Polygonal presentations of semisimple tensor categories - Project Euclid

0 downloads 0 Views 702KB Size Report
of C. The free module Z½Sٹ generated by the set SpecًCق has the ring structure ... defined as the multiplicity of Z-component in the decomposition of X nY, i.e.,.
J. Math. Soc. Japan Vol. 54, No. 1, 2002

Polygonal presentations of semisimple tensor categories By Shigeru Yamagami (Received Jun. 15, 2000)

Abstract. A polygonal description of semisimple tensor categories is presented and the rigidity as well as the associated involutions are analysed in terms of this.

Introduction. The combinatorial structures behind tensor categories play fundamental roles in recent studies of quantum symmetries ([1], [2], [5], [11]). Although it is common to impose the strictness on associativity in tensor categories, which does not lose the information thanks to the coherence theorem, an explicit use of associativity constraints is often convenient in concrete computations. A direct manipulation of such combinatorial data, however, can easily lose the navigation. When tensor categories are semisimple, we can divide the relevant structure into two parts: the skeleton information of fusion rule (algebra) and the remaining flesh part, which provides a kind of (non-linear) cohomological information. Viewing this way, semisimple tensor categories can be reconstructed from hom-vector spaces relating three simple objects. These are then pictorially assigned to edges of a triangle and the general hom-sets are expressed in terms of polygons together with triangular decompositions. This kind of geometrical presentation of tensor categories is useful in actual computations to get perspectives: we shall describe the rigidity as well as the associated involutions (if any exists) and show that the variety of duality isomorphisms comes from choices of ‘‘characters’’ of fusion rules. 1. Polygonal vector spaces. A monoidal category ðC; n; I Þ with a, l and r denoting associativity, left-unit and right-unit constraints respectively is called a tensor category over a field K if hom-sets are K-vector spaces and all the relevant operations are K-linear. In what follows, we shall exculisvely deal with tensor categories over the complex number field C and vector spaces are assumed to be C-linear unless otherwise stated. 2000 Mathematics Subject Classification. Primary 18D15; Secondary 46L37. Key Words and Phrases. tensor category, pentagonal relation, rigidity, duality.

62

S. Yamagami

Given tensor categories C and C 0 , a monoidal functor F : C ! C 0 is called an isomorphism if (i) F gives an isomorphism of vector spaces HomðX ; Y Þ ! HomðF ðX Þ; F ðY ÞÞ for any pair ðX ; Y Þ of objects in C and (ii) each object X 0 in C 0 is isomorphic to F ðX Þ for some object X in C. Two tensor categories C and C 0 are said to be isomorphic if there is an isomorphism between them. An object X in a tensor category is simple if EndðX Þ ¼ C1X . A tensor category C is semisimple if the unit object is simple and every object is isomorphic to a direct sum of simple objects. Given a semisimple tensor category C, we denote by S ¼ SpecðCÞ the set of equivalence classes of simple objects in C, which is referred to as the spectrum of C. The free module Z½S Š generated by the set SpecðCÞ has the ring structure defined by X ½X нY Š ¼ dimðHomðZ; X n Y ÞÞ½Z Š; ½ZŠ A S

i.e., the ring Z½S Š has a special basis SpecðCÞ for which structure constants are non-negative integers. Rings, furnished with such bases, are referred to as fusion algebras although this terminology is usually used in a more restrictive sense. For x ¼ ½X Š, y ¼ ½Y Š and z ¼ ½Z Š in SpecðCÞ, a non-negative integer Nzxy is defined as the multiplicity of Z-component in the decomposition of X n Y , i.e., Nzxy ¼ dim HomðZ; X n Y Þ: The totality fNzxy gx; y; z A Spec is, by definition, the fusion rule of C. Let C be a semisimple tensor category with the spectrum S ¼ SpecðCÞ. For the time being, we fix a specific representative containing the unit object I and regard S as a set consisting of simple objects in C. For a finite family fXj g0a jam of simple objects in S, we set   X1    Xm ¼ HomðX1 n    n Xm ; X0 Þ; X0   X0 ¼ HomðX0 ; X1 n    n Xm Þ; X1    Xm which are finite dimensional vector spaces with the pairing     X1    Xm X0  C f  g 7! h f ; gi A C X0 X1    Xm defined by h f ; gi1X0 ¼ f  g;

63

Polygonal presentations of tensor categories

i.e., we have the natural identification     X1    Xm X0 : ¼ X1    Xm X0   X1    Xm The vector space is graphically represented by a polygon of m þ 1 X0 vertices with edges labeled by the sequence fX0 ; X1 ; . . . ; Xm g clockwise, where the initial label X0 plays a special role and we place it at the bottom edge.   XY Triangular vector spaces then form building blocks of Z X; Y;Z A S   X1    Xm polygonal vector spaces such as in the following sense: We begin X0 with the case m ¼ 3 and consider X1 n X2 n X3 . According to two ways of grouping ðX1 n X2 Þ n X3 and X1 n ðX2 n X3 Þ, we have two natural isomorphisms of vector spaces       X1 X2 X12 X3 X1 X2 X3 0 n ! ; X12 X0 X0 X12 A S 0 X23 A S



     X2 X3 X1 X23 X1 X2 X3 n ! : X23 X0 X0

We denote this situation graphically as X2

X2 X3 ! X1

X1 X0

X2 X3 ;

X3 ! X1

X1

X0

X2

X0

X3 X0

and the composite isomorphism X2

X2 X3 ! X1

X1 X0

X3 X0

is referred to as an associativity transformation. For the case m ¼ 4, we have five groupings ððX1 X2 ÞX3 ÞX4 ; ðX1 X2 ÞðX3 X4 Þ, ðX1 ðX2 X3 ÞÞX4 ; X1 ððX2 X3 ÞX4 Þ; X1 ðX2 ðX3 X4 ÞÞ with the associated vector spaces

X0

X0

X0

X0

X0

64

S. Yamagami

Figure 1.

where X2

X3

X1

X4

¼ 0 X12 ; X123



     X1 X2 X12 X3 X123 X4 n n X12 X123 X0

X0 and so on. If we apply associativity transformations to squares inside triangulated pentagons, then we have the commutative diagram in Figure 1, i.e., associativity transformations satisfy the pentagonal relation. Generalizing these, groupings in X1    Xm are parametrized by triangular decompositions of an ðm þ 1Þ-polygon, which in turn give rise to triangular decom  X1    Xm positions of the vector space and composed associativity transformaX0 tions provide the same result as long as the initial and final vector spaces are the same (the coherence for associativity transformations).     IY YI Triangular vector spaces of the form and are non-trivial only   X X IX for X ¼ Y and, if this is the case, they admit distinguished vectors lX A   X XI and rX A given by the unit constraints. X The associativity transformation     ðXIÞY X ðIY Þ T: ! Z Z

Polygonal presentations of tensor categories

65

then takes an especially simple form: 

