arXiv:2108.00269v2 [math.QA] 24 Sep 2021 Preliminaries 2 Introduction 1 Contents categories. monoidal Theory; Representation spaces; unu ersnainTer n ai arcsI: matrices Manin and Theory Representation Quantum ∗ . ood n hi cin 11 ...... products . . . . finite . . . with . . . . . a . . . in . groups . . . . and . . Monoids . actions . . their . and . . 2.5 Monoids functors . . and . categories 2.4 schemes Monoidal . affine . and sets 2.3 . Algebraic and spaces 2.2 Vector 2.1 Keywords: [email protected] agaeo ai arcsadoti unu nlge fd of examples some analogues give we quantum Finally, form obtain s representations. We and of parameter product matrices the algebras. Manin quadratic where of generated language product, finitely Manin connected with monoi of a algebras to approach graded this of of application by obtained subcategory. is parameter represe Theory a to t with approach hom-functor category general internal monoidal a symmetric the construct generalise and parameter we a it with do geo To linear ‘non-commutative [Man88]. of in frame in representations of on nttt o ula eerh 490Dba Moscow Dubna, 141980 Research, Nuclear for Institute Joint ecntutQatmRpeetto hoywihdescribe which Theory Representation Quantum construct We tt nvriy”un” 490Dba ocwrgo,Rus region, Moscow Dubna, 141980 ”Dubna”, University State udai lers ai arcs unu rus non-commu groups; quantum matrices; Manin algebras; quadratic nt-iesoa case finite-dimensional lxySilantyev Alexey Abstract 1 ∗ fqatmrepresentations. quantum of er’dvlpdb Manin by developed metry’ a aeoyo oeclass some of category dal unu Representation Quantum ttoso oodi a in a of ntations h aeo adjunction of case the o lt hster nthe in theory this ulate rc u n tensor and sum irect unu analogue quantum s baeoyconsists ubcategory ein Russia region, sia 17 . . . 8 . . 4 . tative 4 2 6 3 Internal hom and representations 19 3.1 Internal (co)hom-functor and its generalisation ...... 19 3.2 (Co)representationsof(co)monoids ...... 23 3.3 Translation of (co)representations under monoidal functors ...... 27

4 Quantum linear spaces 30 4.1 Operations with quadratic algebras ...... 31 4.2 Maninmatrices ...... 37 4.3 Semi-linear spaces and their quantum analogue ...... 39 4.4 Internalhomforquantumlinearspaces ...... 44

5 Quantum 50 5.1 Quantumrepresentations...... 50 5.2 Operationswithquantumrepresentations...... 55 5.3 S∗-embedding of classical representations ...... 64 5.4 T ∗-embedding of classical representations ...... 70

6 Examples 72 6.1 Corepresentations of Mq(m)...... 72 6.2 An extension of the Y (glm)anditscorepresentations ...... 74 6.3 Corepresentations of an extended Y (som)...... 76

Conclusion and further directions 77

1 Introduction

In the end of 80s Yuri Manin generalised the notion of finite-dimensional vector (linear) space to the quantum1 case [Man88]. He introduced the term ‘quantum linear spaces’ for (finitely generated connected) quadratic algebras considered as objects of the opposite category. The usual finite-dimensional vector spaces correspond to the polynomial algebras. Manin defined four binary operations with quadratic algebras, they where denoted by ◦, •, ⊗, ⊗. He also introduced a duality of quadratic algebras, which coincides with the for the quadratic Koszul algebras (see [Man87, Man88, Man91]). The most important binary operation for our purposes is ‘◦’, which we call ‘Manin white product’ or just ‘Manin product’. As it was shown in [Man88] the category of quantum linear spaces with these product is a closed and that the internal hom can be written via the Koszul duality and the binary operation ‘•’, which is called ‘black

1We use the word ‘quantum’ instead of ‘non-commutative’ in the sense of non-commutative spaces by following Manin [Man88] and Drinfeld [Dr], because sometimes the word ‘non-commutative’ is confusing. For example, the term ‘non-commutative group’ does not mean a non-commutative space with a group structure (i.e. a ), but rather a non-abelian group. Moreover, we want to consider the usual ‘commutative’ spaces as a particular case of the non-commutative ones.

