(EMS Series of Lectures in Mathematics) Claudio Carmeli, Lauren Caston, Rita Fioresi - Mathematical Foundations of Supersymmetry (Ems Series of Lectures in Mathematics) - European Mathematical Society
(EMS Series of Lectures in Mathematics) Claudio Carmeli, Lauren Caston, Rita Fioresi - Mathematical Foundations of Supersymmetry (Ems Series of Lectures in Mathematics) - European Mathematical Society
Mathematical
Foundations of
Supersymmetry
Authors:
2010 Mathematics Subject Classification: 58-02; 58A50, 58C50, 14M30, 17A70, 32C11, 81Q60
ISBN 978-3-03719-097-5
The Swiss National Library lists this publication in The Swiss Book, the Swiss national bibliography,
and the detailed bibliographic data are available on the Internet at http://www.helveticat.ch.
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, re-use of illustrations, recitation, broadcasting,
reproduction on microfilms or in other ways, and storage in data banks. For any kind of use permission
of the copyright owner must be obtained.
Contact address:
European Mathematical Society Publishing House
Seminar for Applied Mathematics
ETH-Zentrum FLI C4
CH-8092 Zürich
Switzerland
Phone: +41 (0)44 632 34 36
Email: [email protected]
Homepage: www.ems-ph.org
987654321
Preface
ever, the fact that conventional descriptions of space will fail at very small distances
(Planck length) has been the driving force behind the discoveries of unconventional
models of space that are rich enough to portray the quantum fluctuations of space at
these unimaginably small distances. Supergeometry is perhaps the most highly de-
veloped of these theories; it provides a surprising application and continuation of the
Grothendieck theory and opens up large vistas. One should not think of it as a mere
generalization of classical geometry, but as a deep continuation of the idea of space
and its geometric structure.
Out of the first supergeometric objects constructed by the pioneering physicists
came mathematical models of superanalysis and supermanifolds independently by
F. A. Berezin [10], B. Kostant [49], D. A. Leites [53], and De Witt [25]. The idea
to treat a supermanifold as a ringed space with a sheaf of Z=2Z-graded algebras was
introduced in these early works. Later, Bernstein [22] and Leites [53] used techniques
from algebraic geometry to deepen the study of supersymmetry. In particular, Bern-
stein and Leites accented the functor of points approach from Grothendieck’s theory
of schemes. Interest in SUSY has grown in the past decade, and most recently works
by V. S. Varadarajan [76] and others have continued exploration of this beautiful area
of physics and mathematics and have inspired this work. Given the interest and the
number of people who have contributed greatly to this field from various perspectives,
it is impossible to give a fair and accurate account of all the works related to ours. We
have nevertheless made an attempt and have provided bibliographical references at the
end of each chapter, pointing out the main papers that have inspired our work. We
apologize for any involuntary omissions.
In our exposition of mathematical SUSY, we use the language of T -points to build
supermanifolds up from their foundations in Z=2Z-graded linear algebra (superalge-
bra). The following is a brief description of our work.
In Chapter 1 we begin by studying Z=2Z-graded linear objects. We define super
vector spaces and superalgebras, then generalize some classical results and ideas from
linear algebra to the super setting. For example, we define a super Lie algebra, discuss
supermatrices, and formulate the super trace and determinant (the Berezinian). We
also discuss the Poincaré–Birkhoff–Witt theorem in full detail.
In Chapter 2 we provide a brief account of classical sheaf theory with a section
dedicated to schemes. This is meant to be an introductory chapter on this subject and
the advanced reader may very well skip it.
In Chapter 3 we introduce the most basic geometric structure: a superspace. We
present some general properties of superspaces which lead into two key examples of
superspaces, supermanifolds and superschemes. Here we also introduce the notion of
T -points which allows us to treat our geometric objects as functors; it is a fundamental
tool to gain geometric intuition in supergeometry.
Chapters 4–9 lay down the full foundations of C 1 -supermanifolds over R. In
Chapter 4, we give a complete proof of foundational results like the chart theorem
and the correspondence between morphisms of supermanifolds and morphisms of the
superalgebras of their global sections. In Chapter 5 we discuss the local structure
Introduction ix
of morphisms proving the analog of the inverse function, submersion and immersion
theorems. In Chapter 6 we prove the local and global Frobenius theorem on super-
manifolds. In Chapters 7 and 8 we give special attention to super Lie groups and their
associated Lie algebras, as well as look at how group actions translate infinitesimally.
We then use infinitesimal actions and their characterizations to build the super Lie sub-
group, subalgebra correspondence. Finally in Chapter 9 we discuss quotients of Lie
supergroups.
Chapters 10, 11 expand upon the notion of a superscheme which we introduce
in Chapter 3. We immediately adopt the language of T -points and give criteria for
representability: in supersymmetry it is often most convenient to describe an ob-
ject functorially, and then show that it is representable. We explicitly construct the
Grassmannian functorially, then use the representability criterion to show that it is a
superscheme. Chapter 10 concludes with an examination of the infinitesimal theory of
superschemes.
We continue this exploration in Chapter 11 from the point of view of algebraic
supergroups and their Lie algebras. We discuss the linear representations of affine
algebraic supergroups; in particular we show that all affine supergroups are realized as
subgroups of the general linear supergroup.
We have made an effort to make this work self-contained and suggest that the reader
begins with Chapters 1–3, but Chapters 4–9 and Chapters 10–11 are somewhat disjoint
and may be read independently.
Contents
Preface v
Introduction vii
3 Supergeometry 45
3.1 Superspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Supermanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.3 Superschemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.4 The functor of points . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4 Differentiable supermanifolds 54
4.1 Superdomains and their morphisms . . . . . . . . . . . . . . . . . . . 54
4.2 The category of supermanifolds . . . . . . . . . . . . . . . . . . . . 59
4.3 Local and infinitesimal theory of supermanifolds . . . . . . . . . . . 65
4.4 Vector fields and differential operators . . . . . . . . . . . . . . . . . 70
4.5 Global aspects of smooth supermanifolds . . . . . . . . . . . . . . . 74
4.6 The functor of points of supermanifolds . . . . . . . . . . . . . . . . 79
4.7 Distributions with finite support . . . . . . . . . . . . . . . . . . . . 82
xii Contents
B Categories 259
B.1 Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
B.2 Sheafification of a functor . . . . . . . . . . . . . . . . . . . . . . . . 262
B.3 Super Nakayama’s lemma and projective modules . . . . . . . . . . . 266
B.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
Bibliography 277
Index 283
1
Z=2Z-graded linear algebra
The theory of manifolds and algebraic geometry are ultimately based on linear algebra.
Similarly the theory of supermanifolds needs super linear algebra, which is linear
algebra in which vector spaces are replaced by vector spaces with a Z=2Z-grading,
namely, super vector spaces. The basic idea is to develop the theory along the same
lines as the usual theory, adding modifications whenever necessary. We therefore first
build the foundations of linear algebra in the super context. This is an important starting
point as we later build super geometric objects from sheaves of super linear spaces.
Let us fix a ground field k, char.k/ ¤ 2; 3.
Since any element may be expressed as the sum of homogeneous elements, it suffices
to consider only homogeneous elements in the statement of definitions, theorems, and
proofs.
Definition 1.1.3. The superdimension of a super vector space V is the pair .p; q/
where dim.V0 / D p and dim.V1 / D q as ordinary vector spaces. We simply write
dim.V / D pjq.
From now on we will simply refer to the superdimension as the dimension when it
is clear that we are working with super vector spaces. If dim.V / D pjq, then we can
find a basis fe1 ; : : : ; ep g of V0 and a basis f1 ; : : : ; q g of V1 so that V is canonically
isomorphic to the free k-module generated by the fe1 ; : : : ; ep ; 1 ; : : : ; q g. We denote
this k-module by k pjq and we will call .e1 ; : : : ; ep ; 1 ; : : : ; q / the canonical basis of
k pjq . The .ei / form a basis of k p D k0pjq and the .j / form a basis for k q D k1pjq .
2 1 Z=2Z-graded linear algebra
Definition 1.1.4. A morphism from a super vector space V to a super vector space W
is a linear map from V to W preserving the Z=2Z-grading. Let Hom.V; W / denote
the vector space of morphisms V ! W .
Thus we have formed the category1 of super vector spaces that we denote by .smod/.
It is important to note that the category of super vector spaces also admits an “inner
Hom”, which we denote by Hom.V; W /; for super vector spaces V; W , Hom.V; W /
consists of all linear maps from V to W ; it is made into a super vector space itself by
the following definitions:
.V ˝ W /0 D .V0 ˝ W0 / ˚ .V1 ˝ W1 /;
.V ˝ W /1 D .V0 ˝ W1 / ˚ .V1 ˝ W0 /:
cV;W W V ˝ W ! W ˝ V
where v ˝ w 7! .1/jvjjwj w ˝ v.
The significance of this definition is as follows. If we are working with the category
of vector spaces, the commutativity isomorphism takes v ˝ w to w ˝ v. In super linear
algebra we have to add the sign factor in front. This is a special case of the general
1
We refer the reader not accustomed to category language to Appendix B.1.
1.1 Super vector spaces and superalgebras 3
principle called the “sign rule” that one finds in some physics and mathematics litera-
ture. The principle says that in making definitions and proving theorems, the transition
from the usual theory to the super theory is often made by just simply following this
principle, which introduces a sign factor whenever one reverses the order of two odd
elements. The functoriality underlying the constructions makes sure that the definitions
are all consistent.
The commutativity isomorphism satisfies the so-called hexagon diagram:
cU;V ˝W
U ˝ V ˝ QW /V ˝W ˝U
QQQ
QQQU;V
c c U;W mmm
mm6
QQQ mm
QQ( mm
mmm
V ˝U ˝W
where, if we had not suppressed the arrows of the associativity morphisms, the diagram
would have the shape of a hexagon.
The definition of the commutativity isomorphism, also informally referred to as the
sign rule, has the following very important consequence. If V1 ; : : : ; Vn are super vector
spaces and and are two permutations of n elements, no matter how we compose
associativity and commutativity morphisms, we always obtain the same isomorphism
from V.1/ ˝ ˝ V.n/ to V.1/ ˝ ˝ V.n/ namely:
where N is the number of pairs of indices i , j such that vi and vj are odd and 1 .i / <
1 .j / with 1 .i / > 1 .j /.
The dual, V , of V is defined as
V ´ Hom.V; k/:
g.v ˝ w/ D f .v; w/ .v 2 V; w 2 W /:
4 1 Z=2Z-graded linear algebra
from which the definition of V ˝njm follows by the universal property. In other words,
we have:
V ˝njm ´ V ˝ V ˝ ˝ V ˝ ….V / ˝ ….V / ˝ ˝ ….V /
where the parity is coming from the tensor product.
In the ordinary setting, an algebra is a vector space A with a multiplication which
is bilinear. We may therefore think of it as a vector space A together with a linear map
A ˝ A ! A. We now define a superalgebra in the same way:
Definition 1.1.7. A superalgebra is a super vector space A together with a multiplica-
tion morphism W A ˝ A ! A.
We then say that a superalgebra A is (super)commutative if
B cA;A D ;
that is, if the product of homogeneous elements obeys the rule
ab D .1/jajjbj ba:
This is an example of the sign rule mentioned earlier. Note that the signs do not
appear in the definition; this is the advantage of the categorical view point which
suppresses signs and therefore streamlines the theory.
Similarly we say that A is associative if
B ˝ id D B id ˝
on A ˝ A ˝ A. In other words if .ab/c D a.bc/. We also say that A has a unit if there
is an even element 1 so that
.1 ˝ a/ D .a ˝ 1/ D a
for all a 2 A, that is if a 1 D 1 a D a.
The tensor product A ˝ B of two superalgebras A and B is again a superalgebra,
with multiplication defined as
.a ˝ b/.c ˝ d / D .1/jbjjcj .ac ˝ bd /:
As an example of associative superalgebra we are going to define the tensor super-
algebra.
1.1 Super vector spaces and superalgebras 5
together with the product defined, as usual, via the ordinary bilinear map r;s W V ˝r
V ˝s ! V ˝.rCs/ ,
r;s .vi1 ˝ ˝ vir ; wj1 ˝ ˝ wjs / D vi1 ˝ ˝ vir ˝ wj1 ˝ wjs :
One can check that T .V / is a well-defined associative superalgebra with unit, which
is noncommutative except when V is even and one-dimensional.
From now on we will assume that all superalgebras are associative and with unit
unless specified. Moreover we shall denote the category of commutative superalgebras
by .salg/.
If we take a super vector space and mod out the odd part, we obtain a classical (that
is, purely even) vector space. In a superalgebra the corresponding object is defined
by taking the quotient by the ideal generated by the odd elements. This allows one to
always refer back to the classical setting.
We denote by JA the ideal in the commutative superalgebra A generated by the odd
elements in A.
Example 1.1.9 (Grassmann coordinates). Let
A D kŒt1 ; : : : ; tp ; 1 ; : : : ; q
where the t1 ; : : : ; tp are ordinary indeterminates and the 1 ; : : : ; q are odd indetermi-
nates, i.e., they behave like Grassmann coordinates:
i j D j i :
(This of course implies that i2 D 0 for all i .) In other words we can view A as
the ordinary tensor product kŒt1 ; : : : ; tp ˝ ^.1 ; : : : ; q /, where ^.1 ; : : : ; q / is the
exterior algebra generated by 1 ; : : : ; q .
As one can readily check, A is a supercommutative algebra. In fact,
˚ P
A0 D f0 C jI j even fI I j I D fi1 < < ir g
where I D i1 i2 : : : ir , jI j D r and f0 ; fI 2 kŒt1 ; : : : ; tp , and
˚P
A1 D jJ j odd fJ J j J D fj1 < < js g :
Note that although the fj g 2 A1 , there are plenty of nilpotents in A0 ; take for example
1 2 2 A0 .
This example is important since any finitely generated commutative superalgebra
is isomorphic to a quotient of the algebra A by a homogeneous ideal.
As one can readily check, JA D .1 ; : : : ; q / and A=JA Š kŒt1 ; : : : ; tp .
6 1 Z=2Z-graded linear algebra
where
p.s/ ´ jf.k; l/ j k > l; vk ; vl odd; s.k/ < s.l/gj:
Let an be the subspace generated by the elements
Symn .V / D V ˝n =an :
Sym.V / D T .V /=a:
for a; b 2 A.
This definition itself is an instance of the sign rule; since a and D are being inter-
changed, the sign factor appears.
The derivations in Hom.smod/ .A; A/ are even (as above) while those contained in
Hom.smod/ .A; A/1 are odd. The set of all derivations of a superalgebra A, denoted
Der.A/, is itself a special type of superalgebra called a super Lie algebra which we
describe in the following section.
Example 1.1.16. Consider the k-linear operators f@=@ti g and f@=@j g of the polyno-
mial superalgebra A D kŒt1 ; : : : ; tp ; 1 ; : : : ; q to itself defined as
and extended to the whole of A according to (1.1). Then f@=@ti ; @=@j g 2 Der.A/, and
we have ² ³
@ @
Der.A/ D spanA ; :
@ti @j
@=@i is to be viewed as an odd derivation.
8 1 Z=2Z-graded linear algebra
It is important to note that in the super category, these conditions are modifications
of the properties of the bracket in a Lie algebra, designed to accommodate the odd
variables.
We shall often use also the term super Lie algebra instead of Lie superalgebra since
both are present in the literature.
Remark 1.2.2. We can immediately extend this definition to the case where L is an
A-module for A a commutative superalgebra, thus defining a Lie superalgebra in the
category of A-modules that we shall discuss in detail in Section 1.3.
Example 1.2.3. Define the associative superalgebra End.V / as the super vector space
Hom.V; V /:
End.V / D Hom.V; V /0 ˚ Hom.V; V /1
with the composition as product. It is a Lie superalgebra with bracket
ŒX; Y D X Y .1/jXjjY j YX;
where the bracket as usual is defined only on homogeneous elements and then extended
by linearity.
Example 1.2.4. In Example 1.1.16 above,
² ³
@ @
Der.A/ D spanA ;
@ti @j
is a super Lie algebra where the bracket is defined, for D1 ; D2 2 Der.A/, to be
ŒD1 ; D2 D D1 D2 .1/jD1 jjD2 j D2 D1 .
1.3 Modules for superalgebras 9
Remark 1.2.5. If the ground field has characteristic 2 or 3 in addition to the antisym-
metry and Jacobi conditions, one requires that Œx; x D 0 for x even if the characteristic
is 2 and Œy; Œy; y D 0 for y odd if the characteristic is 3. For more details on super-
algebras over fields with positive characteristic see [3].
Definition 1.3.1. A left A-module is a super vector space M with a morphism A˝M !
M , a ˝ m 7! am of super vector spaces obeying the usual identities; that is, for all a,
b 2 A and x, y 2 M we have
(1) a.x C y/ D ax C ay,
(2) .a C b/x D ax C bx,
(3) .ab/x D a.bx/,
(4) 1x D x.
m a D .1/jmjjaj a m
Definition 1.3.2. We say that an A-module M is free if it is isomorphic (in the category
of A-modules) to Apjq for some .p; q/.
This definition is equivalent to saying that M contains p even elements fe1 ; : : : ; ep g
and q odd elements f1 ; : : : ; q g such that
We shall also say that M is the free module generated over A by the even elements
e1 ; : : : ; ep and the odd elements 1 ; : : : ; q .
Let T W Apjq ! Arjs be a morphism of free A-modules and write epC1 ; : : : ; epCq
for the odd basis elements 1 ; : : : ; q . Then T is defined on the basis elements
fe1 ; : : : ; epCq g by
X
pCq
T .ej / D ei tji : (1.2)
iD1
though in matrix form each morphism contains blocks of odd elements of A, each
morphism is an even linear map from M to itself since a morphism in the super category
must preserve parity. In order to get a truly supergeometric version of the ordinary
matrix algebra, we must consider all linear maps M to M , i.e., we are interested in
Hom(M; M ). Now we can talk about even and odd matrices. An even matrix T takes
on the block form from above. But the parity of the blocks is reversed for an odd matrix
S ; we get
S1 S2
SD ;
S3 S4
where S1 is a p p odd block, S4 is a q q odd block, S2 is a p q even block, and S3 is
a q p even block. Note that in the case where M D k pjq , the odd blocks are just zero
blocks. We will denote this superalgebra of even and odd .pCq/.pCq/ D pjq pjq
matrices by Mat.Apjq /. This superalgebra is in fact a super Lie algebra where we define
the bracket Œ ; as in Example 1.2.4:
ŒT; S D T S .1/jT jjSj S T
for S; T 2 Mat.Apjq /.
Remark 1.4.1. Note that Mat.Apjq / D Hom.Apjq ; Apjq /. We do not want to confuse
this with what we will later denote as Mpjq .A/, which will functorially only include
the even part of Mat.Apjq /, i.e.,
Mat.Apjq /0 D Mpjq .A/ D Hom.Apjq ; Apjq /:
We shall come back to this key model in Example 3.1.5.
We now turn to the supergeometric extensions of the trace and determinant. Let
T W Apjq ! Apjq be a morphism (i.e., T 2 .Mat.Apjq //0 ) with block form (1.3).
Definition 1.4.2. We define the super trace of T to be
str.T / ´ tr.T1 / tr.T4 /;
where “tr” denotes the ordinary trace.
This negative sign is actually forced upon us when we take a categorical view of
the trace. We will not discuss this here, but we later motivate this definition when we
explore the supergeometric-extension of the determinant.
Remark 1.4.3. The super trace is actually defined for all linear maps. For S 2
Mat.Apjq /1 an odd matrix,
str.S / D tr.S1 / C tr.S4 /:
Note the sign change. Note also that the trace is commutative, meaning that for even
matrices A; B 2 Mat.Apjq /0 , we have the familiar formula
str.AB/ D str.BA/
12 1 Z=2Z-graded linear algebra
that we shall prove in Observation 1.5.8 after we have introduced the notion of Bere-
zinian.
Definition 1.4.4. If M is an A-module, then GL.M / is defined as the group of auto-
morphisms of M and we call it the super general linear group of automorphisms of
M . If M D Apjq the free A-module generated by p even and q odd variables, then
we write GL.M / D GLpjq .A/. We may also use the notation GLpjq .A/ D GL.Apjq /.
but the T4 block is). There is a similar formulation of the Berezinian which requires
that only the T1 block be invertible:
Ber.T / D det.T4 T3 T11 T2 /1 det.T1 /:
So we can actually define the Berezinian on all matrices with either the T1 or the T4
block invertible. Note that in the case where both blocks are invertible (i.e., when the
matrix T is invertible), both formulae of the Berezinian give the same answer as we
shall see after the next proposition.
Proposition 1.5.4. The Berezinian is multiplicative: For S; T 2 GL.Apjq /,
Ber.S T / D Ber.S / Ber.T /:
Proof. We will only briefly sketch the proof here and leave the details to the reader.
First note that any T 2 GL.Apjq / with block form (1.3) may be written as the product
of the following “elementary matrices”:
1 X Y1 0 1 0
TC D ; T0 D ; T D :
0 1 0 Y2 Z 1
If we equate T D TC T0 T , we get a system of equations which lead to the solution
X D T2 T41 ;
Y1 D T1 T2 T41 T3 ;
Y2 D T4 ;
Z D T41 T3 :
It is also easy to verify that Ber.S T / D Ber.S / Ber.T / for S of type fTC ; T0 g
for all T or T of type fT ; T0 g for all S. Let G GLpjq .A/ be the set of elements
S such that Ber.S T / D Ber.S / Ber.T / for all T . One can check right away that G
is a subgroup of GLpjq .A/. To prove our result is it enough to show that matrices of
type TC , T , T0 2 G since they generate GLpjq .A/. By our previous discussion TC ,
T0 2 G, hence we only need to show that Ber.S T / D Ber.S / Ber.T / for S of type
T for all T . Notice that
Ber.S TC T0 T / D Ber.S TC / Ber.T0 / Ber.T /
as we have already seen. Hence, the last case to verify is for
1 0 1 X
SD and T D :
Z 1 0 1
We may assume that both X and Z each have only one non-zero entry since the product
of two matrices of type TC results in the sum of the upper right blocks, and likewise
with the product of two type T matrices. Let xij ; zkl ¤ 0. Then
1 X
ST D
Z 1 C ZX
14 1 Z=2Z-graded linear algebra
and Ber.ST / D det.1 X.1 C ZX /1 Z/ det.1 C ZX /1 . Since both X and Z have
only one non-zero entry, .1 C ZX /1 D .1 ZX /, hence
This is because all the values within the determinants are either upper triangular or
contain an entire column of zeros (X; Z have at most one non-zero entry), the values
xij ; zkl contribute to the determinant only when the product ZX has its non-zero term
on the diagonal, i.e., only when i D j D k D l. Ber.S T / D det.1 XZ/ det.1
ZX/ D .1 xi i zi i /.1 C xi i zi i / D 1. A direct calculation shows that Ber.S / D
Ber.T / D 1.
Corollary 1.5.5. Let T 2 GLpjq .A/. Then
This very much resembles the big cell decomposition in the theory of ordinary algebraic
groups, however here the decomposition holds globally.
Observation 1.5.8. The usual determinant on the general linear group GLn induces
the trace on its Lie algebra, namely the matrices Mn (see Remark 1.4.1). Let be an
even indeterminate, 2 D 0 and let us compute the value Ber.1 C T / 2 AŒ:
Definition 1.6.1. Let g be a Lie superalgebra, T .g/ the tensor superalgebra over the
underlying super vector space of g; i W g ! T .g/ the natural immersion. Let I T .g/
be the two-sided ideal in T .g/ generated by
(As usual in the super setting we give relations only for homogeneous elements.)
We define U.g/ the universal enveloping superalgebra (UESA) of g as
U.g/ ´ T .g/=I:
is commutative:
U.g/
{{=
j
{{
{{
{{
g / A.
Proof. (1) j.g/ generates U.g/ since j D B i and i.g/ generates T .g/.
(2) Since .I / D 0,
(3) Define the superalgebra morphism 0 W T .g/ ! A, 0 .i.X // D .X/ for all
X 2 g. Since by definition 0 .I / D 0 we have that 0 factors as B for some
W U.g/ D T .g/=I ! A. The uniqueness of comes from (1).
Theorem 1.6.4. Let g be a Lie superalgebra and U.g/ its UESA. Let be a represen-
tation of g in a super vector space V . Then extends uniquely to a representation 0
of U.g/ in V .
The fact that an odd element X appears only with exponent 1 is due to the relation
j.ŒX; X/ D 2j.X /2 . The Poincaré–Birkhoff–Witt Theorem says that these monomi-
als form a basis for U.g/.
Let T 0 T .g/ denote the linear super space of the standard monomials, that is,
the monomials of the form
Xk1 ˝ ˝ Xkr ˝ XkrC1 ˝ ˝ XkrCs ;
k1 kr < krC1 < < krCs ;
Xki even; 1 i r; Xkj odd; r C 1 j r C s:
Assume without loss of generality that r D l 1. To ease the notation let Xir D X ,
XirC1 D Y , XirC2 D Z, so that we have t D Xk1 ˝ ˝ X ˝ Y ˝ Z ˝ ˝ Xkp .
Let us first assume that no two of X , Y , Z are odd and equal. After some calculations
one finds that the two possible ways of defining L.t / give the following two results:
Hence we have
Hence
a b D L.: : : ŒX; ŒX; Y 12 ŒŒX; X ; Y : : : / D 0:
Again by the Jacobi identity it follows that ŒX; ŒX; Y C .1/jXjjY j ŒX; ŒY; X C
ŒY; ŒX; X D 2ŒX; ŒX; Y ŒŒX; X ; Y D 0.
To ease the notation, from now on we identify g with its image j.g/ in U.g/.
Next we show that U.g/ has a filtered algebra structure and we relate it to Sym.g/
the symmetric algebra over the super vector space g.
From now on assume that g is finite-dimensional. Let us first recall some basic
definitions.
0in A.i/ , though not all filtered algebras arise in this way. Vice versa to any
filtered superalgebra B, we can associate a graded superalgebra
L Gr.B/ in the following
way. Define Gr.B/.n/ ´ B n =B n1 and Gr.B/ ´ n0 Gr.B/.n/ as vector super
spaces. Let n W B n ! Gr.B/.n/ be the natural projection. Given a 2 Gr.B/.m/ and
b 2 Gr.B/.n/ choose a0 2 B m and b 0 2 B n such that m .a0 / D a and n .b 0 / D b.
Then define the product ab ´ mCn .a0 b 0 / 2 Gr.B/.mCn/ . One can check this is well
defined and that Gr.B/ is a graded superalgebra.
There are two examples of graded and filtered algebras that are of special interest
to us: the tensor superalgebra T .g/ and the UESA U.g/. T .g/ is graded by taking
T .g/.n/ as the tensors
L of degree n. As any graded algebra, T .g/ is also filtered by
taking T .g/n D 0in T .g/.i/ .
Since the ideal I T .g/ generated by the elements X ˝ Y .1/jXjjY j Y ˝ X
ŒX; Y in the tensor superalgebra T .g/ is not homogeneous (i.e. is not generated by
homogeneous elements) we cannot expect U.g/ to be graded. However it is filtered
with U.g/n ´ .T .g/n / (recall W T .g/ ! T .g/=I Š U.g/). The next proposition
clarifies the structure of U.g/ as filtered superalgebra.
1 X
Q0 D 1; Qn .t / D st
nŠ
s2Sn
Lemma 1.6.12. Let the notation be as above and let W T .g/ ! U.g/.
(1) W T .g/ ! U.g/ is a linear isomorphism preserving the filtration, that is,
.T .g/n / D U.g/n .
(2) T .g/ D I C T .g/.
t t D X1 ˝ : : : Xi ˝ XiC1 ˝ ˝ Xn
.1/jXi jjXi C1 j X1 ˝ ˝ XiC1 ˝ Xi ˝ Xn
D X1 ˝ ˝ ŒXi ; XiC1 ˝ Xn C I 2 T .g/n1 C I:
u Qn .u/ 2 T .g/n1 C I:
Let gN be an abelian Lie superalgebra with the same underlying super vector space
as g. Clearly U.g/N D Sym.g/. Let N W T .g/
N ! U.g/. N J D ker N is the ideal in T .g/
generated by the elements
X ˝ Y .1/jXjjY j Y ˝ X:
24 1 Z=2Z-graded linear algebra
Definition 1.6.14. We define the symmetrizer of Sym.g/ onto U.g/ as the unique linear
map S making the following diagram commute:
T .g/
/ U.g/
u :
S uu
u
u
N
uu
uu
Sym.g/:
Theorem 1.6.15. The symmetrizer map S is a linear isomorphism and preserves the
filtration; moreover
1 X
S..X N n // D
N 1 / : : : .X .1/p.s/ Xs.1/ : : : Xs.n/ :
nŠ
s2Sn
Proof. The fact that S is a linear isomorphism and preserves the filtration is clear by
definition. For its expression observe that
1 X
.Qn .X1 ˝ ˝ Xn // D .1/p.s/ Xs.1/ : : : Xs.n/
nŠ
s2Sn
and
N n .X1 ˝ ˝ Xn // D .X
.Q N 1 / : : : .X
N n /:
By the commutativity of the above diagram we are done.
In conclusion of this section, we want to remark that in [22] the authors take a
radically different point of view in proving the statement of the PBW Theorem. They
use the symmetrizer linear isomorphism to transfer the product structure from U.g/ to
a new product structure in Sym.g/, and then they prove that .Sym.g/; / Š U.g/ as
superalgebras. Their proof holds over a field of characteristic zero, however it is more
general in the sense that it holds for a Lie algebra object in an arbitrary tensor category.
Another different and more general proof than the one that we present is found also
in [76]. There, the statement is given for a Lie superalgebra over k commutative ring
with 1, where 2 and 3 are invertible elements. In this case, the technique of the proof
makes essential use of the “Diamond Lemma” by G. Bergman [12].
1.7 Hopf superalgebras 25
A ˝O A
id˝ /A˝k A ˝O A
˝id /k˝A A ˝O A
˝id /A˝A˝A
O O O
Š Š id˝
A
id / A, A
id / A, A
/ A.
A ˝O A
S˝id /A˝A A ˝O A
id˝S /A˝A
O O
A
i / A, A
i / A.
SB D SA ;
.I / I ˝ A C A ˝ I; .I / D 0; S.I / I:
26 1 Z=2Z-graded linear algebra
One can check immediately that the superalgebra A=I inherits naturally a Hopf super-
algebra structure from A.
I is a coideal of a coalgebra A if it is an abelian subgroup of A and
.I / I ˝ A C A ˝ I:
There are many interesting examples of Hopf algebras; we refer the reader to [20],
[59], [72] for a comprehensive treatment.
In particular to any affine algebraic group in ordinary geometry, we can associate
two very important Hopf algebras: O.G/, the commutative Hopf algebra of algebraic
functions on G, and U.Lie.G//, the universal enveloping algebra of the Lie algebra
Lie.G/. In the case of O.G/ the comultiplication and the counit are given as
follows:
.f /.x ˝ y/ ´ f .xy/; .f / D f .1G /:
In the case of U.Lie.G// the comultiplication and the counit are given on the
generators g as follows:
.X / D X ˝ 1 C 1 ˝ X; .X / D 0 for all X 2 g:
Under suitable conditions, these two Hopf algebras are in duality with each other, as
is explained in [20], Ch. 7.
Proposition 1.8.2. Any A0 -multilinear map .f1 .A/; f2 .A// W V1 .A/ V2 .A/ ! V .A/
functorial in A comes from a unique linear morphism f W V1 ˝ V2 ! V .
Proof (Sketch). We start by proving uniqueness. We want to determine f .v1 ˝ v2 /
by the knowledge of the multilinear morphism f .A/ D .f1 .A/; f2 .A//, for A D
kŒ1 ; 2 . Define the element in V .A/:
´
vi if vi even;
wi D
i vi if vi odd:
For concreteness assume that v1 even and v2 odd. Then we have
f .A/.w1 ; w2 / D f .A/.v1 ; 2 v2 / D 2 f .v1 ˝ v2 /:
This formula determines uniquely f .v1 ˝ v2 /. This formula moreover also gives us
existence since it provides a way to define f .v1 ˝ v2 /.
Remark 1.8.3. (1) There is an obvious generalization of the previous result to the case
of linear morphisms f W ˝i2I Vi ! V . The reasoning is similar, and we invite the
reader to consult [22] for more details.
(2) The previous result holds if A is just taken to be an exterior algebra, A D
kŒ 1 : : : n , in other words, we can substantially weaken our hypothesis. This is clear
by looking at the proof.
(3) The same result holds also for modules over some commutative superalgebra
R replacing our field k while A runs over commutative R-superalgebras and V .A/ is
the A0 -module .A ˝R V /0 as before. The proof goes practically unchanged.
Proposition 1.8.2 has the following very useful corollary, which allows us to extend
in a straightforward manner algebraic structures from V .A/ to VA .
Corollary 1.8.4. Let the notation be as above. In order to give a superalgebra structure
on a super vector space V , it is enough to give a functorial structure of A0 -algebra on
V .A/. The superalgebra will be Lie, respectively associative or commutative and so
on, if and only if the V .A/ A0 -algebras are.
Proof. The first statement is clear since a superalgebra structure comes precisely from
the morphisms considered in Proposition 1.8.2. As for the Lie, associative and com-
mutative properties, they rely on commutative diagrams constructed starting from such
morphisms. Proposition 1.8.2 ensures the commutativity of such diagrams by the func-
toriality of the corresponding diagrams for V .A/. We leave to the reader all the checks
involved in this construction.
1.9 References
All of the material appearing here is standard and can be found in one of the references
[56], [76], [10], [49].
2
Sheaves, functors and the geometric point of view
Classically, sheaf theory provides an alternative and very elegant way to look at the-
ories, like differentiable manifolds and algebraic varieties, which originally were in-
troduced using very different methods. On the other hand, in supergeometry, its use
is unavoidable since at the very start we need sheaves in order to define properly any
supergeometric object in differential or algebraic category.
We start our treatment by discussing sheaves of functions in the familiar setting of
differentiable manifolds and algebraic varieties and then we go on to the more abstract
definition of sheaves in general, arriving finally at ringed spaces and locally ringed
spaces, which is our starting point for the development of supergeometry, as we shall
see in the next chapter.
We also give a brief introduction to schemes and their functor of points, discussing
the examples of projective space and Grassmannian variety. The language of the functor
of points and its generalization to supergeometry are extremely important to us since
they provide geometric intuition to an otherwise very abstract setting. In the end, we
recall some results on coherent sheaves, that we shall need in the sequel.
We do not present here proofs for most of our statements, which are all classical and
well known; a treatment of sheaf theory and ringed spaces can be found. for example.
in [43], Ch. II, or in [29], Ch. I, while a comprehensive functorial treatment of algebraic
geometry, via the functor of points, is found in [23].
than the similar generalization of differentiable geometry. This is because the machin-
ery of algebraic geometry was developed to take already into account the presence of
(even) nilpotents and consequently, the language is more suitable to supergeometry.
Let X be an affine algebraic variety in the affine space An over an algebraically
closed field k and let O.X / D kŒx1 ; : : : ; xn =I be its coordinate ring, where the ideal
I is prime. This corresponds topologically to the irreducibility of the variety X . We
can think of the points of X as the zeros of the polynomials in the ideal I in An . X is a
topological space with respect to the Zariski topology, whose closed sets are the zeros
of the polynomials in the ideals of O.X / (see [43], Ch. I, for a complete discussion of
the Zariski topology). For each open U in X , consider the assignment
U 7! OX .U /;
where OX .U / is the k-algebra of regular functions on U . By definition, these are the
functions f W X ! k that can be expressed as a quotient of two polynomials at each
point of U X .
As in the case of differentiable manifolds, our assignment satisfies the properties
(1) and (2) described above. The first property is clear while for the second one we
refer the reader to [43], Ch. II, Section 2.
The assignment U 7! OX .U / is another example of a sheaf and we shall call it the
structure sheaf of the variety X or the sheaf of regular functions. .X; OX / is another
example of a (locally) ringed space.
U V 7! rV;U W F .V / ! F .U /
The suffix “op” denotes as usual the opposite category; in other words, F is a con-
travariant functor from top.M / to .alg/. A presheaf is a sheaf if it satisfies the property
(2) of Definition 2.2.1.
If F is a (pre)sheaf on jM j and U is open in jM j we define F jU , the ( pre)sheaf F
restricted to U , as the functor F restricted to the category of open sets in U (viewed as
a topological space itself). In other words, we restrict our attention to the assignment
V 7! F .V / for just the open sets V in U .
We have defined sheaves as functors with values in the category of commutative
algebras, however, in the same way, we can also define sheaves with values in commuta-
tive rings, groups, sets or other algebraic structures and their super correspondents. Of
course, depending on the structure of the category of arrival, the restriction morphisms
have to be taken as the appropriate morphisms.
From now on, we shall speak only of “presheaf” and of “sheaf”, without further
specifications, whenever our statements are independent of the category of arrival.
A most important object associated to a given presheaf is the stalk at a point.
Definition 2.2.4. Let F be a presheaf on the topological space jM j and let x be a point
in jM j. We define the stalk Fx of F , at the point x, as the direct limit
lim F .U /;
!
32 2 Sheaves, functors and the geometric point of view
where the direct limit is taken for all the U open neighbourhoods of x in jM j. Fx
consists of the disjoint union of all pairs .U; s/ with U open in jM j, x 2 U , and
s 2 F .U /, modulo the equivalence relation: .U; s/ Š .V; t / if and only if there exists
a neighbourhood W of x, W U \ V , such that sjW D t jW .
The elements in Fx are called germs of sections.
F .U /
U
/ G .U /
rU;V rU;V
F .V /
V
/ G .V /.
Equivalently and more elegantly, one can also say that a morphism of presheaves is a
natural transformation between the two presheaves F and G viewed as functors (see
the alternative definition of sheaf in Definition 2.2.3).
A morphism of sheaves is just a morphism of the underlying presheaves.
`
map sOU W U ! x2jU j Fx , sOU .x/ D sx . We give to the étalé space the finest topology
that makes the maps sOU continuous, for all open U jM j and all sections s 2 F .U /.
We define Fet to be the presheaf on jM j:
`
U 7! Fet .U / D fOsU W U ! x2jU j Fx ; sOU .x/ D sx 2 Fx g:
As we shall see, the presheaf Fet is actually a sheaf and it provides an explicit
construction of the sheafification of the presheaf F .
Definition 2.2.7. Let F be a presheaf on jM j. A sheafification of F is a sheaf Fz ,
together with a presheaf morphism ˛ W F ! Fz , such that
˛
(1) any presheaf morphism W F ! G , G a sheaf, factors via ˛, i.e., W F !
Fz ! G ,
(2) F and Fz are locally isomorphic, i.e., there exists an open cover fUi gi2I of jM j
such that F .Ui / Š Fz .Ui / via ˛.
The next proposition establishes the existence of the sheafification of a presheaf.
Proposition 2.2.8. Let F be a presheaf on jM j.
(1) The sheafification Fz of F always exists and it is unique.
(2) The presheaf Fet is a sheaf and it is the sheafification of the presheaf F with
˛ W F ! Fet defined as ˛U .s/.x/ D sx , s 2 F .U /, x 2 U .
In particular F is a sheaf if and only if F Š Fz D Fet .
Proof. See [43], p. 64.
The previous discussion enables us to properly define the quotient of sheaves.
Definition 2.2.9. Let F and G be sheaves of rings on some topological space jM j.
Assume that we have an injective morphism of sheaves G ! F such that G .U /
F .U / for all U open in jM j. We define the quotient F =G to be the sheafification of
the image presheaf: U 7! F .U /=G .U /.
In general, .F =G /.U / ¤ F .U /=G .U /, however they are locally isomorphic by
Definition 2.2.7.
There is a most effective way to define a sheaf and a sheaf morphism, which proves
to be very useful for applications. We just state the results, referring to [29], Ch. I, for
more details.
Definition 2.2.10. Let B be a base for the open sets in the topological space jM j. The
assignment U 7! F .U /, for all U 2 B, is called a B-sheaf if it satisfies the condition
(1) as in Definition 2.2.1 for the open sets in B and the following modification of the
condition (2). For all open U in B, let fUi gi2I be a covering in B of U and ffi gi2I a
family such that fi 2 F .Ui /. If fi jV D fj jV for all V Ui \ Uj , V 2 B, then there
exists a unique f 2 F .U / such that f jUi D fi .
34 2 Sheaves, functors and the geometric point of view
Proposition 2.2.11. Let B be a base for the open sets in the topological space jM j.
(1) Every B-sheaf extends uniquely to a sheaf on jM j.
