Quat Book
Quat Book
Quat Book
John Voight
[email protected]
v.0.9.28
May 20, 2021
Preface
Goal
Quaternion algebras sit prominently at the intersection of many mathematical subjects.
They capture essential features of noncommutative ring theory, number theory, 𝐾-
theory, group theory, geometric topology, Lie theory, functions of a complex variable,
spectral theory of Riemannian manifolds, arithmetic geometry, representation theory,
the Langlands program—and the list goes on. Quaternion algebras are especially
fruitful to study because they often reflect some of the general aspects of these subjects,
while at the same time they remain amenable to concrete argumentation. Moreover,
quaternions often encapsulate unique features that are absent from the general theory
(even as they provide motivation for it).
With this in mind, the main goal in writing this text is to introduce a large subset
of the above topics to graduate students interested in algebra, geometry, and number
theory. To get the most out of reading this text, readers will likely want to have
been exposed to some algebraic number theory, commutative algebra (e.g., module
theory, localization, and tensor products), as well as the fundamentals of linear algebra,
topology, and complex analysis. For certain sections, further experience with objects in
differential geometry or arithmetic geometry (e.g., Riemannian manifolds and elliptic
curves), may be useful. With these prerequisites in mind, I have endeavored to present
the material in the simplest, motivated version—full of rich interconnections and
illustrative examples—so even if the reader is missing a piece of background, it can be
quickly filled in.
Unfortunately, this text only scratches the surface of most of the topics covered
in the book! In particular, some appearances of quaternion algebras in arithmetic
geometry that are dear to me are absent, as they would substantially extend the length
and scope of this already long book. I hope that the presentation herein will serve as a
foundation upon which a detailed and more specialized treatment of these topics will
be possible.
I have tried to maximize exposition of ideas and minimize technicality: sometimes
I allow a quick and dirty proof, but sometimes the “right level of generality” (where
things can be seen most clearly) is pretty abstract. So my efforts have resulted in a level
of exposition that is occasionally uneven jumping between sections. I consider this a
feature of the book, and I hope that the reader will agree and feel free to skip around
(see How to use this book below). I tried to “reboot” at the beginning of each part
and again at the beginning of each chapter, to refresh our motivation. For researchers
i
ii
working with quaternion algebras, I have tried to collect results otherwise scattered
in the literature and to provide some clarifications, corrections, and complete proofs
in the hopes that this text will provide a convenient reference. In order to provide
these features, to the extent possible I have opted for an organizational pattern that is
“horizontal” rather than “vertical”: the text has many chapters, each representing a
different slice of the theory.
I tried to compactify the text as much as possible, without sacrificing completeness.
There were a few occasions when I thought a topic could use further elaboration or has
evolved from the existing literature, but did not want to overburden the text; I collected
these in a supplementary text Quaternion algebras companion, available at the website
for the text at http://quatalg.org.
As usual, each chapter also contains a number of exercises at the end, ranging
from checking basic facts used in a proof to more difficult problems that stretch the
reader. Exercises that are used in the text are marked by I. For a subset of exercises
(including many of those marked with I), there are hints, comments, or a complete
solution available online.
1. For an introductory survey course on quaternion algebras, read just the introduc-
tory sections in each chapter, those labelled with ⊲, and supplement with sections
from the text when interested. These introductions usually contain motivation
and a summary of the results in the rest of the chapter, and I often restrict the
level of generality or make simplifying hypotheses so that the main ideas are
made plain. The reader who wants to quickly and gently grab hold of the basic
concepts may digest the book in this way. The instructor may desire to fill in
some further statements or proofs to make for a one semester course: chapters
1, 2, 11, 25, and 35 could be fruitfully read in their entirety.
2. For a mini-course in noncommutative algebra with emphasis on quaternion
algebras, read just part I. Such an early graduate course would have minimal
prerequisites and in a semester could be executed at a considered pace; it would
provide the foundation for further study in many possible directions.
3. For quaternion algebras and algebraic number theory, read parts I and II. This
course would be a nice second-semester addition following a standard first-
semester course in algebraic number theory, suitable for graduate students in
algebra and number theory who are motivated to study quaternion algebras as
“noncommutative quadratic fields”. For a lighter course, chapters 6, 20, and
21 could be skipped, and the instructor may opt to cover only the introductory
section of a chapter for reasons of time and interest. To reinforce concepts from
algebraic number theory, special emphasis could be placed on chapter 13 (where
iii
local division algebras are treated like local fields) and 18 (where maximal orders
are treated like noncommutative Dedekind domains).
There are also more specialized options, beginning with the introductory sections
in part I and continuing as follows.
4. For quaternion algebras and analytic number theory, continue with the intro-
ductory sections in part II (just chapters 9–17), and then cover part III (at least
through Chapter 29). This course could follow a first-semester course in an-
alytic number theory, enriching students’ understanding of zeta functions and
𝐿-functions (roughly speaking, beginning the move from GL1 to GL2 ). The
additional prerequisite of real analysis (measure theory) is recommended. Op-
tionally, this course could break after chapter 26 to avoid adeles, and perhaps
resume in an advanced topics course with the remaining chapters.
5. For quaternionic applications to geometry (specifically, hyperbolic geometry
and low-dimensional topology), continue with the introductory sections in part
II (through chapter 14), and then cover part IV (optionally skipping Chapter 32).
6. For an advanced course on quaternion algebras and arithmetic geometry, con-
tinue with part II, the introductory sections in part IV, and part V. Chapter 41
could be read immediately after part II. This path is probably most appropriate
for an advanced course for students with some familiarity with modular forms
and some hyperbolic geometry, and chapter 42 is probably only meaningful for
students with a background in elliptic curves (though the relevant concepts are
reviewed at the start).
7. Finally, for the reader who is studying quaternion algebras with an eye to appli-
cations with supersingular elliptic curves, the reader may follow chapters 2–4,
9–10, 13–14, 16 –17, 23, then the main event in chapter 42. For further reading
on quaternion orders and ternary quadratic forms, I suggest chapters 5, 22, and
24.
Sections of the text that are more advanced (requiring more background) or those
may be omitted are labeled with ∗. The final chapter (Chapter 43) is necessarily
more advanced, and additional prerequisites in algebraic and arithmetic geometry are
indicated.
It is a unique feature of quaternion algebras that topics overlap and fold together like
this, and so I hope the reader will forgive the length of the book. The reader may find
the symbol definition list at the end to help in identifying unfamiliar notation. Finally,
to ease in location I have chosen to number all objects (theorem-like environments,
equations, and figures) consecutively.
Companion reading
Several general texts can serve as companion reading for this monograph:
• The lecture notes of Vignéras [Vig80a] have been an essential reference for the
arithmetic of quaternion algebras since their publication.
iv
• The seminal text by Reiner [Rei2003] on maximal orders treats many introduc-
tory topics that overlap this text.
• The book of Maclachlan–Reid [MR2003] gives an introduction to quaternion
algebras with application to the geometry of 3-manifolds.
• The book by Deuring [Deu68] (in German) develops the theory of algebras over
fields, culminating in the treatment of zeta functions of division algebras over
the rationals, and may be of historical interest as well.
• Finally, Pizer [Piz76a] and Alsina–Bayer [AB2004] present arithmetic and al-
gorithmic aspects of quaternion algebras over Q.
Acknowledgements
This book began as notes from a course offered at McGill University in the Winter 2010
semester, entitled Computational aspects of quaternion algebras and Shimura curves. I
would like to thank the members of my Math 727 class for their invaluable discussions
and corrections: Dylan Attwell-Duval, Xander Faber, Luis Finotti, Andrew Fiori,
Cameron Franc, Adam Logan, Marc Masdeu, Jungbae Nam, Aurel Page, Jim Parks,
Victoria de Quehen, Rishikesh, Shahab Shahabi, and Luiz Takei. This course was part
of the special thematic semester Number Theory as Applied and Experimental Science
organized by Henri Darmon, Eyal Goren, Andrew Granville, and Mike Rubinstein
at the Centre de Recherche Mathématiques (CRM) in Montréal, Québec, and the
extended visit was made possible by the generosity of Dominico Grasso, dean of the
College of Engineering and Mathematical Sciences, and Jim Burgmeier, chair of the
Department of Mathematics and Statistics, at the University of Vermont. I am very
grateful for their support.
After a long hiatus, the writing continued while the author was on sabbatical
at the University of California, Berkeley. Several students attended these lectures
and gave helpful feedback: Watson Ladd, Andrew Niles, Shelly Manber, Eugenia
Ros, u, Emmanuel Tsukerman, Victoria Wood, and Alex Youcis. Thanks to Ken Ribet
for sponsoring my visit and helping to organize these lectures. My sabbatical from
Dartmouth College for the Fall 2013 and Winter 2014 quarters was made possible by
the efforts of Associate Dean David Kotz, and I thank him for his support. During this
sabbatical, further progress on the text was made in preparation for a minicourse on
Brandt modules as part of Minicourses on Algebraic and Explicit Methods in Number
Theory, organized by Cécile Armana and Christophe Delaunay at the Laboratoire de
Mathématiques de Besançon in Salins-les-Bains, France.
Thanks also go to the patient participants in my Math 125 Quaternion algebras
class at Dartmouth in Spring 2014 (Daryl Deford, Tim Dwyer, Zeb Engberg, Michael
Firrisa, Jeff Hein, Nathan McNew, Jacob Richey, Tom Shemanske, Scott Smedinghoff,
and David Webb) and my Math 125 Geometry of discrete groups in Summer 2015
(Angelica Babei, Ben Breen, Sara Chari, Melanie Dennis, James Drain, Tim Dwyer,
Jon Epstein, David Freund, Sam Kater, Michael Musty, Nicolas Petit, Sam Schiavone,
Scott Smedinghoff, and Everett Sullivan). I am also grateful for the feedback given
during my course at the Institute for Computational and Experimental Research in
Mathematics (ICERM) at Brown University in Fall 2015 as part of the Semester
v
I can only hope that this book will receive better reviews!
On a more personal note, I have benefited from the insight of incredible teachers
and collaborators throughout my mathematical life, and their positive impact can
hopefully be seen in this manuscript: among many, I would like to give special thanks
to Pete L. Clark, Bas Edixhoven, Benedict Gross, Hendrik Lenstra, and Bjorn Poonen
for their inspiration and guidance over the years.
Many thanks go to the others who offered helpful comments, corrections, answered
my questions, and provided feedback: Konstantin Ardakov, Sarah Arpin, Asher Auel,
Andreas Bächle, John Boxall, Ben Breen, Juliusz Brzezinski, Billy Chan, Pierre
Clare, France Dacar, Lassina Dembélé, Juanita Duque Rosero, Kirsten Eisentraeger,
John Enns, Zak Evans, Steven Galbraith, Darij Grinberg, Joseph Gunther, Jonathan
Hanke, Thorsten Herrig, Aleksander Horawa, Will Jagy, Ariyan Javanpeykar, Bo-
Gwang Jeon, Chan-Ho Kim, Robert Kucharczyk, Joel Laity, Dion Leijnse, Patrick
Lenning, Chao Li, Wen-Wei Li, Benjamin Linowitz, Travis Morrison, Ariel Pacetti,
Lorenz Panny, Jennifer Park, Carl Pomerance, Jose Pujol, Matthieu Rambaud, John
Riccardi, Sam Roven, Nithi Rungtanapirom, Ciaran Schembri, sibilant, Jana Sotakova,
Tom Shemanske, Daniel Smertnig, Jim Stankewicz, Kate Stange, Drew Sutherland,
Nicole Sutherland, Jacob Swenberg, Anne van Weerden, Annalena Wernz, Mckenzie
West, Dana Williams, Jiangwei Xue, and Vitthal Yellambalse. I am indebted to C.
Emily I. Redelmeier for fabulous TeXnical assistance with figures. Special thanks to the
vi
anonymous referees for their constructive feedback and to Sara Chari, Peter Landweber,
Joe Quinn, and Grant Molnar for a careful reading. The errors and omissions that
remain are, of course, my own: please contact me at [email protected] if you find
mistakes or have other suggestions (e.g., additional references, alternate proofs, further
exercises or applications). Collected errata and a corrected version will be posted on
the webpage for the text (http://quatalg.org).
I am profoundly grateful to those who lent encouragement at various times during
the writing of this book, when the going was tough: Srinath Baba, Chantal David,
Matthew Greenberg and Kristina Loeschner, Laurie Johnson at the National Center
for Faculty Development and Diversity (NCFDD), David Michaels, Tom Shemanske,
and my mother Connie Voight. Thank you all! I am also grateful to Marc Strauss and
Loretta Bartolini, editors at Springer, for their professional guidance. Finally, I would
like to offer my deepest gratitude to my partner Brian Kennedy: this book would not
have been possible without his patience, equanimity, and wit. As they say, happiness
is a journey, not a destination; the same I think is true for happy mathematics, and I
am blessed to have had both personal and professional companionship on this one.
Contents
Contents vii
1 Introduction 1
1.1 Hamilton’s quaternions . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Algebra after the quaternions . . . . . . . . . . . . . . . . . . . . . 6
1.3 Quadratic forms and arithmetic . . . . . . . . . . . . . . . . . . . . 10
1.4 Modular forms and geometry . . . . . . . . . . . . . . . . . . . . . 11
1.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
I Algebra 19
2 Beginnings 21
2.1 ⊲ Conventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 ⊲ Quaternion algebras . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 ⊲ Matrix representations . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 ⊲ Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3 Involutions 35
3.1 ⊲ Conjugation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Involutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3 Reduced trace and reduced norm . . . . . . . . . . . . . . . . . . . 38
3.4 Uniqueness and degree . . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 Quaternion algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4 Quadratic forms 47
4.1 ⊲ Reduced norm as quadratic form . . . . . . . . . . . . . . . . . . 47
4.2 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3 Discriminants, nondegeneracy . . . . . . . . . . . . . . . . . . . . . 52
4.4 Nondegenerate standard involutions . . . . . . . . . . . . . . . . . . 54
4.5 Special orthogonal groups . . . . . . . . . . . . . . . . . . . . . . . 55
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
vii
viii CONTENTS
6 Characteristic 2 83
6.1 Separability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2 Quaternion algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.3 ∗ Quadratic forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.4 ∗ Characterizing quaternion algebras . . . . . . . . . . . . . . . . . 87
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7 Simple algebras 93
7.1 ⊲ Motivation and summary . . . . . . . . . . . . . . . . . . . . . . 93
7.2 Simple modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.3 Wedderburn–Artin . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7.4 Jacobson radical . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.5 Central simple algebras . . . . . . . . . . . . . . . . . . . . . . . . 103
7.6 Quaternion algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.7 The Skolem–Noether theorem . . . . . . . . . . . . . . . . . . . . . 106
7.8 Reduced trace and norm, universality . . . . . . . . . . . . . . . . . 110
7.9 Separable algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
II Arithmetic 133
10 Orders 151
10.1 ⊲ Lattices with multiplication . . . . . . . . . . . . . . . . . . . . . 151
10.2 Orders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
10.3 Integrality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
10.4 Maximal orders . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
10.5 Orders in a matrix ring . . . . . . . . . . . . . . . . . . . . . . . . 156
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
15 Discriminants 233
x CONTENTS
31 Selectivity 557
31.1 Selective orders . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
31.2 Selectivity conditions . . . . . . . . . . . . . . . . . . . . . . . . . 560
31.3 ∗ Selectivity setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
31.4 ∗ Outer selectivity inequalities . . . . . . . . . . . . . . . . . . . . 564
31.5 ∗ Middle selectivity equality . . . . . . . . . . . . . . . . . . . . . . 565
31.6 ∗ Optimal selectivity conclusion . . . . . . . . . . . . . . . . . . . . 567
31.7 ∗ Selectivity, without optimality . . . . . . . . . . . . . . . . . . . . 568
31.8 ∗ Isospectral, nonisometric manifolds . . . . . . . . . . . . . . . . . 570
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
Bibliography 823
Index 849
Chapter 1
Introduction
We begin following the historical arc of quaternion algebras and tracing their impact
on the development of mathematics. Our account is selective: for further overview,
see Lam [Lam2003] and Lewis [Lew2006a].
1
2 CHAPTER 1. INTRODUCTION
For at least ten years (on and off), Hamilton had been attempting to model (real)
three-dimensional space with a structure like the complex numbers, whose addition
and multiplication occur in two-dimensional space. Just like the complex numbers had
a “real” and “imaginary” part, so too did Hamilton hope to find an algebraic system
whose elements had a “real” and two-dimensional “imaginary” part. In the early part
of the month of October 1843, his sons Archibald Henry and William Edwin Hamilton,
while still very young, would ask their father at breakfast [Ham67, p. xv]: “Well, papa,
can you multiply triplets?” To which Hamilton would reply, “with a sad shake of the
head, ‘No, I can only add and subtract them’” [Ham67, p. xv]. For a history of the
“multiplying triplets” problem—the nonexistence of division algebra over the reals of
dimension 3—see May [May66, p. 290].
𝑖 2 = 𝑗 2 = 𝑘 2 = 𝑖 𝑗 𝑘 = −1
For more on the history of Hamilton’s discovery, see the extensive and detailed
accounts of Dickson [Dic19] and Van der Waerden [vdW76]. There are also three
main biographies written about the life of William Rowan Hamilton, a man sometimes
referred to as “Ireland’s greatest mathematician”: by Graves [Grav1882, Grav1885,
Grav1889] in three volumes, Hankins [Hankin80], and O’Donnell [O’Do83]. Numer-
ous other shorter biographies have been written [DM89, Lanc67, ÓCa2000]. (Certain
4 CHAPTER 1. INTRODUCTION
aspects of Hamilton’s private life deserve a more positive portrayal, however: see Van
Weerden–Wepster [WW2018].)
• Thia formula (A),.,.. accordingly made the bcui• of that Calculns In the lint
communication on the subject, by tbe preeent writer, to the Royal lriah Aeademy in
1848 ; and the !etten, i, j, I, coatinued to be, for some time, the ore~, pHtlliGr .,..
hl• of the c:aleulus In question. But it was gradually found to be nsefa1 to Incor-
porate with these a few other aotatiolu (auch as K and U, &c.), for rep~ntlag
OportJtitnU 011 QuleMiiOfl•. It was aleo thought to be Instructive to eetablilh the
priAeiplu or that Calculus, on a more !lft't11Mcal (or leu exeluaively ·~
f'*"datiora tban at first ; which was accordingly af'lerwards done, in the volume en-
titled: L«nru 011 QruJtmaiou (Dublin, 1858); and I• again attempted in the pre·
eent work, although with many dilfereneea In the adopted plara of expoeltion 1 and in
the applictJtioro• brought forward, or suppreaeed.
(𝑎 1 + 𝑎 2𝑖 + 𝑎 3 𝑗 + 𝑎 4 𝑘) (𝑏 1 + 𝑏 2𝑖 + 𝑏 3 𝑗 + 𝑏 4 𝑘) = 𝑐 1 + 𝑐 2 𝑖 + 𝑐 3 𝑗 + 𝑐 4 𝑘
𝑣𝑤 = −𝑣 · 𝑤 + 𝑣 × 𝑤 (1.1.7)
analysis that would apply in arbitrary dimension; on the relationship between these
works, Gibbs wrote after learning of the work of Grassmann: “I saw that the methods
wh[ich] I was using, while nearly those of Hamilton, were almost exactly those of
Grassmann” [Whe62, p. 108]. For more on the history of quaternionic and vector
calculus, see Crowe [Cro64] and Simons [Sim2010].
The rivalry between physical notations flared into a war in the latter part of the 19th
century between the ‘quaternionists’ and the ‘vectorists’, and for some the preference
of one system versus the other became an almost partisan split. On the side of
quaternions, James Clerk Maxwell (1831–1879), who derived the equations which
describe electromagnetic fields, wrote [Max1869, p. 226]:
The invention of the calculus of quaternions is a step towards the knowl-
edge of quantities related to space which can only be compared, for its
importance, with the invention of triple coordinates by Descartes. The
ideas of this calculus, as distinguished from its operations and symbols,
are fitted to be of the greatest use in all parts of science.
And Peter Tait (1831–1901), Hamilton’s “chief disciple” [Hankin80, p. 316], wrote in
1890 [Tai1890] decrying notation and attacking Willard Gibbs (1839–1903):
It is disappointing to find how little progress has recently been made
with the development of Quaternions. One cause, which has been spe-
cially active in France, is that workers at the subject have been more
intent on modifying the notation, or the mode of presentation of the
fundamental principles, than on extending the applications of the Calcu-
lus. . . . Even Prof. Willard Gibbs must be ranked as one the retarders of
quaternions progress, in virtue of his pamphlet on Vector Analysis, a sort
of hermaphrodite monster, compounded of the notation of Hamilton and
Grassman.
Game on! On the vectorist side, Lord Kelvin (a.k.a. William Thomson, who formulated
the laws of thermodynamics), said in an 1892 letter to R. B. Hayward about his textbook
in algebra (quoted in Thompson [Tho10, p. 1070]):
Quaternions came from Hamilton after his really good work had been
done; and, though beautifully ingenious, have been an unmixed evil to
those who have touched them in any way, including Clerk Maxwell.
(There is also a rompous fictionalized account by Pynchon in his tome Against the Day
[Pyn2006].) Ultimately, the superiority and generality of vector notation carried the
day, and only certain useful fragments of Hamilton’s quaternionic notation—e.g., the
“right-hand rule” 𝑖 × 𝑗 = 𝑘 in multivariable calculus—remain in modern usage.
So with the introduction of the quaternions, the floodgates of algebraic possibility had
been opened. See Happel [Hap80] for an overview of the early development of algebra
following Hamilton’s quaternions, as well as the more general history given by Van
der Waerden [vdW85, Chapters 10–11].
The day after his discovery, Hamilton sent a letter [Ham1844] describing the
quaternions to his friend John T. Graves (1806–1870). Graves replied on October 26,
1843, with his compliments, but added:
There is still something in the system which gravels me. I have not yet
any clear views as to the extent to which we are at liberty arbitrarily to
create imaginaries, and to endow them with supernatural properties. . . .
If with your alchemy you can make three pounds of gold, why should you
stop there?
Following through on this invitation, on December 26, 1843, Graves wrote to Hamilton
that he had successfully generalized the quaternions to the “octaves”, now called
octonions O, an algebra in eight dimensions, with which he was able to prove that the
product of two sums of eight perfect squares is another sum of eight perfect squares,
a formula generalizing (1.1.6). In fact, Hamilton first invented the term associative in
1844, around the time of his correspondence with Graves. Unfortunately for Graves, the
octonions were discovered independently and published in 1845 by Cayley [Cay1845b],
who often is credited for their discovery. (Even worse, the eight squares identity was
also previously discovered by C. F. Degen.) For a more complete account of this
story and the relationships between quaternions and octonions, see the survey article
by Baez [Bae2002], the article by Van der Blij [vdB60], and the delightful book by
Conway–Smith [CSm2003].
Cayley also studied quaternions themselves [Cay1845a] and was able to reinterpret
them as arising from a doubling process, also called the Cayley–Dickson construction,
which starting from R produces C then H then O, taking the ordered, commutative,
associative algebra R and progressively deleting one adjective at a time. So algebras
were first studied over the real and complex numbers and were accordingly called
hypercomplex numbers in the late 19th and early 20th century. And this theory
8 CHAPTER 1. INTRODUCTION
flourished. Hamilton himself considered the algebra over C defined by his famous
equations (1.1.1), calling them biquaternions. In 1878, Ferdinand Frobenius (1849–
1917) proved that the only finite-dimensional associative real division algebras are R,
C, and H [Fro1878]. This result was also proven independently by C.S. Peirce, the
son of Benjamin Peirce, below. Adolf Hurwitz (1859–1919) later showed that the
only normed finite-dimensional not-necessarily-associative real division algebras are
R, C, H, and O. (The same statement is true without the condition that the algebra
be normed, but currently the proofs use topology, not algebra! Bott–Milnor [BM58]
and Kervaire [Ker58] proved that the (𝑛 − 1)-dimensional sphere {𝑥 ∈ R𝑛 : k𝑥k 2 = 1}
has trivial tangent bundle if and only if there is an 𝑛-dimensional not-necessarily-
associative real division algebra if and only if 𝑛 = 1, 2, 4, 8. The solution to the
Hopf invariant one problem by Adams also implies this result; an elegant and concise
proof using 𝐾-theory, Adams operations, and elementary number theory was given
by Adams–Atiyah [AA66]. See Hirzebruch [Hir91] or Ranicki [Ran2011] for a more
complete account.)
In another attempt to seek a generalization of the quaternions to higher dimension,
William Clifford (1845–1879) developed a way to build algebras from quadratic forms
in 1876 [Cli1878]. Clifford constructed what we now call a Clifford algebra 𝐶 (𝑉)
associated to 𝑉 = R𝑛 (with the standard Euclidean norm); it is an algebra of dimension
2𝑛 containing 𝑉 with multiplication induced from the relation 𝑥 2 = −k𝑥k 2 for all 𝑥 ∈ 𝑉.
We have 𝐶 (R1 ) = C and 𝐶 (R2 ) = H, so the Hamilton quaternions arise as a Clifford
algebra—but 𝐶 (R3 ) is not the octonions. The theory of Clifford algebras tightly
connects the theory of quadratic forms and the theory of normed division algebras
and its impact extends in many mathematical directions. For more on the history of
Clifford algebras, see Diek–Kantowski [DK95].
A further physically motivated generalization was pursued by Alexander Macfar-
lane (1851–1913): he developed a theory of what he called hyperbolic quaternions
[Macf00] (a revised version of an earlier, nonassociative attempt [Macf1891]), with
the multiplication laws
𝑖 2 = 𝑗 2 = 𝑘 2 = 1,
√ √ √ (1.2.1)
𝑖 𝑗 = −1𝑘 = − 𝑗𝑖, 𝑗 𝑘 = −1𝑖 = −𝑘 𝑗, 𝑘𝑖 = −1 𝑗 = −𝑖𝑘.
√
Thought of as an algebra over C = R( −1), Macfarlane’s hyperbolic quaternions
are isomorphic to Hamilton’s biquaternions (and therefore isomorphic to M2 (C)).
Moreover, the restriction of the norm to the real span of the basis 1, 𝑖, 𝑗, 𝑘 in Mac-
farlane’s algebra is a quadratic form of signature (1, 3): this gives a quaternionic
version of space-time, something also known as Minkowski space (but with Macfar-
lane’s construction predating that of Minkowski). For more on the history and further
connections, see Crowe [Cro64].
Around this time, other types of algebras over the real numbers were also being
investigated, the most significant of which were Lie algebras. In the seminal work
of Sophus Lie (1842–1899), group actions on manifolds were understood by looking
at this action infinitesimally; one thereby obtains a Lie algebra of vector fields that
determines the local group action. The simplest nontrivial example of a Lie algebra
is the cross product of two vectors, related to quaternion multiplication in (1.1.7): it
1.2. ALGEBRA AFTER THE QUATERNIONS 9
𝑖 × ( 𝑗 × 𝑘) + 𝑘 × (𝑖 × 𝑗) + 𝑗 × (𝑘 × 𝑖) = 0. (1.2.2)
The Lie algebra “linearizes” the group action and is therefore more accessible. Wilhelm
Killing (1847–1923) initiated the study of the classification of Lie algebras in a series
of papers [Kil1888], and this work was completed by Élie Cartan (1869–1951). We
refer to Hawkins [Haw2000] for a description of this rich series of developments.
In this way, the study of division algebras gradually evolved, independent of
physical interpretations. Benjamin Peirce (1809–1880) in 1870 developed what he
called linear associative algebras [Pei1882]; he provided a decomposition of an algebra
relative to an idempotent (his terminology). The first definition of an algebra over an
arbitrary field seems to have been given by Leonard E. Dickson (1874–1954) [Dic03]:
at first he still called the resulting object a system of complex numbers and only later
adopted the name linear algebra.
The notion of a simple algebra had been discovered by Cartan, and Theodor Molien
(1861–1941) had earlier shown in his terminology that every simple algebra over the
complex numbers is a matrix algebra [Mol1893]. But it was Joseph Henry Maclagan
Wedderburn (1882–1948) who was the first to find meaning in the structure of simple
algebras over an arbitrary field, in many ways leading the way forward. The jewel
of his 1908 paper [Wed08] is still foundational in the structure theory of algebras: a
simple algebra (finite-dimensional over a field) is isomorphic to a matrix ring over a
division ring. Wedderburn also proved that a finite division ring is a field, a result that
like his structure theorem has inspired much mathematics. For more on the legacy of
Wedderburn, see Artin [Art50].
In the early 1900s, Dickson was the first to consider quaternion algebras over
a general field [Dic12, (8), p. 65]. He began by considering more generally those
algebras in which every element satisfies a quadratic equation [Dic12], exhibited a
diagonalized basis for such an algebra, and considered when such an algebra can be
a division algebra. This led him to multiplication laws for what he later called a
generalized quaternion algebra [Dic14, Dic23], with multiplication laws
𝑖 2 = 𝑎, 𝑗 2 = 𝑏, 𝑘 2 = −𝑎𝑏,
(1.2.3)
𝑖 𝑗 = 𝑘 = − 𝑗𝑖, 𝑖𝑘 = 𝑎 𝑗 = −𝑘𝑖, 𝑘 𝑗 = 𝑏𝑖 = − 𝑗 𝑘
with 𝑎, 𝑏 nonzero elements in the base field. (To keep track of these, it is helpful
to write 𝑖, 𝑗, 𝑘 around a circle clockwise.) Today, we no longer employ the adjective
“generalized”—over fields other than R, there is no reason to privilege the Hamiltonian
quaternions—and we can reinterpret this vein of Dickson’s work as showing that every
4-dimensional central simple algebra is a quaternion algebra (a statement that holds
even over a field 𝐹 with char 𝐹 = 2). See Fenster [Fen98] for a summary of Dickson’s
work in algebra, and Lewis [Lew2006b] for a broad survey of the role of involutions
and anti-automorphisms in the classification of algebras.
10 CHAPTER 1. INTRODUCTION
lectures Emmy Noether (1882–1935) considered the even more general crossed product
algebras. Not very long after, in a volume dedicated to Hensel’s seventieth birthday,
Richard Brauer (1901–1977), Hasse, and Noether proved a fundamental theorem for
the structure theory of algebras over number fields [BHN31]: every central division
algebra over a number field is a cyclic algebra. This crucial statement had profound
implications for class field theory, the classification of abelian extensions of a number
field, with a central role played by the Brauer group of a number field, a group
encoding its division algebras. For a detailed history and discussion of these lines,
see Fenster–Schwermer [FS2007], Roquette [Roq2006], and the history of class field
theory summarized by Hasse himself [Hass67].
At the same time, Abraham Adrian Albert (1905–1972), a doctoral student of Dick-
son, was working on the structure of division algebras and algebras with involution,
and he had written a full book on the subject [Alb39] collecting his work in the area,
published in 1939. Albert had examined the tensor product of two quaternion algebras,
called a biquaternion algebra (not to be confused with Hamilton’s biquaternions), and
he characterized when such an algebra was a division algebra in terms of a senary (six
variable) quadratic form. Albert’s classification of algebras with involution was moti-
vated by understanding possible endomorphism algebras of abelian varieties, viewed
as multiplier rings of Riemann matrices and equipped with the Rosati involution: a
consequence of this classification is that quaternion algebras are the only noncommu-
tative endomorphism algebras of simple abelian varieties. He also proved that a central
simple algebra admits an involution if and only if the algebra is isomorphic to its oppo-
site algebra (equivalently, it has order at most 2 in the Brauer group). For a biography
of Albert and a survey of his work, see Jacobson [Jacn74]. Roquette argues convinc-
ingly [Roq2006, §8] that because of Albert’s contributions to its proof (for example,
his work with Hasse [AH32]), we should refer to the Albert–Brauer–Hasse–Noether
theorem in the previous paragraph.
∞
∑︁
𝜃 (𝜏) := exp(2𝜋𝑖𝑛2 𝜏) = 1 + 2𝑞 + 2𝑞 4 + 2𝑞 9 + . . . (1.4.1)
𝑛=−∞
where 𝜏 is a complex number with positive imaginary part and 𝑞 = exp(2𝜋𝑖𝜏). Jacobi
12 CHAPTER 1. INTRODUCTION
where 𝜎 ∗ (𝑛) := 4-𝑑 |𝑛 𝑑 is the sum of divisors of 𝑛 not divisible by 4. In this way,
Í
Jacobi gave an explicit formula for the number of ways of expressing a number as the
sum of four squares. For a bit of history and an elementary derivation in the style of
Gauss and Jacobi, see Ewell [Ewe82].
As a Fourier series, the Jacobi theta function 𝜃 (1.4.1) visibly satisfies 𝜃 (𝜏 + 1) =
𝜃 (𝜏). Moreover, owing to its symmetric description, Jacobi showed using Poisson
summation that 𝜃 also satisfies the transformation formula
√︁
𝜃 (−1/𝜏) = 𝜏/𝑖 𝜃 (𝜏). (1.4.3)
Felix Klein (1849–1925) saw geometry in formulas like (1.4.3). In his Erlangen
Program (1872), he recast 19th century geometry in terms of the underlying group
of symmetries, unifying Euclidean and non-Euclidean formulations. Turning then to
hyperbolic geometry, he studied the modular group SL2 (Z) acting by linear fractional
transformations on the upper half-plane, and interpreted transformation formulas for
elliptic functions: in particular, Klein defined his absolute invariant 𝐽 (𝜏) [Kle1878],
a function invariant under the modular group. Together with his student Robert Fricke
(1861–1930), this led to four volumes [FK1890-2, FK1897, FK12] on elliptic modular
functions and automorphic functions, combining brilliant advances in group theory,
number theory, geometry, and invariant theory (Figure 1.4.4).
At the same time, Henri Poincaré (1854–1912) brought in the theory of linear
differential equations—and a different, group-theoretic approach. In correspondence
with Fuchs in 1880 on hypergeometric differential equations, he writes about the
beginnings of his discovery of a new class of analytic functions [Gray2000, p.177]:
They present the greatest analogy with elliptic functions, and can be
represented as the quotient of two infinite series in infinitely many ways.
Amongst those series are those which are entire series playing the role of
Theta functions. These converge in a certain circle and do not exist outside
it, as thus does the Fuchsian function itself. Besides these functions there
are others which play the same role as the zeta functions in the theory
of elliptic functions, and by means of which I solve linear differential
equations of arbitrary orders with rational coefficients whenever there are
only two finite singular points and the roots of the three determinantal
equations are commensurable.
As he reminisced later in his Science et Méthode [Poi1908, p. 53]:
I then undertook to study some arithmetical questions without any great
result appearing and without expecting that this could have the least con-
nection with my previous researches. Disgusted with my lack of success,
I went to spend some days at the sea-side and thought of quite different
things. One day, walking along the cliff, the idea came to me, always with
the same characteristics of brevity, suddenness, and immediate certainty,
that the arithmetical transformations of ternary indefinite quadratic forms
were identical with those of non-Euclidean geometry.
In other words, like Klein, Poincaré launched a program to study complex analytic
functions defined on the unit disc that are invariant with respect to a discrete group
of matrix transformations that preserve a rational indefinite ternary quadratic form.
Today, such groups are called arithmetic Fuchsian groups, and we study them as unit
groups of quaternion algebras. To read more on the history of differential equations
in the time of Riemann and Poincaré, see the history by Gray [Gray2000], as well as
Gray’s scientific biography of Poincaré [Gray2013].
In the context of these profound analytic discoveries, Erich Hecke (1887–1947)
began his study of modular forms. He studied the Dedekind zeta function, a gener-
alization of Riemann’s zeta function to number fields, and established its functional
equation using theta functions. In the study of similarly defined analytic functions
arising from modular forms, he was led to define the “averaging” operators acting on
spaces of modular forms that now bear his name. In this way, he could interpret the
Fourier coefficients 𝑎(𝑛) of a Hecke eigenform (normalized, weight 2) as eigenvalues
of his operators: he proved that they satisfy a relation of the form
∑︁
𝑎(𝑚)𝑎(𝑛) = 𝑎(𝑚𝑛/𝑑 2 )𝑑 (1.4.5)
𝑑 |gcd(𝑚,𝑛)
and consequently a two-term recursion relation. He thereby showed that the Dirichlet
𝐿-series of an eigenform, defined via Mellin transform, has an Euler product, analytic
continuation, and functional equation.
14 CHAPTER 1. INTRODUCTION
Hecke went further, and connected the analytic theory of modular forms and his
operators to the arithmetic theory of quadratic forms. In 1935–1936, he found that
for certain systems of quaternary quadratic forms, the number of representations of
integers by the system satisfied the recursion (1.4.5), in analogy with binary quadratic
forms. He published a conjecture on this subject in 1940 [Hec40, Satz 53, p. 100]:
that the weighted representation numbers satisfy the Hecke recursion, connecting
coefficients to operators on theta series, and further that the columns in a composition
table always result in linearly independent theta series. He verified the conjecture up
to prime level 𝑞 < 37, but was not able to prove this recursion using his methods of
complex analysis (see his letter [Bra41, Footnote 1]).
The arithmetic part of these conjectures was investigated by Heinrich Brandt
(1886–1954) in the quaternionic context—and so the weave of our narrative is further
tightly sewn. Preceding Hecke’s work, and inspired by Gauss composition of binary
quadratic forms as the product of classes of ideals in a quadratic field, Brandt had
earlier considered a generalization to quaternary quadratic forms and the product of
classes of ideals in a quaternion algebra [Bra28]: he was only able to define a partially
defined product, and so he coined the term groupoid for such a structure [Bra40]. He
then considered the combinatorial problem of counting the ways of factoring an ideal
into prime ideals, according to their classes. In this way, he recorded these counts in
a matrix 𝑇 (𝑛) for each positive integer 𝑛, and he proved strikingly (sketched in 1941
[Bra41], dated 1939, and proved completely in 1943 [Bra43]) that the matrices 𝑇 (𝑛)
satisfy Hecke’s recursion (1.4.5). To read more on the life and work of Brandt, see
Hoehnke–Knus [HK2004]. Today we call the matrices 𝑇 (𝑛) Brandt matrices, and
for certain purposes, they are still the most convenient way to get ahold of spaces of
modular forms.
Martin Eichler (1912–1992) wrote his thesis [Eic36] under the supervision of
Brandt on quaternion orders over the integers, in particular studying the orders that
now bear his name. Later he continued the grand synthesis of modular forms, quadratic
forms, and quaternion algebras, viewing in generality the orthogonal group of a
quadratic form as acting via automorphic transformations [Eic53]. In this vein, he
formulated his basis problem (arising from the conjecture of Hecke) which sought
to understand explicitly the span of quaternionic theta series among classical modu-
lar forms, giving a correspondence between systems of Hecke eigenvalues appearing
in the quaternionic and classical context. He answered the basis problem in affir-
mative for the case of prime level in 1955 [Eic56a] and then for squarefree level
[Eic56b, Eic58, Eic73]. For more on Eichler’s basis problem and its history, see
Hijikata–Pizer–Shemanske [HPS89a].
Having come to recent history, our account now becomes much more abbreviated:
we provide further commentary in situ in remarks in the rest of this text, and we
conclude with just a few highlights. In the 1950s and 1960s, there was subtantial
work done in understanding zeta functions of certain varieties arising from quaternion
algebras over totally real number fields. For example, Eichler’s correspondence was
generalized to totally real fields by Shimizu [Shz65]. Shimura embarked on a deep and
systematic study of arithmetic groups obtained from indefinite quaternion algebras over
totally real fields, including both the arithmetic Fuchsian groups of Poincaré, Fricke,
and Klein, and the generalization of the modular group to totally real fields studied
1.5. CONCLUSION 15
1.5 Conclusion
We have seen how quaternion algebras have threaded mathematical history through to
the present day, weaving together advances in algebra, quadratic forms, number theory,
geometry, and modular forms. And although our history ends here, the story does not!
Quaternion algebras continue to arise in unexpected ways. In the arithmetic setting,
quaternion orders arise as endomorphism rings of supersingular elliptic curves and
have been used in proposed post-quantum cryptosystems and digital signature schemes
(see for example the overview by Galbraith–Vercauteren [GV2018]). In the field of
quantum computation, Parzanchevski–Sarnak [PS2018] have proposed Super-Golden-
Gates built from certain special quaternion algebras and their arithmetic groups that
would give efficient 1-qubit quantum gates. In coding theory, lattices in quaternion
algebras (and more generally central simple algebras over number fields) yield space-
time codes that achieve high spectral efficiency on wireless channels with two transmit
antennas, currently part of certain IEEE standards [BO2013].
Quaternions have also seen a revival in computer graphics, modeling, and anima-
tion [HFK94, Sho85]. Indeed, a rotation in R3 about an axis through the origin can be
represented by a 3 × 3 orthogonal matrix with determinant 1, conveniently encoded in
Euler angles. However, the matrix representation is redundant, as there are only three
16 CHAPTER 1. INTRODUCTION
degrees of freedom in such a rotation. Moreover, to compose two rotations requires the
product of the two corresponding matrices, which requires 27 multiplications and 18
additions in R. Quaternions, on the other hand, represent this rotation with a 4-tuple,
and multiplication of two quaternions takes only 16 multiplications and 12 additions
in R (if done naively). In computer games, quaternion interpolation provides a way
to smoothly interpolate between orientations in space—something crucial for fighting
Nazi zombies. Quaternions are also vital for attitude control of aircraft and spacecraft
[Hans2006]: they avoid the ambiguity that can arise when two rotation axes align,
leading to a potentially disastrous loss of control called gimbal lock.
In quantum physics, quaternions yield elegant expressions for Lorentz transforma-
tions, the basis of the modern theory of relativity [Gir83]. Some physicists are now
hoping to find deeper understanding of these principles of quantum physics in terms
of quaternions. And so, although much of Hamilton’s quaternionic physics fell out of
favor long ago, we have come full circle in our elongated historical arc. The enduring
role of quaternion algebras as a catalyst for a vast range of mathematical research
promises rewards for many years to come.
Exercises
1. Hamilton sought a multiplication ∗ : R3 × R3 → R3 that preserves length:
k𝑣k 2 · k𝑤k 2 = k𝑣 ∗ 𝑤k 2
𝐷 := R + R𝑖 + R 𝑗 ' R3
Algebra
19
Chapter 2
Beginnings
2.1 ⊲ Conventions
Throughout this text (unless otherwise stated), we let 𝐹 be a (commutative) field with
algebraic closure 𝐹 al .
When 𝐺 is a group, and 𝐻 ⊆ 𝐺 is a subset, we write 𝐻 ≤ 𝐺 when 𝐻 is a subgroup
and 𝐻 E 𝐺 when 𝐻 is a normal subgroup; if 𝐺 is abelian (written multiplicatively),
we write 𝐺 𝑛 := {𝑔 𝑛 : 𝑔 ∈ 𝐺} ≤ 𝐺 for the subgroup of 𝑛th powers for 𝑛 ∈ Z>0 .
We suppose throughout that all rings are associative, not necessarily commutative,
with multiplicative identity 1, and that ring homomorphisms preserve 1. In particular, a
subring of a ring has the same 1. For a ring 𝐴, we write 𝐴× for the multiplicative group
of units of 𝐴. An algebra over the field 𝐹 is a ring 𝐵 equipped with a homomorphism
𝐹 → 𝐵 such that the image of 𝐹 lies in the center 𝑍 (𝐵) of 𝐵, defined by
if 𝑍 (𝐵) = 𝐹, we say 𝐵 is central (as an 𝐹-algebra). We write M𝑛 (𝐹) for the 𝐹-algebra
of 𝑛 × 𝑛-matrices with entries in 𝐹.
One may profitably think of an 𝐹-algebra as being an 𝐹-vector space that is also
compatibly a ring. If the 𝐹-algebra 𝐵 is not the zero ring, then its structure map 𝐹 → 𝐵
is necessarily injective (since 1 maps to 1) and we identify 𝐹 with its image; keeping
track of the structure map just litters notation. The dimension dim𝐹 𝐵 of an 𝐹-algebra
𝐵 is its dimension as an 𝐹-vector space.
A homomorphism of 𝐹-algebras is a ring homomorphism which restricts to
the identity on 𝐹. An 𝐹-algebra homomorphism is necessarily 𝐹-linear. An 𝐹-
algebra homomorphism 𝐵 → 𝐵 is called an endomorphism. By convention (and
as usual for functions), endomorphisms act on the left. An invertible 𝐹-algebra
homomorphism 𝐵 −→ ∼ 𝐵 0 is called an isomorphism, and an invertible endomorphism
is an automorphism.
21
22 CHAPTER 2. BEGINNINGS
−1, −1
Example 2.2.3. The R-algebra H := is the ring of quaternions over the real
R
numbers, discovered by Hamilton; we call H the ring of (real) Hamiltonians (also
known as Hamilton’s quaternions).
𝑖 2 = 𝑎, 𝑗 2 = 𝑏, and 𝑖 𝑗 = − 𝑗𝑖 (2.2.6)
with 𝑎, 𝑏 ∈ 𝐹 × .
In other words, once the relations (2.2.6) are satisfied for generators 𝑖, 𝑗, then
automatically 𝐵 has dimension 4 as an 𝐹-vector space, with 𝐹-basis 1, 𝑖, 𝑗, 𝑖 𝑗.
Proof. It is necessary and sufficient to prove that the elements 1, 𝑖, 𝑗, 𝑖 𝑗 are linearly
independent. Suppose that 𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 = 0 with 𝑡, 𝑥, 𝑦, 𝑧 ∈ 𝐹. Using the
relations given, we compute that
𝛼 − (𝑡 + 𝑥𝑖) − (𝑡 + 𝑦 𝑗) − (𝑡 + 𝑧𝑖 𝑗) = −2𝑡 = 0.
Remark 2.2.7. Invertibility of both 𝑎 and 𝑏 in 𝐹 is needed for Lemma 2.2.5: the
commutative algebra 𝐵 = 𝐹 [𝑖, 𝑗]/(𝑖, 𝑗) 2 is generated by the elements 𝑖, 𝑗 satisfying
𝑖 2 = 𝑗 2 = 𝑖 𝑗 = − 𝑗𝑖 = 0 but 𝐵 is not a quaternion algebra.
Remark 2.2.8. In light of Lemma 2.2.5, we will often drop the symbol 𝑘 = 𝑖 𝑗 and
reserve it for other use. (In particular, in later sections we will want 𝑘 to represent
other quaternion elements.) If we wish to use this abbreviation, we will assign 𝑘 := 𝑖 𝑗.
Proof. Injectivity follows by checking ker 𝜆 = {0} on matrix entries, and the homo-
morphism property can be verified directly, checking the multiplication table (Exercise
2.10).
Remark 2.3.3. Proposition 2.3.1 can be turned around to assert the existence of quater-
nion algebras: one can check that the set
√ √
√
𝑡 + 𝑥 √𝑎 𝑏(𝑦 + 𝑧√ 𝑎)
: 𝑡, 𝑥, 𝑦, 𝑧 ∈ 𝐹 ⊆ M2 (𝐹 ( 𝑎))
𝑦−𝑧 𝑎 𝑡−𝑥 𝑎
Proof. Specializing Proposition 2.3.1, we see the map is an injective 𝐹-algebra homo-
morphism, so since dim𝐹 𝐵 = dim𝐹 𝑀2 (𝐹) = 4, the map is also surjective.
The provenance of the map (2.3.2) is itself important, so we now pursue another
(more natural) proof of Proposition 2.3.1.
2.3.8. Let
𝐾 := 𝐹 [𝑖] = 𝐹 ⊕ 𝐹𝑖 ' 𝐹 [𝑥]/(𝑥 2 − 𝑎)
be the (commutative) 𝐹-algebra
√ generated by 𝑖. Suppose first that 𝐾 is a field (so
𝑎 ∉ 𝐹 ×2 ): then 𝐾 ' 𝐹 ( 𝑎) is a quadratic field extension of 𝐹. The algebra 𝐵 has the
structure of a right 𝐾-vector space of dimension 2, with basis 1, 𝑗: explicitly,
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 = (𝑡 + 𝑥𝑖) + 𝑗 (𝑦 − 𝑧𝑖) ∈ 𝐾 ⊕ 𝑗 𝐾
reading functions from right to left as usual. The map 𝜆 is injective (𝜆 is a faithful
representation) since 𝜆 𝛼 = 0 implies 𝜆 𝛼 (1) = 𝛼 = 0.
In the basis 1, 𝑗 we have End𝐾 (𝐵) ' M2 (𝐾), and 𝜆 is given by
𝑖 0 0 𝑏
𝑖 ↦→ 𝜆𝑖 = , 𝑗 ↦→ 𝜆 𝑗 = ; (2.3.10)
0 −𝑖 1 0
these matrices act on column vectors on the left. We then recognize the map 𝜆 given
in (2.3.2).
If 𝐾 is not a field, then 𝐾 ' 𝐹 × 𝐹, and we repeat the above argument but with
√ free module of rank 2 over 𝐾; then projecting onto one of the factors (choosing
𝐵 a
𝑎 ∈ 𝐹) gives the map 𝜆, which is still injective and therefore induces an 𝐹-algebra
isomorphism 𝐵 ' M2 (𝐹).
26 CHAPTER 2. BEGINNINGS
Remark 2.3.11. In Proposition 2.3.1, 𝐵 acts on columns on the left; if instead, one
wishes to have 𝐵 act on the right on rows, give 𝐵 the structure of a left 𝐾-vector space
and define accordingly the right regular representation instead (taking care about the
order of multiplication).
Remark 2.3.14. The left regular representation 2.3.8 is not the only way to embed 𝐵
as a subalgebra of 2 × 2-matrices. Indeed, the “splitting” of quaternion algebras in
this way, in particular the question of whether or not 𝐵 ' M2 (𝐹), is a theme that will
reappear throughout this text. For a preview, see Main Theorem 5.4.4.
2.4 ⊲ Rotations
To conclude this chapter, we return to Hamilton’s original design: quaternions model
rotations in 3-dimensional space. This development is not only historically important
but it also previews many aspects of the general theory of quaternion algebras over
fields. In this section, we follow Hamilton and take 𝑘 := 𝑖 𝑗.
Proposition 2.3.1 provides an R-algebra embedding
H1 := {𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 ∈ H : 𝑡 2 + 𝑥 2 + 𝑦 2 + 𝑧2 = 1}.
(In some contexts, one also writes GL1 (H) = H× and SL1 (H) = H1 .)
As a set, the unit Hamiltonians are naturally identified with the 3-sphere in R4 . As
groups, we have an isomorphism H1 ' SU(2) with the special unitary group of rank
2, defined by
2.4.5. Just as for the complex numbers, every element of H is the sum of its real
part and its pure (imaginary) part. And just like complex conjugation, we define a
(quaternion) conjugation map
:H→H
(2.4.6)
𝛼 = 𝑡 + (𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘) ↦→ 𝛼 = 𝑡 − (𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘)
𝛼 + 𝛼 = tr(𝜆(𝛼)) = 2𝑡
(2.4.7)
k𝛼k 2 := det(𝜆(𝛼)) = 𝛼𝛼 = 𝛼𝛼 = 𝑡 2 + 𝑥 2 + 𝑦 2 + 𝑧 2 .
The notation k k 2 is used to indicate that it agrees the usual square norm on H ' R4 .
The conjugate transpose map on M2 (C) restricts to quaternion conjugation on the
image of H in (2.4.1), also known as adjugation
𝑢 −𝑣 𝑢 𝑣
𝛼= ↦→ 𝜆(𝛼) = .
𝑣 𝑢 −𝑣 𝑢
Thus the elements 𝛼 ∈ H such that 𝜆(𝛼) = 𝜆(𝛼) (i.e., 𝐴∗ = 𝐴, and we say 𝐴 is
Hermitian), are exactly the scalar (real) matrices; and those that are skew-Hermitian,
i.e., 𝐴∗ = −𝐴, are exactly the pure quaternions. The conjugation map plays a crucial
role for quaternion algebras and is the subject of the next chapter (Chapter 3), where
to avoid confusion with other notions of conjugation we refer to it as the standard
involution.
2.4.8. Let
H0 := {𝑣 = 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 ∈ H : 𝑥, 𝑦, 𝑧 ∈ R} ' R3
28 CHAPTER 2. BEGINNINGS
be the set of pure Hamiltonians, the three-dimensional real space on which we will soon
see that the (unit) Hamiltonians act by rotations. (The reader should not confuse 𝑣 ∈ H0
with 𝑣 the entry of a 2 × 2-matrix in a local instantiation above.) For 𝑣 ∈ H0 ' R3 ,
𝑣𝑤 = −𝑣 · 𝑤 + 𝑣 × 𝑤 (2.4.10)
where 𝑣 = 𝑣 1𝑖 + 𝑣 2 𝑗 + 𝑣 3 𝑘 and 𝑤 = 𝑤 1𝑖 + 𝑤 2 𝑗 + 𝑤 3 𝑘, so
𝑣 · 𝑤 = 𝑣1 𝑤1 + 𝑣2 𝑤2 + 𝑣3 𝑤3
and
𝑣 × 𝑤 = (𝑣 2 𝑤 3 − 𝑣 3 𝑤 2 )𝑖 + (𝑣 3 𝑤 1 − 𝑣 1 𝑤 3 ) 𝑗 + (𝑣 1 𝑤 2 − 𝑣 2 𝑤 1 )𝑘.
The formula (2.4.10) is striking: it contains three different kinds of ‘multiplications’!
2.4.13. The group H1 acts on our three-dimensional space H0 (on the left) by conju-
gation:
H1 H0 → H0
(2.4.14)
𝑣 ↦→ 𝛼𝑣𝛼−1 ;
indeed, tr(𝜆(𝛼𝑣𝛼−1 )) = tr(𝜆(𝑣)) = 0 by properties of the trace, so 𝛼𝑣𝛼−1 ∈ H0 . Or
H0 = {𝑣 ∈ H : 𝑣 2 ∈ R ≤0 }
and this latter set is visibly stable under conjugation. The representation (2.4.14) is
called the adjoint representation.
2.4. ⊲ ROTATIONS 29
2.4.15. Let 𝛼 ∈ H1 r {±1}. Then there exists a unique 𝜃 ∈ (0, 𝜋) such that
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 = cos 𝜃 + (sin 𝜃)𝐼 (𝛼) (2.4.16)
where 𝐼 (𝛼) is pure and k𝐼 (𝛼) k = 1: to be precise, we take 𝜃 such that cos 𝜃 = 𝑡 and
𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘
𝐼 (𝛼) := .
|sin 𝜃|
We call 𝐼 (𝛼) the axis of 𝛼, and observe that 𝐼 (𝛼) 2 = −1.
Remark 2.4.17. In analogy with Euler’s formula, we can write (2.4.16) as
𝛼 = exp(𝐼 (𝛼)𝜃).
We are now prepared to identify this action by quaternions with rotations. As
usual, let
O(𝑛) := { 𝐴 ∈ M𝑛 (R) : 𝐴t = 𝐴−1 }
be the orthogonal group of R𝑛 (preserving the standard inner product), and let
SO(𝑛) := { 𝐴 ∈ O(𝑛) : det( 𝐴) = 1} E O(𝑛)
to be the special orthogonal group of rotations of R𝑛 , a normal subgroup of index 2
fitting into the exact sequence
det
1 → SO(𝑛) → O(𝑛) −−→ {±1} → 1.
Proposition 2.4.18. H1 acts by rotation on H0 ' R3 via conjugation (2.4.14): specif-
ically, 𝛼 acts by rotation through the angle 2𝜃 about the axis 𝐼 (𝛼).
Proof. Let 𝛼 ∈ H1 r {±1}. Then for all 𝑣 ∈ H0 ,
k𝛼𝑣𝛼−1 k 2 = k𝑣k 2
by (2.4.9), so 𝛼 acts by a matrix belonging to O(3).
But we can be more precise. Let 𝑗 0 ∈ H0 be a unit vector orthogonal to 𝑖 0 = 𝐼 (𝛼).
Then (𝑖 0) 2 = ( 𝑗 0) 2 = −1 by Lemma 2.4.12(b) and 𝑗 0𝑖 0 = −𝑖 0 𝑗 0 by Lemma 2.4.12(c),
so without loss of generality we may suppose that 𝐼 (𝛼) = 𝑖 and 𝑗 0 = 𝑗. Thus
𝛼 = 𝑡 + 𝑥𝑖 = cos 𝜃 + (sin 𝜃)𝑖 with 𝑡 2 + 𝑥 2 = cos2 𝜃 + sin2 𝜃 = 1, and 𝛼−1 = 𝑡 − 𝑥𝑖.
We have 𝛼𝑖𝛼−1 = 𝑖, and
𝛼 𝑗 𝛼−1 = (𝑡 + 𝑥𝑖) 𝑗 (𝑡 − 𝑥𝑖) = (𝑡 + 𝑥𝑖) (𝑡 + 𝑥𝑖) 𝑗
(2.4.19)
= ((𝑡 2 − 𝑥 2 ) + 2𝑡𝑥𝑖) 𝑗 = (cos 2𝜃) 𝑗 + (sin 2𝜃)𝑘
by the double angle formula. Consequently,
𝛼𝑘𝛼−1 = 𝑖(𝛼 𝑗 𝛼−1 ) = (− sin 2𝜃) 𝑗 + (cos 2𝜃)𝑘
so the matrix of 𝛼 in the basis 𝑖, 𝑗, 𝑘 is
1 0 0
𝐴 = 0 cos 2𝜃 − sin 2𝜃 ® , (2.4.20)
© ª
«0 sin 2𝜃 cos 2𝜃 ¬
a (counterclockwise) rotation (determinant 1) through the angle 2𝜃 about 𝑖.
30 CHAPTER 2. BEGINNINGS
1 → {±1} → H1 → SO(3) → 1.
Proof. The map H1 → SO(3) is surjective, since every element of SO(3) is rotation
about some axis (Exercise 2.15). If 𝛼 belongs to the kernel, then 𝛼 = cos 𝜃 +(sin 𝜃)𝐼 (𝛼)
must have sin 𝜃 = 0 so 𝛼 = ±1.
where 𝜎𝑥 , 𝜎𝑦 , 𝜎𝑧 are the famous Pauli spin matrices. Because of this application to
the spin (a kind of angular momentum) of an electron in particle physics, the group
H1 also goes by the name H1 ' Spin(3).
The extra bit of information conveyed by spin can also be seen by the “belt trick”
[Hans2006, Chapter 2].
2.4.23. We conclude with one final observation, returning to the formula (2.4.10).
There is another way to mix the dot product and cross product (2.4.11) in H: we define
the scalar triple product
H×H×H→R
(2.4.24)
(𝑢, 𝑣, 𝑤) ↦→ 𝑢 · (𝑣 × 𝑤).
Amusingly, this gives a way to “multiply” triples of triples! The map (2.4.24) defines
an alternating, trilinear form (Exercise 2.19). If 𝑢, 𝑣, 𝑤 ∈ H0 , then the scalar triple
product is a determinant
𝑢 𝑢2 𝑢3
© 1
𝑢 · (𝑣 × 𝑤) = det 𝑣 1
ª
𝑣2 𝑣3 ®
«𝑤 1 𝑤2 𝑤3¬
Exercises
Let 𝐹 be a field with char 𝐹 ≠ 2.
2.4. ⊲ ROTATIONS 31
(H1 × H1 ) H → H
(2.4.26)
𝑥 ↦→ 𝛼𝑥 𝛽−1
𝜙 : H1 × H1 → O(4).
(b) Show that 𝜙 surjects onto SO(4) < O(4). [Hint: If 𝐴 ∈ SO(4) fixes 1 ∈ H,
then 𝐴 restricted to H0 is a rotation and so is given by conjugation. More
generally, if 𝐴1 = 𝛼, consider 𝑥 ↦→ 𝛼−1 𝐴𝑥.]
(c) Show that the kernel of 𝜙 is {±1} embedded diagonally, so there is an exact
sequence
[More generally, the universal cover of SO(𝑛) for 𝑛 ≥ 3 is a double cover called
the spin group Spin(𝑛), and so Corollary 2.4.21 shows that Spin(3) ' SU(2)
and this exercise shows that Spin(4) ' SU(2) × SU(2). For further reading, see
e.g. Fulton–Harris [FH91, Lecture 20].]
18. Let 𝜌𝑢, 𝜃 : R3 → R3 be the counterclockwise rotation by the angle 𝜃 about the
axis 𝑢 ∈ R3 ' H0 , with k𝑢k = 1. Prove Rodrigues’s rotation formula: for all
𝑣 ∈ R3 ,
Cdet : M2 (𝐵) → 𝐵
𝛼 𝛽
Cdet = 𝛼𝛿 − 𝛾 𝛽
𝛾 𝛿
34 CHAPTER 2. BEGINNINGS
(a) Show that Cdet is 𝐹-multilinear in the rows and columns of the matrix.
(b) Show that Cdet is not left 𝐵-multilinear in the rows of the matrix.
(c) Give an example showing that Cdet is not multiplicative.
(d) Find a matrix 𝐴 ∈ M2 (H) that is invertible (i.e., having a two-sided inverse)
but has Cdet( 𝐴) = 0. Then find such an 𝐴 with the further property that
its transpose has nonzero determinant but is not invertible.
[Moral: be careful with matrix rings over noncommutative rings! For more on
quaternionic determinants, including the Dieudonné determinant, see Aslaksen
[Asl96].]
Chapter 3
Involutions
3.1 ⊲ Conjugation
The quaternion conjugation map (2.4.6) defined on the Hamiltonians H arises naturally
from the notion of real and pure (imaginary) parts, as defined by Hamilton. This
involution has a natural generalization to a quaternion algebra 𝐵 = (𝑎, 𝑏 | 𝐹) over a
field 𝐹 with char 𝐹 ≠ 2: we define
:𝐵→𝐵
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ↦→ 𝛼 = 𝑡 − (𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗)
Multiplying out, we then verify that
𝛼𝛼 = 𝛼𝛼 = 𝑡 2 − 𝑎𝑥 2 − 𝑏𝑦 2 + 𝑎𝑏𝑧 2 ∈ 𝐹.
The way in which the cross terms cancel, because the basis elements 𝑖, 𝑗, 𝑖 𝑗 skew
commute, is a calculation that never fails to enchant!
But this definition seems to depend on a basis: it is not intrinsically defined.
What properties characterize it? Is it unique? We are looking for a good definition
of conjugation : 𝐵 → 𝐵 on an 𝐹-algebra 𝐵: we will call such a map a standard
involution.
The involutions we consider should have basic linearity properties: they are 𝐹-
linear (with 1 = 1, so they act as the identity on 𝐹) and have order 2 as an 𝐹-linear map.
An involution should also respect the multiplication structure on 𝐵, but we should not
require that it be an 𝐹-algebra isomorphism: instead, like the inverse map (or transpose
map) reverses order of multiplication, we ask that 𝛼𝛽 = 𝛽 𝛼 for all 𝛼 ∈ 𝐵. Finally, we
want the standard involution to give rise to a trace and norm (a measure of size), which
is to say, we want 𝛼 + 𝛼 ∈ 𝐹 and 𝛼𝛼 = 𝛼𝛼 ∈ 𝐹 for all 𝛼 ∈ 𝐵. The precise definition
is given in Definition 3.2.1, and the defining properties are rigid: if an algebra 𝐵 has a
standard involution, then it is necessarily unique (Corollary 3.4.4).
35
36 CHAPTER 3. INVOLUTIONS
𝛼2 − (𝛼 + 𝛼)𝛼 + 𝛼𝛼 = 0
Theorem 3.1.1. Let 𝐵 be a division 𝐹-algebra of degree 2 over a field 𝐹 with char 𝐹 ≠
2. Then either 𝐵 = 𝐾 is a quadratic field extension of 𝐹 or 𝐵 is a division quaternion
algebra over 𝐹.
3.2 Involutions
Throughout this chapter, let 𝐵 be an 𝐹-algebra. For the moment, we allow 𝐹 to be of
arbitrary characteristic. We begin by defining involutions on 𝐵.
(i) 1 = 1;
(ii) 𝛼 = 𝛼 for all 𝛼 ∈ 𝐵; and
(iii) 𝛼𝛽 = 𝛽 𝛼 for all 𝛼, 𝛽 ∈ 𝐵 (the map is an anti-automorphism).
3.2.2. We define the opposite algebra of 𝐵 by letting 𝐵op = 𝐵 as 𝐹-vector spaces but
with multiplication 𝛼 ·op 𝛽 = 𝛽 · 𝛼 for 𝛼, 𝛽 ∈ 𝐵.
One can then equivalently define an involution to be an 𝐹-algebra isomorphism
∼ 𝐵op whose underlying 𝐹-linear map has order at most 2.
𝐵 −→
Remark 3.2.5. Standard involutions go by many other names. The terminology stan-
dard is employed because conjugation on a quaternion algebra is the “standard” ex-
ample of such an involution. Other authors call the standard involution the main
3.2. INVOLUTIONS 37
involution for quaternion algebras, but then find situations where the “main” involu-
tion is not standard by our definition. The standard involution is also called conjugation
on 𝐵, but this can be confused with conjugation by an element in 𝐵× . We will see in
Corollary 3.4.4 that a standard involution is unique, so it is also called the canonical
involution; however, there are other circumstances where involutions can be defined
canonically that are not standard (like the map induced by 𝑔 ↦→ 𝑔 −1 on the group ring
𝐹 [𝐺]).
(𝛼 + 1) (𝛼 + 1) = (𝛼 + 1) (𝛼 + 1) = 𝛼𝛼 + 𝛼 + 𝛼 + 1 ∈ 𝐹
and hence 𝛼 + 𝛼 ∈ 𝐹 for all 𝛼 ∈ 𝐵 as well; it then also follows that 𝛼𝛼 = 𝛼𝛼, since
3.2.9. Suppose char 𝐹 ≠ 2 and let 𝐵 = (𝑎, 𝑏 | 𝐹). Then the map
:𝐵→𝐵
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ↦→ 𝛼 = 𝑡 − 𝑥𝑖 − 𝑦 𝑗 − 𝑧𝑖 𝑗
𝑖 𝑗 = 𝑖 𝑗 = −𝑖 𝑗 = 𝑗𝑖 = (− 𝑗) (−𝑖) = 𝑗 𝑖
(𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗) (𝑡 − 𝑥𝑖 − 𝑦 𝑗 − 𝑧𝑖 𝑗) = 𝑡 2 − 𝑎𝑥 2 − 𝑏𝑦 2 + 𝑎𝑏𝑧2 ∈ 𝐹. (3.2.10)
Remark 3.2.11. Algebras with involution play an important role in analysis, in particu-
lar Banach algebras with involution and 𝐶 ∗ -algebras (generally of infinite dimension).
A good reference is the text by Dixmier [Dix77] (or the more introductory book by
Conway [Con2012]).
38 CHAPTER 3. INVOLUTIONS
trd : 𝐵 → 𝐹
(3.3.1)
𝛼 ↦→ 𝛼 + 𝛼
nrd : 𝐵 → 𝐹
(3.3.2)
𝛼 ↦→ 𝛼𝛼.
Example 3.3.3. For 𝐵 = M2 (𝐹), equipped with the adjugate map as a standard
involution as in Example 3.2.8, the reduced trace is the usual matrix trace and the
reduced norm is the determinant.
3.3.4. The reduced trace trd is an 𝐹-linear map, since this is true for the standard
involution:
for all 𝛼, 𝛽 ∈ 𝐵.
𝐵0 := {𝛼 ∈ 𝐵 : trd(𝛼) = 0}
(3.3.5)
𝐵1 := {𝛼 ∈ 𝐵× : nrd(𝛼) = 1}
for the 𝐹-subspace 𝐵0 ⊆ 𝐵 of elements of reduced trace 0 and for the subgroup
𝐵1 ≤ 𝐵× of elements of reduced norm 1. We observe that 𝐵1 E 𝐵× is normal, by
multiplicativity, indeed we have an exact sequence of groups
nrd
1 → 𝐵1 → 𝐵× −−→ 𝐹 ×
Lemma 3.3.6. If 𝐵 is not the zero ring, then 𝛼 ∈ 𝐵 is a unit (has a two-sided inverse)
if and only if nrd(𝛼) ≠ 0.
Proof. We have
and so
trd(𝛼𝛽) = trd(𝛼𝛽) = trd(𝛽𝛼) = trd(𝛼) trd(𝛽) − trd(𝛽𝛼)
therefore trd(𝛼𝛽) = trd(𝛽𝛼).
Remark 3.3.8. The maps trd and nrd are called reduced for the following reason.
Let 𝐴 be a finite-dimensional 𝐹-algebra, and consider the left regular representation
𝜆 : 𝐴 ↩→ End𝐹 ( 𝐴) given by left multiplication in 𝐴 (cf. Proposition 2.3.1, but over 𝐹).
We then have a (left) trace map Tr : 𝐴 → 𝐹 and (left) norm map Nm : 𝐴 → 𝐹 given
by mapping 𝛼 ∈ 𝐵 to the trace and determinant of the endomorphism 𝜆 𝛼 ∈ End𝐹 ( 𝐴).
When 𝐴 = M2 (𝐹), a direct calculation (Exercise 3.13) reveals that
(algebra trace, reduced trace, and matrix trace, respectively; there is no difference
between left and right), and
for all 𝛼 ∈ 𝐴, whence the name reduced. (To preview the language of chapter 7, this
calculation can be efficiently summarized: as a left 𝐴-module, 𝐴 is the sum of two
simple 𝐴-modules—acting on the columns of a matrix—and the reduced trace and
reduced norm represent ‘half’ of this action.)
3.3.9. Since
𝛼2 − (𝛼 + 𝛼)𝛼 + 𝛼𝛼 = 0 (3.3.10)
identically we see that 𝛼 ∈ 𝐵 is a root of the polynomial
which we call the reduced characteristic polynomial of 𝛼. The fact that 𝛼 satisfies its
reduced characteristic polynomial is the reduced Cayley–Hamilton theorem for an
algebra with standard involution. When 𝛼 ∉ 𝐹, the reduced characteristic polynomial
of 𝛼 is its minimal polynomial, since if 𝛼 satisfies a polynomial of degree 1 then 𝛼 ∈ 𝐹.
Example 3.4.3. The reduced trace and norm on a quadratic algebra are precisely the
usual algebra trace and norm. If char 𝐹 ≠ 2 and 𝐾 ⊇ 𝐹 is a quadratic field extension
of 𝐹, then the standard involution is just the nontrivial element of Gal(𝐾 | 𝐹).
Proof. For any 𝛼 ∈ 𝐵r𝐹, we have from (3.3.10) that dim𝐹 𝐹 [𝛼] = 2, so the restriction
of the standard involution to 𝐹 [𝛼] is unique. Therefore the standard involution on 𝐵
is itself unique.
We have seen that the equation (3.3.10), implying that if 𝐵 has a standard involution
then every 𝛼 ∈ 𝐵 satisfies a quadratic equation, has figured prominently in the above
proofs. To further clarify the relationship between these two notions, we make the
following definition.
Definition 3.4.5. The degree of 𝐵 is the smallest 𝑚 ∈ Z ≥0 such that every element
𝛼 ∈ 𝐵 satisfies a monic polynomial 𝑓 (𝑥) ∈ 𝐹 [𝑥] of degree 𝑚, if such an integer exists;
otherwise, we say 𝐵 has degree ∞.
3.4.6. If 𝐵 has finite dimension 𝑛 = dim𝐹 𝐵 < ∞, then every element of 𝐵 satisfies
a polynomial of degree at most 𝑛: if 𝛼 ∈ 𝐵 then the elements 1, 𝛼, . . . , 𝛼 𝑛 are
linearly dependent over 𝐹. Consequently, every finite-dimensional 𝐹-algebra has a
(well-defined) integer degree, at most 𝑛.
Theorem 3.5.1. Suppose char 𝐹 ≠ 2 and let 𝐵 be a division 𝐹-algebra. Then 𝐵 has
degree at most 2 if and only if one of the following hold:
(i) 𝐵 = 𝐹;
(ii) 𝐵 = 𝐾 is a quadratic field extension of 𝐹; or
(iii) 𝐵 is a division quaternion algebra over 𝐹.
3.5. QUATERNION ALGEBRAS 41
Proof. From Example 3.4.7, we may suppose that 𝐵 ≠ 𝐹 and 𝐵 has degree 2.
Let 𝑖 ∈ 𝐵 r 𝐹. Then 𝐹 [𝑖] = 𝐾 is a (commutative) quadratic 𝐹-subalgebra of the
division ring 𝐵, so 𝐾 = 𝐹 (𝑖) is a field. If 𝐾 = 𝐵, we are done. Completing the square
(since char 𝐹 ≠ 2), we may suppose that 𝑖 2 = 𝑎 ∈ 𝐹 × .
Let 𝜙 : 𝐵 → 𝐵 be the map given by conjugation by 𝑖, i.e., 𝜙(𝛼) = 𝑖 −1 𝛼𝑖. Then 𝜙
is a 𝐾-linear endomorphism of 𝐵, thought of as a (left) 𝐾-vector space, and 𝜙2 is the
identity on 𝐵. Therefore 𝜙 is diagonalizable, and we may decompose 𝐵 = 𝐵+ ⊕ 𝐵−
into eigenspaces for 𝜙: explicitly, we can always write
𝛼 + 𝜙(𝛼) 𝛼 − 𝜙(𝛼)
𝛼= + ∈ 𝐵+ ⊕ 𝐵− .
2 2
We now prove dim𝐾 𝐵+ = 1. Let 𝛼 ∈ 𝐵+ . Then 𝐿 = 𝐹 (𝛼, 𝑖) is a field. Since
char 𝐹 ≠ 2, and 𝐿 is a compositum of quadratic extensions of 𝐹, the primitive element
theorem implies that 𝐿 = 𝐹 (𝛽) for some 𝛽 ∈ 𝐿. But by hypothesis 𝛽 satisfies a
quadratic equation so dim𝐹 𝐿 = 2 and hence 𝐿 = 𝐾. (For an alternative direct proof
of this claim, see Exercise 3.10.)
If 𝐵 = 𝐵+ = 𝐾, we are done. So suppose 𝐵− ≠ {0}. We will prove that
dim𝐾 𝐵− = 1. If 0 ≠ 𝑗 ∈ 𝐵− then 𝑖 −1 𝑗𝑖 = − 𝑗, so 𝑖 = − 𝑗 −1𝑖 𝑗 and hence all elements of
𝐵− conjugate 𝑖 to −𝑖. Thus if 0 ≠ 𝑗1 , 𝑗2 ∈ 𝐵− then 𝑗 1 𝑗 2 centralizes 𝑖 and 𝑗 1 𝑗 2 ∈ 𝐵+ = 𝐾.
Thus any two nonzero elements of 𝐵− are 𝐾-multiples of each other.
Finally, let 𝑗 ∈ 𝐵− r {0}; then 𝐵 = 𝐵+ ⊕ 𝐵− = 𝐾 ⊕ 𝐾 𝑗 so 𝐵 has 𝐹-basis 1, 𝑖, 𝑗, 𝑖 𝑗
and 𝑗𝑖 = −𝑖 𝑗. We claim that trd( 𝑗) = 0: indeed, both 𝑗 and 𝑖 −1 𝑗𝑖 = − 𝑗 satisfy
the same reduced characteristic (or minimal) polynomial of degree 2, so trd( 𝑗) =
trd(− 𝑗) = − trd( 𝑗) so trd( 𝑗) = 0. Thus 𝑗 2 = 𝑏 ∈ 𝐹 × , and 𝐵 is a quaternion algebra by
definition.
Corollary 3.5.4. Let 𝐵 be a division 𝐹-algebra with char 𝐹 ≠ 2. Then 𝐵 has degree
at most 2 if and only if 𝐵 has a standard involution.
Proof. In each of the cases (i)–(iii), 𝐵 has a standard involution; and conversely if 𝐵
has a standard involution, then 𝐵 has degree at most 2 (Example 3.4.7).
42 CHAPTER 3. INVOLUTIONS
Corollary 3.5.6. Let 𝐵 be a division 𝐹-algebra with char 𝐹 ≠ 2. Then the following
are equivalent:
If 𝐵 has finite degree (such as when dim𝐹 𝐵 = 𝑛 < ∞), then 𝐵 is algebraic.
Remark 3.5.10. The theorem of Frobenius (Corollary 3.5.8) extends directly to fields
𝐹 akin to R, as follows. A field is formally real if −1 cannot be expressed in 𝐹 as a
sum of squares and real closed if 𝐹 is formally real and has no formally real proper
algebraic extension. The real numbers R and the field of all real algebraic numbers are
real closed. A real closed field has characteristic zero, is totally ordered, and contains
a square root of each nonnegative element; the field obtained from 𝐹 by adjoining a
root of the irreducible polynomial 𝑥 2 + 1 is algebraically closed. For these statements,
see Rajwade [Raj93, Chapter 15]. Every √ finite-dimensional division algebra over a
real closed field 𝐹 is either 𝐹 or 𝐾 = 𝐹 ( −1) or 𝐵 = (−1, −1 | 𝐹).
Remark 3.5.11. Algebras of dimension 3, sitting somehow between quadratic ex-
tensions and quaternion algebras, can be characterized in a similar way. If 𝐵 is an
R-algebra of dimension 3, then either 𝐵 is commutative or 𝐵 has a standard involution
and is isomorphic to the subring of upper triangular matrices in M2 (R). A similar
statement holds for free 𝑅-algebras of rank 3 over a (commutative) domain 𝑅; see
Levin [Lev2013].
3.5. QUATERNION ALGEBRAS 43
Exercises
Throughout these exercises, let 𝐹 be a field.
I 1. Let 𝐵 be an 𝐹-algebra and let : 𝐵 → 𝐵 be an 𝐹-linear map with 1 = 1. Show
that is an involution if and only if (ii)–(iii) in Definition 3.2.1 hold for a basis
of 𝐵 (as an 𝐹-vector space).
I 2. Let 𝐾 = 𝐹 [𝛼] be a quadratic 𝐹-algebra, with 𝛼2 = 𝑡𝛼 − 𝑛 for (unique) 𝑡, 𝑛 ∈ 𝐹.
Extending linearly, show that there is a unique standard involution : 𝐾 → 𝐾
with the property that 𝛼 = 𝑡 − 𝛼, and show that
trd(𝑥 + 𝑦𝛼) = 2𝑥 + 𝑡𝑦
nrd(𝑥 + 𝑦𝛼) = 𝑥 2 + 𝑡𝑥𝑦 + 𝑛𝑦 2
for all 𝑥 + 𝑦𝛼 ∈ 𝐹 [𝛼].
I 3. Verify that the map in Example 3.2.9 is a standard involution.
4. Determine the standard involution on 𝐾 = 𝐹 × 𝐹 (with 𝐹 ↩→ 𝐾 under the
diagonal map).
I 5. Let 𝐵 be an 𝐹-algebra with a standard involution. Show that 0 ≠ 𝛼 ∈ 𝐵 is a left
zerodivisor if and only if 𝛼 is a right zerodivisor if and only if nrd(𝛼) = 0. In
particular, if 𝐵 is not the zero ring, then 𝛼 ∈ 𝐵 is (left and right) invertible if
and only if nrd(𝛼) ≠ 0.
6. Suppose char 𝐹 ≠ 2, let 𝐵 be a division quaternion algebra over 𝐹, and let
𝐾1 , 𝐾2 ⊆ 𝐵 be subfields with 𝐾1 ∩ 𝐾2 = 𝐹. Show that the 𝐹-subalgebra of 𝐵
generated by 𝐾1 and 𝐾2 is equal to 𝐵. Conclude that if 1, 𝛼, 𝛽 ∈ 𝐵 are 𝐹-linearly
independent, then 1, 𝛼, 𝛽, 𝛼𝛽 are an 𝐹-basis for 𝐵. [Hint: use the involution.]
By way of counterexample, show that these results need not hold for 𝐵 = M2 (𝐹).
7. Show that 𝐵 = M𝑛 (𝐹) has a standard involution if and only if 𝑛 ≤ 2.
8. Let 𝐺 be a finite group. Show that the 𝐹-linear mapÉ induced by 𝑔 ↦→ 𝑔 −1
for 𝑔 ∈ 𝐺 is an involution on the group ring 𝐹 [𝐺] = 𝑔 ∈𝐺 𝐹𝑔. Determine
necessary and sufficient conditions for this map to be a standard involution.
9. Let 𝐵 be an 𝐹-algebra with a standard involution : 𝐵 → 𝐵. In this exercise,
we examine when is the identity map.
(a) Show that if char 𝐹 ≠ 2, then 𝑥 ∈ 𝐵 satisfies 𝑥 = 𝑥 if and only 𝑥 ∈ 𝐹.
(b) Suppose that dim𝐹 𝐵 < ∞. Show that the identity map is a standard
involution on 𝐵 if and only if (i) 𝐵 = 𝐹 or (ii) char 𝐹 = 2 and 𝐵 is a
quotient of the commutative ring 𝐹 [𝑥1 , . . . , 𝑥 𝑛 ]/(𝑥12 − 𝑎 1 , . . . , 𝑥 𝑛2 − 𝑎 𝑛 )
with 𝑎 𝑖 ∈ 𝐹.
10. Let 𝐾 ⊇ 𝐹 be a field which has degree 𝑚 as an 𝐹-algebra in the sense of
Definition 3.4.5. Suppose that char 𝐹 - 𝑚. Show that [𝐾 : 𝐹] = 𝑚, i.e., 𝐾 has
degree 𝑚 in the usual sense. (What happens when char 𝐹 | 𝑚?)
11. Let 𝐵 be an 𝐹-algebra with standard involution. Suppose that 𝜙 : 𝐵 −→∼ 𝐵 is
an 𝐹-algebra automorphism. Show for 𝛼 ∈ 𝐵 that 𝜙(𝛼) = 𝜙(𝛼), and therefore
44 CHAPTER 3. INVOLUTIONS
that trd(𝜙(𝛼)) = trd(𝛼) and nrd(𝜙(𝛼)) = nrd(𝛼). [Hint: consider the map
𝛼 ↦→ 𝜙−1 (𝜙(𝛼)).]
12. In this exercise, we explore further the relationship between algebras of degree
2 and those with standard involutions (Remark 3.5.5).
(a) Suppose char 𝐹 ≠ 2 and let 𝐵 be a finite-dimensional 𝐹-algebra. Show
that 𝐵 has a standard involution if and only if deg𝐹 𝐵 ≤ 2.
(b) Let 𝐹 = F2 and let 𝐵 be a Boolean ring, a ring such that 𝑥 2 = 𝑥 for all
𝑥 ∈ 𝐵. (Verify that 2 = 0 in 𝐵, so 𝐵 is an F2 -algebra.) Prove that 𝐵 does not
have a standard involution unless 𝐵 = F2 or 𝐵 = F2 × F2 , but nevertheless
any Boolean ring has degree at most 2.
I 13. Let 𝐵 = M𝑛 (𝐹), and consider the map 𝜆 : 𝐵 ↩→ End𝐹 (𝐵) by 𝛼 ↦→ 𝜆 𝛼 defined
by left-multiplication in 𝐵. Show that for all 𝛼 ∈ M𝑛 (𝐹), the characteristic
polynomial of 𝜆 𝛼 is the 𝑛th power of the usual characteristic polynomial of 𝛼.
Conclude when 𝑛 = 2 that tr(𝛼) = 2 trd( 𝐴) and det(𝛼) = nrd(𝛼) 2 .
14. Considering a slightly different take on the previous exercise: let 𝐵 be a quater-
nion algebra over 𝐹. Show that the characteristic polynomial of left multipli-
cation by 𝛼 ∈ 𝐵 is equal to that of right multiplication and is the square of the
reduced characteristic polynomial. [Hint: if a direct approach is too cumber-
some, consider applying the previous exercise and the left regular representation
as in 2.3.8.]
15. Let 𝑉 be an 𝐹-vector space and let 𝑡 : 𝑉 → 𝐹 be an 𝐹-linear map. Let 𝐵 = 𝐹 ⊕𝑉
and define the binary operation 𝑥 · 𝑦 = 𝑡 (𝑥)𝑦 for 𝑥, 𝑦 ∈ 𝑉. Show that · induces
a multiplication on 𝐵, and that the map 𝑥 ↦→ 𝑥 = 𝑡 (𝑥) − 𝑥 for 𝑥 ∈ 𝑉 induces
a standard involution on 𝐵. [Such an algebra is called an exceptional algebra
[GrLu2009, Voi2011b].] Conclude that there exists a central 𝐹-algebra 𝐵 with
a standard involution in any dimension 𝑛 = dim𝐹 𝐵 ≥ 3.
I 16. In this exercise, we mimic the proof of Theorem 3.5.1 to prove that a quaternion
algebra over a finite field of odd cardinality is not a division ring, a special case
of Wedderburn’s little theorem: a finite division ring is a field.
Assume for purposes of contradiction that 𝐵 is a division quaternion algebra
over 𝐹 = F𝑞 with 𝑞 odd.
(a) Let 𝑖 ∈ 𝐵 r 𝐹. Show that the centralizer 𝐶 𝐵× (𝑖) = {𝛼 ∈ 𝐵× : 𝑖𝛼 = 𝛼𝑖} of 𝑖
in 𝐵× satisfies 𝐶 𝐵× (𝑖) = 𝐹 (𝑖) × .
(b) Conclude that any noncentral conjugacy class in 𝐵× has order 𝑞 2 + 1.
(c) Derive a contradiction from the class equation 𝑞 4 − 1 = 𝑞 − 1 + 𝑚(𝑞 2 + 1)
(where 𝑚 ∈ Z).
[For the case 𝑞 even, see Exercise 6.16; for fun, the eager reader may wish to
prove Weddernburn’s little theorem for 𝐹 = F2 directly.]
17. Derive Euler’s identity (1.1.6) that the product of the sum of four squares is
again the sum of four squares as follows. Let 𝐹 = Q(𝑥1 , . . . , 𝑥 4 , 𝑦 1 , . . . , 𝑦 4 ) be a
function field over Q in 8 variables and consider the quaternion algebra (−1, −1 |
𝐹). Show (by an explicit universal formula) that if 𝑅 is any commutative ring
3.5. QUATERNION ALGEBRAS 45
and 𝑥, 𝑦 ∈ 𝑅 are the sum of four squares in 𝑅, then 𝑥𝑦 is the sum of four squares
in 𝑅.
18. Suppose char 𝐹 ≠ 2. For an 𝐹-algebra 𝐵, let
𝑉 (𝐵) = {𝛼 ∈ 𝐵 r 𝐹 : 𝛼2 ∈ 𝐹} ∪ {0}.
Let 𝐵 be a division ring. Show that 𝑉 (𝐵) is a vector space (closed under
addition) if and only if 𝐵 = 𝐹 or 𝐵 = 𝐾 is a quadratic field extension of 𝐹 or 𝐵
is a quaternion algebra over 𝐹.
19. Let 𝐵 be an 𝐹-algebra with 𝐹-basis 𝑒 1 , 𝑒 2 , . . . , 𝑒 𝑛 . Let : 𝐵 → 𝐵 be an
involution. Show that is standard if and only if
Quadratic forms
47
48 CHAPTER 4. QUADRATIC FORMS
An isometry 𝑓 ∈ O(𝑄) (𝐹) is special if det 𝑓 = 1, and the special orthogonal group
of 𝑄 is the group of special isometries of 𝑄.
More generally, we have seen that any algebra with a standard involution has a
quadratic form nrd. We say that the standard involution is nondegenerate whenever
the quadratic form nrd is so. Generalizing Theorem 3.1.1, we prove the following (see
Main Theorem 4.4.1 for the proof).
(i) 𝐵 = 𝐹;
(ii) 𝐵 = 𝐾 has dim𝐹 𝐾 = 2 and either 𝐾 ' 𝐹 × 𝐹 or 𝐾 is a field; or
(iii) 𝐵 is a quaternion algebra over 𝐹.
This theorem gives another way of characterizing quaternion algebras: they are
noncommutative algebras with a nondegenerate standard involution.
In Section 2.4, we saw that the unit Hamiltonians H1 act on the pure Hamiltonians
0
H (Section 2.4) by rotations: the standard Euclidean quadratic form (sum of squares)
is preserved by conjugation. This generalizes in a natural way to an arbitrary field,
and so we can understand the group of linear transformations that preserve a ternary
(or quaternary) form in terms of the unit group of a quaternion algebra 𝐵 (Proposition
4.5.10): there is an exact sequence
1 → 𝐹 × → 𝐵× → SO(nrd | 𝐵0 ) (𝐹) → 1
where SO(𝑄) (𝐹) is the group of special (or oriented) isometries of the quadratic form
𝑄 and 𝐵0 := {𝛼 ∈ 𝐵 : trd(𝛼) = 0}.
𝑉
𝑄
/𝐹
o 𝑓 o 𝑢 (4.2.5)
𝑄0
𝑉0 /𝐹
commutes. In a similarity ( 𝑓 , 𝑢), the scalar 𝑢 is called the similitude factor of the
similarity. An isometry of quadratic forms (or isomorphism of quadratic spaces) is a
similarity with similitude factor 𝑢 = 1; we write in this case 𝑄 ' 𝑄 0.
Definition 4.2.6. The general orthogonal group (or similarity group) of the quadratic
form 𝑄 is the group of self-similarities of 𝑄 under composition
GO(𝑄) (𝐹) := {( 𝑓 , 𝑢) ∈ Aut𝐹 (𝑉) × 𝐹 × : 𝑄( 𝑓 (𝑥)) = 𝑢𝑄(𝑥) for all 𝑥 ∈ 𝑉 };
the orthogonal group of 𝑄 is the group of self-isometries of 𝑄, i.e.,
O(𝑄) (𝐹) := { 𝑓 ∈ Aut𝐹 (𝑉) : 𝑄( 𝑓 (𝑥)) = 𝑄(𝑥) for all 𝑥 ∈ 𝑉 }.
50 CHAPTER 4. QUADRATIC FORMS
realizing O(𝑄) (𝐹) ≤ GO(𝑄) (𝐹) as the subgroup of self-similarities with similitude
factor 𝑢 = 1.
4.2.9. Returning to 4.2.2, suppose dim𝐹 𝑉 = 𝑛 < ∞ and char 𝐹 ≠ 2. Then one can
understand the orthogonal group of 𝑄 quite concretely in matrix terms as follows.
Choose a basis 𝑒 1 , . . . , 𝑒 𝑛 for 𝑉 and let [𝑇] be the Gram matrix of 𝑄 with respect to
this basis, so that 2𝑄(𝑥) = 𝑥 t [𝑇]𝑥 for all 𝑥 ∈ 𝑉 ' 𝐹 𝑛 . Then Aut𝐹 (𝑉) ' GL𝑛 (𝐹) and
𝐴 ∈ GL𝑛 (𝐹) belongs to O(𝑄) if and only if
and
GO(𝑄) (𝐹) = {( 𝐴, 𝑢) ∈ GL𝑛 (𝐹) × 𝐹 × : 𝐴t [𝑇] 𝐴 = 𝑢[𝑇]} (4.2.11)
Theorem 4.2.22. Let 𝑉 ' 𝑉 0 be isometric quadratic spaces with orthogonal decom-
positions 𝑉 ' 𝑊1 𝑊2 and 𝑉 0 ' 𝑊10 𝑊20 .
Proof. The proof is requested in Exercise 4.16. For a proof and the equivalence be-
tween Witt cancellation (part (a)) and Witt extension (part (b)), see Lam [Lam2005,
Proof of Theorem I.4.2, p. 14], Scharlau [Scha85, Theorem 1.5.3], or O’Meara
[O’Me73, Theorem 42:17].
Theorem 4.2.22(a) is called Witt cancellation and 4.2.22(b) is called Witt exten-
sion.
A form presented with a basis as in Lemma 4.3.1 is called normalized (or diag-
onal). For a diagonal quadratic form 𝑄, the associated Gram matrix [𝑇] is diagonal
with entries 2𝑎 1 , . . . , 2𝑎 𝑛 .
4.3.2. The determinant det( [𝑇]) of a Gram matrix for 𝑄 depends on a choice of
basis for 𝑉, but by (4.2.3), a change of basis matrix 𝐴 ∈ GL𝑛 (𝐹) operates on [𝑇] by
𝐴t [𝑇] 𝐴, and det( 𝐴t [𝑇] 𝐴) = det( 𝐴) 2 det( [𝑇]), so we obtain a well-defined element
det(𝑇) ∈ 𝐹/𝐹 ×2 independent of the choice of basis.
When it will cause no confusion, we will represent the class of the discriminant in
𝐹/𝐹 ×2 simply by a representative element in 𝐹.
Remark 4.3.4. The extra factor 2−𝑛 is harmless since char 𝐹 ≠ 2, and it allows us to
naturally cancel certain factors 2 that appear whether we are in even or odd dimension—
it will be essential when we consider the case char 𝐹 = 2 (see 6.3.1). The distinction
between even and odd dimensional quadratic spaces is not arbitrary: indeed, this
distinction is pervasive, even down to the classification of semisimple Lie algebras.
𝑉 → Hom(𝑉, 𝐹)
𝑥 ↦→ (𝑦 ↦→ 𝑇 (𝑥, 𝑦))
and 𝑇 is nondegenerate if and only if this map is injective (and hence an isomorphism) if
and only if det(𝑇) ≠ 0. Put another way, 𝑄 is nondegenerate if and only if disc(𝑄) ≠ 0,
and so a diagonal form h𝑎 1 , . . . , 𝑎 𝑛 i is nondegenerate if and only if 𝑎 𝑖 ≠ 0 for all 𝑖.
4.3.9. One can often restrict to the case where a quadratic form 𝑄 is nondegenerate
by splitting off the radical, as follows. We define the radical of 𝑄 to be
Proof of Main Theorem 4.4.1. If 𝐵 = 𝐹, then the standard involution is the identity
and nrd is nondegenerate. If dim𝐹 𝐾 = 2, then after completing the square we may
write 𝐾 ' 𝐹 [𝑥]/(𝑥 2 − 𝑎) and in the basis 1, 𝑥 we find nrd ' h1, 𝑎i. By Example 4.3.5,
nrd is nondegenerate if and only if 𝑎 ∈ 𝐹 × if and only if 𝐾 is a quadratic field extension
of 𝐹 or 𝐾 ' 𝐹 × 𝐹.
Suppose that dim𝐹 𝐵 > 2. Let 1, 𝑖, 𝑗 be a part of a normalized basis for 𝐵 with
respect to the quadratic form nrd. Then 𝑇 (1, 𝑖) = trd(𝑖) = 0, so 𝑖 2 = 𝑎 ∈ 𝐹 × , since
nrd is nondegenerate. Note in particular that 𝑖 = −𝑖. Similarly 𝑗 2 = 𝑏 ∈ 𝐹 × , and
by (4.2.16) we have trd(𝑖 𝑗) = 𝑖 𝑗 + 𝑗𝑖 = 0. We have 𝑇 (1, 𝑖 𝑗) = trd(𝑖 𝑗) = 0, and
𝑇 (𝑖 𝑗, 𝑖) = trd(𝑖(𝑖 𝑗)) = −𝑎 trd( 𝑗) = 0 and similarly 𝑇 (𝑖 𝑗, 𝑗) = 0, hence 𝑖 𝑗 ∈ {1, 𝑖, 𝑗 }⊥ .
If 𝑖 𝑗 = 0 then 𝑖(𝑖 𝑗) = 𝑎 𝑗 = 0 so 𝑗 = 0, a contradiction. Since nrd is nondegenerate, it
follows then that the set 1, 𝑖, 𝑗, 𝑖 𝑗 is linearly independent.
Therefore, the subalgebra 𝐴 of 𝐵 generated by 𝑖, 𝑗 satisfies 𝐴 ' (𝑎, 𝑏 | 𝐹), and if
dim𝐹 𝐵 = 4 we are done. So let 𝑘 ∈ 𝐴⊥ ; then trd(𝑘) = 0 and 𝑘 2 = 𝑐 ∈ 𝐹 × . Thus
𝑘 ∈ 𝐵× , with 𝑘 −1 = 𝑐−1 𝑘. By 4.2.13 we have 𝑘𝛼 = 𝛼𝑘 for any 𝛼 ∈ 𝐴 since 𝑘 = −𝑘.
But then
𝑘 (𝑖 𝑗) = (𝑖 𝑗)𝑘 = 𝑗 𝑖𝑘 = 𝑗 𝑘𝑖 = 𝑘 ( 𝑗𝑖). (4.4.3)
But 𝑘 ∈ 𝐵× so 𝑖 𝑗 = 𝑗𝑖 = −𝑖 𝑗, and this is a contradiction.
Proof. Immediate.
Corollary 4.4.5. Let 𝐵 have a nondegenerate standard involution, and suppose that
𝐾 ⊆ 𝐵 is a commutative 𝐹-subalgebra such that the restriction of the standard
involution is nondegenerate. Then dim𝐹 𝐾 ≤ 2. Moreover, if 𝐾 ≠ 𝐹, then the
centralizer of 𝐾 × in 𝐵× is 𝐾 × .
4.5. SPECIAL ORTHOGONAL GROUPS 55
Proof. The first statement is immediate; the second follows by considering the algebra
generated by the centralizer.
Remark 4.4.6. Algebras with involutions come from quadratic forms, and the re-
sults of this chapter are just one special case of a much more general theory. More
precisely, there is a natural bijection between the set of isomorphism classes of finite-
dimensional simple 𝐹-algebras equipped with an 𝐹-linear involution and the set of
similarity classes of nondegenerate quadratic forms on finite-dimensional 𝐹-vector
spaces. More generally, for involutions that act nontrivially on the base field, one
looks at Hermitian forms. Consequently, there are three broad types of involutions
on central simple algebras, depending on the associated quadratic or Hermitian form:
orthogonal, symplectic, and unitary. Accordingly, algebras with involutions can be
classified by the invariants of the associated form. This connection is the subject
of the tome by Knus–Merkurjev–Rost–Tignol [KMRT98]. In this way the theory of
quadratic forms belongs to the theory of algebras with involution, which in turn is a
part of the theory of linear algebraic groups, as expounded by Weil [Weil60]: see the
survey by Tignol [Tig98] for an overview and further references.
If 𝑛 is odd, we have little choice other than to define GSO(𝑄) (𝐹) := GO(𝑄) (𝐹).
is the group of linear maps preserving length (but not necessarily orientation), whereas
SO(𝑄) (R) is the usual group of rotations of 𝑉 (preserving orientation). Similarly,
GSO(𝑄) (R) consists of orientation-preserving similarities, preserving orientation but
allowing a constant scaling.
In particular, if 𝑛 = 2 then O(2) := O(𝑄) (R) contains
cos 𝜃 sin 𝜃
:=
SO(2) SO(𝑄) (R) = : 𝜃 ∈ R ' R/(2𝜋Z) ' S1
− sin 𝜃 cos 𝜃
(the circle group) with index 2, with a reflection in any line through the origin repre-
senting a nontrivial coset of SO(2) ≤ O(2).
4.5.6. More generally, we may define reflections in O(𝑄) (𝐹) as follows. For 𝑥 ∈ 𝑉
anisotropic (so 𝑄(𝑥) ≠ 0), we define the reflection in 𝑥 to be
𝜏𝑥 : 𝑉 → 𝑉
𝑇 (𝑣, 𝑥)
𝜏𝑥 (𝑣) = 𝑣 − 𝑥.
𝑄(𝑥)
Proof. See Lam [Lam2005, §I.7], O’Meara [O’Me73, §43B], or Scharlau [Scha85,
Theorem 1.5.4]. The proof is by induction on 𝑛, carefully recording the effect of a
reflection in an anisotropic vector.
4.5. SPECIAL ORTHOGONAL GROUPS 57
Since reflections have determinant −1, an isometry 𝑓 is special (and 𝑓 ∈ SO(𝑄) (𝐹))
if and only if it is the product of an even number of reflections.
4.5.8. Now let 𝐵 be a quaternion algebra over 𝐹, and recall (3.3.5) that we have defined
𝐵0 := {𝑣 ∈ 𝐵 : trd(𝐵) = 0}.
Proposition 4.5.10. Let 𝐵 be a quaternion algebra over 𝐹. Then the action (4.5.9)
induces an exact sequence
𝑇 (𝑣, 𝑥) trd(𝑣𝑥)
𝜏𝑥 (𝑣) = 𝑣 − 𝑥=𝑣− 𝑥
𝑄(𝑥) nrd(𝑥) (4.5.12)
= 𝑣 − (𝑣𝑥 + 𝑥𝑣)𝑥 −1 = −𝑥𝑣 𝑥 −1 = 𝑥𝑣𝑥 −1 ,
the final equality from 𝑥 = −𝑥 as 𝑥 ∈ 𝐵0 . The product of two such reflections is thus
of the form 𝑣 ↦→ 𝛼𝑣𝛼−1 with 𝛼 ∈ 𝐵× . Therefore 𝐵× acts by special isometries, and
every special isometry so arises: the map 𝐵× → O(𝑄) (𝐹) surjects onto SO(𝑄) (𝐹).
The kernel of the action is given by those 𝛼 ∈ 𝐵× with 𝛼𝑣𝛼−1 = 𝑣 for all 𝑣 ∈ 𝐵0 , i.e.,
𝛼 ∈ 𝑍 (𝐵× ) = 𝐹 × .
The second statement follows directly by writing 𝐵× = 𝐵1 𝐹 × .
Example 4.5.13. If 𝐵 ' M2 (𝐹), then nrd = det, so det0 ' h1, −1, −1i and (4.5.11)
yields the isomorphism PGL2 (𝐹) ' SO(h1, −1, −1i) (𝐹).
Example 4.5.14. If 𝐹 = R and 𝐵 = H, then det(H) = R>0 = R×2 , and the second
exact sequence is Hamilton’s (Section 2.4).
58 CHAPTER 4. QUADRATIC FORMS
1 → {±1} → H1 × H1 → SO(4) → 1
Proposition 4.5.17. With notation as in 4.5.15, the left action (4.5.16) induces exact
sequences
1 → 𝐹 × → 𝐵× × 𝐵× → GSO(nrd) (𝐹) → 1
(4.5.18)
𝑎 ↦→ (𝑎, 𝑎)
and
is exact.
Proof. For the first statement, we first show that the kernel of the action is the diagonally
embedded 𝐹 × . Suppose that 𝛼𝑣 𝛽−1 = 𝑣 for all 𝑣 ∈ 𝐵; taking 𝑣 = 1 shows 𝛽 = 𝛼, and
then we conclude that 𝛼𝑣 = 𝑣𝛼 for all 𝑣 ∈ 𝐵 so 𝛼 ∈ 𝑍 (𝐵) = 𝐹.
Next, the map 𝐵× × 𝐵× → GSO(nrd) (𝐹) is surjective. If 𝑓 ∈ GSO(nrd) (𝐹)
then nrd( 𝑓 (𝑥)) = 𝑢 nrd(𝑥) for all 𝑥 ∈ 𝐵, so in particular 𝑢 ∈ nrd(𝐵× ). Every such
similitude factor occurs, since the similitude factor of (𝛼, 1) is nrd(𝛼). So it suffices
to show that the map
𝜏𝑥 (𝑣) = −𝑥𝑣 𝑥 −1 .
4.5. SPECIAL ORTHOGONAL GROUPS 59
Example 4.5.19. When 𝐵 = M2 (𝐹), then nrd(𝐵× ) = det(GL2 (𝐹)) = 𝐹 × , giving the
exact sequence
Exercises
Let 𝐹 be a field with char 𝐹 ≠ 2.
Δ: 𝐵 → 𝐹
Δ(𝛼) = trd(𝛼) 2 − 4 nrd(𝛼)
is a quadratic form.
60 CHAPTER 4. QUADRATIC FORMS
𝑉 → Hom𝐹 (𝑉, 𝐹)
𝑥 ↦→ (𝑦 ↦→ 𝑇 (𝑥, 𝑦))
(c) Show that if 𝐾 ⊇ 𝐹 is a inseparable field extension of finite degree, then the
trace form on 𝐾 (as an 𝐹-algebra) is identically zero. On the other hand,
show that if 𝐾/𝐹 is a finite separable field extension (with char 𝐹 ≠ 2)
then the trace form is nondegenerate.
√
(d) Compute the trace form on Q( 5) and Q(𝛼) where 𝛼 = 2 cos(2𝜋/7), so
that 𝛼3 + 𝛼2 − 2𝛼 − 1 = 0.
I 10. Let 𝑄 : 𝑉 → 𝐹 and 𝑄 0 : 𝑉 0 → 𝐹 be quadratic forms over 𝐹 with dim𝐹 𝑉 =
dim𝐹 𝑉 0 = 𝑛 < ∞, and let 𝑇, 𝑇 0 be the associated bilinear forms. Suppose
that there is a similarity 𝑄 ∼ 𝑄 0 with similitude factor 𝑢 ∈ 𝐹 × . Show that
det 𝑇 0 = 𝑢 𝑛 det 𝑇 ∈ 𝐹/𝐹 ×2 .
11. Let 𝑄 : 𝑉 → 𝐹 be a nondegenerate quadratic form with dim𝐹 𝑉 = 𝑛 < ∞.
Let
𝑥 𝑦
M2 (R) 0 = {𝑣 ∈ M2 (R) : tr(𝑣) = 0} = : 𝑥, 𝑦, 𝑧 ∈ R .
𝑧 −𝑥
𝜏𝑣 : 𝑉 → 𝑉
𝑇 (𝑣, 𝑥)
𝜏𝑣 (𝑥) = 𝑥 − 𝑣.
𝑄(𝑣)
Observe that 𝜏𝑣 is 𝐹-linear, and then show that 𝜏𝑣 ∈ O(𝑉) with det 𝜏𝑣 = −1.
[Hint: extend 𝑣 to a basis of the orthogonal complement of 𝑉.] Why is 𝜏𝑣
called a reflection?
(b) If 𝑥, 𝑦 ∈ 𝑉 are anisotropic with 𝑄(𝑥) = 𝑄(𝑦), show that there exists
𝑓 ∈ O(𝑉) such that 𝑓 (𝑥) = 𝑦. [Hint: reflect along either 𝑣 = 𝑥 + 𝑦 or
𝑣 = 𝑥 − 𝑦 as at least one is anisotropic, in the former case postcomposing
with reflection along 𝑥.]
(c) Let 𝑄 0 : 𝑉 0 → 𝐹 be another quadratic form, and let 𝑓 : 𝑉 −→ ∼ 𝑉 0 be an
⊥
isometry. For 𝑊 ⊆ 𝑉, show that 𝑓 (𝑊 ) = 𝑓 (𝑊) . ⊥
(d) Prove Theorem 4.2.22(a). [Hint: reduce to the case where dim𝐹 𝑊1 =
dim𝐹 𝑊10 = 1; apply parts (b) and (c).]
(e) Prove Theorem 4.2.22(b). [Hint: compare the isometry 𝑉 ' 𝑉 0 with the
isometry 𝑔.]
I 17. Prove the following weakened version of the Cartan–Dieudonné theorem (The-
orem 4.5.7): Let (𝑉, 𝑄) be a nondegenerate quadratic space with dim𝐹 𝑉 = 𝑛.
Show that every isometry 𝑓 ∈ O(𝑄) (𝐹) is a product of at most 2𝑛−1 reflections.
[Hint: in the proof of Exercise 4.16(b), note that 𝑓 can be taken to be a product
of at most 2 reflections, and finish by induction.]
Chapter 5
𝐵0 := {𝛼 ∈ 𝐵 : trd(𝛼) = 0}
63
64 CHAPTER 5. TERNARY QUADRATIC FORMS
The map 𝐵 ↦→ nrd | 𝐵0 in Theorem 5.1.1 has inverse defined by the even Clifford
algebra (see section 5.3). The similarity class of a nondegenerate ternary quadratic
form cuts out a well-defined plane conic 𝐶 ⊆ P2 over 𝐹, so one also has a bijection
between isomorphism classes of quaternion algebras over 𝐹 and isomorphism classes
of conics over 𝐹. Finally, keeping track of an orientation allows one to fully upgrade
this bijection to an equivalence of categories (Theorem 5.6.8).
The classification of quaternion algebras over 𝐹 is now rephrased in terms of
quadratic forms, and a more detailed description depends on the field 𝐹. In this vein,
the most basic question we can ask about a quaternion algebra 𝐵 is if it is isomorphic
to the matrix ring 𝐵 ' M2 (𝐹): if so, we say that 𝐵 is split over 𝐹. For example, every
quaternion algebra over C (or an algebraically closed field) is split, and a quaternion
algebra (𝑎, 𝑏 | R) is split if and only if 𝑎 > 0 or 𝑏 > 0.
Ultimately, we will identify six equivalent ways (Main Theorem 5.4.4) to check if
a quaternion algebra 𝐵 is split; in light of Theorem 5.1.1, we isolate the following.
Proposition 5.1.2. 𝐵 is split if and only if the quadratic form nrd | 𝐵0 represents 0
nontrivially.
5.2.2. Recalling (3.3.5), we have the 𝐹-vector space of pure (trace 0) elements of 𝐵
given by 𝐵0 = {1}⊥ . The standard involution restricted to 𝐵0 is given by 𝛼 = −𝛼 for
𝛼 ∈ 𝐵0 , so equivalently 𝐵0 is the −1-eigenspace for . We have 𝐵0 = 𝐹𝑖 ⊕ 𝐹 𝑗 ⊕ 𝐹𝑖 𝑗
and in this basis
nrd | 𝐵0 ' h−𝑎, −𝑏, 𝑎𝑏i (5.2.3)
so that disc(nrd | 𝐵0 ) = (𝑎𝑏) 2 = 1 ∈ 𝐹 × /𝐹 ×2 (cf. Example 4.3.8).
Proposition 5.2.4. Let 𝐵, 𝐵 0 be quaternion algebras over 𝐹. Then the following are
equivalent:
Proof. We follow Lam [Lam2005, Theorem III.2.5]. The equivalence (i) ⇔ (ii)
follows from postcomposing with the standard involution : 𝐵 0 −→ ∼ (𝐵 0) op .
The implication (i) ⇒ (iii) follows from the fact that the standard involution on an
algebra is unique and the reduced norm is determined by this standard involution, so
the reduced norm on 𝐵 is identified with the reduced norm on 𝐵 0.
The implication (iii) ⇒ (iv) follows from Witt cancellation (Theorem 4.2.22); and
(iv) ⇒ (iii) is immediate, since 𝐵 = h1i 𝐵0 and 𝐵 0 = h1i (𝐵 0) 0 so the isometry
extends by mapping 1 ↦→ 1. (Or use Witt extension, Theorem 4.2.22(b).)
So finally we prove (iv) ⇒ (i). Let 𝑓 : 𝐵0 → (𝐵 0) 0 be an isometry of quadratic
spaces. Suppose 𝐵 ' (𝑎, 𝑏 | 𝐹). Since 𝑓 is an isometry, nrd( 𝑓 (𝑖)) = nrd(𝑖) = −𝑎 and
Main Theorem 5.2.5. Let 𝐹 be a field with char 𝐹 ≠ 2. Then the functor 𝐵 ↦→ nrd | 𝐵0
yields an equivalence of categories between
and
By the expression functorial with respect to 𝐹, we mean that this bijection respects
(is compatible with) field extensions: explicitly, if 𝐹 ↩→ 𝐾 is an inclusion of fields,
and 𝐵 is a quaternion algebra with associated ternary quadratic form 𝑄 : 𝐵0 → 𝐹,
then the quaternion algebra 𝐵 𝐾 = 𝐵 ⊗𝐹 𝐾 has associated ternary quadratic form
𝑄 𝐾 : 𝐵0𝐾 = 𝐵0 ⊗𝐹 𝐾 → 𝐾.
(𝐵 𝐾 ) 0 = (𝐵 ⊗𝐹 𝐾) 0 = 𝐵0 ⊗𝐹 𝐾
Remark 5.2.7. We will refine Main Theorem 5.2.5 in section 5.6 by restricting the
isometries to those that preserve orientation.
5.3. CLIFFORD ALGEBRAS 67
𝑉
𝜄 / Clf 𝑄
𝜙
!
𝜄𝐴
𝐴
commutes.
The pair (Clf 𝑄, 𝜄) is unique up to unique isomorphism.
The algebra Clf 𝑄 in Proposition 5.3.1 is called the Clifford algebra of 𝑄.
Proof. Let
∞
Ê
Ten 𝑉 := 𝑉 ⊗𝑑 (5.3.2)
𝑑=0
where
𝑉 ⊗𝑑 := 𝑉 ⊗ · · · ⊗ 𝑉 and 𝑉 ⊗0 := 𝐹,
| {z }
𝑑
so that
Ten 𝑉 = 𝐹 ⊕ 𝑉 ⊕ (𝑉 ⊗ 𝑉) ⊕ . . . .
Then Ten 𝑉 has a multiplication given by tensor product: for 𝑥 ∈ 𝑉 ⊗𝑑 and 𝑦 ∈ 𝑉 ⊗𝑒 we
define
𝑥 · 𝑦 = 𝑥 ⊗ 𝑦 ∈ 𝑉 ⊗ (𝑑+𝑒)
(concatenate, and possibly distribute, tensors). In this manner, Ten 𝑉 has the structure
of an 𝐹-algebra, and we call Ten 𝑉 the tensor algebra of 𝑉.
Let
𝐼 (𝑄) = h𝑥 ⊗ 𝑥 − 𝑄(𝑥) : 𝑥 ∈ 𝑉i ⊆ Ten 𝑉 (5.3.3)
68 CHAPTER 5. TERNARY QUADRATIC FORMS
be the two-sided ideal generated the elements 𝑥 ⊗ 𝑥 − 𝑄(𝑥) for all 𝑥 ∈ 𝑉. Let
Clf 𝑄 = 𝐹 ⊕ 𝐹𝑒 1 ⊕ 𝐹𝑒 2 ⊕ 𝐹𝑒 1 𝑒 2 (5.3.13)
As it will cause no confusion, we may identify 𝑉 with its image 𝜄(𝑉) ↩→ Clf 𝑄.
5.3.15. The reversal map, given by
on pure tensors (and extended 𝐹-linearly) is well-defined, as it maps the ideal 𝐼 (𝑄) to
itself, and so it defines an involution on Clf 𝑄 that we call the reversal involution.
Lemma 5.3.17. The association 𝑄 ↦→ Clf 𝑄 induces a faithful functor from the
category of
quadratic forms over 𝐹, under isometries
to the category of
finite-dimensional 𝐹-algebras with involution, under isomorphisms.
so 𝑓 also induces an 𝐹-algebra map Clf 𝑄 → Clf(𝑄 0). Repeating with the inverse
map, and applying the universal property, we see that these maps are inverse, so define
isomorphisms. The functor is faithful because 𝑉 ⊂ Clf 𝑄, so if 𝑓 : 𝑉 −→ ∼ 𝑉 acts as the
identity on Clf 𝑄 then it acts as the identity on 𝑉, so 𝑓 itself is the identity. (This can
be rephrased in terms of the universal property: see Exercise 5.13.)
5.3.18. The tensor algebra Ten 𝑉 has a natural Z ≥0 grading by degree, and by con-
struction (5.3.4), the quotient Clf 𝑄 = Ten 𝑉/𝐼 (𝑄) retains a Z/2Z-grading
where Clf 0 𝑄 ⊆ Clf 𝑄 is the 𝐹-subalgebra of terms of even degree and Clf 1 𝑄 the
Clf 0 𝑄-bimodule of terms with odd degree. The reversal involution 5.3.15 preserves
Clf 0 𝑄 and so descends to an involution on Clf 0 𝑄.
70 CHAPTER 5. TERNARY QUADRATIC FORMS
We call Clf 0 𝑄 the even Clifford algebra and Clf 1 𝑄 the odd Clifford bimodule
of 𝑄. The former admits the following direct construction: let
∞
Ê
Ten0 𝑉 := 𝑉 ⊗2𝑑
𝑑=0
and let 𝐼 0 (𝑄) := 𝐼 (𝑄) ∩ Ten0 𝑉; then Clf 0 𝑄 ' Ten0 𝑉/𝐼 0 (𝑄).
Example 5.3.20. Continuing Example 5.3.12, we see that the reversal involution fixes
𝑖, 𝑗 and acts as the standard involution on Clf 0 𝑄. So the algebra Clf 𝑄 is not just a
quaternion algebra, but one retaining a Z/2Z-grading.
Lemma 5.3.21. The association 𝑄 ↦→ Clf 0 𝑄 defines a functor from the category of
to the category of
Ten0 𝑉 → Ten0 (𝑉 0)
𝑥 1 ⊗ · · · ⊗ 𝑥 𝑑 ↦→ (𝑢 −1 ) 𝑑/2 𝑓 (𝑥 1 ) ⊗ · · · ⊗ 𝑓 (𝑥 𝑑 ).
so 𝐼 0 (𝑄) maps to 𝐼 0 (𝑄 0), and the induced map Clf 0 𝑄 → Clf 0 (𝑄 0) is an 𝐹-algebra
isomorphism.
5.3.22. Note that unlike the Clifford functor, the even Clifford functor need not be
faithful: for example, the map −1 : 𝐹 2 → 𝐹 2 has 𝑒 1 𝑒 2 ↦→ (−𝑒 1 ) (−𝑒 2 ) = 𝑒 1 𝑒 2 so acts
by the identity on Clf 0 𝑄.
5.3.23. Suppose that char 𝐹 ≠ 2 and let 𝑄(𝑥) = h𝑎, 𝑏, 𝑐i be a nondegenerate ternary
quadratic form. Then the even Clifford algebra Clf 0 𝑄 is given by
Clf 0 𝑄 = 𝐹 ⊕ 𝐹𝑖 ⊕ 𝐹 𝑗 ⊕ 𝐹𝑖 𝑗
5.4. SPLITTING 71
𝑖 2 = −𝑎𝑏, 𝑗 2 = −𝑏𝑐, 𝑖 𝑗 + 𝑗𝑖 = 0.
So
0 −𝑎𝑏, −𝑏𝑐
Clf 𝑄 ' .
𝐹
Letting 𝑘 = 𝑒 3 𝑒 1 , we obtain symmetrically with the other two pairs of generators 𝑗, 𝑘
or 𝑘, 𝑖 that
0 −𝑏𝑐, −𝑎𝑐 −𝑎𝑐, −𝑎𝑏
Clf 𝑄 ' ' .
𝐹 𝐹
−𝑎𝑏, 𝑎𝑏 2
0 𝑎, 𝑏
Clf 𝑄 ' ' . (5.3.24)
𝐹 𝐹
Remark 5.3.25. The even Clifford map does not furnish an equivalence of categories
for the same reason as in 5.3.22; one way to deal with this issue is to restrict the
isometries to those that preserve orientation: we carry this out in section 5.6.
5.4 Splitting
The moral of Main Theorem 5.2.5 is that the problem of classifying quaternion algebras
depends on the theory of ternary quadratic forms over that field (and vice versa). We
now pursue the first consequence of this moral, and we characterize the matrix ring
among quaternion algebras. Suppose that char 𝐹 ≠ 2, but still 𝑄 : 𝑉 → 𝐹 a quadratic
form with dim𝐹 𝑉 < ∞.
A hyperbolic plane 𝐻 is universal, its associated bilinear form has Gram matrix
0 1
in the standard basis, and 𝐻 has normalized form 𝐻 ' h1, −1i.
1 0
Proof. For the implication (⇐), we have an isotropic vector from either one of the
two basis vectors. For the implication (⇒), let 𝑥 ∈ 𝑉 be isotropic, so 𝑥 ≠ 0 and
satisfy 𝑄(𝑥) = 0. Since 𝑄 is nondegenerate, there exists 𝑦 ∈ 𝑉 such that 𝑇 (𝑥, 𝑦) ≠ 0;
rescaling 𝑦, we may assume 𝑇 (𝑥, 𝑦) = 1. If 𝑇 (𝑦, 𝑦) = 2𝑄(𝑦, 𝑦) ≠ 0, replacing
𝑦 ← 𝑦 − 2𝑥/𝑇 (𝑦, 𝑦) gives 𝑇 (𝑦, 𝑦) = 0. Thus 𝑄 restricted to 𝐹𝑥 + 𝐹 𝑦 is isometric
to 𝐻, and in particular is nondegenerate. Therefore letting 𝑉 0 := (𝐹𝑥 + 𝐹 𝑦) ⊥ and
𝑄 0 := 𝑄| 𝑉 0 , we have 𝑉 ' (𝐹𝑥 + 𝐹 𝑦) 𝑉 0 and 𝑄 ' 𝐻 𝑄 0.
Lemma 5.4.3. Suppose 𝑄 is nondegenerate and let 𝑎 ∈ 𝐹 × . Then the following are
equivalent:
(i) 𝑄 represents 𝑎;
(ii) 𝑄 ' h𝑎i 𝑄 0 for some nondegenerate form 𝑄 0; and
(iii) h−𝑎i 𝑄 is isotropic.
Proof. For (i) ⇒ (ii), we take 𝑄 0 = 𝑄| 𝑊 and 𝑊 = {𝑣}⊥ ⊂ 𝑉 where 𝑄(𝑣) = 𝑎. For (ii)
⇒ (iii), we note that h−𝑎i 𝑄 ' h𝑎, −𝑎i 𝑄 0 is isotropic. For (iii) ⇒ (i), suppose
(h−𝑎i 𝑄) (𝑣) = 0, so 𝑄(𝑣) = 𝑎𝑥 2 for some 𝑥 ∈ 𝐹. If 𝑥 = 0, then 𝑄 is isotropic and
by Lemma 5.4.2 represents 𝑎; if 𝑥 ≠ 0, then by homogeneity 𝑄(𝑣/𝑥) = 𝑎 and again 𝑄
represents 𝑎.
on the circumstances, one of these formulations may be more natural than the other.)
Proof. We follow Lam [Lam2005, Theorem 2.7]. The isomorphism (1, 1 | 𝐹) '
M2 (𝐹) in (i) follows from Example 2.2.4. The implication (i) ⇒ (ii) is clear. The
equivalence (ii) ⇔ (iii) follows from the fact that 𝛼 ∈ 𝐵× if and only if nrd(𝛼) ∈ 𝐹 ×
(Exercise 3.5).
We now prove (iii) ⇒ (iv). Let 0 ≠ 𝛼 ∈ 𝐵 be such that nrd(𝛼) = 0. If trd(𝛼) = 0,
then we are done. Otherwise, trd(𝛼) ≠ 0. Let 𝛽 be orthogonal to 1, 𝛼, so that
trd(𝛼𝛽) = 0. We cannot have both 𝛼𝛽 = 0 and 𝛼𝛽 = (trd(𝛼) − 𝛼) 𝛽 = 0, so we may
suppose 𝛼𝛽 ≠ 0. But then nrd(𝛼𝛽) = nrd(𝛼) nrd(𝛽) = 0 as desired.
5.4. SPLITTING 73
To complete the equivalence of the first four we prove (iv) ⇒ (i). Let 𝛽 ∈ 𝐵0
satisfy nrd(𝛽) = 0. Since nrd | 𝐵0 is nondegenerate, there exists 0 ≠ 𝛼 ∈ 𝐵0 such
that trd(𝛼𝛽) ≠ 0. Therefore, the restriction of nrd to 𝐹𝛼 ⊕ 𝐹 𝛽 is nondegenerate
and isotropic. By Lemma 5.4.2, we conclude there exists a basis for 𝐵0 such that
nrd | 𝐵0 ' h1, −1i h𝑐i = h1, −1, 𝑐i; but disc(nrd | 𝐵0 ) = −𝑐 ∈ 𝐹 ×2 by 5.2.2; rescaling,
we may suppose 𝑐 = −1. But then by Proposition 5.2.4 we have 𝐵 ' (1, 1 | 𝐹).
Now we show (iv) ⇒ (v). For 𝛼 ∈ 𝐵0 ,
In the equivalence (vi) ⇔ (vi0), the two statements are identical if 𝑎 ∉ 𝐹 ×2 and
both automatically satisfied if 𝑎 ∈ 𝐹 ×2 .
To conclude, we prove (vi) ⇒ (iii). If 𝑏 = 𝑥 2 − 𝑎𝑦 2 ∈ Nm𝐾 |𝐹 (𝐾 × ), then 𝛼 =
𝑥 + 𝑦𝑖 + 𝑗 ≠ 0 has nrd(𝛼) = 𝑥 2 − 𝑎𝑦 2 − 𝑏 = 0.
Example 5.4.6. The fundamental example of a splitting field for a quaternion algebra
is that C splits the real Hamiltonians H: we have H ⊗R C ' M2 (C) as in (2.4.1).
have 𝐵 ⊗𝐹 𝐾 ' M2 (𝐾) if and only if h−𝑎, −𝑏, 𝑎𝑏i is isotropic over 𝐾, which is to say
there exist 𝑥, 𝑦, 𝑧, 𝑢, 𝑣, 𝑤 ∈ 𝐹 such that
√ √ √
−𝑎(𝑥 + 𝑢 𝑑) 2 − 𝑏(𝑦 + 𝑣 𝑑) 2 + 𝑎𝑏(𝑧 + 𝑤 𝑑) 2 = 0. (5.4.8)
Example 5.4.10. Let 𝑝 be an odd prime and let 𝑎 be a quadratic nonresidue modulo
𝑎, 𝑝
𝑝. We claim that is a division quaternion algebra over Q. By Main Theorem
Q
5.4.4, it suffices to show that the quadratic form h1, −𝑎, −𝑝, 𝑎 𝑝i is anisotropic. So
suppose that 𝑡 2 − 𝑎𝑥 2 = 𝑝(𝑦 2 − 𝑎𝑧2 ) with 𝑡, 𝑥, 𝑦, 𝑧 ∈ Q not all zero. The equation is
homogeneous, so we can multiply through by a common denominator and suppose
that 𝑡, 𝑥, 𝑦, 𝑧 ∈ Z with gcd(𝑡, 𝑥, 𝑦, 𝑧) = 1. Reducing modulo 𝑝 we find 𝑡 2 ≡ 𝑎𝑥 2
(mod 𝑝); since 𝑎 is a quadratic nonresidue, we must have 𝑡 ≡ 𝑥 ≡ 0 (mod 𝑝).
Plugging back in and cancelling a factor of 𝑝 we find 𝑦 2 ≡ 𝑎𝑧 2 ≡ 0 (mod 𝑝), and again
𝑦 ≡ 𝑧 ≡ 0 (mod 𝑝), a contradiction.
If we identify
P(𝐵0 ) := (𝐵0 r {0})/𝐹 × ' P2 (𝐹)
with (the points of) the projective plane over 𝐹, then the vanishing locus 𝐶 = 𝑉 (𝑄) of
𝑄 = nrd | 𝐵0 defines a conic over 𝐹: if we take the basis 𝑖, 𝑗, 𝑖 𝑗 for 𝐵0 , then the conic
𝐶 is defined by the vanishing of the equation
By Lemma 5.4.7, a quadratic field 𝐾 over 𝐹 embeds in 𝐵 if and only if the ternary
quadratic form nrd | 𝐵0 represents 0 over 𝐾. We can also rephrase this in terms of the
values represented by nrd | 𝐵0 .
Remark 5.5.5. Two conics over 𝐹 are isomorphic (as plane curves) if and only if their
function fields are isomorphic (Exercise 5.23).
5.6 Orientations
To conclude, we show that the notion of orientation underlying the definition of special
isometries (as in Example 4.5.5) extends more generally to isometries between two
different quadratic spaces by keeping track of one bit of extra information, refining
Main Theorem 5.2.5. We follow Knus–Murkurjev–Rost–Tignol [KMRT98, Theorem
15.2]. We retain our hypothesis that char 𝐹 ≠ 2.
Let 𝑄 : 𝑉 → 𝐹 be a quadratic space with dim𝐹 𝑉 = 𝑛 odd.
The signed discriminant gives a simpler statement; one could equally well work
with the usual discriminant and keep track of the sign.
the resulting group is independent of the choice, and we recover the same group as in
Definition 4.5.1.
𝑎, 𝑏
5.6.7. Let 𝐵 = be a quaternion algebra over 𝐹. In previous sections, we took
𝐹
nrd | 𝐵0 : 𝐵0 → 𝐹, a nondegenerate ternary quadratic space of discriminant 1. Since
we are working with the signed discriminant, we take instead − nrd | 𝐵0 : 𝐵0 → 𝐹 with
sgndisc(− nrd | 𝐵0 ) = 1; this map has a nice description as the squaring map, since
𝛼2 = − nrd(𝛼) for 𝛼 ∈ 𝐵0 .
We claim that 𝐵0 has a canonical orientation. We have an inclusion 𝜄 : 𝐵0 ↩→ 𝐵
with 𝜄(𝑥) 2 = − nrd(𝑥) for all 𝑥 ∈ 𝐵0 . By the universal property of Clifford algebras,
we get an 𝐹-algebra homomorphism 𝜙 : Clf (𝐵0 ) → 𝐵. We see that 𝜙 is surjective
so it induces an 𝐹-algebra map 𝜋 : 𝑍 (Clf(𝐵0 )) → 𝑍 (𝐵) = 𝐹 (Exercise 2.8). This
defines a unique orientation 𝜁 𝐵 = 𝜁 with 𝜁 − 1 ∈ ker 𝜋, by 5.6.4.
Explicitly, let 𝑖, 𝑗, 𝑘 be the standard basis for 𝐵 with 𝑘 = 𝑖 𝑗. Then nrd(𝑘) = 𝑎𝑏,
and 𝑖, 𝑗, 𝑘 is a basis for 𝐵0 . Let 𝜁 = 𝑖 𝑗 𝑘 −1 = −𝑖 𝑗 𝑘/(𝑎𝑏) ∈ 𝑍 (Clf(𝐵0 )). Then
5.6. ORIENTATIONS 77
(𝑄, 𝜁) ↦→ Clf 0 𝑄
(− nrd | 𝐵0 , 𝜁 𝐵 ) ← 𝐵
and
−𝑎𝑏, 𝑎𝑏 2
Clf 0 (− nrd | 𝐵0 ) = Clf 0 (h𝑎, 𝑏, −𝑎𝑏i) =
𝐹
−𝑎𝑏, 𝑎𝑏 2
𝑎, 𝑏
→
𝐹 𝐹
𝑖 0 , 𝑗0 ↦→ 𝑖 𝑗, 𝑗 𝑘.
∼ 𝐵 yields a natural isomor-
Therefore, the canonical isomorphism Clf 0 (− nrd | 𝐵0 ) −→
phism between these composed functors and the identity functor, giving an equivalence
of categories.
Conversely, let (𝑄, 𝜁) be an oriented ternary quadratic space, let 𝐵 = Clf 0 𝑄, and
consider (− nrd | 𝐵0 , 𝜁 𝐵 ). We define a natural oriented isometry between these two
spaces. We have a natural inclusion 𝑉 ↩→ Clf 𝑄, and we define the linear map
𝑚𝜁 : 𝑉 → 𝐵
𝑣 ↦→ 𝑣𝜁;
𝜖𝑚 𝜁 (𝑒 1 ) = 𝑒 1 (𝑒 1 𝑒 2 𝑒 3 ) = 𝑎𝑒 2 𝑒 3 = 𝑎 𝑗
𝜖𝑚 𝜁 (𝑒 2 ) = 𝑒 2 (𝑒 1 𝑒 2 𝑒 3 ) = −𝑏𝑒 1 𝑒 3 = −𝑘 (5.6.9)
𝜖𝑚 𝜁 (𝑒 3 ) = 𝑒 3 (𝑒 1 𝑒 2 𝑒 3 ) = 𝑐𝑖;
𝑚 𝜁 (𝜁) = 𝑚 𝜁 (𝜖 𝑒 1 𝑒 2 𝑒 3 ) = 𝜖 (𝑒 1 𝜁) (𝑒 2 𝜁) (𝑒 3 𝜁)
= 𝜖 (𝜖 𝑎 𝑗) (−𝜖 𝑘) (𝜖 𝑐𝑖) = (−𝑎𝑐) (𝑖 𝑗 𝑘) = (−𝑎𝑏𝑐)𝑖 𝑗 𝑘 −1 = 𝑖 𝑗 𝑘 −1 = 𝜁 𝐵 .
This natural oriented isometry gives a natural transformation between these composed
functors and the identity functor, and the statement follows.
and recover the same group as in 4.5.4. If 𝑛 is odd, we declare that every similarity is
oriented and let GSO(𝑄) (𝐹) := GO(𝑄) (𝐹).
5.6. ORIENTATIONS 79
Exercises
Throughout, let 𝐹 be a field with char 𝐹 ≠ 2.
1. Let 𝐵, 𝐵 0 be quaternion algebras over 𝐹. Show that if the quadratic forms nrd 𝐵
and nrd 𝐵0 are similar, then they are isometric.
2. Consider the hyperbolic quaternions 𝐻Mac of Macfarlane (1.2.1).
(a) Show that 𝐻Mac is the Clifford algebra of h1, 1, 1i over R.√
(b) Show that 𝐻Mac is isomorphic as an algebra over√C = R( −1) to the even
Clifford algebra of the ternary quadratic form − −1h1, 1, 1i.
3. Prove the implication (vi) ⇒ (v) of Main Theorem 5.4.4 directly.
4. Use Main Theorem 5.4.4(vi) to give another proof that there is no division
quaternion algebra 𝐵 over a finite field 𝐹 = F𝑞 (with 𝑞 odd).
5. (a) Show that the quadratic form 𝑄(𝑥, 𝑦, 𝑧) = 𝑥 2 + 𝑦 2 + 𝑧2 is isotropic over
F 𝑝 for all odd primes 𝑝. Conclude that (−1, −1 | F 𝑝 ) ' M2 (F 𝑝 ). [Hint:
count squares and nonsquares.]
(b) More generally, show that every ternary quadratic form over a finite field
F𝑞 (with 𝑞 odd) is isotropic. [Hint: Reduce to the case of finding a solution
to 𝑦 2 = 𝑓 (𝑥) where 𝑓 is a polynomial of degree 2.] Use Main Theorem
5.4.4(iv) to give yet another proof that there is no division quaternion
algebra 𝐵 over F𝑞 .
(c) Show that over a finite field F𝑞 with 𝑞 odd, there is a unique anisotropic
binary quadratic form up to isometry.
6. Show that if 𝑄 : 𝐹 → 𝐹 is the quadratic form 𝑄(𝑥) = 𝑎𝑥 2 with 𝑎 ∈ 𝐹, then
Clf (𝐹) ' 𝐹 [𝑥]/(𝑥 2 − 𝑎).
−1, 26
7. Show that (−1, 26)Q = 1, i.e., ' M2 (Q).
Q
−1, 𝑝
8. Let 𝑝 be prime. Show that ' M2 (Q) if and only if 𝑝 = 2 or 𝑝 ≡ 1
Q
(mod 4).
9. Show that
−2, −3 −1, −1 −2, −5 −1, −1
' but that ; .
Q Q Q Q
10. Let 𝐵 = (𝑎, 𝑏 | 𝐹) be a quaternion algebra over 𝐹. Give a constructive
(algorithmic) proof of the implication (iv) ⇒ (i) in Main Theorem 5.4.4, as
follows.
Let 𝜖 = 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ∈ 𝐵 satisfy nrd(𝜖) = −𝑎𝑥 2 − 𝑏𝑦 2 + 𝑎𝑏𝑧2 = −𝜖 2 = 0.
(a) Show that there exists 𝑘 ∈ {𝑖, 𝑗, 𝑖 𝑗 } such that trd(𝜖 𝑘) = 𝑠 ≠ 0.
(b) Let 𝑡 := trd(𝑘) and 𝑛 := nrd(𝑘), and let 𝜖 0 := 𝑠−1 𝜖. Let
𝑖 0 := 𝜖 0 𝑘 − (𝑘 + 𝑡)𝜖 0
𝑗 0 := 𝑘 + (−𝑡𝑘 + 𝑛 + 1)𝜖 0 .
80 CHAPTER 5. TERNARY QUADRATIC FORMS
(a) Let 𝐶 0 = Clf 0 𝑄 be the even Clifford algebra of the reduced norm 𝑄. Show
that 𝑍 (𝐶 0 ) ' 𝐹 × 𝐹. [Hint: 𝑍 (𝐶 0 ) is generated by 𝑒 0 𝑒 1 𝑒 2 𝑒 3 .]
(b) Show that 𝐶 0 ' 𝐵 × 𝐵op (' 𝐵 × 𝐵) as 𝐹-algebras.
(c) Prove that if 𝐵 0 is a quaternion algebra over 𝐹 then 𝐵 ' 𝐵 0 are isomorphic
as 𝐹-algebras if and only if the reduced norms 𝑄 ∼ 𝑄 0 are similar as
quadratic spaces.
12. Let 𝑄 : 𝑉 → 𝐹 be a nondegenerate quadratic form. Show that the reversal map
: Clf 0 𝑄 → Clf 0 𝑄 on the Clifford algebra has the property that 𝑥𝑥 ∈ 𝐹 for all
pure tensors 𝑥 = 𝑒 1 𝑒 2 · · · 𝑒 𝑑 , but defines a standard involution on Clf 𝑄 if and
only if 𝑉 = {0} and on Clf 0 𝑄 if and only if dim𝐹 𝑉 ≤ 3.
13. Give another proof of Lemma 5.3.17 using the universal property of the Clifford
algebra.
I 14. Expand (5.4.8) and prove as a consequence that if 𝛼 = 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 and
𝛽 = 𝑢𝑖 + 𝑣 𝑗 + 𝑤𝑖 𝑗, then trd(𝛼𝛽) = 0 (so 𝛼 is orthogonal to 𝛽) and moreover
nrd(𝛼) + 𝑑 nrd(𝛽) = 0.
15. Show that the Hilbert symbol is Galois equivariant, in the following sense: for all
field automorphisms 𝜎 ∈ Aut(𝐹) and all 𝑎, 𝑏 ∈ 𝐹 × , we have (𝜎(𝑎), 𝜎(𝑏)) 𝐹 =
(𝑎, 𝑏) 𝐹 .
16. Let 𝑎, 𝑏, 𝑏 0 ∈ 𝐹 × . Show that there exists an 𝐹-linear isomorphism 𝜙 : (𝑎, 𝑏 |
∼ (𝑎, 𝑏 0 | 𝐹) with 𝜙(𝑖) = 𝑖 0 if and only if 𝑏/𝑏 0 ∈ Nm
𝐹) −→ ×
√ 𝐾 |𝐹 (𝐾 ) where
𝐾 = 𝐹 ( 𝑎). [More generally, see Corollary 7.7.6.]
17. Let 𝑎 ∈ Q× r Q×2 . Show that there are infinitely many distinct isomorphism
classes of conics 𝑥 2 − 𝑎𝑦 2 = 𝑏𝑧 2 for 𝑏 ∈ Q× .
18. Let 𝐾 = 𝐹 (𝑎, 𝑏) with 𝑎, 𝑏 algebraically independent,
transcendental elements.
𝑎, 𝑏
Show that the generic quaternion algebra is a division algebra. [Hint:
𝐾
show the associated ternary quadratic form is anisotropic.]
I 19. Prove Lemma 5.6.1 for general odd 𝑛 as follows.
20. Let 𝑄 : 𝑉 → 𝐹 be a quadratic form. Show that the even Clifford algebra Clf 0 𝑄
with its map 𝜄 : 𝑉 ⊗ 𝑉 → Clf 0 𝑄 has the following universal property: if 𝐴 is an
𝐹-algebra and 𝜄 𝐴 : 𝑉 ⊗ 𝑉 → 𝐴 is an 𝐹-linear map such that
(i) 𝜄 𝐴 (𝑥 ⊗ 𝑥) = 𝑄(𝑥) for all 𝑥 ∈ 𝑉, and
(ii) 𝜄 𝐴 (𝑥 ⊗ 𝑦)𝜄 𝐴 (𝑦 ⊗ 𝑧) = 𝑄(𝑦)𝜄 𝐴 (𝑥 ⊗ 𝑧) for all 𝑥, 𝑦, 𝑧 ∈ 𝑉,
then there exists a unique 𝐹-algebra homomorphism 𝜙 : Clf 0 𝑄 → 𝐴 such that
the diagram
𝜄 /
𝑉 ⊗𝑉 Clf 0 𝑄
𝜙
$
𝜄𝐴
where Clf 1 (𝑄 1 ) ⊗ Clf 1 (𝑄 2 ) has multiplication induced from the full Clif-
ford algebras Clf (𝑄 1 ) and Clf (𝑄 2 ).
(b) Prove that there is a Clf 0 (𝑄 1 𝑄 2 )-bimodule isomorphism
∼ (Clf 0 (𝑄 ) ⊗ Clf 1 (𝑄 )) ⊕ (Clf 1 (𝑄 ) ⊗ Clf 0 (𝑄 ))
Clf 1 (𝑄 1 𝑄 2 ) −→ 1 2 1 2
23. In this exercise, we assume background in algebraic curves. Show that two
conics over 𝐹 are isomorphic (as projective plane curves) if and only if their
function fields are isomorphic. [Hint: conics are anticanonically embedded—
the restriction of 𝒪P2 (−1) to the conic is a canonical sheaf—so an isomorphism
of function fields induces an linear isomorphism of conics.]
Chapter 6
Characteristic 2
In this chapter, we extend the results from the previous four chapters to the neglected
case where the base field has characteristic 2. Throughout this chapter, let 𝐹 be a field
with algebraic closure 𝐹 al .
6.1 Separability
To get warmed up, we give a different notation (symbol) for quaternion algebras that
holds in any characteristic and which is convenient for many purposes.
Example 6.1.2. If 𝐴 ' 𝐹 [𝑥]/( 𝑓 (𝑥)) with 𝑓 (𝑥) ∈ 𝐹 [𝑥], then 𝐴 is separable if and
only if 𝑓 has distinct roots in 𝐹 al .
6.1.3. If char 𝐹 ≠ 2, and 𝐾 is a quadratic 𝐹-algebra, then after completing the square,
we see that the following are equivalent:
(i) 𝐾 is separable;
(ii) 𝐾 ' 𝐹 [𝑥]/(𝑥 2 − 𝑎) with 𝑎 ≠ 0;
(iii) 𝐾 is reduced (𝐾 has no nonzero nilpotent elements);
(iv) 𝐾 is a field or 𝐾 ' 𝐹 × 𝐹.
𝐾 ' 𝐹 [𝑥]/(𝑥 2 + 𝑥 + 𝑎)
83
84 CHAPTER 6. CHARACTERISTIC 2
the 𝐹-algebra with basis 1, 𝑗 as a left 𝐾-vector space and with the multiplication rules
𝑗 2 = 𝑏 and 𝑗 𝛼 = 𝛼 𝑗 for 𝛼 ∈ 𝐾, where is the standard involution on 𝐾 (the nontrivial
element of Gal(𝐾 | 𝐹) if 𝐾 is a field). We will also write (𝐾, 𝑏 | 𝐹) for formatting.
is a quaternion algebra over 𝐹. The point is that we cannot complete the square in
characteristic 2, so the more general notation gives a characteristic-independent way
to define quaternion algebras. In using this symbol, we are breaking the symmetry
between the standard generators 𝑖, 𝑗, but otherwise have not changed anything about
the definition.
𝑖 2 + 𝑖 = 𝑎, 𝑗 2 = 𝑏, and 𝑘 = 𝑖 𝑗 = 𝑗 (𝑖 + 1) (6.2.2)
with 𝑎 ∈ 𝐹 and 𝑏 ∈ 𝐹 × .
Just as when char 𝐹 ≠ 2, we find that the multiplication table for a quaternion
algebra 𝐵 is determined by the rules (6.2.2), e.g.
𝑗 𝑘 = 𝑗 (𝑖 𝑗) = (𝑖 𝑗 + 𝑗) 𝑗 = 𝑏𝑖 + 𝑏 = 𝑘 𝑗 + 𝑏.
𝑎, 𝑏
We denote by or [𝑎, 𝑏 | 𝐹) the 𝐹-algebra with basis 1, 𝑖, 𝑗, 𝑖 𝑗 subject to the
𝐹
𝑎, 𝑏
multiplication rules (6.2.2). The algebra is not symmetric in 𝑎, 𝑏 (explaining
𝐹
the choice of notation), but it is still functorial in the field 𝐹.
If we let 𝐾 = 𝐹 [𝑖] ' 𝐹 [𝑥]/(𝑥 2 + 𝑥 + 𝑎), then
𝑎, 𝑏 𝐾, 𝑏
'
𝐹 𝐹
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ↦→ 𝛼 = 𝑥 + 𝛼 = (𝑡 + 𝑥) + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗
since
𝛼𝛼 = (𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗) ((𝑡 + 𝑥) + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗)
(6.2.7)
= 𝑡 2 + 𝑡𝑥 + 𝑎𝑥 2 + 𝑏𝑦 2 + 𝑏𝑦𝑧 + 𝑎𝑏𝑧2 ∈ 𝐹.
Consequently, one has a reduced trace and reduced norm on 𝐵 as in Chapter 3.
Theorem 6.2.8. Let 𝐵 be a division 𝐹-algebra with a standard involution that is not
the identity. Then either 𝐵 is a separable quadratic field extension of 𝐹 or 𝐵 is a
quaternion algebra over 𝐹.
6.3.1. We begin with the definition of the discriminant. When 𝑛 is even, we sim-
ply define disc(𝑄) = det(𝑇) ∈ 𝐹/𝐹 ×2 —this is equivalent to Definition 4.3.3 when
char 𝐹 ≠ 2, having absorbed the square power of 2.
When 𝑛 is odd, the symmetric matrix 𝑇 always has determinant 0 (Exercise 6.8);
we need to “divide this by 2”. So instead we work with a generic quadratic form, as
follows. Consider the quadratic form
∑︁
𝑄 univ (𝑥1 , . . . , 𝑥 𝑛 ) := 𝑎 𝑖 𝑗 𝑥𝑖 𝑥 𝑗 = 𝑎 11 𝑥12 + 𝑎 12 𝑥1 𝑥 2 + · · · + 𝑎 𝑛𝑛 𝑥 𝑛2 (6.3.2)
1≤𝑖 ≤ 𝑗 ≤𝑛
over the field 𝐹 univ := Q(𝑎 𝑖 𝑗 )𝑖, 𝑗=1,...,𝑛 (now of characteristic zero!) with 𝑎 𝑖 𝑗 transcen-
dental elements. We compute its universal determinant
2𝑎 11 𝑎 12 ··· 𝑎 1𝑛
2𝑎 22 ···
© ª
𝑎 12 𝑎 2𝑛
det( [𝑇 univ ]) = det .
®
.. .. .. ® ∈ 2Z[𝑎 𝑖 𝑗 ] 𝑖, 𝑗 (6.3.3)
.. . ®
. . ®
« 𝑎 1𝑛 𝑎 2𝑛 ··· 2𝑎 𝑛𝑛 ¬
as a polynomial with integer coefficients. We claim all of these coefficients are even:
indeed, reducing modulo 2 and computing the determinant over F2 (𝑎 𝑖 𝑗 )𝑖, 𝑗 , we recall
that the determinant of an alternating matrix of odd size is zero (over any field).
Therefore, we may let
with 𝑎, 𝑏, 𝑐, 𝑢, 𝑣, 𝑤 ∈ 𝐹, then
in all characteristics.
Definition 6.3.6. We say 𝑄 is nondegenerate if disc(𝑄) ≠ 0.
Next, even though not every quadratic form over 𝐹 can be diagonalized, so we will
also make use of one extra form: for 𝑎, 𝑏 ∈ 𝐹, we write [𝑎, 𝑏] for the quadratic form
𝑎𝑥 2 + 𝑎𝑥𝑦 + 𝑏𝑦 2 on 𝐹 2 .
6.4. ∗ CHARACTERIZING QUATERNION ALGEBRAS 87
𝑄 ' [𝑎 1 , 𝑏 1 ] · · · [𝑎 𝑚 , 𝑏 𝑚 ] h𝑐 1 , . . . , 𝑐𝑟 i (6.3.8)
with 𝑎 𝑖 , 𝑏 𝑖 , 𝑐 𝑗 ∈ 𝐹.
Example 6.3.9. The quadratic forms 𝑄(𝑥, 𝑦, 𝑧) = 𝑥 2 +𝑦𝑧+𝑧2 and 𝑄(𝑥, 𝑦, 𝑧) = 𝑥 2 +𝑦 2 +𝑧2
are normalized over F2 , but the quadratic form 𝑄(𝑥, 𝑦, 𝑧) = 𝑥𝑧 + 𝑦𝑧 + 𝑧 2 is not.
In 𝐹/𝐹 ×2 , we have
0, if 𝑟 ≥ 2;
disc(h𝑐 1 , . . . , 𝑐𝑟 i) = 𝑐 1 , if 𝑟 = 1;
1,
if 𝑟 = 0.
Therefore, 𝑄 is nondegenerate if and only if 𝑎 1 · · · 𝑎 𝑚 𝑐 1 · · · 𝑐𝑟 ≠ 0 and 𝑟 ≤ 1.
𝑎, 𝑏
Example 6.3.11. Let 𝐵 = be a quaternion algebra. Then 1, 𝑖, 𝑗, 𝑖 𝑗 is a normal-
𝐹
ized basis for 𝐵, and by (6.2.7),
Theorem 6.4.1. Let 𝐵 be an 𝐹-algebra (with char 𝐹 = 2). Then 𝐵 has a nondegenerate
standard involution if and only if one of the following holds:
(i) 𝐵 = 𝐹;
(ii) 𝐵 = 𝐾 is a separable quadratic 𝐹-algebra; or
(iii) 𝐵 is a quaternion algebra over 𝐹.
88 CHAPTER 6. CHARACTERISTIC 2
Proof. If 𝐵 = 𝐹, then the standard involution is the identity, and nrd is nondegenerate
on 𝐹 because the reduced (half-)discriminant of the quadratic form nrd(𝑥) = 𝑥 2 is 1.
If dim𝐹 𝐵 = 2, then 𝐵 = 𝐾 has a unique standard involution (Lemma 3.4.2). By
6.1.4, we see that the involution is nondegenerate if and only if 𝐾 is separable.
So suppose dim𝐹 𝐵 > 2. Since 𝐵 has a nondegenerate standard involution, there
exists an element 𝑖 ∈ 𝐵 such that 𝑇 (𝑖, 1) = trd(𝑖) ≠ 0. We have 𝑖 ∉ 𝐹 since
trd(𝐹) = {0}. Rescaling we may suppose trd(𝑖) = 1, whence 𝑖 2 = 𝑖 + 𝑎 for some 𝑎 ∈ 𝐹,
and nrd | 𝐹 +𝐹 𝑖 = [1, 𝑎]. (We have started the proof of Lemma 6.3.7, and 1, 𝑖 is part of
a normalized basis, in this special case.)
By nondegeneracy, there exists 𝑗 ∈ {1, 𝑖}⊥ such that nrd( 𝑗) = 𝑏 ≠ 0. Thus
trd( 𝑗) = 0 so 𝑗 = 𝑗 and 𝑗 2 = 𝑏 ∈ 𝐹 × . Furthermore,
0 = trd(𝑖 𝑗) = 𝑖 𝑗 + 𝑗𝑖 = 𝑖 𝑗 + 𝑗 (𝑖 + 1)
disc(nrd | 𝐵0 ) = 𝑏 2 = 1 ∈ 𝐹 × /𝐹 ×2 . (6.4.6)
Theorem 6.4.7. Let 𝐹 be a field with char 𝐹 = 2. Then the functor 𝐵 ↦→ nrd | 𝐵0 yields
an equivalence of categories between
and
Proof. We argue as in Theorem 5.6.8 but with char 𝐹 = 2. The argument here is
easier, because all sign issues go away and there is no orientation to chase: by Exercise
6.12, there is a unique 𝜁 ∈ Clf 1 𝑄 r 𝐹 such that 𝜁 2 = 1. The inclusion 𝜄 : 𝐵0 ↩→ 𝐵
induces a surjective 𝐹-algebra homomorphism Clf 0 (nrd | 𝐵0 ) → 𝐵, so by dimensions
it is an isomorphism; this gives one natural transformation. In the other direction, the
map 𝑚 𝜁 : 𝑉 → 𝐵0 by 𝑣 ↦→ 𝑣𝜁 is again an isometry by (5.6.10), giving the other.
Here is a second direct proof. By 6.4.3, the quadratic form nrd | 𝐵0 has discriminant
1. To show the functor is essentially surjective, let 𝑄 : 𝑉 → 𝐹 be a ternary quadratic
form with discriminant 1 ∈ 𝐹 × /𝐹 ×2 . Then 𝑄 ' h𝑢i [𝑏, 𝑐] for some 𝑢, 𝑏, 𝑐 ∈ 𝐹. We
have disc(𝑄) = 𝑢𝑏 2 = 1 ∈ 𝐹 ×2 so 𝑏 ∈ 𝐹 × and 𝑢 ∈ 𝐹 ×2 . Rescaling the first variable,
we obtain 𝑄 ∼ h1i [𝑏, 𝑐]. Thus by 6.4.3, 𝑄 arises up to isometry from the quaternion
𝑎, 𝑏
algebra with 𝑎 = 𝑐𝑏 −1 .
𝐹
For morphisms, we argue as in the proof of Proposition 5.2.4 but with char 𝐹 = 2. In
one direction, an 𝐹-algebra isomorphism 𝐵 −→ ∼ 𝐵 0 induces an isometry 𝐵0 −→ ∼ (𝐵 0) 0
by uniqueness of the standard involution. Conversely, let 𝑓 : 𝐵 → (𝐵 ) 0 be an
0 0
𝑎, 𝑏
isometry. Let 𝐵 ' . Extend 𝑓 to an 𝐹-linear map 𝐵 → 𝐵 0 by mapping
𝐹
𝑖 ↦→ 𝑏 −1 𝑓 (𝑖 𝑗) 𝑓 ( 𝑗). The map 𝑓 preserves 1: it maps 𝐹 to 𝐹 by Exercise 6.15, since
𝐹 = (𝐵0 ) ⊥ = ((𝐵 0) 0 ) ⊥ , and 1 = nrd(1) = nrd( 𝑓 (1)) = 𝑓 (1) 2 so 𝑓 (1) = 1. We have
𝑓 ( 𝑗) 2 = nrd( 𝑓 ( 𝑗)) = nrd( 𝑗) = 𝑏 and similarly 𝑓 (𝑖 𝑗) 2 = 𝑎𝑏 since 𝑗, 𝑖 𝑗 ∈ 𝐵0 . Thus
1 = trd(𝑖) = 𝑏 −1 trd((𝑖 𝑗) 𝑗) = 𝑏 −1𝑇 (𝑖 𝑗, 𝑗) =
= 𝑏 −1𝑇 ( 𝑓 (𝑖 𝑗), 𝑓 ( 𝑗)) = trd(𝑏 −1 𝑓 (𝑖 𝑗) 𝑓 ( 𝑗)) = trd( 𝑓 (𝑖))
and similarly nrd( 𝑓 (𝑖)) = nrd(𝑖) = 𝑎, thus 𝑓 (𝑖) 2 + 𝑓 (𝑖) + 𝑎 = 0. Finally,
𝑓 (𝑖) 𝑓 ( 𝑗) = 𝑏 −1 𝑓 (𝑖 𝑗) 𝑓 ( 𝑗) 2 = 𝑓 (𝑖 𝑗)
and
𝑓 ( 𝑗) 𝑓 (𝑖) = 𝑏 −1 𝑓 ( 𝑗) 𝑓 (𝑖 𝑗) 𝑓 ( 𝑗) = 𝑏 −1 𝑓 ( 𝑗) ( 𝑓 ( 𝑗) 𝑓 (𝑖 𝑗) + 𝑇 ( 𝑓 ( 𝑗), 𝑓 (𝑖 𝑗)))
= 𝑓 (𝑖 𝑗) + 𝑓 ( 𝑗) = ( 𝑓 (𝑖) + 1) 𝑓 ( 𝑗) = 𝑓 (𝑖) 𝑓 ( 𝑗)
so 𝑓 is an isomorphism of 𝐹-algebras. Therefore the functor is full and faithful,
yielding an equivalence of categories.
Corollary 6.4.8. The maps 𝐵 ↦→ 𝑄 = nrd | 𝐵0 ↦→ 𝐶 = 𝑉 (𝑄) yield bijections
Nondegenerate ternary
Quaternion algebras over 𝐹 quadratic forms over 𝐹
↔
up to isomorphism
with discriminant 1 ∈ 𝐹 × /𝐹 ×2
up to isometry
( Nondegenerate ternary )
↔ quadratic forms over 𝐹
up to similarity
Conics over 𝐹
↔
up to isomorphism
90 CHAPTER 6. CHARACTERISTIC 2
Proof. The remaining parts of the bijection follow as in the proof of Corollary 5.2.6.
Proof. Only condition (v) requires significant modification in the case char 𝐹 = 2; see
Exercise 6.13.
𝜇 = 𝛽𝛾 −1 = nrd(𝛾) −1 𝛽𝛾
Exercises
Throughout these exercises, we let 𝐹 be a field (of any characteristic, unless specified).
𝑎, 𝑏
7. Let char 𝐹 = 2 and let 𝐵 = be a quaternion algebra over 𝐹. Compute the
𝐹
left regular representation 𝜆 : 𝐵 → End𝐾 (𝐵) ' M2 (𝐾) where 𝐾 = 𝐹 [𝑖] as in
2.3.8.
I 8. Suppose char 𝐹 = 2. Let 𝑀 ∈ M𝑛 (𝐹) be a symmetric matrix with 𝑛 odd, and
suppose that all diagonal entries of 𝑀 are zero. Show that det 𝑀 = 0.
I 9. Let char 𝐹 = 2 and let 𝐵 be a division 𝐹-algebra with a standard involution.
Prove that either the standard involution is the identity (and so 𝐵 is classified by
Exercise 3.9), or that the conclusion of Theorem 3.5.1 holds for 𝐵: namely, that
either 𝐵 = 𝐾 is a separable quadratic field extension of 𝐹 or that 𝐵 is a quaternion
algebra over 𝐹. [Hint: Replace conjugation by 𝑖 by the map 𝜙(𝑥) = 𝑖𝑥 + 𝑥𝑖, and
show that 𝜙2 = 𝜙. Then diagonalize and proceed as in the case char 𝐹 ≠ 2.]
I 10. Let char 𝐹 = 2. Show that the even Clifford algebra Clf 0 𝑄 of a nondegenerate
ternary quadratic form 𝑄 : 𝑉 → 𝐹 is a quaternion algebra over 𝐹.
I 11. Prove Lemma 6.3.7, that every quadratic form over 𝐹 with char 𝐹 = 2 has a
normalized basis.
I 12. Let char 𝐹 = 2 and let 𝑄 : 𝑉 → 𝐹 be a quadratic form over 𝐹 with discriminant
𝑑 ∈ 𝐹 × /𝐹 ×2 and dim𝐹 𝑉 = 𝑛 odd. Show that 𝑍 (Clf 𝑄) ' 𝐹 [𝑥]/(𝑥 2 − 𝑑) and
that there is a unique 𝜁 ∈ 𝑍 (Clf 𝑄) ∩ Clf 1 𝑄 such that 𝜁 2 = 1.
I 13. Prove Theorem 6.4.11.
I 14. Let 𝑄 := 𝑄 0 𝑄 00 be an orthogonal sum of two anisotropic quadratic forms over
𝐹 (with 𝐹 of arbitrary characteristic). Show that 𝑄 is isotropic if and only if
there exists 𝑐 ∈ 𝐹 × that is represented by both 𝑄 0 and −𝑄 00.
I 15. Let 𝐵 be a quaternion algebra over 𝐹 (with 𝐹 of arbitrary characteristic). Show
that 𝐹 = (𝐵0 ) ⊥ .
I 16. Prove Wedderburn’s little theorem in the following special case: a quaternion
algebra over a finite field with even cardinality is not a division ring. [Hint: See
Exercise 3.16.]
Chapter 7
Simple algebras
93
94 CHAPTER 7. SIMPLE ALGEBRAS
Just as above, for our quaternionic purposes, we can give a direct proof.
𝑊 = {𝛽 ∈ 𝐵 : 𝛽𝛼2 = 𝛼1 𝛽} (7.1.6)
As shown in the above proof, Corollary 7.1.5 can be seen as a general reformulation
of the rational canonical form from linear algebra.
𝑉 = 𝑉0 ) 𝑉1 ) 𝑉2 ) · · · ) 𝑉𝑟 = {0}
𝐼 = ann(𝑥) := {𝛼 ∈ 𝐵 : 𝛼𝑥 = 0}.
Proof. For the first statement, a submodule of 𝐵/𝐼 corresponds to a left ideal containing
𝐼, so 𝐵/𝐼 is simple if and only if 𝐼 is maximal. Conversely, letting 𝑥 ∈ 𝑉 r {0} we
have {0} ≠ 𝐵𝑥 ⊆ 𝑉 a 𝐵-submodule and so 𝐵𝑥 = 𝑉; and consequently 𝑉 ' 𝐵/𝐼 where
𝐼 = ann(𝑥) and again 𝐼 is a maximal left ideal.
7.2. SIMPLE MODULES 97
Having defined the notion of simplicity for modules, we now consider simplicity
of the algebra 𝐵.
Definition 7.2.8. An 𝐹-algebra 𝐵 is simple if the only two-sided ideals of 𝐵 are {0}
and 𝐵.
Equivalently, 𝐵 is simple if and only if any 𝐹-algebra (or even ring) homomorphism
𝐵 → 𝐴 is either injective or the zero map.
Example 7.2.9 shows that algebras of the form M𝑛 (𝐷) with 𝐷 a division 𝐹-algebra
yield a large class of simple 𝐹-algebras. In fact, these are all such algebras, a fact we
will now prove. First, a few preliminary results.
Proof. We have that ker 𝜙 and img 𝜙 are 𝐵-submodules of 𝑉1 and 𝑉2 , respectively, so
either 𝜙 = 0 or ker 𝜙 = {0} and img 𝜙 = 𝑉2 , hence 𝑉1 ' 𝑉2 .
7.2.14. Let 𝐵 be an 𝐹-algebra and consider 𝐵 as a left 𝐵-module. Then there is a map
where 𝐵op is the opposite algebra of 𝐵 defined in 3.2.2. The map 𝜌 is injective since
𝜌 𝛼 = 0 implies 𝜌 𝛼 (1) = 𝛼 = 0; it is also surjective, since if 𝜙 ∈ End 𝐵 (𝐵) then letting
𝛼 = 𝜙(1) we have 𝜙(𝛽) = 𝛽𝜙(1) = 𝛽𝛼 for all 𝛽 ∈ 𝐵. Finally, it is an 𝐹-algebra
homomorphism, since
7.2.16. Many theorems of linear algebra hold equally well over division rings as they
do over fields, as long as one is careful about the direction of scalar multiplication. For
example, let 𝐷 be a division 𝐹-algebra and let 𝑉 be a left 𝐷-module. Then 𝑉 ' 𝐷 𝑛
is free, and choice of basis for 𝑉 gives an isomorphism End𝐷 (𝑉) ' M𝑛 (𝐷 op ). When
𝑛 = 1, this becomes End𝐷 (𝐷) ' 𝐷 op , as in 7.2.14.
Example 7.2.18. The unique simple left M𝑛 (𝐹)-module (up to isomorphism) is the
space 𝐹 𝑛 of column vectors (Example 7.2.1).
7.2.19. Every algebra can be decomposed according to its idempotents 7.2.15. Let 𝐵
be a finite-dimensional 𝐹-algebra. Then we can write 𝐵 = 𝐼1 ⊕ · · · ⊕ 𝐼𝑟 as a direct
sum of indecomposable left 𝐵-modules: this follows by induction, as the decomposing
procedure must stop because each factor is a finite-dimensional 𝐹-vector space. This
means we may write
1 = 𝑒1 + · · · + 𝑒𝑟
Í
with 𝑒 𝑖 ∈ 𝐼𝑖 . For each 𝛼 ∈ 𝐼𝑖 we have 𝛼 = 𝑖 𝛼𝑒 𝑖 whence 𝛼𝑒 𝑖 = 𝛼 and 𝛼𝑒 𝑗 = 0 for
𝑗 ≠ 𝑖, which implies that
𝑒 2𝑖 = 𝑒 𝑖 , 𝑒𝑖 𝑒 𝑗 = 0 for 𝑗 ≠ 𝑖, and 𝐼𝑖 = 𝐵𝑒 𝑖 .
Remark 7.2.20. The tight connection between 𝐹 and M𝑛 (𝐹) is encoded in the fact that
the two rings are Morita equivalent: there is an equivalence of categories between
𝐹-vector spaces and left M𝑛 (𝐹)-modules. For more on this rich subject, see Lam
[Lam99, §18], Reiner [Rei2003, Chapter 4], and Curtis–Reiner [CR81, §35].
7.3. WEDDERBURN–ARTIN 99
Remark 7.3.2. More precisely, we have defined the notion of left semisimple and
could equally well define right semisimple; below we will see that these two notions
are the same.
Remark 7.3.6. Doing linear algebra with semisimple modules mirrors very closely
linear algebra over a field. We have already seen that every submodule and quotient
module of a semisimple module is again semisimple. Moreover, every module homo-
morphism 𝑉 → 𝑊 with 𝑉 semisimple splits, and every submodule of a semisimple
module is a direct summand. The extent to which this fails over other rings concerns
the structure of projective modules; we take this up in Chapter 20.
Proof. Let 𝐼 ⊆ 𝐵 be a minimal nonzero left ideal, the unique simple left 𝐵-module up
to isomorphism as in Lemma 7.2.17. For all 𝛼 ∈ 𝐵, the left ideal 𝐼𝛼 is a homomorphic
Í
image of 𝐼, so by Schur’s lemma, either 𝐼𝛼 = {0} or 𝐼𝛼 is simple. Let 𝐴 := 𝛼∈𝐵 𝐼𝛼.
Then 𝐴 is a nonzero two-sided (!) ideal of 𝐵, so since 𝐵 is simple, we conclude
𝐴 = 𝐵. Thus 𝐵 is the sum of simple 𝐵-modules, and the result follows from Lemma
7.3.5(a).
The converse of Corollary 7.3.8 is true and is proven as Corollary 7.3.14, a conse-
quence of the Wedderburn–Artin theorem.
In analogy to 7.2.16, we have the following corollary.
Corollary 7.3.9. Let 𝐵 be a simple 𝐹-algebra and let 𝑉 be a left 𝐵-module. Then
𝑉 ' 𝐼 ⊕𝑛 for some 𝑛 ≥ 1, where 𝐼 is a simple left 𝐵-module. In particular, two left
𝐵-modules 𝑉1 , 𝑉2 are isomorphic if and only if dim𝐹 𝑉1 = dim𝐹 𝑉2 .
