<<

An introduction to supermanifolds and

Fr´ed´eric H´elein Institut de Math´ematiques de Jussieu, UMR 7586 Universit´eDenis Diderot – Paris 7, Case 7012, 2 place Jussieu 75251 Paris Cedex 5, France February 14, 2008

The aim of these notes (which were partially presented in a talk given at the Peyresq Summer School on 17–22 June, 2002) is to give an introduction to some mathemat- ical aspects of supersymmetry. Although supersymmetry has a greater and greater impact on mathematics today, this subject remains not very well-known among mathematicians, despite the fact that several mathematical theories has been built for three decades (and many good texts are available now). I believe that, beside the technical complexity of the subject, this is due to the psychological difficulty to feel confortable with a with ‘odd’, i.e. anticommutating coordinates. Hence one of my motivations in writing this text was precisely to try to remove some of these psychological obstacles and I hope to help the interested reader in this direc- tion. Most notions expounded here are standard and can be found in many other references (see for instance [9, 8, 11, 25]). However I also added some (hopefully) original point of view, by using and developping some results in [14], where maps of supermanifolds are analyzed in details and also in the presentation of the super Minkowski spaces in dimensions 3, 4, 6 and 10 in Section 8. Two different theo- ries exists (see Section 1.2.1) and I choosed to follow the one inspired by algebraic geometry (which goes back to F. Berezin). I introduce some terminology, speak- ing of ‘skeletal’ supermanifolds and of supermanifolds ‘with flesh’ (D. Freed writes in [11] ‘with fuzz’) for pedagogical reasons, since we find both kinds of animals in the litterature under the same name. During the maturation and the preparation of theses notes, I have been very much inspired by [8, 11] and I beneficiated from several discusssions with D. Bennequin.

1 Introduction

1.1 Motivations was invented by physicists and mathematicians at the beginning of the seventies as a mathematical framework for the notion of supersymmetry. This notion merged out from two questions:

1 1. Does there exists a symmetry which exchanges Bosons and Fermions ?

2. Are they non trivial extensions of the Poincar´egroup, the symmetry group of Special Relativity ?

It could be tempting a priori to answer ‘no’ to both questions. Indeed first it is recognized for a long time that, in the theory of quantum fields, Bosons are created and annihilated by commutative operators, whereas to Fermions correspond anti- commuting operators, so that it is not straightforward to figure out how these two families of operators could be exchanged. Second a ‘no go’ theorem of S. Coleman and J. Mandula [6] answers negatively to the second question. Note also that a con- nection between the two questions can be expected, since Bosons and Fermions can also be distinguished by the fact that their fields are transformed in two different ways by the action of the Lorentz group. However it turns out that, in some enlarged conceptual framework, the answers to both questions are positive and rely on the same ideas. The key was the intro- duction by A. Salam and J. Strathdee of anticommuting coordinate variables (also called Grassmann variables), which one could heuristically think as classical limits of Fermionic operators. This leads to a theory of or supermanifolds, proposed by J. Wess and B. Zumino and developped by F. Berezin, B. DeWitt, A. Rogers in which the ‘classical limit’ of the symmetry between Bosons and Fermions admits a geometric formulation. In this framework one can also extend the Lie al- gebra of the Poincar´egroup into a super , as done by Y.A. Gol’fand and E.P. Likhtman [13] and D. Volkov and V. Akulov [31]. The mathematical theory of Lie super algebra and Lie super groups was then constructed by V. Kac [15].

1.2 Supermanifolds 1.2.1 A first draft We first need a mathematical definition of supermanifolds. Actually several theories are possible, depending on the point of view we start with and on the applications we have in mind. We hence have different, sometimes non equivalent, notions of supermanifolds which however all share the same underlying idea. The two main approaches are:

1. either we define supermanifolds indirectly by defining their (super)algebras of functions. This point of view, inspired by algebraic geometry, was developped by F. Berezin, D. Le˘ıtes, Y. Manin.

2. or we define supermanifolds as sets with a suitable topology. This point of view, inspired by the differential geometry, was introduced by B. DeWitt and A. Rogers (see section 5).

In both cases the definition of a supermanifold can be decomposed in two steps: m k m first one defines the analogues R | (for m, k N) of R and their open subsets ∈ m k and, second, we fix the rules for gluing together open subsets of R | to get a supermanifold of dimension m k. However variants exist in both approaches. Here we will follow mainly the first point| of view. We start by defining a basic version of it, that we propose to call ‘skeletal supermanifolds’.

2 m k m k Definition 1.1 (Skeletal R | ) The skeletal super space R | is a ‘geometric ob- m k ject’ indirectly defined through its ring ( ∞(R | ), +, ) of smooth functions. This C 1 m· 1 k ring is spanned algebraically over functions in ∞(R ) by k generators θ , , θ m k C ··· ∈ ∞(R | ) which satisfy C i, j, θiθj + θjθi =0 (1) ∀ and such that θ1 θk =0. (2) ··· 6 m k m 1 k In other words, ∞(R | ) := ∞(R )[θ , , θ ], i.e. any smooth function F C C ··· ∈ ∞(W ) can be written in an unique way as the polynomial C k i1 ij F = fi1 i θ θ , ··· j ··· j=0 1 i1<

Definition 1.2 (super vector space) A real super vector space V is a real Z2- graded vector space, i.e. a vector space over R which admits a decomposition V = V 0 V 1. Vectors in V 0 are called even, vectors in V 1 are called odd. ⊕ 0 1 0 If x V V , its parity or its grading gr(x) Z2 is 0 if x V and 1 if x V 1. ∈ ∪ ∈ ∈ ∈ Definition 1.3 (Super algebra) A real super algebra A is a real super vector space A = A0 A1 endowed with an associative bilinear multiplication2 law A A ⊕ × −→ A which maps Aa Ab to Ac, for all a, b, c 0, 1 such that a + b = c mod 2. A super algebra× is super commutative∈{if a,} b 0, 1 , f Aa, g Ab, fg =( 1)abgf. ∀ ∈ { } ∀ ∈ ∀ ∈ − m k m k 0 We see immediately that ∞(R | ) is a super algebra, and that ∞(R | ) (respec- m k 1 C C i tively ∞(R | ) ) is the subspace of even (respectively odd) polynomials in the θ ’s. RelationC (1) means that this super algebra is super commutative3. Examples of super algebras are the exterior algebra (or the Grassmann algebra) Λ∗V of a vec- tor space V or the set of sections of the exterior bundle Λ∗E of a vector bundle over a manifold (actually the latter example is very close to the notion of a M supermanifold, see below). m k With the definition of open subsets of R | at hand, one can define a skeletal supermanifold of dimension m k as follows: it is an ordinary manifold of M | |M| 1This can be rephrased by saying that ∞(Rm|k) is a module over ∞(Rm). 2One can alternatively view the multiplicationC law as a linear mapC from A A to A. 3Moreover (2) garantees that it is the free super commutative super algebra⊗ spanned by θ1, ,θk. · · · 3 dimension m endowed with a of super algebras (the sheaf of functions) M over such that, for any open subset U of Owith a chart U Ω m |M| | | |M| | | −→ | | ⊂ R , the Γ( U , U ) of sections of over U is isomorphic to some 1 | |k O| | OM | | m k ∞(Ω) = ∞( Ω )[θ , , θ ], where Ω is an open subset of R | . Note that the C C | | ··· algebra ∞( U ) ∞( Ω ) of smooth functions on the ordinary open subset U C | | ≃ C | | | | coincides with ∞(Ω)/ , where is the ideal of nilpotent elements of ∞(Ω). C JΩ JΩ C This allows to recover the sheaf of smooth functions from and hence the |M| M manifold . We shall not developp this descriptionO further,O since most of the |M| m k examples studied in this text requires only to understand R | (see [18] for details). Skeletal supermanifolds of dimension m k can be constructed starting from a rank k real vector bundleME over an m-dimensional| ordinary manifold . We |M| just set that the sheaf of functions on is isomorphic to the set Γ( , Λ∗E) M |M| of smooth sections of the exterior algebras bundle Λ∗E over . We denote by |M| S( , E)= this supermanifold and call it a split supermanifold (see e.g. [25]). A theorem|M| byM M. Batchelor [4] asserts that any supermanifold has a split structure. We chosed the terminology ‘skeletal’ because, as we shall see in Section 2, if we use naively Definition 1.1 for studying physical models we run relatively quickly into inconsistencies. Hence we will need to refine this notion and to put ‘flesh’ on the skeletal super . However (in the algebro-geometric theory) a super manifold with flesh can be completely recovered by knowing its skeletal part.

1.2.2 Related algebraic notions Clifford algebras — Let V be a real vector space endowed with a symmetric bilin- ear form B( , ) and let us define the Clifford algebra C(V ) associated to (V, B( , )). · · j 0 R · · We consider the free tensor algebra V := j∞=0 V ⊗ (where V ⊗ := ) and the ideal in V spanned by x y + y x +2B(x, y) x, y V . We let C(V ) I { ⊗ N ⊗ L | ∈ } to be the quotient of V by . If(e1, , ek) is a basis of V , then as a vector space, C(VN) is spanned over RIby all products··· e e , where 0 j k and i1 ··· ij ≤ ≤ 1 i < < i k, whereN we write xy = x y for shortness (with the convention ≤ 1 ··· j ≤ ⊗ that, for j = 0, ei1 eij = 1). As an algebra, C(V ) is spanned by 1, e1, , ek as- suming the conditions··· e e + e e = 2B(e , e ). If for instance B = 0, C(V···) is then i j j i − i j nothing but the Grassmann algebra R[e1, , ek]. A Clifford algebra is naturally a super algebra with C(V ) = C(V )0 C(V···)1, where C(V )0 (respectively C(V )1) is ⊕ spanned by products ei1 eij of an even (respectively odd) number of factors. It is however not super commutative,··· unless B = 0. A simple example is the following: let L be a real Euclidean vector space of dimension 1 (i.e. isomorphic to... R) and note ǫ a normed basis of L (i.e. ǫ, ǫ = 1). As a vector space, the Clifford algebra C(L) is isomorphic to a + bǫ ha, bi R and, because of the relation ǫ2 = 1, we have the algebra isomorphism{ |C(L)∈ C}. − ≃ It looks like a complicated way to define C, however C(L) comes with a natural grading C(L) = C(L)0 C(L)1, where C(L)0 = R and C(L)1 = Rǫ (i.e. ǫ is odd). ⊕ We shall use this construction in Section 8. Clifford algebras play an important role in physical applications since, roughly speaking, the quantization of (observable) functions on an odd supermanifold are elements of a Clifford algebra (the Hilbert space of quantum states being a spinorial representation of the Clifford representation). A Clifford algebra can so be seen as a deformation quantization of a super commutative super algebra described to first

4 order by the bilinear form B (which can be interpreted as a , see [17]).

Lie super algebras — A super Lie algebra g is a super vector space g = g0 g1 ⊕ endowed with a bilinear product law [ , ] : g g g called super Lie bracket which satisfies the following properties. First· · it is×super−→ skewsymmetric: x, y g0 g1, ∀ ∈ ∪ [x, y]+( 1)gr(x)gr(y)[y, x]=0, (3) − and second it satisfies a super Jacobi identity: x, y, z g0 g1 ∀ ∈ ∪ ( 1)gr(x)gr(z)[x, [y, z]]+( 1)gr(y)gr(x)[y, [z, x]]+( 1)gr(z)gr(y)[z, [x, y]]=0. (4) − − − In particular a super Lie algebra is not a super algebra since it is not associative (the latter property is replaced by the super Jacobi identity). A convenient way to take into account the sign rules in these identities consists in tensoring g with 1 L i a Grassmann algebra ΛL = R[η , , η ] (and assuming that the odd variables η supercommute with elements in g):··· for any x, y, z g0 g1, we choose ‘coefficients’ ∈ ∪ α,β,γ ΛL such that αx, βy and γz are all even (equivalentely gr(α) = gr(x), etc.). Then∈ (3) and (4) are equivalent to the usual relations [αx,βy] + [βy,αx]=0 and [αx, [βy,γz]] + [βy, [γz,αx]] + [γz, [αx,βy]] = 0 of a Lie algebra (see 6.2 for an 0 0 0 1 1 § example). In other words (ΛL g) = (ΛL g ) (ΛL g ) is a Lie algebra. Alternatively the super Jacobi⊗ identity⊗ (4) can⊕ be interpre⊗ ted as follows: 1. relation (3) for x, y g0 and (4) for x, y, z g0 means that (g0, [ , ]) is a Lie ∈ ∈ · · algebra;

2. relation (4) for x, y g0 and z g1 means that, x g0, the linear map ∈ ∈ ∀ ∈ ad : g1 g1 x −→ z [x, z] 7−→ satisfies ad (z)=ad (ad (z)) ad (ad (z)), i.e. g0 End(g1), x ad [x,y] x y − y x −→ 7−→ x is a representation of g0;

3. relation (4) for x g0 and y, z g1 means that the (symmetric) bilinear map ∈ ∈ g1 g1 g0 × −→ (y, z) [y, z] 7−→ is g0-equivariant (hence the adjoint representation of g0 on g1 is somehow a ‘square root’ of the adjoint representation of g0 on g0);

4. relation (4) for x, y, z g1 is equivalent to [x, [x, x]] = 0, x g1. ∈ ∀ ∈ The classical Lie algebras are of course examples of super Lie algebra (with g1 = 0 ). Some non trivial Lie algebras come from differential Geometry. For instance { } consider an usual manifold and let Ω∗( ) be the space of smooth sections of M M the exterior bundle Λ∗T ∗ of the cotangent bundle T ∗ (i.e. smooth exterior differential forms on ).M Let ξ be a smooth vector fieldM on . Then there exists M M three natural operators acting on Ω∗( ): the exterior differential d, the interior M product by ξ, ιξ, and the Lie Lieξ. The space g spanned by these three 0 1 operators is a Lie super algebra, where g is spanned by Lieξ and g is spanned

5 by d and ιξ. This is a way to summarize the commutation relations [Lieξ,d] = [Lieξ, ιξ] = 0, [d,d]=2dd = 0, [ιξ, ιξ]=2ιξιξ = 0 and the well-known Cartan identity [d, ιξ] = dιξ + ιξd = Lieξ, an anti-commutator. Another example is the space of multivectors (i.e. sections of the exterior product of the tangent bundle T ) equipped with the Schouten bracket (a generalization of the Lie bracket of vectorM fields), see e.g. [27]. Then the parity of a multivector field is 1 plus its degree mod 2. Super Lie algebra play also an important role in supersymmetric theories since they precisely encode the supersymmetry of a given physical model. Among the most interesting are the extensions of the (trivial) Lie algebra of translations of the Minkowski space-time (this is then the super Lie algebra of super translations of space-time) or of the Poincar´eLie algebra (called the super Poincar´eLie algebra). We present examples of such Lie super algebras in Section 8. Lie super algebras are naturally associated with Lie super groups, which are particular examples of supermanifolds. However a Lie super algebra as defined here is a skeletal object, which needs some extra ‘flesh’ for the construction of the corresponding (what we would call a Lie algebra with flesh corresponds actually to a super Lie module, see [25]). We refer to [11, 25] for quickly accessible expositions of the theory of super Lie algebras and to [8, 7, 29, 15] for more details.

2 The superparticle: a first overview

Recall that supersymmetry was introduced as a hypothetical symmetry of a quantum field theory which would exchange Bosons with Fermions. However most quantum field descriptions of a particle are connected with a classical field description, i.e. a variational problem. So we can expect that the supersymmetry can be read at this classical level as a symmetry group which acts on solutions of the Euler–Lagrange equations of some variational problem: we shall see that this is indeed true.

2.1 The model A simple example is a superparticle evolving in a Riemannian manifold . A first N (and relatively rough) description of this system is through a pair (x, ψ), where

x is a smooth map from R to • N

ψ is a section of the pull-back bundle x∗ (ΠT ), i.e. for all t R, ψ(t) belongs • M ∈ to ΠT . Here ΠT is the space modelled on T with opposite x(t)N x(t)N x(t)N parity, i.e. odd. This means that the set of smooth functions on ΠTx(t) is 1 n 1 n N R[α , , α ] Λ∗T ∗ , where (α , , α ) is a basis of T ∗ . ··· ≃ x(t)N ··· x(t)N In other words (x, ψ) can be understood as a map from R to the split manifold S( , T ∗ ) (up to some difficulties discussed at the end of 3.2.1). NHereN we have used the notation ΠT , where Π is§ a vector space x(t)N which reverses the parity (odd versus even) of a given Z2 graded vector space. This notation is useful because of its concision but one should keep in mind that Π is never a super algebra morphism.

