Bull. Soc. Math.
Fr. 128 (2000), 577–584.
2. W. Fulton and R. Lazarsfeld: On the connectedness of degeneracy loci and special divisors
Acta Math. 146 (1981), no. 3–4, 271–283.
3. R. Lazarsfeld: Some applications of the theory of Positive vector bundles In: Complete Inter-
sections, C.I.M.E., Acireale 1983. Lect. Notes. Math. vol. 1092, Springer, Berlin, Heidelberg,
New York, 1984.
4. C. Okonek, M. Schneider and H. Spindler: Vector bundles on complex projective spaces.
Progress in Mathematics, No. 3, Birkhäuser, Boston, Mass., 1980.
5. H. Tango: On (n-1)-dimensional projective spaces contained in the Grassmann variety
Gr(n, 1). J. Math. Kyoto Univ. 14-3 (1974) 415–460.
Adams Operations
and the Brown-Gersten-Quillen
Spectral Sequence
Alexander Merkurjev
1 Introduction
Let X be a separated scheme of finite type over a field F. We write X(p) for the set
of points in X of dimension p. There is the niveau spectral sequence
1
E p,q = K p+q F(x) ⇒ G p+q (X )
x∈X(p)
ϕ p : CH p (X ) = E p,−p
2
→ G0 (X )(p/p−1).
Alexander Merkurjev
Department of Mathematics, University of California, Los Angeles, CA 90095-1555, USA,
e-mail: [email protected]
2010 Mathematics subject classification. 14C35, 19E08, 19L20.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 305
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 19,
c Springer Science+Business Media, LLC 2010
306 Alexander Merkurjev
2 The Category Al
For a prime integer l, let Z(l) denote the localization of Z with respect to the prime
ideal lZ.
Lemma 2.1. Let f , g ∈ Z(l) [t] be polynomials such that f and g are coprime over Q
and the residues f¯ and ḡ are coprime over Z/lZ. Then f and g are coprime over
Z(l) , i.e., f and g generate the unit ideal in Z(l) [t].
Proof. Let I be the ideal in Z(l) [t] generated by f and g. By assumption, I contains
l k and a polynomial 1 − lh for some k > 0 and h ∈ Z(l) [t]. Then 1 − l k hk ∈ I and
hence 1 ∈ I.
2.1 Definition of Al
2.2 A Decomposition
Let M ∈ Al and let a ≤ b be two integers such that M (a) = M and M (b) = 0. The set
J of all integers j such that a ≤ j < b is called an interval of M.
Let k > 1 be an integer such that the congruence class [k]l of k modulo l is a
generator of (Z/lZ)× . Write Λ = Z/(l − 1)Z. For any congruence class ρ ∈ Λ ,
consider the polynomial
fρ = ∏ (t − k j ) (2.2)
j∈ρ ∩J
f= ∏ fρ = ∏(t − k j ).
ρ ∈Λ j∈J
k − k j ) M ( j) ⊂ M ( j+1) for any j, hence
It follows from (ii) that (ψM
f (ψM
k
) = 0. (2.3)
M= Mρ (2.4)
ρ ∈Λ
Mρ = Ker fρ (ψM
k
) ⊂ Ker fρ (ψM
k
) = Mρ .
We also prove that the decomposition (2.4) does not depend on the choice of the
integer k. Let k > 1 be an integer such that the congruence class [k ]l is a generator
of (Z/lZ)× . Define the polynomials f , fρ , and f in Z(l) [t] as earlier with k replaced
by k . Then M is a direct sum of the submodules Mρ = Ker fρ (ψM k ) over all ρ ∈ Λ .
k
Mρ = Ker fρ (ψM
k
) = Im fρ (ψM
k
) ⊂ Ker fρ (ψM ) = Mρ .
By symmetry, Mρ ⊂ Mρ .
Thus, the submodules Mρ in the decomposition (2.4) depend only on the object
M in the category Al .
Let s : M → N be a morphism in Al . Choose a common interval J for both M
and N and let fρ be the polynomials defined by (2.2). Then s ◦ fρ (ψM
k ) = f (ψ k ) ◦ s,
ρ N
hence
s Mρ = s Ker fρ (ψM k
) ⊂ Ker fρ (ψNk ) = Nρ ,
i.e., s induces a Z(l) -module homomorphism Mρ → Nρ . Thus for every ρ ∈ Λ , we
have a functor from Al to the category of Z(l) -modules taking an object M to Mρ .
Let M be a scheme and Z ⊂ M a closed subscheme. We write KmZ (M) for the
K-groups of M with support in Z and Fγi KmZ (M) for the i-th term of the (finite)
Adams Operations and the Brown-Gersten-Quillen Spectral Sequence 309
gamma-filtration on KmZ (M) (see [11, Sect. 4]). If M is a regular scheme, there exists
a canonical isomorphism
Gm (Z) KmZ (M),
where Gm (Z) is the K-group of the category of coherent sheaves on Z [3, Th.2.14].
For any integer k, there is the Adams operation ψ k on KmZ (M) satisfying the
following properties [11], [7, Sect. 9]:
• ψ k is a group endomorphism of KmZ (M).
• ψ k respects the gamma-filtration Fγi KmZ (M).
(i/i+1)
• ψ k acts as multiplication by ki on the subsequent factor Fγ KmZ (M).
• ψ k ψ k = ψ kk , in particular, ϕ k and ϕ k commute.
Let l be a prime integer. Then the Z(l) -module KmZ (X )Z(l) together with the
gamma-filtration and the Adams operations ψ k on it yield an object of the category
Al . Therefore, KmZ (X)Z(l) decomposes as in (2.4) into a direct sum of submodules
Z
Km (X )Z(l) ρ over all ρ ∈ Λ = Z/(l − 1)Z.
Let Z be a scheme. For any integer k > 0, there is a well-defined map (see [11,
Sect. 4]):
θ k : K0 (Z) → K0 (Z),
natural in Z, satisfying:
• For every exact sequence of vector bundles 0 → E → E → E → 0 over Z, we
have θ k [E] = θ k [E ] · θ k [E ].
• For every line bundle L over Z,
such that
f∗ ψ k (θ k (N) · α ) = ψ k f∗ (α )
for any α ∈ K∗ (Z) and any k.
Corollary 3.2. Suppose that, in addition, N is a trivial bundle. Then θ (N) = kd−p
and hence
f∗ kd−p · ψ k (α ) = ψ k f∗ (α ) .
310 Alexander Merkurjev
in the category Al .
that is isomorphic to the localization exact sequence [10, Sect. 7, Prop. 3.2]:
The homomorphisms in (3.3) commute with the Adams operations by [7, Remark
9.6(1)]. Then, localizing at a prime integer l, we can view (3.3) and (3.4) as se-
quences of morphisms in the category Al .
Let X be a scheme. We embed X into a regular variety M of dimension d as a
closed subscheme. For a pair of integers p and q set
Z\Z
1
E p,q = colim K p+q (M \ Z ) = colim G p+q (Z \ Z ),
where the colimit is taken over all pairs (Z , Z) with Z a closed subscheme of X of
dimension p and Z a closed subscheme of Z of dimension p − 1. Note that for any
such Z one can find a Z such that Z \ Z is regular and the normal bundle of Z \ Z
in M \ Z is trivial. It follows that
1
E p,q = K p+q F(x)
x∈X(p)
Set
D1p,q = colim K p+q
Z
(M) = colim G p+q (Z),
where the colimit is taken over all closed subschemes Z of X of dimension p. Taking
colimits of the exact sequences (3.3) and (3.4) we get the exact sequences:
· · · → E p,q
1
→ D1p−1,q → D1p,q−1 → E p,q−1
1
→ ···
for any s ≥ 1 if p + q ≤ 2.
Proof. By [11, Sect. 2], for a field L and m ≤ 2, we have Fγm Km (L) = Km (L) and
Fγm+1 Km (L) = 0, hence ψ k acts on Km (L) by multiplication by km . Therefore,
Km (L)Z(l) , if ρ = [m]l−1 ;
Km (L)Z(l) ρ =
0, otherwise.
∂ : E p,q
s
→ E p−s,q+s−1
s
be the differential in the spectral sequence (3.6) with p + q ≤ 2. Then the order
ord(∂ ) is finite and if l is a prime divisor of ord(∂ ), then l ≤ p and l − 1 divides
s − 1.
Proof. If s > p, then ∂ = 0 since E p−s,q+s−1
s = 0, so we may assume that s ≤ p.
We claim that if l is a prime integer such that ∂ Z(l) = 0, then l − 1 divides s − 1.
Set ρ = [q + d]l−1 ∈ Λ . By Lemma 3.7, the (nonzero) image of ∂ Z(l) is contained in
(E p−s,q+s−1
s Z(l) )ρ and therefore, ρ + d = [q + d + s − 1]l−1, i.e., l − 1 divides s − 1.
The claim is proved.
Taking l > s, we get ∂ Z(l) = 0 from the claim, i.e., ∂ has finite order.
Let l be a prime divisor of ord(∂ ). Then ∂ Z(l) = 0 and hence by the claim, l − 1
divides s − 1. In particular, l ≤ s ≤ p.
312 Alexander Merkurjev
The spectral sequence is compatible with the Adams operations ψ k , and ψ k acts by
multiplication by kq on E2p,q [7, Th. 9.7].
Theorem 3.11. Let X be a smooth scheme. Let
∂ : Esp,q → Esp+s,q−s+1
be the differential in the spectral sequence (3.10). Then ord(∂ ) is finite and if l is a
prime divisor of ord(∂ ) then l ≤ dim(X) − p and l − 1 divides s − 1.
Proof. The proof is parallel to the one of Theorem 3.8. One remarks that ∂ = 0 if
s > dim(X) − p.
Acknowledgements I would like to thank Marc Levine for useful discussions. The work has been
supported by the NSF grant DMS #0652316.