In other words, this corollary says that if 𝐵 is simple then every left 𝐵-module 𝑉
is free over 𝐵, so has a left basis over 𝐵; if we define the rank of a left 𝐵-module 𝑉
to be cardinality of this basis (the integer 𝑛 such that 𝑉 ' 𝐼 ⊕𝑛 as in Corollary 7.3.9),
then two such modules are isomorphic if and only if they have the same rank.
We now come to one of the main results of this chapter.
by 7.2.16,
End 𝐵 𝐼𝑖⊕𝑛𝑖 ' M𝑛𝑖 (𝐷 𝑖 )
Remark 7.3.11. Main Theorem 7.3.10 as it is stated was originally proven by Wedder-
burn [Wed08], and so is sometimes called Wedderburn’s theorem. However, this term
may also apply to the theorem of Wedderburn that a finite division ring is a field; and
Artin generalized Main Theorem 7.3.10 to rings where the ascending and descending
chain condition holds for left ideals [Art26]. We follow the common convention by
referring to Main Theorem 7.3.10 as the Wedderburn–Artin theorem.
Corollary 7.3.12. Let 𝐵 be a simple 𝐹-algebra. Then 𝐵 ' M𝑛 (𝐷) for a unique
𝑛 ∈ Z ≥1 and a division algebra 𝐷 unique up to isomorphism.
Definition 7.4.1. The Jacobson radical rad 𝐵 of 𝐵 is the intersection of all maximal
left ideals of 𝐵.
We will in Corollary 7.4.6 see that this definition has left-right symmetry. Before
doing so, we see right away the importance of the Jacobson radical in the following
lemma.
Proof. Let 𝐽 = rad 𝐵. Under the natural map 𝐵 → 𝐵/𝐽, the intersection of all maximal
left ideals of 𝐵/rad 𝐵 corresponds to the intersection of all maximal left ideals of 𝐵
containing 𝐽; but rad 𝐵 is the intersection thereof, so rad(𝐵/𝐽) = {0} and by Lemma
7.4.2, 𝐵/𝐽 is semisimple.
ann 𝑉 := {𝛼 ∈ 𝐵 : 𝛼𝑉 = 0}.
Proof. The statement follows by combining 7.4.4 and Lemma 7.4.5: rad 𝐵 is the
intersection of two-sided ideals and so is itself a two-sided ideal.
𝐵 ⊃ 𝐽 ⊇ 𝐽2 ⊇ . . . .
There exists 𝑛 ∈ Z ≥1 such that 𝐽 𝑛 = 𝐽 2𝑛 . We claim that 𝐽 𝑛 = {0}. Assume for the
purposes of contradiction that 𝐼 ⊆ 𝐽 𝑛 is a minimal left ideal such that 𝐽 𝑛 𝐼 ≠ {0}. Let
𝛼 ∈ 𝐼 be such that 𝐽 𝑛 𝛼 ≠ {0}; by minimality 𝐽 𝑛 𝛼 = 𝐼, so 𝛼 = 𝜂𝛼 for some 𝜂 ∈ 𝐽 𝑛 ,
thus (1 − 𝜂)𝛼 = 0. But 𝜂 ∈ 𝐽 𝑛 ⊆ 𝐽 = rad 𝐵. By Lemma 7.4.5, 1 − 𝜂 ∈ 𝐵× is a unit
hence 𝛼 = 0, a contradiction.
Example 7.4.9. Suppose 𝐵 has a standard involution. Then by Lemma 7.4.8 and the
fact that 𝐵 has degree 2, we conclude that rad 𝐵 ⊆ {𝜖 ∈ 𝐵 : 𝜖 2 = 0}. If char 𝐹 ≠ 2
and we define rad(nrd) as in 4.3.9 for the quadratic form nrd, then rad(nrd) = rad 𝐵
(Exercise 7.20).
Corollary 7.4.10. The Jacobson radical rad 𝐵 is the intersection of all maximal right
ideals of 𝐵.
Example 7.5.2. The center 𝑍 (𝐵) of a simple 𝐹-algebra is a field, since it is a simple
commutative 𝐹-algebra. One reaches the same conclusion by applying Corollary
7.3.12 together with 𝑍 (M𝑛 (𝐷)) = 𝑍 (𝐷) (Exercise 7.5).
The category of central simple algebras is closed under tensor product, as follows.
Í
Proof. First, centrality in part (a). Suppose that 𝛾 = 𝑖 𝛼𝑖 ⊗ 𝛽𝑖 ∈ 𝑍 ( 𝐴 ⊗ 𝐵) (a finite
sum). By rewriting the tensor, without loss of generality, we may suppose that 𝛼𝑖
are linearly independent over 𝐹. Í Then by properties of tensor products, the elements
𝛽𝑖 ∈ 𝐵 in the representation 𝛾 = 𝑖 𝛼𝑖 ⊗ 𝛽𝑖 are unique. But then for all 𝛽 ∈ 𝐵,
! !
∑︁ ∑︁ ∑︁ ∑︁
(𝛼𝑖 ⊗ 𝛽𝛽𝑖 ) = (1 ⊗ 𝛽) 𝛼𝑖 ⊗ 𝛽 𝑖 = 𝛼𝑖 ⊗ 𝛽𝑖 (1 ⊗ 𝛽) = (𝛼𝑖 ⊗ 𝛽𝑖 𝛽)
𝑖 𝑖 𝑖 𝑖
but by minimality of 𝑚, the right-hand side is zero, so 𝛽𝛽𝑖 = 𝛽𝑖 𝛽 for all 𝑖. Hence
𝛽𝑖 ∈ 𝑍 (𝐵) = 𝐹 for all 𝑖 and as above 𝛾 = 𝛼 ⊗ 1 for some 0 ≠ 𝛼 ∈ 𝐴. But then
𝐼 ⊇ ( 𝐴 ⊗ 1) (𝛼 ⊗ 1) ( 𝐴 ⊗ 1) = ( 𝐴𝛼𝐴) ⊗ 1 = 𝐴 ⊗ 1
is an isomorphism.
Proof. First, the implication (⇒). Just as in 7.2.14, 𝜙 is a nonzero 𝐹-algebra ho-
momorphism. By Proposition 7.5.3, 𝐵 ⊗𝐹 𝐵op is simple, so 𝜙 is injective. Since
dim𝐹 (𝐵 ⊗𝐹 𝐵op ) = dim𝐹 End𝐹 (𝐵) = (dim𝐹 𝐵) 2 , 𝜙 is an isomorphism.
Now the converse (⇐); suppose 𝜙 is an isomorphism. If 𝐼 is an ideal of 𝐵 then
𝜙(𝐼 ⊗ 𝐵op ) ⊆ End𝐹 (𝐵) is an ideal; but End𝐹 (𝐵) is simple over 𝐹, therefore 𝐼 is trivial.
And if 𝛼 ∈ 𝑍 (𝐵) then 𝜙(𝛼 ⊗ 1) ∈ 𝑍 (End𝐹 (𝐵)) = 𝐹, so 𝛼 ∈ 𝐹.
7.6. QUATERNION ALGEBRAS 105
7.5.5. Among central simple algebras over a field, quaternion algebras have an espe-
cially nice presentation because the quadratic norm form can be put into a standard
form (indeed, diagonalized in characteristic not 2). More generally, one may look at
algebras with a similarly nice presentation, as follows.
Let 𝐹 be a field, let 𝐾 ⊃ 𝐹 be a cyclic extension of 𝐹 of degree 𝑛 = [𝐾 : 𝐹], let
𝜎 ∈ Gal(𝐾 | 𝐹) be a generator, and let 𝑏 ∈ 𝐹 × . For example, if 𝐹 contains a primitive
√
𝑛th√root of √unity 𝜁 ∈ 𝐹 × , and 𝑎 ∈ 𝐹 × r 𝐹 ×𝑛 , then we may take 𝐾 = 𝐹 ( 𝑛 𝑎) and
𝜎( 𝑛 𝑎) = 𝜁 𝑛 𝑎. We then define the cyclic algebra
𝐾, 𝜎, 𝑏
= 𝐾 ⊕ 𝐾 𝑗 ⊕ · · · ⊕ 𝐾 𝑗 𝑛−1
𝐹
to be the left 𝐾-vector space with basis 1, 𝑗, . . . , 𝑗 𝑛−1 and with multiplication 𝑗 𝑛 = 𝑏
and 𝑗 𝛼 = 𝜎(𝛼) 𝑗 for 𝛼 ∈ 𝐾. The definition of a cyclic algebra generalizes that
of 6.1.5, where there is only one choice for the generator 𝜎. A cyclic algebra is a
central simple algebra over 𝐹 of dimension 𝑛2 , and indeed (𝐾, 𝜎, 𝑏 | 𝐾) ' M𝑛 (𝐾).
(See Exercise 7.11.) More generally, we may relax the condition that 𝐺 be cyclic:
there is an analogous construction for any finite Galois extension, yielding a central
simple algebra called a crossed product algebra (and giving an interpretation to a
second cohomology group): see Reiner [Rei2003, §29–30]. There are significant
open problems relating cyclic algebras and crossed products to central simple algebras
in general [ABGV2006].
It is a consequence of the main theorem of class field theory that if 𝐹 is a global
field then every (finite-dimensional) central simple algebra over 𝐹 is isomorphic to a
cyclic algebra.
Remark 7.5.6. The theory of central simple algebras and Brauer groups extends to one
over commutative rings (or even schemes), and this becomes the theory of Azumaya
algebras: see Saltman [Sal99, §2].
Proof. First, (i) ⇒ (ii): if 𝐵 is a quaternion algebra, then 𝐵 is central simple (7.2.11).
The equivalence (ii) ⇔ (iii) follows from the Wedderburn–Artin theorem:
𝑟
∑︁
1 = dim 𝑍 (𝐵) = dim𝐹 𝑍 (𝐷 𝑖 ) ≥ 𝑟
𝑖=1
106 CHAPTER 7. SIMPLE ALGEBRAS
so 𝑟 = 1.
Next we prove (ii) ⇒ (iv). If 𝐵 is central simple, then 𝐵 ⊗𝐹 𝐹 al is a central simple
𝐹 -algebra by Proposition 7.5.3. But by Exercise 2.9, the only division 𝐹 al -algebra
al
Inspired by the proof of this result, we reconsider and reprove our splitting criterion
for quaternion algebras.
Proposition 7.6.2. Let 𝐵 be a quaternion algebra over 𝐹. Then the following are
equivalent:
Proof. The equivalence (i) ⇔ (ii) follows from the Wedderburn–Artin theorem (also
proved in Main Theorem 5.4.4 and Theorem 6.4.11). The implications (i) ⇒ (iii) ⇒
(ii) and (i) ⇒ (iv) ⇒ (ii) are clear.
7.6.3. We showed in Lemma 7.2.17 that a simple algebra 𝐵 has a unique simple left
𝐵-module 𝐼 up to isomorphism, obtained as a minimal nonzero left ideal. If 𝐵 is
a quaternion algebra, this simple module 𝐼 can be readily identified using the above
proposition. If 𝐵 is a division ring, then necessarily 𝐼 = 𝐵. Otherwise, 𝐵 ' M2 (𝐹),
and then 𝐼 ' 𝐹 2 , and the map 𝐵 → End𝐹 (𝐼) given by left matrix multiplication is an
isomorphism.
Proof. By Corollary 7.3.12, we have 𝐵 ' End𝐷 (𝑉) ' M𝑛 (𝐷 op ) where 𝑉 is a simple
𝐵-module and 𝐷 = End 𝐵 (𝑉) is a central 𝐹-algebra. Now the maps 𝑓 , 𝑔 give 𝑉 the
structure of an 𝐴-module in two ways. The 𝐴-module structure commutes with the
𝐷-module structure since 𝐵 ' End𝐷 (𝑉). So 𝑉 has two 𝐴 ⊗𝐹 𝐷-module structures via
𝑓 and 𝑔.
By Proposition 7.5.3, since 𝐷 is central over 𝐹, we have that 𝐴 ⊗𝐹 𝐷 is a simple 𝐹-
algebra. By Corollary 7.3.9 and a dimension count, the two 𝐴 ⊗𝐹 𝐷-module structures
on 𝑉 are isomorphic. Thus, there exists an isomorphism 𝛽 : 𝑉 → 𝑉 of 𝐴⊗𝐹 𝐷-modules;
i.e. 𝛽( 𝑓 (𝛼)𝑥) = 𝑔(𝛼) 𝛽(𝑥) for all 𝛼 ∈ 𝐴 and 𝑥 ∈ 𝑉, and 𝛽(𝛿𝑥) = 𝛿𝛽(𝑥) for all 𝛿 ∈ 𝐷
and 𝑥 ∈ 𝑉. We have 𝛽 ∈ End𝐷 (𝑉) ' 𝐵 and so we can write 𝛽 𝑓 (𝛼) 𝛽−1 = 𝑔(𝛼) for all
𝛼 ∈ 𝐴, as claimed.
Proof. Let 𝜄𝑖 : 𝐴𝑖 ↩→ 𝐵 be the natural inclusions, and apply Main Theorem 7.7.1 to
𝑓 := 𝜄1 and 𝑔 := 𝜄2 ◦ 𝜙: we conclude there exists 𝛽 ∈ 𝐵× such that 𝜄1 (𝛼) = 𝛼 =
𝛽−1 𝜄2 (𝜙(𝛼)) 𝛽 or equivalently 𝜙(𝛼) = 𝛽𝛼𝛽−1 for all 𝛼 ∈ 𝐴1 , as desired.
Proof. The implication (⇐) is immediate. Conversely (⇒), let 𝐴𝑖 := 𝐹 [𝛼𝑖 ] '
𝐹 [𝑥]/( 𝑓𝑖 (𝑥)) where 𝑓𝑖 (𝑥) ∈ 𝐹 [𝑥] are minimal polynomials over 𝐹. Since these
polynomials are irreducible, 𝐴𝑖 is a field hence simple, so Corollary 7.7.2 gives the
result.
Proof. For the implication (⇐), if 𝑏 0/𝑏 = Nm𝐾 |𝐹 (𝛼) with 𝛼 ∈ 𝐾 × , then an isomor-
phism is furnished as left 𝐾-vector spaces by sending 𝑗 ↦→ 𝛼 𝑗.
∼ 𝐵 0 := (𝐾, 𝑏 0 | 𝐹) be an isomorphism
For the implication (⇒), let 𝜙 : (𝐾, 𝑏 | 𝐹) −→
of 𝐹-algebras. If 𝐾 ' 𝐹 × 𝐹 is not a field, then Nm𝐾 |𝐹 (𝐾 × ) = 𝐹 × and the result holds.
So suppose 𝐾 is a field. Then 𝜙(𝐾) ⊆ 𝐵 0 isomorphic to 𝐾 as an 𝐹-algebra, but need
not be the designated one in 𝐵’; however, by the Skolem–Noether theorem, we may
postcompose 𝜙 with an automorphism that sends 𝜙(𝐾) to the designated one, i.e., we
may suppose that 𝜙 is a 𝐾-linear map (taking the algebras as left 𝐾-vector spaces).
Let 𝜙( 𝑗) = 𝛼 + 𝛽 𝑗 0 with 𝛼, 𝛽 ∈ 𝐾. Then
𝐶 𝐵 ( 𝐴) := {𝛽 ∈ 𝐵 : 𝛼𝛽 = 𝛽𝛼 for all 𝛼 ∈ 𝐴}
be the centralizer of 𝐴 in 𝐵.
Proof. First, part (a). We interpret the centralizer as arising from certain kinds of
endomorphisms. We have that 𝐵 is a left 𝐴⊗ 𝐵op module by the action (𝛼⊗ 𝛽) · 𝜇 = 𝛼𝜇𝛽
for 𝛼 ⊗ 𝛽 ∈ 𝐴 ⊗ 𝐵op and 𝜇 ∈ 𝐵. We claim that
𝛾𝛼 = 𝜙(𝛼) = 𝛼𝜙(1) = 𝛼𝛾
7.7. THE SKOLEM–NOETHER THEOREM 109
𝐶 𝐵 ( 𝐴) = End 𝐴⊗𝐵op (𝐵) ' EndM𝑛 (𝐷) (𝑉 𝑟 ) ' M𝑟 (EndM𝑛 (𝐷) (𝑉)) ' M𝑟 (𝐷 op ).
Thus 𝐶 𝐵 ( 𝐴) is simple.
For part (b),
and
dim𝐹 ( 𝐴 ⊗ 𝐵op ) = dim𝐹 𝐴 · dim𝐹 𝐵 = 𝑛2 dim𝐹 𝐷
and finally
dim𝐹 𝐵 = dim𝐹 𝑉 𝑟 = 𝑟 dim𝐹 𝐷 𝑛 = 𝑟𝑛 dim𝐹 𝐷;
putting these together gives dim𝐹 𝐴 · dim𝐹 𝐶 𝐵 ( 𝐴) = 𝑟𝑛 dim𝐹 𝐷 = dim𝐹 𝐵.
Finally, part (c) follows from (a) and (b):
Example 7.7.10. We always have the two extremes 𝐴 = 𝐹 and 𝐴 = 𝐵, with 𝐶 𝐵 (𝐹) = 𝐵
and 𝐶 𝐵 (𝐵) = 𝐹, accordingly.
We note the following structurally crucial corollary of Proposition 7.7.8.
Corollary 7.7.11. Let 𝐵 be a central simple 𝐹-algebra and let 𝐾 be a maximal subfield.
Then [𝐵 : 𝐹] = [𝐾 : 𝐹] 2 .
The following example will hopefully illustrate the role of this notion.
𝑎, 𝑏
Example 7.8.3. For char 𝐹 ≠ 2 and 𝐵 = , in the basis 1, 𝑖, 𝑗, 𝑖 𝑗 we have
𝐹
𝜉 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 (substituting 𝑡, 𝑥, 𝑦, 𝑧 for 𝑥1 , . . . , 𝑥 4 ). We claim that the universal
minimal polynomial is
Indeed,
that 𝜉 satisfies 𝑚 𝐵 (𝜉; 𝜉) = 0 by considering 𝜉 ∈ 𝐵 ⊗𝐹 𝐹 (𝑡, 𝑥, 𝑦, 𝑧) =
we verify
𝑎, 𝑏
and computing that trd(𝜉) = 2𝑡 and nrd(𝜉) = 𝑡 2 − 𝑎𝑥 2 − 𝑏𝑦 2 + 𝑎𝑏𝑧 2 ;
𝐹 (𝑡, 𝑥, 𝑦, 𝑧)
and this polynomial is minimal because 𝜉 ∉ 𝐹 (𝑡, 𝑥, 𝑦, 𝑧) does not satisfy a polynomial
of degree 1 over 𝐹 (𝑡, 𝑥, 𝑦, 𝑧).
7.8.4. If 𝐵 ' 𝐵1 × · · · × 𝐵𝑟 , then in a basis for 𝐵 obtained from the union of bases for
the factors 𝐵𝑖 with universal elements 𝜉𝑖 , we have
In the proofs that follow, we abbreviate by using multi-index notation, e.g. writing
𝐹 [𝑥] := 𝐹 [𝑥 1 , . . . , 𝑥 𝑛 ].
Lemma 7.8.5. We have 𝑚 𝐵 (𝜉; 𝑇) ∈ 𝐹 [𝑥1 , . . . , 𝑥 𝑛 ] [𝑇], i.e., the universal minimal
polynomial has coefficients in 𝐹 [𝑥1 , . . . , 𝑥 𝑛 ].
Proof. For part (a), we consider the map given by left multiplication by 𝜉 on 𝐵⊗𝐹 𝐹 (𝑥).
In the basis 𝑒 1 , . . . , 𝑒 𝑛 , almost by construction we find that the matrix of this map has
coefficients in 𝐹 [𝑥] (it is the matrix of linear forms obtained from left multiplication
by 𝑒 𝑖 ). We conclude that 𝜉 satisfies the characteristic polynomial of this matrix,
which is a monic polynomial with coefficients in 𝐹 [𝑥]. Since 𝑚 𝐵 (𝜉; 𝑇) divides this
polynomial (over 𝐹 (𝑥)) by minimality, by Gauss’s lemma we conclude that 𝑚 𝐵 (𝜉; 𝑇) ∈
𝐹 [𝑥] [𝑇].
Lemma 7.8.7. For any field extension 𝐾 ⊇ 𝐹, we have 𝑚 𝐵 ⊗𝐹 𝐾 (𝜉; 𝑇) = 𝑚 𝐾 (𝜉; 𝑇).
Proof. First, because an 𝐹-basis for 𝐵 is a 𝐾-basis for 𝐵 ⊗𝐹 𝐾, the element 𝜉 (as
the universal element of 𝐵), also serves as a universal element of 𝐵 ⊗𝐹 𝐾. Since
𝐾 (𝑥1 , . . . , 𝑥 𝑛 ) ⊆ 𝐹 (𝑥1 , . . . , 𝑥 𝑛 ), by minimality we have 𝑚 𝐵𝐾 (𝜉; 𝑇) | 𝑚 𝐵 (𝜉; 𝑇). Con-
versely, let 𝐹 (𝑥) [𝜉] ⊆ 𝐵 ⊗𝐹 𝐹 (𝑥) be the subalgebra generated by 𝜉 over 𝐹 (𝑥); then
𝐹 (𝑥) [𝜉] ' 𝐹 (𝑥) [𝑇]/(𝑚 𝐵 (𝜉; 𝑇)). Tensoring with 𝐾 gives
Proof. For (a), we recall (as in the proof of Lemma 7.8.5 that 𝜉 satisfies the character-
istic polynomial of left multiplication on 𝐾, a polynomial of degree 𝑛 = [𝐾 : 𝐹]; on
the other hand, choosing a primitive element 𝛼, we see the specialization 𝑚 𝐾 (𝛼; 𝑇) is
satisfied by 𝛼 so has degree at least 𝑛, so equality holds and 𝑚 𝐾 (𝛼; 𝑇) is the charac-
teristic polynomial. Therefore the universal minimal polynomial is the characteristic
polynomial, and hence the same is true under every specialization.
For (b), it suffices to prove this when 𝐹 = 𝐹 al is algebraically closed, in which case
𝐵 ' M𝑛 (𝐹); by Proposition 7.8.6, we may assume 𝐵 = M𝑛 (𝐹). By 7.8.1, we want
to show that 𝑚 𝐵 (𝛼; 𝑇) for 𝛼 ∈ M𝑛 (𝐹) is the usual characteristic polynomial. But the
universal element (in a basis of matrix units, or any basis) satisfies its characteristic
polynomial of degree 𝑛, and a nilpotent matrix (with 1s just above the diagonal) has
minimal polynomial 𝑇 𝑛 , so we conclude as in the previous paragraph.
Proof. Consider (again) 𝑉 := 𝐹 (𝑥) [𝜉] ⊆ 𝐵 ⊗𝐹 𝐹 (𝑥) the subalgebra generated over
𝐹 (𝑥) by 𝜉; then 𝜉 acts on 𝑉 ' 𝐹 (𝑥) [𝑇]/(𝑚 𝐵 (𝜉; 𝑇)) by left multiplication with
characteristic polynomial 𝑚 𝐵 (𝜉; 𝑇). It follows that the reduced trace and reduced
norm are the usual trace and determinant in this representation, so the announced
properties follow on specialization.
Proof. The center of 𝐵 is a field 𝐾 = 𝑍 (𝐵) and as a 𝐾-algebra, the center 𝑍 (𝐵) =
𝐾 is certainly separable over 𝐾. (Or use Proposition 7.5.3 and Theorem 7.9.4(iii)
below.)
𝐵×𝐵→𝐹
(𝛼, 𝛽) ↦→ trd(𝛼𝛽)
is nondegenerate.
Moreover, if char 𝐹 = 0, then these are further equivalent to:
(v) The bilinear form (𝛼, 𝛽) ↦→ Tr 𝐵 |𝐹 (𝛼𝛽) is nondegenerate.
A separable 𝐹-algebra is sometimes called absolutely semisimple, in view of
Theorem 7.9.4(iii).
Proof. First we prove (i) ⇒ (ii). Let 𝐵𝑖 be a simple component of 𝐵; then 𝑍 (𝐵𝑖 ) is
separable over 𝐹. Let 𝐾𝑖 ⊇ 𝐹 be a separable field extension containing 𝑍 (𝐵𝑖 ) that
splits 𝐵𝑖 , so 𝐵𝑖 ⊗ 𝑍 (𝐵𝑖 ) 𝐾𝑖 ' M𝑛𝑖 (𝐾𝑖 ). Let 𝐾 be the compositum of the fields 𝐾𝑖 . Then
𝐾 is separable, and
The implication (iv) ⇒ (i) holds with char 𝐹 arbitrary: if 𝜖 ∈ rad 𝐵 then 𝛼𝜖 ∈ rad 𝐵
is nilpotent and trd(𝛼𝜖) = 0 for all 𝛼 ∈ 𝐵, and by nondegeneracy 𝜖 = 0.
The final equivalence (iv) ⇔ (v) follows when char 𝐹 = 0 since the algebra trace
pairing on each simple factor is a scalar multiple of the reduced trace pairing.
Exercises
Throughout the exercises, let 𝐹 be a field.
𝑎, 𝑏
I 1. Prove that a quaternion algebra 𝐵 = with char 𝐹 ≠ 2 is simple by a direct
𝐹
calculation, as follows.
(a) Let 𝐼 be a nontrivial two-sided ideal, and let 𝜖 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ∈ 𝐼. By
considering 𝑖𝜖 − 𝜖𝑖, show that 𝑡 + 𝑥𝑖 ∈ 𝐼.
(b) Arguing symmetrically and taking a linear combination, show that 𝑡 ∈ 𝐼,
and conclude that 𝑡 = 0, whence 𝑥 = 𝑦 = 𝑧 = 0.
𝑎, 𝑏
Modify this argument to show that an algebra 𝐵 = is simple when
𝐹
char 𝐹 = 2. [We proved these statements without separating into cases in
7.2.11.]
2. Let 𝐵 be a quaternion algebra over 𝐹, and let 𝐾 ⊆ 𝐵 be an 𝐹-subalgebra that is
commutative. Show that dim𝐹 𝐾 ≤ 2.
3. Let 𝐵 be a quaternion algebra. Exhibit an explicit isomorphism
∼ M (𝐹).
𝐵 ⊗𝐹 𝐵 −→ 4
𝑖 𝑛 = 𝑎, 𝑗 𝑛 = 𝑏, 𝑗𝑖 = 𝜁𝑖 𝑗 .
𝐶 𝐴⊗𝐵 ( 𝐴 0 ⊗ 𝐵 0) = 𝐶 𝐴 ( 𝐴 0) ⊗ 𝐶 𝐵 (𝐵 0).
13. Let 𝐵 be a finite-dimensional 𝐹-algebra. Show that the following are equivalent:
(i) 𝐵 is separable;
(ii) 𝐵 is semisimple and the center 𝐾 = 𝑍 (𝐵) is separable;
(iii) 𝐵 ⊗𝐹 𝐵op is semisimple.
14. Let 𝐺 ≠ {1} be a finite group. Show that the augmentation ideal, the two-sided
ideal generated by 𝑔 − 1 for 𝑔 ∈ 𝐺, is a nontrivial ideal, and hence 𝐹 [𝐺] is not
simple as an 𝐹-algebra.
7.9. SEPARABLE ALGEBRAS 117
15. Let 𝐺 be a finite group of order 𝑛 = #𝐺. Show that 𝐹 [𝐺] is a separable 𝐹-
algebra if and only if char 𝐹 - 𝑛 as follows. [This exercise is known as Maschke’s
theorem.]
(a) Suppose first that char 𝐹 = 0 for a special but quick special case. Compute
Í 𝐹 [𝐺] is separable.
the trace pairing and conclude
(b) If char 𝐹 | 𝑛, show that 𝑁 = 𝑔 ∈𝐺 𝑔 is a nilpotent element in the center of
𝐹 [𝐺], so 𝐹 [𝐺] is not semisimple.
(c) Suppose that char 𝐹 - 𝑛. Let 𝐵 = 𝐹 [𝐺]. Define the map of left 𝐵-modules
by
𝜙 : 𝐵 → 𝐵 ⊗𝐹 𝐵op =: 𝐵e
1 ∑︁
𝜙(1) = 𝑔 ⊗ 𝑔 −1
𝑛 𝑔 ∈𝑔
is true if we consider the separable closure. (We proved this already in Exercise
6.3 for 𝐷 a quaternion algebra.)
Let 𝐹 be a separably closed field, so every nonconstant separable polynomial
with coefficients in 𝐹 has a root in 𝐹. Let 𝐷 be a finite-dimensional central
division algebra over 𝐹 with char 𝐹 = 𝑝. For purposes of contradiction, assume
that 𝐷 ≠ 𝐹.
(a) Prove that dim𝐹 𝐷 is divisible by 𝑝.
(b) Show that the minimal polynomial of each nonzero 𝑑 ∈ 𝐷 has the form
𝑒
𝑥 𝑝 − 𝑎 for some 𝑎 ∈ 𝐹 and 𝑒 ≥ 0.
∼ M (𝐹 al ). Show that
(c) Choose an 𝐹 al -algebra isomorphism 𝜙 : 𝐷 ⊗𝐹 𝐹 al −→ 𝑛
tr 𝜙(𝑥 ⊗ 1) = 0 for all 𝑥 ∈ 𝐷.
(d) Prove that 𝐷 does not exist.
24. Let 𝐾 ⊇ 𝐹 be a separable (possibly infinite) extension, and let 𝑓 (𝑇) ∈ 𝐾 [𝑇].
Suppose that 𝑓 (𝑇) 𝑛 ∈ 𝐹 [𝑇] for some 𝑛 ∈ Z ≥1 . Show that 𝑓 (𝑇) ∈ 𝐹 [𝑇]. [Hint:
when 𝑝 = char 𝐹 | 𝑛, use the fact that 𝑎 𝑝 ∈ 𝐹 implies 𝑎 ∈ 𝐹.]
25. Let 𝐵 be a finite-dimensional 𝐹-algebra, let 𝛼 ∈ 𝐵, and let 𝑓L (𝛼; 𝑇) and 𝑓R (𝛼; 𝑇)
be the characteristic polynomial of left and right multiplication of 𝛼 on 𝐵,
respectively.
(a) If 𝐵 is semisimple, show that 𝑓L (𝛼; 𝑇) = 𝑓R (𝛼; 𝑇).
(b) Give an example where 𝑓L (𝛼; 𝑇) ≠ 𝑓R (𝛼; 𝑇).
I 26. Use the Skolem–Noether theorem to give another solution to Exercise 6.2: if
𝐾 ⊂ 𝐵 is a separable quadratic 𝐹-algebra then 𝐵 ' (𝐾, 𝑏 | 𝐹) for some 𝑏 ∈ 𝐹 × .
27. Give a direct proof of Corollary 7.7.4. [Hint: Use the fact that there is a unique
simple left 𝐵-module.]
28. Let 𝐵 = (𝐾, 𝑏 | 𝐹) be a quaternion algebra. Show that the subgroup of Aut(𝐵)
that maps 𝐾 ⊆ 𝐵 to itself is isomorphic to the group
𝐾 × /𝐹 × ∪ 𝑗 (𝐾 × /𝐹 × ).
Show that the subgroup of Aut(𝐵) that restricts to the identity on 𝐾 (fixing 𝐾
elementwise) is isomorphic to 𝐾 × /𝐹 × .
I 29. Use the Skolem–Noether theorem and the fact that a finite group cannot be writ-
ten as the union of the conjugates of a proper subgroup to prove Wedderburn’s
little theorem: a finite division ring is a field.
I 30. Let 𝐵 be a quaternion algebra over 𝐹. In this exercise, we show that the
commutator subgroup
is precisely
(b) Show that [GL2 (𝐹), GL2 (𝐹)] = SL2 (𝐹) if #𝐹 > 3. [Hint: choose 𝑧 ∈ 𝐹
𝑧 0
such that 𝑧2 − 1 ∈ 𝐹 × , let 𝛾 = , and show that for all 𝑥 ∈ 𝐹 we
0 𝑧 −1
have
1 𝑥(𝑧2 − 1) −1
1 𝑥
= 𝛾,
0 1 0 1
and analogously for the transpose. See 28.3 for a review of elementary
matrices.]
(c) Suppose that 𝐵 is a division algebra. Let 𝛾 ∈ 𝐵1 . Show that there exists
𝛼 ∈ 𝐾 = 𝐹 (𝛾) such that 𝛼𝛼−1 = 𝛾. [Hint: This is a special case of
Hilbert’s theorem 90. Let 𝛼 = 𝛾 + 1 if 𝛾 ≠ −1, and 𝛼 ∈ 𝐵0 r {0} if
𝛾 = −1, with appropriate modifications if char 𝐹 = 2.] Conclude from the
Skolem–Noether theorem that there exists 𝛽 ∈ 𝐵× such that 𝛽𝛼𝛽−1 = 𝛼,
and thus 𝛾 ∈ [𝐵× , 𝐵× ].
31. Show that every ring automorphism of H is inner. (Compare this with ring
automorphisms of C!)
Chapter 8
121
122 CHAPTER 8. SIMPLE ALGEBRAS AND INVOLUTIONS
• 𝐵1 ⊗ 𝐵2 is a division algebra;
• 𝐵1 ⊗ 𝐵2 ' M2 (𝐵3 ) where 𝐵3 is a quaternion division algebra over 𝐹; or
• 𝐵1 ⊗ 𝐵2 ' M4 (𝐹).
We could combine the latter two and just say that 𝐵1 ⊗ 𝐵2 ' M2 (𝐵3 ) where 𝐵3 is a
quaternion algebra over 𝐹, since M2 (M2 (𝐹)) ' M4 (𝐹) as 𝐹-algebras.
𝑎, 𝑏 2
𝑎, 𝑏 1 𝑏 2 𝑎, 𝑏 𝑎, 𝑏
where 𝐵3 = . In particular, ⊗ ' M4 (𝐹), since '
𝐹 𝐹 𝐹 𝐹
M2 (𝐹).
𝐵1 ↩→ 𝐵1 ⊗ 𝐵2
𝛼 ↦→ 𝛼 ⊗ 1
𝐵1 ⊗ 𝐵2 = (𝐵1 ) 𝐾 + (𝐵1 ) 𝐾 𝑗
𝛼1 ⊗ 𝑧 + 𝛼1 ⊗ (𝑡 − 𝑧) = 𝛼1 ⊗ 𝑡 = 0.
Remark 8.2.6. In view of Proposition 8.2.3, we say that two quaternion algebras
𝐵1 , 𝐵2 over 𝐹 are linked if they contain a common quadratic field extension 𝐾 ⊇ 𝐹.
For further discussion of biquaternion algebras and linkage in characteristic 2 (where
one must treat separable and inseparable extensions differently), see Knus [Knu93],
Lam [Lam2002], or Sah [Sah72]. Garibaldi–Saltman [GS2010] study the subfields of
quaternion algebra over fields with char 𝐹 ≠ 2.
From now on, we suppose that char 𝐹 ≠ 2. (For the case char 𝐹 = 2, see Chapman–
Dolphin–Laghbribi [CDL2015, §6].)
124 CHAPTER 8. SIMPLE ALGEBRAS AND INVOLUTIONS
Then dim𝐹 𝑉 = 6, and we may identify 𝑉 = 𝐵10 ⊗ 1−1 ⊗ 𝐵20 . The reduced norm on each
factor separately defines a quadratic form on 𝑉 by taking the difference: explicitly, if
𝐵1 = (𝑎 1 , 𝑏 1 | 𝐹) and 𝐵2 = (𝑎 2 , 𝑏 2 | 𝐹), then taking the standard bases for 𝐵1 , 𝐵2
The quadratic form 𝑄(𝐵1 , 𝐵2 ) : 𝑉 → 𝐹 is called the Albert form of the biquaternion
algebra 𝐵1 ⊗ 𝐵2 .
Proof. The implication (ii) ⇒ (iv) follows by construction 8.2.7. To prove (iv) ⇒
(ii), without loss of generality, we may suppose 𝐵1 , 𝐵2 are division algebras; then
an isotropic vector of 𝑄 corresponds√to elements 𝛼1 ∈ 𝐵1 and 𝛼2 ∈ 𝐵2 such that
𝛼12 = 𝛼22 = 𝑐 ∈ 𝐹 × . Therefore 𝐾 = 𝐹 ( 𝑐) is a common quadratic splitting field.
Remark 8.2.9. Albert’s book [Alb39] on algebras still reads well today. The proof of
Proposition 8.2.3 is due to him [Alb72]. (“I discovered this theorem some time ago.
There appears to be some continuing interest in it, and I am therefore publishing it
now.”) Albert used Proposition 8.2.8 to show that
−1, −1 𝑥, 𝑦
𝐵1 = and 𝐵2 =
𝐹 𝐹
Remark 8.3.7. Just as quaternion algebras are in correspondence with conics (Corollary
5.5.2), with a quaternion algebra split if and only if the corresponding conic has a
rational point (Theorem 5.5.3), similarly the Brauer group of a field has a geometric
interpretation (see e.g. Serre [Ser79, §X.6]): central simple algebras correspond to
Brauer–Severi varieties—for each degree 𝑛 ≥ 1, both are parametrized by the Galois
cohomology set 𝐻 1 (Gal(𝐹 sep | 𝐹), PGL𝑛 ).
thus ∗
is positive, and the norm 𝛼 ↦→ Tr(𝛼∗ 𝛼) is (an integer multiple of) the Frobenius
norm on 𝐵.
We will soon see that every positive involution can be derived from the conjugate
transpose as in 8.4.3. First, we reduce to the case where 𝐵 is a semisimple algebra.
Proof. We give two proofs. First, we appeal to Theorem 7.9.4: since the trace pairing
is positive definite, it is nondegenerate and immediately 𝐵 is semisimple.
8.4. POSITIVE INVOLUTIONS 127
For a second (more general) proof, let 𝐽 = rad 𝐵 be the Jacobson radical of 𝐵.
By Lemma 7.4.2, 𝐵 is semisimple if and only if rad 𝐵 = {0}, and by Lemma 7.4.8,
𝐽 = rad 𝐵 is nilpotent. Suppose for purposes of contradiction that 𝐽 ≠ {0}. Then there
exists 𝑛 > 0 such that 𝐽 𝑛 ≠ {0} but 𝐽 𝑛+1 = {0}. Let 𝜖 ∈ 𝐽 be such that 𝜖 𝑛 ≠ 0 but
𝜖 𝑛+1 = 0. The involution gives an isomorphism 𝐵 → 𝐵op taking maximal left ideals
to maximal right ideals and therefore by Corollary 7.4.6 we conclude 𝐽 ∗ = 𝐽. Thus
𝜖 𝑛 𝜖 ∗ = 0 so Tr(𝜖 𝑛 (𝜖 ∗ ) 𝑛 ) = Tr(𝜖 𝑛 (𝜖 𝑛 ) ∗ ) = 0, contradicting that ∗ is positive.
𝛼† = 𝜇−1 𝛼∗ 𝜇
for all 𝛼 ∈ 𝐵.
Proof. First suppose 𝐵 is central over R. Then the involutions † and ∗ give two R-
algebra maps 𝐵 → 𝐵op . By the Skolem–Noether theorem (Main Theorem 7.7.1), there
exists 𝜇 ∈ 𝐵× such that 𝛼† = 𝜇−1 𝛼∗ 𝜇. Since
a contradiction.
A similar argument holds if 𝐵 has center 𝑍 (𝐵) = C. The restriction of an involution
to 𝑍 (𝐵) is either the identity or complex conjugation; the latter holds for the conjugate
transpose involution, as well as for † : if 𝑧 ∈ 𝑍 (𝐵) then Tr(𝑧𝑧 † ) = 𝑛2 (𝑧𝑧 † ) > 0,
and we must have 𝑧 † = 𝑧. So the map 𝛼 ↦→ (𝛼∗ ) † is a C-linear automorphism, and
again there exists 𝜇 ∈ 𝐵× such that 𝛼† = 𝜇−1 𝛼∗ 𝜇. By the same argument, we have
𝜇∗ = 𝑧𝜇 with 𝑧 ∈ C, but now 𝜇 = (𝜇∗ ) ∗ = 𝑧𝑧𝜇 so | 𝑧| = 1. Let 𝑤 2 = 𝑤/𝑤 = 𝑧; then
(𝑤𝜇) ∗ = 𝑤𝜇∗ = 𝑤𝑧𝜇 = 𝑤𝜇. Replacing 𝜇 by 𝑤𝜇, we may take 𝑧 = 1.
Corollary 8.4.10. The only positive involution on a real division algebra is the stan-
dard involution.
It follows from the spectral theorem that the R-linear endomorphism of 𝐵 given by
left-multiplication by 𝜇 on 𝐵 as an R-algebra is diagonalizable (with real eigenvalues)
via a symmetric matrix. We say 𝜇 is positive definite (for ∗ ) if all eigenvalues of 𝜇
are positive. The map 𝛼 ↦→ Tr(𝛼∗ 𝜇𝛼) defines a quadratic form on 𝐵, and 𝜇 is positive
definite if and only this quadratic form is positive definite.
Lemma 8.4.12. Let 𝜇∗ = 𝜇. Then the involution 𝛼† = 𝜇−1 𝛼∗ 𝜇 is positive if and only
if either 𝜇 or −𝜇 is positive definite.
Remark 8.4.17. Beyond the application to endomorphism algebras, Weil [Weil60] has
given a more general point of view on positive involutions, connecting them to the
classical groups. For more on involutions on finite-dimensional algebras over real
closed fields, see work of Munn [Mun2004].
Remark 8.5.2. The involution † : 𝐵 → 𝐵 is called the Rosati involution (and depends
on a choice of polarization 𝜆 : 𝐴 → 𝐴∨ , where 𝐴∨ is the dual abelian variety).
Now Lemma 8.4.5 and Proposition 8.5.1 imply that 𝐵 is semisimple as a Q-algebra,
with
Ö𝑟
𝐵' M𝑛𝑖 (𝐷 𝑖 )
𝑖=1
𝐴1𝑛1 × · · · × 𝐴𝑟𝑛𝑟
𝐾0 := 𝐾 h†i = {𝑎 ∈ 𝐾 : 𝑎 † = 𝑎}
Lemma 8.5.3. 𝐾0 is a totally real number field, i.e., every embedding 𝐾0 ↩→ C factors
through R, and if † acts nontrivially on 𝐾, then 𝐾 is a CM field, i.e., 𝐾 is a totally
imaginary extension of 𝐾0 .
130 CHAPTER 8. SIMPLE ALGEBRAS AND INVOLUTIONS
𝐷 ⊗Q R ' M2 (R) 𝑛 ,
𝐷 ⊗Q R ' H𝑛 ,
Proof. We have assembled many of the tools needed to prove this theorem, and
hopefully motivated its statement sufficiently well—but unfortunately, a proof remains
just out of reach: we require some results about quaternion algebras over number
fields not yet in our grasp. For a proof, see Mumford [Mum70, Application I, §21] or
Birkenhake–Lange [BL2004, §§5.3–5.5].
To connect a few dots as well as we can right now, we give a sketch in the case
where 𝐾 = 𝐾0 for the reader who is willing to flip ahead to Chapter 14. In this case,
𝐷 is a central division algebra over 𝐾 = 𝐾0 and has a 𝐾0 -linear involution giving an
isomorphism 𝐷 −→ ∼ 𝐷 op of 𝐾 -algebras. Looking in the Brauer group Br(𝐾 ), we
0 0
conclude that [𝐷] = [𝐷 op ] = [𝐷] −1 , so [𝐷] ∈ Br(𝐾0 ) has order at most 2. By class
field theory (see Remark 14.6.10), we conclude that either 𝐷 = 𝐾0 or 𝐷 is a (division)
quaternion algebra over 𝐾0 . If 𝐷 = 𝐾0 , weÎare in case (I), so suppose 𝐷 is a quaternion
algebra over 𝐾0 . We have 𝐷 ⊗Q R ' 𝑣 |∞ 𝐷 𝑣 a direct product of 𝑛 quaternion
algebras 𝐷 𝑣 over R indexed by the real places 𝑣 of 𝐾0 . We have 𝐷 𝑣 ' M2 (R) or
𝐷 𝑣 ' H, and our positive involution induces a corresponding positive involution on
each 𝐷 𝑣 . If there exists 𝑣 such that 𝐷 𝑣 ' H, then by Corollary 8.4.10, the positive
involution on 𝐷 𝑣 is the standard involution, so it is so on 𝐷, and then all components
must have 𝐷 𝑣 ' H as the standard involution is not positive on M2 (R)—and we are in
case (II). Otherwise, we are in case (III), with Proposition 8.4.7 and Example 8.4.15
characterizing the positive involution.
8.5. ∗ ENDOMORPHISM ALGEBRAS OF ABELIAN VARIETIES 131
Exercises
Let 𝐹 be a field.
8. Let 𝐵 ∈ CSA(𝐹) and suppose that 𝐵 has an involution (not necessarily standard).
Show that [𝐵] has order at most 2 in Br 𝐹.
9. Let 𝐾 ⊇ 𝐹 be a field extension. Show that the map 𝐴 ↦→ 𝐴 ⊗𝐹 𝐾 induces
a group homomorphism Br 𝐹 → Br 𝐾. Conclude that the set of isomorphism
classes of central division 𝐹-algebras 𝐷 such that 𝐷 ⊗𝐹 𝐾 ' M𝑛 (𝐾) for some
𝑛 ≥ 1 forms a subgroup of Br 𝐹, called the relative Brauer group Br(𝐾 | 𝐹).
10. In this exercise, we give an example of a central simple algebra of infinite
dimension, called the Weyl algebra.
Suppose char 𝐹 = 0, let 𝐹 [𝑥] be the polynomial ring over 𝐹 in the variable 𝑥.
Inside the enormous algebra End𝐹 𝐹 [𝑥] is the operator 𝑓 (𝑥) ↦→ 𝑥 𝑓 (𝑥), denoted
also 𝑥, and the differentiation operator 𝛿 : 𝐹 [𝑥] → 𝐹 [𝑥]. These two operators
are related by the product rule:
Arithmetic
133
Chapter 9
In many ways, quaternion algebras are like “noncommutative quadratic field exten-
sions”: this is apparent from their very definition, but also
√ from their description as
wannabe 2 × 2-matrices. Just as the quadratic fields Q( 𝑑) are wonderously rich,
so too are their noncommutative analogues. In this part of the text, we explore these
beginnings of noncommutative algebraic number theory.
In this chapter, we begin with some prerequisites from commutative algebra,
embarking on a study of integral structures and linear algebra over domains.
Z ( 𝑝) := {𝑎/𝑏 ∈ Q : 𝑝 - 𝑏} ⊂ Q.
In the localization, we can focus on those aspects of the lattice concentrated at the
prime 𝑝. Extending scalars, 𝑀 ( 𝑝) := 𝑀Z ( 𝑝) ⊆ 𝑉 is a Z ( 𝑝) -lattice in 𝑉, again called the
localization of 𝑀 at 𝑝. These localizations determine the lattice 𝑀 in the following
strong sense (Theorem 9.4.9).
135
136 CHAPTER 9. LATTICES AND INTEGRAL QUADRATIC FORMS
Condition (i) explains (partly) the ‘quadratic’ nature of the map, and part (ii) is the usual
way relating norms (quadratic forms) to bilinear forms. Choosing a basis 𝑒 1 , . . . , 𝑒 𝑛
for 𝑀 ' Z𝑛 , we may then write
9.3 Lattices
Let 𝑉 be a finite-dimensional 𝐹-vector space.
Definition 9.3.1. An 𝑅-lattice in 𝑉 is a finitely generated 𝑅-submodule 𝑀 ⊆ 𝑉 with
𝑀 𝐹 = 𝑉. We refer to a Z-lattice as a lattice.
The condition that 𝑀 𝐹 = 𝑉 is equivalent to the requirement that 𝑀 contains a
basis for 𝑉 as an 𝐹-vector space.
Example 9.3.2. An 𝑅-lattice in 𝑉 = 𝐹 is the same thing as a fractional ideal of 𝑅.
We will be primarily concerned with projective 𝑅-lattices; if 𝑅 is a Dedekind
domain, then a finitely generated 𝑅-submodule 𝑀 ⊆ 𝑉 is torsion free and hence
automatically projective (9.2.4).
9.3.3. If there is no ambient vector space around, we will also call a finitely generated
torsion free 𝑅-module 𝑀 an 𝑅-lattice: in this case, 𝑀 is a lattice in the 𝐹-vector space
𝑀 ⊗𝑅 𝐹 because the map 𝑀 ↩→ 𝑀 ⊗𝑅 𝐹 is injective (as 𝑀 is torsion free).
Remark 9.3.4. Some authors omit the second condition in the definition of an 𝑅-lattice
and say that 𝑀 is full if 𝑀 𝐹 = 𝑉. We will not encounter 𝑅-lattices that are not full
(and when we do, we call them finitely generated 𝑅-submodules), so we avoid this
added nomenclature.
By definition, an 𝑅-lattice can be thought of an 𝑅-submodule that “allows bounded
denominators”, as follows.
Lemma 9.3.5. Let 𝑀 ⊆ 𝑉 be an 𝑅-lattice and let 𝐽 ⊆ 𝑉 be a finitely generated
𝑅-submodule. Then the following statements hold.
(9.3.8)
0 /𝑊 /𝑉 / 𝑉/𝑊 /0
𝑀 = 𝔞1 𝑥1 ⊕ · · · ⊕ 𝔞𝑛 𝑥 𝑛
𝑁 = 𝔟1 𝑥1 ⊕ · · · ⊕ 𝔟𝑛 𝑥 𝑛
The unique fractional ideals 𝔡1 , . . . , 𝔡𝑛 given by Theorem 9.3.9 are called the
invariant factors of 𝑁 relative to 𝑀.
𝑀 = 𝑅𝑥1 ⊕ · · · ⊕ 𝑅𝑥 𝑛−1 ⊕ 𝔞𝑥 𝑛
9.4 Localizations
Properties of a domain are governed in an important way by its localizations, and
consequently the structure of lattices, orders, and algebras can often be understood by
looking at their localizations (and later, completions).
For a prime ideal 𝔭 ⊆ 𝑅, we denote by
the localization of 𝑅 at 𝔭. (We reserve the simpler subscript notation for the completion,
defined in section 9.5.)
where the intersections are over all prime ideals of 𝑅 and all maximal ideals of 𝑅,
respectively.
Let 𝑉 be a finite-dimensional 𝐹-vector space and let 𝑀 ⊆ 𝑉 be an 𝑅-lattice. For a
prime 𝔭 of 𝑅, let
𝑀 (𝔭) := 𝑀 𝑅 (𝔭) ⊆ 𝑉
140 CHAPTER 9. LATTICES AND INTEGRAL QUADRATIC FORMS
𝑀 (𝔭) := 𝑀 ⊗𝑅 𝑅 (𝔭) .
(𝑀 : 𝑥) := {𝑟 ∈ 𝑅 : 𝑟𝑥 ∈ 𝑀 }.
Ñ(i) ⇒ (ii) ⇒ (iii) are direct; for the implication (iii) ⇒ (i), we
Proof. TheÑimplications
have 𝑀 = 𝔭 𝑀 (𝔪) ⊆ 𝔭 𝑁 (𝔪) = 𝑁 by Lemma 9.4.6.
9.4.8. A property that holds if and only if it holds locally (as in Corollary 9.4.7, for
the property that one lattice is contained in another) is called a local property.
9.5 Completions
Next, we briefly define the completion and show that the local-global dictionary holds
in this context as well. (We will consider completions in the context of local fields
more generally starting in chapter 12, so the reader may wish to return to this section
later.) For a general reference on completions (and the induced topology), see e.g.
142 CHAPTER 9. LATTICES AND INTEGRAL QUADRATIC FORMS
map is injective. Moreover, since 𝔭 is maximal, then in fact this inclusion factors via
𝑅 ↩→ 𝑅 (𝔭) ↩→ 𝑅𝔭 inducing isomorphisms (Exercise 9.8)
∼ 𝑅 /𝔭𝑒 𝑅 −→
𝑅/𝔭𝑒 −→ ∼ 𝑅 /𝔭𝑒 𝑅 (9.5.2)
(𝔭) (𝔭) 𝔭 𝔭
In particular, Lemma 9.5.3 implies that in the local-global dictionary for lattices
over a Dedekind domain 𝑅 (Theorem 9.4.9), we may also work with collections of
𝑅𝔭 -lattices (𝑁𝔭 )𝔭 over the completions at primes.
9.6 Index
Continuing with 𝑅 a noetherian domain, let 𝑀, 𝑁 ⊆ 𝑉 be 𝑅-lattices.
The style of Definition 9.6.1, given by a large generating set (9.6.2), is the replace-
ment for being able to work with given bases; this style will be typical for us in what
follows. The determinants det(𝛿) are meant in the intrinsic sense, but can be computed
as the determinant of a matrix upon choosing a basis for 𝑉.
Lemma 9.6.3. The index [𝑀 : 𝑁] 𝑅 is a nonzero 𝑅-module, and if 𝛼 ∈ Aut𝐹 (𝑉) then
[𝛼𝑀 : 𝑁] = det(𝛼) −1 [𝑀 : 𝑁].
Proof. If 𝛿(𝑀) ⊆ 𝑁, then 𝛿(𝑀 (𝔭) ) ⊆ 𝑁 (𝔭) by 𝑅 (𝔭) -linearity, giving the inclusion
(⊇). For (⊆), let 𝛿 ∈ End𝐹 (𝑉) be such that 𝛿(𝑀 (𝔭) ) ⊆ 𝑁 (𝔭) . For any 𝑥 ∈ 𝑀, we
have 𝛿(𝑥) ∈ 𝛿(𝑀) ⊆ 𝛿(𝑀 (𝔭) ) ⊆ 𝑁 (𝔭) , so there exists 𝑦 ∈ 𝑁 and 𝑠 ∈ 𝑅 r 𝔭 such
that 𝑠𝛿(𝑥) = 𝑦 ∈ 𝑁. Let 𝑥1 , . . . , 𝑥 𝑚 generate 𝑀 as an 𝑅-module, and for each 𝑖,
Î
let 𝑠𝑖 ∈ 𝑅 r 𝔭 be such that 𝑠𝑖 𝛿(𝑥 𝑖 ) ∈ 𝑁. Let 𝑠 := 𝑖 𝑠𝑖 . Then 𝑠𝛿(𝑀) ⊆ 𝑁, so
det(𝑠𝛿) = 𝑠 𝑛 det(𝛿) ∈ [𝑀 : 𝑁] 𝑅 , if 𝑛 := dim𝐹 𝑉. Finally, 𝑠 ∈ 𝑅 ×(𝔭) , we conclude that
det 𝛿 ∈ ( [𝑀 : 𝑁] 𝑅 ) (𝔭) , as desired.
Proposition 9.6.8. Suppose that 𝑀, 𝑁 are projective 𝑅-modules. Then [𝑀 : 𝑁] 𝑅 is a
projective 𝑅-module. Moreover, if 𝑁 ⊆ 𝑀 then [𝑀 : 𝑁] 𝑅 = 𝑅 if and only if 𝑀 = 𝑁.
Proof. Let 𝔭 be a prime of 𝑅 and consider the localization ( [𝑀 : 𝑁] 𝑅 ) (𝔭) at 𝔭. Since
𝑀, 𝑁 are projective 𝑅-modules, they are locally free (9.2.1). By Lemma 9.6.4, the
local index [𝑀 (𝔭) : 𝑁 (𝔭) ] 𝑅 (𝔭) is a principal 𝑅 (𝔭) -ideal. By Lemma 9.6.7, we conclude
that [𝑀 : 𝑁] 𝑅 is locally principal, therefore projective.
The second statement follows in a similar way: we may suppose that 𝑅 is local and
thus 𝑁 ⊆ 𝑀 are free, in which case 𝑀 = 𝑁 if and only if a change of basis matrix
from 𝑁 to 𝑀 has determinant in 𝑅 × .
For Dedekind domains, the 𝑅-index can be described as follows.
Lemma 9.6.9. If 𝑅 is a Dedekind domain and 𝑁 ⊆ 𝑀, then [𝑀 : 𝑁] 𝑅 is the product
of the invariant factors (or elementary divisors) of the torsion 𝑅-module 𝑀/𝑁.
Proof. Exercise 9.11.
𝑀
𝑄
/𝐿
o 𝑓 o ℎ (9.7.7)
𝑄0
𝑀0 / 𝐿0
𝑇 : 𝑀 → Hom𝑅 (𝑀, 𝐿)
(9.7.9)
𝑥 ↦→ (𝑦 ↦→ 𝑇 (𝑥, 𝑦))
𝑄 𝐹 : 𝑀 ⊗𝑅 𝐹 → 𝐿 ⊗𝑅 𝐹 ' 𝐹
is nondegenerate, since the kernel can be detected over 𝐹. Recalling the definition
of discriminant (Definition 4.3.3 for char 𝐹 ≠ 2 and Definition 6.3.1 in general), we
conclude that 𝑄 is nondegenerate if and only if disc 𝑄 𝐹 ≠ 0.
146 CHAPTER 9. LATTICES AND INTEGRAL QUADRATIC FORMS
then 𝑄 is primitive if and only if the coefficients 𝑎 𝑖 𝑗 generate the unit ideal 𝑅.
If 𝑅 is a Dedekind domain, then 𝑄(𝑀) ⊆ 𝐿 is again projective (locally at a prime
generated by an element of minimal valuation), so one can always replace 𝑄 : 𝑀 → 𝐿
by 𝑄 : 𝑀 → 𝑄(𝑀) to get a primitive quadratic module; when 𝑅 is a PID, up to
similarity we may divide through by greatest common divisor of the coefficients 𝑎 𝑖 𝑗
in the previous paragraph.
9.7.16. In our admittedly abstract treatment of quadratic modules so far, we have
specifically allowed the codomain of the quadratic map to vary at the same time as the
domain—in particular, we do not ask that they necessarily take values in 𝑅.
Remark 9.7.17. In certain lattice contexts with 𝑅 a Dedekind domain, a quadratic form
with values in a fractional ideal 𝔞 is called an 𝔞-modular quadratic form. Given the
overloaded meanings of the word modular, we do not employ this terminology. In the
geometric context, a quadratic module is called a line-bundle valued quadratic form.
Whatever the terminology, we will see in Chapter 22 that it is important to keep track
of the codomain of the quadratic map just as much as the domain, and in particular we
cannot assume that either is free when 𝑅 is not a PID.
In case (ii), we necessarily have 𝑣(2) > 0 and 𝑣(𝑏 2 − 4𝑎𝑐) = 2𝑣(𝑏).
Example 9.8.2. Suppose 𝑅 = Z2 is the ring of 2-adic integers, so that 𝑣(𝑥) = ord2 (𝑥)
is the largest power of 2 dividing 𝑥 ∈ Z2 . Recall that Z×2 /Z×2
2 is represented by the
elements ±1, ±5, therefore a quadratic form 𝑄 over Z2 is atomic of type (i) above
if and only if 𝑄(𝑥) ' ±𝑥 2 or 𝑄(𝑥) ' ±5𝑥 2 . For forms of type (ii), the conditions
𝑣(𝑏) < 𝑣(2𝑎) = 𝑣(𝑎) + 1 and 𝑣(𝑎)𝑣(𝑏) = 0 imply 𝑣(𝑏) = 0, and so a quadratic
form 𝑄 over Z2 is atomic of type (ii) if and only if 𝑄(𝑥, 𝑦) ' 𝑎𝑥 2 + 𝑥𝑦 + 𝑐𝑦 2 with
ord2 (𝑎) ≤ ord2 (𝑐). Replacing 𝑥 by 𝑢𝑥 and 𝑦 by 𝑢 −1 𝑦 for 𝑢 ∈ Z×2 we may suppose
𝑎 = ±2𝑡 or 𝑎 = ±5 · 2𝑡 with 𝑡 ≥ 0, and then the atomic representative [𝑎, 1, 𝑐] of the
isomorphism class of 𝑄 is unique.
Proof. If 𝑄 is atomic of type (i) then the space underlying 𝑄 has rank 1 and is therefore
indecomposable. Suppose 𝑄 = [𝑎, 𝑏, 𝑐] is atomic of type (ii) and assume for purposes
of contradiction that 𝑄 is decomposable. It follows that if 𝑥, 𝑦 ∈ 𝑀 then 𝑇 (𝑥, 𝑦) ∈ 2𝑅.
Thus we cannot have 𝑣(𝑏) = 0, so 𝑣(𝑎) = 0, and further 𝑣(𝑏) ≥ 𝑣(2) = 𝑣(2𝑎); this
contradicts the fact that 𝑄 is atomic.
𝑄 ' 𝜋 𝑒1 𝑄 1 · · · 𝜋 𝑒𝑛 𝑄 𝑛
Proof. When 𝑅 = 𝐹 is a field with char 𝐹 ≠ 2, we are applying the standard method
of Gram–Schmidt orthogonalization to diagonalize the quadratic form. This argument
can be adapted to the case where 𝑅 = 𝐹 is a field with char 𝐹 = 2, see e.g. Scharlau
[Scha85, §9.4]. For the general case, we make further adaptations to this procedure:
see Voight [Voi2013, Algorithm 3.12] for a constructive (algorithmic) approach.
Exercises
Let 𝑅 be a noetherian domain with field of fractions 𝐹 := Frac 𝑅.
𝑅 ↩→ 𝑅 (𝔪) ↩→ 𝑅𝔪
from the domain into its localization into the completion, inducing isomorphisms
∼ 𝑅 /𝔭𝑒 𝑅 −→
𝑅/𝔭𝑒 −→ ∼ 𝑅 /𝔭𝑒 𝑅
(𝔭) (𝔭) 𝔭 𝔭
for all 𝑒 ≥ 1.
9. Let 𝑉 be a finite-dimensional 𝐹-vector space and let 𝑀, 𝑁 ⊆ 𝑉 be 𝑅-lattices.
9.8. NORMALIZED FORM 149
(a) Show that the index [𝑀 : 𝑁] 𝑅 is a nonzero 𝑅-module. [Hint: use Lemma
9.3.5.]
(b) For 𝛼 ∈ Aut𝐹 (𝑉), show [𝛼𝑀 : 𝑁] = det(𝛼) −1 [𝑀 : 𝑁].
10. Find 𝑅-lattices 𝑀, 𝑁 ⊆ 𝑉 such that [𝑀 : 𝑁] 𝑅 = 𝑅 but 𝑀 ≠ 𝑁.
11. Prove Lemma 9.6.9, as follows. Suppose 𝑅 is a Dedekind domain, and let
𝑁 ⊆ 𝑀 ⊆ 𝑉 be 𝑅-lattices in a finite-dimensional vector space 𝑉 over 𝐹. Prove
that [𝑀 : 𝑁] 𝑅 is the product of the invariant factors (or elementary divisors) of
the torsion 𝑅-module 𝑀/𝑁.
12. Suppose 𝑅 is a Dedekind domain. Let 𝑄 : 𝑉 → 𝐹 be a quadratic form over 𝐹,
let 𝑀 ⊆ 𝑉 be an 𝑅-lattice, and let 𝐿 := 𝑄(𝑀) ⊆ 𝐹 be the 𝑅-submodule of 𝐹
generated by the values of 𝑄. Show that 𝐿 is a fractional 𝑅-ideal.
13. Consider the ternary quadratic form 𝑄(𝑥, 𝑦, 𝑧) = 𝑥𝑦 + 𝑥𝑧 over Z2 . Compute a
normalized form for 𝑄.
14. Consider the following ‘counterexamples’ to Theorem 9.4.9 for more general
integral domains as follows. Let 𝑅 = Q[𝑥, 𝑦] be the polynomial ring in two
variables over Q, so that 𝐹 = Q(𝑥, 𝑦). Let 𝑉 = 𝐹 and 𝐼 = 𝑅.
(a) Show that 𝑦𝑅 has the property that 𝑦𝑅𝔭 ≠ 𝑅𝔭 for infinitely many prime
ideals 𝔭 of 𝑅.
(b) Consider the collection of lattices given by 𝐽𝔭 = 𝑓 (𝑥)𝑅𝔭 if 𝔭 = (𝑦, 𝑓 (𝑥))
where 𝑓 (𝑥) ∈ Q[𝑥] is irreducible and 𝐽𝔭 = 𝑅𝔭 otherwise. Show that
Ñ
𝔭 𝐽𝔭 = (0).
Orders
151
152 CHAPTER 10. ORDERS
10.2 Orders
Throughout, let 𝑅 be a domain with field of fractions 𝐹 := Frac(𝑅), and let 𝐵 be
a finite-dimensional 𝐹-algebra. For further reference about orders (as lattices), see
Reiner [Rei2003, Chapter 2] and Curtis–Reiner [CR81, §§23, 26].
ExampleÉ10.2.3. The matrix algebra M𝑛 (𝐹) has the 𝑅-order M𝑛 (𝑅). The subring
𝑅[𝐺] = 𝑔 ∈𝐺 𝑅𝑔 is an 𝑅-order in the group ring 𝐹 [𝐺].
Example 10.2.4. Let 𝑎, 𝑏 ∈ 𝑅r{0} and consider the quaternion algebra 𝐵 = (𝑎, 𝑏 | 𝐹).
Then O = 𝑅 ⊕ 𝑅𝑖 ⊕ 𝑅 𝑗 ⊕ 𝑅𝑘 is an 𝑅-order, because it is closed under multiplication
(e.g., 𝑖𝑘 = 𝑖(𝑖 𝑗) = 𝑎 𝑗 ∈ O).
OR (𝐼) := {𝛼 ∈ 𝐵 : 𝐼𝛼 ⊆ 𝐼}.
Example 10.2.9. It follows from Lemma 10.2.7 that 𝐵 has an 𝑅-order: the 𝑅-span of
an 𝐹-basis for 𝐵 defines an 𝑅-lattice, so OL (𝐼) is an 𝑅-order. (This is a nice way of
“clearing denominators” from a multiplication table to obtain an order.)
We can read other properties about lattices from their localizations, such as in the
following lemma.
(i) 𝐼 is an 𝑅-order;
(ii) 𝐼 (𝔭) is an 𝑅 (𝔭) -order for all primes 𝔭 of 𝑅; and
(iii) 𝐼 (𝔪) is an 𝑅 (𝔪) -order for all maximal ideals 𝔪 of 𝑅.
Proof. For (i) ⇒ (ii) ⇒ (iii), if 𝐼 is an 𝑅-order then 𝐼 (𝔭) is an 𝑅 (𝔭) -order for all primes
𝔭, hence a fortiori for all maximal ideals 𝔪.
To conclude, we prove Ñ (iii) ⇒ (i), and suppose that 𝐼 (𝔪) is an Ñ 𝑅 (𝔪) -order for all
maximal ideals 𝔪. Then Ñ 𝐼
𝔪 (𝔪) = 𝐼 by Lemma 9.4.6. Thus 1 ∈ 𝔪 𝐼 (𝔪) = 𝐼, and for
all 𝛼, 𝛽 ∈ 𝐼 we have 𝛼𝛽 ∈ 𝔪 𝐼 (𝔪) = 𝐼, so 𝐼 is a subring of 𝐵 and hence an order.
Remark 10.2.11. The hypothesis that 𝑅 is noetherian is used in Lemma 10.2.7, but it is
not actually needed; the fact that OL (𝐼) is an order follows by a process often referred
to as noetherian reduction. A basis of 𝐵 yields a multiplication table, consisting
of finitely many elements of 𝐹; moreover, we know that 𝐼 is finitely generated as an
𝑅-module. Writing these generators in terms of a basis we can express these generators
over the basis using finitely many elements of 𝐹. Let 𝑅0 be the subring of 𝑅 generated
by these finitely elements, with field of fractions 𝐹0 , let 𝐵0 be the 𝐹0 -algebra with the
same multiplication table as 𝐵; let 𝐼0 be the 𝑅0 -submodule generated by the generators
for 𝐼 written over 𝑅0 . Then 𝐵 = 𝐵0 ⊗𝐹0 𝐹 and 𝐼 = 𝐼0 ⊗𝑅0 𝑅. But now 𝑅0 is a finitely
generated commutative algebra over its prime ring (the subring generated by 1), so by
the Hilbert basis theorem, 𝑅0 is noetherian. The argument given then shows that 𝐼0 is
finitely generated as an 𝑅0 -module, whence 𝐼 is finitely generated as an 𝑅-module.
Noetherian reduction applies to many results in this text, but non-noetherian rings
are not our primary concern; we retain the noetherian hypothesis for simplicity of
argument and encourage the interested reader to seek generalizations (when they are
possible).
154 CHAPTER 10. ORDERS
10.3 Integrality
Orders are composed of integral elements, defined as follows. If 𝛼 ∈ 𝐵, we denote by
Í
𝑅[𝛼] = 𝑑 𝑅𝛼 𝑑 the (commutative) 𝑅-subalgebra of 𝐵 generated by 𝛼.
Proof. This lemma is standard; the only extra detail here is to note that in (iii) we do
not need to assume that the subring 𝐴 is commutative: (ii) ⇒ (iii) is immediate taking
𝐴 = 𝑅[𝛼], and for the converse, if 𝐴 ⊆ 𝐵 is a subring that is finitely generated as an
𝑅-module, then 𝑅[𝛼] ⊆ 𝐴 and since 𝑅 is noetherian and 𝐴 is finitely generated as an
𝑅-module, it follows that 𝑅[𝛼] is also finitely generated as an 𝑅-module.
Proof. Let 𝑓 (𝑥) ∈ 𝑅[𝑥] be a monic polynomial that 𝛼 satisfies, and let 𝑔(𝑥) ∈ 𝐹 [𝑥] be
the minimal polynomial of 𝛼. Let 𝐾 be a splitting field for 𝑔(𝑥), and let 𝛼1 , . . . , 𝛼𝑛 be
the roots of 𝑔(𝑥) in 𝐾. Since 𝑔(𝑥) | 𝑓 (𝑥), each such 𝛼𝑖 is integral over 𝑅, and the set
of elements in 𝐾 integral over 𝑅 forms a ring, so each coefficient of 𝑔 is integral over
𝑅 and belongs to 𝐹; but since 𝑅 is integrally closed, these coefficients must belong to
𝑅 and 𝑔(𝑥) ∈ 𝑅[𝑥].
Proof. If O (𝔭) is maximal for each prime 𝔭 then by Corollary 9.4.7 we see that O is
maximal. Conversely, suppose O is maximal and suppose that O (𝔭) ( O0(𝔭) is a proper
containment of orders for some nonzero prime 𝔭. Then the set O0 = 0
Ñ
𝔮≠𝔭 O𝔮 ∩ O (𝔭)
is an 𝑅-order properly containing O by Lemma 10.2.10 and Theorem 9.4.9.
Lemma 10.4.4. Let O ⊂ 𝐵 be an 𝑅-order. Then for all but finitely many primes 𝔭 of
𝑅, we have that O (𝔭) = O ⊗𝑅 𝑅 (𝔭) is maximal.
𝔞 𝑗 𝔞−1
𝑖 ' Hom 𝑅 (𝔞𝑖 , 𝔞 𝑗 ) ⊆ Hom𝐹 (𝐹, 𝐹) ' 𝐹
𝔞2 𝔞−1
𝑅 1
End𝑅 (𝑀) ' ⊆ M2 (𝐹).
𝔞1 𝔞−1
2 𝑅
Proof. As in the proof of Lemma 10.2.7, we conclude that End𝑅 (𝑀)𝐹 = 𝐵. Let
𝛼1 , 𝛼2 , . . . , 𝛼𝑛 be an 𝐹-basis for 𝑉 and let 𝑁 = 𝑅𝛼1 ⊕ · · · ⊕ 𝑅𝛼𝑛 . Thus End𝑅 (𝑁) '
M𝑛 (𝑅) is finitely generated as an 𝑅-module.
By Lemma 9.3.5 there exists nonzero 𝑟 ∈ 𝑅 such that 𝑟 𝑁 ⊆ 𝑀 ⊆ 𝑟 −1 𝑁. Therefore,
if 𝜙 ∈ End𝑅 (𝑀), so that 𝜙(𝑀) ⊆ 𝑀, then
and thus End𝑅 (𝑀) ⊆ 𝑟 −2 End𝑅 (𝑁); since 𝑅 is noetherian, this implies that End𝑅 (𝑀)
is finitely generated as an 𝑅-module and End𝑅 (𝑀) is an 𝑅-order in 𝐵.
10.5. ORDERS IN A MATRIX RING 157
Lemma 10.5.4. Let O ⊆ 𝐵 = End𝐹 (𝑉) be an 𝑅-order. Then O ⊆ End𝑅 (𝑀) for some
𝑅-lattice 𝑀 ⊆ 𝑉. In particular, if O ⊆ 𝐵 is a maximal 𝑅-order, then O = End𝑅 (𝑀)
for some 𝑅-lattice 𝑀.
Proof. The Éisomorphism 𝐵 ' M𝑛 (𝐹) arises from a choice of basis 𝑥1 , . . . , 𝑥 𝑛 for 𝑉;
𝑛
letting 𝑁 = 𝑖=1 𝑅𝑥 𝑖 we have End 𝑅 (𝑁) ' M𝑛 (𝑅). The 𝑅-order M𝑛 (𝑅) is maximal
by Exercise 10.6, since a PID is integrally closed.
By Lemma 10.5.4, we have O ⊆ End𝑅 (𝑀) for some 𝑅-lattice 𝑀 ⊆ 𝑉, so if O is
maximal then O = End𝑅 (𝑀). If 𝑅 is a PID then 𝑀 is free as an 𝑅-module, and we
can write 𝑀 = 𝑅𝑦 1 ⊕ · · · ⊕ 𝑅𝑦 𝑛 ; the change of basis matrix from 𝑥 𝑖 to 𝑦 𝑖 then realizes
End𝑅 (𝑀) as a conjugate of End𝑅 (𝑁) ' M𝑛 (𝑅).
Exercises
Let 𝑅 be a noetherian domain with field of fractions 𝐹.
With the preceding chapters on lattices and orders in hand, we are now prepared
to embark on a general treatment of quaternion algebras over number fields and the
arithmetic of their orders. Before we do so, for motivation and pure enjoyment, in this
chapter we consider the special case of the Hurwitz order. Not only is this appropriate
in a historical spirit, it is also instructive for what follows; moreover, the Hurwitz order
has certain exceptional symmetries that make it worthy of specific investigation.
159
160 CHAPTER 11. THE HURWITZ ORDER
in 𝐵 is the unique order that properly contains Zh𝑖, 𝑗i, and O is maximal.
The order O in (11.1.3) is called the Hurwitz order, and it contains Zh𝑖, 𝑗i with
index 2. Note that if 𝛼 ∈ O, then 𝛼 ∈ Zh𝑖, 𝑗i if and only if trd(𝛼) ∈ 2Z.
11.1.4. We can recast this calculation in terms of the local-global dictionary for
lattices (Theorem 9.1.1). Since O[ 12 ] = Zh𝑖, 𝑗i [ 12 ], for every odd prime 𝑝 we have
O ( 𝑝) = Zh𝑖, 𝑗i ( 𝑝) , and O (2) ) Zh𝑖, 𝑗i (2) .
O/3O → M2 (F3 )
0 −1 1 1
𝑖, 𝑗 ↦→ ,
1 0 1 −1
11.2.2. The group SL2 (F3 ) acts on the left on the set of nonzero column vectors
F23 up to sign, a set of cardinality (9 − 1)/2 = 4. (More generally, SL2 (F 𝑝 ) acts
on P1 (F 𝑝 ) := (F2𝑝 r {(0, 0)})/F×𝑝 , a set of cardinality 𝑝 + 1.) This action yields a
permutation representation SL2 (F3 ) → 𝑆4 ; the kernel of this map is the subgroup
generated by the scalar matrix −1 and so the representation gives an injective group
homomorphism from PSL2 (F3 ) := SL2 (F3 )/{±1} into 𝑆4 . Since 𝐴4 ≤ 𝑆4 is the
unique subgroup of size 24/2 = 12, we must have PSL2 (F3 ) ' 𝐴4 , giving an exact
sequence
1 → {±1} → O× → 𝐴4 → 1. (11.2.3)
11.2.4. We can also visualize the group O× and the exact sequence (11.2.3), thinking
of the Hamiltonians as acting by rotations (section 2.4). Recall there is an exact
sequence (Corollary 2.4.21)
−i − j + k
i+j+k
j
i −i + j − k y
i−j−k
x
Figure 11.2.7: Symmetries of a tetrahedron, viewed quaternionically
Inside the cube in R3 with vertices (±1, ±1, ±1) = ±𝑖 ± 𝑗 ± 𝑘, we can find four
inscribed tetrahedra, for example, the tetrahedron 𝑇 with vertices
𝑖 + 𝑗 + 𝑘, 𝑖 − 𝑗 − 𝑘, −𝑖 + 𝑗 − 𝑘, −𝑖 − 𝑗 + 𝑘.
Then the elements ±𝑖, ± 𝑗, ±𝑘 act by rotation about the 𝑥, 𝑦, 𝑧 axes by an angle 𝜋 (so
interchanging points with the same 𝑥, 𝑦, 𝑧 coordinate). The element ±𝜔 = ±(−1 + 𝑖 +
𝑗 + 𝑘)/2 rotates by the angle 2𝜋/3 fixing the point (1, 1, 1) and cyclically permuting the
other three points, and by symmetry we understand the action of the other elements of
O× . We therefore call O× the binary tetrahedral group. Following Conway–Smith
[CSm2003, §3.3], we also write 2𝑇 = O× for this group; the notation 𝐴 e4 is also used.
The subgroup 𝑄 8 E 2𝑇 is normal (as it is characteristic, consisting of all elements
of O of order dividing 4), and so we can write 2𝑇 = 𝑄 8 o h𝜔i where h𝜔i ' Z/3Z acts
on 𝑄 8 by conjugation, cyclically rotating the elements 𝑖, 𝑗, 𝑘. Finally, the group 2𝑇
has a presentation (Exercise 11.7)
We conclude by noting that the difference between the Lipschitz and Hurwitz
orders is “covered” by the extra units.
Lemma 11.2.9. For every 𝛽 ∈ O, there exists 𝛾 ∈ O× such that 𝛽𝛾 ∈ Zh𝑖, 𝑗i.
11.3. ⊲ EUCLIDEAN ALGORITHM 163
11.3.1. The Hurwitz order also has a left (or right) Euclidean algorithm generalizing
the commutative case, as follows. There is an embedding 𝐵 ↩→ 𝐵 ⊗Q R ' H, and
inside H ' R4 the Hurwitz order sits as a (Z-)lattice equipped with the Euclidean
inner product, so we can think of the reduced norm by instead thinking of distance.
In the Lipschitz order, we see by rounding coordinates that for all 𝛾 ∈ 𝐵 there exists
𝜇 ∈ Zh𝑖, 𝑗i such that nrd(𝛾 − 𝜇) ≤ 4 · (1/2) 2 = 1—a farthest point occurs at the
center (1/2, 1/2, 1/2, 1/2) of a unit cube. But this is precisely the point where the
Hurwitz quaternions occur, and it follows that for all 𝛾 ∈ 𝐵, there exists 𝜇 ∈ O such
that nrd(𝛾 − 𝜇) < 1. (In fact, we can take nrd(𝛾 − 𝜇) ≤ 1/2; see Exercise 11.8.)
Lemma 11.3.2 (Hurwitz order is right norm Euclidean). For all 𝛼, 𝛽 ∈ O with 𝛽 ≠ 0,
there exists 𝜇, 𝜌 ∈ O such that
𝛼 = 𝛽𝜇 + 𝜌 (11.3.3)
and nrd(𝜌) < nrd(𝛽).
A similar statement to Lemma 11.3.2 holds for division on the left, i.e., in (11.3.3)
we may take 𝛼 = 𝜇𝛽 + 𝜌 (with possibly different elements 𝜇, 𝜌 ∈ O, of course).
Proposition 11.3.4. Every right ideal 𝐼 ⊆ O is right principal, i.e., there exists 𝛽 ∈ 𝐼
such that 𝐼 = 𝛽O.
Proof. Let 𝐼 ⊆ O be a right ideal. If 𝐼 = {0}, we are done. Otherwise, there exists an
element 0 ≠ 𝛽 ∈ 𝐼 with minimal reduced norm nrd(𝛽) ∈ Z>0 . We claim that 𝐼 = 𝛽O.
For all 𝛼 ∈ 𝐼, by the left Euclidean algorithm in Lemma 11.3.2, there exists 𝜇 ∈ O
such that 𝛼 = 𝜇𝛽 + 𝜌 with nrd(𝜌) < nrd(𝛽); but 𝜌 = 𝛼 − 𝛽𝜇 ∈ 𝐼, so by minimality,
nrd(𝜌) = 0 and 𝜌 = 0, hence 𝛼 = 𝛽𝜇 ∈ 𝛽O as claimed.
164 CHAPTER 11. THE HURWITZ ORDER
𝐼 = {𝛼 ∈ O : 𝛼 mod 𝑝 ∈ 𝐼 mod 𝑝}
11.4. ⊲ UNIQUE FACTORIZATION 165
Remark 11.4.2. Once we have developed a suitable theory of norms, the proof that
nrd(𝛽) = 𝑝 above will be immediate: if we define N (𝐼) := #(O/𝐼) then N (𝐼) = 𝑝 2 by
construction, and it turns out that N (𝐼) = nrd(𝛽) 2 .
Theorem 11.4.3 (Lagrange). Every integer 𝑛 ≥ 0 is the sum of four squares, i.e., there
exist 𝑡, 𝑥, 𝑦, 𝑧 ∈ Z such that 𝑛 = 𝑡 2 + 𝑥 2 + 𝑦 2 + 𝑧2 .
where 𝛾1 , . . . , 𝛾𝑟 ∈ O× .
the latter isomorphism by Hamilton’s original (!) motivation for quaternion algebras
(Corollary 2.4.21). Therefore Γ/{±1} ⊆ SO(3) is a finite rotation group, and these
groups have been known since antiquity.
Cases (iii)–(v) are the symmetry groups of the corresponding Platonic solids and
are called exceptional rotation groups.
Proof. Let 𝐺 ≤ SO(3) be a finite subgroup with #𝐺 = 𝑛 > 1; then 𝐺 must consist of
rotations about a common fixed point (its center of gravity), which we may take to be
the origin. The group 𝐺 then acts on the unit sphere, and every nonidentity element
of 𝐺 acts by rotation about an axis, fixing the poles of its axis on the sphere. Let 𝑉 be
the subset of these poles in the unit sphere; the set 𝑉 will soon be the vertices of our
(possibly degenerate) polyhedron. Let
Since each 𝑔 ∈ 𝐺 r {1} has exactly two poles, we have #𝑋 = 2(𝑛 − 1). On the other
hand, we can also count organizing by orbits. Choose a representative set 𝑣 1 , . . . , 𝑣 𝑟
of poles, one from each orbit of 𝐺 on 𝑉, and let
𝑛𝑖 = # Stab𝐺 (𝑣 𝑖 ) = #{𝑔 ∈ 𝐺 : 𝑔𝑣 𝑖 = 𝑣 𝑖 }
be the order of the stabilizer: this group is a cyclic subgroup about a common axis.
Then
𝑟 𝑟 𝑟
∑︁ ∑︁ 𝑛 ∑︁ 1
2𝑛 − 2 = #𝑋 = #(𝐺𝑣 𝑖 ) (𝑛𝑖 − 1) = (𝑛𝑖 − 1) = 𝑛 1− ,
𝑖=1
𝑛
𝑖=1 𝑖 𝑖=1
𝑛𝑖
Since 𝑛 > 1, we have 1 ≤ 2 − 2/𝑛 < 2; and since each 𝑛𝑖 ≥ 2, we have 1/2 ≤
1 − 1/𝑛𝑖 < 1. Putting these together, we must have 𝑟 = 2, 3.
If 𝑟 = 2, then (11.5.3) becomes 2 = 𝑛/𝑛1 + 𝑛/𝑛2 , with 𝑛/𝑛𝑖 = #(𝐺𝑣 𝑖 ) ≥ 1, so
𝑛1 = 𝑛2 = 𝑛, there is only one axis of rotation, and 𝐺 is cyclic.
If 𝑟 = 3, then the only possibilities for (𝑛1 , 𝑛2 , 𝑛3 ) with 𝑛1 ≤ 𝑛2 ≤ 𝑛3 are
(2, 2, 𝑐), (2, 3, 3), (2, 3, 4), (2, 3, 5); the corresponding groups have sizes 2𝑐, 12, 24, 60,
respectively, and can be identified with 𝐷 2𝑐 , 𝐴4 , 𝑆4 , 𝐴5 by a careful but classical anal-
ysis of orbits. See Armstrong [Arm88, Chapter 19], Grove–Benson [GB2008, §2.4],
or Conway–Smith [CSm2003, §3.3].
11.5.4. The octahedral group 𝑆4 pulls back to the binary octahedral group 2𝑂 ⊆ H1
of order 24 · 2 = 48, whose elements act by rigid motions of the octahedron (or dually,
the cube). We make identifications following 11.2.4, shown in Figure 11.5.5.
−i − j + k
−i + j + k
k
−i
i−j+k
−j
j
i −i + j − k y
−k
i−j−k
x i+j−k
±𝑖 ± 𝑗 ± 𝑘, ±𝜏𝑖 ± 𝜏 −1 𝑗, ±𝜏 𝑗 ± 𝜏 −1 𝑘, ±𝜏𝑘 ± 𝜏 −1 𝑖
√
where 𝜏 = (1 + 5)/2 is the golden ratio. The elements of order 5 are given by
conjugates and powers of the element 𝜁 = (𝜏 + 𝜏 −1𝑖 + 𝑗)/2, which acts by rotation
about a face. The group 2𝐼 can be presented as
For further references, see Conway–Sloane [CSl88, §8.2], who describe the binary
icosahedral group in detail, calling it the icosian group.
We now consider the related possibilities over Q. (We will return to a general
classification
in section 32.4.) To put ourselves in a situation
like (11.2.6), let 𝐵 =
𝑎, 𝑏 𝑎, 𝑏
be a quaternion algebra over Q such that 𝐵 ⊗Q R = ' H: in this case,
Q R
we say that 𝐵 is definite. By Exercise 2.4, 𝐵 is definite if and only if 𝑎, 𝑏 < 0. Let
O ⊆ 𝐵 be an order in 𝐵; we would like to understand its unit group.
Proof. We may take 𝐵 = (𝑎, 𝑏 | Q) with 𝑎, 𝑏 < 0. Consider the reduced norm
nrd : 𝐵 → Q, given by nrd(𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗) = 𝑡 2 + |𝑎|𝑥 2 + | 𝑏|𝑦 2 + |𝑎𝑏|𝑧2 , so
nrd(𝐵× ) ⊆ Q×>0 . At the same time, nrd(O× ) ⊆ Z× = {±1}, so we conclude O× = O1 .
This group is finite because the restriction nrd | O of the reduced norm to O ' Z4 defines
a (still) positive definite quadratic form, so there are only finitely many elements of O
of any fixed reduced norm. (For a geometric perspective, viewing the elements of O1
as lattice points on an ellipsoid in R4 ), see Proposition 17.5.6.)
11.5.10. Among the (nontrivial) cyclic groups, only subgroups of order 2, 4, 6 are
possible over Q. Indeed, a generator satisfies a quadratic equation with integer coeffi-
cients and so belongs to the ring of integers of an imaginary quadratic field; and only
two imaginary
√ quadratic fields have units other than ±1, namely,√the Eisenstein order
Z[(−1 + −3)/2] of discriminant −3 and the Gaussian order Z[ −1] of discriminant
−4 with groups of size 4, 6, respectively. (The more precise question of whether or
not there is a unit of specified order is a question of embedding numbers, the subject
of Chapter 30.)
170 CHAPTER 11. THE HURWITZ ORDER
11.5.11. Next, suppose that O× /{±1} is dihedral, and let 𝑗 ∈ O× r {±1} act by
inversion (equivalently, conjugation) on a cyclic group (of order 2, 3, by 11.5.10),
generated by an element 𝑖. Let 𝐾 = Q(𝑖). Since 𝑗 acts by inversion, we have 𝑗 2 ∈ Q,
and since
𝑗 ∈ O× we have 𝑗 2 = −1. It follows that 𝑗 𝛼 = 𝛼 𝑗 for all 𝛼 ∈ 𝐾. Thus
𝐾, −1
𝐵' , and we have two possibilities:
Q
(i) If 𝑖 has order 4, then 𝐵 ' (−1, −1 | Q) and O contains the order generated
by 𝑖, 𝑗. This is the case treated in section 11.2: O is the Lipschitz order, and
O× ' 𝑄 8 is the quaternion group of order 8.
(ii) Otherwise, 𝑖 = 𝜔 has order 6, and 𝐵 ' (−3, −1 | Q). By Exercise 11.12(a),
we have (−3, −1 | Q) ; (−1, −1 | Q). By an argument similar to Lemma
11.1.2—and boy, there is more of this to come in Chapter 32—we see that
O = Z + Z𝜔 + Z 𝑗 + Z𝜔 𝑗 (11.5.12)
is maximal. The group O× /{±1} ' 𝐷 6 is a dihedral group of order 6, and the
group O× is generated by 𝜔, 𝑗 with relations 𝜔3 = 𝑗 2 = −1 and 𝑗𝜔 = 𝜔−1 𝑗; in
other words, O× ' 𝐶3 o 𝐶4 is the semidirect product of the cyclic group 𝐶3 of
order 3 by the action of the cyclic group 𝐶4 with a generator acting by inversion
on 𝐶3 . Because 𝑖 2 = −1 is central, we also have an exact sequence
1 → 𝐶6 → O× → 𝐶2 → 1
where 𝐶6 ' h𝜔i and 𝐶2 ' h 𝑗i/{±1}. This group is also called the binary
dihedral or dicyclic group of order 12, denoted 2𝐷 6 .
11.5.13. To conclude, suppose that O× /{±1} is exceptional. Each of these groups
contain a dihedral group, so the argument from 11.5.11 applies: the only new group
we see is the (binary) tetrahedral group obtained from the Hurwitz units (section 11.2).
Here is another proof: the group 𝑆4 contains an element of order 4 and 𝐴5 an element
of order 5, and these lift to elements of order 8, 10 in O× , impossible.
We have proven the following theorem.
Theorem 11.5.14. Let 𝐵 = (𝑎, 𝑏 | Q) be a quaternion algebra over Q with 𝑎, 𝑏 < 0,
and let O ⊆ 𝐵 be an order. Then O× is either cyclic of order 2, 4, 6, quaternion of
order 8, binary dihedral of order 12, or binary tetrahedral of order 24.
Moreover, O× is quaternion, binary dihedral, or binary tetrahedral if and only
if O is isomorphic to the Lipschitz order, the order (11.5.12), or the Hurwitz order,
respectively.
Proof. Combine 11.5.10, 11.5.11, and 11.5.13.
Exercises
I 1. Check directly that the Hurwitz order
1+𝑖+ 𝑗 + 𝑘
O = Z + Z𝑖 + Z 𝑗 + Z
2
11.5. FINITE QUATERNIONIC UNIT GROUPS 171
−1, −1
is indeed an order in 𝐵 = .
Q
−1, −1
2. Let 𝐵 = , and let O ⊆ 𝐵 be the Hurwitz order. For the normalizer
Q
𝑁 𝐵× (O) := {𝛼 ∈ 𝐵× : 𝛼−1 O𝛼 = O}
show the equality 𝑁 𝐵× (Zh𝑖, 𝑗i) = 𝑁 𝐵× (O). [Hint: consider units and their
traces.]
3. (a) Show that the Lipschitz order Zh𝑖, 𝑗i is the unique suborder of the Hurwitz
order O with index 2 (as abelian groups).
(b) Show that
Zh𝑖, 𝑗i = {𝛼 ∈ O : trd(𝛼) is even}.
4. Check that the map
O/3O → M2 (F3 )
0 −1 1 1
𝑖, 𝑗 ↦→ ,
1 0 1 −1
(cf. (11.2.8)).
I 8. Let
Λ = Z4 + Z( 12 , 12 , 21 , 12 ) ⊂ R4
be the image of the Hurwitz order O under the natural embedding O ↩→ H ' R4 .
Show that for every 𝑥 ∈ R4 , there exists 𝜆 ∈ Λ such that k𝜆k 2 ≤ 1/2. [Hint:
without loss of generality we may take 0 ≤ 𝑥𝑖 ≤ 1/2 for all 𝑖; then show we may
take 𝑥 1 + 𝑥2 + 𝑥3 + 𝑥4 ≤ 1; conclude that the maximum value of k𝑥k 2 with these
conditions occurs at the point ( 12 , 12 , 0, 0).]
9. Let 𝐵 be a definite quaternion algebra over Q and let O ⊆ 𝐵 be an order. Show
that O is left Euclidean if and only if O is right Euclidean (with respect to a
norm 𝑁).
10. Let O ⊂ 𝐵 := (−1, −1 | Q) be the Hurwitz order.
(a) Consider the natural ring homomorphism O → O/2O = O ⊗Z F2 giving
the reduction of the algebra O modulo 2. Show that O/2O is an F2 -
algebra, that #(O/2O) = 16, and that (O/2O) × ' 𝐴4 is isomorphic to
172 CHAPTER 11. THE HURWITZ ORDER
and
1 0
𝜋= , 𝑎 = 0, 1, . . . , 𝑝 − 1.
𝑎 𝑝
[Hint: use column operations.]
(b) Repeat (a) but with SL2 (Z ( 𝑝) ) acting on the set of matrices 𝜋 ∈ M2 (Z ( 𝑝) )
with det(𝜋) = 𝑝, with the same conclusion.
(c) Show that the number of (left or) right ideals of O of reduced norm 𝑝 is
equal to 𝑝 + 1.
(d) Accounting for units, conclude that the number of ways of writing an odd
prime 𝑝 as the sum of four squares is equal to 8( 𝑝 + 1).
15. In the following exercise, we consider a computational problem, suitable for
those with some background in number theory algorithms (see e.g. Cohen
[Coh93]).
(a) Show that one can find 𝑥, 𝑦, 𝑧 ∈ Z such that 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑝𝑚 with 𝑝 - 𝑚
in probabilistic polynomial time in log 𝑝.
(b) Describe the right Euclidean algorithm as applied to 𝛼 = 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 and
𝑝 to obtain 𝜋 ∈ O with nrd(𝜋) = 𝑝. Adjust as in Lemma 11.2.9 to find a
solution to 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑝 with 𝑥, 𝑦, 𝑧 ∈ Z. Estimate the running time of
this algorithm.
Chapter 12
In this chapter, we classify quaternion algebras over local fields using quadratic forms;
this generalizes the classification of quaternion algebras over R.
where each 𝑎 𝑖 ∈ {0, . . . , 𝑝 − 1} are the digits of 𝑎. We continue “to the left” because
a decimal expansion is a series in the base 1/10 < 1 and instead we have a base 𝑝 > 1.
Put a bit more precisely, we define the 𝑝-adic absolute value on Q by defined by
|0| 𝑝 := 0 and
|𝑐| 𝑝 := 𝑝 −𝑣𝑝 (𝑐) for 𝑐 ∈ Q× , (12.1.2)
where 𝑣 𝑝 (𝑐) is the power of 𝑝 occurring in 𝑐 in its unique factorization (taken to be
negative if 𝑝 divides the denominator of 𝑐 written in lowest terms). Then the field Q 𝑝
is the completion of Q with respect to | | 𝑝 , that is to say, Q 𝑝 is the set of equivalence
classes of Cauchy sequences of rational numbers, and it obtains a topology induced by
the metric 𝑑 𝑝 (𝑥, 𝑦) = |𝑥 − 𝑦| 𝑝 . We have |𝑎| 𝑝 = 𝑝 𝑘 for 𝑎 as in (12.1.1) with 𝑎 −𝑘 ≠ 0.
175
176 CHAPTER 12. TERNARY QUADRATIC FORMS OVER LOCAL FIELDS
Of course, all of the information in the 𝑝-adic absolute value is encoded in the 𝑝-adic
valuation 𝑣 𝑝 : Q → R ∪ {∞}.
Inside Q 𝑝 is the ring Z 𝑝 of 𝑝-adic integers, the completion of Z with respect to
| | 𝑝 : the ring Z 𝑝 consists of those elements of Q 𝑝 with 𝑎 𝑛 = 0 for 𝑛 < 0. (The ring Z 𝑝
might be thought of intuitively as Z/𝑝 ∞ Z, if this made sense.)
Equipped with their topologies, the ring Z 𝑝 is compact and the field Q 𝑝 is lo-
cally compact. These statements can be understood quite easily by viewing Z 𝑝 in a
slightly different way, as a projective limit with respect to the natural projection maps
Z/𝑝 𝑛+1 Z → Z/𝑝 𝑛 Z:
Z 𝑝 = lim Z/𝑝 𝑛 Z
←−−
𝑛
( ∞
Ö
) (12.1.3)
𝑛 𝑛
= 𝑥 = (𝑥 𝑛 ) 𝑛 ∈ Z/𝑝 Z : 𝑥 𝑛+1 ≡ 𝑥 𝑛 (mod 𝑝 ) for all 𝑛 ≥ 1 .
𝑛=1
Happily, this proposition can be proved using some rather direct manipulations
with quadratic forms and gives a very “hands on” feel; it is also suggests the arguments
we use for a more general result. The main input we need is a quadratic Hensel’s
lemma, or more precisely, the following consequence.
Proof. Let 𝑎 ∈ Q×𝑝 and let 𝑚 := 𝑣 𝑝 (𝑎). Then 𝑎 = 𝑏 𝑝 𝑚 with 𝑏 := 𝑎/𝑝 𝑚 ∈ Z×𝑝 , and by
squaring 𝑎 ∈ Q×2 ×2 ×2
𝑝 if and only if 𝑏 ∈ Z 𝑝 and 𝑚 is even. We claim that 𝑏 ∈ Z 𝑝 if and
×
only if its reduction 𝑏 modulo 𝑝 is a square in (Z/𝑝Z) . With the forward implication
immediate, suppose 𝑏 ≡ 𝑐2 (mod 𝑝) with 𝑐 ∈ Z×𝑝 , then 𝑏/𝑐2 ∈ 1 + 𝑝Z 𝑝 . But squaring
is a bijection on 1 + 𝑝Z 𝑝 , by expanding the square root as a convergent series (see
Exercise 12.1) and using that 𝑝 ≠ 2. Thus 𝑏/𝑐2 ∈ Z×2 𝑝 , and the result follows.
• When (𝑏, 𝑐) = (𝑒, 𝑒), we have after rescaling 𝑥 2 + 𝑦 2 − 𝑒𝑧2 . We claim this form is
always isotropic. Indeed, the form reduces to a nondegenerate ternary quadratic
form over F 𝑝 . Such a form is always isotropic by a delightful counting argument
(Exercise 5.5(b), or a second chance in Exercise 12.6!). Lifting, there exist
𝑥, 𝑦, 𝑧 ∈ Z 𝑝 , not all zero modulo 𝑝, such that 𝑥 2 ≡ −𝑦 2 + 𝑒𝑧 2 (mod 𝑝). Since
𝑒 is a nonsquare, we have 𝑝 - 𝑥 (arguing similarly as in the first paragraph).
Let 𝑑 := −𝑦 2 + 𝑒𝑧2 ∈ Q×𝑝 . By the possibilities in Lemma 12.1.7, we must have
𝑑 ∈ Q×2 2
𝑝 ; solving 𝑥 = 𝑑 for 𝑥 ∈ Q 𝑝 then shows that 𝑄 is isotropic.
• The case (𝑒, 𝑝) is our desired form.
• In the third case (eep!), we substitute 𝑥 ← 𝑒𝑥 and divide by 𝑒 to obtain the form
−𝑦 2 + 𝑒𝑥 2 − 𝑝𝑧 2 . We claim that there is an isometry h−1, 𝑒i ' h1, −𝑒i: indeed,
in the first bullet we showed that the quadratic form h−1, 𝑒, −1i is isotropic,
so −𝑥 2 + 𝑒𝑦 2 represents 1; using this representation as the first basis vector,
extending to a basis, and diagonalizing, we conclude that h−1, 𝑒i ' h1, 𝑏i. By
discriminants, we have −𝑒 = 𝑏 up to squares. This brings us back to the first
case.
• In cases ( 𝑝, 𝑝) or (𝑒 𝑝, 𝑒 𝑝), replacing 𝑥 ← 𝑝𝑥 and dividing gives the quadratic
forms 𝑥 2 + 𝑦 2 − 𝑝𝑧2 and 𝑥 2 + 𝑦 2 − 𝑒 𝑝𝑧 2 . If −1 ∈ Z×2
𝑝 , then the form is isotropic;
otherwise, we may take 𝑒 = −1 and we are back to cases (𝑒, 𝑝), (𝑒, 𝑒 𝑝).
• In the final case (keeping pep!), we substitute 𝑥 ← 𝑝𝑥 and divide by −𝑝 to get
𝑦 2 + 𝑒𝑧2 − 𝑝𝑥 2 . If −1 ∈ Z×2 𝑝 , then by substitution we change the middle sign
to return to the first case. Otherwise, we may take 𝑒 = −1, and the form is
isotropic, a contradiction.
Although direct, the proof we just gave has the defect that quadratic forms behave
differently in characteristic 2, and so one may ask for a proof that works uniformly in
all characteristics: we give such a proof in the next chapter by extending valuations.
One of the nice applications of this classification is that it gives a necessary
condition for two quaternion algebras to be isomorphic. Let 𝐵 = (𝑎, 𝑏 | Q) be a
quaternion algebra over Q and consider its scalar extension 𝐵 𝑝 = 𝐵 ⊗Q Q 𝑝 ' (𝑎, 𝑏 |
Q 𝑝 ). If 𝐵 0 is another quaternion algebra over Q and 𝐵 ' 𝐵 0, then this implies
𝐵 𝑝 ' 𝐵 0𝑝 for all primes 𝑝, and of course the same is true over R. Perhaps surprisingly,
it turns out that the collection of all of these tests is also sufficient: if 𝐵, 𝐵 0 become
isomorphic over R and over Q 𝑝 for all primes 𝑝, then in fact 𝐵 ' 𝐵 0 are isomorphic
over Q! This profound and powerful principle—detecting global isomorphism from
local isomorphisms, a local-global principle—will be examined in chapter 14.
Our motivation for local fields is as follows: we want a topology compatible with
the field operations in which the field is Hausdorff and locally compact (every element
has a compact neighborhood), analogous to what holds over the real and complex
numbers. And to avoid trivialities, we will insist that this topology is not the discrete
topology (where every subset of 𝐹 is open). To carry this out, we begin with some
basic definitions.
Definition 12.2.1. A topological group is a group equipped with a topology such that
the group operation and inversion are continuous. A homomorphism of topological
groups is a group homomorphism that is continuous.
A topological ring is a ring 𝐴 equipped with a topology such that the ring oper-
ations (addition, negation, and multiplication) are continuous. A homomorphism of
topological rings is a ring homomorphism that is continuous. A topological field is a
field that is also a topological ring in such a way that division by a nonzero element is
continuous.
One natural way to equip a ring with a topology is by way of an absolute value. To
get started, we consider such notions first for fields. Throughout this section, let 𝐹 be
a field.
| | : 𝐹 → R ≥0
such that:
Example 12.2.5. The fields R and C are topological fields with respect to the usual
archimedean absolute value.
180 CHAPTER 12. TERNARY QUADRATIC FORMS OVER LOCAL FIELDS
Remark 12.2.6. A field with absolute value is archimedean if and only if it satisfies
the archimedean property: for all 𝑥 ∈ 𝐹 × , there exists 𝑛 ∈ Z such that |𝑛𝑥| > 1. In
particular, a field 𝐹 equipped with an archimedean absolute value has char 𝐹 = 0.
Example 12.2.7. Every field has the trivial (nonarchimedean) absolute value, defined
by |0| = 0 and |𝑥| = 1 for all 𝑥 ∈ 𝐹 × ; the trivial absolute value induces the discrete
topology on 𝐹.
Example 12.2.10. Let 𝑘 be a field and 𝐹 = 𝑘 (𝑡) the field of rational functions over
𝑘. For 𝑓 (𝑡) = 𝑔(𝑡)/ℎ(𝑡) ∈ 𝑘 (𝑡) r {0} with 𝑔(𝑡), ℎ(𝑡) ∈ 𝑘 [𝑡], define 𝑣( 𝑓 (𝑡)) :=
deg ℎ(𝑡) − deg 𝑔(𝑡). Then 𝑣 is a discrete valuation on 𝐹.
Given the parallels between them, it should come as no surprise that a valuation
gives rise to an absolute value on 𝐹 by defining
Example 12.2.12. The trivial valuation is the valuation 𝑣 satisfying 𝑣(0) = ∞ and
𝑣(𝑥) = 0 for all 𝑥 ∈ 𝐹 × . The trivial valuation gives the trivial absolute value on 𝐹.
Two valuations 𝑣, 𝑤 are equivalent if there exists 𝑎 ∈ R>0 such that 𝑣(𝑥) = 𝑎𝑤(𝑥)
for all 𝑥 ∈ 𝐹; equivalent valuations give the same topology on a field. A nontrivial
discrete valuation is equivalent after rescaling (by the minimal positive element in the
value group) to one with value group Z, since a nontrivial discrete subgroup of R is
cyclic; we call such a discrete valuation normalized.
12.2. LOCAL FIELDS 181
12.2.13. Given a field 𝐹 with a nontrivial discrete valuation 𝑣, the valuation ring is
𝑅 := {𝑥 ∈ 𝐹 : 𝑣(𝑥) ≥ 0}. We have 𝑅 × = {𝑥 ∈ 𝐹 : 𝑣(𝑥) = 0} since
for all 𝑥 ∈ 𝐹 × . The valuation ring is a local domain with unique maximal ideal
𝔭 := {𝑥 ∈ 𝐹 : 𝑣(𝑥) > 0} = 𝑅 r 𝑅 × .
Recall that a topological space is locally compact if each point has a compact
neighborhood (every point is contained in a compact set containing an open set).
Definition 12.2.14. A local field is a Hausdorff, locally compact topological field with
a nondiscrete topology.
Theorem 12.2.15. A field 𝐹 with absolute value is a local field if and only if 𝐹 is one
of the following:
Proof. See Neukirch [Neu99, Chapter II, §5], Cassels [Cas86, Chapter 4, §1], or Serre
[Ser79, Chapter II, §1].
Although a local field is only locally compact, the valuation ring is itself compact,
as follows.
𝑥 ∈ 𝐹 × with |𝑥| = 𝛿 > 0. The image | 𝐹 × | ⊆ R>0 is discrete, so there exists 0 < 𝜖 < 𝛿
such that | 𝑦| < 𝛿 implies | 𝑦| ≤ 𝛿 − 𝜖 for all 𝑦 ∈ 𝐹. Thus an open ball is a closed ball
since 𝑥 ∈ 𝐹 × and 𝛿 > 0 were arbitrary, the only connected subset containing 0 is {0}.
Next, we show 𝑅 is compact. There is a natural continuous ring homomorphism
∞
Ö
𝜙: 𝑅 → 𝑅/𝔭𝑛
𝑛=1
where each factor 𝑅/𝔭𝑛 is equipped with the discrete topology Ñ and𝑛 the product is
given the product topology. The map 𝜙 is injective, since ∞ 𝑛=1 𝔭 = {0} (every
nonzero element has finite valuation). The image of 𝜙 is obviously closed. Therefore
𝑅 is homeomorphic onto its closed image. But by Tychonoff’s theorem, the product
Î∞
𝑛=1 𝑅/𝔭 of compact sets is compact, and a closed subset of a compact set is compact,
𝑛
thus 𝑅 is compact.
Proof. The result is straightforward to prove using Taylor expansion or the same
formulas as in Newton’s method.
Lemma 12.2.18 (Hensel’s lemma). Let 𝐹 be a nonarchimedean local field with valu-
ation 𝑣 and valuation ring 𝑅, and let 𝑓 (𝑥 1 , . . . , 𝑥 𝑛 ) ∈ 𝑅[𝑥 1 , . . . , 𝑥 𝑛 ] with 𝑛 ≥ 1.
Let 𝑎 ∈ 𝑅 𝑛 have 𝑚 := 𝑣( 𝑓 (𝑎)) and suppose that
𝜕𝑓
𝑚 > 2 min 𝑣 (𝑎) ≥ 0.
𝑖 𝜕𝑥𝑖
𝑎 ∈ 𝑅 𝑛 such that 𝑓 (e
Then there exists e 𝑎 ) = 0 and
𝑎≡𝑎
e (mod 𝔭𝑚 ).
Proof. One can reduce from several variables to the one variable version of Hensel’s
lemma (Lemma 12.2.17): see Exercise 12.11.
Remark 12.2.19. With essentially the same proof, Hensel’s lemma holds more gener-
ally for 𝑅 a complete DVR (without the condition on the residue field) and becomes
axiomatically the property of Henselian rings.
12.3. CLASSIFICATION VIA QUADRATIC FORMS 183
Recall definitions and notation for quadratic forms over 𝑅 provided in section 9.7,
we embark on a proof in the case where char 𝑘 ≠ 2, beginning with the following
lemma.
184 CHAPTER 12. TERNARY QUADRATIC FORMS OVER LOCAL FIELDS
so the vector of partial derivatives ((𝜕𝑄/𝜕𝑥𝑖 ) (𝑎))𝑖 = [𝑇]𝑎 is just the matrix product
of the Gram matrix with the vector 𝑎 = (𝑎 𝑖 )𝑖 . Working modulo 𝔭, we have disc 𝑄 𝑘 =
2−𝑛 det[𝑇] ∈ 𝑘 × , using that 2 ∈ 𝑘 × , so the kernel of [𝑇] mod 𝔭 is zero. Since
𝑎 has nonzero reduction, we conclude that [𝑇]𝑎 also has nonzero reduction, which
means that min𝑖 𝑣((𝜕𝑄/𝜕𝑥𝑖 ) (𝑎)) = 0. Therefore the hypotheses of Hensel’s lemma
are satisfied with 𝑚 = 1, and we conclude there exists a nonzero e 𝑎 ∈ 𝑀 such that
𝑄(e𝑎 ) = 0 and so 𝑄 is isotropic.
Proof of Theorem 12.3.4 (char 𝑘 ≠ 2). Let 𝑄 ' h𝑎, −𝑏, −𝑐i be an anisotropic ternary
quadratic form over 𝐹. Then 𝑄 is nondegenerate. After rescaling and a change of basis
(Exercise 12.8), we may suppose that 𝑎 = 1 and 0 = 𝑣(𝑏) ≤ 𝑣(𝑐). If 𝑣(𝑏) = 𝑣(𝑐) = 0
then the quadratic form modulo 𝔭 is nonsingular, so by Lemma 12.3.5 it is isotropic
and by Lemma 12.3.6 we conclude 𝑄 is isotropic, a contradiction.
We are left with the case 𝑣(𝑏) = 0 and 𝑣(𝑐) = 1. By Lemma 12.3.9, we may
suppose 𝑏 = 1 or 𝑏 = 𝑒 where 𝑒 is a nonsquare in 𝑘. If 𝑏 = 1, then the form is
12.3. CLASSIFICATION VIA QUADRATIC FORMS 185
Proof of Theorem 12.3.4 (char 𝑘 = 2). We first claim that the form [1, 1, 𝑡] h𝜋i is
anisotropic; this follows from a straightforward modification of the argument as in the
proof when char 𝑘 ≠ 2 above.
We now show this form is the unique one up to similarity. Suppose that 𝑄 is a
ternary anisotropic form over 𝑅. Let 𝑥 ∈ 𝑉 be nonzero; since 𝑄 is anisotropic, we
may scale 𝑥 so that 𝑎 := 𝑄(𝑥) ∈ 𝑅. Since dim 𝑉 ≥ 3, there exists nonzero 𝑦 0 ∈ 𝑉
such that 𝑇 (𝑥, 𝑦 0) = 0; rescale 𝑦 0 so that 𝑄(𝑦 0) ∈ 𝑅. Let 𝑦 := 𝑥 + 𝑦 0. Then 𝑇 (𝑥, 𝑦) =
𝑇 (𝑥, 𝑥 + 𝑦 0) = 𝑎 and 𝑏 := 𝑄(𝑦) ∈ 𝑅, so 𝑄 on this basis is 𝑎𝑥 2 + 𝑎𝑥𝑦 + 𝑏 2 ' [𝑎, 𝑎, 𝑏].
We compute that disc [𝑎, 𝑎, 𝑏] = 𝑎(𝑎 − 4𝑏) ≡ 𝑎 2 (mod 𝔭), so disc [𝑎, 𝑎, 𝑏] ∈ 𝑅 ×
and [𝑎, 𝑎, 𝑏] is nonsingular; completing to a basis with a nonzero element in the
orthogonal complement and rescaling, we may suppose without loss of generality that
𝑄 = [𝑎, 𝑎, 𝑏] h𝑐i with 𝑎, 𝑏, 𝑐 ∈ 𝑅 and both 𝑣(𝑎) = 0, 1 and 𝑣(𝑐) = 0, 1, with
disc 𝑄 = 𝑎(𝑎 − 4𝑏)𝑐. This leaves four cases.
Having exhausted the cases and the reader, the result now follows.
Corollary 12.3.12. Let 𝐹 be a nonarchimedean local field with valuation ring 𝑅 and
uniformizer 𝜋 ∈ 𝑅. Let 𝐵 be a quaternion algebra over 𝐹.
If char 𝑘 ≠ 2, then 𝐵 is a division algebra if and only if
𝑒, 𝜋
𝐵' , where 𝑒 ∈ 𝑅 × is nontrivial in 𝑘 × /𝑘 ×2
𝐹
and if char 𝐹 = char 𝑘 = 2, then 𝐵 is a division algebra if and only if
𝑡, 𝜋
𝐵' , where 𝑡 ∈ 𝑅 is nontrivial in 𝑘/℘(𝑘).
𝐹
12.4. HILBERT SYMBOL 187
( , ) 𝐹 : 𝐹 × × 𝐹 × → {±1}
𝑎, 𝑏
by the condition that (𝑎, 𝑏) 𝐹 = 1 if and only if the quaternion algebra ' M2 (𝐹)
𝐹
is split.
The Hilbert symbol is well-defined as a map
𝐹 × /𝐹 ×2 × 𝐹 × /𝐹 ×2 → {±1}
(Exercise 2.4). By Main Theorem 5.4.4(v), we have (𝑎, 𝑏) 𝐹 = 1 if and only if the
Hilbert equation 𝑎𝑥 2 + 𝑏𝑦 2 = 1 has a solution with 𝑥, 𝑦 ∈ 𝐹: this is called Hilbert’s
criterion for the splitting of a quaternion algebra.
𝑎, 𝑏
Remark 12.4.2. The similarity between the symbols and (𝑎, 𝑏) 𝐹 is intentional;
𝐹
but they are not the same, as the former represents an algebra and the latter takes the
value ±1.
In some contexts, the Hilbert
symbol (𝑎, 𝑏) 𝐹 is defined to be the isomorphism class
𝑎, 𝑏
of the quaternion algebra in the Brauer group Br(𝐹), rather than ±1 according
𝐹
to whether or not the algebra is split. Conflating these two symbols is not uncommon
and in certain contexts it can be quite convenient, but we warn that it can lead to
confusion and caution against referring to a quaternion algebra or its isomorphism
class as a Hilbert symbol.
Lemma 12.4.3. Let 𝑎, 𝑏 ∈ 𝐹 × . Then the following statements hold:
(a) (𝑎𝑐2 , 𝑏𝑑 2 ) 𝐹 = (𝑎, 𝑏) 𝐹 for all 𝑐, 𝑑 ∈ 𝐹 × .
(b) (𝑏, 𝑎) 𝐹 = (𝑎, 𝑏) 𝐹 .
(c) (𝑎, 𝑏) 𝐹 = (𝑎, −𝑎𝑏) 𝐹 = (𝑏, −𝑎𝑏) 𝐹 .
(d) (1, 𝑎) 𝐹 = (𝑎, −𝑎) 𝐹 = 1.
(e) If 𝑎 ≠ 1, then (𝑎, 1 − 𝑎) 𝐹 = 1.
188 CHAPTER 12. TERNARY QUADRATIC FORMS OVER LOCAL FIELDS
Proof. Statements (a)–(c) follow from Exercise 2.4. For (d), the Hilbert equation
𝑥 2 + 𝑎𝑦 2 = 1 has the obvious solution (𝑥, 𝑦) = (1, 0). And h𝑎, −𝑎i is isotropic (taking
(𝑥, 𝑦) = (1, 1)) so is a hyperbolic plane and represents 1 as in the proof of Main
Theorem 5.4.4, or we argue
by Exercise 2.4. For part (e), by Hilbert’s criterion (𝑎, 1 − 𝑎) 𝐹 = 1 since the quadratic
equation 𝑎𝑥 2 +(1−𝑎)𝑦 2 = 1 has the solution (𝑥, 𝑦) = (1, 1). Finally, part (f): the Hilbert
equation 𝑎𝑥 2 + 𝑏𝑦 2 = 1 has a solution with 𝑥, 𝑦 ∈ 𝐹 if and only if 𝜎(𝑎)𝑥 2 + 𝜎(𝑏)𝑦 2 = 1
has such a solution.
Remark 12.4.4. Staring at the properties in Lemma 12.4.3 and seeking to axiomatize
them, the study of symbols like the Hilbert symbol leads naturally to the definition of
𝐾2 (𝐹). In its various formulations, algebraic 𝐾-theory (𝐾 for the German “Klasse”,
following Grothendieck) seeks to understand certain kinds of functors from rings to
abelian groups in a universal sense, encoded in groups 𝐾𝑛 (𝑅) for 𝑛 ∈ Z ≥0 and 𝑅 a
commutative ring: see e.g. Karoubi [Kar2010]. For a field 𝐹, we have 𝐾0 (𝐹) = Z and
𝐾1 (𝐹) = 𝐹 × . By a theorem of Matsumoto [Mat69] (see also Milnor [Milno71]), the
group 𝐾2 (𝐹) is the universal domain for symbols over 𝐹:
(The tensor product over Z views 𝐹 × as an abelian group and therefore a Z-module.)
The map 𝑎 ⊗ 𝑏 ↦→ (𝑎, 𝑏) 𝐹 extends to a map 𝐾2 (𝐹) → {±1}, a Steinberg symbol, a
homomorphism from 𝐾2 (𝐹) to a multiplicative abelian group. The higher 𝐾-groups
are related to deeper arithmetic of commutative rings. For an introduction, see Weibel
[Weib2013] and Curtis–Reiner [CR87, Chapter 5].
We now turn to be quite explicit about the values of the Hilbert symbol. We begin
with the case where 𝐹 is archimedean. If 𝐹 = C, then the Hilbert symbol is identically
1. If 𝐹 = R, then (
1, if 𝑎 > 0 or 𝑏 > 0;
(𝑎, 𝑏)R = (12.4.5)
−1, if 𝑎 < 0 and 𝑏 < 0.
Proof (char 𝑘 ≠ 2). This lemma can be read off of the direct computation below
(12.4.9), recording what was computed along the way in the proof of Theorem
12.3.4.
12.4.8. Since the Hilbert symbol is well-defined up to squares, the symbol (𝑎, 𝑏) 𝐹 is
determined by the values with 𝑎, 𝑏 ∈ {1, 𝑒, 𝜋, 𝑒𝜋} where 𝑒 is a nonsquare in 𝑘 × . Let
𝑠 = (−1) (#𝑘−1)/2 , so that 𝑠 = 1, −1 according as −1 is a square in 𝑘. Then:
(𝑎, 𝑏) 𝐹 1 𝑒 𝜋 𝑒𝜋
1 1 1 1 1
𝑒 1 1 −1 −1 (12.4.9)
𝜋 1 −1 𝑠 −𝑠
𝑒𝜋 1 −1 −𝑠 𝑠
where 𝑞 = #𝑘 and
𝑐
= 0, ±1 ≡ 𝑐 (𝑞−1)/2 (mod 𝔭) (12.4.11)
𝔭
is the Legendre symbol: see Exercise 12.16. (Multiplicativity can also be read off of
the formula 12.4.10.)
12.4.12. The following easy criteria follow from 12.4.9 (or (12.4.10)):
12.4.13. To compute the Hilbert symbol for a local field 𝐹 with char 𝐹 = 0 and
char 𝑘 = 2 is significantly more involved. But we can at least compute the Hilbert
symbol by hand for 𝐹 = Q2 .
To begin, the group Q×2 /Q×2
2 is generated by −1, −3, 2, so representatives are
{±1, ±3, ±2, ±6}. We recall Hilbert’s criterion: (𝑎, 𝑏) 𝐹 = 1 if and only if 𝑎𝑥 2 +𝑏𝑦 2 = 1
has a solution with 𝑥, 𝑦 ∈ 𝐹.
If 𝑎, 𝑏 ∈ Z are odd, then
By the determination above, we see that (−3, 𝑏) = −1 for 𝑏 = ±2, ±6 and (2, 2)2 =
(−1, 2)2 = 1 the latter by Hilbert’s criterion, as −1 + 2 = 1; knowing multiplicativity
(Lemma 12.4.6), we have uniquely determined all Hilbert symbols, in particular, for
𝑎 ∈ Z odd we have
2
(𝑎, 2)2 = (−1) (𝑎 −1)/8 . (12.4.15)
It is still useful to compute several of these symbols individually, in the same manner
as (12.4.13) (working modulo 8): see Exercise 12.17. We summarize the results here:
(𝑎, 𝑏)2 1 −3 −1 3 2 −6 −2 6
1 1 1 1 1 1 1 1 1
−3 1 1 1 1 −1 −1 −1 −1
−1 1 1 −1 −1 1 1 −1 −1
3 1 1 −1 −1 −1 −1 1 1 (12.4.16)
2 1 −1 1 −1 1 −1 1 −1
−6 1 −1 1 −1 −1 1 −1 1
−2 1 −1 −1 1 1 −1 −1 1
6 1 −1 −1 1 −1 1 1 −1
Remark 12.4.17. Analogously, one can define a symbol [𝑎, 𝑏) 𝐹 for the splitting of
quaternion algebras for 𝐹 a local field with char 𝐹 = 2. This symbol is no longer called
the Hilbert symbol, but many properties
remain: in particular, there is still an analogue
𝑎, 𝑏
of the Hilbert equation, and is split if and only if 𝑏𝑥 2 + 𝑏𝑥𝑦 + 𝑎𝑏𝑦 2 = 1 has a
𝐹
solution with 𝑥, 𝑦 ∈ 𝐹.
Exercises
1. Let 𝑝 be an odd prime. Show
(a) Show the equality
∞
∑︁
(1 − 4𝑥) 1/2 = 1 − 𝐶𝑛 𝑥 𝑛 ∈ Z[[𝑥]]
𝑛=1
2. Recall that a topological space is T1 if for every pair of distinct points, each
point has an open neighborhood not containing the other.
(a) Show that a topological space 𝑋 is T1 if and only if {𝑥} is closed for all
𝑥 ∈ 𝑋.
(b) Let 𝐺 be a topological group. Show that 𝐺 is Hausdorff if and only if 𝐺
is T1 .
3. In this exercise we prove some basic facts about topological groups. Let 𝐺 be a
topological group.
(a) Let 𝐻 ≤ 𝐺 be a subgroup. Show that 𝐻 is open if and only if there exists
ℎ ∈ 𝐻 and an open neighborhood of ℎ contained in 𝐻.
(b) Show that if 𝐻 ≤ 𝐺 is an open subgroup, then 𝐻 is closed.
(c) Show that a closed subgroup 𝐻 ≤ 𝐺 of finite index is open.
(d) Suppose that 𝐺 is compact. Show that an open subgroup 𝐻 ≤ 𝐺 is of finite
index, and that every open subgroup contains an open normal subgroup.
I 4. Let 𝐺 be a topological group. Let 𝑈 3 1 be an open neighborhood of 1.
(a) Show that there exists an open neighborhood 𝑉 ⊆ 𝑈 of 1 ∈ 𝑉 such that
𝑉 2 = 𝑉 · 𝑉 ⊆ 𝑈. [Hint: Multiplication is continuous.]
(b) Similarly, show that there exists an open neighborhood 𝑉 ⊆ 𝑈 of 1 ∈ 𝑉
such that 𝑉 −1𝑉 ⊆ 𝑈.
5. Let 𝐺 be a topological group and let 𝐻 ≤ 𝐺 be a closed subgroup. Equip 𝐺/𝐻
with the quotient topology. Show that 𝐺/𝐻 is Hausdorff. [Hint: Use Exercise
12.4(b).]
I 6. Let 𝑘 be a finite field and let 𝑄 : 𝑉 → 𝑘 be a ternary quadratic form. Show that 𝑄
is isotropic. [Hint: Reduce to the case of finding a solution to 𝑦 2 = 𝑓 (𝑥) where
𝑓 is a polynomial of degree 2. If #𝑘 is odd, count squares and the number of
distinct values taken by 𝑓 (𝑥) in 𝑘. Second approach: reduce to the case where
#𝑘 is odd, and show that 𝑥 2 + 𝑦 2 represents a nonsquare, since the squares
cannot be closed under addition!] [This repeats Exercise 5.5!]
I 7. Let 𝑘 be a finite field with char 𝑘 ≠ 2 and let 𝑒 ∈ 𝑘 × . Show directly that there is
an isometry h−1, 𝑒i ' h1, −𝑒i.
I 8. Let 𝑅 be a DVR with field of fractions 𝐹, let 𝑎, 𝑏, 𝑐 ∈ 𝐹 be nonzero and
let 𝑄 = h𝑎, −𝑏, −𝑐i. Show that 𝑄 is similar over 𝐹 to h1, −𝑏 0, −𝑐 0i with
0 = 𝑣(𝑏 0) ≤ 𝑣(𝑐 0). [Hint: first get 𝑣(𝑎), 𝑣(𝑏), 𝑣(𝑐) ∈ {0, 1}.]
I 9. Let 𝑘 be a finite field with even cardinality. Show that #𝑘/℘(𝑘) = 2, where ℘(𝑘)
is the Artin-Schreier group.
10. By Theorem 12.2.15, a complete archimedean local field is isomorphic to R or
C. Extend this classification to division algebras as follows.
The notion of absolute value (Definition 12.2.2) extends to a division algebra
without modification, as does the notion of archimedean and nonarchimedean.
√︁
(a) Show that H has an absolute value |𝛼| = nrd(𝛼) for 𝛼 ∈ H.
(b) Let 𝐷 be a division algebra equipped with an absolute value | |. Show
that if | | is archimedean, then char 𝐷 = 0 and if the restriction of | | to its
192 CHAPTER 12. TERNARY QUADRATIC FORMS OVER LOCAL FIELDS
Suppose 𝑚 > 2𝑣(det D 𝑓 (𝑎)). Show that there exists e 𝑎 ∈ 𝑅 𝑛 such that
𝑓𝑖 (𝑎) = 0 for all 𝑖 = 1, . . . , 𝑛 and e
𝑎 ≡ 𝑎 (mod 𝔭 ). [Hint: by Taylor
𝑚
I 17. Show that the table of Hilbert symbols (12.4.16) is correct by considering the
equation 𝑎𝑥 2 + 𝑏𝑦 2 ≡ 1 (mod 8).
18. Prove a descent for the Hilbert symbol, as follows. Let 𝐾 be a finite extension
of the local field 𝐹 with char 𝐹 ≠ 2 and let 𝑎, 𝑏 ∈ 𝐹 × . Show that (𝑎, 𝑏) 𝐾 =
(𝑎, Nm𝐾 |𝐹 (𝑏)) 𝐹 = (Nm𝐾 |𝐹 (𝑎), 𝑏) 𝐹 .
Chapter 13
In this chapter, we approach the classification of quaternion algebras over local fields
in a second way, using valuations.
𝑣(Nm𝐾 |Q 𝑝 (𝑥))
𝑤(𝑥) := . (13.1.2)
[𝐾 : Q 𝑝 ]
𝑤 : 𝐵 → R ∪ {∞}
𝑣(nrd(𝛼)) (13.1.3)
𝛼 ↦→ .
2
The valuation ring
O := {𝛼 ∈ 𝐵 : 𝑤(𝛼) ≥ 0} (13.1.4)
195
196 CHAPTER 13. QUATERNION ALGEBRAS OVER LOCAL FIELDS
is the unique (!) maximal 𝑅-order in 𝐵, consisting of all elements of 𝐵 that are integral
over Z 𝑝 . The set
𝑃 := {𝛼 ∈ 𝐵 : 𝑤(𝛼) > 0} (13.1.5)
is the unique maximal two-sided (bilateral) ideal of O.
Using the unique extension of the valuation, we obtain the following main result
of this chapter (a special case of Theorem 13.3.11).
The method of proof used in this classification can also be used to classify central
division algebras over local fields in much the same manner.
13.2 Valuations
To begin, we briefly review extensions of valuations; for further reading, see the
references given in section 12.2.
Let 𝑅 be a complete DVR with valuation 𝑣 : 𝑅 → Z ≥0 ∪ {∞}, field of fractions
𝐹, maximal ideal 𝔭 generated by a uniformizer 𝜋 (with 𝑣(𝜋) = 1), and residue field
𝑘 := 𝑅/𝔭. Then 𝑅 is an integrally closed PID (every ideal is a power of the maximal
ideal 𝔭), and 𝑅 = {𝑥 ∈ 𝐹 : 𝑣(𝑥) ≥ 0}. Let | | 𝑣 be an absolute value attached to 𝑣, as in
(12.2.11).
Let 𝐾 ⊇ 𝐹 be a finite separable extension of degree 𝑛 := [𝐾 : 𝐹]. Then in fact 𝐾
is also a nonarchimedean local field; more precisely, we have the following lemma.
𝑣(Nm𝐾 |𝐹 (𝑥))
𝑤(𝑥) := . (13.2.2)
[𝐾 : 𝐹]
The integral closure of 𝑅 in 𝐾 is the valuation ring
𝑆 := {𝑥 ∈ 𝐾 : 𝑤(𝑥) ≥ 0}.
Proof. See e.g. Neukrich [Neu99, Chapter II, Theorem (4.8)], Cassels [Cas86, Chapter
7, Theorem 1.1], or Serre [Ser79, Chapter II, §2, Proposition 3].
By the same token using (13.2.2), there exists a unique absolute value | | 𝑤 on 𝐾
which restricts to | | 𝑣 on 𝐹; we pass freely between these two formulations.
13.2. VALUATIONS 197
a uniformizer for 𝐹.
In general, there is a (unique) maximal unramified subextension 𝐾un ⊆ 𝐾, and the
extension 𝐾 ⊇ 𝐾un is totally ramified.
𝐾
𝑒
𝐾un
𝑓
𝐹
We say that 𝑒 = [𝐾 : 𝐾un ] is the ramification degree and 𝑓 = [𝐾un : 𝐹] the inertial
degree, and the fundamental equality
𝑛 = [𝐾 : 𝐹] = 𝑒 𝑓 (13.2.4)
holds.
13.2.5. Suppose that 𝐹 is a local field (equivalently, the residue field 𝑘 is a finite field).
Then for all 𝑓 ∈ Z ≥1 , there is a unique unramified extension of 𝐹 of degree 𝑓 and such
a field corresponds to the unique extension of the residue field 𝑘 of degree 𝑓 . In an
unramified extension 𝐾 ⊇ 𝐹 of degree [𝐾 : 𝐹] = 𝑓 , we have Nm𝐾 |𝐹 (𝐾 × ) = 𝑅 × 𝜋 𝑓 Z ,
so 𝑏 ∈ Nm𝐾 |𝐹 (𝐾 × ) if and only if 𝑓 | 𝑣(𝑏).
If char 𝑘 ≠ 2, then by Hensel’s lemma, the unramified extension of degree 2 is
given by adjoining a square root of an element of 𝑅 which reduces to the unique
nontrivial class in 𝑘 × /𝑘 ×2 ; if char 𝑘 = 2, then the unramified extension of degree 2
is given by adjoining a root of the polynomial 𝑥 2 + 𝑥 + 𝑡 where 𝑡 ∈ 𝑅 reduces to an
element which is nontrivial in the Artin-Schreier group 𝑘/℘(𝑘) (recalling 12.3.11).
Before proceeding further, we describe local fields by their defining polynomials—
we will need this later in the study of norms and strong approximation.
Lemma 13.2.6 (Krasner’s lemma). Let 𝐾 ⊇ 𝐹 be a finite, Galois extension with
absolute value | | 𝑤 . Let 𝛼, 𝛽 ∈ 𝐾, and suppose that for all 𝜎 ∈ Gal(𝐾 | 𝐹) with
𝜎(𝛼) ≠ 𝛼, we have
|𝛼 − 𝛽| 𝑤 < |𝛼 − 𝜎(𝛼)| 𝑤 . (13.2.7)
Then 𝐹 (𝛼) ⊆ 𝐹 (𝛽).
Intuitively, we can think of Krasner’s lemma as telling us when 𝛽 is closer to 𝛼
than any of its conjugates, then 𝐹 (𝛽) contains 𝛼. It is for this reason that we state the
lemma in terms of absolute values (instead of valuations).
Proof. Let 𝜎 ∈ Gal(𝐾 | 𝐹 (𝛽)) have 𝜎(𝛼) ≠ 𝛼. Then by the ultrametric inequality,
|𝜎(𝛼) − 𝛼| 𝑤 = |𝜎(𝛼) − 𝛽 + 𝛽 − 𝛼| 𝑤 ≤ max(|𝜎(𝛼) − 𝛽| 𝑤 , | 𝛽 − 𝛼| 𝑤 )
(13.2.8)
= max(|𝜎(𝛼 − 𝛽)| 𝑤 , | 𝛽 − 𝛼| 𝑤 ) = |𝛼 − 𝛽| 𝑤 ,
198 CHAPTER 13. QUATERNION ALGEBRAS OVER LOCAL FIELDS
the final equality a consequence of (13.2.2) and the fact that Galois conjugates have the
same norm. This contradicts the existence of 𝜎, so 𝜎(𝛼) = 𝛼 for all 𝜎 ∈ Gal(𝐾 | 𝐹 (𝛽)).
By Galois theory, we conclude that 𝐹 (𝛼) ⊆ 𝐹 (𝛽).
Proof. Since 𝑓 (𝑥) is separable, its discriminant disc( 𝑓 ) is nonzero. The discriminant
is a polynomial function in the coefficients, so by continuity (multivariate Taylor
expansion), there exists 𝛿1 > 0 such that if 𝑔(𝑥) = 𝑥 𝑛 + · · · + 𝑏 0 ∈ 𝐹 [𝑥] has |𝑎 𝑖 − 𝑏 𝑖 | 𝑣 <
𝛿1 for all 𝑖, then |disc(𝑔) − disc( 𝑓 )| 𝑣 < |disc( 𝑓 )| 𝑣 ; by the ultrametric inequality,
we conclude that for such 𝑔(𝑥) we have |disc(𝑔)| 𝑣 = |disc( 𝑓 )| 𝑣 so in particular
disc(𝑔) ≠ 0.
Let 𝑔(𝑥) be as in the previous paragraph; then 𝑔(𝑥) is separable. We first consider
the caseÎ where 𝑓 (𝑥) is irreducible. Let 𝐾Î⊇ 𝐹 be a splitting field for the polynomials
𝑛 𝑛
𝑓 (𝑥) = 𝑖=1 (𝑥 − 𝛼𝑖 ) ∈ 𝐾 [𝑥] and 𝑔(𝑥) = 𝑖=1 (𝑥 − 𝛽𝑖 ). Let | | 𝑤 on 𝐾 extend | | 𝑣 . Let
Finally, let
𝑛
Ö 𝑛
Ö
𝜌(𝑔) = 𝜌(𝑏 0 , . . . , 𝑏 𝑛−1 ) := 𝑔(𝛼𝑖 ) = (𝛼𝑖 − 𝛽 𝑗 ). (13.2.11)
𝑖=1 𝑖, 𝑗=1
[𝐹 (𝛼𝑖 ) : 𝐹] = 𝑛 ≤ [𝐹 (𝛽 𝑗 ) : 𝐹] ≤ 𝑛
as desired.
The case when 𝑓 (𝑥) = 𝑓1 (𝑥) · · · 𝑓𝑟 (𝑥) is reducible follows by repeating the above
argument on each factor, and finishing using the continuity of multiplication among
the coefficients: the details are requested in Exercise 13.17.
13.3. CLASSIFICATION VIA EXTENSIONS OF VALUATIONS 199
𝑤 : 𝐷 → R ∪ {∞}
𝑣(Nm𝐷 |𝐹 (𝛼)) 𝑣(nrd(𝛼)) (13.3.1)
𝛼 ↦→ = ,
[𝐷 : 𝐹] 𝑛
where the equality follows from the fact that Nm𝐷 |𝐹 (𝛼) = nrd(𝛼) 𝑛 (see section 7.8).
Lemma 13.3.2. The map 𝑤 is the unique valuation on 𝐷 extending 𝑣, i.e., the following
hold:
Proof. Since 𝐷 is a division ring, statement (i) is immediate. Statement (ii) follows
from the multiplicativity of nrd and 𝑣. To prove (iii), we may suppose 𝛽 ≠ 0 and so
𝛽 ∈ 𝐷 × . We have
O := {𝛼 ∈ 𝐷 : 𝑤(𝛼) ≥ 0}
Proposition 13.3.4. The ring O is the unique maximal 𝑅-order in 𝐷, consisting of all
elements of 𝐷 that are integral over 𝑅.
We first show the inclusion (⊇) of (13.3.5), and suppose 𝛼 ∈ 𝐷 is integral over
𝑅. Since 𝑅 is integrally closed, by Lemma 10.3.5 the coefficients of the minimal
polynomial 𝑓 (𝑥) ∈ 𝐹 [𝑥] of 𝛼 belong to 𝑅. Since 𝐷 is a division ring, 𝑓 (𝑥) is
irreducible and hence the reduced characteristic polynomial 𝑔(𝑥) is a power of 𝑓 (𝑥)
200 CHAPTER 13. QUATERNION ALGEBRAS OVER LOCAL FIELDS
and thus has coefficients in 𝑅. Up to sign, the constant coefficient of 𝑔(𝑥) is nrd(𝛼),
so 𝑤(𝛼) = 𝑣(nrd(𝛼)) ≥ 0, hence 𝛼 ∈ O.
Next we prove (⊆) in (13.3.5). Suppose 𝛼 ∈ O, so that 𝑤(𝛼) ≥ 0, and let
𝐾 = 𝐹 (𝛼). Let 𝑓 (𝑥) ∈ 𝐹 [𝑥] be the minimal polynomial of 𝛼. We want to conclude
that 𝑓 (𝑥) ∈ 𝑅[𝑥] knowing that 𝑤(𝛼) ≥ 0. But the restriction of 𝑤 to 𝐾 is the unique
extension of 𝑣 to 𝐾, and this is a statement about the extension 𝐾 ⊇ 𝐹 of fields and
therefore follows from the commutative case, Lemma 13.2.1.
We can now prove that O is an 𝑅-order. Scaling an element of 𝐷 × by an appropriate
power of 𝜋 gives it positive valuation, so O𝐹 = 𝐷. To conclude, we must show that O
is finitely generated as an 𝑅-module. Recall that 𝐷 is a central division algebra over 𝐹,
hence a separable 𝐹-algebra, so we may apply Lemma 10.3.7: every 𝛼 ∈ O is integral
over 𝑅 and O is a ring, and the lemma implies that O is an 𝑅-order.
Finally, it follows immediately that O is a maximal 𝑅-order: by Corollary 10.3.3,
every element of an 𝑅-order is integral over 𝑅, and O contains all such elements.
Remark 13.3.6. For a quaternion division algebra 𝐷, we can argue more directly in
the proof of Proposition 13.3.4 using the reduced norm: see Exercise 13.4.
13.3.7. It follows from Proposition 13.3.4 that O is a finitely generated 𝑅-submodule
of 𝐷. But 𝑅 is a PID so in fact O is free of rank [𝐷 : 𝐹] as an 𝑅-module. We have
O× = {𝛼 ∈ 𝐷 : 𝑤(𝛼) = 0} (13.3.8)
since 𝑤(𝛼−1 ) = −𝑤(𝛼), and in particular 𝛼 ∈ O× if and only if nrd(𝛼) ∈ 𝑅 × .
Consequently,
𝑃 := {𝛼 ∈ 𝐷 : 𝑤(𝛼) > 0} = O r O× (13.3.9)
is the unique maximal two-sided (bilateral) ideal of O, as well as the unique left or
right ideal of O. Therefore O is a noncommutative local ring, a noncommutative
ring with a unique maximal left (equivalently, right) ideal.
13.3.10. Let 𝛽 ∈ 𝑃 have minimal (positive) valuation 𝑤(𝛽) > 0. Then for all
0 ≠ 𝛼 ∈ 𝑃 we have 𝑤(𝛼𝛽−1 ) = 𝑤(𝛼) − 𝑤(𝛽) ≥ 0 so 𝛼𝛽−1 ∈ O and 𝛼 ∈ O𝛽. Arguing
on the other side, we have also 𝛼 ∈ 𝛽O. Thus 𝑃 = O𝛽 = 𝛽O = O𝛽O.
Arguing in the same way, we see that every one-sided ideal of O is in fact two-sided,
and every two-sided ideal of O is principally generated by any element with minimal
valuation hence of the form 𝑃𝑟 for some 𝑟 ∈ Z ≥0 .
We are now prepared to give the second proof of the main result in this chapter
(Main Theorem 12.3.2). We now add the hypothesis that 𝐹 is a local field, so that 𝑘 is
a finite field.
Theorem 13.3.11. Let 𝐹 be a nonarchimedean local field. Then the following state-
ments hold.
(a) There is a unique division quaternion algebra 𝐵 over 𝐹, up to 𝐹-algebra iso-
morphism given by
𝐾, 𝜋
𝐵' ,
𝐹
where 𝐾 is the unique quadratic unramified (separable) extension of 𝐹.
13.3. CLASSIFICATION VIA EXTENSIONS OF VALUATIONS 201
(b) Let 𝐵 be as in (a). Then the valuation ring of 𝐵 is O ' 𝑆 ⊕ 𝑆 𝑗, where 𝑆 is the
integral closure of 𝑅 in 𝐾. Moreover, the ideal 𝑃 = O 𝑗 is the unique maximal
ideal; we have 𝑃2 = 𝜋O, and O/𝑃 ⊇ 𝑅/𝔭 is a quadratic extension of finite
fields.
Proof. We begin with existence in part (a), and existence: we prove that 𝐵 = (𝐾, 𝜋 | 𝐹)
is a division algebra. We recall that 𝐵 is a division ring if and only if 𝜋 ∉ Nm𝐾 |𝐹 (𝐾 × )
by Main Theorem 5.4.4 and Theorem 6.4.11. Since 𝐾 ⊇ 𝐹 is unramified, we have
Nm𝐾 |𝐹 (𝐾 × ) = 𝑅 × 𝜋 2Z by 13.2.5. Putting these together gives the result.
Continuing with (a), we now show uniqueness. Let 𝐵 be a division quaternion
algebra over 𝐹. We refer to 13.3.10, and let 𝑃 = O𝛽. Then 𝑤(𝛽) ∈ 12 Z>0 , so
𝑤(𝛽) ≤ 𝑤(𝜋) = 𝑣(𝜋) = 1 ≤ 2𝑤(𝛽) = 𝑤(𝛽2 ); (13.3.12)
we conclude that 𝛽O = 𝑃 ⊇ 𝜋O ⊇ = 𝑃2 𝛽2 O. The map 𝛼 ↦→ 𝛼𝛽 yields an
isomorphism O/𝑃 −→∼ 𝑃/𝑃2 of 𝑘-vector spaces, so
4 = dim 𝑘 (O/𝜋O) ≤ dim 𝑘 (O/𝑃2 ) = dim 𝑘 (O/𝑃) + dim 𝑘 (𝑃/𝑃2 ) = 2 dim 𝑘 (O/𝑃)
(13.3.13)
and thus dim 𝑘 (O/𝑃) ≥ 2, with equality if and only if 𝜋O = 𝑃2 .
As in (13.3.9), we have O r 𝑃 = O× , so the ring O/𝑃 is a division algebra over
𝑘 and hence a finite division ring. By Wedderburn’s little theorem (Exercise 7.29),
we conclude that O/𝑃 is a finite field! So there exists 𝑖 ∈ O such that its reduction
generates O/𝑃 as a finite extension of 𝑘. But 𝑖 satisfies its reduced characteristic
polynomial, a monic polynomial of degree 2 with coefficients in 𝑅, so its reduction
satisfies a polynomial of degree 2 with coefficients in 𝑘. Since 𝑖 is a generator, we
conclude [O/𝑃 : 𝑘] ≤ 2. Together with the conclusion of the previous paragraph,
we conclude that [O/𝑃 : 𝑘] = dim 𝑘 (O/𝑃) = 2, in other words O/𝑃 is a (separable)
quadratic field extension of 𝑘. It then follows from 13.2.5 that 𝐾 := 𝐹 (𝑖) is the unique,
unramified (separable) quadratic extension of 𝐹. Therefore equality holds in (13.3.13)
and 𝑃2 = 𝜋O. Since 𝛽2 O = 𝑃2 = 𝜋O, we have 𝑤(𝛽) = 1/2.
By Exercise 6.2 or 7.26, there exists 𝑏 ∈ 𝐹 × such that 𝐵 ' (𝐾, 𝑏 | 𝐹). Recalling the
first paragraph above, since 𝐵 is a division algebra, we have 𝑏 ∉ Nm𝐾 |𝐹 (𝐾 × ) = 𝑅 × 𝜋 2Z .
Applying Exercise 6.4, we may multiply 𝑏 by a norm from 𝐾 × , so we may suppose
𝑏 = 𝜋, and therefore 𝐵 ' (𝐾, 𝜋 | 𝐹). This concludes the proof of (a).
We turn now to (b), with 𝐵 = (𝐾, 𝜋 | 𝐹) = 𝐾 + 𝐾 𝑗 with 𝑗 2 = 𝜋. Let 𝛼 = 𝑢 + 𝑣 𝑗 ∈ 𝐵
with 𝑢, 𝑣 ∈ 𝐾. Then nrd(𝛼) = Nm𝐾 |𝐹 (𝑢) − 𝜋 Nm𝐾 |𝐹 (𝑣) = 𝑥 − 𝜋𝑦 with 𝑥, 𝑦 ∈ 𝐹 and
𝑣(𝑥) even and 𝑣(𝜋𝑦) odd (as norms from 𝐾). By the ultrametric inequality, we have
𝑤(𝛼) = 𝑣(nrd(𝛼)) ≥ 0 if and only if 𝑣(𝑥), 𝑣(𝑦) ≥ 0 if and only if 𝑢, 𝑣 ∈ 𝑆 (as 𝑆 is the
valuation ring of 𝐾). Therefore O = 𝑆 + 𝑆 𝑗. Since 𝑗 2 = 𝜋, we have 𝑤( 𝑗) = 1/2, so
𝑗O = 𝑃. The remaining statements were proven in the course of proving (a).
Corollary 13.3.14. Let 𝐹 ≠ C be a local field. Let 𝐾 be the unramified quadratic
extension of 𝐹, with h𝜎i = Gal(𝐾 | 𝐹). Then the 𝐹-subalgebra
𝑢 𝜋𝑣
𝐵= : 𝑢, 𝑣 ∈ 𝐾 ⊂ M2 (𝐾)
𝜎(𝑣) 𝜎(𝑢)
is the unique division quaternion algebra over 𝐹 (up to isomorphism).
202 CHAPTER 13. QUATERNION ALGEBRAS OVER LOCAL FIELDS
Proof. Using Theorem 13.3.11(a), we split 𝐵 over 𝐾 as in 2.3.4. (We may also put 𝜋
below the diagonal as in 2.3.12.)
13.4 Consequences
We now observe a few consequences of Theorem 13.3.11.
13.4.2. Let 𝐵 be a division quaternion algebra over 𝐹. In analogy with the case of
field extensions (13.2.4), we define the ramification index of 𝐵 over 𝐹 as 𝑒(𝐵|𝐹) = 2
since 𝑃2 = 𝜋O, and the inertial degree of 𝐵 over 𝐹 as 𝑓 (𝐵|𝐹) = 2 since 𝐵 contains
the unramified quadratic extension 𝐾 of 𝐹, and note the equality
as in the commutative case. (Viewed in this way, 𝐵 is obtained from first an unramified
extension and then a “noncommutative” ramified extension.)
Remark 13.4.3. Theorem 13.3.11, the fundamental result describing division quater-
nion algebras over a local field, is a special case of a more general result as follows.
Let 𝑅 be a complete DVR with maximal ideal 𝔭 = 𝜋𝑅 and 𝐹 := Frac(𝑅).
Let 𝐷 be a (finite-dimensional) division algebra over 𝐹, and let O ⊆ 𝐷 be the
valuation ring and 𝑃 ⊂ O the maximal ideal. Then 𝑃 𝑒 = 𝔭O for some 𝑒 ≥ 1, called the
ramification index; the quotient O/𝑃 is a division algebra over the field 𝑘 = 𝑅/𝔭, and
we let the inertial degree be 𝑓 = dim 𝑘 (O/𝑃). Then 𝑒 𝑓 = dim𝐹 𝐷 = 𝑛2 ; moreover,
if 𝑘 is finite (𝐹 is a local field), then 𝑒 = 𝑓 = 𝑛. For a proof, see Exercise 13.11; or
consult Reiner [Rei2003, Theorems 12.8, 13.3, 14.3]. However, the uniqueness of 𝐷
up to 𝐹-algebra isomorphism no longer holds. If 𝐹 is a local field, then the possibilities
for 𝐷 are classified up to isomorphism by a local invariant inv 𝐷 ∈ ( 𝑛1 Z)/Z ' Z/𝑛Z.
These patch together to give a global result: see Remark 14.6.10.
This classification can be further extended to an arbitrary central simple algebra
𝐵 ' M𝑛 (𝐷) over 𝐹: see Reiner [Rei2003, §17–18].
Splitting of local division quaternion algebras over extension fields is given by the
following simple criterion.
Proposition 13.4.4. Let 𝐵 be a division quaternion algebra over a local field 𝐹, and
let 𝐿 be a separable field extension of 𝐹 of finite degree. Then 𝐿 is a splitting field for
𝐵 if and only if [𝐿 : 𝐹] is even.
13.4. CONSEQUENCES 203
We repeat now Lemma 12.4.6, giving a proof that works without restriction on
characteristic.
Corollary 13.4.6. If char 𝐹 ≠ 2, then the Hilbert symbol defines a symmetric, nonde-
generate bilinear form on 𝐹 × /𝐹 ×2 .
Proof. Let 𝐾 := 𝐹 [𝑥]/(𝑥 2 − 𝑎). The Hilbert symbol gives a well-defined map of sets
𝐹 × /𝐹 ×2 → {±1}
𝑏 ↦→ (𝑎, 𝑏) 𝐹
and we may conclude as in Lemma 12.4.6 if we show that this is a nontrivial group
homomorphism.
First we show it is nontrivial. By Corollary 13.4.5, the field 𝐾 embeds in the
division quaternion algebra 𝐵, so by Exercise 2.5, there exists 𝑏 such that 𝐵 ' (𝑎, 𝑏 |
𝐹), whence (𝑎, 𝑏) 𝐹 = −1.
Next, we show it is a homomorphism. We appeal to Main Theorem 5.4.4. We have
(𝑎, 𝑏) 𝐹 = 1 if and only if 𝑏 ∈ Nm𝐾 |𝐹 (𝐾 × ). So we reduce to showing that if (𝑎, 𝑏) 𝐹 =
(𝑎, 𝑏 0) 𝐹 = −1 for 𝑏, 𝑏 0 ∈ 𝐹 × , then (𝑎, 𝑏𝑏 0) 𝐹 = 1. But by Corollary 7.7.6, since there
is a unique division quaternion algebra, we conclude that 𝑏/𝑏 0 ∈ Nm𝐾 |𝐹 (𝐾 × ); thus
𝑏𝑏 0 = (𝑏 0) 2 (𝑏/𝑏 0) ∈ Nm𝐾 |𝐹 (𝐾 × ) and (𝑎, 𝑏𝑏 0 | 𝐹) ' M2 (𝐹) so (𝑎, 𝑏𝑏 0) 𝐹 = 1, as
claimed.
Remark 13.4.7. The proof of Corollary 13.4.6 (pairing with any 𝐹 × /𝐹 ×2 ) shows that
Conversely, if we know (13.4.8) then the properties of the Hilbert symbol are immedi-
ate. Although it was not hard to prove (13.4.8) when char 𝑘 ≠ 2, to establish its truth
when char 𝑘 = 2, one is led to study higher ramification groups (e.g. Serre [Ser79,
Chapter XV]) eventually leading to local class field theory.
The norm groups played an important role in the proof above, so we conclude by
recording the image of the reduced norm nrd(𝐵×𝑣 ) ⊆ 𝐹𝑣× .
13.5.1. 𝐹 is a locally compact topological field (by definition) but 𝐹 is not itself
compact. The subgroup 𝐹 × = 𝐹 r {0} is given the topology induced from the
embedding
𝐹 × ↩→ 𝐹 × 𝐹
𝑥 ↦→ (𝑥, 𝑥 −1 );
it turns out here that this coincides with the subspace topology 𝐹 × ⊆ 𝐹 (see Exercise
13.15(a)). Visibly, 𝐹 × is open in 𝐹 so 𝐹 × is locally compact.
If 𝐹 is nonarchimedean, with valuation ring 𝑅 and valuation 𝑣, then 𝐹 × is totally
disconnected and further
𝑅 × = {𝑥 ∈ 𝑅 : 𝑣(𝑥) = 0} ⊂ 𝑅
13.5.2. As an 𝐹-vector space, 𝐵 has a unique topology compatible with the topology on
𝐹 as all norms on a topological vector 𝐹-space extending the norm on 𝐹 are equivalent
(the sup norm is equivalent to the sum of squares norm, etc.): see Exercise 13.13. In
particular, two elements are close in the topology on 𝐵 if and only if their coefficients
are close with respect to a (fixed) basis: for example, two matrices in M𝑛 (𝐹) are close
if and only if all of their coordinate entries are close. (Of course, the precise notion
of “close” depends on the choice of norm.) Consequently, 𝐵 is a complete, locally
compact topological ring, taking a compact neighborhood in each coordinate.
13.5.3. The group 𝐵× is a topological group, with the topology given by the embedding
𝐵× 3 𝛼 ↦→ (𝛼, 𝛼−1 ) ∈ 𝐵 × 𝐵. This topology coincides with the subspace topology
(see Exercise 13.15(b)). From this, we can see that 𝐵× is locally compact: the norm
Nm 𝐵 |𝐹 : 𝐵× → 𝐹 × is a continuous map, so 𝐵× = Nm−1 ×
𝐵 |𝐹 (𝐹 ) is open in 𝐵, and an
open subset of a Hausdorff, locally compact space is locally compact in the subspace
topology (Exercise 13.15(c)).
13.5.8. Suppose 𝐵 = 𝐷 is a division algebra. Then the valuation ring O is the maximal
compact subring of 𝐵, for the same reason as in the commutative case (see 13.5.5,
details requested in Exercise 13.18(a)). There is a filtration
O ⊃ 𝑃 ⊃ 𝑃2 ⊃ . . .
206 CHAPTER 13. QUATERNION ALGEBRAS OVER LOCAL FIELDS
O× ⊃ 1 + 𝑃 ⊃ 1 + 𝑃 2 ⊃ . . . . (13.5.9)
As in the second proof of Main Theorem 12.3.2, the quotient O/𝑃 is a finite extension
of the finite residue field 𝑘, so (O/𝑃) × is a finite cyclic group. The maximal two-sided
ideal 𝑃 is principal, generated by an element 𝑗 of minimal valuation, and multiplication
by 𝑗 𝑛 gives an isomorphism O/𝑃 −→ ∼ 𝑃 𝑛 /𝑃 𝑛+1 of 𝑘-vector spaces (or abelian groups)
for all 𝑛 ≥ 1.
Furthermore, for each 𝑛 ≥ 1, there is an isomorphism of groups
∼ (1 + 𝑃 𝑛 )/(1 + 𝑃 𝑛+1 )
O/𝑃 ' 𝑃 𝑛 /𝑃 𝑛+1 −→
(13.5.10)
𝛼 ↦→ 1 + 𝛼.
𝐵1 := {𝛼 ∈ 𝐵 : nrd(𝛼) = 1};
some authors also write SL1 (𝐵) := 𝐵1 . Then 𝐵1 is a closed subgroup of 𝐵× , since the
reduced norm is continuous.
If 𝐵 is a division ring and 𝐹 is archimedean, then 𝐵 ' H and 𝐵1 ' H1 ' SU(2) is
compact (it is identified with the 3-sphere in R4 ). In a similar way, if 𝐵 is a divison ring
and 𝐹 is nonarchimedean, then 𝐵1 is compact: for 𝐵 has a valuation 𝑣 and valuation
ring O, and if 𝛼 ∈ 𝐵 has nrd(𝛼) = 1 then 𝑣(𝛼) = 0 and 𝛼 ∈ O, and consequently
𝐵1 ⊆ O× is closed in a compact set so compact.
If 𝐵 is not a division ring, then either 𝐵 is the product of two algebras or 𝐵 is a
matrix ring over a division ring, and in either case 𝐵1 is not compact.
Remark 13.5.13. The locally compact division algebras over a nonarchimedean field
are necessarily totally disconnected. On the other hand, it is a theorem of Pontryagin
[War89, Theorem 27.2] that if 𝐴 is a connected locally compact division ring, then 𝐴
is isomorphic as a topological ring to either R, C, or H.
Exercises
−1, −1
1. Let 𝐵 := .
Q2
(a) Show that 𝐵 is a division ring that is complete with respect to the discrete
valuation 𝑤 defined by 𝑤(𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗) = 𝑣(𝑡 2 + 𝑥 2 + 𝑦 2 + 𝑧2 ) for
𝑡, 𝑥, 𝑦, 𝑧 ∈ Q2 .
13.5. SOME TOPOLOGY 207
1 + 𝑖 + 𝑗 + 𝑖𝑗
O := Z2 ⊕ Z2𝑖 ⊕ Z2 𝑗 ⊕ Z2 ⊂𝐵
2
is the valuation ring of 𝐵.
2. Let 𝐵 be a division quaternion algebra over a nonarchimedean local field 𝐹.
Give another proof that the unramified quadratic extension 𝐾 of 𝐹 embeds in 𝐵
as follows.
Suppose it does not: then for all 𝛼 ∈ O, the extension 𝐹 (𝛼) ⊇ 𝐹 is ramified,
so there exists 𝑎 ∈ 𝑅 such that 𝛼 − 𝑎 ∈ 𝑃 ∩ 𝐾 (𝛼); let 𝑃 = 𝑗O and write
𝛼 = 𝛼0 = 𝑎 + 𝑗 𝛼1 , and iterate to conclude that 𝛼 = ∞
Í
𝑛=0 𝑎 𝑛 𝑗 with 𝑎 𝑛 ∈ 𝑅. But
𝑛
(a) Show that every one-sided (left or right) ideal of O is a power of the
maximal ideal 𝑃 and hence is two-sided.
(b) Let
[O, O] := h𝛼𝛽 − 𝛽𝛼 : 𝛼, 𝛽 ∈ Oi
be the commutator ideal [O, O] of O, the two-sided ideal generated by
commutators of elements of O. Show that 𝑃 = [O, O].
9. Let 𝐹 be a nonarchimedean local field, let 𝐵 = M2 (𝐹) and O = M2 (𝑅). Show
that there are 𝑞 + 1 right O-ideals of norm 𝔭 corresponding to the elements of
P1 (𝑘) or equivalently the lines in 𝑘 2 .
10. Give another proof of Lemma 13.4.9 using quadratic forms.
11. Let 𝐹 be a nonarchimedean local field with valuation ring 𝑅, maximal ideal 𝔭,
and residue field 𝑘. Let 𝐷 be a division algebra over 𝐹 with dim𝐹 𝐷 = 𝑛2 ,
with valuation ring O and maximal two-sided ideal 𝑃. Show that O/𝑃 is finite
extension of 𝑘 of degree 𝑛, and 𝑃 𝑛 = 𝔭O (cf. Remark 13.4.3).
12. Show that (13.5.10) is an isomorphism of (abelian) groups.
13. Let 𝐹 be a field with absolute value | | and 𝑉 a finite-dimensional 𝐹-vector space.
(a) Let 𝑥 1 , . . . , 𝑥 𝑛 be a basis for 𝑉, and define
k𝑎 1 𝑥1 + · · · + 𝑎 𝑛 𝑥 𝑛 k := max(|𝑎 1 |, . . . , |𝑎 𝑛 |)
for 𝑎 𝑖 ∈ 𝐹. Show that 𝑉 is a metric space with distance 𝑑 (𝑥, 𝑦) = k𝑥 − 𝑦k.
(b) Show that the topology on 𝑉 is independent of the choice of basis in (a).
(c) Finally, show that if 𝐹 is complete with respect to | |, then 𝑉 is also
complete.
14. Let 𝐹 be a topological field. Show that the coarsest topology (fewest open sets)
in which multiplication on M𝑛 (𝐹) is continuous is the coordinate topology.
15. Let 𝐹 be a local field.
(a) The group 𝐹 × has the structure of topological group under the embedding
𝑥 ↦→ (𝑥, 𝑥 −1 ) ∈ 𝐹 × 𝐹 (under the subspace topology in 𝐹 × 𝐹). Show that
this topology coincides with the subspace topology 𝐹 × ⊆ 𝐹.
(b) More generally, let 𝐵 be a finite-dimensional 𝐹-algebra. Show that the
the topology on 𝐵× induced by 𝛼 ↦→ (𝛼, 𝛼−1 ) ∈ 𝐵 × 𝐵 coincides with the
subspace topology 𝐵× ⊆ 𝐵.
(c) Show that an open subset of a Hausdorff, locally compact space is locally
compact in the subspace topology.
16. Let 𝐹 be a finite extension of Q2 . Show that (−1, −1) 𝐹 = (−1) [𝐹 :Q2 ] .
17. Finish the proof of Lemma 13.2.9.
18. Let 𝐷 be a division algebra over a nonarchimedean local field 𝐹. We recall (see
13.5.2) that 𝐷 is a complete, locally compact topological ring.
(a) Verify (as in 13.5.5) that O is the maximal compact subring of 𝐵.
(b) Show that 𝐵× /𝐹 × is a compact topological group.
Chapter 14
In this chapter, we discuss quaternion algebras over global fields and characterize them
up to isomorphism.
14.1 ⊲ Ramification
To motivate the classification of quaternion algebras over Q, we consider by analogy a
classification of quadratic fields. We restrict to the following class of quadratic fields
for the best analogy.
√
Definition 14.1.1. A quadratic field 𝐹 = Q( 𝑑) of discriminant 𝑑 ∈ Z is mildly
ramified if 8 - 𝑑.
√
A quadratic field 𝐹 is mildly ramified if and only if 𝐹 = Q( 𝑚) where 𝑚 ≠ 1 is
√ then 𝑑 = 𝑚 or 𝑑 = 4𝑚 according as 𝑚 = 1, 3 (mod 4).
odd and squarefree;
Let 𝐹 = Q( 𝑑) be a mildly ramified quadratic field of discriminant 𝑑 ∈ Z and
let 𝑅 be its ring of integers. A prime 𝑝 ramifies in 𝐹, i.e. 𝑝𝑅 = 𝔭2 for a prime ideal
𝔭 ⊂ 𝑅, if and only if 𝑝 | 𝑑.
But a discriminant 𝑑 can be either positive or negative; to put this bit of data on
the same footing, we define the set of places of Q to be the primes together with the
symbol ∞, and we make √ the convention that ∞ ramifies in 𝐹 if 𝑑 < 0 and is unramified
if 𝑑 > 0. Let 𝐹 = Q( 𝑑) be a mildly ramified quadratic field, and let Ram(𝐹) be the
set of places that ramify in 𝐹. The set Ram(𝐹) determines 𝐹 up to isomorphism, since
the discriminant of 𝐹 is the product of the odd primes in Ram(𝐹), multiplied by 4 if
2 ∈ Ram(𝐹) and by −1 if ∞ ∈ Ram(𝐹). (For bookkeeping reasons, in this context it
would probably therefore be better to consider 4 and −1 as primes, but we will resist
the inducement here.) However, not every finite set of places Σ occurs: the product 𝑑
corresponding to Σ is a discriminant if and only if 𝑑 ≡ 0, 1 (mod 4). We call this a
parity condition on the set of ramifying places of a mildly ramified quadratic field:
209
210 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
Corollary 14.2.3. Let 𝐵 be a quaternion algebra over Q. Then the set Ram 𝐵 is finite
of even cardinality.
The law of Hilbert reciprocity, as it turns out, is a core premise in number theory:
it is equivalent to the law of quadratic reciprocity
𝑝 𝑞 𝑝−1 𝑞−1
= (−1) 2 2 (14.2.4)
𝑞 𝑝
Proof of Proposition 14.2.1. Since each local Hilbert symbol is bilinear, it suffices to
prove the statement when 𝑎, 𝑏 ∈ Z are equal to either −1 or a prime number. The
Hilbert symbol is also symmetric, sowe may interchange
𝑎, 𝑏.
𝑎, 𝑏 −1, −1
If 𝑎 = 𝑏 = −1, then 𝐵 = = is the rational Hamiltonians, and
Q Q
(−1, −1)∞ = (−1, −1)2 = −1 and (−1, −1) 𝑣 = 1 if 𝑣 ≠ 2, ∞, by the computation of
the even Hilbert symbol (12.4.13). Similarly, the cases with 𝑎 = −1, 2 follow from the
supplement (14.2.5), and are requested in Exercise 14.1.
𝑝, 𝑝 −1, 𝑝
So we may suppose 𝑎 = 𝑝 and 𝑏 = 𝑞 are primes. If 𝑝 = 𝑞 then '
Q Q
and we reduce to the previous case, so we may suppose 𝑝 ≠ 𝑞. Since 𝑝, 𝑞 > 0, we
have ( 𝑝, 𝑞)∞ = 1. We have ( 𝑝, 𝑞)ℓ = 1 for all primes ℓ - 2𝑝𝑞, and
𝑞 𝑝
( 𝑝, 𝑞) 𝑝 = (𝑞, 𝑝) 𝑝 = and ( 𝑝, 𝑞)𝑞 =
𝑝 𝑞
by 12.4.12. Finally,
i.e., ( 𝑝, 𝑞)2 = (−1) ( 𝑝−1) (𝑞−1)/4 , again by the computation of the even Hilbert symbol
(12.4.13). Thus the product becomes
( 𝑝−1) (𝑞−1)/4 𝑝 𝑞
Ö
( 𝑝, 𝑞) 𝑣 = (−1) =1
𝑣
𝑞 𝑝
by quadratic reciprocity.