6 The Lagrangian for the superparticle is

1 2 1 [x, ψ] := x˙ + ψ, x˙ ψ dt, L R 2| | 2h ∇ i Z  

i where, if ψ(t) = ψi(t) ∂ , ψ(t) = dψ (t)+Γi (x(t))x ˙ j (t)ψk(t) ∂ (the Γi ’s ∂xi ∇x˙ dt jk ∂xi jk being Christoffel’s symbols). We already see here that the ψi’s should be anticom- 1 1 d muting quantities: if not then we would have 2 ψ, x˙ ψ = 4 dt ψ, ψ , so that the second term in the Lagrangian would be unuseful.h ∇ i h i The Euler–Lagrange system of equations is

x˙ = 1 R(ψ, ψ)x ˙ ∇x˙ 2 (5) ψ = 0.  ∇x˙ l i j k ∂ Again R(ψ, ψ)x ˙ = R(x)ijk ψ ψ x˙ ∂xl does not necessarily vanish precisely because ψi and ψj anticommute. When contemplating this Euler–Lagrange system we see a difficulty due to our too naive approach. Indeed in the first Euler–Lagrange equation l l j k 1 l i j k x¨ +Γjk(x)x ˙ x˙ = 2 R(x)ijk ψ ψ x˙ , the left hand side is real valued whereas the right hand side contains the bilinear quantity ψiψj. However ψiψj cannot be real (although it is even) because it is nilpotent (it squares to zero). Hence

Remark 2.1 The coordinates xi of the map x cannot be real valued because of (5).

The action enjoys a particular symmetry (which corresponds actually to a supersymmetry)L as follows: we consider the transformation 1 ητ acting on the − Q field (x, ψ) defined by: x x ηψ 7−→ − (6) ψ ψ + ηx.˙  7−→ Here an important point is that η which plays the role of an infinitesimal parameter is odd. Otherwise this transformation would exchange odd fields with even ones. Note however that x cannot not be mapped into an ordinary map by (6) because the components of ηψ cannot be real; hence this confirms the conclusion claimed in Remark 2.1. The ‘vector’ ( ηψ(t), ηx˙(t)) can be interpreted as a vector tangent to ΠT to (x(t), ψ(t)). Let us− compute the effect of this infinitesimal transformation N on the action : we need to compute the term ητQ δ (x, ψ) in the development (x ηψ,ψ +Lηx˙)= (x, ψ) ητ δ (x, ψ). ObservingL that (6) implies that L − L − Q L

x˙ x˙ η x˙ ψ ψ 7−→ −ψ +∇η x,˙  ∇x˙ 7−→ ∇x˙ ∇x˙ we obtain that 1 1 ητQ δ (x, ψ) = x,˙ η x˙ ψ + ηx,˙ x˙ ψ + ψ, η x˙ x˙ dt L R h ∇ i 2 h− ∇ i 2 h − ∇ i Z1  d  = η ψ, x˙ dt 2 R dt h i 1 Z = η [ ψ(+ ), x˙(+ ) ψ( ), x˙( ) ] . 2 h ∞ ∞ i−h −∞ −∞ i

7 This quantity does not vanish in general, but is exact. So the transformation (6) should formally change infinitesimally a solution of the Euler–Lagrange equations into another one4. This symmetry gives rise to a conserved quantity (a version of Noether theorem): R let us introduce a smooth compactly supported function χ 0∞( ) and, instead of (6), apply to (x, ψ) the infinitesimal transformation ∈ C

x x ηχψ (7) ψ 7−→ ψ −+ ηχx,˙  7−→ with x˙ x˙ η (χ x˙ ψ +χψ ˙ ) ψ 7−→ −ψ + η∇(χ x˙ +χ ˙x˙) .  ∇x˙ 7−→ ∇x˙ ∇x˙ Then we have 1 1 (χητQ) δ (x, ψ) = χ x,˙ η x˙ ψ + ηx,˙ x˙ ψ + ψ, η x˙ x˙ dt L R h ∇ i 2 h− ∇ i 2 h − ∇ i Z  1  + x,˙ ηχψ˙ + ψ, ηχ˙x˙ dt R h i 2 h − i 1 Z  d 3 d = η χ ψ, x˙ dt η χ ψ, x˙ dt 2 R dt h i − 2 R dt h i Z d Z = η χ ψ, x˙ dt. − R dt h i Z In all these computations we have used that η, ψi and ψ˙ i anticommute. We hence conclude that the quantity ψ, x˙ is conserved, i.e. Noether’s theorem holds here5, hence the symmetry which generatesh i our fields transformation should have a geomet- rical interpretation. Eventually this ‘Noether’s charge’ has an interesting physical meaning, since the corresponding quantum operator is nothing but the Dirac oper- ator on (see [1, 11, 25, 33]). N Also we can interpret the transformation acting on the set of fields

Q := ητ :(x, ψ) ( ηψ,ηx˙) − Q 7−→ − as a tangent vector field to the infinite dimensional set of fields (x, ψ):

∂ ∂ Q(x, ψ)= η ψ(t) +x ˙(t) dt R − ∂x(t) ∂ψ(t) Z  

It is interesting to choose two different anticommuting variables η1 and η2 and to compute the commutator of the ‘vector fields’ Q1 := η1τQ and Q2 := η2τQ. This can be done by computing the action of Q Q : − − 1 ◦ 2

Q2 Q1 x η2ψ η1(η2x˙) x˙ − − = η1η2 , ψ 7−→ η2x˙ 7−→ η1( η2 x˙ ψ) − x˙ ψ      − ∇   ∇  4The question whether this infinitesimal transformation can be exponentiated is however more delicate and should be done in terms of a map defined on R1|1, see [25] 5in a way similar to the fact that the time translation symmetry t t + ε induces the fields 7−→ transformation (x, ψ) (x εx,˙ ψ ε x˙ ψ) which is related, through Noether’s theorem, to the 7−→ 1 − 2 1 − ∇ conservation of the energy x˙ + ψ, x˙ ψ . 2 | | 2 h ∇ i 8 and the action of Q2 Q1 (vice-versa). Then by using the fact that the Levi–Civita connection is torsion-free,◦ we can write that [Q , Q ](x, ψ) = Q (Q (x, ψ)) ∇ 1 2 1 2 − Q2 (Q1(x, ψ)) and deduce that

[η τ , η τ ](x, ψ) = [Q , Q ](x, ψ)= 2η η (x, ψ). 1 Q 2 Q 1 2 − 1 2∇x˙

Factoring out by η1 and η2 (and taking into account the fact that the odd numbers ηi anticommute also with the odd operator τQ) we deduce that

[τ , τ ](x, ψ)=2 (x, ψ). Q Q ∇x˙

Here the notation [τQ, τQ] does not recover a commutator, but an anticommutator ([τQ, τQ]=2τQτQ). We conclude that the action of τQ on fields should be related somehow to an infinitesimal geometric symmetry which behaves like the square root d of the time translation generator 2 dt .

2.2 The superfield formulation As we shall see in more details in Section 6 there is an alternative, more elegant way to picture geometrically this symmetry by viewing the multiplet (x, ψ) as the 1 1 components of a single superfield Φ = x + θψ, from the supertime R | on which the ring of functions is ∞(R)[θ] (i.e. θ is here an odd coordinate) to . Then C 1 ∂NΦ the Lagrangian action can be written as [Φ] := 2 R1|1 dtdθ DΦ, ∂t , where ∂ ∂ L − D := ∂θ θ ∂t . And the transformation (6) results from an infinitesimal translation R1 1 − ∂ ∂ R R in | by the vector ητQ, where τQ := ∂θ + θ ∂t . However another limitation of our temporary definition of a supermanifold is that:

1 1 Remark 2.2 The superfield Φ does not belong to ∞(R | , ) according to the def- 1 1 C N inition of maps in ∞(R | ), since ψ is odd. C 2.3 Conclusion We have seen a first example of a supersymmetric model, an opportunity to test the concept of a skeletal supermanifold, and some points must be clarified: 1. to give a sense to (5) and (7), which requires in our example to change the mathematical definition of x and also symmetrically of ψ (Remark 2.1);

2. to extend the definition of a map on a supermanifold in such a way that the superfield Φ = x + θψ makes sense (Remark 2.2);

3. a last question, not yet discussed: can we figure out what is a point in a supermanifold ? Let us first consider the second point: one of the more natural ideas which comes in mind to answer it would be to assume that the components ψi of ψ are not R-valued but takes values in, say R[η1], where η1 is an odd parameter which is different from θ (to ensure that θη1 = 0). However this would imply that the right 6 1 hand side of the first equation of (5) is 2 R(ψ, ψ)x ˙ = 0, because η1η1 = 0. This would solve the first point, but in a stupid way and the superparticle equation would have

9 little interest since the equation on x would not be coupled to ψ. Hence a refined solution is to assume that the components of ψ take values in, say R[η1, η2], where θ, η1, η2 are three different odd parameters. Then the right hand side of (5) would be proportional to η1η2. This means that the components of x should not be real but with values in R[η1η2]. Now we observe that there are no a priori reason for using only two odd parame- ters η1, η2. Hence we can answer the two first questions by assuming that all the fields x and ψ are R[η , , η ]-valued, where L is arbitrary, possibly infinite and that x 1 ··· L is an even polynomial on (η1, , ηL), whereas ψ is an odd polynomial. There are different ways to implement this··· idea. We here privilege the point of view inspired by algebraic geometry, i.e. to view any odd parameter as a coordinate function on some ‘space’. Hence the ring R[η1, , ηL] has to be thought as the ring of functions 0 L ··· 1 1 0 L 1 1+L on some space R | and the field Φ = x+θψ as a map on R | R | R | . Then × ≃ the conditions that x (respectively ψ) is an even (respectively odd) function of the ηi’s can be encapsulated by requiring that Φ is an even function of (θ, η1, , ηL). 1 1 0 L ···0 All that show that we need a definition of the set ∞(R | R | , ) of even 1 1 0 L C × N maps from R | R | to . It should be the set of Φ such that, for any local coordinate y : × U NR, one can make sense of y Φ and such that y Φ 1 1 0 LN ⊃0 −→ ◦ ◦ ∈ ∞(R | R | , R) . A precise statement of this property will be given in 3.1. C × § More generally the mathematical description of a σ-model6 on a supermanifold , i.e. a field defined on and with values on a (possibly super) manifold will M M 0 L N be obtained through an even smooth map from R | to , where the choice of the value of L has to be precised and its meaningM × can be interpretedN in different ways. Hence we are led to another definition of a supermanifold, that we propose to call ‘supermanifolds with flesh’, see 2.4. § This change of paradigm changes also the content of the last question, which relies on our intuitive picture of a supermanifold. We invite the Reader to build his own vision, after reading the next sections (see also [11, 14]).

2.4 A new definition

m k 0 L m k Definition 2.1 A supermanifold with flesh is a product | R | , where | is a skeletal supermanifold and L N is ‘sufficiently large’.M × M ∈ m k A map with flesh on the supermanifold | and with values in an ordinary n ℓ Mm k 0 L n ℓ manifold | is an even smooth map from | R | to | . N M × N m k 0 L We postpone the precise definition of what is an even smooth map from | R | n ℓ M × to | to the next paragraph (3.1). Moreover we stayed vague about the the precise N value of L and there are several ways to implement its role. Either one let L to be infinite and we precise some topology: this point of view • is connected to the theory of G∞-functions and G∞-manifolds developped by B. DeWitt and A. Rogers and which will be briefly expounded in Section 5. Or one looks for the minimal value of L which is needed in order to make the • computations consistent (we need at least L>k): this is connected to the

6This includes the important case of superstrings, where is a supermanifold of even dimension 2. M

10 theory of GH∞-functions (an intermediate between the G∞ and H∞ theories, see Section 5), see [24].

Or we let the value of L to be free and arbitrary: it means that a map with • m k n ℓ m k flesh Φ : | | represents the collection of all even maps Φ : | 0 L M n ℓ −→ N M × R | | , for L N. Equivalentely, as explained in the solution of the −→ N ∈ m k n ℓ Fall Problem 2 in [34], a ‘map’ Φ : | | should be considered M −→ N 0 L as a functor from the category of odd vector spaces R | (or more generally m k 0 L n ℓ superspaces) to even skeletal maps Φ : | R | | . M × −→ N The strategy to understand maps with flesh between supermanifolds is then based on two steps:

1. first define smooth maps (and in particular even ones) between skeletal super- manifolds. This is the subject of the next Section;

2. then specialize the preceding analysis to maps on the skeletal super manifold m k 0 L m k | R | and deduce a description of maps with flesh on | . M × M Note that we may adapt Definition 2.1 of a map between supermanifolds to sections of fiber bundles or connections.

3 Maps between skeletal supermanifolds

m q n ℓ In this section we define maps between skeletal supermanifolds | and | (of M N respective dimensions m q and n ℓ and where, in the following applications, we may |m q | m k 0 L have q = k + L and | = | R | ). The strategy (as in [11, 8]) is the M M × m q n ℓ following: rather than defining a map Φ : | | directly we set how it M −→ N acts on functions by the pull-backf (or composition) operation f Φ∗f = f Φ. 7−→ ◦ We will essentially use the fact that Φ∗ should be a morphism between the two m q n ℓ superalgebras ∞( | , R) and ∞( | , R). C M C N 3.1 The morphism property

m q n ℓ m q Given two skeletal supermanifolds | and | , a smooth map Φ : | n ℓ M N M −→ | , is defined by duality through the induced morphism of algebras N n ℓ m q Φ∗ : ∞( | , R), +, . ∞( | , R), +, . C N −→ C M A Φ∗A,  7−→  where we can think secretly that Φ∗A = A Φ. Recall also that we denote by m q 0 n ℓ 0 ◦ ∞( | , R) and ∞( | , R) the subalgebras of even elements of respectively C Mm q C Nn ℓ ∞( | , R) and ∞( | , R). Then we will say that Φ∗ is an even morphism (or C M C N m q n ℓ n ℓ 0 equivalentely that Φ : | | is an even map) if Φ∗ maps ∞( | , R) to m q 0 M −→n ℓ N 1 m q 1 C N ∞( | , R) and maps ∞( | , R) to ∞( | , R) . C TheM important thing isC toN check the morphismC M property, i.e. that

Φ∗(1) = 1, where 1 is the constant equal to one (the unit in algebraic terms) (8) n ℓ λ,µ R, A, B ∞( | , R), Φ∗(λA + µB)= λΦ∗A + µΦ∗B (9) ∀ ∈ ∀ ∈ C N 11 and n ℓ A, B ∞( | , R), Φ∗(AB)=(Φ∗A)(Φ∗B). (10) ∀ ∈ C N These conditions impose severe constraints [11]. In the following we analyze their consequences through some examples.