References
Summary We give several proofs that a stably free module of even rank need not
be self-dual.
1 Introduction
In this note we prove that the projective module corresponding to the universal
unimodular vector of odd size > 3 is not isomorphic to its dual. We give three
proofs which depend on topological facts, and also two algebraic arguments, one
using Suslin matrices and one using Riemann–Roch algebra.
For any commutative ring R, we set
Theorem 1.1. Let A = A2n−1(Z). Let xi denote the class of Xi , and yi denote the
class of Yi in A, for all i. Let v = (x1 , . . . , x2n−1 ), v∗ = (y1 , . . . , y2n−1 ). Then for
Madhav V. Nori
Department of Mathematics, The University of Chicago, Chicago, IL 60637, USA,
e-mail: [email protected]
Ravi A. Rao
School of Mathematics, Tata Institute of Fundamental Research, Mumbai 400 005, India,
e-mail: [email protected]
Richard G. Swan
Department of Mathematics, The University of Chicago, Chicago, IL 60637, USA,
e-mail: [email protected]
2010 Mathematics subject classification. Primary: 13C10, 19A13. Secondary: 14C25, 19B14,
19L10, 55Q05, 55R25.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 315
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 20,
c Springer Science+Business Media, LLC 2010
316 Madhav V. Nori, Ravi A. Rao, and Richard G. Swan
n > 2, v, v∗ are not in the same GL2n−1 (A) orbit, i.e., the projective module P
(of rank 2n − 2) defined by v is not isomorphic to the projective module P∗ defined
by v∗ .
Thus, one gives examples of non-self-dual stably free modules of rank 2m for
every even 2m ≥ 4. We now explain why these examples do not work when the rank
r is odd or when r = 2.
Let P be the projective An (R)-module of rank n − 1 corresponding to the uni-
modular row v = (x1 , . . . , xn ). Then P∗ = HomR (P, R) is isomorphic to the projective
R-module corresponding to the unimodular row w = (y1 , . . . , yn ).
It can be seen easily that P and P∗ are isomorphic when rank P is odd; in
fact, the rows v, w are in the same elementary orbit by a lemma of Roitman, see
[9, Lemma 1].
In [1, Prop. 4.4, Cor. 4.5], Bass showed that any projective R-module P of rank
two admits a symplectic structure if and only if the invertible module ∧2 P is free.
In particular, if P ⊕ R is free then P is isomorphic to its dual.
Let Vectn (R) denote the set of isomorphism classes of projective R-modules of rank
n, where R is a commutative ring. The free module Rn serves as the base point
c(n) of the set Vectn (R). The map P → R ⊕ P defines f : Vectn−1 (R) → Vectn (R).
The kernel of f , i.e., f −1 c(n), consists of the isomorphism classes of projective
modules P such that there is an isomorphism T : R ⊕ P ∼ = Rn . In such a situation,
with (1, 0) ∈ R ⊕ P, we obtain x = (x1 , x2 , ..., xn ) = T (1, 0) ∈ Rn . With p : R ⊕ P → R
denoting the projection to the first factor, we also obtain y = p ◦ T −1 : Rn → R.
Identifying HomR (Rn , R) with Rn in the standard manner, we get y ∈ Rn so that
∑ni=1 xi yi = 1. The set of (x, y) ∈ Rn × Rn satisfying this identity is Dn (R), the set of
ring homomorphisms An (Z) → R.
Note also that every (x, y) ∈ Dn (R) gives rise to the projective module ker(y) :
Rn → R, which is naturally isomorphic to Rn /Rx. This gives rise to g : Bn (R) →
Vectn−1 (R) and we obtain an exact sequence of pointed sets
Now let Z be a compact Hausdorff space and let C(Z) be the ring of continuous
C-valued functions on Z. By [14] we can identify Vectk (C(Z)) with the set of iso-
morphism classes of complex rank k bundles on Z. To simplify notation, we denote
Vectk (C(Z)) by Vectk (Z) in this situation. The case of immediate interest is Z = SY
the suspension of a compact Hausdorff space Y . Here, by employing clutching
functions (cf. [7, 10]), we see that Vectk (SY ) is identified with [Y ; GLk (C)], the set
of homotopy classes of continuous maps Y → GLk (C). In addition, ring homor-
phisms An (Z) → C(Z) are readily identified with continuous maps Z → S2n−1(C)
Non-self-dual Stably Free Modules 317
When Y is a sphere, the above is contained in the long exact sequence of homotopy
groups associated to the fiber bundle GLn(C) → S2n−1 (C) with fiber GLn−1(C) given
by g ∈ GLn (C) → (x, y), where x is the first column of g and y is the first row of g−1 .
When Y is the real sphere S2n−2, the preceding arrow in the long exact sequence is
π2n−1 (GLn (C)) → [SY ; S2n−1(C)]. Regarding the real (2n − 1)-sphere S2n−1 as the
unit sphere in Cn , the embedding S2n−1 → S2n−1 (C) given by i(x) = (x, x̄) is readily
checkedtobeahomotopyequivalence. Thus, [SY ; S2n−1 (C)]isnaturallyidentifiedwith
Z. Thanksto Bottperiodicity, thegroup π2n−1(GLn (C)) = Vectn (S2n )isalso identified
canonically with Z. The image of the arrow π2n−1 (GLn (C)) → [SY ; S2n−1 (C)] is
(n − 1)!Z, since π2n−2 (GLn−1 (C)) = Z/(n − 1)!Z and π2n−2(GLn (C)) = 0 by [3,4].
Thus with Z = SY = S2n−1, we see that a, b : S2n−1 → S2n−1 (C) give rise to isomorphic
rank (n − 1) complex vector bundles on S2n−1 if and only if deg(a) ≡ deg(b)
mod (n − 1)!.
With i, j : S2n−1 → S2n−1(C) given by i(x) = (x, x̄) and j(x) = (x̄, x), we see
that deg( j) = (−1)n deg(i) = (−1)n , so the corresponding rank (n − 1) projective
modules over C(S2n−1) are isomorphic to each other if and only if 1 − (−1)n is
divisible by (n − 1)!, in other words, if n is even or n ≤ 3. In particular, when n does
not satisfy these restrictions, the projective modules given by the unimodular rows
x and y are not isomorphic in the universal situation.
Note that π2n−2 U(n) = 0 by [4], and so ∂ , and therefore f∗ , is onto and γ gener-
ates π2n−1BU(n − 1) = Z/(n − 1)! by [3].
Now E ∗ is defined by z1 , . . . , zn , and so is c∗ (E) where c : S2n−1 −→ S2n−1 is
given by (z1 , . . . , zn ) −→ (z1 , . . . , zn ). Therefore, E ∗ is classified by:
c f
S2n−1 −→ S2n−1 −→ BU(n − 1)
We now give an alternative proof of Theorem 1.1 using only elementary homotopy
theory in addition to Bott’s calculations.
Regard A = An (C) as the ring of complex polynomial functions on the (2n − 1)-
sphere S2n−1 in Cn by sending xi to zi and yi to z̄i .
Let f = ( f1 , . . . , fn ) be a unimodular row over A. It defines a map S2n−1 → S2n−1
sending z to f (z)/ f (z). We write deg f for the degree of this map and write [ f ]
for its homotopy class in π2n−1 (S2n−1) = Z. So [ f ] = deg f ι where ι is the class of
the identity map, which generates π2n−1 (S2n−1).
Let P( f ) be the projective module defined by f so that A f ⊕ P( f ) = An .
Theorem 4.1. If v and w are unimodular rows of length n over An (C) and P(v) ≈
P(w) then deg v ≡ deg w mod (n − 1)!.
Proof. By taking the direct sum of P(v) ≈ P(w) with Av ≈ Aw (by v → w) we get
an element g of GLn (A) such that gv = w cf. [8, Prop. I.4.8]. Regarding elements
of A as functions on S2n−1 we have g : S2n−1 → GLn (C) and, passing to homotopy
classes we get [g] ∈ π2n−1(GLn (C)) = π2n−1 (U(n)) and ([g], [v]) maps to [w] under
the map
As this map is linear we have p∗ [g] + [v] = [w] where p : U(n) → S2n−1 by g → gs
where s is a base point. Now p is a fibration with fiber U(n − 1) [10]. Using Bott’s
calculations from [3, 4], its homotopy sequence
p∗
π2n−1 (U(n)) −→ π2n−1 (S2n−1) → π2n−2 (U(n − 1)) → π2n−2(U(n))
Non-self-dual Stably Free Modules 319
Corollary 4.2. If v is a unimodular row of length n over An (C) and P(v) is free then
deg v ≡ 0 mod (n − 1)!
Corollary 4.4. If v and w are unimodular rows of length n over An (C) and P(v)∗ ≈
P(w) then deg w ≡ (−1)n deg v mod (n − 1)!.
Proof. Suppose v · u = 1. Then P(v)∗ ≈ P(u) [8, Prop. I.4.10]. The map u/u is
homotopic to the map v̄/v̄ by (tu + (1 − t)v̄)/(tu + (1 − t)v̄). Note that tu +
(1 − t)v̄ is never 0 because its inner product with ū is tu2 + 1 − t > 0. Now the
map v̄ is the composition of v with the map of S2n−1 to itself sending (z1 , . . . , zn ) to
its conjugate. This map reverses the sign of the n imaginary components and so has
degree (−1)n . Therefore, deg u = (−1)n degv and the theorem applies.
In particular, P(v) ≈ P(v)∗ implies that deg v ≡ (−1)n deg v mod (n − 1)! so if n
is odd, deg v ≡ 0 mod (n − 1)!/2. If v = (x1 , . . . , xn ) then deg v = 1 so, for odd n,
P(v) ≈ P(v)∗ implies that n ≤ 3.