 XY TðrX n xÞ ¼ lX n x for x A : Z To facilitate the visual notation further, given a triangular decomposition of a polygon and a family of vectors in the accompanied triangular vector spaces we denote its tensor product by putting the vectors inside the belonging triangles: for example, in the triangular decomposition       X1 X2 X3 X1 X2 X12 X3 ¼ 0 n X0 X12 X0 X12 A S with x A HomðX1 X2 ; X12 Þ and h A HomðX12 X3 ; X0 Þ, the associated vector hðx n 1X3 Þ is denoted by X2 X1 x

h

X3 :

X0 We now discuss the reverse process of the construction according to [13]. Suppose that we are given a set S with a distinguished element 1 and a family of   x1 x2 finite-dimensional vector spaces indexed by triplets ðx 0 ; x1 ; x2 Þ in the set S x0 (called triangular vector spaces) and a family of isomorphisms         x2 x3 x1 x2 x1 x23 x12 x3 x1 x2 x3 Tx 0 ! 0 n : 0 n x0 x x x 0 23 12 x23 A S x12 A S indexed by quadruplets ðx 0 ; x1 ; x2 ; x3 Þ in the set S (called associativity transformations) which satisfies the pentagonal identity. Furthermore, we assume that     x1 1x there are distinguished vectors lx A such that , rx A x x       Clx if x ¼ y, Crx if x ¼ y, x1 1x ¼ ¼ f0g otherwise, f0g otherwise, y y and 

 xy Tðrx n sÞ ¼ l y n s; s A s     ðx1Þy xð1yÞ for the associativity transformation T : ! . s s

66

S. Yamagami

The totality of these data is referred to as a monoidal system. Given a monoidal system    xy ; ; Tsxyz ; ls ; rs z s; x; y; z A S we define the dual system by    xy xyz   ; ; Ts ; ls ; rs z s; x; y; z A S     xy xy where is the dual vector space of , z z     1s s1   ; rs A ls A s s are specified by hls ; ls i ¼ 1 ¼ hrs ; rs i and         xy yz tz xt T ¼ Tsxyz : 0 n !0 n s s tAS t tAS t is defined to be the transposed inverse of T ¼ Tsxyz . Given a monoidal system, consider a family X ¼ fX ðsÞgs A S of finitedimensional C-vector spaces with X ðsÞ ¼ f0g except for finitely many s A S. Let Y ¼ fY ðsÞgs A S be another such family. Thinking of these as objects and defining hom-sets by HomðX ; Y Þ ¼ 0 HomðX ðsÞ; Y ðsÞÞ sAS

with the pointwise composition, we obtain a (semisimple) category CðSÞ. Using the assumed triangular vector spaces, the tensor product operation in CðSÞ is introduced by   xy ðX n Y ÞðsÞ ¼ 0 X ðxÞ n Y ðyÞ n s x; y A S for objects X, Y in CðSÞ and ð f n gÞðsÞ ¼ 0 f ðxÞ n gðyÞ n 1 : x; y A S

   xy  xy 0 0 0 X ðxÞ n Y ðyÞ n ! 0 X ðxÞ n Y ðyÞ n s s x; y A S x; y A S 

for f A HomðX ; X 0 Þ and g A HomðY ; Y 0 Þ.

67

Polygonal presentations of tensor categories

The unit object I in CðSÞ is defined to be  C if s ¼ 1, IðsÞ ¼ f0g otherwise. Associativity and unit constraints are defined by aX ; Y ; Z ðsÞ ¼ 0 1X ðxÞ n 1Y ð yÞ n 1ZðzÞ n Tsx; y; z : ðX n Y Þ n Z ! X n ðY n ZÞ x; y; z

with 







xy ððX n Y Þ n ZÞðsÞ ¼ 0 X ðxÞ n Y ðyÞ n ZðzÞ n t x; y; z; t yz ðX n ðY n ZÞÞðsÞ ¼ 0 X ðxÞ n Y ðyÞ n ZðzÞ n t x; y; z; t

 tz  n ; s 



xt n s



;

and  1x  lX ðxÞ : X ðxÞ n C x n lx 7! x A X ðxÞ; x   x1 C x n rx 7! x A X ðxÞ rX ðxÞ : X ðxÞ n x 

with  1x  ðI n X ÞðxÞ ¼ X ðxÞ n ; x 



x1 ðX n IÞðxÞ ¼ X ðxÞ n x



:

It is immediate to see that these in fact satisfy the axioms of tensor category: So far we have defined a semisimple tensor category CðS; TÞ. If we identify an element x A S with the object X in CðS; TÞ defined by  C if s ¼ x, X ðsÞ ¼ f0g otherwise, then the vector space Homðx n y; zÞ is naturally identified with the triangular   xy vector space and we recover the associativity transformation T from the z associativity constraint for CðS; TÞ. Two monoidal systems ðS; TÞ, ðS 0 ; T 0 Þ with the unit elements 1 and 1 0 are said to be equivalent if we can find a bijection f : S ! S 0 such that fð1Þ ¼ 1 0 and     xy fðxÞfð yÞ a family of isomorphisms fzxy : ! which makes the diagram in z fðzÞ Fig. 2 commutative and satisfies fx1x ðlx Þ ¼ lfðxÞ ;

fxx1 ðrx Þ ¼ rfðxÞ :

68

S. Yamagami

Figure 2.

An equivalence ffzxy g is called a gauge transformation if it is associated to the identity map of the index set S. Summarizing these, we have the following. Proposition 1.1 (Reconstruction). Two tensor categories CðS; TÞ and CðS ; T 0 Þ are isomorphic if and only if ðS; TÞ and ðS 0 ; T 0 Þ are equivalent. 0

Example 1.2. Semisimple tensor categories with the fusion rule given by a countable group G are parametrized by elements in the third cohomology group H 3 ðGÞ up to gauge transformations. Proof. In fact, non-trivial triangular vector spaces are 1-dimensional and     g; h g; h given by ; g; h A H. With a choice of associative bases 0 0 ½g; h A gh gh such that ½1; g ¼ lg and ½g; 1 ¼ rg , associativity transformations are specified as T : ½g1 ; g2  n ½g1 g2 ; g3  7! cðg1 ; g2 ; g3 Þ½g2 n g3  n ½g1 ; g2 g3 ; where cðg1 ; g2 ; g3 Þ is a three cocycle of G. The pentagonal identity and the unit constraint condition for T take the form cðg2 ; g3 ; g4 Þcðg1 g2 ; g3 ; g4 Þ1 cðg1 ; g2 g3 ; g4 Þcðg1 ; g2 ; g3 g4 Þ1 cðg1 ; g2 ; g3 Þ ¼ 1 and cðg; 1; hÞ ¼ 1 respectively. The Kelley’s result on unit constraint conditions (see [7]) is then reduced to cð1; g; hÞ ¼ cðg; h; 1Þ ¼ 1, which is a direct consequence of the cocycle condition if we take g2 ¼ 1 or g3 ¼ 1.