Hilbert reciprocity has several aesthetic advantages over the law of quadratic reci-
procity. For one, it is simpler to write down! Also, Hilbert believed that his reciprocity
law is a kind of analogue of Cauchy’s integral theorem, expressing an integral as a
sum of residues (Remark 14.6.4). The fact that a normalized product over all places is
trivial also arises quite naturally: if we define for 𝑥 ∈ Q× and a prime 𝑝 the normalized
absolute value
|𝑥| 𝑝 := 𝑝 −𝑣𝑝 ( 𝑥) ,
and |𝑥| ∞ the usual archimedean absolute value, then
Ö
|𝑥| 𝑣 = 1
𝑣
by unique factorization in Z; this is called the product formula for Q, for obvious
reasons.
From the tight relationship between quaternion algebras and ternary quadratic
forms, we obtain the following corollary.
Corollary 14.2.6. Let 𝑄 be a nondegenerate ternary quadratic form over Q. Then the
set of places 𝑣 such that 𝑄 𝑣 is anisotropic is finite and of even cardinality.
14.2. ⊲ HILBERT RECIPROCITY OVER THE RATIONALS 213
Proof. In the bijection implied by Main Theorem 5.2.5, the quadratic form 𝑄 cor-
responds to a quaternion algebra 𝐵 = (𝑎, 𝑏 | Q), and by Main Theorem 5.4.4, 𝑄 is
isotropic if and only if 𝐵 is split if and only if (𝑎, 𝑏)Q = 1. By functoriality, the same
is true over each completion Q 𝑣 for 𝑣 a place of Q, and therefore the set of places
𝑣 where 𝑄 𝑣 is isotropic is precisely the set of ramified places in 𝐵. The result then
follows by Hilbert reciprocity.
To conclude this section, we show that every allowable product of Hilbert symbols
is obtained.
Proposition 14.2.7. Let Σ be a finite set of places of Q of even cardinality. Then there
exists a quaternion algebra 𝐵 over Q with Ram 𝐵 = Σ.
Remark 14.2.8. Albert [Alb34, Theorem 2, Theorem 3] already sought to simplify
the presentation of a quaternion algebra by a series of transformations, the content
of which is contained in Proposition 14.2.7; this was further investigated by Latimer
[Lat35].
Just as with Hilbert reciprocity, Proposition 14.2.7 touches on a deep statement in
number theory concerning primes, due to Dirichlet.
Theorem 14.2.9 (Infinitude of primes in arithmetic progression). Given 𝑎, 𝑛 ∈ Z
coprime, there are infinitely many primes 𝑝 ≡ 𝑎 (mod 𝑛).
Proof. See e.g. Serre [Ser73, Chapter VI] or Apostol [Apo76, Chapter 7]. We will
prove this theorem in Exercise 26.11 as a consequence of the analytic class number
formula.
Remark 14.2.10. Theorem 14.2.9 seems to require analysis. (For algebraic proofs in
special cases, see e.g., Neukirch [Neu99, Exercise I.10.1] and Lenstra–Stevenhagen
[LS91].) Ram Murty [Mur88] showed that a “Euclidean proof” of the infinitude of
primes 𝑝 ≡ 𝑎 (mod 𝑛) is possible if and only if 𝑎 2 ≡ 1 (mod 𝑛), and Paul Pollack
[Pol2010] has shown that Schnizel’s Hypothesis H gives a heuristic for this. This
crucial role played by analytic methods motivates part III of this monograph.
We now prove Proposition 14.2.7 assuming Theorem 14.2.9.
Î
Proof. Let 𝐷 := 𝑝 ∈Σ 𝑝 be the product of the primes in Σ, and let 𝑢 := −1 if ∞ ∈ Σ
and 𝑢 := 1 otherwise. Let 𝐷 ♦ := 𝑢𝐷. We consider quaternion algebras of the form
♦ ♦
𝑞 ,𝐷
𝐵=
Q
and (
1 (mod 8), if 2 - 𝐷;
𝑞♦ ≡ (14.2.12)
5 (mod 8), if 2 | 𝐷.
There exists a prime satisfying the conditions (14.2.11)–(14.2.12) by Theorem 14.2.9,
since the condition to be a quadratic nonresidue is a congruence condition on 𝑞 ♦ and
hence on 𝑞 modulo 𝑝.
We now verify that 𝐵 has Ram 𝐵 = Σ. We have (𝑞 ♦ , 𝐷 ♦ )∞ = 𝑢 by choice of signs
and (𝑞 ♦ , 𝐷 ♦ ) 𝑝 = 1 for all 𝑝 - 2𝑑𝑞. We compute that
♦
𝑞
(𝑞 ♦ , 𝐷 ♦ ) 𝑝 = = −1 for all odd 𝑝 | 𝐷
𝑝
by (14.2.11). For 𝑝 = 2, we find that (𝑞 ♦ , 𝐷 ♦ )2 = −1 or (𝑞 ♦ , 𝐷 ♦ )2 = 1 according as
2 | 𝐷 or not by the computation of the even Hilbert symbol (12.4.13). This shows that
Σ ⊆ Ram 𝐵 ⊆ Σ ∪ {𝑞}.
The final symbol (𝑞 ♦ , 𝐷 ♦ )𝑞 is determined by Hilbert reciprocity (Proposition 14.2.1):
since #Σ is already
♦ ♦even, we must have (𝑞 ♦ , 𝐷 ♦ )𝑞 = 1. Therefore the quaternion
𝑞 ,𝐷
algebra 𝐵 := has Σ = Ram 𝐵.
Q
Example 14.2.13. Let 𝐵 = (𝑎, 𝑏 | Q) be a quaternion algebra of prime discriminant
𝐷 = 𝑝 over Q. Then:
(i) For 𝐷 = 𝑝 = 2, we take 𝑎 = 𝑏 = −1;
(ii) For 𝐷 = 𝑝 ≡ 3 (mod 4), we take 𝑏 = −𝑝 and 𝑎 = −1;
(iii) For 𝐷 = 𝑝 ≡ 1 (mod 4), we take 𝑏 = −𝑝 and 𝑎 = −𝑞 where 𝑞 ≡ 3 (mod 4) is
𝑞
prime and = −1.
𝑝
−𝑝 𝑞
In case (iii), by qudaratic reciprocity =− = 1 so indeed 𝐵 is not ramified at
𝑞 𝑝
𝑝. In the proof of Theorem 14.2.7 above, we would have required the more restrictive
condition 𝑞 ≡ 3 (mod 8), but we can look again at the table of even Hilbert symbols
(12.4.16): since 𝑏 = −𝑝 = −1, 3 (mod 8), we may take 𝑎 = −𝑞 = 1, −3 (mod 8) freely,
so 𝑞 ≡ 3 (mod 4).
Similarly, for discriminant 𝐷 the product of two (distinct) primes:
(i) For 𝐷 = 2𝑝 with 𝑝 ≡ 3 (mod 4), we take 𝑎 = −1 and 𝑏 = 𝑝;
(ii) For 𝐷 = 2𝑝 with 𝑝 ≡ 5 (mod 8), we take 𝑎 = 2 and 𝑏 = 𝑝;
(iii) For 𝐷 = 𝑝𝑞 with 𝑝 ≡ 𝑞 ≡ 3 (mod 4), we take 𝑎 = −1 and
𝑏 = 𝑝𝑞;
𝑞
(iv) For 𝐷 = 𝑝𝑞 with 𝑝 ≡ 1 (mod 4) or 𝑞 ≡ 1 (mod 4) and ≠ 1, we take 𝑎 = 𝑝
𝑝
and 𝑏 = 𝑞.
For other explicit presentations of quaternion algebras over Q with specified dis-
criminant, see Alsina–Bayer [AB2004, §1.1.2]. See Example 15.5.7 for some explicit
maximal orders.
14.3. ⊲ HASSE–MINKOWSKI THEOREM OVER THE RATIONALS 215
Proposition 14.3.1. Let 𝐵, 𝐵 0 be quaternion algebras over Q. Then the following are
equivalent:
(i) 𝐵 ' 𝐵 0;
(ii) Ram 𝐵 = Ram 𝐵 0;
(iii) 𝐵 𝑣 ' 𝐵 0𝑣 for all places 𝑣 ∈ Pl(Q); and
(iv) 𝐵 𝑣 ' 𝐵 0𝑣 for all but one place 𝑣.
Corollary 14.3.2. Let 𝐵 be a quaternion algebra over Q. Then 𝐵 ' M2 (Q) if and
only if 𝐵 𝑝 ' M2 (Q 𝑝 ) for all primes 𝑝.
Proof. Apply Proposition 14.3.1 (i) ⇔ (iv) with 𝐵 0 = M2 (Q) checking all but the
archimedean place.
By the equivalence between quaternion algebras and quadratic forms (see Chapter
5, specifically section 5.2), the statement of Proposition 14.3.1 is equivalent to the
statement that a ternary quadratic form over Q is isotropic if and only if it is isotropic
over all (but one) completions. In fact, the more general statement is true—and again
we come in contact with a deep result in number theory.
𝑎𝑥 2 + 𝑏𝑦 2 + 𝑐𝑧 2 = 0
has a nontrivial solution 𝑥, 𝑦, 𝑧 ∈ Q if and only if 𝑎, 𝑏, 𝑐 do not all have the same sign
and
−𝑎𝑏, −𝑏𝑐, −𝑎𝑐 are quadratic residues modulo |𝑐|, |𝑎|, | 𝑏|, respectively.
Proof. First, the conditions for solvability are necessary. The condition on signs is
necessary for a solution in R. If 𝑎𝑥 2 + 𝑏𝑦 2 + 𝑐𝑧2 = 0 with 𝑥, 𝑦, 𝑧 ∈ Q not all zero, then
scaling we may suppose 𝑥, 𝑦, 𝑧 ∈ Z satisfy gcd(𝑥, 𝑦, 𝑧) = 1; if 𝑝 | 𝑐 then 𝑝 - 𝑦 (else
𝑝 | 𝑥 and 𝑝 | 𝑧, contradiction), so (𝑥/𝑦) 2 ≡ (−𝑏/𝑎) (mod |𝑐|) and −𝑏𝑎 is a quadratic
residue modulo |𝑐|; the other conditions hold by symmetry.
216 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
because | 𝑏| ≥ 2.
Now write 𝑏 0 = 𝑏 00𝑢 2 with 𝑏 00, 𝑢 ∈ Z and 𝑏 00 squarefree. Then | 𝑏 00 | ≤ | 𝑏 0 | < | 𝑏|
and 𝑏 00 is a norm if and only if 𝑏 0 is a norm. With these manipulations, we propagate
the hypothesis that |𝑎| is a square modulo | 𝑏 00 | and | 𝑏 00 | is a square modulo |𝑎|.
Therefore, the induction hypothesis applies to the equation 𝑎𝑥 2 + 𝑏 00 𝑦 2 = 𝑧2 , and the
proof is complete.
Proof. If 𝑄 is isotropic, then 𝑄 𝑣 is isotropic for all 𝑣. For the converse, suppose that
𝑄 𝑣 is isotropic for all places 𝑣 of Q. As in the proof of Legendre’s Theorem 14.3.4,
we may suppose 𝑄(𝑥, 𝑦, 𝑧) = 𝑎𝑥 2 + 𝑏𝑦 2 − 𝑧2 . The fact that 𝑄 is isotropic over R
implies that 𝑎, 𝑏 are not both negative. Now let 𝑝 | 𝑎 be odd. The condition that 𝑄 𝑝
is isotropic is equivalent to (𝑎, 𝑏) 𝑝 = (𝑏/𝑝) = 1; putting these together, we conclude
that 𝑏 is a quadratic residue modulo |𝑎|. The same holds for 𝑎, 𝑏 interchanged, so
(14.3.5) holds and the result follows.
(i) 𝑡 ∈ 𝑡 𝑝 Q×2
𝑝 for all primes 𝑝 | 𝑑,
(ii) 𝑡 and 𝑡 ∞ have the same sign, and
(iii) 𝑝 - 𝑡 for all primes 𝑝 - 𝑑 except possibly for one prime 𝑞 - 𝑑.
Now the quadratic form h𝑎, 𝑏, −𝑡i is isotropic for all 𝑝 | 𝑑 and at ∞ by construction
and at all primes 𝑝 - 𝑑 except 𝑝 = 𝑞 since 𝑝 - 𝑎𝑏𝑡. Therefore, by case 𝑛 = 3 (using
the “all but one” in Corollary 14.3.6), the form h𝑎, 𝑏, −𝑡i is isotropic.
On the other side, if 𝑛 = 4, then the form h𝑡i 𝑄 0 is isotropic by the same argument.
If 𝑛 ≥ 5, then we apply the induction hypothesis to 𝑄 0: the hypothesis holds, since 𝑄 0
is isotropic at ∞ and all 𝑝 | 𝑑 by construction, and for all 𝑝 - 𝑑 the completion 𝑄 0𝑝 is
a nondegenerate form in ≥ 3 variables over Z 𝑝 so is isotropic by the results of section
12.3, using Hensel’s lemma to lift a solution modulo the odd prime 𝑝.
Putting these two pieces together, we find that 𝑄 is isotropic over Q.
Corollary 14.3.7. Let 𝑄, 𝑄 0 be quadratic forms over Q in the same number of vari-
ables. Then 𝑄 ' 𝑄 0 if and only if 𝑄 𝑣 ' 𝑄 0𝑣 for all places 𝑣.
Proof. The implication (⇒) is immediate. We prove (⇐) by induction on the number
of variables, the case of 𝑛 = 0 variables being clear. By splitting the radical (4.3.9),
we may suppose that 𝑄, 𝑄 0 are nondegenerate. Let 𝑎 ∈ Q× be represented by 𝑄. Since
𝑄 𝑣 ' 𝑄 0𝑣 the quadratic form h−𝑎i 𝑄 0 is isotropic at 𝑣 for all 𝑣, so 𝑄 0 represents 𝑎
(Lemma 5.4.3). In both cases, we can write 𝑄 ' h−𝑎i 𝑄 1 and 𝑄 0 ' h−𝑎i 𝑄 10 for
quadratic forms 𝑄 1 , 𝑄 10 in one fewer number of variables. Finally, by Witt cancellation
(Theorem 4.2.22), from 𝑄 𝑣 ' 𝑄 0𝑣 we have (𝑄 1 ) 𝑣 ' (𝑄 10 ) 𝑣 for all 𝑣, so by induction
𝑄 1 ' 𝑄 10 , and thus 𝑄 ' 𝑄 0.
Proof of Proposition 14.3.1. The implications (i) ⇒ (ii) and (iii) ⇒ (iv) are immediate.
For the implication (ii) ⇒ (iii): either 𝑣 ∈ Ram 𝐵, in which case 𝐵 𝑣 ' 𝐵 0𝑣 is the unique
division algebra over Q 𝑣 (Theorem 12.1.5), or 𝑣 ∉ Ram 𝐵, in which case 𝐵 𝑣 '
M2 (Q 𝑣 ) ' 𝐵 0𝑣 by definition. For the implication (iv) ⇒ (i), recalling Theorem 5.1.1,
218 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
Proof of Main Theorem 14.1.3. The map 𝐵 ↦→ Ram 𝐵 has the desired codomain, by
Hilbert reciprocity (Proposition 14.2.1); it is surjective by Proposition 14.2.7; and it
is injective by Corollaries 14.3.6 and 14.3.7. The second bijection (with squarefree
integers) is immediate.
Proof. Looking modulo 8, we see that the provided condition is necessary (Exercise
14.3(a)). Conversely, suppose 𝑛 > 0 is not of the form 𝑛 = 4𝑎 (8𝑏 + 7), or equivalently
that −𝑛 ∉ Q×22 (Exercise 14.4). We may suppose 𝑎 = 0, 1.
Let 𝐵 = (−1, −1 | Q) be the rational Hamiltonians. We have Ram 𝐵 = {2, ∞},
which is to say the associated ternary quadratic form 𝑥 2 + 𝑦 2 + 𝑧2 is isotropic over Q 𝑝
for all odd primes 𝑝. Consider the quadratic form 𝑄(𝑥, 𝑦, 𝑧, 𝑤) = 𝑥 2 + 𝑦 2 + 𝑧 2 − 𝑛𝑤 2 .
Then 𝑄 is isotropic over R since 𝑛 > 0, and isotropic over all Q 𝑝 with 𝑝 odd taking
𝑤 = 0. The form is also isotropic over Q2 (Exercise 14.3), lifting a solution modulo
8 via Hensel’s lemma. By the Hasse–Minkowski theorem (Theorem 14.3.3), 𝑄 is
isotropic over Q, so there exist 𝑥, 𝑦, 𝑧, 𝑤 ∈ Q not all zero such that 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛𝑤 2 .
We must have 𝑤 ≠ 0 by positivity, and dividing through we get 𝑥, 𝑦, 𝑧 ∈ Q not all zero
such that 𝑥 2 + 𝑦 2 + 𝑧 2 = 𝑛. Let 𝛼 = 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ∈ 𝐵. Then 𝛼2 + 𝑛 = 0 and 𝛼 ∈ 𝐵 is
integral.
Let O0 ⊂ 𝐵 be a maximal order containing 𝛼, and let O be the Hurwitz order. By
Proposition 11.3.7, O0 is conjugate to O; after conjugating, we may suppose 𝛼 ∈ O.
But trd(𝛼) = 0, so necessarily 𝛼 ∈ Zh𝑖, 𝑗i and 𝑥, 𝑦, 𝑧 ∈ Z with nrd(𝛼) = 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛
as desired.
See also Exercise 14.5 for a variant of the proof of the three-square theorem staying
in the language of quaternions.
14.4. GLOBAL FIELDS 219
𝐾 ⊗𝐹 𝐹𝑣 ' 𝐾1 × · · · × 𝐾𝑟 (14.4.5)
where 𝑓 (𝑥) = 𝑓1 (𝑥) · · · 𝑓𝑟 (𝑥) ∈ 𝐹𝑣 [𝑥] is the factorization of 𝑓 (𝑥) into irreducibles in
𝐹𝑣 [𝑥], distinct because 𝑓 is separable. Thus each 𝐾𝑖 is a local field by the classification
220 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
𝐾 ↩→ 𝐾 ⊗𝐹 𝐹𝑣 → 𝐾𝑖
becomes Ö
k𝑥k 𝑣 = 1. (14.4.7)
𝑣 ∈Pl 𝐹
14.4.8. The set of places Pl(Q) of Q consists of the archimedean real place, induced
by the embedding Q ↩→ R and the usual absolute value |𝑥| ∞ , and the set of nonar-
chimedean places indexed by the primes 𝑝 given by the embeddings Q ↩→ Q 𝑝 , with
the preferred absolute value
|𝑥| 𝑝 = 𝑝 −𝑣𝑝 ( 𝑥) .
The statement of the product formula for 𝑥 ∈ Q is
Ö
|𝑥| ∞ 𝑝 −𝑣𝑝 ( 𝑥) = 1; (14.4.9)
𝑝
rearranging
Í as over Q, but now also taking the logarithm in base 𝑝, (14.4.9) is equivalent
to 𝑓 (deg 𝑓 ) ord 𝑓 (𝑥) = deg 𝑥 which follows from unique factorization in F 𝑝 [𝑡].
14.4.12. More generally, let 𝐾 ⊇ 𝐹 be a finite, separable extension of global fields.
Let 𝑣 be a place of 𝐹 with a preferred absolute value and let 𝑤 be a place of 𝐾 above
𝑣. Then the preferred absolute value for 𝑤 is the unique one extending 𝑣, namely
𝑤 :𝐹𝑣 ]
|𝑥| 𝑤 = |Nm𝐾𝑤 |𝐹𝑣 (𝑥)| 1/[𝐾
𝑣
for all 𝑥 ∈ 𝐾, a consequence of (14.4.5) [Jan96, Chapter II, Theorem 5.2; Cas78, §11,
Theorem, p. 59; Neu99, Chapter II, Corollary (8.4)].
In particular, if 𝐹 satisfies the product formula (14.4.6) with respect to preferred
absolute values, then so does 𝐾, since
Ö Ö Ö Ö
|𝑥| 𝑚
𝑤
𝑤
= |𝑥| 𝑚𝑤
𝑤 = |Nm𝐾 |𝐹 (𝑥)| 𝑚
𝑣 = 1.
𝑣
(14.4.14)
𝑤 𝑣 𝑤 |𝑣 𝑣
Remark 14.4.15. The definitions for the preferred absolute values are pretty dry—
sorry! But we will see later that they are natural from the perspective of Haar
measure: see section 29.3 and ultimately (29.6.3).
We will also make use of the following notation in many places in the text. Let 𝐹
be a global field.
Definition 14.4.16. A set S ⊆ Pl 𝐹 is eligible if S is finite, nonempty, and contains all
archimedean places of 𝐹.
Definition 14.4.17. Let S be an eligible set of places. The ring of S-integers in 𝐹 is
the set
𝑅 (S) := {𝑥 ∈ 𝐹 : 𝑣(𝑥) ≥ 0 for all 𝑣 ∉ S}. (14.4.18)
A global ring is a ring of S-integers in a global field for an associated eligible set S.
The expression (14.4.18) makes sense, since if 𝑣 ∉ S then by hypothesis 𝑣 is
nonarchimedean. When no confusion can result, we will abbreviate 𝑅 = 𝑅 (S) for a
global ring 𝑅.
Example 14.4.19. If 𝐹 is a number field and S consists only of the archimedean places
in 𝐹 then 𝑅 (S) is the ring of integers in 𝐹, the integral closure of Z in 𝐹, also denoted
𝑅 (S) = Z𝐹 . If 𝐹 is a function field, corresponding to a curve 𝑋, then 𝑅 (S) is the ring
of all rational functions with no poles outside S. (So in all cases, it is helpful to think
of the ring 𝑅 (S) as consisting of those elements of 𝐹 with “no poles outside S”.)
222 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
Proof. Write 𝐵 = (𝐾, 𝑏 | 𝐹). Since 𝐹 has only finitely many archimedean places, we
may suppose 𝑣 is nonarchimedean. The extension 𝐾 ⊇ 𝐹 is ramified at only finitely
many places, so we may suppose that 𝐾 ⊇ 𝐹 is unramified at 𝑣. Finally, 𝑣(𝑏) = 0 for
all but finitely many 𝑣, so we may suppose 𝑣(𝑏) = 0. But then under these hypotheses,
𝐵 𝑣 = (𝐾 𝑣 , 𝑏 | 𝐹𝑣 ) is split, by Corollary 13.4.1.
Motivated by the fact that the discriminant of a quadratic field extension is divisible
by ramifying primes, we make the following definition.
Definition 14.5.4. The 𝑅-discriminant of 𝐵 is the 𝑅-ideal
Ö
disc𝑅 (𝐵) = 𝔭⊆𝑅
𝔭∈Ram 𝐵
𝔭∉S
14.5.8. If 𝑣 is a complex place, then 𝑣 is necessarily split, since the only quaternion
algebra over C is M2 (C); therefore, if 𝐵 is a totally definite quaternion algebra over a
number field 𝐹, then 𝐹 is totally real.
In other words, if 𝐵 is a quaternion algebra over a global field, then the set of places
of 𝐹 where 𝐵 is ramified is finite and of even cardinality, this set uniquely determines
𝐵 up to isomorphism, and every such set occurs.
Proof. We give a proof in section 26.8, which itself relies on an analytic result (The-
orem 26.8.19) proven in Chapter 29.
Alternatively, this statement can also be viewed a direct consequence of a (hard-
earned) fundamental exact sequence in class field theory: see Remark 14.6.10.
Recall the definition of the Hilbert symbol (as in section 12.4), computed explicitly
for 𝑣 an odd nonarchimedean place (12.4.9): for a place 𝑣 of 𝐹, we abbreviate
(𝑎, 𝑏) 𝐹𝑣 = (𝑎, 𝑏) 𝑣 . We also recall Lemma 14.5.3 that (𝑎, 𝑏) 𝑣 = 1 for all but finitely
many places 𝑣.
Corollary 14.6.2 (Hilbert reciprocity). Let 𝐹 be a global field with char 𝐹 ≠ 2 and let
𝑎, 𝑏 ∈ 𝐹 × . Then Ö
(𝑎, 𝑏) 𝑣 = 1. (14.6.3)
𝑣 ∈Pl 𝐹
Proof. Immediate from Main Theorem 14.6.1: Hilbert reciprocity is equivalent to the
statement that # Ram 𝐵 is even.
Remark 14.6.4. Stating the reciprocity law in the form (14.6.3) is natural from the
point of view of the product formula (14.4.6). And Hilbert reciprocity can be rightly
seen as a law of quadratic reciprocity for number fields (as we saw in section 14.2 for
𝐹 = Q). (For more, see Exercise 14.15.)
Hilbert saw his reciprocity law (Corollary 14.6.2) as an analogue of Cauchy’s inte-
gral theorem [Hil32, p. 367–368]; for more on this analogy, see Vostokov [Vos2009].
Corollary 14.6.5 (Local-global principle for quaternion algebras). Let 𝐵, 𝐵 0 be quater-
nion algebras over 𝐹. Then the following are equivalent:
(i) 𝐵 ' 𝐵 0;
224 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
Proof. For the equivalence (i) ⇔ (ii) ⇔ (iii), combine Main Theorem 14.6.1 and the
fact that for a noncomplex place 𝑣 there is a unique division algebra over 𝐹𝑣 . The
equivalence (iii) ⇔ (iv) follows from the parity constraint, since if 𝑣 is a place and
Ram 𝐵 r {𝑣} = Σ, then 𝑣 ∈ Ram 𝐵 or not according as #Σ is odd or even.
Proof. The equivalence (i) ⇔ (ii) is a consequence of Corollary 14.6.5: they are both
equivalent to Ram 𝐵 𝐾 = ∅, since 𝐾 splits 𝐵 if and only if 𝐵 ⊗𝐹 𝐾 ' M2 (𝐾) if and only
if Ram(𝐵 ⊗𝐹 𝐾) = ∅ if and only if for all places 𝑤 of 𝐾 we have 𝐵 ⊗𝐹 𝐾 𝑤 ' M2 (𝐾 𝑤 ).
The equivalence (i) ⇔ (iii) was given by Lemmas 5.4.7 and 6.4.12.
The implication (iii) ⇒ (iv) is clear. For the implication (iv) ⇒ (v), if 𝑣 ∈ Ram 𝐵,
then 𝐵 𝑣 is a division algebra; so if 𝐾 𝑣 is not a field, then we cannot have 𝐾 𝑣 ↩→ 𝐵 𝑣 .
Finally, for (v) ⇒ (ii), let 𝑤 ∈ Pl 𝐾 with 𝑤 | 𝑣 ∈ Pl 𝐹. If 𝑣 ∉ Ram 𝐵 then already 𝐹𝑣
splits 𝐵; otherwise, 𝑣 ∈ Ram 𝐵 and 𝐾 𝑣 = 𝐾 𝑤 is a field with [𝐾 𝑤 : 𝐹𝑣 ] = 2, so by
Proposition 13.4.4, 𝐾 𝑤 splits 𝐵.
14.6.8. The equivalences (iii) ⇔ (iv) ⇔ (v) in Proposition 14.6.7 hold also for the
separable 𝐹-algebra 𝐾 = 𝐹 × 𝐹: for there is an embedding 𝐹 × 𝐹 ↩→ 𝐵 if and only if
𝐵 ' M2 (𝐹).
We also record the statement of the Hasse–Minkowski theorem over global fields,
generalizing Theorem 14.3.3.
14.7. THEOREMS ON NORMS 225
Proof. The same comments as in the proof of Main Theorem 14.6.1 apply: we give
a proof in section 26.8. But see also O’Meara [O’Me73, §§65–66] for a standalone
class field theory proof for the case when 𝐹 is a number field.
This local-global principle for isotropy of quadratic forms is also called the Hasse
principle. For a historical overview of the Hasse principle, and more generally Hasse’s
contributions in the arithmetic theory of algebras, see Fenster–Schwärmer [FS2007].
Remark 14.6.10. The fact that quaternion algebras are classified by their ramification
set (Main Theorem 14.6.1) over a global field 𝐹 is a consequence of the following
theorem from class field theory: there is an exact sequence
Ê
0 → Br(𝐹) → Br(𝐹𝑣 ) → Q/Z → 0
𝑣 (14.6.11)
Í
( [ 𝐴 𝑣 ]) 𝑣 ↦→ 𝑣 inv 𝑣 [ 𝐴 𝑣 ]
where the first map is the natural diagonal inclusion [ 𝐴] ↦→ ( [ 𝐴 ⊗ 𝑣 𝐹𝑣 ]) 𝑣 and the
second map is the sum of the local invariant maps inv 𝑣 : Br(𝑘 𝑣 ) → Q/Z from Remark
13.4.3. The class of a quaternion algebra 𝐵 in a Brauer group over a field is 2-torsion
by 8.3.4, and the local invariant inv 𝑣 𝐵 𝑣 is equal to 0, 1/2 according as 𝐵 𝑣 is split or
ramified, and in this way we recover the main classification theorem. (In this sense, the
discriminant of a quaternion algebra captures the Brauer class of a quaternion algebra
at the finite places, and the ramification set captures it fully.) The exact sequence
(14.6.11) is sometimes called the fundamental exact sequence of global class field
theory: see Milne [Milne-CFT, §VIII.4] or Neukirch–Schmidt–Wingberg [NSW2008,
Theorem 8.1.17].
14.7.2. Let Ω ⊆ Ram 𝐵 be the set of ramified (necessarily real) archimedean places
in 𝐵. (If 𝐹 is a function field, then Ω = ∅.) Let
be the group of elements that are positive for the embeddings 𝑣 ∈ Ω. For Ω the set of
× and call such elements totally positive.
all real places, we write simply 𝐹>0
Proof. We suppose that char 𝐹𝑣 ≠ 2 and leave the other case as an exercise (Exercise
14.22). If 𝑛 𝑣 ∉ 𝐹𝑣×2 , then we can take 𝑡 𝑣 = 0; √this includes the case where 𝑣 is a
real place. So suppose 𝑛 𝑣 ∈ 𝐹𝑣×2 . Let 𝐾 𝑤 = 𝐹𝑣 ( 𝑑 𝑣 ) ⊇ 𝐹𝑣 be a separable quadratic
extension, and in particular 𝑑 𝑣 ∉ 𝐹𝑣×2 . The quadratic form h1, −4𝑛 𝑣 i ' h1, −1i over
𝐹𝑣 is a hyperbolic plane (Definition 5.4.1) so universal; let 𝑥 𝑣 , 𝑦 𝑣 ∈ 𝐹𝑣 be such that
𝑥 2𝑣 − 4𝑛 𝑣 𝑦 2𝑣 = 𝑑 𝑣 . We cannot have 𝑦 𝑣 = 0, else 𝑥 2𝑣 = 𝑑 𝑣 ∈ 𝐹𝑣×2 . Let 𝑡 𝑣 = 𝑥 𝑣 /𝑦 𝑣 . Then
𝑥 2 − 𝑡 𝑣 𝑥 + 𝑛 𝑣 has discriminant 𝑑 𝑣 and so is separable and irreducible.
For the second statement, multiplying by a square we may suppose without loss of
generality that 𝑑 𝑣 ∈ 𝑅, so the equality 𝑡 2𝑣 = 𝑑 𝑣 + 4𝑛 𝑣 implies 𝑣(𝑡 𝑣 ) ≥ 0 and 𝑡 𝑣 ∈ 𝑅.
Proof. If 𝐹𝑣 is nonarchimedean, the lemma follows from Corollary 13.2.9 and the
fact that 𝐹 is dense in 𝐹𝑣 . The case where 𝐹𝑣 is archimedean is straightforward: see
Exercise 14.23.
The same argument can be applied to several local fields at once, as follows.
14.7. THEOREMS ON NORMS 227
Proof. We repeat
Î the argument of Lemma 14.7.6, using weak approximation (i.e., 𝐹
is dense in 𝑣 𝐹𝑣 ; look ahead to Lemma 28.7.1 and the adjacent discussion) for all
𝑣 ∈ Σ to find 𝑡, 𝑛.
Proof of Main Theorem 14.7.4. Let 𝑛 ∈ 𝐹>×Ω 0 . We will construct a separable quadratic
extension 𝐾 ⊇ 𝐹 with 𝐾 ↩→ 𝐵 such that 𝑛 ∈ Nm𝐾 |𝐹 (𝐾 × ). To this end, by Proposition
14.6.7, it is enough to find 𝐾 ⊇ 𝐹 such that 𝐾 𝑣 is a field for all 𝑣 ∈ Ram 𝐵.
By Lemma 14.7.5, for all 𝑣 ∈ Ram 𝐵, there exists 𝑡 𝑣 ∈ 𝐹𝑣 such that the polynomial
𝑥 2 − 𝑡 𝑣 𝑥 + 𝑛 ∈ 𝐹𝑣 [𝑥] is separable and irreducible over 𝐹𝑣 ; here if 𝑣 ∈ Ω is real we
use that 𝑣(𝑛) > 0. By Corollary 14.7.8, there exists 𝑡 ∈ 𝐹 such that 𝑥 2 − 𝑡𝑥 + 𝑛
irreducible over each 𝐹𝑣 . Let 𝐾 be the extension of 𝐹 obtained by adjoining a root
of this polynomial. Then 𝐾 𝑣 is a field for each ramified 𝑣, and 𝑛 ∈ Nm𝐾 |𝐹 (𝐾 × ) as
desired.
Exercises
I 1. Complete the proof of Hilbert reciprocity (Proposition 14.2.1) in the remaining
cases (𝑎, 𝑏) = (−1, 2), (2, 2), (−1, 𝑝), (2, 𝑝). In particular, show that
−1, 2 2, 2
' ' M2 (Q)
Q Q
and
𝑎
(𝑎, 𝑝)2 = (𝑎, 𝑝) 𝑝 =
𝑝
for 𝑎 = −1, 2 and all primes 𝑝 (cf. 12.4.13).
2. Derive the law of quadratic reciprocity (14.2.4) and the supplement (14.2.5)
from the statement of Hilbert reciprocity (Proposition 14.2.1).
I 3. Let 𝑛 ∈ Z>0 .
(a) Suppose 𝑛 is of the form 𝑛 = 4𝑎 (8𝑏 + 7) with 𝑎, 𝑏 ∈ Z. Show that there is
no solution to 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛 with 𝑥, 𝑦, 𝑧 ∈ Z. [Hint: Look modulo 8.]
(b) Suppose 𝑛 is not of the form 𝑛 = 4𝑎 (8𝑏 + 7) with 𝑎, 𝑏 ∈ Z. Show that there
is a solution to 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛 with 𝑥, 𝑦, 𝑧 ∈ Z2 . [Hint: lift a solution
modulo 8 using Hensel’s lemma.]
I 4. Let 𝑛 ∈ Z be nonzero. Show that 𝑛 is a square in Q2 if and only if 𝑛 is of the
form 𝑛 = 4𝑎 (8𝑏 + 1) with 𝑎, 𝑏 ∈ Z.
228 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
√
5. Let 𝑛 > 0 have −𝑛 ∉ Q×2
2 . Let 𝐵 = (−1, −1 | Q) and let 𝐾 = Q( −𝑛). Show that
𝐾 splits 𝐵. [Hint: Use the local-global principle for embeddings (Proposition
14.6.7).] Conclude that there exists 𝛼 ∈ 𝐵 such that 𝛼2 = −𝑛, and conclude as
in Theorem 14.3.8 that 𝑛 is the sum of three squares.
6. Let 𝐹 be a number field. Show that every totally positive element of 𝐹 is a sum
of four squares of elements of 𝐹.
7. Show that the law of Hilbert reciprocity (Proposition 14.2.1) implies the law of
quadratic reciprocity; with the argument given in section 14.1, this completes
the equivalence of these two laws.
8. In the proof of Legendre’s theorem (Theorem 14.3.4), we reduced to the case
𝑎, 𝑏 > 0 and 𝑐 = −1. Show that this reduction is valid.
9. In this exercise, we generalize the proof of Proposition 14.2.7 to give a more
general construction of quaternion quaternion algebras. Let 𝐷 be a squarefree
positive integer and let 𝑢 = −1 if 𝐷 has an odd number of prime divisors,
otherwise 𝑢 := 1.
√
(a) For 𝑏 ∈ Z squarefree, show that 𝐾 := Q( 𝑏) embeds in a quaternion
algebra of discriminant 𝐷 if and only if:
• 𝑏 < 0 if 𝐵 is definite;
• 𝑏 . 1 (mod 8) if 2 | 𝐷; and
𝑏
• for all odd primes 𝑝 | 𝐷, we have ≠ 1.
𝑝
(b) Suppose 𝑏 satisfies the conditions in (a) but with the
further requirement
𝑏
that 𝐷 | 𝑏, i.e., in the third condition we require = 0. Let 𝑞 be an odd
𝑝
prime such that 𝑞 ♦ := 𝑢𝑞 has:
♦
𝑞
• = −1 for all odd 𝑝 | 𝐷;
𝑝♦
𝑞
• = 1 for all odd 𝑝 | (𝑏/𝐷); and
𝑝
• 𝑞 ♦ ≡ 1, 5 (mod 8) according as 2 - 𝐷 or 2 | 𝐷.
Note there exist infinitely many such primes 𝑞 by the infinitude of primes
𝑞♦, 𝑏
in arithmetic progression. Then show that 𝐵 := has disc 𝐵 = 𝐷
Q
and Ram 𝐵 = Σ.
I 10. Let S ⊆ Pl(Q) be eligible. For each 𝑣 ∈ S, let 𝑡 𝑣 ∈ Q×𝑣 be given. Show that there
exists 𝑡 ∈ Q× such that 𝑡 ∈ 𝑡 𝑣 Q×2
𝑣 for all 𝑣 ∈ S and 𝑣 𝑝 (𝑡) = 0 for all 𝑝 ∉ S r {∞}
except (possibly) for one prime 𝑝 = 𝑞.
√
11. Let 𝐹 = Q( 𝑑) be a real quadratic field. Find 𝑎, 𝑏 ∈ Q× (depending on 𝑑) such
that (𝑎, 𝑏 | 𝐹) is a division ring unramified at all finite places.
12. Let 𝐹 be a global field with char 𝐹 ≠ 2 and let 𝐵 be a quaternion algebra over
𝐹. Let 𝐿 ⊇ 𝐹 be a finite extension. An extension 𝐾 ⊇ 𝐹 is linearly disjoint
with 𝐿 over 𝐹 if the multiplication map 𝐾 ⊗𝐹 𝐿 −→ ∼ 𝐾 𝐿 is an isomorphism of
14.7. THEOREMS ON NORMS 229
𝐹-algebras.
Show that there exists a splitting field 𝐾 ⊇ 𝐹 for 𝐵 such that 𝐾 is linearly disjoint
with 𝐿 over 𝐹.
13. Show that the notion of discriminant of a quaternion algebra as the product of
ramified primes is not such a great notion when 𝑅 is an arbitrary Dedekind
domain, as follows.
Let 𝑅 = Q[𝑡]; then 𝑅 is a Dedekind domain. Let 𝐹 = Frac 𝑅 = Q(𝑡). Let
𝐵0 = (𝑎, 𝑏 | Q) be a division quaternion algebra over Q and let 𝐵 = 𝐵0 ⊗Q 𝐹 =
(𝑎, 𝑏 | Q(𝑡)). Show that there are infinitely primes at which 𝐵 is “ramified”:
for every prime 𝔭 = (𝑡 − 𝑐)𝑅, show that the algebra 𝐵𝔭 is a division quaternion
algebra over 𝐹𝔭 ' Q((𝑡)). [Hint: See Exercise 13.7.]
14. Using Hilbert reciprocity, one can convert the computation of an even Hilbert
symbol to the computation of several odd Hilbert symbols, as follows.
Let 𝐹 be a number field, let 𝔭 | (2), and let 𝑎, 𝑏 ∈ 𝐹 × . Show that there exist
(computable) 𝑎 0, 𝑏 0 ∈ 𝐹 × such that the following hold:
(i) (𝑎, 𝑏)𝔭 = (𝑎 0, 𝑏 0)𝔭 ; and
(ii) ord𝔮 (𝑎 0) = ord𝔮 (𝑏 0) = 0 for all 𝔮 | (2) with 𝔮 ≠ 𝔭.
Conclude that
Ö
(𝑎, 𝑏)𝔭 = (𝑎 0, 𝑏 0) 𝑣 .
𝑣 ∈Pl 𝐹
𝑣 odd
15. Let 𝐹 be a number field with ring of integers Z𝐹 . We say an ideal 𝔟 ⊆ Z𝐹 is odd
is odd, and 𝑏 ∈ Z𝐹 is odd if (𝑏) is odd. For 𝑎 ∈ Z𝐹 and 𝔟 ⊆ Z𝐹 odd,
if Nm(𝔟)
𝑎
let be the generalized Jacobi symbol, extending the generalized Legendre
𝔟
𝑎 𝑎
symbol by multiplicativity, and write := for 𝑎, 𝑏 ∈ Z𝐹 r {0} with
𝑏 𝑏Z𝐹
𝑏 odd.
16. In this exercise, we give a constructive proof of the surjectivity of the map
𝐵 ↦→ Ram 𝐵 in Main Theorem 14.6.1 in the spirit of the proof of Proposition
14.2.7 (assuming two analytic results).
Let 𝐹 be a number field, and let Σ ⊆ Pl 𝐹 be a finite set of noncomplex places
of 𝐹 of even cardinality. Let 𝑅 = Z𝐹 be the ring of integers of 𝐹.
230 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
Î
(a) Let 𝔇 := 𝔭∈Σ 𝔭 be the product of the primes corresponding to nonar-
chimedean places in S. Using weak approximation (see Lemma 28.7.1),
show there exists 𝑎 ∈ 𝔇 such that:
• 𝑣(𝑎) < 0 for all real places 𝑣 ∈ Σ and 𝑣(𝑎) > 0 for all real places
𝑣 ∉ Σ, if there are any; and
• 𝑎𝑅 = 𝔇𝔟 with 𝔇 + 𝔟 = 𝑅 and 2𝑅 + 𝔟 = 𝑅.
In the special case where 𝑅 has narrow class number 1 (that is, every ideal
𝔞 ⊆ 𝑅 is principal 𝔞 = (𝑎) and generated by an element 𝑎 ∈ 𝑅 such that
𝑣(𝑎) > 0 for every real places 𝑣), show that we may take (𝑎) = 𝔇 and
𝔟 = 𝑅.
(b) Show that there exists 𝑡 ∈ 𝑅 coprime to 8𝑎𝑅 such that the following hold:
𝑡
• For all primes 𝔭 | 𝔇 with 𝔭 - 2𝑅, we have = −1;
𝔭 √
• For all primes 𝔭 | 𝔇 with 𝔭 | 2𝑅, the extension 𝐹𝔭 ( 𝑡) is the quadratic
𝑡
unramified extension of 𝐹𝔭 , so = −1 in the sense of the general-
𝔭
ized Kronecker symbol;
𝑡
• For all primes 𝔮 | 𝔟 we have = 1; and
𝔮
• For all prime powers 𝔯 k 8𝑅 with 𝔯 - 𝐷, we have 𝑡 ≡ 1 (mod 𝔯𝑒 ).
𝑒
Show that there exists a quaternion algebra 𝐵 over 𝐹 with the property that
Ram(𝐵 𝐾 ) = Σ𝐾 . (We say that the quaternion algebra associated to the set
Σ𝐾 descends to 𝐹.)
(d) As a special case, what do (a) and (c) say when [𝐾 : 𝐹] = 2?
(e) Restate (a) and (b) in terms of the kernel of the map Br(𝐹) [2] → Br(𝐾) [2]
induced by [𝐵] ↦→ [𝐵 𝐾 ] (see Remark 14.6.10).
19. Let 𝑅 be a global ring with 𝐹 = Frac 𝑅, and let 𝐾 ⊇ 𝐹 be a finite Galois
extension with 𝑆 the integral closure of 𝑅 in 𝐾. Let 𝐵 be a quaternion algebra
over 𝐹 and consider 𝐵 𝐾 = 𝐵 ⊗𝐹 𝐾. Then Gal(𝐾 | 𝐹) acts naturally on 𝐵 𝐾 via
𝜎(𝛼 ⊗ 𝑥) = 𝛼 ⊗ 𝜎(𝑥). (This action is not by 𝐾-algebra isomorphism!)
Show that there exists a maximal 𝑆-order O ⊆ 𝐵 𝐾 stable under Gal(𝐾 | 𝐹), i.e.,
𝜎(O) = O for all 𝜎 ∈ Gal(𝐾 | 𝐹).
I 20. Let 𝐹 be a global field, let 𝑣 1 , . . . , 𝑣 𝑟 be places of 𝐹, and for each 𝑣 𝑖 suppose
we are given the condition ramified, split, or inert. Show that there exists a
separable quadratic extension 𝐾 ⊇ 𝐹 that 𝐾 𝑣𝑖 satisfies the given condition for
each 𝑖. [Hint: follow the proof of Main Theorem 14.7.4.]
21. Let 𝐹 be a global field, let 𝐵1 , 𝐵2 , . . . , 𝐵𝑟 be quaternion algebras over 𝐹, and let
𝐵 := 𝐵1 ⊗ 𝐵2 ⊗ · · · ⊗ 𝐵𝑟 . Recalling section 8.2, show (in as many ways as you
can) that 𝐵 ' M2𝑟−1 (𝐵 0) for a quaternion algebra 𝐵 0 over 𝐹. (Recalling 8.3, by
Merkurjev’s theorem this shows the class of every element in the 2-torsion of
the Brauer group Br(𝐹) [2] is represented by a quaternion algebra.)
I 22. Let 𝐹𝑣 be a local field with char 𝐹𝑣 = 2. Let 𝑛 ∈ 𝐹𝑣 . Show that there exists
𝑡 ∈ 𝐹𝑣 such that 𝑥 2 − 𝑡𝑥 + 𝑛 is separable and irreducible.
I 23. Prove Lemma 14.7.6 for 𝑣 an archimedean place.
24. In this advanced exercise following up on Exercise 9.15, we consider features of
quaternion algebras and orders in the case of a global function field, assuming
background in algebraic geometry.
Let 𝑋 be a smooth, projective, geometrically integral curve over a finite field 𝑘;
then 𝑋 is a separated, integral Dedekind scheme. Let 𝒪𝑋 be its structure sheaf.
Let 𝐹 be its function field, and let 𝐵 be a quaternion algebra over 𝐹. Define a
sheaf of 𝒪𝑋 -orders in 𝐵, or simply an 𝒪𝑋 -order in 𝐵, to be an 𝒪𝑋 -lattice ℬ in
𝐵 such that for each open set 𝑈 ⊆ 𝑋, the 𝒪𝑋 (𝑈)-lattice ℬ(𝑈) is a subring of 𝐵.
We recall the local-global dictionary for 𝒪𝑋 -lattices (Exercise 9.15(c)).
In parts (a)–(c) we work out an example: let 𝑋 = P1 with function field 𝐹 = 𝑘 (𝑡)
where div 𝑡 = (0) − (∞). Suppose that char 𝑘 ≠ 2, and let 𝑢 ∈ 𝑘 × r 𝑘 ×2 . Let 𝐵
be the quaternion algebra with Ram(𝐵) = {(0), (∞)}.
(a) Show that 𝐵 = (𝑢, 𝑡 | 𝐹).
232 CHAPTER 14. QUATERNION ALGEBRAS OVER GLOBAL FIELDS
(b) Let 𝑈 = Spec 𝑘 [𝑡] = 𝑋 r {∞}. Show that there exists a unique 𝒪𝑋 -order ℬ
in 𝐵 with ℬ(𝑈) = 𝑘 [𝑡] + 𝑘 [𝑡]𝑖 + 𝑘 [𝑡] 𝑗 + 𝑘 [𝑡]𝑖 𝑗 and stalk ℬ(∞) a maximal
𝒪𝑋 , (∞) -order. Describe explicitly ℬ(∞) and ℬ(Spec 𝑘 [1/𝑡]) as orders in
𝐵.
(c) With ℬ from (b), show that ℬ(𝑋) = 𝑘 [𝑖].
Restoring generality, let ℬ be an 𝒪𝑋 -order such that ℬ(𝑈) is a maximal 𝒪𝑋 (𝑈)-
order in 𝐵 for all affine open sets 𝑈.
(d) Show that ℬ(𝑋) has a zero divisor if and only if ℬ(𝑋) ' M2 (𝑘) if and
only if 𝐵 ' M2 (𝐹).
(e) Show that ℬ(𝑋) is a 𝑘-algebra with a nondegenerate standard involution.
(f) Suppose that 𝐵 is a division algebra. Show that either ℬ(𝑋) = 𝑘 or
ℬ(𝑋) ' 𝑘 2 is the quadratic extension of 𝑘.
(g) Still supposing that 𝐵 is a division algebra, show that if ℬ(𝑋) = 𝑘 2 , then
every ramified place of 𝐵 has odd degree. [Hint: show that 𝐵 ' (𝐾, 𝑏 | 𝐹)
where 𝐾 = 𝐹 𝑘 2 is the constant field extension of 𝐹 of degree 2, and
𝑏 ∈ 𝐹 × r 𝑘 × . Compute the Hilbert symbol at 𝑣 ∈ Ram(𝐵) to show 𝑣(𝑏) is
odd.]
Chapter 15
Discriminants
(normalized with an extra factor of 2 at the complex places), then the volume of Z𝐹
in this embedding is the absolute determinant of the matrix with columns 𝜄(𝑥𝑖 ), and
its square is defined to be the absolute discriminant of 𝐹. Replacing the dot product
in the definition of 𝑀 in (15.1.1) with the trace form (𝑥, 𝑦) ↦→ Tr𝐹 /Q (𝑥𝑦), we see that
the absolute discriminant is a positive integer. A prime 𝑝 is ramified in 𝐹 if and only
if it divides the discriminant, so this volume also records arithmetic properties of 𝐹.
More generally, whenever we have a symmetric bilinear form 𝑇 : 𝑉 × 𝑉 → 𝐹 on
a finite-dimensional 𝐹-vector space 𝑉, there is a volume defined by the determinant
det(𝑇 (𝑥𝑖 , 𝑥 𝑗 ))𝑖, 𝑗 : and when 𝑇 arises from a quadratic form 𝑄, this is volume is the
discriminant of 𝑄 (up to a normalizing factor of 2 in odd degree, see 6.3.1). In
particular, if 𝐵 is a finite-dimensional algebra over 𝐹, there is a bilinear form
𝐵×𝐵→𝐹
(𝛼, 𝛽) ↦→ Tr 𝐵 |𝐹 (𝛼𝛽)
(or, when 𝐵 is semisimple, the bilinear form associated to the reduced trace trd)
and so we obtain a discriminant—a “squared” volume—measuring in some way the
complexity of 𝐵. As in the commutative case, discriminants encode ramification.
In this chapter, we establish basic facts about discriminants, including how they
behave under inclusion (measuring index) and localization. To illustrate, let 𝐵 be a
233
234 CHAPTER 15. DISCRIMINANTS
disc(O) = (4𝑎𝑏) 2 ;
indeed, this is the discriminant of the quadratic form h1, −𝑎, −𝑏, 𝑎𝑏i, the reduced norm
restricted to O. If 𝑎, 𝑏 < 0, i.e. 𝐵 is definite, then the reduced√ norm is a Euclidean
norm on 𝐵∞ = 𝐵 ⊗Q R ' H; normalizing with an extra factor 2, the discriminant is
square of the covolume of the lattice O ⊂ 𝐵∞ . For example, the Lipschitz√order Zh𝑖, 𝑗i
(11.1.1) has disc(Zh𝑖, 𝑗i) = 42 , the square of the covolume of the lattice ( 2Z) 4 ⊆ R4 .
If O0 ⊇ O, then disc(O) = [O0 : O] 2 disc(O0); in particular O0 = O if and only if
disc(O0) = disc(O). It follows that the discriminant of an order is always a square, so
we define the reduced discriminant discrd(O) to be the positive integer square root,
and discrd(O) 2 = disc(O). The discriminant of an order measures how far the order
is from being a maximal order. We will show (Theorem 15.5.5) that O is a maximal
order if and only if discrd(O) = disc 𝐵, where disc 𝐵 is the (squarefree) product of
primes ramified in 𝐵.
In an extension of Dedekind domains, the different of the extension is an ideal
whose norm is the discriminant of the extension (see Neukirch [Neu99, §III.2]).
The different is perhaps not as popular as its discriminant cousin, but it has many
nice properties, including easy-to-understand behavior under base extension. Similar
conclusions holds in the noncommutative context (presented in section 15.6).
15.2 Discriminant
For further reference on discriminants, see Reiner [Rei2003, §10, §14].
Let 𝑅 be a noetherian domain and let 𝐹 = Frac 𝑅. Let 𝐵 be a semisimple algebra
over 𝐹 with dim𝐹 𝐵 = 𝑛. For elements 𝛼1 , . . . , 𝛼𝑛 ∈ 𝐵, we define
Let 𝐼 ⊆ 𝐵 be an 𝑅-lattice.
Remark 15.2.4. When working over Z, it is common to take the discriminant instead to
be the positive generator of the discriminant as an ideal; passing between these should
cause no confusion.
Although Definition 15.2.2 may look unwieldly, it works as well in the commutative
case as in the noncommutative case. Right away, we see that if O ⊆ O0 are 𝑅-orders,
then disc(O0) | disc(O).
The function 𝑑 itself transforms in a nice way under a change of basis, as follows.
Proof. The matrix 𝑀 writing any other 𝛽1 , . . . , 𝛽𝑛 ∈ 𝐼 in terms of the basis has
𝑀 ∈ M𝑛 (𝑅) so det(𝑀) ∈ 𝑅, and therefore 𝑑 (𝛽1 , . . . , 𝛽𝑛 ) ∈ 𝑑 (𝛼1 , . . . , 𝛼𝑛 )𝑅 by
Lemma 15.2.5.
15.2.8. More generally, if 𝐼 is completely decomposable with
𝐼 = 𝔞1 𝛼1 ⊕ · · · ⊕ 𝔞𝑛 𝛼𝑛
2 0 0 0
0 2𝑎 0 0 ®
© ª
𝑑 (1, 𝑖, 𝑗, 𝑖 𝑗) = det ® = −(4𝑎𝑏) 2 .
0 0 2𝑏 0 ®
«0 0 0 −2𝑎𝑏 ¬
15.2.13. Equation (15.2.6) and the fact that 𝐼 (𝔭) = 𝐼 ⊗𝑅 𝑅 (𝔭) implies the equality
on localizations and for the same reason an equality for the completions disc(𝐼𝔭 ) =
disc(𝐼)𝔭 . In other words, the discriminant respects localization and completion and
can be computed locally. Therefore, by the local-global principle (Lemma 9.4.6),
Ù
disc(𝐼) = disc(𝐼 (𝔭) ).
𝔭
disc(𝐼) = [𝐽 : 𝐼] 2𝑅 disc(𝐽).
Proof. For the first statement, we argue locally, and combine (15.2.6) and Lemma 9.6.4.
For the second statement, clearly disc(𝐽) ⊆ disc(𝐼), and if 𝐼 = 𝐽 then equality holds;
and conversely, from disc(𝐼) = [𝐽 : 𝐼] 2𝑅 disc(𝐽) = disc(𝐽) we conclude [𝐽 : 𝐼] 𝑅 = 𝑅,
hence 𝐽 = 𝐼 by Proposition 9.6.8.
Remark 15.2.16. We defined the discriminant for semisimple algebras so that it is given
in terms of the reduced trace. This definition extends to an arbitrary finite-dimensional
𝐹-algebra 𝐵, replacing the reduced trace by the algebra trace Tr 𝐵 |𝐹 . If 𝐵 is a central
simple 𝐹-algebra of dimension 𝑛2 , then 𝑛 trd = Tr 𝐵 |𝐹 so when 𝑛 ∈ 𝐹 × one can recover
the discriminant as we have defined it here from the more general definition; but if
𝑛 = 0 ∈ 𝐹 then the discriminant of 𝐵 computed with the algebra trace will be zero.
{𝑑 (𝑥1 , . . . , 𝑥 𝑛 ; 𝑓 ) : 𝑥1 , . . . , 𝑥 𝑛 ∈ 𝑀, 𝑓 ∈ 𝐿 ∨ }. (15.3.4)
In particular, since 𝑀, 𝐿 are projective and therefore locally free over 𝑅, the discrimi-
nant of 𝑄 is locally free and hence a projective 𝑅-ideal.
238 CHAPTER 15. DISCRIMINANTS
The notions in this section extend more generally to an arbitrary algebra 𝐵 with a
standard involution.
Lemma 15.5.1. Let O ⊆ O0 be 𝑅-orders. Then O = O0 if and only if disc O = disc O0.
First, we ensure the existence of maximal orders (cf. 10.4.2) using the discriminant.
Proof. The algebra 𝐵 has at least one 𝑅-order O as the left- or right-order of a lattice
10.2.5. If O is not maximal, then there exists an order O0 ) O with disc(O0) ) disc(O)
by Lemma 15.5.1. If O0 is maximal, we are done; otherwise, we can continue in
this way to obtain orders O = O1 ( O2 ( . . . and an ascending chain of ideals
disc(O1 ) ( disc(O2 ) ( . . . of 𝑅; but since 𝑅 is noetherian, the latter stabilizes after
finitely many steps, and the resulting order is then maximal, by Lemma 15.2.15.
Using the discriminant as a measure of index, we can similarly detect when orders
are maximal. We recall (10.4.3) that the property of being maximal is a local property,
so we begin with the local matrix case.
Theorem 15.5.5. Let 𝑅 be a global ring with field of fractions 𝐹, let 𝐵 be a quaternion
algebra over 𝐹, and let O ⊆ 𝐵 be an 𝑅-order. Then O is maximal if and only if
discrd(O) = disc𝑅 (𝐵). (15.5.6)
Proof. Suppose that O is maximal. Then O𝔭 is maximal for all primes 𝔭 of 𝑅. If
𝐵𝔭 ' M2 (𝐹𝔭 ) is split, then by Lemma 15.5.3, discrd O𝔭 = 𝑅𝔭 ; if 𝐵𝔭 is a division
algebra, then discrd O𝔭 = 𝔭𝑅𝔭 . Since discriminants are defined locally, we conclude
that Ö
discrd(O) = 𝔭 = disc𝑅 (𝐵)
𝔭∈Ram 𝐵rS
if 𝑅 as a global ring is the ring of S-integers.
In the other direction, if (15.5.6) holds, we choose O0 ⊇ O be a maximal 𝑅-
superorder and conclude that disc(O) = disc𝑅 (𝐵) 2 = disc(O0) so O = O0 is maximal
by Lemma 15.5.1.
Example 15.5.7. We recall Example 14.2.13, giving an explicit description of quater-
nion algebras 𝐵 = (𝑎, 𝑏 | Q) of prime discriminant 𝐷 = 𝑝. We now exhibit an explicit
maximal order in each of these algebras.
For 𝑝 = 2, we have 𝐵 = (−1, −1 | Q) and take O ⊆ 𝐵 the Hurwitz order.
For 𝑝 ≡ 3 (mod 4), we took 𝐵 = (−𝑝, −1 | Q). The order O := Zh(1 + 𝑖)/2, 𝑗i =
𝑆 ⊕ 𝑆 𝑗 with 𝑆 := Z[(1 + 𝑖)/2] has discrd O = 𝑝 by 15.2.12, so O is maximal by
Theorem 15.5.5.
For 𝑝 ≡ 1 (mod 4), we had 𝐵 = (−𝑝, −𝑞 | Q) where 𝑞 ≡ 3 (mod 4) is prime and
𝑞
= −1, so that by qudaratic reciprocity −𝑝𝑞 = −𝑞 𝑝 = 1. In this case, let 𝑐 ∈ Z be
𝑝
such that 𝑐2 ≡ −𝑝 (mod 𝑞). Then
1+ 𝑗 𝑖(1 + 𝑗) (𝑐 + 𝑖) 𝑗
O := Z ⊕ Z ⊕Z ⊕Z
2 2 𝑞
is a maximal order: one checks that O is closed under multiplication (in particular, the
basis elements are integral), and then that disc O = 𝑝. The order Zh𝑖, (1 + 𝑗)/2i ⊆ O
has the larger reduced discriminant 𝑝𝑞, hence the need for a denominator 𝑞 in the
fourth element.
For further discussion of explicit maximal orders over Z, see Ibukiyama [Ibu82,
pp. 181–182] or Pizer [Piz80a, Proposition 5.2]. For a more general construction, see
Exercise 15.5.
15.6 Duality
To round out the chapter, we relate the discriminant and trace pairings to the dual and
the different. For a detailed, general investigation of the dual in the context of other
results for orders, see Faddeev [Fad65].
We continue with the hypothesis that 𝑅 is a domain with 𝐹 = Frac 𝑅. Let 𝐵 be an
𝐹-algebra with 𝑛 := dim𝐹 𝐵 < ∞. As the trace pairing will play a significant role in
what follows, we suppose throughout that 𝐵 is separable (in particular, semisimple)
as an 𝐹-algebra with reduced trace trd. Let 𝐼, 𝐽 be 𝑅-lattices in 𝐵.
242 CHAPTER 15. DISCRIMINANTS
Lemma 15.6.2.
(a) If 𝐼 ⊆ 𝐽 then 𝐼 ♯ ⊇ 𝐽 ♯ .
(b) For all 𝛽 ∈ 𝐵× , we have (𝛽𝐼) ♯ = 𝐼 ♯ 𝛽−1 .
(c) If 𝔭 ⊆ 𝑅 is prime, then (𝐼 (𝔭) ) ♯ = (𝐼 ♯ ) (𝔭) and the same with the completion.
Proof. For parts (a) and (b), see Exercise 15.15. The proof of part (c) is similarly
straightforward.
15.6.3. Suppose that 𝐼 is free over 𝑅 with basis 𝛼1 , . . . , 𝛼𝑛 . Since the trace pairing on
𝐵 is nondegenerate (Theorem 7.9.4), there exists a dual basis 𝛼𝑖♯ ∈ 𝐵 to 𝛼𝑖 under the
reduced trace trd, so that trd(𝛼𝑖♯ 𝛼 𝑗 ) = 0, 1 according as 𝑖 ≠ 𝑗 or 𝑖 = 𝑗.
Then 𝐼 ♯ is free over 𝑅 with basis 𝛼1♯ , . . . , 𝛼𝑛♯ : if 𝛽 = 𝑏 1 𝛼1♯ + · · · + 𝑏 𝑛 𝛼𝑛♯ with
𝑏 1 , . . . , 𝑏 𝑛 ∈ 𝐹, then 𝛽 ∈ 𝐼 ♯ if and only if trd(𝛼𝑖 𝛽) = 𝑏 𝑖 ∈ 𝑅 for all 𝑖.
Remark 15.6.10. The content of Proposition 15.6.7 is that although one can always
construct the module dual, the trace pairing concretely realizes this module dual as a
lattice. (And we speak of bimodules in the proposition because Hom𝑅 (𝐼, 𝔞) does not
come equipped with the structure of 𝑅-lattice in 𝐵.) This module duality, and the fact
that 𝐼 is projective over 𝑅, can be used to give another proof of Lemma 15.6.5.
The dual asks for elements that pair integrally under the trace. We might also ask
for elements that multiply one lattice into another, as follows.
Definition 15.6.11. Let 𝐼, 𝐽 be 𝑅-lattices. The left colon lattice of 𝐼 with respect to
𝐽 is the set
(𝐼 : 𝐽) L := {𝛼 ∈ 𝐵 : 𝛼𝐽 ⊆ 𝐼}
and similarly the right colon lattice is
(𝐼 : 𝐽) R := {𝛼 ∈ 𝐵 : 𝐽𝛼 ⊆ 𝐼}.
Note that (𝐼 : 𝐼) L = OL (𝐼) is the left order of 𝐼 (and similarly on the right). The
same proof as in Lemma 10.2.7 shows that (𝐼 : 𝐽) L and (𝐼 : 𝐽) R are 𝑅-lattices.
(𝐼𝐽) ♯ = (𝐼 ♯ : 𝐽) R = (𝐽 ♯ : 𝐼) L .
244 CHAPTER 15. DISCRIMINANTS
OL (𝐼) = (𝐼 : 𝐼) L = ((𝐼 ♯ ) ♯ : 𝐼) L = (𝐼 𝐼 ♯ ) ♯
codiff(O) := O♯ .
The major role played by the codifferent is its relationship to the discriminant, as
follows.
Proof. For a prime 𝔭 ⊆ 𝑅 we have disc(O) (𝔭) = disc(O (𝔭) ) and [O♯(𝔭) : O (𝔭) ] 𝑅 (𝔭) =
( [O♯ : O] 𝑅 ) (𝔭) , and so to establish the equality we may argue locally. Since O (𝔭) is
Í
free over 𝑅 (𝔭) , we reduce to the case where O is free over 𝑅, say O = 𝑖 𝑅𝛼𝑖 . Then
O♯ = 𝑖 𝑅𝛼𝑖♯ with 𝛼1♯ , . . . , 𝛼𝑛♯ ∈ 𝐵 the dual basis, as in 15.6.3.
Í
The ideal disc(O) is principal, generated by 𝑑 (𝛼1 , . . . , 𝛼𝑛 ) = det(trd(𝛼𝑖 𝛼 𝑗 ))𝑖, 𝑗 ; at
the same time, the 𝑅-index [O♯ : O] 𝑅 is generated by det(𝛿) where 𝛿 is the change of
basis from 𝛼𝑖♯ to 𝛼𝑖 . But 𝛿 is precisely the matrix (trd(𝛼𝑖 𝛼 𝑗 ))𝑖, 𝑗 (Exercise 15.14), and
the result follows.
Exercises
Unless otherwise specified, let 𝑅 be a noetherian domain with field of fractions 𝐹.
𝑎, 𝑏
1. Let char 𝐹 = 2 and let be a quaternion algebra over 𝐹 with 𝑎, 𝑏 ∈ 𝑅 and
𝐹
𝑏 ≠ 0. Show that O = 𝑅 + 𝑅𝑖 + 𝑅 𝑗 + 𝑅𝑖 𝑗 is an 𝑅-order in 𝐵 and compute the
(reduced) discriminant of O.
2. Let 𝐵 = M𝑛 (𝐹) and O = M𝑛 (𝑅) with 𝑛 ≥ 1. Show that disc(O) = 𝑅. [Hint:
Compute directly on a basis {𝑒 𝑖 𝑗 }𝑖, 𝑗 of matrix units, which satisfy 𝑒 𝑖 𝑗 𝑒 𝑖0 𝑗 0 = 𝑒 𝑖 𝑗 0
if 𝑗 = 𝑖 0, otherwise zero.]
3. Suppose 𝑅 is a global ring, so 𝐹 is a global field; let 𝐵 be a quaternion algebra
over 𝐹 and let O ⊆ 𝐵 be an 𝑅-order. Prove that for all primes 𝔭 ⊆ 𝑅, we have
O𝔭 ' M𝑛 (𝑅𝔭 ) if and only if 𝔭 - disc O.
4. Let 𝐵 := (𝐾, 𝑏 | 𝐹) be a quaternion algebra over a field 𝐹 with 𝑏 ∈ 𝐹 × . Let
𝑆 ⊆ 𝐾 be an 𝑅-order with 𝔡 := disc(𝑆); let 𝔟 ⊆ 𝐾 be a fractional 𝑆-ideal (which
can be but need not be invertible), and finally let O := 𝑆 ⊕ 𝔟 𝑗.
(a) Show that O is an 𝑅-order if and only if Nm𝐾 |𝐹 𝔟 ⊆ 𝑏 −1 𝑅.
(b) Compute that discrd O = 𝔡(Nm𝐾 |𝐹 𝔟)𝑏.
5. In this exercise, we consider a construction of maximal orders as crossed
♦ products
𝑞 ,𝑏
in the simplest case over Q, continuing Exercise 14.9. Let 𝐵 := be a
Q
quaternion algebra of discriminant 𝐷, where 𝑏 ∈ Z is squarefree with 𝐷 | 𝑏 and
♦
√︁ with 𝑞 = ±𝑞 ≡ 1 (mod 4), the minus sign if 𝐵 is indefinite.
𝑞 is an odd prime
Let 𝐾 := Q( 𝑞 ♦ ) be the quadratic field of discriminant 𝑞 ♦ . Let 𝑆 ⊆ 𝐾 be the
ring of integers of 𝐾, so disc 𝑆 = 𝑞 ♦ .
♦
𝑞
(a) Show that for all odd primes 𝑝 | (𝑏/𝐷), we have = 1. Conclude
𝑝
there exists an ideal 𝔟 ⊆ 𝑆 such that Nm 𝔟 = 𝑏/𝐷.
(b) Let 𝔮 ⊆ 𝑆 be the unique prime above 𝑞, and let
O := 𝑆 ⊕ (𝔮𝔟) −1 𝑗
[Hint: reduce to the case where trd(𝛼) = trd(𝛽) = 0, noting the invariance of
𝑠.]
10. Let 𝐵 be a quaternion algebra over 𝐹. Define 𝑚 : 𝐵 × 𝐵 × 𝐵 → 𝐹 by
𝑚(𝛼1 , 𝛼2 , 𝛼3 ) := trd( [𝛼1 , 𝛼2 ]𝛼3 ) for 𝛼𝑖 ∈ 𝐵. If 𝛽𝑖 = 𝑀𝛼𝑖 for some 𝑀 ∈ M3 (𝐹),
show that
𝑚(𝛽1 , 𝛽2 , 𝛽3 ) = det(𝑀)𝑚(𝛼1 , 𝛼2 , 𝛼3 ).
11. Let 𝐵 be a quaternion algebra over 𝐹. Give another proof that
𝑚(𝛼1 , 𝛼2 , 𝛼3 ) 2 = 𝑑 (1, 𝛼1 , 𝛼2 , 𝛼3 )
Much like a space can be understood by studying functions on that space, often the
first task to understand a ring 𝐴 is to understand the ideals of 𝐴 and modules over 𝐴
(in other words, to pursue “linear algebra” over 𝐴). The ideals of a ring that are easiest
to work with are the principal ideals—but not all ideals are principal, and various
algebraic structures are built to understand the difference between these two. In this
chapter, we consider these questions for the case where 𝐴 is a quaternion order.
249
250 CHAPTER 16. QUATERNION IDEALS AND INVERTIBILITY
with both of these products compatible. If a lattice 𝐼 has a two-sided inverse, then this
inverse is uniquely given by
𝐼 −1 := {𝛼 ∈ 𝐵 : 𝐼𝛼𝐼 ⊆ 𝐼}
(defined so as to simultaneously take care of both left and right): we always have that
𝐼 𝐼 −1 ⊆ OL (𝐼), but equality is needed for right invertibility, and the same on the left.
Let O ⊆ 𝐵 be an order. A left fractional O-ideal is a lattice 𝐼 ⊆ 𝐵 such that
O ⊆ OL (𝐼); we similarly define on the right. For a maximal order, all lattices are
invertible (Proposition 16.6.15(b)).
16.2. LOCALLY PRINCIPAL, COMPATIBLE LATTICES 251
and similarly on the left; by Main Theorem 16.1.3 (iv) ⇔ (iv0), when 𝐼 is locally
principal, the left and right absolute norms coincide (called then just absolute norm)
and are related to the reduced norm by N (𝐼) = nrd(𝐼) 2 .
𝐵 = 𝐹𝐽 = 𝐹 (𝑟 𝐽) ⊆ 𝐹 𝐼𝐽
so equality holds.
When multiplication of two lattices matches up their respective left and right
orders, we give it a name.
We will also sometimes just say that the product 𝐼𝐽 is compatible to mean that 𝐼 is
compatible with 𝐽. The relation “is compatible with” is in general neither symmetric
nor transitive. (Looking ahead to groupoids in Chapter 17, we might also say that the
two lattices are composable.)
16.2.6. 𝐼 has the structure of a right OR (𝐼)-module and 𝐽 the structure of a left OL (𝐽)-
module. When OR (𝐼) = OL (𝐽) = O, that is, when 𝐼 is compatible with 𝐽, it makes
sense to consider the tensor product 𝐼 ⊗O 𝐽 as an 𝑅-module. The multiplication map
𝐵 ⊗ 𝐵 𝐵 −→∼ 𝐵 defined by 𝛼 ⊗ 𝛽 ↦→ 𝛼𝛽 restricts to give an isomorphism 𝐼 ⊗ 𝐽 −→ ∼ 𝐼𝐽
O
as 𝑅-lattices. In this way, multiplication of compatible lattices can be thought of as a
special case of the tensor product of modules.
(i) 𝐼 is integral;
(ii) For all 𝛼, 𝛽 ∈ 𝐼, we have 𝛼𝛽 ∈ 𝐼;
(iii) 𝐼 ⊆ OL (𝐼), so 𝐼 is a left ideal of OL (𝐼) in the usual sense;
(iii0) 𝐼 ⊆ OR (𝐼); and
(iv) 𝐼 ⊆ OL (𝐼) ∩ OR (𝐼).
If 𝐼 is integral, then every element of 𝐼 is integral over 𝑅.
Proof. The equivalence (i) ⇔ (ii) follows immediately. For (i) ⇔ (iii), we have 𝐼 𝐼 ⊆ 𝐼
if and only if 𝐼 ⊆ OL (𝐼) by definition of OL (𝐼), and the same argument gives (i)
⇔ (iii0), and this then gives (i) ⇔ (iv). The final statement follows from Lemma
10.3.2.
In light of Lemma 16.2.8, we need not define notions of left integral or right
integral.
For an 𝑅-lattice 𝐼, there exists nonzero 𝑑 ∈ 𝑅 such that 𝑑𝐼 is integral, so every
𝑅-lattice 𝐼 = (𝑑𝐼)/𝑑 is fractional in the sense that it is obtained from an integral lattice
with denominator.
Definition 16.2.9. Let O ⊆ 𝐵 be an 𝑅-order. A left fractional O-ideal is a lattice
𝐼 ⊆ 𝐵 such that O ⊆ OL (𝐼); similarly on the right.
If O, O0 ⊆ 𝐵 are 𝑅-orders, then a fractional O, O0-ideal is a lattice 𝐼 that is a left
fractional O-ideal and a right fractional O0-ideal.
Remark 16.2.10. A left ideal 𝐼 ⊆ O in the usual sense is an integral left O-ideal
in the sense of Definition 16.2.9 if and only if 𝐼 𝐹 = 𝐵, i.e., 𝐼 is a (full) 𝑅-lattice.
(Same for right and two-sided ideals.) If 𝐼 is nonzero and 𝐵 is a division algebra, then
automatically 𝐼 is full and the two notions coincide.
Indeed, suppose 𝐼 ⊆ O is a left ideal of O (in the usual sense). Then O ⊆ OL (𝐼)
so in particular 𝐼 has the structure of an 𝑅-module, and since O is finitely generated as
an 𝑅-module and 𝑅 is noetherian, it follows that 𝐼 is finitely generated. Consequently,
a left ideal 𝐼 ⊆ O is a left fractional O-ideal if and only if 𝐼 𝐹 = 𝐵.
Definition 16.2.11. Let 𝐼 be a left fractional O-ideal. We say that 𝐼 is sated (as a left
fractional O-ideal) if O = OL (𝐼). We make a similar definition on the right and for
two-sided ideals.
Example 16.2.12. By Lemma 15.6.16, codiff(O) is a two-sided sated O-ideal.
Remark 16.2.13. Our notion of sated is sometimes called proper: we do not use this
already overloaded term, as it conflicts with the notion of a proper subset.
Lemma 16.3.2. The reduced norm nrd(𝐼) is a fractional ideal of 𝐹: i.e., it is finitely
generated as an 𝑅-module.
Proof. We first give a proof when 𝐵 has a standard involution, and nrd is a quadratic
form. Since 𝐼 is an 𝑅-lattice we have 𝐼 𝐹 = 𝐵; since nrd(𝐵) ≠ {0}, we have nrd(𝐼) ≠
{0}. And 𝐼 is generated by finitely many 𝛼𝑖 as an 𝑅-module; the 𝑅-module nrd(𝐼) is
then generated by the values 𝑎 𝑖𝑖 = nrd(𝛼𝑖 ) and 𝑎 𝑖 𝑗 = nrd(𝛼𝑖 + 𝛼 𝑗 ) − nrd(𝛼𝑖 ) − nrd(𝛼 𝑗 ),
since then Í
Í Í
nrd 𝑖 𝑐 𝑖 𝛼𝑖 = 𝑖, 𝑗 𝑎 𝑖 𝑗 𝑐 𝑖 𝑐 𝑗 ∈ 𝑖, 𝑗 𝑅𝑎 𝑖 𝑗
for all 𝑐 𝑖 ∈ 𝑅.
Now for the general case. Replacing 𝐼 by 𝑟 𝐼 with 𝑟 ∈ 𝑅 nonzero, we may suppose
that 𝐼 is integral, and hence nrd(𝐼) ⊆ 𝑅. Since 𝐼 is a lattice, there exists 𝑟 ∈ 𝐼 ∩ 𝑅
with 𝑟 ≠ 0. For all 𝔭 such that ord𝔭 (𝑟) = 0, we have 1 ∈ 𝐼 (𝔭) so nrd(𝐼 (𝔭) ) = 𝑅 (𝔭) .
For each of the finitely many primes 𝔭 that remain, we choose an element 𝛼 ∈ 𝐼 such
that ord𝔭 (nrd(𝛼)) is minimal; then nrd(𝛼) generates nrd(𝐼 (𝔭) ), and by the local-global
dictionary, these finitely many elements generate nrd(𝐼).
16.3.3. For a prime 𝔭 of 𝑅 we have nrd(𝐼) (𝔭) = nrd(𝐼 (𝔭) ), so by the local-global
property of lattices (Lemma 9.4.6),
Ù Ù
nrd(𝐼) = nrd(𝐼) (𝔭) = nrd(𝐼 (𝔭) ). (16.3.4)
𝔭 𝔭
Now suppose that 𝐼, 𝐽 are lattices. Then nrd(𝐼𝐽) ⊇ nrd(𝐼) nrd(𝐽). However, we
need not have equality, as the following example indicates.
𝑅 𝑎 −1 𝑅
𝑅 𝑎𝑅
OR (𝐼) = and OL (𝐽) = −1 ,
𝑎𝑅 𝑅 𝑎 𝑅 𝑅
The issue present in Example 16.3.6 is that the product is not as well-behaved
for noncommutative rings as for commutative rings; we need the elements coming
between 𝐼 and 𝐽 to match up.
Lemma 16.3.7. Suppose that 𝐼 is compatible with 𝐽 and that either 𝐼 or 𝐽 is locally
principal. Then nrd(𝐼𝐽) = nrd(𝐼) nrd(𝐽).
16.4. ALGEBRA AND ABSOLUTE NORM 255
Proof. By the local-global property for norms (16.3.4) and since localization com-
mutes with multiplication, i.e.,
𝐼𝐽 = (𝛼O)𝐽 = 𝛼(O𝐽) = 𝛼𝐽
and so nrd(𝐼𝐽) = nrd(𝛼) nrd(𝐽) = nrd(𝐼) nrd(𝐽) by 16.3.5. The case where 𝐽 is
principal follows in the same way.
Lemma 16.3.8. Let 𝐼 be locally principal and let 𝛼 ∈ 𝐼. Then 𝛼 generates 𝐼 if and
only if nrd(𝛼)𝑅 = nrd(𝐼).
Remark 16.4.2. The definition of algebra norm by necessity depends on the choice of
domain 𝑅; indeed, 𝐼 is an 𝑅-lattice.
16.4.5. Recalling the proof of Proposition 16.4.3 and the definition of 𝑅-index, we
always have the containment
and the same on the right; by Propostion 16.4.3, equality is equivalent to 𝐼 being locally
principal.
To conclude this section, we suppose for its remainder that 𝐹 is a local with
valuation ring 𝑅 or a global number field. Then the reduced norm is also related to the
absolute norm, an absolute measure of size, as follows.
16.4.6. For a fractional ideal 𝔞 of 𝑅, we define the absolute norm N (𝔞) to be
and if 𝔞 ⊆ 𝑅 then
N (𝔞) = #(𝑅/𝔞),
so this norm is also called the counting norm.
We extend this definition to elements 𝑎 ∈ 𝐹 × by defining N (𝑎) := N (𝑎𝑅).
16.4.8. Similarly, if 𝐼 ⊆ 𝐵 is a locally principal 𝑅-lattice, we define the absolute norm
of 𝐼 to be
N (𝐼) := [OL (𝐼) : 𝐼] Z = [OR (𝐼) : 𝐼] Z , (16.4.9)
the latter equality by taking 𝑅 = Z in Proposition 16.4.3. If 𝐼 is integral then
Remark 16.4.11. The absolute norm may also be defined for a global function field,
but there is no canonical ‘ring of integers’ as above.
𝐼 𝐼 −1 𝐼 ⊆ 𝐼.
Proof. Statement (a) follows as in the proof of Lemma 10.2.7, and the inclusion is by
the definition of 𝐼 −1 . For statement (b), if 𝛼 ∈ O, then O𝛼O ⊆ O since O is an order;
conversely, if O𝛼O ⊆ O, then taking 1 ∈ O on left and right we conclude 𝛼 ∈ O.
258 CHAPTER 16. QUATERNION IDEALS AND INVERTIBILITY
Proof. The implication (i) ⇒ (ii) is clear. For (ii) ⇒ (i), we need to check the
compatibility of the product: but since 𝐼 −1 𝐼 = OR (𝐼) we have OL (𝐼 −1 ) ⊆ OR (𝐼), and
from the other direction we have the other containment, so these are equal.
The implication (i) ⇒ (iii) is clear. For (iii) ⇒ (i), suppose that 𝐼 0 is an inverse
to 𝐼. Then 𝐼 = 𝐼 𝐼 0 𝐼 so 𝐼 0 ⊆ 𝐼 −1 by definition. Therefore 𝐼 ⊆ 𝐼 𝐼 −1 𝐼 ⊆ 𝐼 and equality
holds throughout. Multiplying by 𝐼 0 on the left and right then gives
𝐼 −1 = (𝐼 0 𝐼)𝐼 −1 (𝐼 𝐼 0) = 𝐼 0 𝐼 𝐼 0 = 𝐼 0 .
Again the implication (i) ⇒ (iv) is immediate. To prove (iv) ⇒ (ii), we need to
show that 𝐼 𝐼 −1 = OL (𝐼) and 𝐼 −1 𝐼 = OR (𝐼); we show the former. By compatibility,
OR (𝐼 −1 ) = OL (𝐼) = O. If 𝐼 𝐼 −1 = 𝐽 then 𝐽 = 𝐼 𝐼 −1 = O(𝐼 𝐼 −1 )O = O𝐽O, so 𝐽 ⊆ O is a
two-sided ideal of O containing 1 hence 𝐽 = O.
Lemma 16.5.9. 𝐼 is invertible if and only 𝐼 (𝔭) is invertible for all primes 𝔭.
A compatible product with an invertible lattice respects taking left (and right)
orders, as follows.
Proof. We always have OL (𝐼) ⊆ OL (𝐼𝐽) (even without 𝐽 invertible). To show the
other containment, suppose that 𝛼 ∈ OL (𝐼𝐽), so that 𝛼𝐼𝐽 ⊆ 𝐼𝐽. Multiplying by 𝐽 −1 ,
we conclude 𝛼𝐼 ⊆ 𝐼 and 𝛼 ∈ OL (𝐼).
𝑝, 𝑝
Example 16.5.12. Let 𝑝 ∈ Z be prime. Let 𝐵 := and
Q
O := Z ⊕ 𝑝Z𝑖 ⊕ 𝑝Z 𝑗 ⊕ Z𝑖 𝑗
𝐼 := 𝑝 2 Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑖 𝑗 .
𝐼 −1 = 𝑝Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑖 𝑗 (16.5.13)
and
1 1
OL (𝐼 −1 ) = OR (𝐼 −1 ) = Z + Z𝑖 + Z 𝑗 + Z𝑖 𝑗 = Z + O; (16.5.14)
𝑝 𝑝
so in the product
𝐼 𝐼 −1 = 𝐼 −1 𝐼 = 𝑝Z ⊕ 𝑝Z𝑖 ⊕ 𝑝Z 𝑗 ⊕ Z𝑖 𝑗 ( O (16.5.15)
𝐼 2 = 𝐼 𝐼 = 𝐼 𝐼 = 𝑝Z ⊕ 𝑝Z𝑖 ⊕ 𝑝Z 𝑗 ⊕ Z𝑖 𝑗 (16.5.16)
and the same on the right. In other words, if we are going to call out an invertible
fractional ideal by labelling actions on left and right, then we require these labels to
be the actual orders that make the inverse work.
Remark 16.5.19. Example 16.1.1 suggested the ‘real issue’ with noninvertible modules
for quadratic orders: as an abelian group,
√
𝔣 = 𝑓 Z + 𝑓 𝑑Z = 𝑓 · 𝑆(𝑑),
260 CHAPTER 16. QUATERNION IDEALS AND INVERTIBILITY
𝑆(𝔞) := {𝑥 ∈ 𝐾 : 𝑥𝔞 ⊆ 𝔞};
the ring 𝑆(𝔞) is an order√of 𝐾 and so is also called the order of 𝔞. In the example
above, 𝑆(𝔣) = 𝑆( 𝑓 (Z + 𝑑 𝐾 Z)) = 𝑆(𝑑 𝐾 ) ) 𝑆(𝑑). It turns out that every lattice in
𝐾 is invertible as an ideal of its multiplicator ring [Cox89, Proposition 7.4], and this
statement plays an important role in the theory of complex multiplication. (Sometimes,
an ideal 𝔞 ⊆ 𝑆 is called proper or regular if 𝑆 = 𝑆(𝔞); both terms are overloaded in
mathematics, so we will mostly resist this notion.)
Unfortunately, unlike the quadratic case, not every lattice 𝐼 ⊂ 𝐵 is projective as
a left module over its left order (or the same on the right): this is necessary, but not
sufficient. In Chapter 24, we classify orders O with the property that every lattice 𝐼
having OL (𝐼) = O is projective as an O-module: they are the Gorenstein orders.
Remark 16.5.20. Invertible lattices give rise to a Morita equivalence between their
corresponding left and right orders: see Remark 7.2.20.
Main Theorem 16.6.1. Let 𝑅 be a Dedekind domain with field of fractions 𝐹, and let
𝐵 be a finite-dimensional 𝐹-algebra with a standard involution. Then an 𝑅-lattice 𝐼 is
invertible if and only if 𝐼 is locally principal.
Remark 16.6.2. We can relax the hypothesis that 𝑅 is a Dedekind domain and in-
stead work with a Prüfer domain, a generalization of Dedekind domains to the non-
noetherian context.
We have already seen (Corollary 16.5.10) that the implication (⇒) in Main Theo-
rem 16.6.1 holds without the hypothesis of a standard involution; the reverse implica-
tion is the topic of this section. This implication is not in general true if this hypothesis
is removed (but is true again when 𝐵 is commutative); see Exercise 16.18(a).
Remark 16.6.3. The provenance of the hypothesis that 𝑅 is a Dedekind domain is the
following: if 𝔞 ⊂ 𝑅 is not invertible as an 𝑅-module, and O ⊂ 𝐵 is an 𝑅-order, then
𝔞O is not invertible as an 𝑅-lattice. To make the simplest kind of arguments here, we
would like for all (nonzero) ideals 𝔞 ⊆ 𝑅 to be invertible, and this is equivalent to the
requirement that 𝑅 is a Dedekind domain (see section 9.2).
Throughout this section, let 𝑅 be a Dedekind domain with field of fractions 𝐹, let
𝐵 be a finite-dimensional 𝐹-algebra, and let 𝐼 ⊂ 𝐵 be an 𝑅-lattice. The following
concept will be useful in this section.
Proof. We have that 𝛼 ∈ 𝐼 is integral over 𝑅 if and only if trd(𝛼) ∈ 𝑅 and nrd(𝛼) ∈ 𝑅
(by Corollary 10.3.6, since 𝑅 is integrally closed) if and only if nrd(𝛼) ∈ 𝑅 and
nrd(𝛼 + 1) = nrd(𝛼) + trd(𝛼) + 1 ∈ 𝑅.
By Lemma 16.6.7, the standard involution gives a bijection between the set of
lattices 𝐼 with OL (𝐼) = O and the set of lattices with OR (𝐼) = O.
16.6.9. Suppose that 𝑅 is a DVR (e.g., a localization of 𝑅 at a prime ideal 𝔭). We will
show how to reduce the proof of Main Theorem 16.6.1 to that of a semi-order.
Since 𝑅 is a DVR, the fractional 𝑅-ideal nrd(𝐼) ⊆ 𝑅 is principal, generated by an
element with minimal valuation: let 𝛼 ∈ 𝐼 achieve this minimum reduced norm. Then
the 𝑅-lattice 𝐽 = 𝛼−1 𝐼 now satisfies 1 ∈ 𝐽 and nrd(𝐽) = 𝑅. Thus 𝐽 is a semi-order,
and 𝐽 is (locally) principal if and only if 𝐼 is (locally) principal.
Proof of Main Theorem 16.6.1. The proof is due to Kaplansky [Kap69, Theorem 2].
The statement is local; localizing, we may suppose 𝑅 is a DVR. By 16.6.9, we reduce
to the case where 𝐼 is a semi-order. In particular, we have 1 ∈ 𝐼. Let 𝛼1 , . . . , 𝛼𝑛 be an
𝑅-basis for 𝐼.
We claim that
𝐼 𝑛+1 = 𝐼 𝑛 (16.6.10)
Since 1 ∈ 𝐼, we have 𝐼 𝑛 ⊆ 𝐼 𝑛+1 . It suffices then to prove that a product of 𝑛 + 1 basis
elements of 𝐼 lies in 𝐼 𝑛 . By the pigeonhole principle, there must be a repeated term 𝛼𝑖
among them. We recall the formula (4.2.16)
for all 𝛼, 𝛽 ∈ 𝐵. We can use this relation to “push” the second instance of the repeated
element until it meets with its mate, at the expense of terms lying in 𝐼 𝑛 . More precisely,
in the 𝑅-module 𝐼 2 /𝐼, by (16.6.11),
𝛼𝑖 𝛼 𝑗 ≡ −𝛼 𝑗 𝛼𝑖 (mod 𝐼)
for all 𝜇, 𝜈 appropriate products of basis elements. Therefore we may suppose that
the repetition 𝛼𝑖2 is adjacent; but then 𝛼𝑖 satisfies a quadratic equation and 𝛼𝑖2 =
trd(𝛼𝑖 )𝛼𝑖 − nrd(𝛼𝑖 ) ∈ 𝐼, so in fact the product belongs to 𝐼 𝑛 , and the claim follows.
Now suppose 𝐼 is invertible; we wish to show that 𝐼 is principal. From the
equality 𝐼 𝑛+1 = 𝐼 𝑛 , we multiply both sides of this equation by (𝐼 −1 ) 𝑛 and obtain
𝐼 = O = OL (𝐼) = OR (𝐼). In particular, 𝐼 is principal, generated by 1.
16.6.14. In the presence of a standard involution, we can write the inverse in another
way: if 𝐼 is invertible, then
𝐼 −1 = 𝐼 nrd(𝐼) −1 .
We similarly define left invertible and left inverse. Applying the same reasoning
as in Lemma 16.5.9, we see that one-sided invertibility is a local property.
Remark 16.7.2. For rings, the (left or) right inverse of an element need not be unique
even though a two-sided inverse is necessarily unique. Similarly, left invertibility does
not imply right invertibility for lattices in general, and so the one-sided notions can be
a bit slippery: see Exercise 16.18(b).
Remark 16.7.3. The compatibility condition in invertibility is important to avoid
trivialities.
Consider
Example 16.3.6: we have 𝐼𝐽 = M2 (𝑅) = OL (𝐼), and if we let
𝑏𝑅 𝑏𝑅
𝐽 = for any nonzero 𝑏 ∈ 𝑅, the equality 𝐼𝐽 = M2 (𝑅) remains true. Not
𝑅 𝑅
every author requires compatibility in the definition of (sided) invertibility.
A natural candidate for the right inverse presents itself: if 𝐼 𝐼 0 = OL (𝐼), then 𝐼 0
maps 𝐼 into OL (𝐼) on the right. We recall the definition of the colon lattices (Definition
15.6.11). Let 𝐼 0 := (OL (𝐼) : 𝐼) R . Then 𝐼 𝐼 0 ⊆ OL (𝐼) by definition; however, in general
equality need not hold and the product need not be compatible. Similarly, since
𝐼 𝐼 −1 𝐼 ⊆ 𝐼 we have 𝐼 𝐼 −1 ⊆ OL (𝐼), but again equality need not hold.
The sided version of Proposition 16.5.8 also holds.
264 CHAPTER 16. QUATERNION IDEALS AND INVERTIBILITY
Proof. This is just a sided restriction of the proof of Proposition 16.5.8. For example,
to show (ii) ⇒ (i), we always have 𝐼 𝐼 −1 𝐼 ⊆ 𝐼 so 𝐼 𝐼 −1 ⊆ OL (𝐼); if 𝐼 0 is a right inverse to
𝐼, then 𝐼 𝐼 0 𝐼 = OL (𝐼)𝐼 = 𝐼 and 𝐼 0 ⊆ 𝐼 −1 , and therefore 𝐼 𝐼 −1 ⊇ 𝐼 𝐼 0 = OL (𝐼). Therefore
𝐼 𝐼 −1 = OL (𝐼) and 𝐼 −1 is a right inverse for 𝐼.
Returning to the setting of the previous section, however, we can show that the
one-sided notions of invertibility are equivalent to the two-sided notion.
Lemma 16.7.5. Suppose 𝐵 has a standard involution. Then an 𝑅-lattice 𝐼 is left
invertible if and only if 𝐼 is right invertible if and only if 𝐼 is invertible.
Proof. We will show that if 𝐼 is right invertible then 𝐼 is left invertible; the other
implications follow similarly. By localizing, we reduce to the case where 𝑅 is a DVR.
By the results of 16.6.9, we may suppose that 𝐼 is a semi-order, so that OL (𝐼) =
OR (𝐼) = O and 𝐼 = 𝐼. Suppose 𝐼 𝐼 0 = O. Then 𝐼 0 𝐼 = O = O, and 𝐼 0 is compatible with
𝐼 since
O = OR (𝐼) = OL (𝐼 0) = OR (𝐼 0)
as desired.
Corollary 16.7.6. Suppose 𝑅 is a Dedekind domain and that 𝐵 has a standard in-
volution. Then an 𝑅-lattice 𝐼 is right invertible with 𝐼 𝐼 0 = OL (𝐼) if and only if
𝐼 0 = (OL (𝐼) : 𝐼) R = 𝐼 −1 .
A similar statement holds for the left inverse; in particular, this shows that a right
inverse is necessarily unique.
Proof. The implication (⇒) is immediate, so we prove (⇐). Let O = OL (𝐼). Then
O = 𝐼 𝐼 0 ⊆ 𝐼 (O : 𝐼) R ⊆ O
Proof. Main Theorem 16.6.1 proves (i) ⇔ (ii). For the equivalence (ii) ⇔ (iii) ⇔
(iii0), apply Lemma 16.7.5. Finally, the equivalence (i) ⇔ (iv) ⇔ (iv0) is supplied by
Proposition 16.4.3.
16.8.3. If codiff(O) is locally principal (section 16.2), then so is diff (O), and by
Proposition 16.4.3 we have
Proposition 16.8.5. If codiff(O) is right invertible, then all sated left fractional O-
ideals are right invertible. Similarly, if codiff(O) is left invertible, then all sated right
fractional O-ideals are left invertible.
Corollary 16.8.7. Suppose that 𝐵 has a standard involution. Then the following are
equivalent:
Proof. Combine Proposition 16.8.5 and (16.8.6) with Lemma 16.7.5 and Corollary
16.7.6.
Proof. See Kaplansky [Kap69, Theorem 10] or Brzezinski [Brz82, Theorem 3.4].
Exercises
Unless otherwise specified, throughout these exercises let 𝑅 be a Dedekind domain
with field of fractions 𝐹, let 𝐵 be a finite-dimensional 𝐹-algebra, and let 𝐼 ⊆ 𝐵 be an
𝑅-lattice.
√
1. Let 𝑑 ∈ Z be a nonsquare discriminant, and let 𝑆(𝑑) = Z[(𝑑 + 𝑑)/2] be the
quadratic ring of discriminant 𝑑.
√
(a) Suppose that 𝑑 = 𝑑 𝐾 𝑓 2 with 𝑓 > 1. Show that the ideal ( 𝑓 , 𝑑) of 𝑆(𝑑)
is not invertible. √
(b) Consider 𝑑 = −12, and 𝑆 = 𝑆(−12) = Z[ −3]. Show that every invertible
ideal of 𝑆 is principal (so 𝑆 has class number 1), but that 𝑆 is not a PID.
I 2. Show that if 𝐼 = OL (𝐼)𝛼 with 𝛼 ∈ 𝐵× , then OR (𝐼) = 𝛼−1 OL (𝐼)𝛼.
I 3. Show that if 𝐽 is an 𝑅-lattice in 𝐵 and 𝜇 ∈ 𝐵× , then 𝜇𝐽 = 𝐽 if and only if
𝜇 ∈ OL (𝐽) × .
I 4. Show that if 𝛼 ∈ 𝐵 then nrd(𝛼𝐼) = nrd(𝛼) nrd(𝐼). Conclude that if 𝐼 is a
principal 𝑅-lattice, generated by 𝛼 ∈ 𝐼, then nrd(𝐼) = nrd(𝛼)𝑅.
5. Let 𝛼1 , . . . , 𝛼𝑛 generate 𝐼 as an 𝑅-module. Give an explicit example where
nrd(𝐼) is not generated by nrd(𝛼𝑖 ) (cf. Lemma 16.3.2). Moreover, show that for
an 𝑅-lattice 𝐼, there exists a set of 𝑅-module generators 𝛼𝑖 such that nrd(𝐼) is in
fact generated by nrd(𝛼𝑖 ).
16.8. INVERTIBILITY AND THE CODIFFERENT 267
((𝐼 : 𝐽) L : 𝐾) R = ((𝐼 : 𝐾) R : 𝐽) L .
I 10. Let 𝐼, 𝐽 ⊆ 𝐵 be 𝑅-lattices and suppose that 𝐼 is compatible with 𝐽. Show that 𝐼𝐽
is invertible (with (𝐼𝐽) −1 = 𝐽 −1 𝐼 −1 ) if 𝐼, 𝐽 are invertible, but the converse need
not hold.
11. Let 𝐼, 𝐽 ⊆ 𝐵 be 𝑅-lattices, and suppose that 𝐽 is invertible. Show that (𝐼 : 𝐽) L =
𝐼𝐽 −1 and (𝐼 : 𝐽) R = 𝐽 −1 𝐼.
12. Let 𝑝 be prime, let 𝐵 = ( 𝑝, 𝑝 | Q), and let O := Zh𝑖, 𝑗i = Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑖 𝑗.
(a) Let 𝐼 = {𝛼 ∈ O : 𝑝 | nrd(𝛼)}. Show that 𝐼 = 𝑝Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑖 𝑗.
(b) Show 𝐼 = O𝑖 + O 𝑗, that O is a two-sided O-ideal, and that [O : 𝐼] = 𝑝.
(c) Show that 𝐼 ( 𝑝) ≠ O ( 𝑝) 𝛼 for all 𝛼 ∈ 𝐼 ( 𝑝) . [Hint: show that if 𝛼 ∈ 𝐼, then
𝑝 2 | [O : O𝛼].]
(d) Compute that OL (𝐼) = Z + Z𝑖 + Z 𝑗 + Z(𝑖 𝑗/𝑝) ) O, and that 𝐼 = OL (𝐼)𝑖 =
OL (𝐼) 𝑗.
(e) Compute codiff(O) and diff (O) and show they are invertible.
[Compare Lemurell [Lem2011, Remark 6.4].]
I 13. Let 𝐾 be a separable quadratic field extension of 𝐹 and let 𝐼 ⊆ 𝐾 be an 𝑅-lattice.
Let O = OL (𝐼) = OR (𝐼).
(a) Show that 𝐼 𝐼 = 𝐼 𝐼 = nrd(𝐼)O. [Hint: argue as in Proposition 16.6.15.]
(b) Conclude that 𝐼 is invertible as a O-module.
14. Show that if 𝐼, 𝐽 ⊆ 𝐵 are locally principal (hence invertible) 𝑅-lattices, then
[𝐼 : 𝐽] 𝑅 = [𝐽 −1 : 𝐼 −1 ] 𝑅 .
(b) Now let 𝔟 ⊂ 𝐾 be a fractional 𝑆-ideal. Show that the following are
equivalent:
(i) 𝔟 is a locally principal 𝑆-ideal;
(ii) 𝔟 is invertible as a fractional 𝑆-ideal, i.e., there exists a fractional ideal
𝔟−1 such that 𝔟𝔟−1 = 𝑆 (necessarily 𝔟−1 = (𝑆 : 𝔟));
(iii) There exists 𝑑 ∈ 𝐾 × such that 𝑑𝔟 + 𝔣 ∩ 𝑆 = 𝑆; and
(iv) 𝔟 is proper, i.e., 𝑆 = O(𝔟) = {𝑥 ∈ 𝐾 : 𝑥𝔟 ⊆ 𝔟}.
I 17. Let O ⊆ 𝐵 be an 𝑅-order.
(a) Let 𝛼 ∈ 𝐵× . Show that 𝐼 = O𝛼 is a lattice with OL (𝐼) = OR (𝐼) = O if
and only if 𝛼 ∈ 𝐵× and O𝛼 = 𝛼O. Conclude that the set of invertible
two-sided principal lattices 𝐼 with OL (𝐼) = OR (𝐼) = O forms a group.
(b) Show that the normalizer of O,
𝑁 𝐵× (O) = {𝛼 ∈ 𝐵× : 𝛼O𝛼−1 = O}
𝜋𝑅 𝜋𝑅 𝑅
𝐼 = 𝜋𝑅 𝑅 ® ⊂ 𝐵 = M3 (𝐹)
© ª
𝜋𝑅
«𝑅 𝑅 𝑅¬
𝜋𝑅 𝜋𝑅 𝑅
𝐼 = 𝜋 2 𝑅 𝜋2 𝑅 𝑅® ⊂ 𝐵 = M3 (𝐹)
© ª
« 𝑅 𝑅 𝑅¬
Having investigated the structure of lattices and ideals in Chapter 16, we now turn to
the study of their isomorphism classes.
269
270 CHAPTER 17. CLASSES OF QUATERNION IDEALS
Unfortunately, the class set Cls O does not have the structure of a group: only
a pointed set, with distinguished element [O] R . One problem is the compatibility
of multiplication discussed in the previous chapter. But even if we allowed products
between incompatible lattices, the product need not be well-defined: the lattices 𝐼𝐽 and
𝐼𝛼𝐽 for 𝛼 ∈ 𝐵× need not be in the same class, because of the failure of commutativity.
(This is the reason we write ‘Cls’ instead of ‘Cl’, as a reminder that it is only a class
set.) In Chapter 19, we will describe the structure that arises naturally instead: a
partially defined product on classes of lattices, a groupoid.
In any case, using the same method of proof (geometry of numbers) as in the
commutative case, we will show that there exists a constant 𝐶 (depending on O)
such that every class in Cls O is represented by an integral ideal 𝐼 ⊆ O with N (𝐼) =
#(O/𝐼) ≤ 𝐶. As a consequence, we have the following fundamental theorem.
Theorem 17.1.1. Let 𝐵 be a quaternion algebra over Q and let O ⊂ 𝐵 be an order.
Then the right class set Cls O is finite.
Accordingly, we call # Cls O ∈ Z ≥1 the (right) class number of O.
Right class sets pass between orders as follows. Let O, O0 ⊂ 𝐵 be orders. If
O ' O0 are isomorphic as rings, then of course this isomorphism induces a bijection
Cls O −→∼ Cls O0. In fact, O ' O0 if and only if there exists 𝛼 ∈ 𝐵× such that
O = 𝛼−1 O𝛼 by the Skolem–Noether theorem; for historical reasons, we say that
0
Cls O → Typ O
(17.1.3)
[𝐼] R ↦→ class of OL (𝐼)
is a surjective map of sets, so the type set is finite: in other words, up to isomorphism,
there are only finitely many types of orders in the genus of O. All maximal orders in 𝐵
are in the same genus, so in particular there are only finitely many conjugacy classes
of maximal orders in 𝐵. In this way, the right class set of O also organizes the types
of orders arising from O.
The most basic question about the class number is of course its size (as a function
of O). In the case of quadratic fields, the behavior of the class group depends in a
significant way on whether the field is imaginary or real: for negative discriminant
17.2. MATRIX RING 271
√︁
𝑑 < 0, the Brauer–Siegel theorem provides that # Cl 𝑆 is approximately of size | 𝑑|;
in contrast, for positive discriminant 𝑑 > 0, one typically sees a small class group and
a correspondingly large fundamental unit, but this statement is notoriously difficult to
establish unconditionally.
The same dichotomy is at play in the case of quaternion algebras, and to state the
cleanest results we suppose that O is a maximal order. Let 𝐷 := disc 𝐵 = discrd(O)
be the discriminant of 𝐵. If 𝐵 is definite, which is to say ∞ ∈ Ram 𝐵, then 𝐵 is like
an imaginary quadratic field 𝐾: the norm is positive definite. In this case, # Cls O is
approximately of size 𝐷, a consequence of the Eichler mass formula, the subject of
Chapter 25. On the other hand, if 𝐵 is indefinite, akin to a real quadratic field, then
# Cls O = 1, this time a consequence of strong approximation, the subject of Chapter
28. Just as in the commutative case, estimates on the size of the class number use
analytic methods and so must wait until we have developed the required tools.
17.2.1. Let 𝑅 be a PID with field of fractions 𝐹, and let 𝐵 = M𝑛 (𝐹). By Corollary
10.5.5, every maximal order of 𝐵 = M𝑛 (𝐹) is conjugate to M𝑛 (𝑅). Moreover, every
two-sided ideal of M𝑛 (𝑅) is principal, generated by an element 𝑎 ∈ 𝐹 × (multiplying
a candidate ideal by matrix units, as in Exercise 7.5(b)), so the group of fractional
two-sided M𝑛 (𝑅)-ideals is canonically identified with the group of fractional 𝑅-ideals,
itself isomorphic to the free abelian group on the (principal) nonzero prime ideals of
𝑅.
Just as in the two-sided case, the right class set for M𝑛 (𝑅) is trivial.
Proposition 17.2.2. Let 𝑅 be a PID with field of fractions 𝐹, and let 𝐵 = M𝑛 (𝐹). Let
𝐼 ⊆ 𝐵 be an 𝑅-lattice with either OL (𝐼) or OR (𝐼) maximal. Then 𝐼 is principal, and
both OL (𝐼) and OR (𝐼) are maximal.
Proof. For each prime 𝔭 of 𝑅, we have that 𝑅𝔭 is a DVR and one of two possibilities:
either 𝐵𝔭 ' M2 (𝐹𝔭 ), in which case we can apply Lemma 17.2.2 to conclude 𝐼𝔭 is
principal, or 𝐵𝔭 is a division algebra, and we instead apply 13.3.10 to conclude that 𝐼𝔭
is principal.
Proof. For (i) ⇒ (ii). If 𝐼 ∼R 𝐽 then 𝐽 = 𝛼𝐼 with 𝛼 ∈ 𝐵× , so OR (𝐽) = OR (𝐼) and the
map left-multiplication by 𝛼 gives a right O-module isomorphism 𝐼 −→∼ 𝐽. Conversely,
∼
for (i) ⇐ (ii), suppose that 𝜙 : 𝐼 −→ 𝐽 is an isomorphism of right O-modules. Then
𝜙 𝐹 : 𝐼 ⊗𝑅 𝐹 = 𝐵 −→ ∼ 𝐽 ⊗ 𝐹 = 𝐵 is an automorphism of 𝐵 as a right 𝐵-module.
𝑅
Then as in Example 7.2.14, such an isomorphism is obtained by left multiplication by
𝛼 ∈ 𝐵× , so by restriction 𝜙 is given by this map as well.
Next, for (i) ⇒ (iii), suppose 𝛼𝐼 = 𝐽 with 𝛼 ∈ 𝐵× . Then
In view of 17.3.2, we will soon abbreviate Cls O := ClsR O and drop the subscript
R from the classes, when no confusion can result.
Remark 17.3.5. The notation Cl O is also used for the class set, but it sometimes means
instead the stably free class group or some other variant. We use “Cls” to emphasize
that we are working with a class set.
17.3.6. The set ClsR O has a distinguished element [O] R ∈ ClsR O, so it has the
structure of a pointed set (a set equipped with a distinguished element of the set).
However, in general it does not have the structure of a group under multiplication:
for classes [𝐼] R , [𝐽] R , we have [𝛼𝐽] R = [𝐽] R for 𝛼 ∈ 𝐵× but we need not have
[𝐼𝛼𝐽] R = [𝐼𝐽] R , because of the lack of commutativity.
17.3.7. An argument similar to the one in Proposition 17.2.2, either arguing locally or
with pseudobases (9.3.7), yields the following [CR81, (4.13)].
Let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅, and let 𝐼 ⊆ 𝐵 be an 𝑅-lattice with
OL (𝐼) = M𝑛 (𝑅). Then there exists 𝛽 ∈ GL𝑛 (𝐹) and fractional ideals 𝔞1 , · · · , 𝔞𝑛 such
that
𝐼 = M𝑛 (𝑅) diag(𝔞1 , . . . , 𝔞𝑛 ) 𝛽 (17.3.8)
where diag(𝔞1 , . . . , 𝔞𝑛 ) is the 𝑅-module of diagonal matrices with entries in the given
fractional ideal. The representation (17.3.8) is called the Hermite normal form of
the 𝑅-module 𝐼, because it generalizes the Hermite normal form over a PID (allowing
coefficient ideals).
By 9.3.10, the Steinitz class [𝔞1 · · · 𝔞𝑛 ] ∈ Cl 𝑅 is uniquely defined. Switching to
the right, this yields a bijection
∼ Cls (M (𝑅))
Cl 𝑅 −→ R 𝑛
(17.3.9)
[𝔞] ↦→ [diag(𝔞, 1, . . . , 1) M𝑛 (𝑅)] R
Proof. If O, O0 are of the same type, then they are isomorphic (under conjugation).
∼ O0 is an isomorphism of 𝑅-algebras, then extending scalars
Conversely, if 𝜙 : O −→
to 𝐹 we obtain 𝜙 𝐹 : O𝐹 = 𝐵 −→ ∼ 𝐵 = O0 𝐹 an 𝐹-algebra automorphism of 𝐵. By
the theorem of Skolem–Noether (Corollary 7.7.4), such an automorphism is given by
conjugation by 𝛼 ∈ 𝐵× , so O, O0 are of the same type.
274 CHAPTER 17. CLASSES OF QUATERNION IDEALS
∼ O0 induces a bijection
17.4.3. If O, O0 are of the same type, then an isomorphism O −→
∼ 0
Cls O −→ Cls O of pointed sets. By Lemma 17.4.2, such an isomorphism is provided
by conjugation O0 = 𝛼−1 O𝛼 for some 𝛼 ∈ 𝐵× . The principal lattice 𝐼 = O𝛼 = 𝛼O0
has OL (𝐼) = O and OR (𝐼) = O0.
Generalizing 17.4.3, the class sets of two orders are in bijection if they are con-
nected, in the following sense.
Definition 17.4.4. O is connected to O0 if there exists a locally principal fractional
O, O0-ideal 𝐽 ⊆ 𝐵, called a connecting ideal.
The relation of being connected is an equivalence relation on the set of 𝑅-orders.
If two 𝑅-orders O, O0 are of the same type, then they are connected by a principal
connecting ideal (17.4.3).
Definition 17.4.5. We say that O, O0 are locally of the same type or locally iso-
morphic if O𝔭 and O𝔭0 are of the same type (i.e., O𝔭 ' O𝔭0 ) for all primes 𝔭 of
𝑅.
Lemma 17.4.6. The 𝑅-orders O, O0 are connected if and only if O, O0 are locally
isomorphic.
Proof. Let 𝐽 be a connecting ideal, a locally principal fractional O, O0-ideal. Then for
all primes 𝔭 of 𝑅 we have 𝐽𝔭 = O𝔭 𝛼𝔭 with 𝛼𝔭 ∈ 𝐵𝔭 , and consequently O𝔭0 = OR (𝐼𝔭 ) =
𝛼𝔭−1 O𝔭 𝛼𝔭 . Therefore O is locally isomorphic to O0.
Conversely, if O, O0 are locally isomorphic, then for all primes 𝔭 of 𝑅 we have
O𝔭 = 𝛼𝔭−1 O𝔭 𝛼𝔭 with 𝛼𝔭 ∈ 𝐵𝔭 . Since 𝑅 is a Dedekind domain, O𝔭0 = O𝔭 for all but
0
finitely many primes 𝔭, so we may take 𝛼𝔭 ∈ O𝔭 = O𝔭0 for all but finitely many primes
𝔭. Therefore, there exists an 𝑅-lattice 𝐼 with 𝐼𝔭 = O𝔭 𝛼𝔭 by the local-global principle
for lattices, and 𝐼 is a locally principal fractional O, O0-ideal.
[𝐼] R ↦→ [𝐼𝐽] R
[𝐼 𝐽 ] R ← [𝐼 0] R
0 −1
ClsR O → Typ O
(17.4.14)
[𝐼] R ↦→ class of OL (𝐼)
Remark 17.4.15. The fibers of the map (17.4.14) is given by classes of two-sided
ideals: see Proposition 18.5.10.
The main result of Minkowski’s geometry of numbers is the following convex body
theorem.
The following proposition can be seen as a generalization of what was done for the
Hurwitz order (11.3.1).
17.6. ⊲ EXAMPLE 277
8
N (𝐼) ≤ discrd(O)
𝜋2
and the right class set Cls O is finite.
𝑎, 𝑏
Proof. Let 𝐵 = , with 𝑎, 𝑏 ∈ Z<0 . Since 𝐵 is definite, there is an embedding
Q
𝐵 ↩→ 𝐵∞ = 𝐵 ⊗Q R ' H. Inside 𝐵∞ ' R4 with Euclidean norm nrd, the order O
sits as a Euclidean lattice. The set O1 is therefore a discrete subset of the compact set
𝐵∞1 ' H1 , so it is finite.
Explicitly, we identify
𝐵∞ −→ ∼ R4
√ √︁ √︁ √︁ (17.5.7)
𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ↦→ 2 𝑡, 𝑥 |𝑎|, 𝑦 |𝑏|, 𝑧 |𝑎𝑏|
𝑋 = {𝑥 ∈ R4 : k𝑥k ≤ 𝑐}.
1
N (𝛼𝐽) = Nm 𝐵 |Q (𝛼) N (𝐽) = nrd(𝛼) 2 N (𝐽) = k𝛼k 4 N (𝐽)
4 (17.5.9)
1 8
≤ 𝑐4 N (𝐽) = 2 discrd(O).
4 𝜋
Since 𝛼 is nonzero and 𝐵 is a division algebra, 𝛼 ∈ 𝐵× . Since 𝛼 ∈ 𝐽 −1 , the integral
right fractional O-ideal 𝐼 = 𝛼𝐽 ⊆ O is as desired.
If 𝐼 ⊆ O has N (𝐼) = #(O/𝐼) ≤ 𝐶 for 𝐶 ∈ Z>0 , then 𝐶O ⊆ 𝐼 ⊆ O hence there are
only finitely many possibilities for 𝐼, and the second statement follows.
17.6 ⊲ Example
We pause for an extended example. We steal the following lemma from the future.
278 CHAPTER 17. CLASSES OF QUATERNION IDEALS
Lemma 17.6.1. Let 𝑒 ∈ Z ≥0 . Then every principal right M2 (Z 𝑝 )-ideal 𝐼 with nrd(𝐼) =
𝑝 𝑒 is of the form 𝐼 = 𝛼 M2 (Z 𝑝 ) where
𝑢
𝑝 0 𝑣
𝛼∈ : 𝑢, 𝑣 ∈ Z ≥0 , 𝑢 + 𝑣 = 𝑒, and 𝑐 ∈ Z/𝑝 Z . (17.6.2)
𝑐 𝑝𝑣
Proof. The lemma follows from the theory of invariant factors: a more general state-
ment is proven in Lemma 26.4.1.
−1, −23
Example 17.6.3. Let 𝐵 = , and let
Q
1+ 𝑗 1+ 𝑗
O = Z + Z𝑖 + Z + Z𝑖 .
2 2
We have discrd(O) = disc 𝐵 = 23, so O is a maximal order, and 𝛽 = (1 + 𝑗)/2 satisfies
𝛽2 − 𝛽 + 6 = 0. For convenience, let 𝛼 = 𝑖, so O = Zh𝛼, 𝛽i. Then
𝛼𝛽 + 𝛽𝛼 = 𝛼. (17.6.4)
O ↩→ M2 (Z2 )
0 −1 1 0
𝛼, 𝛽 ↦→ , .
1 0 0 𝑏0
labelled by the corresponding nonzero column. If one of these three ideals is principal,
then it is generated by an element of reduced norm 2. We have
nrd(𝑡 + 𝑥𝛼 + 𝑦𝛽 + 𝑧𝛼𝛽)
= 𝑡 2 + 𝑡𝑦 + 𝑥 2 + 𝑥𝑧 + 6𝑦 2 + 6𝑧2
2 2 (17.6.6)
1 1 23 2 23 2
= 𝑡+ 𝑦 + 𝑥+ 𝑧 + 𝑦 + 𝑧 .
2 2 4 4
17.7. FINITENESS OF THE CLASS SET: OVER NUMBER RINGS 279
(We have 𝛼𝐼 (1:0) ≠ 𝐼 (1:0) precisely because 𝛼 ∉ OL (𝐼).) In this way, we have found
exactly one new right ideal class, [𝐼2 ] = [𝐼 (1:0) ]. We compute its left order to be
𝑖(1 + 3 𝑗)
O2 := OL (𝐼2 ) = Z + 𝛽Z + Z + (2𝑖 𝑗)Z ; O
4
and we also have a new type [O2 ] ≠ [O1 ] ∈ Typ O.
In a similar way, we find 4 right ideals of reduced norm 3, and exactly one new
right ideal class, represented by the right ideal 𝐼3 = 3O + (𝛼 + 1) 𝛽O. For example, we
find that the right ideal 𝐼 0 = 3O + 𝛽O is not principal using (17.6.6): letting
1 0
(𝐼 0 : 𝐼2 ) L = 𝐼 0 𝐼2−1 = 𝐼 𝐼2
2
and we find a shortest vector
(1 − 𝛽)/2 ∈ (𝐼 0 : 𝐼2 ) L ,
so [𝐼 0] = [𝐼2 ].
Repeating this with ideals of reduced norm 4 (Exercise 17.8), we conclude that
and letting O3 := OL (𝐼3 ), checking it is not isomorphic to the previous two orders, we
have
Typ O = {[O1 ], [O2 ], [O3 ]}.
Main Theorem 17.7.1. Let 𝐹 be a number field, let S ⊆ Pl 𝐹 be eligible and 𝑅 = 𝑅 (S)
be the ring of S-integers in 𝐹. Let 𝐵 be a quaternion algebra over 𝐹, and let O ⊆ 𝐵
be an 𝑅-order in 𝐵. Then the class set Cls O and the type set Typ O is finite.
We call # Cls O the (right) class number of O. (By 17.3.2, the left class number
suitably defined is equal to the right class number.) This result will be drastically
improved upon in Part III of this text from analytic considerations; the proof in this
section, using the geometry of numbers, has the advantage that is easy to visualize, it
works in quite some generality, and it is the launching point for algorithmic aspects.
280 CHAPTER 17. CLASSES OF QUATERNION IDEALS
17.7.2. Before we begin, two quick reductions. The finiteness of the type set follows
from finiteness of the right class set by Lemma 17.4.13. And if 𝑅 = Z𝐹 is the ring of
integers of 𝐹, then the general case follows from the fact that the map
Cls O → Cl(O ⊗𝑅 𝑅 (S) )
(17.7.3)
[𝐼] ↦→ [𝐼 ⊗𝑅 𝑅 (S) ]
is surjective for an eligible set S.
Let 𝐹 be a number field of degree 𝑛 = [𝐹 : Q], let 𝑅 = Z𝐹 be the ring of integers
in 𝐹, and let 𝐵 be a quaternion algebra over 𝐹.
17.7.4. Suppose that 𝐹 has 𝑟 real places and 𝑐 complex places, so that 𝑛 = 𝑟 + 2𝑐.
Then Ö
𝐹 ↩→ 𝐹∞ = 𝐹 ⊗Q R ' 𝐹𝑣 ' R𝑟 × C𝑐 . (17.7.5)
𝑣 |∞
Taking the basis 1, 𝑖 for C, we obtain 𝐹∞ ' R𝑛 , and then in the embedding (17.7.5),
the ring of integers 𝑅 ' Z𝑛 sits discretely inside 𝐹∞ ' R𝑛 as a Euclidean lattice.
𝑎, 𝑏
17.7.6. Suppose 𝐵 = and let 1, 𝑖, 𝑗, 𝑘 be the standard basis for 𝐵 with 𝑘 = 𝑖 𝑗,
𝐹
so 𝐵 = 𝐹 ⊕ 𝐹𝑖 ⊕ 𝐹 𝑗 ⊕ 𝐹 𝑘 ' 𝐹 4 as 𝐹-vector spaces. Then
4
𝐵 ↩→ 𝐵∞ = 𝐵 ⊗Q R ' 𝐵 ⊗𝐹 𝐹∞ ' 𝐹∞ (17.7.7)
in this same basis. Via (17.7.5) in each of the four components, the embedding (17.7.7)
then gives an identification 𝐵∞ ' (R𝑛 ) 4 ' R4𝑛 .
The order 𝑅h𝑖, 𝑗, 𝑘i = 𝑅+𝑅𝑖+𝑅 𝑗 +𝑅𝑘 is discrete in 𝐵∞ exactly because 𝑅 is discrete
in 𝐹. But then implies that an 𝑅-order O is discrete in 𝐵∞ , since [O : 𝑅h𝑖, 𝑗, 𝑘i] Z < ∞.
Therefore O ↩→ R4𝑛 has the structure of a Euclidean lattice.
In the previous section, the real vector space 𝐵∞ was Euclidean under the reduced
norm. In general, that need no longer be the case. Instead, we find a positive definite
quadratic form 𝑄 : 𝐵∞ → R that majorizes the reduced norm in the following sense:
we require that
|Nm𝐹 /Q (nrd(𝛼))| ≤ 𝑄(𝛼) 𝑛 (17.7.8)
for all 𝛼 ∈ 𝐵 ⊆ 𝐵∞ .
Remark 17.7.9. With respect to possible majorants (17.7.8): in general, there are
uncountably many such choices, and parametrizing majorants arises in a geometric
context as part of reduction theory. As it will turn out, the only “interesting” case to
consider here is 17.7.10, by strong approximation (see Theorem 17.8.3).
17.7.10. Let 𝐵 be a totally definite (Definition 14.5.7) quaternion algebra over 𝐹, a
totally real number field. Then the quadratic form
𝑄: 𝐵 → Q
∑︁
𝛼 ↦→ Tr𝐹 /Q (nrd(𝛼)) = 𝑣(nrd(𝛼)) (17.7.11)
𝑣 |∞
17.7. FINITENESS OF THE CLASS SET: OVER NUMBER RINGS 281
Then 𝑄 is a positive definite quadratic form on 𝐵∞ , again called the absolute reduced
𝑎, 𝑏
norm (relative to 𝑎, 𝑏); it depends on the choice of representation 𝐵 = .
𝐹
Nevertheless, (17.7.16) and the arithmetic-geometric mean yield
1 ∑︁
|Nm𝐹 /Q (nrd(𝛼))| 1/𝑛 ≤ 𝑚 𝑣 |𝑣(nrd(𝛼))|
𝑛
𝑣 |∞
(17.7.18)
1 ∑︁
≤ 𝑚 𝑣 𝑄 𝑣 (𝛼) = 𝑄(𝛼).
𝑛
𝑣 |∞
Proof. If 𝐵 ' M2 (𝐹), then we appeal to 17.3.7, where such a bound comes from the
finiteness of Cl 𝑅. So we may suppose that 𝐵 is a division ring.
Let
𝑋 = {(𝑥𝑖 )𝑖 ∈ R4𝑛 : 𝑄(𝛼) ≤ 1}. (17.7.20)
Then 𝑋 is closed, convex, and symmetric.
Let O be an 𝑅-order in 𝐵 and let 𝐽 be an invertible right fractional O-ideal. As in
(17.5.8), counting cosets gives
Let 1/4𝑛
covol(𝐽 −1 )
𝑐 := 2 . (17.7.22)
vol(𝑋)
Then vol(𝑐𝑋) = 𝑐4𝑛 vol(𝑋) = 24𝑛 covol(𝐽 −1 ). By Minkowski’s theorem, there exists
0 ≠ 𝛼 ∈ 𝐽 −1 ∩ 𝑐𝑋, so 𝑄(𝛼) ≤ 𝑐2 . By (17.7.18),
1 𝑛 𝑐2𝑛
|Nm𝐹 /Q (nrd(𝛼))| ≤ 𝑄(𝛼) ≤ .
𝑛𝑛 𝑛𝑛
Consequently
𝑐4𝑛
N (𝛼𝐽) = |Nm𝐹 /Q (nrd(𝛼))| 2 N (𝐽) ≤ N (𝐽)
𝑛2𝑛
24𝑛 N (𝐽) −1 covol(O) 24𝑛 covol(O) (17.7.23)
= N (𝐽) =
𝑛2𝑛 vol(𝑋) 𝑛2𝑛 vol(𝑋)
= 𝐶 N (discrd(O))
with
24𝑛 covol(O)
𝐶 := . (17.7.24)
𝑛2𝑛 vol(𝑋) N (discrd(O))
17.8. EICHLER’S THEOREM 283
Remark 17.7.25. For an explicit version of the Minkowski bound in the totally definite
case, with a careful choice of compact region, see Kirschmer [Kir2005, Theorem
3.3.11].
Lemma 17.7.26. For all 𝐶 > 0, there are only finitely many integral right O-ideals
with N (𝐼) ≤ 𝐶.
Proof of Main Theorem 17.7.1. Combine Proposition 17.7.19, the reductions in 17.7.2,
and Lemma 17.7.26.
Remark 17.7.27. The finiteness statement (Main Theorem 17.7.1) can be generalized
to the following theorem of Jordan–Zassenhaus. Let 𝑅 be a Dedekind domain with
𝐹 = Frac(𝑅) a global field, let O ⊆ 𝐵 be an 𝑅-order in a finite-dimensional semisimple
algebra 𝐵, and let 𝑉 be a left 𝐵-module. Then there are only finitely many isomorphism
classes 𝐼 ⊆ 𝐵 with O ⊆ OL (𝐼). Specializing to 𝑉 = 𝐵 a quaternion algebra, we recover
the Main Theorem 17.7.1. For a proof, see Reiner [Rei2003, Theorem 26.4]; see also
the discussion by Curtis–Reiner [CR81, §24].
Definition 17.8.1 introduces a longer (and rather opaque) phrase for something
that we already had a word for, but its use is prevalent in the literature. There are two
options: either 𝐵 is totally definite (𝐹 is a totally real field and all archimedean places
of 𝐹 are ramified in 𝐵) or 𝐵 is indefinite and satisfies the Eichler condition.
284 CHAPTER 17. CLASSES OF QUATERNION IDEALS
17.8.2. Recall 14.7.2 that we define Ω ⊆ Ram 𝐵 to be the set of real ramified places
of 𝐵 and 𝐹>×Ω 0 to be the positive elements for 𝑣 ∈ Ω.
We now define the group ClΩ 𝑅 as
the group of fractional ideals of 𝐹 under multiplication
modulo
the subgroup of nonzero principal fractional ideals
generated by an element in 𝐹>×Ω 0
If Ω is the set of all real places of 𝐹, then ClΩ 𝑅 = Cl+ 𝑅 is the narrow (or strict) class
group. On the other hand, if Ω = ∅, then ClΩ 𝑅 = Cl 𝑅. In general, we have surjective
group homomorphisms Cl+ 𝑅 → ClΩ 𝑅 and ClΩ 𝑅 → Cl 𝑅. In the language of class
field theory, ClΩ 𝑅 is the class group corresponding to the cycle given by the product
of the places in Ω.
Theorem 17.8.3 (Eichler; strong approximation). Let 𝐹 be a number field and let 𝐵
be a quaternion algebra over 𝐹 that satisfies the Eichler condition. Let O ⊆ 𝐵 be a
maximal Z𝐹 -order. Then the reduced norm induces a bijection
∼ Cl 𝑅
Cls O −→ Ω
(17.8.4)
[𝐼] ↦→ [nrd(𝐼)].
where Ω ⊆ Ram 𝐵 is the set of real ramified places in 𝐵.
Proof. Eichler’s theorem is addressed by Reiner [Rei2003, §34], with a global proof
of the key result [Rei2003, Theorem 34.9] falling over several pages. We will instead
prove a more general version of this theorem as part of strong approximation, when
idelic methods allow for a more efficient argument: see Corollary 28.5.17.
Eichler’s theorem says that when 𝐵 is not totally definite, the only obstruction for
an ideal to be principal in a maximal order is that its reduced norm fails to be (strictly)
principal in the base ring. In particular, we have the following corollary.
Corollary 17.8.5. If # Cl+ 𝑅 = 1, then # Cls O = 1: i.e., every right O-ideal of a
maximal order in an indefinite quaternion algebra is principal.
Proof. Immediate from Eichler’s theorem and the fact that Cl+ 𝑅 surjects onto ClΩ 𝑅,
by 17.8.2.
∼ Cl Z .
Corollary 17.8.6. There is a bijection Cls M2 (Z𝐹 ) −→ 𝐹
Proof. Immediate from Eichler’s theorem; we proved this more generally for a matrix
ring (17.3.9) using the Hermite normal form.
17.8.7. It is sensible for the class group ClΩ 𝑅 to appear by norm considerations. Let
𝑣 ∈ Ω; then 𝐵 𝑣 ' H, and if 𝛼 ∈ 𝐵× then 𝑣(nrd(𝛼)) > 0, as the reduced norm is
positive.
The class sets of totally definite orders are not captured by Eichler’s theorem, and
for good reason: they can be arbitrarily large, a consequence of the Eichler mass
formula (Chapter 25).
17.8. EICHLER’S THEOREM 285
Exercises
Unless otherwise specified, throughout these exercises let 𝑅 be a Dedekind domain
with field of fractions 𝐹, let 𝐵 be a quaternion algebra over 𝐹, and let O ⊆ 𝐵 be an
𝑅-order.
1. Argue for Proposition 17.2.2 directly in a special case as follows. Let 𝐼 ⊆ M2 (𝐹)
be a lattice with OR (𝐼) = M2 (𝑅).
(a) By considering 𝐼 ⊗𝑅 𝐹 show that
𝐹 𝐹 0 0
𝐼⊆ M2 (𝑅) ⊕ M2 (𝑅).
0 0 𝐹 𝐹
Cls O → Cls O0
[𝐼] ↦→ [𝐼O0]
(a) Show that this statement is local, i.e., the statement is true over 𝑅 if and
only if it is true over 𝑅𝔭 for all primes 𝔭 of 𝑅.
(b) Suppose 𝑅 is a DVR. Show that the statement is true if 𝐵 is a division
algebra.
(c) Suppose 𝑅 is a DVR with maximal ideal 𝔭, and that 𝐵 ' M2 (𝐹). Show that
there is a unique 𝛼 ∈ Or𝔭O such that O0 = 𝛼−1 O𝛼 up to left multiplication
by O× , and conclude that 𝐼 = O𝛼 is the unique integral connecting O, O0
ideal of minimal reduced norm. [Hint: 𝑁GL2 (𝐹 ) (M2 (𝑅)) = 𝐹 × GL2 (𝑅).]
(d) Continuing (c), show that nrd(𝛼) = [O : O ∩ O0]. [Hint: the statement
is equivalent under left or right multiplication of 𝛼 by O× ' GL2 (𝑅), so
consider invariant factors.] [For another perspective, see section 23.5.]
286 CHAPTER 17. CLASSES OF QUATERNION IDEALS
𝐼 0 = 10O0 + (1 − 2𝑖 + 𝑗)O0
𝐼 0 𝑗 (𝐼 0) −1 = 5O0 + (𝑖 + 3 𝑗 + 𝑘)O0
is not principal.
12. The finiteness of the class group (see Reiner [Rei2003, Lemma 26.3]) can be
proven replacing the geometry of numbers with just the pigeonhole principle, as
follows. Let 𝐵 be a division algebra over a number field 𝐹 with ring of integers
𝑅, and let O ⊆ 𝐵 be an 𝑅-order.
(a) To prove the finiteness of Cls O, show that without loss of generality we
may take 𝐹 = Q.
(b) Show that Nm 𝐵 |Q (𝑥1 𝛼1 + · · · + 𝑥 𝑛 𝛼𝑛 ) ∈ Q[𝑥 1 , . . . , 𝑥 𝑛 ] is a homogeneous
polynomial of degree 𝑛.
(c) Show that there exists 𝐶 ∈ Z>0 such that for all 𝑡 > 0 and all 𝑥 ∈ Z𝑛 with
|𝑥𝑖 | ≤ 𝑡, we have |Nm 𝐵 |𝐹 (𝑥1 𝛼1 + · · · + 𝑥 𝑛 𝛼𝑛 )| ≤ 𝐶𝑡 𝑛 .
(d) Let 𝐼 ⊆ O be a lattice. Let 𝑠 ∈ Z be such that
𝑠 𝑛 ≤ N (𝐼) = #(O/𝐼) ≤ (𝑠 + 1) 𝑛 .
Í
Using the pigeonhole principle, show that there exists 𝛼 = 𝑖 𝑥 𝑖 𝛼𝑖 ∈ 𝐼
with 𝑥 𝑖 ∈ Z and |𝑥𝑖 | ≤ 2(𝑠 + 1) for all 𝑖.
(e) Show that N (𝛼O) ≤ 2𝑛 (𝑠 + 1) 𝑛 𝐶, and conclude that
#(𝐼/𝛼O) ≤ 4𝑛 𝐶.
𝑀O ⊆ 𝐼 0 ⊆ O
288 CHAPTER 17. CLASSES OF QUATERNION IDEALS
In this chapter, we treat maximal orders like noncommutative Dedekind domains, and
we consider the structure of two-sided ideals (and their classes), in a manner parallel
to the commutative case.
𝐼𝐽 ⊆ 𝑃 ⇒ 𝐼 ⊆ 𝑃 or 𝐽 ⊆ 𝑃.
289
290 CHAPTER 18. PICARD GROUP
Theorem 18.1.2. Let 𝑅 be a Dedekind domain with field of fractions 𝐹 = Frac 𝑅, let
𝐵 be a simple 𝐹-algebra and let O ⊆ 𝐵 be a maximal 𝑅-order. Then the following
statements hold.
Let Idl(O) be the group of invertible two-sided fractional O-ideals. Put another
way, Theorem 18.1.2 says that if O is maximal, then Idl(O) is isomorphic to the free
abelian group on the set of nonzero prime two-sided O-ideals under multiplication.
We now consider classes of two-sided ideals, in the spirit of section 17.1. Two
candidates present themselves. On the one hand, inside the group Idl(O) of invertible
fractional two-sided O-ideals, the principal fractional two-sided O-ideals (those of
the form O𝛼O = O𝛼 = 𝛼O for certain 𝛼 ∈ 𝐵× ) form a subgroup PIdl(O), and we
could consider the quotient. On the other hand, for a commutative ring 𝑆, the Picard
group Pic(𝑆) is defined to be the group of isomorphism classes of rank one projective
(equivalently, invertible) 𝑆-modules under the tensor product. When 𝑆 is a Dedekind
domain, there is a canonical isomorphism Cl 𝑆 Pic(𝑆).
For simplicity, suppose now that 𝑅 = Z. In this noncommutative setting, we
analogously define Pic O to be the group of isomorphism classes of invertible O-
bimodules (over Z) under tensor product. If 𝐼, 𝐽 ∈ Idl(O), then 𝐼, 𝐽 are isomorphic as
O-bimodules if and only if 𝐽 = 𝑎𝐼 with 𝑎 ∈ Q× , and this yields an isomorphism
Let
𝑁 𝐵× (O) = {𝛼 ∈ 𝐵× : 𝛼O = O𝛼}
be the normalizer of O in 𝐵. By the Skolem–Noether theorem,
generated by (unique) prime two-sided O-ideals with reduced norm 𝑝 | 𝐷, and there
is an exact sequence
Cls O → Typ O
[𝐼] R ↦→ class of OL (𝐼)
We recall that this map is surjective. The fibers are given by Theorem 18.1.3 (see
Proposition 18.5.10): the fiber above the isomorphism class of O0 is in bijection with
the set PIdl(O0)\ Idl(O0).
Remark 18.1.5. The structure of Pic O is more complicated when O is not necessarily
a maximal order: in general, the group Pic O is finite but it may be nonabelian (see
Exercise 18.6); worse still, in general the subgroup PIdl(O) may not be a normal
subgroup in Idl(O).
𝐼𝐽 ⊆ 𝑃 ⇒ 𝐼 ⊆ 𝑃 or 𝐽 ⊆ 𝑃.
Example 18.2.3. The zero ideal 𝑃 = {0} is prime: see Exercise 18.2.
Proposition 18.2.7.
Proof. We follow Reiner [Rei2003, Theorem 22.3]. The implication (⇒) in (a) follows
from Lemma 18.2.6. Conversely, let 𝑃 be a nonzero prime two-sided O-ideal, and let
𝔭 = 𝑃 ∩ 𝑅. We show 𝔭 is a nonzero prime. By 18.2.1, 𝑃 is an 𝑅-lattice, so 𝔭 ≠ {0};
since 1 ∉ 𝑃, we have 𝔭 ≠ 𝑅 and 𝔭 is nontrivial. If 𝑎, 𝑏 ∈ 𝑅, then 𝑎𝑏 ∈ 𝔭 implies
(𝑎O) (𝑏O) ⊆ 𝑃; since 𝑃 is prime, we have 𝑎O ⊆ 𝑃 or 𝑏O ⊆ 𝑃, so 𝑎 ∈ 𝔭 or 𝑏 ∈ 𝔭.
Now let 𝐽/𝑃 = rad(O/𝑃) be the Jacobson radical of O/𝑃 (see section 7.4). By
Lemma 7.4.8, the ideal 𝐽/𝑃 is nilpotent; by (18.2.5), we conclude 𝐽/𝑃 = {0}. Thus
O/𝑃 is semisimple by Lemma 7.4.2 and thus is a product of simple 𝑅/𝔭-algebras by
the Wedderburn–Artin theorem (Main Theorem 7.3.10). But the simple components
of O/𝑃 are two-sided ideals that annihilate one another; again by (18.2.5), there can
be only one component, and O/𝑃 is simple. Thus O/𝑃 has no nontrivial ideals, and
𝑃 is maximal.
Proof. If not, then the set of ideals which do not contain such products is nonempty;
since O is noetherian, there is a maximal element 𝑀. Since 𝑀 cannot itself be prime,
there exist ideals 𝐼, 𝐽, properly containing 𝑀, such that 𝐼𝐽 ⊆ 𝑀. But both 𝐼, 𝐽 contain
products of prime ideals, so the same is true of 𝑀, a contradiction.
18.3 Invertibility
We now consider invertibility first in the general context of orders, then for maximal
orders. The general theory of maximal orders over Dedekind domains in simple
algebras was laid out by Auslander–Goldman [AG60]. One of the highlights of this
theory are the classification of such orders: they are endomorphism rings of a finitely
generated projective module over a maximal order in a division algebra. For a quite
general treatment of maximal orders, see the book by Reiner [Rei2003]; in particular,
the ideal theory presented here is also discussed in Reiner [Rei2003, §§22–23].
Lemma 18.3.1. Let 𝐽 be a two-sided O-ideal, not necessarily invertible. If 𝐽 ( O,
then 𝐽 −1 ) O.
18.3. INVERTIBILITY 293
Proof. The 𝑅-lattice 𝐽 −1 has 𝐽 −1 ⊇ O and OL (𝐽 −1 ) ⊆ OR (𝐽) = O and the same result
holds interchanging left and right.
We follow Reiner [Rei2003, Lemma 23.4] (who calls the proof “mystifying”).
Assume for the purposes of contradiction that 𝐽 −1 = O. Since 𝐽 ( O, there exists a
maximal two-sided O-ideal 𝑀 ⊇ 𝐽. Thus 𝑀 −1 ⊆ 𝐽 −1 = O. By Lemma 18.2.6, 𝑀
is prime. Let 𝑎 ∈ 𝑅 ∩ 𝐽 −1 be nonzero. By Lemma 18.2.8, 𝑎O contains a product of
prime two-sided O-ideals, so
𝑀 ⊇ 𝑎O ⊇ 𝑃1 𝑃2 · · · 𝑃𝑟 ,
with each 𝑃𝑖 prime. We may suppose without loss of generality that 𝑟 ∈ Z>0 is minimal
with this property. Since 𝑃1 · · · 𝑃𝑟 ⊆ 𝑀 and 𝑀 is prime, we must have 𝑃𝑖 ⊆ 𝑀, so
𝑃𝑖 = 𝑀 by Proposition 18.2.7. Thus
𝑀 ⊇ 𝑎O ⊇ 𝐽1 𝑀 𝐽2
Using this lemma, we arrive at the following proposition for maximal orders.
Of course, one can also swap left for right in the statement of Proposition 18.3.2.
Using the standard involution, we proved Proposition 18.3.2 when 𝐵 is a quaternion
algebra (Proposition 16.6.15(b)).
𝐽𝐽 −1 = 𝐼 𝐼 −1 𝐽 −1 ⊆ O,
OR (𝐼 −1 ) ⊇ OL (𝐼) = O; (18.3.3)
Proof. For (b), invertibility follows from Proposition 18.3.2. For (b) without unique-
ness, assume for purposes of contradiction that there is a two-sided ideal of O that is
not the product of prime ideals; then there is a maximal counterexample 𝐽. Since 𝐽 is
not prime, there exists a prime 𝑄 with 𝐽 ( 𝑄 ( O, so 𝐽 ⊂ 𝐽𝑄 −1 ( O. If 𝐽 = 𝐽𝑄 −1 ,
so by cancelling 𝑄 = O, a contradiction. Therefore 𝐽𝑄 −1 = 𝑃1 · · · 𝑃𝑟 is the product of
primes by maximality, and 𝐽 = 𝑃1 · · · 𝑃𝑟 𝑄 is the product of primes, a contradiction.
We now prove (a). If 𝑃, 𝑄 ⊆ O are distinct nonzero prime two-sided ideals, and
we let 𝑄 0 = 𝑃−1 𝑄𝑃, then 𝑄 0 ⊆ 𝑃−1 O𝑃 = O is prime and 𝑃𝑄 0 = 𝑄𝑃 ⊆ 𝑄, so 𝑃 ⊆ 𝑄
or 𝑄 0 ⊆ 𝑄; but equality would hold in each case by maximality, and since 𝑃 ≠ 𝑄, we
must have 𝑄 0 = 𝑄, and multiplication is commutative.
Finally, uniqueness of the factorization in (b) follows as in the commutative case.
If 𝑃1 · · · 𝑃𝑟 = 𝑄 1 · · · 𝑄 𝑠 , then 𝑃1 = 𝑄 𝑖 for some 𝑖; multiplying by 𝑃1−1 and repeating
the argument, we find that {𝑃1 , . . . , 𝑃𝑟 } = {𝑄 1 , . . . , 𝑄 𝑠 }, and the result follows.
Corollary 18.3.5. With hypotheses as in Theorem 18.3.4, the group Idl(O) is isomor-
phic to the free abelian group on the set of nonzero prime ideals.
With these arguments in hand, we have the following foundational result for quater-
nion orders.
Theorem 18.3.6. Suppose that 𝑅 is a Dedekind domain. Let 𝐵 be a quaternion algebra
over 𝐹 and let O ⊆ 𝐵 be a maximal 𝑅-order. Then the map
is a bijection.
Moreover, if 𝑅 is a global ring, then there is an exact sequence
Ö
0 → Idl(𝑅) → Idl(O) → Z/2Z → 0
𝔭 |𝔇 (18.3.8)
𝔞 ↦→ O𝔞O
Proof. The map (18.3.7) is defined by Proposition 18.2.7, and it is surjective because
𝔭O ⊆ 𝑃 is contained in a maximal therefore prime ideal.
Next we show that the map is injective. Let 𝑃 be a prime ideal, and work with
completions at a prime 𝔭. Then 𝑃𝔭 = 𝑃 ⊗𝑅 𝑅𝔭 ⊆ O𝔭 is a maximal ideal of O𝔭 . If
𝐵𝔭 ' M2 (𝐹𝔭 ), so O𝔭 ' M2 (𝑅𝔭 ), then the only maximal two-sided ideal is 𝔭O𝔭 ; if
instead 𝐵𝔭 is a division algebra, then there is a unique maximal two-sided ideal 𝑃𝔭
with 𝑃𝔭2 = 𝔭O𝔭 by Theorem 13.3.11. We can also describe this uniformly, by the proof
of Proposition 18.2.7: in all cases, we have 𝑃𝔭 = rad(O𝔭 ).
18.4. PICARD GROUP 295
Idl(𝑅) → Idl(O)
𝔞 ↦→ O𝔞O = 𝔞O
Remark 18.3.9. Many of the theorems stated in this section (and chapter) hold more
generally for hereditary orders: this notion is pursued in Chapter 21. To see what this
looks like in a more general context, see Curtis–Reiner [CR81, §26B]. A very general
context in which one can make an argument like in section 18.3 was axiomatized by
Asano; for an exposition and several references, see McConnell–Robson [McCR87,
Chapter 5].
Definition 18.4.1. The Picard group of O over 𝑅 is the group Pic𝑅 (O) of isomorphism
classes of invertible O-bimodules over 𝑅 under tensor product.
Remark 18.4.2. Some authors also write Picent(O) = Pic 𝑍 (O) (O) when considering
the Picard group over the center of O, the most important case. To avoid additional
complication, in this section we suppose that 𝐵 is central over 𝐹, so Pic𝑅 (O) =
Picent(O).
we claim that the kernel of this map is PIdl(𝑅) E Idl(O). By Lemma 19.5.1, every
isomorphism class of invertible O-bimodule is represented by an invertible 𝑅-lattice
𝐼 ⊆ 𝐵, unique up to scaling by 𝐹 × , and if 𝑎 ∈ 𝐹 × then 𝑎O = O if and only if
296 CHAPTER 18. PICARD GROUP
is exact. One might profitably take (18.4.6) as the definition of Pic𝑅 (O).
18.4.7. If O0 is locally isomorphic O (so they are in the same genus), then there is a
O, O0-connecting ideal 𝐽, and the map
Idl(O) → Idl(O0)
𝐼 ↦→ 𝐽 −1 𝐼𝐽
Although this sequence need not split, it does show that the Picard group of the maximal
order O is not far from the class group Cl 𝑅, the difference precisely measured by the
primes that ramify in 𝐵.
In general, for a quaternion 𝑅-order O we have the following result.
Proposition 18.4.10. Pic𝑅 (O) is a finite group.
Proof. If O is maximal, we combine (18.4.9) with the finiteness of Cl 𝑅 and the fact
that there are only finitely many primes 𝔭 dividing the discriminant 𝔇.
Now let O be an 𝑅-order. Then there exists a maximal 𝑅-order O0 ⊇ O. We argue
as in Exercise 17.3. We define a map of sets:
By the first paragraph, by finiteness of Pic𝑅 (O0), after rescaling we may suppose 𝐼 0
is one of finitely many possibilities. But there exists nonzero 𝑟 ∈ 𝑅 such that 𝑟O0 ⊂ O,
so
𝐼 0 = O0 𝐼O0 ⊆ (𝑟 −1 O)𝐼 (𝑟 −1 O) = 𝑟 −2 𝐼 ⊆ 𝑟 −2 𝐼 0
so 𝑟 2 𝐼 0 ⊆ 𝐼 ⊆ 𝐼 0; since 𝐼 0/𝑟 2 𝐼 0 is a finite group, this leaves only finitely many
possibilities for 𝐼.
Remark 18.4.11. The study of the Picard group is quite general. It was studied in
detail by Fröhlich [Frö73]; see also Curtis–Reiner [CR87, §55].
Remark 18.5.6. The moral of Proposition 18.5.3 is that, unlike the commutative
case where the two notions coincide, the two notions of “isomorphism classes of
invertible bimodules” and “ideals modulo principal ideals” are in general different for
a quaternion order. These notions coincide precisely when 𝑁 𝐵× (O)/𝐹 × ' O× /𝑅 × , or
equivalently (by the Skolem–Noether theorem) that every 𝑅-algebra automorphism of
O is inner, which is to say Aut 𝑅 (O) = Inn𝑅 (O) = O× /𝑅 × .
18.5.7. Unfortunately, the subgroup PIdl(O) ≤ Idl(O) need not be normal in general
(Exercise 17.11), so statements like Proposition 18.5.3 depend on the order O having
good structural properties. If O is a maximal order, then Idl(O) is abelian, so the result
holds in this case.
In general, from the proof but using cosets one still obtains the equality
Remark 18.5.9. If O, O0 are connected, then Pic𝑅 (O) ' Pic𝑅 (O0) by 18.4.7 but this
isomorphism need not respect the exact sequence (18.5.5). Each order O “balances” the
contribution of this group between the normalizer 𝑁 𝐵× (O)/(𝐹 × O× ) and the quotient
Idl(O)/PIdl(O)—and these might be of different sizes for O0. We will return to
examine more closely this structure in section 28.9, when strong approximation allows
us to be more precise in measuring the discrepancy.
We conclude with an application to the structure of (right) class sets. We examine
from Lemma 17.4.13 the fibers of the surjective map (17.4.14)
Cls O → Typ O
[𝐼] ↦→ class of OL (𝐼).
Refreshing our notation, let 𝐵 be a central simple 𝐹-algebra and let O ⊂ 𝐵 an 𝑅-order.
Proposition 18.5.10. The map 𝐼 ↦→ [𝐼] induces a bijection
Proof. Let O0 be an order of the same type as O. Since (17.4.14) is surjective, there
exists [𝐼] ∈ Cls O such that OL (𝐼) ' O0. We are free to replace O0 by an isomorphic
order, so we may suppose OL (𝐼) = O0. For all [𝐼 0] ∈ Cls O with OL (𝐼 0) ' O0 (running
over the fiber), since OL (𝛼𝐼 0) = 𝛼O0 𝛼−1 for 𝛼 ∈ 𝐵× we may suppose without loss of
generality that the representative 𝐼 0 has OL (𝐼 0) = O0.
We then define a map
Proof. For the first equality, combine Lemma 17.4.13 and Proposition 18.5.10, com-
puting the size of the fibers. For the second, substitute (18.5.8) and use 18.4.7.
Proof. We choose representatives and take the fibers of the map (17.4.14).
Remark 18.5.14. When PIdl(O) E Idl(O), then in Proposition 18.5.10 we have written
the class set Cls O as a disjoint union of abelian groups. The fact that the bijection is
noncanonical is due to the fact that we choose a connecting ideal, so without making
choices we obtain only a disjoint union of principal homogeneous spaces (i.e., torsors)
under the groups PIdl(O0)\ Idl(O0).
Exercises
Unless otherwise specified, let 𝑅 be a Dedekind domain with field of fractions 𝐹 =
Frac 𝑅, let 𝐵 be a simple finite-dimensional 𝐹-algebra, and let O ⊆ 𝐵 be an 𝑅-order.
(d) Prove that [𝑃2 ] ∈ Pic𝑅 (O) has order 4, and conclude that the sequence
(18.4.9) does not split.
(e) Show that we may take
√ √
O = 𝑅 + 𝔭−1 −1
2 ( −6 + 𝑖) + 𝑅 𝑗 + 𝔭2 ( −6 + 𝑖) 𝑗
√
as the maximal order,
√ and then that 𝐼 is generated by 𝑖 and −6𝑖 𝑗/2, and
finally that 𝐼 2 = ( −6 + 𝑖)/2.
6. Let 𝐵 = M𝑛 (𝐹) with 𝑛 ≥ 2, let O = M𝑛 (𝑅), let 𝔭 ⊆ 𝑅 be prime with 𝑘 = 𝑅/𝔭,
and let O(𝔭) = 𝑅 + 𝔭O.
(a) Show that O(𝔭) is an order of reduced discriminant 𝔭3 .
(b) Show that O× ' GL𝑛 (𝑅) normalizes O(𝔭) ⊆ O, so that
O(𝔭) × E O× ' GL𝑛 (𝑅),
and that the map
O× ↩→ Idl(O)
𝛾 ↦→ O𝛾 = 𝛾O
induces an injective group homomorphism PGL𝑛 (𝑘) ↩→ Idl(O). Con-
clude that Idl(O) is not an abelian group.
7. Show that Theorem 18.3.4 holds more generally for 𝐵 a semisimple 𝐹-algebra
(but still O ⊆ 𝐵 maximal). [Hint: Decompose 𝐵 into a product of simple
𝐹-algebras.]
8. Let O be maximal, and let 𝑃1 , . . . , 𝑃𝑟 ⊆ O be distinct prime two-sided ideals.
Let
Ö𝑟 Ö 𝑟
𝑓
𝐼 := 𝑃𝑖𝑒𝑖 and 𝐽 := 𝑃𝑖 𝑖
𝑖=1 𝑖=1
with 𝑒 𝑖 , 𝑓𝑖 ∈ Z.
(a) Prove that 𝐼 ⊆ O if and only if 𝑒 𝑖 ≥ 0 for all 𝑖 = 1, . . . , 𝑛, and in this case
there is a ring isomorphism
𝑟
Ê
O/𝐼 ' O/𝑃𝑖𝑒𝑖 .
𝑖=1
Brandt groupoids
In this chapter, we study the relationship between multiplication and classes of quater-
nion ideals.
The set Cl(𝑑) of SL2 (Z)-classes of forms in Q (𝑑) is finite, by reduction theory (see
section 35.2): every form in Q (𝑑) is equivalent under the action of SL2 (Z) to a unique
reduced form, of which there are only finitely many. To study this finite set, Gauss
defined a composition law on Cl(𝑑), giving Cl(𝑑) the structure of an abelian group
by an explicit formula. Gauss’s composition law on binary quadratic forms can be
understood using 2 × 2 × 2 Rubik’s cubes, by a sublime result of Bhargava [Bha2004a].
Today, we see this composition law as a consequence of a natural bijection between
Cl(𝑑) and a set equipped with an obvious group structure. Let 𝑆 = 𝑆(𝑑) be the quadratic
ring of discriminant 𝑑. Define the narrow class group Cl+ (𝑆) as
301
302 CHAPTER 19. BRANDT GROUPOIDS
modulo
the subgroup of nonzero principal fractional ideals
generated by a totally positive element
(i.e., one that is positive in every embedding into R, so if 𝑑 < 0 then this is no
condition). (Alternatively, Cl+ (𝑆) can be thought of as the group of isomorphism
classes of oriented, invertible 𝑆-modules, under a suitable notion of orientation.) Then
there is a bijection between Cl(𝑑) and Cl+ (𝑆): explicitly, to the class of the quadratic
form 𝑄 = 𝑎𝑥 2 + 𝑏𝑥𝑦 + 𝑐𝑦 2 ∈ Q (𝑑), we associate the class of the ideal
√
−𝑏 + 𝑑
𝔞 = 𝑎Z + Z ⊂ 𝑆(𝑑). (19.1.2)
2
√
Conversely, the quadratic form is recovered as the norm form on 𝐾 = Q( 𝑑) restricted
to 𝔞: √
−𝑏 + 𝑑
Nm𝐾 /Q 𝑎𝑥 + 𝑦 = 𝑎𝑥 2 + 𝑏𝑥𝑦 + 𝑐𝑦 2 (19.1.3)
2
where 𝑐 = (𝑏 2 − 𝑑)/(4𝑎) ∈ Z.
Much of the same structure can be found in the quaternionic case, with several
interesting twists. It was Brandt who first asked if there was a composition law for
(integral, primitive) quaternary quadratic forms: it would arise naturally from some
kind of multiplication of ideals in a quaternion order, with the analogous bijection
furnished by the reduced norm form. Brandt started writing on composition laws
for quaternary quadratic forms in 1913 [Bra13], tracing the notion of composition
back to Hermite, who observed a kind of multiplication law (bilinear substitution) for
quaternary forms 𝑥02 + 𝐹 (𝑥1 , 𝑥2 , 𝑥3 ) in formulas of Euler and Lagrange. He continued
on this note during the 1920s [Bra24, Bra25, Bra28, Bra37], when it became clear that
quaternion algebras was the right framework to place his composition laws; in 1943,
he developed this theme significantly [Bra43] and defined his Brandt matrices (that
will figure prominently in Chapter 41.
However, in the set of invertible lattices in 𝐵 under compatible product, one cannot
always multiply! However, this set has the structure of a groupoid: a nonempty set
with an inverse function and a partial product that satisfies the associativity, inverse, and
identity properties whenever they are defined. Groupoids now figure prominently in
category theory (a groupoid is equivalently a small category in which every morphism
is an isomorphism) and many other contexts; see Remark 19.3.11.
Organizing lattices by their left and right orders, which by definition are connected
and hence in the same genus, we define
Brt(O) = {𝐼 : 𝐼 ⊂ 𝐵 invertible 𝑅-lattice and OL (𝐼), OR (𝐼) ∈ Gen O}; (19.1.4)
visibly, Brt(O) depends only on the genus of O. Organizing lattices according to the
genus of orders is sensible: after all, we only apply the composition law to binary
quadratic forms of the same discriminant, and in the compatible product we see
precisely those classes whose left and right orders are connected.
√ In other words,
the set of invertible lattices in the quadratic field 𝐾 = Q( 𝑑) has the structure of a
groupoid if we multiply only those lattices with the same multiplicator ring.
19.1. ⊲ COMPOSITION LAWS AND IDEAL MULTIPLICATION 303
19.2 Example
−2, −37
Consider the quaternion algebra 𝐵 := with standard basis 1, 𝑖, 𝑗, 𝑘 = 𝑖 𝑗,
Q
and the maximal order O of reduced discriminant 37 defined by
1+𝑖+ 𝑗 2+𝑖+ 𝑘
O := Z + 𝑖Z + Z+ Z. (19.2.1)
2 4
The type set Typ O of orders connected to O has exactly two isomorphism classes,
represented by O1 = O and
3 − 7𝑖 + 𝑗 2 − 3𝑖 + 𝑘
O2 := Z + 3𝑖Z + Z+ Z.
6 4
These orders are connected by the O2 , O1 -connecting ideal
3−𝑖+ 𝑗 2 + 3𝑖 − 𝑘 3−𝑖+ 𝑗
𝐼 := 3Z + 3𝑖Z + Z+ Z = 3O + O.
2 4 2
There are isomorphisms
Pic(O) ' Pic(O2 ) ' Z/2Z
with the nontrivial class in Pic(O1 ) represented by the principal two-sided ideal 𝐽1 =
𝑗O = O 𝑗 with 𝑗 ∈ 𝑁 𝐵× (O), and the nontrivial class in Pic(O2 ) represented by the
nonprincipal (but invertible) ideal
111 − 259𝑖 + 𝑗
𝐽2 := 𝐼𝐽1 𝐼 −1 = 37O2 + O2 .
6
In particular,
ClsR (O1 ) = {[O1 ], [𝐼], [𝐽2 𝐼]} and ClsR (O2 ) = {[O2 ], [𝐽2 ], [𝐼]},
[IJ2 ] = [J1 I]
Figure 19.2.2: BrtCl O, for discrd O = 37
The Brandt class groupoid
The quadratic forms nrd𝐼 and nrd𝐼 are isometric but not by an oriented isometry.
(𝑎 ∗ 𝑏) ∗ 𝑐 = 𝑎 ∗ (𝑏 ∗ 𝑐).
(b) [Inverses] For all 𝑎 ∈ 𝐺, there exists 𝑎 −1 ∈ 𝐺 such that 𝑎 ∗ 𝑎 −1 and 𝑎 −1 ∗ 𝑎 are
defined (but not necessarily equal).
(c) [Identity] For all 𝑎, 𝑏 ∈ 𝐺 such that 𝑎 ∗ 𝑏 is defined, we have
(𝑎 ∗ 𝑏) ∗ 𝑏 −1 = 𝑎 and 𝑎 −1 ∗ (𝑎 ∗ 𝑏) = 𝑏. (19.3.3)
for all 𝑎, 𝑏 ∈ 𝐺.
19.3.4. Let 𝐺 be a groupoid. Then the products in the identity law (19.3.3) are defined
by the associative and inverse laws, and it follows that 𝑒 = 𝑎 ∗ 𝑎 −1 , the left identity of
𝑎, and 𝑓 = 𝑎 −1 ∗ 𝑎 the corresponding right identity of 𝑎, satisfy 𝑒 ∗ 𝑎 = 𝑎 = 𝑎 ∗ 𝑓
306 CHAPTER 19. BRANDT GROUPOIDS
for all 𝑎 ∈ 𝐺. (We may have that 𝑒 ≠ 𝑓 , i.e., the left and right identities for 𝑎 ∈ 𝐺
disagree.) The right identity of 𝑎 ∈ 𝐺 is the left identity of 𝑎 −1 ∈ 𝐺, so we call the set
{𝑒 = 𝑎 ∗ 𝑎 −1 : 𝑎 ∈ 𝐺}
19.3.5. Equivalently, a groupoid is a small category (the class of objects in the category
is a set) such that every morphism is an isomorphism: given a groupoid, we associate
the category whose objects are the elements of the set 𝑆 := {𝑒 = 𝑎 ∗ 𝑎 −1 : 𝑎 ∈ 𝐺} of
identity elements in 𝐺 and the morphisms between 𝑒, 𝑓 ∈ 𝑆 are the elements 𝑎 ∈ 𝐺
such that 𝑒 ∗ 𝑎 and 𝑎 ∗ 𝑓 are defined (see Proposition 19.3.9 below). Conversely, to
a category in which every morphism is an isomorphism, we associate the groupoid
whose underlying set is the union of all morphisms under inverse and composition.
Example 19.3.6. The set of homotopy classes of paths in a topological space 𝑋 forms
a groupoid under composition: the paths 𝛾1 , 𝛾2 : [0, 1] → 𝑋 can be composed to a
path 𝛾2 ◦ 𝛾1 : [0, 1] → 𝑋 if and only if 𝛾2 (0) = 𝛾1 (1).
Example 19.3.7. A disjoint union of groups is a groupoid, with the product defined if
and only if the elements belong to the same group; the set of identities is canonically
in bijection with the index set of the disjoint union.
Proof. The set 𝐺 (𝑒, 𝑒) is nonempty, has the identity element 𝑒 ∈ 𝐺, and if 𝑎 ∈ 𝐺 (𝑒, 𝑒)
then 𝑎∗𝑎 −1 = 𝑎 −1 ∗𝑎 = 𝑒. If 𝑒, 𝑓 are identity elements, since 𝐺 is connected there exists
𝑎 ∈ 𝐺 (𝑒, 𝑓 ), so 𝑎 −1 ∈ 𝐺 ( 𝑓 , 𝑒) and the map 𝐺 (𝑒, 𝑒) → 𝐺 ( 𝑓 , 𝑓 ) by 𝑥 ↦→ 𝑎 ∗ 𝑥 ∗ 𝑎 −1
is an isomophism of groups. Similarly, the set 𝐺 (𝑒, 𝑓 ) has a right, simply transitive
action of 𝐺 (𝑒, 𝑒) under right multiplication by ∗.
19.3.10. The moral of Proposition 19.3.9 is that the only two interesting invariants of
a connected groupoid are the number of identity elements (objects in the category) and
the group of elements with a common left and right identity (the automorphism group
of every one of the objects). A connected groupoid is determined up to isomorphism
of groupoids by these two properties.
Remark 19.3.11. After seeing its relevance in the context of composition of quaternary
forms, Brandt set out general axioms for his notion of a groupoid [Bra27, Bra40].
(Brandt’s original definition of groupoid is now called a connected groupoid.) This
notion has blossomed into an important structure in mathematics that sees quite general
use, especially in homotopy theory and category theory. It is believed that the groupoid
axioms influenced the work of Eilenberg–Mac Lane [EM45] in the first definition of
a category: see e.g., Brown [Bro87] for a survey, Bruck [Bruc71] for context in the
theory of binary structures, as well as the article by Weinstein [Wein96].
Groupoids exhibit many facets of mathematics, arising naturally in functional anal-
ysis (𝐶 ∗ -algebras) and group representations, as Figure 19.3.12 indicates (appearing
in Williams [Will2001, p. 21], and attributed to Arlan Ramsay).
group
actions
groups
sets
equivalence
groupoids relations
Proof. For the associative law, suppose 𝐼, 𝐽, 𝐾 are invertible 𝑅-lattices with 𝐼𝐽 and
(𝐼𝐽)𝐾 compatible products. Then OR (𝐼) = OL (𝐽) = OL (𝐽𝐾) and OR (𝐼𝐽) = OR (𝐽) =
OL (𝐾) by Lemma 16.5.11, so the products 𝐽𝐾 and 𝐼 (𝐽𝐾) are compatible. Multipli-
cation is associative in 𝐵, and it follows that 𝐼 (𝐽𝐾) = (𝐼𝐽)𝐾. Inverses exist exactly
because we restrict to the invertible lattices.
The law of identity holds as follows: if 𝐼, 𝐽 are invertible 𝑅-lattices such that 𝐼𝐽 is
a compatible product, then (𝐼𝐽)𝐽 −1 is a compatible product since OR (𝐼𝐽) = OR (𝐽) =
OL (𝐽 −1 ), and by associativity
Proof. By Proposition 19.4.1, the identity elements correspond to orders, and two
orders are connected if and only if there is a (invertible, equivalently locally principal)
connecting ideal if and only if they are in the same genus, as in section 17.4. The
second statement follows immediately.