(2) If G and H are two sheaves on jM j and for all U in B we have a collection of
morphisms
U W G .U / ! H .U /
Proposition 2.2.12. Let fUi gi2I be an open covering of the topological space jM j.
Assume the following:
(1) We have defined sheaves FUi for all i .
(2) Ui Uj W FUi jUi \Uj ! FUj jUi \Uj are isomorphisms satisfying the compatibility
conditions
In the same way we can define analytic real or complex manifolds as locally ringed
spaces locally isomorphic to .Rn ; HRn / or .Cn ; HCn /, where HRn and HCn denote
the sheaves of analytic functions on Rn or Cn respectively (we leave to the reader as
an exercise their definition, see [37] for more details).
It is important to keep this point of view in mind since this is the path we are going
to take to define supermanifolds.
At this point, it is clear that we could also give an equivalent definition of algebraic
variety in order to fit it into the framework of locally ringed spaces. However, as we shall
see in the next section, we prefer to give a far reaching generalization of this picture,
namely the notion of scheme, which turns out to be the best for our supergeometric
applications.
2.3 Schemes
The concept of scheme is a step towards a further abstraction. We shall start by defining
affine schemes and then we proceed to the definition of schemes in general. Our
treatment is necessarily very short, for all the details and the complete story, we refer
the reader to [29], Ch. I, and [43], Ch. II.
Let us start by associating to any commutative ring A its spectrum, that is the
topological space Spec A. As a set, Spec A consists of all the prime ideals in A. For
each subset S A we define as closed sets in Spec A:
One can check that this actually defines a topology on Spec A called the Zariski topol-
ogy.
If X is an affine variety, defined over an algebraically closed field, and O.X / is
its coordinate ring, we have that the points of the topological space underlying X
are in one-to-one correspondence with the maximal ideals in O.X /. So we notice
immediately that Spec O.X / contains far more than just the points of the topological
space of X; in fact it contains also all the subvarieties of X , whose information is
encoded by the prime ideals in O.X /. This tells us that the notion of scheme, we are
about to introduce, is not just a generalization of the concept of algebraic variety, but it
is something deeper, containing more information about the geometric objects we are
interested in.
36 2 Sheaves, functors and the geometric point of view
Uf 7! Af
defines a B-sheaf on the topological space Spec A and it extends uniquely to a sheaf of
commutative rings on Spec A, called the structure sheaf and denoted by OA . Moreover
the stalk at a point p 2 Spec A, OA;p is the localization Ap of the ring A at the prime p.
Proof. Direct check. (See also [29], I-18.)
Hence, given a commutative ring A, Proposition 2.3.1 tells us that the pair
.Spec A; OA / is a locally ringed space that we call Spec A, the spectrum of the ring A.
By an abuse of notation we shall use the word spectrum to mean both the topological
space Spec A and the locally ringed space Spec A, the context making clear which one
we are talking about.
We are finally ready for a definition of scheme. While the differentiable manifolds
are locally modelled, as ringed spaces, by .Rn ; CR1n /, the schemes are geometric objects
modelled by the spectrums of commutative rings.
Definition 2.3.2. We define an affine scheme to be a locally ringed space isomorphic to
Spec A for some commutative ring A. We say that X is a scheme if X D .jX j; OX / is
a locally ringed space, which is locally isomorphic to affine schemes. In other words,
for each x 2 jX j, there exists an open set Ux jX j such that .Ux ; OX jUx / is an affine
scheme. A morphism of schemes is just a morphism of locally ringed spaces.
Observation 2.3.3. (1) There is an equivalence of categories between the category
of affine schemes .aschemes/ and the category of commutative rings .rings/. This
equivalence is defined on the objects by
(2) Since any affine variety X is completely described by the knowledge of its
coordinate ring O.X /, (the ring of regular functions on the whole variety), we can
associate uniquely to an affine variety X the affine scheme Spec O.X /. As we noted
previously, the two notions of X as algebraic variety or as a scheme are different,
however they describe the same geometrical object from two different points of view.
Similarly to any algebraic variety (not necessarily affine) we can associate uniquely
a scheme. Moreover a morphism between algebraic varieties determines uniquely a
morphism between the corresponding schemes. In the language of categories, we say
we have a fully faithful functor from the category of algebraic varieties to the category
of schemes. For more details see [43], Ch. II, Proposition 2.3, and [29], Ch. I. We shall
show this proposition in Chapter 10 in the more general setting of supergeometry.
In the next example we describe the simplest example of a non-affine scheme: the
projective space.
Example 2.3.4 (Proj S). Let S be a graded commutative k-algebra, k a field, i.e.,
L
S D i0 Si , Si Sj SiCj , S0 D k. The elements in Si are called homogeneous
elements of degree i . We define Proj S as the set of all relevant homogeneous
L prime
ideals (i.e., prime ideals generated by homogeneous elements, p ¤ i>0 Si ). Proj S
is a topological space with the closed sets defined as
V .I / D fp 2 Proj S j p I g:
Wf D j Proj Sj V .f /;
cover Proj S.
It is not hard to see, using Proposition 2.2.12, that the structure sheaves on each
.Proj S /i extend uniquely and compatibly to a sheaf on Proj S to give a locally ringed
space that we denote by Proj S. By its very construction the locally ringed space Proj S
is locally isomorphic to the affine schemes Spec S Œxi1 0 , hence it is a scheme.
If S D kŒx0 ; : : : ; xn is the graded polynomial ring, we also write P n for the scheme
Proj S and we call it the projective space of dimension n. (For more details on this
construction see [29], p. 97.)
38 2 Sheaves, functors and the geometric point of view
Hence hX .T / are the T -points of the scheme X . The restriction of hX to affine schemes
is not in general representable. However, since, as we noticed in Observation 2.3.3,
the category of affine schemes is equivalent to the category of commutative rings we
a
have that such restriction gives a new functor hX :
a
hX W .rings/ ! .sets/; a
hX .A/ D Hom.Spec A; X / D A-points of X :
We are going to see a proof of this result in the more general setting of supergeometry
in Chapter 4, Section 4.5.
We are now going to state Yoneda’s lemma, a basic categorical result. As an imme-
diate consequence, we have that the functor of points of a scheme (resp. differentiable
manifold) does determine the scheme (resp. differentiable manifold) itself.
Proof. We briefly sketch it, leaving the details to the reader. Let ˛ W hX ! F . We
can associate to ˛, ˛X .idX / 2 F .X /. Vice versa, if p 2 F .X /, we associate to p,
˛ W hX ! F such that
˛Y W Hom.Y; X / ! F .Y /; f 7! F .f /p:
Corollary 2.4.8. Two schemes (resp. manifolds) are isomorphic if and only if their
functors of points are isomorphic.
The advantages of using the functorial language are many. Morphisms of schemes
are just maps between the sets of their A-points, respecting functorial properties. This
often simplifies matters, allowing us to leave the sheaves machinery in the background.
The problem with such an approach, however, is that not all the functors from .schemes/
to .sets/ are the functors of points of a scheme, i.e., they are representable. The next
theorem establishes an important criterion. We shall state the theorem for functors from
.rings/ to .sets/ leaving to the reader, as a simple exercise, to write a similar statement
for functors from .schemes/ to .sets/.
(2) F admits a cover by open affine subfunctors. This means that there exists
a family Ui of subfunctors of F , i.e., Ui .R/ F .R/ for all R 2 .rings/, Ui D
hSpec Ui , with the property that for all natural transformations f W hSpec A ! F , the
functors f 1 .Ui /, defined as f 1 .Ui /.R/ D f 1 .Ui .R//, are all representable,
i.e., f 1 .Ui / D hVi , and the Vi form an open covering of Spec A.
In Chapter 10 we are going to see a complete proof of this statement in the more
general setting of superschemes.
This theorem states the conditions we expect for F to be the functor of points of a
scheme. Namely, locally, F must look like the functor of points of a scheme (property
(2)), moreover F must be a sheaf, that is F must have a gluing property that allows
us to patch together the open affine cover we are given in the hypothesis (property (1)
and (2) together).
A similar criterion holds for the functor of points in the differential category and
more general for a locally ringed space and we are going to discuss it in Chapter 9.
We conclude this section by examining the two important examples of projective
space and Grassmannian variety using the functorial point of view.
Examples 2.4.10. (1) Projective space. Let us revisit Example 2.3.4 of projective space
as a scheme using the functor of points point of view. Define the functor h W .rings/ !
.sets/, where h.A/ is the set of projective submodules of AnC1 of rank n. Equivalently,
by duality, we have that h.A/ consists of the morphisms ˛ W AnC1 ! L, where L has
rank 1, modulo the equivalence relation ˛ ˛ 0 if and only if ker.˛/ D ker.˛ 0 /.
To complete the definition we need to specify h on morphisms W A ! B.
Given a morphism W A ! B, we can give to B an A-module structure by setting
ab D .a/b; a 2 A; b 2 B:
h. /.f / W B m Š Am ˝A B ! L ˝A B:
We want to show that h is the functor of points of the projective space, in other
words h D haP n . We briefly sketch how to check property (2) in Theorem 2.4.9; as for
property (1) and in general for more details, we refer the reader to [23], Ch. 1, §1.
Property (2) says that we need to cover h by open affine subfunctors vi . The functors
vi are defined as follows. For A local, vi .A/ is the set of ˛ such that ˛.0; : : : ; ai ; : : : ; 0/
is invertible. As an exercise one can show that vi corresponds to an open affine subfunc-
tor of h, and it corresponds to the functor of points of an affine space of dimension n. If
A is local we have the following nice characterization of the A-points of the projective
space (see [29], Ch. III, §2).
42 2 Sheaves, functors and the geometric point of view
Gr. /.f / W B m D Am ˝A B ! L ˝A B:
vI .A/ D f˛ W Am ! L j ˛ B I is invertibleg:
case, we are going to study in detail this same example in the supergeometric setting in
Chapter 10. For more details on this and the previous example in the ordinary setting,
see [23], Ch. I, §1, and [29], Ch. VI.
Uf 7! Mf ;
2.6 References
For the theory of sheaves and schemes see [37], [43], [29], [38] and the fundamental
paper by Serre that originated the theory [70]. For a complete treatment of the functor
of points of a scheme see [23]; however, a good summary of most of the properties
needed is found in [29]. As for all the algebraic statements (e.g. the definition of
localization, its properties and so on), we refer to the classical textbooks [51], [2].
3
Supergeometry
3.1 Superspaces
A unified way to look at the categories of ordinary differentiable manifolds or algebraic
schemes, is to think of an object as a pair, consisting of a topological space together
with a sheaf of functions defined on it. Such a pair is often referred to as a ringed
space. For ordinary manifolds, for example, the sheaf of functions is the sheaf of the
C 1 functions while for ordinary algebraic varieties it is the sheaf of regular functions,
as we have seen in the previous chapter. We are going to generalize this point of
view, discussed in detail for the ordinary setting in the previous chapter, introducing
supermanifolds and superschemes in the framework of ringed spaces.
Definition 3.1.2. A superspace is a super ringed space S with the property that the
stalk OS;x is a local super ring for all x 2 jS j.
invertible.
In the special case M D Rp , we define the superspace
From now on, with an abuse of notation, Rpjq denotes both the super vector space Rp ˚
Rq and the superspace .Rp ; CR1p Œ 1 ; : : : ; q /, the context making clear which one we
mean. Rpjq plays a key role in the definition of supermanifold since it is the local model.
If t 1 ; : : : ; t p are global coordinates for Rp we shall speak of t 1 ; : : : ; t p ; 1 ; : : : ; q as
a set of global coordinates for the superspace Rpjq .
We are going to study the example of Rpjq in detail in the next chapter.
Example 3.1.5 (Supermatrices Mpjq and the general linear supergroup GLpjq ). Let
2 2
Mpjq D Rp Cq j2pq . This is the superspace corresponding to the super vector space
of .pjq pjq/-matrices, the underlying topological space being Mp Mq , the direct
3.1 Superspaces 47
product of p p and .q q/-matrices (see also Remark 1.4.1). As a super vector space
we have
² ³ ² ³ ² ³
A B A 0 0 B
Mpjq D ; .Mpjq /0 D ; .Mpjq /1 D ;
C D 0 D C 0
3.2 Supermanifolds
A supermanifold is a specific type of superspace, which we describe via a local model.
Because we always keep an eye on the physics literature from which supersymmetry
springs, the supermanifolds of interest to us are the C 1 -supermanifolds over R. Nev-
ertheless, all the definitions we give here hold also in the context of analytic real or
complex supermanifolds, as we shall see more explicitly at the end of the next chapter.
Let CU1 be the sheaf of C 1 -functions on the domain U Rp . We define the su-
perdomain U pjq to be the superspace .U; CU1 Œ 1 ; : : : ; q / where CU1 Œ 1 ; : : : ; q D
V
CR1p jU ˝ . 1 ; : : : ; q /. Most immediately, the superspaces Rpjq are superdomains
with sheaf CR1p Œ 1 ; : : : ; q .
Definition 3.2.1. A supermanifold M D .jM j; OM / of dimension pjq is a superspace
that is locally isomorphic to Rpjq . In other words, given any point x 2 jM j, there
exists a neighborhood V jM j of x with q odd indeterminates j so that
X X X
q
f .t; / D f0 .t /C fi .t / i C fij .t / i j C D f0 .t /C fI .t / I ; (3.2)
i i<j jI jD1
JM D OM;1 C OM;1
2
D hOM;1 iOM :
This is an ideal sheaf in OM and defines a natural subspace of M we shall call Mred
or Mz (not to be confused with the notation for coherent sheaves in Proposition 2.5.1),
the reduced manifold associated with M where
z D .jM j; OM =JM /:
M
z is a completely even superspace, and hence lies in the ordinary category
Note that M
1
of ordinary C -manifolds, i.e., it is locally isomorphic to Rp . The quotient map from
OM ! OM =JM defines the inclusion morphism M z ,! M . The subspace M z should
not be confused with the purely even superspace .jM j; OM;0 /, which is not a C 1 -
manifold, since the structure sheaf still contains nilpotents.
We now examine closed submanifolds in the super category.
3.3 Superschemes
A superscheme is an object in the category of superspaces that generalizes the notion of
an ordinary scheme, which we have introduced and discussed in the previous chapter.
As we shall see in Chapter 10, superschemes are also characterized by local models,
called in this case affine superschemes. They are superspaces locally isomorphic to the
spectrum of superrings, in analogy with the ordinary setting.
Because any non-trivial supercommutative ring has non-zero nilpotents, we need
to redefine what we mean by a reduced superscheme.
Example 3.3.3 (The affine superspace). Let Am be the ordinary affine space of di-
mension m over a field k. Am consists of the topological space k m , that is the vector
space k m with the Zariski topology, and the sheaf OAm of regular functions on k m . On
k m we define the sheaf OAmjn of superalgebras in the following way. Given U k m
open, V
OAmjn .U / D Ok m .U / ˝ . 1 ; : : : ; n /
where 1 ; : : : ; n are odd (anticommuting) indeterminates. One may readily check that
.k m ; OAmjn / is a reduced superscheme, which we here denote by Amjn .
Let us now revisit the Example 3.1.5 taking the algebraic point of view.
alg
Example 3.3.4 (Algebraic supermatrices Mpjq and the algebraic general linear super-
2 2
group GLpjq ). Let Mpjq D Ap Cq j2pq be the affine superspace corresponding to the
alg alg
super vector space of .pjq pjq/-matrices with entries in the field k. The underly-
alg alg alg alg
ing topological space of Mpjq is the product Mp Mq , where Mp denotes the set
of .p p/-matrices with entries in k, with the Zariski topology. The super ring of
alg
global sections of Mpjq is kŒt ij ; kl , with 1 i; j p or p C 1 i; j p C q
and 1 k p, p C 1 l p C q or p C 1 k p C q, 1 l p. The
conditions det.tij /1i;j p ¤ 0 and det.tij /pC1i;j pCq ¤ 0 define a Zariski open set
alg alg alg
U in Mp Mq , hence we have a superspace GLpjq D .U; OMpjq jU /. One can check
immediately that this is a superscheme. From now on we shall drop the suffix alg to
improve readability, the context making clear if we are considering the general linear
supergroup GLpjq in the algebraic or in the differential context.
S.T / D Hom.T; S /:
where .sspaces/ denotes the category of superspaces and the superscript op as usual
refers to the opposite category (see Appendix B for more details).
By a common abuse of notation the superspace S and the functor of points of S are
denoted by the same letter. Whenever it is necessary to make a distinction, we shall
write hS for the functor of points of the superspace S.
We have defined the functor of points of a superspace; clearly we can also define the
functor of points of a supermanifold or a superscheme, just by changing the category
we start from.
Definition 3.4.2. Let .smflds/ and .sschemes/ denote respectively the categories of
supermanifolds and superschemes introduced above. We define the functor of points
of the supermanifold M to be the functor
Lemma 3.4.3 (Yoneda’s lemma). Let M and N be two superspaces (resp. superman-
ifolds or superschemes). There is a bijection from the set of morphisms ' W M ! N
to the set of maps 'T W M.T / ! N.T /, functorial in T . In particular M and N are
isomorphic if and only if their functors of points are isomorphic.
Remark 3.4.4. To ease the notation we write OT .T / or simply O.T / for the global
sections of a superspace T .
Examples 3.4.6. (i) Let T be just an ordinary topological point viewed as a super-
manifold, i.e., T D R0j0 D .R0 ; R/. By definition a T -point of a manifold M is a
morphism W R0j0 ! M . consists of a continuous map jj W R0 ! jM j, which cor-
responds to the choice of a point x in the topological space jM j, and a sheaf morphism
W OM ! .R/, which assigns to a section its value in x. Then a T -point of M is
an ordinary topological point of jM j.
(ii) Let M be the supermanifold Rpjq and let T be a supermanifold. By the previous
proposition we have that a T -point of M corresponds to a morphism
Then, in this case, a T -point of M is a choice of p even and q odd global sections
on T . This is made more clear in Chapter 4 by Theorem 4.1.11. Thus Rpjq .T / D
p q
OT;0 .T / ˚ OT;1 .T /.
(iii) Let X be the superscheme Amjn as in Example 3.3.3 and let T be an affine
superscheme. By definition, a T -point of X is a morphism of schemes W T !
Amjn , which again, by the previous proposition, corresponds to a super ring mor-
phism W O.Amjn / ! O.T /, that is W kŒx1 ; : : : ; xm ; 1 ; : : : ; n ! O.T /, where
kŒx1 ; : : : ; xm ; 1 ; : : : ; n denotes the polynomial algebra in the even (commuting) in-
determinates x1 ; : : : ; xm and in the odd (anticommuting) ones 1 ; : : : ; n . Hence, as
before, amounts to a choice of m even global sections in O.T / and n odd ones:
We already see the power of T -points in these examples. The first example (T D
R0j0 ) gives us complete topological information while the second (M D Rpjq ) allow us
to talk about coordinates on supermanifolds. The third example shows how remarkably
the differential and the algebraic categories resemble each other under the functorial
treatment.
We plan to fully explore all these topics in the next chapters.
3.5 References
For our brief introduction and overview of supergeometry we send the reader to the
works by Berezin [10], Kostant [49], Manin [56] and the notes of Bernstein [22]. In
such papers, especially at the beginning, such exposition is done with more details. In
particular in [56] there is a discussion of what a superscheme is, and, though implicitly,
the problem of representability makes its appearance. In [49], in 2.15, there is a
discussion of Proposition 3.4.5.
4
Differentiable supermanifolds
Definition 4.1.2. We define the smooth superdomain of dimension .pjq/ to be the super
ringed space U pjq D .U; OU pjq /, where U is an open subset of Rp and OU pjq .V / D
4.1 Superdomains and their morphisms 55
V V V
C 1 .V / ˝ q for each V open subset of U , with q D . 1 ; : : : ; q /. pjq is called
the superdimension of the superdomain. We denote by O.U pjq / the global sections of
the sheaf OU pjq .
A morphism between two superdomains V njm and U pjq is a morphism of super
ringed spaces.
Remark 4.1.3. It is a consequence of Lemma 4.1.9 or of a similar ordinary result
(see Example 3.1.3) that each superdomain is actually a superspace. Moreover we
will show that morphisms between smooth superdomains are automatically superspace
morphisms (see Lemma 4.2.3).
In the following we will make frequent use of the multi-index notation. An n-
tuple IPD .i1 ; : : : ; in / of non-negative integers is called a multi-index. We define
jI j ´ jnD1 ij and I Š ´ i1 Ši2 Š : : : in Š. Moreover we will use the shorthand
@jI j @jI j
´ :
@x I @x1i1 : : : @xnin
Remark 4.1.7. The reader should notice that here Rpjq denotes a completely differ-
ent object than the super vector space Rpjq introduced in Chapter 1. We will see in
Example 4.6.3, in what sense these two objects can be identified.
If ft i gpiD1 are coordinates in C 1 .U /, and f j gjqD1 is a system of linearly inde-
V
pendent algebraic generators of q , then the set ft i ; j g is called a system of (super)
coordinates on U pjq . The assignment of a superdomain U pjq together with a system
of super coordinates is called a superchart or chart for short. We notice that on U pjq
there is a canonical chart, consisting of the canonical coordinates inherited from Rpjq .
We now want to discuss morphisms between superdomains in more detail. We start
with an example, that will lead us to the formulation of the Chart Theorem.
Example 4.1.8. Consider the supermanifold R1j2 with a morphism W R1j2 ! R1j2 .
On R1j2 we have global coordinates t , 1 , 2 and so we may express any section
f as in (3.2):
f D f .t; 1 ; 2 / D f0 .t / C f1 .t / 1 C f2 .t / 2 C f12 .t / 1 2 :
t 7! t ´ t C 1 2 ;
1 7! 1 ´ 1 ; (4.1)
7!
2 2
´ :2
f D f .t ; 1 ; 2 /
D f0 .t C 1 2 / C f1 .t C 1 2 / 1 C f2 .t C 1 2 / 2 C f12 .t C 1 2 / 1 2 :
And so we must only make sense of fI .t C 1 2 /. The key is that we take a formal
Taylor series expansion; the series of course terminates due to the nilpotence of the the
odd coordinates:
fI .t C 1 2 / D fI .t / C 1 2 fI0 .t /:
It is easy to check that this in fact gives a homomorphism of superalgebras. For
g; h 2 C 1 .R/, .gh/ D gh C 1 2 .gh/0 D g h . Notice moreover that in order to
determine the sheaf morphism, it is enough to specify the images of the global sections
since the full sheaf map is determined by restrictions of the global coordinates, as we
shall see in complete and detailed generality later. So in this example, in fact, the
morphism induced by equations (4.1) is unique via the above construction.
4.1 Superdomains and their morphisms 57
This fact is indeed true in general. As we shall see, the Chart Theorem states that
a morphism between superdomains is determined by images of local coordinates
under the sheaf morphism . Its proof requires a preliminary result on polynomial
approximations of smooth sections, which is of interest by itself.
Lemma 4.1.9 (Hadamard’s lemma). Let U pjq be a superdomain, f 2 O.U pjq / and
x 2 jU pjq j D U . Then for each k 2 N there exists a polynomial Pk;x of degree k in
t 1 t 1 .x/; : : : ; t p t p .x/, 1 ; : : : ; q , such that
f Pk;x 2 IxkC1
X 1 @j j fI
fI .t/ D fI .t0 / C .t t0 / C hI;rC1 .t /.t t0 /rC1 :
Š @t t0
j jr
One can readily check that this satisfies the requirements of the statement.
Now we turn to the second part of the statement. Let h D f g 2 IxkC1 for all
x 2 jU pjq j with k q.PWe want to show that h D 0. Since I D 0 for jI j > q for all
x, it follows that h D I;jI jq hI I , where hI 2 Ixr for some r > 0. By definition,
hI .x/ D 0 for all x 2 U , hence hI is identically zero and we have h D 0.
58 4 Differentiable supermanifolds
Remark 4.1.10. Before turning to the proof of the Chart Theorem we stress that the
above lemma has a number of consequences that deserve attention, and that will be
discussed more deeply in Section 4.3.
Theorem 4.1.11 (Chart Theorem). Let U Rpjq and V Rmjn be open superdomains
(with canonical charts). There is a bijection between
(i) the set of morphisms W V ! U and
(ii) the set of systems of p even functions t i and q odd functions j in O.V / such
that .t 1 .m/; : : : ; t p .m// 2 jU j for all m 2 jV j.
Proof. It is clear that if we have a morphism we can uniquely associate to it a set of
p even and q odd functions in O.V /. In fact, we take t i ´ .t i /, j ´ . j /,
where ft i ; j g is the canonical chart in O.U /.
Assume now that (ii) holds. We denote by ft i ; j g and fx r ; s g coordinates on U
and V , respectively. It is clear that we have a continuous map jj W jV j ! jU j. We V now
need to show V that there exists a unique superalgebra morphism W C 1 .jU j/˝ p !
C 1 .jV j/ ˝ n such that .t i / D t i and . j / D j . Let us start with the
existence of such . It is enough V to show that there exists a superalgebra morphism
W C 1 .jU j/ ! C 1 .jV j/ ˝ n , since we know the image of the polynomial odd
P
generators j . Let t i D t i Cni with t i 2 C 1 .jV j/, ni ´ i I
jI j>1 tI and define
the pullback, through a formal Taylor expansion, by
X 1 @ f ˇ
ˇ
.f / ´ ˇ n ; (4.2)
Š @t ˇ z
t
where we are using a multi-index notation as in Lemma 4.1.9 and by jtz we mean to
substitute any instance of ti with tzi .
We show that it is a superalgebra morphism. Indeed,
X 1 @ .f g/ ˇˇ
.f g/ D ˇ n
Š @t ˇ z
t
X 1 @˛ f ˇ ˇ
ˇ @ ˛ g ˇ
D ˇ ˇ
Š ˛ @t ˛ ˇ z @t ˛ ˇ z n
˛< t t
X 1 @ f ˇˇ ˛ ˇ
1 @ˇ g ˇˇ ˇ
D ˇ n˛ n :
˛Š @t ˛ ˇtz ˇŠ @t ˇ ˇtz
˛;ˇ
We
Vq now come to uniqueness.
Vn Suppose that 1 and 2 are two morphisms C 1 .jU j/ ˝
1
! C .jV j/ ˝ such that 1 .t i / D 2 .t i / D t i , 1 . j / D 2 . j /VD j .
They clearly coincide on polynomial sections. If x 2 jU j, f 2 C 1 .jU j/ ˝ q , due
to Lemma 4.1.9, we can write, for k > q, f D Pk;x C h with h 2 IxkC1 . Hence
.1 2 /.Pk;x C h/ D .1 2 /.h/. It is easily checked that i .Ixk / Iyk with y
such that jj.y/ D x, so that the result follows in view of Lemma 4.1.9.
4.2 The category of supermanifolds 59
V V
Given the superalgebra morphism W C 1 .jU j/ ˝ q ! C 1 .jV j/ ˝ n and an
1
open
Vq subset jW j jU j, reasoning
Vn as before we can define a morphism W W C .jW j/˝
1 1
! C .jj .jW j// ˝ by
t i jW ! t i jj j1 .jW j/ ; j jW ! j jj j1 .jW j/ :
In this way it is easy to check that we have a sheaf morphism. We leave to the reader
the easy check of all details.
Remark 4.1.12. Equation (4.2) in the proof of the previous proposition gives the
recipe for calculating the pullback of a generic section from the pullback of the super
coordinates.
Notice also that because the expansion (4.2) involves an arbitrary number of deriva-
tives, there is no straightforward way to make sense of C k -supermanifolds.
The next lemma shows that a morphism of supermanifolds as super ringed spaces
is automatically a morphism of superspaces, hence of supermanifolds. This result is
especially important for applications.
Lemma 4.2.3. Let M and N be supermanifolds, and let W M ! N , ´ .jj; /
be a super ringed spaces morphism. Then
(i) .f /.x/ D f .jj.x// for all f 2 ON .U /,
(ii) defines a supermanifold morphism.
Proof. Let us start with the proof of (i). Suppose that .f /.x/ ¤ f .jj.x//. If
g D f f .jj.x// 2 ON .U /, this means that .g/.x/ ¤ 0. g is not invertible
in any neighbourhood of jj.x/, however, since is a superalgebra morphism and
.g/.x/ ¤ 0, we have a contradiction.
In order to prove (ii), we only need to prove that is a local morphism in the sense
that if Jj j.x/ denotes the maximal ideal of the stalk OM;j j.x/ then
.Jj j.x/ / Jx :
Since Jx identifies with the germs of sections that evaluated at x are zero, (ii) follows
directly from (i).
Remark 4.2.4. We can state the following more general version of the Chart Theorem
that is proven in the same way as its local version 4.1.11.
Theorem 4.2.5 (Global Chart Theorem). Let U Rpjq be a superdomain and M a
supermanifold. There is a one-to-one correspondence between the morphisms M ! U
and the set of p C q-uples of p even functions t i and q odd functions j on M such
that .t 1 .x/; : : : ; t p .x// 2 jU j for all x 2 jM j.
Proof. One way is clear, namely the fact that for each morphism W M ! U we have
the family of sections detailed above. Suppose now we are given a family of sections
t i , j on M such that .t 1 .x/; : : : ; t p .x// 2 jU j for all x 2 jM j. We want to define
a morphism W M ! U . For each point x 2 jM j, if .Vx ; OM jVx / is a chart around
x, we have a morphism x W Vx ! U which is uniquely determined by the Chart
Theorem for superdomains. More precisely x is the morphism corresponding to the
assignment .x i / D t i jVx , . j / D j jVx , where x i , j are local coordinates
for the chart Vx . Now, if f is a section in O.U /, we can consider the pullbacks x .f /
for each x 2 jM j. Clearly x .f /jVx \Vy D y .f /jVx \Vy for each x and y in jM j.
Due to the sheaf property there exists a unique section .f /, hence we have defined
uniquely our morphism .
Remark 4.2.6. Any section f 2 OM .U / can now be interpreted as a morphism
f W U ! R1j1 , thus recovering the intuitive meaning of sections as functions on open
subsets of the supermanifold. In fact as one can readily see, by the Chart Theorem an
even function f W U ! R1j0 corresponds to the choice of an even section in OM .U /,
4.2 The category of supermanifolds 61
Proposition 4.2.7. Let fUi gi2I be an open covering of jM j then there exists a family
fgj g of sections such that
(1) gj 2 O.M /0 ,
(2) fsupp gj gj 2J is a locally finite refinement of fUi gi2I and supp gj is compact for
each j ,
P
(3) gj D 1 and gzj 01 for each j .
There exists also a family fhi gi2I such that
(i) hi 2 O.M /0 ,
(ii) supp hi Ui for all i 2 I of fUi gi2I ,
P
(iii) hi D 1 and hzi 0 for each i .
Proof. If fUi gi2I is an open cover of jM j, then there exists a locally finite refinement
fVj gj 2J where each Vj is an open subset with compact closure (see [78]). Without loss
of generality assume that both Ui and Vj are supercharts coverings. Hence for each Vj
there exists a superchart Ui.j / containing it such that VSj Ui.j / . For each j define
1
See Remark 4.2.6 for the notation.
62 4 Differentiable supermanifolds
an even section fj 2 OM .Ui.j / /0 such that supp fj D supp fzj VSj and fzj > 0 for
each x 2 Vj . Each fj can be identified with a global section fP j 2 O.M / by defining
c
it to be zero in Ui.j /
. 2
By the local finiteness condition f D fj is a well-defined
section in O.M /0 such that fQ > 0. Hence f is invertible and gj ´ fj is the required
f
partition of unity.
The second part of the proposition P is proved as follows. For each i 2 I let Ji ´
fj 2 J j Vj Ui g and define gi0 D k2Ji gk . Due to local finiteness the sum is well
g0
defined and supp gi0 Ui . The sections hi ´ Pi 0
gi
do the job.
JM .U / ´ JOM .U / ´ ff 2 OM .U / j f is nilpotentg:
Next we want to show that the quotient sheaf OM =JM possesses the property
.OM =JM /.U / D OM .U /=JM .U /; in other words, we do not need to take any sheafi-
fication to obtain the quotient sheaf. (See Definition2.2.7 for the notion of sheafifica-
tion.)
.smflds/ ! .mflds/:
64 4 Differentiable supermanifolds
such that
(1) fij is an isomorphism of ringed superspaces,
(2) jfij j D idUij .
Definition 4.2.15. Let the notation be as above. We say that a supermanifold .jM j; OM /
is obtained by glueing the supermanifolds .Ui ; Oi /, if for each i there exists a sheaf
isomorphism
i W .Ui ; OM jUi / ! .Ui ; Oi /
with ji j D idUi and such that (with some abuse of notation)
fij D i j1
conditions are satisfied, but this calculation mimics that in the ordinary category and
is left to the reader. So M N is a .p C r/j.q C s/-dimensional supermanifold with
B
M N D M z Nz . As in the ordinary category, OM N ¤ OM ˝ ON ; instead we
must take the completion of the tensor product to get an equality.
We shall return to the important concept of product of supermanifolds in Section 4.5.
OM;x D lim OM .U /
!
where the direct limit is taken over all open sets U containing x. As for the ordinary
case, also for supermanifolds the stalk can be characterized as follows:
OM;x Š O.M /= :
Here
f g if there exists an open set U M such that f jU D gjU :
as one can readily check from the definitions (see [29], p. 14 for more details). This
characterization explains why the concept of a stalk at x is well suited for the study of
those properties of supermanifolds depending on the behavior of sections on arbitrarily
small neighborhoods of the point x.
If Œf is a germ at x, it makes sense to evaluate Œf at x in the same way as we
evaluate the sections. We need also to define the natural morphism
x W O.M / ! OM;x ; f 7! Œf x :
The next lemma provides another useful characterization of the stalk at a point of
a supermanifold.
Lemma 4.3.1. If mx D ff 2 O.M / j there exists an open neighbourhood U of x
such that fU D 0g, then
OM;x ' O.M /=mx :
Proof. Consider the morphism x W O.M / ! OM;x defined above. x is surjective.
In fact, if f is a representative of the germ Œf x , we can think of it as a section in
OM .U /, where U is an appropriate neighborhood of x. If V is an open subset such
that V Vx U and 2 O.M / is a section such that supp U and jV D 1, then
f is a global representative of Œf x . It is clear that the kernel of x consists of the
elements of O.M / that are zero in a neighborhood of x, i.e., of mx .
66 4 Differentiable supermanifolds
In Section 4.1 we have discussed Hadamard’s Lemma 4.1.9, which holds also in
the context of germs of functions. We simply restate it, leaving the proof to the reader
as a simple exercise.
Lemma 4.3.2 (Local Hadamard lemma). Let M D .jM j; OM / be a supermanifold,
dim.M / D .mjn/, x a point in jM j, t i , j a local coordinate system around x. Let f
and g be sections defined in a neighbourhood U of x.
(1) If Œf 2 OM;x , then for each k 2 N there exists a polynomial Pk;x of degree k
in Œti t i .x/, Œ j such that
Œf Pk;x 2 JxkC1
where Jx D fŒf 2 OM;x j Œf .x/ D 0g.
(2) Moreover, if Œf and Œg are germs at x such that, for a given k n, Œf Œg 2
JykC1 for each y in a neighbourhood of x, then Œf D Œg.
OM;x is a local ring, that is, OM;x contains a unique homogeneous maximal ideal
Jx . Now we want to clarify the structure of Jx .
Proposition 4.3.3. Let Œf 2 OM;x and dim.M / D .mjn/.
(1) If Œf .x/ D 0, then Œf 2 .Œt i t i .x/; Œ j / OM;x . In particular we have
Jx D .Œt i t i .x/; Œ j / OM;x ; OM;x D R ˚ Jx :
Hence OM;x is a local superalgebra with maximal ideal mx D Jx .
(2) If f 2 O.M /, k > n and Œf x 2 Jxk for all x in an open set U , then f jU D 0.
Proof. This follows immediately from Lemma 4.3.2.
At this point one may be tempted to expand in formal power series the germs at a
point x. This is fact makes sense, provided we exert some care (as one should also do
for the ordinary case). We are going to come back to this point in Remark 4.3.12.
Remark 4.3.4. This gives another, and independent, proof of the fact that each su-
perdomain is a superspace. Indeed the lemma essentially establishes that each stalk
OU pjq ;x has a unique maximal ideal given by the sections whose value at x is zero.
More precisely we have the decomposition
OU pjq ;x D R ˚ Jx
with Jx D .Œt i t i .x/; Œ j / OU pjq ;x , where Jx is the ideal of germs that are zero
at x.
This allows us to characterize the value of a section f 2 OU pjq .V / at a point x 2 V
as the image of f under the natural morphism
OU pjq .V / ! OU pjq ;x ! OU pjq ;x =Jx Š R;
where Jx denotes the maximal ideal in the stalk OU pjq ;x .
4.3 Local and infinitesimal theory of supermanifolds 67
The next results characterize the ideals of finite codimensions in the stalks. We
shall use them in Section 4.7.
Proof. First let us notice that, due to Proposition 4.3.3, Jx is a finitely generated OM;x -
module. Consider then the chain of ideals
Since I has finite codimension and Jxk © JxkC1 , we have that there exists k such that
Jxk C I D JxkC1 C I:
Hence
Jxk JxkC1 C I:
So we have
Jxk D .JxkC1 C I / \ Jxk D JxkC1 C I \ Jxk :
Due to Nakayama’s lemma (see Appendix B) we have I \ Jxk D Jxk .
v W OM;x ! R
such that
v.f g/ D v.f /g.x/ C .1/jvjjf j f .x/v.g/:
Remark 4.3.7. Notice that the sign .1/jvjjf j f is immaterial. In fact the sign appears
only when f and v are odd, but in this case f .x/ D 0.
Definition 4.3.8. The super vector space of all tangent vectors at a point x 2 jM j is
called the super tangent space at x and is denoted by Tx .M /.
Any supermanifold morphism W M ! N induces a stalk morphism
The differential .d/x is an even linear map of super vector spaces and for this
reason, once we fix homogeneous bases for such vector spaces, it corresponds to a
diagonal block matrix; in other words, it will not contain much information about
the behaviour of the odd variables. For example, if is the morphism discussed in
Example 4.1.8, one readily checks that .d/x is the identity for all x. In order to better
study infinitesimally the odd directions, we need the concept of Jacobian that we shall
introduce in the next section and fully discuss in Chapter 5.
In the next two propositions we provide very useful characterizations of the tangent
space to a supermanifold.
@ ˇ
ˇ @ ˇ
ˇ ˇ k ˇ l
i .Œt k / D ıik ; .Œ j / D 0; @ ˇ Œt D 0; @ ˇ Œ D ıj l :
@t x @t i x @ j x @ j x
Proof. Let g be a germ in OM;x . Notice that if g is a constant germ or it is in Jx2 then
v.g/ D 0. Due to Proposition 4.3.3, we can write
X X
gD gI .x/ C Œt i t i .x/gi;I C hI .t i /Œt i t i .x/2 Œ I
I i
with gi;I 2 RP
and hI .t i /Œt i t i .x/2 2 Jx2 . Hence, with a mild abuse of notation, we
have v.g/ D gi;0 v.Œt i / C gi .x/v.Œ i /. Hence if we consider the difference
X ˇ ˇ
@ ˇ @ ˇ
V ´v v.Œt / i ˇˇ C v.Œ j / j ˇˇ ;
i
@t x @ x
Proof. Since x is surjective, the map ˛ is injective. To show that it is surjective, let
us take w 2 Der x .O.M /; R/. We want to show that w.g/ D 0 for all g 2 mx , so
that w induces a derivation v W O.M /=mx Š OM;x ! R, with ˛.v/ D w. So let us
take g 2 mx . By definition of mx , there exists an open subset U jM j such that
gjU D 0. We can find a section f 2 O.M / such that supp f U and f .x/ D 1, by
Corollary 4.2.10. So we have
This shows that w descends to a derivation v of the stalk OM;x and that ˛.v/ D w.
Observation 4.3.11. From the Propositions 4.3.9, 4.3.10 we immediately have the
following facts:
(1) For any tangent vector v 2 Tx M and any neighbourhood U of x, there exists a
unique derivation that we still denote by v W OM .U / ! R.
(2) If .t i ; j / are local coordinates in U , any derivation v W OM .U / ! R is deter-
mined once we know v.t i / and v. j /.
Elements of Jxr will be denoted by O.r/. Due to the Proposition 4.3.3, Jxr is generated
as an ideal by the products
f D f .x/ C O.1/
X
D f .x/ C ai1 :::im ;j1 ;:::;js Œ.t 1 t 1 .x//i1 : : :
1<i1 CCim Csr
@s @i1 :::im f
with ai1 :::im ;j1 ;:::;js D 1
i1 Š:::im Š @ j1 :::@ js @.t 1 /i1 :::@.t m /im
.