3.2 Some examples The goal of this paragraph is to experiment maps from a skeletal supermanifold according to the previous definition.

m 0 0 ℓ 3.2.1 Maps Φ : R | R | −→ (From an even vector space to an odd vector space.) This amounts to look at 1 ℓ m 1 ℓ morphisms Φ∗ from R[θ , , θ ] to ∞(R ). Any function F R[θ , , θ ] writes ··· C ∈ ··· ℓ ℓ ℓ i 1 i1 i2 i1 ij F = F + Fiθ + Fi1i2 θ θ + = Fi1 i θ θ , ∅ 2 ··· ··· j ··· i=1 i1,i2=1 j=0 1 i1, ,i ℓ X X X ≤ X··· j ≤

where the Fi1 i ’s are real constants. Thus, because of (9), it suffices to characterize ··· j i1 i all pull-back images Φ∗(θ θ j ). For j = 0, (8) implies that Φ∗F = F . For ∅ ∅ i1 ···i 2 i1 i 2 j 1 we remark that (θ θ j ) = 0. We deduce by (10) that (Φ∗(θ θ j )) = ≥ i1 i 2 ··· i1 i m ··· i1 i Φ∗ ((θ θ j ) ) = 0 and hence, since Φ∗(θ θ j ) ∞(R ), this forces Φ∗(θ θ j )= ··· ··· ∈ C ··· 0. Hence Φ∗F = F = F (0), i.e. the pull-back image of F by Φ is a constant function, ∅

m 0|p R R

m 0 p Figure 1: A map from R to R |

0 ℓ whose value is equal to F , which can be interpreted as the value of F ‘at 0 R | ’. ∅ In other words Φ looks like a constant function and its constant ‘value’, if it∈ would make sense, is 0; we can hence conclude that Φ 0. This reflects the fact that 0 is 0 ℓ ≡ the only ‘classical’ point in R | . m 0 n ℓ The study of morphisms of the type Φ : R | R | is completely similar: one −→ m n finds that such morphisms are of the form Φ∗F = F ϕ, where ϕ : R R ◦ 1 1 −→ is an ordinary smooth map. In particular if Φ = x + θψ maps R to R | or, more generally, to ΠT , the morphism property forces ψ = 0. M

0 1 n 0 3.2.2 Maps Φ : R | | −→ N ≃N (From an ‘odd line’ to an (even) standard manifold.) This the opposite situation. We look for morphisms Φ∗ : ∞( ) R[θ]. Any such map is characterized by C N −→ two functionals a, b : ∞( ) R such that, f ∞( ), Φ∗f = a(f)+ b(f)θ. C N −→ ∀ ∈ C N Condition (9) implies that a and b are linear functionals. Then Condition (10) amounts to a(fg)= a(f)a(g), (11)

12 b(fg)= a(f)b(g)+ b(f)a(g). (12) The first relation (11) implies that there exists some point m such that a(f)= ∈N f(m), f. Then the second one (12) reads b(fg) = f(m)b(g)+ b(f)g(m), which ∀7 implies that b is a derivation. So that there exists ξ Tm such that b(f)= dfm(ξ), f. Thus ∈ N ∀

ξ

0|1 m R N

0 1 Figure 2: A map from R | to N

Φ∗f = f(m)+ df (ξ)θ, f ∞( ). m ∀ ∈ C N Hence Φ is characterized by a point in the tangent bundle T . N

0 2 n 0 3.2.3 Maps Φ : R | | −→ N ≃N (From an ‘odd plane’ to an (even) standard smooth manifold.) The analysis is 1 2 similar. We analyze morphisms Φ∗ : ∞( ) R[θ , θ ]: they are characterized C N −→ by four linear functionals a, b1, b2,c : ∞( ) R such that f ∞( ), Φ∗f = 1 2 1 2 C N −→ ∀ ∈ C N a(f)+ b1(f)θ + b2(f)θ + c(f)θ θ . Relation (10) for two functions f,g ∞( ) gives that ∈ C N 1 2 1 2 Φ∗(fg)= a(fg)+ b1(fg)θ + b2(fg)θ + c(fg)θ θ should be equal to 1 2 1 2 Φ∗fΦ∗g = (a(f)+ b1(f)θ + b2(f)θ + c(f)θ θ ) 1 2 1 2 (a(g)+ b1(g)θ + b2(g)θ + c(g)θ θ ) × 1 2 = a(f)a(g)+(a(f)b1(g)+ b1(f)a(g)) θ +(a(f)b2(g)+ b2(f)a(g)) θ +(a(f)c(g)+ c(f)a(g)+ b (f)b (g) b (f)b (g)) θ1θ2. 1 2 − 2 1 Again, by identifying coefficients of 1, θ1 and θ2, we deduce that there exists some point m such that a(f) = f(m) and there exist two vectors ξ1,ξ2 Tm ∈ N ∈ 1N2 such that b1(f) = dfm(ξ1) and b2(f) = dfm(ξ2). However by identifying the θ θ coefficients we obtain: c(fg)= f(m)c(g)+ c(f)g(m)+ df (ξ )dg (ξ ) df (ξ )dg (ξ ). m 1 m 2 − m 2 m 1 From fg = gf we deduce that the left hand side should be symmetric in f and g. However the right hand side is symmetric in f and g only if

(df dg) (ξ ,ξ )=0, f,g ∞( ). ∧ m 1 2 ∀ ∈ C N This is possible only if ξ and ξ are linearly dependant8, i.e. ξ T , λ ,λ R, 1 2 ∃ ∈ mN ∃ 1 2 ∈ s.t. ξ = λ ξ and ξ = λ ξ. If so we then conclude for c that ζ T such that 1 1 2 2 ∃ ∈ mN c(f)= dfm(ζ). Hence 1 2 1 2 Φ∗f = f(m)+(λ1θ + λ2θ )dfm(ξ)+ θ θ dfm(ζ). 7assuming that Φ∗ is continuous with respect to the 1 topology 8Such a conclusion does not look very natural: it is relatedC to the fact that we did not assume that Φ is even.

13 m q 3.3 The structure of a map of an open subset of R | We here generalize and precise the observations of the previous section. These results were obtained in [14]. We assume in the following that Ω is an open subset of Rm and is an ordinary manifold of dimension n. | | FirstN we introduce the following notations: for any positive integer j we let Iq(j) := (i , i ) [[1, q]]j i < < i , we denote by I = (i , i ) Iq(j) { 1 ··· j ∈ | 1 ··· j} 1 ··· j ∈ a multi-index and we write ηI := ηi1 ηij . We use also the convention Iq(0) = q q [q/···2] q [(q 1)/2] q . We let Iq := Iq(j), I := Iq(2j), I := − Iq(2j + 1) and I := {∅} ∪j=0 0 ∪j=0 1 ∪j=0 2 [q/2]Iq j=1 (2j). ∪ We denote by π : Ω the canonical projection map and consider | |×N −→ N the vector bundle π∗T : the fiber over each point (x, m) Ω is the tangent N Iq ∈| |×N space Tm . For any I 2, we choose a smooth section ξI of π∗T over Ω and we considerN the R[η∈1, , ηq]0-valued vector field N | |×N ··· I Ξ := ξIη . I Iq X∈ 2

Alternatively Ξ can be seen as a smooth family (Ξx)x Ω of smooth tangent vector 1 q 0 ∈| | fields on with coefficients in R[η , , η ] . So each Ξx defines a first order N ··· 1 q 0 differential operator which acts on the algebra ∞( ) R[η , , η ] , i.e. the set of smooth functions on with values in R[η1, C , ηNq]0,⊗ by the··· relation N ··· 1 q 0 I f ∞( ) R[η , , η ] , Ξ f = ((ξ ) f)η . ∀ ∈ C N ⊗ ··· x I x I Iq X∈ 2 We now define (letting Ξ0 = 1)

[q/2] ∞ Ξn Ξn eΞ := = , n! n! n=0 n=0 X X which can be considered again as a smooth family parametrized by x Ω of 1 ∈q 0 | | differential operators of order at most [q/2] acting on ∞( ) R[η , , η ] . Now we choose a smooth map ϕ : Ω and we considerC theN map⊗ ··· | | −→N 1 ϕ : Ω Ω × | | −→ | |×N x (x, ϕ(x)) 7−→ which parametrizes the graph of ϕ. Lastly we construct the following linear operator 1 q 0 on ∞( ) ∞( ) R[η , , η ] : C N ⊂ C N ⊗ ··· Ξ ∞( ) f (1 ϕ)∗ e f ∞(Ω), C N ∋ 7−→ × ∈ C defined by 

[q/2] n Ξ Ξx (Ξ)x x Ω , (1 ϕ)∗ e f (x) := e f (ϕ(x)) = f (ϕ(x)). ∀ ∈| | × n! n=0     X Ξ Ξ We will often abusively write (1 ϕ)∗ e f (x) ϕ∗ e f (x). We observe that × ≃ actually, for any x Ω , we only need to define Ξ on a neighbourhood of ϕ(x) in  x  , i.e. it suffices to∈| define| the section Ξ on a neighbourhood of the graph of ϕ in NΩ (or even on their Taylor expansion in m at order [q/2] around ϕ(x)). | |×N 14 Ξ Theorem 3.1 [14] The map f (1 ϕ)∗ e f is a morphism from ∞( ) to 0 q 0 7−→ × C N ∞( Ω R | ) , i.e. satisfies assumptions (9) and (10). C | | ×  0 q 0 Conversely for any morphism Φ∗ : ∞( ) to ∞( Ω R | ) , there exists a map ϕ : Ω and a R[η1, , ηq]0-valuedC vectorN fieldC |Ξ |on × along ϕ (i.e. a section | | −→N 1 q···0 N of π∗(T R[η , , η ] ) over a neighbourhood of the graph of ϕ in Ω ) N ⊗N ··· Ξ | |×N such that Φ∗ : f (1 ϕ)∗ e f . 7−→ × Furthermore one can assume that the vector fields ξ used to build Ξ commute,  I i.e. that [ξI,ξJ ]=0, for all I,J. Moreover if we assume that the image of ϕ is contained in the domain V of a single coordinate chart y : V Rn, then it is −→ possible to construct the vector fields ξI in such a way that ξIξJ y =0.

4 Maps with flesh between supermanifolds

Recall that, according to the definition given in 2.4 (following the point of view § m k n ℓ adopted in [10] and [11]), a map with flesh Φ : | | is defined through n ℓ m k M0 L −→ N m k 1 L even morphisms Φ∗ : ∞( | ) ∞( | R | ) ∞( | ) R[η , , η ]. C N −→ C M × ≃ C M ⊗ ··· 4.1 Examples The aim is here to illustrate the differences between a map with flesh and a skeletal one.

0 1 0 2 n 0 4.1.1 Even maps Φ from R | R | to | × N ≃N 0 1 It corresponds to a map with flesh from R | to (compare with example 3.2.2.) 0 1 0 2 N We look at even maps Φ from R | R | to an ordinary manifold , i.e. at × 0 2 0 0 2 N 1 2 even morphisms Φ∗ : ∞( ) R[θ] ∞(R | ) , where ∞(R | ) = R[η , η ]. C N −→ ⊗ C C Setting Φ f = A(f)+ θB(f), where A(f) is even and B(f) is odd, we again find ∗  that (using that A(f) and A(g) commute with all other terms):

A(fg)= A(f)A(g) and B(fg)= A(f)B(g)+ B(f)A(g). (13)

1 2 1 2 We also know that A(f) = a(f)+ η η a12(f) and B(f) = η b1(f)+ η b2(f). The first relation in (13) gives thus

a(fg)= a(f)a(g) and a12(fg)= a(f)a12(g)+ a12(f)a(g),

and hence there exists m and ζ T such that a(f) = f(m) and a (f) = ∈ N ∈ mN 12 dfm(ζ). The second relation gives

1 2 1 2 η [a(f)b1(g)+ b1(f)a(g)] + η [a(f)b2(g)+ b2(f)a(g)] = η b1(fg)+ η b2(fg).

Hence ξ ,ξ T such that b (f)= df (ξ ) and b (f)= df (ξ ). So ∃ 1 2 ∈ mN 1 m 1 2 m 2 1 2 1 2 Φ∗f = (f(m)+ η η dfm(ζ))+ θ (η dfm(ξ1)+ η dfm(ξ2)) 1 2 1 2 = f(m)+ η η dfm(ζ)+ θdfm (η ξ1 + η ξ2) 1 2 1 2 = ((1+ η η ζ)(1+ θ(η ξ1 + η ξ2)) f)(m),

15 where 1 stands for the identity operator acting on ∞( ) and we note, e.g., ζf := 0 2 C N1 2 1 2 Lieζ f. If we introduce the R | valued vector fields Ξ := η η ζ and Ξ1 := η ξ1 +η ξ2, ∅ we can write the preceding relation as:

Φ∗f = ((1+Ξ )(1+ θΞ1) f)(m). (14) ∅ We also write (14) symbolically9 as:

Φ= m (1+Ξ )(1+ θΞ1) . (15) ∅ 1 1 0 1 A variant consists in adding a time variable t R, i.e. to substitute R | to R | . ∈ Then we would obtain by a similar reasoning that a morphism Φ∗ from ∞( ) to 1 2 0 C N ( ∞(R)[θ, η , η ]) has the form C

Φ∗f = ϕ∗ ((1+Ξ )(1+ θΞ1) f) , (16) ∅

where ϕ ∞(R, ) and Ξ and Ξ1 have the same meaning as before. We write ∈ C N ∅ this symbolically as

Φ= ϕ (1+Ξ )(1+ θΞ1)=(ϕ + ϕ Ξ )+ θ (ϕ Ξ1) . (17) ∅ ∅

0 2 0 2 n 0 4.1.2 Even maps Φ from R | R | to | × N ≃N 0 2 It corresponds to a map with flesh from R | to (compare with 3.2.3.) We look at 1 2 N 1 2 0 even morphisms Φ∗ : ∞( ) (R[θ , θ ] R[η , η ]) . One finds that m , C N −→ ⊗ ∃ ∈ N ζ,χ,ξ , π T (for 1 α, β 2) such that ∃ αβ ∈ mN ≤ ≤ 1 2 1 1 2 2 1 2 Φ∗f = (f(m)+ η η dfm(ζ))+ θ dfm (η ξ11 + η ξ12)+ θ dfm (η ξ21 + η ξ22) 1 2 1 2 + θ θ (dfm(χ)+ η η (P f(m)+ dfm(π))) . (18) Here P is the second order differential operator P := ζχ ξ11ξ22 + ξ12ξ21, where we have used arbitrary extensions of the vectors ζ,χ,ξ , π − T to vectors fields αβ ∈ mN defined on a neighbourhood of m (for example we could choose the extensions in such a way that [ζ, χ] = [ξ11,ξ22] = [ξ12,ξ21] = 0). Clearly the operator P depends on the way we have extended the vectors fields. However two different choices of extensions lead to two operators P and P ′ such that P f(m) P ′f(m) = df (V ), − m for some V Tm . Hence changing the extensions is just equivalent to change π T . The∈ rightM hand side of (18) can factorized as ∈ mN 1 1 1 2 Φ∗f = (1+Ξ ) 1+ θ Ξ1 1+ θ Ξ2 1+ θ θ Ξ12 f (m), (19) ∅ 1 2 1 2  1 2   1 2 where Ξ := η η ζ, Ξ1 := η ξ11 + η ξ12, Ξ2 := η ξ11 + η ξ12 and Ξ12 := χ + η η π. ∅ We can summarize this result by the symbolic relation:

1 1 1 2 Φ= m (1+Ξ ) 1+ θ Ξ1 1+ θ Ξ2 1+ θ θ Ξ12 , (20) ∅ 9These notations are actually inspired by a system of notations proposed by M. Kawski and H. Sussmann [16].