As a consequence of Lemma 19.4.2, the subset of 𝑅-lattices whose (left or) right
order belong to a specified genus of orders is a connected subgroupoid.
Definition 19.4.3. Let O ⊆ 𝐵 be an 𝑅-order. The Brandt groupoid of (the genus of)
O is
In the next section, we consider a variant that considers classes of lattices, giving
rise to a finite groupoid.
be the set of homothety classes of 𝑅-lattices in 𝐵 with left order O and right order
O0; equivalently, by Lemma 19.5.1, Pic(O, O0) is the set of isomorphism classes of
invertible O, O0-bimodules over 𝑅. In particular, Pic𝑅 (O) = Pic𝑅 (O, O).
We have Pic𝑅 (O, O0) ≠ ∅ if and only if O is connected to O0.
Let O ⊂ 𝐵 be an order and let O𝑖 be representative orders for the type set Typ O.
We define Ä
BrtCl O := Pic𝑅 (O𝑖 , O 𝑗 ).
𝑖, 𝑗
Theorem 19.5.5. Let 𝑅 be a Dedekind domain with field of fractions 𝐹, and let 𝐵 be
a quaternion algebra over 𝐹. Let O ⊂ 𝐵 be an order. Then the set BrtCl O has the
structure of a finite groupoid that, up to isomorphism, is independent of the choice of
the orders O𝑖 .
In particular, by Theorem 19.5.5 BrtCl O depends only on the genus of O up to
groupoid isomorphism. We call the set BrtCl O the Brandt class groupoid of (the
genus of) O.
310 CHAPTER 19. BRANDT GROUPOIDS
nrd𝐼 : 𝐼 → 𝐿 = nrd(𝐼)
𝐼
nrd 𝐼
/𝐿
o 𝛼·𝛽 o 𝑎𝑏 (19.6.3)
𝐽 = 𝛼𝐼 𝛽
nrd 𝐽
/ 𝑎𝑏𝐿
𝐼
nrd 𝐼
/ 𝐿 = 𝑎𝑅
o 𝑎−1 (19.6.5)
𝑎−1
𝐼
nrd 𝐼
/𝑅
disc(nrd𝐼 ) = disc(O)
Remark 19.6.8. The Brandt groupoid is connected as a groupoid. This can also be
viewed in the language of quadratic forms: a connected class of orders is equivalently
a genus of integral ternary quadratic forms, and this is akin to a resolvent for the
quaternary norm forms. We refer to Chapter 23 for further development.
Exercises
1. Verify the computational details in the example of section 19.2.
2. Let 𝐵 = (−1, −11 | Q) with disc 𝐵 = 11 and O = Zh𝑖, (1 + 𝑗)/2i a maximal
order. Compute BrtCl O, in a manner analogous to the example of section 19.2.
3. Let 𝐺 be a groupoid.
(a) Show that if 𝑎, 𝑏, 𝑐 ∈ 𝐺 and both 𝑎 ∗ 𝑏 and 𝑎 ∗ 𝑐 are defined, then
𝑏 ∗ 𝑏 −1 = 𝑐 ∗ 𝑐−1 (and both are defined).
(b) Show that for all 𝑎 ∈ 𝐺 we have (𝑎 −1 ) −1 .
4. Let 𝐺 be a group acting on a nonempty set 𝑋. Let
𝐴(𝐺, 𝑋) = {(𝑔, 𝑥) : 𝑔 ∈ 𝐺, 𝑥 ∈ 𝑋 }.
Show that 𝐴(𝐺, 𝑋) has a natural groupoid structure with (𝑔, 𝑥) ∗ (ℎ, 𝑦) = (𝑔ℎ, 𝑦)
defined if and only if 𝑥 = ℎ𝑦. What are the identity elements?
5. Show that in a homomorphism 𝜙 : 𝐺 → 𝐺 0 of groupoids, the set of identity
elements of 𝐺 maps to the set of identity elements of 𝐺 0.
6. Let C be a small category. Show that there is a unique maximal subcategory
that is a groupoid. [Hint: Discard all nonisomorphisms.]
7. Let 𝑋 be a set and let ∼ be an equivalence relation on 𝑋, thought of as a
subset 𝑆 ⊆ 𝑋 × 𝑋. Equip 𝑆 with the partial binary operation ∗ defined by
(𝑥, 𝑦) ∗ (𝑦, 𝑧) = (𝑥, 𝑧) for (𝑥, 𝑦), (𝑦, 𝑧) ∈ 𝑆 (and (𝑥, 𝑦) ∗ (𝑤, 𝑧) is not defined if
𝑦 ≠ 𝑤). Show that 𝑆 is a groupoid. [This shows that “equivalence relations are
groupoids”, cf. (19.3.12).]
8. Let 𝐹 be a field and let GL(𝐹) = ∞
Ð
𝑛=1 GL𝑛 (𝐹). Show that GL(𝐹) has a natural
structure of groupoid, sometimes called the general linear groupoid over 𝐹.
9. Show that the reduced norm is a homomorphism from the groupoid of invertible
𝑅-lattices in 𝐵 to the group(oid) of fractional 𝑅-ideals in 𝐹.
10. Let 𝑋 be a nonempty topological space, and let 𝑥, 𝑦 ∈ 𝑋. Recall that a path
from 𝑥 to 𝑦 is a continuous map 𝜐0 : [0, 1] → 𝑋 with 𝜐(0) = 𝑥 and 𝜐(1) = 𝑦.
We say that paths 𝜐0 , 𝜐1 : [0, 1] → 𝑋 from 𝑥 to 𝑦 are homotopic if there exists
a continuous map 𝐻 : [0, 1] × [0, 1] → 𝑋 such that 𝐻 (0, 𝑠) = 𝑥 and 𝐻 (1, 𝑠) = 𝑦
for all 𝑠 ∈ [0, 1] and 𝐻 (𝑡, 0) = 𝜐0 and 𝐻 (𝑡, 1) = 𝜐1 (𝑡) for all 𝑡 ∈ [0, 1]. [So
each 𝐻 (𝑡, 𝑠) for fixed 𝑡 ∈ [0, 1] is a path from 𝑥 to 𝑦, and this set of paths varies
continuously.]
(a) Check that being homotopic defines an equivalence relation on the set of
continuous paths from 𝑥 to 𝑦.
19.6. QUADRATIC FORMS 313
(b) Check that paths can be composed (going at twice speed) and that compo-
sition of paths is well-defined on homotopy classes.
(c) Show that composition of homotopy classes of continuous paths is asso-
ciative.
Let Π(𝑋) be the category whose objects are the points of 𝑋 and with mor-
phisms to be the set of homotopy classes of continuous paths from 𝑥 to 𝑦 under
composition.
(d) Show that Π(𝑋) is a category.
(e) Show that Π(𝑋) is a groupoid, called the fundamental groupoid of 𝑋.
(f) Finally, for all 𝑥 ∈ 𝑋, show the set of all morphisms from 𝑥 to 𝑥 in Π(𝑋) is
a group (the more familiar fundamental group 𝜋1 (𝑋, 𝑥) with base point
𝑥).
11. Continuing the previous exercise, show that if 𝑋 is path-connected, then Π(𝑋)
is equivalent as a category to a groupoid with one object. [Hint: choose a point
𝑥 ∈ 𝑋, look at the group(oid) 𝜋1 (𝑋, 𝑥).]
Chapter 20
One can also tease apart left and right invertibility if desired; in the quaternion
context, these are equivalent anyway because of the standard involution (Main Theorem
20.3.9).
Given our efforts to understand invertible lattices, one may think that Theorem
20.1.1 is all there is to say. However, two issues remain. First, there may be finitely
generated (projective) O-modules that are not lattices, and they play a structurally
important role for the order O. Second, and this point is subtle: there may be lattices
𝐼 ⊆ 𝐵 that are projective as a left O-module, but with OL (𝐼) ) O; in other words, such
lattices are invertible over a larger order, even though they still have good properties
as modules over the smaller order.
315
316 CHAPTER 20. INTEGRAL REPRESENTATION THEORY
The two left O-modules 𝐼1 , 𝐼2 are visibly projective, and they are not isomorphic:
intuitively, an isomorphism would have to be multiplication on the left by a 2×2-matrix
that commutes with multiplication O, and so it must be scalar. More precisely, suppose
𝜙 ∈ HomO (𝐼1 , 𝐼2 ) is an isomorphism of left O-modules. Extending scalars, we have
Q
Q𝐼1 = Q𝐼2 = =: 𝑉,
Q
and the extension of 𝜙 gives an element in Aut 𝐵 (𝑉) where 𝐵 = M2 (Q) = EndQ (𝑉),
so commutes with the action of 𝐵 and is therefore central: which is to say 𝜙 is a scalar
matrix, and that is absurd.
The lattice 𝐼 = M2 (Z) is invertible as lattice, since it is an order (!); and it is a
two-sided fractional O-ideal, but it is not sated. We claim that 𝐼 is also a projective
O-module: this follows from the fact that M2 (Z) ' 𝐼2⊕2 as a left O-module, so M2 (Z)
is isomorphic to a direct summand of O ⊕2 .
In this chapter, we establish some basic vocabulary of modules in the language of
the representation theory of an order. In the case of algebras over a field, we defined
a Jacobson radical as a way to measure the failure of the algebra to be semisimple.
Similarly, for every ring 𝐴, we define the Jacobson radical rad 𝐴 to be the intersection
of all maximal left ideals of 𝐴: it again measures the failure of left indecomposable
modules to be simple. There is a left-right symmetry to rad 𝐴, and in fact rad 𝐴 ⊆ 𝐴
is a two-sided 𝐴-ideal.
Locally, the Jacobson radical plays a key role. Suppose 𝑅 is a complete DVR with
unique maximal ideal 𝔭. Then 𝔭 = rad O since it is the maximal ideal. Moreover, we
will see that 𝔭O ⊆ rad O, so O/rad O is a finite-dimensional semisimple 𝑘-algebra.
Much of the structure of O-modules is reflected in the structure of modules over the
quotient O/rad O (see Lemma 20.6.8).
Remark 20.1.4. In representation theory, generally speaking, to study the action of a
group on some kind of object (vector space, simplicial complex, etc.) one introduces
some kind of group ring and studies modules over this ring. The major task becomes
to classify such modules. For example, let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅,
and let O be an 𝑅-order in a finite-dimensional 𝐹-algebra 𝐵. A finitely generated
integral representation of O is a finitely generated O-module that is projective as
20.2. PROJECTIVE MODULES 317
Proposition 20.2.2. Let 𝑃 be a finitely generated left 𝐴-module. Then the following
are equivalent:
(i) 𝑃 is projective;
(ii) There exists a finitely generated left 𝐴-module 𝑄 such that 𝑃 ⊕ 𝑄 is free as a left
𝐴-module.
(iii) Every surjective homomorphism 𝑓 : 𝑀 → 𝑃 (of left 𝐴-modules) has a splitting
𝑔 : 𝑃 → 𝑀 (i.e., 𝑓 ◦ 𝑔 = id 𝑃 );
(iv) Every diagram
𝑃
𝑞
𝑝
~
𝑀
𝑓
/𝑁 /0
of left 𝐴-modules with exact bottom row can be extended as indicated, with
𝑝 = 𝑓 ◦ 𝑞; and
(v) Hom 𝐴 (𝑃, −) is a (right) exact functor.
Proof. See Lam [Lam99, Chapter 2]. In statement (v), given a short exact sequence
0→𝑄→𝑀 →𝑁 →0
is exact; the condition for 𝑃 to be projective is that Hom 𝐴 (𝑃, −) is right exact, so the
full sequence
is short exact.
É
20.2.5. A finite direct sum 𝑃 = 𝑖 𝑃𝑖 of finitely generated 𝐴-modules is projective
if and only if each summand
Î 𝑃 𝑖 is projective: indeed, the functor Hom𝑅 (𝑃, −) is
naturally isomorphic to 𝑖 Hom𝑅 (𝑃𝑖 , −), so we apply condition (v) of Proposition
20.2.2.
20.2.6. Localizing Proposition 20.2.2(v), and using the fact that a sequence is exact
if and only if it is exact locally (Exercise 20.1(a)), we see that 𝑃 is projective as a left
O-module if and only if 𝑃 (𝔭) is projective as a left O (𝔭) -module for all primes 𝔭 ⊆ 𝑅
Definition 20.2.7. A left O-lattice is an 𝑅-lattice 𝑀 that is a left O-module.
We make a similar definition on the right.
20.2.8. A left O-lattice 𝑀 is locally free of rank 𝑟 ≥ 1 if 𝑀𝔭 ' O𝔭⊕𝑟 as left O-modules
for all primes 𝔭 ⊆ 𝑅. If follows from 20.2.5 and 20.2.6 that a locally free O-lattice is
projective.
𝑔:𝐼→𝑀
∑︁
𝛽 ↦→ 𝛽𝛼𝑖∗ 𝑒 𝑖 ;
𝑖
the map 𝑔 is defined because for all 𝛽 ∈ 𝐼, we have 𝛽𝛼𝑖∗ ∈ 𝐼 𝐼 −1 , and as always 𝐼 𝐼 −1 𝐼 ⊆ 𝐼
so 𝐼 𝐼 −1 ⊆ OL (𝐼). The map 𝑔 is a splitting of 𝑓 since
∑︁ ∑︁
( 𝑓 ◦ 𝑔) (𝛽) = 𝛽𝛼𝑖∗ 𝛼𝑖 = 𝛽 𝛼𝑖∗ 𝛼𝑖 = 𝛽.
𝑖 𝑖
Remark 20.3.6. The proof of Theorem 20.3.3 follows what is sometimes called the
dual basis lemma for a projective module: see Lam [Lam99, (2.9)], Curtis–Reiner
[CR81, (3.46)], or Faddeev [Fad65, Proposition 18.2].
320 CHAPTER 20. INTEGRAL REPRESENTATION THEORY
(ii) 𝐼 is invertible;
(iii) 𝐼 is left invertible;
(iii0) 𝐼 is right invertible;
(v) 𝐼 is projective as a left OL (𝐼)-module; and
(v0) 𝐼 is projective as a right OR (𝐼)-module.
Proof. The equivalences (ii) ⇔ (ii0) ⇔ (iii) are from Main Theorem 16.7.7 (proven in
Lemma 16.7.5). Theorem 20.3.3(a) gives (v) ⇒ (iii) and (v0) ⇒ (iii0), and Theorem
20.3.3(b) gives (ii) ⇒ (v), (v0).
Example 20.3.10. Consider again Example 16.5.12. The lattice 𝐼 has OL (𝐼) =
OR (𝐼) = O (so has the structure of a sated O, O-bimodule) but 𝐼 is not invertible; from
Main Theorem 20.3.9, it follows that 𝐼 is not projective as a left or right O-module.
𝐼 = ann(𝑥) := {𝛼 ∈ 𝐴 : 𝛼𝑥 = 0}.
Definition 20.4.3. The Jacobson radical rad 𝐴 is the intersection of all maximal left
ideals of 𝐴. The ring 𝐴 is Jacobson semisimple if rad 𝐴 = {0}.
Lemma 20.4.4. The Jacobson radical rad 𝐴 is the intersection of all annihilators of
simple left 𝐴-modules; rad 𝐴 ⊆ 𝐴 is a two-sided 𝐴-ideal.
Proof. The same proof as in Lemma 7.4.5 and Corollary 7.4.6 applies, mutatis mu-
tandis.
Example 20.4.5. If 𝐴 is a commutative local ring, then rad 𝐴 is the unique maximal
ideal of 𝐴.
Example 20.4.6. Let 𝑅 be a complete DVR with maximal ideal 𝔭 = rad 𝑅. Let
𝐹 = Frac 𝑅 and let 𝐷 be a division algebra over 𝐹. Let O ⊆ 𝐷 be the valuation ring,
the unique maximal 𝑅-order (Proposition 13.3.4). Then O has a unique two-sided
ideal 𝑃 by 13.3.10, and so rad O = 𝑃.
Lemma 20.4.7. 𝐴/rad 𝐴 is Jacobson semisimple.
Proof. Let 𝐽 = rad 𝐴. Since 𝐽 𝑀 = {0} for each simple left 𝐴-module 𝑀, we may view
each such 𝑀 as a simple left 𝐴/𝐽-module. Now let 𝛼 ∈ 𝐴 be such that 𝛼+𝐽 ∈ rad( 𝐴/𝐽);
then (𝛼 + 𝐽) 𝑀 = {0}, so 𝛼𝑀 = {0} and 𝛼 ∈ 𝐽; thus rad( 𝐴/𝐽) = {0}, and 𝐴/𝐽 is
Jacobson semisimple.
Proof. Let 𝛽 ∈ rad 𝐴, let 𝛼10 , 𝛼20 ∈ 𝐴 0; since 𝜙 is surjective, there exist preimages
𝛼1 , 𝛼2 ∈ 𝐴. By Lemma 20.4.8, 1 − 𝛼1 𝛽𝛼2 ∈ 𝐴× and
Proof. We have a surjection 𝜙 : 𝐴 → 𝐴/𝐼. For (a), we get 𝜙(rad 𝐴) ⊆ rad( 𝐴/𝐼) = {0}
from Corollary 20.4.10, so rad 𝐴 ⊆ 𝐼. For (b), we get rad( 𝐴)/𝐼 ⊆ rad( 𝐴/𝐼) from the
surjection, and applying (a) to ( 𝐴/𝐼)/(rad( 𝐴)/𝐼) we get rad( 𝐴/𝐼) ⊆ rad( 𝐴)/𝐼.
𝑥 1 = 𝛽1 𝑥 1 + · · · + 𝛽 𝑛 𝑥 𝑛
Proof. If 𝐼 does not contain rad 𝐴, then 𝐼 + rad 𝐴 is a two-sided ideal of 𝐴 containing
rad 𝐴 and properly containing 𝐼. Since 𝐼 is maximal, we have 𝐼 + rad 𝐴 = 𝐴. By (the
corollary to) Nakayama’s lemma, we get 𝐼 = 𝐴, a contradiction.
Remark 20.5.4. The order O as a free 𝑅-module has a natural topology induced from
the 𝔭-adic topology on 𝑅; 𝐽 is topologically nilpotent if and only if 𝐽 𝑟 → {0} in this
topology.
Corollary 20.5.5. Let 𝐼 ⊆ O be a two-sided ideal. Then the following are equivalent:
(a) 𝐼 ⊆ rad O;
(b) 𝐼 𝑟 ⊆ rad O for some 𝑟 > 0; and
(c) 𝐼 is topologically nilpotent.
0 0
É𝑛
Proof. By hypothesis, we can write 𝑖=1 𝑀𝑖 = 𝑁 ⊕ 𝑁 , with 𝑁 a finitely generated
left O-module. By the Krull–Schmidt theorem (Theorem 20.6.2), if we write 𝑁, 𝑁 0 as
the direct sums of indecomposable modules, the conclusion follows.
324 CHAPTER 20. INTEGRAL REPRESENTATION THEORY
20.6.4. We saw in 7.2.19 that idempotents govern the decomposition of the 𝐹-algebra 𝐵
into indecomposable left 𝐵-modules. The same argument shows that a decomposition
O = 𝑃 1 ⊕ · · · ⊕ 𝑃𝑟 (20.6.5)
into a direct sum of indecomposable left O-modules corresponds to an idempotent
decomposition 1 = 𝑒 1 + · · · + 𝑒𝑟 , with the 𝑒 𝑖 a complete set of primitive orthogonal
idempotents. Moreover, each 𝑃𝑖 = O𝑒 𝑖 is a projective indecomposable left O-module.
Conversely, if 𝑃 is a projective indecomposable finitely generated left O-module,
then 𝑃 ' 𝑃𝑖 for some 𝑖: taking a set of generators we have a surjective O-module
homomorphism O𝑟 → 𝑃, and since 𝑃 is projective we have 𝑃 ⊆ O𝑟 a direct summand,
so Corollary 20.6.3 applies.
Consequently, if 𝑃 is a projective left O-lattice, then 𝑃 ' 𝑃1⊕𝑛1 ⊕ · · · ⊕ 𝑃𝑟⊕𝑛𝑟 with
𝑛𝑖 ≥ 0 for 𝑖 = 1, . . . , 𝑟.
20.6.6. The decomposition of an order into projective indecomposables is a nice way
to keep track of other orders, as follows. We extend our notation slightly, and define
OL (𝑀) := {𝛼 ∈ 𝐵 : 𝛼𝑀 ⊆ 𝑀 }
Proof. The proof requires a bit of fiddling with idempotents, but is otherwise straight-
forward—so it makes a good exercise (Exercise 20.7).
20.7. ∗ STABLE CLASS GROUP AND CANCELLATION 325
We finish our local study over 𝑅 a complete DVR with composition series for
modules over an order.
20.6.11. If 𝑀 has a finite composition series, then its length ℓ(𝑀) is well-defined,
independent of the series. For example, taking 𝑅 = 𝐹 and O = 𝐵, a finitely generated
𝐵-module is a finite-dimensional 𝐹-vector space, so every composition series is finite
and every 𝐵-module 𝑉 has a well-defined length ℓ(𝑉).
20.7.1. Recall that the group Cl 𝑅 records classes of fractional ideals, or what is
more relevant here, isomorphism classes of projective modules of rank 1. Here is
another way to see the group law on Cl 𝑅: given two such fractional ideals 𝔞, 𝔟 up to
isomorphism, there is an isomorphism of 𝑅-modules
𝔞 ⊕ 𝔟 ' 𝑅 ⊕ 𝔞𝔟,
𝐼 ⊕ 𝐼0 ' 𝐽 ⊕ O (20.7.5)
of left O-modules.
Proof. We may suppose without loss of generality that 𝐼, 𝐼 0 ⊆ O. Then we have exact
sequences of left O-modules
𝜙
→ O → O/𝐼 → 0
0→𝐼−
𝜙0
0 → 𝐼 0 −−→ O → O/𝐼 0 → 0
𝜙 + 𝜙0 : 𝐼 ⊕ 𝐼 0 → O (20.7.6)
obtained by summing the two natural inclusions. We just showed that (𝜙 + 𝜙 0)𝔭
is surjective for all primes 𝔭, so it follows that 𝜙 + 𝜙 0 is surjective: the cokernel
𝑀 := coker(𝜙 + 𝜙 0) has 𝑀𝔭 = {0} for all 𝔭, and 𝑀 = {0}. Moreover, since O is
projective as a left O-module, the map 𝜙 + 𝜙 0 splits (or note that the map splits locally
for every prime 𝔭, so it splits globally, Exercise 20.1(b)). If we let 𝐽 := ker(𝜙 + 𝜙 0),
we then obtain an isomorphism
𝐼 ⊕ 𝐼 0 ' 𝐽 ⊕ O. (20.7.7)
But by the Krull–Schmidt theorem (Theorem 20.6.2) and Exercise 20.8, we can cancel
one copy of O𝔭 from both sides! We conclude that 𝐽𝔭 ' O𝔭 as left O-modules and
therefore by Lemma 17.3.3 that 𝐽𝔭 is (right) principal.
20.7. ∗ STABLE CLASS GROUP AND CANCELLATION 327
Lemma 20.7.9. Let 𝐼, 𝐼 0 ⊆ O be locally principal integral left O-ideals, and suppose
for every prime 𝔭 ⊆ 𝑅 either 𝐼𝔭 = O𝔭 or 𝐼𝔭0 = O𝔭 . Then
𝐼 ⊕ 𝐼 0 ' O ⊕ 𝐽, where 𝐽 = 𝐼 ∩ 𝐼 0 .
𝐽𝔭 = O𝔭 𝛼𝔭 𝛼𝔭0 = O𝔭 𝛼𝔭0 𝛼𝔭 .
and as above 𝐼 ⊕ 𝐼 0 ' 𝐽 ⊕ O with 𝐽 = 𝐼 ∩ 𝐼 0. The final statement follows from the
hypothesis that either 𝐼𝔭 = O𝔭 or 𝐼𝔭0 = O𝔭 , since then 𝛼𝔭 ∈ O×𝔭 or 𝛼𝔭0 ∈ O×𝔭 .
𝐽 ⊕ O ⊕𝑟 ' 𝐽 0 ⊕ O ⊕𝑟
for some 𝑟 ≥ 0.
Let [𝐽] St denote the stable isomorphism class of a left O-ideal 𝐽 and let StCl O be
the set of stable isomorphism classes of left O-ideals in 𝐵.
Proposition 20.7.11. StCl O is an abelian group under the binary operation (20.7.5),
written [𝐼] St + [𝐼 0] St = [𝐽] St , with identity [O] St .
𝐽𝔭 = O𝔭 𝛼𝔭 𝛼𝔭−1 𝛽𝔭 = O𝔭 𝛽𝔭
0 → 𝑃 → 𝑄 → 𝑃0 → 0
since such a sequence splits. The group 𝐾0 ( 𝐴) is sometimes called the projective
class group of 𝐴. (The group 𝐾0 ( 𝐴) is the Grothendieck group of the category P ( 𝐴).)
Then for 𝑃, 𝑄 ∈ P (O), we have [𝑃] = [𝑄] ∈ 𝐾0 (O) if and only if 𝑃, 𝑄 are stably
isomorphic [CR87, Proposition 38.22]. Moreover, there is a natural map
𝐾0 (O) → 𝐾0 (𝐵)
[𝑃] ↦→ [𝐹 ⊗𝑅 O],
and we let 𝑆𝐾0 (O) be its kernel, called the reduced projective class group of O. The
abelian group 𝑆𝐾0 (O) is generated by elements [𝑃] − [𝑄] where 𝑃, 𝑄 ∈ P (O) and
𝐹 ⊗𝑅 𝑃 ' 𝐹 ⊗𝑅 𝑄. Finally, we have an isomorphism [CR87, Theorem 49.32]
∼ 𝑆𝐾 (O)
StCl O −→ 0
(20.7.14)
[𝐼] St ↦→ [𝐼] − [O].
20.7. ∗ STABLE CLASS GROUP AND CANCELLATION 329
In other words, after all of this work—at least for maximal orders—the reduced
projective class group and the stable class group coincide. (For a more general
order, one instead compares to a maximal superorder via the natural extension maps
StCl O → StCl O0.)
The stable class group was first introduced and studied by Swan [Swa60, Swa62]
in this context in the special case where O = Z[𝐺] is the group ring of a finite group
𝐺.
Definition 20.7.15. We say that O has stable cancellation (or the simplification
property) if stable isomorphism implies isomorphism, i.e., if whenever 𝐼, 𝐼 0 are locally
principal left O-ideals with 𝐼 ⊕ O𝑟 ' 𝐼 0 ⊕ O𝑟 for 𝑟 ≥ 0, then in fact 𝐼 ' 𝐼 0.
If we had defined stable isomorphism and cancellation for locally free O-modules,
we would arrive at the same groups and condition, so stable cancellation is also called
the locally free cancellation.
From now on, suppose that 𝑅 is a global ring with 𝐹 = Frac 𝑅, and O ⊂ 𝐵 is a
maximal 𝑅-order in a quaternion algebra 𝐵 over 𝐹. We recall section 17.8, and the
class group ClΩ 𝑅, where Ω ⊆ Ram 𝐵 is the set of real ramified places.
Proof. See Fröhlich [Frö75, Theorem 2, §X], Swan [Swa80, Theorem 9.4], or Curtis–
Reiner [CR87, Theorem 49.32]; we will sketch a proof of a more general version of
this theorem in section 28.10, when we have idelic methods at our disposal.
20.7.18. Since 𝐵 is a quaternion algebra, the notions of invertible and locally principal
coincide. Then there is a surjective map of sets
ClsL O → StCl O
(20.7.19)
[𝐼] L ↦→ [𝐼] St .
Suppose further that 𝐹 = Frac 𝑅 is a number field and 𝑅 is a global ring. Then ClsL O
is a finite set, by Main Theorem 17.7.1; consequently, the stable class group StCl O is
a finite abelian group. However, the map (20.7.19) of sets need not be injective.
The order O has stable cancellation if and only if the map (20.7.19) is injective
(equivalently, bijective).
20.7.20. Suppose that 𝐵 satisfies the Eichler condition. Then by Eichler’s theorem
∼ Cl 𝑅 compatible
(Theorem 17.8.3), the reduced norm also gives a bijection Cls O −→ Ω
with the surjective map ClsL O → StCl O (20.7.19) which must therefore also be a
bijection.
330 CHAPTER 20. INTEGRAL REPRESENTATION THEORY
What remains, then, is the case where 𝐵 is definite. We restrict attention to the
case where the base field 𝐹 is a number field, hence a totally real field, and we work
with 𝑅-orders O ⊆ 𝐵, where 𝑅 = Z𝐹 is the ring of integers of 𝐹. Vignéras [Vig76b]
initiated the classification of definite quaternion orders with stable cancellation, and
showed that there are only finitely many such orders. Hallouin–Maire [HM2006] and
Smertnig [Sme2015] extended this classification to certain classes of orders, and the
complete classification was obtained by Smertnig–Voight [SV2019].
Remark 20.7.23. Jacobinski [Jaci68] was the first to consider the stable class group for
general orders in the context of his work on genera of lattices; his cancellation theorem
states more generally that if 𝐵 is a central simple algebra over 𝐹 and 𝐵 is not a totally
definite quaternion algebra, then every 𝑅-order O ⊆ 𝐵 has stable cancellation. This
result was reformulated by Fröhlich [Frö75] in terms of ideles and further developed
by Fröhlich–Reiner–Ullom [FRU74]. Swan [Swa80] related cancellation to strong
approximation in the context of 𝐾-groups.
Brzezinski [Brz83b] also defines the spinor class group of an order, a quotient of
its locally free class group; this group measures certain invariants phrased in terms of
quadratic forms.
Remark 20.7.24. More generally, a ring 𝐴 in which every stably free right 𝐴-module
is free is called a (right) Hermite ring by some authors: for further reference and
comparison of terminology, see Lam [Lam2006, Section I.4]. If O has locally free
cancellation, then O is Hermite; however, the converse does not hold in general—
a counterexample is described in detail by Smertnig [Sme2015]. Smertnig–Voight
[SV2019] show that there are exactly 375 definite quaternion 𝑅-orders with the Hermite
property up to isomorphism.
Exercises
Throughout these exercises, let 𝑅 be a noetherian domain with 𝐹 = Frac 𝑅, let 𝐵 be a
finite-dimensional 𝐹-algebra, let O ⊆ 𝐵 be an 𝑅-order, and let 𝐽 = rad O.
In this chapter, we consider hereditary orders, those with the simplest kind of module
theory; we characterize these orders in several ways, including showing they have an
extremal property with respect to their Jacobson radical.
We could define also right hereditary, but left hereditary and right hereditary are
equivalent for an 𝑅-order O, and so we simply say hereditary. We have O hereditary
if and only if every O-submodule of a projective finitely generated O-module is
projective—that is to say, projectivity is inherited by submodules. Moreover, being
hereditary is a local property.
Maximal orders are hereditary (Theorem 18.1.2), and one motivation for hereditary
orders is that many of the results from chapter 18 on the structure of two-sided ideals
extend from maximal orders to hereditary orders (Theorem 21.4.9).
333
334 CHAPTER 21. HEREDITARY AND EXTREMAL ORDERS
Hereditary orders are an incredibly rich class of objects, and they may be charac-
terized in a number of equivalent ways (Theorem 21.5.1). We restrict to the complete
local case, and suppose now that 𝑅 is a complete DVR with unique maximal ideal 𝔭
and residue field 𝑘 = 𝑅/𝔭.
Just as maximal orders are defined in terms of containment, we say O is extremal
if whenever O0 ⊇ O and rad O0 ⊇ rad O, then O0 = O. If O is not extremal, then
O0 := OL (rad O) ) O (21.1.3)
(i) O is hereditary;
(ii) 𝐽 is projective as a left O-module;
(ii0) 𝐽 is projective as a right O-module;
(iii) OL (𝐽) = O;
(iii0) OR (𝐽) = O;
(iv) 𝐽 is invertible as a (sated) two-sided O-ideal; and
(v) O is extremal.
The fact that hereditary orders are the same as extremal orders is quite remarkable,
and gives tight control over the structure of hereditary orders: extremal orders are
equivalently characterized as endomorphism algebras of flags in a suitable sense, and
so we have the following important corollary for quaternion algebras.
It is no surprise that we meet again the order from Example 20.1.2! The reader who
is willing to accept Corollary 21.1.5 can profitably move on from this chapter, as the
ring of upper triangular matrices is explicit enough to work with in many cases. That
being said, the methods we encounter here will be useful in framing investigations of
orders beyond the hereditary ones.
21.2.1. Our motivation comes from the following: we canonically associate a super-
order as follows. Let 𝐽 := rad O and O0 := OL (𝐽). Then O0 ⊇ O. By Corollary
20.5.5, 𝐽 𝑟 ⊆ 𝔭O ⊆ 𝔭O0 for some 𝑟 > 0, and then 𝐽 ⊆ rad O0.
We can think of extremal orders as like maximal orders, but under certain inclu-
sions.
Lemma 21.2.7. Suppose that 𝐵 is a division algebra over 𝐹 and let O ⊆ 𝐵 be extremal.
Then O is maximal.
Proof. Recall 13.3.7. The valuation ring O0 ⊇ O has the unique maximal two-sided
ideal 𝐽 0 = rad O0 = O0 \ (O0) × , so O0/𝐽 0 is a field. We have (21.2.6) 𝐽 0 ⊆ 𝐽, but then
𝐽/𝐽 0 = {0} ⊆ O0/𝐽 0 thus 𝐽 = 𝐽 0. Thus O0 radically covers O, and since O is extremal,
O = O0.
Remark 21.2.8. We stop short in our explicit description of local extremal orders in
section 21.2: we gave a construction in 21.3.1 only for 𝐵 ' M𝑛 (𝐹). The results extend
to 𝐵 ' M𝑛 (𝐷) where 𝐷 is a division algebra over 𝐹 by considering lattices in a free
left 𝐷-module: see Reiner [Rei2003, Theorem 39.14].
21.3.1. Let 𝑉 be a finite-dimensional 𝐹-vector space and let 𝐵 = End𝐹 (𝑉); then 𝑉 is
a simple 𝐵-module. Let 𝑀 ⊆ 𝑉 be an 𝑅-lattice. By Lemma 10.5.4, Λ := End𝑅 (𝑀) is
a maximal 𝑅-order; we have rad Λ = 𝔭Λ.
Choosing a basis for 𝑀, we get Λ ' M𝑛 (𝑅) ⊆ M𝑛 (𝐹) ' 𝐵, and rad Λ = M𝑛 (𝔭).
Now let 𝑍 := 𝑀 ⊗𝑅 𝑘 = 𝑀/𝔭𝑀. Then 𝑍 is a finite-dimensional vector space over
𝑘. Let
E : {0} = 𝑍0 ( 𝑍1 ( · · · ( 𝑍 𝑠−1 ( 𝑍 𝑠 = 𝑍
be a (partial) flag, a strictly increasing sequence of 𝑘-vector spaces. We define
Proof. That OL (E) is an order follows in the same way as the proof of Lemma 10.2.7.
For the statement on the radical: let 𝐽 = {𝛼 ∈ Λ : 𝛼𝑍𝑖 ⊆ 𝑍𝑖−1 }. Then 𝐽 ⊆ OL (E)
21.3. ∗ EXPLICIT DESCRIPTION OF EXTREMAL ORDERS 337
each factor is simple, so the sum is (Jacobson) semisimple; therefore 𝐽 ⊆ rad OL (E)
and equality holds.
Example 21.3.4. If we take the trivial flag OL (E) : {0} = 𝑍0 ( 𝑍1 = 𝑍, then
OL (E) = Λ, so this recovers the construction of maximal orders.
Example 21.3.5. Let E be the complete flag of length 𝑠 = 𝑛 + 1 = dim𝐹 𝑉, where
each quotient has dim 𝑘 (𝑍𝑖+1 /𝑍𝑖 ) = 1. Then there exists a basis 𝑧1 , . . . , 𝑧 𝑛 of 𝑍 so
that 𝑍𝑖 has basis 𝑧1 , . . . , 𝑧 𝑛−𝑖 ; We lift this to basis to 𝑥 1 , . . . , 𝑥 𝑛 of 𝑀 (by Nakayama’s
lemma), and in this basis, we have
𝑅 𝑅 𝑅 ... 𝑅
𝔭
© ª
𝑅 𝑅 ... 𝑅®
OL (E) = 𝔭 𝔭
®
𝑅 ... 𝑅®
.. .. .. ..
.
.. ®®
. . . .®
«𝔭 𝔭 𝔭 ... 𝑅¬
consisting of matrices which are upper triangular modulo 𝔭, and
𝔭 𝑅 𝑅 ... 𝑅
0 1 ... 0
𝔭 𝔭
© ª
𝑅 ... 𝑅® ©. .. .. ª®
. ..
rad OL (E) = 𝔭 𝔭 𝔭
® .
... 𝑅 ® = OL (E) . . .® (21.3.6)
.. .. .. .. ®® 0 0 1®
.. ®
...
. . . . .®
«𝜋 0 ... 0¬
«𝔭 𝔭 𝔭 ... 𝔭¬
where the latter is taken to be a block matrix with lower left entry 𝜋 and top right entry
the (𝑛 − 1) × (𝑛 − 1) identity matrix.
Other choices of flag give an order which lie between OL (E) and Λ: we might
think of them as being block upper triangular orders.
Now for the punch line of this section.
Proposition 21.3.7. Let O ⊆ 𝐵 be an 𝑅-order. Then O is extremal if and only if
O = OL (E) for a flag E.
Proof. Let O = OL (E). Let 𝐽 = rad O; we seek to apply Proposition 21.2.3, so we
show that O = OL (𝐽). By Lemma 21.3.3, we have 𝐽 𝑀𝑖 = 𝑀𝑖−1 so OL (𝐽) 𝑀𝑖−1 =
OL (𝐽)𝐽 𝑀𝑖 = 𝐽 𝑀𝑖 = 𝑀𝑖−1 for 𝑖 = 1, . . . , 𝑠. Since 𝑀0 = 𝔭𝑀 ' 𝑀, we conclude
OL (𝐽) = OL (E) = O by definition.
Conversely, suppose O is extremal with 𝐽 = rad O. Let 𝑠 be minimal so that
𝐽 𝑠 = 𝔭O. We may embed O ⊆ Λ for some Λ, and we take the flag
E : {0} = 𝐽 𝑠 𝑍 ( 𝐽 𝑠−1 𝑍 ( · · · ( 𝐽 𝑍 ( 𝑍.
338 CHAPTER 21. HEREDITARY AND EXTREMAL ORDERS
21.4.2. We could similarly define right hereditary, but since an order O is left and
right noetherian, it follows that O is left hereditary if and only if O is right hereditary:
see Exercise 21.8. When 𝐵 is a quaternion algebra, the standard involution inter-
changes and left and right, so the two notions are immediately seen to be equivalent.
Accordingly, we say hereditary for either sided notion.
Example 21.4.3. In the generic case 𝐹 = 𝑅 and O = 𝐵, we note that every semisimple
algebra 𝐵 over a field 𝐹 is hereditary: by Lemma 7.3.5, every 𝐵-module is semisimple
hence the direct sum of simple 𝐵-modules equivalently maximal left ideals, by Lemma
7.2.7.
Lemma 21.4.5. Let O be hereditary, and let 𝑃 be a finitely generated projective left
O-module. Then every submodule 𝑀 ⊆ 𝑃 is isomorphic as a left O-module to a finite
direct sum of finitely generated left O-ideals; in particular, 𝑀 is projective.
0 → ker 𝜙 → 𝑀 → 𝑀 ∩ 𝐸 → 0
and projectivity, we find that 𝑀 ' (𝑀 ∩ 𝐸) ⊕ ker 𝜙 where ker 𝜙 ⊆ O is a left ideal of
O. By induction, 𝑀 ∩ 𝐸 is projective, so the same is true of 𝑀.
21.4. HEREDITARY ORDERS 339
Proof. The implication (⇒) is Lemma 21.4.5; for the implication (⇐), O is projective
(free!) as a left O-module and every left ideal is a O-submodule 𝐼 ⊆ O, so by hypothesis
𝐼 is projective.
One of the desirable aspects of hereditary orders is that many of the results from
chapter 18 on the structure of two-sided ideals extend from maximal orders to heredi-
tary orders. Indeed, section 18.2 made no maximality hypothesis (we held out as long
as we could!).
However, if one supposes that every two-sided ideal is invertible (as a lattice),
then the argument can proceed: this is the class of hereditary orders, and is treated in
Theorem 21.4.9.
Remark 21.4.11. The module theory for hereditary noetherian prime rings, generaliz-
ing hereditary orders, has been worked out by Levy–Robson [LR2011].
Proof. We proved this in Theorem 18.1.2, but here is another proof using Theorem
21.5.1: the property of being maximal is local, and a maximal order is extremal.
𝔭𝑀 = 𝑀0 = 𝐽 𝑠 𝑀 ( 𝐽 𝑠−1 𝑀 ( · · · ( 𝑀𝑠−1 = 𝐽 𝑀 ( 𝑀𝑠 = 𝑀,
Exercises
1. Show that a Dedekind domain is hereditary (cf. Exercise 9.5).
√
2. Let 𝑅 = Z, let 𝐵 = 𝐾 = Q( 𝑑) with 𝑑 the discriminant of 𝐾, and let 𝑆 ⊆ Z𝐾 be
an order. Show that 𝑆 is hereditary if and only if 𝑆 is maximal.
3. Let 𝑅 bea DVR with maximal ideal 𝔭 = 𝜋𝑅 and 𝐹 = Frac 𝑅 with char 𝐹 ≠ 2.
1, 𝜋
Let 𝐵 = and O = 𝑅h𝑖, 𝑗i the standard order. Show directly that rad O =
𝐹
O 𝑗 = 𝑗O, and conclude that O is hereditary (but not a maximal order).
I 4. Give a self-contained proof of Theorem 21.4.9 following Theorem 18.3.4.
(Where does the issue with invertibility arise?)
5. Let 𝑅 be a complete DVR and let O be a hereditary 𝑅-order. Show that O is
hereditary if and only if rad O is an invertible (sated) two-sided O-ideal.
I 6. In this exercise, we prove Lemma 21.2.4 following Reiner [Rei2003, Exercise
39.2]. We adopt the notation from that section, so in particular 𝑅 be a complete
DVR with maximal ideal 𝔭 = rad(𝑅). Let O be an 𝑅-order and let O0 ⊆ 𝐵 be
an 𝑅-order containing O. Let 𝐽 0 = rad O0.
(a) Show that O + 𝐽 0 is an 𝑅-order.
(b) Show that O + 𝐽 0 radically covers O. [Hint: let 𝐽 = rad O, and claim that
𝐽 +𝐽 0 ⊆ rad(O+𝐽 0). For 𝑟 large, show 𝐽 𝑟 ⊆ 𝔭O so (𝐽 +𝐽 0) 𝑟 ⊆ 𝔭O0 +𝐽 0 and
3
(𝐽 0) 𝑟 ⊆ 𝔭O0, and then making 𝑟 even larger show (𝐽 + 𝐽 0) 𝑟 ⊆ 𝔭(O + 𝐽 0).
Conclude using Corollary 20.5.5.]
(c) If further 𝐽 0 ⊆ O, show that 𝐽 0 ⊆ 𝐽.
7. Let 𝑅 be a Dedekind domain with 𝐹 = Frac(𝑅), let 𝐵 be finite-dimensional
𝐹-algebra, and let O ⊆ 𝐵 be a hereditary order. Let 𝑃 be a finitely generated
projective O-module. Show that 𝑃 is indecomposable if and only if 𝑉 := 𝑃 ⊗𝑅 𝐹
is simple as a 𝐵-module.
I 8. Let 𝑅 be a Dedekind domain, and let O ⊆ 𝐵 be an 𝑅-order in a finite-dimensional
𝐹-algebra. Show that O is left hereditary (every left O-ideal is projective) if
and only if it is right hereditary (every right O-ideal is projective). [See Reiner
[Rei2003, Theorem 40.1].]
9. Consider the ring
𝑎 0
𝐴 := : 𝑎 ∈ Z, 𝑏, 𝑐 ∈ Q .
𝑏 𝑐
Show that every submodule of a projective left 𝐴-module is projective, but the
same is not true on the right.
342 CHAPTER 21. HEREDITARY AND EXTREMAL ORDERS
10. Let 𝑅 be a Dedekind domain. Let 𝐵 be a separable 𝐹-algebra, and let 𝐵 '
𝐵1 × · · · × 𝐵𝑟 be its decomposition into simple components, with 𝐵𝑖 = 𝐵𝑒 𝑖 for
central idempotents 𝑒 𝑖 . Let 𝐾𝑖 be the center of 𝐵𝑖 , and let 𝑆𝑖 be the integral
closure of 𝑅 in 𝐾𝑖 .
a) Let O ⊆ 𝐵 be a hereditary 𝑅-order. Show that O ' O1 × · · · × O𝑟 where
O𝑖 = O𝑒 𝑖 , and each O𝑖 is a hereditary 𝑅-order in 𝐵𝑖 .
b) Conversely, if O𝑖 ⊆ 𝐵𝑖 is a hereditary 𝑅-order, then O1 × · · · × O𝑟 is a
hereditary 𝑅-order in 𝐵.
[Hint: use the fact that hereditary orders are extremal.]
11. For the following exercise, we consider integral group rings. Let 𝐺 be a finite
group of order 𝑛 = #𝐺 and let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅.
Suppose that char 𝐹 - 𝑛. Then 𝐵 := 𝐹 [𝐺] is a separable 𝐹-algebra by Exercise
7.15. Let O = 𝑅[𝐺].
a) Let O0 ⊇ O be an 𝑅-superorder of O in 𝐵. Show that
O ⊆ O0 ⊆ 𝑛−1 O.
Tr 𝐵 |𝐹 (𝛼𝑔) = 𝑛𝑎 𝑔 ∈ 𝑅.
12. Give an explicit description like Example 21.3.5 for OL (E) when dim𝐹 𝑉 = 3, 4.
13. Let 𝑅 be a Dedekind domain, and let O ⊆ 𝐵 be an 𝑅-order in a finite-dimensional
simple 𝐹-algebra. Show that O is maximal if and only if O is hereditary and
rad O ⊆ O is a maximal two-sided ideal.
Chapter 22
In this chapter, we classify orders over a Dedekind domain in terms of ternary quadratic
forms; this is the integral analogue to what we did over fields in Chapter 5.
343
344 CHAPTER 22. TERNARY QUADRATIC FORMS
with (half-)discriminant
𝑁 := 4𝑎𝑏𝑐 + 𝑢𝑣𝑤 − 𝑎𝑢 2 − 𝑏𝑣 2 − 𝑐𝑤 2 ≠ 0,
𝑖 𝑗 𝑘 = 𝑗 𝑘𝑖 = 𝑘𝑖 𝑗 = 𝑎𝑏𝑐.
The 𝑅-order O defined by (22.1.2) is called the even Clifford algebra Clf 0 (𝑄) of
𝑄—its algebra structure is obtained by restriction from the even Clifford algebra of
𝑄 𝐹 —and the reduced discriminant of O is discrd(O) = (𝑁). At least one of the
minors
𝑢 2 − 4𝑏𝑐, 𝑣 2 − 4𝑎𝑐, 𝑤 2 − 4𝑎𝑏
of the Gram matrix of 𝑄 in the standard basis is nonzero since 𝑄 is nondegenerate, so
for example if 𝑤 2 − 4𝑎𝑏 ≠ 0 and char 𝐹 ≠ 2, completing the square we find
𝑤 2 − 4𝑎𝑏, −𝑎𝑁
O⊂𝐵' .
𝐹
(O♯ ) 0 = {𝛼 ∈ O♯ : trd(𝛼) = 0}
be the trace zero elements in the dual of O with respect to the reduced trace pairing.
Then we associate the ternary quadratic form
explicitly, we have
𝑁𝑖 ♯ = 𝑗 𝑘 − 𝑘 𝑗, 𝑁 𝑗 ♯ = 𝑘𝑖 − 𝑖𝑘, 𝑁 𝑘 ♯ = 𝑖 𝑗 − 𝑗𝑖
22.1. ⊲ QUATERNION ORDERS AND TERNARY QUADRATIC FORMS 345
where 1, 𝑖, 𝑗, 𝑘 is an 𝑅-basis of O, so
𝑁 nrd♯ (𝑥, 𝑦, 𝑧) = nrd 𝑥( 𝑗 𝑘 − 𝑘 𝑗) + 𝑦(𝑘𝑖 − 𝑖𝑘) + 𝑧(𝑖 𝑗 − 𝑗𝑖) . (22.1.4)
It is then a bit of beautiful algebra to verify that 𝑁 nrd♯ has discriminant 𝑁 and that
(22.1.3) furnishes an inverse to the even Clifford map.
Just as in the case of fields, the translation from quaternion orders to ternary
quadratic forms makes the classification problem easier: we replace the potentially
complicated notion of finding a lattice closed under multiplication in a quaternion
algebra with the simpler notion of choosing coefficients of a quadratic form.
To conclude this introduction, we state a more general bijective result stated in
terms of lattices. Let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅, let 𝑄 𝐹 : 𝑉 → 𝐹
be a nondegenerate ternary quadratic form. If 𝑀 ⊆ 𝑉 is an 𝑅-lattice, and 𝔩 ⊆ 𝐹 is
a fractional ideal of 𝑅 such that 𝑄(𝑀) ⊆ 𝔩, then we have an induced quadratic form
𝑄 : 𝑀 → 𝔩; we call such a form a quadratic module in 𝑉. Given a fractional ideal
𝔞 ⊆ 𝐹, the twist by 𝔞 of the quadratic module 𝑄 : 𝑀 → 𝔩 in 𝑉 is the quadratic module
𝔞𝑀 → 𝔞2 𝔩. A twisted similarity between quadratic modules 𝑄, 𝑄 0 in 𝑉 is a similarity
between 𝑄 and a twist of 𝑄 0. From these notions in hand, we have the following
theorem (a special case of Main Theorem 22.5.7).
By functorial with respect to 𝑅, we mean the same thing as in Corollary 5.2.6, but
with respect to any homomorphism 𝑅 → 𝑆 of Dedekind domains. In particular, the bi-
jection in Theorem 22.1.5 is compatible with the bijections obtained over localizations
of 𝑅, including the bijection over 𝐹 between quaternion algebras and nondegener-
ate ternary quadratic forms previously obtained. In the language of quadratic forms
(Definition 9.7.13), after some additional work (nailing down the difference between
similarity and isometry), we conclude: if the ternary quadratic module 𝑄 corresponds
to the quaternion order O, then there is a bijection
Cl 𝑄 ↔ Typ O, (22.1.6)
i.e. the type number of a quaternion order is the same as the class number of the
corresponding ternary quadratic form.
Remark 22.1.7. If we restrict the correspondence to primitive modules 𝑄 : 𝑀 → 𝔩 (i.e.,
𝑄(𝑀) = 𝔩), then we need only remember the underlying lattice 𝑀, and on the right-
hand side we obtain precisely the Gorenstein orders; these orders will be introduced
in 24.1.1 and this correspondence is proven in section 24.2.
346 CHAPTER 22. TERNARY QUADRATIC FORMS
22.2.1. Let
∞
Ê
Ten0 (𝑀; 𝐿) := (𝑀 ⊗ 𝑀 ⊗ 𝐿 ∨ ) ⊗𝑑 .
𝑑=0
0
Now Ten (𝑀; 𝐿) has a natural tensor multiplication law (rearranging tensors), so
Ten0 (𝑀; 𝐿) is a graded 𝑅-algebra. Let 𝐼 0 (𝑄) be the two-sided ideal of Ten0 (𝑀; 𝐿)
defined by
unfortunately, Ten(𝑀; 𝐿) does not have a natural tensor multiplication law, because
there is no natural map 𝑀 ⊗ 𝑀 → 𝑀 ⊗ 𝑀 ⊗ 𝐿 ∨ . But see 22.2.16 below for the odd
part.
We conclude that the 𝑅-lattice in Clf 0 (𝑄 𝐹 ) defined by the image of 𝑅 3 is closed under
multiplication—something that may also be verified directly—and so Clf 0 (𝑄) is an
𝑅-order in Clf 0 (𝑄 𝐹 ).
22.2.9. If 𝑀 ' 𝑅 𝑛 is free with basis 𝑒 1 , . . . , 𝑒 𝑛 and 𝐿 = 𝑅𝑔 is free, then Clf 0 (𝑄) is a
free 𝑅-module with basis
Theorem 22.2.11. The association 𝑄 ↦→ Clf 0 (𝑄) is a functor from the category of
quadratic 𝑅-modules under similarities to the category of projective 𝑅-algebras with
involution under isomorphism. Moreover, this association is functorial with respect to
𝑅.
We call the association 𝑄 ↦→ Clf 0 (𝑄) in Theorem 22.2.11 the even Clifford
functor.
22.2.12. The statement “functorial with respect to 𝑅” means the following: given
a ring homomorphism 𝑅 → 𝑆, there is a natural transformation between the even
Clifford functors over 𝑅 and 𝑆. Explicitly, given a ring homomorphism 𝑅 → 𝑆 and a
quadratic module 𝑄 : 𝑀 → 𝐿, we have a quadratic module 𝑄 𝑆 : 𝑀 ⊗𝑅 𝑆 → 𝐿 ⊗𝑅 𝑆,
and Clf 0 (𝑄) ⊗𝑅 𝑆 Clf 0 (𝑄 𝑆 ) in a way compatible with morphisms in each category.
In particular, this recovers the identification in Example 22.2.5 arising from 𝑅 ↩→ 𝐹.
Remark 22.2.13. The association 𝑄 ↦→ Clf (𝑄) of the full Clifford algebra is a functor
from the category of quadratic 𝑅-modules under isometries to the category of 𝑅-
algebras with involution under isomorphism that is also functorial with respect to 𝑅.
See Bischel–Knus [BK94].
Proof of Theorem 22.2.11. The construction in 22.2.1 yields an 𝑅-algebra that is pro-
jective as an 𝑅-module; we need to define an association on the level of morphisms.
∼ 𝑀0
Let 𝑄 0 : 𝑀 0 → 𝐿 0 be a quadratic module and ( 𝑓 , ℎ) be a similarity with 𝑓 : 𝑀 −→
∼ 0 0
and ℎ : 𝐿 −→ 𝐿 satisfying 𝑄 ( 𝑓 (𝑥)) = ℎ(𝑄(𝑥)). We mimic the proof of Lemma
5.3.21. We define a map via
and since
𝑔(𝑄(𝑥)) = 𝑔(ℎ−1 (𝑄 0 ( 𝑓 (𝑥)))) = (ℎ−1 ) ∗ (𝑔) (𝑄 0 ( 𝑓 (𝑥))),
we conclude that 𝐼 0 (𝑄) is mapped to 𝐼 0 (𝑄 0). Repeating with the inverse similarity
( 𝑓 −1 , ℎ−1 ), and composing to get the identity, we conclude that the induced map
Clf 0 (𝑄) → Clf 0 (𝑄 0) is an 𝑅-algebra isomorphism.
Functoriality in the sense of 22.2.12 then follows directly.
Remark 22.2.14. In his thesis, Bichsel [Bic85] constructed an even Clifford algebra of
a line bundle-valued quadratic form on an affine scheme using faithfully flat descent.
A related and more general construction was given by Bischel–Knus [BK94]. Several
other constructions are available: see Auel [Auel2011, §1.8] and the references therein.
The direct tensorial construction given above is given for ternary quadratic modules
by Voight [Voi2011a, (1.10)] and in general by Auel [Auel2011, §1.8] and with further
detail in Auel [Auel2015, §1.2]; for a comparison of this direct construction with
others, see Auel–Bernardara–Bolognesi [ABB2014, §1.5, Appendix A].
Remark 22.2.15. Allowing the quadratic forms to take values in a invertible module
is essential for what follows and for many other purposes: for an overview, see the
introduction to Auel [Auel2011].
22.2.16. Let
∞
Ê
Ten1 (𝑀; 𝐿) := 𝑀 ⊗𝑑 ⊗ (𝐿 ∨ ) ⊗ b𝑑/2c = 𝑀 ⊕ (𝑀 ⊗ 𝑀 ⊗ 𝑀 ⊗ 𝐿 ∨ ) ⊕ . . . .
𝑑=1
𝑑 odd
Then Ten1 (𝑀; 𝐿) is a graded Ten0 (𝑀; 𝐿)-bimodule under the natural tensor multi-
plication. Let 𝐼 1 (𝑄) be the 𝑅-submodule of Ten1 (𝑀; 𝐿) generated by the image of
multiplication of 𝐼 0 (𝑄) by 𝑀 on the left and right: then 𝐼 1 (𝑄) is the Ten0 (𝑀; 𝐿)-
bisubmodule generated by the set of elements of the form
with 𝑥, 𝑦 ∈ 𝑀 and 𝑔 ∈ 𝐿 ∨ .
We define the odd Clifford bimodule as
Visibly, Clf 1 (𝑄) is a bimodule for the even Clifford algebra Clf 0 (𝑄).
22.2.17. When 𝐿 = 𝑅, we can combine the construction of the even Clifford algebra
and its odd Clifford bimodule to construct a full Clifford algebra, just as in section
5.3 over a field: see Exercise 22.7. This direct tensorial construction does not extend
in an obvious way when 𝐿 ≠ 𝑅, as we would need to define a multiplication map
𝑀 ⊗ 𝑀 → 𝑀 ⊗ 𝑀 ⊗ 𝐿∨.
22.3. EVEN CLIFFORD ALGEBRA OF A TERNARY QUADRATIC MODULE 349
22.2.18. We will employ exterior calculus in what follows: this is a convenient method
for keeping track of our module maps in a general setting. Let 𝑀 be an 𝑅-module and
let 𝑟 ≥ 1. The 𝑟th exterior power of 𝑀 (over 𝑅) is
𝑀 := 𝑀 ⊗𝑟 /𝐸𝑟
Ó𝑟
𝑅 ⊕ 𝑀 ⊗ 𝑀 ⊗ 𝐿∨
0
Clf (𝑄) ' (22.3.3)
𝐼 0 (𝑄)
𝑥 ⊗ 𝑥 ⊗ 𝑔 − 1 ⊗ 𝑔(𝑄(𝑥))
for 𝑥 ∈ 𝑀 and 𝑔 ∈ 𝐿 ∨ .
We now explicitly give the even Clifford algebra of a ternary quadratic module in
the free case; this could also be taken as the definition when 𝑅 is a PID and 𝑀 = 𝑅 3 .
350 CHAPTER 22. TERNARY QUADRATIC FORMS
22.3.4. Let 𝑀 = 𝑅 3 with standard basis 𝑒 1 , 𝑒 2 , 𝑒 3 be equipped with the quadratic form
𝑄 : 𝑀 → 𝑅 defined by
with 𝑎, 𝑏, 𝑐, 𝑢, 𝑣, 𝑤 ∈ 𝑅. Then
By 22.2.9, we have
Clf 0 (𝑄) = 𝑅 ⊕ 𝑅𝑖 ⊕ 𝑅 𝑗 ⊕ 𝑅𝑘
where
𝑖 := 𝑒 2 𝑒 3 , 𝑗 := 𝑒 3 𝑒 1 , 𝑘 := 𝑒 1 𝑒 2 .
The reversal involution acts by
𝑖 = 𝑒 3 𝑒 2 = 𝑇 (𝑒 2 , 𝑒 3 ) − 𝑖 = 𝑢 − 𝑖,
𝑖 2 = 𝑢𝑖 − 𝑏𝑐 𝑗 𝑘 = 𝑎𝑖
2
𝑗 = 𝑣 𝑗 − 𝑎𝑐 𝑘𝑖 = 𝑏 𝑗 (22.3.7)
2
𝑘 = 𝑤𝑘 − 𝑎𝑏 𝑖 𝑗 = 𝑐𝑘
For example,
𝑖 2 = (𝑒 2 𝑒 3 ) (𝑒 2 𝑒 3 ) = 𝑒 2 (𝑒 3 𝑒 2 )𝑒 3 = 𝑒 2 (𝑢 − 𝑒 2 𝑒 3 )𝑒 3 = 𝑢𝑒 2 𝑒 3 − 𝑒 22 𝑒 23 = 𝑢𝑖 − 𝑏𝑐
and
𝑗 𝑘 = (𝑒 3 𝑒 1 ) (𝑒 1 𝑒 2 ) = 𝑎𝑒 3 𝑒 2 = 𝑎𝑖.
The remaining multiplication laws can be computed in the same way, or by using the
reversal involution and (22.3.7): we compute
𝑎𝑖 = 𝑗 𝑘 = 𝑘 𝑗 = (𝑤 − 𝑘) (𝑣 − 𝑗) = 𝑣𝑤 − 𝑤 𝑗 − 𝑣𝑘 + 𝑘 𝑗
𝑘 𝑗 = −𝑣𝑤 + 𝑎𝑖 + 𝑤 𝑗 + 𝑣𝑘
𝑖𝑘 = −𝑢𝑤 + 𝑤𝑖 + 𝑏 𝑗 + 𝑢𝑘 (22.3.8)
𝑗𝑖 = −𝑢𝑣 + 𝑣𝑖 + 𝑢 𝑗 + 𝑐𝑘
𝑖 𝑗 𝑘 = 𝑗 𝑘𝑖 = 𝑘𝑖 𝑗 = 𝑎𝑏𝑐. (22.3.9)
22.3. EVEN CLIFFORD ALGEBRA OF A TERNARY QUADRATIC MODULE 351
Example 22.3.10. It is clarifying to work out the diagonal case. Let 𝐵 = (𝑎, 𝑏 | 𝐹)
with 𝑎, 𝑏 ∈ 𝑅, and let
O = 𝑅h𝑖, 𝑗i = 𝑅 + 𝑅𝑖 + 𝑅 𝑗 + 𝑅𝑖 𝑗 ⊂ 𝐵
𝑖2 = 𝑎 𝑗 𝑘 = 𝑏𝑖 = −𝑏𝑖
𝑗2 = 𝑏 𝑘𝑖 = 𝑎 𝑗 = −𝑎 𝑗 (22.3.11)
2
𝑘 = −𝑎𝑏 𝑖 𝑗 = −𝑘 = 𝑘.
𝑄(𝑥, 𝑦, 𝑧) = 𝑥𝑦 − 𝑧 2
so (𝑎, 𝑏, 𝑐, 𝑢, 𝑣, 𝑤) = (0, 0, −1, 0, 0, 1); then disc(𝑄) = −𝑐𝑤 2 = 1. Then the even
Clifford algebra Clf 0 (𝑄) = 𝑅 + 𝑅𝑖 + 𝑅 𝑗 + 𝑅𝑘 has multiplication table
𝑖2 = 0 𝑗𝑘 = 0
2
𝑗 =0 𝑘𝑖 = 0 (22.3.13)
𝑘2 = 𝑘 𝑖 𝑗 = −𝑘 = 𝑘 − 1.
We find an isomorphism of 𝑅-algebras
∼ 𝑀 (𝑅)
Clf 0 (𝑄) −→ 2
(22.3.14)
0 1 0 0 0 0
𝑖, 𝑗, 𝑘 ↦→ , , .
0 0 −1 0 0 1
22.3.15. Returning to the free quadratic form 22.3.4, the group GL3 (𝑅) acts naturally
on 𝑀 by change of basis, and this induces an action on Clf 0 (𝑄) by 𝑅-algebra auto-
morphism by functoriality. Explicitly, for 𝜌 ∈ GL3 (𝑅), the action on the basis 𝑖, 𝑗, 𝑘
is by the adjugate adj(𝜌) of 𝜌, the 3 × 3 matrix whose entries are the 2 × 2 minors of
𝜌. The verification is requested in Exercise 22.2.
22.3.16. Let 𝐹 = Frac 𝑅. By base extension, we have a quadratic form 𝑄 𝐹 : 𝑉 → 𝐹
where 𝑉 = 𝑀 ⊗𝑅 𝐹, and by functoriality 22.2.12 with respect to the inclusion 𝑅 ↩→ 𝐹,
we have an inclusion Clf 0 (𝑄) ↩→ Clf 0 (𝑄 𝐹 ) realizing Clf 0 (𝑄) as an 𝑅-order in the
𝐹-algebra Clf 0 (𝑄 𝐹 ).
Lemma 22.3.17. The reversal involution is a standard involution on Clf 0 (𝑄 𝐹 ).
Proof. To check that the involution is standard, we could appeal to Exercise 3.19, but
we find it more illustrative to exhibit the involution on a universal element, yielding a
rather beautiful formula. We choose a basis for 𝑉 and work with the presentation for
Clf 0 (𝑄 𝐹 ) as in 22.3.4.
Let 𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 with 𝑡, 𝑥, 𝑦, 𝑧 ∈ 𝐹. Then 𝛼 = 2𝑡 + 𝑢𝑥 + 𝑣𝑦 + 𝑤𝑧 − 𝛼, and
we find that
𝛼2 − (𝛼 + 𝛼)𝛼 + 𝛼𝛼 = 𝛼2 − 𝜏(𝛼)𝛼 + 𝜈(𝛼) = 0
352 CHAPTER 22. TERNARY QUADRATIC FORMS
where
𝜏(𝛼) = 2𝑡 + 𝑢𝑥 + 𝑣𝑦 + 𝑤𝑧
𝜈(𝛼) = 𝑡 2 + 𝑢𝑡𝑥 + 𝑣𝑡𝑦 + 𝑤𝑡𝑧
(22.3.18)
+ 𝑏𝑐𝑥 2 + (𝑢𝑣 − 𝑐𝑤)𝑥𝑦 + (𝑢𝑤 − 𝑏𝑣)𝑥𝑧
+ 𝑎𝑐𝑦 2 + (𝑣𝑤 − 𝑎𝑢)𝑦𝑧 + 𝑎𝑏𝑧 2
so that the reversal map 𝛼 ↦→ 𝜏(𝛼) − 𝛼 defines a standard involution.
Proof. The construction of the even Clifford algebra is functorial with respect to
localization, and the statement itself is local, so we may suppose that 𝑀 = 𝑅 3 , 𝐿 = 𝑅
are free with the presentation for O = Clf 0 (𝑄) as in 22.3.4.
We refer to section 15.4 and Lemma 15.4.7: we compute
2 𝑢 𝑣 𝑤
𝑢 𝑢 2 − 2𝑏𝑐
© ª
𝑐𝑤 𝑏𝑣 ®
𝑑 (1, 𝑖, 𝑗, 𝑘) = 2
𝑣 − 2𝑎𝑐
®
𝑣 𝑐𝑤 𝑎𝑢 ® (22.3.21)
2
𝑤 − 2𝑎𝑏 ¬
«𝑤 𝑏𝑣 𝑎𝑢
= −(4𝑎𝑏𝑐 + 𝑢𝑣𝑤 − 𝑎𝑢 − 𝑏𝑣 − 𝑐𝑤 2 ) 2 = − disc(𝑄) 2
2 2
so disc(O) = disc(𝑄) 2 𝑅, and the result follows by taking square roots (as ideals).
𝑀 = 𝔞𝑒 1 ⊕ 𝔟𝑒 2 ⊕ 𝔠𝑒 3 and 𝐿=𝔩
where 𝑖, 𝑗, 𝑘 satisfy the multiplication table (22.3.7). We can verify directly that O is
closed under multiplication: for example, if 𝛼 ∈ 𝔟𝔠𝔩−1 so 𝛼𝑖 ∈ O, then
(𝛼𝑖) 2 = 𝑢𝛼𝑖 − 𝛼2 𝑏𝑐 ∈ O
We have √ √ √
discrd(O) = 4(9) (4 + 10)𝔭−3 = (2, 10) 5 (3, 2 + 10) 2
and in particular 𝔭 - discrd(O), and
√ √
−3(4 + 10), −3(4 + 10)
O⊂𝐵=
𝐹
√
with disc 𝐵 = (2 + 𝑤)𝑅, so Ram 𝐵 = {(2, 10), 𝔭, ∞1 , ∞2 } where ∞1 , ∞2 are the two
real places of 𝐹.
354 CHAPTER 22. TERNARY QUADRATIC FORMS
Main Theorem 22.4.1. Suppose that 𝑅 is a PID. Then the association 𝑄 ↦→ Clf 0 (𝑄)
induces a discriminant-preserving bijection
Nondegenerate ternary quadratic Quaternion orders over 𝑅
↔ (22.4.2)
forms over 𝑅 up to similarity up to isomorphism
Remark 22.4.3. The bijection can also be rephrased in terms of the orbits of a group
(following Gross–Lucianovic [GrLu2009]). The group GL3 (𝑅) has a natural twisted
action on quadratic forms by (𝑔𝑄) (𝑥, 𝑦, 𝑧) = (det 𝑔) (𝑄(𝑔 −1 (𝑥, 𝑦, 𝑧) t )), i.e., the usual
action with an extra scalingÓfactor of det 𝑔 ∈ 𝑅 × . This is the natural action on the
𝑅-module Ó Sym2 ((𝑅 3 ) ∨ ) ⊗ 3 𝑅 3 , or equivalently on the set of quadratic modules
3
𝑄 : 𝑅 → 3 𝑅 3 . Main Theorem 22.1.1 states that the nondegenerate orbits of this
action are in functorial bijection with the set of isomorphism classes of quaternion
orders over 𝑅.
We prove this theorem in a few steps. Throughout this section, let 𝑅 be a PID.
First, we prove that the map (22.4.2) is surjective, or equivalently that the even
Clifford functor is essentially surjective from the category of nondegenerate ternary
quadratic forms to the category of quaternion orders.
Proof. We work explicitly with the multiplication table, hoping to make it look like
(22.3.7).
Let O be a quaternion 𝑅-order. Since 𝑅 is a PID, O is free as an 𝑅-module. We
need a slight upgrade from this, a technical result supplied by Exercise 22.1: in fact,
O has an 𝑅-basis containing 1.
So let 1, 𝑖, 𝑗, 𝑘 be an 𝑅-basis for O. Since every element of O is integral over 𝑅,
satisfying its reduced characteristic polynomial of degree 2 over 𝑅, we have
𝑖 2 = 𝑢𝑖 + 𝑙
𝑗2 = 𝑣 𝑗 + 𝑚
𝑘 2 = 𝑤𝑘 + 𝑛
𝑗 𝑘 0 = 𝑗 (𝑘 − 𝑞) = 𝑟 − 𝑎𝑖 + 𝛼𝑘 = (𝑟 + 𝛼𝑞) − 𝑎𝑖 + 𝛼𝑘 0 .
22.4. OVER A PID 355
𝑗 𝑘 = 𝑟 − 𝑎𝑖 + 𝛼𝑘
𝑘𝑖 = 𝑠 − 𝑏 𝑗 + 𝛽𝑖
𝑖 𝑗 = 𝑡 − 𝑐𝑘 + 𝛾 𝑗
As before, the other products can be calculated using the standard involution: for
example, we have
so
𝑖𝑘 = (𝑠 + 𝛽𝑢 − 𝑏𝑣 − 𝑢𝑤) + (𝑤 − 𝛽)𝑖 + 𝑏 𝑗 + 𝑢𝑘. (22.4.5)
But now from these multiplication laws, we compute that the trace of left multiplication
𝑖 is Tr(𝑖) = 0 + 𝑢 + 𝛾 + 𝑢 = 2𝑢 + 𝛾. But in a quaternion algebra, we have Tr(𝑖) =
2 trd(𝑖) = 2𝑢, so we must have 𝛾 = 0. By symmetry, we find that 𝛼 = 𝛽 = 0. Finally,
associativity implies relations on the structure constants in the multiplication table:
we have
𝑗 (𝑘 𝑘) = ( 𝑗 𝑘)𝑘
−𝑛 𝑗 = (𝑟 − 𝑎𝑖) (𝑤 − 𝑘) = 𝑟𝑤 − 𝑎𝑤𝑖 − 𝑟 𝑘 + 𝑎𝑖𝑘 (22.4.6)
−𝑛 𝑗 = (𝑟𝑤 + 𝑎𝑠 − 𝑎𝑏𝑣 − 𝑎𝑢𝑤) + 𝑎𝑏 𝑗 + (𝑎𝑢 − 𝑟)𝑘
using (22.4.5) with 𝛽 = 0; so equality of coefficients of 𝑗, 𝑘 implies 𝑟 = 𝑎𝑢 and
𝑛 = −𝑎𝑏. By symmetry, we find 𝑠 = 𝑏𝑣, 𝑡 = 𝑐𝑤 and 𝑚 = −𝑎𝑐, 𝑛 = −𝑎𝑏, so we have
the following multiplication table:
𝑖 2 = 𝑢𝑖 − 𝑏𝑐 𝑗 𝑘 = 𝑎𝑖
2
𝑗 = 𝑣 𝑗 − 𝑎𝑐 𝑘𝑖 = 𝑏 𝑗
𝑘 2 = 𝑤𝑘 − 𝑎𝑏 𝑖 𝑗 = 𝑐𝑘
This matches precisely the multiplication table (22.3.7) for the even Clifford algebra
of the quadratic form 𝑄(𝑥, 𝑦, 𝑧) = 𝑎𝑥 2 + 𝑏𝑦 2 + 𝑐𝑧2 + 𝑢𝑦𝑧 + 𝑣𝑥𝑧 + 𝑤𝑥𝑦.
22.4.7. More generally, if 𝑅 is a domain and O is a quaternion 𝑅-order such that O
is free as an 𝑅-module with basis 1, 𝑖, 𝑗, 𝑘, then the proof of Proposition 22.4.4 shows
O has a basis 1, 𝑖, 𝑗, 𝑘 satisfying the multiplication laws (22.3.7) of an even Clifford
algebra; we call such a basis a good basis for O. Moreover, we have seen that given
a basis 1, 𝑖, 𝑗, 𝑘, there exist unique 𝜂(𝑖), 𝜂( 𝑗), 𝜂(𝑘) ∈ 𝑅 (in fact, certain coefficients of
the multiplication table) such that
is a good basis.
356 CHAPTER 22. TERNARY QUADRATIC FORMS
To conclude, we need to show that if two quaternion 𝑅-orders are isomorphic, then
they correspond to similar ternary quadratic forms. To this end, we define an inverse.
22.4.8. Let O ⊆ 𝐵 be a quaternion 𝑅-order with 𝑅-basis 1, 𝑖, 𝑗, 𝑘. Let 𝑁 ∈ 𝑅 be such
that (𝑁) = discrd(O); then 𝑁 ≠ 0 and is well-defined up to multiplication by 𝑅 × . Let
1♯ , 𝑖 ♯ , 𝑗 ♯ , 𝑘 ♯ be the dual basis (see 15.6.3); then trd(𝑖 ♯ ) = trd(1 · 𝑖 ♯ ) = 0 and similarly
for 𝑗 ♯ , 𝑘 ♯ , so
(O♯ ) 0 = {𝛼 ∈ O♯ : trd(𝛼) = 0} = 𝑅𝑖 ♯ + 𝑅 𝑗 ♯ + 𝑅𝑘 ♯ .
well-defined up to similarity (along the way, we chose a basis and a generator for
discrd(O)).
Example 22.4.10. We return to Example 22.3.10. The 𝑅-order O has reduced dis-
criminant 𝑁 = 4𝑎𝑏. The (rescaled) dual basis is
𝑁𝑖 ♯ = 2𝑏𝑖, 𝑁 𝑗 ♯ = 2𝑎 𝑗, 𝑁 𝑘 ♯ = −2𝑘
We see that trd(𝑁𝑖 ♯ ) = 0 and the same with 𝑗 ♯ , 𝑘 ♯ . We recall the alternating trilinear
form 𝑚 (defined in 15.4.2). By (22.3.20) we have
𝑚(𝑖, 𝑗, 𝑘) = −𝑁 = trd(𝑖( 𝑗 𝑘 − 𝑘 𝑗)) = − trd(𝑖( 𝑗 𝑘 − 𝑘 𝑗)) = trd(𝑖(𝑁𝑖 ♯ ))
and
𝑚( 𝑗, 𝑗, 𝑘) = 0 = trd( 𝑗 ( 𝑗 𝑘 − 𝑘 𝑗)) = − trd( 𝑗 (𝑁𝑖 ♯ ))
and similarly trd(𝑘 (𝑁𝑖 ♯ )) = 0. The other equalities follow similarly, and this verifies
the dual basis (22.4.13). In particular, we have trd(𝑁𝑖 ♯ ) = trd(𝑁 𝑗 ♯ ) = trd(𝑁 𝑘 ♯ ) = 0.
We then compute the quadratic form on this basis and claim that
satisfies
(𝑁1♯ ) 2 − 𝑁 (𝑁1♯ ) + 𝑁 (𝑎𝑏𝑐 + 𝑢𝑣𝑤) = 0 (22.4.18)
and
Finally, we officially combine our work to prove the main theorem of this section.
Remark 22.4.19. Just as in section 5.5, we may ask about embeddings of a quadratic
ring in an order. However, moving from the rational to the integral is a bit tricky, and
the issue of embeddings is a theme that will return with gusto in Chapter 30. In that
context, it will be more natural to look at a different ternary quadratic form to measure
embeddings; just as in the case of trace zero, it is related to but not the same as the one
obtained in the above bijection.
22.5.4. Second, we extend the definition of the inverse in 22.4.8 using the reduced
norm to the noetherian domain 𝑅 as follows. Let O ⊆ 𝐵 be a quaternion 𝑅-order.
22.5. TWISTING AND FINAL BIJECTION 359
Lemma 22.5.6. For a quaternion order O, the map nrd♯ (O) defines a ternary
quadratic module with values in discrd(O) −1 .
Proof. The reduced norm defines a quadratic map, so we only need to verify that the
codomain is valid. To this end, we may check locally since reduced norm and reduced
discriminant commute with localization. Reducing to the local case, suppose 𝑅 is now
a local Dedekind domain hence a PID. Choosing a basis, we verified in (22.4.14) that
𝑁 nrd♯ (O) ⊆ 𝑅, where discrd(O) = (𝑁); the result follows.
are mutually inverse, discriminant-preserving bijections that are also functorial with
respect to 𝑅.
𝔡 := 𝔞𝔟𝔠𝔩−1 ' 𝑀 ⊗ 𝐿∨ .
Ó3
(22.5.10)
360 CHAPTER 22. TERNARY QUADRATIC FORMS
Then
𝔡(O♯ ) 0 = 𝔞𝑖 ♯ ⊕ 𝔟 𝑗 ♯ ⊕ 𝔠𝑘 ♯ . (22.5.11)
We claim that the reduced norm on 𝔡(O♯ ) 0 has values in 𝑁 −1 𝔩 and is similar to 𝑄. The
claim follows from the same calculation in the proof of Proposition 22.4.12, namely
that nrd(𝑥𝑖 ♯ + 𝑦 𝑗 ♯ + 𝑧𝑘 ♯ ) = 𝑁 −1 𝑄 𝐹 (𝑥, 𝑦, 𝑧)! We conclude that nrd♯ (O) is twisted
similar to 𝑄.
Proof. We first claim that the even Clifford map induces an injection Gen 𝑄 → Gen O,
giving an injection Cl 𝑄 → Typ O. Indeed, let 𝑄 0 ∈ Gen 𝑄, so 𝑄 0 : 𝑀 0 → 𝔩0 is locally
isometric to 𝑄. Let O0 := Clf 0 (𝑄 0). Since 𝑄 0, 𝑄 are locally isometric, they are locally
similar, so O0, O are locally isomorphic by Main Theorem 22.4.1, so O0 ∈ Gen O.
And if 𝑄 0 ' 𝑄 are isometric, again they are (twisted) similar, so by Main Theorem
22.5.7 we have O0 ' O.
To finish, we need to show that the even Clifford map is surjective. We pass from
similarity classes to isometry classes in the same way as in the proof of Corollary
5.2.6. To this end, let O0 ∈ Gen(O). Let 𝑄 0 := 𝔡 nrd♯ (O0) as in (22.5.11). By the
same rescaling argument given in Corollary 5.2.6, applying a similarity to 𝑄 0 we may
further suppose that disc 𝑄 𝐹0 = disc 𝑄 𝐹 ∈ 𝐹 × /𝐹 ×2 . By surjectivity in Main Theorem
22.5.7, for every prime 𝔭 of 𝑅, there exists a twisted similarity from 𝑄 0(𝔭) to 𝑄 (𝔭) over
𝑅 (𝔭) —and since each 𝑅 (𝔭) is a PID, by Example 22.5.2, these are in fact similarities.
Taking such a similarity and considering it as a similarity over 𝐹, again repeating the
same argument as at the end of Corollary 5.2.6, we conclude that 𝑄 0 and 𝑄 are locally
isometric, so 𝑄 0 ∈ Gen 𝑄. Finally, if O0 ' O, repeating these arguments one more
time over 𝑅 (first to go from twisted similar to similar, then to note the similarity gives
rise to an isometry) we conclude that 𝑄 0 ' 𝑄.
Remark 22.5.13. The correspondence between ternary quadratic forms and quaternion
orders has a particularly rich history. Perhaps the earliest prototype is due to Hermite
[Herm1854], who examined the product of automorphs of ternary quadratic forms.
Early versions of the correspondence were given by Latimer [Lat37, Theorem 3], Pall
[Pall46, Theorems 4–5], and Brandt [Bra43, §3ff] over Z by use of explicit formulas.
Various attempts were made to generalize the correspondence to Dedekind do-
mains, with the thorny issue being how to deal with a nontrivial class group. Eichler
[Eic53, §14, p. 96] gave such an extension. Peters [Pet69, §4] noted that Eichler’s cor-
respondence was not onto due to class group issues, and he gave a rescaled version that
gives a bijection for Gorenstein orders. Eichler’s correspondence was further tweaked
by Nipp [Nip74, §3], who opted for a different scaling factor that is not restricted to a
class of orders, but his correspondence fails to be onto [Nip74, p.536].
These correspondences were developed further by Brzezinski [Brz80, §3], [Brz85,
§3], where he connected the structure of orders to relatively minimal models of the cor-
responding integral conic; see also Remark 24.3.11. He revisited the correspondence
22.5. TWISTING AND FINAL BIJECTION 361
again in the context of Gorenstein orders [Brz82, §3] and Bass orders [Brz83b, §2].
Lemurell [Lem2011, Theorem 4.3] gives a concise account of the correspondence of
Brzezinski over a PID (the guts of which are contained in [Brz82, (3.2)]).
More recently, Gross–Lucianovic [GrLu2009, §4] revisited the correspondence
over a PID or local ring, and they extended it to include quadratic forms of nonzero
discriminant and without restricting to Gorenstein orders; this extension is impor-
tant for automorphic reasons, connected to Fourier coefficients of modular forms
on PGSp(6), as developed by Lucianovic in his thesis [Luc2003]. Balaji [Bal2007,
Theorem 3.1] studied degenerations of ternary quadratic modules in the context of
orthogonal groups and Witt invariants and showed that the even Clifford functor is
bijective over a general scheme. Finally, Voight [Voi2011a, Theorem B] gave a gen-
eral and functorial correspondence without any of the above restrictive hypotheses,
including the functorial inverse to the even Clifford functor provided above.
Remark 22.5.14. In the most general formulation of the correspondence, allowing arbi-
trary ternary quadratic modules discriminant over all sorts of rings, Voight [Voi2011a,
Theorem A] characterizes the image of the even Clifford functor, as follows. Let 𝐵
be an 𝑅-algebra that is (faithfully) projective of rank 4 as an 𝑅-module. Then 𝐵 is
a quaternion ring if 𝐵 ' Clf 0 (𝑄) for a ternary quadratic module 𝑄. Then 𝐵 is a
quaternion ring if and only if 𝐵 has a standard involution and for all 𝑥 ∈ 𝐵, the trace
of left (or right) multiplication by 𝑥 on 𝐵 is equal to 2 trd(𝑥).
For example, if we take the quadratic form 𝑄 : 𝑅 3 → 𝑅 defined by 𝑄(𝑥, 𝑦, 𝑧) = 0
identically, the multiplication table on Clf 0 (𝑄) gives the commutative ring
Clf 0 (𝑄) ' 𝑅[𝑖, 𝑗, 𝑘]/(𝑖, 𝑗, 𝑘) 2 .
One can see this as a kind of deformation of a quaternion algebra (in an algebro-
geometric sense), letting 𝑎, 𝑏 → 0.
Exercises
I 1. Let 𝑅 be a PID or local noetherian domain. Let 𝐴 be an 𝑅-algebra that is free
of finite rank as an 𝑅-module. Show that 𝐴 has an 𝑅-basis including 1. [Hint:
show that the quotient 𝐴/𝑅 is torsion-free, hence free; since free modules are
projective, the sequence 0 → 𝑅 → 𝐴 → 𝐴/𝑅 → 0 splits, giving 𝐴 ' 𝑅 ⊕ 𝐴/𝑅.]
I 2. For a free quadratic ternary form (as in 22.3.4), show that a change of basis
𝜌 ∈ GL3 (𝑅) acts on 𝑖, 𝑗, 𝑘 ∈ Clf 0 (𝑄) by the adjugate matrix adj(𝜌) ∈ GL3 (𝑅)
(where the entries of adj(𝜌) are the 2 × 2-cofactors of 𝜌 and 𝜌 adj(𝜌) = det(𝜌)).
3. Let 𝑅 be a domain and let Pic 𝑅 be the group of isomorphism classes of invertible
𝑅-modules (equivalently, classes of fractional 𝑅-ideals in 𝐹). Show that up to
twisted similarity, the target of a quadratic module only depends on its class in
Pic 𝑅/2 Pic 𝑅. [See Example 9.7.5.]
4. Finish the direct verification in Example 22.3.24 that O is closed under multi-
plication.
I 5. Let 𝑅 be a Dedekind domain, and let O be a quaternion 𝑅-order. Show that
there exist 𝑖, 𝑗, 𝑘 ∈ O and 𝔞, 𝔟, 𝔠 ⊂ 𝐹 fractional 𝑅-ideals such that O = 𝑅 + 𝔞𝑖 +
362 CHAPTER 22. TERNARY QUADRATIC FORMS
𝔟 𝑗 + 𝔠𝑘 and such that 1, 𝑖, 𝑗, 𝑘 satisfy the multiplication rules (22.3.7) for some
𝑎, 𝑏, 𝑐, 𝑢, 𝑣, 𝑤 ∈ 𝐹—called a good pseudobasis for O. [Hint: revisit what goes
into 22.4.7; a simple observation will suffice!]
6. Let 𝑄 : 𝑅 3 → 𝑅 be a ternary quadratic form and let O := Clf 0 (𝑄). Show that
O = Z + Z𝑝𝑖 + Z 𝑗 + Z𝑖 𝑗
for an odd prime 𝑝 (that is 𝑝𝑖, not 𝜋!). Let 𝑆 = Z[ 𝑝𝑖] ⊆ O. Show that O
is not projective as a left 𝑆-module.
(c) Show that the property that O is projective as an 𝑆-module is a local
property (over primes of 𝑅).
(d) In light of (c), suppose that 𝑅 is a PID, and write O in a good basis (22.3.7).
Suppose that 𝑆 = 𝑅[𝑖] with 𝑖 2 = 𝑢𝑖 − 𝑏𝑐. Show that O is projective as an
𝑆-module if and only if the quadratic form 𝑏𝑥 2 + 𝑢𝑥𝑦 + 𝑐𝑦 2 represents a
unit.
(e) Using (d), conclude in general that if 𝑆 has conductor coprime to discrd O,
show that O is projective as an 𝑆-module.
12. Let 𝑅 be a global ring with 𝐹 := Frac 𝑅, let 𝐵, 𝐵 0 be quaternion algebras over
𝐹, and let O ⊆ 𝐵 and O0 ⊆ 𝐵 0 be 𝑅-orders. Consider the quaternary quadratic
forms 𝑄 := nrd | O : O → 𝑅 and similarly 𝑄 0 on O0.
(a) Show that 𝑄 is isometric to 𝑄 0, then there is an isomorphism of 𝐹-algebras
∼ 𝐵 0.
𝐵 −→
22.5. TWISTING AND FINAL BIJECTION 363
13. Let 𝑄 : 𝑀 → 𝐿 be a quadratic module. Show that the even Clifford algebra
Clf 0 (𝑄) with its map 𝜄 : 𝑀 ⊗ 𝑀 ⊗ 𝐿 ∨ → Clf 0 (𝑄) has the following universal
property: if 𝐴 is an 𝑅-algebra and 𝜄 𝐴 : 𝑀 ⊗ 𝑀 ⊗ 𝐿 → 𝐴 is an 𝑅-module
homomorphism such that
(i) 𝜄 𝐴 (𝑥 ⊗ 𝑥 ⊗ 𝑓 ) = 𝑓 (𝑄(𝑥)) for all 𝑥 ∈ 𝑀 and 𝑓 ∈ 𝐿 ∨ , and
(ii) 𝜄 𝐴 (𝑥 ⊗ 𝑦 ⊗ 𝑓 )𝜄 𝐴 (𝑦 ⊗ 𝑧 ⊗ 𝑔) = 𝑓 (𝑄(𝑦))𝜄 𝐴 (𝑥 ⊗ 𝑧 ⊗ 𝑔) for all 𝑥, 𝑦, 𝑧 ∈ 𝑀 and
𝑓 , 𝑔 ∈ 𝐿∨,
then there exists a unique 𝑅-algebra homomorphism 𝜙 : Clf 0 (𝑄) → 𝐴 such that
the diagram
𝑀 ⊗ 𝑀 ⊗ 𝐿∨
𝜄 / Clf (𝑄)
𝜙
&
𝜄𝐴
𝐴
commutes. Conclude that the pair (Clf 0 (𝑄), 𝜄) is unique up to unique isomor-
phism.
Chapter 23
Quaternion orders
23.1.1 (Maximal orders). The nicest orders are undoubtedly the maximal orders, those
not properly contained in another order. An order is maximal if and only if it is locally
maximal (Lemma 10.4.3), i.e. 𝑝-maximal for all primes 𝑝; globally, an order O is
maximal if and only if 𝑁 = 𝐷 (i.e., 𝑀 = 1).
We have either 𝐵 ' M2 (Q 𝑝 ) or 𝐵 is a division algebra over Q 𝑝 (unique up
to isomorphism). If 𝐵 is split, then a maximal order is isomorphic (conjugate) to
M2 (Z 𝑝 ), and the corresponding ternary quadratic form is the determinant 𝑥𝑦 − 𝑧 2
(see Example 22.3.12). If instead 𝐵 is division, then the unique maximal order is the
valuation ring, with corresponding anisotropic form 𝑥 2 − 𝑒𝑦 2 + 𝑝𝑧 2 for 𝑝 ≠ 2, where
𝑒 ∈ Z is a quadratic nonresidue modulo 𝑝 (and for 𝑝 = 2, the associated form is
𝑥 2 + 𝑥𝑦 + 𝑦 2 + 2𝑧2 ).
Maximal orders have modules with good structural properties: all lattices 𝐼 ⊂ 𝐵
with left or right order equal to a maximal order O are invertible (Theorem 18.1.2).
There is a combinatorial structure, called the Bruhat–Tits tree, that classifies
maximal orders in M2 (Q 𝑝 ) (as endomorphism rings of lattices, up to scaling): the
Bruhat–Tits tree is a 𝑝 + 1-regular tree (see section 23.5).
365
366 CHAPTER 23. QUATERNION ORDERS
positive integer is the sum of four squares—is properly contained inside the Hurwitz
order (Chapter 11).
23.1.2 (Hereditary orders). More generally, we say that O is hereditary if every left
or right fractional O-ideal (i.e., lattice 𝐼 ⊆ 𝐵 with left or right order containing O) is
invertible. Maximal orders are hereditary, and being hereditary is a local property. A
hereditary Z 𝑝 -order O 𝑝 ⊆ 𝐵 𝑝 is either maximal or
Z𝑝 Z𝑝 𝑎 𝑏
O𝑝 ' = : 𝑎, 𝑏, 𝑐, 𝑑 ∈ Z 𝑝 ⊆ M2 (Q 𝑝 ) ' 𝐵 𝑝
𝑝Z 𝑝 Z 𝑝 𝑝𝑐 𝑑
with associated ternary quadratic form 𝑥𝑦 − 𝑝𝑧2 . Thus O ⊂ 𝐵 is hereditary if and only
if discrd(O) = 𝐷 𝑀 is squarefree, so in particular gcd(𝐷, 𝑀) = 1.
Hereditary orders share the nice structural property of maximal orders: all lattices
𝐼 ⊂ 𝐵 with hereditary left or right order are invertible. The different ideal diff O 𝑝 is
generated by any element 𝜇 ∈ O 𝑝 such that 𝜇2 ∈ 𝑝Z 𝑝 .
23.1.3 (Eichler orders). More generally, we can consider orders that are “upper trian-
gular modulo 𝑀” with gcd(𝐷, 𝑀) = 1 (i.e., avoiding primes that ramify in 𝐵). The
order
Z𝑝 Z𝑝
⊆ M2 (Z 𝑝 )
𝑝𝑒 Z 𝑝 Z 𝑝
is called the standard Eichler order of level 𝑝 𝑒 in M2 (Q 𝑝 ). A Z 𝑝 -order O 𝑝 ⊆
M2 (Q 𝑝 ) is an Eichler order if O 𝑝 is isomorphic to a standard Eichler order. The
ternary quadratic form associated to an Eichler order of level 𝑝 𝑒 is 𝑥𝑦 − 𝑝 𝑒 𝑧 2 .
Globally, we say O ⊂ 𝐵 is a Eichler order of level 𝑀 if discrd(O) = 𝑁 = 𝐷 𝑀
with gcd(𝐷, 𝑀) = 1 and O 𝑝 is an Eichler order of level 𝑝 𝑒 for all 𝑝 𝑒 k 𝑀. In
particular, O 𝑝 is maximal at all primes 𝑝 | 𝐷. Every hereditary order is Eichler, and
an Eichler order is hereditary if and only if its level 𝑀 (or reduced discriminant 𝑁) is
squarefree. A maximal Z 𝑝 -order O 𝑝 ⊆ M2 (Z 𝑝 ) is an Eichler order of level 1 = 𝑝 0 .
Eichler orders play a crucial role in the context of modular forms, as we will see in the
final part of this monograph.
This local description of Eichler orders also admits a global description. The
standard Eichler order O 𝑝 can be written
𝑝 −𝑒 Z 𝑝
Z Z𝑝 Z
O𝑝 = 𝑒 𝑝 = M2 (Z 𝑝 ) ∩ 𝑒 𝑝 = M2 (Z 𝑝 ) ∩ 𝜛 −1 M2 (Z 𝑝 )𝜛
𝑝 Z𝑝 Z𝑝 𝑝 Z𝑝 Z𝑝
a generator of the group 𝑁GL2 (Q 𝑝 ) (O 𝑝 )/Q×𝑝 O×𝑝 ' Z/2Z, and 𝜛 2 = 𝑝 𝑒 . The different
diff O 𝑝 is the two-sided ideal generated by 𝜛.
From the local-global dictionary, it follows that O ⊂ 𝐵 is Eichler if and only if O
is the intersection of two (not necessarily distinct) maximal orders.
23.2. MAXIMAL ORDERS 367
23.2.1. We make the following convention. When we say “𝑅 is local”, we mean that
𝑅 is a complete DVR, and in this setting we let 𝔭 = 𝜋𝑅 be its maximal ideal, and
𝑘 = 𝑅/𝔭 the residue field. When we want to return to the general context, we will say
“𝑅 is Dedekind”.
23.2.2. Being maximal is a local property (Lemma 10.4.3), so the following are
equivalent:
We recall (Theorem 15.5.5) that an order over a global ring 𝑅 is maximal if and
only if discrd(O) = disc𝑅 (𝐵). Local maximal 𝑅-orders have the nice local description.
23.2.3. Suppose 𝑅 is local and that 𝐵 ' M2 (𝐹) is split. Then by Corollary 10.5.5,
every maximal 𝑅-order in M2 (𝐹) is conjugate to M2 (𝑅) by an element of GL2 (𝑅),
i.e. O ' M2 (𝑅). We have discrd(O) = 𝑅. All two-sided ideals of O are powers of
rad(O) = 𝔭O, and
O/rad(O) ' M2 (𝑘).
The associated ternary quadratic form is similar to 𝑄(𝑥, 𝑦, 𝑧) = 𝑥𝑦 − 𝑧2 , by Example
22.3.12 and the classification theorem (Main Theorem 22.1.1). Finally,
23.2.5. Suppose 𝑅 is local but now that 𝐵 is a division algebra. Then the valuation
ring O ⊂ 𝐵 is the unique maximal 𝑅-order by Proposition 13.3.4.
Suppose further that the residue field 𝑘 is finite (equivalently, that 𝐹 is a local
field). Then Theorem 13.3.11 applies, and we have
𝐾, 𝜋
O ' 𝑆 ⊕ 𝑆𝑗 ⊆
𝐹
𝑖 2 = 𝑢𝑖 − 𝑏 𝑗 𝑘 = −𝜋𝑖
2
𝑗 =𝜋 𝑘𝑖 = 𝑏 𝑗 (23.2.6)
2
𝑘 = 𝑏𝜋 𝑖𝑗 = 𝑘
realizing the basis as a good basis in the sense of 22.4.7; the associated ternary
quadratic form is
Finally, since the valuation ring is the unique maximal order and conjugation
respects integrality, we have
𝑁 𝐵× (O) = 𝐵× . (23.2.8)
23.2.9. All lattices 𝐼 ⊂ 𝐵 with left or right order equal to a maximal order O are
invertible, by Theorem 18.1.2, proven in Proposition 18.3.2. (We also gave a different
proof of this fact in Proposition 16.6.15(b).)
The classification of two-sided ideals and their classes follows from that of hered-
itary orders: see 23.3.19.
23.3.1. Suppose 𝑅 is local. By Main Theorem 21.1.4 and Corollary 21.1.5, the
following are equivalent:
(i) O is hereditary;
(ii) rad O is projective as a left (or right) O-module;
(iii) OL (rad O) = OR (rad O) = O;
(iv) rad O is invertible as a (sated) two-sided O-ideal; and
23.3. HEREDITARY ORDERS 369
[O : 𝐼] 𝑅 = nrd(𝐼) 2 = nrd(𝛼) 2 𝑅 | [O : 𝐽] 𝑅 = 𝔭2 .
But
𝑅 𝔭 0 1 0 𝑅 𝑅 𝔭
= * ;
𝔭 𝔭 0 0 0 𝔭 𝔭 𝔭
we get a contradiction with the other possibility by multiplying instead on the left. A
third “pure matrix multiplication” proof is also requested in Exercise 23.1.
The second statement was already proven in (23.3.10).
𝐽 = [O, O] (23.3.16)
23.3.17. We now classify the left O-lattices, up to isomorphism. Each such O-lattice
is projective, since O is hereditary.
By the Krull-Schmidt theorem (Theorem 20.6.2), every O-lattice can be written as
the direct sum of indecomposables, so it is enough to classify
the indecomposables;
𝐹
and we did just this in 21.5.3. Explicitly, we have 𝑉 = ' 𝐹 2 the simple 𝐵 =
𝐹
𝑅 𝑅
M2 (𝐹)-module, and we take the O-lattice 𝑀 = ⊂ 𝑉. We have 𝐽 𝑀 = and
𝑅 𝔭
𝔭
𝐽2 𝑀 = = 𝜋𝑀, and 𝑀, 𝐽 𝑀 give a complete set of indecomposable left O-modules.
𝔭
As expected, O = 𝐽 𝑀 ⊕ 𝑀 is a decomposition of O into projective indecomposable
left O-modules.
The preceding local results combine to determine global structure. Now let 𝑅 be
a global ring.
Lemma 23.3.18. O is hereditary if and only if discrd(O) is squarefree.
Proof. We argue locally; and then we use the characterization (iv), the computation
of the reduced discriminant (23.3.2), and the same argument as in Theorem 15.5.5 to
finish.
23.3.19. Let O be a hereditary (possibly maximal) 𝑅-order. By Theorem 21.4.9, we
know that the group Idl(O) is an abelian group generated by the prime (equivalently,
maximal) invertible two-sided ideals. We claim that the map
{Prime two-sided invertible O-ideals} ↔ {Prime ideals of 𝑅}
(23.3.20)
𝑃 ↦→ 𝑃 ∩ 𝑅
is a bijection, generalizing Theorem 18.3.6. If 𝔭 - 𝔑 then we have 𝑃 = 𝔭O; and
if 𝔭 | 𝔇 then we have a prime two-sided ideal 𝑃 = O ∩ rad(O𝔭 ) with 𝑃2 = 𝔭O.
Otherwise, 𝔭 | 𝔑 but 𝔭 - 𝔇, so O𝔭 is hereditary but not maximal; from the local
description in Lemma 23.3.13, we get a prime ideal 𝑃 = O ∩ rad(O𝔭 ) with 𝑃2 = 𝔭O
as in the ramified case. This proves (23.3.20), and that the sequence
Ö
0 → Idl(𝑅) → Idl(O) → Z/2Z → 0 (23.3.21)
𝔭 |𝔑
is exact.
Taking the quotient by PIdl(𝑅), we obtain the exact sequence
Ö
0 → Cl 𝑅 → Pic𝑅 (O) → Z/2Z → 0. (23.3.22)
𝔭 |𝔑
In particular, if 𝔇 = (1), then Pic𝑅 (O) ' Cl 𝑅. Finally, the group of two-sided ideals
modulo principal two-sided ideals is related to the Picard group by the exact sequence
(18.5.5):
0 → 𝑁 𝐵× (O)/(𝐹 × O× ) → Pic𝑅 (O) → Idl(O)/PIdl(O) → 0
𝛼𝐹 × O× ↦→ [𝛼O] = [O𝛼]
372 CHAPTER 23. QUATERNION ORDERS
(This exact sequence is sensitive to O even within its genus: see Remark 18.5.9.)
Proof. We follow Hijikata [Hij74, 2.2(i)]. Apologies in advance for all of the explicit
matrix multiplication!
We prove (i) ⇒ (ii) ⇔ (iii) and then (ii) ⇒ (iv) ⇒ (i). The implications (ii) ⇒
(iii) and (iv) ⇒ (i) are immediate.
So first (i) ⇒ (ii). Suppose O = O1 ∩O2 . All maximal orders in 𝐵 are 𝐵× -conjugate
to M2 (𝑅), so there exist 𝛼1 , 𝛼2 ∈ 𝐵× such that O𝑖 = 𝛼𝑖−1 M2 (𝑅)𝛼𝑖 for 𝑖 = 1, 2.
Conjugating by 𝛼1 , we may suppose 𝛼1 = 1 and we write 𝛼 = 𝛼2−1 for convenience, so
O ' M2 (𝑅) ∩ 𝛼 M2 (𝑅)𝛼−1 . Scaling by 𝜋, we may suppose 𝛼 ∈ M2 (𝑅) r 𝜋 M2 (𝑅).
By row and column operations (Smith normal form, proven as part of the structure
theorem for finitely generated modules over a PID), there exist 𝛽, 𝛾 ∈ GL2 (𝑅) such
that
1 0
𝛽𝛼𝛾 =
0 𝜋𝑒
is in standard invariant form with 𝑒 ≥ 0. Then
𝔭−𝑒
1 0 𝑅 𝑅 1 0 𝑅
𝛽𝛼 M2 (𝑅)𝛼−1 𝛽−1 = = 𝑒 (23.4.5)
0 𝜋𝑒 𝑅 𝑅 0 𝜋 −𝑒 𝔭 𝑅
so
𝔭−𝑒
𝑅 𝑅 𝑅 𝑅 𝑅
O' ∩ 𝑒 = 𝑒 . (23.4.6)
𝑅 𝑅 𝔭 𝑅 𝔭 𝑅
23.4. EICHLER ORDERS 373
𝑅 0
To show (iii) ⇒ (ii), we may suppose O ⊇ = 𝑅𝑒 11 + 𝑅𝑒 22 with 𝑒 11 =
0 𝑅
1 0 0 0 𝐹 0
, 𝑒 22 = . Then O𝑒 11 ⊆ . Let 𝜋𝑖 𝑗 be the projection onto the
0 0 0 1 𝐹 0
𝑖 𝑗-coordinate. Then
𝑅 0
𝜋11 (O𝑒 11 ) = trd(O𝑒 11 )𝑒 11 ⊆ 𝑅 =
0 0
for some 𝑎 ∈ Z. Arguing again with the other matrix unit 𝑒 22 we conclude that
𝑅 𝔭𝑏
O= 𝑎 (23.4.7)
𝔭 𝑅
with 𝑎, 𝑏 ∈ Z. Multiplying
𝑅 𝔭𝑏 𝑅 𝔭𝑏 𝑅 + 𝔭𝑎+𝑏 𝔭𝑏
=
𝔭𝑎 𝑅 𝔭𝑎 𝑅 𝔭𝑎 𝑅 + 𝔭𝑎+𝑏
Proof. The corollary is local, so we may apply Proposition 23.4.3(iii) to every super-
order.
23.4.12. First two basic facts about the Eichler order O of level 𝔭𝑒 : We have
23.4.13. Let
0 1
𝜛= ∈ O.
𝜋𝑒 0
Then
O = M2 (𝑅) ∩ 𝜛 −1 M2 (𝑅)𝜛
as in (23.4.6); by Proposition 23.4.3 these two orders are the uniquely determined pair
of maximal orders containing O. We have 𝜛 2 = 𝜋 𝑒 , and so 𝜛 ∈ 𝑁 𝐵× (O). It follows
(and can be checked directly) that 𝐼 = O𝜛 = 𝜛O is a two-sided O-ideal. If 𝑒 = 0, we
have 𝐼 = O.
where 𝑓 = min(𝑒, 2), so by Corollary 20.5.5, 𝐽 ⊆ rad O; on the other hand, the
quotient
O/𝐽 ' 𝑘 × 𝑘 (23.4.17)
is semisimple, so rad O ⊆ 𝐽 by Corollary 20.4.11(a).
However, the radical is not an invertible (sated) two-sided O-ideal unless O is
hereditary (𝑒 = 1), by 23.3.1. Indeed, we verify that
𝔭−1
𝑅
OL (𝐽) = 𝑒−1 = OR (𝐽) (23.4.18)
𝔭 𝑅
23.4.19. Suppose that 𝑅 is a global ring. Let disc𝑅 𝐵 = 𝔇 and let O be an Eichler
order with discrd O = 𝔑. If 𝔭 | 𝔇, then 𝐵𝔭 has a unique maximal order, so (as an
‘intersection’) O𝔭 is necessarily the maximal order. If 𝔭 - 𝔇, and ord𝔭 𝔑 = 𝑒 ≥ 0,
then O𝔭 is isomorphic to the standard Eichler order of level 𝔭𝑒 .
We have 𝔑 = 𝔇𝔐 with 𝔐 ⊆ 𝑅 and we just showed that 𝔐 is coprime to 𝔇. We
call 𝔐 the level of the Eichler order O. The pair 𝔇, 𝔐 (or 𝔇, 𝔑) determines a unique
genus of Eichler 𝑅-orders, i.e., this data uniquely determines the isomorphism class
of O𝔭 for each 𝔭.
Putting together Proposition 23.4.14 together with 23.3.19 for the remaining primes
where the order is maximal, we have an exact sequence
Ö
0 → Idl(𝑅) → Idl(O) → Z/2Z → 0 (23.4.20)
𝔭 |𝔑
Remark 23.4.22. Eichler [Eic56a] developed his orders in detail for prime level, and
employed them in the study of modular correspondences [Eic56c] and the trace formula
[Eic73] in the case of squarefree level. (As mentioned in Remark 24.1.5, Eichler’s
376 CHAPTER 23. QUATERNION ORDERS
earlier work [Eic36, §6] over Q included some more general investigations, but his later
work seemed mostly confined to the hereditary orders.) Hijikata [Hij74, §2.2] later
studied these orders in the attempt to generalize Eichler’s result beyond the squarefree
case; calling the orders split (like our name, residually split). Pizer [Piz73, p. 77] may
be the first who explicitly called them Eichler orders.
Remark 23.4.23. An 𝑅-order O ⊆ M𝑛 (𝐹) that contains a GL𝑛 (𝐹)-conjugate of the
diagonal matrices diag(𝑅, . . . , 𝑅) (equivalently, containing 𝑛 orthogonal idempotents)
is said to be tiled. By 23.4.3, an order O ⊆ M2 (𝐹) is tiled if and only if it is
Eichler. Tiled orders also go by other names (including graduated orders) and they
arise naturally in many contexts, including representation theory [Pl83] and modular
forms.
Lemma 23.5.2. We have End𝑅 (𝐿) = End𝑅 (𝑀) if and only if there exists 𝑎 ∈ 𝐹 × such
that 𝑀 = 𝑎𝐿.
With this lemma in mind, we make the following definition (recalling this definition
made earlier in the context of algebras).
𝜋𝐿 ( 𝐿 0 ( 𝐿. (23.5.6)
Proof. We have 𝐿/𝜋𝐿 ' 𝑘 2 , and so the lattices 𝐿 0 satisfying (23.5.6) are in bijection
with 𝑘-subspaces of dimension 1 in 𝐿/𝜋𝐿; such a subspace is given by a choice of
generator up to scaling, so there are exactly (𝑞 2 − 1)/(𝑞 − 1) = 𝑞 + 1 such, and each
vertex has 𝑞 + 1 adjacent vertices. The graph is connected: given two vertices, we may
choose representative lattices 𝐿, 𝐿 0 such that 𝐿 0 ⊆ 𝐿 as in 23.5.4. The quotient 𝐿/𝐿 0
is cyclic, so by induction the lattices 𝐿 𝑖 = 𝜋 𝑖 𝐿 + 𝐿 0 for 𝑖 = 0, . . . , 𝑓 have 𝐿 𝑖 adjacent
to 𝐿 𝑖+1 , and 𝐿 0 = 𝐿 and 𝐿 𝑓 = 𝐿 0, giving a path from [𝐿] to [𝐿 0].
The following argument comes from Dasgupta–Teitelbaum [DT2008, Proposition
1.3.2]. Suppose there is a nontrivial cycle in T
𝜋 𝑣 𝐿 = 𝐿 𝑠 ( 𝐿 𝑠−1 ( · · · ( 𝐿 1 ( 𝐿 0 = 𝐿 (23.5.9)
If 𝛼 ∈ GL2 (𝐹), then multiplying (23.5.13) on the left by 𝛼 shows that 𝛼𝐿, 𝛼𝐿 0 are
adjacent.
23.5.14. The tree T has a natural notion of distance 𝑑 between two vertices, given
by the length of the shortest path between them, giving each edge of T length 1.
Consequently, we have a notion of distance 𝑑 (O, O0) between every two maximal
orders O, O0 ⊆ 𝐵.
Proof. The statement on endomorphism rings comes from Example 10.5.2; we may
suppose up to homothety that 𝐿 0 has basis 𝑥1 , 𝜋 𝑒 𝑥2 ; the maximal lattices as in 23.5.4 are
given by 𝐿 𝑖 = 𝑅𝑥 1 + 𝔭𝑖 𝑅𝑥 2 with 𝑖 = 0, . . . , 𝑒, so the distance is 𝑑 ( [𝐿], [𝐿 0]) = 𝑒.
23.5.16. Importantly, now, we turn to Eichler orders: they are the intersection of two
unique maximal orders, and so correspond to a pair of vertices in T , or equivalently a
path. By Lemma 23.5.15, the standard Eichler order of level 𝔭𝑒 corresponds to a path
of length 𝑒, and by transitivity the same is true of every Eichler order. The normalizer
𝜛 of an Eichler order 23.4.13 acts by swapping the two vertices. Each vertex of the
path corresponds to the 𝑒 + 1 possible maximal superorders.
In this way, the Bruhat–Tits tree provides a visual way to keep track of many
calculations with Eichler orders.
Exercises
Unless otherwise specified, let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅 and let
O ⊆ 𝐵 be an 𝑅-order in a quaternion algebra 𝐵.
0 1
(c) Conclude that 𝑁 𝐵× (O)/(𝐹 × O× ) is generated by 𝜛 = .
𝜋 0
[See also the matrix proof by Eichler [Eic56a, Satz 5], which uses a normal form
for one-sided ideals simplifying the above computation.]
I 2. Let 𝐾 ⊇ 𝐹 be a field extension, let 𝑆 be the integral closure of 𝑅 in 𝐾, and let
𝔞, 𝔟 ⊂ 𝐹 be fractional ideals of 𝑅. Show that 𝔞 = 𝔟 if and only if 𝔞𝑆 = 𝔟𝑆.
I 3. Extend Lemma 13.4.9, and show that if 𝑅 is a DVR and O is an Eichler 𝑅-order
then nrd(O× ) = 𝑅 × .
I 4. Suppose 𝑅 is local and O is a hereditary order. Show that if 𝜇 ∈ O has 𝜇2 = 𝜋,
then 𝜇 generates 𝑁 𝐵× (O)/(𝐹 × O× ).
5. Suppose 𝑅 is local and O ⊆ 𝐵 = M2 (𝐹) is the intersection of two maximal
orders. Give another (independent) proof that O is isomorphic to a standard
Eichler order which replaces matrix calculations in Proposition 23.4.3 with some
representation theory as follows.
(a) Write O ' M2 (𝑅) ∩ O0. Let 𝑒 11 be the top-left matrix unit and let
𝐼 = M2 (𝑅)𝑒 11 . Show 𝐼 0 = O0 𝑒 11 is an 𝑅-lattice in 𝑉 = M2 (𝐹)𝑒 11 ' 𝐹 2 .
(b) Use elementary divisors to show that there exists an 𝑅-basis 𝑥1 , 𝑥2 of 𝐼
such that 𝑥1 , 𝜋 𝑒 𝑥2 is an 𝑅-basis for 𝐼 0.
1 0
(c) Show that the corresponding change of basis matrix 𝛼 = has
0 𝜋𝑒
0
𝐼 = 𝛼𝐼, and use this to identify O with the standard Eichler order of level
𝔭𝑒 .
[See Brzezinski [Brz83a, Proposition 2.1].]
I 6. Let 𝑅 be local and let O be the standard Eichler order of level 𝔭𝑒 for 𝑒 ≥ 1. Let
𝐽 = rad O. Show that
𝔭−1
𝑅
OL (𝐽) = 𝑒−1 = OR (𝐽).
𝔭 𝑅
I 7. Prove Lemma 23.5.1. [Hint: use direct matrix methods or the theory of invariant
factors.]
8. Let 𝑅 be local and let O be a hereditary quaternion 𝑅-order. Show that rad O =
[O, O] is the commutator (cf. Exercise 13.8) and that diff O = rad O.
9. Let 𝑅 be local. Let O, O0 ⊆ 𝐵 be maximal 𝑅-orders. Recall that O, O0 are
vertices in the Bruhat–Tits tree. Define the distance dist(O, O0) to be the
distance in the Bruhat–Tits tree between the respective vertices. Show that
[O : O ∩ O0] = dist(O, O0) = [O0 : O ∩ O0].
10. Let 𝑅 be local, and let O be an Eichler order of level 𝔭𝑒 . Consider the graph
whose vertices are 𝑅-superorders O0 ⊇ O in 𝐵 and with a directed edge whenever
the containment O0 ) O is proper and minimal. What does this graph look like?
[Hint: use the Bruhat–Tits tree; it helps to draw the Eichler orders of the same
level at the same height.]
Chapter 24
In this chapter, we continue our tour of quaternion orders with some more advanced
species.
24.1.1 (Gorenstein and Bass orders). Although Eichler orders may lose the property
that all of its ideals are invertible, we may still insist on the invertibility of its dual.
Recall (Definition 15.6.15) that the codifferent of an order is the lattice codiff(O) = O♯
obtained as the dual of the trace pairing over 𝑅. We say O is Gorenstein if codiff(O)
is invertible, or equivalently (Corollary 16.8.7) every sated left or right fractional O-
ideal (lattice 𝐼 ⊆ 𝐵 with left or right order equal to O) is invertible. Hereditary orders
are Gorenstein, since for a hereditary order every left or right fractional O-ideal (not
necessarily sated) is invertible.
Being Gorenstein is a local property because invertibility is so. An Eichler order is
Gorenstein, but there are Gorenstein orders that are not Eichler. An order is Gorenstein
if and only if its associated ternary quadratic form is primitive, i.e. the greatest common
divisor of its coefficients is 1, or equivalently its values generate Z.
We say O is Bass if every superorder O0 ⊇ O (including O0 = O) is Gorenstein. A
Bass order is Gorenstein, but not always conversely. The fact that every superorder is
Gorenstein reflects into good structural properties of a Bass order. Most importantly,
a Z 𝑝 -order O is Bass if and only if it contains either Z 𝑝 × Z 𝑝 or the ring of integers
in a quadratic extension 𝐾 ⊇ Q 𝑝 (these order are sometimes called primitive; we call
them basic). This embedded subalgebra makes it possible to calculate explicitly with
the order, with important applications to the arithmetic of modular forms, a topic we
pursue in the final part of this book.
381
382 CHAPTER 24. QUATERNION ORDERS: SECOND MEETING
Proof. For the equivalence (i) ⇔ (ii), because codiff (O) is sated it follows from
Theorem 20.3.3 that O is Gorenstein if and only if codiff (O) is projective as an O-
bimodule. For (ii) ⇔ (iii), by Proposition 15.6.7, we have an isomorphism codiff (O) '
Hom𝑅 (O, 𝑅) of O-bimodules over 𝑅. Finally, (i) ⇔ (iv) follows from Corollary
16.8.7.
24.2.4. We call in for relief as well Main Theorem 20.3.9: the equivalent sided notions
(on the left and right) in Proposition 24.2.3 are also all equivalent. In particular, a
suitably defined notion of left Gorenstein or right Gorenstein would also be equivalent.
The Gorenstein condition can be detected on the level of norms as follows.
Lemma 24.2.5. We have
[O : O♯ ] 𝑅 ⊇ Nm 𝐵 |𝐹 (O♯ ) = nrd(O♯ ) 2 ,
with equality if and only if O♯ is locally principal. But by Lemma 15.6.17, we have
[O♯ : O] 𝑅 = disc(O) = discrd(O) 2 , and combining these gives the result.
Our next main result connects the Gorenstein condition to a property of the corre-
sponding ternary quadratic module. Let 𝑄 : 𝑀 → 𝐿 be a ternary quadratic module.
We follow Gross–Lucianovic [GrLu2009, Propositions 6.1–6.2], and consider the odd
Clifford bimodule.
Proposition 24.2.6. Left multiplication gives a pairing
𝑒 1 𝑒 2 𝑒 3 − 𝑢𝑒 1 − 𝑣𝑒 2 − 𝑤𝑒 3 , 𝑒 1 , 𝑒 2 , 𝑒 3 ; (24.2.9)
for example,
so
𝑒 2 𝑒 3 (𝑒 1 𝑒 2 𝑒 3 − 𝑢𝑒 1 − 𝑣𝑒 2 − 𝑤𝑒 3 ) ≡ 𝑒 1 (𝑒 2 𝑒 3 ) 2 − 𝑢𝑒 1 𝑒 2 𝑒 3
≡ 𝑒 1 (𝑢𝑒 2 𝑒 3 ) − 𝑢𝑒 1 𝑒 2 𝑒 3 = 0 (mod 𝑀).
Before we begin the proof, we make a definition and then consider the key ingredient
of the proof: how rescaling the module affects the even Clifford algebra.
Definition 24.2.17. In (24.2.16), we call 𝔣(O) ⊆ 𝑅 the conductor and Gor(O) the
Gorenstein saturation of O.
Remark 24.2.18. Brzezinski [Brz83a] writes 𝔟(O) instead of 𝔣(O) and calls it the
Brandt invariant. Other authors call Gor(O) the Gorenstein closure of O, but this
terminology may be confusing: Gor(O) is not necessarily the smallest Gorenstein
order containing O (see Exercise 24.11).
The fact that Clf 0 (𝑄 [𝔞]) = 𝑅 + 𝔞O is visible from the construction of the even Clifford
algebra (22.2.3); it is also visible from the description (22.3.25) in Example 22.3.24.
In the other direction, we have
Proof of Proposition 24.2.15. We argue using ternary quadratic modules: our proof
amounts to replacing a potentially imprimitive form with a primitive form, following
Theorem 24.2.10.
Let 𝑄 = 𝜓O : 𝑀 → 𝐿 be the ternary quadratic module associated to O, well-
defined up to twisted similarity. We may take 𝐿 = 𝔩 ⊆ 𝐿 ⊗𝑅 𝐹 ' 𝐹 and we do so for
concreteness; accordingly, suppose 𝔩 is a fractional ideal of 𝑅. Then 𝑄(𝑀) = 𝔫 ⊆ 𝔩 is a
finitely generated nonzero 𝑅-submodule; since 𝑅 is a Dedekind domain, 𝔫 is invertible.
Let
𝔣 = 𝔣(O) := 𝔫𝔩−1 ⊆ 𝑅.
Let Gor(𝑄) = 𝑄(𝔣) : 𝑀 → 𝔫 be the primitive ternary quadratic module obtained
by restricting the codomain. Then Clf 0 (Gor(𝑄)) is a Gorenstein order by Theorem
24.2.10, and
Clf 0 (𝑄) = 𝑅 + 𝔣(O) Clf 0 (Gor(𝑄));
by (24.2.20), so we let Gor(O) := Clf 0 (Gor(𝑄)). Uniqueness follows directly from
(24.2.20): if O = 𝑅 + 𝔞O0 and O0 is Gorenstein, then 𝑄(𝔞−1 ) : 𝑀 → 𝔞𝔩 is primitive
and 𝑄(𝑀) = 𝔫 = 𝔞𝔩, thus 𝔞 = 𝔣(O) and O0 = Clf 0 (𝑄(𝔞)) = Gor(O).
To prove the remaining statements, we recall Corollary 22.4.15 to get
O = 𝑅 + discrd(O)O♯ O♯
24.2. GORENSTEIN ORDERS 387
as in (24.2.21) is equivalent to
nrd((O♯ ) 0 ) = 𝔣 discrd(O) −1
Proposition 24.2.23. Let 𝐾 ⊇ 𝐹 be a finite extension and let 𝑆 be the integral closure
of 𝑅 in 𝐾. Then O is a Gorenstein 𝑅-order if and only if O ⊗𝑅 𝑆 is a Gorenstein
𝑆-order.
Δ:𝐵→𝐹
(24.3.4)
Δ(𝛼) = trd(𝛼) 2 − 4 nrd(𝛼)
that computes the discriminant of 𝐹 [𝛼] = 𝐹 [𝑥]/(𝑥 2 − trd(𝑥)𝑥 + nrd(𝑥)) in the basis
1, 𝛼. The form factors through a map Δ : 𝐵/𝐹 → 𝐹.
𝑎
Suppose that #𝑘 < ∞. For 𝑎 ∈ 𝑅, let denote the generalized Kronecker sym-
𝔭
bol, defined to be 0, 1, −1 according as if 𝐹 [𝑥]/(𝑥 2 − 𝑎) is ramified, split (isomorphic
𝑎
to 𝐹 × 𝐹), or inert. If char 𝑘 ≠ 2, then is the Legendre symbol. We then have the
𝔭
following characterization (Exercise 24.4):
O Δ(𝛼)
(a) = ∗ if and only if takes on all of the values −1, 0, 1 for 𝛼 ∈ O.
𝔭 𝔭
O Δ(𝛼)
(b) = 𝜖 if and only if takes the values {0, 𝜖 } for 𝛼 ∈ O.
𝔭 𝔭
24.3. EICHLER SYMBOL 389
Lemma 24.3.6. The order O is Eichler if and only if O is maximal or residually split.
24.3.7. Residually inert orders in a division quaternion algebra 𝐵 over Q 𝑝 were studied
by Pizer [Piz76b], and he described them as follows. Let 𝐾 = Q 𝑝2 be the unramified
extension of Q 𝑝 . Then 𝐾 ↩→ 𝐵. Consider the left regular representation 𝐵 → M2 (𝐾);
it has image
𝑧 𝑤
, with 𝑧, 𝑤 ∈ 𝐾 (24.3.8)
𝑝𝑤 𝑧
The valuation ring of 𝐵 consists of those with 𝑧, 𝑤 ∈ Z 𝑝2 where Z 𝑝2 is the valuation
ring of 𝐾. Pizer then considers those orders with 𝑧 ∈ Z 𝑝2 and 𝑝 𝑟 | 𝑤. He [Piz80a,
Remark 1.5, Proposition 1.6] connected the residually inert and residually split orders
by noting the striking resemblance between (24.3.8) and the standard Eichler order,
remarking:
Thus O0𝑝 and O 𝑝 [the Eichler order and the Pizer order] are both essentially
𝑅 𝑅
subrings of 2𝑟 +1 fixed by certain (different!) Galois actions induced
𝑝 𝑅
by the Galois group of 𝐿/Q 𝑝 and thus they can be viewed as twisted
versions of each other. Hence O and O0 are locally isomorphic at all
primes 𝑞 ≠ 𝑝 while at 𝑝 they are almost isomorphic. Thus it should not
be too surprising that there are close connections between [them].
Pizer works explicitly and algorithmically [Piz80a] with residually inert orders, with
applications to computing modular forms of certain nonsquarefree level.
24.3.9. We can also interpret the Eichler symbol in terms of the reduction of the
associated ternary quadratic form 𝑄.
390 CHAPTER 24. QUATERNION ORDERS: SECOND MEETING
24.3.10. It follows from 24.3.9 that the ternary quadratic form associated to a residually
inert order is similar to
just as in (23.2.7); we have 𝑒 odd if and only if 𝐵 is a division algebra and 𝑒 even if
and only if 𝐵 is split.
The residually ramified orders do not admit such a simple classification; we will
pursue them further in the coming sections.
Remark 24.3.11. Let 𝑅 be a DVR, let O be a Gorenstein 𝑅-order, and let 𝑄 be a ternary
quadratic form over 𝑅 representing the similarity class associated to O. Then 𝑄 is
primitive and so defines an integral model C of a conic 𝐶 ⊆ P2 over 𝐹 = Frac 𝑅. By
discriminants, this conic has good reduction if and only if O is maximal. Moreover, we
saw by direct calculation that this conic has the simplest kind of bad reduction—regular,
with just one node over 𝑘—if and only if O is hereditary. This is no coincidence: in
fact, C is normal if and only if C is Bass. See Brzezinski [Brz80] for more on the
relationship between integral models of conics and quaternion orders and the follow-up
work [Brz85] where the increasing sequence of Bass orders ending in an hereditary
order corresponds to a sequence of elementary blowup transformations.
Another motivation to study the Eichler symbol is that it controls the structure of
unit groups, as follows.
𝑞(𝑞 − 1) 2 (𝑞 + 1), if (O | 𝔭) = ∗;
𝑞 2 (𝑞 − 1) 2 ,
if (O | 𝔭) = 1;
[O× : 1 + 𝔭O] = 2 2
𝑞 (𝑞 − 1),
if (O | 𝔭) = −1;
𝑞 3 (𝑞 − 1),
if (O | 𝔭) = 0.
24.4. CHAINS OF ORDERS 391
24.4.3. We recall our motivation to study extremal orders 21.2.1: O♮ radically covers
O, and by Proposition 21.2.3 we have O♮ = O if and only if O is extremal. By Theorem
21.5.1, O is extremal if and only if O is hereditary. Iterating, we have a canonically
associated chain of orders
O = O0 ( O1 = O♮ ( · · · ( O𝑟 (24.4.4)
b𝑒/2c with quotients dim 𝑘 (O𝑖 /O𝑖+1 ) = 2. On the Bruhat–Tits tree, the Eichler order
corresponds to a path of length 𝑒 by 23.5.16, and O♮ is the path of length 𝑒 − 2 obtained
by plucking away the vertices on the ends. (If desired, one can refine this chain by
squeezing in an extra Eichler order in between each step.)
24.4.6. If O = 𝑅 + 𝔭O0 for an order O0, then O♮ = O0, by Exercise 24.5.
In general, write O = 𝑅 +𝔭 𝑓 Gor(O) where 𝔭 𝑓 = 𝔣(O) is the Gorenstein conductor
of O and Gor(O) is the Gorenstein saturation as in Proposition 24.2.15. Then the chain
of radical idealizers begins
O ( O1 = 𝑅 + 𝔭 𝑓 −1 Gor(O) ( · · · ( O 𝑓 = Gor(O).
For each 𝑖, we have dim 𝑘 (O𝑖 /O𝑖+1 ) = 3.
We next consider the chain of superorders over a (local) residually inert order.
Proposition 24.4.7. Let O be a residually inert 𝑅-order. Then the following statements
hold.
(a) rad O = rad O♮ ∩ O = 𝔭O♮ .
(b) Suppose O is not maximal. Then O♮ is the unique minimal superorder of O.
Moreover, O♮ is residually inert and we have [O♮ : O] 𝑅 = 𝔭2 .
Proof. We begin with the first part of (a), and we show rad O♮ ∩ O = rad O. As in
21.2.1, O♮ is a radical cover so rad O ⊆ rad O♮ ∩ O. But arguing as in the proof
of Lemma 21.2.4, we know that rad O♮ is topologically nilpotent as a O0-ideal and
𝔭𝑟 O0 ⊆ 𝔭O for large 𝑟, so rad O♮ ∩ O is topologically nilpotent as a O-ideal, and
rad O♮ ∩ O ⊆ rad O.
We therefore have a map
O/rad O ↩→ O♮ /rad O♮ ; (24.4.8)
since O/rad O is a quadratic field, we must have (O♮| 𝔭) = ∗, −1, i.e., O♮
is either
maximal or residually inert—and in the latter case, (24.4.8) is an isomorphism.
Let 𝑖 ∈ O generate ℓ = O/rad O as a quadratic extension of 𝑘. Let 𝐾 = 𝐹 (𝑖), and
let 𝑆 be the integral closure of 𝑅 in 𝐾. Then 𝐾 is (separable and) unramified over 𝐹.
We claim that O𝑆 = O ⊗𝑅 𝑆 is residually split. Indeed, we have an isomorphism of
𝑘-algebras
∼ O /𝔭O
(O/𝔭O) ⊗ 𝑘 ℓ −→ 𝑆 𝑆
O ( O1 = O♮ ( · · · ( O𝑟
has length 𝑟 = b𝑒/2c, with O𝑟 maximal, and dim 𝑘 (O𝑖 /O𝑖+1 ) = 2 for all 𝑖; accordingly,
the case 𝑒 even occurs exactly when 𝐵 ' M2 (𝐹) and 𝑒 odd occurs exactly when 𝐵 is a
division algebra.
We conclude the section showing that under certain hypotheses, the radical idealizer
is a minimal (proper) superorder. The results are due to Drozd–Kirichenko–Roiter
[DKR67, Propositions 1.3, 10.3]; we follow Curtis–Reiner [CR81, Exercises 37.5,
37.7].
Definition 24.4.10. Let 𝐼 0 ⊆ 𝐼 be left fractional O-ideals in 𝐵. We say 𝐼 0 is (left)
hypercharacteristic in 𝐼 if for every left O-module homomorphism 𝜙 : 𝐼 0 → 𝐼 we
have 𝜙(𝐼 0) ⊆ 𝐼 0.
Lemma 24.4.11. The map O0 ↦→ 𝐼 0 = (O0) ♯ gives an inclusion-reversing bijection
from the set of 𝑅-superorders O0 ⊇ O to the set of right hypercharacteristic 𝑅-
sublattices 𝐼 0 ⊆ O♯ .
Proposition 24.4.12. Let 𝑅 be local and let O be a Gorenstein 𝑅-order that is not
maximal and such that O is indecomposable as a left O-module. Then there is a unique
minimal 𝑅-superorder O0 ) O and O0 = O♮ .
24.5.2. Since the Gorenstein condition is local by 24.2.2, the Bass condition is also
local. Moreover, an Eichler (i.e., residually split) order is Bass, because Eichler
orders are Gorenstein by 24.2.14, and every superorder of an Eichler order is Eichler
(Corollary 23.4.10).
For the rest of this section, we investigate the local structure of Bass orders, and
we suppose that 𝑅 is local. We do not use the following proposition, but we state it for
context.
The residually inert orders, those with Eichler symbol (O | 𝔭) = −1, give a source
of Bass orders, following Brzezinski [Brz83a, §3].
Combining
24.5.5. 24.5.2 with Proposition 24.5.4, we see that if O is not a Bass order,
O
then = 0.
𝔭
Another rich source of Bass orders are the basic orders.
Remark 24.5.7. Local basic orders in a division quaternion algebra were studied by
Hijikata–Pizer–Shemanske [HPS89b]; they gave the global orders O such that O (𝔭) is
basic if 𝔭 ∈ Ram(𝐵) and Eichler if 𝔭 ∉ Ram(𝐵) the name special. The remaining types
of local basic orders (residually ramified and residually inert for the matrix ring) were
studied by Brzezinski [Brz90], and further worked on by Jun [Jun97]. The role of the
embedded maximal quadratic 𝑅-algebra 𝑆 is that one can compute embedding numbers
for them (see for example the epic work of Hijikata–Pizer–Shemanske [HPS89a]), and
therefore compute explicitly with the trace formula.
Other authors use the term primitive instead of basic, but this quickly gets confusing
as the word primitive is used for the ternary quadratic forms and the two notions do not
coincide. The following propositions show that the sound of the word basic conveys
the right meaning.
𝑖 2 = −𝑏𝑐 𝑗 𝑘 = −𝑖
2
𝑗 =𝑐 𝑘𝑖 = 𝑏 𝑗 (24.5.9)
2
𝑘 =𝑏 𝑖 𝑗 = 𝑐𝑘
field and (O | 𝔭) ≠ 0, and then by 24.5.5 we know O is Bass. Now at least one of the
products 𝑏𝑐, 𝑎𝑐, 𝑎𝑏 has valuation 1 because O contains a maximal 𝑅-order; without
loss of generality we may take, after rescaling, 𝑎 = −1 and ord𝔭 (𝑏) = 1 ≤ ord𝔭 (𝑐)
with 𝑆 = 𝑅[𝑘]. Therefore 𝑄 is primitive and O is Gorenstein.
We now compute that rad O = h𝜋, 𝑖, 𝑗, 𝑘i and so
O♮ = OL (rad O) = 𝑅 + 𝔭−1𝑖 + 𝑅 𝑗 + 𝑅𝑘
Proposition 24.5.10. Suppose 𝐹 = Frac(𝑅) is a number field. Then for the 𝑅-order
O, the following are equivalent:
(i) O is basic;
(ii) O (𝔭) is basic for all primes 𝔭 of 𝑅; and
(iii) O is Bass.
In other words, for orders in a quaternion algebra over a number field, being basic
is a local property and it is equivalent to being Bass.
Remark 24.5.11. One can similarly define basic orders for a general Dedekind domain
𝑅, but the preceding results are not known in this level of generality.
24.5.12. We established several other important features along the way in Proposition
24.5.8 that we now record. Suppose 𝑅 is local with 2 ∈ 𝑅 × and suppose that O is a
residually ramified Bass (i.e., basic) order. Then the quadratic form associated to O
is similar to h−1, 𝑏, 𝑐i with ord𝔭 (𝑏) = 1 ≤ ord𝔭 (𝑐), and so the multiplication table
(24.5.9) holds. The unique superorder O♮ has [O♮ : O] 𝑅 = 𝔭 and associated ternary
quadratic form h−1, 𝑏, −𝑐/𝑏i (see Exercise 24.13).
♮
O O
Corollary 24.5.13. If O is not hereditary, then
♮ = .
𝔭 𝔭
O O
If = 1, then O is Eichler, and so too are its superorders. If = −1 and O♮ is
𝔭 𝔭
♮
O O
not maximal, then = −1 by Proposition 24.4.7(b). For the case = 0, we
𝔭 𝔭
appeal to 24.5.12.
24.6. TREE OF ODD BASS ORDERS 397
We can repackage what we have done for basic orders to give another description
in terms of its hereditary closure.
Proof. The statement follows by induction using the explicit descriptions of these
orders in 24.3.7, 24.5.12, and 24.5.12. See Brzezinski [Brz90, Proposition 1.12].
discrd (O)
B ' M2 (F ) split B ramified
=
R
M2 (R)
det p
p2
p3
p4
O
= −1 0 1 ∗ −1 0
p
res. inert res. split res. inert
hπ 2e i ⊥ NmKun |F (Eichler) hπ 2e+1 i ⊥ NmKun |F
hπ e i ⊥ H
Figure 24.6.1: Tree of local Bass orders, odd characteristic residue field
Each vertex of this graph represents an isomorphism class of Bass order; there is an
edge between two vertices if and only if there is a minimal containment between them.
398 CHAPTER 24. QUATERNION ORDERS: SECOND MEETING
In each case, such a containment is given by the radical idealizer except when the order
is residually split (in which case it hops by two, skipping the minimal superorder).
For the trees when 2 is a uniformizer in 𝑅, and many other explicit calculations,
see Lemurell [Lem2011, §5], as well as Pacetti–Sirolli [PS2014, §5].
Exercises
Unless otherwise specified, let 𝑅 be a Dedekind domain with 𝐹 = Frac 𝑅 and let
O ⊆ 𝐵 be an 𝑅-order in a quaternion algebra 𝐵.
codiff(O) (codiff(O) 2 ) ♯ = O
O
8. Suppose that O is Gorenstein with ≠ 1. Let O0 be the hereditary closure
𝔭
of O. Show that 𝑁 𝐵× (O) ≤ 𝑁 𝐵× (O0), and further that equality holds when O is
residually inert.
9. Let 𝑅 be local and let O be a residually ramified quaternion 𝑅-order that is
Gorenstein but not a Bass order. Let O♮ be the unique minimal order containing
O. Show that 𝔣(O♮ ) = 𝔭. [See Brzezinski [Brz83a, Lemma 4.4].]
10. Let 𝑅 be local with 2 ∈ 𝑅 × , and let 𝑄 = h−1, 𝑏, 𝑐i : 𝑅 3 → 𝑅 with 0 = ord𝔭 (𝑏) <
ord𝔭 (𝑐). Let O = Clf 0 (𝑄) be its even Clifford algebra. Show that
O 𝑏
= .
𝔭 𝔭
11. Let 𝑅 be local, let O = M2 (𝑅), let O1 be the standard Eichler order of level
𝔭𝑒 and let O2 = 𝑅 + 𝔭𝑒 O. Show that O2 ( O1 ( O, that Gor(O2 ) = O, but
Gor(O1 ) = O1 ; conclude that the Gorenstein saturation is not (necessarily) the
smallest Gorenstein superorder.
12. Let 𝑅 be local and let O be the standard Eichler order of level 𝔭𝑒 with 𝑒 ≥ 2.
Show that O contains no integrally closed quadratic 𝑅-order that is a domain
(even though O contains 𝑅 × 𝑅).
I 13. Let 𝑅 be local with 2 ∈ 𝑅 × .
the proof of Proposition 24.5.8) that O is a local Bass order
(a) Show(using
O
with = 0 if and only if its corresponding ternary quadratic form is
𝔭
similar to h−1, 𝑏, 𝑐i with ord𝔭 (𝑏) = 1 ≤ ord𝔭 (𝑐).
(b) In case (a), show that the minimal overorder O0 corresponds to the (simi-
larity class of) ternary quadratic form h−1, 𝑏, −𝑐/𝑏i.
Part III
Analysis
401
Chapter 25
403
404 CHAPTER 25. THE EICHLER MASS FORMULA
The Eichler mass formula does not quite give us a formula for the class number—
rather, it gives us a formula for a “weighted” class number. That being said, we remark
that 𝑤 𝐽 ≤ 24 (see Theorem 11.5.14), and very often 𝑤 𝐽 = 1 (i.e., OL (𝐽) = {±1}). In
order to convert the Eichler mass formula into a formula for the class number itself,
one needs to understand the unit groups of left orders: this can be understood either
as a problem in representation numbers of ternary quadratic forms or of embedding
numbers of quadratic orders into quaternion orders, and we will take this subject up in
earnest in Chapter 30.
Over Q, the Eichler mass formula was first proven by Hey [Hey29, II, (80)], a
Ph.D. student of Artin, along the same lines as the proof sketched below. This formula
was also stated by Brandt [Bra28, §67]. We gradually warm up to this theorem
by considering a broader analytic context. We see the analytic class number for an
imaginary quadratic field as coming from the residue of its zeta function, and we then
pursue a quaternionic generalization.
as the prototypical such function; this series converges for Re 𝑠 > 1, by the comparison
test. By unique factorization, there is an Euler product
Ö −1
1
𝜁 (𝑠) = 1− (25.2.3)
𝑝
𝑝𝑠
where the product is over all primes 𝑝. The function 𝜁 (𝑠) can be meromorphically
continued to the right half-plane Re 𝑠 > 0 using the fact that the sum
∞
∑︁ (−1) 𝑛
𝜁2 (𝑠) =
𝑛=1
𝑛𝑠
so that
1
𝜁 (𝑠) = 𝜁2 (𝑠)
21−𝑠 −1
25.2. ⊲ IMAGINARY QUADRATIC CLASS NUMBER FORMULA 405
and the right-hand side makes sense for Re 𝑠 > 0 except for possible poles where
21−𝑠 = 1. For real values of 𝑠 > 1, we have
∫ ∞ ∫ ∞
1 d𝑥 d𝑥 𝑠
= ≤ 𝜁 (𝑠) ≤ 1 + =
𝑠−1 1 𝑥𝑠 1 𝑥𝑠 𝑠−1
so
1 ≤ (𝑠 − 1)𝜁 (𝑠) ≤ 𝑠;
therefore, as 𝑠 approaches 1 from above, we have lim𝑠&1 (𝑠 − 1)𝜁 (𝑠) = 1, so 𝜁 (𝑠) has
a simple pole at 𝑠 = 1 with residue
25.2.5. For the quadratic field 𝐾, modeled after (25.2.2) we define the Dedekind zeta
function by
∑︁ 1
𝜁 𝐾 (𝑠) := (25.2.6)
𝔞⊆Z
N(𝔞) 𝑠
𝐾
where N(𝔞) := #(Z𝐾 /𝔞) is the absolute norm, the sum is over all nonzero ideals of
Z𝐾 , and the series is defined for 𝑠 ∈ C with Re 𝑠 > 1. (We recall that N(𝔞) is the
positive generator of Nm𝐾 |Q (𝔞), so we could equivalently work with the algebra norm,
if desired.)
We can also write the Dedekind zeta function as a Dirichlet series
∞
∑︁ 𝑎𝑛
𝜁 𝐾 (𝑠) = (25.2.7)
𝑛=1
𝑛𝑠
so that ∑︁
𝜁 𝐾 (𝑠) = 𝜁 𝐾 , [𝔟] (𝑠). (25.2.9)
[𝔟] ∈Cl(𝐾 )
In general, for [𝔟] ∈ Cl(𝐾), we have [𝔞] = [𝔟] if and only if there exists 𝑎 ∈ 𝐾 × such
that 𝔞 = 𝑎𝔟, but since 𝔞 ⊆ Z𝐾 , in fact
𝑎 ∈ 𝔟−1 = {𝑎 ∈ Z𝐾 : 𝑎𝔟 ⊆ Z𝐾 };
406 CHAPTER 25. THE EICHLER MASS FORMULA
where
(1 − 𝑇) 2 , if 𝑝 splits in 𝐾;
𝐿 𝑝 (𝑇) := 1 − 𝑇, if 𝑝 ramifies in 𝐾; and (25.2.13)
1 − 𝑇 2,
if 𝑝 is inert in 𝐾.
25.2. ⊲ IMAGINARY QUADRATIC CLASS NUMBER FORMULA 407
𝐿 𝑝 (𝑇) = (1 − 𝑇) (1 − 𝜒( 𝑝)𝑇).
where −1
Ö 𝜒( 𝑝) ∑︁ 𝜒(𝑛)
𝐿 (𝑠, 𝜒) := 1− 𝑠 = . (25.2.16)
𝑝
𝑝 𝑛
𝑛𝑠
The function 𝐿(𝑠, 𝜒) is in fact holomorphic for all Re 𝑠 > 0; this follows from the
Í
fact that the partial sums 𝑛 ≤𝑥 𝜒(𝑛) are bounded and the mean value theorem. So in
particular the series
1 1 1 1 1
𝐿(1, 𝜒) = 1 − + − + − +...
3 5 7 9 11
−1
Ö (−1) ( 𝑝−1)/2 𝜋
= 1− = = 0.7853 . . . .
𝑝 ≥3
𝑝 4
Remark 25.2.18. The fact that 𝐿(1, 𝜒) ≠ 0, and its generalization to complex characters
𝜒, is the key ingredient to prove Dirichlet’s theorem on primes in arithmetic progression
(Theorem 14.2.9), used in the classification of quaternion algebras over Q. The
arguments to complete the proof are requested in Exercise 26.11.
408 CHAPTER 25. THE EICHLER MASS FORMULA
Remark 25.2.19. To approach the class number problem of Gauss, we would then seek
lower bounds on 𝐿 (1, 𝜒) in terms of the absolute discriminant | 𝑑|. Indeed, the history
of class number problems is both long and beautiful. The problem of determining all
positive definite binary quadratic forms with small class number was first posed by
Gauss [Gau86, Article 303]. This problem was later seen to be equivalent to finding all
imaginary quadratic fields of small class number (as in section 19.1). It would take al-
most 150 years of work, with important work of Heegner [Heeg52] and culminating in
the results of Stark [Sta67] and Baker [Bak71], to determine those fields with class num-
ber 1: there are exactly nine, namely 𝑑 = −3, −4, −7, −8, −11, −19, −43, −67, −163.
See Goldfeld [Gol85] or Stark [Sta2007] for a history of this problem. For more
specifically on the analytic class number formula for imaginary quadratic fields, see
the survey by Weston [Wes] as well as the book by Serre [Ser73, Chapter VI].
where the sum over all invertible (nonzero, integral) right O-ideals and
(By Main Theorem 16.1.3, we have N(𝐼) the totally positive generator of nrd(𝐼) 2 , so
we could equivalently work with the reduced norm.)
Let 𝑎 𝑛 be the number of invertible right O-ideals of reduced norm 𝑛 > 0 (with
positive generator chosen, as usual). Then N(𝐼) = Nm(𝐼) = nrd(𝐼) 2 by 16.4.10, so
∞
∑︁ 𝑎𝑛
𝜁O (𝑠) = 2𝑠
. (25.3.2)
𝑛=1
𝑛
To establish an Euler product for 𝜁O (𝑠), in due course we will give a kind of factoriza-
tion formula for right ideals of O—but by necessity, writing an ideal as a compatible
product will involve the entire set of orders connected to O! A direct consequence of
the local-global dictionary for lattices (Theorem 9.4.9) is that
𝑎 𝑚𝑛 = 𝑎 𝑚 𝑎 𝑛 (25.3.3)
whenever 𝑚, 𝑛 are coprime. Next, we will count the ideals of a given reduced norm
𝑞 = 𝑝 𝑒 a power of a prime: the answer will depend on the local structure of the order
25.3. ⊲ EICHLER MASS FORMULA: OVER THE RATIONALS 409
with 𝜁O, 𝑝 (𝑇) ∈ 1 + 𝑇Z[𝑇]. In particular, 𝜁O (𝑠) only depends on the genus (local
isomorphism classes) of O.
For simplicity, we first consider the case where O is a maximal order. Since there
is a unique genus of maximal orders, the zeta function is independent of the choice of
O and so we will write 𝜁 𝐵 (𝑠) := 𝜁O (𝑠) for O maximal. Then by a local count, we will
show that (
2 1, if 𝑝 | 𝐷;
𝜁 𝐵, 𝑝 (𝑇) = (1 − 𝑇 ) · (25.3.5)
1 − 𝑝𝑇 2 , if 𝑝 - 𝐷.
From (25.3.5),
Ö 1
𝜁 𝐵 (𝑠) = 𝜁 (2𝑠)𝜁 (2𝑠 − 1) 1− . (25.3.6)
𝑝 |𝐷
𝑝 2𝑠−1
In particular, since 𝜁 (𝑠) has a simple pole at 𝑠 = 1 with residue 1 and 𝜁 (2) = 𝜋 2 /6
(Exercise 25.1),
𝜋2 Ö
1
res𝑠=1 𝜁 𝐵 (𝑠) = lim (𝑠 − 1)𝜁 𝐵 (𝑠) = 1− (25.3.7)
𝑠&1 12 𝑝
𝑝 |𝐷
(We could also look to cancel the poles of 𝜁 𝐵 (𝑠) in a similar way to define an 𝐿-function
for 𝐵, holomorphic for Re 𝑠 > 0.)
Now we break up the sum (25.3.1) according to right ideal class:
∑︁
𝜁 𝐵 (𝑠) = 𝜁 𝐵, [𝐽 ] (𝑠)
[𝐽 ] ∈Cls O
where ∑︁ 1
𝜁 𝐵, [𝐽 ] (𝑠) := . (25.3.8)
N(𝐼) 𝑠
𝐼 ⊆O
[𝐼 ]=[𝐽 ]
Since [𝐼] = [𝐽] if and only if 𝐼 = 𝛼𝐽 for some invertible 𝛼 ∈ 𝐽 −1 , and 𝜇𝐽 = 𝐽 if and
only if 𝜇 ∈ OL (𝐽) × , we conclude that
1 ∑︁ 1
𝜁 𝐵, [𝐽 ] (𝑠) = (25.3.9)
N(𝐽) 𝑠 −1 ×
N(𝛼) 𝑠
0≠𝛼∈𝐽 /OL ( 𝐽 )
where the sum is taken over the nonzero elements 𝛼 ∈ 𝐽 −1 up to right multiplication
by units OL (𝐽) × in the left order.
In order to proceed, we now suppose that 𝐵 is definite (ramified at ∞) and hence
that #OL (𝐽) × < ∞ (see Lemma 17.7.13); this is the analogue with the case of an
imaginary quadratic field, and each 𝐽 has the structure of a lattice in the Euclidean
space R4 via the embedding
𝐽 ↩→ 𝐵 ↩→ 𝐵 ⊗Q R ' H ' R4 . (25.3.10)
410 CHAPTER 25. THE EICHLER MASS FORMULA
Let 𝑤 𝐽 = #OL (𝐽) × /{±1}. We again argue by counting lattice points to prove the
following proposition.
Proposition 25.3.11. The function 𝜁 𝐵, [𝐽 ] (𝑠) has a simple pole at 𝑠 = 1 with residue
𝜋2
res𝑠=1 𝜁 𝐵, [𝐽 ] (𝑠) = .
𝑤𝐽 𝐷
Proof sketch. From a more general result (Theorem 26.2.12, proven in the next section
and used to prove the analytic class number formula itself), we will show that
1 vol((R4 ) ≤1 )
res𝑠=1 𝜁 𝐵, [𝐽 ] (𝑠) = (25.3.12)
2𝑤 𝐽 N(𝐽) covol(𝐽)
where under (25.3.10) we have
𝜋2
vol((R4 ) ≤1 ) = vol({𝑥 ∈ R4 : |𝑥| ≤ 1}) =
2
and
covol(O) 𝐷/4
covol(𝐽) = = .
N(𝐽) N(𝐽)
Putting all of these facts together,
𝜋2 4N(𝐽) 𝜋2
res𝑠=1 𝜁 𝐵, [𝐽 ] (𝑠) = = . (25.3.13)
4𝑤 𝐽 N(𝐽) 𝐷 𝑤𝐽 𝐷
In particular the pole of each zeta function 𝜁 𝐵, [𝐽 ] (𝑠) is almost independent of the
class [𝐽], with the only relevant term being 𝑤 𝐽 the number of units.
Combining Proposition 25.3.13 with Proposition 25.3.11,
𝜋 2 ∑︁ 𝜋2 Ö
1 1
res𝑠=1 𝜁 𝐵 (𝑠) = = 1− (25.3.14)
𝐷 𝑤𝐽 12 𝑝
[𝐽 ] ∈Cls O 𝑝 |𝐷
Remark 25.3.17. The Eichler mass formula is also very similar to the mass formula
for the number of isomorphism classes of supersingular elliptic curves: this is no
coincidence, and its origins will be explored in section 42.2.
To extend the Eichler mass formula to a more general class of orders, one only
needs to replace the local calculation in 25.3.5 by a count of invertible ideals in the
order. First we treat the important case of Eichler orders (see 23.1.3).
25.4. CLASS NUMBER ONE AND TYPE NUMBER ONE 411
Theorem 25.3.18 (Eichler mass formula, Eichler orders over Q). Let O ⊂ 𝐵 be an
Eichler order of level 𝑀 in a definite quaternion algebra 𝐵 of discriminant 𝐷. Then
∑︁ 1 𝜑(𝐷)𝜓(𝑀)
=
𝑤𝐽 12
[𝐽 ] ∈Cls O
where
Ö Ö 1
𝑒 𝑒−1
𝜓(𝑀) = (𝑝 + 𝑝 )=𝑀 1+ .
𝑝
𝑝𝑒 k 𝑀 𝑝 |𝑀
where
1 + 1/𝑝, if (O | 𝑝) = 1;
1 − 𝑝 −2
𝜆(O, 𝑝) = = 1 − 1/𝑝, if (O | 𝑝) = −1; and (25.3.20)
O −1
1− 𝑝 1 − 1/𝑝 2 ,
if (O | 𝑝) = 0.
𝑝
Main Theorem 25.3.19 was proven by Brzezinski [Brz90, (4.6)] and more generally
over number rings by Körner [Kör87, Theorem 1].
Remark 25.4.2. The primes 𝑝 = 2, 3, 5, 7, 13 in Theorem 25.4.1 are also the primes
𝑝 such that the modular curve 𝑋0 ( 𝑝) has genus 0. This is not a coincidence, and
reflects a deep correspondence between classical and quaternionic modular forms (the
Eichler–Shimizu–Jacquet–Langlands correspondence): see Remark 41.5.13.
412 CHAPTER 25. THE EICHLER MASS FORMULA
The list of all definite quaternion orders (over Z) of class number 1 was determined
by Brzezinski [Brz95]. (Brzezinski mistakenly lists an order of class number 2, and
so he counts 25, not 24; he corrects this later in a footnote [Brz98, Footnote 1].)
The list of orders with class number 1 is given in Table 25.4.4. We provide
𝑁 := discrd(O), 𝐷 := disc(𝐵), we list the Eichler symbols (O | 𝑝) for the relevant
primes 𝑝 ≤ 13, we say if the order is maximal, hereditary (but not maximal), Eich-
ler/residually split or residually inert (but not hereditary), Bass (but not Eichler), or
non-Gorenstein. This list includes the three norm Euclidean maximal orders (Exercise
25.6) of discriminant 𝐷 = 2, 3, 5.
(O | 𝑝)
𝑁 𝐷 2 3 5 7 11 13 Class 𝑄
2 2 −1 ∗ ∗ ∗ ∗ ∗ maximal 𝑥 2 − 𝑥𝑦 − 𝑥𝑧 + 𝑦 2 + 𝑦𝑧 + 𝑧2
3 3 ∗ −1 ∗ ∗ ∗ ∗ maximal 𝑥 2 − 𝑥𝑦 + 𝑦 2 + 𝑧2
4 2 0 ∗ ∗ ∗ ∗ ∗ Bass 𝑥 2 + 𝑦2 + 𝑧2
5 5 ∗ ∗ −1 ∗ ∗ ∗ maximal 𝑥 − 𝑥𝑦 − 𝑥𝑧 + 𝑦 2 + 𝑦𝑧 + 2𝑧2
2
6 2 −1 1 ∗ ∗ ∗ ∗ hereditary 𝑥 2 − 𝑥𝑦 + 𝑦 2 + 2𝑧 2
6 3 1 −1 ∗ ∗ ∗ ∗ hereditary 𝑥 + 𝑥𝑧 + 𝑦 2 − 𝑦𝑧 + 2𝑧2
2
7 7 ∗ ∗ ∗ −1 ∗ ∗ maximal 𝑥 2 − 𝑥𝑧 + 𝑦 2 + 2𝑧2
8 2 −1 ∗ ∗ ∗ ∗ ∗ Bass 𝑥 + 𝑥𝑦 − 𝑥𝑧 + 𝑦 2 − 𝑦𝑧 + 3𝑧2
2
8 2 0 ∗ ∗ ∗ ∗ ∗ Bass 𝑥 2 + 𝑦 2 + 2𝑧2
10 2 −1 ∗ 1 ∗ ∗ ∗ hereditary 𝑥 − 𝑥𝑧 + 𝑦 2 − 𝑦𝑧 + 3𝑧2
2
10 5 1 ∗ −1 ∗ ∗ ∗ hereditary 𝑥 + 𝑥𝑦 + 𝑥𝑧 + 2𝑦 2 + 2𝑦𝑧 + 2𝑧 2
2
12 3 0 −1 ∗ ∗ ∗ ∗ Bass 𝑥 2 + 𝑦 2 + 3𝑧2
13 13 ∗ ∗ ∗ ∗ ∗ −1 maximal 𝑥 − 𝑥𝑦 + 2𝑦 2 + 𝑦𝑧 + 2𝑧 2
2
16 2 0 ∗ ∗ ∗ ∗ ∗ non-Gorenstein 2(𝑥 2 − 𝑥𝑦 − 𝑥𝑧 + 𝑦 2 + 𝑦𝑧 + 𝑧2 )
16 2 0 ∗ ∗ ∗ ∗ ∗ Bass 𝑥 2 + 2𝑦 2 + 2𝑧 2
18 2 −1 −1 ∗ ∗ ∗ ∗ Bass 𝑥 + 𝑥𝑧 + 𝑦 2 − 𝑦𝑧 + 5𝑧2
2
Proof of Theorem 25.4.3. Suppose # Cls O = 1. We apply the mass formula (Main
25.4. CLASS NUMBER ONE AND TYPE NUMBER ONE 413
and therefore 𝜑(𝑁) ≤ 12. By elementary number theory, this implies that
2 ≤ 𝑁 ≤ 16 or 𝑁 = 18, 20, 21, 22, 24, 26, 28, 30, 36, 42.
This immediately gives a finite list of possibilities for the discriminant 𝐷 = disc 𝐵 ∈
{2, 3, 5, 7, 11, 13, 30, 42}, as 𝐷 must be a squarefree product of an odd number of
primes.
By Exercise 17.3, if O ⊆ O0 then there is a natural surjection Cls O → Cls O0,
which is to say an order has at least as large a class number as any superorder. So
we must have # Cls O0 = 1 for a maximal order O0 ⊆ 𝐵, and by Theorem 25.4.1, this
then reduces us to 𝐷 = 2, 3, 5, 7, 13. Because # Cls O0 = 1, the maximal order O0 is
unique up to conjugation, and so fixing a choice of such a maximal order O0 up to
isomorphism we may suppose O ⊆ O0. But now the index [O0 : O] = 𝑀 = 𝐷/𝑁 is
explicitly given, and there are only finitely many suborders of bounded index; for each,
we may compute representatives of the class set in a manner similar to the example in
section 17.6. (We are aided further by deciding what assignment of Eichler symbols
and unit orders would be necessary in each case.)
There is similarly an interest in definite quaternion orders O of type number 1:
these are the orders with the property that the “local-to-global principle applies for
isomorphisms”, i.e., if O𝔭0 ' O𝔭 for all primes 𝔭 then O0 ' O. If an order has class
number 1 then it has type number 1, by Lemma 17.4.13, but one may have # Typ O = 1
but # Cls O > 1. Since an order has the same type number as its Gorenstein saturation,
i.e. Typ O = Typ Gor(O), it suffices to classify the Gorenstein orders with this property.
By the bijection between ternary forms and quaternion orders, this is equivalently
the problem of enumerating one-class genera of primitive ternary quadratic forms.
The list was drawn up by Jagy–Kaplansky–Schiemann [JKS97] (with early work due
to Watson [Wats75]), and has been independently confirmed by Lorch–Kirschmer
[LK2013].
Theorem 25.4.6 (Watson, Jagy–Kaplansky–Schiemann, Lorch–Kirschmer). There
are exactly 794 primitive ternary quadratic forms of class number 1, corresponding
to 794 Gorenstein quaternion orders of type number 1. The largest prime dividing
a discriminant is 23, and the largest (reduced) discriminant is 28 33 72 = 338688.
There are exactly 9 corresponding to maximal quaternion orders: they have reduced
discriminants
𝐷 = 2, 3, 5, 7, 13, 30, 42, 70, 78.
Remark 25.4.7. The generalization of the class number 1 problem to quadratic forms
of more variables was pursued by Watson, who showed that one-class genera do not
exist in more than ten variables [Wats62]. Watson also tried to compile complete lists
in low dimensions, followed by work of Hanke, and recently the complete list has been
drawn up in at least 3 variables over Q by Lorch–Kirschmer [LK2013] and over totally
real fields for maximal lattices by Kirschmer [Kir2014].
414 CHAPTER 25. THE EICHLER MASS FORMULA
Exercises
1. A short and fun proof of the equality 𝜁 (2) = 𝜋 2 /6 is due to Calabi [BCK93].
(a) Expand (1 − 𝑥 2 𝑦 2 ) −1 in a geometric series and integrate termwise over
𝑆 = [0, 1] × [0, 1] to obtain
∫∫
2 2 −1 1 1 1
(1 − 𝑥 𝑦 ) d𝑥 d𝑦 = 1 + 2 + 2 + . . . = 1 − 𝜁 (2).
𝑆 3 5 4
𝜋2
∫∫ ∫∫
(1 − 𝑥 2 𝑦 2 ) −1 d𝑥 d𝑦 = d𝑢 d𝑣 =
𝑆 𝑇 8
1 1 1
𝑓 (𝑚, 𝑛) := 3
+ 2 2
+ 3
𝑚𝑛 2𝑚 𝑛 𝑚 𝑛
for 𝑚, 𝑛 ∈ Z>0 , and prove that
1
𝑓 (𝑚, 𝑛) − 𝑓 (𝑚 + 𝑛, 𝑛) − 𝑓 (𝑚, 𝑚 + 𝑛) = .
𝑚 2 𝑛2
(b) Prove
∞
∑︁ 5
𝜁 (2) 2 = 𝑓 (𝑛, 𝑛) = 𝜁 (4).
𝑛=1
2
𝑘−2
1 1 ∑︁ 1 1
𝑓 (𝑚, 𝑛) = + +
𝑚𝑛 𝑘−1 2 𝑟 =2 𝑚 𝑟 𝑛 𝑘−𝑟 𝑚 𝑘−1 𝑛
25.4. CLASS NUMBER ONE AND TYPE NUMBER ONE 415
so by induction 𝜁 (𝑘) ∈ Q𝜋 𝑘 .
I 3. Prove Lemma 25.2.11 as follows.
(a) Let 𝑃 be a fundamental parallelogram for Λ, and for 𝜆 ∈ Λ let 𝑃𝜆 := 𝑃 + 𝜆.
For 𝑥 > 1, let
𝐷 (𝑥) := {𝑧 ∈ C : | 𝑧| ≤ 𝑥}
and
Show that
𝑁 𝑃 (𝑥) ≤ 𝑁 (𝑥) ≤ 𝑁 𝑃+ (𝑥).
(b) Show that 𝑁 𝑃 (𝑥) ≤ 𝜋𝑥 2 /𝐴 ≤ 𝑁 𝑃+ (𝑥).
(c) Let 𝑙 be the length of a long diagonal in 𝑃. Show that for all 𝜆 ∈ Λ ∩ 𝐷 (𝑥),
we have 𝑃𝜆 ⊆ 𝐷 (𝑥 + 𝑙), so
𝜋(𝑥 + 𝑙) 2
𝑁 (𝑥) ≤ 𝑁 𝑃 (𝑥 + 𝑙) ≤ .
𝐴
Similarly, show that if 𝑃𝜆 ∩ 𝐷 (𝑥 − 𝑙) ≠ ∅ then 𝑃𝜆 ⊆ 𝐷 (𝑥) and 𝜆 ∈ 𝐷 (𝑥),
so
𝜋(𝑥 − 𝑙) 2
≤ 𝑁 𝑃+ (𝑥 − 𝑙) ≤ 𝑁 (𝑥).
𝐴
(d) Conclude that Lemma 25.2.11 holds with 𝐶 := 𝜋(2𝑙 + 𝑙 2 )/𝐴.
I 4. Using the previous exercise, we now prove the analytic class number formula
(Theorem 25.2.12).
(a) Let 𝔟 ⊂ C be a fractional ideal, and let
Show that
∑︁ 𝜋𝑥 √
𝑏 − ≤𝐶 𝑥
𝑛 ≤𝑥 𝑛
𝐴
416 CHAPTER 25. THE EICHLER MASS FORMULA
Show by the comparison test that 𝑓 (𝑠) converges for all Re 𝑠 > 1/2.
(c) Show that
2𝜋
res𝑠=1 𝜁 𝐾 (𝑠) = lim (𝑠 − 1)𝜁 𝐾 , [𝔟] (𝑠) = √︁ .
𝑠&1 𝑤 | 𝑑|
(d) Sum the residues over [𝔟] ∈ Cl(𝐾) to derive the theorem.
I 5. In this exercise, we prove Theorem 25.4.1: if O is a maximal order in a definite
quaternion algebra over Q of discriminant 𝐷, then # Cls O = 1 if and only if
𝐷 = 2, 3, 5, 7, 13. By the Eichler mass formula (Theorem 25.3.15), we have
# Cls O = 1 if and only if
1 𝜑(𝐷)
=
𝑤 12
where 𝑤 = #O/{±1}.
(a) Show (cf. 11.5.13) that if 𝐷 > 3 then 𝑤 ≤ 3.
(b) Show that # Cls O = 1 for 𝐷 = 2, 3. (The case 𝐷 = 2 is the Hurwitz
order and 𝐷 = 3 is considered in Exercise 11.12. In fact, these orders are
Euclidean with respect to the reduced norm: see the next exercise.)
(c) Show that if 𝐷 is a squarefree positive integer with an odd number of prime
factors and 𝜑(𝐷)/12 ∈ {1, 1/2, 1/3}, then 𝐷 ∈ {5, 7, 13, 42}.
(d) Prove that # Cls O = 1 for 𝐷 = 5, 7, 13 (cf. Exercise 17.10).
(e) Show that # Cls O = 2 for 𝐷 = 42.
6. Let O be a definite quaternion order over Z. If O is Euclidean, then # Cls O = 1,
and we saw in 11.3.1 that the Hurwitz order O of discriminant 𝐷 = 2 is Euclidean
with respect to the reduced norm.
(a) Show that if O is norm Euclidean, then O is maximal.