Corollary 4.5. If v = (x1 e1 , . . . , xn en ) and n is odd then P(v) ≈ P(v)∗ implies that
2e1 · · · en ≡ 0 mod (n − 1)!
We now give an alternative algebraic proof of Theorem 1.1 via Suslin matrices,
when rank P > 2 and even. Before proving the theorem we recall the Suslin matrices
and some of their properties.
The Suslin Matrices Sr (v, w). The construction of the Suslin matrices Sr (v, w) is
possible once we have two rows v, w of length (r + 1). It becomes more interesting
if their dot product v, w = v · wt = 1. (The rows are then automatically unimodular
rows, i.e., the coordinates of each row generate the unit ideal.)
Andrei Suslin’s inductive definition: Let v = (a0 , a1 , . . . , ar ) = (a0 , v1 ) with v1 =
(a1 , . . . , ar ) and w = (b0 , b1 , . . . , br ) = (b0 , w1 ) with w1 = (b1 , . . . , br ). Set S0 (v, w) =
a0 , and set
a0 I2r−1 Sr−1 (v1 , w1 )
Sr (v, w) = .
−Sr−1 (w1 , v1 )t b0 I2r−1
320 Madhav V. Nori, Ravi A. Rao, and Richard G. Swan
for r ≥ 1. The Suslin matrices were introduced by Suslin in [11, Sect. 5] to show that
a unimodular row of the form (a0 , a1 , a22 , . . . , arr ) can be completed to an invertible
matrix of determinant one.
Nature of the Suslin Matrices. Suslin defines a sequence of forms Jr ∈ M2r (R)
by the recurrence formulae:
⎧
⎪
⎨1 for r = 0,
Jr = Jr−1 ⊥ −Jr−1 , for r even, and
⎪
⎩
Jr−1 − Jr−1, for r odd.
(The English
translation
incorrectly
says that Jr = Jr−1 Jr−1 when r is odd.) Here
α 0 0 α
α ⊥ β = 0 β , while α β = β 0 .
r(r+1)
It is easy to see that det Jr = 1 for all r, and that Jrt = Jr−1 = (−1) 2 Jr . More-
over, Jr is antisymmetric if r = 4k + 1 and r = 4k + 2, whereas Jr is symmetric for
r = 4k and r = 4k + 3. Suslin noted that the following formulae are valid:
Then they showed that the map v → wt(v) is a Mennicke symbol. Moreover, they
showed that wt(vα ) = wt(v) + ∑ni=0 (−1)i [∧i α ].
Suslin interprets this map and shows that wt(v) = [Sn−1(v, w)] ∈ SK1 (R).
We next recall an important result of Suslin, Theorem 1.2 in [13]. Suslin showed
that, SK1 (An (Z)) Z, with generator [Sn−1((x1 , . . . , xn ), (y1 , . . . , yn ))].
We now reprove Theorem 1.1 using Suslin matrices and their properties outlined
earlier. Let R = A2n−1(Z) and let v = (x1 , . . . , x2n−1 ) be as earlier so wt(v) generates
SK1 (R) = Z. Suppose that vσ = w, for some σ ∈ GL2n−1 (R), with n > 2. Then
2n−1
wt(w) = wt(vσ ) = wt(v) + ∑ (−1)i [∧i σ ].
i=0
As SK1 (R) = Z, [σ ] = [S2n−2 (v, w)]r for some r. Hence [∧i σ ] = r[∧i S2n−2(v, w)].
Therefore,
2n−1 2n−1
∑ (−1)i[∧i σ ] = r ∑ (−1)i [∧i S2n−2 (v, w)] = r wt(x1 , x2 , x23 , . . . , x2n−2
2n−1 )
i=0 i=0
= r(2n − 2)! wt(v).
Thus,
As v is of odd length, [S2n−2 (w, v)] = [S2n−2(w, v)t ], by the relations of Suslin. But
S2n−2(v, w)S2n−2 (w, v)t = I, and so [S2n−2 (v, w)] = [S2n−2 (w, v)]−1 .
Thus, one gets (2 + r(2n − 2)!) wt(v) = 0, a contradiction except when n = 2 and
r = −1.
The result of Corollary 4.3 was extended to all characteristics by Suslin and,
independently, by Mohan Kumar and Nori. Suslin’s proof from [12, 13] used the
methods of Sect. 5. It is clear that it suffices to consider rings An (k) defined over
a field k for these results. We give here another proof of Theorem 1.1 valid in all
characteristics using the method of Mohan Kumar and Nori. An expository account
of this method is presented in [15, Sect. 17]. It is based on a result coming from
Grothendieck’s Riemann–Roch theorem. Let R be a smooth affine domain over a
field k. Let Ai (R) be the Chow group of algebraic cycles of codimension i on Spec R
modulo rational equivalence. We write [R/I] for the class of the cycle Spec R/I.
Filter K0 (R) by letting F i K0 (R) be the subgroup generated by all [R/I] with htI ≥ i.
Then F 0 K0 (R) = K0 (R), F 1 K0 (R) = K
0 (R), and F N K0 (R) = 0 for N 0. Define
gr K0 (R) = F K0 (R)/F K0 (R). The map ϕ : Ai (R) → gri K0 (R) sending [R/I] to
i i i+1
[R/I] is well defined and is clearly onto. Grothendieck defines algebraic Chern
classes and a map ψ : gri K0 (R) → Ai (R) sending [P] − [Ri ] in gri K0 (R) to ci (P) in
Ai (R). The result we will need is the following.
322 Madhav V. Nori, Ravi A. Rao, and Richard G. Swan
k[x1 , . . . , xn , y1 , . . . , yn , z]
B = Bn = .
(∑ xi yi − z(1 − z))
Lemma 6.5. Let I = (y1 , . . . , yn , 1 − z). Then [B/I ] = (−1)n+1 [B/I] in K0 (B).
Proof. Let αi be the automorphism of B which switches xi and yi and let β be the
automorphism of B which switches z and 1 − z, all generators not mentioned being
fixed. Then I = α1 · · · αn β I so it will suffice to show that all αi and β induce
−1 on K
0 (B). It is enough to do the case of α1 and β . Let C = B/(x2 , . . . , xn ) =
x1
B1 [y2 , . . . , yn ]. The sequence 0 → C −
→ C → C/(x1 ) → 0 shows that [C/(x1 )] = 0
but C/(x1 ) = B/I ⊕ B/(β I) so β induces −1 on K
0 (B). Similarly, [C/(z)] = 0 and
y1
we have 0 → C/(z) → B/I ⊕ B/(α1 I) → C/(x1 , y1 , z) → 0. Since 0 → C/(x1 , z) − →
C/(x1 , z) → C/(x1 , y1 , z) → 0, we have [C/(x1 , y1 , z)] = 0 showing that [B/(α1 I)] =
−[B/I].
Theorem 6.6. Let A = An (k), where k is a nonzero commutative ring with unit, n is
odd and n > 3 . Then P(x1 , . . . , xn ) ≈ P(x1 , . . . , xn )∗ .
Proof. For the proof we can assume that k is a field. Recall that P(x1 , . . . , xn )∗ ≈
P(y1 , . . . , yn ) by [8, Prop. I.4.10]. Let J = ((1 − z)x1 , . . . , (1 − z)xn , zy1 , . . . , zyn )
in B. Then Jz = (y1 , . . . , yn )Bz and J1−z = (x1 , . . . , xn )B1−z . We have exact se-
quences 0 → L → Bn1−z → J1−z → 0 and 0 → N → Bnz → Jz → 0. These split over
Bz(1−z) as Jz(1−z) = Bz(1−z) is projective. Now Lz = P(x1 , . . . , xn ) over Bz(1−z) and
N1−z = P(y1 , . . . , yn ) over Bz(1−z) so Lz and N1−z are induced from P(x1 , . . . , xn ) and
P(y1 , . . . , yn ) over A by the map A → Bz(1−z) sending xi to xi /z and yi to yi /(1 − z). If
P(x1 , . . . , xn ) is self dual it follows that Lz ≈ N1−z so by Lemma 6.3, there is a finitely
generated projective B-module Q which maps onto J. Therefore, Lemma 6.2 shows
that [B/J] ≡ 0 mod (n − 1)! in K0 (B). Now z(1 − z) ∈ (x1 , . . . , xn ) ∩ (y1 , . . . , yn ) so
z(1 − z) = z2 (1 − z) + z(1 − z)2 lies in J. Therefore, B/J = B/(J, z) ⊕ B/(J, 1 − z) =
B/I ⊕ B/I . By Lemma 6.5, [B/J] = (1 + (−1)n+1 )[B/I] ≡ 0 mod (n − 1)!. As
[B/I] generates K
0 (B) = Z, we get 2 ≡ 0 mod (n − 1)! for n odd which implies
n ≤ 3.
With a bit more effort a similar proof can be given for Corollary 4.5 over any
nonzero commutative ring with unit, see [17]. This also contains an alternative
approach which avoids the use of the Riemann–Roch theorem.
References
6. A. Grothendieck, La théorie des classes de Chern, Bull. Soc. Math. France 86 (1958),
137–154.
7. D. Husemoller, Fibre bundles, Third edition, Graduate Texts in Mathematics 20, Springer-
Verlag, New York, 1994. ISBN: 0-387-94087-1.
8. T. Y. Lam, Serre’s Problem on Projective Modules, Monographs in Mathematics, Springer-
Verlag, Berlin-Heidelberg-New York, 2006.
9. M. Roitman, On stably extended projective modules over polynomial rings, Proc. Amer. Math.
Soc. 97 (1986), 585–589.
10. N. Steenrod, The Topology of Fibre Bundles. Princeton Mathematical Series, vol. 14. Prince-
ton University Press, Princeton, N.J., 1951.