16 4.2 The general description of maps with flesh

0 L 4.2.1 Even maps Φ from Ω R | to an ordinary manifold × N m k Here Ω is an open subset of R | . This case is a straightforward application of 0 k n Theorem 3.1. Indeed we have Ω Ω R | , where Ω is an open subset of R 0 L 0 q ≃ | | × | | and thus Ω R | Ω R | , with q := k + L. Hence we just need to label 1 ×k 1 ≃L | | × 0 k+L by (θ , , θ , η , , η ) the odd coordinates on R | and to apply Theorem 3.1. ··· ··· k k This requires adapted notations: we note A (0) = and for any j N∗, A (j) := j {∅} ∈ (a1, aj) [[1,k]] a1 < < aj . We denote by A = (a1, aj) an element { ··· ∈ | ··· } ··· [k/2] Ak A a1 aj Ak k Ak Ak Ak of (j) and write θ := θ θ . We let := j=0 (j), 0 := j=0 (2j), [(k 1)/2] ··· [k/2] ∪ ∪ Ak := − Ak(2j + 1), Ak := Ak(2j) and Ak := Ak Ak. Lastly we set 1 ∪j=0 2 ∪j=1 + 1 ∪ 2 AI := AI A Ak, I IL and, defining the length of AI to be the some of the { | ∈ ∈ } lengthes of A and I, we define similarly AI(j), AI0, AI1 and AI2. Hence any (even) 0 L p k function f ∞(Ω R | ) (where Ω is an open subset of R | ) can be decomposed as f = ∈ C θAηI×f , where f ( Ω ), AI AI . AI AI0 AI AI ∞ 0 ∈ ∈ C | | ∀ ∈ Then Theorem 3.1 implies that for any morphism Φ∗ from ∞( ) to ∞(Ω R0 L)0, thereP exists a smooth map ϕ ( Ω , ) and a smooth familyC N (ξ )C of× | ∞ AI AI AI2 ∈ C | | N ∈ sections of π∗T defined on a neighbourhood of the graph of ϕ in Ω such that if Ξ := Nξ θAηI then f ( ), φ f = (1 f) eΞf .| We|×N decompose Ξ AI AI2 AI ∞ ∗ ∗ as ∈ ∀ ∈ C N × P A a a1 a2  Ξ= θ ΞA = Ξ + θ Ξa + θ θ Ξa1a2 + , ∅ k k k ··· A A a A (1) (a1,a2) A (2) X∈ ∈X X∈ k I k I where A A , ΞA = IL ξAI η and A A , ΞA = IL ξAIη . In particular 1 I 1 0 I 2 ∀ ∈ I ∈ ∀ ∈ ∈ Ξ = I IL ξ Iη and we see that ΞA is odd if A is odd and is even if A is even. ∅ 2 ∅ P P Recall that∈ the vector fields (ξ ) can be chosen in such a way that they AI AI AI2 P ∈ commute pairwise. If so the relations [ξAI ,ξA′I′ ] = 0 imply that the vector fields ΞA supercommute pairwise, i.e.

k k jj′ A A (j), A′ A (j′), Ξ Ξ ′ ( 1) Ξ ′ Ξ =0. ∀ ∈ ∀ ∈ A A − − A A k A A′ This is equivalent to the fact that A, A′ A , [θ ΞA, θ ΞA′ ] = 0. This last commutation relation implies that ∀ ∈

Ξ P θAΞ Ξ θAΞ e = e A∈Ak A = e ∅ e A . A Ak Y∈ + Hence

Ξ θAΞ f ∞( ), φ∗f = (1 ϕ)∗ e ∅ e A f . (21) ∀ ∈ C N ×   A Ak Y∈ +   We thus obtain a generalization of (16) and (19). We could alternatively write (21) by the symbolic relation A Φ= ϕ eΞ∅ eθ ΞA , (22) A Ak Y∈ + which generalizes (17) and (20).

17 m k 0 L n ℓ 4.2.2 Even maps from | R | to a supermanifold | M × N m k m k First assume that | = Ω is an open subset of R | . We also suppose that Ω M is sufficiently small, so that its image by Φ is contained in some open subset U n ℓ n ℓ of | on which there exists a local chart Y : U R | . By this we assume N −→ that Y ∗ : ∞(Y (U)) ∞(U) is an isomorphism of super algebras. Hence for C −→ C any function f ∞(U), there exists an unique function F ∞(Y (U)) such that ∈ C 1 n 1 ℓ ∈ C f = Y ∗F . Denote by (z,ζ)=(z , , z ,ζ , ,ζ ) the canonical coordinates on Rn ℓ ··· ··· J | Y (U). Using the decomposition F = J Iℓ FJ ζ and the fact that Y is a morphism⊃ we deduce that ∈ P

J j1 j f = Y ∗F = f ψ = f ψ ψ k , J J ··· J Iℓ J Iℓ X∈ X∈ j j ℓ where fJ := Y ∗FJ , ψ := Y ∗ζ and we recall that I is the set of multi-indices r J =(j1, , jr) [[1,ℓ]] such that j1 < < jr, for 0 r ℓ. Observe that fJ is a smooth real··· valued∈ function defined on ···U of Rn. Then≤ by≤ the morphism property, | |

J j1 jr Φ∗f = (Φ∗f )(Φ∗ψ )= (Φ∗f )(Φ∗ψ ) (Φ∗ψ ). J J ··· J Iℓ J Iℓ X∈ X∈

Hence we conclude that the morphism Φ∗ is completely determined as soon as we j ℓ know all Φ∗ψ ’s, for j [[1, m]], and all pull-back images Φ∗fJ , for J I . However j j ∈ ∈ each χ := Φ∗ψ is an odd function on Ω which can be choosed arbitrarily and, since each fJ is a function on U , its pull-back by Φ is again described by Theorem 3.1. | |⊂|N| m k 0 L n ℓ m k The case of an even map Φ from | R | to | , where | is a su- M × N M permanifold, can be treated similarly: by taking the restriction of Φ to an open m k 0 L subset of | times R | and using a local chart on it, we are left to the previous M situation.

4.3 Application: how to read Φ = x + θψ ?

1 1 We have claimed in 4.1.1 that, when writing Φ = x + θψ for a map from R | 0 2 § 1 2 × R | to an ordinary manifold , one should read ‘ϕ + ϕ η η ζ’ instead of ‘x’ and 1 2 N ‘ϕ (η ξ1 + η ξ2)’ instead of ‘ψ’. By the results expounded in the previous section 1 1 0 L the generalization of this to a map Φ from R | R | to (where L is arbitrary) × N is (using (θ)2 = 0)

Ξ∅ θΞ1 Ξ∅ Ξ∅ Ξ∅ Φ= ϕ e e = ϕ e (1 + θΞ1)= ϕ e + θ ϕ e Ξ1 ,

Ξ∅ Ξ∅  so that one should interpret that ‘x’ = ϕ e and ‘ψ’ = ϕ e Ξ1. We can interpret similarly the decomposition of a supersymmetric σ-model

1 2 1 2 Φ= φ + θ ψ1 + θ ψ2 + θ θ F (23)

m 2 from R | to an ordinary manifold . Understanding this field as an even map Φ m 2 0 L N from R | R | to , we can write by Theorem 3.1 × N

Ξ∅ 1 2 1 2 Φ= ϕ e 1+ θ Ξ1 1+ θ Ξ2 1+ θ θ Ξ12 , (24)    18 meaning that f ∞( ), ∀ ∈ C N

Ξ∅ 1 2 1 2 Φ∗f = ϕ∗ e 1+ θ Ξ1 1+ θ Ξ2 1+ θ θ Ξ12 f . (25)

We can now understand the meaning of the usual expression (23) as follows. Some- how in this relation one implicitely assumes that ϕ takes values in an open set V of on which a coordinate chart y = (y1, ,yn) : V Rn is defined. Then N ··· −→ i the right hand side of (23) is y∗Φ. So let us use (25) with f = y (1 i n) and ≤ ≤ exploit the fact that, thanks to Theorem 3.1, we can chose all vector fields ξI such i i i i i i i i that ξI ξJ y = 0. Then, setting ϕ := y ϕ, Ξ1 := Ξ1ϕ , Ξ2 := Ξ2ϕ , Ξ12 := Ξ12ϕ and φi := ϕi + Ξ ϕi, (25) implies ◦ ∅ i i i i 1 i 2 i 1 2 i Φ Φ∗y = (ϕ + Ξ ϕ )+ θ (Ξ1ϕ )+ θ (Ξ2ϕ )+ θ θ (Ξ12ϕ ) ≃ i 1 ∅ i 2 i 1 2 i (26) = φ + θ Ξ1 + θ Ξ2 + θ θ Ξ12.

i i i i i i i i i A comparison with (23) gives us: φ = ϕ + Ξ ϕ , ψ1 = Ξ1, ψ2 = Ξ2 and F = Ξ12. ∅ One the other hand, for an arbitrary non linear function f ∞( ), (25) developps as ∈ C N Ξ 1 2 1 2 f Φ Φ∗f = ϕ∗ e ∅ f + θ Ξ f + θ Ξ f + θ θ (Ξ Ξ Ξ )f . ◦ ≃ 1 2 12 − 1 2 We hence deduce the relation in local coordinates  n ∂f n ∂f f(φ + θ1ψ + θ2ψ + θ1θ2F ) = f(φ)+ θ1 (φ)ψi + θ2 (φ)ψi 1 2 ∂yi 1 ∂yi 2 i=1 i=1 n X n X ∂f ∂2f + θ1θ2 (φ)F i (φ)ψi ψj . ∂yi − ∂yi∂yj 1 2 i=1 i,j=1 ! X X 4.4 Vector fields on a supermanifold with flesh

m k 0 L 1 m 1 k Consider the R | R | and denote by x , , x , θ , , θ the coordi- m k 1 ×L 0···L ··· nates on R | and by η , , η the coordinates on R | . For 1 i m, the action ∂ Rm ··· ≤ ≤ Rm k R0 L of ∂xi on ∞( ) can be extended in a straightforward way to ∞( | | ): C Rm k R0 L CA I × for any function f ∞( | | ), decompose f = AI AI fAI θ η (see 4.2.1), k ∈ C L × m ∈ § where, A A , I I , fAI ∞(R ) and just set ∀ ∈ ∀ ∈ ∈ C P ∂f ∂f := AI θAηI . (27) ∂xi ∂xi AI AI X∈ ∂ m k 0 L We define the odd derivative as the operator acting on ∞(R | R | ) by the ∂θa C × rule ∂f ∂θA = ( 1)gr(fAI )f ηI, ∂θj − AI ∂θa AI AI X∈ ∂θA where, if A does not contain a, ∂θa = 0 and, if A contains a, say A =(a1, , aγ 1, a, aγ+1 , ak), A ··· − ··· ∂θ γ 1 1 aγ−1 aγ+1 ak Rm k ∂θa =( 1) − θ θ θ θ . More generally, if Ω is an open subset of | , − ··· ··· 0 L 0 L we consider differential operators : ∞(Ω R | ) ∞(Ω R | ) of the form D C × −→ C × m ∂ k ∂ = a + b , D i ∂xi a ∂θa i=1 a=1 X X 19 0 L where a and b belongs to ∞(Ω R | ). In the following we assume that the a ’s i a C × i are even and the ba’s are odd so that is a derivation, i.e. it satisfies the Leibniz D m k 0 L 1 1 0 L rule (fg)=( f)g + f( g), f,g ∞(R | R | ). For example on R | R | D D D ∀ ∈ C ∂ × R0 L 1 × we consider in Section 6 the vector field ∂t and, letting η ∞( | ) (η is odd), ∂ ∂ ∂ ∈ C∂ the (even) vector fields ηD = η ∂θ ηθ ∂t and ητQ = η ∂θ + ηθ ∂t . For the following discussion it− is useful to present an alternative description 0 L 0 m k of the set of even maps ∞(Ω R | ) , where Ω R | . Let q := k + L and 1 k 1 C L × 1 q ⊂ 2 replace (θ , , θ , η , , η ) by (η , , η ) to simplify the notations. Let Λ+∗ 2q−1 1 ··· ··· ··· ≃ R − be the vector subspace of even elements of positive degree of the exterior Rq I 2 algebra Λ∗ , denote by s = s Iq the canonical coordinates on Λ+∗ and define I 2 R 1 q ∈ 2 I ∂ the ‘canonical’ [η , , η ]-valued  vector field ϑ on Λ+∗ by ϑ := I Iq η ∂sI . Then, ··· ∈ 2 f R0 L 0 f 2 as shown in [14], for any ∞(Ω | ) , there exists a mapP ∞( Ω Λ+∗) such that x Ω , f(x, η)∈ = C eϑf×(x, 0), or, by using the canonical∈ C embedding| | × ∀ ∈ | 2| ι : Ω Ω Λ ∗, x (x, 0), | | −→| | × + 7−→  ϑ f = ι∗ e f .

Note that f is not unique: two maps f and f′ produces the same function f if and q only if their difference f f′ lies in the ideal ( Ω ) spanned over ∞( Ω ) by the −I1 Ij I |q | I1 Ij C | | polynomials sI1 sIj ǫ ··· sI , for I , , I I , where ǫ ··· = 0 if there exists ··· − I 1 ··· j ∈ 2 I some index i [[1, q]] which appears at least two times in the list (I1, , Ij) and I1 I ∈ I1 I ··· ǫ ··· j = 1 in the other cases (then ǫ ··· j is the signature of the permutation I ± I (I1, , Ij) I), see [14]. One hence deduces the algebra isomorphism ∞(Ω R0 L···0 7−→ 2 q C × | ) ∞( Ω Λ+∗)/ ( Ω ). Thanks≃ C to| the| × isomorphismI | | f f mod q( Ω ), we can represent each vector 0 L 0 7−→ I | | 2 field X acting on ∞(Ω R | ) by a (non unique) vector field X on Ω Λ ∗, such C × | | × + that 0 L 0 ϑ f ∞(Ω R | ) , Xf = ι∗ e Xf . (28) ∀ ∈ C × Below is a list of examples.  m ∂ 1. Any vector field X = i=1 Xi(x) ∂xi tangent to Ω has a straightforward 0 q | | extension to Ω R | , induced by (27), that we still denote by X. Similarly | | × P 2 X can be extended to Ω Λ+∗ into an unique vector field, also denoted I | | × ∂ R0 L 0 by X, such that Xs = 0 and [X, ∂sI ] = 0. Then f ∞(Ω | ) , ϑ ∀ ∈ C × Xf = ι∗ e Xf , i.e. we can choose X = X. 2. Let X be a vector field as in the previous example and consider Y = η1η2X. Then 0 q 0 1 2 ϑ 12 f ∞( Ω R | ) , η η Xf = ι∗ e s Xf , ∀ ∈ C | | × 12 I/1/2q i.e. we can choose Y = s X. This is a consequence of the following. Let 2 Iq be the subset of multi-indices in 2, where the values 1 and 2 are forbidden for the indices. For I I/1/2q, let 12I Iq be the multi-index obtained by ∈ 2 ∈ 2 concatening 1, 2 and I. Then s12sI = s12I mod q( Ω ). I | | 1 ∂ 1 ∂ 1I ∂ s /1/2q s 3. For Z = η ∂η2 , we have, setting ∗ ∂s2∗ := I I ∂s2I , ∗ ∈ 2 P P ∂f ∂f R0 q 0 1 ϑ s1 f ∞( Ω | ) , η 2 = ι∗ e ∗ 2 , ∀ ∈ C | | × ∂η ∂s ∗ ! X∗ 20 1 ∂ i.e. we can choose Z = s ∗ ∂s2∗ . ∗ We can now make sense ofP odd of even maps into an ordinary manifold in the sense of 4.2.1 as follows. We first present a consequence of Theorem 3.1 N § derived in [14]. Recall that, by Theorem 3.1, any morphism Φ∗ : ∞( ) 0 q 0 Ξ C N −→ ∞( Ω R | ) has the form Φ∗f = (1 ϕ)∗ e f , where, in the decomposition C | | × I × Ξ = I Iq ξIη , we can assume that all vector fields ξI’s commute. Hence we can ∈ 2  integrate simultaneously these vector fields and we obtain a smooth (ordinary) map P 2 F : Ω Λ ∗ such that: | | × + −→ N ∂F F(x, 0) = ϕ(x) and (x, s)= ξ (F(x, s)) . (29) ∂sI I

Conversely one can recover Φ∗ (and hence Φ) from F through the formula

ϑ f ∞( ), Φ∗f = ι∗ e (f F) . (30) ∀ ∈ C N ◦ 0 q Now if X is a tangent vector on Ω R | , we would like to define XΦ = Φ X as a | | × ∗ section of Φ∗T or alternatively (in order to avoid the definition of the fiber bundle N 0 q 0 q Φ∗T over Ω R | ) through the map (Φ,XΦ) from Ω R | to T . We do it N | | × | | × N indirectly by defining α (XΦ), for any section α of T ∗ on . We first consider Φ N N the case where α is exact, i.e. α = df, for some f ∞( ). We then have, by using (30) and (28) ∈ C N

ϑ ϑ ϑ df (XΦ) = X(f Φ) = X(Φ∗f)= Xι∗ e (f F) = ι∗ e X(f F) = ι∗ e df (XF) . Φ ◦ ◦ ◦ F (Note that eϑ and X commute). We deduce from this identity a similar formula for a 1-form gdf, where f,g ∞( ) by writing ∈ C N ϑ ϑ ϑ (g Φ)(df (XΦ)) = ι∗ e (g F) ι∗ e df (XF) = ι∗ e (g F)df (XF) . ◦ Φ ◦ F ◦ F And since any 1-form on T can be written as a finite sum of forms of the type gdf, we obtain: N ϑ αΦ(XΦ) = ι∗ e αF(XF) .