(b) Show that O is Euclidean with respect to the reduced norm if and only if
for all 𝛾 ∈ 𝐵, there exists 𝜇 ∈ O such that nrd(𝛾 − 𝜇) < 1.
(c) Show that if O is maximal, then O is norm Euclidean if and only if
𝐷 = 2, 3, 5.
7. Generalizing the previous exercise, we may ask for the Euclidean ideal classes
in maximal orders. We will show that there are no nonprincipal Euclidean two-
sided ideal classes in maximal definite quaternion orders over Z. [This exercise
was suggested by Pete L. Clark.]
25.4. CLASS NUMBER ONE AND TYPE NUMBER ONE 417
In this chapter, we prove the Eichler mass number for a definite quaternion order over
a totally real field using classical analytic methods.
26.1.3. We saw in Lemma 17.7.13 that for each definite order O, the group O1 of
units of reduced norm 1 is a finite group; we will see in Lemma 26.5.1 that further the
group O× /𝑅 × is finite. For a right O-ideal 𝐽, the automorphism group of 𝐽 (as a right
419
420 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
Main Theorem 26.1.5 (Eichler mass formula). With notation as in 26.1.1, we have
2𝜁 𝐹 (2) 3/2 Ö
mass(Cls O) = 𝑑 ℎ 𝐹 N(𝔑) 𝜆(O, 𝔭). (26.1.6)
(2𝜋) 2𝑛 𝐹 𝔭 |𝔑
26.1.7. The functional equation for the Dedekind zeta function relates 𝑠 to 1− 𝑠, giving
an alternative way of writing (26.1.6) as
2𝜁 𝐹 (2) 3/2 | 𝜁 𝐹 (−1)|
𝑑 = . (26.1.8)
(2𝜋) 2𝑛 𝐹 2𝑛−1
We notice that the Eichler mass formula then implies that 𝜁 𝐹 (−1) ∈ Q.
Remark 26.1.9. More generally, the rationality of the values 𝜁 𝐹 (−𝑛) with 𝑛 ∈ Z>0 is
a theorem of Siegel [Sie69] and Deligne–Ribet [DR80].
Remark 26.1.10. The weighting in the mass is what makes Main Theorem 26.1.5 so
simple. In the (unlikely) situation where 𝑤 𝐽 = 𝑤 O is independent of 𝐽, we would have
a formula for the class number, but more generally we will need to take account of unit
groups by computing embedding numbers of cyclotomic quadratic orders: we will do
this in Chapter 30.
We now make the formula (26.1.6) a bit more explicit for the case of Eichler orders.
26.1.11. Let O be an Eichler order of level 𝔐, so that 𝔑 = 𝔇𝔐 with 𝔇, 𝔐 coprime.
Then
−1, if 𝔭 | 𝔇;
O
= 1, if 𝔭 | 𝔐;
𝔭
∗,
if 𝔭 - 𝔑.
Accordingly, we define the generalized Euler 𝜑-function and Dedekind 𝜓-function by
Ö Ö 1
𝜑(𝔇) := (N(𝔭) − 1) = N(𝔇) 1−
N(𝔭)
𝔭 |𝔇 𝔭 |𝔇
Ö Ö 1
𝜓(𝔐) := N(𝔭) 𝑒−1 (N(𝔭) + 1) = N(𝔐) 1+
N(𝔭)
𝑒 𝔭 k𝔐 𝔭 |𝔐
(recalling 𝔇 is squarefree, with the natural extension 𝜑(𝔇) = #(𝑅/𝔇) × for all 𝔇).
The 𝜓-function computes a unit index: see Lemma 26.6.7.
26.2. ANALYTIC CLASS NUMBER FORMULA 421
The Eichler mass formula (Main Theorem 26.1.5) for Eichler orders then reads as
follows.
Theorem 26.1.12 (Eichler mass formula, Eichler orders). With notation as in 26.1.11,
we have
2𝜁 𝐹 (2) 3/2
mass(Cls O) = 𝑑 ℎ 𝐹 𝜑(𝔇)𝜓(𝔐). (26.1.13)
(2𝜋) 2𝑛 𝐹
Remark 26.1.14. The Eichler mass formula in the form (26.1.13) for maximal orders
was proven by Eichler (working over a general totally real field) using the techniques in
this chapter [Eic38b, Satz 1], and was extended to squarefree level 𝔑 (i.e., hereditary
orders) again by Eichler [Eic56a, §4]. This was extended by Brzezinski [Brz90, (4.6)]
to a general formula over Q and by Körner [Kör87, Theorem 1], using idelic methods.
The classical method to prove Main Theorem 26.1.5 is similar to the one we
sketched over Q in chapter 25, with some added technicalities of working over a
number field. We follow this approach, first proving the formula when O is a maximal
order, and then deducing the general case. We will return in chapter 29 and reconsider
the Eichler mass formula from an idelic point of view, thinking of it as a special case
of a volume formula (for a finite set of “quotient points”). It is hoped that this chapter
will serve to show both the power and limits of classical methods before we build upon
them using idelic methods.
where the sum is over all nonzero ideals of 𝑅 and N(𝔞) = #(𝑅/𝔞) = [𝑅 : 𝔞] is the
absolute norm; we have N(𝔞) = Nm(𝔞) with norm taken from 𝐹 to Q and positive
generator chosen.
26.2.1. The Dedekind zeta function converges for Re 𝑠 > 1 and has an Euler product
Ö −1
1
𝜁 𝐹 (𝑠) = 1− (26.2.2)
𝔭
N(𝔭) 𝑠
422 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
where the product is over all nonzero primes of 𝑅—this follows formally from unique
factorization of ideals after one shows that the pruned product converges.
The Dedekind zeta function has properties analogous to the Riemann zeta function,
which is the special case 𝐹 = Q. In particular, we can extend 𝜁 𝐹 (𝑠) in a manner
analogous to 25.2.1 to Re 𝑠 > 0. For 𝑎 ∈ C we write 𝜁 𝐹∗ (𝑎) for the leading coefficient
in the Laurent series expansion for 𝜁 𝐹 at 𝑠 = 𝑎.
Theorem 26.2.3 (Analytic class number formula). 𝜁 𝐹 (𝑠) has analytic continuation to
Re 𝑠 > 0, with a simple pole at 𝑠 = 1 having residue
2𝑟 (2𝜋) 𝑐
𝜁 𝐹∗ (1) = lim (𝑠 − 1)𝜁 𝐹 (𝑠) = √︁ ℎ 𝐹 Reg𝐹 . (26.2.4)
𝑠→1 𝑤 𝐹 | 𝑑𝐹 |
Remark 26.2.5. The formula (26.2.4) is known as Dirichlet’s analytic class number
formula (even though the original form of Dirichlet’s theorem concerned quadratic
forms rather than classes of ideals, so is closer to Theorem 25.2.12).
Before we finish this section, we review a few ingredients from the proof of the
analytic class number formula (26.2.4) to set up the Eichler mass formula.
26.2.7. We first write the Dedekind zeta function as a sum over ideals in a given ideal
class [𝔟] ∈ Cl(𝑅): we define the partial zeta function
∑︁ 1
𝜁 𝐹 , [𝔟] (𝑠) := (26.2.8)
𝔞⊆𝑅
N(𝔞) 𝑠
[𝔞]=[𝔟]
Now note that [𝔞] = [𝔟] if and only if 𝔞 = 𝑎𝔟 for some nonzero
𝑎 ∈ 𝔟−1 = {𝑥 ∈ 𝐹 : 𝑥𝔟 ⊆ 𝑅},
so there is a bijection between nonzero ideals 𝔞 ⊆ 𝑅 such that [𝔞] = [𝔟] and the set of
nonzero elements in 𝔟−1 /𝑅 × . So
1 ∑︁ 1
𝜁 𝐹 , [𝔟] (𝑠) = . (26.2.10)
N(𝔟) 𝑠 Nm(𝑎) 𝑠
0≠𝑎 ∈𝔟−1 /𝑅 ×
One now reduces to a problem concerning lattice points in a fundamental domain for
the action of 𝑅 × , and examining the residue of the pole at 𝑠 = 1 fits into a more general
framework (invoked again below).
26.2. ANALYTIC CLASS NUMBER FORMULA 423
Suppose that
𝑋 ≤1 := {𝑥 ∈ 𝑋 : 𝑁 (𝑥) ≤ 1} ⊆ R𝑛 (26.2.13)
is a bounded subset with volume vol(𝑋 ≤1 ). Let Λ ⊆ R𝑛
be a (full) Z-lattice in R𝑛 , and
let
∑︁ 1
𝜁Λ,𝑋 (𝑠) := .
𝜆∈𝑋 ∩Λ
𝑁 (𝜆) 𝑠
Then 𝜁Λ,𝑋 (𝑠) converges for Re 𝑠 > 1 and has a simple pole at 𝑠 = 1 with residue
∗ vol(𝑋 ≤1 )
𝜁Λ,𝑋 (1) = lim (𝑠 − 1)𝜁Λ,𝑋 (𝑠) = .
𝑠&1 covol(Λ)
26.2.14. To apply Theorem 26.2.12 for 𝜁 𝐹 , [𝔟] (𝑠), we embed 𝐹 ↩→ 𝐹R ' R𝑟 × C𝑐 and
we equip 𝐹R with the inner product
𝑟
∑︁ 𝑐
∑︁
h𝑥, 𝑦i = 𝑥𝑖 𝑦 𝑖 + 2 Re(𝑥𝑟 + 𝑗 𝑦 𝑟 + 𝑗 ) (26.2.15)
𝑖=1 𝑗=1
for 𝑥 = (𝑥𝑖 )𝑖 , 𝑦 = (𝑦 𝑖 )𝑖 ∈ 𝐹R . This inner product modifies the usual one by rescaling
complex coordinates, and the volume form vol induced by h , i is 2𝑐 times the standard
Lebesgue volume on R𝑟 × C𝑐 . With this convention, we have h𝑥, 1i = Tr𝐹 |Q (𝑥) and
√︁
covol(𝑅) = | 𝑑 𝐹 |.
We then take Λ to be the image of 𝔟−1 , and take 𝑋 to be a cone fundamental
domain for the action of the unit group 𝑅 × . The absolute norm N(𝑥) = |Nm𝐹 |Q (𝑥)|
then satisfies the required homogeneity property, and 𝑋 ≤1 is bounded, so by Theorem
26.2.12,
1 vol(𝑋 ≤1 )
𝜁 𝐹∗ , [𝔟] (1) = . (26.2.16)
N(𝔟) covol(Λ)
We have √︁
covol(𝑅) | 𝑑𝐹 |
covol(Λ) = = . (26.2.17)
N(𝔟) N(𝔟)
It requires a bit more work to compute vol(𝑋 ≤1 ).
2𝑟 (2𝜋) 𝑐 Reg𝐹
vol(𝑋 ≤1 ) = (26.2.19)
𝑤𝐹
424 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
where the first sum is over all (nonzero) integral, invertible right O-ideals 𝐼 and in the
second sum we define
Proof. We use the local-global dictionary for lattices (Theorem 9.4.9). To ease paren-
theses in the notation, we work in the completion, but one can also work just in the
localization. For all 𝔭, we have O𝔭0 = 𝜈𝔭−1 O𝔭 𝜈𝔭 for some 𝜈𝔭 ∈ 𝐵𝔭× , and we may take
𝜈𝔭 = 1 for all but finitely many 𝔭; the element 𝜈𝔭 is well-defined up to left multiplication
by O×𝔭 and right multiplication by O𝔭0× .
26.3. CLASSICAL ZETA FUNCTIONS OF QUATERNION ALGEBRAS 425
Then to an integral, invertible right O-ideal 𝐼, we associate the unique lattice 𝐼 0 such
that 𝐼𝔭0 = 𝜈𝔭−1 𝐼𝔭 𝜈𝔭 ; such a lattice is well-defined independent of the choice of 𝜈𝔭 . By
construction, OR (𝐼𝔭0 ) = O𝔭0 so OR (𝐼 0) = O0. And since 𝐼 is integral, 𝐼𝔭 ⊆ O𝔭 whence
𝐼𝔭0 ⊆ 𝜈𝔭−1 𝐼𝔭 𝜈𝔭 ⊆ O𝔭0 and 𝐼 is locally integral hence integral. Since 𝐼 is invertible,
𝐼 is locally principal, so 𝐼 0 is also locally principal, hence invertible. Finally, again
checking locally, we have nrd(𝐼 0) = nrd(𝐼).
Repeating this argument going from 𝐼 0 to 𝐼, we see that the corresponding sets of
ideals are in bijection, as claimed.
26.3.5. From Lemma 26.3.4, we see that 𝜁O (𝑠) only depends on the genus of O. Since
there is a unique genus of maximal orders in 𝐵, following the number field case we
will write 𝜁 𝐵 (𝑠) = 𝜁O (𝑠) where O is any maximal order.
Our next order of business is to establish an Euler product for 𝜁O (𝑠). We prove a
more general result on the factorization of invertible lattices.
Lemma 26.3.6. Let 𝐼 be an invertible, integral lattice and suppose that nrd(𝐼) = 𝔪𝔫
with 𝔪, 𝔫 ⊆ 𝑅 coprime ideals. Then there exists a unique invertible, integral lattice 𝐽
such that 𝐼 is compatible with 𝐽 −1 with 𝐼𝐽 −1 integral and nrd(𝐽) = 𝔪.
Proof. We use the local-global dictionary for lattices, and we define 𝐽 ⊆ 𝐵 to be the
unique lattice such that
𝐼 = O (𝔭) , if 𝔭 - 𝔪𝔫;
(𝔭)
𝐽 (𝔭) := 𝐼 (𝔭) , if 𝔭 | 𝔪; (26.3.7)
O (𝔭) ,
if 𝔭 | 𝔫.
We have OR (𝐽) = O and nrd(𝐽) = 𝔪, since these statements hold locally. Integrality
and invertibility are local; since these are true for 𝐼 they are true for 𝐽. Finally, we
compute that (𝐼𝐽 −1 ) (𝔭) = O (𝔭) for all 𝔭 - 𝔫 and (𝐼𝐽 −1 ) (𝔭) = 𝐼 (𝔭) for 𝔭 | 𝔫, so 𝐼𝐽 −1
is locally integral and hence integral. The uniqueness of 𝐽 can be verified directly
(Exercise 26.4).
Proof. Write 𝐴𝔫 (O) for the set of integral, invertible right O-ideals 𝐼 with nrd(𝐼) = 𝔫.
Then #𝐴𝔫 (O) = 𝑎 𝔫 (O). According to Lemma 26.3.6, there is a map
We claim that this map is surjective and that each fiber has cardinality 𝑎 𝔪 (O). Indeed,
these statements follow at the same time from the following observation: if 𝐽 ∈ 𝐴𝔫 (O)
with O0 = OL (𝐽), then for each 𝐼 0 ∈ 𝐴𝔪 (O0), we have 𝐼 0 compatible with 𝐽 and
𝐼 = 𝐼 0 𝐽 ∈ 𝐴𝔪 , and conversely; so the fiber of (26.3.10) is identified with 𝐴𝔪 (O0), of
cardinality 𝑎 𝔪 (O0) = 𝑎 𝔪 (O) by Lemma 26.3.4.
26.3.11. From Proposition 26.3.9 and unique factorization of ideals in 𝑅, we find that
𝜁O has an Euler product Ö
𝜁O (𝑠) = 𝜁O𝔭 (𝑠) (26.3.12)
𝔭
where
∞
∑︁ 1 ∑︁ 𝑎 𝔭𝑒 (O)
𝜁O𝔭 (𝑠) := = . (26.3.13)
𝐼𝔭 ⊆O𝔭
N(𝐼𝔭 ) 𝑠
𝑒=0
N(𝔭) 2𝑠
Remark 26.3.14. Zeta functions of semisimple algebras over a number field can be
defined in the same way as in (26.3.2), following Solomon [Sol77]: see the survey on
analytic methods in noncommutative number theory by Bushnell–Reiner [BR85].
Remark 26.3.15. The world of 𝐿-functions is rich and very deep: for a beautiful
survey of the analytic theory of automorphic 𝐿-functions in historical perspective,
see Gelbart–Miller [GM2004]. In particular, we have not given a general definition
of zeta functions (or 𝐿-functions) in this section, but it is generally agreed that the
Selberg class incorporates the minimal essential featuers: definition as a Dirichlet
series, meromorphic continuation to the complex plane, Euler product, and functional
equation. See e.g. Conrey–Ghosh [CG93] and the references therein.
so it suffices to count the number of ideals in the local case. In this section, we carry
out this count for maximal orders.
So let 𝔭 ⊂ 𝑅 be a (nonzero) prime and let 𝑞 := Nm(𝔭). Let O𝔭 ⊂ 𝐵𝔭 be a maximal
order. Let
𝑎 𝔭𝑒 (O𝔭 ) = #{𝐼𝔭 = 𝛼𝔭 O𝔭 ⊆ O𝔭 : nrd(𝐼𝔭 ) = 𝔭𝑒 }
count the number of right integral O𝔭 -ideals of norm 𝔭𝑒 . Since O𝔭 is maximal, every
nonzero ideal is invertible; and because 𝑅𝔭 is a DVR, all such invertible ideals are
principal.
(a) If 𝐵𝔭 is a division ring, then every right integral O𝔭 -ideal is a power of the
maximal ideal and 𝑎 𝔭𝑒 (O𝔭 ) = 1.
26.4. COUNTING IDEALS IN A MAXIMAL ORDER 427
(b) If 𝐵𝔭 ' M2 (𝐹𝔭 ), so that O𝔭 ' M2 (𝑅𝔭 ), then the set of right integral O𝔭 -ideals
of reduced norm 𝔭𝑒 is in bijection with the set
𝑢
𝜋 0 𝑣
: 𝑢, 𝑣 ∈ Z ≥0 , 𝑢 + 𝑣 = 𝑒 and 𝑐 ∈ 𝑅/𝔭
𝑐 𝜋𝑣
and
𝑎 𝔭𝑒 (O𝔭 ) = 1 + 𝑞 + · · · + 𝑞 𝑒 . (26.4.2)
Proof. For (a), if 𝔭 is ramified then by the work of section 13.3, there is a unique
maximal order O𝔭 with a unique (two-sided) maximal ideal 𝐽𝔭 having nrd(𝐽𝔭 ) = 𝔭,
and all ideals of O𝔭 are powers of 𝐽𝔭 .
To prove (b), we appeal to the theory of elementary divisors (applying column
operations, acting on the right). Suppose O𝔭 = M2 (𝑅𝔭 ). Let 𝐼𝔭 = 𝛼𝔭 O𝔭 be a right
integral O𝔭 -ideal of norm 𝔭𝑒 and let 𝜋 be a uniformizer for 𝔭. Then by the theory of
elementary divisors, we can write
𝑢
𝜋 0
𝛼𝔭 =
𝑐 𝜋𝑣
26.4.3. There is an alternate bijection that is quite useful. We say an integral right
O-ideal 𝐼 is 𝔭-primitive if it does not contain 𝔭O (so we cannot write 𝐼 = 𝔭𝐼 0 with 𝐼 0
integral).
For a commutative ring 𝐴, we define the projective line over 𝐴 to be the set
P1 ( 𝐴) := {(𝑥, 𝑦) ∈ 𝐴2 : 𝑥 𝐴 + 𝑦 𝐴 = 𝐴}/𝐴×
Any ideal of the form in the right-hand side of (26.4.4) is a primitive right integral
O𝔭 -ideal with reduced norm 𝔭𝑒 . Conversely, suppose that 𝐼𝔭 = 𝛼𝔭 O𝔭 is primitive. We
have nrd(𝛼𝔭 ) ≡ 0 (mod 𝔭𝑒 ). We find a “standard form” for 𝐼𝔭 by looking at the left
kernel of 𝛼𝔭 . Let
𝐿 := {𝑥 ∈ (𝑅/𝔭𝑒 ) 2 : 𝑥𝛼𝔭 ≡ 0 (mod 𝔭𝑒 )}.
We claim that 𝐿 is a free 𝑅/𝔭𝑒 -module of rank 1. Indeed, 𝐿 is one-dimensional over
𝑅/𝔭 since 𝐼𝔭 is primitive and so 𝛼𝔭 . 0 (mod 𝔭); by Hensel’s lemma, it follows that
𝐿 is also one-dimensional. Therefore, there is a unique generator (𝑎 : 𝑐) ∈ P1 (𝑅/𝔭𝑒 )
for 𝐿. We therefore define an map 𝐼𝔭 ↦→ (−𝑐 : 𝑎) and verify that this furnishes an
inverse to (26.4.4).
428 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
Equivalently,
(
𝜁 𝐹𝔭 (2𝑠), if 𝔭 is ramified;
𝜁 𝐵𝔭 (𝑠) =
𝜁 𝐹𝔭 (2𝑠)𝜁 𝐹𝔭 (2𝑠 − 1), if 𝔭 is split.
Proof. We use Lemma 26.4.1. If 𝐵𝔭 is a division ring, then Lemma 26.4.1(a) applies,
and the result is immediate. For the second case, we compute
∞ ∞
∑︁ 1 + 𝑞 + · · · + 𝑞𝑒∑︁ 1 − 𝑞 𝑒+1
𝜁 𝐵𝔭 (𝑠) = =
𝑒=0
𝑞 2𝑒𝑠 𝑒=0
(1 − 𝑞)𝑞 2𝑒𝑠
∞ ∞
!
1 ∑︁ 1 ∑︁ 1
= −𝑞
1 − 𝑞 𝑒=0 𝑞 2𝑒𝑠 𝑞 (2𝑠−1)𝑒
𝑒=0
(26.4.8)
1 1 𝑞
= −
1 − 𝑞 1 − 1/𝑞 2𝑠 1 − 1/𝑞 2𝑠−1
−1 −1
1 1
= 1 − 2𝑠 1 − 2𝑠−1
𝑞 𝑞
as claimed.
𝜁 𝐹∗ (1) Ö
𝜁 𝐵∗ (1) = lim (𝑠 − 1)𝜁 𝐵 (𝑠) = 𝜁 𝐹 (2) (1 − N(𝔭) −1 ). (26.4.12)
𝑠→1 2
𝔭 |𝔇
Proof. Since 𝜁 𝐹 (𝑠) has only a simple pole at 𝑠 = 1, with residue computed in Theorem
26.2.3, there is a single simple pole of 𝜁 𝐵 (𝑠) at 𝑠 = 1.
O1 := {𝛾 ∈ O× : nrd(𝛾) = 1}
O1 O× nrd 𝑅 ×
1→ → × −−→ ×2 . (26.5.2)
{±1} 𝑅 𝑅
By Dirichlet’s unit theorem, the group 𝑅 × is finitely generated (of rank 𝑟 + 𝑐 − 1), so
the group 𝑅 × /𝑅 ×2 is a finite abelian 2-group. The result follows.
We will examine unit groups in detail in Chapter 32. With this finiteness statement
in hand, we make the following definition.
2
mass(Cls O) = ℎ 𝐹 𝑑 𝐹3/2 𝜑(𝔇)
(2𝜋) 2𝑛
Î
where 𝜑(𝔇) = 𝔭 |𝔇 (N(𝔭) − 1).
Following the strategy in the classical case (to prove the analytic class number
formula), to prove Theorem 26.5.4 we will write 𝜁O (𝑠) as a sum over right ideal
classes and analyze its residue at 𝑠 = 1 by a volume computation.
Then ∑︁
𝜁O (𝑠) = 𝜁O, [𝐽 ] (𝑠).
[𝐽 ] ∈Cls O
We have [𝐼] = [𝐽] if and only if 𝐼 ' 𝐽 if and only if 𝐼 = 𝛼𝐽 for nonzero 𝛼 ∈ 𝐽 −1 .
Since 𝜇𝐽 = 𝐽 if and only if 𝜇 ∈ OL (𝐽) × (Exercise 16.3), it follows that
1 ∑︁ 1
𝜁O, [𝐽 ] (𝑠) = . (26.5.7)
N(𝐽) 𝑠 N(𝛼) 𝑠
0≠𝛼∈𝐽 −1 /OL ( 𝐽 ) ×
Proposition 26.5.10. Let 𝔑 := discrd(O). Then 𝜁O, [𝐽 ] (𝑠) has a simple pole at 𝑠 = 1
with residue
∗ 2𝑛 (2𝜋) 2𝑛 Reg𝐹
𝜁O, [𝐽 ] (1) = .
8𝑤 𝐽 𝑑 𝐹2 N(𝔑)
Proof. We relate residue to volumes using Theorem 26.2.12. We recall (again) 17.7.10:
this gives
𝐽 −1 ↩→ 𝐵 ↩→ 𝐵R := 𝐵 ⊗Q R ' H𝑛 ' R4𝑛
the structure of a Euclidean lattice Λ ⊆ R4𝑛 with respect to the absolute reduced norm.
We take the function 𝑁 in Theorem 26.2.12 to be the absolute norm N (recalling
16.4.8).
26.5. EICHLER MASS FORMULA: MAXIMAL ORDERS 431
We claim that
𝑑 𝐹2 N(𝔑)
covol(O) = . (26.5.11)
2𝑛
By compatible real scaling, it is enough to prove that this relation holds for a single
order O, and we choose the 𝑅-order
O = 𝑅 ⊕ 𝑅𝑖 ⊕ 𝑅 𝑗 ⊕ 𝑅𝑘. (26.5.12)
√ √ 4
The lattice 𝑅 ⊆ 𝐹R has covolume 𝑑 𝐹 , so 𝑅 4 has covolume 𝑑 𝐹 = 𝑑 𝐹2 ; the Z-order
Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑘 has reduced discriminant 4 and covolume 1; and putting these
together, the formula (26.5.11) is verified.
Then (26.5.11) and N(𝐽) = [O : 𝐽] Z = [𝐽 −1 : O] imply that
covol(O) 𝑑 𝐹2 N(𝔑)
covol(Λ) = = 𝑛 . (26.5.13)
N(𝐽) 2 N(𝐽)
Next, the group OL (𝐽) × acts on 𝐽 −1 (and on 𝐵R ); and this group contains 𝑅 × with
finite index 𝑤 𝐽 = [OL (𝐽) : 𝑅 × ], so
1
vol(OL (𝐽) × \𝐵R ) = vol(𝑅 × \𝐵R ). (26.5.14)
𝑤𝐽
Multiplication provides an identification
so
𝑋 ≤1 = 𝑅 × \𝐵R, ≤1 ' (𝐸\𝐹R, ≤1 ) × ({±1}\(H1 ) 𝑛 ) (26.5.15)
where 𝐸 ≤ 𝑅 × is acting by squares. Thus
2𝑛−1 1
vol(𝑅 × \𝐹R, ≤1 ) = Reg𝐹 = Reg𝐹 . (26.5.16)
2(2𝑛 ) 4
Therefore
(2𝜋 2 ) 𝑛 Reg𝐹
vol(𝑋 ≤1 ) = . (26.5.17)
8𝑤 𝐽
From Theorem 26.2.12 together with (26.5.13) and (26.5.17),
Proof of Theorem 26.5.4. We now suppose that O ⊂ 𝐵 is a maximal order, and write
𝜁 𝐵 (𝑠) and 𝜁 𝐵, [𝐽 ] (𝑠). We compare the evaluation of residues given by Corollary
26.4.11 and Proposition 26.5.10. Since 𝜁 𝐵 (𝑠) and each 𝜁 𝐵, [𝐽 ] (𝑠) have simple poles at
𝑠 = 1, we get ∑︁
𝜁 𝐵∗ (1) = ∗
𝜁 𝐵, [𝐽 ] (1).
[𝐽 ] ∈Cls O
432 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
From (26.4.12),
2𝑛
𝜁 𝐹∗ (1) = √ ℎ 𝐹 Reg𝐹
2 𝑑𝐹
since 𝑤 𝐹 = 2 (as 𝐹 is totally real).
Adding the residues from Lemma 26.5.10, we find that
Cancelling, we find
∑︁ 1 2
mass(Cls O) = = 𝜁 (2)𝑑 𝐹3/2 ℎ 𝐹 𝜑(𝔇)
2𝑛 𝐹
(26.5.21)
[𝐽 ] ∈Cls O
𝑤𝐽 (2𝜋)
Remark 26.5.22. For an alternative direct approach in this setting using Epstein zeta
functions, see Sands [San2017].
26.6.1. Let O0 ⊇ O be an 𝑅-superorder, and suppose that there is a prime 𝔭 such that
O𝔮0 = O𝔮 for all primes 𝔮 ≠ 𝔭. We refine the map from Exercise 17.3(b) as follows.
For 𝐼 ⊆ O a right O-ideal, we define the right O0-ideal 𝜌(𝐼) = 𝐼O0 ⊆ O0 obtained by
extension. Then 𝜌 induces a map
Cls O → Cls O0
(26.6.2)
[𝐼] ↦→ [𝐼O0]
that is well-defined and surjective (Exercise 26.5(a)). Let [𝐼 0] ∈ Cls O0 and consider
the set
𝜌 −1 (𝐼 0) = {𝐼 ⊆ O : 𝐼O = 𝐼 0 },
the fiber of the extension map over 𝐼 0.
26.6. EICHLER MASS FORMULA: GENERAL CASE 433
𝐼 h𝜇𝔭 i𝔭 = 𝛽𝔭 𝜇𝔭 O𝔭
and 𝐼 h𝜇𝔭 i𝔮 = 𝐼𝔮 = 𝐼𝔮0 for all 𝔮 ≠ 𝔭, using the local-global dictionary (Theorem 9.4.9).
This defines a right action of O𝔭0× ; it acts simply transitively on 𝜌 −1 (𝐼 0), and the kernel
of this action is visibly the subgroup O×𝔭 . Therefore
#𝜌 −1 (𝐼 0) = [O𝔭0× : O×𝔭 ].
We now look at the classes in the fiber. If 𝜇𝔭 , 𝜈𝔭 ∈ O𝔭0× have [𝐼 h𝜇𝔭 i] = [𝐼 h𝜈𝔭 i] ∈
Cls O, then there exists 𝛼 ∈ 𝐵× such that
𝛼𝐼 h𝜇𝔭 i = 𝐼 h𝜈𝔭 i
as claimed.
In order to apply Proposition 26.6.4, we need to compute the index of unit groups,
a quantity that depends on the (locally defined) Eichler symbol. For a prime 𝔭, we
define
1 + 1/𝑞, if (O | 𝑝) = 1;
1 − Nm(𝔭) −2
𝜆(O, 𝔭) := = 1 − 1/𝑞, if (O | 𝑝) = −1; (26.6.6)
O
Nm(𝔭) −1
1− 1 − 1/𝑞 2 , if (O | 𝑝) = 0.
𝔭
434 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
Proof. We follow Körner [Kör85, §3]. To prove the lemma, we may localize at 𝔭 and
so we drop the subscripts. Let 𝑛 ∈ Z ≥1 be such that 𝔭𝑛 O0 ⊆ 𝔭O. Then
and similarly with O0, all indices taken as abelian groups. Therefore
[1 + 𝔭O0 : 1 + 𝔭𝑛 O0]
= [O0 : O].
[1 + 𝔭O : 1 + 𝔭𝑛 O0]
For the other terms, we recall Lemma 24.3.12. We divide up into the cases, noting
that if(O0 | 𝔭) = −1 then we must have 𝜀 = −1, 0 by classification (Exercise 24.3);
this leaves 6 cases to compute. For example, if (O0 | 𝔭) = ∗ and (O0 | 𝔭) = 1, then
Proof of Main Theorem 26.1.5. We first invoke Theorem 26.5.4 for a maximal order
O0 ⊇ O to get
2 Ö 1
0 3/2
mass(Cls O ) = ℎ 𝐹 𝑑 𝐹 Nm(𝔇) 1− .
(2𝜋) 2𝑛 Nm(𝔭)
𝔭 |𝔇
where Γ(𝑠) is the complex Γ-function (Exercise 26.2). Riemann proved that 𝜉 (𝑠)
extends to a meromorphic function on C and satisfies the functional equation
𝜉 (1 − 𝑠) = 𝜉 (𝑠). (26.8.1)
436 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
(It is also common to multiply 𝜉 (𝑠) by 𝑠(1 − 𝑠) to cancel the poles at 𝑠 = 0, 1.) We
will prove this in section 29.1—the impatient reader is encouraged to flip ahead!
26.8.2. This result extends to Dedekind zeta functions (retaining the notation from
section 26.2). Define
𝜉 𝐹 (1 − 𝑠) = 𝜉 𝐹 (𝑠) (26.8.5)
for all 𝑠 ∈ C. We prove (26.8.5) as Corollary 29.10.3(a) using idelic methods; this
proof will also motivate the completion defined above. For now, we borrow from the
future. The functional equation gives 𝜁 𝐹 (𝑠) meromorphic continuation to C via
𝑠−1/2
| 𝑑𝐹 | Γ(𝑠/2) 𝑟 Γ(𝑠) 𝑐
𝜁 𝐹 (1 − 𝑠) = 𝜁 𝐹 (𝑠) . (26.8.6)
4𝑐 𝜋 𝑛 Γ((1 − 𝑠)/2) 𝑟 Γ(1 − 𝑠) 𝑐
26.8.7. Using the functional equation (26.8.6), we can rewrite (26.2.4) to obtain the
tidier expression
ℎ 𝐹 Reg𝐹
𝜁 𝐹∗ (0) = lim 𝑠−(𝑟 +𝑐−1) 𝜁 𝐹 (0) = ; (26.8.8)
𝑠→0 𝑤𝐹
in particular, 𝜁 𝐹 has a zero at 𝑠 = 0 of order 𝑟 + 𝑐 − 1, the rank of the unit group of 𝑅
by Dirichlet’s unit theorem.
In terms of the completed Dedekind zeta function, we find 𝜉 𝐹 (𝑠) has analytic
continuation to C r {0, 1} with simple poles at 𝑠 = 0, 1 and residues
2𝑟 +𝑐 ℎ 𝐹 Reg𝐹
𝜉 𝐹∗ (0) = 𝜉 𝐹∗ (1) = . (26.8.9)
𝑤𝐹
where Ω ⊆ Ram 𝐵 be the set of real, ramified places in 𝐵. The definition (26.8.12) is
motivated by the simplicity of the functional equation; see also Remark 26.8.15 below.
Written in a different way,
Remark 26.8.15. The completion factors (26.8.12) are not arbitrarily chosen; they have
a natural interpretation from an idelic perspective. Perhaps this serves as a motivation
for working idelically: namely, that it helps to nail down these kinds of quantities! For
more, see section 29.8.
Some properties can be read off easily from (26.8.12).
Proposition 26.8.16 (Analytic continuation, functional equation). Let 𝑚 = # Ram 𝐵.
Then the following statements hold.
(a) 𝜉 𝐵 (𝑠) has meromorphic continuation to C and is holomorphic in C r {0, 1/2, 1}
with simple poles at 𝑠 = 0, 1.
(b) 𝜉 𝐵 (𝑠) satisfies the functional equation
Proof. Statement (a) follows from (26.8.12), recalling that 𝜉 𝐹 (𝑠) is holomorphic in
C r {0, 1} by 26.8.7 with simple poles at 𝑠 = 0, 1. Part (c) follows similarly from (b),
since the other factors in (26.8.12) have a simple zero at 𝑠 = 1/2.
To prove (b), we consider each term in the definition of (26.8.12). The functional
equation (26.8.5) for 𝜉 𝐹 (𝑠) with 𝑠 ← 1 − 𝑠 implies
For
ℓ(𝑠) = 𝑞 𝑠 (1 − 𝑞 1−2𝑠 ) = 𝑞 𝑠 − 𝑞 1−𝑠
and 𝑞 > 0 we have ℓ(1 − 𝑠) = −ℓ(𝑠), so with 𝑞 = Nm(𝔭) the factors 𝔭 | 𝔇 are taken
into account. Finally, 2(1 − 𝑠) − 1 = −(2𝑠 − 1) takes care of 𝑣 ∈ Ω, and (b) follows.
Proof. This theorem was proven by Hey [Hey29, §3] (more generally, for division
algebras over Q) following the same general script as in the proof of the functional
equation for the Dedekind zeta function (26.8.5), as proven first by Hecke: the key
ingredient is Poisson summation. The argument is also given by Eichler [Eic38a, Part
V]. We instead prove this theorem in the language of ideles (Main Theorem 29.2.6),
as it simplifies the calculations—and so for continuity of ideas in the exposition, we
borrow from the future.
Assuming Theorem 26.8.19, we can now deduce the main classification theorem
(Main Theorem 14.6.1) for quaternion algebras over number fields. First, we have
Hilbert reciprocity as an immediate consequence.
Corollary 26.8.22. We have 𝐵 ' M2 (𝐹) if and only if 𝐵 𝑣 ' M2 (𝐹𝑣 ) for all (but one)
places 𝑣 ∈ Pl 𝐹.
Proof. The implication (⇒) is immediate. For the converse (⇐), by Proposition
26.8.16(c), 𝜉 𝐵 (𝑠) has a pole of order 2 − 𝑚 at 𝑠 = 1/2, so if 𝑚 ≤ 1 then 𝜉 𝐵 (𝑠) is not
holomorphic at 𝑠 = 1/2; but then by Theorem 26.8.19, 𝐵 is not a division algebra, so
𝐵 ' M2 (𝐹) (and the order of pole is necessarily 2, and 𝐵 𝑣 ' M2 (𝐹𝑣 ) for all 𝑣).
From this corollary, we are able to deduce the Hasse norm theorem for quadratic
extensions.
Proof. Consider the quaternion algebra 𝐵 = (𝐾, 𝑏 | 𝐹). Then by Main Theorem 5.4.4,
we have 𝑏 ∈ Nm𝐾 |𝐹 (𝐾 × ) if and only if 𝐵 ' M2 (𝐹). By Corollary 26.8.22, this holds
if and only if 𝐵 𝑣 ' M2 (𝐹𝑣 ) for all (but one) places 𝑣. Repeating the application of
Main Theorem 5.4.4, this holds if and only if 𝐵 𝑣 ' M2 (𝐹𝑣 ) for all (but one) 𝑣.
We may similarly conclude the all important local-global principle for quadratic
forms.
Proof. The implication (⇒) is immediate, so we prove (⇐). We may suppose without
loss of generality that 𝑄 is nondegenerate. If 𝑛 = dim𝐹 𝑉 = 1, the theorem is vacuous.
Suppose 𝑛 = 2. Then after scaling we may suppose 𝑄 = h1, −𝑎i, and 𝑄 is isotropic√
if and only if 𝑎 is a square. Suppose for purposes of contradiction that 𝐾 = 𝐹 ( 𝑎)
is a field. Since 𝑄 𝑣 is isotropic for all 𝑣, we have 𝐾 𝑣 ' 𝐹𝑣 × 𝐹𝑣 for all 𝑣, and thus
𝜁 𝐾 (𝑠) = 𝜁 𝐹 (𝑠) 2 . But as Dedekind zeta functions, both 𝜁 𝐹 (𝑠) and 𝜁 𝐾 (𝑠) have poles
of order 1 at 𝑠 = 1 (we evaluated the residue in the analytic class number formula,
Theorem 26.2.3), a contradiction.
Suppose 𝑛 = 3. Again after rescaling we√may suppose 𝑄 = h1, −𝑎, −𝑏i, and 𝑄 is
isotropic if and only if 𝑏 is a norm from 𝐹 [ 𝑎]: then the equivalence follows from
Theorem 26.8.23. √
Next, suppose 𝑛 = 4, and 𝑄 = h1, −𝑎, −𝑏, 𝑐i. Let 𝐾 = 𝐹 ( 𝑎𝑏𝑐). By extension,
𝑄 is isotropic over 𝐾 and all of its completions. But now 𝑄 ' h1, −𝑎, −𝑏, 𝑎𝑏i over
𝐾. Let 𝐵 = (𝑎, 𝑏 | 𝐾). Then by Main Theorem 5.4.4, we have 𝐵 𝑤 ' M2 (𝐾 𝑤 ) for all
𝑤; thus by Corollary 26.8.22 we have 𝐵 ' M2 (𝐾), so 𝐾 splits 𝐵. By 5.4.7, we have
𝐾 ↩→ 𝐵, so there exist 𝑥, 𝑦, 𝑧 ∈ 𝐹 such that
Proof. Apply the same method of proof as in Corollary 14.3.7: see Exercise 26.10.
Proof of Main Theorem 14.6.1, 𝐹 a number field. First, the map 𝐵 ↦→ Ram 𝐵 has the
correct codomain by Hilbert reciprocity (Corollary 26.8.21). Surjectivity follows
by Exercise 14.16 (using Theorem 26.8.26. To conclude, we show injectivity. We
refer to Corollary 5.2.6, giving a bijection between quaternion algebras over 𝐹 up to
isomorphism and ternary quadratic forms of discriminant 1 up to isometry; and we
recall Theorem 12.3.4, that (rescaling) there is a unique anisotropic ternary quadratic
form of discriminant 1 up to isometry. Therefore Corollary 26.8.25 implies that the
map 𝐵 ↦→ Ram 𝐵 is injective, since the set Ram 𝐵 records those places 𝑣 where the
ternary quadratic form attached to 𝐵 is anisotropic.
We will give another proof of Main Theorem 14.6.1 over global fields using
the characterization of idelic norms in Proposition 27.5.15 (avoiding fiddling with
quadratic forms and the use of primes in arithmetic progression).
Remark 26.8.27. For the readers who accept the fundamental exact sequence of class
field theory as in Remark 14.6.10, the arguments above can be run in reverse, and the
analytic statement in Theorem 26.8.19 can be deduced as a consequence.
Exercises
1. Prove Proposition 26.2.18 that
2𝑟 (2𝜋) 𝑐 Reg𝐹
vol(𝑋 ≤1 ) =
𝑤𝐹
𝑟 4 (𝑛) := #{(𝑡, 𝑥, 𝑦, 𝑧) ∈ Z4 : 𝑡 2 + 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛}
so 𝜁 𝐹 (2) 1.
(b) Show that
Nm(𝔑)
𝜑(𝔇)𝜓(𝔐) Nm(𝔑) (log log Nm(𝔐)) .
log log Nm(𝔇)
[Hint: you may need some elementary estimates from analytic number
theory, adapted for this purpose; you may wish to start with the case
𝐹 = Q.]
(c) Conclude (26.8.28).
10. Prove Corollary 26.8.25: if 𝑄, 𝑄 0 are quadratic forms over 𝐹 in the same number
of variables, then 𝑄 ' 𝑄 0 if and only if 𝑄 𝑣 ' 𝑄 0𝑣 for all places 𝑣 ∈ Pl 𝐹. [Hint:
see Corollary 14.3.7.]
I 11. Use Dirichlet’s analytic class number formula to prove the theorem on arithmetic
progressions (Theorem 14.2.9) as follows.
26.8. FUNCTIONAL EQUATION AND CLASSIFICATION 443
(c) Let
∑︁ 1
𝜁O𝑛 (𝑠) := ,
N(𝐼)
𝐼 ⊆O𝑛
the sum over nonzero right ideals of O𝑛 . Show that 𝜁O1 (𝑠) = 𝜁 𝐹 (𝑠) =
(1 − 𝑞 −𝑠 ) −1 .
444 CHAPTER 26. CLASSICAL ZETA FUNCTIONS
Adelic framework
We have already seen that the local-global dictionary is a powerful tool in understanding
the arithmetic of quaternion algebras. In this section, we formalize this connection by
consideration of adeles and ideles.
The basic idea: we want to consider all of the completions of a global field at once.
There are at least two benefits to this approach:
The adelic framework, and its use in class field theory, is a vast topic whose complete
development deserves its own book. We do our best in this chapter to develop this
notation and state what is needed for the case of quaternion algebras. For further
background reading, see Childress [Chi2009] and the references given at the start of
section 27.4
27.1.1. Recall in section 12.1 that for a prime 𝑝 we defined Z 𝑝 = lim Z/𝑝 𝑟 Z as a
←−−𝑟
projective limit, and each Z 𝑝 is compact. We can package these together to make the
direct product ring
Ö
Z :=
b Z𝑝 (27.1.2)
𝑝
445
446 CHAPTER 27. ADELIC FRAMEWORK
We can see bZ itself as projective limit as follows. By the Sun Zi theorem (CRT),
we have an isomorphism
∼ lim Z/𝑛Z
Ö
Z=
b lim Z/𝑝 𝑟 Z −→
←−− ←−−
𝑝 𝑟 𝑛
of topological rings, with the projective limit indexed by positive integers partially
ordered under divisibility; so under this isomorphism, we may identify
Z = lim Z/𝑛Z
b
← −−
𝑛 (27.1.3)
∞
∈ ∞
Î
= (𝑎 𝑛 ) 𝑛=1 𝑛=1 Z/𝑛Z : 𝑎 𝑚 ≡ 𝑎 𝑛 (mod 𝑛) for all 𝑛 | 𝑚
The natural ring homomorphism Z → b Z which takes every element to its reduction
modulo 𝑛 is injective; the image of Z is discrete and dense in b
Z again by the CRT. One
warning is due: b
Z is not a domain.
27.1.4. We now make the ring b Z a bit bigger so that it contains Q as a subring. If
Î
we were to take the ring 𝑝 Q 𝑝 , a product of locally compact rings, unfortunately
we wouldÎno longer have something that is locally compact (see Exercise 27.1): the
product 𝑝 Q 𝑝 is much too big, allowing denominators in every component, whereas
the image of Q will only have denominators in finitely many positions. We should also
keep track of archimedean information at the same time.
With these in mind we define, for each finite set S of primes, the ring
Ö Ö
𝑈S := R × Q𝑝 × Z𝑝 (27.1.5)
𝑝 ∈S 𝑝∉S
equipped with the product topology, so that 𝑈S is locally compact. For example,
𝑈∅ = R × b
Z. (27.1.6)
To assemble these rings together, allowing more denominators and arbitrarily large
sets 𝑆, we take the injective limit of 𝑈S under the natural directed system 𝑈S ↩→ 𝑈S0
for S ⊆ S 0. The resulting object is the restricted direct product of Q 𝑝 relative to Z 𝑝
and is called the adele ring Q of Q:
Ö0 Ö0
Q := R × Q𝑝 = Q𝑣
𝑝 𝑣 ∈Pl Q
Î (27.1.7)
:= R × 𝑥 = (𝑥 𝑝 ) 𝑝 ∈ 𝑝 Q 𝑝 : 𝑥 𝑝 ∈ Z 𝑝 for all but finitely many 𝑝
Î
= 𝑥 = (𝑥 𝑣 ) 𝑣 ∈ 𝑣 Q 𝑣 : |𝑥 𝑣 | 𝑣 ≤ 1 for all but finitely many 𝑣
We declare the sets 𝑈S ⊆ Q with the product topology to be open in Q; and with
this basis of open neighborhoods of 0 (open in 𝑈S for some S), we have given Q the
structureÎof a topological ring. The sets 𝑈S ⊆ Q are also closed. Note that the topology
on Q ⊂ 𝑣 Q 𝑣 is not the subspace topology.
27.1. ⊲ THE RATIONAL ADELE RING 447
for the projection of Q onto the factors away from S. We also write
Ö
QS := Q𝑣 . (27.1.10)
𝑣 ∈𝑆
We embed each of these into Q extending by zero and identify them with their images,
so that Q = Q6S × QS .
Remark 27.1.11. Our notation Q for the adele ring returns to the notation of Weil
[Weil82] but is not standard; more typically, the adele ring is denoted A (which we
find markedly problematic).
Remark 27.1.12. Although Q6S and QS are rings and (via projection) are naturally
quotient rings of Q, they are not subrings because they do not contain 1. This subtlety
should cause no confusion in what follows (especially because we will be focused on
the multiplicative case and working with groups, where there is no issue extending by
the multiplicative identity 1).
We have a natural continuous embedding Q ↩→ Q 𝑣 for all 𝑣 ∈ Pl Q, and this
extends to a diagonal embedding Q ↩→ Q.
Lemma 27.1.13. The diagonal embedding Q ↩→ Q is a continuous injective ring
homomorphism and the image is closed and discrete as a subring of Q.
27.1.15. The proof of Lemma 27.1.14 shows that the natural map R × b Z → Q/Q is
surjective; its kernel is Z diagonally embedded, so we have an isomorphism
∼ (R × b
Q/Q −→ Z)/Z
448 CHAPTER 27. ADELIC FRAMEWORK
···
···
···
···
···
Very often, we will want to tease apart the nonarchimedean and archimedean parts
of the adele ring Q, and will write
Ö0
b := Q {∞} =
Q Q𝑝 ' b
Z ⊗Z Q (27.1.17)
𝑝
b × R.
so that extending by zero we have Q = Q
That is to say, Q× is the restricted direct product of the spaces Q×𝑝 with respect to Z×𝑝 .
The topology is such that for S a finite set of primes, the set
Ö Ö
𝑉𝑆 = R× × Q×𝑝 × Z×𝑝
𝑝 ∈S 𝑝∉S
Remark 27.2.3. Chevalley first used the name élément idéal for elements of Q× , but at
Hasse’s suggestion he abbreviated it to idèle; the name adèle was then shorthand for
an “additive idele”. Anglifying, we drop the accents on these words.
27.2. ⊲ THE RATIONAL IDELE GROUP 449
A direct sum appears because an element of the restricted direct product is a 𝑝-adic
unit for all but finitely many 𝑝. We project Q× onto the product of the first and last
factor, getting a continuous surjective map
Ê
Q× → {±1} × Z. (27.2.8)
𝑝
The idele group of 𝐵 is 𝐵× := 0𝑣 𝐵×𝑣 , the restricted direct product of the topo-
Î
logical groups 𝐵×𝑣 with respect to O×𝑣 ; we similarly define
Ö Ö0
b× :=
O O×𝑝 ≤ 𝐵b× := 𝐵×𝑝 . (27.3.5)
𝑝 𝑝
[𝐼] R ↦→ 𝐵× b
𝛼Ob× .
The most fundamental result in this chapter is the following (see Main Theorem
27.6.14, taking 𝐹 = Q).
Theorem 27.3.7. Let 𝐵 be a division quaternion algebra over Q. Then 𝐵× ≤ 𝐵× is
cocompact and the set 𝐵× \ 𝐵
b× /O
b× is finite.
In particular, combining Lemma 27.3.6 and Theorem 27.3.7, we conclude that the
class set of O is finite, something we proved earlier using the geometry of numbers for
𝐵 definite in section 17.5.
27.4. ADELES AND IDELES 451
27.4.1. We recall notation from section 14.4 for convenience. The set of places of 𝐹
is denoted Pl 𝐹. For a place 𝑣 of 𝐹, we denote by 𝐹𝑣 the completion of 𝐹 at the place
𝑣, with preferred (normalized) absolute value | | 𝑣 so that the product formula holds in
𝐹: see 14.4.12. If 𝑣 is nonarchimedean, we let
be the valuation ring of 𝐹𝑣 , where we write 𝑣 also for the discrete valuation associated
to the place 𝑣. If 𝐹 is a number field, we will sometimes denote an archimedean place
by writing 𝑣 | ∞, and for an archimedean place we just take 𝑅 𝑣 = 𝐹𝑣 . A set S ⊆ Pl 𝐹
of places is eligible if it is finite, nonempty, and contains all archimedean places.
27.4.3. The adele ring of 𝐹 is the restricted direct product of 𝐹𝑣 with respect to 𝑅 𝑣 :
Ö0
𝐹 := 𝐹𝑣
𝑣
Î (27.4.4)
:= {(𝑥 𝑣 ) 𝑣 ∈ 𝑣 𝐹𝑣 : 𝑥 𝑣 ∈ 𝑅 𝑣 for all but finitely many 𝑣}
Î
= {(𝑥 𝑣 ) 𝑣 ∈ 𝑣 𝐹𝑣 : |𝑥 𝑣 | 𝑣 ≤ 1 for all but finitely many 𝑣}
with the restricted direct product topology. The Îtopology is uniquely characterized
(as a topological ring) by the condition that 𝑅 := 𝑣 𝑅 𝑣 (with the product topology)
is open. Accordingly,
Î a subset 𝑈 ⊆ 𝐹 is open if and only if for all 𝑎 ∈ 𝐹, the set
(𝑎 + 𝑈) ∩ 𝑣 𝑅 𝑣 is open in the product topology.
and extending by zero we identify these sets with their images in 𝐹, so that 𝐹 = 𝐹6S ×𝐹S ;
we call 𝐹6S the S-finite adele ring of 𝐹.
452 CHAPTER 27. ADELIC FRAMEWORK
27.4.8. The idele group of 𝐹 is the restricted direct product of 𝐹𝑣× with respect to 𝑅 ×𝑣 :
Ö0
𝐹 × := 𝐹𝑣×
𝑣 (27.4.9)
:= (𝑥 𝑣 ) 𝑣 ∈ 𝑣 𝐹𝑣× : |𝑥 𝑣 | 𝑣 = 1 for all but finitely many 𝑣 .
Î
𝐴× ↩→ 𝐴 × 𝐴
𝑥 ↦→ (𝑥, 𝑥 −1 ).
The justification for calling this the idele class group is given in section 27.5.
27.4.13. With respect to the normalized absolute values 14.4.12, we have a natural
map
𝐹 × → R>0
(27.4.14)
Ö
(𝑥 𝑣 ) 𝑣 ↦→ k𝑥 𝑣 k 𝑣 .
𝑣
When 𝐹 is a number field, the map (27.4.14) is surjective; when 𝐹 is a function field
with constant field F𝑞 , the image is 𝑞 Z , the cyclic subgroup of R>0 generated by 𝑞. Let
𝐹 (1) := {𝑥 = (𝑥 𝑣 ) 𝑣 :
Î
𝑣 k𝑥 𝑣 k 𝑣 = 1} (27.4.15)
so that 𝐹 (1) is the kernel of (27.4.14). Then 𝐹 × ≤ 𝐹 (1) by the product formula (14.4.6).
27.5. CLASS FIELD THEORY 453
Proof. We give a proof in a more general context in Main Theorem 27.6.14 below. Or
see e.g. Cassels [Cas2010, §16, p. 69] for a direct proof.
Theorem 27.4.16 is equivalent (!) to the Dirichlet unit theorem and the finiteness
of the class group in the number field case, and finite generation of the unit group of a
coordinate ring of a curve and the finiteness of the group of rational divisors of degree
zero in the function field case [Cas2010, §§17–18].
Via the projection map 𝐹 (1) → 𝐹6S× , we have 𝐹 × cocompact also in 𝐹6S× .
27.5.1. Let 𝑅 = 𝑅 (S) be a global ring (the ring of S-integers) for the eligible set
S ⊆ Pl 𝐹. Then 𝑅 is a Dedekind domain with field of fractions 𝐹. The class group of
𝑅 admits an idelic description, embodying the definitions above, as follows.
To simplify notation, throughout we abbreviate 𝐹6𝑆 = 𝐹,b as we take the set S to be
fixed. To an invertible fractional ideal 𝔞 ⊆ 𝐹 of 𝑅, we have 𝔞𝔭 = 𝑅𝔭 for all but finitely
many primes 𝔭, so we can consider its idelic image (𝔞𝔭 )𝔭 ⊆ 𝐹 b under the product of
completions. Since 𝔞 is locally principal, we can write each 𝔞𝔭 = 𝑎 𝔭 𝑅𝔭 with 𝑎 𝔭 ∈ 𝐹𝔭× ,
well-defined up to an element of 𝑅𝔭× ; putting these together we obtain an element
b×
𝑎 = (𝑎 𝔭 )𝔭 ∈ 𝐹
b
and
𝔞 = 𝔞𝑅
b b= b
𝑎𝑅b ⊆ 𝐹.
b
We recover 𝔞 = b 𝔞 ∩ 𝐹 from Lemmas 9.4.6 and 9.5.3. Therefore the group of invertible
fractional ideals of 𝑅 is canonically isomorphic to the quotient
b× / 𝑅
Idl 𝑅 ' 𝐹 b× . (27.5.2)
In idelic class field theory, it is often convenient to move between quotients of the
finite ideles and quotients of the full idele group as follows.
Î
Lemma 27.5.6. Let 𝐹∞ := 𝑣 |∞ 𝐹𝑣 and let
×
𝐹∞,>0 := {(𝑎 𝑣 ) 𝑣 ∈ 𝐹∞ : 𝑎 𝑣 > 0 for all 𝑣 real}.
27.5.9. Class field theory relates class groups to abelian extensions. For example, let
Ö Ö
𝐻=𝑅 b× × 𝐹S× = 𝑅 ×𝑣 × 𝐹𝑣× ≤ 𝐶𝐹 := 𝐹 × /𝐹 × .
𝑣∉S 𝑣 ∈S
The main theorem of idelic class field theory for finite extensions is as follows.
27.5. CLASS FIELD THEORY 455
27.5.12. Rewriting the main theorem (Theorem 27.5.10) slightly, we see that if 𝐻 ≤ 𝐹 b×
×
is an open finite-index subgroup containing 𝐹>0 , then there is a finite abelian extension
𝐾 ⊇ 𝐹 with the Artin isomorphism
∼ Gal(𝐾 | 𝐹).
b× /𝐻 −→
𝐹
called the global Artin homomorphism, where 𝐹 ab ⊆ 𝐹 sep is the maximal abelian
extension of 𝐹 in 𝐹 sep .
If 𝐹 is a number field, then 𝜃 is surjective; let 𝐷 𝐹 be the connected component of
1 in 𝐶𝐹 . Then 𝐷 𝐹 is a closed subgroup with
Proof. Let 𝐾 ⊇ 𝐹 be a separable quadratic extension that is inert (an unramified field
extension) at every 𝑣 ∈ Σ: such an extension exists by Exercise 14.20. By the main
theorem of class field theory, we have [𝐶𝐹 : 𝐹 × Nm𝐾 /𝐹 𝐶𝐾 ] = [𝐾 : 𝐹] = 2, where
𝐶𝐹 = 𝐹 × /𝐹 × and similarly 𝐶𝐾 are idele class groups. Therefore
as well.
For each 𝑣 ∈ Σ, let 𝜋 𝑣 be a uniformizer for 𝑅 𝑣 and if 𝑣 is real let 𝜋 𝑣 = −1. Since
𝐾 𝑣 ⊇ 𝐹𝑣 is an unramified field extension, we have 𝜋 𝑣 ∉ Nm𝐾𝑣 /𝐹𝑣 (𝐾 𝑣× ). For 𝑣 ∈ Σ,
let 𝜋 𝑣 = (1, . . . , 1, 𝜋 𝑣 , . . . ) ∈ 𝐹 × . Then 𝜋 𝑣 ∉ Nm𝐾 /𝐹 (𝐾 × ).
456 CHAPTER 27. ADELIC FRAMEWORK
27.6.1. The adele ring of 𝐵 is the restricted direct product of the topological rings 𝐵 𝑣
with respect to O 𝑣 :
Ö0 Î
𝐵 := 𝐵 𝑣 = {(𝛼 𝑣 ) 𝑣 ∈ 𝑣 𝐵 𝑣 : 𝛼 𝑣 ∈ O 𝑣 for all but finitely many 𝑣 ∉ S}
𝑣 ∈Pl(𝐹 )
The topology onÎ𝐵 (as a topological ring) is uniquely characterized by the property
that the subring 𝑣 O 𝑣 is open with the product topology.
By the local-global dictionary for lattices (Theorem 9.4.9), the definition of 𝐵 is
independent of the choice of order O and eligible set S (and base ring 𝑅 = 𝑅 (S) ).
We now turn to the multiplicative structure, the main object of our concern.
27.6.3. The idele group of 𝐵 is the restricted direct product of the topological groups
𝐵×𝑣 with respect to O×𝑣 :
Ö0
𝐵× := 𝐵×𝑣 = (𝛼 𝑣 ) 𝑣 ∈ 𝑣 𝐵×𝑣 : 𝛼 𝑣 ∈ O×𝑣 for all but finitely many 𝑣 ;
Î
𝑣
27.6. NONCOMMUTATIVE ADELES 457
equivalently, 𝐵× is the unit group of 𝐵 with the topology as in 27.4.10. The topology
Î on
𝐵× as a topological group is characterized by the condition that the subgroup 𝑣 O×𝑣
is open with the product topology. Again, 𝐵× is independent of the choice of O and
eligible set S because any two such constructions differ at only finitely many places.
extending by zero, we may identify 𝐵6S with its image in 𝐵. The S-finite adele ring has
a compact open subring Ö
O6S := O 𝑣 ⊆ 𝐵6S . (27.6.6)
𝑣∉S
We similarly define the S-finite idele group with its compact open subgroup
Ö0 Ö
𝐵6S× := 𝐵×𝑣 ⊃ O×𝑣 =: O×6S . (27.6.7)
𝑣∉S 𝑣∉S
When no confusion can result (S is clear from context), we will drop the superscript
b = 𝐵6S and O× = O
and replace with hats, writing simply 𝐵 b× , etc.
6S
Just as in 27.5.1, the ideles provide a convenient way of encoding fractional ideals,
as follows.
Lemma 27.6.8. The set of locally principal, right fractional O-ideals is in bijection
b×/O
with 𝐵 b× via the map 𝐼 ↦→ b
𝛼Ob× , where 𝐼𝔭 = 𝛼𝔭 O𝔭 and b
𝛼 = (𝛼𝔭 )𝔭 ; this map induces
a bijection
ClsR O ↔ 𝐵× \ 𝐵 b×/O b×
(27.6.9)
[𝐼] R ↦→ 𝐵× b
𝛼Ob× .
Proof. Let 𝐼 be a locally principal right fractional O-ideal, so 𝐼𝔭 = 𝛼𝔭 O𝔭 for all primes
𝔭 of 𝑅, with 𝛼𝔭 well-defined up to right multiplication by an element of O×𝔭 , so to 𝐼
we associate (𝛼𝔭 O×𝔭 )𝔭 = b𝛼Ob× ∈ 𝐵 b×/Ob× . Conversely, given b 𝛼 ∈ 𝐵b×/Ob× we recover
𝐼 =b𝛼Ob ∩ 𝐵 from Lemmas 9.4.6 and 9.5.3.
The equivalence relation defining the (right) class set is given by left multiplication
by 𝐵× , so the second statement follows.
Remark 27.6.10. We recall by Main Theorem 16.6.1 that for 𝐵 a quaternion algebra,
an 𝑅-lattice 𝐼 ⊂ 𝐵 is locally principal if and only if it is invertible.
k k : 𝐵× → R>0
(27.6.12)
Ö
𝛼 = (𝛼 𝑣 ) 𝑣 ↦→ |nrd(𝛼 𝑣 )| 𝑣
𝑣
458 CHAPTER 27. ADELIC FRAMEWORK
and we define
𝐵 (1) = kerk k = {𝛼 = (𝛼 𝑣 ) 𝑣 :
Î
𝑣 |nrd(𝛼 𝑣 )| 𝑣 = 1} (27.6.13)
𝐵1 = {𝛼 ∈ 𝐵 : nrd(𝛼) = 1}
𝐵1 = {𝛼 ∈ 𝐵 : nrd(𝛼) = 1}
satisfying 𝐵1 ≤ 𝐵1 ≤ 𝐵 (1) .
The following theorem is fundamental (see Fujisaki [Fuj58, Theorem 8.3], Weil
[Weil82, Lemma 3.1.1]).
Main Theorem 27.6.14 (Fujisaki’s lemma). Let 𝐵 be a finite-dimensional division
algebra over a global field 𝐹. Then the following statements hold.
(a) 𝐵× ≤ 𝐵 (1) is cocompact.
b× is cocompact and the set
(b) For any eligible set S, the subgroup 𝐵× ≤ 𝐵6S× = 𝐵
𝐵× \ 𝐵
b×/O
b× is finite.
Proof. The natural place to prove this result is after some more serious analysis has
been done—but it is too important to wait for this. The small amount of input needed,
which can be seen as an (ineffective) idelic version of the Minkowski convex body
theorem (Theorem 17.5.5), is as follows. There exists a compact subset 𝐸 such that
for all 𝛽 ∈ 𝐵 (1) ,
For the proof of (27.6.15), see Exercise 29.11: in a nutshell, there is a measure 𝜇 on 𝐵 in
which 𝜇(𝐵\𝐵) < ∞, and a compact 𝐸 with 𝜇(𝐸) satisfies 27.6.15, as 𝜇(𝛽𝐸) = 𝜇(𝐸).
We first quickly prove part (b), assuming part (a). We have O b× open in 𝐵 b× , so the
open cover {𝐵× b 𝛼Ob× } b× can be reduced to a finite cover, whence the double coset
b∈ 𝐵
𝛼
space is finite.
We now turn to prove (a), which we do in steps.
Step 1: Setup. From the set 𝐸 granted above, we define
𝑋 := 𝐸 − 𝐸 = {𝜂 − 𝜂 0 : 𝜂, 𝜂 0 ∈ 𝐸 }. (27.6.16)
To check the claim, let 𝛽 ∈ 𝐵 (1) . By Step 2, 𝛽𝑋∩𝐵× ≠ ∅ and (similarly) 𝑋 𝛽−1 ∩𝐵× ≠ ∅.
Therefore there exist 𝜈, 𝜈 0 ∈ 𝑋 and 𝛽, 𝛽 0 ∈ 𝐵× such that
Therefore
𝛽 0 𝛽 = (𝜈 0 𝛽−1 ) (𝛽𝜈) = 𝜈 0 𝜈 ∈ 𝐵× ∩ 𝑋 𝑋.
{𝜈 ∈ 𝐵 (1) : (𝜈, 𝜈 −1 ) ∈ 𝐾 }
is compact; then, by the claim in Step 3, this set surjects onto 𝐵× \𝐵 (1) . We conclude
that 𝐵× \𝐵 (1) is compact, completing the proof of part (a).
Remark 27.6.21. Corollary 27.6.20 covers all quaternion algebras 𝐵. This finiteness
statement generalizes to the theorem of Jordan–Zassenhaus: see Remark 17.7.27.
27.6.22. The idelic point of view (Lemma 27.6.8) also makes it clear why the class
number is independent of the order within its genus (Lemma 17.4.11): the idelic
description only depends on the local orders, up to isomorphism.
27.6.23. The genus of an order and its type set (see section 17.4) can be similarly
described. Let O be an 𝑅-order, and let O0 ∈ Gen O be an order in the genus of O.
Then O0 is locally isomorphic to O, so there exists b𝜈 ∈ 𝐵 b× such that b b𝜈 −1 = O
𝜈 Ob b0 ,
well defined up to right multiplication by an element of the normalizer 𝑁 𝐵b× ( O).
b We
0
recover O = O 0
b ∩ 𝐵, so this gives a bijection
b×/𝑁 b× ( O).
Gen O ↔ 𝐵 b
𝐵
Two such orders are isomorphic if and only if there exists 𝛽 ∈ 𝐵× such that 𝛽O𝛽−1 = O0,
so we have a bijection
Typ O ↔ 𝐵× \ 𝐵 b×/𝑁 b× ( O).
𝐵
b (27.6.24)
27.6.26. Referring to section 18.5 locally, we see that the group of locally principal
two-sided O-ideals Idl(O) is in bijection with
b× \𝑁 b× ( O)/
O b O b× = 𝑁 b× ( O)/
b O b× = O
b× \𝑁 b× ( O)
b (27.6.27)
𝐵 𝐵 𝐵
where
𝑁 𝐵b× ( O)
b = {b b× : b
𝛼∈𝐵 𝛼Ob = Ob
b𝛼 }
b× = 𝑁 b× ( O)/ b× = 𝑁 𝐵× (O) O
b× \𝑁 b× ( O).
𝑁 𝐵× (O)\𝑁 𝐵b× ( O)/
b O
𝐵
b 𝑁 𝐵× (O) O
𝐵
b
27.7.1. Since S contains all archimedean places, by the local norm calculation (Lemma
13.4.9), we have nrd( 𝐵b× ) = 𝐹 b× . By the Hasse–Schilling theorem on norms (Main
Theorem 14.7.4), we have nrd(𝐵× ) = 𝐹>×Ω 0 , where Ω ⊆ Ram 𝐵 is the set of real
ramified places and 𝐹>×Ω 0 is the set of elements positive at all 𝑣 ∈ Ω (recalling 14.7.2).
Therefore, the reduced norm (in each component) yields a surjective map
nrd : 𝐵× \ 𝐵
b×/O
b× → 𝐹 × \𝐹
>Ω 0
b×/nrd( O
b× ). (27.7.2)
b× ) = 𝑅
then nrd( O b× .
Proof. In Lemma 13.4.9, we saw that if O𝔭 is maximal, then nrd(O×𝔭 ) = 𝑅𝔭× ; for
the finitely many remaining 𝔭 ⊆ 𝑅, the 𝑅𝔭 -order O𝔭 is of finite index in a maximal
superorder, so nrd(O×𝔭 ) is a finite index open subgroup of 𝑅𝔭× . Putting these together,
we conclude nrd( O b× ) is a finite index open subgroup of 𝑅
b× .
But 𝐹b×/ 𝑅
b× 𝐹 × ' ClΩ 𝑅 is a finite group and therefore
>Ω 0
b× : nrd( O
[𝐹 b× )𝐹 × ] = [ 𝐹
b× : 𝑅
b× 𝐹 × ] [ 𝑅
b× 𝐹 × : nrd( O
b× )𝐹 × ] < ∞.
>Ω 0 >Ω 0 >Ω 0 >Ω 0
27.7.4. Let
𝐺 (O) := 𝐹>×Ω 0 nrd( O
b× ) ≤ 𝐹
b× . (27.7.5)
From Lemma 27.7.3, 𝐺 (O) is a finite-index open subgroup containing 𝐹>0 × . By class
field theory 27.5.12, there exists a class field 𝐾 for 𝐺 (O), i.e., there exists a finite
abelian extension 𝐾 ⊇ 𝐹 and an Artin isomorphism
∼ Gal(𝐾 | 𝐹).
Cl𝐺 (O) 𝑅 = 𝐹 × /𝐺 (O) −→ (27.7.6)
The group Cl𝐺 (O) 𝑅 only depends on the genus of O: if O0 ∈ Gen O then O0 is locally
isomorphic to O, so there exists b b× such that O
𝜈∈𝐵 b0 = b
𝜈 −1 Ob b0× ) = nrd( O
b𝜈 so nrd( O b× ).
Example 27.7.7. Suppose 𝐹 is a number field and S is the set of archimedean places
of 𝐹, so that 𝑅 = Z𝐹 is the ring of integers in 𝐹. Suppose further that O is maximal.
Then 𝐺 (O) = 𝐹>×Ω 0 𝑅
b× . Recalling 17.8.2, let Ω be the set of ramified (necessarily real)
archimedean places of 𝐵, and let ClΩ Z𝐹 := 𝐹>×Ω 0 \𝐹 b× / 𝑅
b× , equivalently, ClΩ Z𝐹 is the
group of fractional ideals of Z𝐹 modulo the subgroup of principal fractional ideals
generated by elements in 𝐹>×Ω 0 . Then Cl𝐺 (O) Z𝐹 = ClΩ Z𝐹 by definition, so we have
a surjective map
nrd : Cls O → Cl𝐺 (O) Z𝐹 = ClΩ Z𝐹 .
The two extreme cases: if 𝐵 is unramified at all real places, then Ω = ∅, and
ClΩ Z𝐹 = Cl Z𝐹 is the class group; if 𝐵 is ramified at all real places, then ClΩ Z𝐹 =
Cl+ Z𝐹 is the narrow class group.
Exercises
Unless otherwise specified, let 𝐹 be a global field and let 𝐵 be a quaternion algebra
over 𝐹.
1. If we take the direct product instead of the restricted direct product in the
definition of the adele ring, we lose local compactness. More precisely, let
Î
{𝑋𝑖 }𝑖 ∈𝐼 be a collection of nonempty topological spaces. Show that 𝑋 = 𝑖 ∈𝐼 𝑋𝑖
is locally compact if and only if each 𝑋𝑖 is locally compact and all but finitely
many 𝑋𝑖 are compact.
2. Review the language of group actions and fundamental sets (section 34.1).
a) Equip Q with the discrete topology. We have a group action Q Q (by
Z × [0, 1] ⊆ Q
addition). Show that b b × R is a fundamental set for this action.
[Hint: Review the arguments in Lemmas 27.1.13–27.1.14.]
b) Similarly, show that Q× Q× and that bZ× × R>0 ⊆ Qb × R is a fundamental
set for this action.
3. For a prime 𝑝, let 𝑝b = ( 𝑝, 1, . . . , 1, 𝑝, 1, . . . ) ∈ Q be the adele which is equal to
𝑝 in the 𝑝th and ∞th component and 1 elsewhere.
a) Show that the sequence 𝑝b, ranging over primes 𝑝, does not converge in
Q× ; conclude that Q× is not compact.
b) However, show that the sequence 𝑝b has a subsequence converging to the
identity in the quotient Q× /Q× .
Î
Z = lim Z/𝑛Z ' 𝑝 Z 𝑝 .
4. Recall that b
←−−𝑛
𝛼= ∞
Í
(a) Prove that each b 𝛼∈b Z has a unique representation as b 𝑛=1 𝑐 𝑛 𝑛! where
𝑐 𝑛 ∈ Z and 0 ≤ 𝑐 𝑛 ≤ Î𝑛.
(b) Prove that b Z× ' bZ× ∞ 𝑛=1 Z/𝑛Z as profinite groups. [Hint: Consider the
product of the 𝑝-adic logarithm maps and use the fact that for every prime
power 𝑝 𝑒 there are infinitely many primes 𝑞 such that 𝑝 𝑒 k (𝑞 − 1).]
(c) Prove for every 𝑛 ∈ Z>0 that the natural map Z/𝑛Z → b Z/𝑛bZ is an isomor-
phism.
(d) Prove that there is a bijection from Z ≥0 to the set of open subgroups of b Z
mapping 𝑛 ↦→ 𝑛b Z.
5. We recall from 27.1.15 the solenoid Sol = (R × b Z)/Z (with Z embedded diago-
nally, and given the quotient topology).
(a) Prove that Sol is a compact, Hausdorff, and connected topological group.
(b) Prove that Sol ' Q/Q as topological groups.
(c) Prove that Sol ' lim R/ 𝑛1 Z with respect to the directed system 𝑛 ∈ Z ≥1
←−−𝑛
under divisibility.
(d) Show that the group of path components of Sol is isomorphic to b Z/Z, and
conclude that Sol is not path connected. [Hint: the neutral path component
is the image of {0} × R ⊆ Q b × R.]
27.7. REDUCED NORMS 463
6. Recall that by definition, a set 𝑈 ⊆ 𝐹 is open if and only if 𝑈 ∩ 𝐹6S is open in 𝐹6S
S ⊆ Pl 𝐹. Show that
for all eligible Î Î 𝑈 ⊆ 𝐹 is open if and only if for all 𝑎 ∈ 𝐹
that (𝑎 + 𝑈) ∩ 𝑣 𝑅 𝑣 is open in 𝑣 𝑅 𝑣 .
7. Show that the topology on 𝐹 × agrees with the subspace topology induced on
𝐹 × ↩→ 𝐹 × 𝐹 by 𝑥 ↦→ (𝑥, 𝑥 −1 ).
8. Show that if S is eligible and O ⊆ 𝐵 is an 𝑅 (S) -order, then
Î
𝐵 = {(𝑥 𝑣 ) 𝑣 ∈ 𝑣 𝐵𝑣 : 𝑥 𝑣 ∈ O 𝑣 for all but finitely many 𝑣}
and
𝐵× = (𝑥 𝑣 ) 𝑣 ∈ 𝑣 𝐵×𝑣 : 𝑥 𝑣 ∈ O×𝑣 for all but finitely many 𝑣
Î
and therefore that this definition is independent of the choice of O (and S).
I 9. Prove Lemma 27.5.6.
10. Returning to 27.5.13, let 𝐷 𝐹 ≤ 𝐶𝐹 be the connected component of 1 in the idele
class group of 𝐹. Show that 𝐷 𝐹 is a closed subgroup with
where 𝑟 is the number of real places of 𝐹 and 𝑐 the number of complex places.
Interpret this isomorphism explicitly for 𝐹 a quadratic field for both 𝐹 real and
imaginary: what ‘explains’ the factors that appear?
I 11. Let 𝐵 be a finite-dimensional 𝐹-algebra.
(a) Show that 𝐵 is discrete and closed in 𝐵. [Hint: 𝐹 is discrete in 𝐹 by the
product formula.]
(b) Show that 𝐵 is cocompact in 𝐵 (under the diagonal embedding), i.e., that
𝐵/𝐵 is compact.
(c)
12. Give a fundamental system of neighborhoods of 0 in 𝐵 b× .
b and of 1 in 𝐵
I 13. Let 𝐴 be a topological ring.
(a) Suppose that 𝐴× ⊆ 𝐴 has the induced topology. Give an example to show
that the map 𝑥 ↦→ 𝑥 −1 on 𝐴× is not necessarily continuous.
(b) Now embed
𝐴× ↩→ 𝐴 × 𝐴
𝑥 ↦→ (𝑥, 𝑥 −1 )
Î
(b) Prove that if O is an 𝑅 (S) -order in 𝐵, then O is discrete in 𝐵S = 𝑣 ∈S 𝐵𝑣 .
15. Let 𝐹 be a global field and let 𝐾 be a finite separable extension of 𝐹.
Î
(a) Show that 𝐾 ' 𝐹 ⊗𝐹 𝐾. [Hint: Use the fact that 𝐹𝑣 ⊗𝐹 𝐾 ' 𝑤 𝐾 𝑤
where 𝑤 runs over the places above 𝑣.]
(b) Show that
Î
𝐾 = (𝑥 𝑤 ) 𝑤 ∈ 𝑤 𝐾 𝑤 : |Nm𝐾𝑤 /𝐹𝑣 𝑥 𝑣 | 𝑣 ≤ 1 for almost all 𝑣
Strong approximation
28.1 ⊲ Beginnings
We have already seen in several places in this book how theorems about quaternion
algebras over global fields are often first investigated locally, and then a global result is
recovered using some form of approximation. Approximation provides a way to trans-
fer analytic properties (encoded in congruences or bounds) into global elements. In
this chapter, we develop robust approximation theorems and investigate their arithmetic
applications.
We begin by reviewing weak and strong approximation over Q, taking a breath in
preparation for the idelic efforts to come.
28.1.1. The starting point is the Sun Zi theorem (CRT): given a finite, nonempty set
S of primes, and for each 𝑝 ∈ S an exponent 𝑛 𝑝 ∈ Z ≥1 and an element 𝑥 𝑝 ∈ Z/𝑝 𝑛 𝑝 Z,
there exists 𝑥 ∈ Z such that 𝑥 ≡ 𝑥 𝑝 (mod 𝑝 𝑛 𝑝 ) for all 𝑝 ∈ S. These congruences can
be equivalently formulated in the 𝑝-adic metric by lifting Î to 𝑥 𝑝 ∈ Z 𝑝 and asking for
|𝑥 −𝑥 𝑝 | < 𝑝 −𝑛 𝑝 for 𝑝 ∈ S; or equivalently, the map Z → 𝑝 ∈S Z 𝑝 has dense image for
any finite, nonempty set of primes S (giving the target the product topology). We may
therefore think of the CRT as an approximation theorem, in the sense that it allows us
to find an integer that simultaneously approximates a finite number of 𝑝-adic integers
arbitrarily well.
465
466 CHAPTER 28. STRONG APPROXIMATION
strong approximation below; for a short, uniform proof (which reproves the CRT), see
Exercise 28.1.
Case 1. Suppose ∞ ∉ S and 𝑥 𝑝 ∈ Z 𝑝 for all 𝑝 ∈ S. Then (28.1.3) holds by the CRT,
as in 28.1.1; in fact, we have infinitely many 𝑥 ∈ Z for this purpose.
Case 2. Suppose ∞ ∉ S, but 𝑥 𝑝 ∈ Q 𝑝 for 𝑝 ∈ S. We employ continuity of
multiplication to reduce to the previous case, as follows. We consider the least
common denominator:
Ö
𝑑 := 𝑝 max(0,−𝑣𝑝 ( 𝑥 𝑝 )) ∈ Z>0 . (28.1.4)
𝑝 ∈S
Then 𝑑𝑥 𝑝 ∈ Z 𝑝 for all 𝑝 ∈ S. By the case just established by the CRT, there exists
𝑥 0 ∈ Z such that |𝑥 0 − 𝑑𝑥 𝑝 | 𝑝 < 𝜖 | 𝑑| 𝑝 for all 𝑝 ∈ S, so taking 𝑥 := 𝑥 0/𝑑 and dividing
through we conclude that (28.1.3) holds.
Case 3. To conclude, suppose ∞ ∈ S. We employ an additive translation: we find
a rational number close to 𝑥∞ and add to it a small solution to the previous case, as
follows. Since Q ⊆ R is dense, there exists 𝑦 ∈ Q such that | 𝑦 − 𝑥∞ | ∞ < 𝜖/2. Let
𝑦 𝑝 := 𝑥 𝑝 − 𝑦. From case 2, we find 𝑦 0 ∈ Q such that | 𝑦 0 − 𝑦 𝑝 | 𝑝 < 𝜖 for all 𝑝 ∈ S. By
case 1, there exist infinitely many 𝑚 ∈ Z such that |1 − 𝑚| 𝑝 < min(1, | 𝑦 0 − 𝑦 𝑝 | 𝑝 ) for
all 𝑝 ∈ S; for such 𝑚, we have 𝑚 ≡ 1 (mod 𝑝) so |𝑚| 𝑝 = 1 and
| 𝑦 0 − 𝑚𝑦 𝑝 | 𝑝
| (𝑦 0/𝑚) − 𝑦 𝑝 | 𝑝 = = | 𝑦 0 − 𝑦 𝑝 + (1 − 𝑚)𝑦 𝑝 | 𝑝 = | 𝑦 0 − 𝑦 𝑝 | 𝑝 < 𝜖 (28.1.5)
|𝑚| 𝑝
and
|𝑥 − 𝑥∞ | ∞ ≤ |𝑥 − 𝑦| ∞ + | 𝑦 − 𝑥∞ | ∞ = | 𝑦 0/𝑚| ∞ + 𝜖/2 < 𝜖 (28.1.7)
proving (28.1.3).
28.1.9. Written out in the standard basis of open sets, strong approximation is equiv-
alent to: given a finite set T ⊆ Pl Q disjoint from S, elements 𝑥 𝑣 ∈ Q 𝑣 for 𝑣 ∈ T, and
𝜖 > 0, there exists 𝑥 ∈ Q such that |𝑥 − 𝑥 𝑣 | 𝑣 < 𝜖 for all 𝑣 ∈ T and 𝑥 ∈ Z 𝑝 for all
𝑝 ∉ S t T with 𝑝 ≠ ∞.
Weak approximation follows from strong approximation by forgetting S and weakly
approximating 𝑥 𝑣 for 𝑣 ∈ T. Indeed, the difference between the ‘weak’ and the
‘strong’ is meaningful here. In weak approximation, we satisfy only a finite number
of conditions, with no control over the rational number at places 𝑣 ∉ S. By contrast, in
strong approximation, the role of the set S is switched, and have specified conditions
28.2. ⊲ STRONG APPROXIMATION FOR SL2 (Q) 467
Proof of Theorem 28.1.8. We prove the statement in its formulation 28.1.9. Naturally,
we return to the proof of weak approximation. Without loss of generality (proving a
stronger statement), we may assume #S = 1.
If S = {∞}, we apply step 2 of weak approximation over the set T: the result
𝑥 = 𝑥 0/𝑑 already has 𝑥 ∈ Z𝑞 for 𝑞 ∉ T, since then 𝑞 - 𝑑.
So suppose S = {ℓ} with ℓ prime. We return to case 3. To define 𝑦, we note instead
that Z[1/𝑝] ⊆ R is dense, so we may take 𝑦 ∈ Z[1/ℓ], so in particular 𝑦 ∈ Z𝑞 for
𝑞 ≠ ℓ. We just showed that we may take 𝑦 0 ∈ Z𝑞 for 𝑞 ∉ T. And for the integers 𝑚, we
claim we may take 𝑚 = ℓ 𝑘 for 𝑘 ∈ Z ≥0 : indeed, we are applying case 1 (CRT) and, as
in 28.1.1, this asks for 𝑚 ≡ 1 (mod 𝑝 𝑛 𝑝 ) for 𝑝 ∈ T (with 𝑛 𝑝 ∈ Z ≥1 , and ℓ ≠ 𝑝), so
we just need to take 𝑘 to be a common multiple of the orders of ℓ ∈ (Z/𝑝 𝑛 𝑝 Z) × . With
this strengthening, we have 𝑥 = 𝑦 0/𝑚 + 𝑦 ∈ Z𝑞 for all 𝑞 ∉ S t T.
28.1.11. In weak approximation, we can replace the additive group Q with with the
multiplicative group Q× : the image of Q× ↩→ 𝑣 ∈S Q×𝑣 is dense a fortiori.
Î
However, the embedding Q× ↩→ Q×6S is not dense: that is to say, we do not have
strong approximation for Q× . Indeed, taking S = {∞} we have Q× = Q b× ; and since
6S
Z× ∩ Q× = Z× = {±1}, the open set b
b Z× r {±1} is disjoint from Q× . In view of 28.1.10,
the problem is also indicated by the fact that Z× = {±1} does not surject onto (Z/𝑚Z) ×
for 𝑚 ≥ 7.
28.2.1. Recall that 𝐵 𝑣 = M2 (Q 𝑣 ) ' Q4𝑣 has the coordinate topology (see section 13.5);
therefore weak and strong approximation for 𝐵 = M2 (Q) follow from these statements
for Q, and weak approximation for GL2 (Q) follows as the determinant is continuous.
28.2.2. We should not expect the embedding GL2 (Q) ↩→ GL2 ( Q) b to be dense any
more than it was for Q× = GL1 (Q), as in 28.1.11. In fact, we rediscover the same issue
by taking determinants: the map GL2 (Z) → GL2 (Z/𝑚Z) is not surjective, because
det(GL2 (Z)) = ±1 whereas det(GL2 (Z/𝑚Z)) = (Z/𝑚Z) × .
468 CHAPTER 28. STRONG APPROXIMATION
Theorem 28.2.3 is known as strong approximation for the group SL2 (Q). We
give a quick proof of Theorem 28.2.3 in two steps. In preparation, we recall from
27.2.6 that
b × = Q× b
Q Z×
(“denominators can be handled globally”) and prove an analogous decomposition.
GL2 ( Q)
b = GL2 (Q) GL2 (b
Z)
(28.2.5)
SL2 ( Q)
b = SL2 (Q) SL2 (b
Z).
Proof. We begin with the first statement. The inclusion GL2 (Q) GL2 (b Z) ⊆ GL2 ( Q)b
holds; we prove the other containment. Let b 𝛼 ∈ GL2 ( Q). Consider the collection of
b
lattices (𝐿 𝑝 ) 𝑝 with 𝐿 𝑝 = 𝛼 𝑝 Z2𝑝 ⊆ Q2𝑝 . Since 𝛼 𝑝 ∈ GL2 (Z 𝑝 ) for all but finitely many
𝑝, we have 𝐿 𝑝 = Z2𝑝 for all but finitely many 𝑝. By the local-global dictionary for
lattices (Theorem 9.4.9), there exists a unique lattice 𝐿 ⊆ Q2 whose completions are
𝐿 𝑝 . We now rephrase this adelically (and succinctly): letting b 𝐿=b 𝛼bZ2 ⊆ Q b2 , we take
𝐿 = 𝐿 ∩ Q . Choose a basis for 𝐿 and put the columns in a matrix 𝛼, so 𝐿 = 𝛼Z2 .
b 2
𝛼) = det(𝛼) det(b
1 = det(b 𝛾 ) ∈ Q× bZ× = Q b×
× × −1 0
but Q ∩Z = {±1}; multiplying both 𝛼, b
b 𝛾 by on the right and left respectively,
0 1
𝛾 ) = 1, i.e., 𝛼 ∈ SL2 (Q) and b
if necessary, we may take det(𝛼) = det(b 𝛾 ∈ SL2 (b Z).
Proof of Theorem 28.2.6. Let 𝛼 ∈ M2 (Z) be such that 𝛼 maps to the desired matrix
in SL2 (Z/𝑚Z), so in particular det(𝛼) ≡ 1 (mod 𝑚). By the theory of elementary
divisors (Smith normal form), there exist matrices 𝜇, 𝜈 ∈ SL2 (Z) such
that 𝜇𝛼𝜈 is
𝑎 0
diagonal; so without loss of generality, we may suppose that 𝛼 = with 𝑎𝑏 ≡ 1
0 𝑏
(mod 𝑚). Let
0 𝑎 −(1 − 𝑎𝑏)
𝛼 = . (28.2.7)
1 − 𝑎𝑏 𝑏(2 − 𝑎𝑏)
Then 𝛼 0 ≡ 𝛼 (mod 𝑚) and
Remark 28.2.9. The proof of Theorem 28.2.6 extends in two ways. First, we can
replace Z with a PID and the same proof works. Second, arguing by induction, one
can show that the map SL𝑛 (Z) → SL𝑛 (Z/𝑚Z) is surjective for all 𝑛 ≥ 2 and 𝑚 ∈ Z.
We are now ready to prove strong approximation for SL2 (Q).
With the preceding context, we are now ready to state a more general formulation
of strong approximation for indefinite quaternion algebras over Q. The following
theorem is a special case of Main Theorem 28.5.3.
(a) Every locally principal right O-ideal is in fact principal, i.e., # Cls O = 1.
(b) Every order O0 locally isomorphic to O is in fact isomorphic to O, i.e., # Typ O =
1.
(c) For any integer 𝑚, the reduction map O1 → (O/𝑚O) 1 is surjective.
Proof. Specialize Main Theorem 28.5.3, Corollary 28.5.6, and Corollary 28.5.14,
respectively, using 28.5.16.
470 CHAPTER 28. STRONG APPROXIMATION
28.3.2. If 𝐹 is a field, then SL𝑛 (𝐹) is generated by elementary matrices by the theory
of echelon forms (Exercise 28.3).
Lemma 28.3.3. Let 𝑅 be a Euclidean domain. Then SL2 (𝑅) is generated by elemen-
tary matrices.
Remark 28.3.4. Lemma 28.3.3 holds for general 𝑛 ≥ 2, and it follows from the above
by induction: see Exercise 28.4.
This theory of elementary matrices has the following striking consequence.
Proposition 28.3.5. Let 𝑅 be a Dedekind domain. Then for all ideals 𝔪 ⊆ 𝑅, the map
is surjective.
28.4. STRONG APPROXIMATION AND THE IDEAL CLASS SET 471
Proof. We may suppose 𝔪 is nonzero. Then by the CRT, 𝑅/𝔪 is a finite product of local
Artinian principal ideal rings. Therefore 𝑅/𝔪 is Euclidean and by a generalization
of Lemma 28.3.3, the group SL2 (𝑅/𝔪) is generated by elementary matrices: see
Exercise 28.6. Every elementary matrix in SL2 (𝑅/𝔪) lifts to an elementary matrix in
SL2 (𝑅), and the statement follows.
Corollary 28.3.6. Let 𝐹 be a global field, and let 𝑅 ⊆ 𝐹 be a global ring with eligible
set S. Then the image of the map
Ö0
SL2 (𝐹) ↩→ SL2 (𝐹6𝑆 ) = SL2 (𝐹𝑣 )
𝑣∉S
is dense.
Proof. For brevity, we write 𝐹b = 𝐹6𝑆 and 𝑅b = Î 𝑣∉S 𝑅 𝑣 . We first show that
SL2 (𝑅) ↩→ SL2 ( 𝑅)
b is dense. If 𝑈 ⊆ SL2 ( 𝑅)
b is open, then 𝑈 contains a standard
open neighborhood of the form
{ 𝛽b ∈ SL2 ( 𝑅)
b : 𝛽b ≡ 𝛼𝔪 (mod 𝔪)}
for some 𝛼𝔪 ∈ SL2 (𝑅/𝔪) and 𝔪 ⊆ 𝑅. The surjectivity in Proposition 28.3.5 then
implies that 𝑈 ∩ SL2 (𝑅) ≠ ∅.
For the statement itself, we again argue with elementary matrices. Let b 𝛼 = (𝛼 𝑣 ) 𝑣 ∈
SL2 ( 𝐹);
b then 𝛼 𝑣 ∈ SL2 (𝑅 𝑣 ) for all but finitely many 𝑣. For these finitely many 𝑣, we
know SL2 (𝐹𝑣 ) is generated by elementary matrices by Lemma 28.3.2, so by strong
approximation in 𝐹 (Lemma 28.7.2) we can approximate 𝛼 𝑣 by an element of SL2 (𝐹)
that belongs to any open neighborhood of 𝛼 𝑣 ; for the remaining places we apply the
previous paragraph, and we finish using the continuity of multiplication.
nrd : Cls O = 𝐵× \ 𝐵
b×/O
b× → 𝐹 × \𝐹
>Ω 0
b×/nrd( O
b× ) (28.4.2)
We now investigate the injectivity of the reduced norm map (28.4.2). This map is
only a map of (pointed) sets, so first we show that it suffices to look at an appropriate
kernel.
b× , the map b
28.4.3. For all 𝛽b ∈ 𝐵 𝛼O ↦→ b
𝛼O 𝛽b gives a bijection
Cls O = 𝐵× \ 𝐵
b× /O
b× ↔ 𝐵× \ 𝐵
b× /O
b0× = Cls O0
Our investigations will involve the kernels of the reduced norm maps:
𝐵1 := {𝛼 ∈ 𝐵× : nrd(𝛼) = 1} ≤ 𝐵
b1 := {b b× : nrd(b
𝛼∈𝐵 𝛼) = 1} (28.4.4)
and the group 𝐵b1 consists of local solutions at all primes that belong to 𝑅 4𝑣 for almost
all places 𝑣 ∈ Pl 𝐹.
Lemma 28.4.6. Let O ⊆ 𝐵 be an 𝑅-order. Then the reduced norm map (28.4.2) is
injective for all orders O0 ∈ Gen O if and only if 𝐵
b1 ⊆ 𝐵× O
b0× for all O0 ∈ Gen O.
For the converse, since nrd : 𝐵× → 𝐹>×Ω 0 and nrd : O b× → nrd( O b× ) are both
surjective, to show nrd is injective for O we may show that if nrd(b 𝛼) = nrd( 𝛽)b ∈𝐹 b×
then b × × ×
𝛼O = 𝑧 𝛽bO for some 𝑧 ∈ 𝐵 . We consider (b
b b −1 −1
𝛼 𝛽b ) ( 𝛽bO 𝛽b ) = (b
b −1 0
𝛼 𝛽b ) O where
b
as above O0 = 𝐵 ∩ 𝛽bO b 𝛽b−1 ∈ Gen O. Since b 𝛼 𝛽b−1 ∈ 𝐵b1 , by hypothesis b𝛼 𝛽b−1 = 𝑧b𝜇0 =
b𝜇 𝛽b−1 ) where 𝑧 ∈ 𝐵× and 𝜇
𝑧( 𝛽b b∈ O b× , and consequently b 𝛼O b𝜇O
b = 𝑧 𝛽b b = 𝑧 𝛽bO,b and
hence the map is injective.
28.4.7. We have 𝐵× O
b× ∩ 𝐵 b1 if and only if nrd(O× ) = 𝐹 × ∩ nrd( O
b1 = 𝐵1 O
>Ω 0
b× )
(Exercise 28.10).
We should not expect hypothesis of 28.4.8 to hold for all quaternion algebras: see
Exercise 28.7.
28.5. STATEMENT AND FIRST APPLICATIONS 473
28.5.2. If 𝐹 is a number field, then this definition agrees with Definition 17.8.1; and
since a complex place is necessarily split and S contains the archimedean places, if 𝐵
is S-definite over a number field 𝐹 then 𝐹 is a totally real number field.
28.5.4. One can think of strong approximation from the following informal perspective:
if 𝐵1𝑣 is not compact, then there is enough room for 𝐵1 to “spread out” in 𝐵1𝑣 so that
correspondingly 𝐵1 is dense in the S-finite part 𝐵b1 .
1 Î 1
The hypothesis that 𝐵S = 𝑣 ∈S 𝐵 𝑣 is noncompact is certainly necessary for the
conclusion that 𝐵1 is dense in 𝐵 b1 . Indeed, if 𝐵S1 = Î 𝑣 ∈S 𝐵1𝑣 is compact, then since
𝐵1 is discrete in 𝐵1 , the subgroup 𝐵1 𝐵S1 ≤ 𝐵1 is closed in 𝐵1 , and 𝐵1 𝐵S1 ≠ 𝐵1 . On
the other hand, if 𝐵1 is dense in 𝐵b1 , then adding the components for 𝑣 ∈ S we have
𝐵1 𝐵S1 ≤ 𝐵1 dense. This is a contradiction.
We give two proofs of strong approximation over the next two sections. For the
moment, we consider some applications.
Our main motivation for strong approximation is the following proposition. We
recall the class group 27.7.4 associated to O.
nrd : Cls O = 𝐵× \ 𝐵
b× /O
b× → Cl𝐺 (O) 𝑅 = 𝐹 × \𝐹
>Ω 0
b× /nrd(O× )
Corollary 28.5.6. If 𝐵 is S-indefinite and Cl𝐺 (O) 𝑅 is trivial, then Typ O is trivial,
i.e., every order O0 locally isomorphic to O is in fact isomorphic to O.
Proof. The class set Cls O maps surjectively onto Typ O by Lemma 17.4.13, and the
latter is trivial by Theorem 28.5.5.
28.5.7. More generally, we can grapple with the type set of O, measured by a different
(generalized) class group. Recall (27.6.24) that
Typ O ↔ 𝐵× \ 𝐵
b× /𝑁 b× ( O).
𝐵
b
474 CHAPTER 28. STRONG APPROXIMATION
Let
𝐺 𝑁 (O) := 𝐹>×Ω 0 nrd(𝑁 𝐵b× ( O))
b ≤𝐹b× . (28.5.8)
b× ≤ 𝑁 b× ( O),
Since O b we have 𝐺 𝑁 (O) ≥ 𝐺 (O). Define accordingly the class group
𝐵
b× /𝐺 𝑁 (O).
Cl𝐺 𝑁 (O) 𝑅 = 𝐹 (28.5.9)
Corollary 28.5.10. If 𝐵 is S-indefinite, then the reduced norm map induces a bijection
∼ Cl
nrd : Typ O −→ 𝐺 𝑁 (O) 𝑅.
Proof. We take the further quotient by the normalizer in the bijection in Theorem
28.5.5.
b1 = 𝐵1 O
𝐵 b1 and 𝐵 1 = 𝐵 1 O1 . (28.5.13)
Proof. The inclusion 𝐵1 Ob1 ⊆ 𝐵b1 holds, and the converse holds when 𝐵1 is dense in
b1 by 28.4.8. For the second statement, we have 𝐵 = 𝐵
𝐵 b × 𝐵S and O = O
b × 𝐵S , so we
1
take norm 1 units and multiply both sides of (28.5.13) by 𝐵S = 𝑣 ∈S 𝐵1𝑣 .
Î
We now give a name to a large classes of orders where the group 𝐺 (O) governing
principality is explicitly given.
28.5. STATEMENT AND FIRST APPLICATIONS 475
Certain special cases of Theorem 28.5.5 are important in applications. Recall that
Ω ⊆ Ram 𝐵 is the set of real ramified places, and ClΩ 𝑅 as defined in 17.8.2 is class
group associated to Ω, a quotient of the narrow class group.
Corollary 28.5.17. Suppose 𝐹 is a number field and let S be the set of archimedean
places of 𝐹. Suppose 𝐵 is S-indefinite and O ⊆ 𝐵 is locally norm-maximal 𝑅-order.
Then nrd : Cls O → ClΩ 𝑅 is a bijection.
Proof. This is just a restatement of Theorem 28.5.5 once we note that Cl𝐺 (O) 𝑅 =
ClΩ 𝑅 by Example 27.7.7.
Proposition 28.5.18. Let T ⊇ S be a set of primes of 𝑅 that generate Cl𝐺 (O) 𝑅 and
suppose 𝐵 is T-indefinite. Then every class in Cls O contains an integral (invertible
right) O-ideal whose reduced norm is supported in the set T.
Proof. Let 𝑅 (T) denote the (further) localization of 𝑅 at the primes in T. We apply
Theorem 28.5.5 to the order O (T) := O ⊗𝑅 𝑅 (T) : we conclude that there is a bijection
Cls O (T) −→∼ Cl
𝐺 (O (T ) ) 𝑅 (T) . But Cl𝐺 (O (T ) ) 𝑅 (T) is the quotient of Cl𝐺 (O) 𝑅 by the
primes in T, and so by hypothesis is trivial. Therefore if 𝐼 is a right O-ideal, then
𝐼 (T) := 𝐼 ⊗𝑅 𝑅 (T) has 𝐼 (T) = 𝛼O (T) for some 𝛼 ∈ 𝐵× . Let 𝐽 = 𝛼−1 𝐼. Then [𝐽] R = [𝐼] R
and 𝐽𝔭 = O𝔭 for all primes 𝔭 ∉ T and so 𝐽 has reduced norm supported in 𝑇. Replacing
𝐽 by 𝑎𝐽 with 𝑎 ∈ 𝑅 nonzero and supported in T, we may suppose further that 𝐽 ⊆ O
is integral, and the result follows.
Example 28.5.19. Let 𝐵 be a definite quaternion algebra over a totally real (number)
field 𝐹 and let S be the set of archimedean places, so 𝑅 = Z𝐹 . Let O be a locally
norm-maximal 𝑅-order in 𝐵. Suppose that ClΩ 𝑅 = {1} and let 𝔭 ⊆ 𝑅 be a prime of
𝑅 unramified in 𝐵. Then by Proposition 28.5.18, every ideal class in Cls O contains
an integral O-ideal whose reduced norm is a power of 𝔭.
As a special case, we may take 𝐹 = Q: then ClΩ Z = Cl+ Z = {1}. Therefore,
if 𝐵 is a definite quaternion algebra of discriminant 𝐷 over Q, and O ⊆ 𝐵 a locally
norm-maximal order (e.g., an Eichler order), then for a prime 𝑝 - 𝐷, every invertible
right O-ideal class is represented by an integral ideal whose reduced norm is a power
of 𝑝.
476 CHAPTER 28. STRONG APPROXIMATION
Proof. By Main Theorem 14.7.4, there exists 𝛼 ∈ 𝐵× such that nrd(𝛼) = 𝑛. For each
prime 𝔭, the set
𝑈𝔭 = {𝛽𝔭 ∈ 𝐵𝔭1 : trd(𝛽𝔭 𝛼) ∈ 𝑅𝔭 } (28.6.2)
is (closed and) open since trd is continuous, and further 𝑈𝔭 is nonempty: if 𝔭 ∈ Ram 𝐵
then already 𝛼𝔭 is integral
over 𝑅𝔭 and 1 ∈ 𝑈𝔭 , and otherwise 𝐵𝔭 ' M2 (𝐹𝔭 ) and we
0 −𝑛
may suppose 𝛼𝔭 = is in rational canonical form whereby
1 𝑡
0 1 0 −𝑛 1 𝑡
trd = trd = 𝑛 + 1 ∈ 𝑅𝔭 (28.6.3)
−1 0 1 𝑡 0 𝑛
0 1
shows 𝛽𝔭 = ∈ 𝑈𝔭 .
−1 0
b1
Î
b :=
Let 𝑈 𝔭 𝑈𝔭 ∩ 𝐵 ; then 𝑈 is open and nonempty. By strong approximation,
b
there exists 𝛽 ∈ 𝑈 1
b ∩ 𝐵 . Thus trd(𝛽𝛼) ∈ Ñ𝔭 𝑅𝔭 = 𝑅 and nrd(𝛽𝛼) = nrd(𝛼) = 𝑛.
Therefore 𝛽𝛼 is as desired.
Let 𝑅>× Ω 0 := 𝑅 × ∩ 𝐹>×Ω 0 be the subgroup of units that are positive at the places
𝑣 ∈ Ω (the set of real, ramified places in 𝐵).
Proof. Let 𝑢 ∈ 𝑅>× Ω 0 . We repeat the argument of Theorem 28.6.1, with a slight
modification. Let 𝔐 be the level of O.
Let 𝔭 | 𝔐 be a prime that divides the level 𝔐. We choose an isomorphism
𝐵𝔭 ' M2 (𝐹𝔭 ) such that 𝛼𝔭 is in rational canonical form, and let O𝔭0 be the standard
Eichler order in M2 (𝐹𝔭 ) of the same level as O. Define
This is again an open condition because multiplication is continuous, and the calcula-
tion −1
𝑡𝑢 1 0 −𝑢 1 0
=
−1 0 1 𝑡 0 𝑢
shows also that 𝑈𝔭 ≠ ∅.
28.6. FURTHER APPLICATIONS 477
choosing a local generator. The projection map Cls+R O → ClsR O has finite fibers as
𝐵×>0 ≤ 𝐵× has finite index, so since ClsR O is a finite set, so too is Cls+R O.
Corollary 28.6.8. Let 𝐹 be a number field and suppose 𝐵 is S-definite. Then the map
Proof. Let 𝑎 ∈ 𝐹>×Ω 0 be such that 𝑣(𝑎) = 𝜖 𝑣 for all 𝑣 real not in Ω and the class of
𝑎𝑅 is trivial in Cl𝐺 (O) 𝑅: these constraints together impose congruence conditions on
elements in a real cone. By the Hasse–Schilling norm theorem, there exists 𝛼 ∈ 𝐵×
such that nrd(𝛼) = 𝑎. Thus the class of nrd(𝛼O) in Cl+𝐺 (O) 𝑅 = Cl𝐺 (O) 𝑅 is trivial.
But then by (28.6.9) (a consequence of strong approximation), there exists 𝛽 ∈ 𝐵×>0
such that 𝛽O = 𝛼O, and therefore 𝛽 = 𝛼𝛾 with 𝛾 ∈ O× . Since 𝛽 is totally positive, for
all real places 𝑣 ∉ Ω we have sgn(𝑣(nrd(𝛾))) = sgn(𝑣(nrd(𝛼))) = 𝜖 𝑣 , completing the
proof.
For the second statement, we only need to note that Cl+ Z = Cl Z = {1} and recall
that nrd(O× ) ≤ {±1}.
Remark 28.6.11. More generally, let 𝐵 be a central simple algebra over the global field
𝐹. We say 𝐵 satisfies the S-Eichler condition if there exists a place 𝑣 ∈ S such that
𝐵 𝑣 is not a division algebra. (In this text, for quaternion algebras we prefer to use the
term S-indefinite because it readily conveys the notion, but both terms are common.)
If 𝐹 is a number field and S is the set of archimedean places of 𝐹, then 𝐵 satisfies
the S-Eichler condition if and only if 𝐵 is not a totally definite quaternion algebra
(Exercise 28.5). So the condition is a mild condition, and the quaternion algebra case
requires special effort.
When 𝐵 satisfies the S-Eichler condition, then 𝐵1 ↩→ 𝐵 b1 is dense, and for O ⊆ 𝐵
a maximal 𝑅-order, a locally principal right fractional O-ideal 𝐼 ⊆ 𝐵 is principal if
and only if its reduced norm nrd(𝐼) ⊆ 𝑅 is trivial in ClΩ 𝑅, where Ω is the set of real
places 𝑣 ∈ Pl 𝐹 such that 𝐵 𝑣 ' M𝑛 (H), generalizing the quaternion case.
Eichler proved Theorem 28.5.5 and the more general statement of the previous
paragraph [Eic37, Satz 2], also providing several reformulations and applications
[Eic38a, Eic38c]. For a full exposition, see Reiner [Rei2003, §34].
are dense.
Proof. See e.g. Neukirch [Neu99, Theorem II.3.4] or O’Meara [O’Me73, §11E].
Lemma 28.7.2 (Strong approximation forÎ𝐹). Let S ⊆ Pl 𝐹 be a finite nonempty set
of places. Then the image of 𝐹 ↩→ 𝐹6S := 𝑣∉S 𝐹𝑣 is dense.
28.7. FIRST PROOF 479
Proof. See e.g. Neukirch [Neu99, Exercise III.1.1] or O’Meara [O’Me73, §33G].
Î
We recall that 𝐵 is a quaternion algebra over 𝐹 and we write 𝐵S := 𝑣 ∈S 𝐵𝑣 .
are dense.
Î
Proof. By weak approximation for 𝐹 (Lemma 28.7.1), we have 𝐹 dense in 𝑣 ∈S 𝐹𝑣 .
Choosing an 𝐹-basis for 𝐵, we have 𝐵 ' 𝐹 4 as topological 𝐹-vector Î spaces, and so
by approximating in each coordinate, we conclude that 𝐵 is dense in 𝑣 ∈S 𝐵 𝑣 . The
multiplicative case follows a fortiori, restricting open neighborhoods.
Finally we treat 𝐵1 . By Exercise 7.30, we know that 𝐵1 = [𝐵× , 𝐵× ] is the
commutator. Let (𝛾 𝑣 ) 𝑣 ∈
Î 1
𝑣 ∈S 𝐵 𝑣 . Then for each 𝑣 ∈ S, we can write 𝛾 𝑣 =
𝛼 𝑣 𝛽 𝑣 𝛼 𝑣 𝛽 𝑣 with 𝛼 𝑣 , 𝛽 𝑣 ∈ 𝐵 . Then by weak approximation for 𝐵× , we can find a
−1 −1 ×
Next, we need to approximate polynomials: this kind of lemma was first performed
in section 14.7 to prove the Hasse–Schilling theorem of norms, and here we need
another variant.
Proof. We argue as in Lemma 14.7.6 (and Corollary 14.7.8), but instead of weak
approximation we now use strong approximation (Lemma 28.7.2). Our job is a bit
easier because we are only asking for irreducibility, not separability.
Since S is nonempty, by strong approximation for 𝐹 we can find 𝑡 arbitrarily close
𝑡, thus ensuring that the desired inequalities hold and that 𝑓 (𝑥) is irreducible over
to b
𝐹𝑣 for 𝑣 ∈ Σ; to ensure that 𝑓 (𝑥) is separable, we need only avoid the locus 𝑡 2 = 4,
and similarly we may ensure 𝑓 (𝑥) is irreducible.
Proof of Main Theorem 28.5.3. We follow Vignéras [Vig80a, Théorème III.4.3]; see
also Miyake [Miy2006, Theorems 5.2.9–5.2.10] for the case 𝐹 = Q. We show that the
closure of 𝐵1 is equal to 𝐵
b1 . Let b b1 ; we will find a sequence of elements
𝛾 = (𝛾 𝑣 ) 𝑣 ∈ 𝐵
1
of 𝐵 converging to b 𝛾.
Step 1: Setup. We claim it is enough to consider the case where 𝛾 𝑣 = 1 for all but
finitely many 𝑣, by a Cantor-style diagonalization argument. Indeed, for a finite set
T ⊆ Pl 𝐹 disjoint from S, we let b 𝛾 [T ] be the idele obtained from b 𝛾 replacing 𝛾 𝑣 = 1
for 𝑣 ∉ T. Then for a sequence of subsets T eventually containing each place 𝑣, we
have b 𝛾 . Thus, if we can find sequences in 𝐵1 converging to each b
𝛾 [T ] → b 𝛾 [T ] we can
diagonalize to find a sequence converging to b 𝛾 , since Pl 𝐹 is countable.
So we may suppose without loss of generality that 𝛾 𝑣 = 1 for all but finitely many
𝑣. To find a sequence in 𝐵1 converging to b 𝛾 , our strategy in the proof is as follows:
in shrinking open neighborhoods of b 𝛾 we first find an element in 𝐵1 whose reduced
characteristic polynomial is close to an element in the open neighborhood, and then
we conjugate by 𝐵× to get the limits themselves to match.
To this end, let O ⊂ 𝐵 be a reference 𝑅-order, let T ⊆ Pl 𝐹 be a finite set of places
disjoint from S containing the primes 𝑣 where 𝛾 𝑣 ≠ 1 and the ramified places of 𝐵 not
in S. We consider open neighborhoods of the form
Ö Ö
𝑈= 𝛾𝑣 𝑈𝑣 × O1𝑣
𝑣 ∈T 𝑣∉S∪T
and the reduced trace is an open (linear) map, with a closer approximation we may
𝛾 0 ∈ 𝑈 such that trd(𝛽) = trd(b
suppose trd(𝛽) ∈ trd(𝑈), and therefore there exists b 𝛾 0)
so that 𝛽, b0
𝛾 have the same irreducible minimal polynomial. By the Skolem–Noether
𝛼∈𝐵
theorem (Corollary 7.7.3), there exists b b× such that
𝛼−1 b
𝛽=b 𝛾 0b
𝛼. (28.7.5)
𝛼𝑛−1 b
𝐵1 3 𝛽𝑛 = b 𝛾𝑛0 b 𝛼𝑛−1𝑈𝑛 b
𝛼𝑛 ∈ b 𝛼𝑛 . (28.7.6)
𝛾𝑛0 ∈ 𝑈𝑛 , we have b
Since b 𝛾𝑛0 → b
𝛾 , and in particular for 𝑣 ∈ Pl 𝐹 r S ∪ T, we have
0
𝛾𝑛,𝑣 → 𝛾 𝑣 = 1.
Step 4: Harmonizing the sequence. By ‘harmonizing’ the conjugating elements b 𝛼𝑛 ,
we will realize a sequence in 𝐵1 tending to b 𝛾 as desired, in two (subset)steps. First, by
Main Theorem 27.6.14, 𝐵 b×/𝐵× is compact. So restricting to a subsequence, we have
b b ×
𝛼𝑛 = 𝛿 𝑛 𝜇 𝑛 with 𝜇 𝑛 ∈ 𝐵 and b 𝛿𝑛 → b b× . Second, by weak approximation
𝛿 = (𝛿 𝑣 ) 𝑣 ∈ 𝐵
for 𝐵 (Proposition 28.7.3), 𝐵 is dense in 𝑣 ∈T 𝐵×𝑣 , so there is a sequence 𝜈𝑛 from 𝐵×
× Î
such that 𝜈𝑛 → (𝛿 𝑣 ) 𝑣 ∈T .
Step 5: Conclusion. To conclude, we consider the sequence
𝛿−1
𝛾 . For 𝑣 ∈ T, we have 𝜈𝑛,𝑣 b
We claim that this sequence tends to b −1
𝑛,𝑣 → 𝛿 𝑣 𝛿 𝑣 = 1 so
(𝜈𝑛,𝑣 𝛿−1 0 −1
𝑛,𝑣 )𝛾 𝑛,𝑣 (𝛿 𝑛,𝑣 𝜈 𝑛,𝑣 ) → 𝛾 𝑣 . (28.7.8)
0
On the other hand, for 𝑣 ∈ Pl 𝐹 r (S ∪ T), we have 𝛾𝑛,𝑣 → 1, so
(𝜈𝑛,𝑣 𝛿−1 0 −1
𝑛,𝑣 )𝛾 𝑛,𝑣 (𝛿 𝑛,𝑣 𝜈 𝑛,𝑣 ) → 1 = 𝛾 𝑣 . (28.7.9)
Proof. Suppose that 𝐵1𝑣 ⊆ 𝑍 for all 𝑣 ∉ T where T is a finite set. Let b 𝛾 ∈ 𝐵 b1 . If
𝛾 𝑣 = 1 for all 𝑣 ∈ T, then b 𝛾 is a limit of elements in 𝑍 (approximating at a finite
𝛾 ∈ 𝑍. Otherwise, by weak approximation for 𝐵 (Proposition 28.7.3), there
level), so b
exists 𝛾 ∈ 𝐵1 such that 𝛾 is near 𝛾 𝑣 for all 𝑣 ∈ T. Let 𝛽b have 𝛽 𝑣 = 1 for 𝑣 ∈ T and
𝛽 𝑣 = 𝛾 −1 𝛾 𝑣 for 𝑣 ∉ T; then 𝛽b ∈ 𝑍, and b
𝛾 is the limit of the 𝛾 𝛽.
b
Now we consider
𝑍1 := {b
𝛾 ∈ 𝑍 : 𝛾 𝑣 = 1 for all but finitely many 𝑣}. (28.8.2)
b1 is a normal subgroup.
Lemma 28.8.3. 𝑍1 E 𝐵
Proof. Let b𝛾∈𝐵 b1 and let b𝛼 ∈ 𝑍1 with 𝛼 𝑣 = 1 for 𝑣 ∉ T with T a finite set. By weak
approximation for 𝐵 (Proposition 28.7.3), there exists 𝛾 ∈ 𝐵1 with 𝛾 close to 𝛾 𝑣 for
all 𝑣 ∈ T. Therefore 𝛾 −1 b 𝛾 −1 b
𝛼 𝛾 is near b 𝛾 for 𝑣 ∈ T and 𝛾 −1 𝛼 𝑣 𝛾 = 𝛾 −1
𝛼b 𝑣 𝛼 𝑣 𝛾 𝑣 = 1 for
𝑣 ∉ T, so is the limit of such in 𝑍. Therefore b 𝛾 −1 b𝛾 ∈ 𝑍1 , thus b
𝛼b 𝛾 −1 𝑍1 b
𝛾 ⊆ 𝑍1 and
𝑍1 E 𝐵 b1 .
Lemma 28.8.4. Let 𝐹 be an infinite field. Then PSL2 (𝐹) is a simple group.
Proof. See e.g. Grove [Grov2002, Theorem 1.13]. Briefly, the result can be proven
using Iwasawa’s criterion, since SL2 (𝐹) acts doubly transitively on the linear subspaces
of 𝐹 2 : the kernel of the action is the center {±1}, and the stabilizer subgroup of a
standard basis element is the subgroup of upper triangular matrices, whose conjugates
generate SL2 (𝐹).
Idl(O) → Cls O
𝐼 ↦→ [𝐼]
and PIdl(O) is the kernel of this map, the preimage of the trivial class in Cls O. The
composition of this map with the reduced norm gives a group homomorphism:
Proof. By strong approximation (Theorem 28.5.5), the reduced norm gives a bijection
nrd : Cls O → Cl𝐺 (O) (𝑅); thus 𝐼 ∈ Idl(O) is principal, belonging to PIdl(O), if and
only if [nrd(𝐼)] is trivial in Cl𝐺 (O) (𝑅).
Now let O, O0 be locally isomorphic orders (in the same genus) with connecting
O, O0-ideal 𝐽.
Proposition 28.9.7. Suppose that 𝐵 is S-indefinite. Then the map (28.9.6) induces a
commutative diagram
o o o
1 / 𝑁 𝐵× (O0)/(𝐹 × O0× ) / Pic𝑅 (O0) / Idl(O0)/PIdl(O0) /1
Proof. We verify that 𝐽 −1 PIdl(O)𝐽 = PIdl(O0), from which both statements follow;
and this verification comes from Lemma 28.9.4, as
Proposition 28.9.7 says that when 𝐵 is S-indefinite, then the structure of the
normalizer group, the Picard group, and group of ideals modulo principal ideals are
all isomorphic for all orders in a genus. The same is not true when 𝐵 is S-definite;
we always have an isomorphism in the middle, but for locally isomorphic orders,
the Picard group may be distributed differently between the normalizer and the ideal
group.
By chasing a few diagrams, we can be more specific about the structure of Idl(O)
by seeking out primitive ideals.
The definition (28.9.9) makes sense because Idl(𝑅) is indeed a subgroup: since if
𝔞 ∈ Idl(𝑅) then [nrd(𝔞O)] = [𝔞] 2 ∈ (Cl𝐺 (O) 𝑅) 2 .
Idl(𝑅) → Idl(O)
(28.9.12)
𝔞 ↦→ 𝔞O;
(compare to (18.5.5)). From Lemma 28.9.4 and the fact that PIdl(𝑅) ⊆ ker(𝑐), we
obtain an exact sequence
𝑐
1 → 𝑁 𝐵× (O)/(𝐹 × O× ) → Pic𝑅 (O) →
− Cl𝐺 (O) (𝑅). (28.9.14)
We write (Cl 𝑅) [2] ↑O because this is the subgroup of ideal classes that lift to the group
Cl𝐺 (O) (𝑅) [2] (having order dividing 2). Therefore the following diagram commutes,
486 CHAPTER 28. STRONG APPROXIMATION
1 / (Cl 𝑅) [2] ↑O / 𝑁 𝐵× (O)/(𝐹 × O× ) / AL(O)
1 / Cl 𝑅 / Pic𝑅 (O) / Idl(O)/Idl(𝑅) /1
𝑐
1 / (Cl𝐺 (O) 𝑅) 2 / Cl𝐺 (O) (𝑅) / Cl𝐺 (O) (𝑅)/(Cl𝐺 (O) 𝑅) 2 /1
1
(28.9.16)
The second main result of this section is the following.
Proposition 28.9.17. Suppose 𝐵 is S-indefinite. Then there is a (non-canonically)
split exact sequence
1 → (Cl 𝑅) [2] ↑O → 𝑁 𝐵× (O)/(𝐹 × O× ) → AL(O) → 1. (28.9.18)
Proof. The snake lemma implies that the top row of (28.9.16) is exact; the sequence
is split by choosing for each class in AL(O) a generator of a representative ideal.
Here is second, self-contained proof which captures the above discussion. Say that
a two-sided (integral) O-ideal 𝐼 is 𝑅-primitive if 𝐼 is not divisible by any integral ideal
of the form 𝔞O with 𝔞 ( 𝑅. (If 𝐼 is integral but not 𝑅-primitive, with 𝐼 ⊆ 𝔞O and 𝔞 as
small as possible, then 𝔞−1 𝐼 ⊆ O is integral and now 𝑅-primitive.) Let 𝛼 ∈ 𝑁 𝐵× (O).
Then O𝛼O = 𝐼 ∈ Idl(O), so we can factor 𝐼 = 𝔠𝐽 uniquely with 𝔠 a fractional ideal of
𝑅 and 𝐽 an 𝑅-primitive ideal. We have
𝑎𝑅 = nrd(𝛼)𝑅 = nrd(𝐼) = 𝔠2 nrd(𝐽) = 𝔠2 𝔞 (28.9.19)
and so
1 = [(𝑎)] = [𝔠] 2 [𝔞] ∈ Cl𝐺 (O) (𝑅)
and in particular [nrd(𝐽)] ∈ (Cl𝐺 (O) 𝑅) 2 . Therefore there is a map
𝑁 𝐵× (O) → AL(O). (28.9.20)
This map is surjective by strong approximation (see Lemma 28.9.4): if 𝐽 ∈ Idl(O)
has nrd(𝐽) = 𝔞 and [𝔞] = [𝔠−1 ] 2 ∈ (Cl𝐺 (O) 𝑅) 2 , then [nrd(𝔠𝐽)] = 1 ∈ Cl𝐺 (O) (𝑅) so
by Theorem 28.5.5, there exists 𝛼 ∈ 𝐵× such that O𝛼O = 𝔠𝐽 and since 𝔠𝐽 ∈ Idl(O)
we have 𝛼 ∈ 𝑁 𝐵× (O). The map is split by this construction, with a choice of 𝐽
up to Idl 𝑅. The kernel of the map in (28.9.20) consists of 𝛼 ∈ 𝑁 𝐵× (O) such that
O𝛼O = 𝔠O with 𝔠 ∈ Idl(𝑅), and from the preceding paragraph [𝔠2 ] = 1 ∈ Cl𝐺 (O) (𝑅)
so [𝔠] ∈ Cl𝐺 (O) (𝑅) [2]; however, the kernel also contains 𝐹 × O× , whence the class of
𝔠 is well-defined only in Cl(𝑅). Therefore the kernel is canonically identified with
(Cl 𝑅) [2] ↑O .
28.10. ∗ STABLE CLASS GROUP 487
Corollary 28.9.21. Suppose that O is an Eichler order with discrd O = 𝔑 and that 𝐵
is S-indefinite. Then
is an abelian 2-group with rank at most 𝜔(𝔑) + ℎ2 (𝑅), where 𝜔(𝔑) is the number of
prime divisors of 𝔑 and ℎ2 (𝑅) = dimF2 (Cl 𝑅) [2].
Proof. Apply Corollary 28.9.21 with Cl𝐺 (O) (𝑅) = Cl(𝑅) so (Cl 𝑅) [2] ↑O = (Cl 𝑅) [2]
and AL(O) the trivial group by Example 28.9.10.
Proof. We give only a sketch of the proof. For further details, see the references given
for the proof of Theorem 20.7.16.
We first show that the map (28.10.2) is well-defined. Suppose that 𝐼 ⊕ O𝑟 ' 𝐼 0 ⊕ O𝑟
with 𝑟 ≥ 0. If 𝑟 = 0, we are done; so suppose 𝑟 ≥ 1. Extending scalars, we
find an isomorphism 𝜙 : 𝐵𝑟 +1 → 𝐵𝑟 +1 of left 𝐵-modules, represented by an element
𝛾 ∈ GL𝑟 +1 (𝐵) acting on the left. In a similar way, associated to 𝐼 ⊕ O𝑟 we obtain a
class
GL𝑟 +1 (𝐵)b𝛼 GL𝑟 +1 ( O)
b ∈ GL𝑟 +1 (𝐵)\ GL𝑟 +1 ( 𝐵)/GL
b 𝑟 +1 ( O)
b
488 CHAPTER 28. STRONG APPROXIMATION
after we check that the codomain is indeed the image of the reduced norm. Then
𝛼 0) = nrd(𝛾b
nrd(b 𝛼) = nrd(𝛾) nrd(b 𝛼) ∈ 𝐹>×Ω 0 \𝐹
𝛼) = nrd(b b×/nrd( O
b× )
Exercises
Unless otherwise specified, let 𝑅 be a global ring with eligible set S and 𝐹 = Frac 𝑅,
and let 𝐵 be a quaternion algebra over 𝐹, and let O ⊂ 𝐵 be an 𝑅-order.
(b) Using (a), show that Proposition 28.3.5 and Corollary 28.3.6 hold for SL𝑛 .
5. Let 𝐵 be a central simple algebra over the global field 𝐹. We say 𝐵 satisfies the
S-Eichler condition if there exists a place 𝑣 ∈ S such that 𝐵 𝑣 is not a division
algebra. Show that if 𝐹 is a number field and S is the set of archimedean places
of 𝐹, then 𝐵 satisfies the S-Eichler condition if and only if 𝐵 is not a totally
definite quaternion algebra.
6. In this exercise, we provide details in the proof of Proposition 28.3.5, considering
elementary matrices of rings that are not necessarily domains.
Let 𝐴 be a commutative ring (with 1), not necessarily a domain. We say 𝐴 is
Euclidean if there exists a function 𝑁 : 𝐴 → Z ≥0 such that for all 𝑎, 𝑏 ∈ 𝐴 with
𝑏 ≠ 0, there exists 𝑞, 𝑟 ∈ 𝐴 such that 𝑎 = 𝑞𝑏 + 𝑟 and 𝑁 (𝑟) < 𝑁 (𝑏).
In the first parts, we suppose 𝐴 is Euclidean with respect to 𝑁 and show that the
situation is quite analogous to the case when 𝐴 is a domain.
(a) Show that for all 𝑏 ∈ 𝐴 we have 𝑁 (𝑏) ≥ 𝑁 (0) with equality if and only if
𝑏 = 0.
(b) Let 𝑚 = min{𝑁 (𝑏) : 𝑏 ≠ 0}. Show that if 𝑁 (𝑏) = 𝑚 then 𝑏 ∈ 𝐴× .
(c) Show that every ideal of 𝐴 is principal. [We call 𝐴 a principal ideal ring.]
We now consider examples.
(d) Let 𝐴 be an Artinian local principal ideal ring with a unique maximal ideal
𝔪 = 𝜋 𝐴. (For example, we may take 𝐴 = 𝑅/𝔭𝑒 where 𝑅 is a Dedekind
domain, 𝔭 is a nonzero prime ideal, and 𝑒 ∈ Z ≥0 .) Show that for all 𝑥 ∈ 𝐴
with 𝑥 ≠ 0, there exists 𝑢 ∈ 𝐴× and a unique 𝑛 ∈ Z ≥0 such that 𝑥 = 𝑢𝜋 𝑛 .
ConcludeÎ that 𝐴 is Euclidean with 𝑁 (𝑥) = 𝑛.
(e) If 𝐴 ' 𝑟𝑖=1 𝐴𝑖 with each 𝐴𝑖 an Artinian
Í local principal ideal ring, show
that 𝐴 is Euclidean under 𝑁 (𝑥) = 𝑟𝑖=1 𝑁𝑖 (𝜋𝑖 (𝑥)) where 𝑁𝑖 is as given in
(a) for 𝐴𝑖 and 𝜋𝑖 : 𝐴 → 𝐴𝑖 is the projection.
We conclude with the application.
(f) Let 𝐴 be a Euclidean ring. Show that SL2 ( 𝐴) is generated by elementary
matrices. [Hint: Show that the proof in Lemma 28.3.3 carries over.]
7. Consider the quaternion algebra 𝐵 := (−11, −17 | Q) and let O = Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕
Z𝑖 𝑗. Then 𝐵 is definite, so strong approximation does not apply. Indeed, show
that 𝐵b1 * 𝐵× O
b× as follows.
𝜉 (1 − 𝑠) = 𝜉 (𝑠). (29.1.2)
(More generally, we saw in 26.8.2 that the Dedekind zeta function of a number field
can be completed in an analogous manner, again with functional equation.)
In this introductory section, we sketch a proof of the functional equation and see
it as a consequence of Poisson summation (arising naturally in Fourier analysis, still
following Riemann). Then, following Tate we reinterpret this extra factor in a manner
that realizes the zeta function as a zeta integral on an adelic space, giving a uniform
description and making the whole setup more suitable for analysis.
To begin, we the function 𝜉 (𝑠) itself as an integral. Looking at one term in 𝜉 (𝑠),
we have
∫ ∞ ∫ ∞
d𝑥 2 d𝑢
𝜋 −𝑠/2 Γ(𝑠/2)𝑛−𝑠 = 𝜋 −𝑠/2 𝑛−𝑠 𝑒 −𝑥 𝑥 𝑠/2 = 𝑒 − 𝜋𝑛 𝑢 𝑢 𝑠/2 . (29.1.3)
0 𝑥 0 𝑢
491
492 CHAPTER 29. IDELIC ZETA FUNCTIONS
1 ∞
∫
d𝑢
𝜉 (𝑠) = (Θ(𝑢) − 1)𝑢 𝑠/2 (29.1.5)
2 0 𝑢
valid for Re 𝑠 > 1, where the integral converges (so we may justify interchanging
summation and integration). We will soon rewrite this integral so as to extend its
definition to all 𝑠 ∈ C apart from 𝑠 = 0, 1.
Let 𝑓 : R → C be a Schwartz function, so 𝑓 is infinitely differentiable and every
derivative decays rapidly (for all 𝑚, 𝑛 ≥ 0 we have | 𝑓 (𝑚) (𝑥)| = 𝑂 (𝑥 −𝑛 ) as |𝑥| → ∞).
We define the Fourier transform 𝑓 ∨ : R → C of 𝑓 by
∫ ∞
𝑓 ∨ (𝑦) = 𝑒 −2 𝜋𝑖𝑥 𝑦 𝑓 (𝑥) d𝑥. (29.1.6)
−∞
Proof. The condition that 𝑓 is Schwartz ensures good analytic behavior allowing
the interchange of sum and integral, the details of which we elide. Let 𝑔(𝑥) :=
Í∞
𝑚=−∞ Í 𝑓 (𝑥 + 𝑚). Then 𝑔 is periodic with period 1, so by Fourier expansion we have
𝑔(𝑥) = ∞ 𝑛=−∞ 𝑎 𝑛 𝑒
2 𝜋𝑖𝑛𝑥 where
∫ 1 ∫ 1 ∞
∑︁
𝑎 𝑛 := 𝑔(𝑥)𝑒 −2 𝜋𝑖𝑛𝑥 d𝑥 = 𝑓 (𝑥 + 𝑚)𝑒 −2 𝜋𝑖𝑛𝑥 d𝑥
0 0 𝑚=−∞
∞ ∫
∑︁ 1 ∫ ∞
−2 𝜋𝑖𝑛( 𝑥+𝑚)
= 𝑓 (𝑥 + 𝑚)𝑒 d𝑥 = 𝑓 (𝑥)𝑒 −2 𝜋𝑖𝑛𝑥 d𝑥 = 𝑓 ∨ (𝑛).
𝑚=−∞ 0 −∞
Thus
∞
∑︁ ∞
∑︁
𝑓 (𝑚) = 𝑔(0) = 𝑓 ∨ (𝑛).
𝑚=−∞ 𝑛=−∞
2
− 𝜋 𝑥 . Then 𝑓 is Schwartz. By contour integration and us-
∫ ∞ take2 𝑓 (𝑥) = 𝑒
Now
ing −∞ 𝑒 − 𝜋 𝑥 d𝑥 = 1, we conclude that 𝑓 ∨ (𝑦) = 𝑓 (𝑦) for all 𝑦. For 𝑢 > 0, let
29.2. ⊲ IDELIC ZETA FUNCTIONS, AFTER TATE 493
𝑓𝑢 (𝑥) := 𝑓 (𝑢𝑥); then 𝑓𝑢∨ (𝑦) = 𝑢 −1 𝑓1/𝑢 (𝑦) by change of variable. Applying Poisson
summation to 𝑓√𝑢 (𝑥) then gives
1
Θ(𝑢) = √ Θ(1/𝑢). (29.1.9)
𝑢
The equation (29.1.9) implies the functional equation (29.1.2) for 𝜉 (𝑠) as follows:
we split up the integral (29.1.5) as
∫ 1 ∫ ∞
1 1 d𝑢 1 d𝑢
𝜉 (𝑠) = − + Θ(𝑢)𝑢 𝑠/2 + (Θ(𝑢) − 1)𝑢 𝑠/2
𝑠 2 0 𝑢 2 1 𝑢
1 ∞
∫
d𝑢 1 1
𝜉 (𝑠) = (Θ(𝑢) − 1) (𝑢 𝑠/2 + 𝑢 (1−𝑠)/2 ) − − (29.1.10)
2 1 𝑢 𝑠 1−𝑠
∞
∑︁
𝜁 𝑝 (𝑠) := 𝑝 −𝑒𝑠 = (1 − 𝑝 −𝑠 ) −1 ;
𝑒=0
we recover these factors from an integral. First we need a measure to integrate against.
We have Z 𝑝 = lim Z/𝑝 𝑛 Z as a projective limit with compatible projection maps
←−−𝑛
𝜋 𝑛 : Z 𝑝 → Z/𝑝 𝑛 Z. We define the measure on Z 𝑝 as the projective limit of the
counting measures on each Z/𝑝 𝑛 Z with total measure 1, i.e., for a set 𝐸 ⊆ Z 𝑝 we
define
#𝜋 𝑛 (𝐸)
𝜇 𝑝 (𝐸) := lim
𝑛→∞ 𝑝𝑛
494 CHAPTER 29. IDELIC ZETA FUNCTIONS
when this limit exists. The measure extends additively to Q 𝑝 = Z 𝑝 [1/𝑝] and is
invariant under additive translation
the measure is invariant under 𝑥 ← 𝑎𝑥 for 𝑎 ∈ Q×𝑝 as well as under the substitution
𝑥 ← 𝑥 −1 , and with our normalization we have
−1
1
𝜇×𝑝 (Z×𝑝 ) = 1− 𝜇 𝑝 (Z×𝑝 ) = 1.
𝑝
This looks a bit weird at first, because it is not over a subgroup of Q×𝑝 or anything.
Nevertheless, it works! For every nonzero 𝑥 ∈ Z 𝑝 r {0}, we may write 𝑥 = 𝑝 𝑒 𝑥0 with
𝑥0 ∈ Z×𝑝 and 𝑒 ≥ 0, therefore the integral can be written as a sum over the level sets
𝑝 𝑒 Z×𝑝 :
∫ ∞
∑︁
|𝑥| 𝑠𝑝 d𝜇×𝑝 (𝑥) = 𝑝 −𝑒𝑠 𝜇× (Z×𝑝 ) = (1 − 𝑝 −𝑠 ) −1 = 𝜁 𝑝 (𝑠). (29.2.3)
Z 𝑝 r{0} 𝑒=0
It is more common to rewrite this as an integral over Q×𝑝 by letting Ψ 𝑝 be the charac-
teristic function of Z 𝑝 r {0}, so that
∫ ∫
|𝑥| 𝑠𝑝 d𝜇×𝑝 (𝑥) = |𝑥| 𝑠𝑝 Ψ 𝑝 (𝑥) d𝜇×𝑝 (𝑥). (29.2.4)
Z 𝑝 r{0} Q×𝑝
In this chapter, we use this method of Poisson summation and idelic integrals to
prove the basic properties of the zeta function of a central simple algebra over a global
field (Main Theorem 29.10.1). Translated back into classical language, we prove as
a consequence the key result (the crux of which is Theorem 26.8.19) announced in
section 26.8.
Main Theorem 29.2.6. Let 𝐹 be a number field with ring of integers 𝑅 = Z𝐹 and let
𝐵 be a quaternion algebra over 𝐹 with maximal order O. Let
∑︁
𝜁 𝐵 (𝑠) := N (𝐼) −𝑠
𝐼 ⊆O
be the sum over nonzero right O-ideals, where N is the absolute (counting) norm
(16.4.7) and let 𝜉 𝐵 (𝑠) be its completion (26.8.13). Then 𝜉 𝐵 (𝑠) has meromorphic
continuation to C, holomorphic away from {0, 1/2, 1} with simple poles at 𝑠 = 0, 1,
and it satisfies the functional equation
𝜉 𝐵 (1 − 𝑠) = 𝜉 𝐵 (𝑠). (29.2.7)
29.3 Measures
In this section, we define the local measures we will use. As references for this
section and the next, see Bekka–de la Harpe–Valette [BHV2008, Appendices A and
B], Deitmar [Dei2005], Deitmar–Echterhoff [DE2009, Chapters 1 and 3, Appendix
B], Loomis [Loo53], and Ramakrishnan–Valenza [RM99, Chapters 1–4] as references
on harmonic analysis, and Vignéras [Vig80a, §II.4] and Weil [Weil74, Chapter XI] for
the present context.
Let 𝐺 be a Hausdorff, locally compact, second countable topological group. For
example, we may take 𝐺 = SL2 (R) or 𝐺 = SL𝑛 (R) (or more generally a semisimple
real Lie group). A Borel measure on 𝐺 is a countably additive function, with values
in [0, ∞], defined on the 𝜎-algebra generated by open sets in 𝐺 under complement
and finite or countable unions.
We will construct Haar measures explicitly as we need them, so we need not appeal
to the general result of Proposition 29.3.3 beyond the uniqueness statement which is
itself straightforward to establish: see Exercise 29.4.
Example 29.3.4. On 𝐺 = R𝑛 under addition, a left (and right) Haar measure is given
by the usual Lebesgue measure.
Example 29.3.5. Suppose 𝐺 has the discrete topology (all sets are open). Then the
counting measure 𝜇(𝐸) = #𝐸 ∈ Z ≥0 ∪ {∞} is a left (and right) Haar measure.
29.3.6. In general, a left Haar measure need not also be right translation-invariant. For
𝑔 ∈ 𝐺, the measure defined by 𝜇𝑔 (𝐸) := 𝜇(𝐸𝑔) for a Borel set 𝐸 is again a left Haar
measure, so by Proposition 29.3.3, we have 𝜇𝑔 = Δ𝐺 (𝑔)𝜇 for some Δ𝐺 (𝑔) ∈ R>0 .
Since 𝜇 is unique up to scaling, the function Δ𝐺 (𝑔) does not depend on 𝜇.
29.3. MEASURES 497
By 29.3.6, 𝐺 is unimodular if and only if every left Haar measure is a right Haar
measure.
29.3.10. In this paragraph, we briefly review the theory of integration we need; for
further details, see Deitmar–Echterhoff [DE2009, Appendix B.1]. Let 𝑓 : 𝐺 → C
be a complex-valued function on 𝐺. We say 𝑓 is measurable if for all Borel sets
𝐸 ⊆ C, the subset 𝑓 −1 (𝐸) is measurable; we say 𝑓 is real(-valued) if 𝑓 (𝐺) ⊆ R and
is nonnegative if 𝑓 (𝐺) ⊆ R ≥0 .
If 𝐸 is a measurable set then the characteristic function 1𝐸 of 𝐸 (equal to 1 on 𝐸
and 0 outside 𝐸) is defined to have integral
∫ ∫
1𝐸 (𝑥) d𝜇(𝑥) := 𝜇(𝐸) = d𝜇(𝑥). (29.3.11)
𝐺 𝐸
We say that left Haar measures on 𝐺, 𝑁, 𝐻 are compatible if for all integrable functions
𝑓 : 𝐺 → C,
∫ ∫ ∫
𝑓 (𝑥) d𝜇(𝑥) = 𝑓 (𝑧𝑦) d𝜇(𝑧) d𝜇(𝜙(𝑦)). (29.3.14)
𝐺 𝐻 𝑁
Given measures on two terms, there exists a unique compatible measure on the third—
but note, this measure depends on the exact sequence (Exercise 29.7).
for all measurable sets 𝐸. In particular, for all integrable functions 𝑓 on 𝐺 we have
∫ ∫ ∫
𝑓 (𝑥) d𝜇(𝑥) = 𝑓 (𝜙(𝑥)) d𝜇(𝜙(𝑥)) = k𝜙k 𝑓 (𝜙(𝑥)) d𝜇(𝑥). (29.4.3)
𝐺 𝐺 𝐺
Now let 𝐴 be a Hausdorff, locally compact, second countable topological ring and
let 𝜇 be a left Haar measure on the additive group of 𝐴. Since 𝐴 is abelian, 𝐴 is
unimodular and 𝜇 is also a right Haar measure.
d𝜇(𝑎𝑥)
k𝑎k = . (29.4.7)
d𝜇(𝑥)
d𝜇(𝑥)
d𝜇× (𝑥) := (29.4.9)
k𝑥k
Remark 29.4.11. We reserve the term character for continuous group homomorphisms
𝐴 → 𝐾 × , where 𝐾 is a field (of values for the character); this notion makes sense
for any field 𝐾. Some authors call unitary characters just characters, then calling our
characters instead quasi-characters.
Now suppose that there exists 𝜓 ∈ 𝐴∨ such that the map
𝐴 → 𝐴∨
𝑥 ↦→ (𝑦 ↦→ 𝜓(𝑥𝑦))
𝑓∨: 𝐴 → C
∫
𝑓 ∨ (𝑥) = 𝑓 (𝑦)𝜓(𝑥𝑦) d𝜇(𝑦).
𝐴
Proof. The proof of this theorem is beyond the scope of this textbook, and we use it as
a black box; see e.g. Deitmar–Echterhoff [DE2009, Theorem 3.5.8] or Folland [Fol95,
Theorems 4.32–4.33].
29.5.1. The (additive) Haar measure 𝜇 on 𝐵 is the usual (Lebesgue) measure, normal-
ized as follows:Íletting 𝑛 = dimR 𝐵, we choose an R-basis 𝑒 1 , . . . , 𝑒 𝑛 for 𝐵, so that we
may write 𝑥 = 𝑖 𝑥𝑖 𝑒 𝑖 ∈ 𝐵 with 𝑥𝑖 are coordinates on 𝐵, and we define
d𝑥 := | 𝑑 (𝑒 1 , . . . , 𝑒 𝑗 )| 1/2 d𝑥1 · · · d𝑥 𝑛
where 𝑑 is the discriminant defined by (15.2.1) and the reduced trace is taken on 𝐵 as
an R-algebra. By Lemma 15.2.5, we see that this measure is independent of the choice
of basis 𝑒 𝑖 .
Another application of Lemma 15.2.5 then gives the modulus
for all 𝛼 ∈ 𝐵× .
d𝑥 = d𝑥, for 𝑥 ∈ R;
√
d𝑧 = 2 d𝑥 d𝑦, for 𝑧 = 𝑥 + 𝑦 −1 ∈ C;
d𝛼 = 4 d𝑡 d𝑥 d𝑦 d𝑧, for 𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 ∈ H;
Ö
d𝛼 = d𝑥𝑖 𝑗 , for 𝛼 = (𝑥 𝑖 𝑗 )𝑖, 𝑗 ∈ M𝑛 (R);
𝑖, 𝑗
2
Ö √
d𝛼 = 2𝑛 d𝑥 𝑖 𝑗 d𝑦 𝑖 𝑗 , for 𝛼 = (𝑥𝑖 𝑗 + 𝑦 𝑖 𝑗 −1)𝑖, 𝑗 ∈ M𝑛 (C).
𝑖, 𝑗
And:
The modulus for R is the usual absolute value, whereas the modulus for C is the square
of the absolute value (cf. Remark 12.2.3), explaining conventions on normalized
absolute values in the product formula 14.4.6.
d𝜇(𝛼)
d𝜇× (𝛼) := . (29.5.5)
k𝛼k
29.5. ARCHIMEDEAN MEASURES AND ZETA FUNCTIONS 501
𝜓 : 𝐵 → C1
𝛼 ↦→ exp(−2𝜋𝑖 trd(𝛼))
with reduced trace taken on 𝐵 as an R-algebra. We note that 𝜓(𝛼𝛽) = 𝜓(𝛽𝛼) for all
𝛼, 𝛽 ∈ 𝐵.
A fundamental result in standard Fourier analysis (generalizing the case 𝐵 = R, C
and following from it in the same way) is the following proposition.
Proposition 29.5.7. The standard unitary character induces an isomorphism 𝐵 −→ ∼ 𝐵∨
of topological groups, and the measure 𝜇 defined in 29.5.1 is self-dual (with respect
to 𝜓).
In light of Proposition 29.5.7, we will also write 𝜏 = 𝜇 for the measure defined
above.
29.5.8. The exact sequence
nrd
1 → H1 → H× −−→ R>0 → 1
𝐸 = {𝛼 ∈ H× : nrd(𝛼) ≤ 𝜌 2 }
be the punctured ball of radius 𝜌. Let 𝑓 be the function k𝛼k on 𝐸, zero elsewhere. On
the one hand, ∫ ∫
× d𝜏(𝛼)
𝑓 (𝛼) d𝜏 (𝛼) = k𝛼k = 𝜏(𝐸).
H × 𝐸 k𝛼k
Recalling that 𝜏 is 4 times Lebesgue measure, and that the Lebesgue measure of a
sphere of radius 𝜌 has volume 𝜋 2 𝜌 4 /2, we get 𝜏(𝐸) = 2𝜋 2 𝜌 4 . On the other hand, by
compatibility, this integral is equal to
∫ ∫ ∫ 𝜌∫
1 × 2 2𝑟 d𝑟
𝑓 (𝛼1 𝑟) d𝜏 (𝛼1 ) d (𝑟 ) = k𝛼1 𝑟 kd𝜏 1 (𝛼1 ) 2
R>0 H 1 0 H 1 𝑟
𝜌4
∫ 𝜌
= 2𝜏 1 (H1 ) 𝑟 3 d𝑟 = 𝜏 1 (H1 ) .
0 2
We conclude that 𝜏 1 (H1 ) = 4𝜋 2 .
We will need further functions to integrate, so we make the following definition.
When we consider functions on 𝐵, we Í may think of choosing an R-basis 𝑒 1 , . . . , 𝑒 𝑛 of
𝐵, identifying 𝐵 ' R𝑛 and writing 𝑥 = 𝑖 𝑥𝑖 𝑒 𝑖 ∈ 𝐵; our definitions will be independent
of this choice of basis.
502 CHAPTER 29. IDELIC ZETA FUNCTIONS
𝑄(𝑥) = 𝑥 2 , for 𝑥 ∈ R;
2 2 2
√
𝑄(𝑧) = 2| 𝑧| = 2(𝑥 + 𝑦 ), for 𝑧 = 𝑥 + 𝑦 −1 ∈ C;
𝑄(𝛼) = 2 nrd(𝛼), for 𝛼 ∈ H;
∑︁ (29.5.12)
𝑄(𝛼) = 𝑥𝑖2𝑗 , for 𝛼 = (𝑥 𝑖 𝑗 )𝑖, 𝑗 ∈ M𝑛 (R); and
𝑖, 𝑗
∑︁
𝑄(𝛼) = 2 | 𝑧𝑖 𝑗 | 2 , for 𝛼 = (𝑧 𝑖 𝑗 )𝑖, 𝑗 ∈ M𝑛 (C).
𝑖, 𝑗
𝑍R (𝑠) = ΓR (𝑠),
𝑍C (𝑠) = 𝜋ΓC (𝑠),
𝑍H (𝑠) = 𝜋(2𝑠 − 1)ΓR (2𝑠)ΓR (2𝑠 − 1) = 2𝜋 2 ΓR (2𝑠)ΓR (2𝑠 + 1),
𝑍M2 (R) (𝑠) = 𝜋ΓR (2𝑠)ΓR (2𝑠 − 1),
𝑍M2 (C) (𝑠) = 2𝜋 3 ΓC (2𝑠)ΓC (2𝑠 − 1).
Proof. We have ∫ ∞
2 d𝑥
𝑍R (𝑠) = 2 𝑥 𝑠 𝑒− 𝜋 𝑥 ;
0 𝑥
29.6. LOCAL MEASURES: COMMUTATIVE NONARCHIMEDEAN CASE 503
making the substitution 𝑥 ← 𝜋𝑥 2 gives the result. A similar argument with polar
coordinates gives 𝑍C (𝑠). The remaining integrals are pretty fun, so they are left as
Exercise 29.8.
Lemma 29.5.18 explains the provenance of the definitions of ΓR (𝑠), ΓC (𝑠) from
(26.8.3), and ultimately their appearance in (26.8.14). (A quick check on the con-
stant in front is provided by 𝑍 𝐵 (1) = 1; these particular choices of constants follow
convention.) By comparison, Lemma 29.5.18 then shows that in general 𝑍 𝐵Φ (𝑠) has
meromorphic continuation to C.
we have
1
𝜇(𝑅 × ) = 1 − . (29.6.5)
𝑞
504 CHAPTER 29. IDELIC ZETA FUNCTIONS
so that 𝜇× (𝑅 × ) = 1.
Next, we consider the Fourier transform in this context.
29.6.7. We first define an additive homomorphism
h i𝐹 : 𝐹 → R/Z
as follows.
(a) If 𝐹 = Q 𝑝 , we define h𝑥iQ 𝑝 ∈ Q to be such that 0 ≤ h𝑥iQ 𝑝 < 1 and 𝑥 − h𝑥iQ 𝑝 ∈
Z𝑝. Í
(b) If 𝐹 = F𝑞 ((𝑡)) and 𝑥 = 𝑖 𝑎 𝑖 𝑡 𝑖 then we take h𝑥i𝐹 := TrF𝑞 /F 𝑝 (𝑎 −1 )/𝑝.
(c) In general, if 𝐹 ⊇ 𝐹0 is a finite separable extension of fields, then we define
𝜓 𝐹 : 𝐹 → C1
𝜓 𝐹 (𝑥) = exp(2𝜋𝑖h𝑥i𝐹 ).
of topological groups.
Proof. First consider (a), and suppose 𝐹 is a finite extension of Q 𝑝 . We seek to satisfy
(29.4.14); the equation holds up to a constant, so we may choose appropriate 𝑓 and 𝑥.
We choose 𝑓 as the characteristic function of 𝑅 and 𝑥 = 0, so that 𝑓 (0) = 1 and
∫ ∫ ∫ ∫
𝑓 (𝑧)𝜓(𝑦𝑧) d𝜇(𝑧) d𝜇(𝑦) = 𝑓 (𝑧) 𝜓(𝑦𝑧) d𝜇(𝑦) d𝜇(𝑧). (29.6.11)
𝐹 𝐹 𝐹 𝐹
29.7. LOCAL ZETA FUNCTIONS: NONARCHIMEDEAN CASE 505
By character theory,
∫
𝜓(𝑦𝑧) d𝜇(𝑦) = 𝜇({𝑦 ∈ 𝐹 : 𝜓(𝑦𝑧) = 1}) = 𝜇({𝑦 ∈ 𝐹 : Tr𝐹 /Q 𝑝 (𝑦𝑧) ∈ Z 𝑝 })
𝐹
Let 𝑥𝑖 be a Z 𝑝 -basis for 𝑅 with 𝑥𝑖♯ the dual basis, giving a Z 𝑝 -basis for 𝑅 ♯ . By Lemma
15.6.17 we have
disc(𝑅) = [𝑅 ♯ : 𝑅] Z 𝑝
so since 𝜇(𝑅) = 1 by additivity we have
𝜇(𝑅 ♯ ) = |disc(𝑅)| −1 .
29.7.2. For the (multiplicative) Haar measure 𝜇× on 𝐵× , we use the same normalization
factor (29.6.5) as for 𝐹 × , defining the normalized measure
d𝜇(𝛼)
d𝜇× (𝛼) := (1 − 1/𝑞) −1 . (29.7.3)
k𝛼k
𝜓 𝐵 (𝛼) := 𝜓 𝐹 (trd(𝛼))
Proof. The same arguments as in Propositions 29.6.9 and 29.6.10 apply, with appro-
priate modifications.
29.7.9. Having normalized the multiplicative measure 𝜇× and seeing the relevant
modification in (29.7.8), just as in the case of 𝐵 = 𝐹 we define
We may now define the local zeta function in the nonarchimedean context.
Definition 29.7.11. A function 𝑓 : 𝐵 → C is Schwartz–Bruhat if 𝑓 is locally constant
(for every 𝛼 ∈ 𝐵, there exists an open neighborhood 𝑈 3 𝑥 such that 𝑓 |𝑈 is constant)
with compact support. A standard function Ψ on 𝐵 is the characteristic function of
a maximal order.
Taking an open cover of the support, we see that every Schwartz–Bruhat function
can be expressed as a (finite) C-linear combination of characteristic functions of
compact open subsets of 𝐵.
Definition 29.7.12. For a Schwartz–Bruhat function Φ, we define the (local) zeta
function ∫
𝑍 𝐵Φ (𝑠) := k𝛼k 𝑠 Φ(𝛼) d𝜏 × (𝛼). (29.7.13)
𝐵×
We write ∫
𝑍 𝐵 (𝑠) := 𝑍 𝐵Ψ (𝑠) = k𝛼k 𝑠 d𝜏 × (𝛼) (29.7.14)
O∩𝐵 ×
for the zeta function with respect to a standard function Ψ; this is well-defined (inde-
pendent of Ψ) as any two maximal orders are conjugate, and k𝛼k is well-defined on
conjugacy classes by 29.7.1.
We begin with a basic convergence estimate.
Lemma 29.7.15. The function 𝑍 𝐵Φ (𝑠) converges for Re 𝑠 > 1.
29.7. LOCAL ZETA FUNCTIONS: NONARCHIMEDEAN CASE 507
Proof. We reduce to the case where Φ is the characteristic function of a compact open
set 𝑈, since Φ is a (finite) C-linear combination of such. Since 𝑈 is compact, k𝛼k is
bounded on 𝑈, so too is k𝛼k 𝑠−1 for fixed Re 𝑠 > 1; thus
∫ ∫
k𝛼k 𝑠 d𝜏 × (𝛼) = k𝛼k 𝑠−1 d𝜏(𝛼)
𝑈 ∩𝐵 × O∩𝐵 ×
This integral representation recovers the classical zeta function we studied earlier
in the number field case (section 26.4).
29.7.16. Since there is a unique ideal in 𝑅 of absolute norm 𝑞 𝑒 for each 𝑒 ≥ 0, we
have
∞
∑︁ 1
𝜁 𝐹 (𝑠) = 𝑒𝑠
= (1 − 𝑞 −𝑠 ) −1 .
𝑒=0
𝑞
In like fashion, for 𝑒 ≥ 0, let 𝑎 𝔭𝑒 = 𝑎 𝔭𝑒 (O) be the number of (necessarily principal)
right ideals of O of reduced norm 𝔭𝑒 and thereby absolute norm 𝑞 𝑛𝑒 . We then define
∞
∑︁ 1 ∑︁ 𝑎 𝔭𝑒
𝜁 𝐵 (𝑠) := = ,
N(𝐼) 𝑠 𝑞 𝑛𝑒𝑠
𝐼 ⊆O 𝑒=0
(c) If 𝐵 ' M𝑛 (𝐹), then 𝜁 𝐵 (𝑠) converges for Re 𝑠 > 1 − 1/𝑛, and
𝑛−1
Ö
𝜁 𝐵 (𝑠) = 𝜁 𝐹 (𝑛𝑠 − 𝑖).
𝑖=0
Proof. First part (a). For 𝑒 ≥ 0, choose representatives for the 𝑎 𝔭𝑒 = 𝑎 𝔭𝑒 (O) classes
in O/O× with reduced norm 𝔭𝑒 . Every element 𝛼 ∈ O ∩ 𝐵× can be written as the
product of one of the representatives and an element of O× ; since k𝛼k = |nrd(𝛼) 𝑛 | we
have
∫ ∞
× × ×
∑︁ 𝑎 𝔭𝑒
𝑍 𝐵 (𝑠) = 𝑠
k𝛼k d𝜏 (𝛼) = 𝜏 (O ) 𝑛𝑒𝑠
= 𝜏 × (O× )𝜁 𝐵 (𝑠) (29.7.18)
O∩𝐵 ×
𝑒=0
𝑞
508 CHAPTER 29. IDELIC ZETA FUNCTIONS
as claimed. For (b) and (c), we gave a formula for 𝑎 𝔭𝑒 for 𝐵 = M𝑛 (𝐹) in Exercise
26.12 and for 𝐵 a quaternion algebra in Corollary 26.4.7 (which extends to the function
field case without change). Of course, these identities only hold within their respective
domains of convergence, which for 𝜁 𝐹 (𝑠) = (1 − 𝑞 −𝑠 ) −1 is Re 𝑠 > 0, thereby giving
the rest.
In the proof of the functional equation, we will need the following proposition,
recalling duality (section 15.6).
Proposition 29.7.19. For Ψ the characteristic function of O, we have
(
∨ 𝜏(O) = D (𝐵) −1/2 , if 𝛼 ∈ O♯ ;
Ψ (𝛼) =
0, otherwise.
Moreover,
∨
𝑍 𝐵Ψ (𝑠) = D (𝐵) 𝑠−1/2 𝑍 𝐵Ψ (𝑠). (29.7.20)
𝑠−1/2 𝑠 × 𝑠−1/2 Ψ
= D (𝐵) k𝛼k d𝜏 (𝛼) = D (𝐵) 𝑍 𝐵 (𝑠),
O∩𝐵 ×
(29.7.21)
proving the second statement.
Proof. If 𝐵 is a division ring, then O is the valuation ring; let 𝐽 ⊆ O be the maximal
ideal, so O/𝐽 ' F𝑞2 hence 𝜇(𝐽) = 1/𝑞 2 , and then
1 − 1/𝑞 2
𝜇× (O× ) = (1 − 1/𝑞) −1 (𝜇(O) − 𝜇(𝐽)) = = 1 + 1/𝑞
1 − 1/𝑞
29.8. IDELIC ZETA FUNCTIONS 509
Similarly, if 𝐵 ' M2 (𝐹) then O ' M2 (𝑅), and from the exact sequence
29.7.24. Let 𝐵 be a quaternion algebra. Then the reduced norm yields an exact
sequence
nrd
1 → O1 → O× −−→ 𝑅 × → 1
the product over all nonarchimedean places of the local zeta functions 𝜁 𝐵𝑣 (𝑠) defined
in 29.7.16. We studied these (global) zeta functions over number fields in Chapters 25
and 26.
𝑛−1
Ö
𝜁M𝑛 (𝐹 ) (𝑠) = 𝜁 𝐹 (𝑛𝑠 − 𝑖).
𝑖=0
With this in mind, we construct idelic zeta functions as products of local zeta
functions, according to the product measures and characters, as follows.
510 CHAPTER 29. IDELIC ZETA FUNCTIONS
on 𝐵 as the product measure (on the restricted direct product). In particular, for every
eligible set S ⊂ Pl 𝐹 of places and every Î O in 𝐵, the measure 𝜇 restricts to
Î 𝑅 (S) -order
the (convergent) product measure on 𝑣 ∈S 𝐵 𝑣 × 𝑣∉S O 𝑣 —and this uniquely defines
the 𝜇 as a Haar measure on 𝐵.
Definition 29.8.4. The absolute discriminant D (𝐵) of 𝐵 is defined as the product of
the local absolute discriminants over all nonarchimedean places:
Ö
D (𝐵) := D (𝐵 𝑣 ).
𝑣-∞
Î
where 𝑡 𝑣 is a uniformizer at 𝑡, and we again define 𝜓 := 𝑣 𝜓 𝑣 . In both cases,
𝜓(𝛼𝛽) = 𝜓(𝛽𝛼) for all 𝛼, 𝛽 ∈ 𝐵, as this holds for 𝜓 𝑣 for all 𝑣.
In all cases, we define the function
Ö
Φ := Φ𝑣
𝑣
Proof sketch. In the number field case, the statement follows from the local versions
and the product formula; in the function field case, it follows similarly and reduces to
the fact that the sum of the residues is zero.
29.8.12. The Tamagawa measure 𝜏 is then also characterized as the unique Haar
measure on 𝐵 that is self-dual with respect to the Fourier transform associated to the
standard character.
29.8.13. Moving now from additive to multiplicative, we define the normalized Haar
measure Ö
𝜇× = 𝜇×𝑣 (29.8.14)
𝑣
on 𝐵× ,where 𝜇×𝑣
is defined in 29.5.5 for 𝑣 archimedean and (29.7.3) for 𝑣 nonar-
chimedean. Finally, we define the Tamagawa measure on 𝐵× by
Ö
𝜏 × := D (𝐵) −1/2 𝜇× = 𝜏𝑣× , (29.8.15)
𝑣
where 𝜏𝑣× = 𝜇×𝑣 for 𝑣 archimedean and 𝜏𝑣× = D (𝐵 𝑣 ) −1/2 𝜇×𝑣 as in (29.7.10) for 𝑣
nonarchimedean.
With measures finally (!) in hand, we can now idelically integrate functions against
them.
Using 𝜓 to identify 𝐵 with the dual of 𝐵/𝐵 via Proposition 29.8.10 (and attending
carefully to relevant analytic concerns), the symmetrized function is equal to its Fourier
series ∑︁
(ΣΦ) (𝛼) = 𝑎 𝛽 𝜓(−𝛽𝛼) (29.8.19)
𝛽 ∈𝐵
where ∫
𝑎𝛽 = (ΣΦ) (𝛼)𝜓(𝛽𝛼) d𝜏(𝛼)
𝐵/𝐵
∫ ∑︁
= Φ(𝛼 + 𝛾)𝜓(𝛽(𝛼 + 𝛾)) d𝜏(𝛼) (29.8.20)
𝐵/𝐵 𝛾 ∈𝐵
∫
= Φ(𝛼)𝜓(𝛽𝛼) d𝜏(𝛼) = Φ∨ (𝛽)
𝐵
Remark 29.8.21. When 𝐹 is a function field, the alignment in 𝜓 implies the Riemann–
Roch theorem for the curve with function field 𝐹: see Exercise 29.13.
We have
Φ
Ö
𝑍 𝐵 (𝑠) = 𝑍 𝐵Φ𝑣𝑣 (𝑠) (29.8.23)
𝑣 ∈Pl 𝐹
wherever the product is absolutely convergent; the local zeta functions 𝑍 𝐵𝑣 (𝑠) are
defined in (29.5.16) for archimedean 𝑣 and (29.7.13) for nonarchimedean 𝑣.
We can make the comparison to the classical zeta function (29.8.1) explicit up to
an idelic volume, as follows.
𝜏× (O
b b× ) −1 = | 𝑑 𝐹 | 2 𝜁 𝐹 (2)𝜑(𝔇) (29.8.25)
Î
where 𝜑(𝔇) = 𝔭 |𝔇 (Nm(𝔭) − 1). If 𝐹 is a function field, then
Ö
𝜏 × (O× ) −1 = 𝜁 𝐹 (2) (1 − 1/𝑞 𝑣 ).
𝑣 ∈Ram(𝐵)
Proof. For part (a), we use Lemma 29.7.17 for the relationship between zeta functions.
For part (b), we combine Lemmas 29.5.18 and 29.7.17(b) and we use Lemma 29.7.22
for the local computation of measure and the equality
Ö √︁
(1 − N(𝔭) −1 ) D (𝐵) = | 𝑑 𝐹 | 2 𝜑(𝔇)
𝔭 |𝔇
𝐵 (1) = 𝛼 ∈ 𝐵× : k𝛼k = 1 ,
k𝛼k 𝐵 = knrd(𝛼) 𝑛 k 𝐹 .
1 → 𝐵 (1) → 𝐵× →
− k𝐵× k → 1; (29.8.27)
Φ Φ∨ (0)𝜁 𝐹∗ (1)
res𝑠=1 𝑍 𝐵 (𝑠) =
𝑛
where 𝜁 𝐹∗ (1) = res𝑠=1 𝜁 𝐹 (1).
Proof. In a nutshell, we prove this proposition by comparison to the classical zeta
function 𝜁M𝑛 (𝐹 ) (𝑠) of the matrix ring, for which we can compare explicitly with a
Dirichlet series. Î
We may suppose without loss of generality that Φ = 𝑣 Φ 𝑣 , since by definition Φ
is a linear combination of such. Let 𝑆 ⊆ Pl(𝐹) be the (finite) set of places 𝑣 of 𝐹 such
that one of the following holds:
• 𝑣 is archimedean (if 𝐹 is a number field);
• 𝐵 𝑣 ; M𝑛 (𝐹𝑣 ); or
• Φ 𝑣 ≠ Ψ 𝑣 , i.e., Φ 𝑣 is not the standard function.
Then
Φ Φ
Ö Ö
𝑍 𝐵 (𝑠) = 𝑍 𝐵𝑣 (𝑠) 𝑍M𝑛 (𝐹𝑣 ) (𝑠). (29.9.2)
𝑣 ∈𝑆 𝑣∉𝑆
where the first product is finite. By Lemma 29.7.15 for 𝑣 ∈ 𝑆 archimedean and com-
parison to Lemma 29.5.18 for 𝑣 archimedean, the first product is absolutely convergent
for Re 𝑠 > 1. For the second (infinite) product, by Lemma 29.7.17(c), we have
𝑛−1
Ö 𝑛−1
Ö
𝜁M𝑛 (𝐹𝑣 ) (𝑠) = 𝜁 𝐹𝑣 (𝑛𝑠 − 𝑖) = (1 − 𝑞 𝑖−𝑛𝑠 ) −1 (29.9.3)
𝑖=0 𝑖=0
where 𝑞 𝑣 is the size of the residue field of 𝐹𝑣 ; note 𝜁M𝑛 (𝐹𝑣 ) (𝑠) −1 = 𝑖 (1 − 𝑞 𝑖−𝑛𝑠 ) is
Î
holomorphic. Putting these together gives
Ö Ö
𝜏× (O
𝑍M𝑛 (𝐹𝑣 ) (𝑠) = b b× )𝜁M (𝐹 ) (𝑠)
𝑛 𝑍M𝑛 (𝐹𝑣 ) (𝑠) −1 .