69

Polygonal presentations of tensor categories

If we denote by CðG; cÞ the semisimple tensor category associated to a normalized 3 cocycle c on a group G, then it is immediate to see that CðG; cÞ G CðG 0 ; c 0 Þ i¤ there is a group isomorphism f : G ! G 0 such that c and c 0  f are cohomologous in H 3 ðGÞ. r 2. Rigidity. Recall that an object X in a tensor category C is said to be rigid if we can find an object X  and a pair of morphisms e : X n X  ! I , d : I ! X  n X such that the composite morphisms (the associativity isomorphisms being omitted by coherence theorem) 1nd

en1

X ƒƒ! X n X  n X ƒƒ! X ;

dn1

1ne

X  ƒƒ! X  n X n X  ƒƒ! X 

are identities. The object X  is unique up to isomorphisms and is referred to as a dual object of X. The tensor category C is rigid if every object is rigid and isomorphic to a dual of another object. Lemma 2.1 (cf. [6]). For a simple object X in a rigid (semisimple) tensor category, its dual object X  is again simple and X itself is a dual of X  . Proof. If X  is not simple, X  G Y  l Z  by the rigidity of C and then X G Y l Z by the uniqueness of (pre)dual objects. Thus X  is simple if X is so. By Frobenius reciprocity and the semisimplicity of C, we have HomðX ; X  Þ G HomðX  n X ; IÞ G HomðI; X  n X Þ G EndðX Þ; whence there is a non-trivial morphism X ! X  . Since X  is simple as a dual of the simple object X  , they are actually isomorphic. r As a consequence of the above lemma, an involution is defined on the spectrum set S of a rigid tensor category by ½X Š  ¼ ½X  Š, which satisfies the duality relation  1 if x ¼ y  , xy N1 ¼ 0 otherwise. The fusion algebra C½S Š is a *-algebra by extending the involution on the set S (see [12], [4], [14] for more information on fusion algebras in the present context). Assume that in the tensor category CðS; TÞ the fusion set S is furnished with an involution  satisfying the duality relation. For an object X in CðS; TÞ and a morphism f : X ! Y , we then define the object X  by X  ðsÞ ¼ ðX ðs  ÞÞ  ð¼ the dual vector space of X ðs  ÞÞ

70

S. Yamagami

and the morphism tf : Y  ! X  by ð tf ÞðsÞ ¼ tf ðs  Þ : Y ðs  Þ  ! X ðs  Þ  ; where tf ðs  Þ denotes the transposed map of f ðs  Þ. The operation X  , tf gives a contravariant functor from CðS; TÞ into itself. Lemma 2.2 (Local Rigidity). The semisimple tensor category CðS; TÞ is rigid  if and only if S admits an involution  satisfying N1x y ¼ dx; y and hd n rs ; Tðe n ls Þi ¼ hd n ls ; T 1 ðe n rs  Þi 0 0      ss ss with s A S. for 0 0 e A , 00d A 1 1     xy yx Proof. For e A , we have and d A 1 1 ðe n 1Þa 1 ð1 n dÞ ¼ hd n rx ; Tðe n lx Þi1x ; ð1 n eÞaðd n 1Þ ¼ hd n l y ; T

1

ðe n ry Þi1y

with obvious identifications on unit constraints. The condition in question is then equivalent to the rigidity of simple objects in CðS; TÞ, which in turn implies      ss ss the rigidity for arbitrary objects: For each s A S, choose e s A and ds A 1 1 so that hds n rs ; Tðe s n ls Þi ¼ hds n ls ; T 1 ðe s n rs  Þi ¼ 1: For any object X in CðS; TÞ, we then define morphisms eX : X n X  ! I, dX : I ! X  n X by    X ss  C 0 xs n hs n ss 7! hxs ; hs ihe s ; ss i A C ¼ Ið1Þ; 0 X ðsÞ n X ðsÞ n 1 sAS sAS s  s s  I ð1Þ ¼ C C 1 7! 0 dX ðsÞ n ds A 0 X ðsÞ n X ðsÞ n : 1 s sAS 



Here dX ðsÞ ¼

X

xj n xj

j

with fxj g a basis in X ðsÞ and fxj g its dual basis. It is straightforward to check that these in fact give a rigidity pairing between X and X  . r Example 2.3. The tensor category CðG; cÞ associated to a 3 cocycle c of a group G is rigid.

71

Polygonal presentations of tensor categories

Proof. The fusion rule set S is the group G itself and the involution is given  1    g; g 1 g ;g   1 1 1 by g ¼ g for g A G. Let e ¼ ½g; g Š A and d ¼ ½g ; gŠ A 1 1 in Lemma 2.2. Then hd n rg ; Tðe n lg Þi ¼ cðg; g 1 ; gÞ;

hd n lg 1 ; T

1

ðe n rg 1 Þi ¼ cðg 1 ; g; g 1 Þ

1

shows that the rigidity is equivalent to cðg; g 1 ; gÞ ¼ cðg 1 ; g; g 1 Þ 1 ; whence it follows from the cocycle condition of c if we consider the condition dcðg; g 1 ; g; g 1 Þ ¼ 1. r Let CðS; TÞ be a semisimple tensor category described by a monoidal system ðS; TÞ with S a fusion rule set and T a system of associativity transformations on S. Assume that the tensor category CðS; TÞ is rigid, i.e., S admits an involution  satisfying the duality relation and the condition in Lemma 2.2.      ss ss We then choose e s A and extend it to rigidity pairings eX : , ds A 1 1 X n X  ! I, dX : I ! X  n X as described in the proof of Lemma 2.2. The following is immediate from definitions. Lemma 2.4. The contravariant functor ðX ; f Þ 7! ðX  ; tf Þ is compatible with the rigidity pairing fðeX ; dX Þg: eY ð f n 1Þ ¼ eX ð1 n tf Þ or equivalently ð1 n f ÞdX ¼ ð tf n 1ÞdY for f : X ! Y . By the uniqueness of rigidity pairings, we can define isomorphisms cX ; Y : Y n X  ! ðX n Y Þ  by the commutativity of the diagram 

1neY

X n Y n?Y  n X  ƒƒƒ! X n?X  ? ? ?eX ; 1ncX ; Y ? y y

X n Y n ðX n Y Þ  ƒƒƒ! eXY

I

where parentheses are omitted thanks to the coherence theorem. By the above lemma, the family fcX ; Y g is natural in X, Y and hence it is determined by isomorphisms cx; y for x, y A S as cX ; Y ðsÞ ¼ 0 1 n cx; y ðsÞ : ðY  n X  ÞðsÞ ! ðX n Y Þ  ðsÞ x; y A S