Hence XΦ can be represented by the section XF of F∗T or, equivalentely (Φ,XΦ) N 2 is a map from Ω to T which can be represented by the map (F, XF) from Ω Λ ∗ N | |× + to T . N 5 An alternative theory

5.1 Another consequence of the morphism property In this paragraph we still use the same point of view as before but, as a motivation for the following, we expound another consequence of the morphism properties (9) and m k (10). Assume that | is a (skeletal, for simplicity) supermanifold of dimension Mm k m k and let U | an open subset and X : U Ω be a chart, where Ω | ⊂ M m k 1 m −→1 k is an open subset of R | . Let (z,ζ)=(z , , z ,ζ , ,ζ ) be the canonical m k ··· ··· coordinates on R | Ω. As in 4.2.2, for any function f ∞(U) there exists ⊃ § ∈ C an unique function F ∞(Ω), such that f = X∗F and, if we decompose F = I ∈ C i i I Ik FI ζ , for some functions FI ∞( Ω ), and if we set θ := X∗ζ , we obtain ∈ ∈ C | | P 21 I f = X∗F = I Ik (X∗FI )θ . Moreover X∗FI can be estimated by the following: ∈ since FI ∞( Ω ), we can deduce from a Taylor expansion (with integral rest) that, for any z,∈ z C P|Ω|, 0 ∈| | ∂rF (z z )r F (z)= I (z ) − 0 + G (z)(z z )r, I (∂z)r 0 r! I,r − 0 r k r =k+1 |X|≤ | |X i i for some smooth functions GI,r. Using (9) and (10) and setting x := X∗z and i i x0 := X∗z0, this implies r r ∂ FI (x x0) r X∗F = (z ) − + G (y)(x x ) . I (∂z)r 0 r! I,r − 0 r k r =k+1 |X|≤ | |X

Now in the case where z z0 (i.e. x x0) is nilpotent, the last term in the right − − ∂ ∂ hand side vanishes and, setting fI := X∗FI and X i = i , we are left with ∗ ∂x ∂z ∂rf (x x )r f (x)= I (x ) − 0 if x x is nilpotent. (31) I (∂x)r 0 r! − 0 r k |X|≤ 5.2 The theory of DeWitt and Rogers Another theory of supermanifolds was developped by B. DeWitt and A. Rogers [23]. The model space here is m k 0 m 1 k R ′| := Λ ′ Λ ′ , (32) L L × L R0 L′ RL′ N where ΛL′ is a superalgebra isomorphic to ∞( | ) Λ∗ and L′ . (We may C ≃ ∈ choose L′ = , but in the following we assume that L′ is finite for simplicity.) We de- ∞ fine the body map β :ΛL′ R which, to each z ΛL′ , associates the only real num- ber β(z) such that z β(z−→) is nilpotent (we also call∈ σ(z) := z β(z) the soul map). − Rm k Rm 1 − m 1 k We extend β to a map β : L′| which, to each (z , , z ,ζ , ,ζ ) Rm k 1 m −→ 1 k 1 ··· m ··· ∈ L′| , associates (β(z ), , β(z ), β(ζ ), , β(ζ )) (β(z ), , β(z )). Of course, as a set, the right hand··· side in the definition··· (32) looks≃ too la···rge. Its too large di- mension is compensated by Rm k the choice of a (non Hausdorff) topology on it: open subsets of L′| are of • 0 m 1 k m 0 the form Ω (Λ∗ ′ ) (Λ ′ ) , where Ω is an open subset of R and Λ∗ ′ | | × L × L | | L 0 Rm k is the subset of nilpotent elements of ΛL′ . Alternatively open subsets of L′| are inverse images of open subsets of Rm by β.

m k a choice of restrictions on the class of functions on R ′| we are working with. • L Rm k We are going to present classes of functions on L′| which depend only on what is Rm Rm k happening in a formal neighbourhood of inside L′| . These classes of functions will be built starting from the coordinate functions

i Rm k x : L′| ΛL′ (z1, , zm,ζ1, ,ζk) −→ zi ··· ··· 7−→ and i Rm k θ : L′| ΛL′ (z1, , zm,ζ1, ,ζk) −→ ζi. ··· ··· 7−→ 22 We observe that we should at least suppose that L′ k in order to ensure that θ1 θk = 0. ≥ ··· 6 m k The class H∞ — For any open subset Ω R ′| , a function F : Ω Λ ′ belongs ⊂ L −→ L to H∞(Ω) if I F = θ fI (x), (33) I Ik X∈ I i1 ij b k where θ := θ θ , for I = (i1, , ij) and, for each I I , there exists a (real valued !) function···f : β(Ω) R ···such that, for any z Ω,∈ I −→ ∈ ∂rf (β(z)) (z β(z))r f (z) := I − . (34) I (∂β(z))r r! r=(r1, ,rm) X··· b Note that the right hand side is a finite sum, since the components of z β(z) are nilpotent. One can motivate conditions (33) and (34) by Relation (31) obtained− in 5.1. § However the theory of H∞-functions is limited and its use for physical models leads to the same type of inconsistencies as the theory of skeletal algebro-geometric supermanifold.

The class G∞ — Functions f G∞(Ω) can be defined in exactly the same way as ∈ functions in H∞(Ω) with the exception that, in (34), the function fI takes values in ΛL′ instead of R. Hence the set G∞(Ω) is much larger than H∞(Ω) and allows to perform many computations on physical models that are forbidden within H∞(Ω).

The class GH∞ — Since however G∞(Ω) may be too big, one can use as a variant the set GH∞(Ω) of maps where one assumes that the functions fJ in (34) takes values in ΛL, for L < L′: it can be shown that, for many purposes is is enough to choose the integers L and L′ such that 2k < L′ 2L (see [25]). The resulting theory is similar to the algebro-geometric theory of supermanifol≤ ds with flesh.

Comparisons — Each of these notions of superfunctions leads to a particular dif- Rm k ferent notion of supermanifolds, by gluing open subsets of L′| though H∞-, G∞- or GH∞-diffeomorphisms, see [25]. Again most of the ‘physical’ applications requires the use of G∞- or GH∞-supermanifolds. One can compare the various theories as follows:

1. any H∞-supermanifold is a G∞- or a GH∞-supermanifold, a consequence of the trivial facts that any H∞-function is a GH∞-function, and any GH∞- function is a G∞-function;

2. conversely any G∞-supermanifold can be endowed with a (non unique) H∞- superstructure, a consequence of 3. and 4. below;

3. any G∞-supermanifold has a split structure (i.e. is of the form S( , E), where E is a vector bundle over ) [4]; |M| |M| 4. any split manifold S( , E) has a H∞-structure; |M| 5. any H∞-supermanifold is a (skeletal) algebro-geometric supermanifold [25];

23 6. conversely any (skeletal) algebro-geometric supermanifold is a H∞-supermanifold [5]. See [25] (and [3] about 5 and 6) for a complete overview.

6 The superparticle: second version

We now present the superparticle model using the superfield Φ = x + θψ, a map 1 1 from R | to an ordinary Riemannian manifold . We need for that purpose to N know how to integrate a function a supermanifold.

6.1 The Berezin integral The theory for integrating a function on a supermanifold is due to Berezin. In the m k following we just explain how to define such an integral on an open subset of R | . We do not expound the complete theory, which rests on a delicate change of variable formula. We refer to [18, 10, 25] for a complete exposition. m k Let Ω be an open subset of R | and f ∞(Ω) be a smooth function on the I ∈ C R0 L Ik closure of Ω. Let us write f = I Ik θ fI , where fI ∞( Ω ) ∞( | ), I , for some L N. Then we define the∈ Berezin integral∈ of C f over| | ⊗ Ω C by ∀ ∈ ∈ P 1 m 1 k 1 m dx dx dθ dθ f := dx dx f1 k. Ω ··· ··· Ω ··· ··· Z Z| |

So we have selected the highest degree coefficient f1 k in the decomposition of f ··· 0 L and we just integrate it over Ω . Note that this integral takes values in ∞(R | ). | | C It also enjoys the following property

dx1 dxmdθ1 dθkf(x, η + ζ)= dx1 dxmdθ1 dθkf(x, η), (35) ··· ··· ··· ··· ZΩ ZΩ which reflects the invariance by translation in the odd variables of the Berezin in- tegral. Indeed (35) just follows by expanding (η1 + ζ1) (ηk + ζk) and using the definition of the Berezin integral. The ‘infinitesimal’ vers···ion of (35) is: ∂f j [[1,k]], dx1 dxmdθ1 dθk =0. (36) ∀ ∈ ··· ··· ∂θj ZΩ ∂ This can be shown by observing that the action of ∂θj decreases by one the degree of I ∂ 1 k j 1 1 j 1 j+1 k all monomials θ . For instance ∂θj (θ θ )=( 1) − θ θ − θ θ , so that dx1 dxmdθ1 dθk ∂ (θ1 θk)=0.··· − ··· ··· Ω ··· ··· ∂θj ··· R 1 1 6.2 The supertime R | 1 1 The simplest non trivial example of a superspace is the ‘supertime’ R | . We denote by t (the time) the even coordinate and by θ the odd coordinate on it. We will R1 1 ∂ consider three vector fields on | : the time derivative ∂t acting on even functions 1 1 0 L on R | R | which is the extension of the usual operator as described in 4.4 and: × § ∂ ∂ ∂ ∂ D := θ and τ := + θ . ∂θ − ∂t Q ∂θ ∂t 24 Clearly D and τQ are odd operators, i.e. they change the parity of the maps on 1 1 0 L which they act. So in particular, if Φ : R | R | is an even map, DΦ and × −→ N τQΦ are not even. There are then two ways to deal with these derived maps: either we introduce the supermanifold ΠT and we can make sense of these derivatives N directly. Or we allow only to consider ηDΦ and ητQΦ, where η is odd, and we are then in the setting expounded in 4.4 (i.e. we implicitely assume that we are 1 1 0 L § working with maps on R | R | ). Note that D (resp. τQ) can be considered as a × 1 1 left (resp. right) invariant odd vector field on R | (we shall see why in Section 8). ∂ Operators ∂t , D and τQ satisfy the following relations: ∂ ∂ D2 = , τ 2 = , Dτ + τ D =0, −∂t Q ∂t Q Q ∂ ∂ ∂ ∂ D D = τ τ =0. ∂t − ∂t Q ∂t − ∂t Q For example

D2(f(t)+ θg(t)) = Dd g(t) θf˙(t) = f˙(t) θg˙(t). − − −   These relations can be summarized together by introducing the supercommutator [ , ] of two homogeneous operators A and B: if A has parity gr(A)(=0if A is even, · · =1 if A is odd) and B has parity gr(B), then

[A, B] := A B ( 1)gr(A)gr(B)B A. ◦ − − ◦ We thus have ∂ ∂ ∂ ∂ [D,D]= 2 , [τ , τ ]=2 , [D, τ ]= D, = τ , =0. − ∂t Q Q ∂t Q ∂t Q ∂t     1 1 ∂ We see that the space m | ‘spanned’ by ∂t , D and τQ is a Lie super algebra, the 1 1 super Lie algebra of supertranslations of R | . Note that, as pointed out in 1.2.2, § Lie super brackets can be replaced by ordinary Lie brackets, which are easier to 1 1 0 L 1 L handle with, by tensoring m | with ∞(R | )= R[η η ]. One then manipulates ∂ C ··· R 1 L only linear combinations of ∂t , D and τQ with coefficients in [η η ] which are 1 2 ··· even. For instance η D and η τQ are even and their commutator is:

η1D, η2τ = η1D η2τ η2τ η1D = η1η2 (Dτ + τ D)= 2η1η2 [D, τ ] , Q Q − Q − Q Q − Q

which is equal to 0 because  the anticommutator  [D, τQ] vanishes. 1 1 ∂ Note also that if we consider m | R[θ], the space spanned by , D and τ ⊗ ∂t Q over R[θ], then these three generators are not linearly independant, because of the relation τ D =2θ ∂ . Q − ∂t 6.3 The superfield formulation of the superparticle model Now we come back to the superparticle: instead of working on fields (x, ψ) we 1 1 0 L consider even maps Φ = x + θψ from R | R | to , with the meaning discussed × N in 4.3 (i.e. Φ = x(1 + θψ)). Recall that the vector field Q = ητ was defined by § − Q Q(x, ψ)=( ηψ,ηx˙). − 25 It is straightforward to check that this coincides with the action of ητ on Φ since: − Q ∂ ∂ ∂x ∂x ητ (Φ) = η + θ (x + θψ)= η ψ + θ = ηψ θ η . Q ∂θ ∂t ∂t − ∂t       Furthermore the action [x, ψ] can be written in terms of the field Φ as follows. Consider the following action,L given by a Berezin integral (see 6.1): § 1 ∂Φ [Φ] := dtdθ DΦ, . L −2 R1|1 ∂t Z Z   Through the substitution Φ = x + θψ one finds 1 ∂Φ 1 1 θ DΦ, = ψ θx,˙ x˙ + θ ψ = ψ, x˙ + x˙ 2 + ψ, ψ . (37) −2 ∂t −2 h − ∇x˙ i −2 h i 2 | | h ∇x˙ i    Thus we recover

1 2 [Φ] = dt x˙ + ψ, x˙ ψ = [x, ψ], L 2 R | | h ∇ i L Z  The superfield formulation allows an alternative proof of the invariance of [Φ] under the action of Q. For that purpose we need to compute [Φ ητ Φ]. FirstL we have L − Q (using the fact that η anticommutes with D):

∂ ∂Φ ∂Φ ∂∇ D(Φ ητ Φ), (Φ ητ Φ) = DΦ, +η D∇τ Φ, +η DΦ, τ Φ , − Q ∂t − Q ∂t Q ∂t ∂t Q         ∂∇ where D∇ and ∂t denote the covariant derivatives for the pull-back of the Levi- Civita connection on by Φ (or by F, see 4.4). Now we use the anticommutation relation [D, τ ] = 0, theN commutation relation§ ∂ , τ = 0 and the fact that is Q ∂t Q ∇ torsion free to get:   ∂ D(Φ ητ Φ), (Φ ητ Φ) − Q ∂t − Q   ∂Φ ∂Φ ∂ = DΦ, + η τ ∇DΦ, + η DΦ, τ ∇ Φ ∂t − Q ∂t Q ∂t  ∂Φ  ∂Φ    = DΦ, ητ DΦ, . ∂t − Q ∂t     We hence deduce an expression for [Φ ητQΦ]. By a comparison with the relation [Φ ητ Φ] = [Φ] ητ δ [Φ],L we− deduce L − Q L − Q L 1 ∂Φ ητQ δ [Φ] = dtdθητQ DΦ, . L −2 R1|1 ∂t Z Z   At this point we can conclude that ητ induces a symmetry since ητ δ [Φ] − Q Q L reduces to a boundary integral and because of (36). However it is interesting to compute explicitely this term in view of obtaining the Noether charge of this sym- ∂Φ metry. We observe that if we note DΦ, ∂t = f + θg, then using the fact that f is odd (see (37)),

∂f ∂f dtdθητQ(f + θg)= dtdθ ηg θη = dt η . R1|1 R1|1 − ∂t − R ∂t Z Z Z Z   Z 26 Hence, in view of (37), 1 ∂ ητQ δ [Φ] = η dt ψ, x˙ , (38) L 2 R ∂t h i Z and we obtain the same result as in Section 2. We now recover Noether’s theorem. It relies on redoing the preceding calculation but ητQ is now replaced by a smooth compactly supported ‘modulation’ of it, i.e. we compute χητ δ [Φ], where χ Q L ∈ ∞(R). It gives us: Cc (0) (1) [Φ χητQΦ] = [Φ] δ [Φ] δ [Φ], L − L − χητQ − χητQ (0) where we have splitted χητQ δ [Φ] into the sum of δχητQ [Φ], which is a linear (1) L dχ (0) expression in χ, and of δχητQ [Φ], which is a linear expression in dt . However δχητQ [Φ] is simply obtained by inserting χ in the right hand side of (38), i.e.