𝑣∉𝑆 𝑣 ∈𝑆
and so we reduce to showing that 𝜁M𝑛 (𝐹 ) (𝑠) is absolutely convergent for Re(𝑠) > 1.
But multiplying (29.9.3) gives
𝑛−1
Ö
𝜁M𝑛 (𝐹 ) (𝑠) = 𝜁 𝐹 (𝑛𝑠 − 𝑖)
𝑖=0
29.9. CONVERGENCE AND RESIDUE 515
coming from the normalization factors between additive and multiplicative Haar mea-
sure at the nonarchimedean places, where 𝑞 𝑣 is the size of the residue field of 𝐹𝑣 .
Therefore
Φ𝑣
𝑍 𝐵𝑣 (𝑠) 1 − 𝑞 −𝑛𝑠+𝑛−1
𝑣
∫
lim = lim Φ 𝑣 (𝛼 𝑣 ) k𝛼k 𝑠−1 d𝛼 𝑣
𝑠&1 𝜁 𝐹𝑣 (𝑛𝑠 − 𝑛 + 1) 𝑠&1 1 − 𝑞 −1
𝑣 𝐵 ×
𝑣 (29.9.6)
∫
∨
= Φ 𝑣 (𝛼 𝑣 ) d𝛼 𝑣 = Φ 𝑣 (0).
𝐵𝑣
the interchange of the product and the limit is justified by absolute convergence of the
product in a neighborhood of 𝑠 = 1. Therefore
Φ Φ∨ (0)𝜁 𝐹∗ (1)
res𝑠=1 𝑍 𝐵 (𝑠) = Φ∨ (0) res𝑠=1 𝜁 𝐹 (𝑛𝑠 − 𝑛 + 1) =
𝑛
as claimed, finishing the proof of (b).
516 CHAPTER 29. IDELIC ZETA FUNCTIONS
Main Theorem 29.10.1. Let 𝐹 be a global field, let 𝐵 be a central division algebra
over 𝐹 with 𝑛2 = dim𝐹 𝐵. Let Φ be a Schwartz–Bruhat function on 𝐵. Then the
following statements hold.
Φ
(a) The function 𝑍 𝐵 (𝑠) has meromorphic continuation to C. Moreover:
Φ
(i) If 𝐹 is a number field, then 𝑍 𝐵 (𝑠) is holomorphic in C r {0, 1} with simple
poles at 𝑠 = 0, 1 and residues
Φ Φ∨
𝑍 𝐵 (1 − 𝑠) = 𝑍 𝐵 (𝑠). (29.10.2)
(c) We have
𝜏 (1) (𝐵× \𝐵 (1) ) = 𝜁 𝐹∗ (1).
The value of the residue 𝜁 𝐹∗ (1) is given by the analytic class number formula
(Theorem 26.2.3). Taking 𝑛 = 1 gives the following important special case (without
requiring this as input).
(a) If 𝐹 is a number field with absolute discriminant 𝑑 𝐹 , let 𝜉 𝐹 (𝑠) be the completed
Dedekind zeta function as defined in (26.8.4). Then 𝜉 𝐹 (𝑠) is holomorphic in
C r {0, 1}, with simple poles at 𝑠 = 0, 1, and satisfies the functional equation
𝜉 𝐹 (1 − 𝑠) = 𝜉 𝐹 (𝑠).
(b) If 𝐹 is a function field of genus 𝑔 with field of constants F𝑞 , let 𝜉 𝐹 (𝑠) :=
𝑞 (𝑔−1)𝑠 𝜁 𝐹 (𝑠). Then 𝜁 𝐹 (𝑠) is holomorphic in C r {𝑠 : 𝑞 𝑠 = 𝑞 0 , 𝑞 1 }, with simple
poles when 𝑞 𝑠 = 𝑞 0 , 𝑞 1 , and satisfies the functional equation 𝜉 𝐹 (1 − 𝑠) = 𝜉 𝐹 (𝑠).
29.10. MAIN THEOREM 517
Î
Proof. Taking 𝑛 = 1 gives 𝐵 = 𝐹; we take Φ = Ψ = 𝑣 Ψ 𝑣 to be the standard
function.
In the number field case, we have 𝜉 (𝑠) = | 𝑑 𝐹 | 𝑠/2 𝑍 𝐹 (𝑠). By Lemma 29.8.24 and
Lemma 29.5.18 for the archimedean contribution, we see that 𝜉 𝐹 (𝑠) = 𝑐 𝐹 | 𝑑 𝐹 | 𝑠/2 𝑍 𝐹 (𝑠)
where 𝑐 𝐹 is a constant depending on 𝐹. We accordingly conclude holomorphicity
from Main Theorem 29.10.1(a). For the functional equation, by (29.7.20), we have
𝑍 𝐹𝑣 (𝑠) = D (𝐹𝑣 ) 𝑠−1/2 𝑍 𝐹𝑣 (𝑠). Taking the product and reading off Main Theorem
29.10.1(b) gives
The function field case in concluded in a similar but more direct manner.
Main Theorem 29.10.1 extends to the case of a matrix algebra (over a division
algebra), with some additional complications because of the existence of zerodivisors:
see Theorem 29.10.20 for the case 𝐵 = M2 (𝐹). As we will see, the proof is close
to uniform in the number field and function field cases; we have separated these two
cases in the statement for clarity, so in particular the poles in the function field case
occur at 𝑠 ∈ 2𝜋𝑖(log 𝑞)Z and 𝑠 ∈ 1 + 2𝜋𝑖(log 𝑞)Z.
Remark 29.10.4. Weil [Weil82] chooses a different normalization, giving a functional
equation relating 𝑠 to 𝑛 − 𝑠 and with residues Φ∨ (0)𝜁 𝐹∗ (1) at 𝑠 = 0 and −Φ(0)𝜁 𝐹∗ (1)
at 𝑠 = 𝑛.
𝑓+ (𝛼) := 𝜆(k𝛼k −1 ),
(29.10.6)
𝑓− (𝛼) := 𝜆(k𝛼k).
Then 𝑓+ (𝛼−1 ) = 𝑓− (𝛼) and 𝑓+ (𝛼) + 𝑓− (𝛼) = 1 for all 𝛼 ∈ 𝐵× , and so defining
∫
Φ,±
𝑍 𝐵 (𝑠) = 𝑓± (𝛼) k𝛼k 𝑠 Φ(𝛼) d× 𝛼, (29.10.7)
𝐵×
we obtain
Φ Φ,+ Φ,−
𝑍 𝐵 (𝑠) = 𝑍 𝐵 (𝑠) + 𝑍 𝐵 (𝑠). (29.10.8)
Φ,+
We claim that the function 𝑍 𝐵 (𝑠) is holomorphic. Indeed, by Proposition 29.9.1,
Φ Φ,+
𝑍 𝐵 (𝑠) converges absolutely for Re 𝑠 > 1; thus the same is true for 𝑍 𝐵 (𝑠). But if
Φ,+
𝑍 𝐵 (𝑠) converges absolutely at 𝑠 = 𝑠0 , then it does so for all Re(𝑠) ≤ Re(𝑠0 ), so
Φ,+
𝑍 𝐵 (𝑠) is holomorphic in C.
518 CHAPTER 29. IDELIC ZETA FUNCTIONS
Φ,−
For the remaining piece 𝑍 𝐵 (𝑠), we have
∫ !
Φ,−
∑︁
𝑍 𝐵 (𝑠) = 𝑓− (𝛼) k𝛼k 𝑠
Φ(𝛽𝛼) d× 𝛼. (29.10.9)
𝐵 × \𝐵 × 𝛽 ∈𝐵 ×
using that 𝜓 is well-defined on conjugacy classes in the last equality. Plugging this
into Theorem 29.8.17 gives
∑︁ 1 ∑︁ ∨ −1
Φ(𝛽𝛼) = Φ (𝛼 𝛽). (29.10.11)
𝛽 ∈𝐵
k𝛼k 𝛽 ∈𝐵
At this point, we use the hypothesis that 𝐵 is a division algebra over 𝐹. With this
assumption in hand, ∑︁ ∑︁
Φ(𝛽𝛼) = Φ(𝛽𝛼) − Φ(0).
𝛽 ∈𝐵 × 𝛽 ∈𝐵
𝑠 ← 1 − 𝑠, 𝛼 ← 𝛼−1
Φ∨ ,+
in the definition of 𝑍 𝐵 (𝑠) (29.10.7). The Tamagawa measure d× 𝛼 is invariant under
inversion 𝛼 ← 𝛼−1 , so
∫
Φ∨ ,+
𝑍 𝐵 (1 − 𝑠) = 𝑓+ (𝛼−1 ) k𝛼k −(1−𝑠) Φ∨ (𝛼−1 ) d× 𝛼
𝐵×
∫ (29.10.13)
= 𝑓− (𝛼) k𝛼k 𝑠−1 Φ∨ (𝛼−1 ) d× 𝛼.
𝐵×
29.10. MAIN THEOREM 519
(replacing 𝛽−1 ← 𝛽 in the sum): in writing the integral this way, we integrate
over any measurable set 𝐵× that maps injectively under the continuous quotient map
𝐵× → 𝐵× \𝐵× . Combining (29.10.12) and (29.10.14), we obtain
∫
Φ,− Φ∨ ,+
𝑍 𝐵 (𝑠) − 𝑍 𝐵 (1 − 𝑠) = 𝜈(k𝛼k) d× 𝛼 (29.10.15)
𝐵 × \𝐵 ×
where
𝜈(𝑡) := Φ∨ (0)𝑡 𝑠−1 − Φ(0)𝑡 𝑠 𝜆(𝑡).
The function 𝜈 only depends on lengths in 𝐵× . Recalling 29.8.26, in particular the
exact sequence (29.8.27), we obtain
∫ ∫ !
× (1) × (1) ×
𝜈(k𝛼k) d 𝛼 = 𝜏 (𝐵 \𝐵 ) 𝜈(𝑡) d 𝑡 (29.10.16)
𝐵 × \𝐵 × k𝐵 × k
is the volume of any measurable set in 𝐵 (1) that maps injectively under the quotient
map.
When 𝐹 is a number field,
Φ∨ (0)
∫ ∫ 1 ∫ 1
d𝑡 Φ(0) d𝑡
𝜈(𝑡) d× 𝑡 = 𝑡 𝑠−1 − 𝑡𝑠
k 𝐵× k 𝑛 0 𝑡 𝑛 0 𝑡
∨ (29.10.17)
1 Φ (0) Φ(0)
=− + .
𝑛 1−𝑠 𝑠
When 𝐹 is a function field with constant field F𝑞 ,
∞
! ∞
!
Φ∨ (0) 1 ∑︁
∫
−𝑑 (𝑠−1) Φ(0) 1 ∑︁ −𝑑𝑠
𝜈(𝑡) d𝑡 = + 𝑞 − + 𝑞
k𝐵 × k 𝑛 2 𝑑=1 𝑛 2 𝑑=1
(29.10.18)
1 + 𝑞 −𝑠
1 ∨ 1 + 𝑞 𝑠−1
=− Φ (0) + Φ(0) .
2𝑛 1 − 𝑞 𝑠−1 1 − 𝑞 −𝑠
We now finish the proof, beginning with the case that 𝐹 is a number field. Com-
bining (29.10.8), (29.10.15), and (29.10.17), we obtain
Φ Φ,+ Φ∨ ,+
𝑍 𝐵 (𝑠) = 𝑍 𝐵 (𝑠) + 𝑍 𝐵 (1 − 𝑠)
1
∨
Φ (0) Φ(0)
(29.10.19)
− 𝜏 (1) (𝐵× \𝐵 (1) ) + .
𝑛 1−𝑠 𝑠
520 CHAPTER 29. IDELIC ZETA FUNCTIONS
The substitution 𝑠 ← 1 − 𝑠 interchanges Φ and Φ∨ , since (Φ∨ ) ∨ (𝑠) = Φ(𝑠); from this
Φ
we conclude the functional equation (b) and that 𝑍 𝐵 (𝑠) has meromorphic continuation.
Φ,+ Φ∨ ,+ Φ
Since 𝑍 𝐵 (𝑠) and 𝑍 𝐵 (𝑠) are entire, we conclude that 𝑍 𝐵 (𝑠) is holomorphic in C
away from 𝑠 = 0, 1. Proposition 29.9.1(b) gives
Φ Φ∨ (0)𝜁 𝐹∗ (1)
res𝑠=1 𝑍 𝐵 (𝑠) =
𝑛
which proves (a); on the other hand, (29.10.19) tells us
Φ Φ0
Ö 𝑍 𝐵Ψ𝑣𝑣 (𝑠)
𝑍 𝐵 (𝑠) = 𝑍 𝐵0 (𝑠) Ψ0
. (29.10.22)
𝑣 ∈Ram(𝐵0 ) 𝑍 𝐵0𝑣 (𝑠)
𝑣
29.10. MAIN THEOREM 521
𝑍 𝐵Ψ𝑣𝑣 (𝑠)
Ψ0
= 𝜁 𝐹𝑣 (2𝑠 − 1) = (1 − 𝑞 1−2𝑠
𝑣 ) −1
𝑍 𝐵0𝑣 (𝑠)
𝑣
As an important special case, we have the following theorem, recalling the calcu-
lation of D (𝐵) (Example 29.8.5).
Theorem 29.10.23. Let 𝐵 be a quaternion algebra over 𝐹 and let D (𝐵) ∈ Z>0 be the
absolute discriminant of 𝐵. Then the zeta function 𝑍 𝐵 (𝑠) has the following properties.
𝜁 ∗ (1) 𝜁 𝐹∗ (1)
res𝑠=1 𝑍 𝐵 (𝑠) = √︁𝐹 , res𝑠=0 𝑍 𝐵 (𝑠) = − ;
2 D (𝐵) 2
Proof. We apply Main Theorem 29.10.1 and Theorem 29.10.20 with Φ as in 29.8.9
with Φ 𝑣 the characteristic function of a maximal order or the standard function ac-
cording as 𝑣 is nonarchimedean or archimedean, so Φ(0) = 1.
522 CHAPTER 29. IDELIC ZETA FUNCTIONS
In the number field case, we have Φ∨𝑣 = Φ 𝑣 self-dual when 𝑣 is archimedean, and
by (29.7.20) we have
Φ∨
𝑍 𝐵𝑣𝑣 (𝑠) = D (𝐵 𝑣 ) 1/2−𝑠 𝑍 𝐵Φ𝑣 (𝑠)
and Φ∨𝑣 (0) = D (𝐵 𝑣 ) −1/2 when 𝑣 is nonarchimedean. Multiplying these together and
applying Theorem 29.10.1(c) gives
Φ Φ∨ Φ
𝑍 𝐵 (1 − 𝑠) = 𝑍 𝐵 (𝑠) = D (𝐵) 1/2−𝑠 𝑍 𝐵 (𝑠).
In the function field case, a similar argument holds but with the character modified
by a global differential as in 29.8.9: for the relevant additional factor, see Exercise
29.13.
To conclude, we note that when 𝐵 ' M2 (𝐹) then by Lemma 29.8.24 the nonar-
chimedean part of 𝑍 𝐵 (𝑠) is given by 𝜁 𝐹 (2𝑠)𝜁 𝐹 (2𝑠 − 1), and this has a double pole at
𝑠 = 1/2 (or accordingly 𝑞 𝑠−1/2 = 1).
Remark 29.10.24. In her 1929 Ph.D. thesis, Hey [Hey29] defined the zeta function of a
division algebra over Q, proving that it has an Euler product and functional equation—
this was a tour de force in algebraic and analytic number theory, especially at the time!
For more on Hey’s contribution and its role in the development of class field theory,
see the perspective by Roquette [Roq2006, §9], including Zorn’s observation that the
functional equation yields an analytic proof of the classification of division algebras.
Hey’s thesis was never published (though the content appears in Deuring [Deu68, VII,
§8]), and classical treatments of the zeta function for the most part gave way to the
development of Chevalley’s adeles and ideles.
Tate’s thesis [Tate67] (from 1950, published in 1967) is usually given as the
standard reference for the adelic recasting of zeta and 𝐿-functions of global fields as
above: it gives the general definition of a zeta function associated to a local field, an
integrable function, and a quasi-character [Tate67, §2]; see also the Bourbaki seminar
by Weil [Weil66]. But already in 1946, Matchett (also a student of Emil Artin) wrote
a Ph.D. thesis at Indiana University [Mat46] beginning the redevelopment of Hecke’s
theory of zeta and 𝐿-functions in terms of adeles and ideles. At the time, Iwasawa also
contributed to the development of the theory; see his more recently published letter to
Dieudonné [Iwa92].
These results were generalized to central simple algebras over local fields by
Godement [God58a, God58b], Fujisaki [Fuj58], and Weil [Weil82, Chapter III] (and
again Weil [Weil74, Chapter XI]) in the same style as Iwasawa and Tate; and then they
were further generalized (allowing representations) by Godement–Jacquet [GJ72],
providing motivation for the Langlands program.
As an example of the generalizations indicated by Remark 29.10.24, we conclude
with a slightly more general statement.
Theorem 29.10.25. Let 𝐹 be a global field, let 𝐵 be a central division algebra over 𝐹,
let Φ be a Schwartz–Bruhat function on 𝐵, and let 𝜒 : 𝐵× → C be a unitary character
such that 𝜒 restricted to 𝐹 × is trivial. Define
∫
Φ
𝐿 𝐵 (𝑠, 𝜒) := Φ(𝛼) 𝜒(𝛼) k𝛼k 𝑠 d𝜏 × (𝛼).
𝐵×
29.11. TAMAGAWA NUMBERS 523
Φ
Then the function 𝐿 𝐵 (𝑠, 𝜒) is absolutely convergent for Re 𝑠 > 1. If 𝜒 is nontrivial,
Φ
then 𝐿 𝐵 (𝑠, 𝜒) has holomorphic continuation to C and satisfies the functional equation
Φ Φ∨
𝐿 𝐵 (𝑠, 𝜒) = 𝐿 𝐵 (𝑠, 𝜒−1 ).
Proof. Proven in the same way as Main Theorem 29.10.1, just keeping track of the
character 𝜒. The (possible) residues at 𝑠 = 0, 1 (or more generally in the function field
case 𝑞 𝑠 = 𝑞 0 , 𝑞 1 ) are multiplied by
∫
𝜒−1 (𝛼) d𝜏 (1) (𝛼)
𝐵 (1) /𝐹 ×
which is zero when 𝜒 is nontrivial by character theory and therefore the 𝐿-function is
fact holomorphic at these points.
Lemma 29.11.1. Let Ω be the set of real ramified places in 𝐵, and let
Proof. The surjectivity of the reduced norm is locally established, and its kernel is
𝐵1 by definition. The image of 𝐵× under the reduced norm is 𝐹>×Ω 0 by the Hasse–
Schilling theorem of norms (Main Theorem 14.7.4), with kernel 𝐵1 . Moreover, the
natural inclusion
𝐹>×Ω 0 \𝐹>(1)
Ω0
↩→ 𝐹 × \𝐹 (1)
is also surjective by weak approximation. Putting these together, we obtain (29.11.2),
which we might be tempted to summarize in the exact sequence of pointed sets
nrd
1 → 𝐵1 \𝐵1 → 𝐵× \𝐵 (1) −−→ 𝐹 × \𝐹 (1) → 1
and
𝜏 1 (𝐵1 \𝐵1 ) = 1.
Proof. The equality 𝜏 (1) (𝐹 × \𝐹 (1) ) = 𝜁 𝐹∗ (1) is the statement of Main Theorem
29.10.1(c) applied to 𝐵 = 𝐹. Then 𝐵 is a division algebra, the equality 𝜏 (1) (𝐵× \𝐵 (1) ) =
𝜁 𝐹∗ (1) is again Main Theorem 29.10.1(c), and the equality 𝜏 1 (𝐵1 \𝐵1 ) = 1 then follows
from (29.11.2).
To conclude, we make the appropriate modifications in the remaining case, and
suppose that 𝐵 = M2 (𝐹). Then 𝐵1 = SL2 (𝐹) and similarly 𝐵1 = SL2 (𝐹). By the exact
sequence (29.11.2), we may show 𝜏 1 (SL2 (𝐹)\ SL2 (𝐹)) = 𝜏 1 (SL2 (𝐹)/SL2 (𝐹)) = 1.
We will do Fourier analysis on 𝐹 2 , extended from 𝐹, with self-dual measure 𝜏 and
character 𝜓. Let Φ be a Schwartz function on 𝐹 2 . The Fourier transform is
∫
∨
Φ (𝑦) = Φ(𝑥)𝜓(𝑦 t · 𝑥) d𝑥
𝐹2
The group SL2 (𝐹) acts on column vectors 𝐹 2 r {0} by left multiplication, with the
stabilizer of 𝐹 2 r {0} given by SL2 (𝐹). Thus
∫ ∫ !
∑︁
∨
Φ (0) = Φ(𝑥) d𝜏(𝑥) = Φ(𝛼𝑥) − Φ(0) d𝜏 1 (𝛼). (29.11.5)
𝐹2 SL2 (𝐹 )/SL2 (𝐹 )
𝑥 ∈𝐹 2
∫ ∫
© ∑︁ ∨
Φ(𝑥) d𝜏(𝑥) = Φ ((𝛼t ) −1 𝑦) − Φ(0) ® d𝜏 1 (𝛼). (29.11.7)
ª
𝐹2 SL2 (𝐹 )/SL2 (𝐹 )
« 𝑦 ∈𝐹 2 ¬
Replacing Φ ← Φ∨ and 𝛼 ← (𝛼t ) −1 (preserving the measure) in (29.11.7) gives
∫ ∫
∨ © ∑︁
Φ(0) = Φ (𝑥) d𝜏(𝑥) = Φ(𝛼𝑦) − Φ∨ (0) ® d𝜏 1 (𝛼).
ª
𝐹2 SL2 (𝐹 )/SL2 (𝐹 )
« 𝑦 ∈𝐹 2 ¬
(29.11.8)
29.11. TAMAGAWA NUMBERS 525
𝜏(G) := vol(G(𝐹)/G(𝐹)).
For example, in the above we computed the volume for the group G associated to the
group 𝐵1 of quaternions of reduced norm 1.
In the late 1950s, Tamagawa defined the Tamagawa measure [Tam66]. Weil
[Weil82] (based on notes from lectures at Princeton 1959–1960), computed that 𝜏(𝐺) =
1 for 𝐺 a classical semisimple and simply connected group; the conjecture that this
holds in general was known as Weil’s conjecture on Tamagawa numbers. This difficult
conjecture was proven by the efforts of many people: see Scharlau [Scha2009, §2] for
a history. In particular, the calculation of the Tamagawa number of the orthogonal
group of a quadratic form recovers the Smith–Minkowski–Siegel mass formula that
computes the mass of a genus of lattice.
Exercises
1. Let 𝜇 𝑝 be the standard Haar measure on Q 𝑝 .
(a) Let 𝐷 (𝑎, 𝛿) := {𝑥 ∈ Q 𝑝 : |𝑥 − 𝑎| < 𝛿} be the open ball of radius 𝛿 ∈ R>0
around 𝑎 ∈ Q 𝑝 . Compute the measure 𝜇 𝑝 (𝐷 (𝑎, 𝛿)), and repeat with the
closed ball of radius 𝛿 ∈ R ≥0 .
(b) Show that ∫
log 𝑝
log(|𝑥| 𝑝 ) d𝜇 𝑝 (𝑥) = − .
Z𝑝 𝑝−1
2. Let 𝐹 := F𝑞 (𝑇); then 𝐹 is the function field of P1 , a curve of genus 𝑔 = 0 and
1
𝜁 𝐹 (𝑠) = .
(1 − 𝑞 −𝑠 ) (1 − 𝑞 1−𝑠 )
where the map 𝜙 is given by either the quotient by ±1 (equivalently, the absolute
value) or by the map 𝑥 ↦→ 𝑥 2 . Equip R× and R×>0 with the standard Haar measure
d𝑥/|𝑥|. Compute the unique compatible measures on {±1} for the two choices
of 𝜙 and show they differ by a factor 2.
8. Finish the proof of Lemma 29.5.18. [Hint: It may help to use the Iwasawa
decomposition (Proposition 33.4.2 and Lemma 36.2.7).]
I 9. Prove Proposition 29.6.9, as follows. Let 𝐹 be a nonarchimedean local field
and let 𝜓 : 𝐹 → C1 be a nontrivial unitary character of 𝐹, for example the
standard unitary character. For each 𝑥 ∈ 𝐹, let 𝜓 𝑥 : 𝐹 → C1 be defined by
𝜓 𝑥 (𝑦) = 𝜓(𝑥𝑦).
(a) Show that 𝜓 𝑥 is again a unitary character of 𝐹 and that 𝜓 𝑥 is trivial if and
only if 𝑥 = 0.
(b) Show that
Ψ : 𝐹 → 𝐹∨
𝑥 ↦→ 𝜓 𝑥
defines a continuous, injective group homomorphism. [Hint: recall that
𝐹 ∨ inherits a compact-open topology, so a basis of neighborhoods of the
identity is given by { 𝑓 ∈ 𝐹 ∨ : 𝑓 (𝐾) ⊆ 𝑉 } for 𝐾 ⊆ 𝐹 compact and 𝑉 ⊆ 𝐹 ∨
an open neighborhood of 1. Given such 𝐾, 𝑉, show that there exists an
open neighborhood 𝑈 ⊆ 𝐹 of 0 such that 𝑥𝐾 ⊆ 𝜓 −1𝑉 for all 𝑥 ∈ 𝑈.]
(c) Show that Ψ(𝐹) is dense in 𝐹 ∨ . [Hint: if 𝜓 𝑥 (𝑦) = 0 for all 𝑥 ∈ 𝐹, then
𝑦 = 0.]
(d) Prove that Ψ−1 is continuous (on its image). [Hint: work in the other
direction in (b).]
29.11. TAMAGAWA NUMBERS 527
Optimal embeddings
𝑟 4 ( 𝑝) = 8( 𝑝 + 1)
and we upgraded this in Exercise 26.7 to a general formula for 𝑟 4 (𝑛) in terms of the
sum of (odd) divisors of 𝑛.
529
530 CHAPTER 30. OPTIMAL EMBEDDINGS
We may similarly ask for a formula for 𝑟 3 , but it is more difficult both to state and
to prove. Define
prim
𝑟3 (𝑛) = #{(𝑥, 𝑦, 𝑧) ∈ Z3 : 𝑥 2 + 𝑦 2 + 𝑧2 = 𝑛 and gcd(𝑥, 𝑦, 𝑧) = 1} (30.1.2)
Let ℎ(𝑑) = # Pic 𝑆 𝑑 be the class number of the quadratic order of discriminant 𝑑 < 0;
equivalently, ℎ(𝑑) is the number of reduced primitive integral positive definite binary
quadratic forms of discriminant 𝑑.
prim
Gauss [Gau86, Section 291] showed that 𝑟 3 (𝑛) is a constant multiple of ℎ(−4𝑛)
as follows.
Theorem 30.1.3 (Gauss). We have 𝑟 3 (1) = 6, 𝑟 3 (3) = 8, and for 𝑛 ∈ Z ≥0 :
0, if 𝑛 ≡ 0, 4, 7 (mod 8);
prim
𝑟 3 (𝑛) = 12ℎ(−4𝑛), if 𝑛 ≡ 1, 2, 5, 6 (mod 8) and 𝑛 ≠ 1;
8ℎ(−4𝑛) = 24ℎ(−𝑛), if 𝑛 ≡ 3 (mod 8) and 𝑛 ≠ 3.
(One can√ uniformly include the cases 𝑛 = 1, 3 by accounting for the extra roots of
unity in Q( −𝑛).)
Theorem 30.1.3 is a special case of Theorem 30.4.7—see Exercise 30.4—but for
historical and motivational reasons, we also give in the next section an essentially
self-contained proof for the case 𝑛 ≡ 1, 2 (mod 4), following Venkov [Ven22, Ven29].
The main obstacle to generalizing Gauss’s theorem (Theorem 30.1.3) is that quater-
nion orders need not have class number 1: a generalization with this hypothesis
“following the general plan described by Venkov” is given by Shemanske [Shem86].
Another annoyance is the growing technicality of the local computations giving the
explicit constants involved. Both of these issues are in some sense resolved by em-
ploying idelic methods (hence the placement of this chapter in this text) and even the
proof of Gauss’s theorem itself is simplified by these methods (in the next section).
The result is Theorem 30.4.7: representations are spread across the genus of an order,
with the constants given by local factors (computed in this chapter for maximal orders
and then Eichler orders).
30.1.4. For indefinite quaternion orders, strong approximation applies, and we are
almost always able to prove that the contribution to each order in the genus is equal,
with one quite subtle issue known as selectivity: in certain rare circumstances, a
quadratic order embeds in precisely half of the orders in a genus. We pursue selectivity
in the next chapter (Chapter 31): technical and rather extraordinary, it is a subject that
demands care.
Happily, a locally norm-maximal order (such as an Eichler order) over Z is not
selective!
30.2. SUMS OF THREE SQUARES 531
The latter “error term” is accounted for by (finite) cyclic subgroups spread across
representative left orders OL (𝐽)—and this is precisely the contribution computed
idelically above!
:=
√ of Theorem 30.1.3
Proof √ for 𝑛 ≡ 1, 2 (mod 4). Suppose 𝑛 ≡ 1, 2 (mod 4). Let 𝑆
Z[ −𝑛] ⊂ 𝐾 := Q( −𝑛). Then 𝑆 is maximal
√ and ramified at 2, i.e., 𝑆 ⊗ Z2 is the ring
of integers of the field 𝐾 ⊗ Q2 = Q2 ( −𝑛).
Let 𝐵 = (−1, −1 | Q) be the rational Hamiltonians and O ⊂ 𝐵 the Hurwitz order.
We consider the set
𝐾 ↩→ 𝐵
√
−𝑛 ↦→ 𝛼
and for convenience identify 𝐾 with its image. By the Skolem–Noether theorem
(Corollary 7.1.5), every other element 𝛼 0 ∈ 𝑊 is the form 𝛼 0 = 𝛽−1 𝛼𝛽 with 𝛽 ∈ 𝐵×
(but not necessarily conversely!). Let
𝐸 := {𝛽 ∈ 𝐵× : 𝛽−1 𝛼𝛽 ∈ 𝑊 }. (30.2.2)
The set 𝐸 has a right action of the normalizer 𝑁 𝐵× (O) (checking that primitivity is
preserved).
We relate 𝐸 to the group of fractional ideals Idl 𝑆 as follows. Let 𝔟 ⊆ 𝐾 be a
fractional 𝑆-ideal. Since O is (right) Euclidean, 𝔟O = 𝛽O for some 𝛽 ∈ 𝐵× that is
well-defined up to right multiplication by O× . The heart of the proof is the following
claim: the map
Idl 𝑆 → 𝐸/𝑁 𝐵× (O)
(30.2.3)
𝔟 ↦→ 𝛽𝑁 𝐵× (O)
is a well-defined, surjective map of sets. The most efficient (and clear) proof of this
claim is idelic, and we prove it in two steps.
First, the map (30.2.3) is well-defined: that is to say, 𝛽 ∈ 𝐸. Write 𝔟𝑆b = b b so
𝑏 𝑆,
that 𝛽O
b=b 𝑏Ob and 𝛽 = b
𝑏b𝜇 with 𝜇b∈ O ×
b . Then b𝑏 commutes with 𝛼 (in 𝐾b ⊆ 𝐵),
b so
𝛼 0 = 𝛽−1 𝛼𝛽 = 𝜇
b−1b
𝑏 −1 𝛼b b−1 𝛼b
𝜇=𝜇
𝑏b b−1 Ob
𝜇∈𝜇 b−1 Ob
𝜇⊆𝜇 b𝜇 ∩ 𝐵 = O
b ∩ 𝐵 = O.
To show that the map is surjective, we need to establish one other important point
𝜈 ∈ 𝑁 𝐵b× ( O)
comparing the global and the idelic: we claim that there exists b b such that
𝜈 −1 𝛼b
b 𝜈 = 𝛽−1 𝛼𝛽. (30.2.4)
b× 𝑁 b× ( O).
𝛽∈𝐾 b
𝐵
30.3. OPTIMAL EMBEDDINGS 533
b× 𝑁 b× ( O)
𝐾 b× O
b =𝐾 b× ,
𝐵
Importantly, Venkov’s proof of Gauss’s theorem given above is explicit and con-
structive, given at least one representation as a sum of three squares.
The early observation made in the proof above is that the sum of three squares is
the restriction of the reduced norm to the trace zero elements of the Hurwitz order.
One then seeks a similar statement for quadratic forms 𝑄 = nrd | O0 obtained more
generally. (This is almost the same thing as the ternary quadratic form associated to
O itself via the Clifford algebra construction in Chapter 22; the difference is that for
the latter we take the dual of the order and scale, as in (22.1.3).)
Just as in the proof above, for a representation 𝑄(𝑥,√ 𝑦, 𝑧) = 𝑛 corresponding to
𝛼 ∈ O with 𝛼2 +𝑛 = 0, we obtain an embedding 𝑆 = Z[ −𝑛] ↩→ O of a quadratic order
into the quaternion order; conversely, to such an embedding we find a representation. It
is more convenient to work with embeddings, as they possess more structure. Viewed
in this way, we may equivalently restrict the reduced norm from O to the order 𝑆
itself, and then we are asking for the representation of a binary quadratic form by a
quaternary quadratic form.
30.3.1. The set of embeddings of 𝐾 in 𝐵 is identified with the set 𝐾 × \𝐵× , by 7.7.12:
if 𝜙 : 𝐾 ↩→ 𝐵 is another embedding, then by the Skolem–Noether theorem there
exists 𝛽 ∈ 𝐵× such that 𝜙(𝛼) = 𝛽−1 𝛼𝛽 for all 𝛼 ∈ 𝐾 with 𝛽 well-defined up to left
multiplication by 𝐾 × , the centralizer of 𝐾 under conjugation by 𝐵× .
We now turn to the integral theory. Let O ⊆ 𝐵 be a quaternion 𝑅-order, and let
𝑆 ⊆ 𝐾 be a quadratic 𝑅-order; we will be interested in embeddings 𝜙 : 𝑆 ↩→ O. Such
an embedding gives an embedding 𝜙 : 𝐾 ↩→ 𝐵 by extending scalars. We keep the
embeddings for various suborders organized as follows.
534 CHAPTER 30. OPTIMAL EMBEDDINGS
𝜙(𝐾) ∩ O = 𝜙(𝑆).
Let
Emb𝑅 (𝑆, O) := {Optimal embeddings 𝑆 ↩→ O}. (30.3.3)
When no confusion can result, we drop the subscript 𝑅 and write simply Emb(𝑆, O).
Proof. Immediate from the local-global dictionary for lattices (Theorem 9.4.9).
Lemma 30.3.6 says that for an embedding 𝜙, the property of being optimal is a
local property.
We define
𝐸 = 𝐸 𝑆,O := {𝛽 ∈ 𝐵× : 𝛽−1 𝐾 𝛽 ∩ O = 𝛽−1 𝑆𝛽}
(30.3.7)
= {𝛽 ∈ 𝐵× : 𝐾 ∩ 𝛽O𝛽−1 = 𝑆}.
(The set 𝐸 also depends on the fixed embedding 𝐾 ↩→ 𝐵, but this ‘reference’ embed-
ding will remain fixed throughout.) In equation 30.3.7, we see two different ways to
think about embeddings: we either move 𝑆 and see how it fits into O, or we fix 𝐾 and
move O.
30.3.10. Let O1 ≤ Γ ≤ 𝑁 𝐵× (O). Then the image of Γ in 𝑁 𝐵× (O)/𝐹 × has finite index.
For example, we may take Γ = O× . (The scalars do not play a role in the theory of
embeddings: they act by the identity under conjugation.)
For 𝛾 ∈ Γ, and an optimal embedding 𝜙 ∈ Emb(𝑆, O), we obtain a new embedding
via 𝛼 ↦→ 𝛾 −1 𝜙(𝛼)𝛾, i.e., Γ acts on the right on Emb(𝑆, O) by conjugation, and
correspondingly on the right on 𝐸 by right multiplication.
Let
and
𝑚(𝑆, O; Γ) := # Emb(𝑆, O; Γ). (30.3.12)
The quantity 𝑚(𝑆, O; Γ) only depends on the type (isomorphism class) of O (trans-
porting Γ under the isomorphism of orders, of course).
By Lemma 30.3.8, there is a bijection
∼ 𝐾 × \𝐸/Γ.
Emb(𝑆, O; Γ) −→ (30.3.13)
We conclude this section by a comparison: for groups Γ sitting between unit groups
and norm 1 unit groups, we can compare embedding numbers as follows.
Emb(𝑆, O; Γ) → Emb(𝑆, O; O× )
and the lemma amounts to looking at the fibers. From (30.3.13), we turn instead to
𝐾 × \𝐸/Γ → 𝐾 × \𝐸/O× .
Remark 30.3.17. The term optimal goes back at least to Schilling [Schi35], but the
notion was studied in the context of maximal orders as well by Chevalley [Chev34],
Hasse [Hass34], and Noether [Noe34]. The theory of optimal embeddings was devel-
oped thereupon by Eichler [Eic56a, §3]; many key ideas can be seen transparently in
Eichler [Eic73, Chapter II, §§3–5]. For further history up to the present, see Remark
30.6.18.
536 CHAPTER 30. OPTIMAL EMBEDDINGS
30.4.1. By the local-global principle for embeddings (Proposition 14.6.7, and 14.6.8
for the case 𝐾 ' 𝐹 × 𝐹), there exists an 𝐹-algebra embedding 𝐾 ↩→ 𝐵 if and only if
there exist 𝐹𝑣 -algebra embeddings 𝐾 𝑣 ↩→ 𝐵 𝑣 for all 𝑣 ∈ Pl 𝐹 if and only if 𝐾 𝑣 is a
field for all 𝑣 ∈ Ram 𝐵.
30.4.2. The definitions in the previous section extend to each completion. Let 𝐾
b=
𝐾 ⊗𝐹 𝐹 and similarly 𝑆 = 𝑆 ⊗𝑅 𝑅. Let
b b b
Γ = (Γ 𝑣 ) 𝑣 ≤ 𝑁 𝐵b× O
b b
b O;
Emb𝑅b ( 𝑆, bb Γ-conjugacy classes of optimal 𝑆b ↩→ O}
Γ) := {b b (30.4.3)
and
b O;
𝑚( 𝑆, bb b O;
Γ) := # Emb𝑅b ( 𝑆, bbΓ). (30.4.4)
We now show that the number of global embeddings is counted by a class number
times the number of local embeddings. In view of Lemma 30.3.14, we may focus on
Γ=O
the case b b× . As usual, we write Cls O for the right class set of O.
𝐾 × \𝐸/
bO b× → 𝐾
b× \𝐸/
bO b× (30.4.9)
which is a surjective map of pointed sets. The fiber of (30.4.9) over the identity
element is
𝐾 × \𝐾
b× /𝑆b× ' Pic 𝑆. (30.4.10)
We claim that the fibers of (30.4.9) may be similarly identified. Indeed, the fiber over
b× 𝛽bO
𝐾 b× consists of the double cosets 𝐾 ×b b× with 𝐾 ×b
𝜈 𝛽bO 𝜈 ∈ 𝐾 × \𝐾
b× , and
𝐾 ×b b× = 𝐾 × 𝛽bO
𝜈 𝛽bO b×
if and only if
𝜈 𝛽b = 𝜌 𝛽b
b b𝜇
𝜌 −1b b𝜇 𝛽b−1 ∈ 𝐾
𝜈 = 𝛽b b× ∩ 𝛽bO
b× 𝛽b−1 = 𝑆b× (30.4.11)
b if and only if 𝐾 ×b
(since 𝛽b ∈ 𝐸), 𝜈 ⊆ 𝐾 × 𝑆b× , as claimed. From the claim and (30.4.6),
we conclude that
#(𝐾 × \𝐸/
bO b× ) = ℎ(𝑆)𝑚( 𝑆, b O;
b Ob× ). (30.4.12)
b× ∈ 𝐸/
On the other hand, each 𝛽bO bO b× defines a right O-ideal;
b intersecting with 𝐵
and organizing these right ideals by their classes, we will now show that they give rise
to optimal embeddings of the corresponding left order. For brevity, write O𝐼 = OL (𝐼).
There is a map of pointed sets
𝐾 × \𝐸/
bO b× → 𝐵× \ 𝐵
b× /O ∼ Cls O.
b× −→ (30.4.13)
Choose representatives
Ä
𝐵× \ 𝐵
b× /O
b× = 𝐵× b b×
𝛼𝐼 O (30.4.14)
[𝐼 ] ∈Cls O
𝛼𝐼 O
so that 𝐼 = b b ∩ 𝐵. Then
O𝐼 = O 𝐿 (𝐼) = b b𝛼−1 ∩ 𝐵.
𝛼𝐼 Ob𝐼
Let
𝐸 𝐼 = {𝛽 ∈ 𝐵× : 𝐾 ∩ 𝛽O𝐼 𝛽−1 = 𝑆}. (30.4.15)
Now if
b× ∈ 𝐸/
𝛽bO bO b× ,
𝐵× 𝛽bO
b× ⊆ 𝐵× b b× ,
𝛼𝐼 O
538 CHAPTER 30. OPTIMAL EMBEDDINGS
and therefore
b× = (𝛽b
𝛽bO b×
𝛼𝐼 ) O
for some 𝛽 ∈ 𝐵× . If 𝛽 0 ∈ 𝐵× is another, then
𝛽−1 𝛽 0 ∈ b b× b
𝛼𝐼 O 𝛼−1 ×
𝐼 ∩ 𝐵 = O𝐼
𝐾 ∩ 𝛽O𝐼 𝛽−1 = 𝐾 ∩ ( 𝛽b
b𝛼−1 b𝛼−1 ) (b
𝛼𝐼 Ob
𝐼 ) (b 𝐼 𝛼𝐼 𝛽b−1 )
b 𝛽b−1
= 𝐾 ∩ 𝛽bO
∼ ∼
Ä Ä
𝐾 × \𝐸/
bO b× −→ 𝐾 × \𝐸 𝐼 /O×𝐼 −→ Emb(𝑆, O𝐼 ; O×𝐼 )
𝐼 𝐼 (30.4.16)
× bb× × ×
𝐾 𝛽O ↦→ 𝐾 𝛽O𝐼 .
Putting (30.4.12) and (30.4.16) together and counting, the theorem follows.
When O has class number 1, we hit the embedding number on the nose.
𝑚(𝑆, O; O× ) = ℎ(𝑆)𝑚( 𝑆,
b O; b× ).
b O
We recall that the genus of O (Definition 17.4.8) is the set Gen O of 𝑅-orders
O0 ⊆ 𝐵 locally isomorphic to O.
Corollary 30.4.18. If Emb(𝑆𝔭 , O𝔭 ) ≠ ∅ for all primes 𝔭, then there exists an order
O0 ∈ Gen O such that Emb(𝑆, O0) ≠ ∅.
Remark 30.4.19. The possible failure of local optimal embeddings to ‘glue’ to a global
optimal embedding into O is measured by the phenomenon of selectivity, studied in
Chapter 31.
be index of the subgroup principal two-sided fractional O-ideals inside the invertible
ones, studied in section 18.5. By Proposition 18.5.10, the equation (30.4.8) then
becomes
∑︁
ℎ(O0)𝑚(𝑆, O0; O0× ) = ℎ(𝑆)𝑚( 𝑆,b O; b× ).
b O (30.4.21)
[O0 ] ∈Typ O
Remark 30.4.22. The foundational formula (30.4.8) is proven by counting a set two
different ways. As such, it admits a purely combinatorial proof: Brzezinski [Brz89]
shows that it follows from looking at “transitive actions of groups on pairs of sets and
on relations invariant with respect to these actions” [Brz89, p. 199].
We also record the following corollary (in the vein of Remark 30.4.22).
𝐾 × \𝐾
b× /𝑆b× 𝐾 × \𝐸/
bO b×
Proof. In the proof of the trace formula, we identified 𝐾 × \𝐾 b× /𝑆b× ' Pic 𝑆 and estab-
lished an idelic to global bijection in (30.4.16); in (30.4.11), each fiber of the map
(30.4.9) is identified with Pic 𝑆, describing the orbits of the action. (See also Exercise
30.12, where a direct proof is requested.)
30.5.1. Since we are now local, the 𝑅-order 𝑆 is free and so 𝑆 = 𝑅[𝛾] for some (not
unique) 𝛾 ∈ 𝑆. Let 𝑓 𝛾 (𝑥) := 𝑥 2 − 𝑡 𝛾 𝑥 + 𝑛 𝛾 be the minimal polynomial of 𝛾, and let
𝑑 𝛾 := 𝑡 𝛾2 − 4𝑛 𝛾 be the discriminant of 𝑓 𝛾 , equal to the discriminant of 𝑆 in 𝑅/𝑅 ×2 . An
𝑅-algebra embedding from 𝑆 is then determined uniquely by the image of 𝛾.
540 CHAPTER 30. OPTIMAL EMBEDDINGS
1, if 𝐾 ' 𝐹 × 𝐹 is split;
𝐾
:= 0, if 𝐾 ⊇ 𝐹 is a ramified field extension;
𝔭
−1
if 𝐾 ⊇ 𝐹 is an unramified field extension.
𝑚(𝑆, O; 𝑁 𝐵× (O)) = 1
× 𝐾
𝑚(𝑆, O; O ) = 1 − .
𝔭
We recall that 𝑁GL2 (𝐹 ) (M2 (𝑅)) = 𝐹 × GL2 (𝑅), so (a) also includes the case of the
normalizer.
Proof. First, part (a). We have at least one embedding 𝜙 : 𝑆 ↩→ End𝑅 (𝑆) ' M2 (𝑅)
given by the regular representation of 𝑆 on itself (in a basis): in the basis 1, 𝛾, we have
0 −𝑛 𝛾
𝛾 ↦→ (30.5.4)
1 𝑡𝛾
a matrix in rational canonical form. This embedding is optimal, because if 𝑥, 𝑦 ∈ 𝐹
satisfy
𝑥 −𝑛 𝛾 𝑦
𝜙(𝑥 + 𝑦𝛾) = ∈ M2 (𝑅)
𝑦 𝑥 + 𝑡𝛾 𝑦
then 𝑥, 𝑦 ∈ 𝑅 already, so 𝜙(𝐾) ∩ M2 (𝑅) = 𝜙(𝑆).
To finish (a), we need to show that the embedding (30.5.4) is the unique one, up to
conjugation by GL2 (𝑅). So let 𝜓 : 𝑆 ↩→ M2 (𝑅) be another optimal embedding. Then
via 𝜓, the 𝑅-module 𝑀 = 𝑅 2 (column vectors) has the structure of a left 𝑆-module; the
condition that 𝜓 is optimal translates into the condition that the (left) multiplicator ring
30.5. LOCAL EMBEDDING NUMBERS: MAXIMAL ORDERS 541
There are only finitely many superorders 𝑆 0 in the sum, since the integral closure 𝑅𝐾
contains all 𝑆 0 so [𝑅𝐾 : 𝑆 0] 𝑅 | [𝑅𝐾 : 𝑆] 𝑅 —or equally well, compare discriminants.
Applying Proposition 30.5.3, we conclude 𝑚(𝑆, O; O× ) < ∞.
542 CHAPTER 30. OPTIMAL EMBEDDINGS
𝜙 𝜛 : 𝑆 ↩→ O
𝜙 𝜛 (𝛼) = 𝜛 −1 𝜙(𝛼)𝜛.
and
𝑎 1 𝑎 1
𝜈 −1 𝜙(𝛾)𝜈 = = (30.6.5)
𝑏𝑐𝜋 𝑒 𝑑 − 𝑓 𝛾 (𝑎) 𝑡𝛾 − 𝑎
30.6. ∗ LOCAL EMBEDDING NUMBERS: EICHLER ORDERS 543
as desired.
0 1
If 𝑐 ∈ 𝑅× , then we take 𝜈 = 𝜛 = 𝑒 ; now
𝜋 0
𝑑 𝑐
𝜈 −1 𝜙(𝛾)𝜈 = (30.6.6)
𝑏𝜋 𝑒 𝑎
and we apply the previous case.
Finally, if 𝑎 − 𝑑 ∈ 𝑅 × , we may suppose 𝑏 ∈ 𝔭 as well as 𝑐 ∈ 𝔭 if 𝑒 = 0, and then
1 1
we take 𝜈 = to find
0 1
𝑎 − 𝑐𝜋 𝑒 𝑎 − 𝑑 + 𝑏 − 𝑐𝜋 𝑒
𝜈 −1 𝜙(𝛾)𝜈 = (30.6.7)
𝑐𝜋 𝑒 𝑐𝜋 𝑒 + 𝑑
to reduce again to the first case.
1 0
if 𝑥 ≡ 𝑥 0 (mod 𝔭𝑒 ), then let 𝜇 = 0 ; we confirm that
𝑥 −𝑥 1
𝑥0
1
𝜇−1 𝜙(𝛾)𝜇 = = 𝜙 0 (𝛾).
− 𝑓 𝛾 (𝑥 0) 𝑡 𝛾 − 𝑥 0
In preparation for (b) and (c), we note that
𝑡 − 𝑥 − 𝑓 𝛾 (𝑥)/𝜋 𝑒
𝜙 𝜛 (𝛾) = 𝛾 𝑒 . (30.6.10)
𝜋 𝑥
544 CHAPTER 30. OPTIMAL EMBEDDINGS
We now prove (b). If 𝑒 = 0, the statement is true: 𝜛 ∈ O× and all embeddings are
conjugate. Suppose 𝑒 ≥ 1. Since 𝑓 𝛾 (𝑥) ≡ 0 (mod 𝔭𝑒 ) we have
Proof. By Lemma 30.6.3, the set Emb(𝑆, O; O× ) is represented by the set of normal-
ized embeddings and their conjugates under 𝜛. By Lemma 30.6.9(a), the normalized
embeddings according to 𝑥 ∈ 𝑀 (𝑠) are distinct; by (b)–(c), the remaining conjugate
embeddings are new when they lift to 𝑀 (𝑒 + 1).
√ √
Example 30.6.14. Let 𝑅 = Z2 and 𝑆 = Z2 [ −1], so 𝛾 = −1 and 𝑓 𝛾 (𝑥) = 𝑥 2 + 1. We
have (
{1 mod 2}, if 𝑠 = 1;
𝑀 (𝑠) =
∅, if 𝑠 ≥ 2.
Therefore by Proposition 30.6.12, if O is an Eichler order of level 2𝑒 over Z2 , then
(
√ 1, if 𝑒 ≤ 1;
𝑚(Z2 [ −1], O; O× ) = (30.6.15)
0, if 𝑒 ≥ 2.
30.6. ∗ LOCAL EMBEDDING NUMBERS: EICHLER ORDERS 545
(c) If 𝑒 = 𝑓 , then
( 𝑓 −1)/2 , if 𝑓 is odd;
𝑞
×
𝑚(𝑆, O; O ) = 𝑓 /2 + 𝑞 𝑓 /2−1 𝐾
𝑞
1+ , if 𝑓 is even.
𝔭
0, if 𝑓 is odd;
×
𝑚(𝑆, O; O ) = 𝑓 /2−1 (𝑞 + 1) 1 + 𝐾
𝑞
, if 𝑓 is even.
𝔭
Proof. Since the residue field 𝑘 has odd characteristic, we can complete the square
and without loss of generality we may suppose that trd(𝛾) = 0, and
On the other hand, when the residue field 𝑘 has even characteristic, the computa-
tions become even more involved!
Remark 30.6.18. Eichler studied optimal embeddings [Eic38b, §2] very early on,
computing the contribution of units (coming from embeddings of Z[𝜔] and Z[𝑖] in a
maximal order O) in the mass formula. He then [Eic56a, §3] studied more generally
optimal embeddings of quadratic orders into his Eichler orders of squarefree level.
Hijikata [Hij74, §2] studied optimal embeddings in the context of computing traces
of Hecke operators on Γ0 (𝑁) (general 𝑁), with embedding numbers given for certain
orders. See also Eichler’s treatment [Eic73, §3] in the context of the basis problem for
modular forms, as well as Pizer’s presentation [Piz76a, §3]. (See Remark 41.5.13 for
further detail.) Brzezinski [Brz91, Corollary 1.16] (a typo has it appear as Corollary
1.6) gives a general formula for Eichler orders (which is to say, a generalization of
Lemma 30.6.17 to include 𝑞 even)—the proof method is different than the method of
Hijikata above, and the answer is organized a bit differently than Lemma 30.6.17.
But these papers are just the beginning, and there is a cornucopia of further results.
Many of these are obtained in pursuit of progressively more general forms of the trace
formula (see for example the summary of results by Hashimoto [Hash77]) for Eichler
orders. Shimizu [Shz63, §§26–27] considered embedding numbers over totally real
fields in computing the contribution of elliptic elements to formulas for the dimension
of spaces of cusp forms and later for the trace formula [Shz65, §3]. The contributions
of elliptic elements over totally real fields was also pursued by Prestel [Pre68, §5] and
more generally for embeddings by Schneider [Schn75] (and quite explicitly for real
quadratic fields [Schn77]) and Vignéras [Vig76a, §4].
Pizer [Piz76b, §§3–5] considered optimal embeddings for residually split orders
(see 24.3.7) over Q: these were then applied to further cases of the basis problem for
modular forms [Piz76c, Piz80b]. Then Hijikata–Pizer–Shemanske [HPS89b, §§1–5]
developed in a uniform manner the optimal embedding theory for basic orders (they
30.7. GLOBAL EMBEDDING NUMBERS 547
called them special, cf. Remark 24.5.7): the application to the trace formula is then
contained in their monograph [HPS89a]. Brzezinski [Brz90, §3] also obtains recursive
formulas for optimal embedding numbers of a local Bass (equivalently, basic) order
(in characteristic not 2), using an effective description of the automorphism group of
the order.
30.7.1. Our global ring 𝑅 = 𝑅 (T) comes from an eligible set T ⊆ Pl 𝐹. (This set is
usually denoted S, but we do not want any confusion with the quadratic 𝑅-algebra
𝑆 ⊆ 𝐾.)
For all but finitely many places 𝑣 ∉ T, we have O 𝑣 ' M2 (𝑅 𝑣 ) maximal and Γ 𝑣 =
O×𝑣 . By Proposition 30.5.3(a), for such places 𝑣, we have # Emb𝑅𝑣 (𝑆 𝑣 , O 𝑣 ; O×𝑣 ) = 1.
Therefore the number 𝑚( 𝑆, b O;
bb b O;
Γ) = # Emb𝑅b ( 𝑆, bbΓ) of adelic embeddings is given
by the (well-defined, finite) product
Ö
b O;
𝑚( 𝑆, bbΓ) = 𝑚(𝑆 𝑣 , O 𝑣 ; Γ 𝑣 ) (30.7.2)
𝑣∉T
Proof. For all 𝔭 - 𝔑, we have O𝔭 ' M2 (𝑅𝔭 ), so the result follows by combining
Theorem 30.4.7 with 30.7.1.
Proof. As in the proof of Theorem 30.7.3, but with a further appeal to Proposition
30.5.3 to show that all local embeddings numbers are equal to 1.
with the local embedding numbers computed in Proposition 30.5.3(b) for 𝔭 | 𝔇 and
Lemma 30.6.16 for 𝔭 | 𝔐.
Suppose further that 𝐵 is T-indefinite and that # ClΩ 𝑅 = 1; then # Cls O = # ClΩ 𝑅
by Corollary 28.5.17 (an application of strong approximation), so
Ö Ö
𝐾
𝐾
𝑚(𝑆, O; O× ) = ℎ(𝑆) 1− 1+ . (30.7.6)
𝔭 𝔭
𝔭 |𝔇 𝔭 |𝔐
Absent further hypothesis, it is difficult to tease apart the term 𝑚(𝑆, O; O× ) from
the sum over left orders in Theorem 30.4.7. In the next chapter, we will show that the
hypothesis that 𝐵 is T-indefinite is almost enough.
2𝑤 𝑆
𝑚(𝑆, O; O× ) = #{𝜙(𝑆) ⊆ O : 𝜙 ∈ Emb(𝑆; O)} . (30.8.4)
𝑤O
Proof. We count off the group O× /𝑅 × by maximal cyclic subgroups, keeping track
of the trivial class. By Lemma 30.8.2, every nontrivial 𝛼𝑅 × ∈ O× /𝑅 × belongs to
a unique maximal cyclic subgroup of some order 𝑞 ≥ 2: such a subgroup is of the
form 𝜙(𝑆) × /𝑅 × with 𝜙(𝑆) ⊆ O an optimally embedded order, and has 𝑞 − 1 nontrivial
elements. Therefore
∑︁ ∑︁
𝑤O − 1 = (𝑞 − 1)#{𝜙(𝑆) ⊆ O : 𝜙 ∈ Emb(𝑆; O)}.
𝑞 ≥2 𝑆 ⊆𝐾𝑞
[𝑆 × :𝑅 × ]=𝑞
We now recall the Eichler mass formula (Main Theorem 26.1.5), giving an explicit
formula for the weighted class number
∑︁ 1
mass(Cls O) := ,
𝑤𝐼
[𝐼 ] ∈Cls O
where the inner sum is over all quadratic 𝑅-algebras 𝑆 ⊇ 𝑅 such that [𝑆 × : 𝑅 × ] = 𝑞 ∈
Z ≥2 , and ℎ(𝑆) = # Pic 𝑆.
Proof. We apply Proposition 30.8.5 to each order OL (𝐼) for [𝐼] ∈ Cls O and sum. We
obtain
∑︁ 1
1− = # Cls O − mass(Cls O)
𝑤𝐼
[𝐼 ] ∈Cls O
∑︁
1 ∑︁ 1 ∑︁
= 1− 𝑚(𝑆, OL (𝐼); OL (𝐼) × )
2 𝑞 ≥2 𝑞 × ×
[𝑆 :𝑅 ]=𝑞 [𝐼 ] ∈Cls O
∑︁
1 ∑︁ 1 b O;
b Ob× ),
= 1− ℎ(𝑆)𝑚( 𝑆,
2 𝑞 ≥2 𝑞 × ×
[𝑆 :𝑅 ]=𝑞
ℎ(𝑅𝐾 ) Ö
𝐾 1
ℎ(𝑆) = × : 𝑆 × ] N (𝔣)
[𝑅𝐾
1−
𝔭 N (𝔭)
(30.8.11)
𝔭 |𝔣
𝐾
where N is the absolute norm and is given (globally) as in 30.5.2:
𝔭
−1, if 𝔭 is inert in 𝐾;
𝐾
= 0, if 𝔭 is ramified in 𝐾;
𝔭
1,
if 𝔭 splits in 𝐾.
# Cls O ∑︁ 1
# Typ O = + 1− (30.9.1)
# Pic𝑅 O 0
𝑧 O0
[O ] ∈Typ O
where 𝑧O0 = [𝑁 𝐵× (O0) : 𝐹 × O0× ]. But now the structure of the normalizer groups
come into play, and one can give a type number formula similar to the class number
formula 30.8.6 in terms of certain embedding numbers at least for Eichler orders.
Unfortunately, even over Q, these formulas quickly get very complicated! To give a
sense of what can be proven, in this section we provide a type number formula in a
special but interesting case due originally to Deuring [Deu51], and we refer to Remark
30.9.12 for further reference.
Proposition 30.9.2 (Deuring). Let 𝐵 be a definite quaternion algebra over Q with
disc 𝐵 = 𝑝 prime and let O ⊂ 𝐵 be a maximal order. Then # Typ O = 1 for 𝑝 = 2, 3,
and for 𝑝 ≥ 5,
1 1
# Typ O = # Cls O + [ℎ(−𝑝)] + ℎ(−4𝑝)
2 4
where [ℎ(−𝑝)] = ℎ(−𝑝) when 𝑝 ≡ 3 (mod 4) and is 0 otherwise.
Proof. In light of (30.9.1), we begin by considering the Picard group Pic O (with
𝑅 = Z): by 18.4.8, we have an isomorphism Pic(O) ' Z/2Z generated by the unique
right ideal 𝐽 ⊆ O with nrd(𝐽) = 𝑝. The ideal 𝐽 is automatically two-sided and contains
all elements of reduced norm divisible by 𝑝 (see 13.3.7); and thus 𝐽 is principal if and
only if there exists an element 𝛼 ∈ O with nrd(𝛼) = 𝑝 if and only if 𝑧 O = 2.
Therefore, (30.9.1) reads
1 1 ∑︁
# Typ O = # Cls O + #{𝐽 0 ⊆ O0 principal right ideal : nrd(𝐽 0) = 𝑝}.
2 2
[O0 ] ∈Typ O
(30.9.3)
552 CHAPTER 30. OPTIMAL EMBEDDINGS
We now compute this sum in terms of embedding numbers. First, the map 𝛼 ↦→ 𝛼O
gives
#{𝐽 ⊆ O principal right ideal : nrd(𝐽) = 𝑝}
1 (30.9.4)
= #{𝛼 ∈ O : nrd(𝛼) = 𝑝}
2𝑤 O
where 𝑤 O = [O× : Z× ].
Next, we claim that if 𝛼 ∈ O has nrd(𝛼) = 𝑝, then trd(𝛼) = 0, i.e., 𝛼2 + 𝑝 = 0.
Indeed, if 𝑡 = trd(𝛼) then the field 𝐾 = Q(𝛼) has discriminant 𝑡 2 − 4𝑝 < 0 (since
√
𝐵 is definite) so |𝑡| < 2 𝑝. If 𝑡 ≠ 0, then the polynomial 𝑥 2 − 𝑡𝑥 + 𝑝 splits modulo
𝑝, so 𝑝 splits in 𝐾; but 𝐾 𝑝 ↩→ 𝐵 𝑝 and 𝐵 𝑝 is a division algebra, so 𝐾 𝑝 is a field, a
contradiction. Thus 𝑡 = 0.
With these results in hand, we can bring in the theory of embedding numbers. We
have
#{𝛼 ∈ O : nrd(𝛼) = 𝑝} = #{𝛼 ∈ O : 𝛼2 + 𝑝 = 0}
∑︁
= # Emb(𝑆, O). (30.9.5)
√
𝑆 ⊇Z[ − 𝑝]
Combining (30.9.4), (30.9.5), and (30.9.6), and plugging into (30.9.3), we have
1 1 ∑︁ ∑︁
# Typ O = # Cls O + 𝑚(𝑆, O0; O0× ). (30.9.7)
2 4 √
𝑆 ⊇Z[ − 𝑝] [O0 ] ∈Typ O
√
By (30.4.21) (rewriting Theorem 30.4.7), for 𝑆 ⊇ Z[ −𝑝] we have
∑︁
ℎ(O0)𝑚(𝑆, O0; O0× ) = ℎ(𝑆)𝑚( 𝑆,
b O;
b Ob× )
[O0 ] ∈Typ O
where ℎ(O0) = [Idl O0 : PIdl O0]; but ℎ(O0) = 1 whenever 𝑚(𝑆, O0; O0× ) ≠ 0 by the
first paragraph, so we may substitute into (30.9.7) to get
1 1 ∑︁
b× ).
# Typ O = # Cls O + b O;
ℎ(𝑆)𝑚( 𝑆, b O (30.9.8)
2 4 √
𝑆 ⊇Z[ − 𝑝]
The order O is maximal, so the adelic embedding number is the product of local em-
bedding numbers computed in Proposition 30.5.3: there is only a possible contribution
at 𝑝, since 𝑝 ≠ 2 the order 𝑆 is maximal, and 𝐾 is ramified so 𝑚(𝑆 𝑝 , O 𝑝 ; O×𝑝 ) = 1,
b O;
thus 𝑚( 𝑆, b O b× ) = 1.
√
Finally, the order Z[ −𝑝] of discriminant −4𝑝 is maximal whenever 𝑝 ≡ 1 (mod 4)
and the sum becomes simply ℎ(−4𝑝); when 𝑝 ≡ 3 (mod 4), this order is contained in
the maximal order of discriminant −𝑝, so the sum is ℎ(−𝑝) + ℎ(−4𝑝). The result is
proven.
30.9. TYPE NUMBER FORMULA 553
Remark 30.9.9. For an alternate direct proof of Proposition 30.9.2 working with elliptic
curves, see Cox [Cox89, Theorem 14.18].
30.9.10. The sum of class numbers in Proposition 30.9.2 can be rewritten uniformly
in terms of the ring of integers as follows:
1, if 𝑝 ≡ 1 (mod 4);
√
[ℎ(−𝑝)] + ℎ(−4𝑝) = # Cl Q( −𝑝) · 4, if 𝑝 ≡ 3 (mod 8);
2, if 𝑝 ≡ 7 (mod 8).
When 𝑝 ≡ 1 (mod 4), there is nothing to do. For 𝑝 ≡ 3 (mod 4), we have
(
3ℎ(−𝑝), if 𝑝 ≡ 3 (mod 8);
ℎ(−4𝑝) =
ℎ(−𝑝), if 𝑝 ≡ 7 (mod 8);
√
according as 2 is inert or split in 𝐾 = Q( −𝑝).
Remark 30.9.11. In section 42.1–42.2, we relate quaternion algebras to supersingular
elliptic curves; in this language, Proposition 30.9.2 gives rise to a formula for the
number of supersingular elliptic curves defined over F 𝑝 up to isomorphism.
Remark 30.9.12. Eichler [Eic56a, Satz 11] gave a type number formula for definite
hereditary orders over a totally real field; this formula has an error which was corrected
by Peters [Pet69, Satz 14, Satz 15] over fields of class number one and by Pizer [Piz73,
Theorem A] in general. Pizer [Piz76a, Theorem 26] gives a formula for the type
number for (general) Eichler orders over Q. Finally, Vignéras [Vig80a, Corollaire
V.2.6] gives a “structural” type number formula (without explicit evaluation of the
sum) for Eichler orders, and Körner [Kör87, Theorem 3] gives a general type number
formula. For a generalization to totally definite orders in central simple algebras of
prime index over global fields, see Brzezinski [Brz97].
Exercises
1. As in section 30.1, for 𝑘 ≥ 1 let
We gave formulas for 𝑟 3 (𝑛), 𝑟 4 (𝑛). For completeness, observe that 𝑟 1 (𝑛) = 2, 0
according as 𝑛 is a square or not, and give a formula for 𝑟 2 (𝑛) in terms of the
factorization of 𝑛 in the ring Z[𝑖].
I 2. Let 𝐵 be a quaternion algebra over Q, let O ⊂ 𝐵 be an order, let 𝐾 be a quadratic
field with an embedding 𝐾 ↩→ 𝐵 and suppose 𝑆 = 𝐾 ∩ O is the ring of integers
of 𝐾.
(a) Let 𝔟 ⊂ 𝐾 be an invertible fractional 𝑆-ideal. Show that 𝔟O∩𝐾 = 𝔟. [Hint:
since 1 ∈ O, we have 𝔟O ∩ 𝐾 ⊇ 𝔟. For the other inclusion, consider
𝑎2 + 𝑏2 + 𝑐2 + 𝑑 2 = 𝑛
𝑎+𝑏+𝑐+𝑑 =𝑚
Selectivity
In the previous chapter, we saw that (conjugacy classes of) embeddings of a quadratic
order into a quaternion algebra are naturally distributed over the genus of a quaternion
order; in applications, we want to compare the number of embeddings over orders in
a genus. Such a comparison can be thought of as a strong integral refinement of the
local-global principle for embeddings of quadratic fields (Proposition 14.6.7), which
belongs to the more general framework of the Albert–Brauer–Hasse–Noether theorem.
This chapter is quite technical, and it may be skipped on a first reading. To reward
the reader who persists, we conclude this chapter with the construction of isospectral,
nonisometric hyperbolic Riemannian manifolds, following Vignéras.
557
558 CHAPTER 31. SELECTIVITY
These two orders are not isomorphic and up to isomorphism represent the two types
of maximal 𝑅-orders in M2 (𝐹).
We claim that there is an embedding Z𝐾 ↩→ O but no embedding Z𝐾 ↩→ O0.
The first part of the claim is easy: taking the rational canonical form, we take the
embedding
0 √1
𝑤 ↦→ 𝛼 = . (31.1.2)
1 −5
The proof that Z𝐾 6↩→ O0 is more difficult. (The embedding (31.1.2) does not
extend to O0 because of the off-diagonal coefficients; and we cannot conjugate this
embedding in an obvious way because the ideal 𝔭 is not principal.) Such an embedding
would be specified by a matrix
𝑎 𝑏√ 𝑅 𝔭
𝛼0 = ∈ −1
𝑐 −𝑎 + −5 𝔭 𝑅
with √
− det(𝛼 0) = 𝑎 2 − −5𝑎 + 𝑏𝑐 = 1; (31.1.3)
so the content in the second claim is that there is no solution to the quadratic equation
(31.1.3). √
Indeed, suppose there is a solution. Let 𝑓 (𝑥) = 𝑥 2 − −5𝑥 − 1 ∈ Z𝐹 [𝑥], so that
𝑓 (𝑎) + 𝑏𝑐 = 0. We may factor 𝑏Z𝐹 = 𝔭𝔟 with 𝔟 ⊆ Z𝐹 and [𝔟] ∈ Cl Z𝐹 nontrivial;
by parity, there exists a prime 𝔮 | 𝔟 with [𝔮] nontrivial. Factoring 𝑐Z𝐹 = 𝔭−1 𝔠 with
𝔠 ⊆ Z𝐹 , we have 𝑏𝑐Z𝐹 = 𝔟𝔠 ⊆ 𝔮, so 𝑓 (𝑎) = −𝑏𝑐 ≡ 0 (mod 𝔮). But 𝑓 (𝑥) has trivial
discriminant, and modulo a prime 𝔮 ⊆ Z𝐹 it either splits (into distinct linear factors) or
remains irreducible. And by the Artin map, 𝑓 (𝑥) splits modulo 𝔮 if and only if 𝔮 splits
in 𝐾 if and only if the class [𝔮] ∈ Cl Z𝐹 is trivial. Putting these two pieces together,
we have 𝑓 (𝑎) ≡ 0 (mod 𝔮) and 𝑓 (𝑥) is irreducible modulo 𝔮. This is a contradiction,
and there can be no solution.
With this cautionary but illustrative example in hand, we state our main theorem.
We return to the idelic notation of section 30.4. We will consider embeddings in the
context of strong approximation (see Chapter 28).
The following notation will be in use throughout this chapter.
31.1.4. Let 𝑅 = 𝑅 (T) be a global ring with eligible set T and let 𝐹 = Frac 𝑅 be its field
of fractions. Let 𝐵 be a quaternion algebra over 𝐹 and suppose that 𝐵 is T-indefinite.
Let O ⊆ 𝐵 be an 𝑅-order.
Let 𝐾 ⊇ 𝐹 be a separable quadratic 𝐹-algebra and let 𝑆 ⊆ 𝐾 be an 𝑅-order.
Suppose that Emb( 𝑆;b O)
b ≠ ∅, which is to say, for all primes 𝔭 ⊆ 𝑅, the 𝑅𝔭 -algebra 𝑆𝔭
embeds optimally into O𝔭 .
Our struggle will be to understand when the local optimal embeddings glue together
to give a global optimal embedding. As a start, we know by Corollary 30.4.18 that
there exists some order O0 ∈ Gen O in the genus of O (i.e., locally isomorphic to O)
such that Emb(𝑆, O0) ≠ ∅.
31.1. SELECTIVE ORDERS 559
Definition 31.1.5. We say that Gen O is genial for 𝑆 if Emb(𝑆, O0) ≠ ∅ for all
O0 ∈ Gen O. If Gen O is not genial, i.e., there exists O0 ∈ Gen O such that
Emb(𝑆, O0) = ∅,
The core application of the optimal selectivity theorem is the following corollary.
ℎ(𝑆)
𝑚(𝑆, O; O× ) = b O;
𝑚( 𝑆, b× ).
b O
# Cls O
Proof. We combine Main Theorem 31.1.7 and Theorem 30.4.7.
nrd(O× ) = 𝑅>× Ω 0 .
Proposition 31.2.1. Let O be an Eichler order of level 𝔐. Then Condition (OS) holds
if and only if all of the following four conditions hold:
(a) The extension 𝐾 ⊇ 𝐹 and the quaternion algebra 𝐵 are ramified at the same
(possibly empty) set of archimedean places of 𝐹;
(b) 𝐾 and 𝐵 are unramified at all nonarchimedean places 𝑣 ∈ Pl 𝐹;
(c) Every nonarchimedean place 𝑣 ∈ T splits in 𝐾; and
(d) If 𝔭 ⊂ 𝑅 is a nonzero prime and ord𝔭 (𝔐) is odd, then 𝔭 splits in 𝐾.
Proof. We determine the class field 𝐻𝐺 𝑁 (O) obtained from the group 𝐺 𝑁 (O) =
𝐹>×Ω 0 nrd(𝑁 𝐵b× ( O)).
b
Recall we have 𝐺 (O) = 𝐹>×Ω 0 nrd( O
b× ) = 𝐹 × 𝑅
>Ω 0
b× , since O is an Eichler order
and therefore locally norm-maximal, so 𝐻𝐺 (O) is the maximal abelian extension of 𝐹
unramified away from the real places in Ram(𝐵) and such that the remaining places
𝑣 ∈ T split completely.
31.3. ∗ SELECTIVITY SETUP 561
𝐵× \ 𝐵
b× /b ∼ 𝐹 × \𝐹
Γ −→ b× /nrd(b
Γ) = Cl𝐺 (Γ) 𝑅 (31.3.4)
>Ω 0
𝐸/Γ = { 𝛽bb
Γ ∈ 𝐸/
bb b ∈ 𝐺 (Γ)} ⊆ 𝐸/
Γ : nrd( 𝛽) bb Γ.
That is to say, if 𝛽b ∈ 𝐸,
b then there exists 𝛽 ∈ 𝐸 such that 𝛽bb
Γ = 𝛽b
Γ if and only if
b ∈ 𝐺 (Γ).
nrd( 𝛽)
b𝛽 ∩ O
𝐾 b𝛾b = 𝐾
b𝛽 ∩ O
b = 𝑆b𝛽
Lemma 31.3.5 points the way more generally, at least to detect if there is an
embedding in the first place in an order. First, we need to give representatives of the
type set.
31.3.6. Recalling 28.5.7, there is a bijection
Typ O ↔ 𝐵× \ 𝐵
b× /𝑁 b× ( O);
𝐵
b
O0 = b bb𝜈 −1 ∩ 𝐵
b𝜈 −1 ∩ 𝐵 = O
𝜈 Ob
b× in 𝐵× \ 𝐵
𝜈∈𝐵
(yes, the choice of inverse is deliberate), with the class of b b× /𝑁 b× ( O)
b
𝐵
uniquely defined.
In the presence of strong approximation (Corollary 28.5.10), we have a further
bijection
Typ O ↔ Cl𝐺 𝑁 (O) 𝑅
where
𝐺 𝑁 (O) = 𝐺 (𝑁 𝐵× (O)) = 𝐹>×Ω 0 nrd(𝑁 𝐵b× ( O)).
b
31.3.9. By the main theorem of class field theory (Theorem 27.5.10), the Artin map
gives a bijection
∼ Gal(𝐾 | 𝐹) ' Z/2Z.
𝐹 × /𝐹 × Nm𝐾 |𝐹 (𝐾 × ) −→ (31.3.10)
𝐹 × /𝐹 × 𝐹∞,>0
× ∼ 𝐹 × /𝐹 × .
−→ 6𝑆 >0
We have
× ×
𝐹∞,>0 ≤ Nm𝐾 |𝐹 (𝐾∞ ) ≤ Nm𝐾 |𝐹 (𝐾 × )
(the latter embedded at the infinite place with the other components 1), and by the same
argument as in Lemma 27.5.6, the image of 𝐹 × Nm𝐾 |𝐹 (𝐾 × ) under the isomorphism
(27.5.8) is 𝐹>×Σ 0 Nm𝐾 |𝐹 (𝐾6𝑆× ) where Σ ⊆ Pl(𝐹) is the set of places ramified in 𝐾 (going
from real to complex in the extension 𝐹 ⊆ 𝐾) and
𝐹 × /𝐹 × Nm𝐾 |𝐹 (𝐾 × ) → 𝐹
b× /𝐹 × Nm𝐾 |𝐹 ( 𝐾
>Σ 0
b× ). (31.3.12)
𝐹>×Σ 0 Nm𝐾 |𝐹 ( 𝐾
b× ) ≤ 𝐹
b×
with total index at most 2, and index equal to 2 if and only if every nonarchimedean
place 𝑣 ∈ T is split in 𝐾.
Proof. In the projection (31.3.12), we start with a group of order 2; in order to keep
it this size, the projection away from the nonarchimedean places in T must be an
isomorphism, which holds if and only if for all nonarchimedean places 𝑣 ∈ T we must
have 𝑣 split in 𝐾.
𝐹>×Σ 0 nrd( 𝐾
b× ) ≤ 𝐹 × nrd( 𝐾
>Ω 0
b× ) nrd(𝑁 b× ( O))
𝐵
b (31.4.2)
with index at most 2, and equality holds if and only if either the optimal selectivity
condition (OS) holds or there exists a nonarchimedean place 𝑣 ∈ T inert or ramified
in 𝐾.
𝐹>×Σ 0 nrd( 𝐾
b× ) ≤ 𝐹 × nrd( 𝐾
>Ω 0
b× ) nrd(𝑁 b× ( O))
𝐵
b ≤𝐹b× (31.4.3)
we again have total index at most 2. By class field theory and the Galois correspondence
relative to the corresponding tower of class groups, we have
𝐹>×Σ 0 nrd( 𝐾
b× ) = 𝐹 × nrd( 𝐾
>Ω 0
b× ) nrd(𝑁 b× ( O))
𝐵
b <𝐹b×
(so first equality, then strict inequality in (31.4.3)) if and only if 𝐾 ⊆ 𝐻𝐺 𝑁 (O) if and
only if (OS) holds. The result then follows by Lemma 31.3.13.
We next consider the right-most inequality, and we show that it contains the
obstruction to selectivity.
Proof. Define
0 0
𝐸 0 = {𝛽 0 ∈ 𝐵× : 𝐾 𝛽 ∩ O0 = 𝑆 𝛽 }
31.5. ∗ MIDDLE SELECTIVITY EQUALITY 565
∼ 𝐾 × \𝐸 0, and similarly 𝐸
so that Emb𝑅 (𝑆, O0) −→ b0.
Suppose Emb(𝑆, O ) ≠ ∅, represented by 𝛽 ∈ 𝐸 0. Then
0 0
b𝛽0 ∩ O
𝐾 b𝛽0 ∩ O
b0 = 𝐾 bb𝜈 −1 = 𝑆b𝛽0
so 𝛽 0b
𝜈 = 𝛽b ∈ 𝐸.
b Therefore
b𝛽b ∩ O
𝐾 b = 𝑆b𝛽b
thus if 𝛽b0 = 𝛽b
b𝜈 −1 we get
b𝛽b0 ∩ O
𝐾 bb𝜈 −1 = 𝐾
b𝛽b0 ∩ O
b0 = 𝑆b𝛽b0 (31.4.5)
and 𝛽b0 ∈ 𝐸
b0. We have
Proof. The statement about index follows from the layering of the sandwich (31.3.15).
For the second statement, suppose that Emb(𝑆, O0) is nonempty; then by Proposition
31.4.4, we have O0 = O bb𝜈 −1 ∩ 𝐵 with nrd(b𝜈 ) ∈ 𝐹>×Ω 0 nrd( 𝐸).
b If equality holds in
×
(31.5.2), then there exists 𝑎 ∈ 𝐹 , b 𝛼∈𝐾 ×
b , and 𝜂b ∈ 𝑁 b× ( O)
b such that
>Ω 0 𝐵
𝜈 ) = 𝑎 nrd(b
nrd(b 𝛼) nrd(b
𝜂). (31.5.3)
We restore notation from Proposition 31.4.4, and modify the argument in the
converse. We define the map
b0
𝐸→𝐸
(31.5.4)
𝛽 ↦→ 𝛽b0 = b
𝛼 𝛽b𝜈 −1
𝜂b
b𝛽 ∩ O
We argue as in (31.4.5). From 𝛽 ∈ 𝐸, we have 𝐾 b× , so
𝛼∈𝐾
b = 𝑆b𝛽 . We have b
𝛼
b = 𝐾.
𝐾 b b And 𝜂b ∈ 𝑁 b× ( O), so O = O. Therefore
b b 𝜂
b b
𝐵
b𝛽b0 ∩ O
𝐾 bb𝜈 −1 = 𝐾
b𝛽b0 ∩ O
b0 = 𝑆b𝛽b0 (31.5.5)
𝛼 𝛽𝜇b
b 𝜈 −1 O
𝜂b b0 = b
𝛼 𝛽𝜇bb𝜈 −1 = b
𝜂Ob 𝛼 𝛽𝜇Ob𝜈 −1 = b
b𝜂b 𝛼 𝛽Ob𝜈 −1 = b
b𝜂b 𝛼 𝛽b𝜈 −1 O
𝜂b b0 .
This map works as well interchanging the roles of O and O0, and after a little chase,
we verify that the map (31.5.6) is bijective. Taking orbits under 𝐾 × on the left, we
conclude the proof.
b× ) nrd(𝑁 b× ( O))
nrd( 𝐾 b = nrd( 𝐸);
b (31.5.8)
𝐵
the inclusion (≤) was direct, so we prove (≥). The desired equality is now idelic, so
we reduce to checking in the completion at a prime 𝔭. If 𝐵𝔭 is a division algebra, then
O𝔭 is maximal, and 𝑚(𝑆𝔭 , O𝔭 ; 𝑁 𝐵𝔭× (O𝔭 )) = 1 by Proposition 30.5.3(b) we have the
stronger equality 𝐸 𝔭 = 𝑁 𝐵𝔭× (O𝔭 ) = 𝐵𝔭× .
31.6. ∗ OPTIMAL SELECTIVITY CONCLUSION 567
nrd(𝛽𝔭 ) ∈ nrd(𝐾𝔭× )𝑅𝔭× ≤ nrd(𝐾𝔭× ) nrd(O×𝔭 ) ≤ nrd(𝐾𝔭× ) nrd(𝑁 𝐵𝔭× (O𝔭 )) (31.5.9)
as claimed.
It has been a long and pretty technical road, so as refreshment we work through an
example (cf. Maclachlan [Macl2008, §4, Example 1]).
Example 31.6.1. Let 𝐹 be the totally real cubic field Q(𝑏) where 𝑏 3 − 4𝑏 − 1 = 0;
then 𝐹 has discriminant 229 and ring of integers 𝑅 = Z𝐹 = Z[𝑏]. The usual class
group Cl 𝑅 is trivial, but the narrow class group is Cl+ 𝑅 ' Z/2Z, represented by the
ideal 𝔭 = (𝑏 + 1)Z𝐹 of norm 2—the ideal 𝔭 is principal, but there is no generator√that
+ +
is totally positive.
narrow class field 𝐾 = 𝐻 ⊇ 𝐹 is quadratic, with 𝐻 = 𝐹 ( 𝑏).
The
−1, 𝑏
Let 𝐵 = . Then 𝑏 is positive at precisely one real place and negative at
𝐹
the other two, and 𝑏 ∈ Z×𝐹 . Computing the Hilbert symbol at the even primes, we
568 CHAPTER 31. SELECTIVITY
conclude that Ram(𝐵) is equal to two real places. In particular, 𝐵 is indefinite. The
class group Cl𝐺 (O) 𝑅 with modulus equal to these two real places is equal to Cl+ 𝑅, as
we see by the real signs of 𝑏.
Next, we compute representatives of the type set of maximal orders for 𝐵. By strong
approximation (Corollary 28.5.10), we have Typ O in bijection with Cl𝐺 𝑁 (O) 𝑅, so
we need to compute the idelic normalizer: but 𝐵 is unramified at all nonarchimedean
places, and
𝑁 𝐵b× ( O)
b =𝑁 b× b×
b) (M2 ( 𝑅)) = 𝐹 O .
b
GL2 ( 𝐹
𝐺 𝑁 (O) = 𝐹>×Ω 0 𝐹
b×2 𝑅
b× = 𝐹 × 𝑅
>Ω 0
b× = 𝐺 (O).
In other words, the quotient map Cl𝐺 (O) 𝑅 → Cl𝐺 𝑁 (O) 𝑅 is an isomorphism, still a
group of order 2. We conclude that # Typ O = 2.
We compute a maximal order
𝑏2𝑖 + 𝑗 𝑏2 + 𝑖 𝑗
O = O1 = Z 𝐹 ⊕ Z 𝐹 𝑖 ⊕ Z 𝐹 ⊕ Z𝐹 .
2 2
We conjugate this order by an ideal of reduced norm 𝔭 to get the second representative
(𝑏 2 + 𝑏 + 1) + (𝑏 + 1)𝑖 + 𝑗 (𝑏 + 1) + (𝑏 2 + 𝑏 + 1)𝑖 + 𝑖 𝑗
O2 = Z 𝐹 ⊕ Z 𝐹 𝑖 ⊕ Z 𝐹 ⊕ Z𝐹 .
2 2
Therefore these orders represent the two types of maximal orders, and Typ O =
{[O1 ], [O2 ]}.
With all of these elements in place, we can observe selectivity (Main Theorem
31.1.7). We saw that both 𝐾 and 𝐵 are ramified at no nonarchimedean places and
exactly the same set of real places. In particular, the field 𝐾 ↩→ 𝐵 embeds by the
local-global principle. Let 𝑆 = Z𝐾 = Z𝐹 [𝑤] be the maximal order in 𝐾. Then
𝑤 2 − 𝑏𝑤 + 1 = 0. Then Emb( 𝑆;b O)
b ≠ ∅ (Proposition 30.5.3(a)).
The optimal selective condition (OS) holds because we took it so, 𝐾 = 𝐻 + . It
follows that 𝑆 embeds in exactly one of O1 or O2 . We find that
𝑏2 + 𝑖 𝑗
𝛼= ∈ O1
2𝑏
satisfies 𝛼2 − 𝑏𝛼 + 1 = 0 as desired; so 𝑆 embeds in O1 (and not O2 ).
Remark 31.6.2. Without the hypothesis of strong approximation, it is very difficult to
tease apart the contributions from different orders in the genus: indeed, the generating
series for representation numbers for a definite quaternion order give coefficients of
modular forms, discussed in Chapter 41.
Definition 31.7.1. We say that Gen O is selective for 𝑆 if there exists O0 ∈ Gen O
such that there is no embedding 𝑆 ↩→ O0 of 𝑅-algebras.
The difference between Definition 31.1.5 and Definition 31.7.1 is that in the latter,
we do not insist that the embedding is optimal. It may happen that Gen O is selective
for 𝑆, but Gen O is not optimally selective for 𝑆: such a situation arises exactly when
there is an order O0 ∈ Gen O such that 𝑆 embeds in O0 but does not optimally embed
in O0.
31.7.4. If Gen O is optimally selective for 𝑆 but not selective for 𝑆, then 𝑆 ⊆ O is
optimal but there is an order O0 such that 𝜙 0 : 𝑆 ↩→ O0 is an embedding but not an
optimal embedding. Let 𝑆 0 = 𝜙 0 (𝐾) ∩ O0 ) 𝑆. So there exists a prime 𝔭 | [𝑆 0 : 𝑆] 𝑅 ,
and in particular, 𝑆 is not maximal at 𝔭. In particular, if 𝑆 is integrally closed, then
Gen O is selective for 𝑆 if and only if Gen O is optimally selective for 𝑆.
(S) If 𝔭 | disc𝑅 𝑆 and ord𝔭 (𝔐) ≠ ord𝔭 (disc𝑅 𝑆), then 𝔭 splits in 𝐾 ⊇ 𝐹.
570 CHAPTER 31. SELECTIVITY
Proof. Our very setup (section 31.3) is designed to count optimal embeddings, so
to avoid lengthening this chapter, we refer the reader to Chinburg–Friedman [CF99,
Theorem 3.3] for the case of maximal orders, and Chan–Xu [CX2004, Theorem 4.7]
and Guo–Qin [GQ2004, Theorem 2.5] (independently) for Eichler orders.
Remark 31.7.6. The condition Proposition 31.2.1(d) (one part of (OS) for Eichler
orders) is not visible in the selectivity theorem for Eichler orders, but is implied by it:
by (b), the extension 𝐾 ⊇ 𝐹 is unramified so 𝔭 is unramified in 𝐾, thus ord𝔭 (disc𝑅 (𝑆))
is even and necessarily not equal to ord𝔭 (𝔐) if the latter is odd.
31.7.7. Chinburg–Friedman [CF99, Theorem 3.3] prove Theorem 31.7.5 for maximal
orders, and they applied this theorem to embeddings in maximal arithmetic groups
[CF99, Theorem 4.4]. Chinburg–Friedman proved their results in the language of the
Bruhat–Tits tree of maximal orders. This selectivity theorem was then generalized to
Eichler orders by Chan–Xu [CX2004, Theorem 4.7] and Guo–Qin [GQ2004, Theorem
2.5] (independently). Interestingly, while Guo–Qin follow Chinburg–Friedman in their
proof, Chan–Xu instead use results on exceptional spinor genera and their results are
phrased and proven in the language of indefinite integral quadratic forms. (These
results are given for number fields, but the proofs adapt to global fields as pursued
here.)
Some selectivity theorems beyond those for Eichler orders are also known. Arenas-
Carmona [A-C2013, Theorem 1.2] considers more general intersections of maximal
orders. Linowitz [Lin2012, Theorems 1.3–1.4] gives a selectivity theorem for (op-
timal) embeddings into arbitrary orders, subject to some additional technical (copri-
mality) hypotheses. More generally, selectivity theorem have been pursued in the
more context of central simple algebras: see e.g. Linowitz–Shemanske [LS2012] and
Arenas-Carmona [A-C2012].
However, these selectivity results do not prove Main Theorem 31.1.7 on the nose,
either because they deal with selectivity instead of optimal selectivity or do not prove
the more powerful statement that the embedding numbers are in fact equal. On the
latter point, a general setup to establish equality of embedding numbers can be found
in work of Linowitz–Voight [LV2015, §2].
Theorem 31.8.1 (Vignéras). For every 𝑚 ≥ 2, there exist Laplace isospectral and
nonisometric manifolds of dimension 𝑚.
We sketch a proof of this theorem in this section, with attention to the particular
detail of selectivity. In 1994, Maclachlan–Rosenberger [MacRos94] claimed to have
produced a pair of Laplace isospectral, nonisometric hyperbolic 2-orbifolds of genus
0, but then Buser–Flach–Semmler [BFS2008] later showed that these examples were
too good to be true! The subtle issue they found: the phenomenon of selectivity.
Our construction is quaternionic, of course. We consider the situation of sections
38.2–38.3, specifically the setup in 38.2.1 and 38.3.1. Let 𝐹 be a number field with 𝑟
real places and 𝑐 complex places, so that [𝐹 : Q] = 𝑟 + 2𝑐 = 𝑛. Let 𝐵 be a quaternion
algebra over 𝐹 and suppose that 𝐵 is split at 𝑡 real places. We have an embedding
(38.2.2) 𝜄 : 𝐵 ↩→ M2 (R) 𝑡 × M2 (C) 𝑐 . Letting H := (H2 ) 𝑡 × (H3 ) 𝑐 as in (38.2.9) and
P𝐵×>0 := 𝐵×>0 /𝐹 × ,
Theorem 31.8.2. Let O0 ∈ Gen O. Suppose that for every quadratic field 𝐾 ⊇ 𝐹 and
every quadratic 𝑅-order 𝑆 ⊆ 𝐾, we have the equality of embedding numbers
Proof. See Vignéras [Vig80b, Théorème 6]. The statement there is in terms of all
embeddings, not just optimal embeddings—but the total count of conjugacy classes
of embeddings of an quadratic order is a sum of the corresponding count of optimal
embeddings of superorders (as in 31.7), so it is sufficient to have a genial order.
The key ingredient in the proof is the Selberg trace formula, which allows us
to show that the spectra of the Laplace operators agree by the stronger condition of
representation equivalence: for a more general point of view on this deduction, see
Deturck–Gordon [DG89].
The rub is in the equality (31.8.3). The restriction of the equivalence classes to
units of reduced norm 1 is harmless: by Lemma 30.3.14, we have 𝑚(𝑆, O; O1 ) =
𝑚(𝑆, O; O× ) [nrd(O× ) : nrd(𝑆 × )], and as a consequence of strong approximation we
have [nrd(O× ) : nrd(𝑆 × )] = [nrd( Ob× : nrd( 𝑆b× )]. Thus if O0 ∈ Typ O, then (31.8.3)
holds if and only if 𝑚(𝑆, O; O ) = 𝑚(𝑆, O0; O0× ).
×
Proof. Isospectrality follows from Theorem 31.8.2 with Main Theorem 31.1.7. To
show that 𝑋, 𝑋 0 are not isometric, since O0 ; O we know that PO0 is not conjugate
to PO in 𝐵× , but we need this for the groups Γ1 (O0), Γ1 (O) in Isom+ (H ), which is
slightly larger (see Remark 38.2.11). We leave the details to Exercise 31.3 (or see
Linowitz–Voight [LV2015, Proposition 2.24]).
The remainder of the proof of Theorem 31.8.1 involves finding suitable data
𝐹, 𝐵, O. We exhibit a pair for 𝑚 = 2 following Linowitz–Voight [LV2015, Example
5.2], giving a pair of compact hyperbolic 2-manifolds of genus 6 which are isospectral
but not isometric.
Example 31.8.5. Let 𝐹 = Q(𝑤) where 𝑤 is a root of the polynomial 𝑥 4 −5𝑥 2 −2𝑥+1 = 0.
Then 𝐹 is a totally real quartic field with discriminant 𝑑 𝐹 = 5744 = 24 359, Galois
group 𝑆4 , class number # Cl 𝑅 = 1 and narrow class number # Cl+ 𝑅 = 2. Let 𝐵 be
the quaternion algebra over 𝐹 which is ramified at the prime ideal 𝔭13 of norm 13
generated by 𝑏 := 𝑤 3 − 𝑤 2 − 4𝑤 and three of the four real places, with split place
𝑎, 𝑏
𝑤 ↦→ −0.751024 . . .: then 𝐵 = where 𝑎 = 𝑤 3 − 𝑤 2 − 3𝑤 − 1 is a root of
𝐹
𝑥 4 + 8𝑥 3 + 12𝑥 2 − 1.
A maximal order O ⊂ 𝐵 is given by
(𝑤 3 + 1) + 𝑤 2𝑖 + 𝑗 (𝑤 + 1) + (𝑤 3 + 1)𝑖 + 𝑖 𝑗
O = 𝑅 ⊕ 𝑅𝑖 ⊕ 𝑅⊕ 𝑅;
2 2
O has type number 2, so there exists two isomorphism classes of maximal orders
O1 = O and O2 .
31.8. ∗ ISOSPECTRAL, NONISOMETRIC MANIFOLDS 573
We claim that O1 and O2 have no elements of finite order other than ±1. Indeed,
if we had such an element of order 𝑞 then 𝐹 (𝜁2𝑞 ) is a cyclotomic quadratic extension
of 𝐹, whence Q(𝜁2𝑞 ) + ⊆ 𝐹; but 𝐹 is primitive, so the only cyclotomic quadratic
√ √
extensions
√ of 𝐹 √are 𝐾 = 𝐹 ( −1) and 𝐾 = 𝐹 ( −3). But as 𝔭13 splits completely in
𝐹 ( −1) and 𝐹 ( −3), neither field embeds into 𝐵. We conclude that the groups Γ𝑖
are torsion free.
Since 𝐵 is ramified at a finite place, the genus of O is genial by Theorem 31.2.1,
and since Aut(𝐹) is trivial, the hypothesis of Corollary 31.8.4 are satisfied: 𝑋11 , 𝑋21
are Laplace isospectral, but not isometric.
Finally, by Theorem 39.1.13, we have area(𝑋𝑖 ) = 20𝜋, so 𝑔(𝑋𝑖 ) = 6 for 𝑖 = 1, 2.
Fundamental domains for these are given in Figure 31.8.6.
Figure 31.8.6: Fundamental domain for the genus 6 manifolds 𝑋 (Γ11 ),𝑋 (Γ12 )
We obtain a second example by choosing the split real place 𝑤 ↦→ −1.9202 . . ., and
since 𝐹 is not Galois, as in the case of the 2-orbifold pairs 2 and 3, these are pairwise
nonisometric.
For an example with 𝑚 = 3, see Exercise 31.4.
1 1
Exercises
I 1. Prove Main Theorem 31.1.7 in the case 𝐾 ' 𝐹 × 𝐹: to be precise, show that
an 𝑅-order 𝑆 ⊆ 𝐹 × 𝐹 embeds equally in all Eichler 𝑅-orders. [Hint: we must
have 𝐵 ' M2 (𝐹), so reduce to the case where 𝑆 is embedded in the diagonal
and then conjugate.]
2. The following exercise gives insight into the proof of Theorem 31.7.5 on selec-
tivity. Let 𝑅 be local, and let O be an Eichler order of level 𝔭𝑒 .
Let 𝜙 : 𝑆 ↩→ O be an optimal embedding that is normalized and associated to
𝑥 ∈ 𝑅, so represented by
𝑥 1
𝛼=
− 𝑓 𝛾 (𝑥) 𝑡 − 𝑥
574 CHAPTER 31. SELECTIVITY
as in Definition 30.6.8.
(a) Compute 𝜈 −1 𝛼𝜈 for the matrix
𝑎 𝑏
𝜈= .
𝑐 𝑑
𝐵×
𝜏 / 𝐵×
𝜄 𝜄
𝐺
𝜎 /𝐺 𝜈 /𝐺
575
Chapter 32
Unit groups
Having moved from algebra and arithmetic to analysis, and in particular the study of
class numbers, in this part we consider geometric aspects of quaternion algebras, and
the unit group of a quaternion order acting by isometries on a homogeneous space.
i ρ −ρ2
−1 1 −1 1 −1 1
−i ρ2 −ρ
577
578 CHAPTER 32. UNIT GROUPS
of places of 𝐹, and let 𝑅 = 𝑅S be the global ring associated to S, the ring of S-integers
of 𝐹. (As always, the reader may keep the case 𝐹 = Q, S = {∞}, and 𝑅 = Z in mind.)
Further, let 𝐵 be a quaternion algebra over 𝐹, and let O ⊂ 𝐵 be an 𝑅-order.
We are interested in the structure of the group O× . Since 𝑍 (𝐵× ) = 𝐹 × , we have
𝑅 ≤ 𝑍 (O× ) central. We understand the structure of 𝑅 × by Dirichlet’s unit theorem,
×
as follows.
32.2.1. From Dirichlet’s unit theorem (and its extension to S-units and the function
field case), the group 𝑅 × of units is a finitely generated abelian group of rank #S − 1,
so that
𝑅 × ' Z/𝑤Z ⊕ Z#S−1 (32.2.2)
where 𝑤 is the number of roots of unity in 𝐹. (The proof is briefly recalled in 32.3.1.)
The group O× is (in general) noncommutative, so we should not expect a description
like 32.2.1. But to get started, we consider the quotient O× /𝑅 × , and the reduced norm
map which gets us back into 𝑅 × .
32.2.3. We recall the theorem on norms (see section 14.7): as before, let
Ω := {𝑣 ∈ Ram 𝐵 : 𝑣 real} ⊆ Pl 𝐹 (32.2.4)
be the set of real ramified places in 𝐵 (recalling that complex places cannot be ramified),
and
𝐹>×Ω 0 := {𝑥 ∈ 𝐹 × : 𝑣(𝑥) > 0 for all 𝑣 ∈ Ω} (32.2.5)
the set of elements that are positive at the places 𝑣 ∈ Ω. If 𝐹 is a function field, then
Ω = ∅, and 𝑅>× Ω 0 = 𝑅 × . The Hasse–Schilling norm theorem (Main Theorem 14.7.4)
says that nrd(𝐵× ) = 𝐹>×Ω 0 . Letting
32.2.11. In general, the exact sequence (32.2.10) does not split, so the group O× /𝑅 ×
will be a nontrivial extension of O1 /{±1} by an elementary abelian 2-group.
Remark 32.2.14. Some authors write GL1 (O) = O× and SL1 (O) = O1 , and this
notation suggests generalizations. In such situations, it is natural to write PGL1 (O) =
O× /𝑅 × and PSL1 (O) = O1 /{±1}.
32.3.1. To build intuition, suppose 𝐹 is a number field with 𝑟 real places and 𝑐 complex
places. Recall the proof of Dirichlet’s unit theorem: we define a map
𝑅 × → RS
(32.3.2)
𝑥 ↦→ (𝑚 𝑣 log|𝑥| 𝑣 ) 𝑣 .
The kernel of this map is the group of roots of unity (the torsion subgroup of 𝑅 × ).
Í
The image lies inside the trace zero hyperplane 𝑣 ∈S 𝑥 𝑣 = 0 by the product formula
(14.4.6), and it is discrete and cocompact inside this hyperplane, so it is isomorphic
to ZS−1 . In particular, 𝑅 × is finite if and only if #S = 1; since S always contains
the set of archimedean places of size 𝑟 + 𝑐, we see that 𝑅 × is finite if and only if
(𝑟, 𝑐) = (1, 0), (0, 1), so 𝐹 = Q or 𝐹 is an imaginary quadratic field.
Remark 32.3.3. Informally, one might say that 𝑅 × is finite only when the completions
at the places in S provide “no room” for the unit group to become infinite. This is
analogous to the informal case for strong approximation in 28.5.4: if there is a place
𝑣 ∈ S where 𝐵1𝑣 is not compact, then there is enough room for 𝐵1 to “spread out” and
become dense.
32.3. UNITS IN DEFINITE QUATERNION ORDERS 581
32.3.4. Recall (Definition 28.5.1) that 𝐵 is S-definite if S ⊆ Ram(𝐵), i.e., every place
in S is ramified in 𝐵. In particular, if 𝐹 is a number field, then since a complex place
is split and S contains the archimedean places, if 𝐵 is S-definite then 𝐹 is totally real;
in this case, when S is exactly the set of archimedean places, we simply say that 𝐵 is
definite.
32.3.5. Consider the setup in analogy with Dirichlet’s unit theorem 32.3.1. We
consider the embedding of 𝐵 into the completions at all places in S:
Ö
𝐵 ↩→ 𝐵S := 𝐵𝑣 .
𝑣 ∈S
Î
By Exercise 27.14, 𝑅 is discrete in 𝐹S = 𝑣 ∈S 𝐹𝑣 and O is discrete in 𝐵6S (the point
being that in the number field case, S contains all archimedean places). Consequently,
the injections Ö
O× /𝑅 × ↩→ (𝐵S ) × /(𝐹S ) × := 𝐵×𝑣 /𝐹𝑣×
𝑣 ∈𝑆
Ö (32.3.6)
1 1
O ↩→ (𝐵S ) = 𝐵1𝑣
𝑣 ∈𝑆
have discrete image.
Depending on whether the place 𝑣 is nonarchimedean (split or ramified) or
archimedean (split real, ramified real, or complex), we have a different target compo-
nent 𝐵×𝑣 /𝐹𝑣× or 𝐵1𝑣 . The major task of Part IV is to describe these possibilities in detail
and look at the associated symmetric spaces.
We begin with the simplest case, where the unit groups involved are finite.
Proposition 32.3.7. The group O× /𝑅 × is finite if and only if O1 is finite if and only if
𝐵 is S-definite.
Proof. By the exact sequence (32.2.10), the group O× /𝑅 × is finite if and only if the
group O1 is finite.
First, suppose that 𝐵 is S-definite. Then by definition, for each 𝑣 ∈ S, the
completion 𝐵 𝑣 is a division algebra over 𝐹𝑣 . But each 𝐵1𝑣 is compact, from the
topological discussion in section 13.5. Therefore in (32.3.6), the group O1 is a closed,
discrete subgroup of a compact group—hence finite.
Now suppose 𝐵 is not S-definite. Then there is a place 𝑣 0 ∈ S that is unramified; we
will correspondingly find an element of infinite order (like solutions to the quaternion
Pell equation coming from the original Pell’s equation (32.1.2)). We have 𝐵 𝑣0 '
M2 (𝐹𝑣0 ), so there exists 𝛼 ∈ 𝐵 be such that the reduced characteristic polynomial
splits in 𝐹𝑣0 ; we may suppose without loss of generality that 𝐾 = 𝐹 [𝛼] is a field. Let
𝑆 be the integral closure of 𝑅 in 𝐾 (not to be confused with the set S). Then by the
Dirichlet S-unit theorem (32.2.1), the rank of 𝑆 × /𝑅 × is at least 1: the set of places
𝑤 ∈ Pl(𝐾) such that 𝑤 lies above 𝑣 ∈ S contains at least one element from each 𝑣
and two above 𝑣 0 , because it is split. So there is an element 𝛾 ∈ 𝑆 × /𝑅 × of infinite
order. As 𝑅-lattices, the order 𝑆 ∩ O has finite 𝑅-index and hence finite index in 𝑆,
so 𝑆 × /(𝑆 ∩ O) × is a finite group, and therefore a sufficiently high power of 𝛾 lies in
(𝑆 ∩ O) × ⊆ O× , and O× is infinite.
582 CHAPTER 32. UNIT GROUPS
O[1/2] × = h2, 𝑖, 𝑗, 1 + 𝑖i
(32.3.9)
O[1/2] × /Z[1/2] × = h2, 𝑖, 𝑗, 1 + 𝑖i/h−1, 2i ' 𝑄 8 o Z/2Z
(Exercise 32.2).
Finally, if we take S = {5, ∞}, then 𝐵 is no longer S-definite; and O[1/5] × contains
the element 2 + 𝑖 of norm 5 ∈ Z[1/5] × and infinite order.
32.3.10. Suppose 𝐵 is S-definite. Then by Proposition 32.3.7, the group O× /𝑅 × is
finite. Since we have an embedding
O× /𝑅 × ↩→ 𝐵× /𝐹 ×
Proof. Let 𝑣 be an archimedean place of 𝐹. Then the natural map 𝐵× → 𝐵×𝑣 /𝐹𝑣× has
kernel 𝐹𝑣× ∩ 𝐵× = 𝐹 × , so the group homomorphism 𝐵× /𝐹 × ↩→ 𝐵×𝑣 /𝐹𝑣× is injective.
First suppose that 𝑣 is a ramified (real) place, so 𝐵 𝑣 ' H and
By Corollary 2.4.21, we have H1 /{±1} ' SO(3) so Γ is a finite rotation group: these
are classified in Proposition 11.5.2.
32.5. CYCLIC SUBGROUPS 583
In general, we seek to conjugate the group Γ in order to reduce to the case above.
We may prove the lemma after making a base extension of 𝐹, so we may suppose
that 𝑣 is complex, with 𝐵 𝑣 ' M2 (C). Then 𝐵×𝑣 /𝐹𝑣× ' PGL2 (C), and via the injection
𝐵× /𝐹 × ↩→ PGL2 (C) we obtain a finite subgroup Γ ⊆ PGL2 (C). The natural map
SL2 (C) → PGL2 (C) is surjective, as we may rescale every invertible matrix by a
square root of its determinant to have determinant 1, and its kernel is {±1}, giving an
isomorphism PSL2 (C) ' PGL2 (C). We then lift Γ under the projection SL2 (C) →
PSL2 (C) to a finite group (containing −1). We have
Choose an orthonormal basis for h , iΓ and let 𝑇 ∈ SL2 (C) be the change of basis matrix
relative to the standard basis. Then h𝑧, 𝑤iΓ = h𝑇 𝑧, 𝑇 𝑤i and therefore 𝑇Γ𝑇 −1 ⊂ SU(2).
The result now follows from (32.4.2) and the previous case.
Proposition 32.5.1. Let 𝑚 > 2 and let 𝜁 𝑚 ∈ 𝐹 al be a primitive 𝑚th root of unity. Then
P𝐵× contains a cyclic subgroup of order 𝑚 if and only if 𝜁 𝑚 + 𝜁 𝑚 −1 ∈ 𝐹 and 𝐹 (𝜁 )
𝑚
splits 𝐵. Such a cyclic subgroup is unique up to conjugation in P𝐵× .
Now we prove (⇒). Suppose that 𝛾 ∈ 𝐵× has image in P𝐵× of order 𝑚 > 2, so
that 𝛾 𝑚 = 𝑎 ∈ 𝐹 × . We do calculations in the commutative 𝐹-algebra 𝐾 := 𝐹 [𝛾]. Let
𝜍 := 𝛾𝛾 −1 ∈ 𝐾 × . Then
𝜍 𝑚 = 𝛾 𝑚 (𝛾 𝑚 ) −1 = 𝑎𝑎 −1 = 1 (32.5.2)
so 𝜍 𝑚 = 1. If 𝜍 𝑑 = 1 for 𝑑 | 𝑚 then 𝛾 𝑑 = 𝜍 𝑑 𝛾 𝑑 = 𝛾 𝑑 so 𝛾 𝑑 ∈ 𝐹 × and thus 𝑑 = 𝑚;
thus 𝜍 has order 𝑚 in 𝐵× . Applying the standard involution again gives
𝛾 = 𝛾𝜍 = 𝜍𝛾 = 𝜍 𝜍𝛾; (32.5.3)
field in 𝐵, in which case 𝐾 splits 𝐵 by Lemma 5.4.7, or 𝐾 is not a field and 𝐵 ' M2 (𝐹),
in which case 𝐹 already splits 𝐵.
We conclude with uniqueness. Continuing from the previous paragraph, we have
shown that 𝛾 + 𝛾 = (1 + 𝜍)𝛾 ∈ 𝐹 × , so 𝛾 and 1 + 𝜍 generate the same cyclic subgroup
of P𝐵× , where 𝜍 𝑚 = 1. If 𝐾 = 𝐹 (𝜍) is a field, then all embeddings 𝐹 (𝜁 𝑚 ) ↩→ 𝐵 are
conjugate in 𝐵× by the Skolem–Noether theorem (Corollary 7.1.5), and consequently
every two cyclic subgroups of order 𝑚 are conjugate. Otherwise, the reduced charac-
teristic polynomial of 𝜍 factors, so 𝐵 ' M2 (𝐹), and its roots (the eigenvalues of 𝜍)
belong to 𝐹. If the eigenvalues
are repeated, then up to conjugation in GL2 (𝐹), 𝜍 is
1 𝑏
a scalar multiple of with 𝑏 ∈ 𝐹, and therefore has infinite order, impossible.
0 1
1 0
Thus the roots are distinct, and 𝜍 is conjugate to a multiple of and so 𝜆 is a
0 𝜆
primitive 𝑚th root of unity and the cyclic subgroup is unique up to conjugation.
32.5.5. The proof of Proposition 32.5.1 describes the cyclic subgroup explicitly, up to
conjugation (still with 𝑚 > 2):
In contrast to Proposition 32.5.1, there are a great many cyclic subgroups of order
𝑚 = 2 in P𝐵× , described as follows.
32.6. ∗ DIHEDRAL SUBGROUPS 585
Corollary 32.5.7. Let 𝑚 ≥ 1 be odd. Then P𝐵× contains a cyclic subgroup of order
𝑚 if and only if P𝐵× contains a cyclic subgroup of order 2𝑚.
Corollary 32.5.8. P𝐵× contains a cyclic subgroup of order 2𝑚 if and only if P𝐵1
contains a cyclic subgroup of order 𝑚.
Proof. The corollary follows from 32.5.5: the subgroup of P𝐵× of order 2𝑚 generated
by 𝛾2𝑚 yields the subgroup of P𝐵1 of order 𝑚 generated by 𝜁2𝑚 , and vice versa.
Proof. First (a). The implication (⇒) is immediate, so we prove (⇐). Let 𝑚 ≥ 2
and suppose that P𝐵× contains a cyclic subgroup of order 𝑚, generated by the image
of 𝛾 ∈ 𝐵× , and let 𝐾 = 𝐹 [𝛾]. Let 𝑗 ∈ 𝐵 be orthogonal to 𝐾 under nrd. Then
𝑗 2 = 𝑏 ∈ 𝐹 × , and 𝑗 𝛼 = 𝛼 𝑗 = 𝛼−1 𝑗 ∈ P𝐵× for all 𝛼 ∈ 𝐾, so the subgroup h𝛾, 𝑗i is
𝐾, 𝑏
dihedral of order 2𝑚 in P𝐵× , and 𝐵 ' , as in Exercise 6.2.
𝐹
Now (b). We just showed (⇒) in the previous part, so we show (⇐). Let Γ ⊆ P𝐵×
be a dihedral subgroup of order 2𝑚, where Γ = h𝛾, 𝑗i has 𝛾 ∈ 𝐵× generating a cyclic
subgroup of order 𝑚 in P𝐵× and 𝑗 ∈ 𝐵× satisfies
𝑗 −1 𝛾 𝑗 = 𝛾 −1 ∈ 𝐵× /𝐹 × .
586 CHAPTER 32. UNIT GROUPS
𝐾, 𝑏
Let 𝐾 = 𝐹 [𝛾]. We claim that 𝐵 = 𝐾 + 𝐾 𝑗 ' .
𝐹
First we show 𝑗 −1 𝛾 𝑗 = 𝛾. This follows from a direct argument using reduced
norm and trace (see Exercise 32.4), but we have also the following argument. Since
𝐾 = 𝐹 [𝛾] is semisimple (see 32.5.5) and conjugation by 𝑗 acts as an 𝐹-algebra
automorphism of 𝐾 = 𝐹 [𝛾], it is either the identity or the standard involution, and
thus 𝑗 −1 𝛾 𝑗 = 𝛾, 𝛾. But we cannot have 𝑗 −1 𝛾 𝑗 = 𝛾, because then 𝐾 [ 𝑗] ⊆ 𝐵 would be
a commutative subalgebra of dimension ≥ 3, a contradiction.
Now by (4.2.16), expanding the trace gives
trd( 𝑗 𝛾) = 𝑗 𝛾 + 𝛾 𝑗 = 𝑗 𝛾 + 𝛾(trd( 𝑗) − 𝑗)
(32.6.2)
= 𝑗 𝛾 − 𝛾 𝑗 + trd( 𝑗)𝛾 = trd( 𝑗)𝛾.
Lemma 32.6.3. Let 𝑚 > 2. Then the set of dihedral subgroups of order 2𝑚 up to
conjugation in P𝐵× are in bijection with the group
×)
Nm𝐾𝑚 /𝐹 (𝐾𝑚
(32.6.4)
h𝛿i𝐹 ×2
−1 .
where 𝐾𝑚 is as in 32.5.5 and 𝛿 = 2 + 𝜁 𝑚 + 𝜁 𝑚
𝑏𝑏 0 ∈ 𝐹 ×2 h𝛿i.
𝛼−1 𝑗 𝛼 = 𝛼𝛼−1 𝑗 = 𝛽 𝑗
as desired.
32.7. ∗ EXCEPTIONAL SUBGROUPS 587
Remark 32.6.5. One can rephrase Lemma 32.6.3 in terms of a global equivalence
relation, further encompassing the case 𝑚 = 2: see Chinburg–Friedman [CF2000,
Lemma 2.4].
1
Corollary 32.6.6.
group P𝐵 contains a dihedral group of order 2𝑚 > 4 if and
The
𝐾2𝑚 , −1
only if 𝐵 ' .
𝐹
Proof. We first prove (⇒). If P𝐵1 contains a dihedral group Γ = h𝛾, 𝑗i of order 2𝑚
then it contains a cyclic subgroup of order 𝑚 so by Corollary 32.5.8 the group P𝐵×
contains a cyclic subgroup of order 2𝑚, which we may take to be generated by 𝛾2𝑚
as in 32.5.5 with 𝐾2𝑚 = 𝐹 [𝛾2𝑚 ] = 𝐹 [𝛾]; by hypothesis we have 𝑗 2 = − nrd( 𝑗) = −1,
𝐾2𝑚 , −1
and so 𝐵 ' as in the classification in Lemma 32.6.1.
𝐹
Next we prove the converse implication (⇐). We refer to 32.5.5. In case (i) where
𝜁2𝑚 ∈ 𝐹, we have nrd(𝜁2𝑚 −1 𝛾 ) = 𝜁 2 𝜁 = 1 so we may take Γ = h𝜁 −1 𝛾 , 𝑗i; in case
𝑚 2𝑚 𝑚 2𝑚 𝑚
(ii), where 𝜁2𝑚 ∉ 𝐹, we take Γ = h𝜁2𝑚 , 𝑗i.
(a) P𝐵× contains a subgroup isomorphic to 𝐴4 if and only if P𝐵1 contains a subgroup
isomorphic to 𝐴4 if and only if 𝐵 ' (−1, −1 | 𝐹).
(b) P𝐵× contains a subgroup isomorphic to 𝑆4 if and only if it contains a subgroup
isomorphic to 𝐴4 ; and√P𝐵1 contains a subgroup isomorphic to 𝑆4 if and only if
𝐵 ' (−1, −1 | 𝐹) and 2 ∈ 𝐹.
(c) P𝐵× contains a subgroup isomorphic to 𝐴5 if and only if√P𝐵1 contains a subgroup
isomorphic to 𝐴5 if and only if 𝐵 ' (−1, −1 | 𝐹) and 5 ∈ 𝐹.
Any two such exceptional subgroups of P𝐵× (or P𝐵1 ) are conjugate by an element of
𝐵× if and only if they are isomorphic as groups.
Proof. First we prove (a); let Γ ⊆ P𝐵× be a subgroup with Γ ' 𝐴4 . The reduced
norm gives a homomorphism Γ → nrd(Γ) ⊆ 𝐹 × /𝐹 ×2 , but 𝐴4 has no nontrivial
homomorphic image of exponent 2, so nrd(Γ) ⊆ 𝐹 ×2 . Therefore, there is a unique
lift of Γ to 𝐵1 /{±1}, and the map 𝐵1 /{±1} → 𝐵× /𝐹 × is an isomorphism from this
lift to 𝐻. This shows the first implication; its converse follows from the injection
P𝐵1 ↩→ P𝐵× . For the second implication, let 𝑖, 𝑗 ∈ 𝐵1 generate the 𝑉4 -subgroup (the
normal subgroup of index 3 isomorphic to the Klein 4 group) of 𝐴4 in P𝐵1 . Then
𝑖, 𝑗 ∉ 𝐹 × , and 𝑖 2 = − nrd(𝑖) = −1 = 𝑗 2 ; and similarly (𝑖 𝑗) 2 = −1 implies 𝑗𝑖 = −𝑖 𝑗. By
Lemma 2.2.5, we conclude 𝐵 ' (−1, −1 | 𝐹). The converse follows from the Hurwitz
unit group 11.2.4.
588 CHAPTER 32. UNIT GROUPS
For part (b), the implication (⇒) is immediate; the implication (⇐) follows by
taking the Hurwitz units and adjoining the element 1 + 𝑖, as in 11.5.4 but working
modulo scalars. For the second statement, an element of order√4 in 𝐵1 /{±1} lifts to
an element of order 8 in 𝐵1 and therefore has reduced trace ± 2 ∈ 𝐹; the converse
follows again from the explicit construction in 11.5.4.
For part (c), we argue similarly. Since 𝐴5 is generated by its subgroups
√ isomorphic
to 𝐴4 , we may apply (a) to get a lift, and by the reduced trace we get 5 ∈ 𝐹.
The uniqueness statement is requested in Exercise 32.5.
Exercises
Unless otherwise indicated, let 𝐹 be a number field with ring of integers 𝑅, let 𝐵 be a
quaternion algebra over 𝐹 and let O ⊆ 𝐵 be an 𝑅-order.
𝜙 : 𝑆× → 𝑆×
𝑢 ↦→ 𝑢/𝑢
I 5. Prove the uniqueness statement in Proposition 32.7.1: Show that every two
isomorphic exceptional subgroups of P𝐵× are conjugate by an element of 𝐵× ,
and the same for 𝐵1 .
6. Let Γ ≤ O1 be a maximal finite subgroup. Combining results from sections
11.2 and 11.5 and this chapter, prove the following.
(a) Γ is isomorphic to one of the following groups:
• cyclic of order 2𝑚 with 𝑚 ≥ 1,
• binary dihedral (dicyclic) 2𝐷 2𝑚 of order 4𝑚 with 𝑚 ≥ 1,
• binary tetrahedral 2𝑇 of order 24,
• binary octahedral 2𝑂 of order 48, or
• binary icosahedral 2𝐼 of order 120.
In the latter three cases,√we call Γ exceptional. √
(b) If Γ ' 2𝑂 then 𝐹 ⊇ Q( 2) and if Γ ' 2𝐼 then 𝐹 ⊇ Q( 5).
−1, −1
(c) If Γ is exceptional, then 𝐵 ' .
𝐹
7. Continuing with the previous exercise, suppose that 𝐵 is totally definite, so 𝐹 is
totally real. Prove the following statements.
(a) If O1 does not contain an element of order 4, then O1 is √
cyclic.
(b) If O1 is quaternion 𝑄 8 ' 2𝐷 4 or exceptional, then 𝑅[ −1] ↩→ O and
discrd(O) is only divisible by primes
√ dividing 2.
(c) If O1 ' 2𝐷 2𝑚 with 𝑚 ≥ 3, then 𝑅[ −1], 𝑅[𝜁2𝑚 ] ↩→ O and discrd(O) is
−1 .
only divisible by primes dividing 𝜆22𝑚 − 4, where 𝜆2𝑚 := 𝜁2𝑚 + 𝜁2𝑚
8. Continuing further with the previous exercise, we compare O× and O1 .
(a) Show that [O× : O1 ] = 1, 2, 4. [Hint: Use Exercise 32.1.]
(b) Show that if O1 ' 2𝑂, 2𝐼, then O× = O1 .
(c) Show that if O1 ' 2𝑇, then [O× : O1 ] ≤ 2, and equality holds if and only
if there exists 𝛾 ∈ (1 + 𝑖)𝐹 × ∩ O× such that nrd(𝛾) ∉ 𝐹 ×2 .
[For a complete account covering all cases, see Vignéras-Guého [VG74].]
Chapter 33
Hyperbolic plane
591
592 CHAPTER 33. HYPERBOLIC PLANE
a metric when this infimum exists (i.e., there exists a path of finite length 𝑥 → 𝑦) for
all 𝑥, 𝑦 ∈ 𝑋. If the distance on 𝑋 is of the form (33.2.2), we call 𝑋 a length metric
space or a path metric space, and by construction ℓ(𝑔𝜐) = ℓ(𝜐) for all paths 𝜐 and
𝑔 ∈ Isom(𝑋).
Example 33.2.3. The space 𝑋 = R𝑛 with the ordinary Euclidean metric is a path
metric space; it is sometimes denoted E𝑛 as Euclidean space (to emphasize the role of
the metric).
33.2.4. If 𝑋 is a path metric space and 𝜐 achieves the infimum in (33.2.2), then we say
𝜐 is a geodesic segment in 𝑋. A geodesic is a continuous map (−∞, ∞) → 𝑋 such
that the restriction to every compact interval defines a geodesic segment. If 𝑋 is a path
33.2. GEODESIC SPACES 593
metric space such that every two points in 𝑋 are joined by a geodesic segment, we say
𝑋 is a geodesic space, and if this geodesic is unique we call 𝑋 a uniquely geodesic
space.
33.2.6. In the context of differential geometry (our primary concern), these notions
can be made concrete with coordinates. Suppose 𝑈 ⊆ R𝑛 is an open subset. Then
a convenient way to specify the length of a path in 𝑈 is with a length element in
real-valued coordinates. To illustrate, the ordinary metric on R𝑛 is given by the length
element √︃
d𝑠 := d𝑥 12 + · · · + d𝑥 𝑛2 ,
as usual.
More generally, if 𝜆 : 𝑈 → R>0 is a positive continuous function, then the length
element 𝜆(𝑥) d𝑠 defines a metric (33.2.2) on 𝑈, as follows. The associated length
(33.2.7) is symmetric, nonnegative, and satisfies the triangle inequality. To show that
𝜌(𝑥, 𝑦) > 0 when 𝑥 ≠ 𝑦, by continuity 𝜆 is bounded below by some 𝜂 > 0 on a suitably
small 𝜖 ball neighborhood of 𝑥 not containing 𝑦, so every path 𝜐 : 𝑥 → 𝑦 has ℓ(𝜐) ≥ 𝜖𝜂
and 𝜌(𝑥, 𝑦) > 0.
In this context, we also have a notion of orientation, and we may restrict to
isometries that preserve this orientation. We return to this in section 33.8, rephrasing
this in terms of Riemannian geometry.
Remark 33.2.8. The more general study of geometry based on the notion of length in a
topological space (the very beginnings of which are presented here) is the area of metric
geometry. Metric geometry has seen significant recent applications in group theory and
dynamical systems, as well as many other areas of mathematics. For further reading,
see the texts by Burago–Burago–Ivanov [BBI2001] and Papadopoulous [Pap2014].
In particular, geodesic spaces are quite common in mathematics, including com-
plete Riemannian manifolds; Busemann devotes an entire book to the geometry of
geodesics [Bus55]. Uniquely geodesic spaces are less common; examples include
simply connected Riemannian manifolds without conjugate points, CAT(0) spaces,
and Busemann convex spaces.
The following theorem nearly characterizes geodesic spaces.
594 CHAPTER 33. HYPERBOLIC PLANE
H = H2
since
(𝑎𝑧 + 𝑏) (𝑐𝑧 + 𝑑) 𝑎𝑐|𝑧| 2 + 𝑎𝑑𝑧 + 𝑏𝑐𝑧 + 𝑏𝑑
𝑔𝑧 = = (33.3.8)
|𝑐𝑧 + 𝑑| 2 |𝑐𝑧 + 𝑑| 2
we have
det 𝑔
Im 𝑔𝑧 = Im 𝑧. (33.3.9)
|𝑐𝑧 + 𝑑| 2
and so if Im 𝑧 > 0, then Im 𝑔𝑧 > 0 if and only if det 𝑔 > 0. Therefore, the subgroup
GL+2 (R) = {𝑔 ∈ GL2 (R) : det(𝑔) > 0}
preserves the upper half-plane H2 . Moreover, because the action of GL+2 (R) is holo-
morphic, it is orientation-preserving.
The kernel of this action, those matrices acting by the identity as linear fractional
transformations, are the scalar matrices, since (𝑎𝑧 + 𝑏)/(𝑐𝑧 + 𝑑) = 𝑧 identically if
and only if 𝑐 = 𝑏 = 0 and 𝑎 = 𝑑. Taking the quotient we get a faithful action of
PGL+2 (R) = GL+2 (R)/R× on H2 . There is a canonical isomorphism
∼ PSL (R)
PGL+ (R) −→
2 2
1
𝑔 ↦→ √︁ 𝑔.
det(𝑔)
with the same action on the upper half-plane.
33.3.10. The determinant det : PGL2 (R) → R× /R×2 ' {±1} has the inverse image
of +1 equal to PGL+2 (R) both open and closed in PGL2 (R); therefore, any 𝑔 with
det(𝑔) < 0 together with PGL+2 (R) generates PGL2 (R): for example, we may take
−1 0
𝑔= (33.3.11)
0 1
In view of (33.3.9), we extend the action of PGL2 (R) on H2 by defining for 𝑔 ∈
PGL2 (R) and 𝑧 ∈ H2
(
𝑔𝑧, if det 𝑔 > 0;
𝑔·𝑧= (33.3.12)
𝑔𝑧 = 𝑔𝑧, if det 𝑔 < 0.
The elements 𝑔 ∈ PGL2 (R) with det 𝑔 < 0 act anti-holomorphically and so are
orientation-reversing. The matrix 𝑔 in (33.3.11) then acts by 𝑔(𝑧) = −𝑧.
This action also arises naturally from another point of view. Let
H2− = {𝑧 ∈ C : Im 𝑧 < 0}
be the lower half-plane, let H2+ = H2 , and let
H2± = H2+ ∪ H2− = {𝑧 ∈ C : Im 𝑧 ≠ 0} = C − R.
Then PGL2 (R) acts on H2± (it preserves R hence also its complement in C). Complex
conjugation 𝑧 ↦→ 𝑧 interchanges H2+ and H2− , and there is a canonical identification
∼ H2
H2± /h i −→
from which we obtain the action (33.3.12) of PGL2 (R) on H2 .
596 CHAPTER 33. HYPERBOLIC PLANE
Remark 33.3.13. The fact that PSL2 (R) has elements 𝑔 ∈ PSL2 (R) that are matrices
up to sign means that whenever we do a computation with a choice of matrix, implicitly
we are also checking that the computation goes through with the other choice of sign.
Most of the time, this is harmless—but in certain situations this sign plays an important
role!
Let Isom+ (H2 ) ≤ Isom(H2 ) be the subgroup of isometries of H2 that preserve
orientation.
Theorem 33.3.14. The group PSL2 (R) acts on H2 via orientation-preserving isome-
tries, i.e., PSL2 (R) ↩→ Isom+ (H2 ).
Proof. Because the metric is defined by a length element d𝑠, we want to show that
d(𝑔𝑠) = d𝑠 for all 𝑔 ∈ PSL2 (R), i.e.,
|d(𝑔𝑧)| |d𝑧|
=
Im(𝑔𝑧) Im 𝑧
for all 𝑔 ∈ PSL2 (R). Since |d(𝑔𝑧)| = |d𝑔(𝑧)/d𝑧||d𝑧|, it is equivalent to show that
|d𝑔(𝑧)/d𝑧| 1
= (33.3.15)
Im 𝑔𝑧 Im 𝑧
by (33.3.9),
Im 𝑧
Im 𝑔𝑧 = ,
|𝑐𝑧 + 𝑑| 2
so taking the ratio, the two factors |𝑐𝑧 + 𝑑| 2 exactly cancel, establishing (33.3.15).
Finally, the action is holomorphic so (by the Cauchy–Riemann equations) it lands
in the orientation-preserving subgroup.
33.3.17. The action of PSL2 (R) extends to the boundary as follows. We define the
circle at infinity to be the boundary
bd H2 := R ∪ {∞} ⊆ C ∪ {∞}.
H2∗ := H2 ∪ bd H2 .
33.4. CLASSIFICATION OF ISOMETRIES 597
The topology on H2∗ is the same as the Euclidean topology on H2 , and we take a
fundamental system of neighborhoods of the point ∞ to be sets of the form
{𝑧 ∈ H2 : Im 𝑧 > 𝑀 } ∪ {∞}
𝑎𝑖 + 𝑏
We have 𝐾 = StabSL2 (R) (𝑖) since = 𝑖 if and only if 𝑑 = 𝑎 and 𝑐 = −𝑏, and
𝑐𝑖 + 𝑑
𝑎 0
then the determinant condition implies 𝑎 2 + 𝑏 2 = 1. An element acts by
0 1/𝑎
𝑧 ↦→ 𝑎2 𝑧, fixing the origin and stretching along lines through the origin. An element
1 𝑛
acts by the translation 𝑧 ↦→ 𝑧 + 𝑏.
0 1
𝑁 𝐴 ∩ 𝐾 = {1} = 𝑁 ∩ 𝐴.
This map is surjective as follows. Let 𝑔 ∈ SL2 (R), and let 𝑧 = 𝑔(𝑖). Let 𝑛𝑔 =
√
1 − Re 𝑧 1/ 𝑦 0
∈ 𝑁, so that (𝑛𝑔 𝑔) (𝑖) = 𝑦𝑖. Let 𝑎 𝑔 = √ ∈ 𝐴; then
0 1 0 𝑦
(𝑎 𝑔 𝑛𝑔 𝑔) (𝑖) = 𝑖, so 𝑎 𝑔 𝑛𝑔 𝑔 ∈ StabSL2 (R) (𝑖) = 𝐾, and peeling back we get 𝑔 ∈ 𝑁 𝐴𝐾,
proving surjectivity.
Lemma 33.4.4. The group SL2 (R) is generated by the subgroups 𝐴, 𝑁, and the
0 −1
element , which acts on H2 by 𝑧 ↦→ −1/𝑧.
1 0
𝑎 𝑏
Proof. Let ∈ SL2 (R). The lemma follows by performing row reduction on
𝑐 𝑑
the matrix using the given generators. We find that if 𝑐 ≠ 0, then
𝑎 𝑏 1 𝑎/𝑐 0 −1 𝑐 0 1 𝑑/𝑐
=
𝑐 𝑑 0 1 1 0 0 1/𝑐 0 1
and if 𝑐 = 0 then
𝑎 𝑏 𝑎 0 1 𝑏/𝑎
= .
0 1/𝑎 0 1/𝑎 0 1
|Tr(𝐾)| = [0, 2], |Tr( 𝐴)| = [2, ∞), and |Tr(𝑁)| = {2}.
33.4. CLASSIFICATION OF ISOMETRIES 599
Every nonidentity element 𝑔 ∈ PSL2 (R) belongs to one of these three types (even
though 𝑔 need not belong to one of the subgroups 𝑁, 𝐴, 𝐾 individually).
elliptic
a unique fixed point in H2 ,
hyperbolic if and only if 𝑔 has two fixed points on bd H2 ,
a unique fixed point on bd H2 .
parabolic
𝑎 𝑏
Proof. Let 𝑔 = have det(𝑔) = 𝑎𝑑 − 𝑏𝑐 = 1. We look to solve the equation
𝑐 𝑑
𝑎𝑧 + 𝑏
=𝑧
𝑐𝑧 + 𝑑
elliptic
33.5 Geodesics
In this section, we prove two important theorems: we describe geodesics, giving a
formula for the distance, and we classify isometries.
Theorem 33.5.1. The hyperbolic plane H2 is a uniquely geodesic space. The unique
geodesic passing through two distinct points 𝑧, 𝑧 0 ∈ H2 is a semicircle orthogonal to
R or a vertical line, and
| 𝑧 − 𝑧0| + | 𝑧 − 𝑧0|
𝜌(𝑧, 𝑧 0) = log (33.5.2)
| 𝑧 − 𝑧0| − | 𝑧 − 𝑧0|
| 𝑧 − 𝑧 0 |2
cosh 𝜌(𝑧, 𝑧 0) = 1 + . (33.5.3)
2 Im(𝑧) Im(𝑧 0)
33.5. GEODESICS 601
Proof. We first prove the imaginary axis is a geodesic with 𝑧, 𝑧 0 ∈ R>0𝑖. Let 𝜐(𝑡) =
𝑥(𝑡) + 𝑖𝑦(𝑡) : 𝑧 → 𝑧 0 be a path; then
∫ 1 √︁ ∫ 1
(d𝑥/d𝑡) 2 + (d𝑦/d𝑡) 2 d𝑦/d𝑡
ℓ(𝜐) = d𝑡 ≥ d𝑡
0 𝑦(𝑡) 0 𝑦(𝑡)
0 (33.5.4)
𝑧
= log 𝑦(1) − log 𝑦(0) = log
𝑧
with equality if and only if 𝑥(𝑡) = 0 identically and d𝑦/d𝑡 ≥ 0 for all 𝑡 ∈ [0, 1]. This
is achieved for the path
𝜐(𝑡) = (|𝑧|(1 − 𝑡) + |𝑧 0 |𝑡)𝑖;
so 𝜌(𝑧, 𝑧 0) = log | 𝑧 0/𝑧|, and the imaginary axis is the unique geodesic.
For arbitrary points 𝑧, 𝑧 0 ∈ H2 , we apply Lemma 33.4.11. The statement on
geodesics follows from the fact that the image of R>0𝑖 under an element of PSL2 (R)
is either a semicircle orthogonal to R or a vertical line (Exercise 33.6).
The formula (33.5.2) for the case 𝑧, 𝑧 0 ∈ R>0𝑖 follows from (33.5.4) and plugging
in along the imaginary axis; the general case then follows from the invariance of both
𝜌(𝑧, 𝑧 0) and
| 𝑧 − 𝑧0| + | 𝑧 − 𝑧0|
log
| 𝑧 − 𝑧0| − | 𝑧 − 𝑧0|
under 𝑔 ∈ PSL2 (R), checked on the generators in Lemma 33.4.4 (Exercise 33.10).
Finally, the formula (33.5.3) follows directly from formulas for hyperbolic cosine,
requested in Exercise 33.11.
Theorem 33.5.5. We have
Isom(H2 ) ' PGL2 (R)
and
Isom+ (H2 ) ' PGL+2 (R) ' PSL2 (R).
Proof. Let 𝑍 = {𝑡𝑖 : 𝑡 > 0} be the positive part of the imaginary axis. By Theorem
33.5.1, 𝑍 is the unique geodesic through any two points of 𝑍.
Let 𝜙 ∈ Isom(H2 ). Then 𝜙(𝑍) is a geodesic (33.2.5), so by Exercise 33.7, there
exists 𝑔 ∈ PSL2 (R) such that 𝑔𝜙 fixes 𝑍 pointwise. Replacing 𝜙 by 𝑔𝜙, we may
suppose 𝜙 fixes 𝑍 pointwise.
Let 𝑧 = 𝑥 + 𝑖𝑦 ∈ H2 and 𝑧 0 = 𝑥 0 + 𝑖𝑦 0 = 𝜙(𝑧). For all 𝑡 > 0,
𝜌(𝑧, 𝑖𝑡) = 𝜌(𝜙𝑧, 𝜙(𝑖𝑡)) = 𝜌(𝑧 0, 𝑖𝑡).
Plugging this into the formula (33.5.3) for the distance, we obtain
(𝑥 2 + (𝑦 − 𝑡) 2 )𝑦 0 = (𝑥 02 + (𝑦 0 − 𝑡) 2 )𝑦.
Dividing both sides by 𝑡 2 and taking the limit as 𝑡 → ∞, we find that 𝑦 = 𝑦 0, and
consequently that 𝑥 2 = 𝑥 02 and 𝑥 = ±𝑥 0. The choice of sign ± varies continuously over
the connected set H2 and so must be constant. Therefore 𝜙(𝑧) = 𝑧 or 𝜙(𝑧) = −𝑧 for all
𝑧 ∈ H2 . The latter generates PGL2 (R) over PSL2 (R) (33.3.10), and both statements
in the theorem follow.
602 CHAPTER 33. HYPERBOLIC PLANE
d𝑥 d𝑦
d𝐴 = .
𝑦2
Proof. We verify that the hyperbolic area form is invariant. We first check this for the
orientation-reversing isometry
𝑎𝑧 + 𝑏
𝑤 = 𝑔(𝑧) = = 𝑢 + 𝑖𝑣,
𝑐𝑧 + 𝑑
𝑦
with 𝑎𝑑 − 𝑏𝑐 = 1. By (33.3.9), 𝑣 = . We compute that
|𝑐𝑧 + 𝑑| 2
d𝑔 1
(𝑧) = . (33.6.3)
d𝑧 (𝑐𝑧 + 𝑑) 2
Now 𝑔 is holomorphic; so by the Cauchy–Riemann equations, its Jacobian is given by
𝜕 (𝑢, 𝑣) d𝑔 2
= (𝑧) = 1
;
𝜕 (𝑥, 𝑦) d𝑧 |𝑐𝑧 + 𝑑| 4
therefore
d𝑢 d𝑣 𝜕 (𝑢, 𝑣) d𝑥 d𝑦 1 |𝑐𝑧 + 𝑑| 4
d(𝑔 𝐴) = = = d𝑥 d𝑦 = d𝐴.
𝑣2 𝜕 (𝑥, 𝑦) 𝑣 2 |𝑐𝑧 + 𝑑| 4 𝑦2
A hyperbolic triangle is visibly convex; for more on convexity, see Exercises 33.8–
33.9. The following key formula expresses the hyperbolic area of a triangle in terms
of its angles.
Theorem 33.6.8 (Gauss–Bonnet formula). Let 𝑇 be a hyperbolic triangle with angles
𝛼, 𝛽, 𝛾. Then
𝜇(𝑇) = 𝜋 − (𝛼 + 𝛽 + 𝛾).
𝜇(𝑃) = (𝑛 − 2)𝜋 − (𝜃 1 + · · · + 𝜃 𝑛 ).
surface in one place, the curvature is forced to rise somewhere else. If instead one has
a surface 𝑋 with geodesic boundary, then the formula (33.6.11) becomes
∫ ∑︁
𝐾 d𝐴 + (𝜋 − 𝜃 𝑖 ) = 2𝜋 𝜒(𝑋)
𝑋 𝑖
Definition 33.7.1. The hyperbolic unit disc is the (open) unit disc
D2 = {𝑤 ∈ C : |𝑤| < 1}
2|d𝑤|
d𝑠 = .
1 − |𝑤| 2
The hyperbolic unit disc D2 is also called the Poincaré model of planar hyperbolic
geometry. The circle at infinity is the boundary
bd D2 = {𝑤 ∈ C : |𝑤| = 1}.
𝑧−𝑖 𝑤+1
𝜙(𝑧) = , 𝜙−1 (𝑤) = −𝑖 . (33.7.4)
𝑧+𝑖 𝑤−1
upper half-plane and unit disc as each has its advantage, we find it notationally simpler
to avoid this extra decoration. The distance on D2 is
|1 − 𝑤𝑤 0 | + |𝑤 − 𝑤 0 |
𝜌(𝑤, 𝑤 0) = log
|1 − 𝑤𝑤 0 | − |𝑤 − 𝑤 0 |
(33.7.5)
|𝑤 − 𝑤 0 | 2
cosh 𝜌(𝑤, 𝑤 0) = 1 + 2
(1 − |𝑤| 2 ) (1 − |𝑤 0 | 2 )
so that
1 + |𝑤|
𝜌(𝑤, 0) = log = 2 tanh−1 |𝑤|. (33.7.6)
1 − |𝑤|
The map 𝜙 (33.7.3) maps the geodesics in H2 to geodesics in D2 , and as a Möbius
transformation, maps circles and lines to circles and lines, preserves angles, and maps
the real axis to the unit circle; therefore the geodesics in D2 are diameters through the
origin and semicircles orthogonal to the unit circle, as in Figure 33.7.7.
t2 − x2 − y 2 = 1
L2
cone
t2 = x2 + y 2
By pullback, Isom+ (L2 ) ' PSL2 (R). However, other isometries are also apparent:
a linear change of variables that preserves the quadratic form 𝑞 also preserves the
Lorentz metric. Let
−1 0 0
O(2, 1) = {𝑔 ∈ GL3 (R) : 𝑔 𝑡 𝑚𝑔 = 𝑚}, where 𝑚 = 0 1 0®.
© ª
«0 0 1¬
Then
d𝑠2 = 𝑣 𝑡 𝑚𝑣, where 𝑣 = (d𝑡, d𝑥, d𝑦) 𝑡 ,
so if 𝑔 ∈ O(2, 1) then immediately d𝑔𝑠2 = d𝑠2 .
Next, if 𝑔 ∈ O(2, 1), then 𝑔 𝑡 𝑚𝑔 = 𝑚 implies det(𝑔) = ±1. The elements of O(2, 1)
that map the hyperboloid to itself comprise the subgroup
let SO+ (2, 1) ≤ SO(2, 1) be the further subgroup that maps the upper sheet of the
hyperboloid to itself, the connected component of the identity.
Remark 33.7.14. We have proven that there is an isomorphism of Lie groups
for standard coordinates 𝑥𝑖 on R𝑛 , and the matrix (𝜂𝑖 𝑗 ) is symmetric, positive definite,
and differentiable. The metric determines a volume formula as
√︁
d𝑉 = det 𝜂 d𝑥1 · · · d𝑥 𝑛 .
T(H2 ) := {(𝑧, 𝑣) : 𝑧 ∈ H2 , 𝑣 ∈ T𝑧 H2 }
Since isometries of H2 are differentiable, they act on the tangent bundle by differentials
preserving the norm and angle, and therefore Isom(H2 ) acts conformally or anti-
conformally on H2 .
If we restrict to the unit tangent bundle
(Exercise 33.17).
Remark 33.8.3. The natural generalization of Euclid’s geometry is performed on a
Riemannian manifold 𝑋 that is homogeneous, i.e., the isometry group Isom(𝑋) acts
transitively on 𝑋, as well as isotropic, i.e., Isom(𝑋) acts transitively on frames (a basis
of orthonormal tangent vectors) at a point. In this way, homogeneous says that every
point “looks the same”, and isotropic says that the geometry “looks the same in every
33.8. RIEMANNIAN GEOMETRY 609
direction” at a point. Taken together, these natural conditions are quite strong, and
there are only three essentially distinct simply connected homogeneous and isotropic
geometries in any dimension, corresponding to constant sectional curvatures zero,
positive, or negative: these are Euclidean, spherical, and hyperbolic geometry, respec-
tively. Put this way, the hyperbolic plane is the unique complete, simply connected
Riemann surface with constant sectional curvature −1. For more on geometries in this
sense, we encourage the reader to consult Thurston [Thu97].
To conclude this section, we briefly review a few facts from the theory of Rieman-
nian manifolds.
33.8.4. A (topological) 𝑛-manifold is a (second-countable) Hausdorff topological
space 𝑋 locally homeomorphic to R𝑛 , i.e., for every 𝑥 ∈ 𝑋, there exists an open
neighborhood 𝑈 3 𝑥 and a continuous map 𝜙 : 𝑈 ↩→ R𝑛 that is a homeomorphism
onto an open subset; the map 𝜙 : 𝑈 → 𝜙(𝑈) ⊆ R𝑛 is called a chart (at 𝑥 ∈ 𝑋), and an
open cover of charts is called an atlas.
Two charts 𝜙1 : 𝑈1 → R𝑛 and 𝜙2 : 𝑈2 → R𝑛 are (𝐶 ∞ -)smoothly compatible if the
transition map
𝜙12 = 𝜙2 𝜙−1
1 : 𝜙1 (𝑈1 ∩ 𝑈2 ) → 𝜙2 (𝑈1 ∩ 𝑈2 )
(see Figure 33.8.5) is 𝐶 ∞ -smooth (i.e., has continuous partial derivatives of all orders).
An atlas on a manifold is smooth on a manifold if all charts are smoothly compatible.
A smooth manifold is a manifold equipped with a smooth atlas.
X
U2
U1
φ1 φ2
φ12
φ2 (U2 )
φ1 (U1 )
Rn Rn
Figure 33.8.5: A manifold, by its atlas
A morphism of smooth manifolds is a continuous map 𝑓 : 𝑌 → 𝑋 such that for
the atlases {(𝜙𝑖 , 𝑈𝑖 )}𝑖 of 𝑋 and {(𝜓 𝑗 , 𝑉 𝑗 )} 𝑗 of 𝑌 , each map
𝜙𝑖 𝑓 𝜓 −1
𝑗 : 𝜓 𝑗 (𝑉 𝑗 ∩ 𝑓
−1
(𝑈𝑖 )) → 𝜙𝑖 ( 𝑓 (𝑉 𝑗 ) ∩ 𝑈𝑖 )
610 CHAPTER 33. HYPERBOLIC PLANE
Exercises
1. Show that the hyperbolic metric has the same topology as the Euclidean metric
in two ways.
(a) Show directly that open balls nest: for all 𝑧 ∈ H2 and all 𝜖 > 0, there exist
𝜂1 , 𝜂2 > 0 such that
for all 𝑤 ∈ H2 .
(b) Show that the collection of Euclidean balls coincides with the collection
of hyperbolic balls. [Hint: applying an isometry, reduce to the case of
balls around 𝑖 and check this directly; it is perhaps even clearer moving to
the unit disc model.]
33.8. RIEMANNIAN GEOMETRY 611
for suitably nice functions 𝑓 , 𝑔. Using this formula, verify that the curvature of
H2 and D2 is −1.
4. Consider C with the standard metric. Let
Isomℎ (C) = {𝑔 ∈ Isom(C) : 𝑔 is holomorphic} ≤ Isom(C).
Exhibit an isomorphism of groups
ℎ 𝑎 𝑏
Isom (C) ' ∈ GL2 (C) : |𝑎| = 1
0 1
and an isometry of C that is not holomorphic. [Hint: An invertible holomorphic
map C → C is of the form 𝑧 ↦→ 𝑎𝑧 + 𝑏.]
I 5. Show that for every 𝑧, 𝑧 0 ∈ H2 , there exists 𝑔 ∈ PSL2 (R) such that Re 𝑔𝑧 =
Re 𝑔𝑧 0 = 0. [Hint: Work in D2 .]
I 6. Show that the image of R>0𝑖 under an element of PSL2 (R) is either a semicircle
orthogonal to R or a vertical line. [Hint: Look at the endpoints.]
I 7. We consider the action of PSL2 (R) on geodesics in H2 .
(a) Show that PSL2 (R) acts transitively on the set of geodesics in H2 .
(b) Show that given 𝑧, 𝑧 0 ∈ H2 , there exists 𝑔 ∈ PSL2 (R) such that 𝑔𝑧 = 𝑧 0
and such that 𝑔 that maps the geodesic through 𝑧 and 𝑧 0 to itself.
(c) Show that every isometry of H2 that maps a geodesic to itself and fixes
two points on this geodesic is the identity.
I 8. Let 𝑧 1 , 𝑧2 ∈ H2 be distinct. Let
𝐻 (𝑧1 , 𝑧2 ) = {𝑧 ∈ H2 : 𝜌(𝑧, 𝑧1 ) ≤ 𝜌(𝑧, 𝑧2 )}
be the locus of points as close to 𝑧1 as to 𝑧 2 , and let 𝐿(𝑧 1 , 𝑧2 ) = bd 𝐻 (𝑧 1 , 𝑧2 ).
Show that 𝐻 (𝑧 1 , 𝑧2 ) is a convex (Definition 33.6.6) half-plane, and that
𝐿 (𝑧1 , 𝑧2 ) = {𝑧 ∈ H2 : 𝜌(𝑧, 𝑧1 ) = 𝜌(𝑧, 𝑧2 )}
is geodesic and equal to the perpendicular bisector of the geodesic segment from
𝑧1 to 𝑧2 .
612 CHAPTER 33. HYPERBOLIC PLANE
(𝜙𝑔𝜙−1 ) (𝑤) = 𝑤
L2 → D2
𝑥 + 𝑖𝑦
(𝑡, 𝑥, 𝑦) ↦→
𝑡+1
identifies the metrics on L2 and D2 , via pullback.
I 16. Show that the Lorentz metric restricted to the hyperboloid is an honest (Rie-
mannian) metric. [Hint: Show that a tangent vector 𝑣 at a point 𝑝 satisfies
𝑏( 𝑝, 𝑣) = 0, where 𝑏 is the bilinear form associated to 𝑞; then show that the
orthogonal complement to 𝑝 has signature +2.]
33.8. RIEMANNIAN GEOMETRY 613
Our ongoing goal in this part of the text is to understand quotient spaces that locally
look like (products of) hyperbolic spaces. In order to get off the ground, here we put
the previous two chapters in a more general context, seeking to understand nice group
actions on topological spaces and indicating how these fit in with more general notions
in topology. Pathologies exist! Our goal in this chapter is to provide basic context (for
further references see 34.5.4) before turning to the central case of interest: a discrete
group acting properly on a locally compact, Hausdorff topological space.
𝐺×𝑋→𝑋
(34.1.1)
(𝑔, 𝑥) ↦→ 𝑔𝑥
satisfying 1𝑥 = 𝑥 and (𝑔𝑔 0)𝑥 = 𝑔(𝑔 0𝑥) for all 𝑥 ∈ 𝑋 and all 𝑔, 𝑔 0 ∈ 𝐺. A right action
is instead a map 𝑋 × 𝐺 → 𝑋 by (𝑥, 𝑔) ↦→ 𝑥𝑔 satisfying 𝑥1 = 𝑥 and 𝑥(𝑔 0 𝑔) = (𝑥𝑔 0)𝑔
for all 𝑥 ∈ 𝑋 and 𝑔, 𝑔 0 ∈ 𝑔. We will need to consider actions both on the left and the
right; if not specified, a left action is assumed.
We will also sometimes write 𝐺 𝑋 for an action of 𝐺 on 𝑋.
Example 34.1.2. A group 𝐺 acts on itself by left multiplication, the (left) regular
group action of 𝐺. If 𝐻 ≤ 𝐺 is a subgroup, then 𝐻 also acts on 𝐺 by left multiplication.
For example, if 𝑉 is an R-vector space with dimR 𝑉 = 𝑛, and Λ ⊆ 𝑉 is a (full) Z-lattice
in 𝑉, then Λ ' Z𝑛 is a group and Λ acts on 𝑉 by translation.
Another important and related example is the left action of 𝐺 on the set of right
cosets 𝑋 = 𝐺/𝐻 again by multiplication, namely
615
616 CHAPTER 34. DISCRETE GROUP ACTIONS
𝐺\𝑋 0 → 𝐺\𝑋
𝐺𝑥 0 ↦→ 𝐺 𝑓 (𝑥 0),
𝑋0
𝑓
/𝑋
𝜋0 𝜋 (34.1.7)
𝐺\𝑋 0 / 𝐺\𝑋
𝜆𝑔 : 𝑋 → 𝑋
𝑥 ↦→ 𝑔𝑥
34.1.11. The quotient 𝐺\𝑋 is equipped with the quotient topology, so that the quotient
map 𝜋 : 𝑋 → 𝐺\𝑋 is continuous: a subset 𝑉 ⊆ 𝐺\𝑋 is open if and only if 𝜋 −1 (𝑉) ⊆ 𝑋
is open.
The projection 𝜋 is an open map, which is to say ifÐ𝑈 ⊆ 𝑋 is open then 𝜋(𝑈) =
𝐺𝑈 ⊆ 𝐺\𝑋 is open: if 𝑈 is open then 𝜋 −1 (𝜋(𝑈)) = 𝑔 ∈𝐺 𝑔𝑈 is open, so 𝜋(𝑈) is
open by definition of the topology.
𝐺 → 𝐺𝑥 ⊆ 𝑋
𝑔 ↦→ 𝑔𝑥
is continuous (it is the restriction of the action map to 𝐺 × {𝑥}). Let 𝐾 = Stab𝐺 (𝑥).
Then this map factors naturally as
∼ 𝐺𝑥
𝐺/𝐾 −→ (34.1.13)
where we give 𝐺/𝐾 the quotient topology; then (34.1.13) then a bijective continuous
map, a topological upgrade of the orbit–stabilizer theorem. The map (34.1.13) need
not always be a homeomorphism (Exercise 34.5), but we will see below that it becomes
a homeomorphism under further nice hypotheses (Exercise 34.6, Proposition 34.4.11).
(i) cl(int(◊)) = ◊;
(ii) 𝐺◊ = 𝑋; and
(iii) int(◊) ∩ int(𝑔◊) = ∅ for all 1 ≠ 𝑔 ∈ 𝐺.
618 CHAPTER 34. DISCRETE GROUP ACTIONS
The condition (i) ensures our basic intuition about tilings (and avoids fundamental
sets that contain an extraneous number of isolated points); condition (ii) says that ◊
tiles 𝑋; and condition (iii) shows that the tiles only overlap along the boundary, and
there is no redundancy in the interior. If there is a fundamental set for 𝐺 𝑋, then
the action is faithful.
Theorem 34.2.1. Let Γ ≤ PSL2 (R) be a subgroup and equip Γ with the subspace
topology. Then the following are equivalent:
(i) Γ is discrete;
(ii) For all 𝑧 ∈ H2 , we have # StabΓ (𝑧) < ∞ and there exists an open neigborhood
𝑈 3 𝑧 such that 𝛾𝑈 ∩ 𝑈 ≠ ∅ implies 𝛾 ∈ StabΓ (𝑧);
(iii) For all compact subsets 𝐾 ⊆ H2 , we have 𝐾 ∩ 𝛾𝐾 ≠ ∅ for only finitely many
𝛾 ∈ Γ; and
(iv) For all 𝑧 ∈ H2 , the orbit Γ𝑧 ⊆ H2 is discrete and # StabΓ (𝑧) < ∞.
Moreover, if these equivalent conditions hold, then the quotient Γ\H2 is Hausdorff,
and the quotient map 𝜋 : H2 → Γ\H2 is a local isometry at all points 𝑧 ∈ H2 with
StabΓ (𝑧) = {1}.
Definition 34.3.1. We say the action of 𝐺 on 𝑋 is a covering space action if for all
𝑥 ∈ 𝑋, there exists an open neighborhood 𝑈 3 𝑥 such that 𝑔𝑈 ∩ 𝑈 = ∅ for all 𝑔 ∈ 𝐺
with 𝑔 ≠ 1.
34.3. COVERING SPACE AND WANDERING ACTIONS 619
Proof. The implication (ii) ⇒ (i) is immediate; we prove the converse. Let 𝑈 be a
neighborhood of 𝑥 ∈ 𝑋 such that 𝑔𝑈 ∩ 𝑈 ≠ ∅ for only finitely many 𝑔 ∈ 𝐺. We have
# Stab𝐺 (𝑥) < ∞ since 𝑔 ∈ Stab𝐺 (𝑥) implies 𝑥 ∈ 𝑔𝑈 ∩ 𝑈. Let
𝑈𝑖 ∩ 𝑔𝑖 𝑈𝑖 ⊆ 𝑈𝑖 ∩ 𝑔𝑖 𝑊𝑖0 ⊆ 𝑉𝑖 ∩ 𝑊𝑖 = ∅. (34.3.8)
and the latter map Stab𝐺 (𝑥)\𝑈 → 𝐺\𝑋 is a homeomorphism onto its image; we say
𝜋 is a local homeomorphism modulo stabilizers. If the action of 𝐺 is free, then we
recover 34.3.2.
Remark 34.3.10. If 𝐺 has the discrete topology and the condition in Lemma 34.3.6(ii)
holds, then some authors call the action of 𝐺 properly discontinuous. This is probably
because 𝐺 is then as broken (“discontinuous”) as possible: 𝐺 has the discrete topology,
and we should be able to find neighborhoods that pull apart the action of 𝐺. (Klein
[Kle79, p. 321] uses the term discontinuous because “points that are ‘equivalent’ with
respect to [the group] are separated”.) This nomenclature is strange because we still
want the action to be continuous, just by a discrete group. Adding to the potential
confusion is the issue that different authors give different definitions of “properly
discontinuous” depending on their purposes; most of these can be seen to be equivalent
under the right hypotheses on the space, but not all. We avoid this term.
It turns out that a wandering action is too weak a property in this level of generality
for us to work with. However, it is close, and we will shortly see that it suffices with
additional hypotheses on the space 𝑋.
Remark 34.3.11. Let 𝑋 be a topological space, and let 𝐺 be a set of continuous maps
𝑋 → 𝑋. Then there is a natural map 𝐺 ↩→ 𝑋 𝑋 defined by 𝑔 ↦→ (𝑔𝑥) 𝑥 . We give 𝑋 𝑋
the compact-open topology and 𝐺 the subspace topology, so a subbasis of the topology
on 𝐺 is given by
𝑉 (𝐾, 𝑈) = { 𝑓 ∈ 𝐺 : 𝑓 (𝐾) ⊆ 𝑈}
for 𝐾 ⊆ 𝑋 compact and 𝑈 ⊆ 𝑋 open.
If 𝑋 is Hausdorff and locally compact, then the compact-open topology on 𝐺 is
the weakest topology (smallest, fewest open sets) for which the map 𝐺 × 𝑋 → 𝑋
is continuous (also called an admissible topology on 𝐺) [McC2011, §VII, pp. 171–
172]. Under the hypotheses of Exercise 34.4, this implies that the topology of pointwise
convergence and the compact-open topology coincide.
𝐺×𝑋→𝑋×𝑋
(34.4.2)
(𝑔, 𝑥) ↦→ (𝑥, 𝑔𝑥)
is closed.
Proof. The implication (i) ⇔ (ii) follows directly from properties of the quotient map:
the preimage of open neighborhoods separating 𝐺𝑥 and 𝐺 𝑦 under the continuous
projection map have the desired properties, and conversely the pushforward of the
given neighborhoods under the open projection map separate 𝐺𝑥 and 𝐺 𝑦.
To conclude, we prove (i) ⇔ (iii). We use the criterion that a topological space is
Hausdorff if and only if the diagonal map has closed image. The continuous surjective
map 𝜋 : 𝑋 → 𝐺\𝑋 is open, so the same is true for
𝜋 × 𝜋 : 𝑋 × 𝑋 → (𝐺\𝑋) × (𝐺\𝑋).
Therefore the diagonal 𝐺\𝑋 ↩→ (𝐺\𝑋) × (𝐺\𝑋) is closed if and only if its preimage
is closed in 𝑋 × 𝑋. But this preimage consists exactly of the orbit relation
Lemma 34.4.5. Suppose that 𝑌 is locally compact and Hausdorff, and let 𝑓 : 𝑋 → 𝑌
be continuous and quasi-proper. Then 𝑋 is locally compact, and 𝑓 is proper.
Proof. For the first statement, cover 𝑌 with open relatively compact sets 𝑈𝑖 ⊆ 𝐾𝑖 ; then
𝑉𝑖 = 𝑓 −1 (𝑈𝑖 ) is an open cover of 𝑋 by relatively compact sets.
Next, we claim that 𝑓 is in fact already proper; that is to say, we show that 𝑓
is closed. Let 𝑊 ⊆ 𝑋 be a closed set and consider a sequence {𝑦 𝑛 } 𝑛 from 𝑓 (𝑊)
with 𝑦 𝑛 → 𝑦. Let 𝐾 be a compact neighborhood of 𝑦 containing {𝑦 𝑛 }; taking a
subsequence, we may suppose all 𝑦 𝑛 ∈ 𝐾. Let 𝑥 𝑛 ∈ 𝑓 −1 (𝑦 𝑛 ) ∩ 𝑊 be primages. Since
622 CHAPTER 34. DISCRETE GROUP ACTIONS
𝐺×𝑋→𝐺×𝑋×𝑋 →𝑋×𝑋
(𝑔, 𝑥) ↦→ (𝑔, 𝑥, 𝑔𝑥) ↦→ (𝑥, 𝑔𝑥);
the first map is the graph of a continuous map to a Hausdorff space and is closed
(Exercise 34.9); the second (projection) map is closed, as 𝐺 is compact (by a standard
application of the tube lemma). Therefore the composition of these maps is closed.
𝜄 𝑥 : 𝐺/Stab𝐺 (𝑥) → 𝐺𝑥
𝑔 ↦→ 𝑔𝑥 ∈ 𝑋
is a homeomorphism.
34.5. PROPER ACTIONS ON A LOCALLY COMPACT SPACE 623
Proof. For part (a), by Lemma 34.4.1, it is enough to note that by definition the image
of the action map 𝜆 in (34.4.8) is closed. Part (b) follows in the same way, as
This also implies part (c) (cf. 34.1.12): the map 𝜄 𝑥 is bijective and continuous, and
it is also closed (whence a homeomorphism) since 𝜄 𝑥 is a factor of the closed map
𝐺 → 𝐺𝑥.
Finally, for part (d), let 𝜆 : 𝐺 × 𝑋 → 𝑋 × 𝑋 be the action map and let 𝑥 ∈ 𝑋. Then
by definition that 𝜆−1 (𝑥, 𝑥) = Stab𝐺 (𝑥) × {𝑥} ' Stab𝐺 (𝑥), so by definition Stab𝐺 (𝑥)
is compact.
Proof. First, we show (i) ⇒ (ii). Let 𝜆 : 𝐺 × 𝑋 → 𝑋 × 𝑋 be the action map. Let
𝐾 ⊆ 𝑋 be compact. Then
𝜆−1 (𝐾 × 𝐾) = {(𝑔, 𝑥) ∈ 𝐺 × 𝑋 : 𝑥 ∈ 𝐾, 𝑔𝑥 ∈ 𝐾 }
𝐾 ∩ 𝑔𝐿 ⊆ (𝐾 ∪ 𝐿) ∩ 𝑔(𝐾 ∪ 𝐿) ≠ ∅
624 CHAPTER 34. DISCRETE GROUP ACTIONS
𝑊 = {𝑔 ∈ 𝐺 : 𝑔𝑈 ∩ 𝑉 ≠ ∅}
𝐺/Stab𝐺 (𝑥) → 𝐺𝑥
for 𝑛 sufficiently large. Since ℎ 𝑛 𝑥 ≠ 𝑥 for all 𝑛, this contradicts that the orbit 𝐺𝑥 is
discrete, having no limit points.
34.6. SYMMETRIC SPACE MODEL 625
34.5.3. From Lemma 34.3.6 and the implication Theorem 34.5.1(v) ⇒ (i), we see that
proper actions generalize covering space actions when 𝑋 is locally compact metric
space and 𝐺 acts by isometries. In fact, a more general statement is true: if 𝐺 is a
discrete group with a covering space action on 𝑋 such that 𝐺\𝑋 is Hausdorff, then 𝐺
acts properly on 𝑋. The (slightly involved) proof in general is requested in Exercise
34.16.
Remark 34.5.4. Bourbaki discusses proper maps [Bou60, Chapter I, §10] and more
generally groups acting properly on topological spaces [Bou60, Chapter III, §§1,4];
the definition of proper is equivalent to ours as follows. Let 𝑓 : 𝑋 → 𝑌 be continuous,
and say 𝑓 is Bourbaki proper to mean that 𝑓 × id : 𝑋 × 𝑍 → 𝑌 × 𝑍 is closed for every
topological space 𝑍. If 𝑓 is Bourbaki proper, then 𝑓 is proper [Bou60, Chapter I, §10,
Proposition 6]. In the other direction, if 𝑓 is proper then 𝑓 is closed and 𝑓 −1 (𝑦) is
compact for all 𝑦 ∈ 𝑌 , and this implies that 𝑓 is Bourbaki proper [Bou60, Chapter I,
§10, Theorem 1].
34.6.1. Let 𝐺 = SL2 (R). As a matrix group, 𝐺 comes with a natural metric. The
space M2 (R) ' R4 has the usual structure of a metric space, with
2 2 2 2 2 𝑎 𝑏
k𝑔k = 𝑎 + 𝑏 + 𝑐 + 𝑑 , if 𝑔 = ∈ M2 (R).
𝑐 𝑑
We give SL2 (R) ⊂ M2 (R) the subspace metric and PSL2 (R) the quotient metric.
