• 検索結果がありません。

Integration of Vector Fields on Smooth and Holomorphic Supermanifolds

N/A
N/A
Protected

Academic year: 2022

シェア "Integration of Vector Fields on Smooth and Holomorphic Supermanifolds"

Copied!
28
0
0

読み込み中.... (全文を見る)

全文

(1)

Integration of Vector Fields on Smooth and Holomorphic Supermanifolds

St´ephane Garnier and Tilmann Wurzbacher

Received: September 19, 2012 Revised: December 12, 2012

Communicated by Christian B¨ar

Abstract. We give a new and self-contained proof of the existence and unicity of the flow for an arbitrary (not necessarily homogeneous) smooth vector field on a real supermanifold, and extend these results to the case of holomorphic vector fields on complex supermanifolds.

Furthermore we discuss local actions associated to super vector fields, and give several examples and applications, as, e.g., the construction of an exponential morphism for an arbitrary finite-dimensional Lie supergroup.

2010 Mathematics Subject Classification: Primary 58A50, 37C10;

Secondary 57S20, 32C11

Keywords and Phrases: Supermanifolds; Vector fields; Flows; Group actions

1 Introduction

The natural problem of integrating vector fields to obtain appropriate “flow maps” on supermanifolds is considered in many articles and monographs (com- pare, e.g., [17], [2], [19], [14] and [3]) but a “general answer” was to our knowl- edge only given in the work of J. Monterde and co-workers (see [12] and [13]).

Let us consider a supermanifold M = (M,OM) together with a vector field X in TM(M), and an initial condition φ in Mor(S,M), where S = (S,OS) is an arbitrary supermanifold and Mor(S,M) denotes the set of morphisms fromS toM. The case of classical, ungraded, manifolds leads one to consider the following question: does there exist a “flow map” F defined on an open sub supermanifold V ⊂R1|1× S and having values inM and an appropriate

(2)

derivation onR1|1,D=∂t+∂τ+τ(a∂t+b∂τ), where∂t= ∂t,∂τ =∂τ anda, b are real numbers, such that the following equations are fulfilled

D◦F = F◦X

F◦injV{0}×S = φ . (1)

Of course, V should be a “flow domain”, i.e. an open sub supermanifold of

R

1|1 × S such that {0} ×S is contained in the body V of V and for x in S, the set IxR defined by Ix× {x} = (R× {x})∩V is an open interval.

Furthermore injV{0}×S denotes the natural injection morphism of the closed sub supermanifold {0} × S of V into V. Of course, we could concentrate on the caseS=Mandφ= idM, but it will be useful for our later arguments to state all results in this (formally) more general setting.

Though for homogeneous vector fields (X =X0 or X =X1) system (1) does always have a solution, in the general case (X =X0+X1withX06= 0 andX16=

0) the system is overdetermined. A simple example of an inhomogeneous vector field such that (1) is not solvable is given byX =X0+X1 =

∂x∂ξ +

∂ξ∂x

onM=R1|1. The crucial novelty of [13] is to consider instead of (1) the following modified, weakened, problem

(injRR1|1)◦D◦F = (injRR1|1)◦F◦X

F◦injV{0}×S = φ , (2)

where injRR1|1 = injRR1|1×S×S is again the natural injection (and where the above more general derivationDcould be replaced by∂t+∂τ since (injRR1|1) annihi- lates germs of superfunctions of the typeτ·f, f ∈ OR×S).

In [13] (making indispensable use of [12]) it is shown that in the smooth case (2) has a unique maximal solution F, defined on the flow domain V = (V,OR

1|1×S|V), whereV ⊂R×S is the maximal flow domain for the flow of the reduced vector fieldXe =Xf0 onM with initial condition φ. Since thee results of [12] are obtained by the use of a Batchelor model for M, i.e. a real vector bundle E → M such that M ∼= (M,ΓΛE), and a connection on E, we follow here another road, closer to the classical, ungraded, case and also applicable in the case of complex supermanifolds and holomorphic vector fields.

Our new method of integrating smooth vector fields on a supermanifold in Sec- tion 2 consists in first locally solving a finite hierarchy of ordinary differential equations, and is here partly inspired by the approach of [3], where the case of homogeneous super vector fields on compact supermanifolds is treated. We then show existence and unicity of solutions of (2) on smooth supermanifolds and easily deduce the results of [13] from our Lemmata 2.1 and 2.2.

(3)

A second beautiful result of [13] (more precisely, Theorem 3.6 of that reference) concerns the question if the flowF solving (2) fulfills “flow equations”, as in the ungraded case. Hereby, we mean the existence of a Lie supergroup structure onR1|1 such thatF is a local action ofR1|1 onM(in caseS =M, φ= idM).

Again, the answer is a little bit unexpected: in general, given X and its flow F : R1|1× M ⊃ V → M, there is no Lie supergroup structure on R1|1 such that F is a localR1|1-action (with regard to this structure). The condition for the existence of such a structure on R1|1 is equivalent to the condition that (2) holds without the post-composition with (injRR1|1), i.e. the overdetermined system (1) is solvable. Furthermore, both conditions cited are equivalent to the condition that RX0RX1 is a sub Lie superalgebra of TM(M), the Lie superalgebra of all vector fields on M.