11. A. A. Suslin, Stably free modules. (Russian) Math. Sbornik (N.S.) 102 (144), No. 4, 537–550,
632, 1977; English translation in Math. USSR Sbornik Vol 31, 479–491, 1977.
12. , Mennicke symbols and their applications in the K-theory of fields. Algebraic
K-theory, Part I (Oberwolfach, 1980), pp. 334–356, Lecture Notes in Math. 966, Springer,
Berlin-New York, 1982.
13. , K-theory and K-cohomology of certain group varieties. Algebraic K-theory, 53–74,
Adv. Soviet Math. 4, American Mathematical Socity, Providence, RI, 1991.
14. R. G. Swan, Vector bundles and projective modules, Trans. Amer. Math. Soc. 105(1962),
264–277.
15. , Vector Bundles, Projective Modules and the K-theory of Spheres, 432–522, Ch.
XVIII, Algebraic Topology and Algebraic K-theory, ed. W. Browder, Annals of Mathemat-
ics Studies 113, Princeton University Press, Princeton, NJ, 1987.
16. , Some stably free modules which are not self dual. http://www.math.uchicago.edu/
∼swan/stablyfree.pdf
17. , On a theorem of Mohan Kumar and Nori, http://www.math.uchicago.edu/∼swan/
MKN.pdf
18. R. G. Swan, J. Towber, A class of projective modules which are nearly free, J. Algebra 36
(1975), 427–434.
Homotopy Invariance of the Sheaf WNis
and of Its Cohomology
I. Panin
Summary A conjecture of F. Morel states that the motivic group π0,0 (k) of a field
k coincides with the Grothendieck-Witt group GW (k) of quadratic forms over k
provided that char(k) = 2. Morel’s proof of the conjecture is based among others
on the the following result: the Nisnevich sheaf WNis associated with the presheaf
X → W (X) is homotopy invariant and all its Nisnevich cohomology are homotopy
invariant too. A rather short and self-contained proof of the result is given here.
1 Introduction
A conjecture of Morel states that the motivic group π0,0 (k) of a field k coincides
with the Grothendieck-Witt group GW (k) of quadratic forms over k provided that
char(k) = 2. Morel’s proof of the conjecture presented in [M] is based, among
other things, on the the following result: the Nisnevich sheaf WNis associated with
the presheaf X → W (X) is homotopy invariant and all its Nisnevich cohomology
are homotopy invariant too. A rather short and self-contained proof of the result
is given here. It is inspired by Voevodsky’s paper [V] and may be regarded as an
extended version of the author’s preprint [P]. Other proofs are due to Hornbostel
[H] and Fasel [F].
We consider a field of characteristic different from 2. For an affine k-scheme S
we write W (S) for the Witt group of quadratic spaces over the ring k[S] of regular
functions on S. We consider the big Nisnevich site SmNis of k-smooth schemes and
the Nisnevich sheaf WNis associated with the presheaf X → W (S) on SmNis . Fur-
thermore, consider the big Zariski site SmZar of k-smooth schemes and the Zariski
I. Panin
St. Petersburg Branch, V.A. Steklov Institute of Mathematics (POMI), Russian Academy
of Sciences, 191023 St. Petersburg, Russia, e-mail: [email protected]; [email protected]
2010 Mathematics subject classification. 18F20, 14F42.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 325
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 21,
c Springer Science+Business Media, LLC 2010
326 I. Panin
sheaf WZar associated with the presheaf X → W (S) on SmZar . The main aim of the
preprint is to prove the following result.
Main Theorem 1.1. The Nisnevich sheaf WNis is strictly homotopy invariant. That
is, for each k-smooth variety X one has
(1) WNis (X) = WNis (X × A1 ) and
p p
(2) HNis (X ,WNis ) = HNis (X × A1 ,WNis ) for p > 0.
Moreover
(3) WZar = WNis , that is for any k-smooth X one has WZar (X ) = WNis (X ),
i (X ,W ) = H i (X ,W ) for any i > 0.
(4) HZar Zar Nis Nis
In particular, the sheaf WZar is strictly homotopy invariant.
The proof of the theorem is based on certain known results concerning Witt groups:
(1.2) For each field extension k /k one has W (k ) = W (A1k ), see [K] or [OP 00].
(1.3) For a point x ∈ X and the local scheme U = Spec(OX,x ) the Gersten-Witt
complex
res
res
res
0 → W (OX,x ) → W (k(X)) −→ W (k(y)) −→ W (k(z)) −→ · · ·
y∈U (1) z∈U (2)
is exact by [B].
(1.4) For each field extension k /k and each open U ⊂ A1k there is an exact sequence
of Witt groups (see [L, Thm. IX.3.1])
∑ resπx
0 → W (U) → W (k (t)) −−−−→ W (k (x)) → 0.
x∈U
We first provide a proof of the main theorem under the assumption that the ground
field is perfect and infinite. This is done in Sects. 2 to 4. Eventually, in Sects. 5 and 6
we prove the main theorem for an arbitrary ground field. For that we use arguments
based on the behavior of the Scharlau trace for odd-degree extensions of fields.
2 Voevodsky Trick
Suppose that the ground field k is perfect and infinite. Let F be a homotopy in-
variant Nisnevich sheaf on the big Nisnevich site SmNis . Assume furthermore that
for an integer j > 0 the functors HNis
i (−, F ) are homotopy invariant for each i < j.
Finally, assume that for each field extension k /k one has HNis
i (A1 , F ) = 0 for
k Nis
i > 0. The following result is extracted from [V, Sect. 4.6].
Theorem 2.1 (Voevodsky trick). Suppose for each essentially smooth local scheme
U and each essentially smooth divisor D on U the map
j j
HNis (U × A1 , F ) → HNis ((U − D) × A1, F )
Homotopy Invariance of the Sheaf WNis and of Its Cohomology 327
j ji∗
Ker [HNis (X × A1 , F ) −
→0
HNis (X , F )] = 0.
j j α β
{0} → HNis (X , R0 p∗ (F )) −
→ HNis (X × A1, F ) −
→ HNis
0
(X , R j p∗ (F )).
Moreover, R0 p∗ (F ) = F and α = p∗ .
j j i∗ j
Set H̄Nis (X × A1 , F ) = Ker [HNis (X × A1 , F ) −
→0
HNis (X , F )].
Lemma 2.4. Under the hypotheses of Theorem 2.1 for each essentially smooth
closed Z ⊂ X the map
j j
H̄Nis (X × A1 , F ) → H̄Nis ((X − Z) × A1, F )
is injective.
j β
H̄Nis (X × A1 , F ) −−−−→ 0 (X , R j p (F ))
HNis ∗
⏐ ⏐
⏐ ⏐
j β
H̄Nis ((X − Z) × A1, F ) −−−−→ HNis
0
(X − Z, R j p∗ (F )).
The horizontal arrows are both injective by Lemma 2.3. For each x ∈ X and
U = Spec(OX,x ) there exists an essentially smooth divisor D in U containing
j j
Z ∩ U. Moreover, the map HNis (U × A1 , F ) → HNis ((U − D) × A1 , F ) is injec-
tive. Thus for the open inclusion j : X − Z → X the adjunction sheaf morphism
R j p∗ (F ) → j∗ j∗ R j p∗ (F ) is injective. Taking the global sections we get the injec-
tivity of the right-hand side vertical arrow in the last diagram. The lemma follows.
Lemma 2.5. Under the hypotheses of Theorem 2.1 for each closed Z ⊂ X the map
j j
H̄Nis (X × A1 , F ) −
→ H̄Nis ((X − Z) × A1, F )
is injective.
328 I. Panin
j
Proof. Take an element a ∈ H̄Nis (X × A1 , F ) with a|(X−Z)×A1 = 0 and consider the
singular locus Sing(Z) of Z. It is a proper closed subset of Z because the field k is
perfect. Applying Lemma 2.4 to the pair (X (1) , Z (1) ) := (X − Sing(Z), Z − Sing(Z))
and the element a(1) = a|X (1) we conclude that 0 = a|X−Sing(Z) = a(1) . Replace Z by
Sing(Z) and Sing(Z) by Sing(Sing(Z)). Since k is perfect Sing(Sing(Z)) is a proper
closed subset of Sing(Z). Applying Lemma 2.4 to the pair
j j
Proof (of Theorem 2.1). Let a ∈ H̄Nis (X × A1 , F ). As HNis (A1k(X) , F ) = 0 there
exists a closed Z in X such that a vanishes on (X − Z) × A1. Lemma 2.5 completes
the proof of the Theorem.
3 Auxiliary Results
Lemma 3.3. For each field extension k /k the restriction of the presheaf W to the
affine line A1k is already a Zariski sheaf and even a Nisnevich sheaf.
Proof. For any open U ⊆ A1k one has an exact sequence of Witt groups (see (1.4))
∑ resπ
0 → W (U) → W (k (t)) −−−−→
x
W (k (x)) → 0.
x∈U
This sequence and Lemma 3.1 shows that W (U) = WZar (U). By Lemma 4.1
WZar (U) = WNis (U). The lemma follows.
∑ resπ
0 → WZar → W (k (t)) −−−−→
x
ix,∗W (k (x)) → 0.
x∈U
(The middle term here is considered as a constant sheaf.) The sequence is exact by
property (1.4), so it can be considered as a flabby resolution of the sheaf WZar on A1k .
As the sequence of global sections on A1k is exact too, the Zariski cohomology
i (A1 ,W ) vanishes for i > 0. The same arguments work for the Nisnevich
HZar k Zar
sheaf WNis .
Lemma 4.1. The Zariski sheaf WZar coincides with the Nisnevich sheaf WNis .
Proof. It suffices to prove that for an elementary Nisnevich square (see [MV, Prop.