Intuitively, in this metric 𝑔, ℎ ∈ PSL2 (R) are close if there exist matrices representing
𝑔, ℎ (corresponding to a choice of sign) with all four entries of the matrix close in R.
34.6.2. Recall from 34.1.12 that if 𝐺 acts (continuously and) transitively on 𝑋, then
for all 𝑥 ∈ 𝑋, the natural map 𝑔 ↦→ 𝑔𝑥 gives a continuous bijection
∼ 𝐺𝑥 = 𝑋.
𝐺/Stab𝐺 (𝑥) −→
Let 𝑋 = H2 be the hyperbolic plane and let 𝐺 = SL2 (R). Then 𝐺 acts transitively on
𝑋. The stabilizer of 𝑥 = 𝑖 is the subgroup 𝐾 = Stab𝐺 (𝑥) = SO(2) ≤ SL2 (R), so there
626 CHAPTER 34. DISCRETE GROUP ACTIONS
is a continuous bijection
∼ H2 = 𝑋
𝐺/𝐾 = SL2 (R)/SO(2) −→
(34.6.3)
𝑔𝐾 ↦→ 𝑔𝑖.
This formula follows directly from the formula (33.5.3) for distance; the calculation is
requested in Exercise 34.17. It follows that the map 𝐺 → 𝑋 is open, and thus (34.6.3)
is a homeomorphism. In fact, by (34.6.5), if we reparametrize the metric on either
SL2 (R)/SO(2) or H2 by the appropriate factor involving the hyperbolic cosine, the
map (34.6.3) becomes an isometry.
We conclude this section with a view to a more general setting where the above
situation applies.
Remark 34.6.7. More generally, a (globally) symmetric space is a space of the form
𝐺/𝐾 where 𝐺 is a Lie group and 𝐾 ≤ 𝐺 a maximal compact subgroup. Alternatively,
it can be defined as a space where every point has a neighborhood where there is an
isometry of order 2 fixing the point. For more reading on the theory of symmetric
spaces, and the connection to differential geometry and Lie groups, see the book by
Helgason [Hel2001].
Lemma 34.7.1. Let Γ ≤ SL2 (R). Then the following are equivalent:
(i) Γ is discrete;
(ii) If 𝛾𝑛 ∈ Γ and 𝛾𝑛 → 1, then 𝛾𝑛 = 1 for almost all 𝑛; and
(iii) For all 𝑀 ∈ R>0 , the set {𝛾 ∈ Γ : k𝛾k ≤ 𝑀 } is finite.
Proof. The equivalence (i) ⇔ (ii) is requested in Exercise 34.15. The implication (i)
⇔ (iii) follows from the fact that the ball of radius 𝑀 in SL2 (R) is a compact subset of
M2 (R), and a subset of a compact set is finite if and only if it is discrete. Slightly more
elaborately, a sequence of matrices with bounded norm has a subsequence where the
entries all converge; since the determinant is continuous, the limit exists in SL2 (R) so
Γ is not discrete.
In particular, we find from Lemma 34.7.1 that a discrete subgroup of SL2 (R) is
countable.
Proposition 34.7.2. Let Γ ≤ PSL2 (R) be a subgroup (with the subspace topology).
Then Γ has a wandering action on H2 if and only Γ is discrete.
applying the above argument to 𝜙−1 Γ𝜙 shows that the orbit Γ𝑧 is discrete. This
concludes the proof.
Alternatively, here is a self-contained proof that avoids the slightly more involved
topological machinery. We again work in the unit disc D2 . First we prove (⇐).
Since Γ is discrete, there is an 𝜖-neighborhood 𝑈 3 1 with 𝑈 ⊆ PSU(1, 1) such that
𝑈 ∩ Γ = {1}; therefore, if
𝑎 𝑏
𝛾= ∈ Γ \ {1}
𝑏 𝑎
then | 𝑏| > 𝜖 or (without loss of generality) |𝑎 − 1| > 𝜖. We claim that in either case
𝑏
|𝛾(0)| = > 𝜖,
𝑎
628 CHAPTER 34. DISCRETE GROUP ACTIONS
and thus the orbit is discrete. Indeed, if | 𝑏| > 𝜖, then since |𝑎| < 1 anyway immediately
| 𝑏/𝑎| > 𝜖; if |𝑎 − 1| > 𝜖 then |𝑎| < 1 − 𝜖 so |𝑎| 2 < 1 − 𝜖 2 and 1/|𝑎| 2 > 1 + 𝜖 2 , giving
2 2
= 1 − |𝑎| > (1 + 𝜖 2 ) − 1 = 𝜖 2 .
𝑏
𝑎 |𝑎| 2
For (⇒), suppose that Γ is not discrete; then there is a sequence 𝛾𝑛 ∈ Γ \ {1} of
elements such that 𝛾𝑛 → 1. Therefore, for all 𝑧 ∈ H2 , we have 𝛾𝑛 𝑧 → 𝑧 and 𝛾𝑛 𝑧 = 𝑧
for only finitely many 𝑛, so every neighborhood of 𝑧 contains infinitely many distinct
points 𝛾𝑛 𝑧.
plane 𝑋 = C, the punctured plane C× ' C/h𝑢i with 𝑢 ∈ C× , and complex tori C/Λ
where Λ ⊂ C is a lattice with Λ ' Z2 . We will embark on a classification of these tori
up to isomorphism by their 𝑗-invariants in section 40.1.
All other Riemann surfaces are hyperbolic with 𝑋 e = H2 , and so are of the form
2
𝑋 = Γ\H with Γ ≤ PSL2 (R) a torsion-free Fuchsian group.
Remark 34.8.3. Klein and Poincaré conjectured the uniformization theorem for alge-
braic curves over C, with rigorous proofs were given by Poincaré: see Gray [Gray94].
Finally, before departing our topological treatment, we consider quotients of man-
ifolds by the (continuous) action of a group. As we have seen, it is quite restrictive
to suppose that this group action is free: we will still want to take quotients by such
groups. Accordingly, we need to model not spaces that are locally modelled by R𝑛 but
those that are locally modelled by the quotient of R𝑛 by a finite group.
Definition 34.8.4. An 𝑛-orbifold 𝑋 is a (second-countable) Hausdorff topological
space that is locally homeomorphic to a quotient 𝐺\R𝑛 with 𝐺 a finite group acting
(continuously). An atlas for an orbifold 𝑋 is the data
(i) An open cover {𝑈𝑖 }𝑖 ∈𝐼 of charts 𝑈𝑖 closed under finite intersection; and
(ii) For each 𝑖 ∈ 𝐼, an open subset 𝑉𝑖 ⊆ R𝑛 equipped with the (continuous) action
of a finite group 𝐺 𝑖 𝑉𝑖 , and a homeomorphism
∼ 𝐺 \𝑉
𝜙𝑖 : 𝑈𝑖 −→ 𝑖 𝑖
satisfying the atlas axiom: for all 𝑈𝑖 ⊆ 𝑈 𝑗 , there exists an injective group homomor-
phism 𝑓𝑖 𝑗 : 𝐺 𝑖 ↩→ 𝐺 𝑗 and a 𝐺 𝑖 -equivariant map 𝜓𝑖 𝑗 : 𝑉𝑖 ↩→ 𝑉 𝑗 satisfying 𝜙 𝑗 ◦𝜓𝑖 𝑗 = 𝜙𝑖
(see Figure 34.8.5).
Ui
Uj
φj
∼
φi
∼
Vi /Gi
Vj /Gj
ψij
Vi Vj
Gi Gj
Orbifolds were introduced by Thurston [Thu97, Chapter 13], who adds a wealth
of motivation and examples; see also the surveys by Scott [Sco83, §2] and Gordon
[Gor2012] as well as the chapter by Ratcliffe [Rat2006, Chapter 13].
34.8.6. We can further ask that the transition maps 𝑓𝑖 𝑗 in an atlas be smooth to get
a smooth orbifold, preserve a 𝐺 𝑖 -Riemannian metric to get a Riemann orbifold,
etc.; replacing R𝑛 by C𝑛 and smooth by holomorphic, we similarly define a complex
𝑛-orbifold, locally modelled on the quotient 𝐺\C𝑛 with 𝐺 a finite group acting
holomorphically.
Definition 34.8.7. Let 𝑋 be an 𝑛-orbifold.
(a) A point 𝑧 ∈ 𝑋 such that there exists a chart 𝑈𝑖 3 𝑧 with group 𝐺 𝑖 ≠ {1} fixing 𝑧
is called an orbifold point of 𝑋, with stabilizer group (or isotropy group) 𝐺 𝑖 ;
the set of orbifold points of 𝑋 is called the orbifold set of 𝑋.
(b) If 𝑧 ∈ 𝑈𝑖 is an isolated orbifold point and its stabilizer group is cyclic, we call 𝑧
a cone point.
Exercises
Unless otherwise specified, let 𝐺 𝑋 be an action of a group 𝐺 on a set 𝑋.
The product 𝑋 𝑋 has the product topology, and so Isom(𝑋) (and every space of
maps from 𝑋 to 𝑋) has an induced subspace topology. A basis of open sets for
Isom(𝑋) in this topology consists of finite intersections of open balls
𝑉 (𝑔; 𝑥, 𝜖) = {ℎ ∈ Isom(𝑋) : 𝜌(𝑔(𝑥), ℎ(𝑥)) < 𝜖 }.
Equip the group 𝐺 = Isom(𝑋) with the topology of pointwise convergence.
(a) Show that 𝐺 is a topological group.
(b) Show that 𝐺 acts continuously on 𝑋.
5. Let 𝐺 = Z be given the discrete topology, and let 𝐺 𝑋 = R/Z act by
𝑥 ↦→ 𝑥 + 𝑛𝑎 ∈ R/Z for 𝑛 ∈ Z for 𝑎 ∈ R − Q. Show that this action is free and
continuous, and show that for all 𝑥 ∈ 𝑋 the map (34.1.13)
𝐺/Stab𝐺 (𝑥) = 𝐺 → 𝐺𝑥
𝑔 ↦→ 𝑔𝑥
is (continuous and bijective but) not a homeomorphism, giving 𝐺𝑥 ⊆ 𝑋 the
subspace topology.
I 6. Let 𝐺 act (continuously and) transitively on 𝑋. Suppose that 𝐺, 𝑋 are (Hausdorff
and) locally compact, and suppose further that 𝐺 has a countable base of open
sets. Let 𝑥 ∈ 𝑋 and let 𝐾 = Stab𝐺 (𝑥). Show that the natural map 𝐺/𝐾 → 𝑋 is
a homeomorphism.
7. Let 𝐺 𝑋 be a free and wandering action, and let 𝑈 be an open set such
that 𝑔𝑈 ∩ 𝑈 = ∅ for all 𝑔 ≠ 1. Show that the map 𝐺 × 𝑈 → 𝜋 −1 (𝜋(𝑈)) is a
homeomorphism and the restriction 𝜋 : 𝐺 × 𝑈 → 𝜋(𝑈) ' 𝑈 is a (split) covering
map.
8. Let 𝑋 be (Hausdorff and) locally compact, let 𝑥 ∈ 𝑋, and let 𝑈 3 𝑥 be an open
neighborhood. Show that there exists an open neighborhood 𝑉 3 𝑥 such that
𝐾 = cl(𝑉) ⊆ 𝑈 is compact.
I 9. Let 𝑋, 𝑌 be (Hausdorff) topological spaces, let 𝑓 : 𝑋 → 𝑌 be a continuous map,
and let
gr( 𝑓 ) : 𝑋 → 𝑋 × 𝑌
𝑥 ↦→ (𝑥, 𝑓 (𝑥))
be the graph of 𝑓 . Show that 𝑓 is a closed map.
10. One way to weaken the running hypothesis that 𝑋 is Hausdorff in this chapter
is to instead assume only that 𝑋 is locally Hausdorff: every 𝑥 ∈ 𝑋 has an open
neighborhood 𝑈 3 𝑥 such that 𝑈 is Hausdorff.
Show that a weakened version of Lemma 34.3.6(i) ⇒ (ii) is not true with only
the hypothesis that 𝑋 is locally Hausdorff: that is, exhibit a locally Hausdorff
topological space 𝑋 with a (continuous) wandering action of a group 𝐺 such
that 𝜋 : 𝑋 → 𝐺\𝑋 is not a local homeomorphism, and so Lemma 34.3.6(ii) does
not hold. [Hint: Let 𝑋 be the bug-eyed line and 𝐺 ' Z/2Z acting by 𝑥 ↦→ −𝑥
on R× and swapping points in the doubled origin.]
34.8. RIEMANN UNIFORMIZATION AND ORBIFOLDS 633
11. Let 𝐺 = Z and let 𝐺 ↩→ 𝑋 = R2 r {(0, 0)} act by 𝑛 · (𝑥, 𝑦) = (2𝑛 𝑥, 𝑦/2𝑛 ).
In other words, 𝐺 is the group of continuous maps 𝑋 → 𝑋 generated by
(𝑥, 𝑦) ↦→ (2𝑥, 𝑦/2).
(a) Show that the action of 𝐺 on 𝑋 is free and wandering.
(b) Show that the quotient 𝐺\𝑋 is not Hausdorff.
(c) Let 𝐾 = {(𝑡, 1−𝑡) : 𝑡 ∈ [0, 1]}. Then 𝐾 is compact. Show that 𝐾 ∩ 𝑔𝐾 ≠ ∅
for infinitely many 𝑔 ∈ 𝐺. [So Theorem 34.5.1(v) holds but (ii) does not,
and in particular that the action of 𝐺 is not proper. Can you see this directly
from the definition of proper?]
12. Let 𝑋 be a Hausdorff topological space with a continuous action of a Hausdorff
topological group 𝐺. Suppose that the action of 𝐺 is wandering. Show that
for all 𝑥 ∈ 𝑋, there is an open neighborhood 𝑈 3 𝑥 such that the finite group
Stab𝐺 (𝑥) acts on 𝑈 (i.e., 𝑔𝑈 ⊆ 𝑈 for all 𝑔 ∈ Stab𝐺 (𝑥)).
13. Show that a subgroup Γ ≤ R𝑛 is discrete if and only if Γ = Z𝑣 1 + · · · + Z𝑣 𝑚
with 𝑣 1 , . . . , 𝑣 𝑚 ∈ Γ linearly independent over R. As a consequence, show that
Γ ≤ R𝑛 is a lattice if and only if Γ is discrete with 𝑚 = 𝑛.
14. Exhibit an injective group homomorphism SO(𝑛) ↩→ SO(𝑛 + 1) and a homeo-
morphism
S𝑛 ' SO(𝑛 + 1)/SO(𝑛),
where S𝑛 = {𝑥 ∈ R𝑛+1 : k𝑥k 2 = 1} is the 𝑛-dimensional sphere, analogous to
(34.6.3).
I 15. Let 𝐺 be a topological group with a countable system of fundamental open
neighborhoods of 1 ∈ 𝐺 (for example, this holds if 𝐺 is metrizable). Show that
𝐺 is discrete if and only if whenever {𝑔𝑛 } 𝑛 is a sequence from 𝐺 with 𝑔𝑛 → 1,
then 𝑔𝑛 = 1 for all but finitely many 𝑛.
16. Let 𝐺 be a discrete group with a (continuous) covering space action on a
Hausdorff space 𝑋 such that 𝐺\𝑋 is Hausdorff. Show that 𝐺 acts quasi-properly
on 𝑋.
I 17. Show that for 𝑔 ∈ SL2 (R),
19. Show that the stabilizer group of an orbifold point is well-defined up to group
isomorphism, independent of the chart.
20. The group SL2 (R) acts on P1 (R) = R∪ {∞} by linear fractional transformations.
Show that 𝐺 = SL2 (Z) ≤ SL2 (R) is discrete, but 𝐺 does not act properly on
P1 (R). [So discrete groups can act on locally compact spaces without necessarily
acting properly.]
Chapter 35
In this chapter, we introduce the classical modular group PSL2 (Z) ≤ PSL2 (R), a
discrete group acting on the upper half-plane that has received extensive study because
of the role it plays throughout mathematics. We examine the group in detail via a
fundamental domain and conclude with some applications to number theory. This
chapter will serve as motivation and example for the generalizations sought later in
this part of the text.
There are very many references for the classical modular group, including Apostol
[Apo90, Chapter 2], Diamond–Shurman [DS2005, Chapter 2], and Serre [Ser73,
Chapter VII].
The group PSL2 (Z) acts faithfully on the upper half-plane H2 by linear fractional
transformations; equipping H2 with the hyperbolic metric, this action is by orientation-
preserving isometries.
Since Z ⊆ R is discrete, so too is SL2 (Z) ⊆ M2 (Z) ⊆ M2 (R) discrete and therefore
PSL2 (Z) ≤ PSL2 (R) is a Fuchsian group (Definition 34.7.3).
35.1.2. Our first order of business is to try to understand the structure of the classical
modular group in terms of this action. Let
0 −1 1 1
𝑆 := , 𝑇 := ∈ PSL2 (Z).
1 0 0 1
635
636 CHAPTER 35. CLASSICAL MODULAR GROUP
T −1 1 T
T −1 S S TS
ST S ST ST −1 ST −1 S
i
ω −ω 2 = −ω
S
z7→w= z−i
z+i
−−−−−−−→
Figure 35.1.7: ◊ in D2
The following three lemmas describe the relationship of the set ◊ to Γ.
Lemma 35.1.8. For all 𝑧 ∈ H2 , there exists a word 𝛾 ∈ h𝑆, 𝑇i such that 𝛾𝑧 ∈ ◊.
and the number of 𝑐, 𝑑 ∈ Z such that |𝑐𝑧 + 𝑑| < 1 is finite: the set Z + Z𝑧 ⊆ C is a
lattice, so there are only finitely many elements of bounded norm. (Alternatively, the
orbit Γ𝑧 is discrete by Theorem 34.5.1—or the direct argument given in Proposition
34.7.2—therefore, its intersection with the compact set
√
Since Im 𝑧 > Im 𝜔 = 3/2, from (35.1.11) we conclude that 𝑐2 ≤ 4/3 so |𝑐| ≤ 1.
If 𝑐 = 0 then 𝑎𝑑 − 𝑏𝑐 = 𝑎𝑑 = 1 so 𝑎 = 𝑑 = ±1, and then 𝑧 0 = 𝛾𝑧 = 𝑧 ± 𝑏, which
immediately implies 𝑏 = 0 so 𝛾 = 1 as claimed. If instead |𝑐| = 1, then the conditions
together imply 𝑑 = 0; but then |𝑐𝑧 + 𝑑| = | 𝑧| ≤ 1, and since 𝑧 ∈ int(◊) we have | 𝑧| > 1,
a contradiction.
If instead Im 𝑧 0 < Im 𝑧, we interchange the roles of 𝑧, 𝑧 0 and have strict inequality
in (35.1.11); by the same argument and the weaker inequality |Re 𝑧| ≤ 1/2, we then
obtain | 𝑧| < 1, a contradiction.
Proof. Let 𝑧 = 2𝑖 ∈ int(◊). Let 𝛾 ∈ Γ, and let 𝑧 0 = 𝛾𝑧. By Lemma 35.1.8, there exists
𝛾 0 a word in 𝑆, 𝑇 such that 𝛾 0 𝑧 0 ∈ ◊. By Lemma 35.1.10, we have 𝛾 0 𝑧 0 = (𝛾 0 𝛾)𝑧 = 𝑧,
so 𝛾 0 𝛾 = 1 and 𝛾 = 𝛾 0 ∈ h𝑆, 𝑇i.
Although we have worked in PSL2 (Z) throughout, it follows from Lemma 35.1.12
that the matrices 𝑆, 𝑇 also generate SL2 (Z), since 𝑆 2 = −1. See Exercise 35.3 for
another proof of Lemma 35.1.12.
Proof. The statement follows from Lemmas 35.1.8 and 35.1.10 (recalling the definition
of fundamental set, Definition 34.1.14).
35.1.14. If 𝑧 ∈ ◊ has StabΓ (𝑧) ≠ {1}, then we claim that one of the following holds:
The orbit of the limit point ∞ under Γ is P1 (Q) ⊆ bd H2 ; letting H2∗ = H2 ∪ P1 (Q),
there is a homeomorphism
𝑋 = Γ\H2∗ ' P1 (C)
as in Figure 35.1.15.
∞
∞
' ω
i
ω
Away from the orbits Γ𝑖, Γ𝜔 with nontrivial stabilizer, the complex structure on H2
descends and gives the quotient 𝑌 r {Γ𝑖, Γ𝜔} the structure of a Riemann surface. By
studying the moduli of lattices, later we will give an explicit holomorphic identification
𝑗 : 𝑌 → C.
35.1.16. By 34.8.11, the quotient 𝑌 has the structure of a good complex 1-orbifold,
when we keep track of the two nontrivial stabilizers.
Alternatively, we can also give 𝑋 the structure of a compact Riemann surface as
follows. Let 𝑧0 ∈ ◊ ∪ {∞}. If 𝑧0 = ∞, we take the chart 𝑧 ↦→ 𝑒 2 𝜋𝑖𝑧 . Otherwise, let
𝑒 = # StabΓ (𝑧0 ) < ∞, let 𝑤 = (𝑧 − 𝑧0 )/(𝑧 − 𝑧0 ) be the local coordinate as in (33.7.3),
and take the chart 𝑧 ↦→ 𝑤 𝑒 at 𝑧0 .
𝑧 𝑖 := 𝛾𝑖 𝑧0 = (𝛿1 · · · 𝛿𝑖 )𝑧 0 .
640 CHAPTER 35. CLASSICAL MODULAR GROUP
zr = 2i − 1 z0 = 2i
δr = T
z1 = 12 i
(Alternatively, for a proof in the style of Lemma 35.1.10, see Exercise 35.5.)
Remark 35.1.20. Alperin [Alp93] uses the action on the irrational numbers to show
directly that PSL2 (Z) is the free product of Z/2Z and Z/3Z (but note the typo 𝛽(𝑧) =
1 − 1/𝑧 on the first page).
disc(𝑄) := 𝑏 2 − 4𝑎𝑐.
be the set of primitive, positive definite forms of discriminant 𝑑. The group Γ acts on
Q 𝑑 the right by change
of variable: for 𝛾 ∈ Γ, we define (𝑄 𝛾 ) (𝑥, 𝑦) = 𝑄(𝛾(𝑥, 𝑦) t ), so
𝑟 𝑠
that if 𝛾 = then
𝑡 𝑢
Thus
1 𝑏 1
− ≤ Re 𝑧 𝑄 = − <
2 2𝑎 2
so −𝑎 < 𝑏 ≤ 𝑎, or equivalently,
| 𝑏| ≤ 𝑎 and (𝑏 ≥ 0 if | 𝑏| = 𝑎);
and
𝑏2 − 𝑑 𝑐
| 𝑧𝑄 | = = ≥1
4𝑎 2 𝑎
so 𝑎 ≤ 𝑐 and 𝑏 ≥ 0 if equality holds. In sum, every positive definite form 𝑄 is
equivalent to a (SL2 (Z)-)reduced form satisfying
| 𝑏| ≤ 𝑎 ≤ 𝑐 with 𝑏 ≥ 0 if | 𝑏| = 𝑎 or 𝑎 = 𝑐.
We now show that there are only finitely many reduced forms with given discrim-
inant 𝑑 < 0, i.e., that ℎ(𝑑) < ∞. The inequalities | 𝑏| ≤ 𝑎 ≤ 𝑐 imply that
| 𝑑| = 4𝑎𝑐 − 𝑏 2 ≥ 3𝑎 2 ,
√︁
so 𝑎 ≤ |𝑑|/3 and |𝑏| ≤ 𝑎, so there are only finitely many possibilities for 𝑎, 𝑏; and
then 𝑐 = (𝑏 2 − 𝑑)/(4𝑎) is determined. This gives an efficient method to compute the
set Q 𝑑 /Γ efficiently. √ √
Let 𝑆 = Z ⊕ Z[(𝑑 + 𝑑)/2] ⊂ 𝐾 = Q( 𝑑) be the quadratic ring of discriminant
𝑑 < 0. Let Pic(𝑆) be the group of invertible fractional ideals of 𝑆 modulo principal
ideals. Then there is a bijection
Q 𝑑 /Γ ↔ Pic(𝑆)
" √ !#
2 2 −𝑏 + 𝑑
[𝑎𝑥 + 𝑏𝑥𝑦 + 𝑐𝑦 ] ↦→ [𝔞] = 𝑎,
2
(Exercise 35.9). In the same stroke, we have proven the finiteness of the class number
# Pic(𝑆) < ∞.
35.3. MODULI OF LATTICES 643
1
Λ∼ Λ = Z + Z𝜏
𝑧1
where 𝜏 = 𝑧 2 /𝑧1 ∈ H2 .
𝑎𝜏 + 𝑏
𝜏0 = .
𝑐𝜏 + 𝑑
Therefore 𝑔 ∈ SL2 (Z), and since 𝑔 is well-defined as an element of Γ = PSL2 (Z), the
result follows.
𝑌 = Γ\H2 → {Λ ⊂ C lattice}/∼
(35.3.4)
Γ𝜏 ↦→ [Z + Z𝜏];
To avoid confusion, from now on we will now write Γ(1) = PSL2 (Z).
35.4.2. By strong approximation for SL2 (Z) (Theorem 28.2.6), the map SL2 (Z) →
SL2 (Z/𝑁Z) is surjective for all 𝑁 ≥ 1, so there is an exact sequence
𝑌0 (𝑁) := Γ0 (𝑁)\H2
(35.4.7)
𝑋0 (𝑁) := Γ0 (𝑁)\H2∗
From 35.4.2,
Γ(1)/Γ(2) ' PSL2 (Z/2Z) = GL2 (F2 ) ' 𝑆3 (35.4.9)
the nonabelian group of order 6, so in particular [Γ(1) : Γ(2)] = 6.
We can uncover the structure of the group Γ(2) in a manner similar to what we did
for Γ(1) in section 35.1—the details are requested in Exercise 35.10. The group Γ(2)
is generated by
1 2 1 0
and
0 1 2 1
1 2
0 1
1 0
2 1
1 1
2 (−1 + i) 2 (1 + i)
−1 0 1
(In fact, later we will see from more general structural results that Γ(2) is freely gen-
erated by these two elements, so it is isomorphic to the free group on two generators.)
The action Γ(2) H2 is free: by 35.1.14, if 𝛾𝑧 = 𝑧 with 𝑧 ∈ H2 and 𝛾 ∈ Γ(2) ≤
Γ(1), then 𝛾 is conjugate in Γ(1) to either 𝑆, 𝑆𝑇; but Γ(2)
E Γ(1) is normal, so
0 −1 0 −1
without loss of generality either 𝑆 = or 𝑆𝑇 = belongs to Γ(2), a
1 0 1 1
contradiction.
Let 𝑌 (2) := Γ(2)\H2 . Then gluing together the fundamental set, there is a
homeomorphism
𝑌 (2) ' P1 (C) r {0, 1, ∞}.
The limit points of ◊ in bd H2 are the points −1, 0, 1, ∞ and the points −1, 1 are
identified in the quotient (by translation). The orbit of these points under Γ(2) is
P1 (Q) ⊆ bd H2 , so letting H2∗ = H2 ∪ P1 (Q), there is a homeomorphism
where 𝛾 ∈ Γ(1) is a lift, so the map (35.4.12) is obtained as the quotient by GL2 (F2 ).
Finally, the congruence conditions (35.4.6) imply that Γ0 (2) = Γ1 (2) has index 2 in
Γ(2), with the quotient generated by 𝑇, and we obtain a fundamental set by identifying
the two ideal triangles in ◊ above.
Exercises
1. Prove that 𝑌 (1) = SL2 (Z)\H2 has area(𝑌 (1)) = 𝜋/3 by direct integration
(verifying the Gauss–Bonnet formula).
1 0
2. Show that PSL2 (Z) is generated by 𝑇 and 𝑈 = . [So PSL2 (Z) is generated
1 1
by two parabolic elements (of infinite order), just as it is generated by elements
of order two and three.]
3. Prove Lemma 35.1.12 using Lemma 28.3.3 (elementary matrices).
4. In this exercise, we link the fact that PSL2 (Z) is generated by 𝑆, 𝑇 to a kind of
continued fraction via the Euclidean algorithm. [So the reduction algorithm is
a way to visualize the Euclidean algorithm.] Let 𝑎, 𝑏 ∈ Z ≥1 with 𝑎 ≥ 𝑏.
(a) Show that there exist unique 𝑞, 𝑟 ∈ Z such that 𝑎 = 𝑞𝑏 − 𝑟 and 𝑞 ≥ 2 and
0 ≤ 𝑟 < 𝑏.
From (a), define inductively 𝑟 0 = 𝑎, 𝑟 1 = 𝑏, and 𝑟 𝑖−1 = 𝑞 𝑖 𝑟 𝑖 − 𝑟 𝑖+1 with
0 ≤ 𝑟 𝑖+1 < 𝑟 𝑖 ; we then have 𝑟 1 > 𝑟 2 > · · · > 𝑟 𝑡 > 𝑟 𝑡+1 = 0 for some 𝑡 > 0.
b) Show that gcd(𝑎, 𝑏) = 𝑟 𝑡 , and if gcd(𝑎, 𝑏) = 1 then
𝑎 1
= 𝑞1 − .
𝑏 1
𝑞2 −
1
···−
𝑞𝑡
1 𝑏0
𝑊𝐴 =
0 1
with 𝑏 0 ∈ Z. Conclude that h𝑆, 𝑇i = PSL2 (Z). (So how, in the end, does
this procedure to write 𝐴 in terms of 𝑆 and 𝑇 relate to the one given by the
reduction algorithm in Lemma 35.1.8?)
5. In this exercise, we give a “matrix proof” that a complete set of relations satisfied
by 𝑆, 𝑇 in PSL2 (Z) are 𝑆 2 = (𝑆𝑇) 3 = 1.
(a) Show that it suffices to show that no word 𝑆(𝑆𝑇) 𝑒1 𝑆(𝑆𝑇) 𝑒2 . . . 𝑆(𝑆𝑇) 𝑒𝑛
with 𝑒 𝑖 = 1, 2 is equal to 1.
(b) Observe that 𝑆(𝑆𝑇) = 𝑇 and 𝑆(𝑆𝑇) 2 have at least one off-diagonal entry
nonzero and can be represented with a matrix whose entries all have the
same sign.
𝑎 𝑏
(c) Show that if 𝐴 = has at least one off-diagonal entry nonzero and all
𝑐 𝑑
entries of the same sign, then these properties hold also for both 𝑆(𝑆𝑇) 𝐴
and 𝑆(𝑆𝑇) 2 𝐴. Conclude that (a) holds.
[This argument is given by Fine [Fin89, Theorem 3.2.1].]
6. Show that the commutator subgroup Γ0 E Γ = PSL2 (Z) (the subgroup generated
by commutators 𝛾𝛿𝛾 −1 𝛿−1 for 𝛾, 𝛿 ∈ Γ) has index 6 and Γ/Γ0 ' Z/6Z.
7. Compute the class number ℎ(𝑑) and the set of reduced (positive definite) binary
quadratic forms of discriminant 𝑑 = −71.
8. Let Q 𝑑 be the set of primitive,
Ð positive definite binary quadratic forms of
discriminant 𝑑 < 0 and let Q = 𝑑 Q 𝑑 .
(a) Show that the group GL2 (Z) acts naturally on Q by change of variables,
with PGL2 (Z) acting faithfully.
(b) Consider the action of PGL2 (Z) on H2 . Show that every 𝑄 ∈ Q 𝑑 is
equivalent to a GL2 (Z)-reduced form 𝑎𝑥 2 + 𝑏𝑥𝑦 + 𝑐𝑦 2 satisfying
0 ≤ 𝑏 ≤ 𝑎 ≤ 𝑐.
I 9. Let √ √
𝑆 = Z ⊕ Z[(𝑑 + 𝑑)/2] ⊂ 𝐾 = Q( 𝑑)
be the quadratic ring of discriminant 𝑑 < 0. Let Pic(𝑆) be the group of invertible
fractional ideals of 𝑆 modulo principal ideals. Show that the map
Q 𝑑 /Γ → Pic(𝑆)
" √ !#
2 2 −𝑏 + 𝑑
[𝑎𝑥 + 𝑏𝑥𝑦 + 𝑐𝑦 ] ↦→ [𝔞] = 𝑎,
2
Hyperbolic space
In this chapter, we extend the notions introduced for the hyperbolic plane to hyperbolic
space in three dimensions; we follow essentially the same outline, and so our exposition
is similarly brief.
Hyperbolic space is the set H3 equipped with the metric induced by the hyperbolic
length element
|d𝑥| 2 + d𝑦 2 d𝑥12 + d𝑥22 + d𝑦 2
d𝑠2 := = .
𝑦2 𝑦2
36.1.2. The space H3 is the unique three-dimensional (connected and) simply con-
nected Riemannian manifold with constant sectional curvature −1. The volume ele-
ment corresponding to the hyperbolic length element is accordingly
d𝑥 1 d𝑥2 d𝑦
d𝑉 := .
𝑦3
36.1.3. A vertical half-plane in hyperbolic space is a set of points with 𝑦 arbitrary
and the coordinate 𝑥 confined to a line in C. The hyperbolic length element restricted
to every vertical half-plane is (equivalent to) the hyperbolic length element on the
hyperbolic plane. Therefore, H3 contains many isometrically embedded copies of H2 .
36.1.4. The sphere at infinity is the set
bd H3 = P1 (C) = C ∪ {∞}
649
650 CHAPTER 36. HYPERBOLIC SPACE
(analogous to the circle at infinity for H2 ), with the image of C corresponding to the
locus of points with 𝑡 = 0. We then define the completed upper half-space to be
H3∗ := H3 ∪ bd H3 .
36.1.5. The metric space H3 is complete, and the topology on H3 is the same as the
topology induced by the Euclidean metric.
The geodesics in H3 are the Euclidean hemicircles orthogonal to C and vertical half-
lines: every two points lie in a vertical hyperbolic plane (see 36.1.3), so this statement
can be deduced from the case of the hyperbolic plane. (Alternatively, by applying an
element of PSL2 (C) it is enough to show that the vertical axis 𝑍 = {(0, 𝑦) : 𝑦 > 0}
is a geodesic, and arguing as in (33.5.4) we obtain the result.) Accordingly, H3 is a
uniquely geodesic space.
36.2 Isometries
Analogous to the case of H2 , with orientation-preserving isometries given by PSL2 (R)
acting by linear fractional transformations, in this section we identify the isometries
of hyperbolic space H3 as coming similarly from PSL2 (C).
36.2.1. The group PSL2 (C) acts on the sphere at infinity P1 (C) by linear fractional
transformations. We extend this action to H3 (with almost the same definition!) as
follows. We identify
H3 ↩→ H = C + C 𝑗
(𝑥, 𝑦) ↦→ 𝑧 = 𝑥 + 𝑦 𝑗
SL2 (C) × H3 → H3
(36.2.2)
(𝑔, 𝑧) ↦→ 𝑔𝑧 = (𝑎𝑧 + 𝑏) (𝑐𝑧 + 𝑑) −1
𝑎 𝑏
for 𝑔 = ∈ SL2 (C). If 𝑧 = 𝑥 + 𝑦 𝑗, then in coordinates (Exercise 36.3)
𝑐 𝑑
where
k𝑐𝑧 + 𝑑 k 2 = nrd(𝑐𝑧 + 𝑑) = |𝑐𝑥 + 𝑑| 2 + |𝑐| 2 𝑦 2 .
Therefore the image of this map lies in H3 . (Compare this formula with the action of
SL2 (R) in (33.3.8).)
Lemma 36.2.4. The map (36.2.2) defines a group action of SL2 (C) on H3 .
under the equivalence relation (𝛼, 𝛽) ∼ (𝛼𝛾, 𝛽𝛾) for 𝛾 ∈ H× , and we denote by
(𝛼 : 𝛽) ∈ P1 (H) the equivalence class of (𝛼, 𝛽). We verify that the group SL2 (C) acts
on P1 (H) by
𝑎 𝑏
· (𝛼 : 𝛽) = (𝑎𝛼 + 𝑏𝛽 : 𝑐𝛼 + 𝑑𝛽);
𝑐 𝑑
the left action of SL2 (C) commutes with the right action of H× . The restriction of this
action to H3 ↩→ P1 (H) by 𝑧 ↦→ (𝑧 : 1) is
as above.
652 CHAPTER 36. HYPERBOLIC SPACE
We now show that PSL2 (C) acts on H3 by isometries. This can be verified directly
by the formula, with some effort; we prefer to verify this on a convenient set of
generators, and so we are first led already to the following decomposition of SL2 (C)
(cf. Proposition 33.4.2).
36.2.5. Let
𝑎 𝑏
𝐾 := SU(2) = ∈ M2 (C) : |𝑎| 2 + | 𝑏| 2 = 1 ' H1
−𝑏 𝑎
𝑎 0
𝐴 := : 𝑎 ∈ R×>0 ' R
0 1/𝑎
1 𝑏
𝑁 := : 𝑏 ∈ C ' C.
0 1
Proof. We apply the same method as in the proof of Proposition 33.4.2. For surjec-
1 −𝑥
tivity, we let 𝑧 = 𝑔( 𝑗) = 𝑥 + 𝑦 𝑗, let 𝑛𝑔 = ∈ 𝑁 so that (𝑛𝑔 𝑔) ( 𝑗) = 𝑦 𝑗; then let
0 1
√
1/ 𝑦 0
𝑎𝑔 = √ ∈ 𝐴, so (𝑎 𝑔 𝑛𝑔 𝑔) ( 𝑗) = 𝑗 and 𝑎 𝑔 𝑛𝑔 𝑔 ∈ StabSL2 (C) ( 𝑗) = 𝐾.
0 𝑦
Lemma 36.2.8. The group SL2 (C) is generated by the subgroups 𝐴, 𝑁, and the
0 −1
element , which acts on H3 by
1 0
0 −1 1
(𝑧) = −𝑧 −1 = (−𝑥 + 𝑦 𝑗) (36.2.9)
1 0 k𝑧k 2
We are now ready to investigate the consequences of this decomposition for the
geometry of hyperbolic space.
Theorem 36.2.11. The map (36.2.2) defines a faithful, transitive action of PSL2 (C)
on H3 by isometries.
Proof. We use the generators in Lemma 36.2.8. The fact that the action is faithful
3 3
follows directly. For transitivity,we show that H is the orbit of 𝑗. If 𝑧 = 𝑥 + 𝑦 𝑗 ∈ H
1 −𝑥
then we first apply a translation to reduce to the case 𝑧 = 𝑦 𝑗 and then reduce
0 1
to the case of the hyperbolic plane.
Next, we show that PSL2 (C) ↩→ Isom+ (H3 ). Verification that d𝑔(𝑠) = d𝑠 for 𝑔 a
generator in one of the first two cases of Lemma 36.2.8 is immediate, from the definition
of the metric; the third case can be checked directly (Exercise 36.4). Orientation is
preserved in each case.
36.2.12. The group PSL2 (C) acts transitively on geodesics and consequently on pairs
of points at a fixed distance: by the transitive action of PSL2 (C) on H3 , every point
can be mapped to 𝑗; and applying an element of StabSL2 (C) 𝑗 = SU(2), every other
point 𝑢 can be brought to 𝑡 𝑗 with 𝑡 ≥ 1, with log 𝑡 = 𝜌( 𝑗, 𝑢) by the distance in the
hyperbolic plane. It follows that
|𝑧 − 𝑧 0 | 2 |𝑥 − 𝑥 0 | 2 + (𝑦 − 𝑦 0) 2
cosh 𝜌(𝑧, 𝑧 0) = 1 + = 1 + (36.2.13)
2𝑦𝑦 0 2𝑦𝑦 0
by verifying (36.2.13) in the special case where 𝑧 = 𝑗 and 𝑧 0 = 𝑦 𝑗 with 𝑦 > 0, and then
using the preceding transitive action and the fact that the right-hand side of (36.2.13)
is invariant under the action of SL2 (C), verified again using the generators in Lemma
36.2.8.
and
Isom(H3 ) ' PSL2 (C) o Z/2Z (36.2.16)
where the nontrivial element of Z/2Z acts by complex conjugation on PSL2 (C) and
(𝑧, 𝑡) ↦→ (𝑧, 𝑡) on H3 .
𝑍 = {𝑦 𝑗 : 𝑦 > 0} ⊆ H3 .
𝑒𝑖 𝜃 0
isometric rotation fixes 𝑍, and applying such a rotation we may suppose
0 𝑒 −𝑖 𝜃
further that 𝜙 fixes H .
Now let 𝑧 = 𝑥 + 𝑦 𝑗 and 𝜙(𝑧) = 𝑧 0 = 𝑥 0 + 𝑦 0 𝑗. Let 𝑟 + 𝑠 𝑗 ∈ H . Then
𝜌(𝑧, 𝑟 + 𝑠 𝑗) = 𝜌(𝜙(𝑧), 𝜙(𝑟 + 𝑠 𝑗)) = 𝜌(𝑧 0, 𝑟 + 𝑠 𝑗)
so from (36.2.13)
|𝑥 − 𝑟 | 2 + (𝑦 − 𝑠) 2 |𝑥 0 − 𝑟 | 2 + (𝑦 0 − 𝑠) 2
= ;
2𝑠𝑦 2𝑠𝑦 0
letting 𝑠 → ∞ we find that 𝑦 = 𝑦 0 and |𝑥 − 𝑟 | = |𝑥 0 − 𝑟 | for all 𝑟 ∈ R, thus Re 𝑥 = Re 𝑥 0
and Im 𝑥 = ± Im 𝑥 0. By continuity, the sign is determined uniquely by 𝑔, and we
conclude that either 𝑔(𝑧) = 𝑧 or 𝑔(𝑧) = 𝑥 + 𝑦 𝑗, as claimed.
36.2.17. The isometry group PSL2 (C) also admits a ‘purely geometric’ definition via
the Poincaré extension, as follows.
An element 𝑔 ∈ SL2 (C) as a Möbius transformation, induces a biholomorphic
map of the Riemann sphere P1 (C) = C ∪ {∞}. This map can be represented as a
composition of an even number (at most four) inversions in circles in P1 (C), or circles
and lines in C (Exercise 36.2). We have identified P1 (C) = bd H3 as the boundary,
and for each circle in P1 (C) there is a unique hemisphere in H3 which intersects bd H3
in this circle; if this circle is a line, then we take a vertical half-plane. We then lift
the action of 𝑔 ∈ PSL2 (C) one inversion at a time with respect to the corresponding
hemisphere or half-plane. It turns out that the action of this product does not depend
on the choice of the circles.
To verify that PSL2 (C) acts by isometries, we need to know that inversion in a
hemisphere or vertical half-plane is an isometry of H3 ; after observing that the first
two types of generators in Lemma 36.2.8 (stretching and translating) are isometries,
one reduces to the case of checking that inversion in the unit hemisphere, defined by
𝑧
𝑧 ↦→ ,
k𝑧k 2
is an isometry; and this boils down to the same calculation as requested in Exercise
36.4.
36.2.18. We have a similar classification of isometries of H3 as in the case of H2 as
follows. Let 𝑔 ∈ PSL2 (C).
(i) If ± Tr(𝑔) ∈ (−2, 2), then 𝑔 is elliptic: it has two distinct fixed points in bd H3
and fixes every point in the geodesic between them, called its axis, acting by
(hyperbolic) rotation around its axis.
(ii) If ± Tr(𝑔) ∈ R r [−2, 2], then 𝑔 is hyperbolic; if ± Tr(𝑔) ∈ C \ R, then 𝑔 is
loxodromic. (Some authors combine these two cases.) In these cases, 𝑔 has two
fixed points in bd H3 and the line through these two points is stabilized, and 𝑔
has no fixed point in H3 .
(iii) Finally and otherwise, if ± Tr(𝑔) = ±2, then 𝑔 is parabolic: it has a unique
fixed point in bd H3 and no fixed point in H3 .
36.3. UNIT BALL, LORENTZ, AND SYMMETRIC SPACE MODELS 655
bd D3 = {𝑤 ∈ R3 : k𝑤k = 1}.
k𝑤 − 𝑤 0 k 2
cosh 𝜌(𝑤, 𝑤 0) = 1 + 2 . (36.3.6)
(1 − k𝑤k 2 ) (1 − k𝑤 0 k 2 )
In the unit ball model, the geodesics are intersections of D3 of Euclidean circles
and straight lines orthogonal to the sphere at infinity, and similarly geodesic planes
are intersections of D3 with Euclidean spheres and Euclidean planes orthogonal to the
sphere at infinity.
36.3.7. The isometries of D3 are obtained by pushforward from H3 . Explicitly, we
first identify
D3 ↩→ H
(36.3.8)
𝑤 ↦→ 𝑤 1 + 𝑤 2𝑖 + 𝑤 3 𝑗 .
We then define the involution
∗
:H→H
(36.3.9)
𝛼 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 ↦→ 𝑘𝛼𝑘 −1 = 𝑡 + 𝑥𝑖 + 𝑦 𝑗 − 𝑧𝑘
We find that
SU2 (H, ∗ ) ' 𝜙 SL2 (C)𝜙−1
with 𝜙 as in 36.3.5. The group SU2 (H, ∗ ) acts on D3 by
𝑔𝑤 = (𝛼𝑤 + 𝛽) (𝛽∗ 𝑤 + 𝛼∗ ) −1 .
with
d𝑠2 := −d𝑡 2 + d𝑥12 + d𝑥22 + d𝑥32
and orientation-preserving isometries given by the subgroup SO+ (3, 1) ≤ SO(3, 1)
of elements mapping L3 to itself. The relationship between the Lorentz model and
the upper half-space model relies on the exceptional isomorphism of Lie algebras
𝔰𝔬3,1 ' 𝔰𝔩2,C and the double cover SL2 (C) → SO+ (3, 1).
To conclude, we find the symmetric space model of H3 , analogous to section 34.6.
36.3.13. The group 𝐺 := SL2 (C) has the structure of a metric space induced from the
usual structure on M2 (C) ' C4 . Since 𝐺 acts transitively on H3 , and the stabilizer of
𝑗 is 𝐾 = SU(2),
𝐺/𝐾 = SL2 (C)/SU(2) −→ ∼ H3
(36.3.14)
𝑔𝐾 ↦→ 𝑔 𝑗;
from the Iwasawa decomposition (Lemma 36.2.7), there is a homeomorphism
L: R → R
∫ 𝜃 (36.5.2)
L (𝜃) = − log |2 sin 𝑡| d𝑡.
0
36.5.3. The first derivative of the Lobachevsky function is L 0 (𝜃) = − log |2 sin 𝜃| by
the fundamental theorem of calculus, so L attains its maximum value at L (𝜋/6) =
0.50747 . . . and minimum at L (5𝜋/6) = −L (𝜋/6). The second derivative is L 00 (𝜃) =
− cot 𝜃. A graph of this function is sketched in Figure 36.5.4.
L (θ)
1/2
θ
π/2 π
−1/2
Lemma 36.5.5. L (𝜃) is odd, periodic with period 𝜋, and satisfies the identity
𝑛−1
∑︁
L (𝑛𝜃) = 𝑛 L (𝜃 + 𝑗 𝜋/𝑛) (36.5.6)
𝑗=0
for all 𝑛 ∈ Z.
Proof. Since L 0 (𝜃) = − log |2 sin 𝜃| is an even function and L (0) = 0, we conclude
L (𝜃) is an odd function, i.e., L (−𝜃) = −L (𝜃) for all 𝜃 ∈ R.
Let 𝑛 ∈ Z. From
𝑛−1
Ö
𝑧𝑛 − 1 = (𝑧 − 𝑒 2 𝜋𝑖 𝑗/𝑛 )
𝑗=0
36.5. HYPERBOLIC VOLUME 659
which yields
𝑛−1 𝑛−1
1 ∑︁ ∑︁
L (𝑛𝜃) = L (𝜃 + 𝑗 𝜋/𝑛) − L ( 𝑗 𝜋/𝑛) (36.5.8)
𝑛 𝑗=0 𝑗=0
1
L (𝜋) = L (𝜋) − L (0) = L (𝜋)
𝑛
so L (𝜋) = 0.
Now since L 0 (𝜃) is periodic with period 𝜋 and L (0) = L (𝜋) = 0, we conclude
that L (𝜃 + 𝜋) = L (𝜃) is also periodic with period 𝜋. Finally,
𝑛−1
∑︁ 𝑛−1
∑︁ 𝑛−1
∑︁ 𝑛−1
∑︁
L ( 𝑗 𝜋/𝑛) = − L (− 𝑗 𝜋/𝑛) = − L ((𝑛 − 𝑗)𝜋/𝑛) = L ( 𝑗 𝜋/𝑛)
𝑗=0 𝑗=0 𝑗=0 𝑗=0
Í𝑛−1
so 𝑗=0 L ( 𝑗 𝜋/𝑛) = 0, and the result follows from (36.5.8).
With the relevant function having been defined, we now return to our geometric
application.
γ β
γ β
α
α α
β γ
β γ
γ β
α
α
Figure 36.5.12: Ideal tetrahedra and its shadow in C
Proof. We follow Milnor [Milno82, Appendix, Lemma 2]; see also Thurston(–Milnor)
[Thu97, Theorem 7.2.1] and Ratcliffe [Rat2006, Theorem 10.4.10]. We may suppose
without loss of generality that one vertex is at ∞ and the finite face lies on the unit
sphere. Projecting onto the unit disc in the 𝑥-plane, we obtain a triangle inscribed
in the unit circle with angles 𝛼, 𝛽, 𝛾 with 𝛼 + 𝛽 + 𝛾 = 2𝜋. We make the simplifying
assumption that all three angles are acute (the argument for the case of an obtuse angle
is similar). We take the barycentric subdivision of the triangle and add up 6 volumes.
We integrate the volume element d𝑥1 d𝑥 2 d𝑦/𝑦 3 over the region 𝑇 (𝛼) defined by the
inequalities
√︁
𝑦 ≥ 1 − |𝑥| 2 , 0 ≤ 𝑥2 ≤ 𝑥1 tan 𝛼, 0 ≤ 𝑥 1 ≤ cos 𝛼. (36.5.14)
1 d𝑥1 d𝑥2 ∞
∫∫∫ ∫∫
d𝑥 1 d𝑥 2 d𝑦
= −
𝑇 ( 𝛼) 𝑦3 2 𝑦 2 𝑦=√1−| 𝑥 |2
∫∫
1 d𝑥1 d𝑥2
=−
2 0≤𝑥1 ≤cos
0≤𝑥 ≤𝑥 tan 𝛼
𝛼 1 − 𝑥 12 − 𝑥 22
2 1
36.5. HYPERBOLIC VOLUME 661
substituting gives the volume L (𝛼)/2, and summing over the other 5 triangles gives
the result.
We define now a standard tetrahedron for use in computing volumes. Let 𝑇𝛼,𝛾
be the tetrahedron with one vertex at ∞ and the other vertices 𝐴, 𝐵, 𝐶 on the unit
hemisphere projecting to 𝐴 0, 𝐵 0, 𝐶 0 in C with 𝐴 0 = 0 to make a Euclidean triangle
with angle 𝜋/2 at 𝐵 0 and 𝛼 at 𝐴 0. The dihedral angle along the ray from 𝐴 to ∞ is
𝛼. Suppose that the dihedral angle along 𝐵𝐶 is 𝛾. The acute angles determine the
isometry class of 𝑇𝛼,𝛾 , and we call 𝑇𝛼,𝛾 the standard tetrahedron with angles 𝛼, 𝛾.
For an illustration, see Figure 36.5.18.
Corollary 36.5.16. We have
1
vol(𝑇𝛼,𝛾 ) = L (𝛼 + 𝛾) + L (𝛼 − 𝛾) + 2L (𝜋/2 − 𝛼) .
4
Proof. One proof realizes the standard tetrahedron as a signed combination of ideal
tetrahedra, and uses Proposition 36.5.13. A second proof just repeats the integral
(36.5.15) to get
∫ 𝜋/2
1 2 sin(𝜃 + 𝛼)
vol(𝑇𝛼,𝛾 = log
d𝜃
4 𝛾 2 sin(𝜃 − 𝛼)
(36.5.17)
1
= − L (𝜋/2 + 𝛼) − L (𝛼 + 𝛾) − L (𝜋/2 − 𝛼) + L (𝛾 − 𝛼)
4
which rearranges to give the result.
662 CHAPTER 36. HYPERBOLIC SPACE
B γ
C
A0
α
C0 Im x
B0
Re x
The series (36.5.20) converges rather slowly, but twice integrating the Laurent series
expansion for cot 𝜃 as in Exercise 36.12 gives
∞
!
∑︁ | 𝐵2𝑛 | (2𝜃) 2𝑛+1
L (𝜃) = 𝜃 1 − log|2𝜃| +
𝑛=1
4𝑛 (2𝑛 + 1)!
where
∞
𝑥 ∑︁ 𝑥 𝑘 𝑥 1 𝑥2 1 𝑥4
= 𝐵 𝑘 = 1 − + − +...
𝑒 𝑥 − 1 𝑘=0 𝑘! 2 6 2! 30 4!
Definition 36.6.1. The group PSL2 (Z[𝑖]) is called the (full) Picard modular group.
Throughout this section, we write Γ = PSL2 (Z[𝑖]). In order to understand the
structure of the group Γ, we follow the same script as in section 35.1, and first we seek
a fundamental set.
Proposition 36.6.2. Let
Then ◊ is a fundamental set for Γ H3 , and PSL2 (Z[𝑖]) is generated by the elements
1 1 1 𝑖 𝑖 0 0 −1
, , , . (36.6.4)
0 1 0 1 0 −𝑖 1 0
The set ◊ in Proposition 36.6.2 is displayed in Figure 36.6.5.
Re x
Im x
Figure 36.6.5: A fundamental set ◊ for the Picard group PSL2 (Z[𝑖]) H3
Proof. First, we show that for all 𝑧 ∈ H3 , there exists a word 𝛾 in the matrices
3
(36.6.4) such that 𝛾𝑧 ∈ ◊ via an explicit reduction
algorithm. Let 𝑧 = 𝑥 + 𝑦 𝑗 ∈ H .
1 𝑏
Recalling the action (36.2.6), the element for 𝑏 ∈ Z[𝑖] act by translation
0 1
𝑧 ↦→ 𝑧 + 𝑏, so repeatedly applying matrices from the first two among
(36.6.4),
we may
𝑖 0
suppose that |Re 𝑥|, |Im 𝑥| ≤ 1/2. Then applying the element , which acts by
0 −𝑖
𝑧 ↦→ (𝑖𝑧) (−𝑖) −1 = 𝑖 2 (𝑥 − 𝑦 𝑗) = −𝑥 + 𝑦 𝑗, we may suppose Im 𝑥 ≥ 0. Now if 𝑧 ∈ ◊,
2 2
which k𝑧k ≥ 1, we are done. Otherwise, k𝑧k < 1, and we apply the matrix
is to say
0 −1
𝛾= , which by (36.2.9) acts by
1 0
1
𝛾𝑧 = 𝑧 0 = (−𝑥 + 𝑦 𝑗) = 𝑥 0 + 𝑦 0 𝑗
k𝑧k 2
664 CHAPTER 36. HYPERBOLIC SPACE
and 𝑦 0 = 𝑦/k𝑧k 2 > 𝑦, and we repeat. Since Γ𝑧 is discrete, this terminates after finitely
many steps: the set
A slightly more convenient set of generators, together with the gluing relations
they provide on the fundamental set, is given in Figure 36.6.7.
Remark 36.6.6. By a deeper investigation into the structure of the fundamental set ◊,
in chapter 37 we will find a presentation for Γ as
y
γ3
i 0
γ2 = 0 −i
i 1
γ4 = 0 −i
√
2
− 21 + 21 i + 2 j
j
0 −1
γ1 = 1 0
√
1 3
2 + 2 j
√
γ2 1
+ 12 i + 2
2 2 j
γ1 γ4
Im x
γ3
Re x
36.6.8. We compute the volume of this fundamental domain using formulas from the
previous section. First, we use symmetry to triangulate (tetrahedralize) ◊, as in Figure
36.6.9.
T
◊
x2 = Im x
x1 = Re x
and
2
vol(◊) = 4 vol(𝑇) = L (𝜋/4) = 0.30532 . . .
3
∞ ∞
1 ∑︁ sin(𝑛𝜋/2) 1 ∑︁ 𝜒(𝑛)
L (𝜋/4) = = (36.6.15)
2 𝑛=1 𝑛2 2 𝑛=1 𝑛2
where
0, if 2 | 𝑛;
𝜒(𝑛) = 1, if 𝑛 ≡ 1 (mod 4);
−1,
if 𝑛 ≡ −1 (mod 4).
is the nontrivial Dirichlet character modulo 4. We can analytically continue the sum
(36.6.15) to C via the 𝐿-series
∞
∑︁ 𝜒(𝑛)
𝐿(𝑠, 𝜒) =
𝑛=1
𝑛𝑠
for 𝑠 ∈ C with Re 𝑠 > 1 whose general study was the heart of Part III of this text. Here,
we can just observe that 𝐿 (2, 𝜒) = 2L (𝜋/4) = 0.915965 . . ., so
1
vol(◊) = vol(Γ\H3 ) = 𝐿(2, 𝜒) = 0.30532 . . .
3
Exercises
𝑎 𝑏
1. For 𝑧 ∈ H3 and 𝑔 = ∈ PSL2 (C), show that
𝑐 𝑑
as in Figure 36.6.16.
z0
1
𝑔𝑧 = (−𝑥 + 𝑦 𝑗)
k𝑧k 2
𝐿(𝑧1 , 𝑧2 ) = bd 𝐻 (𝑧 1 , 𝑧2 ).
Show that 𝐻 (𝑧1 , 𝑧2 ) is a convex half-space (for every two points in the half-space,
the geodesic between them is contained in the half-space), and that
is geodesic and equal to the perpendicular bisector of the geodesic segment from
𝑧1 to 𝑧2 .
10. Prove the duplication formula for the Lobachevsky function L (𝜃) using the
double angle formula, given that L (𝜋/2) = 0.
I 11. In this exercise, we prove the Fourier expansion
∞
1 ∑︁ sin(2𝑛𝜃)
L (𝜃) = . (36.6.17)
2 𝑛=1 𝑛2
Fundamental domains
We have seen in sections 35.1 and 36.6 that understanding a nice fundamental set for
the action of a discrete group Γ is not only useful to visualize the action of the group
by selecting representatives of the orbits, but it is also instrumental for many other
purposes, including understanding the structure of the group itself. In this chapter, we
pursue a general construction of nice fundamental domains for the action of a discrete
group of isometries.
As the group Γ will not vary, we suppress the dependence on Γ and often write
simply ◊(𝑧 0 ) = ◊(Γ; 𝑧0 ).
37.1.2. The set ◊(𝑧0 ) is an intersection
Ù
◊(𝑧0 ) = 𝐻 (𝛾; 𝑧0 ) (37.1.3)
𝛾 ∈Γ
where
In particular, since each 𝐻 (𝛾; 𝑧0 ) is closed, we conclude from (37.1.3) that ◊(𝑧 0 ) is
closed.
671
672 CHAPTER 37. FUNDAMENTAL DOMAINS
H(γ; z0 )
L(γ; z0 )
z0
γ −1 z0
◊(z0 )
Let 𝑧 ∈ H2 . Visibly
so
𝜌(𝑧, 2𝑖) ≤ 𝜌(𝑆𝑧, 2𝑖) ⇔ |𝑧| ≥ 1 (37.1.10)
and 𝐻 (𝑆; 2𝑖) = {𝑧 ∈ H2 : |𝑧| ≥ 1}. To see this geometrically, the geodesic between
2𝑖 and 𝑆(2𝑖) = (1/2)𝑖 is along the imaginary axis with midpoint at 𝑖, and so the
perpendicular bisector 𝐿(𝑆; 2𝑖) is the unit semicircle.
The containment (⊆) in (37.1.8) then follows directly from (37.1.9)–(37.1.10).
Conversely, we show the containment (⊇) for the interior—since ◊(2𝑖) is closed, this
implies the full containment. Let 𝑧 ∈ H2 have |Re 𝑧| < 1/2 and |𝑧| > 1, and suppose
that 𝑧 ∉ ◊(2𝑖); then there exists 𝛾 ∈ PSL2 (Z) such that 𝑧 0 = 𝛾𝑧 has 𝜌(𝑧 0, 2𝑖) < 𝜌(𝑧, 2𝑖),
without loss of generality (replacing 𝑧 0 by 𝑆𝑧 0 or 𝑇 𝑧 0) we may suppose |Re 𝑧 0 | ≤ 1/2
and |𝑧 0 | ≥ 1; but then by Lemma 35.1.10, we conclude that 𝑧 0 = 𝑧, a contradiction.
(Note that the same argument works with 𝑧0 = 𝑡𝑖 for all 𝑡 ∈ R>1 .)
With this example in hand, we see that Dirichlet domains have quite nice structure.
To make this more precise, we upgrade our notion of fundamental set (Definition
34.1.14) as follows.
Definition 37.1.11. A fundamental set ◊ for Γ is locally finite if for each compact set
𝐾 ⊂ H2 , we have 𝛾𝐾 ∩ ◊ ≠ ∅ for only finitely many 𝛾 ∈ Γ.
A fundamental domain for Γ H2 is a fundamental set ◊ ⊆ H2 such that
𝜇(bd ◊) = 0.
Theorem 37.1.12. Let 𝑧 0 ∈ H2 satisfy StabΓ (𝑧0 ) = {1}. Then the Dirichlet domain
◊(Γ; 𝑧0 ) is a connected, convex, locally finite fundamental domain for Γ with geodesic
boundary.
φ(ω)
φ(i) 0 = φ(2i) φ(∞)
φ(−ω 2 )
37.2.5. We now pursue a tidy description of the set (37.2.4). From (37.2.1), we have
𝜌(𝑤, 0) ≤ 𝜌(𝑔𝑤, 0) if and only if
𝑑𝑤 + 𝑐
|𝑤| ≤ ; (37.2.6)
𝑐𝑤 + 𝑑
expanding out (37.2.6) and with a bit of patience (Exercise 37.5), we see that this is
equivalent to simply
|𝑐𝑤 + 𝑑| ≥ 1.
But we can derive this more conceptually, as follows. The hyperbolic metric (Definition
33.7.1) on D2 is invariant, so
|d𝑤| |d(𝑔𝑤)|
d𝑠 = = = d(𝑔𝑠);
(1 − |𝑤|) 2 (1 − |𝑔𝑤|) 2
so, by the chain rule,
2
d𝑔 1 − |𝑔𝑤|
d𝑤 (𝑤) = 1 − |𝑤| .
Therefore
d𝑔 1
|𝑤| ≤ |𝑔𝑤| ⇔ d𝑤 (𝑤) = |𝑐𝑤 + 𝑑| 2 ≤ 1
⇔ |𝑐𝑤 + 𝑑| ≥ 1. (37.2.7)
d𝑔
The equivalence (37.2.7) also gives 𝜌(𝑤, 0) = 𝜌(𝑔𝑤, 0) if and only if (𝑤) = 1,
d𝑤
i.e., 𝑔 acts as a Euclidean isometry at the point 𝑤 (preserving the length of tangent
vectors in the Euclidean metric). So we are led to make the following definition.
Definition 37.2.8. The isometric circle of 𝑔 is
d𝑔
𝐼 (𝑔) = 𝑤 ∈ C : (𝑤) = 1 = {𝑤 ∈ C : |𝑐𝑤 + 𝑑| = 1}.
d𝑤
We have 𝑐 = 0 if and only if 𝑔(0) = 0 if and only if 𝑔 ∈ StabPSU(1,1) (0), and in this
case, 𝑔𝑤 = (𝑑/𝑑)𝑤 with |𝑑/𝑑| = 1 is rotation about the origin. Otherwise, 𝑐 ≠ 0, and
then 𝐼 (𝑔) is a circle with radius 1/|𝑐| and center −𝑑/𝑐 ∈ C.
37.2.9. Summarizing, for all 𝑔 ∈ PSU(1, 1), we have
<
ext(𝐼 (𝑔)),
𝜌(𝑤, 0) = 𝜌(𝑔𝑤, 0) according as 𝑤 ∈ 𝐼 (𝑔),
> int(𝐼 (𝑔)).
In particular, Ù
◊(Γ; 0) = cl ext 𝐼 (𝛾).
𝛾 ∈Γ−StabΓ (0)
{𝛾 ∈ Γ : ◊ ∩ 𝛾◊ ≠ ∅}.
(twice applying the defining property of ◊), so equality holds and (viz. 37.1.2) 𝑧 ∈
bd ◊. Since the boundary of ◊ is geodesic, to understand generators we should
organize the matching provided along the geodesic boundary of ◊.
We will continue to pass between H2 and D2 , as convenient.
37.3. ⊲ GENERATORS AND RELATIONS 677
ideal vertex
vertex
37.3.5. Let 𝑆 denote the set of sides of ◊. We define a labeled equivalence relation
on 𝑆 by
𝑃 = {(𝛾, 𝐿, 𝐿 ∗ ) : 𝐿 ∗ = 𝛾(𝐿)} ⊂ Γ × (𝑆 × 𝑆). (37.3.6)
We say that 𝑃 is a side pairing if 𝑃 induces a partition of 𝑆 into pairs, and we denote
by 𝐺 (𝑃) the projection of 𝑃 to Γ. Since (𝛾, 𝐿, 𝐿 ∗ ) ∈ 𝑃 implies (𝛾 −1 , 𝐿 ∗ , 𝐿) ∈ 𝑃, we
conclude that 𝐺 (𝑃) is closed under inverses.
𝛾 −1 𝐿 = ◊ ∩ 𝛾 −1 ◊ = 𝐿 ∗ ≠ 𝐿
δ2 ◊
δ1 ◊
···
δ1 ◊
w ◊
◊ = δ0 ◊
δn ◊
Figure 37.3.9: Standard picture
L∗1 = ◊ ∩ γ1 ◊ ν = γ ν
2 1 1
L1 = ◊ ∩ γ1−1 ◊ γ2 L2
γ1
ν = ν1 ..
◊ .
γm γm−1
L∗m
L∗m−1
Lm νm
y
γ1−1 γ2−1 ◊ = δ2
δm ◊ = γ1−1 · · · γm
−1
◊
···
δm γ1−1 ◊ = δ1
◊
···
Let 𝐿 1 be the side containing 𝑣 1 traveling clockwise from the interior. Then by the
side pairing (Lemma 37.3.7), there is a paired side 𝐿 1∗ with 𝐿 1∗ = 𝛾1 𝐿 1 and 𝛾1 ∈ 𝐺 (𝑃).
(In fact, then 𝐿 1 = ◊ ∩ 𝛾1−1 ◊ and 𝐿 1∗ = ◊ ∩ 𝛾1 ◊.) Let 𝑣 2 = 𝛾1 𝑣 1 . Then 𝑣 2 is a vertex
of ◊, and so is contained in a second side 𝐿 2 . Continuing in this way, with 𝑣 𝑖+1 = 𝛾𝑖 𝑣 𝑖 ,
by local finiteness we find after finitely many steps a final side 𝐿 ∗𝑚 with next vertex
𝑣 𝑚 = 𝑣1.
In terms of the standard picture (Figure 37.3.9), we see that 𝛿1 = 𝛾1−1 and by
induction 𝛿𝑖 = (𝛾𝑖 · · · 𝛾1 ) −1 , since (𝛾𝑖 · · · 𝛾1 ) (𝑣 1 ) = 𝑣 𝑖+1 . Thus 𝛾𝑖+1 = 𝛿𝑖+1
−1 𝛿 for
𝑖
𝑖 = 0, . . . , 𝑚 − 1. Let 𝛿 = 𝛿 𝑚 . Then 𝛿(𝑣) = 𝑣, and 𝛿 acts by counterclockwise
hyperbolic rotation with fixed point 𝑣—and 𝑚 is the smallest nonzero index with
this property. It follows that for all 0 ≤ 𝑗 ≤ 𝑛, writing 𝑗 = 𝑞𝑚 + 𝑖 with 𝑞 ≥ 0
𝑞
and 0 ≤ 𝑖 < 𝑚 we have 𝛿 𝑗 = 𝛿 𝑚 𝛿𝑖 , and in particular that 𝑚 | (𝑛 + 1). Similarly,
◊ ∩ Γ𝑣 1 = {𝑣 1 , . . . , 𝑣 𝑚−1 }.
Let 𝑒 = (𝑛 + 1)/𝑚. Then 𝛿 𝑒 = 1, and we call this relation the vertex cycle
relation for 𝑣. If 𝑣 0 = 𝛾𝑣, then the vertex cycle relation for 𝑣 0 is the conjugate relation
(𝛾 −1 𝛿𝛾) 𝑒 = 𝛾 −1 𝛿 𝑒 𝛾 = 1. Let 𝑅(𝑃) be the set of (conjugacy classes of) cycle relations
arising from Γ-orbits of vertices in ◊.
Example 37.3.12. We compute the set 𝑅(𝑃) of cycle relations for Γ = PSL2 (Z). The
two Γ-orbits of vertices for ◊ are represented by 𝑖 and 𝜌. The vertex 𝑖 is a fixed point
of 𝛿1 = 𝑆, and we obtain the vertex cycle relation 𝑆 2 = 1, and 𝑆 = 𝑆 −1 .
At the vertex 𝜌, we have a picture as in Figure 37.3.13.
δ1 ◊ = T −1 ◊ δ0 ◊ = ◊
T = γ1
L1 L∗1 δ
δ2 ◊ = T −1 S◊
L∗2 i L2
ρ = ν1 S = γ2 ν2 = −ρ2 ST −1 S◊
Figure 37.3.13: The cycle relation at 𝜌
z3
.. z2
.
z1
zk−2 z0 = zk
zk−1
Let 𝑉 be the intersection of the Γ orbit of the vertices of ◊ with the interior of the
loop; by local finiteness, this intersection is a finite set, and we proceed by induction
on its cardinality. The proof boils down to the fact that this loop retracts onto the loops
around vertices obtained from cycle relations, as H2 is simply connected.
If the path from 𝑧0 → 𝑧 1 crosses the same side as the path 𝑧 𝑘−1 → 𝑧 𝑘 = 𝑧0 , then
𝑧1 = 𝑧 𝑘−1 and so 𝛾 𝑘−1 = 𝛾1 , since StabΓ (𝑧0 ) = {1} (see Figure 37.3.16).
z1 = zk−1
z0 = zk
Conjugating the relation by 𝛾1 and repeating if necessary, we may suppose that 𝛾 𝑘−1 ≠
𝛾1 , so 𝑧 𝑘−1 ≠ 𝑧1 ; the set 𝑉 is conjugated so it remains the same size. In particular, if 𝑉
is empty, then this shows that the original relation is conjugate to the trivial relation.
Otherwise, the path 𝑧0 → 𝑧 1 crosses a side and there is a unique vertex 𝑣 on this
side that is interior to the loop (working counterclockwise): see Figure 37.3.17.
v z1 z10
zk−1 zk−1
z0 = zk z0 = zk
The cycle relation 𝛿 𝑒 = 1 at 𝑣 traces a loop around 𝑣, and without loss of generality we
may suppose 𝛿 𝑒 = 𝛼𝛾1 with 𝛼 a word in 𝐺 (𝑃). Therefore, substituting this relation
into the starting relation, we obtain a relation 𝛾2 · · · 𝛾 𝑘 (𝛿 𝑒 𝛼) with one fewer interior
vertex; the result then follows by induction.
37.4. DIRICHLET DOMAINS 681
𝛾 0𝑥 ∈ ◊ ∩ 𝛾 0 𝛾 −1 ◊
𝑓 : 𝑋 → Γ∗ \Γ
(37.4.4)
𝑥 ↦→ Γ∗ 𝛾
Γ∗ = 𝑓 (𝑥) = 𝑓 (𝛾 −1 𝑥) = Γ∗ 𝛾
We now seek a locally finite fundamental set with other nice properties: we will
choose in each Γ-orbit the closest points to a fixed point 𝑥 0 ∈ 𝑋. So we first must
understand the basic local properties of intersections of these half-spaces (as in 37.1.2).
37.4.5. For 𝑥1 , 𝑥2 ∈ 𝑋, define the closed Leibniz half-space
and
bd 𝐻 (𝑥1 , 𝑥2 ) = 𝐿 (𝑥1 , 𝑥2 ) = {𝑥 ∈ 𝑋 : 𝜌(𝑥, 𝑥 1 ) = 𝜌(𝑥, 𝑥 2 )}
is called the hyperplane bisector (or equidistant hyperplane or separator) between
𝑥1 and 𝑥 2 .
Remark 37.4.8. In this generality, unfortunately hyperplane bisectors are not neces-
sarily geodesic (Exercise 37.10).
Definition 37.4.9. A set 𝐴 ⊆ 𝑋 is star-shaped with respect to 𝑥 0 ∈ 𝐴 if for all 𝑥 ∈ 𝐴,
the geodesic between 𝑥 and 𝑥0 belongs to 𝐴.
A set 𝐴 ⊆ 𝑋 that is star-shaped is path connected, so connected.
Lemma 37.4.10. A Leibniz half-plane 𝐻 (𝑥1 , 𝑥2 ) is star-shaped with respect to 𝑥1 .
Proof. Let 𝑥 ∈ 𝐻 (𝑥 1 , 𝑥2 ) and let 𝑦 be a point along the unique geodesic from 𝑥 to 𝑥1 .
Then
𝜌(𝑥 1 , 𝑦) + 𝜌(𝑦, 𝑥) = 𝜌(𝑥 1 , 𝑥).
If 𝑦 ∉ 𝐻 (𝑥1 , 𝑥2 ), then 𝜌(𝑥 2 , 𝑦) < 𝜌(𝑥1 , 𝑦), and so
each half-space is closed and star-shaped with respect to 𝑥0 , so the same is true of
◊(𝑥0 ). In particular, ◊(𝑥0 ) is connected.
37.4. DIRICHLET DOMAINS 683
Lemma 37.4.13. If 𝐴 ⊂ 𝑋 is a bounded set, then 𝐴 ⊆ 𝐻 (𝛾; 𝑥0 ) for all but finitely
many 𝛾 ∈ Γ.
In particular, for all 𝑥 ∈ 𝑋 we have 𝑥 ∈ 𝐻 (𝛾; 𝑥 0 ) for all but finitely many 𝛾 ∈ Γ.
By Theorem 34.5.1, the orbit Γ𝑥0 is discrete and # StabΓ (𝑥0 ) < ∞; since closed balls
are compact by assumption, there are only finitely many 𝛾 ∈ Γ such that
𝜌(𝛾𝑥0 , 𝑥0 ) = 𝜌(𝑥 0 , 𝛾 −1 𝑥0 ) ≤ 2𝑟
γ4−1 x0
γ1−1 x0
x
A γ −1 x0
ρ (x, x0 )
r ρ x, γ −1 x0
x0
γ3−1 x0
γ2−1 x0
37.4.15. Arguing in a similar way as in Lemma 37.4.13, one can show (Exercise 37.8):
if 𝐾 is a compact set, then 𝐾 ∩ 𝐿 (𝛾; 𝑥0 ) ≠ ∅ for only finitely many 𝛾 ∈ Γ.
int ◊(𝑥0 ) = {𝑥 ∈ ◊ : 𝜌(𝑥, 𝑥 0 ) < 𝜌(𝛾𝑥, 𝑥 0 ) for all 𝛾 ∈ Γ r StabΓ (𝑥0 )}.
It follows from Lemma 37.4.16 that 𝑥 ∈ bd ◊(𝑥 0 ) if and only if there exists
𝛾 ∈ Γ r StabΓ (𝑥0 ) such that 𝜌(𝑥, 𝑥 0 ) = 𝜌(𝛾𝑥, 𝑥 0 ).
684 CHAPTER 37. FUNDAMENTAL DOMAINS
the intersection over finitely many 𝛾𝑖 ∈ Γ with 𝛾𝑖 ∉ StabΓ (𝑥0 ) (see Figure 37.4.17).
x
x
U
Figure 37.4.17: The neighborhood 𝑈
Thus Ù
𝑈 ∩ int(◊) = 𝑈 ∩ int 𝐻 (𝑥0 , 𝛾𝑖−1 𝑥0 ).
𝑖
The lemma then follows from (37.4.7).
Proof. Abbreviate ◊ = ◊(𝑥 0 ). We saw that ◊ is (closed and) star-shaped with respect
to 𝑥0 in 37.4.12. Ð
Now we show that 𝑋 = 𝛾 𝛾◊. Let 𝑥 ∈ 𝑋. The orbit Γ𝑥 is discrete, so the distance
a contradiction.
37.5. HYPERBOLIC DIRICHLET DOMAINS 685
Finally, we show that 𝑋 is locally finite. It suffices to check this for a closed disc
𝐾 ⊆ 𝑋 with center 𝑥0 and radius 𝑟 ∈ R ≥0 . Suppose that 𝛾𝐾 meets ◊ with 𝛾 ∈ Γ; then
by definition there is 𝑥 ∈ ◊ such that 𝜌(𝑥 0 , 𝛾 −1 𝑥) ≤ 𝑟. Then
𝜌(𝑥 0 , 𝛾 −1 𝑥 0 ) ≤ 𝑟 + 𝑟 = 2𝑟.
γK
x
x ∈ γK ∪ ◊
x0
γ −1 x γ −1 x ∈ K
γ −1 x0
K
Figure 37.4.21: ◊ is locally finite and star-shaped
For the same reason as in Lemma 37.4.13, this can only happen for finitely many
𝛾 ∈ Γ.
Remark 37.4.22. Dirichlet domains are sometimes also called Voronoi domains.
37.5.2. In Theorem 37.4.18, the hypothesis that StabΓ (𝑧0 ) = {1} is a very mild
hypothesis. If 𝐾 is a compact set, then since ◊ is locally finite, the set of points 𝑧 ∈ 𝐾
with StabΓ (𝑧) ≠ {1} is a finite set of points when H = H2 and a finite set of points
together with finitely many geodesic axes when H = H3 .
686 CHAPTER 37. FUNDAMENTAL DOMAINS
In spite of 37.5.2, we prove a slightly stronger and more useful version of Theorem
37.4.18, as follows. If Γ0 = StabΓ (𝑧 0 ) is nontrivial, we modify the Dirichlet domain
by intersecting with a fundamental set for Γ0 ; the simplest way to do this is just to
choose another point which is not fixed by an element of Γ0 and intersect.
Theorem 37.5.3. Let 𝑧0 ∈ H , let Γ0 = StabΓ (𝑧 0 ), and let 𝑢 0 ∈ H be such that
StabΓ0 (𝑢 0 ) = {1}. Then
◊(Γ; 𝑧 0 ) ∩ ◊(Γ0 ; 𝑢 0 )
is a connected, convex, locally finite fundamental domain for Γ with geodesic boundary
in H .
Proof. Abbreviate
◊ = ◊(Γ; 𝑧0 ) ∩ ◊(Γ0 ; 𝑢 0 ).
First, we show that 𝑧 0 ∈ ◊: we have 𝑧 0 ∈ ◊(Γ; 𝑧 0 ), and by Theorem 37.4.18,
◊(Γ0 ; 𝑢 0 ) is a fundamental set for Γ0 so there exists 𝛾0 ∈ Γ0 such that 𝛾0 𝑢 0 = 𝑢 0 ∈
◊(Γ0 ; 𝑢 0 ).
Ð
Now we show that ◊ is a fundamental set for Γ. First we show H = 𝛾 𝛾◊. Let
𝑧 ∈ H , and let 𝛾 ∈ Γ be such that 𝜌(𝛾𝑧, 𝑧0 ) is minimal as in (37.4.19). Let 𝛾0 ∈ Γ0
such that 𝛾0 (𝛾𝑧) ∈ ◊(Γ0 ; 𝑢 0 ). Then
𝜌(𝛾0 (𝛾𝑧), 𝑧0 ) = 𝜌(𝛾𝑧, 𝑧0 )
so 𝛾0 𝛾𝑧 ∈ ◊. And int(◊) ∩ int(𝛾◊) = ∅ for all 𝛾 ∈ Γ r {1}, because if 𝑧, 𝛾𝑧 ∈ int(◊)
with 𝛾 ≠ 1, then either 𝛾 ∉ Γ0 in which case we obtain a contradiction as in (37.4.20),
or 𝛾 ∈ Γ0 − {1} and then we arrive at a contradiction from the fact that ◊(Γ0 ; 𝑢 0 ) is a
fundamental set.
We conclude by proving the remaining topological properties of ◊. We know that
◊ is locally finite, since it is the intersection of two locally finite sets. We saw in 37.1.2
that the Leibniz half-spaces in H2 are convex with geodesic boundary, and the same is
true in H3 by Exercise 36.9. It follows that ◊ is convex, as the intersection of convex
sets. Thus Ø
bd ◊ ⊆ 𝐿 (𝑧0 , 𝛾 −1 𝑧0 )
𝛾 ∈Γr{1}
is geodesic and measure zero, since 𝐿 (𝑧0 , 𝛾 −1 𝑧0 ) intersects a compact set for only
finitely many 𝛾 by 37.4.15.
Put another way, ◊ satisfies the cycle condition if the sum of the interior angles for a
Γ-orbit of vertices as in the standard picture is an integer submultiple of 2𝜋.
Now suppose ◊ ⊆ H3 . Now we work with edges: for an edge ℓ of ◊, let 𝜗(◊, ℓ)
be the interior angle of ◊ at ℓ. We say that ◊ satisfies the cycle condition if for all
edges ℓ of ◊ there exists 𝑒 ∈ Z>0 such that
∑︁ 2𝜋
𝜗(◊, ℓ𝑖 ) = .
ℓ𝑖 ∈Γ𝑒∩◊
𝑒
Lemma 37.6.2. Let ◊ be a Dirichlet domain. Then ◊ satisfies the cycle condition.
Proof. We explain the case where ◊ ⊆ H2 ; the case of H3 is similar. Let 𝑣 be a vertex
of ◊. Referring to the standard picture 37.3.8,
𝑛
∑︁ 𝑛
∑︁
2𝜋 = 𝜗(𝛿 𝑗 ◊, 𝑣) = 𝜗(◊, 𝛿−1
𝑗 𝑣).
𝑗=0 𝑗=0
In 37.3.10, we proved that 𝛿 𝑚 acts by hyperbolic rotation around 𝑣 and has order
𝑒 = (𝑛 + 1)/𝑚, and
𝑛
∑︁ 𝑚−1
∑︁ 𝑚−1
∑︁
𝜗(◊, 𝛿−1
𝑗 𝑣) = 𝑒 𝜗(◊, 𝛿𝑖−1 𝑣) = 𝑒 𝜗(◊, 𝑣 𝑖 )
𝑗=0 𝑖=0 𝑖=0
with ◊ ∩ Γ𝑣 = {𝑣 1 , . . . , 𝑣 𝑚−1 }. Combining these two equations, we see that the cycle
condition is satisfied.
Proof. Unfortunately, it is beyond the scope of this book to give a complete proof of
Theorem 37.6.4. See Epstein–Petronio [EP94, Theorem 4.14] or Ratcliffe [Rat2006,
§11.2, §13.5]; our statement is a special case of the theorem by Maskit [Mas71], but
see Remark 37.6.5.
Remark 37.6.5. The proof of Poincaré’s theorem [Poi1882, Poi1883] has a bit of a
notorious history. From the very beginning, to quote Epstein–Petronio [EP94, §9,
p. 164]:
688 CHAPTER 37. FUNDAMENTAL DOMAINS
It is clear that Poincaré understood very well what was going on. How-
ever, the papers are not easy to read. In particular, the reader of the
three-dimensional case is referred to the treatment of the two-dimensional
case for proofs; this is fully acceptable for a trail-blazing paper, but not
satisfactory in the long term.
There are a number of reasonable published versions of Poincaré’s The-
orem in dimension two. Of these, we would single out the version by
de Rham [dR71] as being particularly careful and easy to read. Most
published versions of Poincaré’s Theorem applying to all dimensions are
unsatisfactory for one reason or another.
This sentiment is echoed by Maskit [Mas71], who presents a proof for polygons with
an extension to polyhedron, with the opening remark:
Epstein and Petronio [EP94, §9, p. 165] then have this to say:
Maskit published a book [Mas88] containing an expanded version of the proof for
polyhedra, to which Epstein and Petronio [EP94, §9, p. 164] review:
Proof. The proof estimates the contribution to the volume from infinite vertices: for
Fuchsian groups, see Siegel [Sie45, p. 716–718], or the expositions of this argument
by Gel’fand–Graev–Pyatetskii-Shapiro [GG90, Chapter 1, §1.5] and Katok [Kat92,
Theorem 4.1.1].
37.7.3. By Theorem 37.5.3 and Siegel’s theorem (Theorem 37.7.2), a Dirichlet domain
◊ for Γ is connected, convex, hyperbolic polygon. In particular, there are 𝑚 ∈
Z ≥0 vertex cycles, which by 37.3.10 correspond to cyclic stabilizer groups of orders
𝑒 1 , . . . , 𝑒 𝑚 ∈ Z ≥1 listed so that 𝑒 1 , . . . , 𝑒 𝑘 ≥ 2, and finitely many 𝛿 ∈ Z ≥0 infinite
vertex cycles, corresponding to 𝛿 stabilizer groups that are infinite cyclic.
Proposition 37.7.4. We have
𝑘 !
2
∑︁ 1
𝜇(Γ\H ) = 2𝜋 (2𝑔 − 2) + 1− +𝛿 .
𝑖=1
𝑒𝑖
Proof. Let ◊ be a Dirichlet domain for Γ with 2𝑛 sides and 𝑛 finite or infinite vertices.
The hyperbolic area of ◊ is given by the Gauss–Bonnet formula 33.6.9: summing
vertex cycles using the cycle condition (Lemma 37.6.2), we have
𝑘
∑︁ 2𝜋
𝜇(◊) = (2𝑛 − 2)𝜋 − .
𝑖=1
𝑒𝑖
2 − 2𝑔 = (𝑘 + 𝛿) − 𝑛 + 1
so
𝑛 − 1 = 2𝑔 − 2 + (𝑘 + 𝛿).
690 CHAPTER 37. FUNDAMENTAL DOMAINS
Therefore
𝑘
! 𝑘 !
∑︁ 1 ∑︁ 1
𝜇(◊) = 2𝜋 (2𝑔 − 2) + (𝑘 + 𝛿) − = 2𝜋 (2𝑔 − 2) + 1− +𝛿
𝑒
𝑖=1 𝑖 𝑖=1
𝑒𝑖
as claimed.
z−i
i z 7→ w = z+i
w
π/6 π/6
z
Figure 37.8.1: The hyperbolic triangle 𝑇 with angles 𝜋/2, 𝜋/6, 𝜋/6
By symmetry, we may define the side-pairing 𝑃 as shown in Figure 37.8.1. This
polygon satisfies the cycle condition, so by the Poincaré polygon theorem (Theorem
37.6.4), there exists a Fuchsian group Δ ⊂ PSL2 (R) generated by the two side pairing
elements in 𝑃 and with fundamental domain 𝑇. In this section, we construct this group
explicitly and observe some interesting arithmetic consequences.
37.8.2. We seek the position of the vertex 𝑧 ∈ H2 corresponding to 𝑤 ∈ D2 . Zooming
in, we obtain the picture as in Figure 37.8.3.
w
1 π/6
c
π/4
− dc π/12 0
2
|cw + d| = 1
p
d2 −c2 = 1 ⇒ d = 1 + c2
Figure 37.8.3: Finding the position of 𝑧
37.8. THE (6, 4, 2)-TRIANGLE GROUP 691
The edge containing 𝑤 and its complex conjugate is defined by an isometric circle
|𝑐𝑤 + 𝑑| = 1 with 𝑑 2 − 𝑐2 = 1 and by symmetry 𝑐, 𝑑 ∈ R>0 . With the angles as labeled,
we find that 𝑤 = (−𝑑 + 𝑒 𝜋𝑖/12 )/𝑐, and since arg(𝑤) = 3𝜋/4 we compute that
√︁
1 + 𝑐2 − cos(𝜋/12) = sin(𝜋/12) (37.8.4)
√ √︁
so 𝑐2 = 2 cos(𝜋/12) sin(𝜋/12) = sin(𝜋/6) √ = 1/2 thus 𝑐 = 1/ 2 and 𝑑 = √3/2, so
this isometric circle is defined by |𝑤 + 3| 2 = √ 2, and 𝑤 = (−1 + 𝑖)/(1 + 3). By
coincidence, we find that 𝑧 = 𝑧2 = (−1 + 𝑖)/(1 + 3)√= 𝑤 as well.√ The intersection of
this circle with the imaginary axis is the point 𝑧3 = ( 3 − 1)𝑖/ 2.
The unique element mapping the sides meeting at 𝑖 is obtained by pulling back
rotation by −𝜋/4 in the unit disc model; it is given by the matrix
−1
1 −𝑖 𝑒 − 𝜋𝑖/4
0 1 −𝑖 1 1 −1
𝛿4 = = √ (37.8.5)
1 𝑖 0 𝑒 𝜋𝑖/4 1 𝑖 2 1 1
with
𝑧−1
𝛿4 (𝑧) = , for 𝑧 ∈ H2 ,
𝑧+1
and 𝛿44 = 1 ∈ PSL2 (R). From a similar computation, we find that the other side pairing
element acting by hyperbolic rotation around 𝑧 3 is
√
1 0√ −1 + 3
𝛿2 = √ (37.8.6)
2 −1 − 3 0
z1 = i
1 −1
δ4 = · √12
1 1
√
0√ −1 + 3 √1
δ2 = ·
−1 − 3 0 2
√
π/6 3−1
√
2
i = z3
−1+i z4 √ √
√
1+ 3
= z2 −1 1 1 + √3 1 − √3
δ6 = (δ2 δ4 ) = 2 1+ 3 −1 + 3
Throwing
37.8.12. away primes from the previous paragraph, we have the algebra
−1, 3
𝐵 := of discriminant disc 𝐵 = 6 and order
Q
1 + 𝑖 + 𝑗 + 𝑖𝑗
O := Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑘, 𝑘= (37.8.13)
2
with 𝑘 2 − 𝑘 − 1 = 0. (By 23.1.1, since discrd(O) = disc 𝐵 = 6, we conclude that O is
a maximal order in 𝐵.)
This algebra came with the embedding
𝜄∞ : 𝐵 ↩→ M2 (R)
√
0 −1 3 0
√
𝑖, 𝑗 ↦→ ,
1 0 0 − 3
√ √
𝑡 + 𝑦 √3 −𝑥 + 𝑧√ 3
𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑖 𝑗 ↦→ = 𝑔.
𝑥+𝑧 3 𝑡−𝑦 3
37.9. UNIT GROUP FOR DISCRIMINANT 6 693
◊ = ◊(Γ; 0) ∩ {𝑤 ∈ D2 : Re 𝑤 ≤ 0}
3 1 1 1 1 3 1 1
𝑔 = ± ± 𝑖 ± 𝑗 ± 𝑖𝑗 or 𝑔 = ± ± 𝑖 ± 𝑗 ± 𝑖 𝑗.
2 2 2 2 2 2 2 2
√︁
All of these elements
√ have the radius of 𝐼√(𝑔) equal to 2/3 = 0.82 . . ., and the centers
are ±1.15 ± 0.57 −1 and ±0.57 ± 1.15 −1. The corresponding external domain is
given in Figure 37.9.3.
the external domain. The corresponding external domain looks like Figure 37.9.5.
It follows that the ◊ is cut out by the left half with side pairing as in Figure 37.9.6.
γ2
γ1
γ3
Figure 37.9.6: ◊ ⊂ D2
696 CHAPTER 37. FUNDAMENTAL DOMAINS
Pulling back to the upper half-plane, we obtain the domain in Figure 37.9.7.
√
ν1 = −1
γ1 = i
√ √
− 3+ −1
2 = ν2 γ2 = −1 + i + k γ3 = −1 + i − j + k
√
−1+√−1
ν6
= ν3
1+ 3 ν5
√ √
ν4 = (2 − 3) −1
Figure 37.9.7: ◊ ⊂ H2
The corresponding tessellation of the upper half-plane looks like the one in Figure
37.9.8.
0
Figure 37.9.8: Tessellation of H2 by Γ1
In addition to this side pairing, we check the cycle relations on the four orbits of
vertices: the fixed points 𝑣 1 , 𝑣 3 , 𝑣 5 and the orbit 𝑣 2 , 𝑣 4 , 𝑣 6 . In conclusion,
The order of the stabilizers tell us the angles at each vertex, and so by the Gauss–Bonnet
formula (Theorem 33.6.8) we compute that the area is
37.9.10. Finally, the quotient 𝑋 (Γ) = Γ\H2 has the structure of a good complex 1-
orbifold (see 34.8.14), a Riemann surface but with finitely many orbifold points. Since
the fundamental domain ◊ is compact, via the continuous surjective projection map
◊ → 𝑋 (Γ) we see that 𝑋 (Γ) is compact, and 𝜇(𝑋 (Γ)) = 𝜇(◊) = 2𝜋/3. This orbifold
‘folds’ up to a surface with topological genus 0, so the signature 34.8.13 of 𝑋 (Γ) is
(0; 2, 2, 3, 3): see Figure 37.9.11.
ν2 , ν4 , ν6
orbifold signature
ν3 ν5
(0; 2, 2, 3, 3)
ν1
We have seen that the norm 1 group contains the (2, 4, 6)-triangle group Δ, so we
have a map 𝑋 (Γ) → 𝑋 (Δ) = Δ\H2 ; by Gauss–Bonnet, we have
we see
[Δ : Γ] = 𝜇(𝑋 (Γ))/𝜇(𝑋 (Δ)) = (2𝜋/3)/(𝜋/6) = 4,
4
as is visible from Figure 37.9.12 for Γ ≤ Δ.
Remark 37.9.13. The discriminant 6 quaternion algebra has been a favorite to study,
and the fundamental domain described above is also given by e.g. Alsina–Bayer
698 CHAPTER 37. FUNDAMENTAL DOMAINS
Exercises
1. Let Γ = PSL2 (Z). Describe the Dirichlet domain ◊(𝑧) for an arbitrary 𝑧 ∈ H2
with Im 𝑧 > 1.
2. Let Γ = PSL2 (Z[𝑖]) (cf. section 36.6). Show that
and that
𝑖 0
StabΓ (2 𝑗) = ' Z/2Z
0 −𝑖
so ◊(Γ; 2 𝑗) is a union of two copies of a fundamental set for Γ.
3. LetΓ be the cyclic
Fuchsian group generated by the isometry 𝑧 ↦→ 4𝑧, represented
2 0
by ∈ PSL2 (R). Give an explicit description of the Dirichlet domain
0 1/2
◊(Γ; 𝑖) ⊂ H2 and its image ◊(Γ; 0) ⊂ D2 with 𝑖 ↦→ 0.
𝑎 𝑏
4. Let 𝑔 = ∈ PSL2 (R) satisfy 𝑔𝑖 ≠ 𝑖. Let 𝐻 be the perpendicular bisector
𝑐 𝑑
between 𝑖 and 𝑔𝑖.
(a) Show that k𝑔k 2 > 2.
(b) Under the map H2 → D2 taking 𝑖 ↦→ 0, show that the perpendicular
bisector between 𝑖 and 𝑔𝑖 is the isometric circle 𝐼 (𝑔 −1 ) inside D2 .
k𝑔k 2 − 2
(c) Show that 𝐻 is the half-circle of square radius 2 centered at
(𝑐 + 𝑑 2 − 1) 2
𝑎𝑐 + 𝑏𝑑
∈ R, where k𝑔k 2 = 𝑎 2 + 𝑏 2 + 𝑐2 + 𝑑 2 . [Hint: as a check along
𝑐 + 𝑑2 − 1
2
the way, 𝐻 is described by the equation | (𝑑 − 𝑖𝑐)𝑧 + 𝑖(𝑎 + 𝑖𝑏)| = | 𝑧 + 𝑖|.]
5. Let
𝑑 𝑐
𝑔= ∈ PSU(1, 1) D2
𝑐 𝑑
with 𝑐, 𝑑 ∈ C satisfying |𝑑| 2 − |𝑐| 2 = 1. Show directly that |𝑔𝑤| = |𝑤| for
𝑤 ∈ D2 if and only if
|𝑐𝑤 + 𝑑| = 1
by expanding out and simplifying.
6. Show that for all 𝑔 ∈ PSU(1, 1), we have 𝑔𝐼 (𝑔) = 𝐼 (𝑔 −1 ), where 𝐼 (𝑔) is the
isometric circle of 𝑔. (Equivalently, show that if 𝑔 ∈ PSL2 (R) that 𝑔𝐿 (𝑔; 𝑧0 ) =
𝐿(𝑔 −1 ; 𝑧 0 ) for all 𝑧0 ∈ H2 .)
I 7. Let ◊ be a Dirichlet domain for a Fuchsian group Γ. Show that in every compact
set, there are only finitely many sides and finitely many vertices of ◊.
37.9. UNIT GROUP FOR DISCRIMINANT 6 699
◊(Γ; 𝑧 0 ) ∩ ◊(Γ0 ; 𝑢 0 )
is a locally finite fundamental set for Γ that is star-shaped with respect to 𝑥0 and
whose boundary consists of hyperplane bisectors.
10. Consider the egg of revolution, a surface of revolution obtained from convex
curves with positive curvature as in Figure 37.9.14.
top
geodesic
bisector
bottom
An egg of revolution has the structure of a geodesic space with the induced
metric from R3 . Show that the separator between the top and bottom, a circle of
revolution, is not geodesic.
[In fact, Clairaut’s relation shows that the geodesic joining two points in the
same circle of revolution above crest in the 𝑥-axis never lies in this circle of
revolution.]
11. In this exercise, we consider a Fuchsian group constructed from a regular quadri-
lateral.
(a) Show that for every 𝑒 ≥ 2, there exists a regular (equilateral and equian-
gular) quadrilateral ◊ ⊂ D2 , unique up to isometry, with interior angle
𝜋/(2𝑒).
Conclude from Poincaré’s theorem that there is a Fuchsian group, unique
up to conjugation in PSL2 (R), with fundamental domain ◊ and side pairing
as in Figure 37.9.15.
(b) Give a presentation for this group for all 𝑒 ≥ 2, and find explicit matrix
700 CHAPTER 37. FUNDAMENTAL DOMAINS
In this chapter, we now apply our topological and geometric interpretation of discrete
groups in our case of interest: quaternion unit groups.
The wonderful thing that gives life to this part of our monograph is this: the same
thing works if we replace M2 (Q) with a quaternion algebra 𝐵 over a global field! In
this section, we derive key aspects of this program for quaternion algebras over Q in a
self-contained way, before embarking on a more general study in the remainder of the
chapter.
In section 32.1, we already dealt with the case when the quaternion
algebra was
𝑎, 𝑏
definite, finding a finite unit group; so here we take 𝐵 = indefinite. Without
Q
loss of generality (for convenience of presentation), we may
further
suppose 𝑎, 𝑏 > 0
𝑎, 𝑏
and both 𝑎, 𝑏 ∈ Z. To be indefinite means that 𝐵 ⊗Q R = ' M2 (R); we obtain
R
such an embedding via a conjugate of the left-regular representation (2.3.2)
𝜄∞ : 𝐵 ↩→ M2 (R)
√ √ √
(38.1.1)
𝑡 + 𝑥 𝑎√ (𝑦 + 𝑧 𝑎) 𝑏
𝑡 + 𝑥𝑖 + 𝑦 𝑗 + 𝑧𝑘 ↦→ √ √
(𝑦 − 𝑧 𝑎) 𝑏 𝑡 − 𝑥 𝑎.
This embedding is not unique, but another embedding would correspond to post-
composition by an R-algebra automorphism of M2 (R), which by the Skolem–Noether
701
702 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
is a closed, locally finite fundamental domain for Γ1 (O) with geodesic sides by Theo-
rem 37.1.12.
In this way, we have completed our task: starting with an indefinite quaternion
algebra 𝐵 over Q, we constructed a Fuchsian group Γ1 (O) generalizing PSL2 (Z) acting
on the upper half-plane. We pursue a more general construction in this chapter (con-
sidering an indefinite quaternion over a number field) and its geometry and arithmetic
properties in the remainder of this book.
38.2.1. Suppose that 𝐵 is split at 𝑡 real places and ramified at the remaining 𝑟 − 𝑡 real
places, so that
∼ M (R) 𝑡 × H𝑟 −𝑡 × M (C) 𝑐 .
𝐵 ↩→ 𝐵R := 𝐵 ⊗Q R −→ (38.2.2)
2 2
Let
𝜄 = (𝜄1 , . . . , 𝜄𝑡+𝑐 ) : 𝐵 → M2 (R) 𝑡 × M2 (C) 𝑐 (38.2.3)
704 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
denote the map (38.2.2) composed with the projection onto the matrix ring factors.
Then as long as 𝑡 + 𝑐 > 0, we have
𝜄(𝐵× ) ⊂ GL2 (R) 𝑡 × GL2 (C) 𝑐 . (38.2.4)
We have 𝑡 = 𝑐 = 0 if and only if 𝐹 is totally real and 𝐵 is totally definite; in this
case, the geometry disappears, and we are reduced to considering finite unit groups
(see section 32.3). There is still more to say for this case, and we will return to it in
chapter 41 in the context of modular forms. But in this part of the text we are following
geometric threads, so
suppose from now on that 𝑡 + 𝑐 > 0; (38.2.5)
this is another way of saying that 𝐵 is indefinite or equivalently satisfies the Eichler
condition (Definition 28.5.1).
38.2.6. We now restrict to a subgroup acting faithfully via orientation-preserving
isometries. Recall that
×
𝐹>0 = {𝑥 ∈ 𝐹 × : 𝑣(𝑥) > 0 for all real places 𝑣}
is the group of totally positive elements of 𝐹. Let
𝐵×>0 := {𝛼 ∈ 𝐵× : nrd(𝛼) ∈ 𝐹>0
×
} (38.2.7)
be the group of units of 𝐵 of totally positive reduced norm. Then 𝐹 × ⊂ 𝐵×>0 because
× for all 𝑎 ∈ 𝐹 × . Let
nrd(𝑎) = 𝑎 2 ∈ 𝐹>0
P𝐵×>0 := 𝐵×>0 /𝑍 (𝐵×>0 ) = 𝐵×>0 /𝐹 × (38.2.8)
be the quotient by the center; then 𝜄 induces an inclusion
P𝜄(𝐵×>0 ) ⊂ PSL2 (R) 𝑡 × PSL2 (C) 𝑐
(we have PSL2 (C) ' PGL2 (C) and PGL+2 (R) ' PSL2 (R), rescaling by the determi-
nant). Therefore, the group P𝜄(𝐵×>0 ) acts on
where 𝑆 𝑚 denotes the symmetric group. Since PSL2 (R) ' Isom+ (H2 ) ≤ Isom(H2 )
has index 2 and the same for H3 , we conclude that
[Isom(H ) : PSL2 (R) 𝑡 × PSL2 (C) 𝑐 ] = 2𝑡+𝑐 𝑡!𝑐!.
The orientation-preserving subgroup Isom+ (H ) ≤ Isom(H ) has index 2, contain-
ing for example elements which reverse orientation in two components and preserve
orientation in the others; if the orientation is reversed in two components H2 , then the
resulting isometry is orientation-preserving but not holomorphic in these components.
38.3 Discreteness
We now seek discrete subgroups, as follows.
38.3.1. Let 𝑅 = Z𝐹 be the ring of integers of 𝐹 and let O ⊂ 𝐵 be an 𝑅-order. Let
O1 := {𝛾 ∈ O× : nrd(𝛾) = 1} ≤ O× (38.3.2)
be the subgroup of units of reduced norm 1, and let
PO1 := O1 /𝑍 (O1 ) = O1 /{±1}
(38.3.3)
Γ1 (O) := P𝜄(O1 )
Then by (38.2.10), we have Γ1 (O) < Isom+ (H ).
Definition 38.3.4. A subgroup Γ ≤ Isom+ (H ) is arithmetic if Γ is commensurable
with Γ1 (O) for a quaternion algebra 𝐵 and an order O ⊆ 𝐵 (with respect to some
embedding 𝜄).
If Γ is commensurable with Γ1 (O) for an order O then it is commensurable with
Γ1 (O0) for every other order O0, since every two orders have finite 𝑅-index, thus finite
index—so we could equally well compare to one fixed (e.g. maximal) order. The class
of arithmetic groups contains in particular the quaternionic unit groups with which
we started, but contains other groups of interest (including subgroups and discrete
supergroups with finite index).
Remark 38.3.5. There is a more general definition of arithmetic group which reduces
to this one; see section 38.5.
Right away, we show that arithmetic groups are discrete. To do so, we will need
two short lemmas.
Lemma 38.3.6. Let 𝐾, 𝑋 be Hausdorff topological spaces, and suppose 𝐾 is compact.
Let 𝜋 : 𝐾 × 𝑋 → 𝑋 be the projection, and let 𝑌 ⊆ 𝐾 × 𝑋 be discrete and closed. Then
𝜋(𝑌 ) ⊆ 𝑋 is discrete.
Proof. First, 𝑌 has no limit points: a limit point of 𝑌 would belong to 𝑌 , but then 𝑌 is
discrete so every point is isolated point. For the same reason, every subset of 𝑌 is also
(discrete and) closed: a limit point of the subset would be a limit point of 𝑌 .
Now let 𝑥 ∈ 𝜋(𝑌 ) and let 𝑌𝑥 = 𝑌 r 𝜋 −1 (𝑥). Then 𝑌𝑥 ⊆ 𝐾 × 𝑋 is closed. The set
(𝐾 × 𝑋) r 𝑌𝑥 is open and contains 𝐾 × {𝑥}, so by the tube lemma, it contains an open
set 𝐾 × 𝑈. Then 𝑈 3 𝑥 is the desired neighborhood.
706 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
38.4.1. The arithmetic discrete groups arising from the case 𝐵 = M2 (𝐹) of the matrix
ring are of particular interest: they include the case 𝐹 = Q giving rise to the classical
modular group Γ = PSL2 (Z) studied in chapter 35 as well as the case 𝐹 = Q(𝑖) giving
rise to the Picard group Γ = PSL2 (Z[𝑖]) examined in section 36.6 and more generally
the Bianchi groups. Let 𝐵 = M2 (𝐹). Then necessarily 𝑟 = 𝑡 (the matrix ring is
already split!), so H = (H2 ) 𝑟 × (H3 ) 𝑐 , and the embedding 𝜄 in (38.2.3) has the simple
description
𝜄 : M2 (𝐹) ↩→ M2 (𝐹)R ' M2 (R) 𝑟 × M2 (C) 𝑐
𝑎 𝑏 𝑎𝑣 𝑏𝑣 (38.4.2)
𝛼= ↦→
𝑐 𝑑 𝑐𝑣 𝑑𝑣 𝑣
where we embed matrices componentwise, indexed by the archimedean places of 𝐹.
Much is written about the compactification of 𝑌 (Γ), and unfortunately it would
take us too far afield to fully treat this important topic: the groups so obtained behave
differently in several respects than the case where 𝐵 is a division algebra.
Proof. We follow Zassenhaus [Zas72, §1]; see also Kleinert [Klt2000, Theorem 1.1].
First, the group Γ is commensurable with 𝜄(O1 ) for an 𝑅-order O; by comparison
under maps of finite index, we may suppose Γ = Γ1 (O).
Second, we claim that it suffices to show that O1 \𝐵R1 is compact. Indeed, recall
(38.3.9) that
O1 ↩→ 𝐵R1 ' (H1 ) 𝑟 −𝑡 × SL2 (R) 𝑡 × SL2 (C) 𝑐 .
From the symmetric space models (34.6.3) and (36.3.14), we have homeomorphisms
∼ H2
SL2 (R)/SO(2) −→
SL (C)/SU(2) −→∼ H3 ;
2
are given by
(H1 ) 𝑟 −𝑡 × SO2 (R) 𝑡 × SU(2) 𝑐
and therefore compact, and the claim follows.
Now from (38.2.2), we have
choosing an R-basis for 𝐵R , we have 𝐵R ' R4𝑛 and under the standard metric, this
gives O ' Z4𝑛 the (non-canonical) structure of a Euclidean lattice. Let 𝑋 ⊆ 𝐵R
be a compact, convex, symmetric subset with volume vol(𝑋) > 24𝑛 covol(O). (The
precise value of this covolume will not figure in the argument, and anyway would
depend on the choice of Euclidean structure; all such structures induce the same
topology.) Therefore, by Minkowski’s convex body theorem (Theorem 17.5.5), there
exists nonzero 𝛼 ∈ O ∩ 𝑋. Moreover, for all 𝑔 = (𝑔 𝑣 ) 𝑣 ∈ 𝐵R1 we have
vol(𝑔𝑋) = vol(𝑋)
Î
since 𝑣 det(𝑔 𝑣 ) = 1, and 𝑔𝑋 is again compact, convex, and symmetric, so similarly
there exists nonzero 𝛼𝑔 ∈ O ∩ 𝑔𝑋.
We will show that the quotient space O1 \𝐵R1 is sequentially compact. To this end,
let 𝑔𝑛 be a sequence from 𝐵R1 . By the previous paragraph, there exist 𝛼𝑛 ∈ O such that
𝛼𝑛 = 𝑔𝑛 𝑥 𝑛 with 𝑥 𝑛 ∈ 𝑋 nonzero. Since 𝑋 is compact, we can restrict to a subsequence
such that 𝑥 𝑛 → 𝑥 ∈ 𝑋 converges.
The reduced norm nrd : 𝐵 → 𝐹 extends by scaling to a continuous function
nrd : 𝐵R → 𝐹R . Since 𝑋 ⊆ 𝐵R is bounded so too is nrd(𝑋) ⊆ 𝐹R bounded. But
nrd(𝛼𝑛 ) = nrd(𝑔𝑛 ) nrd(𝑥 𝑛 ) = nrd(𝑥 𝑛 ) ∈ nrd(𝑋),
and the values nrd(𝛼𝑛 ) ∈ nrd(O) ⊆ Z𝐹 lie in a discrete subset, so there are only
finitely many possibilities for nrd(𝛼𝑛 ). Moreover, the left ideals 𝐼𝑛 = O𝛼𝑛 have
N(𝐼𝑛 ) = Nm𝐹 |Q (nrd(𝐼𝑛 )) 2 = Nm𝐹 |Q (nrd(𝛼𝑛 )) 2 (38.4.4)
bounded, so there are only finitely many possibilities for 𝐼𝑛 by Lemma 17.7.26. Thus,
by the pigeonhole principle and restricting to a subsequence, we may suppose 𝐼𝑛 =
O𝛼𝑛 = O𝛼1 and nrd(𝛼𝑛 ) = nrd(𝛼1 ) for all 𝑛. Therefore 𝛼𝑛 = 𝛾𝑛 𝛼1 with 𝛾𝑛 ∈ O1 for
all 𝑛.
To conclude, we note that since 𝐵 is a division algebra,
nrd(𝑥 𝑛 ) = nrd(𝛼𝑛 ) = nrd(𝛼1 ) ≠ 0
so
𝑥 𝑛−1 = 𝑥 𝑛 nrd(𝑥 𝑛 ) −1 = 𝑥 𝑛 nrd(𝛼1 ) −1 → 𝑥 nrd(𝛼1 ) −1 = 𝑥 −1
converges. Therefore
𝛾𝑛−1 𝑔𝑛 = 𝛾𝑛−1 𝛼𝑛 𝑥 𝑛−1 = 𝛼1 𝑥 𝑛−1 → 𝛼1 𝑥 −1
converges as well. Therefore the quotient O1 \𝐵R1 is sequentially compact, and therefore
compact.
Remark 38.4.5. Main Theorem 38.4.3 was proven by Hey [Hey29, Hilfssatz 4] in her
1929 Ph.D. thesis (see also Remark 29.10.24) in the case where 𝐵 is a division algebra
over Q. In particular, there is no need to suppose that 𝐵 is central, so it contains the
Dirichlet unit theorem as a consequence: see Exercise 38.6.
After treating the decomposition of the adelic coset space, we will also see Hey’s
theorem as an essentially direct consequence of Fujisaki’s lemma (Main Theorem
27.6.14): see Theorem 38.7.21.
38.5. ∗ ARITHMETIC GROUPS, MORE GENERALLY 709
Proof. Let 𝑥 0 ∈ 𝑋 (Γ) satisfy StabΓ (𝑥0 ) = {1}. Then by Theorem 37.4.18, the
Dirichlet domain ◊ = ◊(Γ; 𝑥0 ) is a locally finite fundamental set for Γ. By Main
Theorem 38.4.3, 𝑋 (Γ) is compact, so too is ◊. Then since ◊ is locally finite, we
conclude that 𝛾◊ ∩ ◊ ≠ ∅ for only finitely many 𝛾 ∈ Γ. But then by Theorem 37.4.2,
the group Γ is generated by such elements and Γ is finitely generated.
◊ ∩ 𝛾◊ ∩ 𝛾 0◊
with 𝛾 ≠ 𝛾 0 provide a finite generating set of relations (generalizing the vertex cycle
relations): see e.g. Raghunathan [Rag72, Theorem 6.15].
Proof. Applying 38.5.2, we have 𝜌 : 𝐵 ↩→ M4 (𝐹) realizing 𝐵1 ' G (𝐹) ≤ GL4 (𝐹) as
a linear algebraic group by appropriate polynomial equations. Under
Î this embedding,
O := 𝜌(𝐵) ∩ M4 (Z𝐹 ) is a Z𝐹 -order, and thus O1 ' G (Z𝐹 ) and 𝑣 G (𝐹𝑣 ) ' 𝑣 𝐵1𝑣 ,
Î
as required in the definition.
710 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
In view of Definition 38.3.4, we consider now the converse: when does the more
general definition give rise to discrete subgroups of two-by-two matrices?
38.5.4. Let G be a linear algebraic group over 𝐹, and suppose that G𝐹 ' GL2,𝐹 , so
there is a chance to obtain discrete groups of symmetries like the ones considered
above. (For a complete treatment, we should consider SL2 as well as the group PGL2 ,
but the arguments are similar.) We say then that G is an 𝐹-form of the algebraic group
GL2 .
Lemma 38.5.5. Let G be an 𝐹-form of GL2 . Then there exists a quaternion algebra
𝐵 over 𝐹, unique up to 𝐹-algebra isomorphism, such that G (𝐹) ' 𝐵× .
Proof. This is a basic result in non-abelian Galois cohomology, and it would take us
too far afield to prove it here: see e.g. Milne [Milne2017, Theorem 20.3.5] and more
generally Serre [Ser79, Chapter X].
Lemma 38.5.5 explains that more general notions of arithmetic groups do not
create anything new beyond our quaternionic definition.
Remark 38.5.6. In this context, there is a criterion for compactness, generalizing
Main Theorem 38.4.3 (conjectured by Godement): A discrete subgroup of G (R)
is cocompact if and only if the reductive part of the connected component of G is
anisotropic over 𝐹. If G is semisimple, then cocompactness is equivalent to asking
that every element of G (𝐹) is semisimple. This criterion was proven by Borel–
Harish-Chandra [BHC62] and Mostow–Tamagawa [MT62]; Godement [God62] (with
Weil) extended the method of Mostow–Tamagawa and simplified the proof by working
directly on adele groups. See also Platonov–Rapinchuk [PR94, §4.5].
𝐵 = M2 (Q) = M2 ( Q)
b × M2 (R) = 𝐵
b × 𝐵∞ .
The order O = M2 (Z) is maximal in 𝐵, and we have the adelic order O b = M2 ( bZ) ⊂
M2 ( Q). (We have seen that 𝐵 = M2 (Q) ≤ 𝐵 = M2 (Q) sits discretely and the quotient
b
𝐵\𝐵 is compact; like the adelic quotient 𝐹\𝐹 itself, this is not very interesting.)
Similarly, we have
𝐵× = GL2 (Q) = 𝐵 ×
b × 𝐵∞ = GL2 ( Q)
b × GL2 (R). (38.6.2)
38.6. ∗ MODULAR CURVES, SEEN IDELICALLY 711
𝐵× \ 𝐵
b×/O
b× = GL2 (Q)\GL2 ( Q)/GL
b 2 (b
Z) = {1} (38.6.3)
by (38.6.3). Therefore, every element of GL2 (Q) is represented in the double coset by
an element of the form (1, 𝛼∞ ) with 𝛼∞ ∈ GL2 (R); and the element 𝛼∞ is well-defined
up the action of the group of pairs (𝛾, 𝜇
b) ∈ GL2 (Q) × GL2 (b
Z) satisfying
0
𝛾(1, 𝛼∞ ) 𝜇
b = (𝛾 𝜇
b, 𝛾𝛼∞ ) = (1, 𝛼∞ )
38.6.6. At this point, we are no stranger to the quotient GL2 (Z)\ GL2 (R)! We have
studied in detail the related quotient SL2 (Z)\ SL2 (R): by the symmetric space de-
scription, we have an isometry
∼ H2
SL2 (R)/SO(2) −→
(38.6.7)
𝑔 SO(2) ↦→ 𝑔𝑖.
(any matrix of negative determinant interchanges the upper and lower half-planes, so
we maintain a bijection). Then we can take the quotient on the left by GL2 (Z) to get
an identification
from (38.6.7), and we find again the classical modular curve we considered in section
35.1. (In section 40.1, we will see another version of this in that the space 𝑌 (1)
parametrizes complex lattices Λ ⊆ C up to scaling by C× .)
712 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
= 𝐵× \( 𝐵
b× /O
b× × 𝐵× /𝐾∞ ) = 𝐵× \𝐵× /𝐾;
∞
𝜄 : 𝐵 → M2 (R) 𝑡 × M2 (C) 𝑐 .
We have defined
H = (H2 ) 𝑡 × (H3 ) 𝑐
where we may also write
H ± = (H2± ) 𝑡 × (H3 ) 𝑐
where similarly
𝑌 Sh (Γ) = 𝐵× \(( 𝐵
b× /b
Γ) × H ± ) (38.7.8)
where 𝐵× acts on H ± via 𝜄 and on 𝐵 b× /b
Γ by left multiplication. Since there can
hopefully no confusion about this action, removing parentheses we will write
𝑌 Sh (Γ) = 𝐵× \( 𝐵
b× × H ± )/b
Γ.
714 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
Therefore Cl+𝐺 (Γ) 𝑅 is a (narrow) class group of 𝐹 associated to the group Γ; as such,
it is a finite abelian group that surjects onto the strict class group Cl+ 𝑅.
38.7.15. By 38.7.12, the space 𝑌 Sh (Γ) is the disjoint union of finitely many connected
components indexed by the group Cl+𝐺 (Γ) 𝑅. We identify these connected components
explicitly as follows.
Let the ideals 𝔟 ⊆ 𝑅 form a set of representatives for Cl+𝐺 (Γ) 𝑅, and let b
𝔟 = 𝔟 ⊗Z b
Z
be their adelification; then for each 𝔟, there exists 𝑏 ∈ 𝑅 generating 𝔟, so
b b b
𝑏𝑅
b b ∩ 𝑅 = 𝔟.
For simplicity, choose 𝔟 = 𝑅 and b𝑏 =b1 for the representatives of the trivial class.
By surjectivity of the map (38.7.14), for each b b× such that
𝑏 there exists 𝛽b ∈ 𝐵
nrd( 𝛽) = 𝑏. Therefore
b b
Ä
𝑌 Sh (Γ) = 𝐵×>0 ( 𝛽bb
Γ × H ). (38.7.16)
𝔟
For each 𝔟, let
Γ 𝛽b−1 ∩ 𝐵×>0
Γ𝔟 := 𝛽bb (38.7.17)
Then we have a natural bijection
𝐵×>0 ( 𝛽bb
Γ × H ) ↔ Γ𝔟 \H
(38.7.18)
( 𝛽bb
Γ, 𝑧) ↦→ 𝑧.
Letting
𝑌 (Γ𝔟 ) := Γ𝔟 \H
we see that each 𝑌 (Γ𝔟 ) is a connected orbifold of dimension 2𝑡 + 3𝑐. We abbreviate
𝑌 (Γ) = 𝑌 (Γ (1) ) for the trivial class. Putting these together, we have
Ä
𝑌 Sh (Γ) = 𝑌 (Γ𝔟 ) (38.7.19)
𝔟
Remark 38.7.20. When 𝑐 = 0, i.e., 𝐹 is totally real, then 𝑌 Sh (Γ) can be canonically
given the structure of an algebraic variety defined over a number field, by work of
Shimura [Shi67] and Deligne [Del71]; in this case we upgrade 𝑌 Sh (Γ) to a quater-
nionic Shimura variety. The theory of Shimura varieties is both broad and deep—see
Milne [Milne-SV] and the references therein. We will begin the study quaternionic
Shimura varieties of dimension 1 in Chapter 43, touching upon the theory of canonical
models in section 43.8.
The above description has hopefully provided a more transparent way to understand
arithmetic orbifolds. For example, we can prove an idelic version of Hey’s theorem
(Main Theorem 38.4.3) as follows.
Proof. We appeal to Fujisaki’s lemma (Main Theorem 27.6.14): the quotient 𝐵× \𝐵 (1)
is cocompact, under the important hypothesis that 𝐵 is a division algebra. Let 𝐾∞ be
as in 38.7.6, and consider the inclusion followed by the projection
Exercises
In these exercises, we maintain the notation in this chapter: let 𝐹 be a number field
with 𝑟 real places and 𝑐 complex places, degree 𝑛 = [𝐹 : Q], and ring of integers 𝑅,
and let O be an 𝑅-order in a quaternion algebra 𝐵 over 𝐹.
√
1. Let 𝛼 ∈ R r Q. Show that Z[𝛼] is not discrete in R. (Taking e.g. 𝛼 = 𝑑, this
gives a reason to worry about discreteness of number fields when we project.)
2. Embed O1 diagonally in SL2 (R) 𝑟 −𝑡 × SL2 (C) 𝑐 . Show that a (further) projection
to a proper factor is not discrete.
𝑎, 𝑏
3. Let 𝐵 = , and let 𝑣 be a split real place of 𝐵. Show that O1 ↩→ 𝐵1𝑣 '
𝐹
SL2 (R) if and only if 𝐹 is totally real and for all nonidentity real places 𝑣 0, we
have 𝑣 0 (𝑎) < 0 and 𝑣(𝑏) < 0.
4. In this exercise, we give a direct argument for the discreteness of an arithmetic
Fuchsian group. Suppose 𝐹 is totally real, let 𝑣 be a split place of 𝐵, consider
𝐹 ↩→ 𝑣(𝐹) ⊆ R as a subfield of R, and suppose that 𝐵 is ramified at all other
(nonidentity) real places. Prove that O1 ⊆ SL2 (R) is discrete.
5. Consider the regular representation 𝜌 : 𝐵 ↩→ M4 (𝐹) (Exercise 2.11). Describe
the image explicitly in terms of polynomial equations in matrix entries. Conclude
that 𝐵× and 𝐵1 are also described by polynomial equations.
716 CHAPTER 38. QUATERNIONIC ARITHMETIC GROUPS
that log 𝑅 1 \ log 𝐹R1 is compact, and therefore log 𝑅 1 has rank 𝑟 + 𝑐 − 1 as
an abelian group (written additively).
(d) Conclude that 𝑅 × has rank 𝑟 + 𝑐 − 1 as an abelian group (written multi-
plicatively).
Chapter 39
Volume formula
39.1 ⊲ Statement
We saw in (35.1.5) that the hyperbolic area of the quotient SL2 (Z)\H2 can be computed
directly from the fundamental domain
as ∫
d𝑥 d𝑦 𝜋
area(SL2 (Z)\H2 ) = area(◊) = = . (39.1.1)
◊ 𝑦2 3
In fact, given a Fuchsian or Kleinian group Γ, the hyperbolic area or volume of the
quotient Γ\H (where H = H2 , H3 , respectively) can be computed without recourse
to a fundamental domain: it is given in terms of the arithmetic invariants of the order
and quaternion algebra that give rise to Γ.
To begin, we consider the already interesting case where the quaternion algebra is
defined over Q.
717
718 CHAPTER 39. VOLUME FORMULA
Example 39.1.5. Recall the case 𝑋01 (6) from section 38.1. We confirm that the
hyperbolic area computed from the fundamental domain agrees with the formula in
Theorem 39.1.2:
𝜋 2𝜋
area(𝑋01 (6, 1)) = 𝜑(6)𝜓(1) = .
3 3
The expression (39.1.3) is quite similar to the Eichler mass formula (Theorem
25.3.15). Indeed, the method of proof is the same, involving the zeta function of the
order O. For the case of a definite quaternion algebra, when the unit group was finite,
we used Theorem 26.2.12 to relate the mass to a residue of the zeta function (see
in particular Proposition 26.5.10); for an indefinite quaternion algebra, to carry this
out in general would involve a multivariable integral whose evaluation is similar to
the proof of Proposition 26.2.18 (the commutative case)—not an appealing prospect.
That being said, this type of direct argument with the zeta function was carried out by
Shimizu [Shz65, Appendix] over a totally real field 𝐹.
We prefer instead to use idelic methods; we already computed the normalized
Tamagawa measure 𝜏 1 (𝐵1 \𝐵1 ) = 1 (see Theorem 29.11.3), and this number has
everything we need! It is a much simpler computation to relate this to the volume of
the hyperbolic quotient: we carry this out in section 39.3.
The following notation will be used throughout.
39.1.6. Let 𝐹 be number field of degree 𝑛 = [𝐹 : Q] with 𝑟 real places and 𝑐 complex
places, so 𝑟 + 2𝑐 = 𝑛. Let 𝐵 be a quaternion algebra over 𝐹 of discriminant 𝔇 that is
split at 𝑡 real places. Suppose that 𝐵 is indefinite (i.e., 𝑡 + 𝑐 > 0).
Let 𝑅 = Z𝐹 be the ring of integers of 𝐹. Let O ⊆ 𝐵 be an 𝑅-order of reduced
discriminant 𝔑 that is locally norm-maximal. Let
H := (H2 ) 𝑡 × (H3 ) 𝑐
and let
Γ1 (O) ≤ PSL2 (R) 𝑡 × PSL2 (C) 𝑐 H
be the discrete group associated to the group PO1 = O1 /{±1} of units of O of reduced
norm 1.
O
For a prime 𝔭 | 𝔑 with Nm(𝔭) = 𝑞, let ∈ {−1, 0, 1} be the Eichler symbol
𝔭
(Definition 24.3.2), and let
1 + 1/𝑞, if (O | 𝑝) = 1;
1 − Nm(𝔭) −2
𝜆(O, 𝔭) := = 1 − 1/𝑞, if (O | 𝑝) = −1; (39.1.7)
O
Nm(𝔭) −1
1− 1 − 1/𝑞 2 , if (O | 𝑝) = 0.
𝔭
2(4𝜋) 𝑡 Ö
vol(Γ1 (O)\H ) = 2 2
𝜁 𝐹 (2)𝑑 𝐹3/2 Nm(𝔑) 𝜆(O, 𝔭). (39.1.9)
(4𝜋 ) (8𝜋 )
𝑟 𝑐
𝔭 |𝔑
39.1. ⊲ STATEMENT 719
Main Theorem 39.1.8 is proven as Main Theorem 39.3.1. Since 𝜁 𝐹 (2) ≈ 1, for
algebras 𝐵 with fixed signature 𝑟, 𝑠, 𝑡, we can roughly estimate
39.1.11. The constant factor in (39.1.9) is written to help in remembering the formula;
multiplying out, we have
2(4𝜋) 𝑡 1
2 2
= 2𝑟 +3𝑐−2𝑡−1 2𝑟 +2𝑐−𝑡 .
(4𝜋 ) (8𝜋 )
𝑟 𝑐 2 𝜋
Note that the right-hand side of (39.1.9) is independent of the choice of order O in
the genus of O, as it depends only on O.
b
Remark 39.1.12. If O is not locally norm-maximal, then one corrects the formula
by inserting factors [𝑅𝔭× : nrd(O×𝔭 )] as in (39.2.10); since 𝑅𝔭× ⊆ O×𝔭 we have 𝑅𝔭×2 ⊆
nrd(O×𝔭 ), so these factors are at most 2 for each 𝔭.
An important special case of Main Theorem 39.1.8 is the case where O is an
Eichler order, generalizing Theorem 39.1.2.
Theorem 39.1.13. Suppose that O = O0 (𝔐) is an Eichler order of level 𝔐 and
disc 𝐵 = 𝔇, so 𝔑 = 𝔇𝔐. Write Γ10 (𝔐) = Γ1 (O) and Γ1 (1) for the group associated
to a maximal order O(1) ⊇ O0 (𝔐).
Then
2(4𝜋) 𝑡
vol(Γ10 (𝔐)\H ) = 𝜁 𝐹 (2)𝑑 𝐹3/2 𝜑(𝔇)𝜓(𝔐)
(4𝜋 2 ) 𝑟 (8𝜋 2 ) 𝑐
where
Ö 1
×
𝜑(𝔇) = #(Z𝐹 /𝔇) = Nm 𝔇 1−
Nm 𝔭
𝔭 |𝔇
Ö (39.1.14)
1 1 1
𝜓(𝔐) = [Γ (1) : Γ0 (𝔐)] = Nm 𝔐 1+ .
Nm 𝔭
𝔭 k𝔐
𝑒
We will convert the volume (39.2.1) into the desired form by separating out the
contribution from the finite places and the infinite (real) ramified places: what remains
is the volume of the orbifold we seek, which we then renormalize from the adelic to
the standard hyperbolic volume.
We define
O := O b × 𝐵∞ ⊆ 𝐵.
Then
O1 = {𝛾 ∈ O× : nrd(𝛾) = 1} = O× ∩ 𝐵1
and
O1 = O
b1 × 𝐵1 .
∞
We have an embedding O1 ↩→ 𝐵∞
1 , so
1 = 𝜏 1 (O1 \O1 ) = b
𝜏1 (O
b1 )𝜏 1 (O1 \𝐵1 ).
∞ ∞ (39.2.4)
so that
b b1 ) −1 = | 𝑑 𝐹 | 3/2 𝜁 𝐹 (2)𝜑(𝔇).
𝜏1 (O (39.2.6)
So by (39.2.6), we conclude that
1
𝜏∞ (O1 \𝐵∞
1
) = 𝑑 𝐹3/2 𝜁 𝐹 (2)𝜑(𝔇). (39.2.7)
39.2. VOLUME SETUP 721
𝜏1 (O
b c01 ) = [O
c01 : O
b1 ]b
𝜏1 (O
b1 )
so similarly
1
𝜏∞ (O1 \𝐵∞
1
) = 𝑑 𝐹3/2 𝜁 𝐹 (2)𝜑(𝔇) [ O
c01 : O
b1 ]. (39.2.9)
If O is locally norm-maximal, then further
c01 : O
[O c0× : O
b1 ] = [O b× ] (39.2.10)
(as in (26.6.8)).
Next, we relate the measures on SL2 (R) and SL2 (C) to the corresponding measures
on H2 and H3 .
39.2.12. Recall the symmetric space identification
The hyperbolic plane H2 is equipped with the hyperbolic measure 𝜇, the unique
measure invariant under the action of PSL2 (R); the group SL2 (R) has the Haar measure
𝜏 1 , also invariant under the left action of SL2 (R). Therefore, the identification (39.2.13)
relates these two measures up to a constant 𝑣(SO(2)) that gives a total measure to SO(2)
(normalizing its Haar measure). Ditto for H3 and SU(2), with a constant 𝑣(SU(2)).
Lemma 39.2.14. We have
Proof. One could compute the relevant constant by doing a (compatible) integral, but
we prefer just to refer to an example where both sides are computed. We consider a
fundamental domain for the action of SL2 (Z) SL2 (R) that is invariant under SO(2):
for example, we can lift a fundamental domain for PSL2 (Z) H2 under (39.2.13).
The difference between SL2 (Z) and PSL2 (Z) = SL2 (Z)/{±1} is annoyingly relevant
here! We have
PSL2 (Z)\H2 = PSL2 (Z)\ SL2 (R)/SO(2);
if we let SO(2)2 ' SO(2)/{±1} be the rotation group acting by 2𝜃 instead of 𝜃, then
we can lift from PSL2 (Z) to SL2 (Z) and
On the bottom we have 𝜇(SL2 (Z)\H2 ) = 𝜋/3 by (35.1.5) and on the top we have
𝜏 1 (SL2 (Z)\ SL2 (R)) = 𝜁 (2) = 𝜋 2 /6 by (39.2.9). Plugging in, we compute
𝜋 2 /6
𝑣(SO(2)) = 2 = 𝜋. (39.2.16)
𝜋/3
Similarly,
𝜏 1 (SL2 (Z[𝑖])\ SL2 (C)) 43/2 𝜁Q(𝑖) (2)
𝑣(SU(2)) = 2 =2 = 8𝜋 2 (39.2.17)
𝜇(PSL2 (Z[𝑖])\H3 ) (2/𝜋 2 )𝜁Q(𝑖) (2)
by Example 39.1.16 and again (39.2.9).
Proof. To summarize the previous section, we started with 𝜏 1 (𝐵1 \𝐵1 ) and concluded
𝜏 1 (O1 \O1 ) = 1 by strong approximation; we factored this into finite and infinite parts,
with the finite part computed in terms of the order and the infinite part. Then
Ö Ö Ö
1 1 1 1 1 1
O \𝐵∞ = O \ 𝐵𝑣 ' 𝐵𝑣 × O \ 𝐵1𝑣 . (39.3.3)
𝑣 |∞ 𝑣 ∈Ω 𝑣 |∞
𝑣∉Ω
Each term contributes to the volume. For the first product, for each of the 𝑟 − 𝑡 places
𝑣 ∈ Ω we have 𝐵1𝑣 ' H1 and we computed in Lemma 29.5.9 that 𝜏 1 (H1 ) = 4𝜋 2 . For
the remaining terms, we employ the comparison formula between measures (Lemma
39.2.14), and are plagued by the same factor 2 coming from the fact that Γ1 (O) arises
from PO1 /{±1}. Putting these together, the decomposition (39.3.3) yields a volume
1
1
𝜏∞ (O1 \𝐵∞
1
) = (4𝜋 2 ) 𝑟 −𝑡 𝜋 𝑡 (8𝜋 2 ) 𝑐 vol(Γ1 \H )
2
(39.3.4)
(4𝜋 2 ) 𝑟 (8𝜋 2 ) 𝑐
= vol(Γ1 \H ).
2(4𝜋) 𝑡
From (39.2.9) and (39.3.4) we conclude
2(4𝜋) 𝑡
vol(Γ1 \H ) = 𝜇(O1 \𝐵∞ 1
)
(4𝜋 2 ) 𝑟 (8𝜋 2 ) 𝑐
(39.3.5)
2(4𝜋) 𝑡
= 𝑑 3/2 𝜁 𝐹 (2)𝜑(𝔇) [ O
c01 : O
b1 ].
(4𝜋 2 ) 𝑟 (8𝜋 2 ) 𝑐 𝐹
39.4. GENUS FORMULA 723
Finally, the computation (39.2.11) of the local index completes the proof.
Remark 39.3.6. A similar proof works for the case where 𝐹 is a function field or
where Γ is an S-arithmetic group, but in both cases still under the hypothesis that 𝐵 is
S-indefinite for an eligible set S (playing the role of the archimedean places above).
Example 39.3.7. Suppose 𝐹 is totally real, and we take 𝐵 = M2 (𝐹) and O = M2 (Z𝐹 ).
Then H = (H2 ) 𝑛 and
Rewriting this slightly, for 𝑞 ∈ Z ≥2 , let 𝑚 𝑞 be the number of elliptic cycles of order 𝑞.
Then
𝜇(𝑌 (Γ)) ∑︁ 1
= 2𝑔 − 2 + 𝑚𝑞 1 − +𝛿 (39.4.2)
2𝜋 𝑞 ≥2
𝑞
× = 2𝑞}
{O1 -conjugacy classes of optimal embeddings 𝜙 : 𝑆 ↩→ O with 𝑆tors
↓
{Elliptic cycles of Γ of order 2𝑞}.
In the notation of 30.3.10, we have shown that
1 ∑︁
𝑚𝑞 = 𝑚(𝑆, O; O1 ). (39.4.4)
2
𝐾𝑞 ⊃𝑆 ⊇𝑅 [𝜁2𝑞 ]
× =2𝑞
#𝑆tors
Our next major ingredient is the theory of selectivity, treated in chapter 31.
39.4.5. We claim that 𝐾𝑞 does not satisfy the selectivity condition (OS), defined in
31.1.6. If O is an Eichler order, we may appeal to Proposition (31.2.1) and condition
(a): since 𝐹 is totally real, 𝐾𝑞 is totally complex, and 𝐵 is split at a real place, condition
(a) fails.
Therefore by Main Theorem 31.1.7(a), Gen O is not optimally selective. By
Corollary 31.1.10, for every 𝑅-order 𝑆 ⊆ 𝐾𝑞 , we have
ℎ(𝑆)
𝑚(𝑆, O; O× ) = b O;
𝑚( 𝑆, b× )
b O (39.4.6)
# ClΩ 𝑅
where we have substituted # Cls O = # ClΩ 𝑅 (by Corollary 28.5.17).
The adelic embedding numbers 𝑚( 𝑆, b O;
b Ob× ), a product of (finitely many) local
embedding numbers by 30.7.1, are computed in section 30.6.
We will need one lemma relating units to class numbers.
Lemma 39.4.7. We have [𝑅>× Ω 0 : 𝑅 ×2 ] = 2[ClΩ 𝑅 : Cl 𝑅].
Proof. The index [𝑅>× Ω 0 : 𝑅 ×2 ], which does not depend on 𝑆, is related to class
numbers as follows. For each real place 𝑣, define sgn 𝑣 : 𝐹 × → {±1} by the real sign
sgn 𝑣 (𝑎) = sgn(𝑣(𝑎)) at 𝑣. Let
sgnΩ : 𝐹 × → {±1}Ω
𝑎 ↦→ (sgn 𝑣 (𝑎)) 𝑣
collect the signs at the places 𝑣 ∈ Ω. Then we have an exact sequence
1 → {±1}Ω /sgnΩ 𝑅 × → ClΩ 𝑅 → Cl 𝑅 → 1 (39.4.8)
where the map on the left is induced by mapping a tuple of signs in {±1}Ω to the
principal ideal generated by any 𝑎 ∈ 𝐹 × with the given signs. We have a second
(tautological) exact sequence
sgnΩ
1 → 𝑅>× Ω 0 /𝑅 ×2 → 𝑅 × /𝑅 ×2 −−−→ {±1}Ω → {±1}Ω /sgnΩ 𝑅 × → 1 (39.4.9)
of elementary abelian 2-groups (or F2 -vector spaces). Combining (39.4.9) with
(39.4.8), and noting that [𝑅 × : 𝑅 ×2 ] = 2𝑟 by Dirichlet’s unit theorem and #Ω = 𝑟 − 1
by hypothesis, we conclude that
[𝑅>× Ω 0 : 𝑅 ×2 ] = 2𝑟 −#Ω [ClΩ 𝑅 : Cl 𝑅] = 2[ClΩ 𝑅 : Cl 𝑅].
39.4. GENUS FORMULA 725
Definition 39.4.10. For a quadratic 𝑅-order 𝑆 ⊆ 𝐾, the Hasse unit index is defined
by
𝑄(𝑆) := [Nm𝐾 |𝐹 (𝑆 × ) : 𝑅 ×2 ].
1 ∑︁ ℎ(𝑆) b O; b× )
b O
𝑚𝑞 = 𝑚( 𝑆,
ℎ(𝑅) 𝑆 ⊂𝐾𝑞
𝑄(𝑆)
× =2𝑞
#𝑆tors
1 ∑︁
𝑚𝑞 = 𝑚(𝑆, O; O1 ). (39.4.13)
2 𝑆 ⊂𝐾𝑞
× =2𝑞
#𝑆tors
Finally, plugging (39.4.16) into (39.4.13) and cancelling a factor 2 gives the result.
Corollary 39.4.17. The signature of a Shimura curve depends only on the discriminant
𝔇 and level 𝔐.
726 CHAPTER 39. VOLUME FORMULA
Proof. The ambiguity corresponds to a choice of Eichler order of level 𝔐 and choice
of split real place; when 𝐹 = Q, there is no ambiguity in either case. So we may
suppose that Γ has no parabolic cycles. Then we simply observe that the formula
(Proposition 39.4.12) for the number of elliptic cycles depends only on O.
b
Theorem 39.4.19. Let 𝑌 1 (O) = Γ1 (O)\H2 . Then 𝑌 1 (O) is an orbifold with genus 𝑔
where
4 3/2
∑︁ 1
2𝑔 − 2 = 𝜁 𝐹 (2)𝑑 𝐹 𝜑(𝔇)𝜓(𝔐) − 𝑚𝑞 1 − −𝛿
(4𝜋 2 ) 𝑟 𝑞 ≥2
𝑞
Proof. Combine the volume formula (Main Theorem 39.3.1) with (39.4.2).
Theorem 39.4.20. Let 𝐷 = disc 𝐵 > 1 and let O ⊆ 𝐵 be an Eichler order of level 𝑀,
so 𝑁 = 𝐷 𝑀 = discrd O with 𝐷 squarefree and gcd(𝐷, 𝑀) = 1.
Then 𝑋 1 (O) = Γ1 (O)\H2 is an orbifold with genus 𝑔 where
𝜑(𝐷)𝜓(𝑀) 𝑚 2 2𝑚 3
2𝑔 − 2 = − −
6 2 3
where the embedding numbers were computed in Example 30.7.7:
Ö Ö
−4
−4
1− 1+ , if 4 - 𝑀;
×
𝑚 2 = 𝑚(Z[𝑖], O; O ) = 𝑝 |𝐷 𝑝
𝑝 |𝑀
𝑝
0, if 4 | 𝑀.
Ö
Ö −3 −3
1− 1+ , if 9 - 𝑀;
×
𝑚 3 = 𝑚(Z[𝜔], O; O ) = 𝑝 |𝐷 𝑝
𝑝 |𝑀
𝑝
0, if 9 | 𝑀;
Example 39.4.21. Suppose 𝐷 = 6 and 𝑀 = 1, so we are in the setting of Then
𝑚 2 = 𝑚 3 = 2 and
𝜙(6) 4 1
2𝑔 − 2 = − 1 − = = −2
6 3 3
so 𝑔 = 0; this confirms that 𝑋 1 (O) has signature (0; 2, 2, 3, 3) as in 37.9.10.
39.4. GENUS FORMULA 727
Exercises
1. Let 𝐹 be the function field of a curve 𝑋 over F𝑞 of genus 𝑔. Let 𝐵 be a quaternion
algebra over 𝐹.
(a) Let 𝑣 ∈ Pl 𝐹 be place that is split in 𝐵. Let S = {𝑣}, let 𝑅 = 𝑅 (S) , and
let O ⊆ 𝐵 be an 𝑅-order. Let T be the Bruhat–Tits tree associated to
𝐵 𝑣 ' M2 (𝐹𝑣 ). Via the embedding 𝜄 : 𝐵 ↩→ 𝐵 𝑣 , show that the group
Γ1 (O) = P𝜄(O1 ) T acts on T by left multiplication as a discrete group
acting properly.
(b) Continuing as in (a), compute the measure of Γ1 (O)\T using the methods
of section 39.3.
2. Generalizing the previous exercise, let 𝐹 be a global field, let 𝐵 be a quaternion
algebra over 𝐹, let S be an eligible set and suppose that 𝐵 is S-indefinite. Let
𝑅 = 𝑅 (S) and let O ⊆ 𝐵 be an 𝑅-order. Define a symmetric space H on
which PO1 acts as a discrete group acting properly, and compute the measure of
Γ1 (O)\H .
Part V
Arithmetic geometry
729
Chapter 40
In this chapter, we introduce modular forms on the classical modular group. This
chapter will provide motivation as well as important examples for generalizations in
this last part of the text.
𝑌 = Γ\H2 → {Λ ⊂ C lattice}/∼
(40.1.1)
Γ𝜏 ↦→ [Z + Z𝜏].
In particular, the set of homothety classes has a natural structure of a Riemann surface,
and we seek now to make this explicit. We show that there are natural, holomorphic
functions on the set of lattices that allow us to go beyond the bijection 40.1.1 to realize
the complex structure on 𝑌 explicitly.
Let Λ ⊂ C be a lattice. To write down complex moduli, we average over Λ in a
convergent way, as follows.
Definition 40.1.2. The Eisenstein series of weight 𝑘 ∈ Z>2 for Λ is
∑︁ 1
𝐺 𝑘 (Λ) = .
𝜆∈Λ
𝜆𝑘
𝜆≠0
731
732 CHAPTER 40. CLASSICAL MODULAR FORMS
The number of pairs (𝑚, 𝑛) with 𝑟 ≤ |𝑚𝜏 + 𝑛| < 𝑟 + 1 is the number of lattice points in
an annulus of area 𝜋(𝑟 + 1) 2 − 𝜋𝑟 2 = 𝑂 (𝑟), so there are 𝑂 (𝑟) such points; and thus the
series (40.1.4) is majorized by (a constant multiple of) 𝑟∞=1 𝑟 1−𝑘 , which is convergent
Í
for 𝑘 > 2.
thus |𝐺 𝑘 (𝑧)| ≤ |𝐺 𝑘 (𝜔)| and so by the Weierstrass 𝑀-test, 𝐺 𝑘 (𝑧) is holomorphic for
𝑧 ∈ ◊: by Morera’s
theorem, uniform convergence implies holomorphicity. But now
𝑎 𝑏
for all 𝛾 = ∈ Γ, we claim that
𝑐 𝑑
(and note this does not depend on the choice of sign): indeed,
1 𝑐𝑧 + 𝑑
= (40.1.9)
𝑚 + 𝑛(𝛾𝑧) (𝑏𝑛 + 𝑑𝑚) + (𝑎𝑛 + 𝑐𝑚)𝑧
40.1.11. In this (somewhat long) paragraph, we connect the theory of Eisenstein series
above to the theory of elliptic curves.
Let Λ ⊂ C be a lattice. We define the Weierstrass ℘-function (relative to Λ) by
1 ∑︁ 1 1
℘(𝑧) = ℘(𝑧; Λ) = 2 + − . (40.1.12)
𝑧 𝜆∈Λ
(𝑧 − 𝜆) 2 𝜆2
𝜆≠0
We have
1 1 |𝑧|(2|𝜆| + |𝑧|) 1
(𝑧 − 𝜆) 2 − 𝜆2 ≤ |𝜆| 2 (|𝜆| − |𝑧|) 2 = 𝑂 |𝜆| 3 (40.1.13)
so as above we see that ℘(𝑧) is absolutely convergent for all 𝑧 ∈ C r Λ and uniformly
convergent on compact subsets, and so defines a holomorphic function on CrΛ. Since
∞
1 1 ∑︁ 𝑧𝑛
− = (𝑛 + 1) (40.1.14)
(𝑧 − 𝜆) 2 𝜆2 𝑛=1 𝜆 𝑛+2
by differentiating the geometric series, we find
∞
1 ∑︁ 1
℘(𝑧) = 2
+ (𝑘 − 1)𝐺 𝑘 (Λ)𝑧 𝑘 = 2 + 3𝐺 4 (Λ)𝑧 4 + 5𝐺 6 (Λ)𝑧6 + . . . . (40.1.15)
𝑧 𝑘=3
𝑧
and
d℘
𝑥(𝑧) = ℘(𝑧; Λ) and 𝑦(𝑧) = (𝑧; Λ).
d𝑧
Then the image of the map
C/Λ → P2 (C)
(40.1.17)
𝑧 ↦→ (𝑥(𝑧) : 𝑦(𝑧) : 1)
𝑦 2 = 4𝑥 3 − 𝑔4 𝑥 − 𝑔6 ;
𝑗 : H2 → C (40.1.18)
∼ C (Theorem
obtained in this way that defines a bijective holomorphic map Γ\H2 −→
40.3.8).
40.1.19. Eisenstein series can also be thought of as weighted averages over the (cosets
of the) group PSL2 (Z) as follows.
Let Γ∞ ≤ Γ = PSL2 (Z) be the stabilizer of ∞; then Γ∞ is the infinite cyclic
group
1 1 𝑎 𝑏
generated by 𝑇 = . We consider the cosets Γ∞ \Γ: for 𝑡 ∈ Z and 𝛾 = we
0 1 𝑐 𝑑
𝑎 + 𝑡𝑐 𝑏 + 𝑡𝑑
have 𝑇 𝑡 𝛾 = with the same bottom row. Thus the function (𝑐𝑧 + 𝑑) 2
𝑐 𝑑
is well-defined on the coset Γ∞ 𝛾. Thus we can form the sum
∑︁ 1 ∑︁ 1
𝐸 𝑘 (𝑧) = (𝑐𝑧 + 𝑑) −𝑘 = , (40.1.20)
2 𝑐,𝑑 ∈Z
(𝑐𝑧 + 𝑑) 𝑘
Γ∞ 𝛾 ∈Γ∞ \Γ
gcd(𝑐,𝑑)=1
the factor 2 coming from the choice of sign in PSL2 (Z). Since every nonzero (𝑚, 𝑛) ∈
Z2 can be written (𝑚, 𝑛) = 𝑟 (𝑐, 𝑑) with 𝑟 = gcd(𝑚, 𝑛) > 0 and gcd(𝑐, 𝑑) = 1, we find
that
𝐺 𝑘 (𝑧) = 𝜁 (𝑘)𝐸 𝑘 (𝑧).
40.2. ⊲ EISENSTEIN SERIES AS MODULAR FORMS 735
with 𝑎 𝑛 ∈ C and 𝑎 𝑛 = 0 for all but finitely many 𝑛 < 0, then we say that 𝑓 is
meromorphic at ∞; if further 𝑎 𝑛 = 0 for 𝑛 < 0, we say 𝑓 is holomorphic at ∞.
More generally, let Γ ≤ PSL2 (Z) be a subgroup of finite index. For 𝛾 ∈ PSL2 (Z),
we define
𝑓 [𝛾] 𝑘 (𝑧) := 𝚥 (𝛾; 𝑧) −𝑘 𝑓 (𝛾𝑧). (40.2.14)
Then 𝑓 [𝛾] 𝑘 (𝑧) is weight 𝑘 invariant under the group 𝛾 −1 Γ𝛾. We say that 𝑓 is mero-
morphic at the cusps if for every 𝛾 ∈ PSL2 (Z), the function 𝑓 [𝛾] 𝑘 is meromorphic
at ∞, in the above sense. Since 𝑓 is weight 𝑘 invariant, to check if 𝑓 is meromorphic
at the cusps, it suffices to take representatives of the finite set of cosets Γ\ PSL2 (Z).
(The name cusp comes from the geometric description at ∞ coming from the parabolic
stabilizer group, recalling Definition 33.4.5.)
Finally, we say that 𝑓 is holomorphic at the cusps if 𝑓 [𝛾] 𝑘 (𝑧) is holomorphic at
∞ for all 𝛾 ∈ PSL2 (Z), and vanishes at the cusps if 𝑓 [𝛾] 𝑘 (∞) = 0 for all 𝛾.
Definition 40.2.15. Let 𝑘 ∈ 2Z and let Γ ≤ PSL2 (Z) be a subgroup of finite index. A
meromorphic modular form of weight 𝑘 is a meromorphic map 𝑓 : H2 → C that is
weight 𝑘 invariant under Γ and meromorphic at the cusps. A meromorphic modular
function is a meromorphic modular form of weight 0.
A (holomorphic) modular form of weight 𝑘 is a holomorphic map 𝑓 : H2 → C
that is weight 𝑘 invariant under Γ and holomorphic at the cusps. A cusp form of
weight 𝑘 is a holomorphic modular form of weight 𝑘 that vanihses at the cusps.
Let 𝑀𝑘 (Γ) be the C-vector space of modular forms of weight 𝑘 for Γ, and let
𝑆 𝑘 (Γ) ⊆ 𝑀𝑘 (Γ) be the subspace of cusp forms.
where
∞
∑︁ 1
𝜁 (𝑘) =
𝑛=1
𝑛𝑘
and ∑︁
𝜎𝑘−1 (𝑛) = 𝑑 𝑘−1 .
𝑑 |𝑛
𝑑>0
40.2. ⊲ EISENSTEIN SERIES AS MODULAR FORMS 737
grouping together terms in the second step. The fact that 𝐺 𝑘 is holomorphic at ∞ then
follows by definition.
40.2.22. We accordingly define the normalized Eisenstein series by
1
𝐸 𝑘 (𝑧) = 𝐺 𝑘 (𝑧)
2𝜁 (𝑘)
(see also 40.1.19). We have
∞
2𝑘 ∑︁
𝐸 𝑘 (𝑧) = 1 − 𝜎𝑘−1 (𝑛)𝑞 𝑛 (40.2.23)
𝐵 𝑘 𝑛=1
Remark 40.2.24. The notion of Eisenstein series extends in a natural way to the Bianchi
groups PSL2 (Z𝐹 ) where 𝐹 is an imaginary quadratic field: see Elstrodt–Grunewald–
Mennicke [EGM98, Chapter 3].
The form 𝑓 has only finitely many zeros or poles in 𝑌 , i.e., only finitely many
Γ-orbits of zeros or poles: since 𝑓 is meromorphic at ∞, there exists 𝜖 > 0 such that
𝑓 has no zero or pole with 0 < |𝑞| < 𝜖, so with
log(1/𝜖)
Im 𝑧 > 𝑀 = ;
2𝜋
but the part of ◊ with Im 𝑧 ≤ 𝑀 is compact, and since 𝑓 is meromorphic in H2 , it has
only finitely many zeros or poles in this part as well.
40.3.2. In a similar way, the order of the stabilizer 𝑒 𝑧 := # StabΓ (𝑧) is well defined on
the orbit Γ𝑧, since points in the same orbit have conjugate stabilizers. By 35.1.14,
3, if Γ𝑧 = Γ𝜔;
𝑒 𝑧 = 2, if Γ𝑧 = Γ𝑖; (40.3.3)
1,
otherwise.
Proposition 40.3.4. Let 𝑓 : H2 → C be a meromorphic modular form of weight 𝑘 for
Γ = PSL2 (Z), not identically zero. Then
∑︁ 1 𝑘
ord∞ ( 𝑓 ) + ord𝑧 ( 𝑓 ) = (40.3.5)
2
𝑒𝑧 12
Γ𝑧 ∈Γ\H
The sum (40.3.5) has only finitely many terms, by 40.3.1, and the stabilizers are
given in 40.3.2.
Proof. See Serre [Ser73, §3, Theorem 3]: the proof consists of performing a contour
1 d𝑓
integration on the boundary of ◊. Alternatively, this statement can be seen
2𝜋𝑖 𝑓
as a manifestation of the Riemann–Roch theorem: see Diamond–Shurman [DS2005,
§3.5].
𝐸 4 (𝑧) 3 − 𝐸 6 (𝑧) 2
Δ(𝑧) = = 𝑞 − 24𝑞 2 + 252𝑞 3 − 1472𝑞 4 + . . . (40.3.7)
1728
𝐸 4 (𝑧) 3 1
𝑗 (𝑧) = = + 744 + 196884𝑞 + 21493760𝑞 2 + . . . (40.3.9)
Δ(𝑧) 𝑞
Proof. The function 𝑗 is weight 0 invariant under Γ as the ratio of two forms that are
weight 12 invariant. Since 𝐸 4 is holomorphic in H2 , and Δ is holomorphic and has
no zeros in H2 , the ratio is holomorphic in H2 ; and 𝑗 (𝑧) has a simple pole at 𝑧 = ∞,
corresponding to a simple zero of Δ at 𝑧 = ∞. From 40.3.6, we have 𝑗 (𝑖) = 1728, and
𝑗 (𝑧) − 1728 has a double zero at 𝑧 = 𝑖, and 𝑗 (𝑧) has a triple zero at 𝑧 = 𝜔.
To conclude that 𝑗 is bijective, we show that 𝑗 (𝑧) − 𝑐 has a unique zero Γ𝑧 ∈ 𝑌 .
If 𝑐 ≠ 0, 1728, this follows immediately from Proposition 40.3.4; if 𝑐 = 0, 1728, the
results follow for the same reason from the multiplicity of the zero.
740 CHAPTER 40. CLASSICAL MODULAR FORMS
Remark 40.3.10. The definition of 𝑗 (𝑧) is now standard, but involves some choices.
In some circumstances (including the generalization to abelian surfaces, see 43.5.7),
it is more convenient to remember the values of the Eisenstein series themselves, as
follows. To 𝑧 ∈ H , we associate the pair (𝐸 4 (𝑧), 𝐸 6 (𝑧)) ∈ C2 ; if 𝛾 ∈ Γ and 𝑧 0 = 𝛾𝑧,
then
(𝐸 4 (𝑧 0), 𝐸 6 (𝑧 0)) = (𝛿4 𝐸 4 (𝑧), 𝛿6 𝐸 6 (𝑧))
where 𝛿 = 𝚥 (𝛾; 𝑧) ∈ C× . We therefore define the weighted projective (4, 6)-space by
where
(𝐸 4 , 𝐸 6 ) ∼ (𝛿4 𝐸 4 , 𝛿6 𝐸 6 )
for 𝛿 ∈ C× . We write equivalence classes (𝐸 4 : 𝐸 6 ) ∈ P(4, 6) (C). The map
under multiplication has the structure of a (graded) C-algebra; we call 𝑀 (Γ) the ring
of modular forms for Γ.
Theorem 40.3.11. We have 𝑀 (Γ) = C[𝐸 4 , 𝐸 6 ], i.e., every modular form for Γ =
PSL2 (Z) can be written as a polynomial in 𝐸 4 , 𝐸 6 .
More generally, one can study modular forms for congruence subgroups (section
35.4) of PSL2 (Z) in an explicit way, as the following example illustrates.
Example 40.3.12. At the end of section 35.4, we examined a fundamental domain for
the group Γ(2), defined by (35.4.8). As with 𝑋 (1), the homeomorphism (35.4.11) can
be given by a holomorphic map
∼ P1 (C)
𝜆 : 𝑋 (2) −→
obtained from Eisenstein series for Γ(2), analogous to 𝑗 (𝑧). The map 𝜆 satisfies
𝜆(𝛾𝑧) = 𝜆(𝑧) for all 𝛾 ∈ Γ(2) and in particular is invariant under 𝑧 ↦→ 𝑧 + 2. One can
compute its Fourier expansion in terms of 𝑞 1/2 = 𝑒 𝜋𝑖𝑧 as:
𝜆(𝑧) = 16𝑞 1/2 − 128𝑞 + 704𝑞 3/2 − 3072𝑞 2 + 11488𝑞 5/2 − 38400𝑞 3 + . . . . (40.3.13)
Since 𝑗 (𝑧) induces a degree 6 = [Γ(2) : Γ(1)] map 𝑋 (2) → 𝑋 (1), we find the
relationship
(𝜆2 − 𝜆 + 1) 3
𝑗 = 256 2 . (40.3.14)
𝜆 (𝜆 − 1) 2
From (40.3.14) (and the first term), the complete series expansion (40.3.13) can be
obtained recursively.
As a uniformizer for a congruence subgroup of PSL2 (Z), the function 𝜆(𝑧) has
a moduli interpretation (cf. 40.1.11): there is a family of elliptic curves over 𝑋 (2)
equipped with extra structure. Specifically, given 𝜆 ∈ P1 (C) \ {0, 1, ∞}, the corre-
sponding elliptic curve with extra structure is given by the Legendre curve
𝐸 𝜆 : 𝑦 2 = 𝑥(𝑥 − 1) (𝑥 − 𝜆),
equipped with the isomorphism (Z/2Z) 2 −→ ∼ 𝐸 [2] determined by sending the standard
generators to the 2-torsion points (0, 0) and (1, 0). The map 𝑗 is the map that forgets
this additional torsion structure on a Legendre curve and remembers only isomorphism
class.
𝑄(𝑥) ≥ 𝑐(𝑥12 + · · · + 𝑥 𝑚
2
).
Thus 𝑟 𝑄 (𝑛) = 𝑂 (𝑛 𝑘 ), and the series Θ𝑞 (𝑧) is majorized by (a constant multiple of)
Í ∞
𝑛=1 𝑛 𝑞 , so converges to a holomorphic function.
𝑘 𝑛
Let [𝑇] be the Gram matrix for the symmetric bilinear form associated to 𝑄;
then [𝑇] ∈ M𝑚 (Z) is an integral symmetric matrix with even diagonal entries. Let
𝑑 = det 𝑄 = det[𝑇] ∈ Z. Then 𝑑𝐴−1 ∈ M𝑚 (Z) is the adjugate matrix: it is again
symmetric.
Definition 40.4.3. The least positive integer 𝑁 ∈ Z>0 such that 𝑁 𝐴−1 is integral with
even diagonal entries is called the level of 𝑄.
We recall the definition of the congruence subgroups 35.4.5.
Theorem 40.4.4. The theta series Θ𝑄 (𝑧) is a modular form of weight 𝑘 for Γ1 (𝑁).
Proof. Unfortunately, in this generality the proof would take us too far afield. Fun-
damentally, the transformation formula for Θ𝑄 follows from Poisson summation and
careful computations: see Eichler [Eic73, §I.3, Proposition 2], Miyake [Miy2006,
Corollary 4.9.5], or Ogg [Ogg69, Chapter VI].
40.4.5. We can be a bit more specific about the transformation group for Θ𝑄 (𝑧) as
(−1) 𝑘 det 𝑄
follows. To 𝑄, we associate the character 𝜒 defined by 𝜒(𝑛) = . Then
𝑛
𝑎 𝑏
Θ𝑄 (𝛾𝑧) = 𝜒(𝑑) (𝑐𝑧 + 𝑑) 𝑘 Θ𝑄 (𝑧) for all 𝛾 = ∈ Γ0 (𝑁);
𝑐 𝑑
Lemma 40.5.2.
A system of representatives of O1 \O𝑛 is given by the set of matrices
𝑎 𝑏
of the form with 𝑎𝑑 = 𝑛, 𝑎 > 0, and 0 ≤ 𝑏 < 𝑑.
0 𝑑
Proof. The lemma follows as in Lemma 26.4.1(b) using the theory of elementary
divisors, but applying row operations (acting on the left).
By the condition of automorphy 𝑓 (𝛾𝑧) = 𝚥 (𝛾; 𝑧) 𝑘 𝑓 (𝑧) and the cocycle relation
(40.2.5), the Hecke operators are well-defined and preserve weight 𝑘 invariance.
∑︁
𝑏𝑚 = 𝑑 𝑘−1 𝑎 𝑚𝑛/𝑑 2 (40.5.8)
𝑑 |gcd(𝑚,𝑛)
𝑑>0
Applying (40.5.9), we see that if 𝑓 ∈ 𝑀𝑘 (Γ) has 𝑓 (∞) = 0 (equivalently, 𝑎 0 = 0), then
the same is true for 𝑇 (𝑛) 𝑓 . Repeating this for the functions 𝑓 [𝛾] 𝑘 with 𝛾 ∈ PSL2 (Z)
(as in (40.2.14)), we conclude that the operators 𝑇 (𝑛) act on the space of cusp forms
𝑆 𝑘 (Γ) ⊂ 𝑀𝑘 (Γ).
744 CHAPTER 40. CLASSICAL MODULAR FORMS
𝑇 ( 𝑝)𝑇 ( 𝑝 𝑟 ) = 𝑇 ( 𝑝 𝑟 +1 ) + 𝑝 𝑘−1𝑇 ( 𝑝 𝑟 −1 ).
Theorem 40.5.12. The Hecke operators 𝑇 (𝑛) for gcd(𝑛, 𝑁) = 1 on 𝑀𝑘 (Γ0 (𝑁))
generate a commutative, semisimple Z-algebra.
Proof. See e.g. Diamond–Shurman [DS2005, Theorem 5.5.4]. Briefly, we treat Eisen-
stein series separately and work with cusp forms 𝑆 𝑘 (Γ0 (𝑁)). To prove that the operators
are semisimple, we would need to show that the Petersson inner product
∫
h 𝑓 , 𝑔i = 𝑓 (𝑧)𝑔(𝑧)𝑦 𝑘 d𝜇(𝑧)
Γ\H2
is well-defined, positive, and nondegenerate, and then verify that the operators are
normal with respect to this inner product.
By Theorem 40.5.12 and linear algebra, there exists a C-basis 𝑓𝑖 (𝑧) of 𝑀𝑘 (Γ0 (𝑁))
consisting of simultaneous eigenfunctions for all 𝑇 (𝑛).
Exercises
I 1. Let 𝑓 : 𝑈 → C be a function that is meromorphic in an open neighborhood
𝑈 ⊇ C with 𝑧 ∈ 𝑈, and let 𝐶 be a contour along an arc of a circle of radius 𝜖 > 0
centered at 𝑧 contained in 𝑈 with total angle 𝜃. Show that
∫
d𝑓
lim = 𝜃𝑖 ord𝑧 ( 𝑓 ).
𝜖 →0 𝐶 𝑓
to show
∞ ∑︁∞
∑︁ 𝑧 𝑘
𝑧 cot 𝑧 = 1 − 2 .
𝑘=2 𝑛=1
𝑛𝜋
𝑘 even
for 𝑘 ∈ 2Z ≥1 .
4. We defined Eisenstein series 𝐺 𝑘 (𝑧) for 𝑘 ≥ 4, and found 𝐺 𝑘 (𝑧) ∈ M 𝑘 (SL2 (Z))
are modular forms of weight 𝑘 for SL2 (Z). The case 𝑘 = 2 is also important,
though we must be a bit more careful in its analysis. Let
∑︁ ∑︁ 1
𝐺 2 (𝑧) := .
𝑐 ∈Z 𝑑 ∈Z
(𝑐𝑧 + 𝑑) 2
(𝑐,𝑑)≠(0,0)
where 𝑞 = 𝑒 2 𝜋𝑖𝑧 .
(b) Show that 𝐺 2 (𝑧 + 1) = 𝐺 2 (𝑧) and
−1 2𝜋𝑖
𝐺2 = 𝐺 2 (𝑧) − .
𝑧 𝑧
where 𝑞 := 𝑒 2 𝜋𝑖𝑧 and 𝑧 ∈ H2 . [The function 𝜗(𝑧) is a theta series for the univari-
ate quadratic form 𝑥 ↦→ 𝑥 2 .] Let 𝑟 4 (𝑛) be the number of ways of representing
𝑛 ≥ 0 as the sum of 4 squares.
(a) Show that
∞
∑︁
Θ𝑄 (𝑧) := 𝜗(𝑧) 4 = 1 + 𝑟 4 (𝑛)𝑞 𝑛 = 1 + 8𝑞 + 12𝑞 2 + . . . .