T
One should be aware that there are germs in the intersection k Jxk . These are
the germs of sections whose germ is in Fx ˝ ƒ where Fx is the ideal of germs of flat
functions at x, i.e., those functions with zero derivatives at all orders. Of course we
cannot obtain any information about flat functions with power series expansion.
70 4 Differentiable supermanifolds
vN W OM N .Um V / / ON .V /
SSS pp
SSS p
SSS pp
SSS
SS) ppppp
wp
ORN .R V /
for any open V jN j (this is easily seen locally by using coordinates, and then by
patching using local uniqueness) so that
vN .a ˝ b/ D v.a/b
where a and b are local functions of M and N , respectively.
72 4 Differentiable supermanifolds
One may similarly “extend” vector fields: let V be a vector field on M , and denote
by 1 the identity operator on the sheaf of the supermanifold. Then we extend V to a
derivation .V ˝ 1/ on M N by forcing V to act trivially on N . If .t; / and .x; / are
local coordinates on M and N respectively, V has the coordinate expression as in (4.3).
Then the extension .V ˝ 1/ has the same coordinate expression on M N described
by coordinates .t; x; ; /, i.e., it is identically zero on .x; /. Again the extension is
made to be unique by patching and using local uniqueness.
The fact that we can define vN just by describing vN .a ˝ b/ will be fully justified
in Section 4.5.
We now want to consider the effect of a derivation on a pullback.
Proposition 4.4.7 (Chain rule). Let U pjq and V mjn be superdomains and denote col-
lectively by ua and v b the supercoordinates over U pjq and V mjn , respectively. If
W U pjq ! V mjn is a morphism, we have
X @ .v b /
@ .f / @f
a
D : (4.4)
@u @ua @v b
b
Proof. The proof makes use of Lemma 4.1.9 and goes along the same line as that of
P b @f
Lemma 4.4.4. Let us write D.f / ´ @ @u.f a
/
and D 0 .f / ´ b @ @u.va / @v b and
0 0
consider the derivation D D . Clearly D D annihilates the supercoordinates v b
and hence all polynomials in them. If f is a generic section in O.V mjn / and x 2 V
then, in view of Lemma 4.1.9, there exists a polynomial Pk;x with k D n C 1 and
h 2 IxnC2 such that f D Pk;x C g. Hence .D D 0 /.f / D .D D 0 /.g/. On the
other hand it is easy to check that .D D 0 /.g/ belongs to IxnC1 . Since x is arbitrary
in V , we can conclude that D D D 0 .
In the ordinary setting equation (4.4) is rewritten in matrix form in terms of the
Jacobian matrix and we have that the composition of morphisms of ordinary manifolds
translates into matrix multiplication of their associated Jacobians. In the super setting
this requires some care due to the appearance of signs. Let us start with a definition of
the Jacobian supermatrix or Jacobian for short.
Proof. The proof goes along the same lines as the classical one and is based on
Lemma 4.4.4 together with an induction argument. We leave this to the reader as
an exercise.
B
pK;D .f / ´ sup j.D.f //.x/j; f 2 OM .U /;
x2K
4.5 Global aspects of smooth supermanifolds 75
where fKni gi;n2N is a countable family of compact subsets covering U with each Kni
contained in a superchart.
(2) For all open coverings fUi g of U , OM .U / has the coarsest topology that makes
the restriction OM .U / ! OM .Ui / continuous.
C 1 .U / y̋ C 1 .V / ' C 1 .U V /;
M N D .jM j jN j; OM N /;
76 4 Differentiable supermanifolds
where the sheaf OM N is defined as follows. For each rectangular open set U V
M N define the superalgebra
OM N .U V / ´ OM .U / Ő ON .V /:
Due to Proposition 2.2.11, since the rectangular sets are a base for the product topology,
it extends to a sheaf over jM j jN j that is clearly locally trivial.
Involving again Proposition 2.2.11 (but this time item (2)), a similar reasoning can
be used for the morphisms.
The next proposition formalizes the important facts about the product of two su-
permanifolds.
Proposition 4.5.4. (1) O.M / ˝ O.N / is dense in O.M N / when they are endowed
with the projective tensor topology.
(2) If i W Mi ! Ni , i D 1, 2 are supermanifold morphisms, then the pullback of
the map 1 2 W M1 M2 ! N1 N2 is given by 1 y̋ 2 which is in turn completely
determined by 1 ˝ 2 .
Notice that this proposition allows us to simplify dramatically the definition of
the pullback morphism on the sheaf on the product of two supermanifolds. We have
already used this in Definition 4.4.6 and we are going to use it again many times in the
text, mostly without mention.
We now turn to the problem of how to recover the topological space jM j of a
supermanifold .jM j; OM / by the superalgebra of its global sections. We start with a
simple proposition.
Proposition 4.5.5. Every x 2 jM j defines the superalgebra morphism evx W O.M / !
R given by evaluation at x, and ker evx D Jx is a maximal ideal in O.M /, with
Proposition 4.5.7 (Super Milnor exercise). All maximal ideals in O.M / are of the
form Jx for some x 2 jM j.
Proof. We have already noticed that each Jx is a maximal ideal in O.M /. Conversely,
let I be a maximal ideal in O.M / and denote IQ the corresponding ideal in C 1 .M z /,
Q 1 z Q
i.e., I ´ ff 2 C .M / j there exists g 2 I such that jgj D f g. I is a maximal ideal
in C 1 .M /. It cannot be the whole C 1 .M / since otherwise I would contain the unit.
Due to the classical Milnor exercise (see [58]) we have that there exists x 2 jM j such
that IQ D Jzx , where Jzx denotes the ideal of smooth functions vanishing at x. Let Jx
be the preimage of Jzx . Clearly I Jx , hence due to maximality I D Jx .
From now on, due to the previous proposition, we can identify the real spectrum
MaxSpecR .O.M // with Hom.O.M /; R/, hence throughout this section we shall iden-
tify a point x with a maximal ideal Jx and with the morphism evx .
We now want to give to MaxSpecR .O.M // a topological space structure.
For each point xN 2 jM j, for each n 2 N, each n-tuple of elements f1 ; : : : ; fn 2
O.M / and each real number , we define the subset
of MaxSpecR .O.M //. As one can readily check, these subsets define a base for a
topology on MaxSpecR .O.M //.
is a homeomorphism.
Proof. This is a classical result; we nevertheless include it for completeness. The fact
that this map is a bijection follows immediately from the previous proposition. To
show that it is a homeomorphism, we need to show that it is open and the preimage of
an open set is open. It is enough to perform this check on the open sets of the basis.
Let U D B .evxN I f1 ; : : : ; fn / be an open subset in MaxSpecR .O.M //; since each fi
is a smooth section, we have 1 .U / is open.
Let now V be an open subset in jM j. We want to show that .V / is open. Let
evx 2 .V / (i.e., x 2 V ). Choose U to be an open subset of jM j such that U
Ux V . Let f 2 O.M / be a section such that f jU D 1 and f jVx c D 0. It is clear
that B 1 .evx I f / .V /, hence any point evx in .V / has an open neighbourhood
2
entirely contained in .V /.
We now want to see how it is possible to reconstruct the whole sheaf OM starting
from the supercommutative algebra of global sections O.M /.
We will use the localization technique, borrowed from the algebraic setting. We
briefly recall a few facts about it, referring the reader to [2], Ch. 3, for all the details.
78 4 Differentiable supermanifolds
The idea is the following: we want to invert some elements in the superalgebra of
sections of the structure sheaf of a supermanifold. Since our rings are not commutative
but supercommutative, we must exert some care.
Consider an open set U jM j and define the subset of sections in O.M / that are
invertible over U . More precisely, put
where the equivalence relation is given by: .s; f / .s 0 ; f 0 / if and only if there exists
s 00 2 U such that
s 00 .s 0 f sf 0 / D 0:
By construction U1 O.M / is an O.M /0 -module, however U1 O.M / is also a super-
algebra if we define addition and multiplication by
The next result is crucial for the reconstruction of the structure sheaf of the super-
manifold from the superalgebra of its global sections.
Proposition 4.5.9. The map
f jU
` W U1 O.M / ! OM .U /; .s; f / 7! ;
sjU
is a superalgebra isomorphism.
Proof. Let us first prove that ` is injective. Suppose hence that fsjjUU is zero. This
means that f jU D 0. Let s 2 U be such that sjUx c D 0, then sf D 0, and injectivity
is proved.
We now come to surjectivity. Let f 2 OM .U /. We want to determine k 2 O.M /
and h 2 SU such that f D kjU = hjU . Let S fUi g be a collection of open sets with
Si UiC1 and
compact closure such that U i Ui D U . Let fgi g be sections in O.M /
such that gi jUi D 1, supp gi UiC1 and 0 jgi j 1 (see Corollary 4.2.10).
Consider
1
X 1
X
1 gi f 1 gi
kD and hD ;
2 1 C qO i .gi / C qO i .gi f /
i 2 1 C qO i .gi / C qO i .gi f /
i
iD1 iD1
where
qO i ´ maxfqi1 ; : : : ; qi i g:
4.6 The functor of points of supermanifolds 79
jM j U 7! U1 O.M / Š OM .U /:
Remark 4.5.11. (1) It is important to remark that, as it happens already in the ordinary
case, OM .U / is in general larger than O.M /jU , that is, there are sections on U that do
not come as the restriction of global sections (e.g., 1=x 2 C1 .R / is not the restriction
of any global section on R).
(2) The superalgebra O.M / does not embed into its localization U1 O.M / Š
OM .U /. This has nothing to do with the odd nilpotents, but it is a phenomenon we
already observe at the ordinary level. Thanks to the partition of unity, we can have
non-zero global sections which are zero on an open set.
Summarizing all of the results of this and the previous section, we can state the
following theorem, which is one of the main results of this chapter.
F W .smflds/ ! .salg/
Proof. The fact that F is a functor is straightforward. The bijectivity on the objects is
a consequence of Proposition 4.5.8 or in any case is coming from the classical result.
The fully faithfulness amounts to Proposition 4.6.1.
Examples 4.6.3. (1) The super vector space Rmjn as superdomain. As we have seen
in Chapter 3, Section 3.4, if T is any supermanifold, we have that the T -points of the
supermanifold Rmjn are given by
where the first Rmjn is the functor of points of the supermanifold Rmjn , while the second
Rmjn is the real super vector space of superdimension mjn, Rmjn D Rm ˚ Rn . In other
words, a T -point of the supermanifold Rmjn consists of an m C n-uple of m even and n
odd global sections of O.T /. We realize that the notation Rmjn is used here to denote
three different objects: the superdomain, its functor of points and the real super vector
space; however this abuse of notation is justified by the above identification, it is widely
spread, and we shall make sure that the context clarifies at each time which one we are
referring to.
(2) The general linear supergroup. Let T be a supermanifold. In Example 3.1.5, we
2 2
have described the super vector space of matrices Mmjn Š Rm Cn j2mn . Reasoning as
in (1), we have that Mmjn .T / can be identified with the endomorphisms of the O.T /0 -
module Rmjn .T / Š .O.T / ˝ Rmjn /0 , by rearranging the m2 C n2 even sections and
the 2mn odd ones in a matrix form with diagonal block matrices with even entries and
off diagonal matrices with odd entries:
² ³
A B
Mmjn .T / D D .Mmjn /0 ˝ O.T /0 ˚ .Mmjn /1 ˝ O.T /1 :
C D
Here A D .aij /, B D .ˇil /, C D .kj /, D D .dkl / with aij ; dkl 2 O.T /0 , and
ˇil ; kj 2 O.T /1 . .Mmjn /0 and .Mmjn /1 denote respectively the even and odd part
of the super vector space of matrices Mmjn (again the same symbol Mmjn has several
different meanings).
Notice that in doing this, we reobtain Mat.Amjn /0 as discussed in Chapter 1, Sec-
tion 1.4, with A D O.T /.
Define F .T / as the group of automorphisms of Rmjn .T /. Clearly F .T / Mmjn .T /.
We want to show that F D GLmjn , i.e., F is the functor of points of the general linear
supergroup, the supermanifold defined in Chapter 3, Sections 3.1–3.2. In that example,
82 4 Differentiable supermanifolds
GLmjn is defined as the open submanifold of the supermanifold Mmjn , whose reduced
space consists of diagonal block mCnmCn real matrices with non-zero determinant:
² ³
P 0
j GLmjn j D j det.P / ¤ 0; det.Q/ ¤ 0 :
0 Q
1
By Proposition 4.5.9 we have O.GLmjn / D jGL O.Mmjn /.
mjn j
Hence
1
GLmjn .T / D Hom.SjGL mjn j
O.Mmjn /; O.T //
Hom.O.Mmjn /; O.T // D Mmjn .T /:
1
The elements in Hom.jGL O.Mmjn /; O.T // correspond to .mjnmjn/-matrices
mjn j
with entries in O.T / whose diagonal blocks are invertible once reduced. We leave it to
the reader as a simple exercise to check that those correspond to .mjn mjn/-matrices
with entries in O.T / whose diagonal blocks are invertible (recall that t C n 2 O.T /,
n nilpotent, is invertible if and only if t is invertible).
Notice that while in general, as we pointed out in Remark 4.5.11, O.M / does not
embed into U1 O.M /, it does in this special case, and this is because here U D jGLmjn j
is dense in jMmjn j.
Hence
² ³
A B
GLmjn .T / D j A; D invertible Mmjn .T /:
C D
This is what we have defined as GLmjn .A/ in Definition 1.4.4 for A D O.T /.
It is well known that the importance of the Sweedler dual stems from the fact that it
inherits a coalgebra structure from the algebra structure of O.M /. Indeed if an algebra
A is not finite-dimensional then the adjoint of the multiplication
h X; a ˝ bi ´ hX; a bi:
.A ˝ A/ © A ˝ A :
The reader can easily check that the support of a distribution is a closed subset of
jM j. As we are going to see presently, the finite supported distributions are identified
with the Sweedler dual. Before this, we need the characterizations of the finite support
distributions and the finite codimension ideals.
Proof (Sketch). (1) Let U be a domain where the local coordinates .t i ; j / make sense
and temporarily denote by OM .U /? the space of distributions with finite support. Since
supp U , we have that is completely determined by jOM .U / . Since U is a
domain, OM .U / Š C 1 .U / ˝ ^. j /, hence OM .U /? ' C 1 .U /? ˝ ^. j /? Since
^. j / is finite-dimensional, its topological dual coincides with its dual as a vector
space. Moreover, since by Schwarz’s theorem (see [64], p. 150) we have that finitely
supported distributions at x are identified with differential operators evaluated at x, we
immediately obtain the result.
(2) Locally at each point of its support, can be written as a point supported
distribution. Using a suitable partition of unity we obtain the global expression of as
the sum of differential operators (that is, point supported distributions) at each point of
the support.
Lemma 4.7.4. Let I be an ideal of finite codimension in O.M /. Then there exists a
finite set x1 ; : : : ; xn in jM j and an n-tuple of integers p1 ; : : : ; pn such that
Proof. Let U be an open subset of jM j and write OM .Ux / ´ O.M /jU . In particular
C 1 .Ux / D C 1 .M /jU identifies with the restriction of smooth functions to the closed
subset Ux .
If J is a finite codimension ideal in OM .Ux /, let JQ CM 1 x
.U / denote the associated
Q
reduced ideal. Clearly J is a finite codimension ideal in CM .U /, and JQ D CM
1 x 1 x
.U / if
and only if J D OM .Ux /. If f 2 CM .Ux /, let Zf ´ f .0/, and let
1 1
T
ZJ ´ ZJQ ´ ZfQ :
f 2J
It is well known (see, for example, [58]) that ZJ is a finite subset fx1 ; : : : ; xn g of jM j,
that ZJ D ; if and only if JQ D CM 1 x
.U /, and that
that gjUx c D f jUx c and let h 2 O.M / be such that hjVx c D 0 and hjU D 1. Then
f D f h D g h 2 I.
We are now ready to end the proof by showing that there exist integers p1 ; : : : ; pn
such that Jxp11 \ \ Jxpnn I . Indeed notice that Ix i ´ x i .I / is a finite codimen-
p
sional ideal of OM;x i and hence, due to Proposition 4.3.5, contains Jx ii for some integer
pi . Let now f 2 Jxp11 \ \ Jxpnn and let gi 2 I be such that Œf x i D Œgi x i . Denote
by Vi the open subsets of jM j such that f jVi D gi jVi , and let Ui be another family of
open sets such that Ui U Si Vi . There exists a partition of unity fh1 ; : : : ; hn ; hg sub-
P
ordinated to the open cover fV1 ; : : : ; Vn ; US1 \ \ USn g. Then f D f hi C f h
c c
The next proposition identifies the finite support distributions with the Sweedler
dual.
Proof. Let a D f h. We are going to show that if a ¤ 0, then there exists a point
supported distribution x with x .a/ ¤ 0. Let U be an open coordinate neighborhood
such that X
ajU D aI I ¤ 0:
I
Clearly there exist aJ and x 2 U such that aJ .x/ ¤ 0. Consider the element x of
O.M /ı given by @ @J x , then x .a/ ¤ 0.
86 4 Differentiable supermanifolds
OM jVx Š Ok pjq jU :
exact same definition for tangent vector and tangent bundle to a real or a complex
supermanifold and we can prove in the same way all of the theorems relative to the local
structure of morphisms, like the inverse function theorem, the submersion, immersion
and the constant rank morphism theorems.
There are however some important differences at the very core of the theory: par-
titions of unity are not available for real or complex analytic (super)manifolds, so one
must exert extreme care in generalizing all the results that make use of them. Moreover,
contrary to the smooth case, the superalgebra of the global sections of the structure
sheaf of a (super)manifold tells us very little, in general, about the structure of the
(super)manifold and in general does not allow us to retrieve all the information about
the (super)manifold itself.
The main goal of this section is to understand how it is possible to define real
structures and real forms in supergeometry. A major character in this game is the
complex conjugate of a super manifold.
Let us first review quickly the classical setting.
Let M D .jM j; HM / be a complex manifold. The complex conjugate of M is the
manifold M D .jM j; HM / where HM is the sheaf of the antiholomorphic functions
on M , which are immediately defined once we have HM and the complex structure on
M . We have the C-antilinear sheaf morphism
HM ! HM ; f 7! fN: (4.5)
In the super context it is not possible to speak directly of antiholomorphic functions
and for this reason we need the following generalization of complex conjugate super
manifold.
Definition 4.8.2. Let M D .jM j; OM / be a complex super manifold. We define a
complex conjugate of M as a complex super manifold M D .jMj; OM /, where now
OM is just a supersheaf, together with a super ringed space C-antilinear isomorphism
with M , which is (4.5) on the reduced supermanifold. This means that we have an
isomorphism of topological spaces jM j Š jMj and a C-antilinear sheaf isomorphism
OM ! OM ; f 7! fN: (4.6)
By an abuse of notation we shall also write f 7! fN for the inverse of the morphism
(4.6).
Example 4.8.3. A complex conjugate of M D C1j1 D .C; OC Œ / is M D .C; OC N
x Œ /,
where OC and OC x denote respectively the sheaf of holomorphic and antiholomorphic
functions on C. In fact the C-antilinear isomorphism is
OM ! OM ; z 7! z;
N 7! N :
Notice that while zN has the meaning of being the complex conjugate of z, N is simply
another odd variable that we introduce to define the complex conjugate.
Practically one can think of the complex conjugate super manifold as a way of
giving a meaning to fN, the complex conjugate of a super holomorphic function.
88 4 Differentiable supermanifolds
Remark 4.8.6. Observe that classically M D .jM j ; OM / has a real manifold struc-
ture. In coordinates, the map is z 7! z.
N Let us look at as a real differentiable map
from M to M , seen as real manifolds. Since this is a local question, we look at in a
neighborhood with coordinates z and zN (see Remark 4.8.4). We have
W M ! M; N 7! .z;
.z; z/ N z/:
We are ready to discuss examples of real structures and forms in the super context.
W M.R/ ! M.R/
4.9 References
The idea of supercalculus goes back to Berezin’s seminal work [10]. The concept
of supermanifold was however introduced later by Berezin and Leites in [11] and in
the beautiful work [49] by Kostant. In particular [49] can be considered as the first
systematic and rigorous treatment of the foundations of smooth supergeometry. The
approach there is very algebraic in nature with a particular attention to the coalgebra
structure of the structure sheaf of a supermanifold. Section 4.7 provides a bridge
between Kostant’s approach and ours.
Our treatment of superdomains, in particular Lemma 4.1.9 and the Chart Theo-
rem 4.1.11, follows quite closely [53]. The content of Section 4.5 and our treatment of
the localization procedure in the super-setting is found also in [6]. In Section 4.7, we
generalize the proof of Proposition 4.7.5 given in [58].
For further reading and a more modern treatment of smooth supergeometry the
reader can consult [56], [6], [22], [76].
5
The local structure of morphisms
The goal of this chapter is to study the local structure of supermanifold morphisms. As
in the ordinary setting we have the inverse function theorem and locally we can classify
the morphisms according to the rank of their differentials, so that we can formulate the
supergeometric versions of the immersion and submersion theorems. There is however
an important difference with the ordinary differentiable manifolds theory, which arises
when we discuss the constant rank morphisms, which are the natural generalization
of the immersion and submersion morphisms. There is in fact a difficulty with the
definition of rank of the Jacobian of a morphism: as we shall see this notion is not
always well defined. This is a supergeometric peculiarity, ultimately linked to the
graded nature of super vector spaces. The constant rank morphisms are crucial for a
complete treatment of submanifolds and it is the key tool for their explicit constructions
as we shall see in the section on submanifolds.
All of these are well-known results that appeared in several papers, including [53],
[56], [76]. Nevertheless we provide complete proofs of them in the effort to make the
text self-contained.
Proof. Since the statement is local we can assume both M and N to be superdomains
U pjq and V pjq respectively. The superdimensions must be equal since the differential
is bijective.
Denote by x i , j and t r , s supercoordinates on U pjq and V pjq , respectively. By
the Chart Theorem 4.1.11, W O.V pjq / ! O.U pjq / is given by
´ P
.t r / D jP j0 Pr .x 1 ; : : : ; x p / P ;
P
. s / D jQj1 ˆsQ .x 1 ; : : : ; x p / Q :
5.1 The inverse function theorem 91
are non-singular. We can hence apply the P classical inverse function theorem to find
an open subset W pjq
U pjq
where f0 ; k Bj k j g define a new supercoordinate
i
i Nj
system that we denote by fxN ; g. If we denote by the corresponding change of
coordinates we have
´ P
. B /.t i / D xN i C jP j2 NPi .xN 1 ; : : : ; xN p / N P ;
P
. B /. j / D N j C jQj3
x̂ j .xN 1 ; : : : ; xN p / N Q :
Q
The next proposition clarifies the relation between local superdiffeomorphism and
superdiffeomorphism.
0 0 0 0
Remark 5.2.2. It is important to notice that the constant rank is not defined for all
supermatrices. For example, for the supermatrix Z D 0 00 2 M1j11j1 we cannot
define any constant rank; in fact, as one can readily check, it is not possible with a
basis change, that is, with left and right multiplication by elements of the general linear
supergroup, to transform Z into a diagonal matrix with 1 or 0 on the diagonal, as the
definition of constant rank requires. However, once the constant rank is defined, it
coincides with the rank as we defined it in Chapter 1, Section 1.5. We leave this check
to the reader as an exercise.
We can now give the following definitions.
5.2 Immersions, submersions and the constant rank morphisms 93
y i 7! t i ; j 7! j ; z a ; b 7! 0:
ftN gaD1 , f gj D1 , fN gbD1 around jj.x/ such that the restriction of the map to
a p j n b q
t i 7! t i ; j 7! j ; tNa ; N b 7! 0:
Proof. The proof mimics the classical one. We sketch it briefly. The equivalence
between (1) and (2) is true by definition, while (3) implies (1) comes from our previous
discussion.
Let .W , fy i g, fj g/ and .Q; fx r g; f s g/ be supercharts near x and jj.x/, respec-
r m
tively. Possibly relabeling the supercoordinates we can suppose that @ @y.xi / r;iD1
s n
and @ @ .j / s;j D1 are non-singular at x.
94 5 The local structure of morphisms
and V , ft i gm
iD1 , f gj D1 around jj.x/ such that the restriction of the map to U
j n
Constant rank morphisms. Due to the fact that the definition of the constant rank
morphisms involves the Jacobian matrix rather than the differential, the discussion is
more involved.
Proposition 5.2.6. Let M and N be supermanifolds, dim M D pjq, dim N D mjn.
Suppose that W M ! N is a constant rank morphism of rank ˛jˇ in a neighbor-
hood of x 2 jM j. Then there exist charts U ' U ˛jˇ U p˛jqˇ with coordinates
.y i ; z j ; r ; s / and V ' V ˛jˇ V m˛jnˇ with coordinates .y i ; t k ; r ; l / containing
x and jj.x/, respectively, such that has the form
y i 7! y i ; r 7! r ;
t k 7! 0; l 7! 0:
Proof. Since the statement is local, we can work on superdomains. Suppose that we
have a morphism W U pjq ! V mjn and let .u; /, .v; / be local coordinates in
U pjq and V mjn . By the Chart Theorem we have that is described by the pullbacks
v i ´ .v i / and j ´ . j /. It is not restrictive to suppose that 0 belongs both
to U pjq and to V mjn and that v i .0/ D j .0/ D 0. Since J has constant rank,
possibly relabeling the coordinates, we can assume that the matrices
i ˛ j ˇ
@v @
r
and
@u i;rD1 @ s j;sD1
5.2 Immersions, submersions and the constant rank morphisms 95
are non-singular. In order to keep the notation minimal let w and generically denote
the supercoordinates in a superdomain. Hence the morphism
v i 7! w i ˝ 1; 1 i ˛;
v !
i
7 g .w ˝ 1; 1 ˝ w ;
j 1 j r
˝ 1; 1 ˝ s
/; ˛ C 1 j m;
!
r
7 r
˝ 1; 1 r ˇ;
s 7! s .w 1 ˝ 1; 1 ˝ w j ; r
˝ 1; 1 ˝ s
/1; ˇ C 1 s n:
Notice that correspondingly also the usual form of the matrices in the general linear
supergroup will change. In other words, when we write matrices in GL.V .R//, where
we choose for V a homogeneous basis where the even elements do not all come at the
beginning, then we no longer have matrices with even entries in the diagonal blocks
and odd entries in the off diagonal blocks. So these matrices represent elements in
GL.V .R//, but they are not in GLkjl .R/, for a given commutative superalgebra R.
The claim is now that, in order for the matrix J B 1 to have constant rank ˛jˇ, it
@g @g
is necessary and sufficient that the submatrix @w @
2 @2
@ is zero. Suppose indeed that
@w2 @2
96 5 The local structure of morphisms
I˛;ˇ 0
M ´ is a matrix of constant rank ˛jˇ, where I˛;ˇ is the identity matrix with
A B
˛ C ˇ rows and columns. Hence there exists G and G x such that GM x G D I˛;ˇ 0 . In
0 0
G1 G2
particular there exists G D G3 G4 such that M G has only the first ˛ C ˇ columns
non-zero. In our case we have
G1 G2
MG D :
AG1 C BG3 AG2 C BG4
g i .w i ˝ 1; r
˝ 1/; ˛ C 1 i m;
.w ˝ 1;
j i r
˝ 1/; ˇ C 1 j n:
U3 ´ f.v 1 ; : : : ; v m ; 1 ; : : : ; n / j .v 1 ; : : : ; v ˛ ; 0; : : : ; 0; 1 ; : : : ; ˇ ; 0; : : : ; 0/ 2 U2 g
via
w i ˝ 1 7! v i ; r
˝ 1 7! r ;
1 ˝ w j 7! v ˛Cj g .˛Cj / .v i ; j /; 1˝ s
7! ˇ Cs .ˇ Cs/ .v i ; j /:
1
It is a simple check to show that ' is invertible and that ' B B has the required
form.
1 2
jU W U ! V ! W:
5.3 Submanifolds
In this section we shall use the results on the local structure of morphisms discussed
in the previous section in order to formulate a theory of smooth submanifolds of su-
permanifolds.1 As we stressed from the very beginning, even though many classical
results and constructions carry over to the super setting, the arguments present some ex-
tra subtleties and we invite the reader to go to [54] for some insight on all the problems
that can arise.
As in the classical theory, submanifolds of a given supermanifold M are defined as
pairs .N; j / where N is a supermanifold and j W N ! M is an injective morphism with
some regularity property. We will distinguish two kinds of submanifolds according to
the properties of the morphism j .
Definition 5.3.1. We say that .N; j / is an immersed submanifold if j W N ! M is
an injective immersion, in other words, if jj j W jN j ! jM j is injective and .dj /m is
injective for all m 2 jM j.
As in the ordinary setting, we can strengthen this notion by introducing the notion
of embedding.
Definition 5.3.2. We say that j W N ! M is an embedding if it is an immersion and
if jj j W jN j ! jM j is a homeomorphism onto its image. We say that .N; j / is an
embedded submanifold if j is an embedding. We say that .N; j / is a closed embedded
submanifold if it is an embedded submanifold and jj j.jN j/ is a closed subset of jM j.
Remark 5.3.3. Notice that the morphism j W M z ! M , where M
z is the reduced
manifold associated with M , is a closed embedding.
Closed embedded submanifolds have the following nice characterization in terms
of the properties of the pullback.
Proposition 5.3.4. Let N and M be supermanifolds. A map j W N ! M is a closed
embedding if and only if j W O.M / ! O.N / is surjective.
Proof. Suppose first that j is a closed embedding and let f 2 O.N /. Denote by
fU˛ g˛2A an open covering of jN j. Since jj j is a homeomorphism onto its image,
there exist open sets fU˛0 g˛2A in jM j such that jj j.U˛ / D U˛0 \ jj j.jN j/. Possibly
passing to a refinement of the open cover, it is not restrictive to suppose each U˛0 to
be a superchart with coordinates x i , y j , p , q such that j .y j / D j .q / D 0
(see Proposition 5.2.4). In particular the map jU 0 W OM .U˛0 / ! ON .U˛ / is surjective.
˛
Let g˛ be a section in OM .U˛0 / such that jU 0 .g˛ / D f˛ ´ f jU˛ . Since j.jN j/ is
˛
a closed subset of jM j, we can consider the open cover of jM j given by fU˛0 ; ˛ 2
A; jM j n jj j.jN j/g and let fh˛ g denote a partition of unity subordinated to the given
1
To ease the terminology we prefer to use the term “submanifold” instead of the more appropriate, but
cumbersome “subsupermanifold”.
98 5 The local structure of morphisms
cover. PClearly g˛ h˛ belongs to O.M / and, due to the local finitenessPof the cover,
g ´ ˛ g˛ h˛ is a well-defined section in O.M / so that j .g/ D j .h˛ /f˛ .
On the other hand it is very easy to check that thePfamily fj .h˛ /g defines a partition
of unity on N subordinated to fU˛ g, hence f D j .h˛ /f˛ . So we have proved the
surjectivity of j .
Suppose now that the pullback j W O.M / ! O.N / is surjective. It is immediate
that both (see notation of Chapter 4, Section 4.5)
jj j W jN j ! jM j; .dj /x W Tx N ! Tjj j.x/ M;
are injective. The fact that jj j is a homeomorphism onto its image is a consequence of
the same result being true for the ordinary setting (see [78], Ch. I).
Proposition 5.3.4 characterizes closed embeddings among all possible supermani-
fold morphisms W N ! M in terms of surjectivity of . As a consequence we have
O.N / ' O.M /=JN where JN ´ ker . Hence the ideal JN O.M / completely
characterizes the closed embedded submanifold N . The next proposition singles out
some important properties of JN .
Proposition 5.3.5. Let j W N ! M be a closed embedding and let JN D ker.j /.
Then if m 2 jM j and Jm is the corresponding maximal ideal in O.M /, JN Jm if
and only if m 2 jj j.jN j/. Moreover:
(1) If m 2 jM j is such that JN Jm , then there exist homogeneous f1 ; : : : ; fn in
JN such that Œf1 ; : : : ; Œfn generate JN;m and .df1 /m ; : : : ; .dfn /m are linearly
independent at m, where JN;m denotes the ideal generated by the image of JN
in the stalk OM;m .
(2) If ffi gi2N is a family in JN suchPthat any compact subset of M intersects only
a finite number of supp fi , then i fi belongs to JN .
Proof. We give a sketch leaving the details to the reader. Since j is a closed embedding,
due to Proposition 5.3.4, O.N / ' O.M /=JN so that there is a bijective correspondence
between maximal ideals in O.N / and maximal ideals in O.M / containing JN . We
now consider .1/. Let x 2 N . Since j is an immersion, there exist supercharts
.U; x i ; j / 3 x and .U W; x i ; y r ; j ; s / 3 jj j.x/ such that j is expressed as an
injection, where the coordinates y r and s are sent to zero. Let U 0 W 0 3 x be an open
subset of U W and consider a section g such that gjU 0 W 0 D 1 and gjU W c D 0.
The global section g allows us to extend the local coordinates to obtain global sections;
in fact the global sections fy r g; s gg have the required properties to be the fi ’s.
We now go about the proof of (2). We leave it to the reader to check that if
ffi g are sections in O.M / such P that fsupp fi g is a locally finite covering, then the
possibly infinite sum f D i fi is a well-defined element in O.M /. Now notice
that JN .U / ´ ker jjU j
is an ideal sheaf, hence to check if f 2 JN D JN .jN j/ it
is enough to check that f jUi 2 JN .Ui / for an open cover fUi g of jN j. Since any
point has a compact neighbourhood and since the hypothesis in (2) holds, we obtain
the result.
5.3 Submanifolds 99
Definition 5.3.6. Let I be an ideal in O.M /. I is called a regular ideal if the following
holds.
(1) If m 2 jM j is such that I Jm , then there exist homogeneous f1 ; : : : ; fn
in I such that Œf1 ; : : : ; Œfn generate Im and .df1 /m ; : : : ; .dfn /m are linearly
independent at m, where Im denotes the ideal generated by the image of I in the
stalk OM;m .
(2) If ffi gi2N is a family in I such
Pthat any compact subset of M intersects only a
finite number of supp fi , then i fi belongs to I .
Remark 5.3.7. If the sum ffi gi2N is finite, the second condition is trivial since it is
clear that a finite sum of elements in I still belongs to the ideal I . When the sum
is infinite, the sum is still well defined since at any point we are summing a finite
number of fi ’s. However the fi ’s that we are summing can vary from point to point
and consequently in general we cannot assume that their sum still lies in I .
The next proposition says that each regular ideal I in O.M / is of the form JN for
a closed embedded subsupermanifold N . This is a converse to Proposition 5.3.5.
Proposition 5.3.8. Let M be a supermanifold and suppose that J is a regular ideal in
O.M /. Then there exists a unique closed embedded supermanifold .N; j / such that J
is the associated regular ideal, i.e., O.N / D O.M /=J .
Proof. Define the set
jN j ´ fx 2 jM j j J Jx g:
T
Clearly jN j D fjf j1 .0/ j f 2 J g so that jN j is closed in jM j. Since J is a regular
ideal, for each x 2 jN j there exist homogeneous sections f 1 ; : : : ; f p ; 1 ; : : : ; q in
J , depending on x, of even and odd parity, respectively, such that f.df i /x , .d j /x g
are linearly independent and Œf i , Œ j generate the space of germs Jx . Because of the
inverse function theorem, there exist sections y r , s in O.M / such that ff i ; y r ; j ; s g
is a coordinate system in a neighborhood U x of x in jM j.
Define now U 0x ´ U x \ jN j and
O.U 0x / ´ C 1 .y 1 ; : : : ; y n / ˝ ƒ.1 ; : : : ; m /:
U 0x ´ fz 2 jM j j f i .z/ D 0g:
ˇx W OM .U x / ! O.U 0x /; f i ; j 7! 0; y r 7! y r ; s 7! s :
100 5 The local structure of morphisms
We can thus define transition functions N xy between the various O.U x0 / through the
diagram
N
O.U 0x / _ _ _ _/ O.U 0y /
xy
O
˛x ˇy
OM .U x /
xy
/ OM .U y /.
It is evident that N xx D id and it is easy to prove that they obey the other glueing
conditions as specified in Chapter 2, Section 2.2. We call N the supermanifold obtained
in this way. It is clear that it is a closed embedded submanifold.
Proof. The correspondence between (1) and (2) is given by Propositions 5.3.8 and
5.3.5. As for (2) and (3), it is clear that if J is a quasi-coherent ideal sheaf, then its
global sections define a regular ideal. Vice versa if I is a regular ideal in O.M /, then for
each point m 2 jM j there exists an open set U containing x and homogeneous sections
.y i ; s / in OM .U / that generate a regular ideal I.U / in OM .U /. These sections are
those computed during the proof of Proposition 5.3.5. The assignment U 7! I.U / is
a sheaf of ideals on jM j.
The next concept we want to introduce is transversality. The main way to define a
submanifold of a given manifold is through equations. We want to give a criterion to
establish when we can give a supermanifold structure to the set of points which are the
zeros of a certain set of equations.
then
Tm M D .d1 /n1 Tn1 N1 C .d2 /n2 Tn2 N2 : (5.1)
The maps 1 and 2 are then said to be transversal at m.
Notice in particular that if m is not in j1 j.jN1 j/ or not in j2 j.jN2 j/, then 1 and
2 are automatically transversal at m. Notice also that the sum in equation (5.1) needs
not to be direct.
Proposition 5.3.12. Let W L ! M be a supermanifold morphism and let .N; j / be
a closed embedded submanifold of M , O.N / D O.M /=JN . Suppose that and j
are transversal and denote by J the ideal in O.L/ generated by .JN /. Then J is a
regular ideal corresponding to a closed submanifold N 0 L.
Proof. Let us first notice that x 2 jLj is such that J Jx if and only if jj.x/ 2 jN j.
Indeed .f /.x/ D 0 for all f 2 JN if and only if f .jj.x// D 0 for all f 2 JN . If
x 2 jN j then there exists a superchart U; x i ; xN j ; r ; N q such that N \ U is determined
by xN j D N q D 0. By using an appropriate global section, we can suppose that xN j ; N q
are defined on the whole M . We claim that .xN j /; .N q / satisfy the properties
of .1/ of Proposition 5.3.5. They are clearly in J and they also generate Jx . It
remains only to prove that their differentials are linearly independent, but this is an
easy consequence
@ of the transversality
P condition. Indeed let Xi 2 Tx L be such that
i
.d/x Xi D @xN i j j.x/ . If ci .d .xN // D 0 for some non-null sequence of numbers
P P P @
.ci / then hXp ; ci .d .xN i //i D ci h .Xp /; xN i i D ci h @xN p j j.x/ ; x i i D 0,
which is clearly impossible.
Proposition 5.3.13. Let W L ! M be a supermanifold morphism and suppose that
m 2 jM j. Suppose that for each x in jj1 .m/ there exists a neighborhood where
is a constant rank morphism.
(1) If J denotes the ideal in O.L/ generated by .Jm /, then J distinguishes a
closed supersubmanifold L0 of L (Jm the ideal in O.M / corresponding to the point m).
(2) If .L0 ; j / denotes the closed embedded submanifold distinguished by J, then
Tx L0 ' ker.d/x
X i; j D 1; : : : ; m;
Ber D det.xkl /1 det.xij ik x
kl
lj /;
k; l D m C 1; : : : ; m C n;
k;l
and x kl denotes the element of the inverse of the matrix .xkl /. One can readily check
that this is a submersion, hence a constant rank morphism at all points of jj1 .1/. By
Proposition 5.3.13 we have that jj1 .1/ has a supermanifold structure and one can
readily check that this is SLmjn , the supermanifold whose T -points are the matrices in
GLmjn .T / with Berezinian equal to 1.
The differential at the identity .d/id can be calculated and it is the morphism
5.4 References
With respect to the classical material we refer the reader to [47], [1], [78] for an
exhaustive treatment. Almost all the supergeometry material presented in this chapter
is a reworking of [53]. More recent treatments of the same material can be found in
[56], [76].
6
The Frobenius theorem
Proof. Let x 2 jM j. We set ourselves up to use the previous result, namely we will find
vector fields fXi ; ‚j g which generate D in an open neighborhood U of x so that the
vectors corresponding to evaluating fXi ; ‚j g at x are a basis for Tx .M /, the tangent
space of M at x. Lemma 6.1.2 then tells us that fXi ; ‚j g are in fact OM;x -linearly
independent in VecM;x .