72

S. Yamagami

with 

y x  ðY  n X  ÞðsÞ ¼ 0 Y ðyÞ n X ðxÞ n s x; y A S 





;

 xy ðX n Y Þ ðsÞ ¼ 0 Y ð yÞ n X ðxÞ n  : s x; y A S 







Here the isomorphism 

y x  cx; y ðsÞ : s





xy !  s



is specified in the follwoing way: Starting with a vector a n b n s in the vector space          xy y x ss ððxyÞðy  x  ÞÞð1Þ ¼ 0  n n ; s 1 sAS s the evaluation by the morphism 1ncx; y

e xy

ðxyÞðy  x  Þ ƒƒƒ! ðxyÞðxyÞ  ƒƒƒ! 1

gives

ha; cx; y ðsÞbihes  ; ss i;

whereas the evaluation by the morphism (see Fig. 3) 1na

a

1

1nðe y n1Þ

ðxyÞðy  x  Þ ƒƒƒƒƒ! xð yðy  x  ÞÞ ƒƒƒƒƒ! xððyy  Þx  Þ ƒƒƒƒƒ! xð1x  Þ 1nl

ex

ƒƒƒƒƒ! xx  ƒƒƒƒƒ! 1

gives the expression

hð1 n T

1

ÞðT n 1Þðey n lx  n ex Þ; a n b n si:

Equating these, we get an explicit formula which determines cx; y ðsÞ: ha; cx; y ðsÞbi ¼ hð1 n T

1

ÞðT n 1Þðey n lx  n ex Þ; a n b n es  i;

where es  A Homð1; s sÞ is specified by hes  ; es  i ¼ 1.

Figure 3.

73

Polygonal presentations of tensor categories

If we change fe s gs A S into fe s0 ¼ fðs  Þ 1 e s gs A S with fðs  Þ A C  and cx; y into cx;0 y , then cx;0 y satisfies ha; cx;0 y ðsÞbifðsÞ

1

¼ fð y  Þ 1 fðx  Þ 1 hð1nT

1

ÞðT n1Þðey n lx  nex Þ; a n b n es  i

and the comparison with the equation for cx; y yields cx;0 y ðsÞ ¼

fðsÞ cx; y ðsÞ: fðx  Þfð y  Þ

If we define a natural family of isomorphisms ffX : X  ! X  g by fX : 0 xðsÞ 7! 0 fðsÞxðsÞ; sAS

sAS

then it intertwines between cX ; Y and cX0 ; Y : cX ; Y

  Yn ? X ƒƒƒ! ðX n ?YÞ ? ? ?fX nY fY nfX ? : y y

ðX n Y Þ  Y  n X  ƒƒƒ! 0 cX ; Y

In other words, the monoidal functor ðX  ; tf ; cX ; Y Þ is unique up to natural equivalences. For an object X in CðS; TÞ, we defined dX  : I ! X  n X  , whereas we have t

eX

 

cX 1; X 

I ƒƒƒ! ðX n X Þ ƒƒƒ! X  n X  :

Lemma 2.5. For an object X in CðS; TÞ, t

eX ¼ cX ; X   dX  : I ! ðX n X  Þ  :

Proof. If we regard t eX as an element in ðX n X  Þ  ð1Þ ¼ ðX n X  Þð1Þ  , then t

 xx  eX ¼ 0 dX ðxÞ  n ex A 0 X ðxÞ n X ðxÞ n ; 1 xAS xAS 



where the second dual of (finite-dimensional) vector spaces are identified with the original ones. Similarly dX  is identified with an element in ðX  n X  Þð1Þ by    xx  dX  ¼ 0 dX ðxÞ  n dx  A 0 X ðxÞ n X ðxÞ n 1 xAS xAS and then cX ; X   dX  takes the form  xx  0 dX ðxÞ  n cx; x  ðdx  Þ A 0 X ðxÞ n X ðxÞ n : 1 xAS xAS 



74

S. Yamagami



 xx  Thus we need to show ex ¼ cx; x  ðdx  Þ in . According to the morphism 1 1ncx; x 

the vector

e xx 

ðx n x  Þ n ðx n x  Þ ƒƒƒƒ! ðx n x  Þ n ðx n x  Þ  ƒƒƒƒ! I; 

xx dx  n dx  n r1 A 1

  



xx n 1

  



11 n 1



0

B B HB Bx @

x

1

x

C C C C x A

1

is mapped by 1 n cx; x  n 1 to the vector         11 xx xx  dx  n cx; x  ðdx  Þ n r1 A n n 1 1 1 and then evaluated by exx  , resulting in the scalar hdx  ; cx; x  ðdx  Þi:

On the other hand, the vector dx  n dx  n r1 is transformed by an associativity transformation into dx  n Tðdx  n r1 Þ, which is equal to 1 0 x x C          B C B xx x 1 xx  C B d x  n rx  n d x  A n n HB C  x 1 1 x @x A 1

by Kelly’s theorem ([7]) and then again by an associativity transformation into 1 0  x x C          B C B xx 1x x x  1 C: B H n n T ðdx  n rx Þ n dx  A B  C 1 1 x x A @x 1

The last vector is evaluated according to the morphism ð1ne x  Þn1

rx n1

ex

ðx n ðx  n xÞÞ n x  ƒƒƒƒƒ ƒ! ðx n 1Þ n x  ƒƒƒƒƒ ƒ! x n x  ƒƒƒƒƒ ƒ! I

into the scalar

hex n lx  ; T

1

ðdx  n rx Þihex ; dx  i ¼ hex ; dx  i:

(In the last line, we used the local rigidity.) Comparing these, we have ex ¼ cx; x  ðdx  Þ, proving the assertion.

r

Polygonal presentations of tensor categories

75

3. Duality. In this section, we assume that the antimonoidal functor f 7! tf is supplemented to an involution by duality isomorphisms fdX : X ! X  gX A ObjectðCÞ , i.e., the family fdX g is natural in X and multiplicative in the sense that the following diagrams commute, dX

 X ? ƒƒƒ! X? ? ? ? tt f f? y y ;

Y ƒƒƒ! Y  dY

Xn ?Y ? d? y

d nd

 ƒƒƒ! X  n ? Y ? ?c ; y

ðX n Y Þ ƒƒƒ! ðY  n X  Þ  t c

and tdX ¼ dX 1 : X  ! X  (see [16] for more information on this). By the naturality in X, d takes the form X ðsÞ C x 7! Ds x  A X ðsÞ ¼ X  ðsÞ for s A S with Ds A C  , where x 7! x  denotes the natural identification X ðsÞ ¼ X ðsÞ . The multiplicativity of d is then reduced to the commutativity of     Dx Dy xy xy ƒƒƒ! s s ? ? ? ? ?cy  ; x  ðsÞ Ds ? y y      xy y x ƒƒƒ! t cx; y ðsÞ s s