3.1.4])
j˜
X̃ ←−−−− Ũ
⏐ ⏐
⏐
p
⏐q
j
X ←−−−− U
the sequence
p∗ + j ∗ q∗ − j̃ ∗
0 → WZar (X ) −−−→ WZar (X̃ ) ⊕ WZar (U) −−−→ WZar (Ũ) (1)
is exact. Recall that the square is elementary if it is Cartesian with an open inclusion
j and an etale morphism p such that p induces an isomorphism of the reduced closed
subschemes X̃ − Ũ and X − U.
The maps j∗ and j˜∗ are injective by Corollary 3.2. So to prove the exactness of
the sequence (1) it suffices to check that for a pair (ã, b) ∈ WZar (X̃ ) ⊕ WZar (U) with
j˜∗ (ã) = q∗ (b) there exist an a ∈ WZar (X ) such that p∗ (a) = ã and j∗ (a) = b.
Claim: The element b ∈ WZar (U) ⊂ W (k(X )) belongs to the subgroup WZar (X ).
Given the claim, set a = b ∈ WZar (X ). Since j˜∗ is injective this a is the required
element.
It remains to check that the claim holds. To do that it suffices to check that for
any codimension-one point x ∈ X the element b belongs to W (OX,x ) (see Lemma
3.1). This is the case if x ∈ U. So we may assume that x ∈ Z ∩ X and consider the
cartesian square
330 I. Panin
j˜
Spec(OX̃ ,x ) ←−−−− Spec(k(X̃ ))
⏐ ⏐
⏐
p
⏐q
j
Spec(OX,x ) ←−−−− Spec(k(X ))
which is the pull back of the square earlier by means of the morphism
Spec(OX,x ) → X. The last square gives rise to a diagram of the Witt groups.
j˜∗ resπ̃
0 / W (OX̃ ,x ) / W (k(X̃ )) / W (k(x)) /0
O O O
p∗ q∗ id
j∗
0 / W (OX,x ) / W (k(X )) resπ
/ W (k(x)) /0
The rows in the diagram are short exact sequences. The residue maps are defined
by local parameters in the discrete valuation ring OX,x and OX̃,x . Choosing a local
parameter π ∈ OX,x we get a local parameter π̃ = p∗ (π ) ∈ OX̃ ,x . This choice of
local parameters makes the last diagram commutative. Since q∗ (b) = j˜∗ (ã) one has
resπ̃ (q∗ (b)) = 0. Thus resπ (b) = 0 and b belongs to W (OX,x ). The claim follows,
whence the lemma.
Lemma 4.2. For a k-smooth scheme X and a closed subset Z ⊂ X let HZi (−,WNis )
be the Nisnevich sheaf associated with the presheaf U → HZ∩U
i (U,WNis ). For each
smooth principal divisor i : D → X one has:
(1) The sheaves HD1 (X ,WNis ) and i∗ (WNis ) on X are isomorphic.
(2) HDi (X,WNis ) = 0 for i = 1.
Proof. This proof is the most difficult one in this section. One can give a self-
contained proof based on the trace method used in the proof of [OP 99, Thm. B].
However, now we give a proof by a reference to [BGPW, Proof of Thm. 4.4].
Let GWC(X) and GWC(X, D) be the Gersten-Witt complex of X and of X with
the supports on D, respectively [BGPW, Def. 3.1]. A choice of the equation defining
the divisor D identifies the Gersten-Witt complexes GWC(X , D) and GWC(D)[1]
by [BGPW, Lemma 3.3]. Thus, H i (GWC(X , D)) = H i−1 (GWC(D)) for all i ≥ 0.
The Gersten-Witt complex GWC(X) (respectively GWC(D)) is the complex of the
global sections of a flabby resolution of the Nisnevich sheaf WNis on X (respectively
on D) [BGPW, Proof of Lemma 4.2]. Thus
Thus one has HDi (X ,WNis ) = H i−1 (D,WNis ). In the case of essentially k-smooth
local Henselian X and i = 1 the group H i−1 (D,WNis ) vanishes by [B, Theorem
4.2] (use property (1.3)). This proves the second assertion of the lemma. The
first one follows from the functoriality of the isomorphisms HD∩U 1
(U,WNis ) =
H (D ∩U,WNis ) = H (U, i∗WNis ), where U runs over Zariski open subsets of X .
0 0
Homotopy Invariance of the Sheaf WNis and of Its Cohomology 331
Proof (of the main theorem in the case of a perfect and infinite ground field). The
outline of our proof follows [V, Proof of Thm. 5.6 or better to say the arguments on
pages 122–123]. We check first that the Nisnevich sheaf WNis is homotopy invariant.
For a smooth X consider the diagram
J∗ can
WNis (X × A1 ) −−−−→ WNis (A1k(X) ) ←−−−− W (A1k(X) )
⏐ ⏐ ⏐
⏐
i∗0,X i∗0,k(X) ⏐ ⏐i∗
0,k(X)
j∗ can
WNis (X ) −−−−→ WNis (k(X)) ←−−−− W (k(X ))
with pull-back mappings J ∗ and j∗ . The maps “can” are isomorphisms by Lemma
3.3. The right-hand side i∗0,k(X) is an isomorphism by (1.2). Thus, the middle arrow
i∗0,k(X) is an isomorphism too. The map J ∗ is injective by Corollary 3.2. Thus, the
map i∗0,X is injective too. Obviously i∗0,X is surjective. Thus, it is an isomorphism
and p∗ : WNis (X ) → WNis (X × A1 ) is an isomorphism too. The homotopy invariance
follows.
We now prove the homotopy invariance of HNis 1 (−,W ). The Leray spectral
Nis
sequence for the projection p : X × A → X gives rise to a short exact sequence of
1
groups
α β
0 → HNis
1
(X, R0 p∗ (WNis )) −
→ HNis
1
(X × A1 ,WNis ) −
→ HNis
0
(X , R1 p∗ (WNis )).
By the homotopy invariance of the sheaf WNis the group HNis 1 (X , R0 p (W )) co-
∗ Nis
∗ ∗
incides with HNis (X,WNis ) and α = p . To show that p is an isomorphism it
1
suffices to check that the Nisnevich sheaf R1 p∗ (WNis ) vanishes. For that it re-
mains to show that for a Henselian essentially smooth local scheme X one has
1 (X × A1 ,W ) = 0.
HNis Nis
1 (A1
Since HNis k(X) ,WNis ) = 0 by Corollary 3.4, it suffices to check that the map
1
HNis (X × A1 ,WNis ) → HNis
1
(A1k(X) ,WNis )
is injective. By the Voevodsky trick 2.1 it suffices to check that for a smooth divisor
i : D → X the map
1
HNis (X × A1 ,WNis ) → HNis
1
((X − D) × A1,WNis )
is injective. For that it suffices to verify that the boundary map
∂ 1
0
HNis ((X − D) × A1,WNis ) −−
→ HD×A
A 1
1 (X × A ,WNis )
1
is an isomorphism, where HZ1 (−,WNis ) is the Nisnevich sheaf associated with the
presheaf U → HZ∩U
1 (U,W ).
Nis
and the sheaf HZ0 (WNis ) := HZ0 (−,WNis ) vanishes by Corollary 3.2.
and the homotopy invariance of the sheaf WNis prove the lemma.
Lemma 4.5. For the Henselian local scheme X the boundary map ∂ : WNis (X − D)
∂
−
→ HD1 (X ,WNis ) is surjective.
1 (X ,W ). It van-
Proof. In fact, the next term in the localization sequence is HNis Nis
ishes, because X is local for the Nisnevich topology.
∂A1 γA1
WNis ((X − D) × A1) / H 1 1 (X × A1 ,WNis ) / H 0 (X × A1 , H 1 1 (WNis ))
O D×A O O D×A
p∗ p∗ p∗
∂ γ
WNis (X − D) / H 1 (X ,WNis ) / H 0 (X , H 1 (WNis )).
D D
The maps γ and γA1 are isomorphisms by Lemma 4.3. The right-hand side map p∗
is an isomorphism by Lemma 4.4. Thus, the middle arrow p∗ is an isomorphism
too. As ∂ is a surjection the map ∂A1 is surjective too. The homotopy invariance of
1 (−,W ) is proved.
HNis Nis
Now prove the homotopy invariance of HNis i (−,W ) for i > 1. We proceed by
Nis
induction on i. So assuming that the homotopy invariance holds for all i < j prove
it for i = j. The Leray spectral sequence for the projection p and the inductive
hypothesis give the following short exact sequence
j αj β
{0} → HNis (X, R0 p∗ (WNis )) −
→ HNis (X × A1 ,WNis ) −
→ HNis
0
(X , R j p∗ (WNis )).
Homotopy Invariance of the Sheaf WNis and of Its Cohomology 333
j
By the homotopy invariance of the sheaf WNis the group HNis (X , R0 p∗ (WNis )) coin-
cides with HNis (X ,WNis ) and α = p∗ . To show that p∗ is an isomorphism it suffices
j
to check that the Nisnevich sheaf R j p∗ (WNis ) vanishes. For that it remains to show
j
that for a Henselian essentially smooth scheme X one has HNis (X × A1 ,WNis ) = 0.
j
As HNis (Ak(X) ,WNis ) = 0 by Corollary 3.4, it suffices to check that the map
1
j j
HNis (X × A1 ,WNis ) → HNis (A1k(X) ,WNis )
is injective. By the Voevodsky trick 2.1 it suffices to check that for an essentially
smooth divisor i : D → X on an essentially smooth Henselian scheme X the map
j j
HNis (X × A1 ,WNis ) → HNis ((X − D) × A1,WNis )
is injective. The localization sequence shows that it remains to check the vanishing
j
of the group HD×A1 (X × A1 ,WNis ) in this case. The spectral sequence
q p+q
2
E p,q = H p (X × A1 , HD×A1 (−,WNis )) ⇒ HD×A1 (X × A ,WNis )
1
j
shows that the group HD×A1 (X × A1 ,WNis ) vanishes if for all pairs (p, q) with p +
j q
q = j the group HD×A 1 (X × A ,WNis ) vanishes. The sheaves HD×A1 (WNis ) vanish
1
By the inductive hypothesis, H j−1 (D × A1 ,WNis ) = H j−1 (D,WNis ) and the last
group vanishes because D is local Henselian and j − 1 > 0. Thus for local essen-
j
1 (X × A ,WNis ) vanishes. The inductive
tially smooth Henselian X the group HD×A 1
Assuming the lemma, we complete the proof of the strict homotopy invariance
as follows. Consider the commutative diagram
Ex
i (X × A1 ,W ) −−−−→ H i ((X × A1 ) ,W )
HNis Nis Nis K Nis
⏐ ⏐
∗ ⏐
i0,X
⏐ ∗
i0,XK
i (X ,W ) ex
HNis Nis −−−−→ i (X ,W ),
HNis K Nis
where the maps Ex and ex are induced by the scalar extension K/k. The map i∗0,XK is
an isomorphism because the main theorem was proved in Sect. 3 for a perfect field.