Let fXi ; ‚j g be vector fields on Dx defined on an open neighborhood U of x so
that their tangent vectors at x are a basis for Dx =mx Dx Tx .M /. Because D is
a distribution, it is locally free, and there exists D VecM .U / so that VecM;y D
Dy ˚ Dy for all y 2 U .
Now let fYk ; „l g 2 Dx defined on U be such that their tangent vectors at x are a
basis for Dx =mx Dx Tx .M /. Since Tx .M / D Dx =mx Dx ˚Dx =mx Dx , it follows
that fXi ; Yk ; ‚j ; „l g evaluated at x is a basis for Tx .M /. In fact, this is true for all
points y 2 U by construction, hence by Lemma 6.1.2 they are linearly independent.
To see that fXi ; ‚j g generate D, let Dy0 D hfXi ; ‚j gi be the submodule of VecM;y
generated by these vector fields. Similarly let Dy0 D hfYk ; „l gi. Then Dy0 Dy and
Dy0 Dy , but we also know that Dy0 ˚ Dy0 D VecM;y D Dy ˚ Dy , hence the
generation.
This proves also that dim.Dy =my Dy / is invariant for y 2 U , if M is connected.
The next definitions are crucial for the statement of the local Frobenius theorem.
Definition 6.1.4. Let M D .jM j; OM / be a connected supermanifold, that is, jM j is
connected, and let D be a distribution on M . We define the rank of a distribution as
rank.D/ ´ dim.Dx =mx Dx /:
The previous lemma ensures the rank is well defined (see also Appendix B, Sec-
tion B.3).
Definition 6.1.5. We say that a distribution D is involutive if it is stable under the
bracket on VecM , i.e., for vector fields X , Y in D the bracket ŒX; Y is also a vector
field in D.
Definition 6.1.6. We say that a distribution D is integrable if, for any x 2 jM j, there
exists (locally) a submanifold N of M whose tangent bundle can be identified locally
with D, that is,
VecN jU D DjU ; x 2 U;
for a suitable neighbourhood U of x.
It is clear that an integral distribution is involutive; the Frobenius theorem in its
local and global versions establishes a converse for this fact.
Lemma 6.1.3 and the definition of rank of a distribution allow us to make some
crucial change of coordinates calculations at the coordinate chart level to shape an
involutive distribution. We now prove a series of lemmas which will demonstrate, by
construction, the local Frobenius theorem on supermanifolds.
6.1 The local super Frobenius theorem 105
Remark 6.1.7. The following lemmas pertain to the local Frobenius theorem and are
local results, thus it suffices to consider the case M D Rpjq in a coordinate neighbor-
hood of the origin, i.e., U pjq .
Next is an explicit local calculation of the previous lemma.
Lemma 6.1.8. Let D be an involutive distribution. Then there exist linearly indepen-
dent supercommuting vector fields which span D.
Proof. Let fX1 ; : : : ; Xr ; 1 ; : : : ; s g be a basis for D, in other words, fX1;x ; : : : ; Xr;x ;
1;x ; : : : ; s;x g form a basis for the free OM;x -module Dx . Let .t; / D .t 1 ; : : : ; t p ,
1 ; : : : ; q / be a local set of coordinates. Then we can express the vector fields
X @ X @ X @ X @
Xj D aj i i
C ˛j l l ; k D ˇki i
C bkl l : (6.1)
@t @ @t @
i l i l
with entries in OM .U / (for a suitable domain U ) and whose rows are the vector fields
X1 ; : : : ; Xr , 1 ; : : : ; s generating the distribution D. T has rank rjs since the fXi ; j g
are linearly independent. This is to say that the submatrix .a/ has rank r and rank.b/
= s. Then by renumerating coordinates .t; /, we may assume that
T D .T0 j /;
@ Xp
@ X
q
@
Xj D C aji C ˛j l l ;
@t j @t i @
iDrC1 lDsC1
(6.2)
@ X
q
@ Xp
@
k D k C bkl l C ˇj i i :
@ @ @t
lDsC1 iDrC1
X
r X
s
ŒXj ; Xk D fi Xi C 'l ; l
iD1 lD1
106 6 The Frobenius theorem
where the fi are even functions and the 'l are odd functions. Then, by (6.2), fi is
the coefficient of the @t@i term in the vector field ŒXj ; Xk . However, again by (6.2), it
is clear that ŒXj ; Xk has only @t@i terms for i > r, and so we have fi D 0 for all i .
Similarly, ŒXj ; Xk has only @@ l terms for l > s, hence also 'l D 0 for all l.
The cases ŒXj ; k D 0 and Œl ; k D 0 follow by using the same argument above.
We now prove the local super Frobenius theorem in the case of an even distribution
of rank 1j0.
Lemma 6.1.9. Let X be an even vector field. There exist local coordinates so that
@
XD :
@t 1
Proof. Let J be the ideal generated by the odd functions on Rpjq . Then, since X
is even, X maps J to itself. Thus X induces a vector field, and hence an integrable
distribution, on the reduced space Rp . Then we may apply the classical Frobenius
theorem to get a coordinate system where X D @t@1 (mod J). So we may assume that
@ X @ X @
XD 1
C ai i C ˛j j ;
@t @t @
i2 j
where the ai are even, the ˛j are odd, and ai ; ˛j 2 J. That the ai are even
P implies that
ai 2 J 2 . Moreover, we can find an even matrix .bj k / so that ˛j D k bj k k (mod
J 2 ), and so modulo J 2 we have
@ X @
XD 1
C bj k k j :
@t @
j;k
and we choose g.t / so that it satisfies the matrix differential equation and initial con-
dition
@g
D gb; g.0/ D I:
@t 1
By (6.3) we may then assume that, modulo J 2 ,
@
XD :
@y 1
6.1 The local super Frobenius theorem 107
Next we claim that if X D @t@1 (mod J k ), then X D @t@1 (mod J kC1 ). Since J is
nilpotent, this claim will imply the result for the 1j0-case.
Again, let .t; / 7! .y; / be a change of coordinates so that y i D t i C ci and
D j C j for ci ; j 2 J k suitably chosen in such a way that they depend on t1
j
@ X @ X @
XD 1
C hi i
C 'u u
@t @t u
@
i2
@ci @l
D hi ; D 'l ;
@t 1 @t 1
X
p
@ X q
@
Xr D fi C 'k k
@t i @
iD1 kD1
for some even functions fi and odd functions 'k . The assumption ŒXr ; Xj D 0 gives
X @ X @ X @ X @
fi i
; Xj C 'k k
; Xj .Xj fi / i
.Xj 'k / k D 0:
@t @ @t @
108 6 The Frobenius theorem
We know that Œ @t@i ; Xj is a linear combination of @t@l for l < r, which means that
Xj fi D 0 for all j > r 1. Because the coefficients of the Xj are “upper triangular”
for j r 1, we see that fi depends only on .t r ; : : : ; t p ; 1 ; : : : ; q / for i r. We
also have Œ @ @k ; Xj D 0 for all k, and so Xj 'k D 0 for all j as well. We can then
again conclude that the 'k depend only on .t r ; : : : ; t p ; 1 ; : : : ; q / as well.
Now we can rewrite Xr as follows:
X
r1
@ X
p
@ Xq
@
Xr D fi i C fl l C 'k k :
@t @t @
iD1 lDr kD1
„ ƒ‚ …
D Xr0
@ Xr1
@
Xr D r
C fi0 i
@t @t
iD1
(where the fi0 are the fi above under the change of coordinates prescribed by Lem-
ma 6.1.9).
In fact, the above lemma proves the local Frobenius theorem in the case when D
is a purely even distribution (i.e., of rank rj0). For the most general case we need
one more lemma, which establishes the local Frobenius theorem in the case of an odd
distribution of rank 0j1.
Lemma 6.1.11. Say is an odd vector field so that 2 D 0 and that spanfg is a
distribution. Then there exist coordinates so that locally D @ @1 .
Proof. As we have previously remarked, since we want a local result, it suffices to
assume that is a vector field on Rpjq near the origin. Let us say .y; / are coordinates
on Rpjq . Then
X @ X @
D ˛i .y; / i C aj .y; / j ;
@y @
i j
where the ˛i are odd, the aj are even, and we may assume that a1 .0/ ¤ 0.
Now consider the map
@
D C Z
@
for some even vector field Z. Recall that 1 D aO 1 where aO 1 is an even invertible
section. Hence D 1 A from some invertible even section A.
Then we see that under the change of coordinates given by ,
@
1 A Z D ;
„ƒ‚…
@ 0
DZ
where Z denotes the pullback of Z by and Z 0 is some even vector field (since both
A and Z are even). Now,
2 D 0 H) . @
@
1 Z 0 /2 D 0
H) . @
@ 2
/ @ .1 Z 0 / .1 Z 0 / @
@
C .1 Z 0 /2 D 0
„ƒ‚… @ „ ƒ‚ …
D0 D0
0
H) aO 1 Z C 1 @
@
Z0 1
Z 0 @
@
D0
H) aO 1 Z 0 D 0
H) Z 0 D 0;
so we really have @
@
D under the change of coordinates.
Now we can prove the full local Frobenius theorem.
110 6 The Frobenius theorem
@ @ @ @
;:::; r ; ;:::; s:
@t 1 @t @ 1 @
Proof. Let fX1 ; : : : ; Xr ; 1 ; : : : ; s g be a basis of vector fields for the distribution D.
By Lemma 6.1.8 we may assume that these basis elements supercommute, so then
D 0 D spanfX1 ; : : : ; Xr g is a subdistribution, and by Lemma 6.1.10 we get that there
exist coordinates so that Xi D @t@i .
We then use the fact that Œ1 ; Xi D 0 for all i to see that 1 depends only on
coordinates .t rC1 ; : : : ; q / (as in the proof of Lemma 6.1.10). In fact, this is not
completely accurate. If we express 1 as in (6.1), we see that it is only the ˇik and
blk which depend only on the coordinates .t rC1 ; : : : ; q /. However, we can always
kill off the first r @=@t i terms by subtracting appropriate linear combinations of the
fX1 D @=@t 1 ; : : : ; Xr D @=@t r g.
Since 21 D 0, by Lemma 6.1.11, we may change only the coordinates .t rC1 ; : : : ; q /
and express 1 D @ @1 . For 2 we apply the same idea: that Œ2 ; Xi D 0 and Œ2 ; 1 D
0 again shows that 2 depends only on coordinates .t rC1 ; : : : ; t p ; 2 ; : : : ; q /, and ap-
plying Lemma 6.1.11 once again gives 2 D @ @2 . The same argument then applies
successively to 3 ; : : : ; s .
We are now in a position to state and prove the global Frobenius theorem on super-
manifolds.
˚@
of @t
; @@ . Now let U jM j and define the following presheaf on jM j:
z x \ U there exists
.U / D hff 2 OM .U / j for all y 2 M
Vy U such that f jVy 2 C 1 .z/Œgi:
6.3 References
The local Frobenius theorem for even distributions was proved by Giachetti–Ricci in
[35], while the version for odd distributions was considered in [50] by Koszul. The
first proof of the local and global super Frobenius theorem we are aware of is the one
contained in [19]. For other more modern treatments we refer the reader to [56], [22],
[76].
7
Super Lie groups
In this chapter we want to take a closer look at supermanifolds with a group structure:
Lie supergroups or super Lie groups. As in the ordinary setting, a super Lie group is
defined as a supermanifold together with the multiplication and inverse morphisms,
that satisfy the usual properties expressed in terms of certain commutative diagrams.
To any Lie supergroup, we can naturally associate a Lie superalgebra, consisting of the
left-invariant vector fields, As in the ordinary setting, the Lie superalgebra is identified
with the tangent superspace to the supergroup at the identity.
We can equivalently approach this theory using the language of the super Harish-
Chandra pairs (SHCP for short), which is historically how it was originally developed
by Kostant and Koszul [49], [50]. The SHCP allows us to recover the supergroup
structure of a Lie supergroup G only by knowing its reduced ordinary Lie group G z and
its associated Lie superalgebra g. In fact an even stronger statement is true: there is
an equivalence of categories between the SHCP and the super Lie groups. As we are
going to see in the next chapter, this equivalence extends to the categories of actions
of the SHCP and the super Lie groups, respectively.
W G G ! G; i W G ! G; e W R0j0 ! G;
called multiplication, inverse, and unit respectively satisfying the commutative dia-
grams
id
GGG /GG G; GG G; GG
O www
hidG ;ei GG hidG ;ii www GG
w GG w GG
w GG w GG
ww # ww eO #
G GG
idG
/ G, / G,
id
w ; G GG w ;
GG ww GG ww
GG ww GG ww
O G i G#
he;id ww hi;idG i G# ww
GG / G, GG GG
7.1 Super Lie groups 113
where eO denotes the composition of the identity e W R0j0 ! G with the unique map
G ! R0j0 . Moreover, h ; i denotes the map . / B dG , dG W G ! G G being
the canonical diagonal map.
We may of course interpret all these maps and diagrams in the language of T -points,
which gives us (for any supermanifold T ) morphisms T W G.T / G.T / ! G.T /,
etc. that obey the same commutative diagrams. In other words, if G is an SLG then
the functor T ! G.T / D Hom.T; G/ takes values in the category of set theoretical
groups. Conversely, Yoneda’s lemma says that if the functor T ! G.T / takes values
in the category of set theoretical groups, then G is actually a super Lie group. This
leads us to an alternative definition of a super Lie group.
Definition 7.1.2. A supermanifold G is a super Lie group if for any supermanifold T ,
G.T / is a group, and for any supermanifold S and morphism T ! S, the corresponding
map G.S / ! G.T / is a group homomorphism.
In other words, a supermanifold G is a super Lie group if and only if its functor of
points T 7! G.T / is a functor into the category of groups.
Remark 7.1.3. Let us notice that to each super Lie group G is associated a Lie group
z It is defined as the underlying manifold G
G. z with the “reduced morphisms”
z G
j jW G z ! G;
z z ! G;
jij W G z z
jej W R0 ! G:
.t; / .t 0 ; 0 / D .t C t 0 C 0 ; C 0 / (7.1)
where the coordinates .t; / and .t 0 ; 0 / represent two distinct T -points for some su-
permanifold T . It is then clear by the formula (7.1) that the group axioms are satisfied.
We are going to return with more details in Example 7.2.4.
Remark 7.1.5. Notice that the properties required in Definition 7.1.1 translate into
properties of the morphisms on the global sections, W O.G/ ! O.G G/,
i W O.G/ ! O.G/, that make O.G/ “almost” a Hopf superalgebra (see Chapter 1,
Section 1.7). One word of caution: since O.G G/ Š O.G/ Ő O.G/, strictly speaking
O.G/ is not a Hopf superalgebra but a topological Hopf superalgebra, meaning that
since we are taking a completion of the tensor product, we are allowed to consider
infinite sums (for a definition of topological Hopf algebras see [20]). In fact an SLG
114 7 Super Lie groups
d `h B X D X B `h :
We want to interpret this in the super category by saying that a left-invariant vector
field on G is invariant with respect to the group law “on the left”.
Let G be a super Lie group with group law W G G ! G and let us denote by 1
the identity at the level of sheaf morphisms.
Let VecG denote the vector fields on G (see Chapter 4, Section 4.4).
Since the bracket of left-invariant vector fields is left-invariant, as one can readily
check, the left-invariant vector fields are a super Lie subalgebra of VecG , which we
denote by g.
is the super Lie algebra associated with the super Lie group G, and we write g D Lie.G/
as usual.
Proof. To prove (1) for the left-invariant vector fields, we need to show that
Œ1 ˝ ..1 ˝ Xe / / B D B Œ.1 ˝ Xe / :
This is a simple check that uses the coassociativity of , that is, .1 ˝ / D
. ˝ 1/ . In fact
Œ1 ˝ .1 ˝ Xe / B B D .1 ˝ 1 ˝ Xe /Œ.1 ˝ / B
D .1 ˝ 1 ˝ Xe /Œ. ˝ 1/ B
D B Œ.1 ˝ Xe / :
As for (2) we notice that the injectivity of the map (7.2) is immediate. Let us
thus focus on the surjectivity. Suppose that X is a left-invariant vector field, i.e.,
.1 ˝ X/ D X . Apply 1 ˝ e to this equality to obtain
.1 ˝ e /.1 ˝ X /
D .1 ˝ e /
X;
from which we get X D .1 ˝ Xe / since .1 ˝ e /
D 1. So we are done.
We can use the previous proposition to endow Te G with the structure of a super
Lie algebra and to identify it with g. From now on we shall use such an identification
freely without an explicit mention.
Example 7.2.4. We want to calculate the left-invariant vector fields on R1j1 by means
of the group law (7.1),
.t; / .t 0 ; 0 / D .t C t 0 C 0 ; C 0 /;
from Example 7.1.4. In terms of W O.R1j1 / ! O.R1j1 R1j1 / the group law reads
.t / D t ˝ 1 C 1 ˝ t C ˝ ; . / D ˝ 1 C 1 ˝ :
@ @ @
XD ; Y D C : (7.4)
@t @t @
A quick check using the definition shows that (7.4) are in fact left-invariant. The
Lie superalgebra structure can now be easily computed: Lie.R1j1 / D spanfX; Y g with
brackets
ŒX; X D 0; ŒX; Y D 0; ŒY; Y D 2X:
Proposition 7.2.5. Let G and H be super Lie groups and let W G ! H be a morphism
of super Lie groups. The map
.d/e W g ! h
is a super Lie algebra homomorphism.
Proof. The only thing to check is that .d/e preserves the super Lie bracket. We
leave this to the reader as an easy exercise, recalling that the bracket has always to be
computed on the left-invariant vector fields.
Corollary 7.2.6. The even part of the super Lie algebra Lie.G/ canonically identifies
z
with the ordinary Lie algebra Lie.G/.
z ! G and the
Proof. This is immediate considering the canonical inclusion j W G
previous proposition.
z acts on G in a natural
We end this section showing that the reduced Lie group G
way.
aW G M ! M
such that
(1) eT m D m,
(2) g1 .g2 m/ D .g1 g2 / m.
In other words,
O 1M i D 1M ;
a B he; (7.5a)
a B . 1M / D a B .1G a/; (7.5b)
118 7 Super Lie groups
a B . 1M / D a B .1G a/ B . 1M /;
lgN f D .evgN ˝ 1/ B
.f /; a f D .j ˝ 1/ B
.f /;
rgN f D .1 ˝ evgN / B
.f /:
h .Xg Yh /; s ˝ ti D hXg ˝ Yh ; .s t /i
D hXg ˝ Yh ; .s/ .t /i
D h Xg Yh ; s ˝ t i:
The fact that S has the property of the antipode is a similar check.
We end this section by pointing out an interesting decomposition of O.G/ı which
parallels a general structure theorem for Hopf algebras (see again [72] for the classical
case).
Let us examine more closely the product law in O.G/ı . Fix a distribution Xg in
O.G/ı . Each such element can be canonically written as the convolution product of a
distribution at the identity e of the super Lie group with the “element” g of jGj:
Thus each element in O.G/ı can be written as a product evg Xe , where Xe 2 O.G/ı
denotes the distribution at e given by
Xe ´ .dlg 1 /g Xg :
Notice that here Xe is used in a more general sense than in Section 11.2.4, where it
denoted an element in Te G.
The next proposition characterizes the distributions with support at e.
Proposition 7.3.6. The space of distributions supported at the identity e of the super
Lie group G is isomorphic, as a super Hopf algebra, to the enveloping algebra U.g/
of the super Lie algebra g associated with the supergroup G.
7.3 The Hopf superalgebra of distributions 121
Proof. It goes exactly as in the classical case, but we shall nevertheless briefly outline
it to make the text self-contained. As we have seen in Chapter 4, Section 4.7, the
finitely supported distributions can be identified with differential operators. Hence a
distribution Xe supported at the identity can be identified with a derivation of type
X ˇ ˇ
@ ˇˇ @ ˇˇ
Xe D cI;J I ˇ ;
@x e @ J ˇe
I;J
where .xi ; j / are local coordinates around the identity e 2 jGj and I and J are multi-
indices. Every distribution Xe described above corresponds to a unique left-invariant
differential operator via the linear map
Xe 7! .1 ˝ Xe / ;
very much in the same way as tangent vectors at the identity correspond to left-invariant
vector fields. The linear map described above is a linear isomorphism; the proof
of this fact resembles very closely what we did in Proposition 7.2.3. Finally a left-
invariant differential operator can be uniquely identified with an element of the universal
enveloping superalgebra of g, a conclusion from dimension considerations and using
the PBW theorem.
Using decomposition (7.6), we can now define a linear map
z ˝ U.g/;
W O.G/ı ! RG Xg 7! g ˝ .dlg 1 /g Xg ;
where we notice that .dlg 1 /g Xg is in fact a distribution supported at the identity, hence
identified with an element in U.g/ (see previous proposition). RG z denotes the group
algebra of Gz and consists of the formal finite sums of elements in G z with coefficients
in R. By Proposition 7.3.6 and equation (7.6), the above map is a bijection.
All the algebraic structures defined over O.G/ı can thus be transported to corre-
sponding structures over RG z ˝ U.g/, as the next proposition formalizes.
Proposition 7.3.7. RG z ˝ U.g/ is a super Hopf algebra. The coalgebra structure is
z and U.g/, namely
the one induced by the coalgebra structures of RG
zW
RG g 7! g ˝ g; gW Xe 7! Xe ˝ 1 C 1 ˝ Xe for all Xe 2 g;
while the algebra structure is given by
.g ˝ X /.gN ˝ Y / D .g gN ˝ .gN 1 X /Y /; (7.7)
where gN 1 X ´ evgN 1 X evgN .
Proof. For the coalgebra structure it is enough to check on the generators:
.g ˝ X / D .g/ .X /
D .g ˝ g/.1 ˝ X C X ˝ 1/
D .g ˝ 1/ ˝ .g ˝ X / C .g ˝ X / ˝ .g ˝ 1/:
122 7 Super Lie groups
Notice that
1
. .g ˝ X // D .evg 1/ ˝ .evg X / C .evg X / ˝ .evg 1/:
.g ˝ X /.gN ˝ Y / ´ B Œ 1 .g ˝ X / 1 .gN ˝ Y /
D .Xg YgN /
D .evg X evgN Y /
D .evg gN evgN 1 X evgN Y /
D .evg gN ..gN 1 X / Y //;
.g ˝ X /.gN ˝ Y / D .g gN ˝ .gN 1 X /Y /:
Notice that we have defined a morphism that to each g 2 jGj associates the super
linear morphism X ! gX. This is of fundamental importance and in the ordinary case
z on U.g/. We are thus led to the following
reduces to the adjoint representation of G
definition.
Definition 7.3.8. Let G be a super Lie group and g the corresponding super Lie algebra.
The morphism
z ! Aut.g/;
Ad W G Ad.g/X ´ .evg ˝ X ˝ evg 1 /.1 ˝
/
;
cg .f / D .g ˝ 1 ˝ g 1 /.
˝ 1/
.f /;
Proposition 7.3.10. Let the notation be as above. Then d Ad D adjg0 , in other words,
Since formula (7.7) resembles the one for semi-direct products, it is customary
z ˝ U.g/, endowed with the Hopf algebra structure just
in the literature to denote RG
described, by
RG z Ì U.g/: (7.8)
This is also called the smash product of RG z with U.g/, but we shall not define smash
products in general, referring the interested reader to [72] for more details.
So far we have associated with each super Lie group G the super Hopf algebra
O.G/ı of finite support distributions over G, and we have seen that it can be written as
z Ì U.g/:
O.G/ı ' RG
Definition 7.4.1. Suppose that .G0 ; g/ are respectively a Lie group and a super Lie
algebra such that
(1) g0 ' Lie.G0 /,
(2) G0 acts on g via a representation such that .G0 /jg0 D Ad and the differential
of acts on g as the adjoint representation, that is,
d .X /Y D ŒX; Y :
Then the pair .G0 ; g/ is called a a super Harish-Chandra pair (SHCP). We shall
write .G0 ; g; / when we want to stress the action .
Definition 7.4.2. Let .G0 ; g; / and .H0 ; h; / be SHCP. A morphism between them
is a pair . 0 ; / such that
(1) 0W G0 ! H0 is a Lie groups homomorphism,
(2) W g ! h is a super Lie algebra homomorphism,
(3) 0 and are compatible in the sense that
Example 7.4.3. If G is a super Lie group, the pair .G; z g/ given by the reduced Lie
group of G and the super Lie algebra g is a super Harish-Chandra pair with respect to the
z on g as defined in Definition 7.3.8. Moreover, given a morphism
adjoint action of G
W G ! H of super Lie groups, determines the morphism of the corresponding
super Harish-Chandra pairs
.jj; .d/e /:
In the next observation we relate the SHCP with the Kostant construction described
in the previous section.
Observation 7.4.4. Given an SHCP pair, we can construct the super Hopf algebra
given by the semidirect product RG0 Ì U.g/ in which G0 is endowed with a Lie group
structure and the product is given by
from the category of super Lie groups to the category of super Harish-Chandra pairs.
The most important result of this chapter is the following:
Theorem 7.4.5. The category of super Lie groups is equivalent to the category of
super Harish-Chandra pairs.
Roughly speaking this theorem says that each problem in the category of super
Lie groups can be reformulated as an equivalent problem in the language of SHCP.
Before embarking on the proof, let us outline the path that we shall follow. We show
the following.
(i) Given a SHCP .G0 ; g/ there exists a super Lie group G whose associated SHCP
is isomorphic to .G0 ; g/.
(ii) Given a morphism of SHCP . 0 ; / W .G0 ; g/ ! .H0 ; h/ there exists a unique
morphism of the corresponding super Lie groups from which . 0 ; / arises.
iii) Due to points (i) and (ii) we have a functor
K W .shcps/ ! SGrp:
In order to prove the theorem we have to show that K B H ' 1SGrp and H B K '
1.shcps/ . This means that, for each G 2 SGrp and .G0 ; g/ 2 .shcps/, .K B H /.G/ ' G
and .H B K/..G0 ; g// ' .G0 ; g/, and moreover the diagrams
/G / .G0 ; g/
.K B H /.G/ .H B K/..G0 ; g//
.KBH /. / .H BK/./ (7.9)
.K B H /.H / / H, .H B K/..H0 ; h// / .H0 ; h/
Lemma 7.4.7. OG , together with the restriction maps (7.11), is a sheaf of superalge-
bras.
Proof. The fact that OG is a presheaf is a routine check. As for the glueing property,
let fU˛ g be open sets covering a fixed open set U and f˛ elements in OG .U˛ / such
that f˛ jU˛ \Uˇ D fˇ jU˛ \Uˇ for each ˛, ˇ. We want to define f 2 OG .U /. For
each X 2 U.g/, the f˛ .X / 2 C 1 .U˛ / glue together to give gX 2 C 1 .U /. Define
f .X/ D gX . Then U.g0 /-linearity is immediate.
Lemma 7.4.8. The antisymmetrizer
1 X
W ƒ.g1 / ! U.g/; X1 ^ ^ Xp 7! .1/jj X.1/ X.p/ ;
pŠ
2Sp
Proof. Both assertions are consequences of the PBW theorem and we leave their proofs
to the reader.
Proposition 7.4.9. .G0 ; OG / is a supermanifold that is globally split, i.e., for each
open subset U G0 there is an isomorphism of superalgebras
V V
OG .U / ' Hom. .g1 /; C 1 .U // ' C 1 .U / ˝ .g1 / :
Proof. In view of Lemma 7.4.7, it only remains to prove the local triviality of the sheaf.
For this purpose we define the map
V
U W OG .U / ! Hom. .g1 /; C 1 .U //; f ! f B :
U .f1 f2 / D mBf1 ˝f2 B U.g/ B D mBf1 ˝f2 B. ˝ / U.g/ D U .f1 /U .f2 /:
So far we have used only the fact that CG10 .U / is a left U.g0 /-module.
The next proposition uses more heavily the structure of G0 and the representation
. It exhibits explicitly the structure of a super Lie group in terms of the corresponding
SHCP.
Proposition 7.4.10. .G0 ; OG / is a super Lie group with respect to the morphisms
multiplication W G G ! G, inverse i W G ! G and unit e W k ! G, which are
defined in terms of their pullbacks , i , e as
Œ .f /.X; Y /.g; h/ D Œf ..h1 X /Y /.gh/;
Œi .f /.X / .g 1 / D Œf .g 1 Xx /.g/; (7.12)
e .f / D Œf .1/.e/;
.e ˝ 1/
.f / D .1 ˝ e /
.f / D f:
7.4 Super Harish-Chandra pairs 129
Œ.m B .1 ˝ i / B
/.f /.X /.g/ D Œ .f /.1 ˝ .g//.1 ˝ SU.g0 / / .X /.g; g 1 /
D Œf .m. .g/ ˝ .g//.1 ˝ SU.g0 / / .X //.e/
D Œf . .g/m..1 ˝ SU.g0 / / .X //.e/
D Œf .g/.X /.e/:
(4) We first check that the vector field .DXL f /.Y / ´ .1/jXj f .YX / is left-
invariant, that is, .1 ˝ DXL / D DXL :
Œ.1 ˝ DXL / .f /.Y; Z/.g; h/ D .1/jXj
.f /.Y; ZX /.g; h/
jXj
D .1/ f .h1 Y ZX /.gh/
and
Œ DXL f .Y; Z/.g; h/ D DXL .f /.h1 Y Z/.gh/
D .1/jXj f .h1 Y ZX /.gh/:
Hence we have a well-defined linear map from the Lie superalgebra g and the left-
invariant vector fields X 7! DXL . In order to show that it is an isomorphism it is
enough to prove injectivity for dimension considerations. If X ¤ Y are two elements
in g, with DXL .f / D DYL .f / for all f , then f .X / D f .Y / for all X and Y , reaching
a contradiction.
(5) For right-invariant vector fields the arguments are very similar and the proofs
are left to the reader.
7.4 Super Harish-Chandra pairs 131
Notice that we have introduced the notation DXL , DXR for the actions of left and right-
invariant vector fields, which is different from the one used in the previous sections.
With this approach it is very natural to recover a super Lie group morphism from
an SHCP one. Suppose that . 0 ; / is a morphism from .G0 ; g/ to .H0 ; h/, and that
f 2 OH .U /. Define .f / via the diagram
U.g/ / U.h/
.f / f
CG0 . 0 .U // o
1 1
CH10 .U /.
0
It is not difficult to prove that this defines a super Lie group morphism with associated
SHCP morphism . 0 ; /. Indeed we have the following proposition.
Proposition 7.4.12. The map
W HomU.h0 / .U.h/; C 1 .H0 // ! HomU.g0 / .U.g/; C 1 .G0 //;
f 7! .f / ´ 0 B f B ;
defines a morphism of super Lie groups whose reduced morphism is 0 and whose
differential at the identity is .
Proof. It is immediate that is well defined. We check that the above defined map is
a map of superalgebras and that it is a super Lie group morphism. It is a superalgebra
morphism since, by Remark 7.4.6,
.f1 f2 /.X /.g/ ´ ŒmC 1 .G0 / .f1 ˝ f2 / .X /. 0 .g//
D Œ.f1 ˝ f2 /. ˝ / .X /. 0 .g/; 0 .g//
D Œ .f1 / .f2 /.X /.g/:
and
i .
.f //.X /.g 1 / D f . .g 1 Xx //. 0 .g//
D f. 0 .g/
1 x
.X //. 0 .g//:
We can finally end the proof of Theorem 7.4.5, proving the following proposition.
Proposition 7.4.13. Let the notation be as above. Define the functors
where G and .G0 ; g; / are objects and , are morphisms of the corresponding
categories ( for the notation relative to see Proposition 7.4.12).
Then H and K define an equivalence between the categories of super Lie groups
and super Harish-Chandra pairs.
x moreover
Proof. We first check that .H B K/.G0 ; g/ ' .G0 ; g/. Clearly G0 D jGj,
the two equalities
for each X; Y 2 g and g 2 G0 tell us that the Lie superalgebras Lie.G/ and g are the
same and so is the action of G0 on g.
We now turn to check that G x D .K B H /.G/ ' G. In order to do this we shall
x
build a morphism W G ! G and prove that it is an SLG isomorphism.
7.4 Super Harish-Chandra pairs 133
also that if fXi g is a basis of g, then DXLi jg ´ evg DXLi , g 2 jGj D jGj, x forms
x
a basis of Tg .G/.
D .1/jXj Œ.1 ˝ X /
G .s/.g/ (since is an SLG morphism)
jXj
D .1/ ŒDXL .s/.g/
D Œ. .s//.X /.g/:
L
(Notice that D.d /e .X/
is on H and DXL on G.)
In the next example we show explicitly how to recover the group structure of a
super Lie group starting from its associated SHCP.
Example 7.4.14. We consider the super Lie group G D GL.1j1/. In the language of
T -points, GL.1j1/.T / is the set of invertible matrices x21 x12 with multiplication
x1 1 y x y C 1 x1 1 C 1 y2
1 1
D 1 1 2
; (7.13)
2 x2 2 y2 2 y1 C x2 2 x2 y2 C 2 1
7.4 Super Harish-Chandra pairs 135
@ @ @ @
X1 D x1 C 2 ; X2 D x2 C 1 ;
@x1 @2 @x2 @1
@ @ @ @
‚1 D x1 2 ; ‚2 D x2 1 ;
@1 @x2 @2 @x1
ŒXi ; Xj D 0; Œ‚i ; ‚i D 0;
ŒXi ; ‚j D .1/ iCj
‚j ; Œ‚1 ; ‚2 D X1 X2 :
y10
The element h D 0 y2 2 jGj acts through the adjoint representation on gl.1j1/1 as
follows:
h ‚1 D y1 ‚1 y21 ; h ‚2 D y2 ‚2 y11 :
Using the theory developed in the previous section, we now want to reconstruct
the multiplication map of G in terms of the corresponding SHCP. Introduce the linear
operators
V
fi W z
.gl.1j1/1 / ! C 1 .G/; 1 7! yi ; ‚1 ; ‚2 ; ‚1 ^ ‚2 7! 0
and
V
'i W z
.gl.1j1/1 / ! C 1 .G/; ‚i 7! 1; 1; ‚j ¤i ; ‚1 ^ ‚2 7! 0:
V
The maps ffi ; 'i g are going to be our (global) coordinates on O.G/ D Hom. .gl.1j1//,
z They extend in a natural way to U.g0 /-linear maps from U.g0 / ˝ ƒ.g1 / to
C 1 .G//.
C .G/, which we denote by the same letter. We write fO (resp. ')
1 z
O for the composition
f B 1 (resp. 'i B 1 ).
We want to calculate the pullbacks:
.
.fi //.X; Y /.g; h/ ´ fOi .h1 .X /.Y //.gh/
(7.14)
D fi .O 1 .h1 .X /.Y ///.gh/;
. .'i //.X; Y /.g; h/ ´ 'Oi h1 .X /.Y //.gh/
(7.15)
D 'i .O 1 .h1 .X /.Y ///.gh/:
In order to perform the computations we first need to compute the elements O 1 .h1
.X/.Y //. The next table collects them.
136 7 Super Lie groups
H
HHY 1 ‚1 ‚2 ‚1 ^ ‚2
X H
1 1 ‚1 ‚2 ‚1 ^ ‚2
y21 y1 .‚1 ^ ‚2 y11 y2
‚1 y11 y2 ‚1 0 2
.X1 C X2 /‚1
21 .X1 C X2 //
y21 y1 .‚1 ^ ‚2 y21 y1
‚2 y21 y1 ‚2 0 2
.X1 C X2 /‚2
21 .X1 C X2 //
‚1 ^ ‚2 ‚1 ^ ‚2 12 .X1 C X2 /‚1 1
2
.X1 C X2 /‚2 1
4
.X1 C X2 /2
From this and using Definitions (7.14) and (7.15), we can calculate easily the
various pullbacks. Let us do it in detail in the case of f1 . The pullback table of
. .f1 //.X; Y /..x1 ; x2 /; .y1 ; y2 // is the following.
HH Y
HH 1 ‚1 ‚2 ‚1 ^ ‚2
X H
1 x1 y1 0 0 0
‚1 0 0 12 x1 y1 0
‚2 0 12 y21 x1 y12 0 0
‚1 ^ ‚2 0 0 0 1
x y
4 1 1
The link with the form of the multiplication morphism as given in equation (7.13)
is established by the isomorphism
x1 D f1 1 C '12'2 ; x2 D f2 1 '12'2 ; i D fi 'i :
The SHCP give us immediately the correspondence between Lie subgroups of a
supergroup G and Lie subalgebras of g D Lie.G/.
Proposition 7.4.15. Suppose that G is a connected super Lie group with super Lie
algebra g, and let h g be a Lie subalgebra of g. There exists a unique immersed
subgroup .H; j / whose Lie superalgebra is h.
Proof. Since G z is a connected Lie group, there exists a unique immersed Lie subgroup
z Q Q
.H0 ; j / with j W Hz ! G. z Let H denote the super Lie group associated with the super
Harish-Chandra pair .H z ; h/, and define
Proposition 7.5.3. There are three distinct types of one-parameter subgroups asso-
ciated with a homogeneous element in g of dimension 1j0, 0j1, and 1j1, respectively.
Their functors of points are:
(a) for any X 2 g0 we have
xX .T / D fexp.tX / j t 2 O.T /0 g D R1j0 .T / D Hom.C 1 .x/; O.T //I
(b) for any Y 2 g1 , ŒY; Y D 0, we have
xY .T / D fexp.#Y / D 1C#Y j # 2 O.T /1 g D R0j1 .T / D Hom.RŒ ; O.T //I
(c) for any Y 2 g1 , Y 2 ´ ŒY; Y =2 6D 0, we have
xY .T / D fexp.t Y 2 C #Y / j t 2 O.T /0 ; # 2 O.T /1 g
D R1j1 .T / D Hom.C 1 .x/Œ ; O.T //:
In cases (a) and (b) the multiplication structure is trivial, while in case (c) it is
given by .t; #/ .t 0 ; # 0 / D .t C t 0 # # 0 ; # C # 0 /.
Proof. The case (a), namely when X is even, is clear. When instead X is odd we have
two possibilities: either ŒX; X D 0 or ŒX; X ¤ 0. The first possibility corresponds,
by the Frobenius theorem, to a 0j1-dimensional subgroup whose functor of points is,
as is easily seen, representable and of the form (b). Let us now examine the second
possibility.
The Lie subalgebra hX i generated by X is of dimension 1j1 by (7.16). Hence by the
Frobenius theorem it corresponds to a Lie subgroup of the same dimension, isomorphic
to R1j1 .
Now we compute the group structure on this R1j1 , using the usual functor of points
notation to give the operation of the supergroup. For any commutative superalgebra
A, we have to calculate t 00 2 A0 , # 00 2 A1 such that
exp.tX 2 C #X / exp.t 0 X 2 C # 0 X / D exp.t 00 X 2 C # 00 X /
where t; t 0 2 A0 , #; # 0 2 A1 . The direct calculation gives
exp.tX 2 C #X / exp.t 0 X 2 C # 0 X /
D .1 C #X / exp.tX 2 / exp.t 0 X 2 /.1 C # 0 X /
D .1 C #X / exp..t C t 0 /X 2 /.1 C # 0 X /
D exp..t C t 0 /X 2 /.1 C #X /.1 C # 0 X /
D exp..t C t 0 /X 2 /.1 C .# C # 0 /X # # 0 X 2 /
D exp..t C t 0 /X 2 /.1 # # 0 X 2 /.1 C .# C # 0 /X /
D exp..t C t 0 /X 2 / exp.# # 0 X 2 /.1 C .# C # 0 /X /
D exp..t C t 0 # # 0 /X 2 /.1 C .# C # 0 /X /
D exp..t C t 0 # # 0 /X 2 C .# C # 0 /X /;
140 7 Super Lie groups
7.6 References
Historically the concept of supermanifold was introduced in order to formalize the no-
tion of super Lie group. In fact, although super Lie algebras were already an object of
study by the physicists who originally introduced them to describe super-time infinites-
imal symmetries (see [36], [80], [34], [65]), the geometrical global objects encoding
such infinitesimal structure were not introduced until 1975 in [11].
The basic reference where super Lie groups are systematically studied for the first
time is again [49]. In particular, in [49], Kostant introduced Lie–Hopf algebras and
stated the analogue of Theorem 7.4.5. In the notes by Deligne and Morgan [22] there is
a brief discussion of super Harish-Chandra pairs that we fully develop in Section 7.4.
The realization of the structure sheaf of a super Lie group we give in terms of the
corresponding super Harish-Chandra pair is due to Koszul and appeared in [50].
The material in Section 7.5 appeared first, to our knowledge, in [31].