After discussing Lie supergroup structures and right invariant vector fields on

R

1|1, as well as local Lie group actions in the category of supermanifolds in general, we show in Section 3 the equivalence of the above three conditions, already given in [13]. We include our proof here notably in order to be able to apply it in the holomorphic case in Section 5 (see below) by simply indicating how to adapt it to this context. Let us nevertheless observe that our result is slightly more general since we do not need to ask for any normalization of the supercommutators betweenX1and X0 resp. X1, thus giving the criterion some extra flexibility in applications.

In Section 4, we give several examples of vector fields on supermanifolds, homo- geneous and inhomogeneous, and explain their integration to flows. Notably, we construct an exponential morphism for an arbitrary finite-dimensional Lie supergroup, via a canonically defined vector field and its flow. We comment here also on the integration of what are usually called “(infinitesimal) super- symmetries” in physics, i.e., purely odd vector fields having non-vanishing self-commutators.

Finally, in Section 5 we adapt our method to obtain flows of vector fields (compare Section 2 and notably Lemma 2.1) to the case of holomorphic vector fields on holomorphic supermanifolds. To avoid monodromy problems one has, of course, to take care of the topology of the flow domains, and maximal flow domains are -as already in the ungraded holomorphic case- no more unique.

Otherwise the analogues of all results in Section 2 and 3 continue to hold in the holomorphic setting.

Throughout the whole article we will work in the ringed space-approach to supermanifolds (see, e.g., [9], [10], [11] and [15] for detailed accounts of this approach). Given two supermanifolds M = (M,OM) and N = (N,ON), a “morphism” φ = (φ, φe ) : M → N is thus given by a continuous map φe : M → N between the “bodies” of the two supermanifolds and a sheaf

(4)

homomophismφ:ON →φeOM. The topological spaceM comes canonically with a sheafCM =OM/J, whereJ is the ideal sheaf generated by the germs of odd superfunctions, such that (M,CM) is a smooth real manifold. Thenφe is a smooth map from (M,CM) to (N,CN). Let us recall that a (super) vector field on M = (M,OM) is, by definition, an element of the Lie superalgebra TM(M) = (DerR(OM))(M) and that X always induces a smooth vector field Xe on (M,CM). For p in M and f +Jp ∈ (CM)p = (OM/J)p one defines Xep(f +Jp) = X0(f)(p), whereX0 is the even part ofX and forg ∈(OM)p, g(p)∈Ris the value ofg in the pointpofM.

2 Flow of a vector field on a real supermanifold

In this section we give our main result on the integration of general (i.e. not necessarily homogeneous) vector fields by a new method, avoiding auxiliary choices of Batchelor models and connections, as in [12]. Our more direct approach is inspired, e.g., by [3], where the case of homogeneous vector fields on compact manifolds is treated, and it can be adapted to the holomorphic case (see Section 4).

For the sake of readability we will often use the following shorthand: if P is a supermanifold, we write injRR1|1 for injRR1|1×P×P. Furthermore, the canonical coordinates of R1|1 will be denoted by t and τ, with ensueing vector fields

t=∂t and∂τ = ∂τ .

Lemma 2.1. Let U ⊂ Rm|n andW ⊂ Rp|q be superdomains, X ∈ TW(W) be a super vector field on W (not necessarily homogeneous) and φ∈Mor(U,W), and t0R. Let furthermore H :V →W be the maximal flow ofXe ∈ X(W), i.e. ∂t◦H=H◦X, subject to the initial conditione H(t0,·) =φe:U →W. Let nowV be(V,OR

1|1×U|V)and(t, τ)the canonical coordinates on R1|1, then there exists a unique F:V → W such that

(injRR1|1)◦(∂t+∂τ)◦F = (injRR1|1)◦F◦X and (3)

F◦injV{t0}×U = φ. (4)

Moreover,Fe:V →W equals the underlying classical flow mapH of the vector fieldXe with initial condition φ.e

Proof. Let (ui) = (xi, ξr) and (wj) = (yj, ηs) denote the canonical coordinates onRm|nandRp|q, respectively. Then there exist smooth functionsajJ∈ C

R

p(W) such that

X = Xp+q j=1

X

J

ajJ(y)ηJ

!

wj,

(5)

where J = (β1, . . . , βq) runs over the index set {0,1}q and ηJ = Yq s=1

ηβss. We then have, of course,

X0=X

j

 X

|J|=|wj|

ajJ(y)ηJ

∂wj resp. X1=X

j

 X

|J|=|wj|+1

ajJ(y)ηJ

∂wj.