(0) 1 ∂ δχητQ [Φ] = η dtχ ψ, x˙ . 2 R ∂t h i Z On the other hand

(1) 1 ∂Φ dχ δχητQ [Φ] = dtdθ Dχ ητQΦ, + DΦ, ητQΦ . 2 R ∂t dt h i Z     ∂Φ dχ We compute then Dχ ητQΦ, ∂t = θη dt ψ, x˙ and DΦ, ητQΦ = 2θη ψ, x˙ , so that − h i h i − h i (1) 3 dχ δχητQ [Φ] = dt η ψ, x˙ . 2 R dt h i Z By integrating by part the last expression we then conclude that χητQ δ [Φ] = η dtχ ∂ ψ, x˙ , as in Section 2. L − R ∂t h i R 3 2 7 Superfields on R |

(1,2) 2 3 2 7.1 The M | R | ≃ 3 2 Another example of superspace is R | , a supersymmetric extension of the Minkowski space of dimension 1 + 2. We denote the even coordinates by t, x and y and the odd coordinates by θ1 and θ2. Beside the translation vector fields ∂/∂t, ∂/∂x and ∂/∂y, we define the (right invariant) supertranslation generators ∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂ τ := + θ1 + + θ2 , τ := + θ1 + θ2 , (39) 1 ∂θ1 ∂t ∂x ∂y 2 ∂θ2 ∂y ∂t − ∂x     which satisfy the supercommutation relations ∂ ∂ ∂ ∂ ∂ [τ , τ ]=2 + , [τ , τ ]=2 , [τ , τ ]=2 . (40) 1 1 ∂t ∂x 1 2 ∂y 2 2 ∂t − ∂x     It may be convenient to embedd the Minkowski space R3 in the space M(2, R) of 2 2 real matrices with the coordinates v11, v12, v21 and v22 though the map × R3 M(2, R) −→ v11 v12 (t + x)/2 y/2 (t, x, y) v = = , 7−→ v21 v22 y/2 (t x)/2    −  27 3 12 21 i.e. to view R as the hyperplane v v = 0 or the space h2(R) of symmetric 2 2 real matrices. We then recover the− Minkowski norm by the relation t2 x2 y2×= 4 det v. Functions on R3 can then be identified with functions on M−(2, R−) annihilated by ∂ ∂ and we have ∂v12 − ∂v21 ∂ 1 ∂ ∂ ∂ 1 ∂ ∂ ∂ 1 ∂ ∂ = + , = , = + . ∂t 2 ∂v11 ∂v22 ∂x 2 ∂v11 − ∂v22 ∂y 2 ∂v12 ∂v21       ∂ ∂ Then, setting ∂a := ∂θa and ∂ab := ∂vab , (39) can be written

b τa := ∂a + θ ∂ab, for a, b =1, 2.

Similarly (40) becomes [τa, τb]=2∂ab. We define also the (left invariant) odd vector fields

D := ∂ θb∂ , for a, b =1, 2. (41) a a − ab

which satisfy the supercommutation relations [Da, τb] = 0 and

[D ,D ]= 2∂ . (42) a b − ab 7.2 An example of supersymmetric model As an elementary example of a supersymmetric model we consider ‘scalar’ fields Φ 3 2 on R | which are critical points of the functional

3 2 1 ab [Φ] := d xd θ ǫ DaΦDbΦ + Φ∗h , A R3|2 4 Z   3 2 1 2 α where d x := dtdxdy, d θ := dθ dθ and h ∞(R, R). We note Φ = φ + θ ψα + 1 α β 1 2 1 2 ∈ C 2 ǫαβθ θ F = φ + θ ψ1 + θ ψ2 + θ θ F . Then, by setting ( ψ) = ∂ ψ ∂ ψ and ( ψ) = ∂ ψ ∂ ψ , D 1 12 1 − 11 2 D 2 22 1 − 12 2 we have D Φ= ψ θ1∂ φ + θ2 (F ∂ φ)+ θ1θ2( ψ) , 1 1 − 11 − 21 D 1 D Φ= ψ θ1 (F + ∂ φ) θ2∂ φ + θ1θ2( ψ) . 2 2 − 12 − 22 D 2 We hence find that, denoting ψ ψ := ǫabψ ( ψ) = ψ ( ψ) ψ ( ψ) , D a D b 1 D 2 − 2 D 1 1 1 ∂φ ∂φ 1 ∂φ 1 1 d2θ ǫabD Φ D Φ = 2+ ψ ψ + F 2 4 a · b 2 ∂y11 ∂y22 − 2 ∂y12 2 D 2 Z   2  2  2 1 ∂φ 1 ∂φ 1 ∂φ 1 1 = + ψ ψ + F 2. 2 ∂t − 2 ∂x − 2 ∂y 2 D 2      

The computation of Φ∗h can be done as in 4.3: it is better to understand the 1 2§ 1 2 decomposition of Φ as Φ = φ(1 + θ ψ1)(1 + θ ψ2)(1 + θ θ F ), where ψ1, ψ2 and F are derivation operators. This gives us

1 2 1 2 1 2 Φ∗h = h φ +(h′ φ) θ ψ + θ ψ + θ θ F (h′′ φ)θ θ ψ ψ , ◦ ◦ 1 2 − ◦ 1 2  28 so that 1 2 dθ dθ Φ∗h =(h′ φ)F (h′′ φ)ψ ψ . ◦ − ◦ 1 2 Z Hence we conclude that

2 3 1 ∂φ 1 2 1 1 2 [Φ] = d x φ + ψ ψ h′′(φ)ψ1ψ2 + F + h′(φ)F , A R3 2 ∂t − 2|∇ | 2 D − 2 Z "   #

2 2 where φ 2 := ∂φ + ∂φ . Another way to write this action is |∇ | ∂x ∂y    2 1 3 ∂φ 2 2 [Φ] = d x φ + ψ ψ 2h′′(φ)ψ1ψ2 (h′(φ)) + [Φ], A 2 R3 ∂t − |∇ | D − − Q Z "  #

1 3 2 where [Φ] := 2 R3 d x(F +h′(φ)) . This makes clear how to eliminate the auxiliary field FQ. The Euler–Lagrange system of equations is R

φ + h′′(φ)h′(φ)+ h′′′(φ)ψ1ψ2 = 0 ψ (h′′ φ)ψ = 0 (43)  D − ◦  F + h′(φ) =0,

2 2 2 where  := ∂ ∂ ∂ . We see that the equation on F is not coupled with (∂t)2 − (∂x)2 − (∂y)2 the other equations (F has no influence on φ et ψ).

7.3 Bogomolny’i–Prasad–Sommerfeld solutions

α 1 α β It is interesting to look at fields Φ = φ + θ ψα + 2 ǫαβθ θ F which satisfies the condition (cos ατ + sin ατ ) Φ = 0 1 2 · ψ = 0 (44)   F + h′(φ) =0, where α R. ∈  Proposition 7.1 Any super field Φ which satisfies (44) is a solution of the Euler– Lagrange system (43).

Proof — The first condition in (44) gives us (by taking into account the fact that ψ = 0) ∂φ ∂φ ∂φ θ1 cos α + + sin α F ∂t ∂x ∂y −   ∂φ ∂φ  ∂φ  +θ2 sin α + cos α + F = 0. ∂t − ∂x ∂y      So that Φ satisfies: ∂φ ∂φ ∂φ cos α + cos α + sin α = sin αF ∂t ∂x ∂y  ∂φ ∂φ ∂φ  sin α sin α + cos α = cos αF.  ∂t − ∂x ∂y −  29 ∂ ∂ ∂ ∂ Denote X := cos 2α ∂x + sin 2α ∂y and Y := sin 2α ∂x + cos2α ∂y . By multiplying the first equation by cos α and the second one− by sin α, we obtain ∂φ + Xφ =0. (45) ∂t By multiplying the first equation by sin α and the second one by cos α, we obtain − Yφ = F which, taking into account the relation F + h′(φ) = 0, gives us −

Yφ = h′(φ). (46)

2 Then on the one hand equation (45) implies ∂ X2 φ = ∂ X ∂ + X φ = (∂t)2 − ∂t − ∂t 0, or equivalentely     ∂2φ ∂2φ ∂2φ ∂2φ cos2 2α 2cos2α sin 2α sin2 2α =0. (47) (∂t)2 − (∂x)2 − ∂x∂y − (∂y)2

2 On the other hand equation (46) gives Y φ = Y (h′ φ)=(h′′ φ)Yφ =(h′′ φ)(h′ φ), ◦ ◦ ◦ ◦ or equivalentely

2 2 2 2 ∂ φ ∂ φ 2 ∂ φ sin 2α 2cos2α sin 2α + cos 2α = h′′(φ)h′(φ). (48) (∂x)2 − ∂x∂y (∂y)2

By substracting (47) and (48) we find that

φ + h′′(φ)h′(φ)=0.

∂φ ∂φ Particularly interesting cases are: (a) if α = π/4, then we find ∂t + ∂y = 0 and ∂φ ∂φ ∂φ ∂φ ∂x = h′(φ); (b) if α = π/4, then we find ∂t ∂y = 0 and ∂x = h′(φ). If we − − ∂φ ∂φ − further impose the condition ∂t = ∂y = 0 we then find the so-called Bogomolny’i solutions: they are stationary solutions which minimizes the energy

2 2 1 ∂φ ∂φ 2 E[φ] := dx + +(h′ φ) 2 R ∂t ∂x ◦ Z "    # among stationary fields with the same asymptotic data at infinity, a consequence of the identity

1 ∂φ 2 ∂φ 2 ∂ E[φ]= dx + h′(φ) 2 (h φ) . 2 R ∂t ∂x ∓ ± ∂x ◦ Z "    # Moreover the energy is equal to the ‘topological charge’ [h(φ( )) h(φ( ))]. The interest of these solutions is also that, by imposing± a invariance∞ − condition−∞ on Φ, we necessarily obtain a dynamical solution. However the condition of being invariant by a supersymmetry translation is stable by quantization, i.e. it makes sense to speak of quantum states which are invariant by such a supertranslation. Such states are called BPS states, after Bogomolny’i, Prasad and Sommerfeld.

30 8 Other super Minkowski spaces

We present here the super Minkowski spaces (with minimal supersymmetry) and the associated super Poincar´egroups for the dimensions 1 +2, 1+3, 1+5 and 1+9: they admit particularly beautiful representations since each of these dimensions is associated with respectively: R, C, the quaternions H and the octonions (or Cayley numbers) O. They are the only normed division algebras and have increasing dimensions and complexity: R and C are commutative fields, H is a non commutative field and O is not associative nor commutative. A good presentation of octonions can be found in [2] (see also 8.5). In the following we set K = R, C, H or O and § k := dim K.

8.1 The Lorentz group in dimensions 3, 4, 6 and 10 We first observe that the Minkowski space R1,1+k can be identified with R1,1 K × which, itself, can be identified with the space h2(K) of Hermitian 2 2 matrices over K by × h : R1,1 K h (K) × −→ 2 t + x z (49) (t, x, z) 1 . 7−→ 2 z t x  −  The Minkowski scalar product on R1+1 K has a simple expression by using h, since × t2 x2 z 2 = 4 det h(t, x, z). − −| | This leads to the identification Spin(1, 1+ k) = SL(2, K). For k = 1 or 2 (i.e. respectively K = R or K = C), this is easy to see. Indeed, if K = R or C, it is straightforward to check that, for any g SL(2, K), ∈

h2(K) Rg : h (K) h (K) 2 −→ 2 (50) m gmg†, 7−→ where g := gt, makes sense and is a linear transformation which preserves the † h2(K) Minkowski norm det m. Moreover the map g Rg is a 2:1 group morphism 7−→ which is the Spin covering of SO0(1, 1+k). The advantage of this construction is that it comes naturally with the spinor representation: the natural action of SL(2, K) on K2. The cases where K = H or O are similar but require some care, since there are no direct definitions of SL(2, H) and SL(2, O). Let us consider first the situation at the level of Lie algebras. We denote by tf(2, K) the vector space of trace free 2 2 matrices over K. We naively expect that sl(2, K), the Lie algebra of SL(2, K), × coincides with tf(2, K) and it is so for K = R or C but not for K = H and O. 10 11 Moreover dimRtf(2, H)= 12 < 15 = dim Spin(1, 5) = dimso(1, 5) and, similarly dimRtf(2, O) = 24 < 45 = dim Spin(1, 9) = dimso(1, 9). Actually, for K = H or O we need to define sl(2, K), the Lie algebra of SL(2, K), as the Lie algebra of endomorphisms over R of K2 spanned by the action of tf(2, K) or, equivalentely, to

10The difference between the dimensions is 3 = dimso(3) = dimso(Im H), see below. 11The difference between the dimensions is 21 = dimso(7) = dimso(Im O), see below.

31 define sl(2, K) as the Lie algebra of real endomorphisms of h2(K) spanned by linear maps h2(K) ρσ : h (K) h (K) 2 −→ 2 m 1 (σm + mσt) , 7−→ 2 K K h2( ) 1 for all values of σ tf(2, ). (Note that σ ρσ is ( 2 times) the linearization h2(K) ∈ h2(K) 7−→ h2(K) of g Rg at g = 1.) So let ρ := ρσ σ tf(2, K) EndR(h (K)) 7−→ tf(2,K) { | ∈ } ⊂ 2 and let us define sl(2, K) to be the Lie subalgebra of (EndR(h2(K)), [ , ]) spanned by K h2( ) K · · ρtf(2,K). Then we claim that sl(2, ) coincides with so(1, 1+ k). This can be checked as follows. Let (u , ,u ) be an orthonormal basis of Im K. Then a basis of the 2 ··· k Minkowski space h2(K) is (e 1, , ek), where − ···

1 0 1 0 0 1 0 uj e 1 := , e0 := , e1 := , ej := − 0 1 0 1 1 0 u 0    −     − j 

and where 2 j k. Note that (in conflict with the usual convention) e 1 − is time-like, whereas≤ ≤ the other vectors are space-like. Then, for 0 j k, ≤ ≤ we define the infinitesimal boost Bj to be the endomorphism of h2(K) such that (e 1, ej) (ej, e 1) (all the other vectors are maps to 0) and, for 0 i < j k, − − the infinitesimal7−→ rotation A to be defined by (e , e ) (e , e )≤ (idem).≤ Re- ij i j 7−→ j − i call that ((Bj)0 j k, (Aij)0 i

8.2 The super translation Lie algebras and Lie groups We will now see that the spinor and the vector representations of so(1, 1+ k) can (1,1+k) 2k 2 be glued together to obtain the Lie super algebra m | h2(K) ΠK (Π is the parity inversion functor) of super translations of the super≃ Minkowski× space (1,1+k) 2k 2 M | . First we construct ΠK as follows: consider the Clifford algebra C(L)=

12In fact these linear maps keep fixed the imaginary part of z in (49).