8
Actions of super Lie groups
In this chapter, we want to focus our attention on the concept of action of a super
Lie group G on a supermanifold M . When G acts on M , if we fix a topological
point p 2 jM j, the orbit morphism G.T / 3 g 7! g p 2 M.T / is a constant rank
morphism. This nice property gives us the representability of the stabilizer functor and
hence allows us to show right away the representability of all the supergroup functors
for the classical supergroups like SLmjn and Ospmjn , the orthosymplectic supergroup
(see Example 8.4.8).
In the previous chapter we have seen how the concept of super Lie group is essen-
tially captured by the super Harish-Chandra pair associated with it, so that we have an
equivalence of categories between super Lie groups and SHCP. It is only natural to ask
how the concept of action of a super Lie group translates in the language of SHCP, to
give an equivalent approach to the theory of homogeneous spaces.
aW G M ! M
or equivalently:
a B . idM / D a B .idG a/; (8.1a)
O idM i D idM :
a B he; (8.1b)
If we have an action a of G on M , then we say that G acts on M , or that M is a
G-supermanifold.
Remark 8.1.3. Equations (8.1a) and (8.1b) correspond to the commutativity of the
diagrams
idM
GGM /GM M ' R0j0
eid /GM
MMM
MMM ww
idG a MMM www
ww a
a
MMM
id
& w{ w
GM
a / M, M
and formalize the usual properties satisfied by classical actions. As we already pointed
out in Remark 7.2.8, we could also define right actions by introducing the flip morphism
in equation (8.1a).
Let p 2 jM j. We can interpret, as usual, a point p in the topological space jM j as
an element of M.R0j0 /, i.e., as morphism pR0j0 W R0j0 ! M with jpR0j0 j W R0 ! jM j
assigning to the only element in R0 the point p and pR 0j0 being the evaluation at p.
Since we always have the unique morphism T ! R0j0 for any supermanifold T , by
functoriality we also have a morphism M.R0j0 / ! M.T / and we can define pT the
image of pR0j0 under such a morphism. We also define pO W M ! M as the morphism
pOT W M.T / ! M.T /, m 7! pT .
Definition 8.1.4. If p 2 jM j and g 2 jGj define the morphisms
ap W G ! M; ag W M ! M
in the functor of points notation as
ap W G.S / ! M.S/; g 7! g pS ;
ag W M.S/ ! M.S/; m 7! gS m:
Equivalently:
ap ´ a B hidG ; pi;
O ag ´ a B hg;
O idM i;
idG p gidM
G 'GR 0j0 /GM M ' R0j0 /GM
LLL w MMM w
LLLap ww MMM g ww
LLL www MaMM www
LL& w a MMM w a
{w
w & w{ w
M, M.
The maps ap and ag satisfy the relations
8.1 Actions of super Lie groups on supermanifolds 143
1
• ag B ag D idM for all g 2 jGj,
• ag B ap D ap B `g for all g 2 jGj and p 2 jM j.
We call ap the orbit morphism.
The next proposition states some nice properties of the maps ag and ap . In particular
we have that ap is a constant rank morphism. In the super context, this is a more delicate
result than its classical counterpart since, as we have seen in Chapter 5, the concept
of constant rank itself is more subtle. As we shall see, this result is of fundamental
importance in the proof of the existence of the stabilizer subgroup (see Section 8.4).
Proposition 8.1.5. Let G be a Lie supergroup acting on a supermanifold M via the
action a. Then
(1) ag is a superdiffeomorphism for all g 2 jGj,
(2) ap has constant rank for all p 2 jM j.
1
Proof. The first item follows at once from ag B ag D idM for all g 2 jGj.
Let us consider the second point. Suppose that M is a supermanifold of dimension
O n/
.m; O and G is a super Lie group of dimension .k;O q/.
O
Let g be the super Lie algebra of G and let Jap be the Jacobian matrix of ap in a
neighbourhood of a point g 2 jGj. Since
jJap j.g/ D .dap /g D .dag /p .dap /e .d`g 1 /g
and ag and `g 1 are diffeomorphisms, jJap j.g/ has rank dim g dim ker .dap /e for
each g 2 jGj. Recall that if X 2 g we denote by DXL ´ .1 ˝ X / the left-invariant
vector field associated with X. Using equation (8.1a) we have, for each X 2 ker .dap /e ,
DXL ap D .1 ˝ X /
.1 ˝ evp /a
D .1 ˝ X ˝ evp /. ˝ 1/a (8.2)
D .1 ˝ .dap /e .X //a D 0:
If ft i ; j g and fx k ; l g are coordinates in a neighbourhood U of e, and in a neigh-
bourhood V jap j.U / of p, respectively, then
0 k .x k /
1
@ap .x / @ap
Jap D @ @a@t i
. l /
@ j A2M
@a . l / O nj
m; O qO .OG .U //:
O k;
p p
@t i @ j
We want to find a matrix A 2 GLkj O qO .OG .U // such that Jap A has a certain set of
column equal to zero. We are going to use equation (8.2).
Let mjn D dim ker .dap /e and let fXu g and f„v g be bases of g0 and g1 such that
Xu ; „v 2 ker .dap /e for u m and v n. Let
X @ X @ X @ X @
DXLu D au;i i
C ˇu;j j ; L
D„ v
D v;i i
C dv;j j
@t @ @t @
i j i j
144 8 Actions of super Lie groups
au;i v;i
(with au;i ; dv;j 2 OG .U /0 and ˇu;j ; v;i 2 OG .U /1 ) and A D ˇu;j dv;j . Clearly,
the vector fields Xu ; „v being linearly independent at each point, we have that the
reduced matrix jAj is invertible so that A 2 GLkj O qO .OG .U //.
Now use equation (8.2) to conclude that the matrix
!
DXLu ap .x k / D„ L k
a .x /
Jap A D v p
DXLu ap . l / D„ L l
a . /
v p
0 1
0 ˛1 0 ˇ1
B0 ˛2 0 ˇ2 C
B C:
@0 1 0 ı1 A
0 2 0 ı2
Since jJap j has rank .kO m; qO n/, we can suppose that ˛1 and ı1 are invertible.
As
in Proposition 5.2.6, we can rearrange the matrix so that it takes the form wz 00 , with
z 1 0 I 0
z invertible and the matrix G D wz 1 I such that GJ D 0 0 .
We can then conclude that Jap has constant rank in U and, by translation, in all
of jGj.
aW G M ! M
and the first statement follows from the fact that the assignment DXL ! DXR is a super
Lie algebra anti-isomorphism.
It thus remains to prove equation 8.4. This is proved by
aW U.Lie.G// ! U.VecM /:
Proof. We use the universal property of the universal enveloping algebra and extend
the anti-morphism by mapping basis to basis. We can characterize the extension also
by the relation (8.4): for v1 ; v2 ; : : : ; vk 2 Te .G/ and f a local section of OM , we have
Since the category of super Lie group is equivalent to the category of SHCP, one
could ask whether assigning an SHCP action on a supermanifold M is equivalent to
assigning an action of the corresponding super Lie group on M . The answer is positive
as we shall presently see.
8.3 Actions of super Harish-Chandra pairs 147
z g/ on M .
define an action of the associated SHCP .G;
We are going to discuss the converse of the previous proposition. In other words, we
want to understand how to recover the action of a super Lie group G on a supermanifold
M from the knowledge of the action of its associated SHCP.
In analogy with the classical case, one could use the super Frobenius theorem to
reconstruct a local action of G from its infinitesimal action (8.5). However, we want
to take a different path and avoid the use of the super Frobenius theorem, proceeding
instead to the explicit reconstruction of the global action.
Assume that we have an action a of G on M , and let .a; a / be the corresponding
action of its SHCP as in Proposition 8.3.2. If f 2 O.M /, then by definition
Using equation (8.1a) and the form of the left-invariant vector fields as given in the
previous chapter, we have
This suggests the definition of the action a and proves its uniqueness, as we shall
see in the next proposition.
Proposition 8.3.3. Let .G;z g/ be the SHCP associated with the super Lie group G and
z g/ on a supermanifold M . Then there is a unique action
let .a; / be an action of .G;
a W G M ! M of the super Lie group G on M whose reduced and infinitesimal
actions are .a; /. a is explicitly given by
z y̋ O.M //;
a W O.M / ! HomU.g0 / .U.g/; C 1 .G/
(8.6)
f 7! ŒX 7! .1/jXj .1C 1 .G/
z ˝ .X //a .f /:
148 8 Actions of super Lie groups
Proof. Let us check that a .f / is U.g0 /-linear. For all X 2 U.g/ and Z 2 g0 we
have, using the fact that a is an anti-homomorphism,
In this case the representation sends each element of gl.1j1/ into the correspond-
ing right-invariant vector field.
The action can be reconstructed using equation (8.6); a simple calculation shows that
.t 1 / D t 1 x 1 .1 C 1 2 / C t 1 2 1 ; .t 2 / D t 2 x 2 .1 C 1 2 / C t 2 1 2 ;
. 1 / D t 1 1 .1 C 1 2 / C t 1 x 2 1 ; . 2 / D t 2 2 .1 1 2 / C t 2 x 1 2 :
The usual form of the multiplication map (as given in Example 7.4.14) is obtained
using the isomorphism
1 2
t 1 7! t 1 .1 C 1 2 /; 1 7! I t 2 7! t 2 .1 C 1 2 /; 2 7! :
t1 t2
Definition 8.4.3. Given two objects X and Y in a category and two arrows ˛ and ˇ
between them, an equalizer is a universal pair .E; / that makes the following diagram
commute:
/ ˛ /
E X /Y:
ˇ
One can also define the notion of coequalizer in the same way, by reversing all the
arrows.
Let p 2 jM j and denote by pO the morphism
Š ip
pO W G !
R0j0
! M;
where Š W G ! R0j0 and ip W R0j0 ! M denote respectively the unique morphism from
G to a point supermanifold and the canonical injection of p in M .
We notice that it is not a priori clear that such an equalizer exists. We are going to
establish a proposition relating the two Definitions 8.4.2 and 8.4.4 of stabilizer functor
and stabilizer subgroup, showing that the stabilizer subgroup always exists and its
functor of points is precisely the stabilizer functor.
Before this, we need a category-theoretic lemma.
Lemma 8.4.5. Let C be a category. The Yoneda embedding, i.e., the injection C 3
X ! Hom.; X/, preserves the equalizers.
Proof. For this proof, we temporarily adopt the following notational convention. If X
is an object in C , the corresponding representable functor is denoted by hX .
e
Suppose that E ! X Y is an equalizer in a category C . Then we need to
show that for any equalizing diagram F ! hX hY , the natural transformation
a W F ! hX factors uniquely via hE ! hX .
For any object Z of C and any x 2 F .Z/, aZ .x/ is an arrow Z ! X that
composes equally with X Y and therefore factors uniquely via e. Therefore define
8.4 The stabilizer subgroup 151
bZ .x/ W Z ! E, bZ .x/ 2 hE .Z/ to be the unique map such that e B bZ .x/ D aZ .x/.
Such b is functorial in Z. Certainly he B b D a by construction.
F @
a / hX
@@ | >
@@b e |||
@@ |
@ |||
hE .
For uniqueness, if b 0 is some other natural transformation with he Bb 0 D a, then for any
0
object Z and any x 2 F .Z/ we have e B bZ .x/ D aZ .x/ D e B bZ .x/. This implies
0
that bZ .x/ D bZ .x/ since e is a monomorphism. Hence b D b 0 and uniqueness is
established.
Lemma 8.4.6. If A and B are two (super)algebras and ˛, ˇ are morphisms between
them, their coequalizer is the algebra C D B=J , where J is the ideal .˛.a/ ˇ.a/ j
a 2 A/.
ˇ
C o B oo A:
˛
Proposition 8.4.7. Let G be a Lie supergroup acting on the supermanifold M and let
p 2 jM j.
(1) The diagram
ap
G // M
pO
jG ap
p
/G // M :
Gp
pO
(Again here hG and hM denote the functor of points of the supermanifolds G and
M .) Since (see Lemma 8.4.5) the Yoneda embedding preserves equalizers and due to
uniqueness, it follows that F Gp ' hGp .
(4) The first statement is clear since jGj ' G.R0j0 / as set-theoretical groups.
Moreover, since jGp B ap .f / is a constant for all f 2 O.M /, we have gp ker dap ,
and equality holds for dimension considerations.
We are ready for some important examples.
Examples 8.4.8. (1) Consider the action (expressed with the functor of points notation):
The stabilizer of the point 1 2 jR1j0 j coincides with all the matrices in GLmjn .T / with
Berezinian equal to 1, that is, the special linear supergroup SLmjn .T /. By Proposi-
tion 8.4.7 we have immediately that SLmjn is representable and that it is a super Lie
subgroup of GLmjn .
8.5 References 153
a W GLmjn B ! B; .g; . ; // ! .g ; g /;
where B is the super vector space of all the symmetric bilinear forms on Rmjn . Consider
the point in B that corresponds to the standard bilinear form in B:
0 1
0 Ip 0 0 0
BIp 0 0 0 0C
B C
ˆDB B 0 0 1 0 0 C if m D 2p C 1; n D 2q;
C
@0 0 0 0 Iq A
0 0 0 Iq 0
or 0 1
0 Ip 0 0
BIp 0 0 0C
ˆDB @0 0
C if m D 2p; n D 2q:
0 Iq A
0 0 Iq 0
The stabilizer of the point ˆ is the supergroup functor Ospmjn . Again this is a Lie
supergroup by Proposition 8.4.7.
For more details on the Lie superalgebras and the bilinear forms see Appendix A.
8.5 References
The action of the super Harish-Chandra pairs appears in [22]. The representability of
the stabilizer functor in Proposition 8.4.7 is stated in [22]; however a complete proof
of this statement appears only in [4].
9
Homogeneous spaces
G M ! M; g; m 7! g m;
Proposition 9.1.4. Suppose that M is a G-superspace, for a fixed action a. Then the
following facts are equivalent:
(1) a is transitive.
(2) jaj W jGj jM j ! jM j is transitive,
.dap /e W g ! Tp .M / is surjective for one p (hence for all p 2 jM j).
(3) If q denotes the odd dimension of G, then
is surjective.
(4) The sheafification of the functor (see Appendix B) im ap W .smflds/op ! .sets/
Proof. (1) () (2). This is an immediate consequence of Proposition 8.1.5 and our
previous remark.
(1) H) (3). If 2 M.R0jq / D Hom.R0jq ; M /, let jj 2 jM j be the image of the
reduced map associated with (with a small, but usual, abuse of notation we denote
by the same symbol jj both the morphism W R0 ! jM j and the point p which is its
image jj.R0 / 2 jM j). The pullback depends only on the restriction of the sections
of OM to an arbitrary neighbourhood of jj. This is an easy exercise, which we leave
to the reader. If ap is a surjective submersion, there exists a local right inverse s of ap
defined in a neighbourhood of jj. By the locality of , s B is a well-defined element
of G.R0jq / and, moreover,
(3) H) (2). Suppose that ap;R0jq surjective. Looking at the reduced part of each
morphism in ap;R0j0 G.R0jq / , we have that ap;R0j0 D jap j W jGj ! jM j is surjective.
As a consequence of the ordinary manifold theory result in [48], Theorem 5.14, jaj is a
classical transitive action and jajp is a submersion. Let now m 2 jM j and ft i ; j g be
coordinates in a neighbourhood U of it. Consider the element 2 M.R0jq / defined
by
V
W OM .U / ! O.R0jq / D .1 ; : : : ; q /; t i 7! jt i j.m/; j 7! j :
By surjectivity of ap;R0jq , there exists 2 G.R0jq / such that ap;R0jq . / D :
B ap .t i / D jt i j.m/;
B ap . j / D j :
This implies that Tm .M /1 is in the image of .dap /j j . Since, by our previous con-
siderations, jap j is a submersion, Tm .M /0 is in the image as well. Hence, due to
Proposition 8.1.5 we are done.
(1) H) (4). Let us suppose that ap is a surjective submersion. Let m 2 jM j
and g 2 jap j1 .m/ (jap j is surjective, so it exists). Since ap is a submersion there
exists V jGj with coordinates X1 ; : : : ; XpCq (dim G D pjq) and W jM j with
coordinates Y1 ; : : : ; YmCn (dim M D mjn) such that
ap .Yi / D Xi :
Let t 2 U jT j for a generic supermanifold T and ˛ W U ! M such that m D j˛j.t /.
We can assume that j˛j.U / W . If ˛ .Yi / D fi 2 OT .U /, then ˇ W U ! V defined
by ´
fi if i m C n;
ˇ .Xi / D
0 otherwise;
satisfies ap B ˇ D ˛. Then Œ˛ 2 .im ap;T / t , hence .im ap;T / t D M t . This shows that
A
im ap D M (see Proposition B.2.9).
A
(4) (H (1). Let us suppose that im ap D M . Taking T D R0j0 we have that jap j
must be surjective. Let us now assume that T D M and m 2 jM j. There exists U 3 m
and W U ! G such that ap B D idU . Then ap must be a submersion at j j.m/,
and this is true everywhere since ap has constant rank. Indeed for all g 2 jGj,
.dap /g B .dlgG /1 D .dlgM /x B .dap /1
0
where the isomorphisms lgG and lgM are the left actions of g on G and M , respectively.
Remark 9.1.5. Notice that in the statement of .4/ in Proposition 9.1.4, it is too restric-
tive to require the transitivity of aT for each T , i.e., im ap .T / D M.T /. In fact, in
the ordinary setting, this would imply that we can lift every morphism T ! M to a
morphism T ! G. As a consequence, we have the existence of a global section of the
fibration G ! M (take T D M and the identity map), which is not true in general. It
is hence necessary to take the sheafification of the image functor, as we do in point (4).
9.2 Homogeneous spaces: The classical construction 157
G=Gx0 ! M; gGx0 7! g x0 ;
projection. It is easy to check that this is a sheaf. We need to show that this defines a
manifold structure on G=H and that G acts on the corresponding ringed space naturally.
To this end, it is enough to show that, for some open subset V in G=H containing .1/,
there is a homeomorphism of V with a smooth manifold W such that OjV goes
over OW under . We obtain such W as a submanifold of G containing 1 such that
V D .W / and 1 .V / Š W H in the following way. The map w; h 7! wh of
W H into G is a diffeomorphism of W H onto 1 .V / (which is open in G)
commuting with the right actions of H (y 7! yh) on G and .w; k/ 7! .w; kh/ on
W H.
The crucial existence of W , based ultimately on the local Frobenius theorem, is
equivalent to the existence of a local section V G=H ! W H G. Moreover,
as mentioned above, one has W H Š W H . The sheaf OjV can hence be identified
with the sheaf of C 1 functions on W . This proves that we have defined a manifold
structure on G=H . The uniqueness and the universal nature of this structure are not
difficult to prove (see [75], Ch. 2).
What we have described above is the key idea to the proof of Theorem 9.2.2, and
we shall see in Section 9.3 that, despite the different context, the existence of a local
section plays a crucial role also in the construction of a quotient of Lie supergroups in
very much the same way.
On the other hand, for any open subset W jGj, invariant under right translations by
elements of jH j, we define
If jH j is connected we have
Oinv .W / D Oh0 .W /;
as one can readily check by looking infinitesimally at the condition rx0 f D f for all
x0 2 jH j. For any open set jW j jX j D jGj=jH j with jW j D 01 .jW j/ we put
OX .jW j/ D Oinv .W / \ Oh .W / OG .W /:
It is clear that the left action of the group jGj on jX j leaves OX invariant and so
it is enough to prove that there is an open neighborhood jW j of the topological point
jj.1/ 1N in jGj=jH j with the property that .jW j; OX jjW j / is a super domain, i.e.,
isomorphic to an open submanifold of Rpjq .
We will do this using the local Frobenius theorem (see Chapter 6). Also, we identify
as usual g with the space of all left-invariant vector fields on G, thereby identifying the
tangent space of G at every point canonically with g itself.
On G we have a distribution spanned by the vector fields in h. We denote it by Dh .
On each jH j-coset x0 jH j we have a supermanifold structure which is a closed
submanifold of G. It is an integral supermanifold of Dh , i.e., the tangent space at any
point is the subspace h at that point. By the local Frobenius theorem there is an open
neighborhood U of 1 and coordinates xi , 1 i n, and ˛ , 1 ˛ m, on U such
that Dh is spanned on U by @=@xi , @=@˛ (1 i r, 1 ˛ s). Moreover, from the
theory on jGj we may assume that the slices L.c/ ´ f.x1 ; : : : ; xn / j xj D cj ; r C1
j ng are open subsets of distinct jH j-cosets for distinct c D .crC1 ; : : : ; cn /. These
slices are therefore supermanifolds with coordinates xi , ˛ , 1 i r, 1 ˛ s.
We have a submanifold W 0 of U defined by xi D 0 with 1 i r and ˛ D 0 with
1 ˛ s. The map jj W jGj ! jX j may be assumed to be a diffeomorphism of
jW 0 j with its image jW j in jX j, and so we may view jW j as a superdomain, say W .
The map jj is then a diffeomorphism of W 0 with W . What we want to show is that
W Š .jW j; OX jjW j /.
160 9 Homogeneous spaces
W0H !
G; w; h ! wh;
Proof. The map in question is the informal description of the map B .iW 0 iH /
where iM refers to the canonical inclusion M ,! G of a sub-supermanifold of G into
G, and W G G ! G is the multiplication morphism of the Lie supergroup G. We
shall use such informal descriptions without comment from now on.
It is classical that the reduced map jj is a diffeomorphism of jW 0 j jH j onto the
open set U D jW j0 jH j. This uses the fact that the cosets wjH j are distinct for distinct
w 2 jW 0 j. It is thus enough to show that d is surjective at all points of jW 0 j jH j.
For any h 2 jH j, right translation by h (on the second factor in W 0 H and simply rh
on G) is a super diffeomorphism commuting with and so it is enough to prove this
at .w; 1/. If X 2 g is tangent to W 0 at w and Y 2 h, then
Hence the range of d is all of g since from the coordinate chart at 1 we see that the
tangent spaces to W 0 and wjH j at w are transversal and span the tangent space to G
at w which is g. This proves the lemma.
Proof. To ease the notation we drop the open set in writing a sheaf superalgebra, that
is, we will write OX instead of OX .U /.
We want to show that for any g in OX jU , g is of the form f ˝ 1 and that the
map g 7! f is bijective with OW 0 . Now intertwines DZ (Z 2 h) with 1 ˝ DZ and
so .1 ˝ DZ / g D 0. Since the DZ span all the super vector fields on jH j, it follows
using charts that for any p 2 jH j we have g D fp ˝ 1 locally around p for some
fp 2 OW 0 . Clearly fp is locally constant in p. Hence fp is independent of p if jH j
is connected. If we do assume that jH j is connected, the right invariance under jH j
shows that fp is independent of p. In the other direction it is obvious that if we start
with f ˝ 1, it is the image of an element of OX jU .
Proof. By the previous lemmas we know that .jX j; OX / is a super manifold at 1. N The
left invariance of the sheaf under jGj shows this to be true at all points of jX j.
9.3 Homogeneous superspaces for super Lie groups 161
1
ˇ
GX /X
commutes.
Proof. Let ˛ ´ B W G G ! X . The action of jGj on jX j shows that such a map
jˇj exists at the reduced level. So it is a question of constructing the pullback map
ˇ W OX ! OGX
such that
.1 / B ˇ D ˛ :
Now is an isomorphism of OX onto the sheaf OG restricted to a sheaf on X
(W 7! OG .jj1 .W //), and so to prove the existence and uniqueness of ˇ it is a
question of proving that ˛ and .1 / have the same image in OGG . It is easy to
see that .1 / has as its image the subsheaf of sections f killed by 1 ˝ DZ .Z 2 h/
and invariant under 1 rh .h 2 jH j/. It is not difficult to see that this is also the image
of ˛ .
We tackle now the question of the uniqueness of X .
Proposition 9.3.6. Let X 0 be a super manifold with jX 0 j D jX j and let 0 be a
morphism G ! X 0 . Suppose that
(1) 0 is a submersion,
(2) the fibers of 0 are the supermanifolds which are the cosets of H .
Then there is a natural isomorphism X ' X 0 .
Proof. Indeed, from the local description of submersions as projections it is clear
that, for any open jW j jX j, the elements of 0 .OX 0 .jW j// are invariant under
rh ; .h 2 jH j/ and killed by DZ .Z 2 h/. Hence we have a natural map X 0 ! X
commuting with and 0 . This is a submersion, and by dimension considerations it
is clear that this map is an isomorphism.
We have proved the following result:
Theorem 9.3.7. Let G be a Lie supergroup and H a closed Lie subgroup. There exist
a supermanifold X D .jGj=jH j; OX / and a morphism W G ! G=H such that the
following properties are satisfied:
162 9 Homogeneous spaces
1
ˇ
GX / X.
Moreover, the pair .X; / subject to the properties (1), (2), and (3) is unique up to
isomorphism. The isomorphism between two choices is compatible with the actions,
and it is also unique.
Proof. This is an immediate consequence of the previous lemmas and propositions.
j˛;ˇ iˇ
i˛
hX˛ / F.
164 9 Homogeneous spaces
hT˛ ´ hX˛ F hT / hT
t˛ t
i˛
hX˛ / F.
Since fhX˛ g is an open cover of F , the fT˛ g form an open cover of T . Since by
Yoneda’s lemma: Hom.hT˛ ; hX˛ / Š Hom.T˛ ; X˛ / we obtain a family of morphisms:
t˛ W T˛ ! X˛ X . The morphisms t˛ glue together to give a morphism t 0 W T ! X ,
hence t 0 2 hX .T /. So we define T .t / D t 0 .
Next we construct another natural transformation T W hX .T / ! F .T /, which
turns out to be the inverse of .
Assume that we have f 2 hX .T /, i.e., f W T ! X . Let T˛ D f 1 .X˛ /. We
immediately obtain morphisms g˛ W T˛ ! X˛ F . By Yoneda’s lemma, g˛ corre-
sponds to a natural transformation g˛ W hT˛ ! hX˛ . Since F is local, the morphisms
i˛ g˛ W hT˛ ! hX˛ ! F glue together to give a morphism g W hT ! F , i.e., an
element g 2 F .T /. Define T .f / D g.
9.4 The functor of points of a quotient supermanifold 165
One can directly check that and are indeed natural transformations inverse to
each other, hence F Š hX .
Remark 9.4.4. The representability criterion we just proved, holds as it is, also for the
category of analytic supermanifolds discussed in Section 4.8. The proof is the same,
and the reader is invited to check that at no point did we make any use of the fact that
a sheaf has any local C 1 property. In fact Theorem 9.4.3 is a categorical result, the
reader may also consult [77], Ch. 1, for a more general setting of this statement. We
are going to revisit this same proposition for the case of superschemes in Chapter 10.
However, we are going to employ a slightly different category and for this reason the
statement is only apparently not the same. All we say in Theorems 9.4.3 and 10.3.7 is
contained in [77], where the treatment is the most general possible.
We now turn to examine the functor of points of the quotient of a Lie supergroup
G by a closed subgroup H . We are interested in a characterization of the functor of
points of X D G=H in terms of the functor of points of the supergroups G and H .
A
Theorem 9.4.5. Let G and H be supergroups as above and let G=H be the sheafifica-
A
tion of the functor: T ! G.T /=H.T /. Then G=H is representable and is the functor
of points of the homogeneous space supermanifold X D G=H constructed above.
Proof. In order to prove this result, we shall use the uniqueness property which char-
A A
acterizes the homogeneous space G=H (see Theorem 9.3.7). So we only need to prove
that G=H is representable, i.e., G=H D hX for a supermanifold X , and that X satisfies
A
the three properties detailed in Theorem 9.3.7.
A
To prove that G=H is representable we use the criterion in Theorem 9.4.3. The fact
A
that G=H has the sheaf property is clear by its very definition. So it is enough to prove
there is a open supermanifold subfunctor of G=H around the origin (by translation we
can transport such open supermanifold subfunctor to obtain a neighbourhood at every
point). But this is given by hW , with W Š W 0 H constructed as in the previous
section.
We now turn to the properties (1), (2) and (3). (1) and (3) are left to the reader as
an exercise. The fact is a submersion, that is property (2), comes by looking at it in
the local coordinates given by W .
Let us fix a topological point X0 D diag. 1 ; : : : ; mCn / 2 jgj, where the real
numbers 1 ; : : : ; mCn are all distinct. Then one sees immediately that the stabilizer
166 9 Homogeneous spaces
subgroup H at X0 is given by
80 19
ˆ
< a1 ::: 0 >
=
B :: C G.T /:
H.T / D @ ::: : A>
:̂ ;
0 ::: amCn
Thus we have defined .g Fi /.T˛ / for a base of open sets. Since the compatibil-
ity conditions in Definition 2.2.10 are satisfied, g Fi is a B-sheaf and we can use
Proposition 2.2.11 to obtain a sheaf on T that we denote by g Fi .
Let us fix the flag F D fOT2j0 OT2j1 g 2 F `.T /. The stabilizer subgroup at F is
the subgroup H of G given in terms of the functor of points by
80 19
ˆ
ˆ g11 g12 g13 g14 15 > >
ˆB
ˆ >
<Bg21 g22 g23 g24 25 C C=
>
H.T / D B B 0 0 g33 g34 0 C C> 2 G.T /:
ˆ
ˆ
ˆ@ 0 0 g43 g44 0 A> >
>
:̂ ;
0 0 53 54 g55
In fact, one can check directly that H stabilizes F and that the functor T 7! H.T /
is the functor of points of a closed subgroup of G. Hence locally, we can identify the
set F `.T / with the set G.T /=H.T / so that the sheafification of both functors coincide
(see Proposition B.2.9). Since F ` is already a sheaf, by its very definition, we have
that F ` is the functor of points of the quotient supermanifold G=H .
Next we want to describe the big cell inside F `. This is an open submanifold, which,
as we shall see, has a special importance in physics since it is the complexification of
the super Minkowski space.
Locally at a point t 2 jT j, an element F1 F2 in F `.T / consists of two free
4j1
submodules of OT;t or rank 2j0 and 2j1, and is thus described by a pair of subspaces
4j1
hv1 ; v2 i hw1 ; w2 ; w3 i with vi ; wj 2 OT;t of suitable parity. Let U1 .T / and U2 .T /
be the set of subsheaves of OT4j1 of ranks 2j0 and 2j1 expressed locally as
8 0 1 0 1 9
ˆ
ˆ v11 v12 >
>
ˆ*B C B C+
ˆ >
>
< Bv21 C Bv22 C ˇ =
Bv31 C ; Bv32 C ˇ detˇ v 11 v12
U1 .T / t D B C B C invertible ;
ˆ
ˆ @v41 A @v42 A v21 v22 >
>
ˆ >
>
:̂ ;
v51 v52
8 0 1 0 1 0 1 9
ˆ
ˆ w11 w12 w13 0 1 >
>
ˆ
ˆ * w C Bw C Bw C+ >
< B B 21 C B 22 C B 23 C ˇˇ w11 w12 w13 >
=
U2 .T / t D Bw C ; Bw C ; B w C ˇ Ber @ w w w A invertible :
ˆ B 31 C B 32 C B 33 C 21 22 23
>
ˆ
ˆ @w41 A @w42 A @w43 A w51 w52 w53 >
>
:̂ >
;
w51 w52 w53
In order to simplify the notation, from now on we shall write everything locally
without further mention.
Using the identification F `.T / Š .G=H /.T /, which locally amounts to having
F `.T / Š G.T /=H.T /, we can write an element in U1 .T / U2 .T / uniquely as
00 1 0 11
I I 0
@@AA ; @B ˇ AA 2 U1 .T / U2 .T /;
˛ 0 1
168 9 Homogeneous spaces
where I is the identity, A and B are .22/-matrices with even entries, and ˛ D .˛1 ; ˛2 /,
ˇ t D .ˇ1 ; ˇ2 / are rows with odd entries in OT;t .
We define the big cell U.T / inside F `.T / as the pairs .u; v/ 2 U1 .T / U2 .T /.
Notice that an element of U1 .T / is inside U2 .T / if
A D B C ˇ˛: (9.1)
So a flag in the big cell U is completely described by the triplet .A; ˛; ˇ/. We see also
that U is an affine 4j4 superspace, i.e., the functor U is representable and is the functor
of points of a 4j4 superspace. Equation (9.1) is also known as the twistor relation in
the physics literature. Since there are no relations among A, ˛ and ˇ, we can take them
as local coordinates in a neighbourhood of the identity in jGj=jH j.
In these coordinates, the flag F corresponding to the identity that we fixed at the
beginning becomes
00 1 0 11
I I 0
@@ 0 A ; @ 0 0AA .0; 0; 0/:
0 0 1
The big cell U1 U2 is obtained by requiring the invertibility of the determinant
.1; 2/ and the Berezinian .1; 2; 5/ (obtained by taking respectively columns .1; 2/ and
rows .1; 2/ and columns .1; 2; 5/ and rows .1; 2; 5/) in the generic expression of a flag
in F `. Clearly one can repeat the same argument and obtain a total of six different big
cells by requiring the invertibility of the determinant and Berezinians:
We leave to the reader the easy check that these big cells cover the whole of F `. Hence
we can apply the representability criterion in Theorem 9.4.3, thus proving that F ` is
an analytic supermanifold.
We now want to write explicitly the morphism W G ! F `, .g/ D g F in these
coordinates and see it is a submersion. In a suitable open subset near the identity of
the group we can take an element g 2 G.T / as
gij i5
gD ; i; j D 1; : : : ; 4:
5j g55
where
g31 g32 g11 g12
1 D 51 52 ; W D ; ZD ;
g41 g42 g21 g22
1 D 15 ; 2 D 35 ; d D .g55 Z 1 1 /1 ;
25 45
1 1
V D W g55 2 1; Y D Z g55 1 1:
g 7! .W Z 1 ; 1Z
1
; .2 W Z 1 1 /d /:
At this point one can compute the super Jacobian and verify that at the identity it is
surjective, so is a submersion. This gives an equivalent and independent proof of the
fact that F ` is the quotient G=H (see Theorem 9.3.7).
The subgroup of G leaving invariant the big cell is the set of matrices in G of the
form 0 1
L 0 0
@NL R RA ;
d' 0 d
with L, N , R being 2 2 even matrices, and odd 1 2 matrix, ' a 2 1 odd matrix
and d a scalar. This is then what the physicists call the complex Poincaré supergroup
and its action on the big cell can be written as
If the odd part is zero, then the action reduces to the one of the classical Poincaré group
on the ordinary Minkowski space.
We see that the big cell of the flag supermanifold F `.2j0; 4j0I 4j1/ can be inter-
preted as the complex super Minkowski space time, the flag being its superconformal
compactification.
We now turn to the construction of the real Minkowski superspace which is ex-
tremely important in physics. We start with a real form for the supergroup G. In order
to obtain such a form (see Section 4.8 for the definition) we need a natural transforma-
tion from G to its complex conjugate G. x Define the natural transformation (T being a
supermanifold) as
x D D i
G.T / ! G.T /; g D !g D :
d i dN
Remark 9.5.2. It is important to notice that we are under the following convention: if
and are odd variables, then
D N N : (9.2)
This convention is opposed to the one used most commonly in physics, namely
D N N ;
but as it is explained in [22], it is the one that makes sense functorially. According to
this convention, then for matrices X , Y with odd entries
.X Y /T D .Yx /T .Xx /T :
Proposition 9.5.4. The topological space G consisting of the points fixed by has a
real supermanifold structure and the supersheaf is composed of those functions f 2 OG
such that .f / D f .
The involution we chose has important physical properties and we invite the reader
to consult [32] for a complete treatment of them.
Also, it is easy to check that it reduces to a conjugation on the Poincaré supergroup
and we can compute explicitly such conjugation as follows:
0 1 0 1
L 0 0 L1 0 0
g D @ M R RA ; g 1 D @R1 ML1 C 'L1 R1 d 1 A ;
d' 0 d 'L1 0 d 1
0 1
1
R 0 0
B C
g D @L 1 M R 1 L 1 ' L 1 j L 1 ' A :
j dN 1 0 dN 1
It follows that the fixed points are those that satisfy the conditions
1 1
L D R ; D j' ; ML1 D .ML1 / j L ' 'L1 : (9.3)
9.5 The super Minkowski and super conformal spacetime 171
To get a more familiar form for the reality conditions, we observe that the last
equation in (9.3) can be cast as
1 1
M 0 L1 ML1 C j L ' 'L1 ; M 0 D M 0 :
2
This is just an odd translation and amounts to multiplying g on the right by the group
element 0 1
I 0 0
1
g 0 D @ 12 jR1 L ' 'L1 I 0A :
0 0 1
We want now to compute the real form of the big cell. The first thing to observe is
that the real form is well defined on the quotient space G=H (the superflag), where H
is the group described previously.
Notice that a point of the big cell .A; ˛; ˇ/ can be represented by an element of the
group 0 1
I 0 0
g D @A I ˇ A
˛ 0 1
since 0 1 0 1 0 1 0 1 0 1
I I I 0 I 0 I 0
g @ 0 A D @AA ; g @0 0A D @A ˇ A @A ˇ˛ ˇA :
0 ˛ 0 1 ˛ 1 0 1
We first compute the inverse
0 1
I 0 0
g 1 D @A C ˇ˛ I ˇ A ;
˛ 0 1
and then 0 1
I 0 0
g D @A ˛ ˇ I j˛ A :
jˇ 0 1
The element g is already in the desired form, so the real points are given by
A D A j˛ ˛; ˇ D j˛ :
We can make a convenient change of coordinates,
1
A0 A C j˛ ˛;
2
so the reality condition is
A0 D A0 ;
and we recover the usual form for the (purely even) Minkowski space time.
172 9 Homogeneous spaces
9.6 References
This treatment of quotients of supergroups, via their functor of points and, equivalently,
through the direct construction of the sheaf of invariant functions, is found in [32], [4],
[5]. As for Example 9.4.6 see [76], [56].
10
Supervarieties and superschemes
The aim of this and the next chapter is to lay down the foundations of algebraic su-
pergeometry, using the machinery of sheaves and schemes together with their functor
of points that we have described in detail in the classical setting in the previous chap-
ters. The presence of nilpotents in the structural sheaf of an ordinary scheme makes its
generalization to supergeometry very natural and allows us to carry most definitions
and results from the ordinary to the super setting, almost with no changes. There are,
however, some remarkable differences that we shall point out along with our treatment.
The most striking one is the rigidity of the projective superspace, which does not con-
tain as many subvarieties as its classical counterpart. We shall, in fact, see that the
Grassmannian superscheme is not in general a projective supervariety. This fact has
some deep consequences, for instance, the strategy for the construction of the quotient
of algebraic supergroups has to be suitably modified. We shall however not pursue this
point further in the text.
We start by giving the definition of superscheme and its functor of points and we
examine some important examples, including affine and projective superspaces and
Grassmannian superschemes. We then describe a representability criterion, which al-
lows us to single out among functors from the category of superalgebras to the category
of sets those which are the functors of points of a superscheme. This criterion is espe-
cially useful in supergeometry since the functor of points is often the only way we can
handle supergeometric objects. Taking this point of view, we discuss the infinitesimal
theory in the algebraic super-setting. In particular we study superderivations and their
relations with the tangent space at a rational point of a superscheme over a field, making
some concrete calculations to exemplify our definitions. Our study of the infinitesimal
theory is of capital importance for the theory of algebraic supergroups that we shall
treat in the next chapter.
We denote their category by .salg/. For a superalgebra A, let JA denote the ideal gen-
erated by the odd elements, i.e., JA D hA1 iA . Write Ar for the quotient A=JA . We say
that Ar is reduced or super reduced even if it may contain some (even) nilpotents.
In Chapter 3 we have introduced the categories of superspaces and of superschemes.
Let us briefly recall these notions.
A superspace X D .jX j; OX / is a topological space jX j together with a sheaf of
superalgebras OX such that OX;x is a local superalgebra, i.e., it has a unique two-sided
maximal homogeneous ideal.
The sheaf of superalgebras OX is a sheaf of OX;0 -modules, where OX;0 is the sheaf
over jXj defined as OX;0 .U / ´ OX .U /0 for all U open in jX j. Notice that also the
sheaf OX;1 , defined as OX;1 .U / D OX .U /1 , is also a sheaf of OX;0 -modules.
Given two superspaces X D .jXj; OX / and Y D .jY j; OY /, a morphism f W X !
Y of superspaces is given by a pair f D .jf j; f / such that
(1) jf j W X ! Y is a continuous map;
(2) f W OY ! f OX is a map of sheaves of superalgebras on jY j, that is, for U
open in jY j there exists a family of morphisms fU W OY .U / ! OX .jf j1 .U //
compatible with restrictions;
(3) the map of local superalgebras fp W OY;jf j.p/ ! OX;p is a local morphism, i.e.,
sends the maximal ideal of OY;jf j.p/ to the maximal ideal of OX;p .