Here,|J|equalsβ1+· · ·+βq mod 2 and|wj|is the parity of the coordinate func- tionwj. The morphismF determines and is uniquely determined by functions fIj, gIj∈ CR×Rm(V) fulfilling for eachj∈ {1, . . . , p+q}

F(wj) = X

|I|=|wj|

fIj(t, x)ξI+ X

|I|=|wj|+1

gjI(t, x)τ ξI

(and fIj = 0 if |I| 6=|wj|, gjI = 0 if |I| 6=|wj|+ 1) as is well-known from the standard theory of supermanifolds (compare, e.g., Thm. 4.3.1 in [20]). Here and in the sequel I = (α1, . . . , αn) is an element of the set {0,1}n and ξI stands for the product ξα11·ξ2α2· · ·ξnαn. The notation |I| again denotes the parity ofI, i.e. |I|=α1+· · ·+αn mod 2.

Equation (3) is equivalent to the following equations:

(injRR1|1)◦∂t◦F = (injRR1|1)◦F◦X0 (5) (injRR1|1)◦∂τ◦F = (injRR1|1)◦F◦X1 (6) Applying (5) to the canonical coordinate functions onW, we get the following system, which is equivalent to (5):

X

|I|=|wj|

tfIj·ξI = X

|J|=|wj|

(ajJ) ˇFJ) for all j in{1, . . . , p+q}, (7)

and (6) is equivalent to X

|I|=|wj|+1

gjI·ξI = X

|J|=|wj|+1

(ajJ) ˇFJ) for all j in{1, . . . , p+q}, (8)

where ˇF :=F◦injRR1|1 : ˇV := (V,OR×U|V)→ W. Let us immediately observe that the underlying smooth map of ˇF equals ˜F, the smooth map underlying the morphismF.

Moreover the initial condition (4) is equivalent to X

|I|=|wj|

fIj(t0, x)ξI(wj) for allj in {1, . . . , p+q}. (9)

We are going to show that (7) and (9) uniquely determine the functions fIj on V, i.e. the morphism ˇF. Then the functions gIj are unambiguously given by

(6)

(8) onV, and the morphismF is fully determined.

Let us develop Equation (7) for a fixedj:

X

|I|=|wj|

tfIj·ξI = X

|J|=|wj| J=(β1,...,βq)

(ajJ) Yq s=1

s)βs (10)

and thus X

|I|=|wj|

tfIj·ξI = X

|J|=|wj| J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

. (11)

For fixed j this is an equation of Grassmann algebra-valued maps in the variables t and xthat can be split in a system of scalar equations as follows.

For K = (α1, . . . , αn) ∈ {0,1}n, we will denote the coefficient hK in front of ξK of a superfunctionh=P

MhM(t, x)ξM ∈ OR

m+1|ncompactly by (h|ξK) in the sequel of this proof.

Let us first describe the coefficients for ˇF(ajJ) in (11):

( ˇF(ajJ)|ξK) = 0 if|K|= 1 (12) and, if|K|= 0,

( ˇF(ajJ)|ξK) =ajJ◦Fe ifK= (0, . . . ,0), and

( ˇF(ajJ)|ξK) = Xp µ=1

(∂yµajJ)(Fe(t, x))·fKµ +R

ajJ,(fIν)ν,deg(I)<deg(K)

(13)

if deg(K)>0.

Here forI= (α1, . . . , αn), deg(I) =α1+· · ·+αn, and -more importantly-R= Rj,J,Kis a polynomial function inajJand its derivatives in they-variables up to orderqincluded, and in the functions{fIν|1≤ν ≤p+q,0≤deg(I)<deg(K)}.

Equation (12) is obvious since ajJ is an even function, whereas equation (13) can be deduced from standard analysis on superdomains. More precisely, leta be a smooth function on Rp andψ:Rm+1|nRp|q a morphism (of course to be applied toa=ajJ, ψ= ˇF). Then we can developψ(a) as follows (compare the proof of Theorem 4.3.1 in [20]):

ψ(a) = X

γ

1

γ!(∂γa)(ψe(y1), . . . ,ψe(yp))· Yp µ=1

(yµ)−ψe(yµ))γµ

= a(ψ(t, x)) +e Xp µ=1

(∂yµa)(ψ(t, x))e ·

X

M6=0

fMµ ·ξM

+

(7)

1 2

Xp µ′′=1

(∂yµyµ′′a)(ψ(t, x))e ·

 X

M6=0

fMµ ·ξM

·

 X

M′′6=0

fMµ′′′′·ξM′′

+. . . ,

where X

M6=0

fMµ ·ξM = ψ(yµ)−ψe(yµ) with fMµ depending on t and x. We observe that the last RHS is a finite sum since we work in the framework of finite-dimensional supermanifolds.

In order to get a contribution to (ψ(a)|ξK) we can either extract fKµ from the “linear term” or from products coming from the higher order terms in the above development. Thus

(a)|ξK) = Xp µ=1

(∂yµa)(ψ(t, x))e ·fKµ +R(a,(fIν)ν,deg(I)<deg(K)), whereR is a polynomial as described after Equation (13).