The map Ex is injective by the lemma. Thus i∗0,X is injective too. Clearly i∗0,X is
surjective. Thus i∗0,X is an isomorphism and p∗ is an isomorphism too. The strict
homotopy invariance follows. It remains to prove the lemma.
Proof (of Lemma 5.1). We may replace the field extension K/k by a finite purely
inseparable extension l/k and even by an extension with l = k(a1/p ) for some a ∈ k.
For such an l, set α = a1/p and consider the k-linear map tr: l → k given by α i → 0
for i = 0, 1, . . . , p − 2 and α p−1 → 1. For a k-smooth variety X set
Tr ◦ ex : HNis
i
(X ,WNis ) → HNis
i
(X ,WNis )
coincides with the multiplication by 1. Thus it is the identity and the map ex :
i
HNis (X ,WNis ) → HNis
i
(Xl ,WNis ) is injective. The lemma follows.
Let p > 0 be the characteristic of the finite field k. Let l be an odd prime number
different from p. Let K/k be an infinite Galois extension with Galois group equal
to the l-adic integers Zl . The field K is perfect and infinite. Thus, the main theorem
holds for varieties over K. Now applying literally arguments from Sect. 5 we get
the main theorem in the case of smooth varieties over the field k. Whence the main
theorem.
Acknowledgements The author is very grateful for the financial support to the Ellentuck Fund of
the year 2004, to the RTN-Network HPRN-CT-2002-00287, to the grant “Research in the funda-
mental domains of the modern mathematics” and to the DFG-RFBR grant 09-01-91333-NNIO-a.
References
[B] P. Balmer. Witt cohomology, Mayer-Vietoris, homotopy invariance, and the Gersten
conjecture, K-Theory, 23 (2001), 15–30.
[BGPW] P. Balmer, S. Gille, I. Panin, Ch. Walter. Witt cohomology, Mayer-Vietoris, homotopy
invariance, and the Gersten conjecture, K-Theory, 23 (2001), 15–30.
[K] M. Karoubi Localisation de formes quadratiques.II, Ann. Sci. Éc. Norm. Sup. vol.8
(1975), 99–155.
[F] J. Fasel. Groupes de Chow-Witt, Mém. Soc. Math. Fr (N.S.) No.113 (2008), viii+197.
[H] J. Hornbostel. A1 -representability of Hermitian K-theory and Witt groups, Topology
44 (2005), no.3, 661–687.
[L] T.Y. Lam. Introduction to quadratic forms over fields, Graduate Studies in Math.,
vol. 67, American Mathematical Society, Providence, RI, 2004.
[M] F. Morel. A1 -algebraic topology over a field. www.math.uiuc.edu/K-theory/0806,
2006.
[MV] F. Morel, V. Voevodsky. A1 -homotopy theory of schemes, Publ. IHES. 90 (2001),
45–143.
[OP 99] M. Ojanguren, I. Panin. A Purity Theorem for the Witt Group, Ann. Sci. Ecole Norm.
Sup., Serie 4, 32 (1999), 71–86.
[OP 00] . The Witt group of Laurent polynomials. Enseign. Math. (2) 46 (2000), no.
3–4, 361–383.
[P] I. Panin. Homotopy invariance of the sheaf WNis and of its cohomology. www.math.
uiuc.edu/K-theory/0715, 2004.
[V] V. Voevodsky. Cohomological theory of presheaves with transfers, In “Cycles, Trans-
fers and Motivic Homology Theories”, by Vladimir Voevodsky, Eric M. Friedlander,
and Andrei Suslin. pp. 87–135.
Imbedding Quasi-split Groups
in Isotropic Groups
M.S. Raghunathan
Summary Borel and Tits have shown that in any isotropic absolutely almost simple
simply connected linear algebraic group G over a field k contains a k-split semisim-
ple k-subgroup containing a maximal k-split torus T of G as well as arbitrarily cho-
sen k-points in the root-groups (with respect to T ) corresponding to the roots in a
simple system of the reduced system of k-roots of G. In this paper we prove that
one can in fact imbed somewhat larger quasi-split groups in isotropic groups when
the characteristic of k is different from 2.
1 Introduction
M.S. Raghunathan
School of Mathematics, Tata Institute of Fundamental Research, Homi Bhabha Road, Mumbai
400 005, India, e-mail: [email protected]
2010 Mathematics subject classification. 20G.
J.-L. Colliot-Thélène et al. (eds.), Quadratic Forms, Linear Algebraic Groups, 337
and Cohomology, Developments in Mathematics 18, DOI 10.1007/978-1-4419-6211-9 22,
c Springer Science+Business Media, LLC 2010
338 M.S. Raghunathan
unipotent k-subgroup of G normalised by T ; it has for its Lie algebra uα , the sum of
all the root-spaces corresponding to the roots of the form n.α , n a positive integer. It
is known that for α in Φ , the only positive integral multiples of α that can be roots
are α and 2α . A root α is multipliable if and only if 2α is a root. Note that if α is not
multipliable, uα = gα and the group Uα is abelian. The set Δ contains at most one
multipliable root. With this notation we can formulate the main result of this note.
Remarks. Here RK/k is the Weil restriction of scalars functor from the category
of K-varieties to the category of k-varieties. If uβ = uβ with β nonmultipliable or
if uβ is in U2β , we recover a theorem of Borel-Tits [1, Theorem 7.2]; in this case
moreover, k = K. However the proof in Borel-Tits for this special case is much
more satisfactory: unlike the proof here it makes no use of the classification of
semisimple groups over k (Tits [4]); nor is there a restriction on the characteristic.
Actually, the proof given later for the classical groups works also in characteristic 2
for groups of type other than Bn , n > 2 and Dn , n > 3. The proof for the exceptional
cases is reduced to that of classical groups, sometimes of the earlier two types. It
seems likely that the result is true in the case of characteristic 2 as well but would
need more delicate arguments. The special case of Theorem 1 when the k-rank of G
is 1 is dealt with in Raghunathan [3, Appendix]. One expects that this theorem may
be helpful in obtaining refined results about the classification of isotropic groups.
2 Classical Groups
2.1 Notation
We continue with the notation introduced earlier. Further L will denote a field
containing k and for an algebraic k-group H, H(L) will denote the L-points of H.
An algebraic k-group is uniquely determined if its L-points are specified for every
field L containing k. In the sequel we will often define a k-subgroup of G by simply
describing its L-points for all L containing k. Also, characters on a k-torus will be
defined by specifying their values on L-points of the torus.
Imbedding Quasi-split Groups in Isotropic Groups 339
2.2 Type 1 An
The assumption char(k) = 2 is not needed here. Every absolutely almost simple
simply connected k-group G of (inner) type 1 An (n, a positive integer, is the rank of
G over an algebraic closure of k) of k-rank l has the following description. There
is a central division algebra D over k of degree d, a divisor of n + 1 with l + 1 =
(n + 1)/d, such that G is isomorphic over k to SLl+1 (D). For a field L containing k,
the group G(L) is the Special Linear group SL(l + 1, D ⊗ L) of (l + 1) × (l + 1)-
matrices with entries in D ⊗ L of (Dieudonné) determinant 1.
Let T be the torus in G for which the L-points T (L) are the diagonal matrices in
G(L) with entries in L; it is a maximal k-split torus of G. The roots of G with respect
to T are the characters αi, j , 1 ≤ i = j ≤ l + 1, defined as follows: for the diagonal
matrix t˜ in T (L) with diagonal entries t1 ,t2 , . . . ,tl put αi, j (t˜) = tit −1
j . A positive
system is given by the collection {αi, j | i < j} and the corresponding simple system
Δ = {αi,i+1 | 1 ≤ i ≤ l}. For i = j, 1 ≤ i, j ≤ l +1, the group Uαi, j (L) is then precisely
{1 + b.Ei, j | b ∈ D⊗ L} where Ei, j is the matrix whose (i, j)-th entry is 1 and all other
entries are zero. The elements uα , α ∈ Δ , are thus of the form 1 + bi .Ei,i+1 with bi in
D, 1 ≤ i ≤ l. For 1 ≤ i ≤ l, set ti = bi .bi+1 .bi+2 ....bl+1 . Let t˜ be the diagonal matrix
whose i-th diagonal entry is ti . Then if we replace the uα by their conjugates under
t˜−1 , we see that they are all in SL(l + 1, k). Let β = αi,i+1 and uβ = 1 + b.Ei,i+1 with
b ∈ D (after the conjugation by b̃−1 ).
Let K be the field generated by k and b. Then we take H to be the K-group
SLl+1,K and we have an obvious morphism of RK/k (H) in G satisfying the required
conditions. This proves the theorem for groups of inner type An .