2 Manin product’. This gives a construction of the internal hom-object as a quadratic generated by entries of some . It turned out that the description of the internal hom-object for the polynomial algebras gives some non-trivial commutation relations for the entries of this matrix. The matri- ces (over a non-commutative in general) satisfying these commutation relations were called ‘Manin matrices’ in [CF]. A lot of their interesting properties and applications were found [CF, CM, CFR, RST]. The properties and applications of q-analogues of Manin ma- trices were considered in [GLZ, CFRS, IO]. The super-version and Manin matrices for the series B,C,D were considered in [Man89, Man91, Man92, MR] and [Molev] respectively. The notion of Manin matrices was generalised in [S]. It was shown that any quadratic algebra can be given by a matrix idempotent A acting in a tensor product of a . This algebra was denoted by XA(K), where K is a basic field. Then it was defined a Manin matrix for a pair of idempotents A and B and showed how this notion can be associated to the pair of quadratic algebras XA(K) and XB(K). The main aim of this paper is to generalise Representation Theory to the case of a quantum representation space. The objects which are to be represented (such as algebras, groups, algebraic groups) are generalised up to ‘quantum algebras’ (these include quantum groups). To do it we formulate a general representation theory in a monoidal category with some generalisation of internal hom-functor. For such category the representations of a monoid can be functorially identified with left actions of this monoid on some objects. In a particular case this approach gives Quantum Representation Theory: we need to take the monoidal category of quadratic algebras with the Manin product ‘◦’ together with some extension of this category, where the category of quadratic algebras plays the role of parameter subcat- egory. This extension is needed to include the representations of quantum groups. As a ‘classical’ representation can be given by a square matrix in a basis of the representation space, a quantum representation is given by a Manin matrix for the pair of idempotents A and B, where A = B. This work is mainly devoted to the finite-dimensional case. In the quantum level this means that the algebras A we consider are finitely generated and, as consequence, all the graded components Ak of these algebras are finite-dimensional vector spaces. For wider gen- erality we sometimes include the infinite-dimensional (infinitely generated) case into consid- eration, but here we do not define quantum representations on infinitely generated quadratic algebras. The paper is organised as follows. In Section 2 we give basic notations and definitions including theory of monoids and their actions. In Section 3 we describe the general approach to construct representation theory in a monoidal category by using the generalised internal hom. Since it is convenient to work in an opposite category we also define the dual notions: generalised internal cohom, corepresentations etc. Section 4 prepares the necessary notions and facts for the quantum representation theory: the Manin’s concept of quantum linear spaces, operations with quadratic algebras, semi-linear spaces and its quantum version, the

3 internal cohom for quadratic algebras and its generalisation. Section 5 is devoted to the description of the quantum representations, to the consideration of classical representations as quantum ones and to binary operations with the quantum representations. In Section 6 we give some examples of quantum representations. Acknowledgements. The author is grateful to A. Isaev, Y. Manin for useful references and to E. Patrin for discussions. The author particularly appreciates advice and guidance from V. Rubtsov.

2 Preliminaries

2.1 Vector spaces and algebras Fist of all, let us fix some terms and notations that we use in this work. 2.1.1. Basic field. Let K be a field such that char K =6 2. All the vector spaces will be over K. For two vector spaces V and W we denote by V ⊗ W = V ⊗K W their tensor product over K. For briefness we say ‘algebra’ for a unital associative algebra over K and we suppose that algebra map unity to unity. The unity of an algebra A we denote by 1 or 1A. 2.1.2. Notations for in a category. We denote the composition of mor- phisms f : Y → Z and g : X → Y in a category C by f · g. We do not use the symbol ‘◦’ for composition, because it will be used for the Manin product. For a natural transformation θ between functors F and G we sometimes omit the subscript of its components θX : FX → GX. In particular, for a f : Y → Z we denote by ∗ f∗ and f the maps

f∗ : Hom(X,Y ) → Hom(X,Z), f∗(g)= f · g : X → Z, g : X → Y, (2.1) f ∗ : Hom(Z,X) → Hom(Y,X), f ∗(h)= h · f : Y → X, h: Z → X. (2.2)

They are the components of the natural transformations Hom(−,Y ) → Hom(−,Z) and Hom(Z, −) → Hom(Y, −) respectively. 2.1.3. Basis. A basis of a vector space V is a map e: I → V from a set I (which may be infinite) such that for any vector space W and any map f : I → W there exists a unique h: V → W satisfying h · e = f. The basis e: I → V is denoted by (ei)i∈I or (ei), where ei is the value of e on i ∈ I. In the case of the set I = {1,...,n} we also use n the notation (ei)i=1. One can check that (ei)i∈I is a basis of V iff the elements ei ∈ V are pairwise different, the set {ei | i ∈ I} = e(I) is linear independent and any vector v ∈ V is K a finite linear combination of the form i αiei where αi ∈ . Notice that the basis (ei) is ′ not just a set, because the permutationP of the elements ei gives another basis (ei) with the ′ same set {ei} = {ei}.