Furthermore, for an element J = (β1, . . . , βq) with |J| = 0 we have for deg(K)>0

 Yq s=1

X

|L|=1

fLp+sξL

βs ξK

=R

(fIj)j,deg(I)<deg(K)

. (14)

And for an elementJ = (β1, . . . , βq) with|J|= 1 we get for deg(K)>0

 Yq s=1

 X

|L|=1

fLp+sξL

βs ξK

=















fKp+l+R

(fIj)j,deg(I)<deg(K)

if deg(J) = 1 andl∈ {1, . . . , q}

such thatβss,l ∀s, R

(fIj)j,deg(I)<deg(K)

if deg(J)>1.

(15)

Obviously, the coefficient ofξK of the LHS of Equation (7) is given by

 X

|I|=|wj|

tfIj·ξI ξK

=

tfKj if|K|=|wj|

0 if|K|=|wj|+ 1 for 1≤j≤p+q.

Taking into account the above descriptions of the ξK-coefficients, we will show the existence (and uniqueness) of the solution functions {fIj|1 ≤ j ≤ p+q, I ∈ {0,1}n} for (t, x) ∈ V by induction on deg(I) and upon observing that all ordinary differential equations occuring are (inho- mogeneous) linear equations for the unknown functions.

(8)

Let us start with deg(I) = 0 that is I = (0,· · ·,0). The “0-level” of the equations (11) and (9) is ∂tf(0,···,0)j = aj(0,···,0) ◦ Fe andf(0,···j ,0)(t0, x) = yj◦φ(x) for alle j such that |wj|= 0.We remark thatf(0,···j ,0)is simplyyj◦Fe and aj(0,···,0) is Xe(yj). Thus Fe is the flow of Xe with initial condition φe at t=t0, i.e.,Fe=H onV. Thus the claim is true forI= (0,· · · ,0).

Suppose k > 0 and that the functions fIj are uniquely defined on V for all j and all I such that deg(I) < k. Let K be such that deg(K) = k. Let us distinguish the two possible parities of k in order to determine fKj for all j.

Recall that fKj = 0 if the parities ofKandj are different.

If k is even, i.e., |K| = 0, we only have to consider j such that |wj| = 0.

Putting (13) and (14) together, we find in this case

tfKj =



 X

|J|=0 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

ξK



=



 X

deg(J)=0 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

+ X

|J|=0 deg(J)>0 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

ξK







=







(aj(0,···,0)) + X

|J|=0 deg(J)>0 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

ξK







= Xp µ=1

yµaj(0,···,0)◦Fe

fKµ +R

(ajJ)J,(fIν)ν,deg(I)<deg(K)

.

Moreover, the initial condition gives fKj(t0, x) = (φ(yj)|ξK), for all j in {1, . . . , p}. Since the ajJ are the (given) coefficients of the vector field X and the functionsfIν with deg(I)< kare known by the induction hypothe- sis, we have a unique local solution functionfKj. Since the ordinary differential equation for fKj is linear its solution is already defined for all (t, x) ∈ V. Thus in the case that k is even fKj is unambiguously defined on V for all j∈ {1, . . . , p+q}and for allK with deg(K) =k.

(9)

Now, if k is odd, i.e.,|K|= 1, we only have to considerj such that|wj|= 1.

Using (13) and (15), we find in this case:

tfKj =



 X

|J|=1 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs ξK



=



 X

deg(J)=1 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

+

+ X

|J|=1 deg(J)>1 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

ξK







=

 Xq s=1

(aj1s,···qs))

X

|L|=1

fLp+sξL

+ X

|J|=1 deg(J)>1 J=(β1,...,βq)

(ajJ) Yq s=1

X

|L|=1

fLp+sξL

βs

ξK







= Xq s=1

aj

1s,···,δqs)◦Fe

fKp+s+R

(ajJ)J,(fIν)ν,deg(I)<deg(K)

.

Moreover, the initial condition gives

fKj(t0, x) = (φ(wj)|ξK) for allj in{p+ 1, . . . , p+q}.

It follows as in the case of|K|= 0, thatfKj exists uniquely for all (t, x)∈V, for allj∈ {1, . . . , p+q} and for allK with deg(K) =k.

We conclude that the functions {fIj|1 ≤j ≤p+q, I ∈ {0,1}n} are uniquely defined on the whole of V. Since the {gIj|1 ≤ j ≤ p+q, I ∈ {0,1}n} are determined by Equation (8) from the {fIj|1 ≤ j ≤ p+q} via comparison of coefficients, the morphism F :V → W is uniquely determined.

We now consider the global problem of integrating a vector field on a super- manifold. In order to prove that there exists a unique maximal flow of a vector field, the following lemma will be crucial.

(10)

Lemma 2.2. Let M = (M,OM) and S = (S,OS) be supermanifolds, X a vector field in TM(M)andφin Mor(S,M). Then

(i) there exists an open sub supermanifoldV = (V,OR

1|1×S|V)ofR1|1×Swith V open inR×S such that{0} ×S⊂V and for all xinS,(R× {x})∩V is an interval, and a morphismF :V → Msatisfying:

(injRR1|1)◦(∂t+∂τ)◦F = (injRR1|1)◦F◦X and (16)

F◦injV{0}×S = φ . (17)

(ii) Let furthermoreF1:V1→ MandF2:V2→ Mbe morphisms satisfying (16) and (17) whereVi = (Vi,OR

1|1×S|Vi)withVi open inR×S such that {0} ×S⊂Vi and for allxinS,(R× {x})∩Viis an interval, fori= 1,2.