2.3 Type Bn , n ≥ 2
In this case, the assumption char(k) = 2 is used. The group G is the Spin Group of a
nondegenerate quadratic form q of Witt-index l on a (2n + 1)-dimensional k-vector
space V . Let B = Bq be the bilinear form associated to q: for v, w in V , 2.B(v, w) =
q(v + w) − q(v) − q(w). Fix a basis {ei | 1 ≤ i ≤ 2n + 1} of V such that the following
hold:
1. For 1 ≤ i, j ≤ l and 2n + 2 − l ≤ i, j ≤ 2n + 1, B(ei , e j ) = 0.
2. For 1 ≤ i, j ≤ l, B(ei , e2n+2− j ) = δi j
3. For 1 ≤ i ≤ l or 2n + 2 − l ≤ i ≤ 2n + 1 and l + 1 ≤ j ≤ 2n + 1 − l, B(ei , e j ) = 0.
340 M.S. Raghunathan
The basis enables us to identify G (L) (for G the Special Orthogonal Group of q)
as a subgroup of GL(2n + 1, L). Let π : G → G be the natural map. We take for
T the inverse image under π of the subgroup T of G for which T (L) consists of
diagonal matrices t˜ (in G (L)) with the i-th entry ti = 1 if l + 1 ≤ i ≤ 2n + 1 − l and
t2n+2−i = ti−1 with ti = 0 for 1 ≤ i ≤ l. For 1 ≤ i < l, let αi be the character on T
defined by setting αi (t˜) = ti /ti+1 for t˜ in T (L). Also define αl by αl (t˜) = tl . By
composing the αi with π we get characters on T which also we denote by αi . Then
Δ = {αi | 1 ≤ i ≤ l} is a simple system for an order on X (T ). It is easily seen that
Δ has no multipliable root.
For 1 ≤ i < l, one has uαi (e j ) = e j if j = i + 1 or 2n + 2 − i while uαi (ei+1 ) =
ei+1 + ei .xi and uαi (e2n+2−i ) = e2n+2−i − e2n+1−i.xi with xi in k. Next, uαl (e j ) = e j
for 1 ≤ j ≤ l and 2n + 1 − l < i ≤ 2n. Also, one has uαl (e j ) = e j + el .y j for l + 1 ≤
j ≤ 2n + 1 − l with y j in k and uαl (e2n+2−l ) = e2n+2−l + v + el .c with v in E and
c in k. Changing the basis {ei | l + 1 ≤ i ≤ 2n + 1 − i} of E, if necessary, we can
assume that v = el+1 . It follows then that uαl (e2n+2−l ) = e2n+2−l + el+1 + el .c and
y j = 0 for l + 2 ≤ j ≤ 2n + 1 − l.
Now if β is one of the αi with i < l then (since Uαi is 1-dimensional), the Special
Orthogonal Group of the restriction of q to the k-linear span W of the {ei | 1 ≤ i ≤ l}
and the {ei | 2n + 2 − l ≤ i ≤ 2n + 1} is split and the corresponding Spin Group is the
requisite quasi-split subgroup. If β = αl and uβ (e2n+2−l ) = e2n+2−l + w + el .c , let
W be the k-linear span of {ei | 1 ≤ i ≤ el+1 }, w and {ei | e2n+2−l ≤ i ≤ 2n + 1. The
Special Orthogonal Group of the quadratic form q restricted to W is then quasi-split
and its Spin Group is the required H.
Suppose now that G is a classical group but is not of inner type An or Bn . We assume
that G is not quasi-split. (The proof later works also if char(k) = 2, provided that G
is not of type Dn , n > 3.) One has, in these cases the following uniform description
of G. There is an extension field k of k of degree at most 2, a central division
algebra D of degree d over k with a k-linear involution σ (which is nontrivial if
D = k ) whose restriction to k is trivial or the Galois conjugation according as k = k
or [k : k] = 2, a right vector space V over D (of dimension N, say, where N is
necessarily even when k = k = D) and a nondegenerate σ -hermitian symmetric
bilinear form h on V such that G is the simply connected cover of the Special Unitary
Group SU(h) of h. Further G is of k-rank l if and only if h has Witt-index l. Our
conventions are such that h : V × V → D satisfies the following conditions: h is
biadditive and h(v.x, w.y) = σ (x).h(v, w).y for v, w in V and x, y in D. Note that
when k = k = D, G is the spin group of a quadratic form in an even number N of
variables. Also when G is of type Cn , k = k and we assume as we may that D = k:
if k = D, the group is split which we have excluded.
Imbedding Quasi-split Groups in Isotropic Groups 341
Now V admits a basis {ei | 1 ≤ i ≤ N} (over D) such that the following conditions
hold. Set vi = el+i for 1 ≤ i ≤ m where m = N − 2l and for 1 ≤ i ≤ l, fi = eN+1−i .
We then have
1. h(ei , e j ) = h( fi , f j ) = 0 for 1 ≤ i, j ≤ l.
2. h(ei , f j ) = δi j for 1 ≤ i, j, ≤ l.
3. h(ei , v j ) = h( fi , v j ) = 0 for 1 ≤ i ≤ l and 1 ≤ j ≤ m.
4. h(vi , v j ) = 0 for 1 ≤ i, j ≤ m with i = j.
Further, for any w in the D-linear span W of {vi | 1 ≤ i ≤ m}, h(w, w) = 0.
In the sequel, elements of SU(h)(L) will be regarded as matrices with entries in
D ⊗k L with respect to this basis. We define a maximal k-split torus T in SU(h)
as follows. An element of T (L) is a (D ⊗ L)-linear automorphism b̃ of V ⊗ L for
which we have b̃(ei ) = ei .bi with bi in L; further bi = 1 for l + 1 ≤ i ≤ l + m and for
1 ≤ i ≤ l, bN+1−i = σ (bi )−1 .
For 1 ≤ i < l, let αi denote the character on T defined by setting αi (b̃) = bi /bi+1
for b̃ in T (L). Let αl (b̃) = bl or b2l according as m > 0 or m = 0. Let π : G → SU(h)
be the natural projection and T the inverse image of T in G under π . Then T
is a maximal k-split torus in G. We treat characters on T as characters on T by
composing them with π . Then Δ = {αi | 1 ≤ i ≤ l} is a simple system of k-roots of
G with respect to T for a suitable order on X (T ). We note that if m > 0, the root αl
is multipliable (and 2αl is a root). Also, if m = 0, no root in Δ is multipliable.
Suppose now that i is such that 1 ≤ i < l and uαi = 1. Then one has uαi (e j ) =
e j for j = i + 1 or N − i + 1 while uαi (ei+1 ) = ei+1 + ei .xi for some xi = 0 in D
and uαi (eN−i+1 ) = eN−i+1 − eN−i .σ (xi ). Next, when m = 0 one has uαl (ei ) = ei
if i = l + 1 and uαl (el+1 ) = el+1 + el .xl with xl ∈ D. Also in this case no root is
multipliable. If m > 0, 2αl is a root and we have uαl (ei ) = ei unless l + 1 ≤ i ≤ l +
m+ 1 while uαl (el+ j ) = el+ j + el .y j for 1 ≤ j ≤ m and uαl (el+m+1 ) = el+m+1 + w0 +
el .c with y j ∈ D and w0 a vector in the D-linear span W of {ei | l + 1 ≤ i ≤ l + m}.
Assume now that m = 0 so that N = 2l. Let b̃ be the diagonal matrix in U(h)
whose i-th entry bi = 1.x1 .x2 ....xi−1 for 1 ≤ i ≤ l (and bN+1−i = σ (bi )−1 ). Replacing
the uαi by their conjugates under b̃ we see that we may assume that xi = 1 for
1 ≤ i ≤ l − 1. We set xl = x. Relative to the basis {ei = ei .x | 1 ≤ i ≤ l} ∪ { fi | l + 1 ≤
i ≤ N}, the elements uαi are seen to be in GL(N, k). The element x is a symmetric
or antisymmetric element in D and the assignment of x−1 .σ (b).x to each b in D
is again an involution σ and h(v, w) = x−1 .h(v, w), for v, w in V , is a hermitian
or antihermitian form on V with respect to the involution σ . If x is symmetric
(respectively, antisymmetric) for σ , h is symmetric or antisymmetric (respectively,
antisymmetric or symmetric) according as h is symmetric or antisymmetric.
Now G is also the Special Unitary Group SU( h). One now has h(ei , ej ) =
h( f , f ) = 0, for 1 ≤ i, j ≤ l and h(e , f ) = δ . The form h is thus obtained as
i j i j ij
an extension of a split nondegenerate symmetric or antisymmetric bilinear form B
(over k) on W = kN , the k-span of {ei | 1 ≤ i ≤ l} and { f j | 1 ≤ j ≤ l} (we have
N = 2l, since m = 0) as a hermitian-symmetric or antisymmetric form to V = DN .
The elements uαi are now easily seen to be k-points in the quasi-split group over k
that preserves B.
342 M.S. Raghunathan
3 Exceptional Groups
r r
2E 16 ri
r
6,2 BC2 r r
2E 16 ri ri r
r
6,2 G2 r r
r
31
E7,2 BC2 r i
r r r r i
r
r
28
E7,3 C3 r
i r
i r r r i
r
r
9
E7,4 F4 r i
r r i
r i
r i
r
r
78
E8,2 G2 r
i r
i r r r r r
r
66
E8,2 BC2 r
i r r r r r i
r
r
28
E8,4 F4 r
i r
i r
i r r r r
i
Table 1 Tits indices for non-quasi-split exceptional simple groups of k-rank ≥ 2, together with the
type of the relative root system.