A superscheme X is a superspace .jX j; OX / such that OX;1 is a quasi-coherent sheaf
of OX;0 -modules. OX is called the structure sheaf of X. A morphism of superschemes
is a morphism of the corresponding superspaces.
For any open U jX j we define the superscheme U D .jU j; OX jjU j /, called an
open subscheme in the superscheme X .
The most important example of a superscheme is given by the spectrum of a super-
algebra A. It consists of the spectrum of the even part A0 together with a certain sheaf
of superalgebras on it. Let us see this construction in detail.
Definition 10.1.1 (The superscheme Spec A). Let A be an object of .salg/. We have
Spec.A0 / D Spec.Ar / since the algebras Ar and A0 differ only by nilpotent elements.
Let us consider OA0 the structure sheaf of Spec.A0 / (see Chapter 2 for more details).
The stalk of this sheaf at the prime p 2 Spec.A0 / is the localization of A0 at p. As for
any superalgebra, A is a module over A0 . So, according to the classical construction
detailed in Chapter 2, Section 2.5, we have a sheaf AQ (that we shall denote also by OA )
of OA0 -modules over Spec A0 , with stalk
˚
Ap D fg j f 2 A; g 2 A0 p ;
the localization of the A0 -module A over each prime p 2 Spec.A0 /. This localization
has a unique two-sided maximal ideal which consists of the maximal ideal in the local
ring .Ap /0 and the generators of .Ap /1 as A0 -module.
As one can easily check, AQ (or OA ) is a sheaf of superalgebras and .Spec A0 ,
OA / is a superscheme that we denote by Spec A. In fact we just showed that it is a
10.1 Basic definitions 175
superspace since OA;p is local; moreover, by its very definition, we have that OA;1 is a
quasi-coherent sheaf of OA;0 -modules since it is a sheaf of OA0 D OA;0 -modules.
Notice that on the basic open sets
Uf D fp 2 Spec A0 j .f / 6 pg; f 2 A0 ;
where JX is the ideal sheaf U 7! JOX .U / with JOX .U / the ideal generated by the odd
nilpotents in OX .U /.
We will call X r D .jX j; OXr / the reduced space associated to the superspace
X D .jX j; OX /. This is a locally ringed space in the classical sense. The scheme
.jX j; OXr / is called the reduced scheme associated to X . Notice that the reduced
scheme associated to a given superscheme may not be reduced, i.e., OXr .U /, U open in
jXj, can contain nilpotents. This is because OXr .U / is obtained by taking the quotient
of OX .U / by the ideal generated only by the odd nilpotents.
176 10 Supervarieties and superschemes
z be as above.
Theorem 10.1.4. Let M be an A-module for a superalgebra A, and let M
Then:
(1) Mz has a natural structure of OA -module.
(2) .Mz /p Š Mp for all p 2 Spec A0 , i.e., the stalk at any prime p of the sheaf M
z
coincides with the localization of the A0 -module M at p.
(3) .Mz /.Spec A0 / D M , i.e., the global sections of the sheaf coincide with the
A-module M .
! .Spec A0 /
IQ I:
.xi ai ; j k /; i D 1; : : : ; m; j; k D 1; : : : ; n;
.p1 ; : : : ; pr ; j k /; j; k D 1; : : : ; n;
where .p1 ; : : : ; pr / is a prime ideal in kŒx1 ; : : : ; xm . In other words the prime ideals
in kŒx1 ; : : : ; xm ; 1 ; : : : ; n 0 are generated by the prime ideals in kŒx1 ; : : : ; xm and
the even nilpotent ideal . i j ; i j /.
At the prime ideal p 2 Spec kŒx1 ; : : : ; xm ; 1 ; : : : ; n 0 the stalk of the structure
sheaf of Amjn is
˚
OAmjn ;p D fg j f 2 OAmjn .Am /; g 2 OAmjn .Am /0 ; g … p :
(2) Supervariety over the sphere S 2 . Consider the polynomial superalgebra gener-
ated over an algebraically closed field k, kŒx1 ; x2 ; x3 ; 1 ; 2 ; 3 , and the ideal
The local ring of kŒX 0 at the maximal ideal m0 is the ring of fractions
˚
.kŒX 0 /m D fg j f; g 2 kŒX 0 ; g … m :
178 10 Supervarieties and superschemes
The stalk of the structure sheaf at m is the localization of kŒX as a kŒX 0 -module, that
is,
kŒX m D f mg
j m 2 kŒX ; g 2 kŒX 0 ; g … mg:
Notice that if a1 ¤ 0 (not all ai are zero simultaneously), then x1 is invertible in the
localization and we have
1
1 D .x2 2 C x3 3 /;
x1
so f 2 ; 3 g generate kŒX m as an OkŒX0 -module.
X is an example of a closed subscheme of the affine superspace A3j3 . Notice also
that the reduced scheme corresponding to X is the unitary sphere S 2 A3 .
We are now going to see that the category of affine superschemes is equivalent to
the category of superalgebras; in other words, the superscheme X D Spec A and the
superalgebra A contain the same information. This is a generalization of the equiva-
lence between the category of affine schemes and the category of algebras (for a proof
of the classical result see Chapter 2).
Let .aschemes/ denote the category of affine superschemes. Let us define the functor
F W .salg/op ! .aschemes/ on the objects as F .A/ D Spec A. In order to define F
on the morphisms, let W A ! B and 0 D jA0 . We need to give a morphism
f D F ./ W Spec B ! Spec A. On the topological space we have immediately
jf j W Spec B0 ! Spec A0 defined as jf j.p/ D 01 .p/. For the sheaf morphism, we
need to give a family of morphisms,
fU W OA .U / ! OB .jf j1 .U //; U open in Spec A0 ;
`
commuting with restrictions. If a 2 OA .U /, i.e., a W U ! x2U OA;x .U / (see
Definition 2.2.6), define fU .a/ W p 7! p .a.jf j.p/// and p W A 1 .p/ ! Bp , p 2
0
Spec B0 . One can check that this determines a sheaf morphism f W OA ! f OB and
that such f is local.
We now claim that F gives an equivalence between the category of superalgebras
and the category of affine superschemes. In order to show that F realizes an equivalence
of categories, we need to prove that there is a one-to-one correspondence between
objects and morphisms in the categories. We are going to do this by establishing an
inverse functor G. On the objects the inverse functor is given by
G W .aschemes/ ! .salg/; Spec A 7! OA .A0 / Š A:
On the morphisms G is given as follows. Let f W Spec B ! Spec A be a morphism.
Since the global sections of the structure sheaves coincide with the rings B and A
respectively, we obtain immediately a morphism from A to B, so we set
G.f / D fSpec A0 W OA .Spec A0 / Š A ! OB .Spec B0 / Š B:
We leave it to the reader to verify that F and G are functors and moreover that
F G Š id and G F Š id.
We have proven the following proposition.
10.2 The functor of points 179
Examples 10.2.9. (1) Affine superspace revisited. Let A 2 .salg/ and let V D V0 ˚ V1
be a free supermodule (over k). Let .smod/ denote the category of k-modules. Define
V .A/ D .A ˝ V /0 D A0 ˝ V0 ˚ A1 ˝ V1 :
Notice that following a very common abuse of notation we are using the same letter V
to denote both the super vector space V and its functor of points. The abuse will soon
become worse when we shall also use V to denote the superscheme whose functor
of points is V . In general this functor is not representable. However, if V is finite-
dimensional, the functor is indeed representable and in fact we have
where Sym.V / denotes the symmetric algebra over the dual space V . Recall that V
V maps V !
is the set of linear
V k not necessarily preserving the parity, and Sym.V / D
Sym.V0 / ˝ V1 , where V1 denotes the exterior algebra over the ordinary space
V1 .
Let us fix a basis for V and let dim V D pjq. The functor V is represented by
kŒV D kŒx1 ; : : : ; xp ; 1 ; : : : ; q ;
(2) Supermatrices revisited. Let A 2 .salg/. Define Mmjn .A/ as the set of endo-
morphisms of the A-supermodule Amjn . Choosing coordinates we can write
² ³
a ˛
Mmjn .A/ D ;
ˇ b
where a, b are m m, n n blocks of even elements in A and ˛, ˇ are m n, n m
blocks of odd elements in A.
This is the functor of points of an affine supervariety represented by the commutative
superalgebra: kŒM.mjn/ D kŒxij ; kl where xij ’s and kl ’s are respectively even
and odd variables with 1 i; j m or m C 1 i; j m C n, 1 k m,
m C 1 l m C n or m C 1 k m C n, 1 l m.
Notice that Mmjn Š hAm2 Cn2 j2mn .
Now consider the invertible matrices in Mmjn .A/. This clearly gives us another
functor
.salg/ ! .sets/; A 7! invertible matrices in Mmjn .A/:
As we shall see in the next chapter, this is the functor of points of the superscheme
GLmjn introduced in Example 3.3.4, though this fact is far from being immediate.
This example raises a very natural question: how can we determine whether a func-
tor F W .salg/ ! .sets/ is the functor of points of a superscheme? We can reformulate
this question in an equivalent way: can we give a representability criterion for functors
.sschemes/ ! .sets/? These questions have a positive answer in the classical setting
and a comprehensive treatment can be found in [23], Ch. I, and in [29], Ch. VI. In the
next section we are going to see how the classical argument can be replicated, with
small changes, in the supergeometric environment.
The sets Uf D Spec.Af /0 D fp 2 Spec A0 j.f / 6 pg are the basic open sets in
the topological space X D Spec A0 (see Chapter 3). In fact by definition the open sets
in the Zariski topology of Spec A0 are
UI D fp 2 Spec A0 j I 6 pg
Definition 10.3.1. Let F W .salg/ ! .sets/ be a functor. Let ffi gi2I be a family of
elements in A0 , .fi ; i 2 I / D A0 and let i W A ! Afi and ij W Afi ! Afi fj be the
natural morphism. We say that the functor F is local or that it is a sheaf in the Zariski
topology if for any family f˛i gi2I , ˛i 2 F .Afi /, such that F .ij /.˛i / D F .j i /.˛j /
for all i and j , there exists a unique ˛ 2 F .A/ with F .i /.˛/ D ˛i .
The examples of local functors are many; the one that interests us most is the
following.
Proposition 10.3.2. The functor of points hX of a superscheme X is local.
Proof. We briefly sketch the proof since it is the same as in the ordinary case. Let the
notation be as above. Consider a collection of maps ˛i 2 hX .Afi / which map to the
same element in hX .Afi fj /. Each ˛i consists of a continuous map j˛i j W Spec Afi 0 !
jXj and a family of ˛i;U W OX .U / ! OAfi .j˛j1 .U // respecting restrictions. The fact
the j˛i j glue together is clear. The gluing of the ˛i ’s to give ˛ W Spec A ! X depends
on the fact that OX and OAfi are sheaves.
We now want to define the second ingredient for our representability criterion,
namely the notion of open subfunctor of a functor F W .salg/ ! .sets/. If we assume
F to be the functor of points of superscheme X, there is a sensible notion of an open
subfunctor of F D hX , namely, we could simply define it as the functor of points
of an open superscheme U X . However since we are precisely interested in a
characterization of those F that come from superschemes, we have to carefully extend
the notion of open subfunctor to subfunctors of functors which do not come necessarily
as the functor of points of a superscheme.
Definition 10.3.3. Let U be a subfunctor of a functor F W .salg/ ! .sets/; this means
that we have a natural transformation U ! F such that U.A/ ! F .A/ is injective for
all A. We say that U is an open subfunctor of F if for all A 2 .salg/, given any natural
transformation f W hSpec A ! F , the subfunctor f 1 .U / coincides with hV for some
open V in Spec A, where
W Spec R ! V D f 1 .U / ! Spec A:
186 10 Supervarieties and superschemes
So fR ./ 2 hU .R/ if and only if 2 hV .R/. We leave to the reader all the routine
checks.
We want to define the notion of an open cover of a functor.
Definition 10.3.5. Let F W .salg/ ! .sets/ be a functor. F is covered by the open
subfunctors .Ui /i2I , if and only if for any affine superscheme Spec A and natural
transformation f W hSpec A ! F we have that the fibered product hSpec A F Ui Š hVi
and .Vi /i2I is an open cover of Spec A. (For the definition of fibered product see the
Appendix B).
Notice that, by the very definition of open subfunctor, the functor hSpec A F Ui is
always representable. In fact, it is equal to f 1 .Ui / which is by definition the functor
of points of an open and affine Vi in Spec A. Moreover, as before, we have that if F is
the functor of points of a superscheme X , i.e., F D hX , the family of open subfunctors
S
.Ui /i2I covers F if and only if Ui D hWi and the Wi cover X , that is jX j D jWi j.
S
Remark 10.3.6. Notice also that asking F .A/ D Ui .A/ for all superalgebras A is
far too restrictive. This is already a phenomenon we observe at the ordinary level. For
example let us consider F D hSpec Z . F admits fU1 D hSpec ZŒ1=2 ; U2 D hSpec ZŒ1=3 g
as open affine cover, in fact fSpec ZŒ1=2; Spec ZŒ1=3g cover Spec Z. However
F .Z/ ¤ U1 .Z/ [ U2 .Z/ since F .Z/ D id, while U1 .Z/ D U2 .Z/ D ;.
We are ready to state the main result of this section that allows us to characterize,
among all the functors from .salg/ to .sets/, those which are the functors of points of
superschemes.
Theorem 10.3.7. A functor
F W .salg/ ! .sets/
is the functor of points of a superscheme X , i.e., F D hX if and only if
(1) F is local,
(2) F admits a cover by affine open subfunctors.
Proof. The proof of this result is similar to that in the ordinary case detailed in [23],
Ch. I, §1, 4, 4.4, but given its importance we shall rewrite it. We first observe that if
hX is the functor of points of a superscheme, by Proposition 10.3.2 it is local and by
Definition 10.3.5 it admits a cover by open affine subfunctors.
Let us now assume that F satisfies the properties (1) and (2) of Theorem 10.3.7. We
need to construct a superscheme X D .jX j; OX / such that hX Š F . The construction
of the topological space jX j is the same as in the ordinary case. Let us sketch it.
Let fhX˛ g˛2A be the affine open subfunctors that cover F .2 Define hX˛ˇ D hX˛ F
hXˇ , (X˛ˇ will correspond to the intersection of the two open affine X˛ and Xˇ in
2
We are now assuming that the open affine subfunctors of F are equal and not just isomorphic to hX˛ .
We leave to the reader the checks necessary to verify all the statements in the slightly more general setting
required by the theorem.
10.3 A representability criterion 187
the superscheme X , that we are now constructing). Notice also that hX˛ F hXˇ is
representable, as we remarked after Definition 10.3.5.
We have the commutative diagram
jˇ;˛
hX˛ˇ D hX˛ F hXˇ / hXˇ
j˛;ˇ iˇ
i˛
hX˛ / F.
As a set we define `
jX j ´ jX˛ j= ;
˛2A
t˛ t
i˛
hX˛ / F.
Since the hX˛ are an open affine cover of F , the fT˛ g form an open affine cover
of Spec A. Since by Yoneda’s lemma Hom.hT˛ ; hX˛ / Š Hom.T˛ ; X˛ /, we obtain a
188 10 Supervarieties and superschemes
Proof. The proof is a simple exercise that makes use of Proposition 10.1.9 and the
universal properties of the tensor product and the fibered product.
Corollary 10.3.9. Fibered products exist in the category of superschemes. The fibered
product X Z Y , for superschemes X , Y , Z,
X Z Y /Y
g
f
X / Z,
hXi ˛ .A/ hZi .A/ hYjˇ .A/ D .hXi ˛ .A/ hZ .A/ hY .A// \ .hX .A/ hZ .A/ hYjˇ .A//;
also hXi ˛ hZi hYjˇ is an open subfunctor of F . Hence fhXi ˛ hZi hYjˇ g form an open
covering of F by open subfunctors. These functors are affine since if Xi˛ D Spec Ri˛ ,
Yjˇ D Spec Sjˇ and Zi D Spec Ti , hXi ˛ hZi hYjˇ D hSpec Ri ˛ ˝Ti Sjˇ
Hence by the representability criterion, Theorem 10.3.7, F is the functor of points
of a superscheme, which is X Z Y , as one can readily check.
We end this section with some remarks on an equivalent approach to the problem
discussed in Theorem 10.3.7.
As we have seen in the previous section, the functor of points of a superscheme X
can be also equivalently defined as the functor hX W .sschemes/op ! .sets/, hX .T / D
Hom.T; X/. Hence the question whether a generic functor F W .sschemes/op ! .sets/
is (isomorphic to) the functor of points of a superscheme, can be reformulated by asking
the conditions for F to be representable.
In such a setting, we can give an equivalent definition for F to be local, formulated
as a proposition, whose proof is a straightforward check, based on the results of the
previous section.
Let F W .sschemes/op ! .sets/ be a functor. Since the category of affine super-
schemes is equivalent to the category of commutative superalgebras, we can regard
F restricted to the category of affine superschemes as a functor from the category of
superalgebras to the category of sets. We shall call such functor F a .
Proposition 10.3.10. Let F W .sschemes/op ! .sets/ be a functor. Then F a is local if
and only if the following condition is satisfied. Let X be a superscheme, Xi an open
cover of X. Assume that there is a family of elements fi 2 F .Xi / such that fij D fj i
for all i; j , where fij is the image of fi under the natural map F .Xi / ! F .Xi \ Xj /
induced by the inclusion Xi \Xj Xi . Then there exists a unique f 2 F .X / mapping
to each fi under the natural map F .X / ! F .Xi /.
The analogue of the notion of an open affine cover of such F (see Definition 10.3.5)
makes sense as it is in this setting, hence we are ready for the Representability Theorem.
Theorem 10.3.11 (Representability criterion for superschemes). A functor
F W .sschemes/op ! .sets/
By its very definition this functor is local. This is perhaps best seen by using the
correspondence detailed in Proposition 10.3.10:
In other words, we are considering the functor Gr extended to the category of su-
perschemes: Gr W .sschemes/op ! .sets/ (by a common abuse of notation we use
the same symbol to denote it). If fTi g is an open cover of the superscheme T , as-
sume that we have a family of subsheaves Fi 2 Gr.Ti /, Fi OTmjn i
D OTmjn jTi ,
with Fi jTi \Tj D Fj jTi \Tj . Then it is clear the Fi ’s glue together to give a subsheaf
F OTmjn on T , still locally constant of rank rjs, hence F 2 Gr.T /.
We have shown that Gr is the functor of points of a superscheme that we will call
the Grassmannian superscheme (or superGrassmannian for short) of rjs subspaces in
an mjn-dimensional space.
S.n/d D S nCd :
Clearly also S.n/ is an S -module (recall that upper indices refer to the Z-grading while
lower indices to the Z2 -grading).
Let M be a graded S-module. Then M is also an S0 -module, hence we can follow
z . One
the classical recipe (see Chapter 2, Section 2.5) and build the sheaf on Proj S0 , M
can check right away that M z .U / is an OS .U /-module for all open sets U in Proj S0 .
We summarize the properties of M z in the following theorem.
(2) .Mz /p Š Mp for all homogeneous prime ideals p 2 Proj S0 , i.e., the stalk at any
prime p of the sheaf Mz coincides with the localization of the S0 -module M at p.
(3) For all homogeneous f 2 S n , M f
A
z jU D M.f / , where Uf is the basic open
set in Proj S0 corresponding to f , M.f / are the elements of degree zero in the
localization Mf (this is commonly called projective localization).
Proof. We leave the proof to the reader as an exercise: it is done precisely as in the
ordinary setting that is found in [43], Ch. II, Section 5.
At this point one could define supercoherent sheaves as sheaves modeled after M z,
in analogy to what we have done in Capter 2, but we shall not take this direction.
We now turn to the generalization of the Serre’s twisting sheaf in order to charac-
terize the morphisms from a superscheme to the projective superspace.
e
Definition 10.5.2. We define OS .n/ as the sheaf S.n/. Note that OS .1/ is called
Serre’s twisting sheaf.
OS .n/ is a locally free OS -module of rank 1. The proof is essentially the same as
the classical one (see [43], Ch. II, Section 5).
Proof. This is an exercise, once one follows the proof for the ordinary setting; see [43],
Ch. II, Section 7.
For a generic sheaf of rings F , let F denote the sheaf of invertible sections, i.e.,
F .U / are the invertible sections in F .U /. The first cohomology group H 1 .jX0 j; OX0 /
for an ordinary scheme X0 classifies the equivalence classes of line bundles on X0 ([43],
194 10 Supervarieties and superschemes
Ch. III, Section 4). The next proposition tells us that the same happens also for super
line bundles. We leave the proof of this result to the reader since it is a precise replica
of the classical one.
We now turn to the main goal of this section: we want to show that the Grassmannian
superscheme cannot in general be embedded into any projective superspace. This
statement appears for the analytic category in [56]; the reasoning here is the same, we
include it for completeness. We start with some general considerations on super line
bundles on superschemes.
Let X be a superscheme and let X0 be its reduced scheme, i.e., X D .jX j; OX /,
X0 D .jXj; OX0 / D .jX j; OX =/ where is the ideal sheaf generated by the odd
nilpotents.
Consider the exact sequence of sheaves
0 ! ! OX ! OX0 ! 0:
! H 1 .jX j; OX / !
H 1 .jX j; OX0 /
! H 2 .jX j; / ! :
which is the stabilizer of the element he1 ; 1 i 2 X.A/. This leads to the usual identifi-
cation of points in X.A/ with matrices in GL2j2 .A/=P .A/.
The open covering of jX j D P 1 P 1 described above induces the following open
covering of X by the open affine subfunctors Uij :
80 19 80 19
ˆ
ˆ 1 0 >
> ˆ
ˆ u1 >
>
<B C= <B C=
B u0 C B 1 0 C
U00 .A/ D @ A> ; U10 .A/ D ˆ@ 0 A> ;
ˆ 0 1 > 1 >
:̂ ; :̂ ;
v0 v0
80 19 80 19
ˆ
ˆ 1 0 >
> ˆ
ˆ u1 >
>
<B C = <B C =
B u0 C B 1 0 C
U01 .A/ D @ A> ; U .A/ D A> :
ˆ v1 >
11
ˆ@ v1 >
:̂ ; :̂ ;
0 1 0 1
Assume that we have a line bundle on X and let us see how sections transform
under coordinate changes. Generic sections on jU00 j and jU11 j respectively are given
by
Let us see the behaviour of the transition function on the intersection jU00 j \ jU11 j. A
straightforward calculation shows that (see also Chapter 1, Section 1.5):
1
u0 .u0 v01 /1 u1 .v0 u1 /1 /
D 0 0 :
v0 v01 .u0 v01 /1 .v0 u1
0 /
1
This allows us to view a generic element in U00 .A/ as an element in U11 .A/:
0 1 0 1
1 0 .u0 v01 /1 u1 1 1
0 .v0 u0 /
Bu0 C B 1 0 C
B C 7! B 1 C:
@0 1 A @v0 .u0 v01 /1 .v0 u1
0 /
1 A
v0 0 1
We have that sections transform in the following way:
where x is an invertible element in jU00 j \ jU11 j. This means that in order to have a
polynomial function the invertible element x must contain a power of the Berezinian.
When we consider this transition function, together with all the requirements of the
other variables, which we do not compute here, one sees that u0 and v0 appear with
exponents with opposite sign. This means that one copy of P 1 can get embedded (the
one corresponding to the positive sign) while the other cannot.
With this particular affine covering, X cannot have any projective embedding since
its reduced variety X0 does not admit such embedding.
In the course of our calculation, we have computed the generic element in
H 1 .U; OX / using the covering U D fU00 ; U01 ; U10 ; U11 g of jX j and Čech coho-
mology. Since H 1 .U; OX / D H 1 .jX j; OX /, the argument we give for the special
covering is actually a generic argument and shows that the image of the map consists
only of line bundles with no global sections, hence they will not correspond to line bun-
dles that classically projectively embed into X0 D P 1 P 1 . Hence X D Gr.1j1; 2j2/
does not admit any projective embedding.
however shall restrict our attention to the rational points since, as we shall see, they
are in one-to-one correspondence with the k-points of the superscheme and in most
applications these are the only interesting points to consider.
It is important not to confuse the points of the topological space jX j with the ele-
ments in hX .A/ for a generic A 2 .salg/. These are called A-points of the superscheme
X. The next observation clarifies the relationship between the points of jX j and hX ,
the functor of points of X .
Tx X D Der.OX;x ; k/;
where k is viewed as an OX;x -module via the identification k Š OX;x =mX;x , with
mX;x the maximal ideal in OX;x .
The next proposition gives an equivalent definition for the tangent space.
2
Note that mX;x =mX;x is an OX;x -supermodule which is annihilated by mX;x , hence
it is a k D OX;x =mX;x -supermodule i.e., a super vector space.
Proof. Let D 2 Der.OX;x ; k/. Since D is zero on k and OX;x D k ˚ mX;x , we have
that D is determined by its restriction to mX;x , DjmX;x . Moreover, since mX;x acts as
zero on k Š OX;x =mX;x , one can check that
W Der.OX;x ; k/ ! Hom.smod/ .mX;x =mX;x
2
; k/; D 7! DjmX;x ;
is well defined.
Now we construct the inverse. Let ˛ W mX;x ! k, ˛.mX;x
2
/ D 0. Define
D˛ W OX;x D k ˚ mX;x ! k; D˛ .a; f / D ˛.f /:
This is a well-defined superderivation.
1
Moreover one can check that the map ˛ 7! D˛ is .
The next proposition provides a characterization of the tangent space that is useful
for explicit calculations.
Proposition 10.6.8. Let X D .jX j; OX / be a supervariety x 2 jX j a rational closed
point. Let U be an affine neighbourhood of x, mx OX .U / the maximal ideal
corresponding to x. Then
Tx X Š Hom.smod/ .mX;x =mX;x
2
; k/ Š Hom.smod/ .mx =m2x ; k/:
Proof. The proof is the same as in the ordinary case and is based on the fact that
localization commutes with exact sequences.
Let us compute explicitly the tangent space in an example.
Example 10.6.9. Consider the affine supervariety represented by the coordinate ring
CŒx; y; ; =.x C y/:
Notice that the reduced variety is the affine plane.
Since C is algebraically closed, all closed points are rational. Consider the closed
point P D .1; 1; 0; 0/ Š mP D .x 1; y 1; ; / CŒx; y; ; =.x C y/, where
we identify .x0 ; y0 ; 0; 0/ with maximal ideals in the ring of the supervariety, as we do
in the commutative case. By Proposition 10.6.8, the tangent space at P consists of all
the functions ˛ W mP ! k, ˛.mP2 / D 0.
A generic f 2 mP lifts to the family of f D f1 C f2 .x C y/ 2 CŒA2j2 D
CŒx; y; ; with f1 .1; 1; 0; 0/ D 0 and where f2 is any function in CŒA2j2 D
CŒx; y; ; . Thus f can be formally expanded in power series around P :
@f1 @f1
f D @x
.P /.x 1/ C @y
.P /.y 1/ C . @f
@
1
.P / C f2 .P //
C . @f
@
1
.P / C f2 .P // C higher order terms:
10.6 The infinitesimal theory 199
Define
@f1 @f1 @f1 @f1
XD @x
.P /; Y D @y
.P /; „D @
.P /; ED @
.P /:
These are coordinates for the super vector space MP =MP2 , MP D .x 1; y 1; ; /
CŒx; y; ; . A basis for the dual space .Mp =Mp2 / , which is the tangent space to A2j2 ,
consists of functions sending the coefficient of one of the x 1, y 1, , to a non-zero
element and the others to zero. We get equations for the tangent space .mP =mP2 / as
a subspace of .MP =MP2 / :
„ E D 0:
So we have described the tangent space .mP =mP2 / as a subspace of .MP =MP2 / ,
the tangent space to the affine superspace Amjn .
There is yet another way to compute the tangent space, in the case X is an affine
supervariety. Before we examine this construction, we must understand first the notion
of differential of a function and differential of a morphism. We start by defining the
value and the differential for the germs of sections.
f .x/ ´ .f /:
Notice that f f .x/ 2 mX;x , where we interpret f .x/ 2 k OX;x . We also define
the differential of f at x to be
We leave to the reader the simple check that this definition is independent from the
chosen open neighbourhood U of x.
200 10 Supervarieties and superschemes
.dxi /P D x xi0 ; .d j /P D j; i D 1; : : : ; m; j D 1; : : : ; n:
W Mx =M2x mx =m2x :
between superspaces. Let us recall the following simple fact of linear algebra.
If a W V1 ! V2 is a surjective linear map between finite-dimensional vector spaces
V1 , V2 and b W V2 V1 is the injective linear map induced by a on the dual vector
spaces, then s 2 im.b/ if and only if sjker.a/ D 0.
We apply this observation to the maps and the differential .d˛/x ,
Observe that ker. / D f.df /x j f 2 I g. By identifying Amjn D k mjn with its double
dual .k mjn / we obtain the result.
Remark 10.6.16. In the notation of the previous proposition, if I D .f1 ; : : : ; fr / one
can readily check that
thus obtaining a quick and effective method to calculate the tangent space to a super-
variety.
Let us revisit Example 10.6.9 and see how the calculation is made using Proposi-
tion 10.6.15.
Example 10.6.17. Consider again the supervariety represented by the superalgebra
Hence by Proposition 10.6.15 the tangent space is the subspace of k 2j2 given by the
equation
D 0:
10.7 References
The definition of superscheme together with its functor of points appears in the work
of Manin [56], where also the examples of the Grassmannian and flag superschemes
are described in detail in the complex analytic category. The representability criterion
Theorem 10.3.7 appears in [23], Ch. I, for the ordinary setting and is used, however
not formally proved, in the work by Manin [56].
11
Algebraic supergroups
In this chapter we restrict our attention to the superschemes that have an additional
structure, namely the group structure, and thus are called supergroup schemes or sim-
ply supergroups for short. The simplest way to introduce this extra structure is the
requirement for the functor of points to be group-valued, that is, if X is a superscheme,
we are asking that hX .A/ be a group for each superalgebra A.
As a general rule in supergeometry, the functor of points is valued in the ordinary
categories like sets, groups, vector spaces or Lie algebras and the supernature of the
geometrical object stems from the category we start from, in our case superalgebras
or superschemes. As it happens for the ordinary setting, when the superscheme G is
affine, G is a supergroup if and only if its representing superalgebra O.G/ is a Hopf
superalgebra. We shall discuss in detail the example of the general and special linear
supergroups together with their Hopf superalgebras.
As for the ordinary setting, we can associate very naturally to any supergroup its
Lie superalgebra, which is a Lie algebra-valued functor, identified with the functor of
points of the super tangent space of the supergroup at the identity. We also introduce
the concept of representation of a supergroup and we prove the following two important
results: we show that any affine supergroup can be embedded into some GL.V / for
a suitable V , and then we prove the representability of the stabilizer functor for the
action of a supergroup on a superscheme. This gives the representability of the classical
algebraic supergroups corresponding to the classical Lie superalgebras (see Appendix A
for their list).
Our treatment follows very closely [23], Ch. II; most of the classical statements go
unchanged to the super-setting and we shall point out the differences when they arise.
G W .salg/ ! .sets/;
id
GG / G.
ide /
GG o
eid
G ekK ek G
KK ss
KK s
KK ss
ss
KK
K% ysss
G.
.id;i/
G /GG
ek e / G.
The supergroup functors together with their morphisms, that is the natural trans-
formations that preserve , e and i, form a category.
If G is the functor of points of a superscheme X , i.e., G D hX , in other words
G.A/ D Hom.Spec A; X /, we say that X is a supergroup scheme. An affine su-
pergroup scheme X is a supergroup scheme which is an affine superscheme, that is
X D Spec O.X / for some superalgebra O.X /. To make the terminology easier we
will drop the word “scheme” when speaking of supergroup schemes, whenever there
is no danger of confusion.
As we shall presently see, the functor of points of an affine supergroup is represented
by a superalgebra, which has the additional structure of a Hopf superalgebra. We refer
the reader to Chapter 1, Section 1.7, for their definitions and main properties.
204 11 Algebraic supergroups
where A is the multiplication in the superalgebra A. One can immediately check that
the multiplication is a morphism, that is,
(though hidden, the sign rule plays a crucial role here). This multiplication in hG .A/
gives rise to in O.G/ as its associated comultiplication, as one can readily see.
The unit eA and the inverse iA in hG .A/ are defined by
A
eA D A B W O.G/ !
k ! A; iA .x/ D x B S;
where A is the unit in A. We leave to the reader the routine checks of Definition 11.1.1.
Vice versa, if G is a supergroup scheme, then we can define the comultiplication
W O.G/ ! O.G/ ˝ O.G/ as the dual of the multiplication 2 Hom.G G; G/
using the identification
(one can readily check that O.G G/ Š O.G/ ˝ O.G/). Similarly one defines the
counit and the antipode and S as the duals of unit e and inverse i .
A careful look shows that formally the diagrams defining a supergroup functor
are essentially the same as those defining a Hopf superalgebra, with arrows reversed.
We leave to the reader the routine verification that hG satisfies all the diagrams in
Definition 11.1.1 if and only if O.G/ satisfies the diagrams in Definition 1.7.1.
The equivalence between the categories of affine supergroups and commutative
Hopf superalgebras is an immediate consequence of the previous discussion and Propo-
sition 10.1.9.
Let us now examine some important examples of supergroup schemes and their
associated Hopf superalgebras.
Examples 11.1.3. (1) Supermatrices Mmjn . In Remark 1.4.1 we have introduced the
functor of points of supermatrices
a ˛
Mmjn W .salg/ ! .sets/; A 7! ;
ˇ b
11.1 Supergroup functors and supergroup schemes 205
where a and b are m m, n n block matrices with entries in A0 , while ˛ and ˇ are
mn, nm block matrices with entries in A1 . As we have seen in Example 3.3.4, Mmjn
is a representable functor, represented by the superalgebra of polynomials kŒxij ; kl
for suitable indices i , j , k, l. The functor Mmjn is group-valued, in fact any Mmjn .A/
has an additive group structure, where the addition is simply defined as the addition of
matrices. Hence by the previous proposition kŒxij ; kl is a Hopf superalgebra, where
the comultiplication , the counit and antipode S are given by
We leave to the reader the verification that kŒxij ; kl together with , and S is a
Hopf superalgebra.
(2) The general linear supergroup GLmjn . Let A 2 .salg/. Let us define GLmjn .A/
as GL.Amjn / (see Chapter 1) to be the set of automorphisms of the A-supermodule
Amjn . Choosing the standard basis we can write
² ³
a ˛
GLmjn .A/ D Mmjn .A/:
ˇ b
where xij ’s, U , V and kl ’s are respectively even and odd variables with 1 i; j m
or m C 1 i; j m C n, 1 k m, m C 1 l m C n or m C 1 k m C n,
1 l m and
X
d1 D .1/l.s/ x1;s.1/ : : : xm;s.m/ ;
s2Sm
X
d2 D .1/l.t/ xmC1;mCt.1/ : : : xmCn;mCt.n/ :
t2Sn
X
mCn
.aij / D aik ˝ akj ; .aij / D ıij :
kD1
We need to specify , and S also on the generators d11 and d21 and then to
check that they are well defined with respect to the ideal of the relations d11 d1 D 1,
d21 d2 D 1:
X
2mnC2
i i
.d11 / D .1/i1 .d11 / ˝ .d11 / . .d1 / d11 ˝ d11 /i1 ;
iD1
X
2mnC2
i i
.d21 / D .1/i1 .d21 / ˝ .d21 / . .d2 / d21 ˝ d21 /i1 ;
iD1
.d11 / D .d21 / D 1:
We leave to the reader the tedious verification that , are well defined and satisfy
respectively the properties of the comultiplication and counit as in Definition 1.7.1 and
they correspond to the group structure of the functor GLmjn .
As for the antipode, the definition turns out to be more complicated and this is
a consequence of the fact that the inverse for a supermatrix has a far more compli-
cated formula than the inverse of an ordinary matrix. For such formulas, we refer the
interested reader to [30].
(3) The special linear supergroup SLmjn . For a superalgebra A, let us define
SLmjn .A/ to be the subgroup of GLmjn .A/ consisting of matrices with Berezinian
equal to 1. This is the functor of points of an affine supergroup and it is represented by
the Hopf superalgebra
1 1
O.SLmjn / D kŒxij ; kl Œd1 ; d2 =.Ber 1/;
where the comultiplication, counit and antipode are inherited naturally from the ones
in GLmjn .
for which OG;1G =m1G Š k, where m1G is the maximal ideal in the local superring
OG;1G , in the following way. By definition, the unit e is a morphism of superschemes
e W ek ! G, hence corresponds to a pair of morphisms e D .jej; e /, jej W jek j ! jGj,
e W OG ! e Oek D k. Define 1G ´ jej.jek j/ 2 jGj. This is a rational point, in
fact OG;1G =m1G Š k. Moreover notice that, by the very definition of e, 1G has the
property of a unit for the topological group jGj.
such that
• a.xy/ D .ax/.ay/,
• .ab/x D a.bx/
for all a; b 2 Ok .A/ and x; y 2 M.A/,
For example if M is the functor of points of a super vector space V , that is, M.A/ D
.A ˝ V /0 , we can define the operation .x; y/ 7! x C y that associates to a pair of
elements in the A0 -module M.A/ their sum. Then M has a natural structure of an
Ok -module, which is the multiplication of elements in M.A/ by scalars in A0 .
208 11 Algebraic supergroups
Definition 11.2.2. Let g be a super vector space. We say that the functor
Remark 11.2.3. If the super vector space g is finite-dimensional, then the functor Lg
is representable and we have Lg D hg . In fact
where .smod/ denotes the category of supermodules (over k in this case) and Sym.g /
the symmetric algebra over g . Notice also that in this case Lg is the functor of points
of an affine superscheme represented by the superalgebra Sym.g /.
We now want to see that the usual notion of Lie superalgebra, as we defined in
Chapter 1, Section 1.2, is equivalent to this functorial definition. We want to show that
g is a Lie superalgebra if and only if the functor Lg is Lie algebra-valued.
Let us first recall the definition of Lie superalgebra given in Chapter 1, Section 1.2.
Definition 11.2.4. Let g be a super vector space. We say that g is a Lie superalgebra
if there exists a bilinear map Œ ; W g g ! g called a superbracket such that
(a) Œx; y D .1/p.x/p.y/ Œy; x,
(b) Œx; Œy; z C .1/p.x/p.y/Cp.x/p.z/ Œy; Œz; x C .1/p.x/p.z/Cp.y/p.z/ Œz; Œx; y
for all x; y; z 2 g.
Proof. This is an immediate consequence of the even rules principle, detailed in Chap-
ter 1, Section 1.8. Nevertheless, given the importance of this construction, we want to
see explicitly how the superbracket on g and the bracket on Lg .A/ correspond to each
other.
If we have a Lie superalgebra Lg there is always, by definition, a super vector
space g associated to it. Moreover by the even rules, there is a unique Lie superalgebra
structure on the A-module LA g D A ˝ g, whose even part is the Lie algebra Lg .A/.
11.2 Lie superalgebras 209
Clearly the bracket fv; wg 2 g does not depend on a; b 2 A. One can verify that it is
a superbracket. Let us see, for example, the antisymmetry property. Observe first that
if a ˝ v 2 .A ˝ g/0 , then p.v/ D p.a/ since .A ˝ g/0 D A0 ˝ g0 ˚ A1 ˝ g1 . So
we can write
Œa ˝ v; b ˝ w D .1/p.b/p.v/ ab ˝ fv; wg:
On the other hand,
Œb ˝ w; a ˝ v D .1/p.a/p.w/ ba ˝ fw; vg
D .1/p.a/p.w/Cp.a/p.b/ ab ˝ fw; vg
D .1/2p.w/p.v/ ab ˝ fw; vg
D ab ˝ fw; vg:
By comparing the two expressions we get the antisymmetry of the superbracket. For
the super Jacobi identity the calculation is the same.
A similar calculation also shows that, given a super Lie algebra g, the functor Lg
is Lie algebra-valued.
Hence the previous proposition shows that a Lie algebra-valued functor Lg ac-
cording to Definition 11.2.2 is equivalent to a super Lie algebra g. With an abuse of
language we will refer to both g and Lg as “Lie superalgebra”.