4 2.1.4. Bimodules. Let R and S be algebras. By an (R, S)-bimodule we understand a vector space M that has a structures of left R- and right S-module agreed with the structure of the vector space. In other words, M is an (R, S)-bimodule in the ring- theoretic sense satisfying the condition αm = mα ∀ α ∈ K, m ∈ M. We also use the term R-bimodule or two-sided R-module for (R, R)-bimodule. For any vector space V the tensor product R⊗V has a structure of R-bimodule: r1(r2 ⊗v)=(r1 ⊗m)r2 =(r1r2)⊗v, r1,r2 ∈ R, v ∈ V .

2.1.5. Grading. Let N0 be the monoid of natural numbers (including 0). We call an N0-graded vector space simply graded vector space. The k-th component of a graded vector space V is denoted by Vk, so that V = Vk. Bya graded algebra we mean an algebra N kL∈ 0 A with a structure of graded vector space such that AkAl ⊂ Ak+l for any k,l ∈ N0. The component A0 is an algebra, so A has a structure of an A0-bimodule. Each component Ak is its A0-submodule. Note that the tensor product R ⊗A of an algebra R with a graded algebra A is a graded algebra with the components (R ⊗A)k = R ⊗Ak. 2.1.6. Connected affinely generated graded algebras. A graded algebra A is called connected if A0 = K. The most interesting graded algebras in the projective (non-com- mutative) geometry are the connected algebras generated by the subspace A1. Let us call a graded algebra A affinely generated if it is generated by the subspace A0 ⊕A1, so the connected graded algebras A generated by A1 are exactly the connected affinely generated graded algebras. Let us call an affinely generated graded algebra A semi-connected if it is isomorphic to the graded algebra R ⊗ B for an algebra R and a connected affinely generated ∼ ∼ graded algebra B (since B0 = K we have R = A0, so that A = A0 ⊗ B). 2.1.7. Quadratic algebras. By a quadratic algebra over an algebra R we mean an algebra isomorphic to TRM/I where

TRM := R ⊕ M ⊕ M ⊗R M ⊕ M ⊗R ···⊗R M (2.3)  Mk>3 k  | {z } is the tensor algebra of an R-bimodule M and I is a (two-sided) ideal of TRM generated by a subset of M ⊗R M. Any quadratic algebra A = TRM/I has a structure of graded algebra ⊗k ⊗k ⊗k with the components A0 = R, A1 = M, Ak = M /(I∩M ), where M = M ⊗R···⊗RM. Hence any quadratic algebra is affinely generated graded algebra. In a narrow sense the notion ‘quadratic algebra’ is used for the quadratic algebras over K. In this case the two-sided module M is a vector space V and αv = vα for any α ∈ K ⊗k and v ∈ V . Such a quadratic algebra is isomorphic to TV/I, where TV = TKV = V k∈N0 is the tensor algebra of the vector space V and I is the ideal of TV generated by a subspaceL R ⊂ V ⊗ V . These quadratic algebras are exactly the connected quadratic algebras. Any semi-connected quadratic algebra over R is isomorphic to R ⊗ (TV/I) ∼= TRM/I where V is a vector space, M = R ⊗ V , I and I are the ideals of TV and TRM generated by a subspace R ⊂ V ⊗2 (over K and over R respectively).

5 2.1.8. Notations for some categories. Set the category of sets, Vect the category of vector spaces (K-modules), FVect the category of finite-dimensional vector spaces, GrVect the category of graded vector spaces, Alg the category of algebras, FAlg the category of finite-dinetional algebras, CommAlg the category of commutative algebras, GrAlg the category of graded algebras, QA the category of quadratic algebras over K,

FQA its full subcategory of A ∈ QA such that A1 ∈ FVect, ∼ QAsc the category of quadratic algebras A = A0 ⊗ B where B ∈ QA, ∼ FQAsc the category of quadratic algebras A = A0 ⊗ B where B ∈ FQA.