Then F1|V12=F2|V12 onV12= (V12,OR

1|1×S|V12), whereV12=V1∩V2. Proof. (i) Letφe:S→M denote the induced map of the underlying classical manifolds. Given now s in S and coordinate domains Us of s and Ws of φ(s), isomorphic to superdomains ˇe UsRm|n resp. WˇsRp|q, by Lemma 2.1 we get solutions of (16) and (17) near s (upon reducing the size of Us if necessary): R1|1× S ⊃R1|1× Us⊃ Vs Fs

−→ Ws⊂ M.IfVs1∩ Vs2 6=∅(compare Figure 1) we know, again by Lemma 2.1, that Fs1 and Fs2 coincide on this intersection. Thus, by taking the unionVofVsfor allsinS, we get a morphism F :R1|1× S ⊃ V → M such thatF|Vs =Fs for all s, and fulfilling (16) and (17).

Figure 1

(ii) We define A as the set of points (t, x) ∈ V12 such that there exists ǫ = ǫ(t,x)>0 andU =U(t,x)an open sub supermanifold ofS, such that its bodyU containsxand forV=V(t,x)= (V(t,x),OR

1|1×S) = (]−ǫ, t+ǫ[×U,OR

1|1×S) we haveF1|V =F2|V. Of course, if t <0 the interval will be of the type ]t−ǫ, ǫ[

(See Figure 2). The claim of the Lemma is now equivalent to A=V12. The set Ais obviously open.

(11)

Figure 2

By an easy application of Lemma 2.1, A contains{0} ×S. The assumptions imply that for allx∈S, the setIxR, defined by (R×{x})∩V12=Ix×{x}, is an open interval containing 0. The definition ofAimplies that the setJx⊂Ix

such that (R× {x})∩A =Jx× {x} is an open interval containing 0 as well.

Assuming now that A 6= V12, then there exists a point (t, x0) ∈ V12\A such that Jx0 6=Ix0. Without loss of generality we can assume thatt >0 and that for 0≤t < t, (t, x0)∈A. LetU0 be an open coordinate neighborhood ofx0

inS andδ >0 such that, withV0:=]t−δ, t+δ[×U0⊂V12,H(V0)⊂W, where W= (W,OM|W) is a coordinate patch ofMandH is the maximal flow ofXe as in Lemma 2.1. Chooset0∈]t−δ, t[. Then (t0, x0)∈Aand thus there exists ǫ >0 andU an open sub supermanifold ofU0= (U0,OS|U0) containingx0such that

F1|V =F2|V, whereV =]−ǫ, t0+ǫ[×R0|1× U ⊂ V12. (18) On V =]t−δ, t+δ[×R0|1× U ⊂ V12,F1 andF2 are defined and for i= 1,2 the maps Fi ◦injV{t0}×U coincide by (18) (Compare Figure 3 for the relative positions of the underlying topological spaces of these open sub supermanifolds ofR1|1× S).

Figure 3

(12)

By Lemma 2.1 we have F1|V = F2|V. Thus F1 = F2 on V ∪ V, and we conclude that (t, x0)∈A. This contradiction shows that V12=A.

Remarks. (1) Obviously, Lemma 2.2 holds true for an arbitrary t0R re- placingt0= 0.

(2) Let us call a “flow domain for X with initial condition φ ∈ Mor(S,M) (with respect tot0R)” a domainV ⊂R1|1× S such that {t0} ×S ⊂V and for allsin S, (R× {s})∩V is connected, i.e. an interval (times{s}) and such that a solution F (a “flow”) of (16) and (17) exists onV. By the preceding lemma there exists such “flow domains”.

Theorem2.3. LetMandSbe supermanifolds,X be a vector field inTM(M), φin Mor(S,M)andt0inR. Then there exists a unique mapF :V → Msuch that

(injRR1|1)◦(∂t+∂τ)◦F = (injRR1|1)◦F◦X and F◦injV{t0}×S = φ ,

where V = (V,OR

1|1×U|V) is the maximal flow domain for X with the given initial condition.

Moreover, Fe:V →M is the maximal flow ofXe ∈ X(M)subject to the initial condition φeatt=t0.

Proof. The proof of the theorem follows immediately from the Lemmata 2.1 and 2.2, upon taking the union of all flow domains and flows forX as defined

in the preceding remark.

3 Supervector fields and localR1|1-actions

Given a vector field on a classical, ungraded, manifold, the flow map Fe (for S=M,φe= idM) is always a local action ofRwith its usual (and unique up to isomorphism) Lie group structure, the standard addition. The flow maps for vector fields described in the preceding section (taking hereS=M,φ= idM), do not always have the analogous property of being local actions ofR1|1 with an appropriate Lie supergroup structure. Two characterizations of those vector fieldsX =X0+X1that generate a localR1|1-action were found by J. Monterde and O. A. S´anchez-Valenzuela. We will give in this section a short proof of a slightly more general result, whose condition (iii) seems to be more easily veri- fied in practice than those given in [13] (compare Thm. 3.6 and its proof there).