32 a + bǫ a, b, R defined in 1.2.2, where ǫ is an odd variable such that ǫ2 = 1, { | 2 ∈ }2 1 § 1 − and let ΠK := K R C(L) (recall that C(L) is the subspace of odd elements of ⊗ 2 C(L)). Second we embedd h (K) ΠK as a subspace of M(5, K) R C(L), where 2 × ⊗ M(5, K) is the space of 5 5 matrices over K, by the mapping × 1 1 0 0 ǫλ1 2 (t + x) 2 z 1 1 0 0 ǫλ2 z (t x) 1 t + x z  2 2 −  , ǫλ , ǫλ 00 0 ǫλ ǫλ . 2 z t x 1 2 7−→ 1 2   −    000 0 0     000 0 0      It is convenient to define, for λ ,λ K, the odd matrices 1 2 ∈

0 0 λ1 0 0 00 0 0 0 000 00 0 0 λ2 0 0 λ1   λ2   Q1 := ǫ 00 0 λ1 0 , Q2 := ǫ 00 0 0 λ2 . (51)  000 00   00 0 0 0       000 00   00 0 0 0          Moreover we introduce the notation 0 0 0 v11 v12 0 0 0 v21 v22 v11 v12   := 000 0 0 v21 v22   000 0 0     000 0 0      and define 1 0 0 1 0 0 0 0 X := , X := , X := , X := . 11 0 0 12 0 0 21 1 0 22 0 1     We then have the following anti-commutation relations: λ,µ K, a, b, =1, 2, ∀ ∈ ∀ [Qλ, Qµ] := QλQµ + QµQλ = λµX µλX . (52) a b a b b a − ab − ba In this computation we have assumed that λ and µ are even variables (in particular they commute with ǫ, although λ and µ may not commute together if K is not commutative). Alternatively if, for any ζ K, we write ∈ (ζ) := (Reζ)(X + X )/2=(ζ + ζ)(X + X )/4 ℜ(ab) ab ba ab ba and (ζ) := (Imζ)(X X )/2=(ζ ζ)(X X )/4 ℑ[ab] ab − ba − ab − ba (note that (ζ)= (ζ) and (ζ)+ (ζ) = 0), we have ℜ(ab) ℜ(ba) ℑ[ab] ℑ[ba] [Qλ, Qµ]= 2 (λµ)+ (λµ) . (53) a b − ℜ(ab) ℑ[ab] K R  Now let (uα)1 α k be an orthonormal basis of over such that u1 = 1 (hence u , ,u are≤ imaginary)≤ and denote: 2 ··· k := (1), for (a, b)=(1, 1), (1, 2) or (2, 2) ; • ℜ(ab) ℜ(ab) 33 1 0 uα α := [12](uα)= 2 for α =2, ,k; • ℑ ℑ uα 0 ··· −  Qα := Quα ; • a a Γ[αβ]γ such that (u u u u )/2=Γ[αβ]γu and ǫ = ǫ = 0, ǫ = ǫ = 1. • α β − β α γ 11 22 12 − 21 Then (53) can also be written as

[Qα, Qβ]= 2 δαβ +Γ[αβ]γǫ . (54) a b − ℜ(ab) abℑγ  α Note that all matrices (ab) and α commute together and with the Qa ’s. All ℜ ℑ (1,1+k) 2k that can be summarized by saying that the vector space m | spanned by α (ab), α and Qa is a super Lie algebra, the super Lie algebra of super translations ℜon theℑ super Minkowski space of dimension (1, 1+ k) 2k. The even part has the | basis (11), (12), (22), 2, , k and coincides with the trivial Lie algebra of ℜ ℜ ℜ ℑ ··· ℑ α α translations. A basis of the odd part is (Q1 , Q2 )α=1, ,k.  ··· The super Minkowski space itself can be identified with the super Lie group (1,1+k) 2k (1,1+k) 2k M | obtained by exponentiating m | . This requires some extra flesh on (1,1+k) 2k 1 q m | : we introduce other odd variables η , , η (some of these variables may a R 1 ··· q be labelled θα later on) such that Λq := [η , , η ] is a free super commutative super algebra (i.e. in particular ηiηj + ηjηi =··· 0), which all anticommute with ǫ. For that purpose, it suffices to construct R[ǫ, η1, , ηq] R[ǫ] Λ as the Clifford ··· ≃ ⊗ q algebra C(L Eq) over the vector space L Eq with basis (ǫ , ǫ , ǫ ) and the ⊕ ⊕ 0 1 ··· q bilinear map B such that B(ǫ0, ǫ0) = 1 and B(ǫa, ǫb)=0if(a, b) = (0, 0). Then (1,1+k) 2k 0 6 m | Λq is a Lie algebra and can be exponentiated as follows. Consider (ab) ⊗ α a α (1,1+k) 2k 0 (ab) α 0 a 1 any v (ab) + v α + θαQa m | Λq , where v , v Λq and θα Λq. ℜ  ℑ ∈ ⊗ ∈ γ γ ∈ In the following we use repeatedly that, as matrices, XabXcd = XabQc = Qc Xab = α β γ  Qa Qb Qc = 0. This implies

(ab) α a α v +v α+θ Q (ab) α a α 1 a α b β e ℜ(ab) ℑ α a =1+ v + v + θ Q + (θ Q ) θ Q . ℜ(ab) ℑα α a 2 α a β b   or, if we set for shortness V := v(ab) + vα and Θ := θa Qα, ℜ(ab) ℑα α a 1 eV +Θ =1+ V +Θ+ Θ2. 2

(1,1+k) 2k We can thus evaluate the product of two elements of M | in the coordinates (ab) α a α given by the exponential chart. Let W +Ψ := w (ab) + w α + ψαQa be another (1,1+k) 2k 0 ℜ ℑ element of m | Λ , then ⊗ q  1 1 eV +ΘeW +Ψ = 1+ V +Θ+ Θ2 1+ W +Ψ+ Ψ2 2 2    1 1 = 1+(V + W )+(Θ+Ψ)+ Θ2 +ΘΨ+ Ψ2. 2 2 However the quadratic terms can be written 1 1 1 Θ2 +2ΘΨ+Ψ2 = (Θ+Ψ)2 + [Θ, Ψ] . 2 2 2  34 We hence deduce 1 exp (V + Θ)exp(W + Ψ) = exp (V + W )+(Θ+Ψ)+ [Θ, Ψ] , (55) 2   where, by using (54),

[Θ, Ψ] = θa Qα, ψb Qβ = θa ψb Qα, Qβ =2θa ψb δαβ +Γ[αβ]γǫ . α a β b − α β a b α β ℜ(ab) abℑγ h i h i  A particular case of (55) is for W = 0: we can then write

eV +ΘeΨ = eV +Θ+DΨ(V +Θ), where 1 D (V +Θ)=Ψ+ [Θ, Ψ] = ψb Qβ θa δαβ +Γ[αβ]γǫ . Ψ 2 β b − α ℜ(ab) abℑγ   We remark that

∂ ∂ ∂ D (V +Θ) = ψb θa δαβ +Γ[αβ]γǫ (V + Θ), Ψ β ∂θb − α ∂v(ab) ab ∂vγ β  !

so that DΨ can be identified with a differential operator. Letting ∂ ∂ ∂ Dα := θb δαβ +Γ[αβ]γǫ , a ∂θa − β ∂v(ab) ab ∂vγ α   b β [αβ] [βα] we have DΨ = ψβDb . We can also introduce the notations: ∂ = ∂ := k [αβ]γ ∂ ∂ −α ∂ γ=2 Γ γ for 1 α, β k, ∂(ab) = ∂(ba) := (ab) for 1 a, b 2 and ∂a := a ∂v ≤ ≤ ∂v ≤ ≤ ∂θα for 1 a 2 and 1 α k and set P ≤ ≤ ≤ ≤ Dα = ∂α θb (δαβ∂ + ǫ ∂[αβ]), a a − β (ab) ab or Dα = ∂α (θ1 ∂ + θ2 ∂ + θ2 ∂[αβ]) and Dα = ∂α (θ1 ∂ + θ2 ∂ θ1 ∂[αβ]). 1 1 − α (11) α (12) β 2 2 − α (12) α (22) − β b β Φ V +Θ V +Θ+τΦ(V +Θ) Similarly, for Φ := φβQb , we have e e = e , where τΦ(V +Θ) = 1 a α Φ+ 2 [Φ, Θ]. Thus we can write τΦ = φατa , where

α α b αβ [αβ] τa = ∂a + θβ(δ ∂(ab) + ǫab∂ ).

α α Operators Da and τa are respectively the left invariant and the right invariant (1,1+k) 2k Φ V +Θ Ψ Φ V +Θ Ψ vector fields on M | . By writing e e e = e e e , we deduce that [τ α,Dβ] = 0, for all a, b =1, 2 and α, β =1, ,k. We moreover have a b ···   α β αβ [αβ] τa , τb =2 δ ∂(ab) + ǫab∂ h i  and Dα,Dβ = 2 δαβ∂ + ǫ ∂[αβ] . a b − (ab) ab h i 

35 (λ1,λ2) λ1 λ2 Remark 8.1 Let us set Q := Q1 + Q2 . Then Relation (52) is equivalent to the fact that, for any (λ ,λ ) K2, 1 2 ∈ 2 (λ1,λ2) (λ1,λ2) (λ1,λ2) (λ1,λ2) λ1 λ1λ2 Q , Q = 2X , where X := | | 2 . − λ2λ1 λ2 | |    We observe that X(λ1,λ2) is a null vector (a vector in the light cone := m 2C { 2 ∈ h2(K) det m = 0 ) with a nonnegative time coordinate t = λ1 + λ1 0. These| two properties} are important since they are the reason for| the| fact| | that≥ supersymmetric Yang–Mills theories exists only in dimensions (1, 1+ k) (see e.g. [11], 5). Moreover this construction provides us with an orientation of the time, an important§ property of supersymmetric theories. Also the set /R of real lines C contained in , with its canonical conformal structure, can be identified with the sphere Sk (theC heavenly sphere at infinity) or, equivalentely with the projective line KP . Indeed the map ϕ : 0 Sk R K defined by (t, x, z) (x/t, z/t) induces a conformal diffeomorphismC\{ } −→ between⊂ ×/R and Sk. Note also that7−→ the image 2 2 C (λ1,λ2) λ1 λ1 λ1λ2 of X by ϕ is | |2−| |2 , 2 2 , i.e. is nothing but the image by the Hopf λ1 + λ1 λ1 + λ1 | | | | | | | | fibration of (λ1,λ2).  These observations are related to the R-symmetries: these are automorphisms of the super Lie algebra of supertranslations which leaves all space-time translations 2 h2(K) invariant (i.e. they act only on ΠK ). It is clear from the previous discussion that these symmetries can be identified with the automorphisms of K2 which maps each fiber of the Hopf fibration onto itself. Hence

if K = R, the R-symmetry group is 1 ((λ ,λ ) (λ ,λ )); • {± } 1 2 7−→ ± 1 2 if K = C, the R-symmetry group is U(1) ((λ ,λ ) (eiθλ , eiθλ ), for θ • 1 2 7−→ 1 2 ∈ R);

if K = H, the R-symmetry group is Spin 3 SU(2) ((λ1,λ2) (λ1α,λ2α), • for α H such that u =1); ≃ 7−→ ∈ | | if K = H, the R-symmetry group is 1 ((λ ,λ ) (λ ,λ ))13. • {± } 1 2 7−→ ± 1 2 8.3 The super Poincar´egroup

(1,1+k) 2k Lastly we can extend the super Lie group M | (respectively its super Lie (1,1+k) 2k algebra m | ) by gluing it with the spin cover of the Lorentz group Spin(1, 1+ k) SL(2, K) (respectively with so(1, 1+ k) sl(2, K)). We obtain the super ≃ ≃ 13 2 2 Indeed assume that there exists some real endomorphism (ϕ1, ϕ2) : O O such that 2 2 −→ 2 λ1, λ2 O, (i) ϕ1(λ1, λ2) = λ1 , (ii) ϕ1(λ1, λ2)ϕ2(λ1, λ2) = λ1λ2 and (iii) ϕ2(λ1, λ2) = ∀ 2 ∈ | | | | | | λ2 . Then: (a) we remark that (i) implies in particular that ϕ1(0, λ2) = 0 and, since ϕ1 is | | linear, this forces that ϕ1(λ1, λ2)= ϕ1(λ1); (b) similarly (iii) implies that ϕ2(λ1, λ2)= ϕ2(λ2); (c) furthermore (i) and (iii) implies that ϕ1 and ϕ2 are real isometries of O; (d) by testing (ii) with λ2 = 1, we deduce that we must have ϕ1(λ)= λα1, where α = ϕ2(1); (e) similarly by testing (ii) with λ1 = 1 we deduce that ϕ2(λ)= λα2, where α = ϕ1(1); (f) by testing (ii) with λ1 = λ2 =1 we obtain that α1 = α2 = α; (g) we can set w.l.g. α = cos θ+u2 sin θ and choose λ1 = u3 and λ2 = u5, where (u2,u3,u5) forms an orthonormal family of imaginary octonions such that u2u3 u5, then ⊥ one checks that the relation (u3α)(u5α)= u3u5 is possible only if α = 1. ±

36 Poincar´egroup P (1, 1+k 2k). A matricial representation of an element g P (1, 1+ k 2k) is | ∈ | S θ1ǫ (t + x)/2 z/2 θ2ǫ z/2 (t x)/2  1 −2  g = 0 0 1 θ ǫ θ ǫ ,  0 0 0   (S ) 1   0 0 0 † −      K 0 t 1 2 K 1 0 K 0 where S SL(2, ) Λq and S† = S , θ , θ Λq, t, x Λq and z Λq. ∈ ⊗ a ∈ ⊗ ∈ ∈ ⊗ Note that the conjugate θ of θa involves the real structure of K, i.e., if θa = u η, K 1 a ⊗ for u and η Λq, then θ = u η. In particular we recover the action of an ∈ ∈ (1,1+k) 2k ⊗ m m 1 element S SL(2, K) on M | by e ge g− , where ∈ 7−→ 0 θ1ǫ (t + x)/2 z/2 S 0 0 θ2ǫ z/2 (t x)/2  1 −2  g = 01 0 and m = 0 0 0 θ ǫ θ ǫ .  1  0 0 (S†)−  0 0 0   0     0 0 0      8.4 Specializations

(1,2) 2 3 2 8.4.1 The super Minkowski space M | R | ≃ It corresponds to the simplest case, when K = R (see 7.1). §

(1,3) 4 4 4 8.4.2 The super Minkowski space M | R | ≃ (1,3) 4 It corresponds to the case k = 2, i.e. K = C. We then obtain M | , the simplest non trivial supersymmetric extension of the standard Minkowsi space M 1,3. In (1,3) 4 order to connect our presentation of the super Lie algebra m | with the standard C (1,3) 4 C one, it is useful to embedd it in its complexification √ 1 R m | . Here √ 1 − ⊗ − denotes a copy of C where the square root of 1 is denoted by √ 1 instead of 14 C (1−,3) 4 −(1,3) 4 u2 = i . For any A = B + √ 1C √ 1 R m | , where B,C m | , we note √ 1 − ∈ − ⊗ ∈ A − := B √ 1C. Then we set, for a =1, 2, − − √ 1 Q := Q1 √ 1Qi /√2 and Q := Q1 + √ 1Qi /√2= Q − . a a − − a a˙ a − a a We then have the relations 

[Q , Q ]= Q , Q˙ = 0 and Q , Q˙ = 2X ˙ , a b a˙ b a b − ab a˙ 15 Xab+Xba  iXab iXba  ab˙ a where Xab˙ = 2 √ 1 −2 . We denote by x , θ and θ (for a =1, 2) (1,3)−4 − ab˙ the coordinates on m | in the basis Xab˙ , Qa, Qa˙ . We write ∂ab˙ := ∂/∂x , ∂a := a a˙ ∂/∂θ and ∂a˙ := ∂/∂θ the partial derivatives with respect to these coordinates. 14 One may set √ 1= i = u2 without trouble here, however the next cases, where K = H or O require more care. − 15 Again we may set √ 1= i: we would then have X ˙ = Xab. − ab

37 The coordinates16 (xab˙ ) are connected to the coordinates (t, x, z1 + iz2) R R C by ∈ × × x11˙ x12˙ 1 t + x z1 + iz2 = . x21˙ x22˙ 2 z1 iz2 t x    − −  This implies that ∂ ˙ = ∂ + ∂ ∂ ˙ = ∂ 1 i∂ 2 11 t x 12 z − z ∂ ˙ = ∂ 1 + i∂ 2 ∂ ˙ = ∂ ∂ . 21 z z 22 t − x Lastly the left and right translations in these coordinates read respectively DΨ = a a˙ a a˙ ψ Da + ψ Da˙ and τΦ = ψ τa + ψ τ a˙ with: ˙ b b D = ∂ θ ∂ ˙ , D = ∂ θ ∂ , a a − ab a˙ a˙ − ba˙ and ˙ b b τa = ∂a + θ ∂ab˙ , τ a˙ = ∂a˙ + θ ∂ba˙ . Very often physicists consider chiral fields, i.e. fields Φ satisfying the constraints

D1˙ Φ= D2˙ Φ=0.

(1,5) 8 6 8 8.4.3 The super Minkowski space M | R | ≃ It corresponds to the case k = 4, i.e. K = H. Here there is an alternative system of coordinates based on the exceptional isomorphism SL(4, C) Spin(6)C. This comes from the identificaton Λ2C4 C6: the natural action of SL≃(4, C) on C4 induces an ≃ C action of SL(4, C)onΛ2C4 which can be identified with the action of Spin(6) on C6. 4 Let (e1, e2, e3, e4) be the canonical basis of C and (ea eb)1 a

∂ = ∂ 3 i∂ 4 ; ∂ = ∂ + ∂ ; ∂ = ∂ 1 i∂ 2 12 z − z 13 t x 14 z − z ∂ = ∂ 1 + i∂ 2 ; ∂ = ∂ ∂ 23 z z 24 t − x ∂34 = ∂z3 + i∂z4 .

16This system of notations is similar to the notations introduced by R. Penrose for the twistor theory, see the text by P. Baird in this volume.