G of k-rank the cardinality of Δ . The intersection of T with this group has for
its identity connected component a maximal k-split torus in the group. We will
now establish the following result which yields our main theorem for exceptional
groups by an obvious induction argument. The lemma itself will be proved by a
case-by-case check using Tits’ classification together with an induction on dim G and
assuming the result for k-rank 1 groups: the case when the k-rank is 1 is dealt with in
Raghunathan [3]. Table 1 lists the Tits indices for nonquasi-split exceptional groups
of k-rank ≥ 2. The facts that are used in proving Lemma 2 for these exceptional
groups can be read off easily from the Tits indices.
344 M.S. Raghunathan
3.2
We now spell out the induction argument needed to deduce Theorem 1 for excep-
tional groups using Lemma 2. We assume that Theorem 1 holds for all exceptional
groups of dimension < dim G over all fields. Let then G be exceptional. Accord-
ing to the lemma, the uαi , 1 ≤ i ≤ l and uβ are all in the centraliser G of X or
S as the case may be. Let G1 be the commutator subgroup of G . Then G1 is
semisimple simply connected and of dimension strictly lower than that of G. Fur-
ther G and G1 have T as a common maximal k-split torus. Moreover, as the uαi ,
1 ≤ i ≤ l and uβ are contained in G1 as well, it follows that G and G1 have the
same k-root system viz. Φ . As Φ is irreducible, the group G1 is k-simple, i.e., it
has no proper connected normal algebraic subgroup defined over k. It follows that
G1 is of the form RK/k (G1 ) where K/k is a finite separable extension of k and G1
is an absolutely almost simple group over K [4, 3.1.2]. Either G1 is classical or
it is exceptional and dimK (G1 ) < dimk (G); thus (in either case) there exists a fi-
nite separable extension L/K, a quasi-split group H1 over L, and a K-morphism
F : H1 = RL/K (H ) → G1 with a split torus mapping isomorphically onto a maximal
split torus and with the uαi , 1 ≤ i ≤ l and uβ in the image of F1 . We can now take
for H the group RK/k (H1 ) = RL/k (H ).
The k-rank of the group is 2. One reads off from the Tits index that both the simple
roots are nonmultipliable. The centraliser Z(T ) of a maximal k-split torus T is of
the form T.M where the isotropic kernel M is a group of type D4 – it can in fact
be identified (assuming that G is simply connected) with the Spin Group Spin(q) of
an anisotropic quadratic form q in eight variables over k; moreover the groups Uα
for α in Δ are k vector spaces with the action of M on them being the two half-spin
representations. Alternatively, one may think of one of them as the natural represen-
tation of SO(q) and the other as one of the half-spin representations. We assume (as
we may) that β = α1 and that the representation on Uβ is the natural representation
of SO(q). If uα1 and uβ are not linearly independent we are in the special case dealt
with in Borel-Tits [1]. We assume then that they are linearly independent. Let M
Imbedding Quasi-split Groups in Isotropic Groups 345
denote the centraliser of uα1 in M. Then one sees immediately that it can be identi-
fied with Spin(q ) where q is the restriction of q to the (codimension 1) k-subspace
V of Uα1 which is the orthogonal complement of uα1 in Uα1 . Now the representation
of M on Uα2 is the spin representation of this Spin Group Spin(q ).
3.4
We now examine the spin representation of M more closely. Fix a maximal torus S
in M . Then over an algebraic closure k̄ of k, the representation space W = Uα2
decomposes into 1-dimensional eigenspaces for S. We fix a simple system of
(absolute) roots (of M with respect to S) {θi | 1 ≤ i ≤ 3} and let B+ denote the Borel
subgroup determined by this simple system. Let λ+ denote the highest weight in W
for this simple root-system and e a nonzero highest weight vector. Let X denote the
orbit of e under M and f : M → X the orbit map. One checks easily that the map
has rank 7 (everywhere) on M . Now if v is any nonzero vector in W (k̄), the set
k̄∗ .B+ (v) contains e in its closure. A semicontinuity argument now shows that the
rank of the orbit map for any nonzero vector in W (k̄) is greater than equal to 7.
Now there is a nondegenerate anisotropic quadratic form q (over k) on W which
is invariant under M, hence under M . Thus the orbit of v is contained in the set
{w ∈ W | q(w) = q(v)}, a subvariety of codimension 1 (hence of dimension 7). It
follows now that the isotropy subgroup Hv in M for any v = 0 in M (k) is defined
over k and has dimension 14. Moreover over k̄ the isotropy groups at all points
in Ω = {w ∈ W | q(w) = 0} are all conjugates. One concludes from this that the
group Hv is semisimple of type G2 ; and setting H = Huα2 , we see that H is an
anisotropic group of type G2 over k. Moreover, the representation of H on V is a
k-form of the Weyl module corresponding to the unique irreducible representation
of dimension 7 of the split G2 in characteristic 0. This representation is irreducible
except when char(k) = 2. (When char(k) = 2, the vector uα1 is invariant under M
and hence under H and the representation on the quotient 6-dimensional space V is
irreducible.)
For a detailed analysis of the earlier facts, see for example [2].
3.5
Now we can analyse the orbits of H acting on V entirely analogously to what we did
for the spin representation of M to arrive at the following conclusion. The isotropy
subgroup I of H at uβ is defined over k and has dimension 8. As any maximal
connected unipotent subgroup of H (over k̄) has dimension 6, we see that I contains
a nontrivial k-torus C. This group C evidently centralises the uαi , i = 1, 2 and uβ .
This settles the case of 1E6,2
28 .
346 M.S. Raghunathan
In these cases too, the k-rank is 2 and both the simple k-roots α1 and α2 are nonmul-
tipliable. The anisotropic kernel M commutes with one of the Uαi , i = 1, 2, say Uα1
which is of dimension 1. It follows that M is the anisotropic kernel of the k-rank 1
group G corresponding to α2 . By the result for k-rank 1 groups, there is a nontrivial
torus S in M or a semisimple element X in m which is not central in G . Hence the
lemma in this case.
3.7 Groups of Type 2 E6,2
16
31
3.8 Groups of Type E7,2
We take α1 to be the multipliable root; then α2 is the circled root at the (short) end
of the Dynkin diagram. The anisotropic kernel M is k-isomorphic to the product
Spin(q) × Spin(q ) where q (respectively, q ) is an anisotropic quadratic form in
Imbedding Quasi-split Groups in Isotropic Groups 347
28 and E 9
3.9 Groups of Type E7,3 7,4
The k-rank of the groups of the first type is 3. In this case the anisotropic kernel
M centralises one of the Uαi , 1 ≤ i ≤ 3 which we take to be α1 . We see that M is
also the anisotropic kernel of the k-rank 2 subgroup H determined by α2 and α3 ;
and since dim H < dim G, by the induction hypothesis there is a nontrivial k-torus S
centralising uαi for i = 2, 3. Thus, we have a torus centralising T as well as the three
unipotent elements.
The groups of the second type have rank 4. The anisotropic kernel here cen-
tralises two of the Uαi , say Uα1 and Uα2 which are both of dimension 1; thus M is
also the anisotropic kernel of the k-rank 2 group H corresponding to the pair of roots
α3 and α4 . By the induction hypothesis there is a torus in M that centralises uαi for
i = 3, 4 and uβ . Hence the result for groups of the second type as well.
348 M.S. Raghunathan
78 and E 28
3.10 Groups of Type E8,2 8,4
In both these cases there is a k-root α1 , say, such that Uα1 is of dimension 1 and
commutes with the anisotropic kernel M. It follows that M is the anisotropic kernel
of the k-group H corresponding to the subset {αi | i = 1} as well. Further, β = αi for
some i = 1. As dim H < dim G, the induction hypothesis yields the desired result.
66
3.11 Groups of Type E8,2
Here the k-rank is 2 and one of the roots, say α1 , is multipliable. The anisotropic
kernel M is Spin(q), the Spin group of an anisotropic quadratic form q on a k-vector
space of dimension 12 which may be taken to be Uα1 /U2α1 , which we denote by V
in the sequel. The subgroup of M that fixes uα1 is the group M = Spin(q ) where
q is the restriction of q to a codimension 1 k-subspace V of V . The representation
of M on Uα2 is then the spin representation of Spin(q ). An analysis of the orbit
structure of the action of Spin(q ) in the spin representation shows that the orbit map
for any orbit is a separable morphism and there are three kinds of orbits (over the
algebraic closure of k): the trivial orbit of 0; the orbits for which the isotropy is the
unique codimension-1 subgroup of a parabolic subgroup conjugate to the parabolic
subgroup determined by the spin representation; and generic orbits for which the
isotropy is the semisimple part of the Levi subgroup of a parabolic subgroup con-
jugate to the parabolic subgroup determined by the spin representation. Since M is
anisotropic over k, it follows that the isotropy H at uα2 for the action of M on Uα2
is defined over k and semisimple of type A4 over an algebraic closure of k. Let S be
a maximal k-torus in H. Let U be the inverse image in Uα1 of the 1-dimensional
k-subspace spanned by the image v of uα1 in V under the natural map Uα1 → V . The
group U2α is of dimension 1 and (hence), M acts trivially on this group. It follows
that S acts trivially on U . Thus S centralises uαi for i = 1, 2.
Once again a more detailed analysis of the action of M on the Uαi can be found
in [2].
Acknowledgements My thanks are due to Skip Garibaldi who pointed out a gap in the proof
in 3.3 in an earlier version.
References
1. A. Borel and J. Tits, Groupes réductifs. Publ. Math. IHES 27 (1965), 55–151.
2. J. Igusa, A classification of spinors up to dimension twelve. Am. J. Math., 92 (1970) 997–1028.
3. M.S. Raghunathan, On the congruence subgroup problem, II. Inv. Math. 85 (1986), 73–117.
4. J. Tits, Classification of semisimple algebraic groups. Algebraic groups and discontinuous
groups, Proc. Symp. Pure Math. 9 (1966), 33–62.