Remark 11.2.6. Given a super vector space g one may also define a Lie superalgebra
to be the representable functor Dg W .salg/ ! .sets/ so that
This is a Lie algebra-valued functor once we define a Lie bracket on each Mmjn .A/:
One word of warning: the super vector space M.mjn/ is not Mmjn .k/. In fact Mmjn .k/
consists only of the even part of the super Lie algebra M.mjn/ and contains just the
diagonal block matrices with entries in k, while M.mjn/, as a vector space, consists of
.m C n m C n/-matrices with entries in k and has superdimension m2 C n2 j2mn.
The purpose of the next two sections is to naturally associate a Lie superalgebra
Lie.G/ to a supergroup G.
Definition 11.3.1. Consider the homomorphism G.p/ W G.A.// ! G.A/. For each
G there is a supergroup functor
1O /1
O
fA
G.A/ / H.A/
O O
G.p/ H.p/
fA./
G.A.// / H.A.//
O O
fA./ jLie.G/.A/
Lie.G/.A/ / Lie.H /.A/
O O
1 / 1.
We hence define: Lie.f /.A/ D fA./ jLie.G/.A/ . The following proposition is im-
mediate.
Proposition 11.3.2. Lie is a functor from the category of supergroup functors to the
category of groups.
Examples 11.3.3. (1) The general linear Lie superalgebra. We want to determine the
functor Lie.GLmjn / for k a field. Consider the morphism
p C p 0 q C q 0 p q
GLmjn .p/ W GLmjn .A.// ! GLmjn .A/; 7 ! ;
r C r 0 s C s 0 r s
where In is an nn identity matrix. The functor Lie.GLmjn / is clearly group-valued and
can be identified with the (additive) group functor Mmjn defined as (see Example 11.2.7)
det.In s 0 / det.Im C p 0 / D 1;
which gives
tr.p 0 / tr.s 0 / D 0:
Hence
Lie.SLmjn /.A/ D fX 2 Mmjn .A/ j str.X / D 0g;
where str is the supertrace, i.e., str a dˇ D tr.a/ tr.d /.
In the previous examples we have discovered that Lie.G/ comes with a built-in
Ok -module structure. This is actually true in general, that is, for all superalgebras
A, we have that Lie.G/.A/ is an A0 -module in a functorial manner. In fact, let
ua W A./ ! A./ be the endomorphism, ua .1/ D 1, ua ./ D a, for a 2 A0 . ua
induces G.ua / W G.A.// ! G.A.//, that by compatibility conditions gives a well-
defined morphism Lie.G/.ua / W Lie.G/.A/ ! Lie.G/.A/. Hence there is a natural
transformation Ok Lie.G/ ! Lie.G/ such that
for any superalgebra A. One can immediately check for GLmjn .A/ and its subgroups
that the morphism .a; x/ 7! ax corresponds to the multiplication of all the entries of
the matrix x by a 2 A0 .
We summarize our discussion with a proposition, the proof of which is a simple
check that we leave to the reader.
Proposition 11.3.4. Let G be a supergroup functor. Then Lie.G/ is an Ok -module
with respect to the morphism .a; x/ 7! ax detailed above. That is:
11.4 Lie.G/ for a supergroup scheme G 213
where fU is the composition of a natural map from OX .U / to the direct limit OX;x and
the projection OX;x ! OX;x =mX;x2
(with x 2 U ).
as Ok -modules.
We now want to build an inverse for . Let z 2 ker.hG .p// hG .A.//, i.e.,
hG .p/z D 1G.A/ , where
Spec A./ z /G
;
O www
ww
w
ww
Spec p
ww 1G.A/
Spec A,
Lie.G/ Š T1G G;
that is, Lie.G/ is identified as a super vector space with the tangent space to G at the
identity.
Proof. Immediate from Theorem 11.4.3 and Proposition 10.6.8.
defined in the previous section, is Lie algebra-valued and is represented by a super affine
space, which carries a Lie superalgebra structure.
Our treatment of these topics again follows closely the classical discussion that is
found in [23], Ch. II. We nevertheless find it necessary to repeat all of the arguments,
given our different scope and in order to make the text self-contained.
We now want to define a natural transformation Œ ; W Lie.G/ Lie.G/ ! Lie.G/
which has the properties of a bracket.
Let GL.Lie.G//.A/ be the (multiplicative) group of linear automorphisms of the
super vector space Lie.G/ and let End.Lie.G//.A/ be the (additive) group of linear
endomorphisms of Lie.G/.A/:
GL.Lie.G//.A/ D f W Lie.G/.A/ ! Lie.G/.A/ j invertible g End.Lie.G//.A/:
The natural Ok -module structure of Lie.G/ makes the two supergroup functors
(one multiplicative the other additive)
GL.Lie.G// W .salg/ ! .sets/; End.Lie.G// W .salg/ ! .sets/
valued in the A0 -linear morphisms of Lie.G/.A/.
As in the ordinary setting, we have an identification between the Lie superalge-
bra of the general linear supergroup Lie.GL.Lie.G/// and the Lie superalgebra of
endomorphisms End.Lie.G//, as we shall see in the next proposition.
Proposition 11.5.1. Let V be a super vector space. Then
Lie.GL.V // D End.V /:
Proof. Let us first define a natural transformation W End.V / ! Lie.GL.V // and
then we show that it is an isomorphism. Let f 2 End.V /. If is an even indeterminate
2 D 0, we have 1 C f 2 End.V /.A.// and actually 1 C f 2 GL.V /.A.// since it
is invertible, its inverse being 1 f . Clearly 1 C f 2 ker GL.p/ D Lie.GL.V //.A/,
so we define .f / D 1 C f . We have immediately that is functorial and A0 -linear.
Now we show that is an isomorphism. Recall in general that if V is a super vector
space we have
End.V /.A.// ´ End.V .A./// Š End.V .A// ˝ A./ D End.V /.A/ ˝ A./:
Hence
End.V /.A.// Š End.V /.A/ ˚ End.V /.A/;
GL.V /.A.// Š GL.V /.A/ ˚ End.V /.A/;
which gives us the result.
As in the ordinary Lie theory, we can define the adjoint morphisms Ad and ad, which
again play a key role. Though we have already defined Ad for supergroup functors, we
prefer to repeat the definition in the context of supergroup schemes, since the notation
will be different.
11.5 The Lie superalgebra of a supergroup scheme 217
Ad W hG ! GL.Lie.G//;
Ad.g/.x/ D hG .i /.g/xhG .i /.g/1 ; g 2 hG .A/; x 2 Lie.G/.A/:
Our goal is to prove that Œ ; is a Lie bracket for all superalgebras A in a functorial
manner or, equivalently, that Lie.G/ is a Lie algebra-valued functor.
Observation 11.5.3. Let c.g/ W hG .A/ ! hG .A/ be the conjugation by an element
g 2 jGj: c.g/.x/ D gxg 1 . Then
Ad.g/ D Lie.c.g//:
Lie.G/.A/
Lie.c.g//
/ Lie.G/.A/
c.g/A./
hG .A.// / hG .A.//
O O
c.g/A
hG .A/ / hG .A/,
Since hGLmjn .i/ W GLmjn .A/ ! GLmjn .A.// is an inclusion if we identify GLmjn .A/
with its image, we can write
Definition 11.5.5. Let W A./ ! B.˛/ be a superalgebra morphism such that ./ D
˛ and , ˛ are two even indeterminates with square zero. Let x 2 Lie.G/.A/
hG .A.//. Define e ˛x D hG ./.x/ 2 hG .B.˛//.
f .e ˛x / D e ˛ Lie.f /x :
Lie.f /
hG .A.// Lie.G/.A/ / Lie.H /.A/ hH .A.//
x / Lie.f /.x/
hG . / hH . /
fB.˛/
hG .B.˛// / hH .B.˛/
/ e ˛ Lie.f /x .
e ˛x
Remark 11.5.7. One can readily check that if G D GL.V /, Lie.G/ D End.V /, we
have
e x D 1 C x:
11.5 The Lie superalgebra of a supergroup scheme 219
Lemma 11.5.9. Let the notation be as above and let , 0 be two even elements with
square zero. Then
0 0 0
e x e y e x e y D e Œx;y 2 hG .A.; 0 //:
.e x / D e Lie. /x
D id C Lie. /x:
Hence
id C 0 Lie. /Œx; y
D .id C Lie. /x/.id C 0 Lie. /y/.id Lie. /x/.id 0 Lie. /y/;
which gives us
is a Lie bracket for all A. Lie.G/ is a Lie algebra functor and is represented by a Lie
superalgebra (still denoted by Lie.G/).
Notice that in the definition of superderivation d W kŒG ! k the sign .1/p.d /p.f /
can be omitted: in fact, whenever p.f / D 1, we have .f / D 0 since is a morphism
and therefore it preserves the parity.
We denote by Der.kŒG; kŒG/ the set of superderivations D W kŒG ! kŒG and b
Der.kŒG; k/ the set of superderivations d W kŒG ! k.
Proposition 11.6.2. The set L.G/ of left-invariant derivations of kŒG is a Lie super-
algebra with bracket
. B D1 D2 /f D . B D1 /.D2 f /
D .id ˝ D1 B B D2 /.f /
D ..id ˝ D1 / B .id ˝ D2 / B /.f /
D ..id ˝ D1 D2 / B /.f /:
Similarly,
Hence
B ŒD1 ; D2 D .id ˝ ŒD1 ; D2 / B ;
as we wanted to show.
In Chapter 7 we have seen the identification between the left-invariant vector fields
and the Lie superalgebra of a Lie supergroup. In the algebraic context the next theo-
rem provides an analogy establishing a one-to-one correspondence between the left-
invariant derivations and the Lie superalgebra of an affine algebraic supergroup.
Theorem 11.6.3. Let G be an affine supergroup scheme. Then we have natural bijec-
tions among the sets:
(1) L.G/ the left-invariant derivations in Der.kŒG; kŒG/,
(2) Der.kŒG; k/,
(3) Lie.G/.
Proof. Let us examine the correspondence between (1) and (2). We want to construct
an invertible map W Der.kŒG; k/ ! L.G/. Let d 2 Der.kŒG; k/. Define .d / D
.id ˝ d / . Such .d / is left-invariant; in fact,
. B .d //.f / D .
B id ˝ d B /.f /
P .1/
D . f d.f .2/ //
P
D .id ˝ id ˝ d /. .f .1/ / ˝ f .2/ /
D ..id ˝ id ˝ d / B . ˝ id/ B /.f /;
P
where we use the Sweedler notation: .f / D f .1/ ˝ f .2/ . On the other hand we
have
.id ˝ .d // .f / D ..id ˝ id ˝ d / B .id ˝ / B /.f /
D .id ˝ .id ˝ d B / B /.f /
which is the same as before since by the coassociativity axiom in a Hopf superalgebra
. ˝ id/ B D .id ˝ / B .
Moreover, .d / is a derivation. In fact, since d is a derivation, it follows that
.d /.fg/ D .id ˝ d / B .fg/
P
D f .1/ g .1/ d.f .2/ g .2/ /
P
D f .1/ g .1/ .d.f .2/ /.g .2/ / C .1/p.d /p.f / .f .2/ /d.g .2/ //:
We can rewrite the last expression as (m denotes the multiplication)
P
m. .f .1/ ˝ d.f .2/ //.g .1/ ˝ .g .2/ ///
C .1/p.d /p.f / m..f .1/ ˝ .f .2/ //.g .1/ ˝ d.g .2/ ///
D .id ˝ d / .f /.id ˝ / .g/ C .1/p.d /p.f / .id ˝ / .f /.id ˝ d / .g/
D .d /.f /g C .1/p.d /p.f / f .d /.g/;
11.7 Linear representations 223
since we recall that, by the Hopf superalgebra axioms, .id ˝ / B D id. Then
.d / 2 Der.kŒG; kŒG/. Vice versa, if D 2 L.G/ define .D/ D B D. We leave
to the reader, as an easy exercise, the check that is a derivation. One can check
also that is the inverse of . We now want a correspondence between (b) and (c).
By Theorem 11.4.3 we have Lie.G/ D Hom.smod/ .T1G .G/ ; A/ D Der.OG;1G ; k/.
Observe that, as in the ordinary case,
Der.OG;1G ; k/ D Der.kŒG; k/;
that is, the derivation on the localization of the ring kŒG is determined by the derivation
on the ring itself.
Definition 11.7.3. Let V be a super vector space. Then V is said to be a left G-comodule
if there exists a linear map
V W V ! kŒG ˝ V;
Here fej g is the canonical homogeneous basis for the super vector space k mjn . We
identify the morphism g 2 GLmjn .A/ D Hom.salg/ .kŒGLmjn ; A/ with the matrix with
entries gij D g.xij / 2 A, where xij ’s are the generators of kŒGLmjn .
This corresponds to the comodule map
P
k mjn W k ! kŒGLmjn ˝ k mjn ; ej 7! xij ˝ ei :
mjn
P
Vice versa, the comodule map ej 7! xij ˝ ei corresponds to the representation
that is, X X X
V .xj / ˝ aj D xj ˝ .aj / D xj ˝ bij ˝ ai :
j j i;j
Hence X
V .xi / D xj ˝ bij :
j
Theorem 11.7.9. Let G be an affine supergroup variety. Then there exists a closed
embedding of algebraic supergroups
G GLmjn
Vy
W Vy ! Vy ˝ kŒG; W
W W ! W ˝ kŒH :
The second map is induced by the first one, in fact
G .I / D I ˝ kŒG C kŒG ˝ I
since I is a Hopf ideal. It is simple to check from the definitions that such a comodule
map corresponds to the following action of G on the super vector space kŒG:
.g f /.x/ D f .xg/; g; x 2 G.A/; f 2 kŒG;
where customarily we denote x.f / by f .x/ and as usual xg D m B .x ˝ g/ B (see
Proposition 11.1.2).
By Proposition 11.7.8 we can find a finite-dimensional subspace V invariant under
the action of G and containing all the generators of kŒG and a subspace W D I \ V
containing the generators of I , hence fixed by H .
11.8 The algebraic stabilizer theorem 227
We now want to describe some important applications of this result, namely, the
representability of the supergroup functors corresponding to the classical Lie superal-
gebras (for a complete description of such superalgebras we refer to Appendix A).
1. A.mjn/ series. A.mjn/ consists of the matrices in the super vector space
M.m C 1jn C 1/ with supertrace zero. This is a Lie superalgebra with the bracket
induced by the one in M.m C 1jn C 1/.
Consider the morphism
where B is the super vector space of all the symmetric bilinear forms on k mj2n . Con-
sider the point in B:
0 1
0 Ip 0 0 0
BIp 0 0 0 0C
B C
ˆDB
B0 0 1 0 0C C if m D 2p C 1, n D 2q;
@0 0 0 0 Iq A
0 0 0 Iq 0
or
0 1
0 Ip 0 0
BIp 0 0 0C
ˆDB
@0
C if m D 2p, n D 2q:
0 0 Iq A
0 0 Iq 0
We define the stabilizer of the point ˆ to be the supergroup functor Ospmj2n . Again
this is an algebraic supergroup by Theorem 11.8.3.
In order to compute its Lie superalgebra, let us consider a generic matrix M D
I CX
W I CZ in M.mjn/ Š Lie.GLmjn /. We are assuming m D 2p even since for
Y
m odd the calculation is very much the same. The condition for M to belong to
11.8 The algebraic stabilizer theorem 229
Lie.Osp2pj2q / is M t ˆM D ˆ, which is
0 1
0 Ip 0 0
I C X t BIp 0C
W t B 0 0 C I C X Y
Y t I C Z t @0 0 0 Iq A W I C Z
0 0 Iq 0
0 1
0 Ip 0 0
BIp 0 0 0C
DB
@0
C:
0 0 Iq A
0 0 Iq 0
After a tedious calculation one finds that the generic matrix M in Lie.Osp2pj2q / has
the form 80 19
ˆ
ˆ a b x x1 > >
<B =
B ct at
t
y y1 C C ;
ˆ@ y1 x1 d e A> >
:̂ ;
y x f d
t t
where the matrices b and c are skewsymmetric and f and e are symmetric and the
entries of all the matrices a, b, c, d , e, f , x, x1 , y, y1 , u, v are in k. This shows that
Lie.Osp2pj2q / D osp.2pj2q/, in fact the matrices in osp.2pj2q/ have precisely the
form prescribed above.
So we have
3. P .n/ series. Define the algebraic supergroup Sp njn , as we did for Ospmjn by
taking antisymmetric bilinear forms instead of symmetric ones. Consider the action
11.9 References
In [44] Kac proved a classification theorem for simple Lie superalgebras that we have
summarized in Appendix A. The description of the supergroup functors, corresponding
to the classical super series of Lie superalgebras introduced by Kac, appeared in [22],
p. 70; however no representability statement of the supergroups was proved there. In
[14], [15] all the classical supergroup functors together with the Hopf superalgebras
representing them are described in detail.
A
Lie superalgebras
The purpose of this appendix is to give a brief introduction to the theory of Lie superal-
gebras and, in particular, to the theory of representations of classical Lie superalgebras.
This material is well known and is mostly found in the works of Kac, [44], [45]. This
appendix is self-contained and does not require knowledge of any material of the pre-
vious chapters. However, we assume basic knowledge of semisimple Lie algebras.
We start by describing the classification of simple finite-dimensional complex Lie
superalgebras. We do not provide any details on the proof of the main Theorem A.1.7,
but we give a full description of the classical families of simple Lie superalgebras,
including the root systems, Cartan matrices, and Dynkin diagrams.
We then discuss the finite-dimensional representations of classical Lie superalge-
bras. In the ordinary setting, the finite-dimensional modules for the special linear Lie
algebra sl2 play a key role. Ultimately, for a Lie algebra g, the conditions to impose
to have a finite-dimensional g-module, boil down to the conditions to impose in order
to have finite-dimensionality for the sl2 -modules, corresponding to simple roots. The
situation for classical Lie superalgebras is different. In the ordinary setting, the root
vectors X˛ and X˛ , corresponding to a pair of opposite roots ˛ and ˛, always gen-
erate an sl2 subalgebra inside the given simple Lie algebra. In contrast, given a simple
Lie superalgebra g and a root ˛ of g, the root vectors corresponding to roots propor-
tional to ˛ may generate a subalgebra of g isomorphic to sl2 , osp.1j2/, sl.1j1/, sq.2/,
or even a nilpotent subalgebra. Hence in order to fully understand the representation
theory of Lie superalgebras it is necessary to study in detail and classify all irreducible
representations of all these Lie superalgebras. We shall describe the irreducible rep-
resentations for sl2 , osp.1j2/, sl.1j1/, and then we will give the main theorem on the
classification of finite-dimensional representations of basic classical Lie superalgebras,
describing in detail the case of osp.2m C 1j2n/. This example is very illuminating
since we can see how the extra conditions for the finite-dimensionality of the modules
make their appearance once we take into account all Borel subalgebras at once. In the
ordinary setting this is not necessary since if we fix a Borel subalgebra, all other Borel
subalgebras are conjugate to the fixed one under the action of the Weyl group. For
Lie superalgebras, the Weyl group, still a classical object, appears to be “too small”
and not all Borel subalgebras are conjugate. For this reason, in order to proceed to the
classification of finite-dimensional representations, we need to use the odd reflections
to compensate for the lack of enough symmetries in the Weyl group. In Section A.5
we will see explicitly in the example of an orthosymplectic Lie superalgebra how this
happens.
232 A Lie superalgebras
Œx; Œy; z C .1/jxjjyjCjxjjzj Œy; Œz; x C .1/jyjjzjCjxjjzj Œz; Œx; y D 0
We now want to define the analogue of the special linear Lie algebra.
Definition A.1.1. We define the special linear Lie superalgebra, sl.mjn/ and the pro-
jective special linear Lie superalgebra psl.mjm/ as
Fixing as above, the matrix form for the orthosymplectic Lie superalgebra, which
we shall denote by osp.mjn/, becomes the following:
1. for m D 2p C 1, n D 2q,
80 19
ˆ
ˆ a b u x x1 > >
ˆB
ˆ >
<B c at v y y1 C
C=
>
osp.mjn/ D B B vt
t
ut 0 z C
z1 C I
ˆ
ˆ >
ˆ@ y1 x1t z1t d e A> >
>
:̂ t ;
y t x t z t f d
2. for m D 2p, n D 2q,
80 19
ˆ
ˆ a b x x1 > >
<B =
c at y y1 C
osp.mjn/ D B C ;
ˆ@ y1t x1t d e A> >
:̂ ;
y t x t f d t
where a, b, c, d , e, f , x, x1 , y, y1 , u, v are matrices of appropriate size and the
matrices b and c are skew-symmetric and e and f are symmetric; all matrices
have entries in k.
Notice that if n D 0, then osp.mj0/ is the orthogonal Lie algebra Bp or Dp de-
pending on the parity of m, while if m D 0, osp.0jn/ is the symplectic Lie algebra
Cq .
The decomposition of osp.mjn/ into even and odd parts is
osp.mjn/0 D Bp ˚ Cq if m D 2p C 1, n D 2q;
osp.mjn/0 D Dp ˚ Cq if m D 2p, n D 2q;
osp.mjn/1 D Vm ˝ Vn ;
where again Vm denotes the defining representation of so.m/ and Vn the defining
representation of Cq .
Next we have the Lie superalgebras in the strange series P .n/ and Q.n/.
Definition A.1.4. We define the strange series P .n/ as
² ³
A B
P .n/ D gl.n C 1jn C 1/;
C At
where A 2 sl.n C 1/, B is symmetric and C skew-symmetric.
The strange series Q.n/ is defined as follows. Set
² ³
A B
q.n/ D I
B A
sq.n/ are the matrices in q.n/ with tr.B/ D 0 and Q.n 1/ D psq.n/ D sq.n/=kI2n ,
i.e., ² ³
A B
Q.n 1/ D j B 2 sln =kI2n :
B A
A.1 Classical Lie superalgebras 235
Again one can check that these are well-defined Lie superalgebras with the bracket
induced from gl.njn/.
We have
V 0
P .n/0 D sl.n C 1/; P .n/1 D Sym2 .VnC1 / ˚ 2 .VnC1 /;
Q.n/0 D sl.n C 1/; Q.n/1 D ad.sl.n C 1//;
where VnC1 and ad.sl.n C 1// are respectively the defining representation and the
adjoint representations of sl.n C 1/.
The simple finite-dimensional Lie superalgebras over k fall into several classes. To
introduce these classes we need a few more definitions.
Definition A.1.5. Let g be a Lie superalgebra (always finite-dimensional). We say
that g is simple if g is not abelian and it admits no non-trivial ideals. g is classical if
it is simple and g1 is completely reducible as a g0 -module, where the action is given
by the bracket. g is basic if it is classical and it admits a consistent, non-degenerate,
invariant bilinear form, that is to say, there exists a consistent, non-degenerate, bilinear
form h ; i W g g ! g such that hX; ŒY; Zi D hŒX; Y ; Zi.
Observation A.1.6. If h ; i is a bilinear invariant form on a simple Lie superalgebra
g, then h ; i is either identically zero or non-degenerate. In fact if X 2 kerh ; i, i.e., if
X is such that hX; Y i D 0 for all Y 2 g, we have
0 D hX; ŒY; Zi D hŒX; Y ; Zi:
Hence the kernel of h ; i is an ideal, hence it is all of g or zero.
The simple Lie superalgebras divide into two main types: the classical type, when
the action of g0 on g1 is completely reducible, and the Cartan type. We make a list
of such Lie superalgebras, referring the reader to [44] for all proofs regarding the
classification.
1 Classical type. The classical type subdivides further into type 1 and type 2. Type 1
classical superalgebras are those for which g1 is not irreducible as g0 -module and
type 2 are those for which g1 is an irreducible g0 -module.
Notice that the upper indices refer to the Z-grading, while the lower indices to the
Z2 -grading.
For A.mjn/ and C.n/ we have that g1 and g1 are dual to each other; in fact,
where Vi always denotes the natural representation of the Lie algebra under consider-
ation and Vi0 its dual.
For g D P .n/, g1 and g1 are not dual to each other. We have
V2
g1 D Sym2 VnC1 ; g1 D 0
VnC1 :
1.2 Classical type 2. The type 2 superalgebras are those for which g1 is irreducible,
so that there is no compatible Z-grading. These Lie superalgebras are
2 Cartan type. Let Sym.V / denote the symmetric algebra of the super vector space
V . If V has dimension mjn and we fix a basis, we can identify Sym.V / with A D
kŒx1 ; : : : ; xm ; 1 ; : : : ; n , the polynomial algebra with m even indeterminates and n
odd ones. We define W .mjn/ D Der.A/ as the superalgebra of derivations of A. In
general W .mjn/ is infinite-dimensional, however when m D 0, it is finite-dimensional.
We write W .n/ for W .0jn/:
n X o
W .n/ D aI;j i1 : : : il @j :
aI;j 2k
A.2 Root systems 237
W .n/ D W .n/0 ˚ W .n/1 , where W .n/0 corresponds to those elements with l odd,
while W .n/1 to those with l even. Define ‚.n/ as the associative superalgebra over
A generated by 1 ; : : : ; n with relations i ^ j D j ^ i (i ¤ j ). This
is a superalgebra with grading induced by deg. i / D 1. Furthermore, W .n/ acts on
‚.n/ by derivations. We define S.n/ and Sz.n/ as the subalgebras of W .n/ annihilating
certain elements of ‚.n/ called volume forms. Namely,
In the next section we will discuss the structure of simple Lie superalgebras in more
detail with a particular attention to the families A, B, C , and D.
If g˛ ¤ f0g for ˛ 2 h0 nf0g, we say that ˛ is a root and g˛ is its root space. We say
that a root ˛ is even if g˛ \ g0 ¤ f0g, and odd if g˛ \ g1 ¤ f0g. Notice that with this
definition a root can be both even and odd; this can actually happen, as we shall see in
Example A.2.2. We denote by the set of all roots.
In the same way as in the ordinary case, we have
X
gDh˚ g˛ :
˛2
The set of roots h0 n f0g is the union of even and odd roots: D 0[ 1, where
where by a slight abuse of notation we denote a matrix with only two non-zero entries
by the corresponding letter above. Clearly, q.2/0 D spank fh1 ; h2 ; e; f g and q.2/1 D
spank fhN 1 ; hN 2 ; e;
N fNg. The space h D spank fh1 ; h2 ; hN 1 ; hN 2 g is a Cartan subalgebra of
q.2/. One can readily check that we have two root spaces of dimension 1j1:
g D h ˚ g˛ ˚ g˛ ; 0 D 1:
This somewhat awkward behaviour may suggest excluding Q.n/ from the treat-
ment, however we believe that the right philosophy is to ask questions for all classical
Lie superalgebras. Before we proceed and give some examples of root systems, we
want to make some observations on the Cartan–Killing form.
Observation A.2.3. Let g be a classical Lie superalgebra. In analogy with the Cartan–
Killing form in the ordinary setting, define the bilinear form
where x; y 2 g. As one can easily check, this form is symmetric and consistent.
However, quite differently from what happens in the classical setting, it is not always
non-degenerate, hence zero. In particular its restriction to a Cartan subalgebra of g
may be degenerate.
For example, on sl.mC1jnC1/, if h is the Cartan subalgebra consisting of diagonal
matrices, we have
.x; y/ D 2.m n/ str.xy/
for x D diag.a1 ; : : : ; amCn / and y D diag.a10 ; : : : ; amCn
0
/ in h. When m D n the
form is identically zero. Furthermore, it factors to a form on the corresponding Cartan
subalgebra of psl.mjm/, which is also identically equal to zero.
The fact that the Cartan–Killing form of a classical Lie superalgebra may be de-
generate prompts the definition of basic classical Lie superalgebras.
Definition A.2.4. A Lie superalgebra g is basic classical if g is simple, g0 is reductive,
and g admits a non-degenerate invariant symmetric consistent bilinear form.
For example, A.mjn/, C.n/, and B.mjn/ are basic simple Lie superalgebras, while
P .n/ is not. Notice that even though A.mjm/ has a degenerate Cartan–Killing form,
there exists a non-degenerate invariant symmetric bilinear form on A.mjm/.
The following table summarizes the classification of simple Lie superalgebras to-
gether with information about the existence of an invariant non-degenerate symmetric
bilinear form.
We now give a brief description of the root systems for the basic classical Lie
superalgebras in the families A, B, C , D. As in the ordinary setting every root system
admits a simple system … D f˛1 ; : : : ; ˛r g . The defining property of … is that
every root in is a linear combination of elements of … with integral all non-positive,
or all non-negative, coefficients, and … is minimal among such subsets of . The
elements ˛i of … are called simple roots. Unlike the ordinary setting, the elements of
… are not necessarily linearly independent, e.g., when g is of type A.mjm/. So, once
we fix a simple system, we can write D C t , where C is the set of the positive
roots, that is the roots ˛ D m1 ˛1 C C mr ˛r , mi 0, while C is the set of the
negative roots, that is the roots ˛ D n1 ˛1 C C nr ˛r , ni 0. From a representation
point of view, one may avoid the technicalities related to the superalgebras of type A
by replacing them by gl.mjm/. In the remainder of this appendix we will not discuss
240 A Lie superalgebras
the fine details distinguishing gl.mjm/ from the rest of the superalgebras of type A.
The interested reader will be able to fill in the details or consult [44], [45] for the whole
story.
For all of the families A, B, C , D, the Cartan subalgebras are even h D h0 , and
we can choose h D h0 to be the subalgebra of diagonal matrices. For each family
we shall give the root system and a choice of a simple system among the various
possibilities. Then for such a choice, in the next section we shall build the Cartan
matrix and the Dynkin diagram associated with the simple system. For a complete
treatment, comprehending all of the simple systems, we refer the reader to [44], 2.5.2,
and [33], part 3.
B.mjn/ D osp.2 m C 1j2n/. The Cartan matrix is the subalgebra of the diagonal
matrices:
h D fh D diag.a1 ; : : : ; am ; a1 ; : : : ; am ; 0; b1 ; : : : ; bn ; b1 ; : : : ; bn /g:
Define i ; ıj 2 h as follows: for h 2 h, let i .h/ D ai , i D 1; : : : ; m, and ıj .h/ D bj ,
j D 1; : : : ; n.
Root system, m ¤ 0:
C.n/ D osp.2j2n 2/. The Cartan matrix is again the subalgebra of the diagonal
matrices:
D.mjn/ D osp.2 mj2n/. The Cartan matrix is again the subalgebra of the diagonal
matrices:
We end this section with a summary of the properties of root systems for classical
Lie superalgebras. For the proofs we refer the reader to [44], Sections 2.5.3–2.5.4.
P
Proposition A.2.5. Let g be a classical Lie superalgebra and let g D h ˚ ˛2 g˛ be
the root decomposition of g with respect to a Cartan subalgebra h. Then the following
hold:
242 A Lie superalgebras
• B.mjn/, m ¤ 0.
0 1
2 1 0 ::: 0 ::: 0 ::: 0
B1 2 1 : : : 0 ::: 0 ::: 0C
B C
B :: :: :: :: :: :: :: :: :: C
::
B : : : : : : : : : C
:
B C
B 2 1 0 C
B C
ADB
B0 ::: 1 0 C1 : : : ::: : : :C
C
B0 ::: 0 1 2 1 ::: : : :C
B C
B :: :: :: :: :: :: :: :: :: :: C
B : : : : : : : : : : C
B C
@0 ::: 1 2 1A
0 ::: 0 2 2
• B.0jn/.
0 1
2 1 0 ::: 0 ::: 0 :::
B1 2 1 : : : 0 ::: 0 : : :C
B C
B :: C
A D B ::: ::
:
::
:
::
:
::
:
::
:
::
: : C
B C
@ 0 ::: 1 2 1A
0 ::: 0 2 2
• C.n/.
0 1
0 C1 0 : : : 0 ::: 0 :::
B1 2 1 : : : 0 ::: 0 : : :C
B C
B :: C
A D B ::: ::
:
::
:
::
:
::
:
::
:
::
: : C
B C
@ 0 ::: 1 2 2A
0 ::: 0 1 2
• D.mjn/.
0 1
2 1 0 ::: 0 ::: 0 ::: 0
B1 2 1 : : : 0 ::: 0 ::: 0C
B C
B :: :: :: :: :: :: :: :: :: :: C
B : : : : : : : : : :C
B C
B0 ::: 1 0 C1 : : : ::: ::: 0C
ADB
B0
C
B ::: 0 1 2 1 ::: ::: 0C C
B :: :: :: :: :: :: :: :: :: :: C
B : : : : : : : : : :C
B C
@0 ::: 1 2 0A
0 ::: 1 0 2
244 A Lie superalgebras
pp
:::
::: pNpNp
D.mjn/ NNN
We leave it to the reader, as a simple exercise, to verify that these are indeed the
Dynkin diagrams associated to the Cartan matrices listed above.
The representations of the special linear Lie algebra sl2 . The finite-dimensional
representations of sl2 are well known, we just quote the theorem, referring the reader to
[75], Ch. IV, for more details. Let sl2 D spank fh; e; f g, with Œe; f D h, Œh; e D 2e,
Œh; f D 2f .
hvi D .n 2i /vi ; 0 i n;
ev0 D 0; evi D .n i C 1/vi1 ; 1 i n;
f vn D 0; f vi D .i C 1/viC1 ; 0 i n 1:
Conversely, for each positive integer n, there exists exactly one equivalence class of
irreducible representations of sl2 of dimension n C 1, defined by the action described
above.
The vector v0 is called a highest weight vector and spans the one-dimensional
subspace of V annihilated by e.
The representations of the orthosymplectic Lie algebra osp.1j2/. Let us start with
an explicit description of this Lie superalgebra.
80 1 9
< 0 ˛ ˇ =
osp.1j2/ D @ ˇ A j B 2 sl2 gl.1j2/:
: B ;
˛
A basis is given by
0 1 0 1 0 1
0 0 0 0 0 0 0 0 0
e D @0 0 1A ; f D @0 0 0A ; h D @0 1 0 A ;
0 0 0 0 1 0 0 0 1
246 A Lie superalgebras
0 1 0 1
0 0 1 0 1 0
x D @1 0 0A ; y D @ 0 0 0A ;
0 0 0 1 0 0
where spank fe; f; hg Š sl2 and the other non-zero brackets are
vi D 1 i
iŠ
f v0 for 0 i n; wi D 1 i
iŠ
f yv0 for 0 i n 1:
One can now verify that the basis of osp.1j2/ acts on the set v0 , : : : , vn , w0 , : : : , wn1
as claimed and that this action provides V D span fv0 , : : : , vn , w0 , : : : , wn1 g with a
structure of an irreducible osp.1j2/-module.
Schematically we have the following picture:
and similarly
f f
V0 vn cFo vn1 fMo vn2 :::
FF y
FF y vvv
v MMM
nnnn
v MMyM y
nnn
FF vv MMM nn
F zvv f wnnn
V1 wn1 o wn2 :::
with brackets
Theorem A.4.3. Let V be the irreducible representation of gl.1j1/ with highest weight
. 1 ; 2 /. Then:
(1) If 1 C 2 ¤ 0, V is 1j1-dimensional and spanned by v, the highest weight
vector, and yv.
(2) If 1 C 2 D 0, V is one-dimensional.
Proof. Let v be the highest weight vector of V . We have
which proves the theorem. For 1C 2 ¤ 0 the statement is illustrated by the diagram
x
(
yv j v:
y
to verify that for generic highest weight . 1 ; 2 /, the highest weight space V.1 ;2 /
is an irreducible module of a (suitably chosen) Clifford algebra. In general it is of
dimension 1j1. Since g contains a subalgebra isomorphic to gl2 , a necessary condition
for finite-dimensionality of V is 1 2 to be a non-negative integer. Surprisingly, it
is not enough. The following theorem is proved in [63]:
Theorem A.4.4. Let D . 1 ; 2 / be a weight of q.2/ such that 1 2 is a positive
integer or 1 D 2 D 0. There exists a unique (up to a change of parity) finite-
dimensional irreducible q.2/-module V . / with highest weight . The weights of
V . / are . 1 ; 2 /; . 1 1; 2 C 1/; : : : ; . 2 ; 1 /.
Conversely, every irreducible finite-dimensional q.2/-module V is isomorphic to
V . / for some as above.
For the proof as well as further study of q.n/-modules we refer the reader to the
work of Penkov, [63].
(3) There exists K M. / largest proper submodule such that M. /=K is irre-
ducible (when finite-dimensional M. /=K is isomorphic to the V . / in TheoremA.5.1).
250 A Lie superalgebras
Proof (Sketch). (1) is a direct check. (2) comes from the fact that if 1 ˝ 1 2 F , then
F D M. /. For (3) take K as the sum of all proper submodules. For the complete
proof see [27], Ch. 7.
We have described an alternative way to construct the representations V . /. This is
the path taken by Kac in the construction of the irreducible representations of classical
Lie superalgebras and as we shall see the construction is very much the same as the
ordinary one.
Let g be a classical Lie superalgebra. Fix a Cartan subalgebra and a simple system
and let nC be as above. Let bC D nC ˚ h be a fixed Borel subalgebra and 2 h . As
before we have a one-dimensional representation of bC :
.h C n/ v D .h/v; v D 1 2 k; h 2 h; n 2 nC :
We define
M. / D IndgbC D U.g/ ˝U.bC / k:
Reasoning as in Proposition A.5.6, we have that M. / contains a unique maximal
submodule K such that M. /=K D V . / is irreducible. Since change of parity is not
an isomorphism of g-modules, we will always assume that the highest weight space is
even, and all statements of uniqueness below are up to a change of parity.
Proposition A.5.7. Let the notation be as above.
(1) The vector v is the unique vector in M. / (up to a constant) such that nC v D 0.
(2) V . 1/ Š V. 2/ if and only if 1 D 2.
˛1 D ı1 ı2 ; ˛2 D ı2 ı3 ; : : : ; ˛n1 D ın1 ın ;
˛n D ın 1 ; ˛nC1 D 1 2 ; : : : ; ˛nCm D m :
ƒD 1 ı1 C C n ın C 1 1 C C m m :
A.5 Representations of basic Lie superalgebras 253
ai D i iC1 ; 1 i n 1;
an D n C 1;
anCj D j j C1 ; 1 j m 1;
amCn D 2 n :
1 2 n 0; 1 2 m 0;
are non-negative and integral. This gives the condition (1) in Theorem A.5.9. Now
observe that
The fact n 2 Z0 gives condition (2). Now we want to see how to get condition (3).
In order to clarify the construction we work in B.2j1/, but the reader can immedi-
ately see how easily this can be generalized to B.mjn/.
In this case our weight ƒ and our fixed Borel subalgebra b (i.e., simple system) are
ƒD 1 ı1 C 1 1 C 2 2 ; fı1 1 ; 1 2 ; 2 g:
In the ordinary setting we have that the Weyl group acts transitively on the set of Borel
subalgebras, hence the weight w ƒ is automatically dominant and integral with respect
to the Borel subalgebra w b. Here however, we have fewer symmetries since the Weyl
group is not a super object, but remains ordinary. In particular not all Borel subalgebras
are conjugate with respect to the Weyl group. We can compensate the lack of enough
symmetries of the Weyl group, by using odd reflections. In this way we can reach
every Borel subalgebra starting from a fixed one. For this reason we are going to use
an action of gl.1j1/, whose representations we have studied in Section A.4.
Our goal is to use odd reflections to move the chosen Borel subalgebra to the one
with the associated simple system: f1 2 ; 2 ı1 ; ı1 g, while tracking down how the
weight ƒ transforms. The dominant integral condition on the new weight will give us
the additional conditions (3).
Let us start with an odd reflection along the root ı1 1 . This corresponds to a
gl.1j1/ representation of highest weight 1 C 1 . By Theorem A.4.3 we have only
two cases:
(1) 1 C 1 D 0, i.e., 1 D 1 D 0 (since they are both positive), corresponding to
a 1-dimensional representation,
254 A Lie superalgebras
• Case 1 D 1 D 0, ƒ D 1 ı1 C 1 1 C 2 2 .
i) If 2 D 1 D 0, then ƒ D 0, which is the trivial case.
ii) If 2 C 1 ¤ 0, we reach a contradiction since 1 2 0 and 1 D 1 D 0.
This proves only of course that the condition (3) is necessary. However, using
Theorem 10.5 in [67] we can conclude that it is also sufficient.
Any detailed discussion of the intricacies of the representation theory of Lie super-
algebras is beyond the scope of this Appendix. However, we would like to illustrate
some features on the case of gl.mjn/. In the ordinary setting, the Verma module M. /
for dominant integral has a unique finite-dimensional quotient. This is no longer true
for gl.mjn/. If is an atypical dominant integral weight, i.e., if V . / is an atypical
finite-dimensional module, M. / has more than one finite-dimensional quotient. In
fact, there is a universal finite-dimensional quotient K. / called the Kac module K. /,
first introduced by Kac. The relationship between V . / and K. / somewhat resembles
the relationship between V . / and M. / for infinite-dimensional modules V . / over a
simple Lie algebra. Hence the problem of finding the character of V . / for atypical
is somewhat close to the problem of finding the characters of the irreducible modules
in the category O of highest weight modules over Lie algebras.