38 One also observes that the restriction of B to M1,5 has the signature (5, 1) and, more precisely, 4B(P (t, x, z),P (t, x, z)) = (t2 x2 z 2). Hence the subgroup of SL(4, C) which preserves M1,5 when acting− on Λ−2C4 can−| be| identified with Spin(1, 5). Indeed SL(2, H) can be identified with a real form of SL(4, C). More precisely we can identify17 C4 with H2 through the map T :(u1,u2,u3,u4) (u1 +ju3,u2 +ju4) and define SL(2, H) as the set of transformations g SL(47−→, C) which maps each ∈ 2 quaternionic line into a quaternionic line, i.e. such that, for any (λ1,λ2) H 0 , 1 ∈ \{ } T g T − maps the line (λ α,λ α) α H to another line (µ α,µ α) α ◦ ◦ { 1 2 | ∈ } { 1 2 | ∈ H . It is easy to see that a complex plane in C4 which contains a vector U = (u}1,u2,u3,u4) = 0 is the inverse image of a quaternionic line by T if and only if it 6 1 1 1 3 2 4 3 4 1 2 contains σ(U) := T − (T (U)j)= T − ((u + ju )j, (u + ju )j)=( u , u , u , u ). − − So SL(2, H) is the set of transformations g SL(4, C) which maps any complex plane spanned by (U, σ(U)) to a complex plane∈ of the same type. Lastly we remark that, for any U C4 0 , U σ(U) belongs to M1,5 and the action of SL(2, H) on the pair (U, σ(U∈)) is mapped\{ } ∧ to the action of SO(1, 5) on M1,5. More precisely one 1 2 3 4 C4 ab can compute that, if U =(u ,u ,u ,u ) , then U σ(U)= 1 a

13 14 12 2 1 y y + y j λ1 λ1λ2 P − (U σ(U)) = = | | , ∧ y23 jy34 y24 λ λ λ 2  −   2 1 | 2|  1 3 2 4 where λ1 := u + ju and λ2 := u + ju , i.e. (λ1,λ2) = T (U). Hence we conclude (λ1,λ2) λ1 λ2 that (recall the notation Q := Q1 + Q2 ):

U C4, if (λ ,λ )= T (U), P [Q(λ1,λ2), Q(λ1,λ2)] = 2U σ(U). (56) ∀ ∈ 1 2 − ∧

ǫU 2U σ(U)  In other words the diagram: ΠC7−→−4 ∧/ M1,5 is commutative. Note that, by po- O T P  2 H / h2(H) Π [ , ] · · larization, (56) implies P [Q(λ1,λ2), Q(µ1,µ2)] = U σ(V ) V σ(U), where − ∧ − ∧ T (V )=(µ1,µ2). (1,5) 8 1,5  4 Lastly the identification m | M ΠC is related to an alternative simple ≃ × (1,5) 8 (1,5) 8 representation of the supercommutator on m | : we again complexify m | by C (1,5) 8 embedding it in √ 1 R m | and set − ⊗ q := Q1 √ 1Qi /√2, q := Qj √ 1Qk /√2, 11 1 − − 1 12 − 1 − − 1 q := Q1 √ 1Qi /√2, q := Qj √ 1Qk /√2, 21 2 − − 2 22 − 2 − − 2 q := Qj + √ 1Qk /√2, q := Q1 + √ 1Qi /√2, 31 1 − 1 32 1 − 1  q := Qj + √ 1Qk /√2, q := Q1 + √ 1Qi /√2. 41 2 − 2 42 2 − 2 17Most Authors use the more pleasant identification (u1,u2,u3,u4)  (u1 + u3j, u2 + u4j); however this convention is not that convenient in our context. 7−→

39 aA We denote by θ 1 a 4;A=1,2 the coordinates associated with these vectors and we aA ≤ ≤ write ∂aA := ∂/∂θ  . Then one can deduce from (54) that [q , q ]= 2ǫ e e , A, B =1, 2, a, b =1, 2, 3, 4, aA bB − AB a ∧ b ∀ ∀ where ǫ = ǫ = 1 and ǫ = ǫ = 0. As a consequence 12 − 21 11 22 D = ∂ ǫ θbB∂ and τ = ∂ + ǫ θbB∂ . aA aA − AB ab aA aA AB ab 8.5 Dimensional reductions We have presented only particular examples of super Minkowski spaces, however one can construct new examples from this list by dimensional reduction, i.e. by consid- ering fields which are invariant under some space-time direction. For instance one can construct super extensions of the usual Minkowski space M 1,3 with more than (1,3) 8 4 odd dimensions: M | (related to N = 2 supersymmetric theories), which can (1,5) 8 (1,3) 16 be obtained by reduction of M | and M | (related to N = 4 supersymmetric (1,9) 16 (1,3) 4 theories), which can be obtained by reduction of M | . Our first example M | corresponds to N = 1 supersymmetric theories. Similar reductions leads to various super space-times with 2 even dimensions, which are important for the theory of superstrings (see e.g. [11]). In the following we present alternative representations (1,5) 8 (1,9) 16 of m | and m | which shows how one can produce N = 2 and N = 4 super- symmetry groups over M 1,3 with central extensions.

(1,5) 8 (1,3) 8 From m | to a central extension of m | (1,5) 8 C (1,5) 8 We embedd m | in its complexification √ 1 R m | and set, for a =1, 2: − ⊗ Q := (Q1 √ 1Qi )/√2; Q := (Q1 + √ 1Qi )/√2; a1 a − − a a˙1 a − a Q := (Qj √ 1Qk)/√2; Q := (Qj + √ 1Qk)/√2. a2 a − − a a˙ 2 a − a Then one deduce from (53) that, for a, b =1, 2 and A, B =1, 2,

Q , Q˙ = 2δ √ 1 (i) ; aA bB − AB R(ab) − − I[ab] [Q , Q ] = 2ǫ ǫ (j) √ 1 (k) ,  aA bB ab AB I[12] − − I[12]  √ 1 √ 1 − − and [QaA˙ , QbB˙ ]= [QaA, QbB] and [QaA˙ , QbB]= [QaA, QbB˙ ] , where ǫ11 = ǫ22 = 0 and ǫ = ǫ = 1. However using the notation X ˙ introduced in 8.4.2 and 12 − 21 ab § letting

Z := ǫ j √ 1k and Z := ǫ j + √ 1k , AB − ABI[12] − − AB − AB I[12] − we can write the anticommutation relations as 

Q , Q˙ = 2X ˙ δ [Q , Q ]= 2X δ aA bB − ab AB aA˙ bB − ba˙ AB (57) [Q , Q ]= 2ǫ Z Q , Q˙ = 2ǫ ˙ Z .  aA bB − ab AB aA˙ bB − a˙ b AB Upon reduction to M 1,3 the relations on the first line are the only nontrivial anti- (1,3) 8 commutation relations of m | (since then we set [QaA, QbB] = QaA˙ , QbB˙ = 0). However D. Olive and E. Witten [21] discovered that, for the N = 2 supersymmet-   ric Yang–Mills on M 1,3, the super Lie algebra of Noether charges of the Lagrangian

40 (1,3) 8 (which is of course symmetric under the action of m | ) is a central extension of (1,3) 8 m | , obtained by adding two central charges which correspond the the real and imaginary parts of ZAB. The amazing point is that these charges have a topological interpretation in way similar to the example seen in 7.3 and are related the possi- bility of having an electro-magnetic duality for supersymm§ etric non Abelian gauge theories conjectured in [19].

(1,9) 16 (1,3) 16 From m | to a central extension of m | (1,9) 16 A similar construction starting from m | leads to a central extension of the Lie (1,4) 4 (1,9) 16 super algebra of the N = 4 supersymmetric extension of m | . We embedd m | C (1,9) 16 in √ 1 R m | and set, for a =1, 2: − ⊗ Q := (Q1 √ 1Q2)/√2; Q := (Q1 + √ 1Q2)/√2; a1 a − − a a˙ 1 a − a Q := (Q3 √ 1Q4)/√2; Q := (Q3 + √ 1Q4)/√2; a2 a − − a a˙ 2 a − a Q := (Q6 √ 1Q7)/√2; Q := (Q6 + √ 1Q7)/√2; a3 a − − a a˙ 3 a − a Q := (Q8 √ 1Q5)/√2; Q := (Q8 + √ 1Q5)/√2, a4 a − − a a˙ 3 a − a where, for α = 1, , 8 and a = 1, 2, Qα = Quα and we recall that (u , ,u ) ··· a a 1 ··· 8 denotes an orthonormal basis of O such that u1 = 1. It is also important to precise the multiplication rule which is adopted here (since several conventions can be found in the litterature). We recall it in the following diagram: we should think the 6

4

3 5

8

2 6 7

Figure 3: The product of two imaginary octonions.

oriented line segments in this diagram as being 6 oriented circles (the end point being connected to the source point by another oriented line segment), so that we actually have 7 oriented circles and seven points of intersections. Each point of intersection is labelled by α running from 2 to 8 and represents an imaginary element uα of the basis. The product uαuβ of two different elements uα and uβ is uγ , where γ is the third point on the circle which contains α and β and the sign ±is + (resp. ) if the sequence (α,β,γ) respects (resp. does not respect) the orientation± of the − circle. We note that, for any A =1, , 4, Q =(Qα √ 1Qβ)/√2, where α and ··· aA a − − a β are chosen in such a way that uαuβ = u2. Then we deduce from (53) that, for a, b =1, 2 and A, B =1, , 4, ···

Q , Q˙ = 2X ˙ δ [Q , Q ]= 2X δ aA bB − ab AB aA˙ bB − ba˙ AB (58) [Q , Q ]= 2ǫ Z Q , Q˙ = 2ǫ ˙ Z ,  aA bB − ab AB aA˙ bB − a˙ b AB   where the generators ZAB and ZAB are given by the relations ZAB + ZBA = 0,

41 √ 1 Z = Z − and (setting := for shortness): AB AB I I[12] Z := u + √ 1u Z := u + √ 1u Z := u + √ 1u 12 I − 3 − 4 13 I − 6 − 7 14 I − 8 − 5 Z := u √ 1u Z := u + √ 1u  23 I − 8 − − 5 24 I 6 − 7  Z := u √ 1u .  34 I − 3 − − 4

We remark that, if we set 12 := 34, 13 := 42 and 14 := 23, then Z AB = ZAB = √ 1 ∗ ∗ ∗ ∗ ZAB − . As shown by H. Osborn [22] a reduction of the N = 1 supersymmetric Yang– Mills theory on M (1,9) to M (1,3) produces an N = 4 supersymmmetric Yang–Mills theory. The Lie algebra of Noether charges of this theory is a central extension of (1,9) 16 the pure N = 4 supersymmetry Lie algebra and is isomorphic to m | , the 6 central charges corresponding to the real and imaginary parts of the ZAB’s. These central charges can again be interpretated as topological charges [22].

9 Conclusion

We have essentially not presented the most important thing, namely the applications of the idea of supersymmetry: On the the physical side, its possible role in the quantum field theories for im- • proving the renormalisation and solving the hierarchy problem (i.e. ensuring that all fundamental forces of Nature unify at very high energy, see e.g. [32] III), a possible solution to the Montonen–Olive conjecture about electromag- netic duality for gauge theories by E. Witten and N. Seiberg, the theory of (see e.g. [12] and the text by M. Eigeleh in this volume), the role of anticommuting variables in the BRST (Bechi–Rouet–Stora–Tyutin) theory of quantization of gauge theories (see e.g. [32] II and [17, 25]), etc., without mentioning the theories of superstrings and branes... On the mathematical side, its application by E. Witten to Morse theory (see • e.g. [25]), the index theorem for Dirac operators (see e.g. [33, 1, 25]), the topology of 4-dimensional manifold (the Seiberg–Witten theory), the mirror symmetry for Calabi–Aubin–Yau manifolds [28, 30], etc.

References

[1] O. Alvarez, Lectures on quantum mechanics and the index theorem, in Geome- try and quantum field theory, D. Freed and K. Uhlenbeck, ed., IAS/Park City Series, Volume 1, American Mathematical Society, 1995. [2] J.C. Baez, The octonions, Bull. Amer. Math. Soc., Vol. 39, No. 2, April 2002, 145–206. [3] A. Bahraini, Supersym´etrie et g´eom´etrie complexe, th`ese de Doctorat, Univer- sit´eParis 7, 2004. [4] M. Batchelor, The structure of supermanifolds, Trans. of the Amer. Math. Soc. 253 (1979), 329–338.

42 [5] M. Batchelor, Two approaches to supermanifolds, Trans. of the Amer. Math. Soc. 258 (1980), 257–270.

[6] S. Coleman, J. Mandula (1967), All Possible Symmetries of the S Matrix, Phys. Rev. 159: 1251–1256.

[7] P. Deligne, Notes on spinors, in [9].

[8] P. Deligne, Morgan, Classical fields and supersymmetry, in [9].

[9] P. Deligne, P. Etingof, D. Freed, L. Jeffrey, D. Kazhdan, J. Morgan. D. Morri- son, E. Witten, ed., Quantum fields and strings: a course for mathematicians, Volume 1, American Mathematical Society, 1999.

[10] P. Deligne, D. Freed, Supersolutions, in [9].

[11] D. Freed, Five lectures on supersymmetry, American Mathematical Society, 1999.

[12] P.G.O. Freund, Introduction to supersymmetry, Cambridge Monographs on Mathematical Physics, Cambridge University Press, 1986.

[13] Y.A. Gol’fand, E.P. Likhtman, Zh. Eksp. Teor. Fiz. Pis. Red. 13 (1971), 452.

[14] F. H´elein, A representation formula for maps on supermanifolds, preprint arXiv:math-ph/0603045, to appear in J. Math. Phys.

[15] V. G. Kac, Lie superalgebras, Adv. Math. 26 (1977), 8–96.

[16] M. Kawski, H. Sussmann, Noncommutative power series and formal Lie- algebraic techniques in nonlinear control theory, European Consort. Math. Indus. Teubner, Stuttgart, 1997.

[17] B. Kostant, S. Sternberg, Symplectic reduction, BRS , and infinite dimensional Clifford algebras, Annals of Physics 176 (1987), 49–113.

[18] Y.I. Manin, Gauge field theory and complex geometry, Grundlehren der Math- ematischen Wissenschaften 289, Springer-Verlag, 1988.

[19] C. Montonen, D. Olive, Magnetic monopoles as gauge particles ?, Phys. Lett. 72 (1977), 117–120.

[20] C. Manogue, J. Schray, Finite Lorentz transformations, automorphisms, and division algebras, J. Math. Phys. 34 (1993), 3746.

[21] D. Olive, E. Witten, Supersymmetry algebras that include topological charges, Phys. Lett. B 78 (1978), 97–101.

[22] H. Osborn, Topological charges for N =4 supersymmetric gauge theories and monopoles of spin 1, Phys. Lett. B 83 (1979), 321–326.

[23] A. Rogers, A global theory of supermanifolds, J. Math. Phys. 21 (1980), 1352– 65; Super Lie groups: global topology and local structures, J. Math. Phys. 22(1981), 939–45.

43 [24] A. Rogers, Graded manifolds, supermanifolds and infinite-deimensional Grass- mann algebras, Comm. Math. Phys. 105 (1986), 353–368.

[25] A. Rogers, Supermanifolds: theory and applications, World Scientific 2007.

[26] H. Tachibana, K. Imaeda, Octonions, superstrings and ten-dimensional spinors, Nuovo Cimento B 104 (1989), 91.

[27] W.M. Tulczyjew, The graded Lie algebra of multivector fields and the general- ized Lie derivative of forms, Bull. de l’Acad. Polon. des Sci., S´erie sci. Math., Astr. et Phys. XXII (1974), 937-942.

[28] C. Vafa, E. Zaslov, eds., Mirror symmetry, Clay Mathematical Monographs 1, American Mathematical Society 2003.

[29] V.S. Varadarajan, Supersymmetry for mathematicians: an introduction, Courant Lecture Notes 11, American Mathematical Society 2004.

[30] C. Voisin, Sym´etrie miroir, Panoramas et synth`eses no. 1, Soci´et´e Math´ematique de France 1996.

[31] D. Volkov, V. Akulov, Is the neutrino a Goldstone particle ?, Physics Letters B 46 (1973), 109–110.

[32] S. Weinberg, The quantum theory of fields, Volume II and III, Cambdridge University Press, 2000.

[33] E. Witten, Index of Dirac operators, in [9].

[34] Fall Problems posed by E. Witten, solutions by P. Deligne, D. Freed, L. Jeffrey and S. Wu, in [9].

44