Let us begin by giving a useful two-parameter family of Lie supergroup struc- tures on the supermanifoldR1|1and their right invariant vector fields.

Lemma 3.1. Let a and b be real numbers such that a·b = 0 and µa,b =µ :

(13)

R

1|1×R1|1R1|1 be defined by e

µ(t1, t2) = t1+t2,

µ(t) = t1+t2+aτ1τ2, µ(τ) = τ1+ebt1τ2 . Then

(i) there exists a unique Lie supergroup structure on R1|1 such that the mul- tiplication morphism is given byµa,b,

(ii) the right invariant vector fields on (R1|1, µa,b) are given by the graded vector spaceRD0RD1, where

D0:=∂t+b·τ ∂τ and D1:=∂τ+a·τ ∂t,

and they obey[D0, D0] = 0,[D0, D1] =−bD1 and[D1, D1] = 2aD0. Proof. Both assertions follow by straightforward verifications.

Remarks. (1) It can easily be checked that the above family yields only three non-isomorphic Lie supergroup structures onR1|1, since (R1|1, µa,0) with a6= 0 is isomorphic to (R1|1, µ1,0) and (R1|1, µ0,b) with b6= 0 is isomorphic to (R1|1, µ0,1) and the three multiplicationsµ0,01,0andµ0,1correspond to non- isomorphic Lie supergroup structures onR1|1. Nevertheless it is very convenient to work here with the more flexible two-parameter family of multiplications.

(2) In fact, all Lie supergroup structures onR1|1are equivalent toµ0,01,0or µ0,1. See, e.g., [4] for a direct approach to the classification of all Lie supergroup structures onR1|1.

Definition 3.2. Let G = (G,OG) resp. M = (M,OM) be a Lie supergroup with multiplication morphism µ and unit element e resp. a supermanifold. A

“local action of G onM” is given by the following data:

a collectionΠof pairs of open subsetsπ= (Uπ, Wπ)ofM, whereUπis relatively compact in Wπ, with associated open sub supermanifolds Uπ⊂ Wπ ⊂ M such that {Uπ|π ∈ Π} is an open covering of M, and for all π in Π an open sub supermanifold Gπ⊂ G, containing the neutral element eand a morphism

Φπ:Gπ× Uπ→ Wπ

fulfilling

(1) Φπ◦(e×idUπ) =idUπ, wheree:{pt} → G is viewed as a morphism, (2) Φπ◦(µ×idM) = Φπ◦(idG×Φπ), where both sides are defined, (3) if Uπ∩Uπ 6=∅,Φπ = Φπ on(Gπ∩ Gπ)×(Uπ∩ Uπ).

(14)

Proposition3.3. Let G= (G,OG)resp. M= (M,OM)be a Lie supergroup resp. a supermanifold. Then

(i) a local G-action on M, specified by a set Π and morphisms {(Uπ,Wπ,Gππ)|π ∈ Π}, gives rise to an open sub supermanifold V ⊂ G × Mcontaining{e} × Mand a morphismΦV:V → Msuch that ΦV◦(µ×idM) = ΦV◦(idG×ΦV), (∗) where both sides are defined and such that

Φπ= ΦV on (Gπ× Uπ)∩ V, ∀π∈Π, (∗∗) (ii) an open sub supermanifold V ⊂ G × Mcontaining{e} × Mand a mor- phism Φ :V → Msuch that (∗) is fulfilled, where it makes sense, yields a local G-action on Msuch that (∗∗) holds.

Proof. As in the classical case of ungraded manifolds and Lie groups.

Theorem 3.4. Let M be a supermanifold, X a vector field on M and V ⊂

R

1|1× M the domain of the maximal flowF :V → Msatisfying (injRR1|1)◦(∂t+∂τ)◦F = (injRR1|1)◦F◦X and

F◦injV{0}×M = idM.

Let a andb be real numbers such that a·b= 0. Then the following assertions are equivalent:

(i) the mapF fulfills

(∂t+∂τ+τ(a∂t+b∂τ))◦F=F◦X, (ii) the map F is a local(R1|1, µa,b)-action onM,

(iii) RX0RX1 is a sub Lie superalgebra of TM(M) with commutators [X0, X1] =−bX1 and [X1, X1] = 2aX0.

Proof. Recall thatF fulfills

(injRR1|1)◦∂t◦F = (injRR1|1)◦F◦X0 and (injRR1|1)◦∂τ◦F = (injRR1|1)◦F◦X1. Denoting the projection fromR1|1 toRbyp, we have

id

R

1|1 = p◦(injRR1|1)+τ·p◦(injRR1|1)◦∂τ

which we will write more succinctly as id

R

1|1 = (injRR1|1)+τ·(injRR1|1)◦∂τ. (19)

(15)

Using relation (19) and the equations fulfilled by F we get F◦X = (injRR1|1)◦F◦X+τ·(injRR1|1)◦∂τ◦F◦X

= (injRR1|1)◦(∂t+∂τ)◦F+τ·(injRR1|1)◦F◦X1◦X

= (injRR1|1)◦(∂t+∂τ)◦F + τ·(injRR1|1)◦F

[X1, X0] +X0◦X1+1

2[X1, X1]

.