(note that t cx; y ðsÞ ¼ t ðcx; y ðs  ÞÞ), i.e., Dst cx; y ðsÞ

    xy  y x ¼ Dx Dy cy  ; x  ðsÞ : ! : s s 

If fDs0 gs A S gives a duality d 0 for the antimultiplicativity fcx;0 y g, then it satisfies Ds0t cx;0 y ðsÞ ¼ Dx0 Dy0 cy0  ; x  ðsÞ; which is equivalent to Dx0 Dy0 fðsÞ Ds0 fðs  Þ : ¼ Ds fðx  Þfðy  Þ Dx Dy fðxÞfðyÞ As a solution, we may take Ds0 ¼

fðsÞ Ds : fðs  Þ

76

S. Yamagami

In other words, there is a natural correspondance between duality isomorphisms for di¤erent antimultiplicativities and we may restrict ourselves to the 0 0 case ðX  ; tf ; cÞ ¼ ðX  ; t f ; c 0 Þ for the studies of possibility of rigidity-compatible involutions: the information on involutions is then stacked up to the choice of duality isomorphisms.   xx Remark. Since the vector space is 1-dimensional, the linear map 1       xx xx ! Cx; x  ð1Þ : 1 1 and its transposed map t Cx; x  ð1Þ coincides. Thus we have D1 ¼ Dx Dx  for any x A S. In particular, we see that D1 ¼ 1;

Dx Dx  ¼ 1:

The condition tdX ¼ dX 1 is a consequence of the multiplicativity. Let d and d 0 be dualities based on a common antimultiplicative functor ðX  ; tf ; cÞ. If we define a family of isomorphisms ffX : X ! X g by dX0 ¼ dX  fX , then it is natural and multiplicative: f fX ¼ fY f for f : X ! Y and fX nY ¼ fX n fY . By the naturality, such a family is determined by scalars ffðsÞgs A S , fs A C  , defined by fs ¼ fðsÞ1s and the multiplicativity is reduced to the condition   xy fðsÞ ¼ fðxÞfð yÞ if 0 f0g: s In particular, it satisfies fð1Þ ¼ 1; fðs  Þ ¼ fðsÞ 1 and gives a character of the fusion rule set S. (This is di¤erent from the notion of character of the fusion algebra C½S Š.) Conversely, given a character ffðsÞgs A S of S, it induces a natural and multiplicative family of isomorphisms by fX ðxÞ : X ðxÞ C x 7! fðxÞx A X ðxÞ: By point-wise multiplication, the set GðSÞ of characters of S becomes a commutative group and choices of dualities for the rigidity-compatible antimultiplicative functor ðX  ; tf ; cÞ are parametrized by elements in GðSÞ. Since the antimultiplicativity fcX ; Y g is uniquely determined by the choice feX g, the isomorphism classes of the structure ðfeX g; fdX gÞ is parametrized by the set GðSÞ as well.

77

Polygonal presentations of tensor categories

We now proceed into the description of Frobenius duality ([15], [16], cf. also [3]), which is a family feX : X n X  ! Ig together with an involution ðX  ; tf ; c; d Þ satisfying (i) (Multiplicativity) ð1neY Þn1

ðX n ðY n Y  ÞÞ n X  ƒƒƒƒƒƒ! ? ? ? ? y

ðX n IÞ n X 

rX n1

ƒƒƒƒƒƒ! X n X  ? ? ? ?eX : y

ðX n Y Þ n ðY  n X  Þ ƒƒƒƒƒƒ! ðX n Y Þ n ðX n Y Þ  ƒƒƒƒƒƒ!

I

eX nY

1ncX ; Y

(ii) (Naturality)

f n1

  Xn ? Y ƒƒƒ! Y n ?Y ? ? ? eY : 1n tf ? y y

X n X  ƒƒƒ! eX

I

(iii) (Faithfulness) The map EndðX Þ C f 7! eX ð f n 1Þ is faithful. (iv) (Neutrality) If we define left and right dimensions (denoted by ldimðX Þ and rdimðX Þ respectively) of an object X by the following composites t

eX

 

cX 1; X 

dX 1 n 1

eX

I ƒƒƒƒ! ðX n X Þ ƒƒƒƒ! X  n X  ƒƒƒƒ! X n X  ƒƒƒƒ! I; t

eX 

cX 1 ; X 

dX 1 n 1

eX 

I ƒƒƒƒ! ðX  n X  Þ  ƒƒƒƒ! X  n X  ƒƒƒƒ! X  n X  ƒƒƒƒ! I;

then they coincide (the common scalar is called the dimension of X and     denoted by dimðX Þ). ss Starting with a choice of rigidity pairings e s A , we enlarge it to the 1 family feX g and define an antimultiplicativity fcX ; Y g as discussed before. Then the first three properties, multiplicativity, naturality and faithfulness, are satisfied by just giving a duality fdX : X ! X  g for ðX  ; tf ; cÞ. Thus whether it gives a Frobenius duality depends on the validity of neutrality. Lemma 3.1. A duality family fDs gs A S for a local rigidity family fe s ; ds gs A S      ss ss with e s A gives a Frobenius duality if and only if , ds A 1 1 Ds2 ¼

hds  ; e s i : hds ; es  i

78

S. Yamagami

Proof. By the naturality, it is enough to check the condition for simple X. In that case, by Lemma 2.5, the composite morphisms in the neutrality are reduced to ds 

ds

1

es

1 ƒƒƒ! s s  ƒƒƒ! ss  ƒƒƒ! 1;

and the equality of these is

ds

1nds

es 

1 ƒƒƒ! s s ƒƒƒ! s s  ƒƒƒ! 1

Ds 1 hds  ; e s i ¼ Ds hds ; es  i:

r

As a conclusion of our discussions, we have the following. Theorem 3.2. Isomorphism classes of Frobenius dualities in the tensor category CðS; TÞ is, if it exists, parametrized by characters of S taking values in fG1g. Example 3.3. Let G be a (discrete) group and c be a normalized cyclic 3cocycle (see Appendix A). The associated tensor category CðG; cÞ is reflexive and reflexivity is parametrized by characters of G. For a cyclic cocycle c, d ¼ fDg gg A G itself is a character (generally, the parameter space is a principal homogeneous space of the character group of G ) and the dimension function is given by cðg; g 1 ; gÞ ldimðgÞ ¼ : Dg Moreover, a Frobenius duality is defined by choosing Dg ¼ cðg; g 1 ; gÞ so that dimðgÞ ¼ 1 for g A G.      g g gg 1 Proof. Let e g ¼ ½g; g Š A with the accompanied dg A given 1 1 by dg ¼ cðg; g 1 ; gÞ 1 ½g; g 1 нg 1 ; gŠ  : Given g, h A G, let       h g g; h 1 1 1 1  C ½h ; g Š 7! m½g; hŠ A cg; h ðh g Þ : h 1g 1 gh be the non-trivial part of cg; h . If we start with the vector 

ðgðhh 1 ÞÞg ½h; h Š n ½g; 1Š n ½g; g 1 Š A 1 1 



 1 

;

its evaluation by e g  e h is equal to 1, whereas repetitions of associativity transformations show that it corresponds to the vector     g Þ ðghÞðh 1    cðg; h; h 1 Þcðgh; h 1 ; g 1 Þ ½g; hŠ n ½h 1 ; g 1 Š n ½gh; h 1 g 1 Š A 1