Example A.7.2. Let us define the complex Heisenberg superalgebra n with even (cen-
tral) generator e and odd generators a1 ; : : : ; an , b1 ; : : : ; bn . The only non-zero brackets
are
Œai ; bi D e:
V
Let V D . 1 ; : : : ; n / denote the complex exterior algebra with generators 1 ; : : : ; n .
This is a super vector space, where the even part is generated by products of an even
number of i ’s while the oddVpart, by products of an odd number of the i ’s. We define
a representation ˛ of n on . 1 ; : : : ; n / in the following way:
ai u D @i u; bi u D ˛ i u; e u D ˛u;
where ˛ is a fixed non-zero complex number. One can easily check that ˛ is well
defined, in other words, the brackets are preserved. We now want to show that ˛ is
irreducible and C. .n// the endomorphisms of V supercommuting with the endomor-
phisms induced by this action are multiples of the identity. We will do this in the case
of n D 2, leaving the generalization to the reader as an exercise. For the irreducibility,
notice that the even and odd parts have dimension 2. Hence a proper invariant subspace
must have even or odd dimension equal to 1 and one can check immediately by direct
inspection that this is not possible.
V
Let 2 gl.V /, V D . 1 ; 2 /. By Schur’s lemma, in order to prove D I it
is enough to prove that preserves parity. By contradiction assume that does not.
Again by Schur’s lemma,
0 B
D ;
C 0
where B and C are two by two invertible matrices and we have chosen the (graded)
basis 1, 1 ^ 2 , 1 , 2 . By hypothesis we have
Similarly by commuting with @2 one can see that b21 D 0. This gives us a contra-
diction since by Schur’s lemma, must be invertible.
Let us now turn to another example, which realizes the second possibility of Schur’s
lemma. Consider n0 D n ˚ hci where the odd element c satisfies
Œn; c D 0; Œc; c D e:
V
Consider the vector space V D . 1 ; : : : ; n / ˝ k./, where k./ D T .k 0j1 /=. ˝
˛=2/, and T .k 0j1 / denotes the full tensor algebra over k of the super vector space
k 0j1 with (canonical) basis . As a super vector space k./ D fa C bg Š k 1j1 . Define
on V the action ˛0 :
0 0
˛ .h/.u ˝ v/ D ˛ .h/u ˝v for all h 2 n; ˛ .c/.u ˝ v/ D .1 ˝ /.u ˝ v/:
258 A Lie superalgebras
We leave it to the reader to check that this is a well-defined action and that it is irre-
ducible. If we take n D 2 we can directly check that the odd morphism
0 I2
D
I2 0
0
commutes with the action ˛ for ˛ D 2, thus realizing (2) in Schur’s lemma.
B
Categories
B.1 Categories
We want to make a brief summary of formal properties and definitions relative to
categories. For more details one can see for example [51].
Examples B.1.2. (1) Let .sets/ denote the category of sets. The objects are the sets,
and for any two sets A; B 2 Ob..sets//, the morphisms are the maps from A to B.
260 B Categories
(2) Let G denote the category of groups. Any object G 2 G is a group, and for any
two groups G; H 2 Ob.G /, the set HomG .G; H / is the set of group homomorphisms
from G to H .
Definition B.1.3. A category C 0 is a subcategory of category C if Ob.C 0 / Ob.C/
and HomC 0 .A; B/ HomC .A; B/ for all A; B 2 C 0 , so that the composition law “B”
on C 0 is induced by that on C .
Example B.1.4. The category A of abelian groups and group morphisms is a subcat-
egory of the category of groups G .
Definition B.1.5. Let C1 and C2 be two categories. Then a covariant resp. contravari-
ant functor F W C1 ! C2 consists of:
(1) a map F W Ob.C1 / ! Ob.C2 / and
(2) a map (denoted by the same F ) F W HomC1 .A; B/ ! HomC2 .F .A/; F .B//
resp. F W HomC1 .A; B/ ! HomC2 .F .B/; F .A// so that
(i) F .idA / D idF .A/ and
(ii) F .f B g/ D F .f / B F .g/ resp. F .f B g/ D F .g/ B F .f / for all A; B 2
Ob.C1 /.
By “functor” we always mean covariant functor. A contravariant functor F W C1 !
op op
C2 is the same as a covariant functor from C1 ! C2 , where C1 denotes the opposite
category, i.e., the category where all morphism arrows are reversed.
Definition B.1.6. Let F1 ; F2 be two functors from C1 to C2 . We say that there is a
natural transformation of functors ' W F1 ! F2 if for all A 2 C1 there is a set of
morphisms 'A W F1 .A/ ! F2 .A/ so that for any f 2 HomC1 .A; B/ (B 2 C1 ), the
following diagram commutes:
'A
F1 .A/ / F2 .A/
F1 .f / F2 .f / (B.1)
'B
F1 .B/ / F2 .B/.
Definition B.1.7. We say that two categories C1 and C2 are equivalent if there exist
two functors F W C1 ! C2 and G W C2 ! C1 such that F G Š idC2 , GF Š idC1 ,
where idC denotes the identity functor of a given category, defined in the obvious way,
while F Š F 0 means that the two functors are isomorphic.
If F is a functor from the category C1 to the category C2 for any two objects A,
B 2 C1 , by its very definition, F induces a function (that we denoted previously
with F )
FA;B W HomC1 .A; B/ ! HomC2 .F .A/; F .B//:
Definition B.1.8. Let F be a functor. We say that F is faithful if FA;B is injective, we
say F is full if FA;B is surjective and we say that F is fully faithful if FA;B is bijective.
Next we want to formally define what it means for a functor to be representable.
Let us first define the representation functors.
Definition B.1.9. Let C be a category, A a fixed object in C . We define the two
representation functors HomA , HomA as
where .sets/ denotes the category of sets. On the arrow f 2 Hom.B; C / we have
Definition B.1.10. Let F be a functor from the category C to the category of sets. We
say that F is representable by X 2 C if for all A 2 C, F Š HomA or F Š HomA .
We end our small exposition of categories by constructing the fibered product which
is very important in our supergeometric constructions especially in Chapter 10.
Definition B.1.11. Given functors A, B, C from a category C to the category of .sets/,
and given natural transformations f W A ! C , g W B ! C , the fibered product AC B
is the universal object making the following diagram commute:
T
x
.x;y/
# %/
y
A C B p A
q f
g
B / C.
ensures the gluing of any family of local sections which agree on the intersection of
any two parts of an open affine covering.
We now want to give the definition of local functor in a more general setting, so
that it applies to the functor of points of supermanifolds and superschemes.
Definition B.2.2. Let F W .sspaces/op ! .sets/ be a functor. We say that F is local
or we also say it is a sheaf if it has the following property. For any superspace T
and any open covering fTi g of T let i W Ti ,! T , ij W Ti [ Tj ,! Ti be the natural
morphisms. If we have a family ˛i 2 F .Ti / such that F .ij /.˛i / D F .j i /.˛j /, then
there exists a unique ˛ 2 F .T / with F .i /.˛/ D ˛i .
One can readily check that this implies that, when F is restricted to the category of
the open sets of a fixed superspace, F is a sheaf in the ordinary sense. As we already
noticed, if we restrict F to the category of affine superschemes, this definition agrees
with the previous one.
We leave the following proposition as an exercise to the reader. It is very similar
to the proof in Chapter 10, Section 10.3.
Proposition B.2.3. If X is a superspace, its functor of points is local.
We now turn to the following problem. If we have a presheaf F on a topological
space in the ordinary sense, we can always build its sheafification, which is a sheaf
Fz together with a sheaf morphism ˛ W F ! Fz . This is the (unique) sheaf, which is
locally isomorphic to the given presheaf and has the following universal property: any
presheaf morphism W F ! G , with G a sheaf, factors via ˛ (for more details on this
construction, see Chapter 2, Section 2.2). We now want to give the same construction
in our more general setting.
The existence of sheafification of a functor from the category of algebras to the
category of sets is granted in the ordinary case by [23], Ch. I, §1, no. 4, which is also
nicely summarized in [23], Ch. III, §1, no. 3. The proof is quite formal and one can
carry it to the supergeometric setting. We however prefer to introduce Grothendieck
topologies and the concept of site and to construct the sheafification of a functor through
them. In fact, as we shall see, very remarkably Grothendieck’s treatment is far more
general and it comprehends supergeometry. For more details we refer the reader to the
classical account by Grothendieck [38], [41] and the more modern treatment by Vistoli
[77].
Definition B.2.4. We call a category C a site if it has a Grothendieck topology, i.e., to
every object U 2 C we associate a collection of so-called coverings of U , that is, sets
of arrows fUi ! U g such that the following holds:
(1) If V ! U is an isomorphism, then the set fV ! U g is a covering.
(2) If fUi ! U g is a covering and V ! U is any arrow, then the fibered products
fUi U V g exist and the collection of projections fUi U V ! V g is a covering.
(3) If fUi ! V g is a covering and for each index i we have a covering fVij ! Ui g,
then the collection fVij ! Ui ! U g is a covering of U .
264 B Categories
The category of superschemes and its subcategory of affine superschemes are sites,
by taking for each object the collection of its (affine) coverings, as in the ordinary
setting. Similarly also the category of commutative superalgebras is also a site (for
the existence of fibered products in such categories and for more details on coverings
see Sections 10.3, 9.4). Such Grothendieck topologies, with an abuse of terminology,
are commonly referred to as Zariski topologies (one should prove that all of these
topologies are essentially equivalent). We shall not dwell upon the technicalities of
Grothendieck topologies, referring the reader to [77] where all the many subtleties are
discussed in the fullest detail.
Definition B.2.5. Let C be a site. A functor F W C op ! .sets/ is called a sheaf if for all
objects U 2 C, coverings fUi ! U g and families ai 2 F .Ui /, we have the following.
Let pij1 W Ui U Uj ! Ui , pij2 W Ui U Uj ! Uj denote the natural projections and
assume that F .pij1 /.ai / D F .pij2 /.aj / 2 F .Ui U Uj / for all i , j . Then there exists
a unique a 2 F .U / whose pullback to F .Ui / is ai for every i .
Again one can check that the functor of points of superschemes and supermanifolds
are sheaves in this more general setting for the corresponding sites.
We are ready for the sheafification of a functor in this very general setting.
Definition B.2.6. Let C be a site and let F W C op ! .sets/ be a functor. A sheafification
of F is a sheaf Fz W C op ! .sets/ with a natural transformation ˛ W F ! Fz such that:
(1) For any U 2 C and ; 2 F .U / such that ˛U . / D ˛U ./ in Fz .U /, there is a
covering fi W Ui ! U g with F .i /. / D F .i /./ in F .Ui /.
(2) For any U 2 C and any 2 Fz .U /, there is a covering fi W Ui ! U g and
elements i 2 F .Ui / such that ˛Ui . i / D Fz .i /. / in Fz .Ui /.
The next theorem states the fundamental properties of the sheafification.
Theorem B.2.7 ([77], p. 42). Let C be a site, F W C op ! .sets/ a functor.
(1) If Fz is a sheafification of F with ˛ W F ! Fz , then any morphism W F ! G,
with G a sheaf, factors uniquely through Fz .
(2) F admits a sheafification Fz , unique up to a canonical isomorphism.
We shall use this construction for the following supergeometric categories: C D
.smflds/, C D .sschemes/, C D .aschemes/, or equivalently C op D .salg/.
Observation B.2.8. Let F W C op ! .sets/ be a functor, Fz its sheafification, where C
is one of the supergeometric categories specified above. Let A be an object of C and
FA the restriction of the functor F to the category of the open subobjects of A (see
above). Then FzA is the sheafification of FA in the usual sense, that is, the sheafification
of F as a sheaf defined on the topological space underlying A. In particular, since a
sheaf and its sheafification are locally isomorphic, we have FA;p Š FzA;p , i.e., they
have isomorphic stalks (via the natural map ˛ W F ! Fz ) at any point p and for all
objects A. To simplify the notation we shall drop the suffix A and write just Fp instead
of FA;p .
B.2 Sheafification of a functor 265
Along the same lines, the reader can prove the following proposition:
We want to show that a projective A-module has the property of being locally free,
that is its localization Mp into primes p of A0 is free as an Ap -module. This result
allows us to define the rank of a projective module as it happens in the ordinary case.
We start with a generalization of Nakayama’s lemma.
Lemma B.3.3 (Super Nakayama lemma). Let A be a local supercommutative ring with
maximal homogeneous ideal m. Let E be a finitely generated module for the ungraded
ring A.
(1) If mE D E, then E D 0; more generally, if H is a submodule of E such that
E D mE C H , then E D H .
(2) Let .vi /1ip be a basis for the k-vector space E=mE where k D A=m. Let
ei 2 E be above vi . Then the ei generate E. If E is a supermodule for the super ring
A, and vi are homogeneous elements of the super vector space E=mE, we can choose
the ei to be homogeneous too (and hence of the same parity as the vi ).
(3) Suppose that E is projective, i.e., there is an A-module F such that E˚F D AN ,
where AN is the free module for the ungraded ring A of rank N . Then E (and hence
F ) is free, and the ei above form a basis for E.
Proof. The proofs are easy extensions of the ones in the commutative case. We begin
the proof of (1) with the following observation: if B is a commutative local ring with n
a maximal ideal, then a square matrix R over B is invertible if and only if it is invertible
modulo n over the field B=n. In fact if this is so, det.R/ … n and so is a unit of B.
B.3 Super Nakayama’s lemma and projective modules 267
It is now enough to prove that L has a left inverse P . Then multiplying the above from
the left by P , we get ui D 0 for all i and so E D 0. It is even true that L is invertible.
To prove this, let us consider B D A=JA where JA is the ideal generated by A1 . Since
JA m we have
A ! B D A=JA ! k D A=m:
Let LB (resp. Lk ) be the reduction of L modulo JA (respectively modulo m). Then B
is local and its maximal ideal is m=JA , where Lk is the reduction of LB mod m=JA .
But B is commutative and Lk D I , and so LB is invertible. But then L is invertible.
If more generally we have E D H C mE, then E=H D m.E=H / and so E=H D 0,
which is to say that E D H .
To prove (2), let H be the submodule of E generated by the ei . Then E D mE CH
and so E D H .
We now prove (3). Clearly F is also finitely generated. We have k N D AN =mN D
E=mE ˚ F=mF . Let .wj / be a basis of F=mF and let fj be elements of F above wj .
Then by (2), the ei , fj form a basis of AN , while the ei (resp. fj ) generate E (resp. F ).
Now there are exactly N of the ei ; fj , and so if X denotes the .N N /-matrix with
columns e1 ; : : : ; f1 ; : : : , then for some .N N /-matrix Y over A we have X Y D I .
Hence XB YB D I where the suffix “B” denotes reduction modulo B. However, B is
commutative and so YB XB D I . Thus X has a left inverse over A, which must be Y
so that YX D I . If there is a linear relation among the ei and the fj , and if x is the
column vector whose components are the coefficients of this relation, then Xx D 0.
But this implies that x D YXx D 0. In particular E is a free module with basis .ei /.
Thus M is projective.
(2) That Mp is free for all primes p is equivalent to M Œfi1 being free for
.f1 ; : : : ; fr / D A0 is a standard fact of commutative algebra and can be found in
[28], p. 623, for example.
Definition B.3.5. Let M be a finitely generated projective A-module. We say that M
has rank rjs if Mp Š Arjs for all primes p in A. In the light of the previous proposition
one can show that the rank is locally constant, that is, if Mp Š Arjs , then there exists
an open neighbourhood U of p 2 Spec A for which Mp0 D Arjs , p 0 2 U .
Remark B.3.6. As in the ordinary setting we have a correspondence between projec-
tive A-modules and locally free sheaves on Spec A0 . In this correspondence, given a
projective A-module M , we view M as an A0 -module and build a sheaf of modules
OM on A0 . The global sections of this sheaf are isomorphic to M itself, and locally,
i.e., on the open sets Ufi D fp 2 Spec A0 j.fi / 6 pg, fi 2 A0 ,
OM .Ufi / D M Œfi1 :
More details on this construction in the ordinary setting can be found, for example, in
[43], Ch. II. As for its generalization to the super context, it is straightforward.
B.4 References
For a complete introduction to categories we refer the reader to [55], however a good
review of the main properties can also be found in [51]. As for the sheafification of a
functor we refer to [77] and of course to [38] by Grothendieck. Finally for the super
Nakayama lemma see [76] and for the ordinary setting [57].
C
Fréchet superspaces
Definition C.1.1. A topological vector space V is both a topological space and a vector
space such that the vector space operations are continuous. A topological vector space
is locally convex if its topology admits a basis consisting of convex sets (a set A is
convex if .1 t /x C ty 2 A for all x; y 2 A and t 2 Œ0; 1).
We say that a locally convex topological vector space is a Fréchet space if its
topology is induced by a translation-invariant metric d and the space is complete with
respect to d , that is, all the Cauchy sequences are convergent.
Definition C.1.5. We say that a sequence .xn / in V converges to x in the Fréchet space
topology defined by a family of seminorms if and only if it converges to x with respect
to each of the given seminorms. In other words, xn ! x if and only if pi .xn x/ ! 0
for each i .
Two families of seminorms defined on the locally convex vector space V are said
to be equivalent if they induce the same topology on V .
To construct a Fréchet space, one typically starts with a locally convex topological
vector space V and defines a countable family of seminorms fpk g on V inducing its
topology and such that:
(I) if x 2 V and pk .x/ D 0 for all k 0, then x D 0 (separation property);
(II) if .xn / is a sequence in V which is Cauchy with respect to each seminorm, then
there exists x 2 V such that .xn / converges to x with respect to each seminorm
(completeness property).
The topology induced by these seminorms (as explained above) turns V into a
Fréchet space; property (I) ensures that it is Hausdorff, while the property (II) guaran-
tees that it is complete. A translation-invariant complete metric inducing the topology
on V can then be defined as above.
The most important example of Fréchet space, at least for the application we have
in mind, is the vector space C 1 .U /, the space of smooth functions on the open set
U Rn or more generally the vector space C 1 .M /, where M is a differentiable
C.1 Fréchet spaces 271
manifold. For each open set U Rn (or U M ), for each K U compact and for
each multi-index I , we define
ˇ @jI j ˇ
ˇ ˇ
kf kK;I ´ sup ˇ I
.f / .x/ˇ; f 2 C 1 .U /:
x2K @x
with fixed g 2 C 1 .U /, K U compact, and multi-index I are open sets and together
with their finite intersections form a basis for the topology.
All these constructions and results can be generalized to smooth manifolds, as we
now briefly outline.
Let M be a smooth manifold and let U be an open subset of M . If K is a compact
subset of U and D is a differential operator over U , then
is a seminorm. The family of all the seminorms pK;D with K and D varying among
all compact subsets and differential operators respectively is a separating family of
1
seminorms endowing CM .U / with the structure of a complete locally convex vector
space. Moreover there exists an equivalent countable family of seminorms, hence
1
CM .U / is a Fréchet space. Let indeed fVj g be a countable open cover of U by open
S and let, for each j , fKj;i g be a countable family of compact subsets
coordinate subsets,
of Vj such that i Kj;i D Vj . We have the countable family of seminorms
ˇ @jI j ˇ
ˇ ˇ
pK;I ´ sup ˇ .f / .x/ˇ; K 2 fKj;i g;
x2K @x I
1
inducing the topology. Notice that CM .U / is also an algebra: the product of two
smooth functions being a smooth function. It is thus natural in this context to introduce
the notion of Fréchet algebra.
Definition C.1.6. A Fréchet space V is said to be a Fréchet algebra if its topology
can be defined by a countable family of submultiplicative seminorms, i.e., a countable
family fqi gi2N of seminorms satisfying
If we come back to our prototype of Fréchet space C 1 .Rn /, one can check that it
is also a Fréchet algebra, the countable family of submultiplicative seminorms given
272 C Fréchet superspaces
by
ˇ @jI j ˇ
ˇ ˇ
qKi;m ;j .f / ´ 2 j
sup ˇ I
.f / .x/ˇ:
x2Ki;m ;jI jj @x
Definition C.2.1. (1) We say that a super vector space V D V0 ˚ V1 is a super Fréchet
space if there exist two families of homogeneous seminorms fp0i g and fp1i g defined
on V0 and V1 , respectively, with respect to which V0 and V1 are Fréchet spaces. We
will usually denote by pi a generic seminorm from one of the two families.
(2) Let A D A0 ˚ A1 be a super Fréchet space with respect to the family of
seminorms fpi gi2I . Suppose also that A is a superalgebra with multiplication m. We
say that A is a super Fréchet algebra if the topology is defined by an equivalent family
of submultiplicative seminorms fqi g: qi .ab/ qi .a/qi .b/.
(3) Suppose now that F is a sheaf such that F .U / is a super Fréchet algebra for
each U . We say that F is a super Fréchet sheaf if for each open set U and for any
open cover fUi g of U the topology of F .U / is the initial topology with respect to the
restriction maps F .U / ! F .Ui /.
We now turn to the example most interesting to us, namely the superalgebra of
sections on a supermanifold.
Let M D .jM j; OM / be a supermanifold. Fix now an open subset U jM j. For
each compact subset K U and each differential operator D over U , define
B
pK;D .f / ´ sup j.D.f //.x/j; f 2 OM .U /:
x2K
C.2 Fréchet superspaces 273
As before, one can readily check that each pK;D defines a seminorm. The family
of seminorms obtained by considering all the differential operators and the compact
subsets K covering U endows OM .U / with a Hausdorff locally convex topology (as
before), where the open sets that form a basis for the topology are
ff 2 OM .U / j pK;D .f g/ < "g;
with fixed g 2 OM .U /, K U compact, D 2 Diff.U /, the differential operators on
U , and " > 0, together with their finite intersections.
In complete analogy with the ordinary setting one can prove the following propo-
sition.
Proposition C.2.2. OM is a super Fréchet sheaf.
Before proving it we need some preliminary results that we collect in the next
lemma and proposition.
Lemma C.2.3. Let U be a chart with coordinates t i , j and let ffn g be a sequence in
OM .U /: X
1
fn D fnI I ; fnI 2 CM .U /:
I
(1) ffn g is a Cauchy sequence in OM .U / if and only if the sequence ffnI g is a
Cauchy sequence in C 1 .U / for each I .
(2) OM .U / is complete, moreover it is a super Fréchet algebra.
P j jCjI j
Proof. (1) Let K be a compact subset of U and let D D a ;I @@t I be a super
F
differential operator. We have
X ˇ X X ˇ
ˇ @ ˇ
pK;D . fJ J / D sup ˇ a ;I fJ J
.x/ˇ
x2K @t @ I
J ;I J
X
max.a ;I /pK; @ .fI /;
x2K @t
;I
It is not difficult to check that the family fq˛;Kn g˛;n is equivalent to the family
fpKn ; @ gn; ;I and that they are submultiplicative.
@t @ I
Proposition C.2.4. Let fUi g be an open cover of an open set U jM j, let also fsn g
be a sequence in OM .U /.
(1) fsn g converges to s in OM .U / if and only if fsn jUi g converges to sjUi in OM .Ui /
for each Ui .
(2) fsn g is a Cauchy sequence in OM .U / if and only if fsn jUi g is a Cauchy sequence
in OM .Ui / for each Ui .
(3) OM .U / is complete for each open subset U .
(4) OM .U / is a super Fréchet algebra for each open subset U .
Proof. (1) Clearly if fsn g ! 0, then sn jUi ! 0 for each i . Suppose vice versa that for
each integer i, sn jUi ! 0 and let K be a compact subset of U . There exists a finite
open cover Vj of K such that each Vxj is compact and for each j there exists i such that
Vj Ui (see [26]). Then .1/ follows from the inequality
X
pK;D .sn / pK\Vxj ;Dj x .sn /:
K\Vj
j
(2) If fsn g is a Cauchy sequence in OM .U / and fUi g are charts covering U , then
there exists a family fsUi g such that sn jUi converges to sUi for each i . An easy check
shows that the various sUi glue together and define a section s in OM .U /. Clearly sn
converges to s.
(3) The fact that OM .U / is complete follows by Lemma C.2.3, by considering a
countable open cover fUi g of U by open supercharts and points (1) and (2).
(4) The fact that OM .U / is a super Fréchet algebra follows by considering the
family of seminorms
Proof. (1) is clear because of the way the seminorms, hence the topology, are defined.
(2) Due to Proposition C.2.4 we can suppose that M and N are superdomains
with coordinates t i , j and x r , s , respectively. We have to prove that if ffn g is a
sequence in O.N / converging to zero, then the sequence f .fn /g converges to zero
j˛jCjI j
in O.M /, i.e., for all compact subsets K and all differential operators D D @@t ˛ @ I ,
pD;K . .fn // tends to zero. Let a 2 Diff 0 .M /, the differential operators of degree
zero, namely a 2 O.M /. Then
pa;K . G
.fn // D sup j.a
.f //.x/j n (C.1)
x2K
. sup ja.x/j/. sup jfn .j j.x//j/; (C.2)
x2K x2K
The numbers at the end of each item refer to the pages on which the respective work is cited.
[1] R. Abraham, J. E. Marsden, and T. Ratiu, Manifolds, tensor analysis and applications.
Appl. Math. Sci. 75, Springer-Verlag, New York 1993. 102
[2] M. F. Atiyah and I. G. MacDonald, Introduction to commutative algebra. Addison-Wesley
Publishing Co., Reading, Mass., 1969. 44, 77, 197, 265
[3] Yu. A. Bahturin, A. A. Mikhalev, V. M. Petrogradsky, and M. V. Zaicev, Infinite dimensional
Lie superalgebras. De Gruyter Exp. Math. 7, Walter de Gruyter, Berlin 1992. 9
[4] L. Balduzzi, C. Carmeli, and G. Cassinelli, Super G-spaces. In Symmetry in mathematics
and physics, Contemp. Math. 490, Amer. Math. Soc., Providence, RI, 2009, 159–176. 153,
172
[5] L. Balduzzi, C. Carmeli, and R. Fioresi, Quotients in supergeometry. In Symmetry in
mathematics and physics, Contemp. Math. 490, Amer. Math. Soc., Providence, RI, 2009,
177–187. 172
[6] C. Bartocci, U. Bruzzo, and D. Hernandez-Ruiperez, The geometry of supermanifolds.
Math. Appl. 71, Kluwer Academic Publishers, Dordrecht 1991. 89
[7] M. Batchelor, Graded manifolds and vector bundles: A functorial correspondence. J. Math.
Phys. 26 (1985), 1578–1582. 63
[8] M. Batchelor, Two approaches to supermanifolds. Trans. Amer. Math. Soc. 258 (1980),
257–270.
[9] M. Batchelor, The structure of supermanifolds. Trans. Amer. Math. Soc. 253 (1979), 329–
338.
[10] F. A. Berezin, Introduction to superanalysis. Math. Phys. Appl. Math. 9, D. Reidel Pub-
lishing C., Dordrecht 1987. viii, 15, 27, 53, 89
[11] F. A. Berezin and D. A. Leites, Supermanifolds. Dokl. Akad. Nauk SSSR 224 (1975),
505–508. 89, 140
[12] G. Bergman, The diamond lemma for ring theory. Adv. Math. 29 (1978), 178–218. 24
[13] A. Borel, Linear algebraic groups. Grad. Texts in Math. 126, Springer-Verlag, New York
1991.
[14] H. Boseck, Affine Lie supergroups. Math. Nachr. 143 (1989), 303–327. 230
[15] H. Boseck, Classical Lie supergroups. Math. Nachr. 148 (1990), 81–115. 230
[16] N. Bourbaki, General topology. Chapters 1-4, Elem. Math. (Berlin), Springer-Verlag,
Berlin 1989. 273
[17] N. Bourbaki, Éléments de mathématique. Fasc. XXXIII. Variétés différentielles et analy-
tiques. Fascicule de résultats (Paragraphes 1 à 7), Actualités Scientifiques et Industrielles
1333, Hermann, Paris 1967. 82
278 Bibliography
[18] N. Bourbaki, Lie groups and Lie algebras. Chapters 1–3, Elem. Math. (Berlin), Springer-
Verlag, Berlin 1998.
[19] U. Bruzzo and R. Cianci, Differential equations, Frobenius theorem and local flows on
supermanifolds. J. Phys. A Math. Gen. 18 (1985), 417–423. 111
[20] V. Chari and A. Pressley, A guide to quantum groups. Cambridge University Press, Cam-
bridge 1994. 26, 113
[21] S.-J. Cheng and W. Wang, Dualities for Lie superalgebras. Preprint 2009.
arXiv:1001.0074v2 [math.RT] 255
[22] P. Deligne and J. W. Morgan, Notes on supersymmetry (following Joseph Bernstein).
In Quantum fields and strings; a course for mathematicians, Vol. 1, Amer. Math. Soc.,
Providence, RI; Institute for Advanced Study (IAS), Princeton, NJ, 1999, 41–97. viii, 24,
26, 27, 53, 89, 111, 140, 153, 170, 229, 230
[23] M. Demazure and P. Gabriel, Groupes algébriques. Tome I: Géométrie algébrique, général-
ités, groupes commutatifs, Masson & Cie, Éditeur, Paris; North-Holland Publishing Co.,
Amsterdam, 1970. 28, 41, 42, 43, 44, 183, 186, 188, 197, 201, 202, 210, 216, 262, 263
[24] M. Demazure and A. Grothendieck, Schémas en groupes. I, II, III. Séminaire de géométrie
algébrique du Bois Marie 1962/64 (SGA 3). Dirigé par M. Demazure et A. Grothendieck.
Lecture Notes in Math. 151, 152, 153, Springer-Verlag, Berlin 1970.
[25] B. DeWitt, Supermanifolds. Cambridge Monogr. Math. Phys., Cambridge University
Press, Cambridge 1984. viii
[26] J. Dieudonné, Éléments d’analyse. Tome III: Chapitres XVI et XVII, Cahiers Scientifiques
XXXIII, Gauthier-Villars Éditeur, Paris 1970. 82, 119, 273, 274
[27] J. Dixmier, Enveloping algebras. Grad. Stud. Math. 11, Amer. Math. Soc., Providence,
RI, 1996. 249, 250
[28] D. Eisenbud, Commutative algebra with a view toward algebraic geometry. Grad. Texts
in Math. 150, Springer-Verlag, New York 1995. 268
[29] D. Eisenbud and J. Harris, The geometry of schemes. Grad. Texts in Math. 197, Springer
Verlag, New York 2000. 28, 33, 35, 36, 37, 39, 41, 43, 44, 65, 180, 183, 262
[30] R. Fioresi, On algebraic super-groups and quantum deformations. J. Algebra Appl. 2
(2003), 403–423. 206
[31] R. Fioresi and F. Gavarini, Chevalley supergroups. Mem. Amer. Math. Soc., to appear; see
also arXiv:0808.0785v5 [math.RA]. 140
[32] R. Fioresi, M. A. Lledó, and V. S. Varadarajan, The Minkowski and conformal superspaces.
J. Math. Phys. 48 (2007), no. 11, 113505-1–27. 170, 172
[33] L. Frappat, A. Sciarrino, and P. Sorba, Dictionary on Lie algebras and superalgebras.
Academic Press, Inc., San Diego, CA, 2000. 240
[34] D. Z. Freedman, P. van Nieuwenhuizen, and S. Ferrara, Progress toward a theory of su-
pergravity. Phys. Rev. D 13 (1976), 3214–3218. 140
[35] R. Giachetti and R. Ricci, R-actions, derivations, and Frobenius theorem on graded man-
ifolds. Adv. Math. 62 (1986), 84–100. 111
Bibliography 279
[36] Yu. A. Gol’fand and E. P. Likhtman, Extension of the algebra of Poincaré group generators
and violation of P invariance. JETP Lett. 13 (1971), 323–326. 140
[37] P. Griffiths and J. Harris, Principles of algebraic geometry. Wiley Classics Lib., John Wiley
& Sons, Inc., New York 1994. 35, 44
[38] A. Grothendieck, Fondaments de la géometrie algébrique. Secretariat Mathematique, Paris
1962. 44, 263, 268
[39] A. Grothendieck, Produits tensoriels topologiques et espaces nucléaires. In Séminaire
Bourbaki, Vol. 2, Exp. No. 69, Soc. Math. France, Paris 1995, 193–200. 75
[40] A. Grothendieck, Résumé de la théorie métrique des produits tensoriels topologiques.
Reprint of Bol. Soc. Mat. São Paulo 8 (1953), 1–79; Resenhas 2 (1996), 401–480. 75
[41] A. Grothendieck, Technique de descente et théorèmes d’existence en géometrie algébrique.
I. Généralités. Descente par morphismes fidèlement plats. In Séminaire Bourbaki, Vol. 5,
Exp. No. 190, Soc. Math. France, Paris 1995, 299–327. 263
[42] C. Gruson and V. Serganova, Cohomology of generalized supergrassmannians and char-
acter formulae for basic classical Lie superalgebras. Proc. London Math. Soc. (3) 101
(2010), 852–892. 255
[43] R. Hartshorne, Algebraic geometry. Grad. Texts In Math. 52, Springer-Verlag, New York
1977. 28, 30, 33, 35, 36, 37, 43, 44, 176, 193, 265, 268
[44] V. G. Kac, Lie superalgebras. Adv. Math. 26 (1977), 8–26. 230, 231, 235, 236, 240, 241,
244, 250, 254
[45] V. G. Kac, Representations of classical Lie superalgebras. In Differential geometrical
methods in mathematical physics, II, Lecture Notes in Math. 676, Springer-Verlag, Berlin
1978, 597–626. 231, 240, 250, 255
[46] V. G. Kac, A sketch of Lie superalgebras theory. Comm. Math. Phys. 53 (1977), 31–64.
[47] A. W. Knapp, Advanced real analysis. Cornerstones, Birkhäuser, Boston, Mass., 2005.
102
[48] I. Kolář, P. W. Michor, and J. Slovák, Natural operations in differential geometry. Springer-
Verlag, Berlin 1993. 156
[49] B. Kostant, Graded manifolds, graded Lie theory, and prequantization. In Differential
geometrical methods in mathematical physics, Lecture Notes in Math. 570, Springer-
Verlag, Berlin 1977, 177–306. viii, 27, 53, 89, 112, 124, 140
[50] J.-L. Koszul, Graded manifolds and graded Lie algebras. In Proceedings of the interna-
tional meeting on geometry and physics, Pitagora Editrice, Bologna 1983,71–84. 111,
112, 137, 140
[51] S. Lang, Algebra. Addison Wesley, Reading, Mass., 1984. 4, 44, 259, 265, 268
[52] S. Lang, Real and functional analysis. Grad. Texts in Math. 142, Springer-Verlag, New
York 1993. 90
[53] D. A. Leites, Introduction to the theory of supermanifolds. Russian Math. Surveys 35
(1980), 1–64. viii, 89, 90, 102
[54] D. A. Leites, Quantization and supermanifolds. Appendix 3 in F. Berezin, M. Shubin The
Schrödinger equation, Math. Appl. (Soviet Ser.) 66, Kluwer Academic Publishers Group,
Dordrecht 1991. 97
280 Bibliography
[55] S. Maclane, Categories for the working mathematician, Grad. Texts in Math. 5, Springer-
Verlag, New York 1998. 268
[56] Y. I. Manin, Gauge field theory and complex geometry. Translated by N. Koblitz and J. R.
King, Grundlehren Math. Wiss. 289, Springer-Verlag, Berlin 1997. 27, 53, 89, 90, 102,
111, 172, 194, 201
[57] H. Matsumura, Commutative ring theory. Cambridge Stud. Adv. Math. 8, Cambridge
University Press, Cambridge 1989. 268
[58] P. W. Michor, Topics in differential geometry. Grad. Stud. Math. 93, Amer. Math. Soc.,
Providence, RI, 2008. 76, 77, 84, 89
[59] S. Montgomery, Hopf algebras and their actions on rings. CBMS Regional Conf. Ser. in
Math. 82 Amer. Math. Soc., Providence, RI, 1993. 26
[60] D. Mumford, The red book of varieties and schemes. Lecture Notes in Math. 1358,
Springer-Verlag, Berlin 1988.
[61] J. Nestruev, Smooth manifolds and observables. Grad. Texts in Math. 220, Springer-Verlag,
New York 2003.
[62] I. Penkov and V. Serganova, Generic irreducible representations of finite-dimensional Lie
superalgebras. Internat. J. Math. 5 (1994), 389–419. 245, 251, 252
[63] I. Penkov, Characters of typical irreducible finite-dimensional q.n/-modules. Funct. Anal.
Appl. 20 (1986), 30–37. 248
[64] W. Rudin, Functional analysis. Internat. Ser. Pure Appl. Math., McGraw Hill, New York
1991. 84, 269, 270
[65] A. Salam and J. Strathdee, Super-gauge transformations. Nucl. Phys. B 76 (1974), 477–
482. vii, 140
[66] M. Scheunert, The theory of Lie superalgebras. Lecture Notes in Math. 716, Springer-
Verlag, Berlin 1978.
[67] V. Serganova, Kac-Moody superalgebras and integrability. Preprint. 252, 254, 255
[68] V. Serganova, Kazhdan-Lusztig polynomials and character formula for the Lie superalge-
bra gl.mjn/. Selecta Math. (N.S.) 2 (1996), 607–651. 255
[69] V. Serganova, On generalizations of root systems. Comm. Algebra 24 (1996), 4281–4299.
[70] J.-P. Serre, Faiseaux algebrique coherent. Ann. of Math. 61 (1955), 197–278. 44
[71] R. Steinberg, Lectures on Chevalley groups. Notes prepared by John Faulkner and Robert
Wilson, Yale University, New Haven, Conn., 1968. 249
[72] M. E. Sweedler, Hopf algebras. Math. Lecture Note Ser., W. A. Benjamin, Inc., New York
1969. 26, 83, 120, 123
[73] F. Trèves, Topological vector spaces, distributions and kernels. Academic Press, NewYork
1967. 75, 273
[74] G. M. Tuynman, Supermanifolds and supergroups. Math. Appl. 570, Kluwer Academic
Publishers, Dordrecht 2004.
[75] V. S. Varadarajan, Lie groups, Lie algebras, and their representations. Grad. Texts in Math.
102, Springer-Verlag, New York 1984. 125, 154, 157, 158, 245
Bibliography 281
JA , 5 of ringed spaces, 34
Jacobi identity, 232 of schemes, 36
Jacobian, 68, 72, 92 of superdomains, 55
JM , 49 of supermanifolds, 48
of superspaces, 47, 174
Kostant, B., v, viii, 53, 82, 89, 112, 118 sheaf, 32
Koszul, J.-L., 112 super vector space, 2
morphisms
left translation, 118 between objects in a category, 259
left-invariant, 221 Mred , 49
left-invariant vector fields, 115 multiplication, 112
Leibniz identity, super, 221
Leites, D. A., v, viii, 89 natural transformation, 260
L.G/, 221
Lie algebra, super, 8 Ob.C /, 259
Lie algebra-valued functor, 208 objects in a category, 259
Lie superalgebra, 207, 232 open affine subfunctor, 185
line in root space, 245 open subfunctor, 163, 185
linear representation, 223 open submanifold, 61
local coordinates, 48 open subscheme, 174
local functor, 163, 183, 262 open subspace, 46
local super Frobenius theorem, 103 open supermanifold subfunctor, 163
local superdiffeomorphism, 91 opposite category, 260
locality, 183 orbit morphism, 143
localization, 78 orthosymplectic Lie superalgebra, 233
locally convex, 269 osp.V /, 233
locally finite refinement, 61
locally ringed space, 34 parity, 1
parity
ˇ reversing
ˇ functor, 2
z , 49, 62 @ ˇ @ ˇ
M ,
@t i x @ j x
, 68
Manin, Yu., v, 53 partition of unity, 61
maximal spectrum, 74 P .n/, 234
Milnor exercise, 76 Poincaré–Birkhoff–Witt Theorem, 18
Minkowski space, vii Poincaré group, vii
Minkowski superspace, 166 Poincaré supergroup, 169
module, 9 point supported distribution, 83
Morgan, J. W., v presheaf, 30
morphism product of supermanifolds, 65, 75
Hopf superalgebra, 25 projective localization, 193
of complex analytic supermanifolds, projective module, 266
86 projective space, 37
of locally ringed spaces, 34 projective special linear Lie superalgebra,
of real analytic supermanifolds, 86 232
286 Index