Since

(injRR1|1)◦F◦X0◦X1 = (injRR1|1)◦∂t◦F◦X1

= ∂t◦(injRR1|1)◦F◦X1

= ∂t◦(injRR1|1)◦∂τ◦F

= ∂t◦∂τ◦F

= ∂τ◦∂t◦F, we arrive at

F◦X = (injRR1|1)◦(∂t+∂τ)◦F+ τ·F

[X1, X0] + 1

2[X1, X1]

+τ·∂τ◦∂t◦F. (20) On the other hand, ifaandb are real numbers, we have, again using (19)

(∂t+∂τ+τ(a∂t+b∂τ))◦F

=

(injRR1|1)+τ·(injRR1|1)◦∂τ

◦(∂t+∂τ)◦F +τ·

a·(injRR1|1)◦∂t+b·(injRR1|1)◦∂τ

◦F

= (inj RR1|1)◦(∂t+∂τ)◦F+τ·∂τ◦∂t◦F +τ·F◦(aX0+bX1).

Thus we have

(∂t+∂τ+τ(a∂t+b∂τ))◦F−F◦X

=τ·F

aX0−1

2[X1, X1] +bX1−[X1, X0]

. (21)

Since F satisfies the initial condition (injV{0}×M) ◦ F = idM, τ ·F is injective and thus Equation (21) easily implies the equivalence of (i) and (iii).

We remark that, in this case, we automatically have a · b = 0 since the Jacobi identity implies that [X1,[X1, X1]] = [[X1, X1], X1] + (−1)1·1[X1,[X1, X1]], i.e., 2a·b·X1= [X1,[X1, X1]] = 0.

(16)

Assume now that a and b are real numbers such that (i) satisfied, and let µ = µa,b be as in Lemma 3.1. We have to show that F is a local action of (R1|1, µ).

Let us define

G:=F◦(idR

1|1×F) :R1|1×(R1|1× M)→ M and

H :=F◦(µ×idM) : (R1|1×R1|1)× M ∼=R1|1×(R1|1× M)→ M.

In order to prove that F is a R1|1-action on, we have to show that G = H.

We observe that G is the integral curve ofX subject to the initial condition F ∈Mor(R1|1× M,M).

Let us prove that the morphismH satisfies the following conditions:

(injR1|1×(R1|1×M)

R×(R1|1×M) )◦(∂t1+∂τ1)◦H = (injR1|1×(R1|1×M)

R×(R1|1×M) )◦H◦X(22) H◦injR{0}×(1|1×(R1|1×M)

R

1|1×M) = F. (23)

Then by the unicity of integral curves we haveH =G.

Equation (23) holds true sinceµ◦injR{0}×1|1×R1|1

R

1|1 = idR

1|1.

Defining D := D0+D1 = ∂t1 +∂τ11(a∂t1 +b∂τ1) and writing inj|t1 for injR1|1×(R1|1×M)

R×(R1|1×M) and using right invariance ofD, we arrive at equation (22) as follows

(injR1|1×(R1|1×M)

R×(R1|1×M) )◦(∂t1+∂τ1)◦H

= (inj |t1)◦D◦H

= (inj |t1)◦(((D⊗idR

1|1)◦µ)×idM)◦F

= (inj |t1)◦((µ◦D)×idM)◦F

= (inj |t1)◦(µ×idM)◦F◦X

= (inj |t1)◦H◦X.

Thus we obtain that (i) implies (ii).

Assume now that (ii) is satisfied, i.e., there exists a Lie supergroup structure onR1|1 with multiplicationµsuch that

F◦(idR

1|1×F) =F◦(µ×idM). (24)

SinceFis a flow forX, with initial conditionφ= idM, the LHS of the preceding equality is a flow for X with initial conditionφ=F, (24) implies

(inj|t1)◦(∂t1+∂τ1)◦(µ×idM)◦F= (inj|t1)◦(µ×idM)◦F◦X.(25)

(17)

By Equation (20), the RHS gives for t1= 0:

(inj|t1=0)◦(µ×idM)◦F◦X = F◦X

= (injRR1|1)◦(∂t+∂τ)◦F +τ·

F

[X1, X0] +1

2[X1, X1]

+∂τ◦∂t◦F

.

Moreover, we have by direct comparison

(∂t1+∂τ1)◦µ = (∂t1+∂τ1)(µ(t))·(µ◦∂t)

+(∂t1+∂τ1)(µ(τ))·(µ◦∂τ).