The categories QA and QAsc consist of the connected and semi-connected quadratic algebras respectively. The subcategory FQA ⊂ QA consists of finitely generated quadratic algebras A ∈ QA, while FQAsc ⊂ QAsc consists of algebras A ∈ QAsc finitely generated over A0. Note that a quadratic algebra A = TV/(R) ∈ QA, where (R) is the ideal generated by a subspace R ⊂ V ⊗2, belongs to FQA iff V ∈ FVect. We obtain the following picture:  FVect / Vect O 8 g❖ ♣♣♣ ❖❖❖ ♣♣ ❖❖❖ ♣♣♣ ❖❖ ♣♣ ❖❖❖  ♣♣ FAlg / Alg o _CommAlg GrVect ? 7 f▼▼▼ ♦♦ ▼▼ ♦♦♦ ▼▼▼ ♦♦♦ ▼▼▼ ♦♦♦   ♦ FQAsc / QAsc / GrAlg O O

?  ? FQA / QA ’where ‘֒→’ are inclusions of full subcategories (fully faithful functors) and ‘→’ are ‘forgetful functors (they are faithful, but not full).

2.2 Algebraic sets and affine schemes Let us introduce some notions from algebraic geometry [Ful], [Man70], [Har]. 2.2.1. Algebraic sets. Define (affine) algebraic set (over K) as the set of all the solu- tions of a system of algebraic equations with coefficients in K, that is a set of all n-tuples (x1,...,xn) ∈ Kn satisfying 1 n Fα(x ,...,x )=0, (2.4)

6 n where Fα : K → K are polynomial functions of n variables (with coefficients in K) and α runs over some set of indices. If the system (2.4) is empty then the corresponding algebraic n n set consists of all the n-tuples; it is denoted by A = AK and called affine space. As a set the affine space An coincides with the vector space Kn, but we consider it as an object of another category, where there are more morphisms. Namely, a morphism of affine spaces n m 1 n 1 n 1 n Φ: A → A is a map (x ,...,x ) 7→ (P1(x ,...,x ),...,Pm x ,...,x ) , where Pi are polynomial functions.  2.2.2. Category of algebraic sets. Any algebraic set is a subset of An for a certain n. A morphism between algebraic sets X ⊂ An and Y ⊂ Am is a map ϕ: X → Y induced by a morphism Φ: An → Am in the sense that the diagram

ϕ X _ / Y _ (2.5)

 Φ  An / Am commutes. Denote the category of algebraic sets over K with these morphisms by AlgSet. 2.2.3. Algebra of regular functions. A regular function on an algebraic set X is a morphism X → A1. Since A1 = K, the regular functions on X form an algebra, denote it by A(X). The algebra of regular functions on An is the algebra of polynomials K[x1,...,xn]. By setting Y = Am = A1 = K in the diagram (2.5) we see that a function f : X → K is regular iff there exists a polynomial F ∈ K[x1,...,xn] such that f(x1,...,xn) = F (x1,...,xn) for any (x1,...,xn) ∈ X. We obtain an algebra epimorphism K[x1,...,xn] ։ A(X), so that A(X) ∼= K[x1,...,xn]/I(X), where I(X) ⊂ K[x1,...,xn] is the ideal of polynomials F ∈ K[x1,...,xn] such that F (x1,...,xn)=0 for any (x1,...,xn) ∈ X. 2.2.4. Affine schemes. For any morphism ϕ: X → Y its pull-back ϕ∗ : A(Y ) → A(X), ϕ∗(f)= f·ϕ, is an algebra . It is known that by mapping X 7→ A(X), ϕ 7→ ϕ∗ we obtain a contravariant fully faithful functor AlgSet → CommAlg. This means that the category AlgSet is embedded into CommAlgop. The notion of algebraic set over K is generalised to the notion of affine scheme over K such that these schemes form the category AffSch = CommAlgop. The affine scheme corresponding to an algebra R ∈ CommAlg is denoted by Spec R. (see e.g. [Man70] for details). 2.2.5. Sheaves of modules and vector bundles. Any quasi-coherent sheaf of modules on an affine scheme Spec R is uniquely determined by the R-module of its global sections. Moreover, the category of all quasi-coherent sheaves of modules is equivalent to the category of R-modules. A vector bundle on an affine scheme is a particular case of a (quasi-)coherent sheaf. A trivial vector bundle (free sheaf) on Spec R with a fibre V ∈ FVect corresponds to the R-module R ⊗ V . If V = Kn and R = A(X) for X ∈ AlgSet, then a global section is a morphism X → X × An such that x 7→ x, v(x) for a morphism v : X → An. These sections form an A(X)-module isomorphic toA(X) ⊗ Kn, an element a ⊗ v ∈ A(X) ⊗ Kn gives a section x 7→ x, a(x)v .  7 2.2.6. Quantum affine schemes. It is natural to define the quantum or ‘non-commutative’ version of the category of affine schemes as Algop (see e.g. [Dr]). Thus, quantum affine schemes are algebras R ∈ Alg, their morphisms are reversed algebra homomorphisms. As an analogue of the category of quasi-coherent sheaves on R ∈ Algop one should regard the category of left, right or two-sided R-modules.