𝑛=1
(b) Show that Θ𝑄 (𝑞) ∈ M2 (Γ0 (4)) is a modular form of weight 2 on Γ0 (4).
(c) Show that dimC 𝑀2 (Γ0 (4)) = 2 and dimC 𝑆2 (Γ0 (4)) = 0.
(d) Let
𝐺 2,2 (𝑧) = 𝐺 2 (𝑧) − 2𝐺 2 (2𝑧)
𝐺 2,4 (𝑧) = 𝐺 2 (𝑧) − 4𝐺 2 (4𝑧).
Show that 𝐺 2,2 , 𝐺 2,4 are a basis for M2 (Γ0 (4)). [Hint: use Exercise
40.4(c).]
(e) Show that
∞
3 ∑︁
:=
𝐸 2,2 (𝑧) − 2 𝐺 2,2 (𝑧) = 1 + 24 𝜎 (2) (𝑛)𝑞 𝑛
𝜋 𝑛=1
∞
1 ∑︁
𝐸 2,4 (𝑧) := − 𝐺 2,4 (𝑧) = 1 + 8 𝜎 (4) (𝑛)𝑞 𝑛
𝜋2 𝑛=1
where ∑︁
𝜎 (𝑚) (𝑛) = 𝑑.
𝑚-𝑑 |𝑛
(f) Matching the first few coefficients, show that
Θ𝑄 (𝑧) = 𝐸 2,4 (𝑧).
Conclude that 𝑟 4 (𝑛) = 8𝜎 (4) (𝑛) for all 𝑛 > 0.
Chapter 41
Brandt matrices
In this chapter, we revisit classes of quaternion ideals: organizing ideals of given norm
in terms of their classes, we find modular forms.
(The notation 𝑇 (𝑛) deliberately overloads that of the Hecke operators defined in section
40.5: keep reading to see why!) The Brandt matrix 𝑇 (𝑛) depends on O, but for brevity
we do not include this in the notation. In the 𝑗th column of the Brandt matrix 𝑇 (𝑛),
we look at the subideals of 𝐼 𝑗 with index 𝑛2 and count them in the 𝑖 𝑗th entry according
to the class [𝐼𝑖 ] they belong to. If 𝑛 = 𝑝 is prime and 𝑝 - 𝑁 = disc O, then there are
747
748 CHAPTER 41. BRANDT MATRICES
exactly 𝑝 + 1 such ideals, so the sum of the entries in every column in 𝑇 ( 𝑝) is equal to
𝑝 + 1.
−1, −23
Example 41.1.2. We continue with Example 17.6.3. We have 𝐵 = of
Q
discriminant 23 and a maximal order O with three ideal classes [𝐼1 ], [𝐼2 ], [𝐼3 ]. In
(17.6.5), we found three ideals in 𝐼1 = O: two belong to the class [𝐼2 ] and the third
is principal, belonging to [𝐼1 ]. This gives the first column of the matrix as (1, 2, 0) t .
Computing further, we find
1 1 0
𝑇 (2) = 2 1 3® .
© ª
«0 1 0¬
In a similar manner, we compute
0 1 3 30 28 24
𝑇 (3) = 2 3 0® , 𝑇 (101) = 56 54 60® .
© ª © ª
«2 0 1¬ «16 20 18¬
41.1.3. There is a second and computationally more efficient way to define the Brandt
matrix using representation numbers of quadratic forms. Let 𝑞 𝑖 = nrd(𝐼𝑖 ), let O𝑖 =
OL (𝐼𝑖 ), and let 𝑤 𝑖 = #O×𝑖 /{±1} < ∞. Then
1
𝑇 (𝑛)𝑖 𝑗 = #{𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 : nrd(𝛼)𝑞 𝑖 /𝑞 𝑗 = 𝑛} :
2𝑤 𝑖
𝑄 𝑖 𝑗 : 𝐼 𝑗 𝐼𝑖−1 → Z
𝑞𝑖 (41.1.4)
𝑄 𝑖 𝑗 (𝛼) = nrd(𝛼)
𝑞𝑗
1
𝑇 (𝑛)𝑖𝑖 = #{𝛾 ∈ O𝑖 : nrd(𝛾) = 𝑛}.
2𝑤 𝑖
For 𝑖 = 1, we have 𝑤 1 = 2 and with 𝛾 = 𝑡 + 𝑥𝛼 + 𝑦𝛽 + 𝑧𝛼𝛽 and 𝑡, 𝑥, 𝑦, 𝑧 ∈ Z, by (17.6.6)
nrd(𝛾) = 𝑡 2 + 𝑡𝑦 + 𝑥 2 + 𝑥𝑧 + 6𝑦 2 + 6𝑧2
There is a third way to understand Brandt matrices which is visual and combina-
torial.
41.1. ⊲ BRANDT MATRICES, NEIGHBORS, AND MODULAR FORMS 749
I1 I2
I3
Example 41.1.11. We check that 𝑇 (2)𝑇 (3) = 𝑇 (3)𝑇 (2) from Example 41.1.2; and
𝑤 1 , 𝑤 2 , 𝑒 3 = 2, 1, 3, so we verify that
2 0 0 2 2 0
0 1 0® 𝑇 (2) = 2 1 3®
© ª © ª
«0 0 3¬ «0 3 0¬
Example 41.1.12. Returning one last time to our example, the space 𝑀2 (Γ0 (23)) of
modular forms of weight 2 and level Γ0 (23) has eigenbasis 𝑒 23 , 𝑓+ , 𝑓− where
∞
11 ∑︁ ∗ ∑︁
𝑒 23 (𝑧) = + 𝜎 (𝑛)𝑞 𝑛 , 𝜎 ∗ (𝑛) = 𝑑
12 𝑛=1
𝑑 |𝑛
23-𝑑
and √
± 5+1 2 √ 3
𝑓± (𝑧) := 𝑞 − 𝑞 + 5𝑞 + . . .
2
are cusp forms matching the eigenbasis in Example 41.1.11.
One of the main applications of Brandt matrices is to express the trace of the Hecke
operator in terms of arithmetic data, as follows.
41.2. BRANDT MATRICES 751
√
𝑑
where ℎ(𝑑) is the class number of Q( 𝑑), 𝑤 𝑑 its number of roots of unity, and is
𝑝
the Kronecker symbol. For 𝑑 0 < 0 a discriminant with 𝑑 0 = 𝑑𝑓 2 and 𝑑 fundamental,
define ℎ 𝐷 (𝑑 0) = ℎ 𝐷 (𝑑). Then the trace of the 𝑛th Brandt matrix associated to O is
(
∑︁
2 𝜑(𝐷)/12, if 𝑛 is a square;
tr 𝑇 (𝑛) = ℎ 𝐷 (𝑡 − 4𝑛) +
𝑡 ∈Z
0, otherwise
𝑡 2 <4𝑛
conversely. So equivalently
𝑀2 (O) := Map(Cls O, Z)
to be the set of maps from Cls O to Z (as sets). Then 𝑀2 (O) has the structure of an
abelian group under addition of maps, and it is a free Z-module on the characteristic
functions for Cls O. The 𝔫-Hecke operator is defined to be
again the sum over all invertible right O-ideals 𝐽 ⊆ 𝐼 with condition on the reduced
norm. Visibly, this definition does not depend on the choice of representative 𝐼 in its
right ideal class. And in the basis of characteristic functions for 𝐼𝑖 , the matrix of 𝑇 (𝔫)
is precisely the 𝔫-Brandt matrix.
Brandt matrices may be given in terms of elements instead of ideals. Let 𝑤 𝑖 =
[O×𝑖 : 𝑅 × ]. By Proposition 32.3.7, since 𝐵 is S-definite, the unit index 𝑤 𝑖 < ∞ is
finite.
Lemma 41.2.7. Let 𝔫𝑖 𝑗 = 𝔫 nrd(𝐼 𝑗 )/nrd(𝐼𝑖 ) for 𝑖, 𝑗 = 1, . . . , ℎ. Then following
statements hold.
(a) We have
1 (41.2.8)
# 𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 : nrd(𝛼)𝑅 = 𝔫𝑖 𝑗 /𝑅 ×
=
𝑤𝑖
where we count orbits under right multiplication by O×𝑖 and 𝑅 × , respectively.
(b) If the class of 𝔫𝑖 𝑗 in Cl+ 𝑅 is nontrivial, then 𝑇 (𝔫)𝑖 𝑗 = 0.
× totally positive. Then
(c) Suppose that 𝔫𝑖 𝑗 = 𝑛𝑖 𝑗 𝑅 with 𝑛𝑖 𝑗 ∈ 𝐹>0
1 ∑︁
# 𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 : nrd(𝛼) = 𝑢𝑛𝑖 𝑗
𝑇 (𝔫)𝑖 𝑗 = (41.2.9)
2𝑤 𝑖 × /𝑅 ×2
𝑢𝑅 ×2 ∈𝑅>0
S-definite.
For (c), we just need to organize our generators; the sum in (41.2.9) is finite by
the Dirichlet S-unit theorem. If nrd(𝛼)𝑅 = 𝔫𝑖 𝑗 then nrd(𝛼) = 𝑢𝑛𝑖 𝑗 for some 𝑢 ∈ 𝑅>0 × .
By definition, the adjacency matrix of the 𝔫-Brandt graph is the 𝔫-Brant matrix
𝑇 (𝔫).
41.2.13. Let 𝔭 - 𝔑 be prime and suppose that the class of 𝔭 generates Cl𝐺 (O) 𝑅. Then
by Proposition 28.5.18, we may take the ideals 𝐼𝑖 to have reduced norm a power of
𝔭. Consider the directed graph whose vertices are the right O-ideals whose reduced
norm is a power of 𝔭 with directed edges for each 𝔭-neighbor relation. (This graph is
a regular directed tree by Proposition 41.3.1 below, every vertex has out degree equal
to N𝔭 + 1.) The equivalence relation of belonging to the same right ideal class (left
equivalent by an element of 𝐵× ) respects edges, and the quotient by this equivalence
relation is the 𝔭-Brandt graph.
754 CHAPTER 41. BRANDT MATRICES
Remark 41.2.14. The Brandt graphs have interesting graph theoretic properties: they
are Ramanujan graphs (also called expander graph), having high connectivity and
are potentially useful in communication networks. In the simplest case where 𝐹 = Q
and 𝐵 is the quaternion algebra of discriminant 𝑝, they were first studied by Ihara,
then studied in specific detail by Lubotzky–Phillips–Sarnak [LPS88] and Margulis
[Marg88]; for further reading, see the books by Lubotzky [Lub2010] and Sarnak
[Sar90]. Over totally real fields, see work of Livné [Liv2001] as well as Charles–
Goren–Lauter [CGL2009]. The proof that Brandt graphs are Ramanujan relies on the
Ramanujan–Petersson conjecture, a deep statement proven by Deligne [Del74], giving
bounds on coefficients of modular forms.
Remark 41.2.15. The space of functions on Cls O can itself be understood as a space
of modular forms, a special case of the theory of algebraic modular forms due to Gross
[Gro99]. This general formulation harmonizes with the double coset description given
in section 38.7, via the canonical bijection Cls O ↔ 𝐵× \ 𝐵 b× /O
b× , but without the
geometry!
which gives the matrix product 𝑇 (𝔪𝔫) = 𝑇 (𝔪)𝑇 (𝔫). Repeating the argument inter-
changing the roles of 𝔪 and 𝔫, the result is proven.
41.3. COMMUTATIVITY OF BRANDT MATRICES 755
For prime powers coprime to 𝔑, we have a recursion for the 𝔭𝑟 -Brandt matrices
that is a bit complicated: the uniqueness of factorization fails when the product is a
two-sided ideal, so we must account for this extra term. To this end, we need to keep
track of the effect of multiplication by right ideals of 𝑅 on the class set.
41.3.4. For an ideal 𝔞 ⊆ 𝑅, let 𝑃(𝔞) ∈ Mℎ (Z) be the permutation matrix given by
𝐼𝑖 ↦→ 𝔞𝐼𝑖 . In other words, we place a 1 in the (𝑖, 𝑗)th entry according as [𝔞𝐼 𝑗 ] = [𝐼𝑖 ]
(with 0 elsewhere). The matrix 𝑃(𝔞) only depends on the class [𝔞] ∈ Cl 𝑅: in
particular, if 𝔞 is principal then 𝑃(𝔞) is the identity matrix. Therefore we have a
homomorphism
𝑃 : Cl 𝑅 → GLℎ (Z)
[𝔞] ↦→ 𝑃(𝔞).
We have
𝑃(𝔞𝔟) = 𝑃(𝔞)𝑃(𝔟) = 𝑃(𝔟)𝑃(𝔞)
and in particular 𝑃(𝔞)𝑃(𝔞−1 )
= 1 and the image 𝑃(Cl 𝑅) ⊆ GLℎ (Z) is an abelian
subgroup; however, this map need not be injective. Moreover, for all 𝔞, 𝔫 we have
by commutativity of multiplication by 𝔞.
As in 26.4.3, we say an integral right O-ideal 𝐼 is primitive if we cannot write
𝐼 = 𝔞𝐼 0 with 𝐼 0 integral and 𝔞 ( 𝑅.
Proposition 41.3.6. Let 𝔭 - 𝔑 be prime. Then for 𝑟, 𝑠 ∈ Z ≥0 ,
min(𝑟
∑︁,𝑠)
𝑇 (𝔭𝑟 )𝑇 (𝔭𝑠 ) = N (𝔭) 𝑖 𝑇 (𝔭𝑟 +𝑠−2𝑖 )𝑃(𝔭) 𝑖 . (41.3.7)
𝑖=0
Proof. When 𝑠 = 0, the matrix 𝑇 (1) is the identity and the result holds. We next
consider the case 𝑠 = 1, and will then proceed by induction, and consider the product
𝑇 (𝔭𝑟 )𝑇 (𝔭): its 𝑖 𝑗th entry
ℎ
∑︁
𝑟
(𝑇 (𝔭 )𝑇 (𝔭))𝑖 𝑗 = 𝑇 (𝔭𝑟 )𝑖𝑘 𝑇 (𝔭) 𝑘 𝑗
𝑘=1
counts the number of compatible products of right ideals 𝐽𝑟0 𝐽 0 where 𝐽𝑟0 is an invertible
O𝑖 , O 𝑘 -ideal with nrd(𝐽𝑟0 ) = 𝔭𝑟 and 𝐽 0 is an invertible O 𝑘 , O 𝑗 -ideal with nrd(𝐽 0) = 𝔭.
The issue: these products may not all be distinct when they are imprimitive. If the
product 𝐽𝑟0 𝐽 0 is imprimitive, then we rewrite it as a compatible product
0 0
𝐽𝑟0 𝐽 0 = (𝐽𝑟0 (𝐽 ) −1 )𝐽 0 𝐽 0 = 𝔭(𝐽𝑟0 (𝐽 ) −1 )
756 CHAPTER 41. BRANDT MATRICES
where now 𝐽𝑟0 −1 = 𝔭−1 𝐽𝑟0 𝐽 0 has reduced norm 𝔭𝑟 −1 . This procedure works in reverse
as well.
With apologies for the temporarily annoying notation, define 𝑇prim (𝔭𝑟 +1 ) and
𝑇imprim (𝔭𝑟 +1 ) to be the 𝔭𝑟 -Brandt matrix counting classes of primitive or imprimi-
tive, accordingly. Then
Since there are N (𝔭) + 1 right O-ideals of reduced norm 𝔭, with the previous paragraph
we obtain
𝑇 (𝔭𝑟 )𝑇 (𝔭) = 𝑇prim (𝔭𝑟 +1 ) + ( N (𝔭) + 1)𝑇imprim (𝔭𝑟 +1 )
= 𝑇 (𝔭𝑟 +1 ) + N (𝔭)𝑇imprim (𝔭𝑟 +1 ) (41.3.11)
𝑟 +1 𝑟 −1
= 𝑇 (𝔭 ) + N (𝔭)𝑇 (𝔭 )𝑃(𝔭).
This proves the result for 𝑠 = 1, and it gives (41.3.8) upon rearrangement and shifting
indices.
We now proceed by (an ugly but harmless) induction on 𝑠:
𝑇 (𝔭𝑟 ) 𝑇 (𝔭𝑠+1 ) + N (𝔭)𝑇 (𝔭𝑠−1 )𝑃(𝔭)
𝑠
∑︁
= N (𝔭) 𝑖 𝑇 (𝔭𝑟 +𝑠+1−2𝑖 )𝑃(𝔭) 𝑖 (41.3.12)
𝑖=0
+N (𝔭) 𝑖+1 𝑃(𝔭𝑟 +𝑠+1−2(𝑖+1) )𝑃(𝔭) 𝑖+1
so
𝑠
∑︁
𝑇 (𝔭𝑟 )𝑇 (𝔭𝑠+1 ) = N (𝔭) 𝑖 𝑇 (𝔭𝑟 +𝑠+1−2𝑖 𝑃(𝔭) 𝑖 )
𝑖=0
+N (𝔭) 𝑖+1 𝑃(𝔭) 𝑖+1 𝑃(𝔭𝑟 +𝑠+1−2(𝑖+1) )
𝑠
∑︁ (41.3.13)
− N (𝔭) 𝑖+1 𝑃(𝔭) 𝑖+1𝑇 (𝔭𝑟 +𝑠+1−2(𝑖+1) )
𝑖=0
𝑠+1
∑︁
= N (𝔭) 𝑖 𝑃(𝔭) 𝑖 𝑇 (𝔭𝑟 +𝑠+1−2𝑖 )
𝑖=0
as claimed.
Definition 41.3.14. The Hecke algebra T(O) is the subring of Mℎ (Z) generated by
the matrices 𝑇 (𝔫) with 𝔫 coprime to 𝔑.
Corollary 41.3.15. The ring T(O) is a commutative Z-algebra.
41.4. SEMISIMPLICITY 757
Proof. By Proposition 41.3.1(b), we reduce to showing that 𝑇 (𝔭𝑟 )𝑇 (𝔭𝑠 ) = 𝑇 (𝔭𝑠 )𝑇 (𝔭𝑟 )
for all 𝑟, 𝑠 ≥ 0, and this holds by Proposition 41.3.6: the right-hand side of (41.3.7) is
symmetric under interchanging 𝑟, 𝑠.
√ √
Example 41.3.16. Let 𝐹 = Q( 10) and 𝑅 = Z𝐹 = Z[ 10] its ring of integers.
Then the√class group Cl 𝑅 ' Z/2Z is nontrivial, represented by the class of the ideal
+
√ and the narrow class group Cl 𝑅 ' Cl 𝑅 is no bigger: the fundamental
𝔭2 = (2, 10),
unit is 3 + 10 of norm −1.
Let 𝐵 = (−1, −1 | 𝐹). Since 2 is not split in 𝐹, the ramification set Ram 𝐵 is the
set of real places of 𝐹. A maximal order is given by
1 + 𝑖 + 𝑗 + 𝑖𝑗
O = 𝑅 ⊕ 𝔭−1 −1
2 (1 + 𝑖) ⊕ 𝔭2 (1 + 𝑗) ⊕ 𝑅 .
2
We find that # Cls O = 4, and
0 0 0 1
0 0 3 2®
© ª
𝑇 (𝔭2 ) = ®.
0 2 0 0®
«3 1 0 0¬
In this case, the matrix 𝑃(𝔭2 ) is the identity matrix: for example, we have 𝔭2 O =
(1 + 𝑖)O. Thus
1 1 0 0
6 6 0 0®
© ª
𝑇 (𝔭22 ) 2
= 𝑇 (2𝑅) = 𝑇 (𝔭2 ) − 2 = ®.
0 0 4 4®
«0 0 3 3¬
41.4 Semisimplicity
We now equip the space 𝑀2 (O) = Map(Cls O, Z) with a natural inner product, and
we show that the Hecke operators are normal with respect to this inner product.
h , i : 𝑀2 (O) × 𝑀2 (O) → Z
(41.4.2)
h1 [𝐼 ] , 1 [𝐽 ] i := 𝑤 [𝐼 ] 𝛿 [𝐼 ], [𝐽 ]
where 𝛿 [𝐼 ], [𝐽 ] = 1, 0 according as [𝐼] = [𝐽] or not, and extend linearly. The matrix
of this pairing in the basis of characteristic functions is the diagonal matrix diag(𝑤 𝑖 )𝑖 ,
where 𝑤 𝑖 = [O×𝑖 : 𝑅 × ]. The pairing is symmetric and nondegenerate.
The Hecke operators 𝑇 (𝔫) are normal with respect to the inner product (41.4.2), and
for 𝔫 trivial in Cl+ 𝑅 the operators 𝑇 (𝔫) are self-adjoint.
Proof. We may show the proposition for the Brandt matrices. Let 𝑊 = diag(𝑤 𝑖 )𝑖 define
the inner product on Zℎ with the Brandt matrices acting on the right on row vectors.
Then the inner product is h𝑥, 𝑦i = 𝑥𝑊 𝑦 t and accordingly the adjoint h𝑥𝑇, 𝑦i = h𝑥, 𝑇 ∗ 𝑦i
is defined by
𝑇 ∗ = 𝑊 −1𝑇 t𝑊 (41.4.6)
The transpose of a permutation matrix is its inverse and that OL (𝔫𝐼𝑖 ) = OL (𝐼𝑖 ), so
that the unit groups match up, whence
giving (41.4.4).
For the Brandt matrices, we refer to Lemma 41.2.7(a), giving
1
# 𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 : nrd(𝛼)𝑅 = 𝔫𝑖 𝑗 /𝑅 ×
𝑇 (𝔫)𝑖 𝑗 =
𝑤𝑖
where 𝔫𝑖 𝑗 = 𝔫 nrd(𝐼 𝑗 )/nrd(𝐼𝑖 ). Let
By (41.4.6),
𝑊𝑇 (𝔫) = (Θ(𝔫)𝑖 𝑗 )𝑖, 𝑗 =: Θ(𝔫).
We extend the definition of Θ(𝔫) to include all fractional ideals 𝔫. For each 𝑖, let
𝑖 0 be such that [𝔫−1 𝐼𝑖 ] = [𝐼𝑖0 ], so that 𝔫−1 𝐼𝑖 = 𝛽𝑖 𝐼𝑖0; the induced action is given by the
permutation map 𝑃(𝔫−1 ).
nrd(𝐼 𝑗 )
𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 = 𝐼𝑖−1 𝐼 𝑗 = 𝐼𝑖 𝐼 −1
𝑗 (41.4.9)
nrd(𝐼𝑖 )
𝛼−1 ∈ 𝔫−1 𝐼𝑖 𝐼 −1 −1
𝑗 = 𝛽𝑖 𝐼 𝑖 0 𝐼 𝑗
41.5. EICHLER TRACE FORMULA 759
𝔫−2 𝔫 𝑗𝑖 so nrd(𝛽𝑖−1 𝛼−1 ) = 𝔫 𝑗𝑖0 . We can run the argument in the other direction to
produce an inverse, and we thereby conclude the map is bijective.
The map (41.4.8) together with the action by permutation and 𝑊 t = 𝑊 yields
and thus 𝑇 (𝔫) ∗ = 𝑃(𝔫) ∗𝑇 (𝔫), and substituting (41.4.7) gives (41.4.5).
For the final statement, by (41.3.5) we have 𝑇 (𝔫) commuting with 𝑃(𝔫), so 𝑇 (𝔫)
commutes with 𝑇 (𝔫) ∗ ; and when 𝔫 is narrowly principal, then 𝑃(𝔫−1 ) is the identity
matrix so 𝑇 (𝔫) ∗ = 𝑇 (𝔫).
Corollary 41.4.10. T(O) is a semisimple commutative ring, and there exists a basis
of common eigenvectors (eigenfunctions) for the Hecke operators. Each 𝑇 (𝔫) with 𝔫
narrowly principal has real eigenvalues.
We also recall
ℎ
∑︁ 1
mass(Cls O) :=
𝑖=1
𝑤 𝑖
and that the Eichler mass formula (Main Theorem 26.1.5) gives an explicit formula for
this mass in terms of the relevant arithmetic invariants.
Main Theorem 41.5.2 (Trace formula). If 𝔫 is not narrowly principal, then tr 𝑇 (𝔫) = 0.
× , then
If 𝔫 = 𝑛𝑅 is narrowly principal with 𝑛 ∈ 𝐹>0
(
1 ∑︁ ℎ(𝑆) b b b× mass(Cls O), if 𝔫 = 𝑐2 𝑅, 𝑐 ∈ 𝐹 × ;
tr 𝑇 (𝔫) = 𝑚( 𝑆, O; O ) +
2 𝑤𝑆 0, otherwise
(𝑢,𝑡 ,𝑆)
where 𝑤 𝑆 := [𝑆 × : 𝑅 × ] and the sum is over finitely many triples (𝑢, 𝑡, 𝑆) where:
• 𝑆 ⊇ 𝑅[𝑥]/(𝑥 2 − 𝑡𝑥 + 𝑢𝑛).
Í𝑘
Proof. We have tr 𝑇 (𝔫) = 𝑖=1 𝑇 (𝔫)𝑖𝑖 . By Lemma 41.2.7, since 𝔫𝑖𝑖 = 𝔫 we conclude
tr 𝑇 (𝔫) = 0 if 𝔫 is not narrowly principal. So suppose 𝔫 = 𝑛𝑅 is narrowly principal,
with 𝑛 ∈ 𝐹>0× . Then by (41.2.9) we have
1 ∑︁
𝑤 𝑖 𝑇 (𝔫)𝑖𝑖 = #{𝛼 ∈ O𝑖 : nrd(𝛼) = 𝑢𝑛}. (41.5.3)
2 × /𝑅 ×2
𝑢𝑅 ×2 ∈𝑅>0
1 ∑︁ ∑︁
𝑤 𝑖 𝑇 (𝔫)𝑖𝑖 = #{𝛼 ∈ O𝑖 : trd(𝛼) = 𝑡, nrd(𝛼) = 𝑢𝑛}. (41.5.4)
2 𝑢 𝑡 ∈𝑅
Since 𝐵 is definite, we have disc(𝛼) = 𝑡 2 − 4𝑢𝑛 either zero or totally negative, so the
inner sum is over finitely many 𝑡 ∈ 𝑅 either satisfying 𝑡 2 = 4𝑢𝑛 or 𝑡 2 − 4𝑢𝑛 ∈ 𝐹<0 × .
1 ∑︁ ∑︁ 𝑤𝑖
𝑤 𝑖 𝑇 (𝔫)𝑖𝑖 = 𝑚(𝑆, O𝑖 ; O×𝑖 )
2 𝑢,𝑡 𝑤𝑆
𝑆 ⊇𝑅 [𝑥 ]/( 𝑥 2 −𝑡 𝑥+𝑢𝑛)
( (41.5.5)
1, if 𝔫 = (𝑐𝑅) 2 ;
+
0, otherwise.
1 ∑︁ ℎ(𝑆) b b b×
# Cls O = mass(Cls O) + 𝑚( 𝑆, O; O )
2 𝑤𝑆
(𝑢,𝑡 ,𝑆)
Corollary 41.5.7 gives a different way to prove (and interpret) the Eichler class
number formula (Main Theorem 30.8.6): for the exact comparison, see Exercise 41.2.
ℎ(𝑆) b b b×
ℎO (𝑆) := 𝑚( 𝑆, O; O )
𝑤𝑆
b O;
where the factor 𝑚( 𝑆, b Ob× ) is defined by purely local data, given in section 30.5 for
maximal orders and section 30.6 for Eichler orders. Writing ℎO (𝑑) = ℎO (𝑆 𝑑 ) for the
order of discriminant 𝑑, we arrive at a pleasing formula:
1 ∑︁ ∑︁
tr 𝑇 (𝑛) = ℎO (𝑑) (41.5.10)
2 𝑡 ∈Z
𝑑 𝑓 2 =𝑡 2 −4𝑛<0
1 ∑︁
# Cls O = mass(Cls O) + ℎO (𝑡 2 − 4)
2 𝑡 ∈Z
𝑡 2 <4 (41.5.11)
1
= mass(Cls O) + ℎO (−4) + 2ℎO (−3).
2
For O an Eichler order, after substitution we recover the Eichler class number formula
(Theorem 30.1.5).
762 CHAPTER 41. BRANDT MATRICES
𝑡 𝑑 ℎ(𝑆 𝑑 ) 𝑤 𝑆𝑑 𝑚( 𝑆b𝑑 , O;
b Ob× ) ℎO (𝑑)
0 −12 1 1 0 0
0 −3 1 3 2 2/3
±1 −11 1 1 1 − (−11 | 2) = 2 2
±2 −8 1 1 1 − (−8 | 2) = 1 1
±3 −3 1 3 1 − (−3 | 2) = 2 2/3
Indeed, more generally since the Hurwitz order has # Cls O = 1, the matrix 𝑇 (𝑛) is
a 1 × 1-matrix with 𝑇 (𝑛) = [𝜎(𝑛)] for 𝑛 odd. This observation implies a nontrivial
(and otherwise surprising) relationship between class numbers of imaginary quadratic
orders!
Remark 41.5.13. Brandt [Bra43, §III] defined Brandt matrices in the same paper as
his groupoid; he called them Hecke matrices, as he claimed to follow parallels with
certain operators defined by Hecke. Indeed, Hecke [Hec40, §9, Satz 53] conjectured
that the space of cusp forms of weight 2 on Γ0 ( 𝑝) for 𝑝 prime was spanned by certain
linear combinations of theta series, and it was this observation that motivated Brandt.
(Eichler [Eic56a, footnote 16] says that Brandt should not have named them after
Hecke, since it was really Brandt who interpreted function-theoretic results of Hecke
using pure arithmetic.)
Eichler [Eic56a] proved that the ring generated by the Brandt matrices was a com-
mutative, semisimple ring and proved the trace formula for Brandt matrices [Eic56a,
§6]. In this early work, he already foresaw the application of Brandt matrices to other
base fields: as an application, he used Brandt matrices to give class number relations
between imaginary quadratic fields, and in the function field case these become rela-
tions among divisor class groups for hyperelliptic curves. Eichler [Eic77, Chapter II]
presented the generalization to totally real fields, giving a treatment of Hecke opera-
tors, Brandt matrices, and theta series, and he proved that the Brandt matrices realize
Hecke operators in certain spaces of Hilbert modular forms.
Eichler then later gave a self-contained presentation [Eic73, Chapter II] of the
theory of Brandt matrices over Q, with the intended application the solution to Hecke’s
conjecture (suitably corrected), now known as the basis problem for Γ0 ( 𝑝): to give
bases of linearly independent forms of spaces of modular forms in terms of theta series
of quadratic forms coming from quaternion algebras. This line of work was followed
by generalizations by Hijikata [Hij74] and Hijikata–Saito [HS73] for general Eichler
orders, Pizer [Piz76b, Piz76c] for residually split orders, culminating in a solution over
the rational numbers to the basis problem by Hijikata–Pizer–Shemanske [HPS89a].
The method of proof for the solution to the basis problem is the use of the trace
formula, for which a key ingredient is the theory of optimal embeddings: see Remark
30.6.18.
41.5. EICHLER TRACE FORMULA 763
Indeed, it is much more involved analytically, but one can similarly compute
the trace of the Hecke operator acting on classical spaces of modular forms or more
generally spaces of Hilbert modular forms. These trace formulae are quite complicated,
but one notices that they have a similar shape as the above trace formula; and in fact,
under certain hypotheses and after restricting to an appropriate new subspace, the traces
are equal. But since both rings are semisimple, this implies that the same systems of
eigenvalues for the Hecke operators arise! Such a correspondence was first given by
Eichler, as above; it was generalized to totally real fields by Shimizu [Shz72] using theta
series, and the most general formulation given by Jacquet–Langlands [JL70]. This
correspondence was conjectured to generalize to the principle of Langlands functorial
transfer: for an introduction to this vast area, see Gelbart [Gel84].
In light of the preceding epic remark, we hope we have inspired the reader to pursue
the relationship between Brandt matrices and modular forms! Unfortunately, it would
require another book to respectfully develop this subject.
Remark 41.5.14. Pizer [Piz80a] was the first to give an algorithm for computing clas-
sical modular forms using Brandt matrices (on Γ0 (𝑁) for 𝑁 not a perfect square); see
also the work of Kohel [Koh2001] over Z. This algorithm was generalized to compute
Hilbert modular forms over a totally real field of narrow class number 1 by Socrates–
Whitehouse [SW2005], with algorithmic improvements by Dembélé [Dem2007]. The
assumption on the class number was removed by Dembélé–Donnelly [DD2008]. A
survey of these methods are given by Dembélé–Voight [DV2013, §4, §8].
Exercises
Unless otherwise specified, in these exercises let 𝑅 be a global ring with eligible set
S ⊆ Pl 𝐹, let 𝐵 be an S-definite quaternion algebra over 𝐹 and let O ⊂ 𝐵 be an 𝑅-order
in 𝐵.
1. Extend the definition of the Brandt matrix to include the case 𝔫 = (0) of the zero
ideal, following (41.2.8): define 𝑇 (0)𝑖 𝑗 = 1/𝑤 𝑖 for 𝑖, 𝑗 = 1, . . . , ℎ. Conclude
tr 𝑇 (0) = mass(Cls O).
2. Show that Corollary 41.5.7 agrees with Main Theorem 30.8.6. [Hint: organize
by 𝑞 := [𝑆 × : 𝑅 × ], observe that in 𝑆 ⊇ 𝑅[𝑥]/(𝑥 2 − 𝑡𝑥 + 𝑢) we have necessarily
[𝑆 × : 𝑅 × ] ≥ 2 and each such 𝑆 contains 𝑞 − 1 orders of the form 𝑅[𝑥]/(𝑥 2 −
𝑡𝑥 + 𝑢).]
3. Refine Lemma 41.2.7(c) in a special case as follows. Suppose Cl+ 𝑅 is trivial.
Show that
1
𝑇 (𝔫)𝑖 𝑗 = #{𝛼 ∈ 𝐼 𝑗 𝐼𝑖−1 : nrd(𝛼) = 𝑛𝑖 𝑗 }
𝑤 𝑖,1
where
ℎ(−4𝑝)/2, if 𝑝 ≡ 1 (mod 4);
ℎO (−4𝑝) = ℎ(−𝑝), if 𝑝 ≡ 7 (mod 8);
2ℎ(−𝑝), if 𝑝 ≡ 3 (mod 8) and 𝑝 > 3.
What does this say about the number of maximal orders in 𝐵 up to isomorphism
such that every two-sided ideal is principal?
5. Give another proof of Proposition 41.3.1(b) using the local-global dictionary for
lattices.
6. Prove (41.3.12) using induction and then expand to verify (41.3.13).
−1, −11
7. Let 𝐵 = with disc 𝐵 = 11 and let O = Zh𝑖, 12 ( 𝑗 + 1)i.
Q
(a) Show that O is a maximal order with #O× = 4.
(b) Show that the ternary quadratic form associated to O is similar to 𝑥 2 − 𝑥𝑧 +
𝑦 2 + 3𝑧2 .
(c) Show that Cl O = {[O], [𝐼2 ]} where 𝐼2 = 2O + 12 (1 + 2𝑖 + 𝑗)O. [Hint:
Follow Example 17.6.3.] Along the way, show that
1 3
𝑇 (2) = .
2 0
(d) Pause and show that O2 := OL (𝐼2 ) has #O×2 = 6 and associated ternary
quadratic form 𝑥 2 − 𝑥𝑦 − 𝑥𝑧 + 𝑦 2 + 𝑦𝑧 + 4𝑧2 .
(e) Show that 𝑀2 (O) has two eigenspaces for the Hecke algebra, one spanned
by a form 𝑒 with 𝑇 ( 𝑝) (𝑒) = ( 𝑝 + 1)𝑒 for all 𝑝 ≠ 11, and the other spanned
by a form 𝑓 with 𝑇 (2) ( 𝑓 ) = −2 𝑓 .
(f) Verify the trace formula (41.5.9) for tr 𝑇 (2) = 1 by computing class num-
bers.
[There is a unique normalized cusp form 𝑓 ∈ 𝑆2 (Γ0 (11)) of weight 2 and level
11 with
∞
Ö ∞
Ö ∞
∑︁
𝑓 (𝑞) = 𝑞 (1 − 𝑞 𝑛 ) 2 (1 − 𝑞 11𝑛 ) 2 = 𝑞 − 2𝑞 2 − 𝑞 3 + · · · = 𝑎𝑛 𝑞𝑛
𝑛=1 𝑛=1 𝑛=1
𝑓𝑖 = 𝑞 + 2𝑞 5 − 3𝑞 9 − 6𝑞 13 + 2𝑞 17 + . . .
In the previous chapter, we showed that Brandt matrices for an order in a definite
quaternion algebra 𝐵 contain a wealth of arithmetic. In the special case where disc 𝐵 =
𝑝 is prime, there is a further beautiful connection between Brandt matrices and the
theory of supersingular elliptic curves, arising from the following important result:
there is an equivalence of categories between supersingular elliptic curves over F 𝑝 and
right ideals in a (fixed) maximal order O ⊂ 𝐵. We pursue this important connection
in this chapter for the reader who has a bit more background in algebraic curves.
767
768 CHAPTER 42. SUPERSINGULAR ELLIPTIC CURVES
Proof. We apply Theorem 3.5.1 to conclude that End(𝐸)Q is either Q, a quadratic field,
or a division quaternion algebra. Then by Example 8.4.2, the involution is positive if
and only if End(𝐸)R is R, C, or H, so in the second case we must have an imaginary
quadratic field and in the third case we must have a definite quaternion algebra.
42.1.8. One can often reduce questions about supersingular elliptic curves to ones
where the base field 𝐹 is F 𝑝2 as follows: by Proposition 42.1.7(iii), if 𝐸 is supersingular
then 𝐸 is isomorphic over 𝐹 al to a curve 𝐸 defined over F 𝑝2 .
The following fundamental result is due to Deuring [Deu41]; we give a proof due
to Lenstra [Len96, §3].
Theorem 42.1.9. Let 𝐸 be an elliptic curve over 𝐹 and suppose that rkZ End(𝐸) = 4.
Then 𝐵 = End(𝐸)Q is a quaternion algebra over Q ramified at 𝑝 = char 𝐹 and ∞, and
End(𝐸) is a maximal order in 𝐵.
In particular, if over 𝐹 we have dim End(𝐸) = 4, then automatically 𝐸 has all of
its endomorphisms defined over 𝐹.
42.1. SUPERSINGULAR ELLIPTIC CURVES 769
as abelian groups, and the endomorphism ring of this abelian group is End 𝐸 [𝑛] '
M2 (Z/𝑛Z).
We claim that the structure map O/𝑛O → End 𝐸 [𝑛] is injective, which is to say,
𝐸 [𝑛] is a faithful module over O/𝑛O. Indeed, suppose 𝜙 ∈ O annihilates 𝐸 [𝑛]; then
since multiplication by 𝑛 is separable, by the homomorphism theorem for elliptic curves
[Sil2009, Corollary III.4.11] there exists 𝜓 ∈ O such that 𝜙 = 𝑛𝜓, so 𝜙 ≡ 0 ∈ O/𝑛O,
proving injectivity. But further, since #O/𝑛O = # End 𝐸 [𝑛] = 𝑛4 , the structure map is
an isomorphism.
Since O is a free Z-module, we have
The structure isomorphisms in the previous paragraph are compatible with respect to
powers of ℓ, so with the previous line they provide an isomorphism
∼ lim End 𝐸 [ℓ 𝑛 ] = End 𝐸 [ℓ ∞ ] ' M (Z )
Oℓ −→ Zℓ 2 ℓ
←−−
𝑛
𝑣 : End(𝐸)Q → Q ∪ {∞}
1 (42.1.10)
𝑣(𝑎𝜙) = ord 𝑝 (𝑎) + ord 𝑝 (deg𝑖 𝜙)
2
for 𝑎 ∈ Q and 𝜙 ∈ End(𝐸) is well-defined (since deg𝑖 [ 𝑝] = deg[ 𝑝] = 𝑝 2 ). Factoring
an isogeny into its separable and inseparable parts shows that
𝑣(𝛼) = ord 𝑝 (𝑎) + 𝑣(𝜙) ≥ 0 and 0 ≤ 𝑣(𝜙) ≤ 1/2, since the multiplication by 𝑝 is
purely inseparable; so ord 𝑝 (𝑎) ≥ −1/2 and therefore 𝑎 ∈ Z ( 𝑝) , and hence 𝛼 ∈ O ( 𝑝) .
Finally, since an order is maximal if and only if it is locally maximal, O itself is a
maximal order in the quaternion algebra 𝐵.
In particular, if 𝐸, 𝐸 0 are supersingular elliptic curves over Fal𝑝 , then there exists a
separable isogeny 𝐸 → 𝐸 0.
Proof. We may suppose 𝐸 is defined over a finite field F𝑞 such that 𝐸 has all of its
endomorphisms defined over F𝑞 . Let 𝜋 ∈ O = End(𝐸) be the 𝑞-power Frobenius
endomorphism. Then 𝐵 = O ⊗Z Q is a quaternion algebra over Q. Since End(𝐸) is
defined over F𝑞 , the endomorphism 𝜋 commutes with every isogeny 𝛼 ∈ O, and
so 𝜋 lies in the center of O; since 𝑍 (𝐵) = Q, we have 𝜋 ∈ Z = 𝑍 (O). But
√ √
deg 𝜋 = 𝜋𝜋 = 𝜋 2 = 𝑞 so 𝜋 = ± 𝑞 ∈ Z. Therefore #𝐸 (F𝑞 ) = 𝑞 + 1 ∓ 2 𝑞. Therefore
2
#𝐸 (F𝑞2 ) = 𝑞 + 1 − 2𝑞 = (𝑞 − 1) . 2
𝜙∗𝐼 : Hom(𝐸 𝐼 , 𝐸) → 𝐼
(42.2.8)
𝜓 ↦→ 𝜓𝜙 𝐼
𝜄 : End(𝐸 𝐼 ) ↩→ 𝐵
1 (42.2.10)
𝜄(𝛽) = 𝜙−1
𝐼 𝛽𝜙 𝐼 = (𝜙∨ 𝛽𝜙 𝐼 )
deg 𝜙 𝐼 𝐼
Proof. The equality in (42.2.10) is justified in (42.1.4). The content of the lemma
follows from the identification in the previous Lemma 42.2.7, by transporting structure:
for 𝛽 ∈ End(𝐸 𝐼 ) acting by precomposition, we fill in the diagram
𝜙 𝐼∗
Hom(𝐸 𝐼 , 𝐸) /𝐼
𝛽 𝛽∗ (42.2.11)
𝜙 𝐼∗
Hom(𝐸 𝐼 , 𝐸) /𝐼
to find that
𝛽∗ (𝜓𝜙 𝐼 ) = 𝜓 𝛽𝜙 𝐼 = 𝜓𝜙 𝐼 (𝜙−1
𝐼 𝛽𝜙 𝐼 ) (42.2.12)
and so 𝜄 defines the induced action on 𝐼 by right multiplication, giving an inclusion
𝜄(End(𝐸 𝐼 )) ⊆ OR (𝐼). But End(𝐸 𝐼 ) is a maximal order and OR (𝐼) is an order, so
equality holds.
Next, we show that the isomorphism class of 𝐸 𝐼 depends only on the left ideal
class of 𝐼.
We claim that 𝛽𝐸 [𝐼 𝛽] = 𝐸 [𝐼]. The containment (⊆) is immediate. For the contain-
ment (⊇), let 𝑄 ∈ 𝐸 [𝐼]. Since 𝛽 is surjective (it is nonconstant), there exists 𝑃 ∈ 𝐸 (𝐹)
such that 𝛽(𝑃) = 𝑄. Thus for all 𝛼 ∈ 𝐼, we have (𝛼𝛽) (𝑃) = 𝛼(𝑄) = 0 so 𝑃 ∈ 𝐸 [𝐼 𝛽].
By the claim, we conclude that 𝜙 𝐼 𝛽 = 𝜙 𝐼 𝛽 and 𝐸 𝐼 𝛽 ' 𝐸 𝐼 .
In general, there exists nonzero 𝑚 ∈ Z such that 𝑚𝛽 ∈ O. By the previous
paragraph, we have isomorphisms 𝐸 𝐼 ' 𝐸 𝐼 (𝑚𝛽) = 𝐸 (𝐼 𝛽)𝑚 ' 𝐸 𝐼 𝛽 .
So far, we have shown how to pass from classes of left O-ideals to (isogenous)
supersingular elliptic curves via kernels. We can also go in the other direction.
Proof. Let 𝜙1 : 𝐸 → 𝐸/𝐻1 . Factoring, without loss of generality we may assume that
𝜙1 is either separable or purely inseparable. Suppose first that 𝜙1 is separable, and let
𝑛 = #𝐻2 (𝐹). By the proof of Theorem 42.1.9, the structure map O/𝑛O → End 𝐸 [𝑛]
is faithful. So if 𝐻2 > 𝐻1 , then there exists 𝛼 ∈ O such that 𝛼(𝐻1 ) = {0} but
𝛼(𝐻2 ) ≠ {0}, so 𝐼 (𝐻2 ) ≠ 𝐼 (𝐻1 ). Second, suppose that 𝜙1 is purely inseparable: then
𝐻1 = ker 𝜙𝑟𝑝1 is the kernel of the 𝑟 1 -power Frobenius for some 𝑟 1 > 0, and 𝐼 (𝐻1 ) = 𝑃𝑟1
as in 42.2.4. Then 𝑝 𝑟1 ∈ 𝐼 (𝐻1 ) = 𝐼 (𝐻2 ), so 𝐸 → 𝐸/𝐻2 is also purely inseparable,
and 𝐻2 = ker 𝜙𝑟𝑝2 and 𝐼 (𝐻2 ) = 𝑃𝑟2 . We conclude 𝑟 1 = 𝑟 2 , and then 𝐻1 = 𝐻2 .
Proposition 42.2.16 justifies the use of overloaded notation. Our proof follows
Waterhouse [Wate69, Theorem 3.15].
Proof. We begin with (a). We first prove it in an illustrative special case. Suppose
𝐼 = O𝛽 is a principal left O-ideal. Then 𝐸 [𝐼] = 𝐸 [𝛽] where 𝜙 𝐼 = 𝛽 : 𝐸 → 𝐸, and
But
so equality holds and 𝐸 [𝐼] = 𝐸 [𝐽]. Thus deg 𝜙 𝐼 = deg 𝜙 𝐽 . By (a), we have
Corollary 42.2.21. For every isogeny 𝜙 : 𝐸 → 𝐸 0, there exists a left O-ideal 𝐼 and
an isomorphism 𝜌 : 𝐸 𝐼 → 𝐸 0 such that 𝜙 = 𝜌𝜙 𝐼 . Moreover, for every maximal order
O0 ⊆ 𝐵, there exists 𝐸 0 such that O0 ' End(𝐸 0).
Hom(𝐸 𝐼 0 , 𝐸 𝐼 ) → (𝐼 0 : 𝐼) R = 𝐼 −1 𝐼 0
(42.2.23)
𝜓 ↦→ 𝜙−1
𝐼 𝜓𝜙 𝐼 0 .
42.3. EQUIVALENCE OF CATEGORIES 775
therefore ∑︁
[1] = 𝛼𝑖 𝛽𝑖∨ ∈ End(𝐸). (42.2.25)
𝑖
𝐼𝜙−1
𝐼 Hom(𝐸 𝐼 0 𝐸 𝐼 )𝜙 𝐼 0 = 𝐼
0
(42.2.27)
and thereby the bijective map (42.2.23), using Exercise 16.11 for the relationship to
the colon ideal.
𝜙∗ : Hom(𝐸 0, 𝐸 0 ) → Hom(𝐸, 𝐸 0 )
(42.3.4)
𝜓 ↦→ 𝜓𝜙.
776 CHAPTER 42. SUPERSINGULAR ELLIPTIC CURVES
is indeed bijective.
Remark 42.3.6. See also Kohel [Koh96, Theorem 45], where the categories are en-
riched with a Frobenius morphism.
Corollary 42.3.7 (Deuring correspondence). There is a bijection between isomor-
phism classes of supersingular elliptic curves over 𝐹 and the left class set ClsL O0 .
Under this bijection, if 𝐸 ↔ [𝐼], then End(𝐸) ' OR (𝐼) and Aut(𝐸) ' OR (𝐼) × .
Proof. Take isomorphism classes on both sides of the equivalence in Theorem 42.3.2,
and compare endomorphism groups and automorphism groups. (We had to work
with left O0 -modules in the equivalence of categories, but each isomorphism class of
objects is represented by a left O0 -ideal 𝐼 ⊆ 𝐵.)
42.3.8. From the Eichler mass formula and Corollary 42.3.7 (swapping left for right,
as in Remark 42.3.3), we conclude that
∑︁ 1 ∑︁ 1 𝑝−1
= = (42.3.9)
# Aut(𝐸) #OL (𝐼) × 24
[𝐸 ] [𝐼 ] ∈Cls R O
where the sum on the left is over isomorphism classes of supersingular elliptic curves
over 𝐹 = Fal𝑝 .
Similarly, from the Eichler class number formula (Theorem 30.1.5), the number of
isomorphism classes of supersingular elliptic curves over 𝐹 is equal to
𝑝 − 1 𝜖2 −4 𝜖3 −3
+ 1− + 1− .
12 4 𝑝 3 𝑝
42.4. SUPERSINGULAR ENDOMORPHISM RINGS 777
Remark 42.3.10. We can generalize this setup slightly as follows. Let 𝑀 ∈ Z>0
be coprime to 𝑝, and let 𝐶0 ≤ 𝐸 0 (𝐹) be a cyclic subgroup of order 𝑀. Then
End(𝐸 0 , 𝐶0 ) ' O0 (𝑀) is an Eichler order of level 𝑀 and reduced discriminant
𝑝𝑀 in 𝐵0 . In a similar way as above, one can show that Hom(−, (𝐸 0 , 𝐶0 )) defines
an equivalence of categories between the category of supersingular elliptic curves
equipped with a cyclic 𝑀-isogeny (under isogenies identifying the cyclic subgroups),
to the category of left invertible O0 (𝑀)-modules (under homomorphisms). The mass
formula now reads
∑︁ 1 ∑︁ 1 𝑝−1
= = 𝜓(𝑀).
# Aut(𝐸, 𝐶) #OL (𝐼) 24
[ (𝐸 ,𝐶) ] [𝐼 ] ∈Cls O0 ( 𝑀 )
Lemma 42.4.1. Let O be a maximal order. Then there exist one or two supersingular
elliptic curves 𝐸 up to isomorphism over 𝐹 such that End(𝐸) ' O. There exist two
such elliptic curves if and only if 𝑗 (𝐸) ∈ F 𝑝2 r F 𝑝 if and only if the unique two-sided
ideal of O of reduced norm 𝑝 is not principal.
Proof. In Corollary 42.2.21, we proved that there is always at least one supersingular
elliptic curve 𝐸 with End(𝐸) ' O using a connecting ideal. We now elaborate on this
point, refining our count.
By Corollary 42.3.7, the isomorphism classes of supersingular elliptic curves are
in bijection with the left class set ClsL O0 ; their endomorphism rings are then given by
End(𝐸) ' OR (𝐼) for [𝐼] ∈ ClsL O0 . By Lemma 17.4.13 (interchanging left for right),
the map
ClsL O0 → Typ O0
(42.4.2)
[𝐼] L ↦→ class of OR (𝐼)
is a surjective map of sets. The connecting ideals are precisely the fibers of this
map, and by the bijection of Corollary 42.3.7, there is a bijection between the set of
supersingular elliptic curves 𝐸 with End(𝐸) ' O and the fiber of this map over the
isomorphism class of O.
We now count these fibers. We recall Theorem 18.1.3 with 𝐷 = 𝑝 and the
text that follows (interchanging left for right): the fibers are given by the quotient
group PIdl O\Idl O of the two-sided invertible fractional two-sided O-ideals by the
subgroup of principal such ideals. There is a surjection Pic(O) → PIdl O\Idl O and
Pic(O) ' Z/2Z is generated by the unique maximal two-sided ideal 𝑃 of reduced norm
𝑃. The class of 𝑃 in the quotient is trivial if and only if 𝑃 = O𝜋 is principal.
To conclude we recall 42.2.4: the Frobenius map is the map 𝐸 → 𝐸 𝑃 ' 𝐸 ( 𝑝) . So
𝑃 is principal if and only if 𝐸 ( 𝑝) ' 𝐸 if and only if 𝑗 (𝐸) = 𝑗 (𝐸 ( 𝑝) ) = 𝑗 (𝐸) 𝑝 if and
only if 𝑗 (𝐸) ∈ F 𝑝 .
We dig into this issue a bit further.
42.4.3. Let 𝐸 be a supersingular elliptic curve over Fal𝑝 . Let 𝜔 be a nonzero invariant
differential on 𝐸. Then there is a ring homomorphism
𝑎 : End(𝐸) → Fal𝑝
(42.4.4)
𝜙 ↦→ 𝑎 𝜙 , where 𝜙∗ 𝜔 = 𝑎 𝜙 𝜔
(see Silverman [Sil2009, Corollary III.5.6]) independent of the choice of 𝜔.
In light of 42.4.3, we make the following definitions. Let 𝐵 be a quaternion algebra
over Q of discriminant disc 𝐵 = 𝑝; such an algebra 𝐵 is unique up to isomorphism.
Let O ⊆ 𝐵 be a maximal order in 𝐵; then discrd O = 𝑝.
Definition 42.4.5. An orientation of O is a ring homomorphism O → Fal𝑝 .
42.4.6. We claim that there are two possible orientations of O. In fact, 𝑃 = [O, O]
is the commutator ideal (cf. Exercise 13.8) and an orientation factors through the
commutator. Since nrd(𝑃) = 𝑝, localizing we have O/𝑃 ' O 𝑝 /𝑃 𝑝 ' F 𝑝2 by Theorem
13.3.11(c). The claim follows as there are two possible inclusions F 𝑝2 ↩→ Fal𝑝 .
42.4. SUPERSINGULAR ENDOMORPHISM RINGS 779
Proof. The association has the right target by Theorem 42.1.9 for the order and 42.4.3
for the orientation. This association is (covariantly) functorial with respect to isomor-
phisms. Indeed, if 𝜓 : 𝐸 −→∼ 𝐸 0 is an isomorphism of elliptic curves, then we have an
induced isomorphism
End(𝐸) → End(𝐸 0)
(42.4.9)
𝜙 ↦→ 𝜓𝜙𝜓 −1 .
that is compatible with composition. The isomorphism (42.4.9) is also compatible
with orientations, as follows. Let 𝜔 0 be a nonzero invariant differential on 𝐸 0; then
𝜓 ∗ 𝜔 0 is so on 𝐸. Thus for all 𝜙 ∈ End(𝐸), we have
𝑎 𝜙 𝜓 ∗ 𝜔 0 = 𝜙∗ 𝜓 ∗ 𝜔 0 = 𝜓 ∗ (𝜓 ∗ ) −1 𝜙∗ 𝜓 ∗ 𝜔 0 = 𝜓 ∗ (𝜓𝜙𝜓 −1 ) ∗ 𝜔 0
(42.4.10)
= 𝜓 ∗ (𝑎 0𝜓 𝜙 𝜓−1 𝜔 0) = 𝑎 0𝜓 𝜙 𝜓−1 𝜓 ∗ 𝜔 0
from (42.4.9) is bijective. Any two reduced isomorphisms on the left differ by an
automorphism of 𝐸, and the same on the right, so it suffies to show that the map
is bijective. Let O = End(𝐸), so Aut(𝐸) ' O× . Then Aut(O) ' 𝑁 𝐵× (O)/Q× , and
where the Atkin–Lehner group AL(O) is nontrivial (and isomorphic to Z/2Z) if and
only if 𝑃 = 𝜋O is principal; but conjugation by 𝜋 acts nontrivially on O/𝑃 and fails to
commute with 𝜁 so does not act by an automorphism of (End(𝐸), 𝜁); thus
(We did not need to choose a base object in order to define this equivalence!)
Exercises
1. Let 𝐸, 𝐸 0 be elliptic curves over 𝐹 and suppose 𝐸, 𝐸 0 are isogenous (necessarily
by a nonzero isogeny). Show that 𝐸 is supersingular if and only if 𝐸 0 is
supersingular. [Hint: show dimQ End(𝐸 𝐹 al )Q = dimQ End(𝐸 𝐹0 al )Q ; or show
that deg𝑖 ( [ 𝑝]) = deg𝑖 ( [ 𝑝] 0) where [ 𝑝], [ 𝑝] 0 are multiplication by 𝑝 on 𝐸, 𝐸 0.]
2. Let 𝐸 be an elliptic curve over 𝐹 with char 𝐹 = 𝑝. Show that for all 𝜙, 𝜓 ∈
End(𝐸), we have
QM abelian surfaces
𝜄∞ : 𝐵 → 𝐵 ⊗Q R ' M2 (R)
783
784 CHAPTER 43. ABELIAN SURFACES WITH QM
𝑋 1 := Γ1 (O)\H2(∗) ,
acting by automorphisms of 𝑋 1 .
In any event, the main result of this chapter (Main Theorem 43.6.14) is that this
association is bijective.
43.1. ⊲ QM ABELIAN SURFACES 785
( 𝐴, 𝜄) principally polarized
complex abelian surfaces
1 2
Γ (O)\H ↔
with QM by (O, 𝜇)
(43.1.5)
up to isomorphism
1
Γ (O)𝜏 ↦→ [( 𝐴 𝜏 , 𝜄 𝜏 )]
is a bijection.
43.1.6. We then define modular forms as for the classical modular group. Let 𝑘 ∈
2Z ≥0 . A map 𝑓 : H2 → C is weight 𝑘-invariant under Γ = Γ1 (O) if
𝑎 𝑏
𝑓 (𝛾𝑧) = (𝑐𝑧 + 𝑑) 𝑘 𝑓 (𝑧) for all 𝛾 = ∈ Γ. (43.1.7)
𝑐 𝑑
has the structure of a graded C-algebra under multiplication. (When 𝐷 > 1, there are
no cusps, so vacuously all modular forms are cusp forms.)
It would not be unreasonable for us to have started the book here, with this topic
at front and center. In this chapter, we will do our best to treat the complex analytic
theory in as complete and self-contained a manner as possible, but this is really just the
beginning of the subject, one that is rich, deep, and complicated—worthy of a book
all to itself. For example, the following result is fundamental.
Theorem 43.1.9 (Shimura [Shi67, p. 58]). There exists a projective nonsingular curve
𝑋 1 defined over Q and a biholomorphic map
∼ 𝑋 1 (C).
𝜑 : Γ1 (O)\H2 −→
786 CHAPTER 43. ABELIAN SURFACES WITH QM
The curve 𝑋 1 over Q coarsely represents the functor from schemes over Q to
sets whose values are isomorphism classes of QM abelian schemes, suitably defined.
Moreover, the map 𝜑 respects the field of definition and Galois action on certain
special points called CM points on Γ1 (O)\H2 obtained as fixed points of elements
𝜈 ∈ 𝐵× with Q(𝜈) an imaginary quadratic field. As a result, the curve 𝑋 1 is canonical,
uniquely characterized up to isomorphism, and is so called the canonical model. We
give some indications of this result by example in the next section and more generally
in section 43.8.
43.2 ⊲ QM by discriminant 6
For concreteness, before embarking on our general treatment, we consider in this
section an illustrative example and one of special interest; it is well-studied and
beloved by quaternionic
practitioners, see Remark 43.2.21 for further reference.
−1, 3
Let 𝐵 = be the quaternion algebra of discriminant 6 studied in sections
Q
37.8–37.9. As in 37.8.12, we have a maximal order
1 + 𝑖 + 𝑗 + 𝑖𝑗
O = Z ⊕ Z𝑖 ⊕ Z 𝑗 ⊕ Z𝑘, 𝑘=
2
with 𝑘 2 − 𝑘 − 1 = 0, and an embedding
𝜄∞ : 𝐵 ↩→ M2 (R)
√
0 −1 3 0
√
𝑖, 𝑗 ↦→ ,
1 0 0 − 3
Let Γ1 = 𝜄∞ (O1 )/{±1} ≤ PSL2 (R) and 𝑋 1 = Γ1 \H2 . We computed a compact
Dirichlet fundamental domain ◊ for Γ1 in 37.9.4, with 𝜇(◊) = 2𝜋/3. Further, we saw
explicitly in 37.9.10 (and again by formula in Example 39.4.21) that Γ1 has signature
(0; 2, 2, 3, 3); that is, 𝑋 1 has topological genus 𝑔 = 0 and there are 4 cone points, two
points 𝑧 2 , 𝑧20 ∈ ◊ with stabilizer of order 2 and two 𝑧3 , 𝑧30 ∈ ◊ with order 3 stabilizer.
As in Chapter 40, to exhibit a model for 𝑋 1 we seek modular forms, indeed, we
now describe the full graded ring of (even weight) modular forms (43.1.8). We will
use the following essential proposition.
Proposition 43.2.1. The following statements hold.
(a) Let 𝑓 : H2 → C be a nonzero meromorphic modular form of weight 𝑘 for Γ1 ,
not identically zero. Then
∑︁ 1 𝑘
ord𝑧 ( 𝑓 ) = .
1 1 2
# StabΓ 1 (𝑧) 6
Γ 𝑧 ∈Γ \H
(b) We have
1, if 𝑘 = 0;
1
dimC 𝑀𝑘 (Γ ) = 0, if 𝑘 = 2; (43.2.2)
1 − 𝑘 + 2b𝑘/4c + 2b𝑘/3c,
if 𝑘 ≥ 4.
43.2. ⊲ QM BY DISCRIMINANT 6 787
Proof. See Theorem 43.9.4; for the purposes of this introduction, we provide a sketch
to tide the reader over. For (a), we argue just as in Proposition 40.3.4: we integrate
dlog 𝑓 = d 𝑓 / 𝑓 over the boundary of the fundamental domain ◊ and use the identifica-
tion of sides provided by rotation at their fixed points (elliptic vertices), reversing the
direction of the path so the contributions cancel, and we are left again to sum angles.
The details are requested in Exercise 43.7. For (b), we can get upper bounds on the
dimension using (a), but to provide lower bounds we need to exhibit modular forms,
and these are provided by the Riemann–Roch theorem. For example, for 𝑘 = 2, we
have dimC 𝑀2 (Γ1 ) = 𝑔 = 0 by (40.2.11).
𝑎 2 + 𝑎 20 𝑎 3 + 𝑎 30 𝑘
𝑎1 + + = . (43.2.4)
2 3 6
By part (b), we have dimC 𝑀𝑘 (Γ1 ) = 0 for 𝑘 < 0, and indeed, there are no solutions.
For 𝑘 = 0, there is a unique solution corresponding to the constant functions. For
𝑘 = 2, there are no solutions, as follows. Let 𝑓 (𝑧) ∈ 𝑀2 (Γ1 ). Let 𝛾3 be a generator
for the stabilizer at 𝑧 3 . Then
43.2.6. In weights 8, 10, we have products of forms seen previously. In weight 𝑘 = 12,
we find a third function ℎ12 ∈ 𝑀12 (Γ1 ) spanning together with 𝑓43 , 𝑔62 . Continuing in
788 CHAPTER 43. ABELIAN SURFACES WITH QM
this way, finally in weight 𝑘 = 24, we find 6 functions in a 5 dimensional space, and so
they must satisfy an equation 𝑟 ( 𝑓4 , 𝑔6 , ℎ12 ) ∈ C[ 𝑓4 , 𝑔6 , ℎ12 ], homogeneous of degree
24 if we give 𝑓4 , 𝑔6 , ℎ12 the weights 4, 6, 12.
Proposition 43.2.7. We have
C[ 𝑓4 , 𝑔6 , ℎ12 ]
𝑀 (Γ1 ) ' .
h𝑟 ( 𝑓4 , 𝑔6 , ℎ12 )i
Proof. The bound on the degrees of generators and relations in Theorem 43.9.6 makes
this proposition immediate. It is also possible to give a proof with bare hands: see
Exercise 43.9.
43.2.8. We do not have Eisenstein series available in this setting, but the notion of
taking averages 40.1.19 is still quite sensible:
we
find what are known as Poincaré
𝑎 𝑏
series. Recall 𝚥 (𝛾; 𝑧) = 𝑐𝑧 + 𝑑 for 𝛾 = ∈ SL2 (R). The square 𝚥 (𝛾; 𝑧) 2 is
𝑐 𝑑
well-defined on 𝛾 ∈ PSL2 (R).
For 𝑘 ∈ 2Z ≥2 , we define the Poincaré series
∑︁
𝑃 𝑘 (𝑧) = 𝚥 (𝛾; 𝑧) −𝑘 .
𝛾 ∈Γ1
is finite, and the Poincaré series converges (absolutely) by comparison [Kat85, §1,
Proposition 1]. The equality (43.2.9) follows from the cocycle relation (40.2.5).
Therefore 𝑃 𝑘 (𝑧) ∈ 𝑆 𝑘 (Γ), and in particular we may take 𝑓4 = 𝑃4 and 𝑔6 = 𝑃6 ; with a
bit more computation, one can also show that 𝑃43 , 𝑃62 , 𝑃12 are linearly independent, so
that we may take ℎ12 = 𝑃12 as well.
43.2.10. A convenient and meaningful normalization of the functions above is given
by Baba–Granath [BG2008, §3.1]. √
√ embeddings 𝑆 = Z[ −6] ↩→ O
First, there are exactly two (necessarily optimal)
by Example 30.7.5: we have √ # Cls O = 1 and 𝐾 ( −6) is ramified at 𝑝 = 2, 3 | 𝐷 = 6,
so 𝑚(𝑆, O; O× ) = ℎ(Z[ −6]) = 2. The fixed points of these two embeddings are
distinct points 𝑧6 , 𝑧60 ∈ ◊. Explicitly, we note that
𝜇 = 3𝑖 + 𝑖 𝑗 = −1 + 2𝑖 − 𝑗 + 2𝑘 (43.2.11)
√
We rescale 𝑔6 so that 𝑓43 (𝑧 6 )/𝑔62 (𝑧6 ) = −3, and we choose ℎ12 such that ℎ12 (𝑧6 ) =
ℎ12 (𝑧60 ) = 0, and rescale so that
Γ1 (O) → P2
(43.2.14)
𝑧 ↦→ ( 𝑓43 (𝑧) : 𝑔62 (𝑧) : ℎ12 (𝑧))
𝑤2 𝑤3 𝑤6
𝑓4 − − +
𝑔6 + − −
ℎ12 − + −
𝑗 : 𝑋 1 → P1
16𝑦 2 (43.2.17)
(𝑥 : 𝑦 : 𝑧) ↦→
9𝑥 2
generically 4-to-1. The map 𝑗 can fruitfully be thought of as an analogue of the classical
elliptic 𝑗-invariant, mindful of the above technicalities: it parametrizes principally
polarized complex abelian surfaces that can be equipped with a QM structure.
790 CHAPTER 43. ABELIAN SURFACES WITH QM
The Igusa invariants 43.3.5 of 𝐴 𝑗 where 𝑗 = 𝑗 (𝜏) are given by [BG2008, Proposi-
tion 3.6]
(𝐼2 : 𝐼4 : 𝐼6 : 𝐼10 )
= (12( 𝑗 + 1) : 6( 𝑗 2 + 𝑗 + 1) : 4( 𝑗 3 − 2 𝑗 2 + 1) : 𝑗 3 ) (43.2.18)
∈ P(2, 4, 6, 10).
There exists a genus 2 curve with these Igusa invariants if and only if 𝑗 = 0, −16/27
or the Hilbert symbol
(−6 𝑗, −2(27 𝑗 + 16))Q = 1
is trivial.
Example 43.2.19. The two points with 𝑗 = 0, ∞ are exactly those points which
√ are
not Jacobians of genus 2 curves: these correspond to points with CM by Z[ −1] and
Z[𝜔], and these abelian surfaces are the squares of the corresponding CM elliptic
curves (with the product polarization). Elkies [Elk98, §3] computes equations and
further CM points for discriminant 6.
Example 43.2.20. The case 𝑗 = −16/27 corresponds to a CM point with discriminant
𝐷 = −24 [BG2008, §3.3]: it is the Jacobian of the curve
√ √ √ √
𝑦 2 = (1 + 2)𝑥 6 − 3(7 − 3 2)𝑥 4 − 3(7 + 3 2)𝑥 2 + (1 − 2)
√
isomorphic to the product of the two elliptic curves with CM by Z[ −6] (but not with
the product polarization).
Remark 43.2.21. For further reading to connect some of the dots above, see the article
by Baba–Granath [BG2008], refining the work by Hashimoto–Murabayashi [HM95,
Theorem 1.3] who give an explicit family of genus 2 curves whose Jacobians have QM
by O.
with 𝑎𝑑 − 𝑏𝑐, 𝑒 ∈ 𝐹 × . After such a change of variable, we may suppose without loss
of generality that deg 𝑓 = 6.
Example 43.3.4. Let 𝑋 al be the base change of 𝑋 to 𝐹 al . The automorphism group
Aut(𝑋 al ) is a finite group containing the hyperelliptic involution (𝑥, 𝑦) ↦→ (𝑥, −𝑦).
The possibilities for this group were classified by Bolza [Bol1887, p. 70]: when
char 𝐹 ≠ 2, 3, 5, the group Aut(𝑋 al ) is isomorphic to one of the groups
𝐶2 , 𝑉4 , 𝐷 8 , 𝐶10 , 𝐷 12 , 2𝐷 12 , 𝑆e4
of orders 2, 4, 8, 10, 12, 24, 48. A generic genus 2 curve over 𝐹 al has Aut(𝑋 al ) ' 𝐶2 .
43.3.5. We now seek invariants of the curve defined in terms of a model to classify
isomorphism classes. We factor
6
Ö
𝑓 (𝑥) = 𝑐 (𝑥 − 𝑎 𝑖 )
𝑖=1
where each sum and product runs over the distinct expressions obtained by permuting
the index set {1, . . . , 6}; by Galois theory, we have 𝐼2 , 𝐼4 , 𝐼6 , 𝐼10 ∈ 𝐹. In particular, we
have Ö
𝐼10 = (4𝑐) 10 (𝑎 𝑖 − 𝑎 𝑗 ) 2 = disc(4 𝑓 ) ≠ 0
1≤𝑖< 𝑗 ≤6
Proposition 43.3.7. The genus 2 curves 𝑋 and 𝑋 0 over 𝐹 are isomorphic over 𝐹 al if
and only if
43.3.8. For arithmetic reasons (in particular to deal with problems in characteristic 2),
Igusa [Igu60, pp. 617ff] defined the invariants [Igu60, pp. 621–622]
𝐽2 := 𝐼2 /8,
𝐽4 := (4𝐽22 − 𝐼4 )/96,
𝐽6 := (8𝐽23 − 160𝐽2 𝐽4 − 𝐼6 )/576, (43.3.9)
𝐽8 := (𝐽2 𝐽6 − 𝐽42 )/4,
𝐽10 := 𝐼10 /4096,
now called the Igusa invariants, with (𝐽2 : 𝐽4 : 𝐽6 : 𝐽8 : 𝐽10 ) ∈ P(2, 4, 6, 8, 10) (𝐹 al ).
Visibly, the Igusa–Clebsch invariants determine the Igusa invariants and vice versa.
Remark 43.3.10. One can also take ratios of these invariants with the same weight and
define (three) absolute invariants analogous to the classical 𝑗-invariant of an elliptic
curve, following Cardona–Nart–Pujolas [CNP2005] and Cardona–Quer [CQ2005].
Example 43.3.11. The locus of genus 2 curves with given automorphism group (cf.
Example 43.3.4) can be described explicitly by the vanishing of polynomials in the
Igusa(–Clebsch) invariants. For example, the unique genus 2 curve up to isomorphism
over 𝐹 al with automorphism group 𝐶10 (when char 𝐹 ≠ 5) is the curve defined by the
equation 𝑦 2 = 𝑥(𝑥 5 − 1) with (𝐼2 : 𝐼4 : 𝐼6 : 𝐼10 ) = (0 : 0 : 0 : 1), with automorphism
group generated by (𝑥, 𝑦) ↦→ (𝜁5 𝑥, −𝜁53 𝑦), where 𝜁5 is a primitive fifth root of unity.
for 𝜎 ∈ Aut𝐹 (𝐹 al ). Given a point 𝑃 ∈ P(2, 4, 6, 10) (𝐹 sep ), we define its field of
moduli 𝑀 (𝑃) to be the fixed field of 𝐹 sep under the stabilizer of 𝑃 under this action.
Just as in the case of ordinary projective space, the field 𝑀 (𝑃) is the minimal field
over which 𝑃 is defined.
43.3.13. In this way, given a genus 2 curve, we have associated invariants of the curve
that determine it up to isomorphism over 𝐹 al . We may also ask the inverse problem:
given Igusa invariants (𝐽 𝑘 ) 𝑘 with 𝐽10 ≠ 0, find a genus 2 curve with the desired
invariants. This problem has been solved explicitly by work of Mestre [Mes91] and
Cardona–Quer [CQ2005].
We give a sketch of the generic case of curves whose only automorphism over 𝐹 al
is the hyperelliptic involution, due to Mestre [Mes91]: in brief, the field of moduli may
not be a field of definition for the desired genus 2 curve, but a quadratic extension will
43.4. COMPLEX ABELIAN VARIETIES 793
Let 𝐴 = 𝑉/Λ be a complex torus of dimension 𝑔. Then 𝑉 ' C𝑔 and Λ ' Z2𝑔 so
𝑉/Λ ' (R/Z) 2𝑔 as smooth real manifolds.
we have 𝑃2−1 Π = Ω 1 , and Ω = 𝑃2−1 𝑃1 ∈ GL𝑔 (C). Therefore every complex torus
is isomorphic to a torus of the form C𝑔 /(ΩZ𝑔 + Z𝑔 ) for some Ω ∈ GL𝑔 (C), called the
small period matrix.
β1
α1 α2 β2
Ω1 × 𝐻1 (𝑋, Z) → C
∫
(𝜔, 𝜐) ↦→ 𝜔
𝜐
be the Jacobian of 𝑋. Then Jac 𝑋 is a complex torus of dimension 𝑔. It has big period
matrix Π = 𝑃1 𝑃2 , where
∫ ∫
𝑃1 = 𝜔𝑗 , 𝑃2 = 𝜔𝑗 .
𝛼𝑖 𝑖, 𝑗 𝛽𝑖 𝑖, 𝑗
43.4. COMPLEX ABELIAN VARIETIES 795
By cutting open the Riemann surface along the given paths and applying Green’s
theorem, we verify that the big period matrix Π is indeed a Riemann matrix. Therefore
the Jacobian Jac(𝑋) is an abelian variety of genus 𝑔.
𝑉 ×𝑉 → R
(𝑥, 𝑦) ↦→ 𝐸 R (𝑖𝑥, 𝑦)
43.4.10. Let 𝐸 be a Riemann form for (𝑉, Λ). If we choose a Z-basis for Λ, we get a
period matrix Π and a matrix for 𝐸 which is a Riemann matrix (satisfying conditions
(i)–(ii) of Definition 43.4.5), and conversely.
Proposition 43.4.11. If 𝐸 is a Riemann form for (𝑉, Λ), then the map
𝐻: 𝑉 ×𝑉 → C
(43.4.12)
𝐻 (𝑥, 𝑦) = 𝐸 R (𝑖𝑥, 𝑦) + 𝑖𝐸 R (𝑥, 𝑦)
Proof. This proposition can be checked directly, a bit of linear algebra fun: see
Exercise 43.6.
𝐸 (𝑥 1 + 𝑖𝑥2 , 𝑦 1 + 𝑖𝑦 2 ) = 𝑥2 𝑦 1 − 𝑥 1 𝑦 2
𝐸 R (𝑖(𝑥1 + 𝑖𝑥2 ), 𝑦 1 + 𝑖𝑦 2 ) = 𝑥1 𝑦 1 − (−𝑥2 𝑦 2 ) = 𝑥1 𝑦 1 + 𝑥2 𝑦 2
𝐻 (𝑥, 𝑦) = 𝑥𝑦.
define a Riemann form 𝐸, its associated (symmetric, positive definite) real part, and
its associated (positive definite) Hermitian form 𝐻.
796 CHAPTER 43. ABELIAN SURFACES WITH QM
Definition 43.4.14. A complex torus 𝐴 = 𝑉/Λ equipped with a Riemann form is said
to be polarized.
A homomorphism of polarized complex tori is a homomorphism 𝜙 : 𝐴 → 𝐴 0 of
complex tori that respects the polarizations in the sense that the diagram
Λ×Λ
𝐸 /; Z
( 𝜙, 𝜙)
𝐸0
Λ0 × Λ0
commutes.
By Theorem 43.4.6 and Proposition 43.4.11, a polarized complex torus is an
abelian variety, and accordingly we call it a polarized abelian variety.
We now seek to classify the possibilities for Riemann forms.
43.4.15. There is a normal form for alternating matrices, analogous to the Smith
normal form of an integer matrix, called the Frobenius normal form. Let 𝑀 be a free
Z-module of rank 2𝑔 equipped with an alternating form 𝐸 : 𝑀 × 𝑀 → Z. Then there
exists a basis of 𝑀 such that the matrix of 𝐸 in this basis is
0 𝐷
[𝐸] =
−𝐷 0
Proof. Compute the period matrix with respect to a basis in which the Riemann form
is in Frobenius normal form.
Example 43.4.18. Let 𝐴1 = 𝑉1 /Λ1 and 𝐴2 = 𝑉2 /Λ2 be two polarized abelian varieties,
with Riemann forms 𝐸 1 , 𝐸 2 . Let 𝐴 = 𝐴1 × 𝐴2 = 𝑉/Λ, where 𝑉 = 𝑉1 ⊕ 𝑉2 and
Λ = Λ1 ⊕ Λ2 ⊆ 𝑉1 ⊕ 𝑉2 = 𝑉. Then 𝐴 can be equipped with the product polarization
𝐸 = 𝐸 1 𝐸 2 , defined by
𝐸 (𝑥 1 + 𝑥2 , 𝑦 1 + 𝑦 2 ) = 𝐸 1 (𝑥 1 , 𝑦 1 ) + 𝐸 2 (𝑥2 , 𝑦 2 ).
𝑉∗ × 𝑉 → R
( 𝑓 , 𝑥) ↦→ Im 𝑓 (𝑥)
is nondegenerate, so
Λ∗ := { 𝑓 ∈ 𝑉 ∗ : Im 𝑓 (Λ) ⊆ Z}
is a lattice in 𝑉 ∗ , called the dual lattice of Λ, and the quotient 𝐴∨ := 𝑉 ∗ /Λ∗ is a
complex torus. Double antiduality and nondegeneracy gives a canonical identification
(𝑉 ∗ ) ∗ 𝑉, giving a canonical identification ( 𝐴∨ ) ∨ 𝐴.
Now suppose 𝐴 is polarized with 𝐸 a Riemann form for (𝑉, Λ), and let 𝐻 be the
associated Hermitian form (43.4.12). Then double duality induces a Riemann form
𝐸 ∗ on (𝑉 ∗ , Λ∗ ), so 𝐴∨ is a polarized abelian variety. We have a C-linear map
𝜆 : 𝑉 → 𝑉∗
(43.4.20)
𝑥 ↦→ 𝐻 (𝑥, −)
with the property that 𝜆(Λ) ⊆ Λ∗ . Since the form is nondegenerate, the induced
homomorphism 𝜆 : 𝐴 → 𝐴∨ is an isogeny of polarized abelian varieties. The degree
of the isogeny 𝜆 is equal to the product 𝑑1 · · · 𝑑 𝑔 of the elementary divisors of 𝐸, so in
particular if 𝐸 is principal then 𝜆 is an isomorphism of principally polarized abelian
varieties.
43.4.21. Let 𝐴 = 𝑉/Λ be a principally polarized complex abelian variety with Riemann
∼ 𝐴∨ be the isomorphism of principally polarized abelian varieties
form 𝐸. Let 𝜙 : 𝐴 −→
induced by (43.4.20). Then we define the Rosati involution associated to 𝐸 (or 𝜙) by
†
: End( 𝐴) → End( 𝐴)
(43.4.22)
𝛼 ↦→ 𝛼† = 𝜙−1 𝛼∨ 𝜙
for all 𝑥, 𝑦 ∈ Λ.
The function 𝜒10 is somewhat analogous to the classical function Δ, in the following
sense.
Lemma 43.5.5. Let 𝜏 ∈ ℌ2 . Then 𝜒10 (𝜏) = 0 if and only if 𝐴 𝜏 is decomposable (as a
principally polarized abelian variety).
In other words, the vanishing locus of 𝜒10 is precisely where case (ii) of Theorem
43.5.1 holds and the abelian surface is not isomorphic to the Jacobian of a genus 2
curve (as a principally polarized abelian surface).
Remark 43.5.6. More generally, a (classical) Siegel modular form of weight 𝑘 ∈ 2Z
for the group Γ = Sp4 (Z) is a holomorphic function 𝑓 : ℌ2 → C such that
for all 𝛾 ∈ Sp4 (Z) and 𝜏 ∈ ℌ2 . (By the Koecher principle, such a function is
automatically holomorphic at infinity in a suitable sense, and so the conditions at
cusps 40.2.12 for classical modular forms do not arise here.)
Let 𝑀𝑘 (Γ) be the C-vector space of Siegel modular forms of weight 𝑘; then 𝑀𝑘 (Γ)
is finite-dimensional, 𝑀𝑘É (Γ) = {0} for 𝑘 < 0, and 𝑀0 (Γ) = C consists of constant
functions. Let 𝑀 (Γ) = 𝑘 ∈2Z≥0 𝑀 𝑘 (Γ); then 𝑀 (Γ) has the structure of a graded
C-algebra under pointwise multiplication of functions. Igusa proved that
in analogy with Theorem 40.3.11. Extending the analogy, Igusa also proved that
𝜓4 , 𝜓6 , −4𝜒10 , 12𝜒12 have integer Fourier coefficients with content 1.
43.5.7. The Igusa–Clebsch invariants 43.3.5 can be expressed in terms of the functions
above. The precise relationship was worked out by Igusa [Igu60, p. 620]: we have
𝜒12
𝐼2 = −23 31
𝜒10
𝐼4 = 22 𝜓4
23 𝜓4 𝜒12
𝐼6 = − 𝜓6 − 25
3 𝜒10
𝐼10 = −214 𝜒10
The functions 𝐼4 , 𝐼6 , 𝜒10 , 𝜒12 are holomorphic, but 𝐼2 is meromorphic. In other words,
if 𝑋 is a complex genus 2 curve with Jac 𝑋 = 𝐴 𝜏 for 𝜏 ∈ ℌ2 , then the algebraic
invariants of 𝑋 can be computed in terms of the values of these functions evaluated
at 𝜏. This description is again analogous to the case of elliptic curves (cf. Remark
40.3.10).
𝛼† = 𝜇−1 𝛼𝜇 (43.6.3)
for all 𝛼 ∈ End( 𝐴). The map † defines a Q-antiautomorphism of End( 𝐴), so by the
Skolem–Noether theorem, we must have 𝜇 ∈ End( 𝐴) × .
From now on, let O be a maximal order in 𝐵. (One can relax this hypothesis with
some additional technical complications, but there is enough to wrangle with here
already!) In light of 43.6.2 we make the following definition (cf. Rotger [Rot2004,
§3]).
Definition 43.6.4. A polarization on O is an element 𝜇 ∈ O such that 𝜇2 ∈ Z<0 ; a
polarization is principal if 𝜇2 + 𝐷 = 0.
An isomorphism of polarized orders (O, 𝜇) ' (O0, 𝜇 0) is an isomorphism of orders
∼ O0 such that 𝜙(𝜇) = 𝜇 0.
𝜙 : O −→
802 CHAPTER 43. ABELIAN SURFACES WITH QM
Proof. We may check the desired equality locally. If 𝑝 - 𝐷, then 𝜇 ∈ O×𝑝 and
trd(𝜇O 𝑝 ) = trd(O 𝑝 ) = Z 𝑝 . Otherwise, if 𝑝 | 𝐷, then 𝜇 generates the normalizer
group 𝑁 𝐵×𝑝 (O 𝑝 )/(Q×𝑝 O×𝑝 ) by Exercise 23.4, and trd(𝜇O 𝑝 ) ⊆ 𝑝Z 𝑝 as desired.
For a principally polarized order (O, 𝜇), we define the positive involution
∗
:𝐵→𝐵
(43.6.8)
𝛼∗ = 𝜇−1 𝛼𝜇
𝐵
𝜄 / End( 𝐴)Q
∗ † (43.6.10)
𝐵
𝜄 / End( 𝐴)Q
Writing out (43.6.10), for a QM structure 𝜄 we require that 𝜄(𝛼) † = 𝜄(𝛼∗ ) for all
𝛼 ∈ 𝐵.
polarized abelian surfaces (respecting the polarization) that also respects 𝜄, 𝜄0 in the
sense that the diagram
𝜄0 /
𝐵 End( 𝐴 0)Q
𝜄
𝜙∗
#
End( 𝐴)Q
commutes; an isogeny is a surjective homomorphism with finite kernel.
QM abelian surfaces can be constructed as follows.
43.6.12. Extend 𝜄∞ to a map 𝜄∞ : 𝐵 ↩→ 𝐵C ' M2 (C). Let 𝜏 ∈ H2 . Let
𝜏
Λ 𝜏 := 𝜄∞ (O) ⊂ C2 .
1
This may look worse, but now we use positivity of ∗ applied to 𝑥𝛿:
trd (𝑥𝛿) (𝑥𝛿) ∗ = trd(𝑥𝛿𝜇−1 𝛿𝑥𝜇) = trd(𝜇𝑥𝛿𝜇−1 𝛿𝑥) > 0.
𝜏 𝜏
It follows that 𝐸 𝑖𝑥 ,𝑥 is always either positive definite or negative definite
1 1
(depending on the sign of nrd(𝛿)).
Nm(𝜇)
det(trd((𝜄∞ (𝜇)/𝐷)𝑥𝑖 𝑥 𝑗 ))𝑖, 𝑗 = det(trd(𝑥 𝑖 𝑥 𝑗 ))𝑖, 𝑗
𝐷4
nrd(𝜇) 2 1
= (discrd O) 2 = 2 𝐷 2 = 1.
𝐷4 𝐷
Lemma 43.6.23. The homomorphism 𝜄 𝜏 : O → End( 𝐴 𝜏 )Q satisfies the compatibility
(43.6.10), and 𝐸 𝜏 is the unique compatible principal polarization on 𝐴 𝜏 compatible
with 𝜄 𝜏 .
as desired.
To conclude, any other polarization corresponds to another positive involution of
the form 𝛼 ↦→ 𝜈 −1 𝛼𝜈 as in 43.6.2; scaling, we may take 𝜈 ∈ O such that trd(𝜈O) ⊂
𝐷O♯ . In the proof of Lemma 43.6.22, we see that the degree of the polarization is
equal to nrd(𝜈) 2 /𝐷 2 , so it is principal if and only if nrd(𝜈) = 𝐷, i.e., 𝜈 2 + 𝐷 = 0. But
then by compatibility trd(𝜄∞ (𝜇)𝑥𝑦) = trd(𝜄∞ (𝜈)𝑥𝑦) for all 𝑥, 𝑦 ∈ O, which implies
𝜇 = 𝜈.
806 CHAPTER 43. ABELIAN SURFACES WITH QM
Remark 43.6.27. Lemma 43.6.23 shows that one could be more relaxed in the definition
of QM abelian surface in the following sense. Let 𝐴 be a (not yet polarized) complex
abelian surface, and let 𝜄 : O ↩→ End( 𝐴) be a ring homomorphism. Then there is a
unique principal polarization on 𝐴 such that the induced Rosati involution is compatible
with 𝜇 in the sense of (43.6.10).
Proof of Main Theorem 43.6.14. By Lemmas 43.6.16, 43.6.22, and 43.6.23, the asso-
ciation 𝜏 ↦→ ( 𝐴 𝜏 , 𝜄 𝜏 ) yields a principally polarized complex abelian surface with QM
by (O, 𝜇). By Proposition 43.6.28, the map is surjective.
𝑎 𝑏
Next, we show the map is well-defined up to the action of Γ. Let 𝛾 = ∈Γ
𝑐 𝑑
0
and 𝜏 = 𝛾𝜏. Then
𝛾𝜏 −1 𝑎𝜏 + 𝑏
Λ 𝜏0 = 𝜄∞ (O) = 𝜄∞ (O) (𝑐𝜏 + 𝑑)
1 𝑐𝜏 + 𝑑
(43.6.29)
−1 𝜏 −1
= (𝑐𝜏 + 𝑑) 𝜄∞ (O)𝛾 = (𝑐𝜏 + 𝑑) Λ 𝜏 .
1
43.7.1. By Example 28.6.5, there exists 𝜖 ∈ O× such that nrd(𝜖) = −1. Then 𝜖 2 ∈ O1
and 𝜖 normalizes O1 , so the action of 𝜖 (as in (33.3.12)) defines an anti-holomophic
involution on 𝑋 (Γ) that is independent of the choice of 𝜖: this gives the natural action
of complex conjugation on 𝑋 (Γ).
With respect to this real structure, and in view of Main Theorem 43.6.14, we may
ask if there are any principally polarized abelian surfaces with QM by (O, 𝜇) with
and the QM defined over R. When 𝐵 ' M2 (Q), then the element
both the surface
1 0
𝜖 = acts by complex conjugation, and the real points are those points on
0 −1
the imaginary axis (the points with real 𝑗-invariant). More generally, the answer is
provided by the following special case of a theorem of Shimura [Shi75, Theorem 0].
Theorem 43.8.1 (Shimura [Shi67, Main Theorem I (3.2)]). There exists a projective,
nonsingular curve 𝑋Q1 defined over Q and a holomorphic map
𝜑 : H2 → 𝑋Q1 (C)
43.8.2. The curve 𝑋Q1 is made canonical (unique up to isomorphism) according to the
field of definition of CM points (see 43.7.4).
Let 𝑧 ∈ H2 be a CM point with 𝑆 of discriminant 𝐷. Let 𝐻 ⊇ 𝐾 be the ring class
extension 𝐻 ⊇ 𝐾 with Gal(𝐻 | 𝐾) ' Pic(𝑆). Then 𝜙(𝑧) ∈ 𝑋Q1 (𝐻). Moreover, there is
an explicit law, known as the Shimura reciprocity law, which describes the action of
Gal(𝐻 | 𝐾) on them: to a class [𝔠] ∈ Pic 𝑆, we have
𝜄𝐾 (𝔠)O = 𝜉O (43.8.3)
for some 𝜉 ∈ O, and if 𝜎 = Frob𝔠 ∈ Gal(𝐻 | 𝐾) under the Artin isomorphism, then
For more detail, see Shimura [Shi67, p. 59]; and for an explicit, algorithmic version,
see Voight [Voi2006].
43.8.5. Before continuing, we link back to the idelic, double-coset point of view,
motivated in section 38.6 and given in general in section 38.7.
The difference between a lattice in R2 and a lattice in C is an identification of R2
with C, i.e., an injective R-algebra homomorphism 𝜓 : C → EndR (R2 ); we call 𝜓 a
complex structure.
There is a bijection between the set of complex structures and the set C r R = H2±
as follows. A complex structure 𝜓 : C → EndR (R2 ), by R-linearity is equivalently
given by the matrix 𝜓(𝑖) ∈ GL2(R) satisfying
𝜓(𝑖) 2 = −1. By rational canonical form,
0 1
every such matrix is similar to , i.e., there exists 𝛽 ∈ GL2 (R) such that
−1 0
−1 0 1
𝜓(𝑖) = 𝛽 𝛽,
−1 0
0 1
and 𝛽 is well-defined up to the centralizer of ; this matrix acts by fixing
−1 0
2
𝑖 ∈ H , and it follows that this centralizer is precisely R× SO(2). Finally, we have
GL2 (R)/R× SO(2) −→ ∼ H2± under 𝛽 ↦→ 𝛽𝑖.
810 CHAPTER 43. ABELIAN SURFACES WITH QM
In this way,
𝑋 1 = Γ1 \H2 ↔ O1 \H2±
↔ O1 \(GL2 (R)/R× SO(2)) (43.8.6)
↔ 𝐵× \( 𝐵
b× /O
b× × GL2 (R)/R× SO(2)).
F (−) / Hom(−, 𝑍)
∃!
%
Hom(−, 𝑋)
The above formulas can be proven in a different and straightforward way in the
language of stacky curves. This description gives the following further information on
𝑀 (Γ).
812 CHAPTER 43. ABELIAN SURFACES WITH QM
Exercises
1. Let 𝑔 = 1 and consider a period matrix Π = 𝜔1 𝜔2 with 𝜔1 , 𝜔2 ∈ C. Let
0 1
𝐸 = . Show that in Definition 43.4.5 for 𝐸 that the condition (i) is
−1 0
automatic and condition (ii) is equivalent to Im(𝜔2 /𝜔1 ) > 0.
2. Let Π ∈ Mat
𝑔×2𝑔 (C). Show that Π is a period matrix for a complex torus if and
Π
only if , the matrix obtained by vertically stacking Π on top of its complex
Π
conjugate Π, is nonsingular.
3. Let 𝐸 : Λ × Λ → Z be a Z-bilinear form that satisfies conditions (i) and (ii)
of Definition 43.4.9 (so a Riemann form but without the condition that 𝐸 is
alternating). Show that 𝐸 R (and 𝐸) are alternating.
4. Show that the symmetry (43.6.19) implies the equality
𝜏 𝜏 𝜏 𝜏
𝐸 𝑖𝑥 , 𝑖𝑦 =𝐸 𝑥 ,𝑦
1 1 1 1
[Hint: break up the sum by congruence class according to the floor and
then use geometric series.]
(b) Let 𝑚 2 , 𝑚 3 ∈ Z ≥0 . For 𝑘 ∈ 2Z ≥0 , define
1, if 𝑘 = 0;
𝑐 𝑘 = 𝑔, if 𝑘 = 2;
(𝑘 − 1) (𝑔 − 1) + 𝑚 2 b𝑘/4c + 𝑚 3 b𝑘/3c,
if 𝑘 ≥ 4.
Show that
∑︁ 1 + 𝑔𝑡 2 + 𝑎 4 𝑡 4 + 𝑎 6 𝑡 6 + 𝑎 4 𝑡 8 + 𝑔𝑡 10 + 𝑡 12
𝑐𝑘 𝑡 𝑘 =
𝑘 ∈2Z≥0
(1 − 𝑡 4 ) (1 − 𝑡 6 )
where
𝑎 4 = 3𝑔 + 𝑚 2 + 𝑚 3 − 4
𝑎 6 = 4𝑔 + 𝑚 2 + 2𝑚 3 − 6.
and dimC 𝑀𝑘 = dimC 𝑀𝑘0 for all 𝑘. [Hint: use Exercise 43.8.]
(e) Conclude that the subring of 𝑀 (Γ1 ) generated by 𝑓4 , 𝑔6 , ℎ12 is equal to
𝑀 (Γ1 ). [Hint: Suppose that equality does not hold, and consider the
minimal degree with a new generator; by dimensions, there must be a new
relation, but argue that this relation must be among 𝑓4 , 𝑔6 , ℎ12 , contradict-
ing (b).]
Symbol Definition List
815
816 CHAPTER 43. ABELIAN SURFACES WITH QM
b= 𝐸
𝐸 bb b
𝑆, O
elements conjugating a fixed embedding into adelic optimal em-
beddings 𝑆b ↩→ O,
b 536
𝐸𝑘 , 𝐺 𝑘 Eisenstein series, 731
b O;
Emb( 𝑆, bbΓ) Γ-conjugacy classes of optimal adelic embeddings 𝑆b ↩→ O
set of b b
of the quadratic order 𝑆 into the quaternion order O, 536
b b
Emb𝑅 (𝑆, O) set of optimal embeddings 𝑆 ↩→ O of the quadratic order 𝑆 into
the quaternion order O, 534
E𝑛 Euclidean 𝑛-space, 592
Emb(𝑆, O; Γ) set of Γ-conjugacy classes of optimal embeddings 𝑆 ↩→ O, 534
𝐹 field, 21
𝐹 (1) product norm 1 ideles, 452
𝐹 al algebraic closure of the field 𝐹, 21
𝑓∨ Fourier transform of Schwartz function 𝑓 , 492
𝔣(O) conductor of the quaternion 𝑅-order O, 385
𝐹>×Ω 0 elements of 𝐹 positive at the real places 𝑣 ∈ Ω, 225
𝐹 sep separable closure of 𝐹, 91
Gal(𝐾 | 𝐹) Galois group of the Galois extension of fields 𝐾 ⊇ 𝐹, 24
Γ0 (𝑁), Γ1 (𝑁) standard congruence subgroups of level 𝑁, 644
Γ(𝑁) full congruence subgroup of level 𝑁, 644
ΓR (𝑠), ΓC (𝑠) standard archimedean Γ-factors, 436
Gen O genus of an 𝑅-order O, 274
Gen 𝑄 genus of a quadratic module 𝑄, 146
𝐺 𝑁 (O) idelic reduced norm group attached to the normalizer of O, 474
𝐺 (O) idelic reduced norm group attached to O, 461
GO(𝑄) (𝐹) general orthogonal group of 𝑄, 49
Gor(O) Gorenstein saturation of the quaternion 𝑅-order O, 385
G linear algebraic group, 709
GSO(𝑄) (𝐹) general special orthogonal group of a quadratic form 𝑄, 55
H (real) Hamiltonians, 23
H0 pure Hamiltonians, 28
H1 unit Hamiltonians, 27
H2 upper half-plane, 594
H2± union of upper and lower half-planes, 595
H2∗ completed upper half-plane, 596
H3 upper half-space, 649
H product of hyperbolic planes and spaces, 704
H𝑛 hyperbolic 𝑛-space, 656
𝐻 hyperbolic plane, 71
ℎ(𝑆) class number of 𝑆, 536
𝐼 lattice (over a domain, in an algebra), 152
𝐼2 , 𝐼4 , 𝐼6 , 𝐼10 Igusa invariants, 800
𝐼 (𝛼) axis of 𝛼 ∈ H1 , 29
818 CHAPTER 43. ABELIAN SURFACES WITH QM
[AB2008] Peter Abramenko and Kenneth S. Brown, Buildings: theory and applications,
Grad. Texts in Math., vol. 248, Springer, New York, 2008. 23.5.17
[AA66] J. F. Adams and M. F. Atiyah, 𝐾-theory and the Hopf invariant, Quart. J. Math.
Oxford (2), 17 (1966), 31–38. 1.2
[Alb34] A. A. Albert, Integral domains of rational generalized quaternion algebras,
Bull. Amer. Math. Soc. 40 (1934), 164–176. 1.3, 14.2.8
[Alb39] A. Adrian Albert, Structure of algebras, AMS Colloquium Publications, vol. 24,
Amer. Math. Soc., Providence, RI, 1939. 1.3, 8.2.9
[Alb72] A. A. Albert, Tensor products of quaternion algebras, Proc. Amer. Math. Soc. 35
(1972), no. 1, 65–66. 8.2.9
[AH32] A. Adrian Albert and Helmut Hasse, A determination of all normal division
algebras over an algebraic number field, Trans. Amer. Math. Soc. 34 (1932), no.
3, 722–726. 1.3, 14.6.6
[Alp93] Roger C. Alperin, 𝑃𝑆𝐿 2 (Z) = Z2 ∗ Z3 , Amer. Math. Monthly 100 (1993), no. 4,
385–386. 35.1.20
[AB2004] Montserrat Alsina and Pilar Bayer, Quaternion orders, quadratic forms, and
Shimura curves, CRM Monograph Series, vol. 22, American Math. Soc., Provi-
dence, RI, 2004. iv (document), 1.4, 14.2.13, 37.9.13
[Alt89] Simon L. Altmann, Hamilton, Rodrigues, and the quaternion scandal, Math.
Magazine 62 (1989), no. 5, 291–308. 1.1
[AO2005] Simon L. Altmann and Eduardo L. Ortiz, eds., Mathematics and social utopias in
France: Olinde Rodrigues and his times, Amer. Math. Society, Providence, RI,
2005. 1.1
[ART79] S. A. Amitsur, L. H. Rowen, and J. P. Tignol, Division algebras of degree 4 and
8 with involution, Israel J. Math. 33 (1979), 133–148. 8.3.6
[Apo76] Tom M. Apostol, Introduction to analytic number theory, Undergraduate Texts in
Math., Springer-Verlag, New York-Heidelberg, 1976. 14.2
[Apo90] Tom M. Apostol, Modular functions and Dirichlet series in number theory,
Springer-Verlag, New York, 1990. 35, 40.1
[And2005] James W. Anderson, Hyperbolic geometry, 2nd. ed., Springer-Verlag, London,
2005. 33.1
[A-C2012] Luis Arenas-Carmona, Maximal selectivity for orders in fields, J. Number Theory
132 (2012), no. 12, 2748–2755. 31.7.7
[A-C2013] Luis Arenas-Carmona, Eichler orders, trees and representation fields, Int. J.
Number Theory 9 (2013), no. 7, 1725–1741. 31.7.7
[Art26] Emil Artin, Zur Theorie der hyperkomplexen Zahlen, Abh. Math. Sem. Hambur-
gischen Univ. 5 (1926), 251–260. 7.3.11
823
824 BIBLIOGRAPHY
[AT2008] Alexander Arhangel’skii and Mikhail Tkachenko, Topological groups and related
structures, Atlantist Studies in Math., vol. 1, World Scientific Publishing Co.,
Hackensack, NJ, 2008. 34.1
[Arm88] M. A. Armstrong, Groups and symmetry, Undergrad. Texts in Math., Springer-
Verlag, New York, 1988. 11.5
[Art50] Emil Artin, The influence of J. H. M. Wedderburn on the development of modern
algebra, Bull. Amer. Math. Soc. 56 (1950), no. 1, 65–72. 1.2
[Asl96] Helmer Aslaksen, Quaternionic determinants, Math. Intelligencer 18 (1996), no.
3, Summer 1996, 57–65. 2.20
[AM69] Michael Francis Atiyah and Ian Grant Macdonald, Introduction to commutative
algebra, Reading, Addison-Wesley, 1969. 9.2, 9.5
[Ati67] M. F. Atiyah, 𝐾-theory, W. A. Benjamin, Inc., New York, 1967.
[ABB2014] Asher Auel, Marcello Bernardara, and Michele Bolognesi, Fibrations in complete
intersections of quadrics, Clifford algebras, derived categories, and rationality
problems, J. Math. Pures Appl. (9) 102 (2014), no. 1, 249–291. 22.2.14
[ABGV2006] Asher Auel, Eric Brussel, Skip Garibaldi, and Uzi Vishne, Open problems on
central simple algebras, Transform. Groups 16 (2011), no. 1, 219–264. 7.5.5
[Auel2011] Asher Auel, Clifford invariants of line bundle-valued quadratic forms, MPIM
preprint series, no. 33, January 2011. 22.2.14, 22.2.15
[Auel2015] Asher Auel, Surjectivity of the total Clifford invariant and Brauer dimension, J.
Algebra, 443 (2015), 395–421. 22.2.14
[AG60] Maurice Auslander and Oscar Goldman, Maximal orders, Trans. Amer. Math.
Soc. 97 (1960), 1–24. 15.17, 18.3
[BG2008] Srinath Baba and Håkan Granath, Genus 2 curves with quaternionic multiplica-
tion, Canad. J. Math. 60 (2008), no. 4, 734–757. 43.2.10, 43.2, 43.2.15, 43.2.16,
43.2.16, 43.2.20, 43.2.21
[Bae2002] John C. Baez, The octonions, Bull. Amer. Math. Soc. (N.S.) 39 (2002), 145–205;
Errata for “The octonions”, Bull. Amer. Math. Soc. (N.S.) 42 (2005), 213. 1.2
[Bae2005] John C. Baez, On quaternions and octonions: their geometry, arithmetic, and
symmetries by John H. Conway and Derek A. Smith, Bull. Amer. Math. Soc. 42
(2005), 229–243. 11.1
[Bak71] A. Baker, Linear forms in the logarithms of algebraic numbers, Mathematika 13
(1966), 204–216. 25.2.19
[Bal2007] Venkata Balaji Thiruvalloor Eesanaipaadi, Line-bundle-valued ternary quadratic
forms over schemes, J. Pure Appl. Algebra 208 (2007), 237–259. 22.5.13
[Bas62] Hyman Bass, Injective dimension in Noetherian rings, Trans. Amer. Math. Soc.
102 24.2.24
[Bas63] Hyman Bass, On the ubiquity of Gorenstein rings, Math. Z. 82 (1963), 8–28.
24.2.24
[BHV2008] Bachir Bekka, Pierre de la Harpe, and Alain Valette, Kazhdan’s Property (T),
Cambridge University Press, Cambridge, 2008. 29.3
[BCK93] F. Beukers, E. Calabi, and J. Kolk, Sums of generalized harmonic series and
volumes, Nieuw Arch. Wisk. 11 (1993), 217–224. 25.1
[Bea95] A. Beardon, The geometry of discrete groups, Grad. Texts in Math., vol. 91,
Springer-Verlag, New York, 1995. 33.1, 34.7, 37.3
[BP92] R. Benedetti and C. Petronio, Lectures on hyperbolic geometry, Universitext,
Springer, Berlin, 1992. 36.3.17
[BO2013] Grégory Berhuy and Frédérique Oggier, An introduction to central simple alge-
bras and their applications to wireless communication, Math. Surveys Monogr.,
BIBLIOGRAPHY 825
[Die48] Jean Dieudonné, Sur les groupes classiques, Actualitiés scientifique et indus-
trielles, no. 1040, Paris, Hermann, 1948. 3.5.3
[Die53] Jean Dieudonné, On the structure of unitary groups (II), Amer. J. Math. 75 (1953),
665–678. 3.5.3
[DK95] Adel Diek and R. Kantowski, Some Clifford algebra history, Clifford Algebras
and Spinor Structures, Math. Appl. 321 (1995), 3–12. 1.2
[DM89] Radoslav Dimitrić and Brendan Goldsmith, The mathematical tourist: Sir William
Rowan Hamilton, Math. Intelligencer 11 (1989), no. 2, 29–30. 1.1
[Dix77] Jacques Dixmier, 𝐶 ∗ -algebras, Translated from the French by Francis Jellett,
North-Holland Mathematical Library, vol. 15, North-Holland, Amsterdam, 1977.
3.2.11
[Don2011] Simon Donaldson, Riemann surfaces, Oxford University Press, Oxford, 2011.
33.8.6
[Dra83] P.K. Draxl, Skew fields, Cambridge University Press, Cambridge, 1983. 7.5
[DK68] Yu. A. Drozd and V. V. Kirichenko, Hereditary orders, Math. Inst. Acad. Sci.
Ukranian SSR, translated from Ukrainskii Mat. Zhurnal, 20 (1968), no. 2, 246–
248. 21.4
[DKR67] Yu. A. Drozd, V. V. Kirichenko, and A. V. Roĭter, On hereditary and Bass orders,
Math. USSR-Izvestija, translated from Izv. Akad. Nauk SSSR, 1 (1967), no. 6,
1357–1375. 21.4, 24.2.24, 24.4, 24.5
[DK94] Yurij A. Drozd and Vladimir V. Kirichenko, Finite dimensional algebras, with
an appendix by Vlastimil Dlab, translated by Vlastimil Dlab, Springer-Verlag,
Berlin, 1994. 7.2, 7.9
[EGM2008] Bas Edixhoven, Gerard van der Geer, and Ben Moonen, Modular forms, Modular
forms on Schiermonnikoog, Cambridge University Press, Cambridge, 2008. 1.4
[Eic36] M. Eichler, Untersuchungen in der Zahlentheorie der rationalen Quaternione-
nalgebren, J. Reine Angew. Math. 174 (1936), 129–159. 1.4, 23.4.22, 24.1.5,
24.5
[Eic37] Martin Eichler, Bestimmung der Idealklassenzahl in gewissen normalen einfachen
Algebren, J. Reine Angew. Math. 176 (1937), 192–202. 28.6.11
[Eic38a] M. Eichler, Allgemeine Kongruenzklasseneinteilungen der Ideale einfacher Alge-
bren über algebraischen Zahlkörper und ihre 𝐿-Reihen, J. Reine Angew. Math.
179 (1938), 227–251. 26.8, 28.6.11
[Eic38b] M. Eichler, Über die Idealklassenzahl total definiter Quaternionenalgebren,
Math. Z. 43 (1938), no. 1, 102–109. 26.1.14, 30.6.18
[Eic38c] M. Eichler, Über die Idealklassenzahl hyperkomplexer Systeme, Math. Z. 43
(1938), no. 1, 481–494. 28.6.11
[Eic53] Martin Eichler, Quadratische Formen und orthogonale Gruppen, Grundlehren
Math. Wiss., vol. 63, Springer, New York, 1974. 1.4, 4.2, 4.2.19, 22.5.13
[Eic55-56] Martin Eichler, Lectures on modular correspondences, Tata Institute of Funda-
mental Research, Bombay, 1955–56. 14.5.2, 25.3
[Eic56a] Martin Eichler, Zur Zahlentheorie der Quaternionen-Algebren, J. Reine Angew.
Math. 195 (1956), 127–151; Berichtigung zu der Arbeit “Zur Zahlentheorie der
Quaternionen-Algebren”, J. Reine Angew. Math. 197 (1957), 220. 1.4, 23.4.22,
23.1, 25.3, 26.1.14, 30.3.17, 30.6.18, 30.8.8, 30.9.12, 41.5.13
[Eic56b] Martin Eichler, Über die Darstellbarkeit von Mdoulformen durch Thetareihen,
J. Reine Angew. Math. 195 (1956), 156–171. 1.4
[Eic56c] Martin Eichler, Modular correspondences and their representations, J. Indian
Math. Soc. (N.S.) 20 (1956), 163–206. 23.4.22
BIBLIOGRAPHY 831
[Eic58] Martin Eichler, Quadratische Formen und Modulfunktionen, Acta Arith. 4 (1958),
217–239. 1.4
[Eic73] Martin Eichler, The basis problem for modular forms and the traces of the Hecke
operators, Modular functions of one variable, I (Proc. Internat. Summer School,
Univ. Antwerp, Antwerp, 1972), Lecture Notes in Math., vol. 320, Springer,
Berlin, 1973, 75–151; Correction to: “The basis problem for modular forms and
the traces of the Hecke operators”, Modular functions of one variable, IV (Proc.
Internat. Summer School, Univ. Antwerp, Antwerp, 1972), Lecture Notes in
Math., vol. 476, Springer, Berlin, 1975, 145–147. 1.4, 23.4.22, 30.3.17, 30.6.18,
40.4, 41.5.13
[Eic77] Martin Eichler, On theta functions of real algebraic number fields, Acta Arith. 33
(1977), 269–292. 41.5.13
[EM45] Samuel Eilenberg and Saunders Mac Lane, General theory of natural equiva-
lences, Transactions Amer. Math. Soc. 58 (1945), 231–294. 19.3.11
[Elk98] Noam D. Elkies, Shimura curve computations, Algorithmic number theory (Port-
land, OR, 1998), Lecture Notes in Comput. Sci., vol. 1423, Springer, Berlin,
1998, 1–47. 43.2.19
[EKM2008] Richard S. Elman, Nikita Karpenko, and Alexander Merkurjev, The algebraic
and geometric theory of quadratic forms, vol. 56, Amer. Math. Soc., 2008. 6.3
[EGM98] Jürgen Elstrodt, Fritz Grunewald, and Jens Mennicke, Groups acting on hyper-
bolic space: harmonic analysis and number theory, Springer-Verlag, Berlin,
1998. 36.1, 40.2.24
[EP94] David B.A. Epstein and Carlo Petronio, An exposition of Poincaré’s polyhedron
theorem, Enseign. Math. 40 (1994), 113–170. 37.6, 37.6.5
[Ewe82] John A. Ewell, A simple derivation of Jacobi’s four-square formula, Proc. Amer.
Math. Soc. 85 (1982), no. 3, 323–326. 1.4
[Fad65] D.K. Faddeev, Introduction to multiplicative theory of modules of integral repre-
sentations, Algebraic number theory and representations, Proc. Steklov Institute
of Math., vol. 80 (1965), American Math. Soc., Providence, RI, 1968, 164–210.
15.6, 16.8, 20.3.6
[FD93] Benson Farb and R. Keith Dennis, Noncommutative algebra, Grad. Texts in Math.,
vol. 144, Springer-Verlag, New York, 1993. 7.2
[Fen98] Della Dumbaugh Fenster, Leonard Eugene Dickson and his work in the Arith-
metics of Algebras, Arch. Hist. Exact Sci. 52 (1998), no. 2, 119–159. 1.2
[FS2007] Della D. Fenster and Joachim Schwermer, Beyond class field theory: Helmut
Hasse’s arithmetic in the theory of algebras in 1931, Arch. Hist. Exact Sci. 61
(2007), 425–456. 1.3, 14.6
[Fin89] Benjamin Fine, The algebraic theory of the Bianchi groups, Marcel Dekker, New
York, 1989. 35.5, 36.4.4, 36.8
[Fit2011] Robert W. Fitzgerald, Norm Euclidean quaternionic orders, INTEGERS 11
(2011), no. A58. 17.10
[Fol95] Gerald Folland, A course in abstract harmonic analysis, Studies in Advanced
Mathematics, CRC Press, Boca Raton, FL, 1995. 29.4
[For72] L. R. Ford, Automorphic functions, 2nd. ed., Chelsea, New York, 1972. 33.1,
37.2.9
[FGS2016] Adam Forsyth, Jacob Gurev, and Shakthi Shrima, Metacommutation as a group
action on the projective line over F 𝑝 , Proc. Amer. Math. Soc. 144 (2016), no. 11,
4583–4590.
[FK1890-2] Robert Fricke and Felix Klein, Vorlesungen über die Theorie der elliptischen
832 BIBLIOGRAPHY
Modulfuctionen, vols. 1 and 2, B.G. Teubner, Leipzig, 1890, 1892. 1.4, 1.4.4
[FK1897] Robert Fricke and Felix Klein, Vorlesungen über die Theorie der automorphen
Functionen. Erster Band: Die gruppentheoretischen Grundlagen, B.G. Teubner,
Leipzig, 1897. 1.4
[FK12] Robert Fricke and Felix Klein, Vorlesungen über die Theorie der automorphen
Functionen. Zweiter Band: Die funktionentheoretischen Ausführungen und die
Anwendungen, B.G. Teubner, Leipzig, 1912. 1.4
[Fro1878] F.G. Frobenius, Über lineare Substitutionen und bilineare Formen, J. Reine
Angew. Math. 84 (1878), 1–63. 1.2
[Frö75] A. Fröhlich, Locally free modules over arithmetic orders, J. Reine Angew. Math.
274/275 (1975), 112–124. 20.7, 20.7.23
[Frö73] A. Fröhlich, The Picard group of noncommutative rings, in particular of orders,
Trans. Amer. Math. Soc. 180 (1973), 1–45. 18.4.11
[FRU74] A. Fröhlich, I. Reiner and S. Ullom, Class groups and Picard groups of orders,
Proc. London Math. Soc. (3) 29 (1974), 405–434. 20.7.23
[FT91] A. Fröhlich and M.J. Taylor, Algebraic number theory, Cambridge Stud. Adv.
Math., vol. 27, Cambridge University Press, Cambridge, 1991. 9.3
[Fuj58] Genjiro Fujisaki, On the zeta-function of the simple algebra over the field of
rational numbers, J. Fac. Sci. Univ. Tokyo. Sect. I 7 (1958), 567–604. 27.6,
29.10.24
[FH91] William Fulton and Joe Harris, Representation theory: a first course, Graduate
Texts in Math., vol. 129, Springer-Verlag, New York, 1991. 2.17
[GV2018] Steven D. Galbraith and Frederik Vercauteren, Computational problems in su-
persingular elliptic curve isogenies, Quantum Inf. Process. 17 (2018), no. 10,
Art. 265, 22 pp. 1.5
[Gar2004] Skip Garibaldi, The characteristic polynomial and determinant are not ad hoc
constructions, American Math. Monthly 2004 (111), no. 9, 761–778. 7.8
[GS2010] Skip Garibaldi and David J. Saltman. Quaternion algebras with the same subfields,
Quadratic forms, linear algebraic groups, and cohomology, Springer, New York,
2010, 225–238. 8.2.6, 14.17
[Gau00] Carl Friedrich Gauss, Mutation des Raumes, in Carl Friedrich Gauss Werke, Band
8, König. Gesell. Wissen., Göttingen, 1900, 357–361. 1.1
[Gau86] Carl Friedrich Gauss, Disquisitiones arithmeticae, Translated and with a preface
by Arthur A. Clarke, revised by William C. Waterhouse, Cornelius Greither, and
A. W. Grootendorst, and with a preface by Waterhouse. Springer-Verlag, New
York, 1986. 25.1, 25.2.19, 30.1
[GMMR97] F. W. Gehring, C. Maclachlan, G. J. Martin, and A. W. Reid, Arithmeticity,
discreteness and volume, Trans. Amer. Math. Soc. 349 (1997), 3611–3643. 32.7
[Gel84] Stephen Gelbart, An elementary introduction to the Langlands program,
Bull. Amer. Math. Soc. (N.S.) 10 (1984), no. 2, 177–219. 1.4, 41.5.13
[GM2004] Stephen S. Gelbart and Stephen D. Miller, Riemann’s zeta function and beyond,
Bull. Amer. Math. Soc. (N.S.) 41 (2004), no. 1, 59–112. 26.3.15
[GG90] I.M. Gel’fand, M.I. Graev, and I.I. Pyatetskii-Shapiro, Representation theory
and automorphic functions, trans. K.A. Hirsch, Generalized Functions, vol. 6,
Academic Press, Boston, 1990. 37.7
[Gib1891] Josiah Willard Gibbs, On the rôle of quaternions in the algebra of vectors, Nature
(1891), vol. 43, 511–513. 1.1
[Gir83] P.R. Girard, The quaternion group and modern physics, Eur. J. Phys. 5 (1984),
25–32. 1.5
BIBLIOGRAPHY 833
[GT2010] Olivier Giraud and Koen Thas, Hearing the shape of drums: Physical and math-
ematical aspects of isospectrality, Rev. Modern Phys. 82 (2010), 2213–2255.
31.8
[God58a] Roger Godement, Les fonctions 𝜁 des algèbres simples, I, Séminaire N. Bourbaki,
1958–1960, exp. no. 171, 27–49. 29.10.24
[God58b] Roger Godement, Les fonctions 𝜁 des algèbres simples, II, Séminaire N. Bourbaki,
1958–1960, exp. no. 176, 109–128. 29.10.24
[God62] Roger Godement, Domaines fondamentaux des groupes arithmétiques, 1964
Séminaire Bourbaki, 1962/63, Fasc. 3, No. 257, Secrétariat mathématique, Paris.
38.5.6
[GJ72] Roger Godement and Hervé Jacquet, Zeta functions of simple algebras, Lecture
Notes in Math., vol. 260, Springer-Verlag, Berlin, 1972. 29.10.24
[Gol85] Dorian Goldfeld, Gauss’ class number problem for imaginary quadratic fields,
Bull. Amer. Math. Soc. 13 (1985), 23–37. 25.2.19
[Gor2000] Carolyn S. Gordon, Survey of isospectral manifolds, Handbook of differential
geometry, Vol. I, North-Holland, Amsterdam, 2000, 747–778. 31.8
[Gor2009] Carolyn S. Gordon, Sunada’s isospectrality technique: two decades later, Spectral
analysis in geometry and number theory, Contemp. Math., vol. 484, Amer. Math.
Soc., Providence, RI, 2009, 45–58. 31.8
[Gol71] Larry Joel Goldstein, Analytic number theory, Prentice-Hall, Inc., Englewood
Cliffs, N.J., 1971. 27.4
[Gor2012] Carolyn S. Gordon, Orbifolds and their spectra, Spectral geometry, Proc. Sympos.
Pure Math., vol. 84, Amer. Math. Soc., Providence, RI, 2012, 49–71. 34.8
[GWW92] Carolyn Gordon, David L. Webb, and Scott Wolpert, One cannot hear the shape
of a drum, Bull. Amer. Math. Soc. (N.S.) 27 (1992), no. 1, 134–138. 31.8
[Gor52] D. Gorenstein, An arithmetic theory of adjoint plane curves, Trans. Amer. Math.
Soc. 72 (1952), 414–436. 24.2.24
[Gou97] Fernando Q. Gouvêa, 𝑝-adic numbers: an introduction, Universitext, 2nd. ed.,
Springer-Verlag, Berlin, 1997. 12.1
[Gras1862] Hermann Grassmann, Die Ausdehnungslehre. Vollstaändig und in strenger Form
begründet [Extension theory], Berlin, Wiegand, 1862. 1.1
[Grav1882] Robert Perceval Graves, Life of Sir William Rowan Hamilton, Volume I, Dublin
University Press, 1882. 1.1
[Grav1885] Robert Perceval Graves, Life of Sir William Rowan Hamilton, Volume II, Dublin
University Press, 1885. 1.1, 1.1
[Grav1889] Robert Perceval Graves, Life of Sir William Rowan Hamilton, Volume III, Dublin
University Press, 1889. 1.1, 1.1
[Gray94] Jeremy Gray, On the history of the Riemann mapping theorem, Rendiconti del
Circolo Matematico di Palermo, Serie II, Supplemento 34 (1994), 47–94. 34.8.3
[Gray2000] Jeremy Gray, Linear differential equations and group theory from Riemann to
Poincaré, 2nd ed., Birkhäuser, Boston, 2000. 1.4
[Gray2013] Jeremy Gray, Henri Poincaré: a scientific biography, Princeton University Press,
Princeton, 2013. 1.4
[Gri28] Lois W. Griffiths, Generalized quaternion algebras and the theory of numbers,
Amer. J. Math. 50 (1928), no. 2, 303–314. 1.3
[Gro87] Benedict H. Gross, Heights and the special values of 𝐿-series, Number theory
(Montreal, Que., 1985), CMS Conf. Proc., vol. 7, Amer. Math. Soc., Providence,
RI, 1987, 115–187. 42.3.13
[Gro99] Benedict Gross, Algebraic modular forms, Israel J. Math. 113 (1999), 61–93.
834 BIBLIOGRAPHY
41.2.15
[GrLu2009] Benedict H. Gross and Mark W. Lucianovic, On cubic rings and quaternion rings,
J. Number Theory 129 (2009), no. 6, 1468–1478. 3.15, 22.4, 22.4.3, 22.5.13,
24.2
[Gro85] Emil Grosswald, Representations of integers as sums of squares, Springer-Verlag,
New York, 1985. 30.2
[GB2008] L. C. Grove and C. T. Benson, Finite reflection groups, 2nd ed., Grad. Texts in
Math, vol. 99, Springer-Verlag, New York, 1985. 11.5
[GrLe87] P.M. Gruber and C.G. Lekkerkerker, Geometry of numbers, 2nd ed., North-
Holland, New York, 1987. 17.5
[GQ2004] Xuejun Guo and Hourong Qin, An embedding theorem for Eichler orders, J.
Number Theory 107 (2004), no. 2, 207–214. 31.7, 31.7.7
[Grov2002] Larry C. Grove, Classical groups and geometric algebra, Grad. Studies in Math.,
vol. 39, Amer. Math. Soc., 2002. 4.2, 6.3, 28.8
[Hae84] A. Haefliger, Groupoides d’holonomie et classifiants, Structure transverse des
feuilletages, Toulouse 1982, Astérisque 116 (1984), 70–97. 34.8.15
[HM2006] Emmanuel Hallouin and Christian Maire, Cancellation in totally definite quater-
nion algebras, J. Reine Angew. Math. 595 (2006), 189–213. 20.7
[Ham1843] William R. Hamilton, On a new species of imaginary quantities connected with
a theory of quaternions, Proc. Royal Irish Acad. 2 (1843), 424–434. 1.1
[Ham1844] William R. Hamilton, On quaternions, or on a new system of imaginaries in
algebra: copy of a letter from Sir William R. Hamilton to John T. Graves, Esq.
on quaternions, Philosophical Magazine 25 (1844), 489–495. 1.2
[Ham1853] W. R. Hamilton, Lectures on quaternions, Cambridge, Cambridge University
Press, 1853. 1.1
[Ham1866] W. R. Hamilton, Elements of quaternions, Cambridge, Cambridge University
Press, 1866. 1.1.5, 1.1
[Ham1899] Hamilton’s quaternions, review of Elements of quaternions, Nature, August 24,
1899, 387. 1.1
[Ham67] W. R. Hamilton, The mathematical papers of Sir William Rowan Hamilton, Vol.
III: Algebra, eds. H. Halberstam and R. E. Ingram, Cambridge University Press,
1967. 1.1, 1.1
[Hanlon81] Phil Hanlon, Applications of quaternions to the study of imaginary quadratic
class groups, Ph.D. thesis, California Institute of Technology, 1981. 30.2
[Hankin80] Thomas L. Hankins, William Rowan Hamilton, Johns Hopkins University Press,
1980. 1.1, 1.1
[Hans2006] Andrew J. Hanson, Visualizing quaternions, Morgan Kaufmann, San Francisco,
2006. 1.5, 2.4.22
[Hap80] Dieter Happel, Klassifikationstheorie endlich-dimensionaler Algebren in der Zeit
von 1880 bis 1920, Enseign. Math. (2) 26 (1980), no. 1–2, 91–102. 1.2
[Har63a] Manabu Harada, Hereditary orders, Trans. Amer. Math Soc. 107 (1963), no. 2,
273–290. 20.3, 21.4
[Har63b] Manabu Harada, Structure of hereditary orders over local rings, J. Math. Osaka
City Univ. 14 (1963), 1–22. 21.4
[Har63c] Manabu Harada, Hereditary orders in generalized quaternions 𝐷 𝜏 , J. Math. Os-
aka City Univ. 14 (1963), 71–81. 21.4
[Har1881] A. S. Hardy, Elements of quaternions, Ginn, Heath, and Company, Boston, 1881.
v (document)
[HFK94] John C. Hart, George K. Francis, Louis H. Kauffman, Visualizing quaternion
BIBLIOGRAPHY 835
https://people.math.ethz.ch/~knus/papers/brandt04.pdf. 1.4
[Hun99] Craig Huneke, Hyman Bass and ubiquity: Gorenstein rings, Algebra, 𝐾-theory,
groups, and education (New York, 1997), Contemp. Math., vol. 243, Amer. Math.
Soc., Providence, RI, 1999, 55–78. 24.2.24
[Hur1898] Adolf Hurwitz, Über die Komposition der quadratischen Formen von beliebig
vielen Variablen, Nach. der köbig. Gesell. Wissen. Göttingen, Math.-Physik.
Klasse, 1898, 309–316. 1.3
[Hur1896] Adolf Hurwitz, Über die Zahlentheorie der Quaternionen, Nachrichten der
Gesellschaft der Wissenschaften zu Göttingen, 1896, 314–340. 1.3
[Hur19] Adolf Hurwitz, Vorlesungen über die Zahlentheorie der Quaternionen, Springer-
Verlag, Berlin, 1919. 1.3, 11.1
[Ibu82] Tomoyoshi Ibukiyama, On maximal orders of division quaternion algebras over
the rational number field with certain optimal embeddings, Nagoya Math. J. 88
(1982), 181–195. 15.5.7
[Igu60] Jun-Ichi Igusa, Arithmetic variety of moduli for genus two, Ann. Math. 72 (1960),
no. 3, 612–649. 43.3.5, 43.3, 43.3.8, 43.5.7
[Ive92] Birger Iversen, Hyperbolic geometry, London Math. Soc. Student Texts, vol. 25,
Cambridge University Press, Cambridge, 1992. 33.1, 36.1
[Iwa92] K. Iwasawa, Letter to J. Dieudonné, dated April 8, 1952, Zeta functions in
geometry, N. Kurokawa and T. Sunada, eds., Adv. Stud. in Pure Math. 21 (1992),
445–450. 29.10.24
[Jaci68] H. Jacobinski, Genera and decompositions of lattices over orders, Acta Math.
121 (1968), 1–29. 20.7.23
[Jaci71] H. Jacobinski, Two remarks about hereditary orders, Proc. Amer. Math. Soc 28
(1971), no. 1, 1–8. 21.2, 21.4
[Jacn74] Nathan Jacobson, Abraham Adrian Albert (1905–1972), Bull. Amer. Math.
Soc. 80 (1974) 1075–1100. 1.3
[Jacn2009] Nathan Jacobson, Finite-dimensional division algebras over fields, Springer-
Verlag, Berlin, 2009. 7.2, 8.2
[JL70] Hervé Jacquet and Robert Langlands, Automorphic forms on GL(2), Lecture
Notes in Math., vol. 114, Springer, New York, 1970. 1.4, 41.5.13
[JKS97] William C. Jagy, Irving Kaplansky, and Alexander Schiemann, There are 913
regular ternary forms, Mathematika 44 (1997), no. 2, 332–341. 25.4
[Jah2010] Majid Jahangiri, Generators of arithmetic quaternion groups and a Diophantine
problem, Int. J. Number Theory 6 (2010), no. 6, 1311–1328. 32.1
[Jan96] Gerald J. Janusz, Algebraic number fields, 2nd ed., Grad. Studies in Math., vol. 7,
Amer. Math. Soc., Providence, RI, 1996. 14.4, 14.4.4, 14.4.12, 27.5
[JS87] Gareth A. Jones and David Singerman, Complex functions: an algebraic and
geometric viewpoint, Cambridge University Press, 1987. 33.1, 34.7
[Jun97] Sungtae Jun, On the certain primitive orders, J. Korean Math. Soc. 34 (1997),
no. 4, 791–807. 24.5.7
[Jun08] Heinrich W.E. Jung, Darstellung der Funktionen eines algebraischen Körpers
zweier unabhängigen Veränderlichen 𝑥, 𝑦 in der Umgebung einer Stelle 𝑥 = 𝑎, 𝑦 =
𝑏, J. reine angew. Math. 133 (1908), 289–314. 35.4
[Kac66] Mark Kac, Can one hear the shape of a drum?, Amer. Math. Monthly 73 (1966),
no. 4, part II, 1–23. 31.8
[Kap69] Irving Kaplansky, Submodules of quaternion algebras, Proc. London Math. Soc.
(3) 19 (1969), 219–232 16.6, 16.6, 16.6, 16.8, 16.18, 19.5.8
[Kar2010] Max Karoubi, 𝐾-theory, an elementary introduction, Cohomology of groups and
BIBLIOGRAPHY 837
algebraic 𝐾-theory, Adv. Lect. Math. (ALM), vol. 12, Int. Press, Somerville, MA,
2010, 197–215. 12.4.4
[Kat85] Svetlana Katok, Closed geodesics, periods and arithmetic of modular forms, Inv.
Math. 80 (1985), 469–480. 43.2.8
[Kat92] Svetlana Katok, Fuchsian groups, University of Chicago Press, Chicago, 1992.
33.1, 33.6, 37.3, 37.7
[Kat2007] Svetlana Katok, 𝑝-adic analysis compared with real, Student Math. Library,
vol. 37, Amer. Math. Soc., Providence, RI, 2007. 12.1
[Ker58] M. Kervaire, Non-parallelizability of the n-sphere for 𝑛 > 7, Proc. Nat. Acad. Sci.
44 (1958) 280–283. 1.2
[Kil1888] Wilhelm Killing, Die Zusammensetzung der stetigen endlichen Transformations-
gruppen, Math. Ann. 31 (1888), 252–290; Math. Ann. 33 (1888), 1–48; Math.
Ann. 34 (1889), 57–122; Math. Ann. 36 (1890), no. 2, 161–189. 1.2
[Kir2005] Markus Kirschmer, Konstruktive Idealtheorie in Quaternionenalgebren, Ph.D.
thesis, Universität Ulm, 2005. 17.7.25
[Kir2014] Markus Kirschmer, One-class genera of maximal integral quadratic forms,
J. Number Theory 136 (2014), 375–393 25.4.7
[KL2016] Markus Kirschmer and David Lorch, Ternary quadratic forms over number fields
with small class number, J. Number Theory 161 (2016), 343–361. 26.7, 26.7.4
[KLwww] Markus Kirschmer and David Lorch, The one-class genera of ternary quadratic
forms, website
http://www.math.rwth-aachen.de/~kirschme/orders/. 26.7.4
[Kle1878] Felix Klein, Über die Transformation der elliptischen Funktionen und die Au-
flösung der Gleichungen fünften Grades, Math. Ann. 14 (1878–1879), 111–172.
1.4
[Kle56] Felix Klein, Lectures on the icosahedron and the solution of equations of the
fifth degree, translated by George Gavin Morrice, 2nd. revised edition, Dover
Publications, New York, 1956. 32.4
[Kle79] Felix Klein, Development of mathematics in the 19th century, translated by M.
Ackerman, Lie Groups: History, frontiers and applications, vol. IX, Math Sci
Press, Brookline, Mass., 1979. 34.3.10
[Klt2000] Ernst Kleinert, Units in skew fields, Progress in Math., vol. 186, Birkhäuser
Verlag, Basel, 2000. 28.8, 38.4
[Kna2016] Anthony W. Knapp, Advanced algebra, digital 2nd. edition, East Setauket, New
York, 2016. 27.4
[Kne66a] Martin Kneser, Starke approximation in algebraischen gruppen. I, J. Reine Angew.
Math. 218 (1965), 190–203. 28.7.10, 28.8
[Kne66b] Martin Kneser, Strong approximation, Algebraic Groups and Discontinuous Sub-
groups (Boulder, Colo., 1965), Proc. Sympos. Pure Math., Amer. Math. Soc.,
Providence, RI, 1966, 187–196. 28.7.10, 28.8
[KKOPS86] M. Kneser, M.-A. Knus, M. Ojanguren, R. Parimala, and R. Sridharan, Composi-
tion of quaternary quadratic forms, Compositio Math. 60 (1986), no. 2, 133–150.
19.5.8
[Knu88] Max-Albert Knus, Quadratic forms, Clifford algebras and spinors, Seminários
de Matemática, 1, Universidade Estadual de Campinas, Instituto de Matemática,
Estatística e Ciência da Computaç ã o, Campinas, 1988. 9.7
[Knu93] Max-Albert Knus, Sur la forme d’Albert et le produit tensoriel de deux algebres
de quaternions, Bull. Soc. Math. Belg. 45 (1993), 333–337. 8.2.6
[KMRT98] Max-Albert Knus, Alexander Merkurjev, Markus Rost, and Jean-Pierre Tignol,
838 BIBLIOGRAPHY
The book of involutions, Amer. Math. Soc. Colloquium Publications, vol. 44,
Amer. Math. Soc., Providence, RI, 1998. 4.4.6, 5.6, 5.6.11
[Kob84] Neal Koblitz, 𝑝-adic numbers, 𝑝-adic analysis, and zeta-functions, 2nd edition,
Grad. Texts in Math., vol. 58, Springer-Verlag, New York, 1984. 12.1
[Koh96] David Kohel, Endomorphism rings of elliptic curves over finite fields, Ph.D. thesis,
University of California, Berkeley, 1996. 42.3.6
[Koh2001] David Kohel, Hecke module structure of quaternions, Class field theory: its
centenary and prospect (Tokyo, 1998), ed. K. Miyake, Adv. Stud. Pure Math.,
vol. 30, Math. Soc. Japan, Tokyo, 2001, 177–195. 41.5.14
[KV2003] David R. Kohel and Helena A. Verrill, Fundamental domains for Shimura curves,
J. Théorie des Nombres de Bordeaux 15 (2003), 205–222. 37.9.13
[Kör85] Otto Körner, Über die zentrale Picard-Gruppe und die Einheiten lokaler Quater-
nionenordnungen, Manuscripta Math. 52 (1985), 203–225. 26.6
[Kör87] Otto Körner, Traces of Eichler–Brandt matrices and type numbers of quaternion
orders, Proc. Indian Acad. Sci. (Math. Sci.) 97 (1987), no. 1–3, 189–199. 25.3,
26.1.14, 30.8.8, 30.9.12
[Kur79] Akira Kurihara, On some examples of equations defining Shimura curves and the
Mumford uniformization, J. Fac. Sci. Univ. Tokyo Sect. IA Math. 25 (1979), no.
3, 277–300. 43.2
[Lam99] Tsit-Yuen Lam, Lectures on modules and rings, Grad. Texts in Math., vol. 189,
Springer-Verlag, New York, 1999. 7.2.20, 20.2, 20.2, 20.3.6
[Lam2001] Tsit-Yuen Lam, A first course in noncommutative rings, 2nd. ed., Grad. Texts in
Math., vol. 131, Springer-Verlag, New York, 2001. 7.2
[Lam2002] Tsit-Yuen Lam, On the linkage of quaternion algebras, Bull. Belg. Math. Soc. 9
(2002), 415–418. 8.2.6
[Lam2003] Tsit-Yuen Lam, Hamilton’s quaternions, Handbook of algebra, vol. 3, North
Holland, Amsterdam, 2003, 429–454. 1
[Lam2005] Tsit-Yuen Lam, Introduction to quadratic forms over fields, Grad. Studies in
Math., vol. 67, Amer. Math. Soc., Providence, RI, 2005. 3.5.3, 4.2, 4.2, 4.5, 5.2,
5.4, 8.2.9, 26.8
[Lam2006] Tsit-Yuen Lam, Serre’s problem on projective modules, Springer Monographs in
Mathematics, Springer-Verlag, Berlin, 2006. 20.7.24
[Lamo86] Klaus Lamotke, Regular solids and isolated singularities, Adv. Lectures in Math.,
Friedr. Vieweg and Sohn, Braunschweig, 1986. 32.4
[Lanc67] C. Lanczos, William Rowan Hamilton—An appreciation, Amer. Scientist 55
(1967), 129–143. 1.1
[Lang94] Serge Lang, Algebraic number theory, 2nd ed., Grad. Texts in Math., vol. 110,
Springer-Verlag, Berlin, 1994. 26.2, 26.2, 26.8, 27.4, 27.5
[Lang95] Serge Lang, Introduction to modular forms, appendixes by D. Zagier and Walter
Feit, Grundlehren der Mathematischen Wissenschaften, vol. 222, Springer-Verlag,
Berlin, 1995. 40.1
[Lang82] Serge Lang, Introduction to algebraic and abelian functions, 2nd ed., Grad. Texts
in Math., vol. 89, Springer-Verlag, New York, 1982. 43.6
[Lat26] Claiborne G. Latimer, Arithmetics of generalized quaternion algebras, Amer. J.
Math. (2) 27 (1926), 92–102. 1.3
[Lat35] Claiborne G. Latimer, On the fundamental number of a rational generalized
quaternion algebra, Duke Math. J. 1 (1935), no. 4, 433–435. 14.2.8
[Lat37] Claiborne G. Latimer, The classes of integral sets in a quaternion algebra, Duke
Math. J. 3 (1937), 237–247. 22.5.13
BIBLIOGRAPHY 839
[Lee2011] John M. Lee, Introduction to topological manifolds, 2nd ed., Grad. Texts in Math.,
vol 202, Springer-Verlag, New York, 2011. 34.5
[Lem2011] Stefan Lemurell, Quaternion orders and ternary quadratic forms, 2011,
arXiv:1103.4922. 16.12, 22.5.13, 24.1.5, 24.6
[Len96] H.W. Lenstra, jr., Complex multiplication structure of elliptic curves, J. Number
Theory 56 (1996), 227–241. 42.1
[LS91] H.W. Lenstra and Peter Stevenhagen, Primes of degree one and algebraic cases
of Cebotarev’s theorem, Enseign. Math. 37 (1991), 17–30. 14.2.10
[Lev2013] Alex Levin, On the classification of algebras, M.Sc. thesis, University of Vermont,
2013. 3.5.11
[LR2011] Lawrence S. Levy and J. Chris Robson, Hereditary Noetherian prime rings and
idealizers, Math. Surveys Monogr., vol. 174, American Math. Soc., Providence,
RI, 2011. 21.4.11
[Lew2006a] David W. Lewis, Quaternion algebras and the algebraic legacy of Hamilton’s
quaternions, Irish Math. Soc. Bull. 57 (2006), 41–64. 1
[Lew2006b] David W. Lewis, Involutions and anti-automorphisms of algebras, Bull. London
Math. Soc. 38 (2006), 529–545. 1.2
[LL2012] Wen-Ching Winnie Li and Ling Long, Atkin and Swinnerton-Dyer congruences
and noncongruence modular forms, Algebraic number theory and related topics
2012, RIMS Kôkyûroku Bessatsu, B51, Res. Inst. Math. Sci. (RIMS), Kyoto,
2014, 269–299, arXiv:1303.6228. 35.4.4
[Lin2012] Benjamin Linowitz, Selectivity in quaternion algebras, J. Number Theory 132
(2012), no. 7, 1425–1437. 31.7.7
[LS2012] Benjamin Linowitz and Thomas R. Shemanske, Embedding orders into central
simple algebras, J. Théor. Nombres Bordeaux 24 (2012), no. 2, 405–424. 31.7.7
[LV2015] Benjamin Linowitz and John Voight, Small isospectral and nonisometric orbifolds
of dimension 2 and 3, Math. Z. 281 (2015), no. 1–2, 523–569. 31.7.7, 31.8, 31.4
[Liv2001] Ron Livné, Communication networks and Hilbert modular forms, Applications
of algebraic geometry to coding theory, physics and computation (Eilat, 2001),
NATO Sci. Ser. II Math. Phys. Chem., vol. 36, Kluwer Acad. Publ., Dordrecht,
2001, 255–270. 41.2.14
[LP2010] Nikolai Ivanovich Lobachevsky, Pangeometry, ed. Athanase Papadopoulos, Eu-
ropean Mathematical Society, 2010. 33.1
[Loo53] Lynn H. Loomis, An introduction to abstract harmonic analysis, D. Van Nostrand
Company, Inc., Princeton, 1953. 29.3
[LK2013] David Lorch and Markus Kirschmer, Single-class genera of positive integral
lattices, LMS J. Comput. Math. 16 (2013), 172–186. 25.4, 25.4.7
[Lub2010] Alexander Lubotzky, Discrete groups, expanding graphs and invariant measures,
appendix by Jonathan D. Rogawski, Birkhäuser Verlag, Basel, reprint edition,
2010. 41.2.14
[LPS88] A. Lubotzky, R. Phillips, and P. Sarnak, Ramanujan graphs, Combinatorica 8
(1988), 261–277. 41.2.14
[Luc2003] Mark Lucianovic, Quaternion rings, ternary quadratic forms, and Fourier coef-
ficients of modular forms on PGSp(6), Ph.D. thesis, Harvard University, 2003.
22.5.13
[Macf1891] Alexander Macfarlane, Principles of the algebra of physics, Proc. Amer. Assoc.
Adv. Sci. 15 (1891), 65–112. 1.2
[Macf00] Alexander Macfarlane, Hyperbolic quaternions, Proc. Royal Soc. Edinburgh 23
(1900) 169–180. 1.2
840 BIBLIOGRAPHY
[Mir95] Rick Miranda, Algebraic curves and Riemann surfaces, Grad. Studies in Math.,
vol. 5, Amer. Math. Soc., Providence, RI, 1995. 33.8.6
[Miy2006] Toshitsune Miyake, Modular forms, Translated from the 1976 Japanese original
by Yoshitaka Maeda, Springer Monographs in Math., Springer-Verlag, Berlin,
2006. 28.7, 40.1, 40.4, 40.5
[Moe2002] Ieke Moerdijk, Orbifolds as groupoids: an introduction, Orbifolds in mathematics
and physics (Madison, WI, 2001), Contemp. Math., vol. 310, Amer. Math. Soc.,
Providence, RI, 2002, 205–222. 34.8.15
[MP1997] I. Moerdijk and D. A. Pronk, Orbifolds, sheaves and groupoids, 𝐾-Theory 12
(1997), no. 1, 3–21. 34.8.15
[Mol1893] Theodor Molien, Über Systeme höherer complexer Zahlen, Math. Ann. 41 (1893),
83–156; Berichtigung zu dem Aufsatze ‘Über Systeme höherer complexer Zahlen’,
Math. Ann. 42 (1893), 308–312. 1.2
[MW77] Hugh L. Montgomery and Peter J. Weinberger, Real quadratic fields with large
class number, Math. Ann. 225 (1977), no. 2, 173–176. 26.7.1
[Moore35] Eliakim Hastings Moore, General analysis, Part I, Memoirs Amer. Phil. Soc., vol.
1, Amer. Phil. Soc., Philadelphia, 1935. 3.5.3
[Mor69] L. J. Mordell, Diophantine equations, Pure Applied Math., vol. 30, Academic
Press, London, 1969. 11.4.4
[MT62] G. D. Mostow and T. Tamagawa, On the compactness of arithmetically defined
homogeneous spaces, Ann. of Math. (2) 76 (1962), 446–463. 38.5.6
[Mum70] David Mumford, Abelian varieties, Tata Institute of Fundamental Research Stud-
ies in Math., no. 5, Oxford University Press, Oxford, 1970. 8.5, 43.4
[Mun2004] W. D. Munn, Involutions on finite-dimensional algebras over real closed fields,
J. Austral. Math. Soc. 77 (2004) 123–128. 8.4.17
[Mur88] M. Ram Murty, Primes in certain arithmetic progressions, J. Madras Univ.,
Section B 51 (1988), 161–169. 14.2.10
[Neu99] Jürgen Neukirch, Algebraic number theory, Grundlehren Math. Wiss., vol. 322,
Springer-Verlag, Berlin, 1999. 12.2, 12.2, 13.2, 14.2.10, 14.4, 14.4.4, 14.4.12,
15.1, 26.2, 26.2, 26.8, 27.4, 27.5, 28.7, 28.7
[New72] Morris Newman, Integral matrices, Pure Appl. Math., vol. 45, Academic Press,
New York, 1972. 17.2
[NSW2008] Jürgen Neukirch, Alexander Schmidt, and Kay Wingberg, Cohomology of number
fields, 2nd. ed., Grundlehren Math. Wiss., vol. 323, Springer, Berlin, 2008.
14.6.10
[Nip74] Gordon L. Nipp, Quaternion orders associate with ternary lattices, Pacific J.
Math. 53 (1974), no. 2, 525–537. 22.5.13
[Noe34] Emmy Noether, Zerfallende verschränkte Produkte und ihre Maximalordnungen,
Act. Sci. Ind. Paris 148 (1934), 15 pp. 30.3.17
[ÓCa2000] Fiacre Ó Cairbre, William Rowan Hamilton (1805–1865), Ireland’s greatest math-
ematician, Ríocht na Midhe (Meath Archaeological and Historical Society) 11
(2000), 124–150. 1.1
[ÓCa2010] Fiacre Ó Cairbre, Twenty years of the Hamilton walk, Irish Math. Soc. Bulletin
65 (201), 33–49. 1.1
[O’Do83] Sean O’Donnell, William Rowan Hamilton, Boole Press Limited, 1983. 1.1
[Ogg69] Andrew Ogg, Modular forms and Dirichlet series, W.A. Benjamin, Inc., New
York, 1969. 40.4
[Ogg83] Andrew Ogg, Real points on Shimura curves, Arithmetic and geometry, Vol. I,
Progr. Math., vol. 35, Birkhäuser Boston, Boston, 1983, 277–307. 43.7
842 BIBLIOGRAPHY
varieties, Int. J. Math. Math. Sci. 2004, no. 49–52, 2795–2808. 43.6
[Rud87] Walter Rudin, Real and complex analysis, 3rd. ed., McGraw-Hill, New York,
1987. 29.3.2
[Sah72] Chih-Han Sah, Symmetric bilinear forms and quadratic forms, J. Algebra 20
(1972), 144–160. 8.2.6
[Sal99] David J. Saltman, Lectures on division algebras, CBMS regional conf. series in
math., vol. 94, Amer. Math. Soc., Providence, RI, 1999. 7.5, 7.5.6
[San2017] Jonathan Sands, Zeta-functions and ideal classes of quaternion orders, Rocky
Mountain J. Math. 47 (2017), no. 4, 1277–1300. 26.5.22
[Sar90] Peter Sarnak, Some applications of modular forms, Cambridge Tracts in Math.,
vol. 99, Cambridge University Press, 1990. 41.2.14
[Scha85] Winfried Scharlau, Quadratic and Hermitian forms, Springer-Verlag, Berlin,
1985. 4.2, 4.2, 4.5, 9.7, 9.8
[Scha2009] Rudolf Scharlau, Martin Kneser’s work on quadratic forms and algebraic groups,
Quadratic forms: algebra, arithmetic, and geometry, Contemporary Math.,
vol. 493, Amer. Math. Soc., Providence, RI, 2009, 339–357. 1.3, 28.7.10, 29.11.9
[Schi35] Otto Schilling, Über gewisse Beziehungen zwischen der Arithmetik hyperkom-
plexer Zahlsysteme und algebraischer Zahlkörper, Math. Ann. 111 (1935), no. 1,
372–398. 30.3.17
[Schn75] Volker Schneider, Die elliptischen Fixpunkte zu Modulgruppen in Quaternionen-
schiefkörpern, Math. Ann. 217 (1975), no. 1, 29–45. 30.6.18
[Schn77] Volker Schneider, Elliptische Fixpunkte und Drehfaktoren zur Modulgruppe in
Quaternionenschiefkörpern über reellquadratischen Zahlkörpern, Math. Z. 152
(1977), no. 2, 145–163. 30.6.18
[Sco83] Peter Scott, The geometries of 3-manifolds, Bull. London Math. Soc. 15 (1983),
401–487. 33.1, 34.8
[Ser73] Jean-Pierre Serre, A course in arithmetic, Grad. Texts in Math., vol. 7, Springer-
Verlag, New York, 1973. 14.1, 14.2, 14.3, 25.2.19, 35, 40.1, 40.3
[Ser79] Jean-Pierre Serre, Local fields, Grad. Texts in Math., vol. 67, Springer-Verlag,
New York, 1979. 8.3.7, 12.2, 12.2, 13.2, 13.4.7, 38.5
[Ser96] J-P. Serre, Two letters on quaternions and modular forms (mod p), Israel J. Math.
95 (1996), 281–299. 42.3.14
[Ser2003] Jean-Pierre Serre, Trees, Translated by John Stillwell, corrected 2nd printing,
Springer Monographs in Math., Springer-Verlag, Berlin, 2003. 23.5
[Sha90] Daniel B. Shapiro, Compositions of quadratic forms, de Gruyter Expositions in
Math., vol. 33, Walter de Gruyter, Berlin, 2000. 1.3
[Shem86] Thomas R. Shemanske, Representations of ternary quadratic forms and the class
number of imaginary quadratic fields, Pacific J. Math. 122 (1986), no. 1, 223–250.
30.1
[Shz63] Hideo Shimizu, On discontinuous groups operating on the product of the upper
half planes, Ann. of Math. (2) 77 (1963), no. 1, 33–71. 30.6.18
[Shz65] Hideo Shimizu, On zeta functions of quaternion algebras, Ann. of Math. (2) 81
(1965), 166–193. 1.4, 30.6.18, 39.1
[Shz72] Hideo Shimizu, Theta series and automorphic forms on GL2 , J. Math. Soc. Japan
24 (1972), 638–683. 41.5.13
[Shi67] Goro Shimura, Construction of class fields and zeta functions of algebraic curves,
Ann. of Math. (2) 85 (1967), 58–159. 38.7.20, 43.1.9, 43.8.1, 43.8.2
[Shi71] Goro Shimura, Introduction to the arithmetic theory of automorphic functions,
Kanô Memorial Lectures, no. 1, Publications of the Mathematical Society of
BIBLIOGRAPHY 845
Japan, no. 11, Iwanami Shoten, Tokyo, Princeton University Press, Princeton,
1971. 28.2, 43.9, 43.9
[Shi75] Goro Shimura, On the real points of an arithmetic quotient of a bounded symmetric
domain, Math. Ann. 215 (1975), 135–164. 43.7
[Shi80] Goro Shimura, On some problems of algebraicity, Proceedings of the International
Congress of Mathematicians (Helsinki, 1978), Acad. Sci. Fennica, Helsinki, 1980,
373–379. 1.4
[Shi96] Goro Shimura, in 1996 Steele Prizes, Notices Amer. Math. Soc. 43 (1996), no.
11, 1340–1347.
[Sho85] Ken Shoemake, Animating rotation with quaternion curves, ACM SIGGRAPH
Computer Graphics 19 (1985), 245–254. 1.5
[Sie45] Carl Ludwig Siegel, Some remarks on discontinuous groups, Ann. of Math. (2)
46 (1945), 708–718. 37.7
[Sie65] Carl Ludwig Siegel, Vorlesungen über ausgewählte Kapitel der Functionentheo-
rie, vol. 2, Lecture notes, Universität Göttingen, Germany, 1965. 37.6.5
[Sie69] Carl Ludwig Siegel, Berechnung von Zetafunktionen an ganzzahligen Stellen,
Nachr. Akad. Wiss. Göttingen Math.-Phys. Kl. II 1969 (1969), 87–102. 26.1.9
[Sie89] Carl Ludwig Siegel, Lectures on the geometry of numbers, Springer, Berlin, 1989.
17.5
[Sil2009] Joseph H. Silverman, The arithmetic of elliptic curves, 2nd ed., Grad. Texts in
Math., vol. 106, Springer-Verlag, New York, 2009. 29.13, 40.1, 40.3.10, 42.1,
42.1, 42.1, 42.1, 42.1, 42.2.4, 42.2.4, 42.2, 42.4.3, 43.3.1
[Sim2010] Peter Simons, Vectors and beyond: geometric algebra and its philosophical
significance, dialectica 63 (2010), no. 4, 381–395. 1.1
[Sme2015] Daniel Smertnig, A note on cancellation in totally definite quaternion algebras,
J. Reine Angew. Math. 707 (2015), 209–216. 20.7, 20.7.24
[SV2019] Daniel Smertnig and John Voight, Definite orders with locally free cancellation,
Trans. Lond. Math. Soc. 6 (2019), no. 1, 53–86. 20.7, 20.7.24
[SW2005] Jude Socrates and David Whitehouse, Unramified Hilbert modular forms, with
examples relating to elliptic curves, Pacific J. Math. 219 (2005), no. 2, 333–364.
41.5.14
[Sol77] Louis Solomon, Zeta functions and integral representation theory, Adv. in Math.
26 (1977), 306–326. 26.3.14
[Sta67] H. M. Stark, A complete determination of the complex quadratic fields of class-
number one, Michigan Math. J. 14 (1967), 1–27. 25.2.19
[Sta2007] H. M. Stark, The Gauss class-number problems, Analytic number theory: a tribute
to Gauss and Dirichlet, Clay Math. Proc., no. 7, Amer. Math. Soc., Providence,
RI, 2007, 247–256. 25.2.19
[Sun85] Toshikazu Sunada, Riemannian coverings and isospectral manifolds, Ann. of
Math. (2) 121 (1985), no. 1, 169–186. 31.8
[Swa60] Richard G. Swan, Induced representations and projective modules, Ann. of Math.
(2) 71 (1960), 552–578. 20.7.13
[Swa62] Richard G. Swan, Projective modules over group rings and maximal orders, Ann.
of Math. (2) 76 (1962), 55–61. 20.7.13
[Swa80] Richard G. Swan, Strong approximation and locally free modules. Ring theory
and algebra, III, Lecture Notes in Pure and Appl. Math., vol. 55, Dekker, New
York, 1980, 153–223. 20.7, 20.7.23, 28.8
[Swi74] H.P.F. Swinnerton-Dyer, Analytic theory of abelian varieties, London Math. Soc.
Lecture Note Ser., no. 14, Cambridge University Press, London-New York, 1974.
846 BIBLIOGRAPHY
43.4
[Syl1883] J. J. Sylvester, Lectures on the principles of universal algebra, American J. Math.
6 (1883–1884), no. 1, 270–286. 1.2
[Tai1890] P. G. Tait, An elementary treatise on quaternions, 3rd ed., Cambridge University
Press, Cambridge, 1890. 1.1
[Tam66] Tsuneo Tamagawa, Adèles, Algebraic groups and discontinuous subgroups, Proc.
Sympos. Pure Math., vol. 9, Amer. Math. Soc., Providence, RI, 113–121. 29.11.9
[Tate2010] J. T. Tate, Global class field theory, Algebraic number theory, 2nd. ed., J.W.S.
Cassels and A. Frölich, eds., London Mathematical Society, London, 2010, 163–
203. 27.5, 27.5, 27.5.13
[Tate67] J. T. Tate, Fourier analysis in number fields and Hecke’s zeta functions (Thesis,
1950), Algebraic number theory, 2nd. ed., J.W.S. Cassels and A. Frölich, eds.,
London Mathematical Society, London, 2010, 305–347. 29.2, 29.10.24
[Tho10] Silvanus P. Thompson, The Life of Lord Kelvin, Baron Kelvin of Largs, Vol. II,
Macmillan, London, 1910. 1.1
[Tit79] J. Tits, Reductive groups over local fields, Automorphic forms, representations
and 𝐿-functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis, Ore.,
1977), Part 1, Proc. Sympos. Pure Math., vol. 33, Amer. Math. Soc., Providence,
RI, 1979, 29–69. 23.5.17
[Thu97] William P. Thurston, Three-dimensional geometry and topology, vol. 1, ed. Silvio
Levy, Princeton University Press, Princeton, 1997. 33.8.3, 34.8, 36.5, 36.5
[Tig98] Jean-Pierre Tignol, Algebras with involution and classical groups. European
Congress of Mathematics, Vol. II (Budapest, 1996), Progr. Math., vol. 169,
Birkhäuser, Basel, 1998, 244–258. 4.4.6
[vdB60] F. van der Blij, History of the octaves, Simon Stevin 34 (1960/1961), 106–125.
1.2
[vdW76] B. L. van der Waerden, Hamilton’s discovery of quaternions, Math. Magazine 49
(1976), 227–234. 1.1
[vdW85] B. L. van der Waerden, A history of algebra, Springer-Verlag, Berlin, 1985. 1.2
[vPr2002] Paul van Praag, Quaternions as reflexive skew fields, Adv. Appl. Clifford Algebr.
12 (2002), no. 2, 235–249. 3.5.3
[Ven22] B.A. Venkov, On the arithmetic of quaternion algebras, Izv. Akad. Nauk (1922),
205–220, 221–246. 1.3, 30.1, 30.2
[Ven29] B.A. Venkov, On the arithmetic of quaternion algebras, Izv. Akad. Nauk (1929),
489–509, 532–562, 607–622. 1.3, 30.1, 30.2
[VG74] Marie-France Vignéras-Guého, Simplification pour les ordres des corps de
quaternions totalement définis, C. R. Acad. Sci. Paris Sér. A 279 (1974), 537–540.
32.8
[Vig76a] Marie-France Vignéras, Invariants numériques des groupes de Hilbert, Math.
Ann. 224 (1976), no. 3, 189–215. 30.6.18
[Vig76b] Marie-France Vignéras, Simplification pour les ordres des corps de quaternions
totalement définis, J. Reine Angew. Math. 286/287 (1976), 257–277. 20.7
[Vig80a] Marie-France Vignéras, Arithmétique des algèbres de quaternions, Lecture Notes
in Math., vol. 800, Springer, Berlin, 1980. iii (document), 21.4.10, 28.7, 28.9.23,
29.3, 30.8.8, 30.9.12, 31.1.8, 31.3, 31.8
[Vig80b] Marie-France Vignéras, Variétés riemanniennes isospectrales et non isomét-
riques, Ann. of Math. (2) 112 (1980), no. 1, 21–32. 31.8, 31.8
[Voi2006] John Voight, Computing CM points on Shimura curves arising from cocompact
arithmetic triangle groups, Algorithmic number theory (ANTS VII, Berlin, 2006),
BIBLIOGRAPHY 847
eds. Florian Hess, Sebastian Pauli, Michael Pohst, Lecture Notes in Comp. Sci.,
vol. 4076, Springer, Berlin, 2006, 406–420. 43.8.2
[Voi2011a] John Voight, Characterizing quaternion rings over an arbitrary base, J. Reine
Angew. Math. 657 (2011), 113–134. 22.2.14, 22.5.13, 22.5.14
[Voi2011b] John Voight, Rings of low rank with a standard involution, Illinois J. Math. 55
(2011), no. 3, 1135–1154. 3.5.5, 3.15
[Voi2013] John Voight, Identifying the matrix ring: algorithms for quaternion algebras and
quadratic forms, Quadratic and higher degree forms, Developments in Math., vol.
31, Springer, New York, 2013, 255–298. 9.8
[VW2014] John Voight and John Willis, Computing power series expansions of modular
forms, Computations with modular forms, eds. Gebhard Boeckle and Gabor
Wiese, Contrib. Math. Comput. Sci., vol. 6, Springer, Berlin, 2014, 331–361.
43.9.7
[VZB2015] John Voight and David Zureick–Brown, The canonical ring of a stacky curve,
2015, arXiv:1501.04657. 43.9.6, 43.9
[Vos2009] S.V. Vostokov, The classical reciprocity law for power residues as an analog of
the abelian integral theorem, St. Petersburg Math. J. 20 (2009), no. 6, 929–936.
14.6.4
[Wad86] A. Wadsworth, Merkurjev’s elementary proof of Merkurjev’s theorem, Ap-
plications of algebraic 𝐾-theory to algebraic geometry and number theory,
Part II, (Boulder, Colorado, 1983), Contemporary Mathematics, vol. 55,
Amer. Math. Soc., Providence, RI, 1986, 741–776. 8.3.6
[War89] Seth Warner, Topological fields, North-Holland Math. Studies, vol. 157, North-
Holland Publishing Co., Amsterdam, 1989. 13.5.13
[Was97] Lawrence C. Washington, Introduction to cyclotomic fields, 2nd ed., Grad. Texts
in Math., vol. 83, Springer, New York, 1997. 39.4.11
[Wate69] William C. Waterhouse, Abelian varieties over finite fields, Ann. Scient. École
Norm. Sup., 4th series, 2 (1969), 521–560. 42.2, 42.2
[Wate79] William C. Waterhouse, Introduction to affine group schemes, Grad. Texts in
Math., vol. 66, Springer-Verlag, New York-Berlin, 1979.
[Wats62] G. L. Watson, Transformations of a quadratic form which do not increase the
class-number, Proc. Lond. Math. Soc. (3) 12 (1962), 577–587. 25.4.7
[Wats75] G. L. Watson, One-class genera of positive ternary quadratic forms, II, Mathe-
matika 22 (1975), 1–11. 25.4
[Wed08] J.H. Maclagan Wedderburn, On hypercomplex numbers, Proc. London Math. Soc.
2 (1908), vol. 6, 77–118. 1.2, 7.3.11
[WW2018] Anne van Weerden and Steven Wepster, A most gossiped about genius: Sir
William Rowan Hamilton, BSHM Bull. (2018) 33, no. 1, 2–20. 1.1
[Weib2013] Charles A. Weibel, The 𝐾-book: An introduction to algebraic 𝐾-theory, Grad.
Studies in Math., vol. 145, Amer. Math. Soc., Providence, RI, 2013. 12.4.4
[Weil60] André Weil, Algebras with involution and the classical groups, J. Indian Math.
Soc. 24 (1960) 589–623. 4.4.6, 8.4.17
[Weil66] André Weil, Fonction zêta et distributions, Séminaire Bourbaki (1964-1966), vol.
9, 523–531. 29.10.24
[Weil74] André Weil, Basic number theory, 3rd ed., Springer-Verlag, Berlin, 1974. 12.2,
29.3, 29.10.24
[Weil82] André Weil, Adeles and algebraic groups, Progress in Math., Birkhäuser, Boston,
1982. 27.1.11, 27.6, 29.5, 29.8.8, 29.10, 29.10.4, 29.10.24, 29.11.9
[Wein96] A. Weinstein, Groupoids: unifying internal and external symmetry, Notices
848 BIBLIOGRAPHY
849
850 INDEX
diagonal, 52 similarity, 49
discriminant, 47, 52, 59 quadric, 61
free, 145 quasi-inverse, 257
Gram matrix, 49 quasi-proper, 621
hyperbolic plane, 61, 71, 90 quaternion
integral, 301 imaginary, 27
isometry, 49 pure, 27, 64
isotropic, 51 real, 27
level, 742 scalar, 64
modular, 146 quaternion algebra, 22, 84
multiplicative, 61 S-definite, 581
nondegenerate, 53, 86, 145 absolute discriminant, 506, 510
nonsingular, 61, 145 definite, 169, 210
normalized, 52 descends, 231
orientation, 76 discriminant, 210, 222
orthogonal, 50 indefinite, 210, 222
orthogonal sum, 51 parity condition, 210
Pfister, 61 ramified, 210, 222
primitive, 301 split, 64, 73, 210, 222
radical, 53 splitting field, 73
reduced, 301 standard generators, 23
regular, 145 totally definite, 222
represents, 51 unramified, 210, 222
signed discriminant, 52 quaternion group, 160, 582
totally hyperbolic, 61 quaternion order
totally isotropic, 61 conductor, 386
universal, 51 discriminant, 234
Witt cancellation, 52 good basis, 355
Witt extension, 52 zeta function, 408, 424
Witt index, 61 quaternion ring, 361
quadratic map, 144 quaternionic multiplication (QM), 801
quadratic module, 145 quaternionic multiplication (QM) struc-
associated bilinear map, 144 ture, 784, 802
class set, 146 quaternionic projective line, 651
free, 145 quaternionic Shimura orbifold, 714
isometry, 145 level, 714
primitive, 146 quaternionic Shimura variety, 715
quadratic module in 𝑉, 345 quotient map, 616
quadratic order quotient set, 616
conductor, 250, 267 quotient topology, 617
narrow class group, 301
quadratic reciprocity, 211 radical, 53
number field, 223 radical idealizer, 382, 391
supplement, 211 radically covers, 335
quadratic space, 49 Radon measure, 496
isomorphism, 49 ramification degree, 197
INDEX 859
valuation, 180
discrete, 180
equivalent, 180
extends, 196
normalized, 180
trivial, 180
value group, 180
valuation ring, 181, 199
value group
valuation, 180
vertex, 677
hyperbolic polygon, 603
vertex cycle relation, 679
volume element, 649
Voronoi domains, 685
wandering
group action, 619
Wedderburn’s little theorem, 44, 92
Wedderburn–Artin theorem, 93, 100
Weierstrass ℘-function, 733
Weierstrass equation