Polygonal presentations of tensor categories

79

in the pentagonal vector space. Now the evaluation by e gh  cg; h gives the result mcðg; h; h 1 Þcðgh; h 1 ; g 1 Þ 1 : Therefore we have m¼

cðgh; h 1 ; g 1 Þ : cðg; h; h 1 Þ

Thus the equation for d takes the form cðgh; h 1 ; g 1 Þ cðh 1 g 1 ; g; hÞ Dgh ¼ Dg Dh : cðg; h; h 1 Þ cðh 1 ; g 1 ; gÞ To solve this equation, we assume the cyclic symmetry on the cocycle c. Then cðgh; h 1 ; g 1 Þ ¼ cðh 1 g 1 ; g; hÞ ¼ 1;

cðg; h; h 1 Þ ¼ cðh 1 ; g 1 ; gÞ;

which reduces the equation to Dgh ¼ Dg Dh ;

g; h A G;

i.e., fDg gg A G is a character of G. The left dimension is calculated by dimðgÞ1I ¼ e g ðdg 1 n 1Þdg  ¼ cðg 1 ; g; g 1 Þ 1 Dg 1 :

r

Remark. For a normalized cyclic 3-cocycle c, we have cðgh; ðghÞ 1 ; ghÞ ¼ cðg; g 1 ; gÞcðh; h 1 ; hÞ: Proof. From the cocycle relation dcðg; h; h 1 g 1 ; ghÞ ¼ 1, we have cðgh; h 1 g 1 ; ghÞ ¼ cðh; h 1 g 1 ; ghÞcðg; g 1 ; ghÞcðg; h; h 1 g 1 Þ; which, combined with the cyclicity of c (g1 g2 g3 g4 ¼ 1 implies cðg1 ; g2 ; g3 Þ ¼ cðg2 ; g3 ; g4 Þ 1 ) of the form cðg; h; h 1 g 1 Þ ¼ 1;

cðh; h 1 g 1 ; ghÞ ¼ cðh 1 ; h; h 1 g 1 Þ 1 ;

produces cðgh; h 1 g 1 ; ghÞ ¼

cðg; g 1 ; ghÞ : cðh 1 ; h; h 1 g 1 Þ

By the cocycle relations dcðg; g 1 ; g; hÞ ¼ 1 ¼ dcðh 1 ; h; h 1 ; g 1 Þ, we have cðg; g 1 ; ghÞ ¼ cðg; g 1 ; gÞcðg 1 ; g; hÞ; cðh 1 ; h; h 1 g 1 Þ ¼ cðh 1 ; h; h 1 Þcðh; h 1 ; g 1 Þ;

80

S. Yamagami

whence cðgh; h1 g1 ; ghÞ ¼

cðg; g1 ; gÞ cðg1 ; g; hÞ : cðh1 ; h; h1 Þ cðh; h1 ; g1 Þ

Finally, we apply the cyclicity of c in the form cðg1 ; g; hÞ ¼ cðh1 ; g1 ; gÞ1 ¼ cðh; h1 ; g1 Þ:

r

Example 3.4. Let G be a finite abelian group. Given a symmetric nondegenerate bicharacter s : G  G ! T and a real number t satisfying t 2 ¼ jGj1 , we can define a tensor category Cðs; tÞ such that S ¼ G t fmg with the fusion P rule am ¼ m ¼ ma and m 2 ¼ a A G a other than the group operation among elements in G (see [13]). Then the tensor category Cðs; tÞ is reflexive and there are two choices of reflexivity, both of which give rise to Frobenius dualities. More precisely, we have Da ¼ 1 for a A G, Dm ¼ G1 and the dimension function is given by dimðaÞ ¼ 1;

dimðmÞ ¼

1 A fGjGj 1=2 g: tDm

Proof. Recall that non-trivial triangular vector spaces are given by         a; b a; m m; a m; m ¼ C½a; b; ¼ C½a; m; ¼ C½m; a; ¼ C½a ab m m a and associativity transformations are given by b

b c C ½a; b n ½ab; c 7! ½b; c n ½a; bc A a

a abc

c; abc

b

b m C ½a; b n ½ab; m 7! ½b; m n ½a; m A a

a m

m; m

m

m b C ½a; m n ½m; b 7! sða; bÞ½m; b n ½a; m A a

a

b;

m

m a

a b C ½m; a n ½m; b 7! ½a; b n ½m; ab A m

m m

b; m

81

Polygonal presentations of tensor categories

m

m m C ½a; mŠ n ½bŠ 7! ½a 1 bŠ n ½a; a 1 bŠ A a

a

m;

b

b

a

a m C ½m; aŠ n ½bŠ 7! sða; bÞ½a; mŠ n ½bŠ A m

m

m;

b

b

m

m a C ½ba 1 Š n ½ba 1 ; aŠ 7! ½m; aŠ n ½bŠ A m

m

a;

b

b m

m m C ½bŠ n ½b; mŠ 7!

m

X

aAG

m

t ½aŠ n ½m; aŠ A m sða; bÞ

m: m

Note that, if we denote by f½aŠ  g the dual basis of f½aŠg and so on, then we have 0

B @a 0

B @m 0

1

m

C B     b A C ½a; mŠ n ½m; bŠ 7! ha; bi 1 ½m; bŠ n ½a; mŠ A @ a

m a

b m

B @m

m

0

1

C B     m A C ½m; aŠ n ½bŠ 7! ha; bi 1 ½a; mŠ n ½bŠ A @ m 1

m a

0

0

X C B   tha; bi½aŠ  n ½m; aŠ  A @ m m A C ½bŠ n ½b; mŠ 7!

b m

a

m

m

1

C bA; 1

C mA; 1

C mA:

As seen in [13], the tensor category Cðs; tÞ is rigid. With the choice of pairings  aa  ea ¼ ½a; a Š A ; 1 

1



 mm e m ¼ ½1Š A ; 1

the associated copairings are given by a a da ¼ ½a 1 ; aŠ A 1 





;



mm dm ¼ t 1 ½1Š A 1 



:

82

S. Yamagami

Based on these data, we can calculate the maps cx; y ðsÞ:       a a; b b  ca; b ðb 1 a 1 Þ : C ½b 1 ; a 1 Š 7! ½a; bŠ A ; b 1a 1 ab     ma am  C ½m; a 1 Š 7! ½a; mŠ A ca; m ðmÞ : ; m m      ma a m  1 C ½a ; mŠ 7! ½m; aŠ A cm; a ðmÞ : ; m m     mm mm 1   C ½a Š 7! t½aŠ A cm; m ða Þ : : a a In fact, if we start with the vector 0

m B B ½bŠ  n ½b; mŠ  n ½1Š  A B Bm @

m

1

1

C C C; mC A

the evaluation by e m ð1 n e m n 1Þ gives d b; 1 , whereas the vector is changed by associativity transformations into 1 0 m m C B X C B   1  1 C: sða; bÞ½aŠ n ½a Š n ½a ; aŠ A B t Bm mC A @ aAG 1

If we define mðaÞ A C by cm; m ða  Þð½a 1 Š  Þ ¼ mðaÞ½aŠ, then the last vector goes to        X mm mm a a   t n mðaÞsða; bÞ½aŠ n ½aŠ n ½a 1 ; aŠ A 0 n a 1 a aAG aAG and its evaluation turns out to be X t mðaÞsða; bÞ: aAG

Comparing this with the Kronecker delta d b; 1 , we obtain mðaÞ ¼ Similarly for other cx; y ðs  Þ’s.