Thus, ifµ:R1|1×R1|1R1|1is given by

µ(t) = µ(te 1, t2) +α(t1, t21τ2

µ(τ) = β(t1, t21+γ(t1, t22, and upon using (injR{0}×1|1×R1|1

R

1|1 )◦µ= id

R

1|1, we have (injR{0}×1|1×R1|1

R

1|1)(∂t1+∂τ1)◦µ = ((∂t1eµ)(0, t) +α(0, t)τ)·∂t

+ (β(0, t) + (∂t1γ)(0, t)τ)·∂τ. Using again (19), we have

t◦F = (inj RR1|1)◦F◦X0+τ·∂τ◦∂t◦F and

τ◦F = (inj RR1|1)◦F◦X1. Then the LHS of (25) att1= 0 is

(inj|t1=0)◦(∂t1+∂τ1)◦(µ×idM)◦F

= (∂t1µ)(0, t)e ·(injRR1|1)◦F◦X0

+β(0, t)·(injRR1|1)◦F◦X1

+τ· (∂t1µ)(0, t)e ·∂τ◦∂t◦F +α(0, t)·F◦X0

+(∂t1γ)(0, t)·F◦X1

.

Using the obtained identities for its LHS and RHS, the “τ-part” of Equation (25) att1= 0 gives us:

τ·

F

[X1, X0] + 1

2[X1, X1]

+∂τ◦∂t◦F

=τ·

(∂t1µ)(0, t)e ·∂τ◦∂t◦F+α(0, t)·F◦X0+ (∂t1γ)(0, t)·F◦X1

.

(18)

Sinceµ(te 1,0) =t1, we have (∂t1µ)(0,e 0) = 1 and therefore the preceding equa- tion evaluated att= 0 yields

[X1, X0] +1

2[X1, X1] = (∂t1γ)(0,0)·X1+α(0,0)·X0

finishing the proof that (ii) implies (iii).

4 Examples and applications

(4.1) IfX =X0 is an even vector field, the fact that it integrates to a (local) action of R=R1|0is almost folkloristic. The relatively recent proof of [3] - in the case of compact supermanifolds - is close to our approach. A non-trivial (local) action ofR1|0can obviously be extended to a (local) action of (R1|1, µa,b) if and only if a = 0. Of course, the ensueing action of R1|1 will not even be almost-effective, since the positive-dimensional sub Lie supergroup R0|1 acts trivially.

(4.2) Our preferred example of an even vector field gives rise to the exponential map on Lie supergroups.

Let us first recall that an even vector fieldXon a supermanifoldMcorresponds to a sectionσX of the tangent bundleTM → M(see, e.g., Sections 7 and 8 of [15] for a construction ofTMand a proof of this statement, and compare also the remark after Thm. 2.19 in [5]). Given an auxiliary supermanifoldS and a morphismψ:S → M, one calls fori∈ {0,1}

Derψ(OM(M),OS(S))i:=

{D:OM(M)→ OS(S)|D isR-linear and∀f, g∈ OM(M) homogeneous, D(f ·g) =D(f)·ψ(g) + (−1)i·|f|ψ(f)·D(g)},

the “space of derivations of parity i along ψ”. In category-theoretical terms the tangent bundle TM represents then the functor from supermanifolds to sets given by S 7→ {(ψ, D)|ψ∈Mor(S,M) andD∈Derψ(OM(M),OS(S))0} (compare, e.g., Section 3 of [6]).

Let now G = (G,OG) be a Lie supergroup with multiplication µ=µG and neutral element e. We define X in Der (OG×TeG(G×TeG)) to be the even vector field on G × TeG corresponding to the following section σX

of TG ×T(TeG) ∼= T(G ×TeG) → G ×TeG. We denote the zero-section of TG → G by σ0 and the canonical inclusion TeG → TG by ie. Then σX := (T µ◦(σ0×ie),0), where T µ : TG ×TG ∼= T(G × G) → TG is the tangential morphism associated to the multiplication morphism. (For sim- plicity, we write 0 for the zero-section ofT(TeG)→TeGhere and in the sequel.)

参照

関連したドキュメント

In [9], it was shown that under diffusive scaling, the random set of coalescing random walk paths with one walker starting from every point on the space-time lattice Z × Z converges

Inside this class, we identify a new subclass of Liouvillian integrable systems, under suitable conditions such Liouvillian integrable systems can have at most one limit cycle, and

We describe a generalisation of the Fontaine- Wintenberger theory of the “field of norms” functor to local fields with imperfect residue field, generalising work of Abrashkin for

Shen, “A note on the existence and uniqueness of mild solutions to neutral stochastic partial functional differential equations with non-Lipschitz coefficients,” Computers

Applying the representation theory of the supergroupGL(m | n) and the supergroup analogue of Schur-Weyl Duality it becomes straightforward to calculate the combinatorial effect

From the- orems about applications of Fourier and Laplace transforms, for system of linear partial differential equations with constant coefficients, we see that in this case if

We study the theory of representations of a 2-group G in Baez-Crans 2- vector spaces over a field k of arbitrary characteristic, and the corresponding 2-vector spaces of

Lie algebras of vector fields on the plane were also classified (both in real and complex case) by Sophus Lie [4], so that the description of irreducible Lie alge- bras of vector