2.3 Monoidal categories and functors Here we remind the basic concepts from the theory of monoidal categories [MacLane], [Bor2]. 2.3.1. Monoidal categories. A monoidal category is a category C equipped with a bi- functor ⊗: C × C → C (called a monoidal product), a unit object I ∈ C and natural (X ⊗ Y ) ⊗ Z =∼ X ⊗ (Y ⊗ Z), I ⊗ X ∼= X and X ⊗ I ∼= X satisfying some conditions (see [MacLane] for details). It is called strict monoidal if these isomorphisms are identities. Due to [MacLane, 11.3,Th. 1] any monoidal category is monoidally equivalent to a strict monoidal category, therefore we can suppose without loss of generality that all con- sidered monoidal categories are strict monoidal (if it is not, one can always insert necessary isomorphisms into diagrams and formulae). We write C =(C, ⊗,I) or simply C =(C, ⊗), since the unit object is unique up to an . In the second case we sometimes write IC for the unit object. Note that if C = (C, ⊗) is a monoidal category, then its opposite op op C = (C , ⊗) is also a monoidal category with the same unit object ICop = IC. A sub- category C′ of a monoidal category (C, ⊗,I) is called monoidal subcategory if I ∈ C′ and f ⊗ g ∈ C′ for any morphisms f,g in C′ (in particular, X ⊗ Y ∈ C′ ∀ X,Y ∈ C′). 2.3.2. Examples of monoidal categories.

• If there exists a terminal object and a product of any pair of objects in a category C, then (C, ×) is a monoidal category and its terminal object is its unit object (in this case C is called a category with finite products, see p. 2.5.1). In particular, (Vect, ⊕) and its subcategory (FVect, ⊕) are monoidal categories with unit object 0, where V ⊕ W is the of the vector spaces V and W .

• If there exists an initial object and a coproduct of any pair of objects in a category C, then (C, ∐) is a monoidal category. For instance, the category (CommAlg, ⊗) is monoidal, since the tensor product R ⊗ S of commutative algebras R and S is their coproduct. The opposite monoidal category is (AffSch, ×), where Spec R × Spec S = Spec(R ⊗ S) is the categorical product.

• Also, the category Vect and its subcategory FVect are monoidal with respect to the usual tensor product ⊗. The unit object is the one-dimensional vector space K = K1.

• The tensor product ⊗ gives a structure of monoidal category on GrVect. The monoidal product of two graded vector spaces V and W is the usual tensor product V ⊗ W with k the components (V ⊗ W )k = Vl ⊗ Wk−l. The unit object is the vector space K lL=0 8 with vanishing components of non-zero order: (K)k = δk0K. In this way we obtain the monoidal category (GrVect, ⊗, K).

• The tensor product of two algebras A, B ∈ Alg is the vector space A⊗B with the mul- tiplication (a⊗b)(a′ ⊗b′)=(aa′)⊗(bb′). This gives the monoidal category (Alg, ⊗, K).

• In the same way we obtain the monoidal category (GrAlg, ⊗, K). One can check that all the full subcategories of GrAlg defined above are monoidal subcategories.

• Another useful example of a monoidal product on GrVect is Manin product. For two objects V, W ∈ GrVect it is defined as the graded vector space V ◦ W with the components

(V ◦ W )k = Vk ⊗ Wk. (2.6)

The monoidal category (GrVect, ◦) is equivalent to the direct product (Vect, ⊗). k∈N0 Its unit object is a graded vector space with one-dimensional components,Q that is the polynomial algebra K[u] with the standard grading. Note also that we have the (non-graded) inclusion of vector spaces V ◦ W ⊂ V ⊗ W .