1 ¼ t: tjGj

Polygonal presentations of tensor categories

83

It is now immediate to write down the equations for fDs g: Dab ¼ Da Db ;

Da Dm ¼ Dm ;

Da ¼ Dm2

with the solutions given by Da 1 1;

Dm A fG1g:

The left dimension is then calculated by dimðaÞ ¼ ea ðda 1 n 1Þda  ¼ 1; dimðmÞ ¼ e m ðdm 1 n 1Þdm  ¼

1 ; tDm

which is automatically *-invariant and hence the involution in consideration gives rise to a Frobenius duality. r 4. Positivity. Now we shall restrict ourselves to tensor categories possessing positivity, i.e., C -tensor categories. A category C is a C  -category if each HomðX ; Y Þ is a Banach space with a conjugate-linear involution HomðX ; Y Þ C f 7! f  A HomðY ; X Þ such that k f  f k ¼ k f k2. A tensor category C is, by definition, a C  -tensor category if it is a C  category at the same time in such a way that all the monoidal structures respect the *-operation (the unit and associativity constarints are then unitaries). A monoidal functor F between C  -tensor categories is called a C  -monoidal functor if it preserves the *-operation: F ð f Þ  ¼ F ð f  Þ for f : X ! Y and the multiplicativity mX ; Y : F ðX Þ n F ðY Þ ! F ðX n Y Þ is unitary. Given a semisimple C  -tensor category C with the spetrum set S represented   xy by simple objects, each triangular vector space is a finite-dimensional Hilbert z space with the inner product defined by 

ðxjhÞ1z ¼ hx  ;

x; h A ½xyŠ:     x1 1x The elements (of unit constarints) lx A are then unit and rx A x x vectors. Moreover, the associativity transformations y z ! x

T: x w are unitaries.

y z w

84

S. Yamagami

A monoidal system satisfying these conditions is referred to as a C  -monoidal system. Given a C  -monoidal system ðS; TÞ, we can reconstruct the C  -tensor category C: an object in C is a family X ¼ fX ðsÞgs A S of finite-dimensional Hilbert spaces with X ðsÞ ¼ f0g for all but finitely many s A S. Hom-sets are then defined by HomðX ; Y Þ ¼ 0 HomðX ðsÞ; Y ðsÞÞ; sAS

which is a vector space of linear maps between Hilbert spaces 0s X ðsÞ and 0s Y ðsÞ, whence it admit the norm as well as the *-operation in the obvious manner. It is now immediate to check that the unit and associativity constraints defined before are unitaries. In the C  -tensor category CðS; TÞ, the operations X 7! X  and f 7! tf are defined exactly as in O2. It is then immediate to check the relation ð tf Þ  ¼ t ð f  Þ for a morphism f : X ! Y . We can apply the discussions on rigidity and (Frobenius) duality to C  -tensor categories as well. For C  -tensor cateogries, however, it is natural to require the positivity in Frobenius duality: a Frobenius duality feX : X n X  ! Ig with an involution ðX  ; tf ; c; d Þ is positive if ðdX 1 n 1ÞcX 1; X  t eX ¼ eX (I  being identified with I ). We proved in [16] the existence and the uniqueness of positive Frobenius duality for a rigid C  -tensor category with simple unit object. The key notion there is the balancedness of rigidity pairs: a rigidity pair fe : X n X  ! I; d : I ! X  n X g is said to be balanced if eða n 1Þe  ¼ d  ð1 n aÞd for any a A EndðX Þ: By Lemma 2.5, the positivity is equivalent in the present context to requiring Ds 1 ds  ¼ e s ;

s A S:

For s 0 s  , we can choose balanced pairings e s , es  so that ds  ¼ e s , which     ss ss  forces Ds to be 1 by positivity. For s ¼ s , let e A be a and d A 1 1 balanced rigidity pair. Since both of e  and d are non-trivial vectors in the 1  ss dimensional vector space , they are proportional 1 d ¼ le  ; where l A C satisfies jlj by kdk ¼ kek (the pair ðe; dÞ being balanced). Since the balanced pair ðe; dÞ is unique up to the phase choice ðe iy e; e iy dÞ ðy A RÞ, the phase factor l does not depend on the choice of balanced pairs and is characterisrtic of

Polygonal presentations of tensor categories

85

s ¼ s  . Thus it must coincides with the duality factor Ds for a positive duality. On the other hand, Ds satisfies Ds2 ¼ 1 for s ¼ s  ; the phase factor is either þ1 or 1. Definition 4.1. Let S be the spectrum (fusion rule set) of a rigid C  -tensor category with simple unit object. A self-dual element s ¼ s  in S is said to be real or pseudoreal according to Ds ¼ 1 or Ds ¼ 1, namely l > 0 or l < 0, where l A R is given by     ss ss  d ¼ le ; e A ; dA 1 1 with ðe; dÞ a rigidity pair for s. Now we can describe a positive Frobenius duality in terms of polygonal     ss  is presentations. Given a rigid C -monoidal system, a family e s A 1 sAS called a balanced system if ðe s ; es  Þ is a rigidity pair for s 0Gs  and ðe s ;Ge s Þ is a rigidity pair for s ¼ Gs  , where s ¼ Gs  means that s is real or pseudoreal according to the signature. Theorem 4.2. In a rigid C  -monoidal system, we can always find a balanced system of duality pairings. Given a balanced system fe s gs A S , its canonical extension feX g provides a positive Frobenius duality together with the involution ðX  ; tf ; cX ; Y ; dX Þ, where the duality isomorphism fdX : X ! X  g is specified by  1 if s A S is pseudoreal, Ds ¼ 1 otherwise. Appendix A. Group Cohomology. In the text, we have occasionally used the cyclic normalization of cocycles in group cohomology, which would be a known fact but we have failed in finding literatures; we shall give an account here for completeness. Let G be a discrete group and q

q

   ! C1 ! C 0 ! Z be a projective resolution of the trivial G-module Z. For an abelian group A with a G-action, the cohomology groups are defined by H n ðG; AÞ ¼ H n ðHom G ðC:; AÞÞ; which is independent of the choice of projective resolutions.

86

S. Yamagami

Commonly used is the standard resolution, where Cn ¼

0

Zðg0 ; g1 ; . . . ; gn Þ

g0 ; g1 ;...; gn A G

is a free Z-module with the G-action defined by gðg0 ; g1 ; . . . ; gn Þ ¼ ðgg0 ; gg1 ; . . . ; ggn Þ: The di¤erential and a chain contraction s (satisfying qs þ sq ¼ id ) are given by qðg0 ; g1 ; . . . ; gn Þ ¼

n X ð1Þ i ðg0 ; . . . ; g^i ; . . . ; gn Þ; i¼0

sðg0 ; g1 ; . . . ; gn Þ ¼ ð1; g0 ; g1 ; . . . ; gn Þ: As a ZðGÞ-basis (the so-called bar basis), we can choose jg1 jg2 j    jgn j ¼ ð1; g1 ; g1 g2 ; . . . ; g1 g2    gn Þ 1 1 (ðg0 ; g1 ; . . . ; gn Þ ¼ g0 jg1 0 g1 jg1 g2 j    jgn1 gn j). Note that

qðjg1 jg2 j    jgn jÞ ¼ g1 jg2 j    jgn j þ

n1 X ð1Þ i jg1 j    jgi giþ1 j    jgn j i¼1

þ ð1Þ n jg1 j    jgn1 j and sðgjg1 j    jgn jÞ ¼ jgjg1 j    jgn j: For g0 , g1 ; . . . ; gn A G, define the wedge product by g0 5g1 5  5gn ¼ Pnþ1 ðg0 ; g1 ; . . . ; gn Þ ¼

X 1 eðsÞðgsð0Þ ; gsð1Þ ; . . . ; gsðnÞ Þ; ðn þ 1Þ! s A S nþ1

which is an element in Q nZ C n . These, when parametrized by unordered n þ 1-tuple fg0 ; g1 ; . . . ; gn g, form a Z-basis of the image Dn of C n under the projection Pnþ1 in Q nZ C n and hence X Dn ¼ Zg0 5g1 5  5gn is a free ZðGÞ-submodule of Q nZ C n . Define a ZðGÞ-linear map d : Dn ! Dn1 by d ¼ Pn q. Then by the formula dðg0 5g1 5  5gn Þ ¼

n X i¼0

we know that d 2 ¼ 0.

ð1Þ i g0 5  5 g^i 5   5gn

Polygonal presentations of tensor categories

87

In fact, X  eðsÞqðgsð0Þ ; . . . ; gsðnÞ Þ P s

¼

n X X i¼0

¼

eðsÞð 1Þ i gsð0Þ 5  5 g^sðiÞ 5  5gsðnÞ

s

XX

eðsÞð 1Þ i gsð0Þ 5  5 g^sðiÞ¼j 5  5gsðnÞ

i; j sðiÞ¼j

(letting s ¼ t  ði; jÞ) X X ¼ eðtÞð 1Þ iþ1 i; j tð jÞ¼j

gtð0Þ 5  5gtði 1Þ 5 g^tð jÞ¼j 5gtðiþ1Þ 5  5gtð j 1Þ 5gtðiÞ 5gtð jþ1Þ 5  5gtðnÞ X X ¼ eðtÞð 1Þ iþ1 ð 1Þ j 1 i gtð0Þ 5   5gtð j 1Þ 5 g^tð jÞ 5gtðð jþ1Þ 5   5gtðnÞ i; j tð jÞ¼j

¼

X X

ð 1Þ j g0 5  5 g^j 5   5gn

i; j tð jÞ¼j

¼ ðn þ 1Þ!

n X

ð 1Þ j g0 5  5 g^j 5  5gn :

j¼0

The chain contraction c for the di¤erential is then defined by cðg0 5  5gn Þ ¼ Pnþ2 sðg0 5  5gn Þ ¼ 15g0 5   5gn : Thus we get another free resolution ðD; d Þ (the wedge resolution) of the trivial Gmodule Z. Since the ZðGÞ-linear projections fPnþ1 : C n ! Dn g give a chain homomorphism ðC; qÞ ! ðD; d Þ (see the explicit formula for d ), the induced inclusion of chain complexes d

Hom G ðDn ; AÞ ƒƒƒ! Hom G ðDnþ1 ; AÞ ? ? ? ? ? ? y y Hom G ðC n ; AÞ ƒƒƒ! Hom G ðCnþ1 ; AÞ  q

induces the isomorphisms. As a consequence, to represent cohomology classes, we can choose cocycles, say F A Hom G ðC n ; AÞ, satisfying F ðgsð0Þ ; . . . ; gsðnÞ Þ ¼ eðsÞF ðg0 ; . . . ; gn Þ

88

S. Yamagami

for s A Snþ1 . In the case of lower n, we can explicitly write down the conditions: With f ðg1 ; . . . ; gn Þ ¼ F ð1; g1 ; g1 g2 ; . . . ; g1 g2    gn Þ; we have f ðg; hÞ ¼ f ðgh; h 1 Þ ¼ f ðg1 ; g2 ; g3 Þ ¼

gf ðg 1 ; ghÞ;

g1 f ðg2 ; g3 ; g4 Þ

if g1 g2 g3 g4 ¼ 1, and so on. Remark. When G acts on A trivially, we can deduce f ðg; hÞ ¼ f ðh 1 ; g 1 Þ;

f ðg1 ; g2 ; g3 Þ ¼

f ðg2 ; g3 ; g4 Þ

from the above conditions. References [ 1 ] V. Chari and A. Pressley, A Guide to Quantum Groups, Cambridge Univ. Press, 1995. [ 2 ] D. Evans and Y. Kawahigashi, Quantum Symmetries on Operator Algebras, Oxford University Press, Oxford, 1998. [ 3 ] T. Hayashi and S. Yamagami, Amenable tensor categories and their realizations as AFD bimodules, J. Funct. Anal., 172 (2000), 19–75. [ 4 ] F. Hiai and M. Izumi, Amenability and strong amenability for fusion algebras with applications to subfactor theory, Internat. J. Math., 9 (1998), 669–722. [ 5 ] C. Kassel, Quantum Groups, Springer-Verlag, Berlin-New York, 1995. [ 6 ] D. Kazhdan and H. Wenzl, Reconstructing monoidal categories, Adv. Soviet Math., 16 (1993), 111–136. [ 7 ] G. M. Kelly, On MacLane’s conditions for coherence of natural associativities, commutativities, etc, J. Algebra, 1 (1964), 397–402. [ 8 ] R. Longo and J. E. Roberts, A theory of dimension, K-Theory, 11 (1997), 103–159. [ 9 ] S. MacLane, Categories for the Working Mathematician, Springer-Verlag, Berlin-New York, 1971. [10] N. Saavedra Rivano, Cate´gories Tannakiennes, Lecture Notes in Math., 265, Springer, Berlin, 1972. [11] S. Shnider and S. Sternberg, Quantum Groups, International Press, Boston, 1993. [12] V. S. Sunder, II1 factors, their bimodules and hypergroups, Trans. Amer. Math. Soc., 330 (1992), 227–256. [13] D. Tambara and S. Yamagami, Tensor categories with fusion rules of self-duality for finite abelian groups, J. Algebra, 209 (1998), 692–707. [14] S. Yamagami, Notes on amenability of commutative fusion algebras, Positivity, 3 (1999), 377–388. [15] ———, Frobenius reciprocity in tensor categories, Math. Scand., to appear. [16] ———, Frobenius duality in C  -tensor categories, preprint.

Shigeru Yamagami Department of Mathematical Sciences Ibaraki University Mito, 310-8512, Japan

Suggest Documents