• One can show that for any A, B ∈ GrAlg their Manin product A ◦ B is a subalgebra of A ⊗ B. Thus we obtain the monoidal category (GrAlg, ◦, K[u]) and its monoidal subcategories (the product ‘◦’ was originally defined by Manin for the category FQA).

2.3.3. Monoidal functors. Let C =(C, ⊗) and D =(D, ⊙) be two monoidal categories. The functor F : C → D equipped with a morphism ϕ: ID → FIC and a natural transfor- mation φX,Y : FX ⊙ FY → F (X ⊗ Y ) is called lax monoidal functor if for any X,Y,Z ∈ C the diagrams

id ⊙φY,Z FX ⊙ FY ⊙ FZ / FX ⊙ F (Y ⊗ Z) (2.7)

φX,Y ⊙id φX,Y ⊗Z

 φX⊗Y,Z  F (X ⊗ Y ) ⊙ FZ / F (X ⊗ Y ⊗ Z)

ID ⊙ FX FX FX ⊙ ID FX (2.8)

ϕ⊙id id ⊙ϕ

 φIC,X  φX,IC F (IC) ⊙ FX / F (IC ⊗ X) FX ⊙ F (IC) / F (X ⊗ IC) commute. The functor F : C → D is called colax monoidal if its opposite F op : Cop → Dop is lax monoidal. In other words, this is a triple (F,ϕ,φ), where ϕ: FIC → ID and φ is a natural transformation with components φX,Y : F (X ⊗ Y ) → FX ⊙ FY such that the reversed diagrams (2.7), (2.8) commute, i.e. (φX,Y ⊙ idFZ) · φX⊗Y,Z = (idFX ⊙φY,Z) · φX,Y ⊗Z ,

(ϕ ⊙ idFX ) · φIC,X = idFX = (idFX ⊙ϕ) · φX,IC .

9 A strong monoidal functor is a lax (or colax) monoidal functor (F,ϕ,φ) such that ϕ and φX,Y are isomorphisms. A strong monoidal functor is called strict monoidal functor if they are identities. For instance, the inclusion of a monoidal subcategory C′ ⊂ C into C is a faithful strict monoidal functor. Note that a composition of two lax/colax/strong monoidal functors is a lax/colax/strong monoidal functor. 2.3.4. Examples of monoidal functors.

• Consider the dualisation functor (−)∗ : Vectop → Vect which sends a vector space V ∈ Vect to its dual V ∗, it is defined on linear maps f : V → W by the formula (2.2). For any vector spaces V, W ∈ Vect the tensor product V ∗ ⊗W ∗ is naturally embedded into (V ⊗ W )∗, so the functor (−)∗ : Vectop → Vect has a structure of lax monoidal functor (Vectop, ⊗) → (Vect, ⊗). The opposite functor (Vect, ⊗) → (Vectop, ⊗) is colax monoidal.

• The same functor considered on finite-dimensional vector spaces is a strong monoidal functor (−)∗ : (FVect, ⊗) → (FVectop, ⊗), since the embedding gives the natural isomorphisms V ∗ ⊗ W ∗ ∼= (V ⊗ W )∗ for any V, W ∈ FVect.

• Note that (V ⊕W )∗ ∼= V ∗⊕W ∗ for any V, W ∈ Vect, so we obtain strong monoidal con- travariant functors (−)∗ : (Vect, ⊕) → (Vect, ⊕)and(−)∗ : (FVect, ⊕) → (FVect, ⊕).

• The functor T : Vect → GrAlg, which gives the tensor algebra TV of a vector space V ∈ Vect, is a strong monoidal functor T : (Vect, ⊗) → (GrAlg, ◦). The isomor- phisms φV,W : TV ◦ T W −→∼ T (V ⊗ W ) have the graded components

⊗k ⊗k ⊗k (φV,W )k : V ⊗ W −→∼ (V ⊗ W ) , (2.9)

v1 ⊗···⊗ vk ⊗ w1 ⊗···⊗ wk 7→ (v1 ⊗ w1) ⊗···⊗ (vk ⊗ wk).

It gives also the strong monoidal functors

T : (Vect, ⊗) → (QA, ◦), T : (FVect, ⊗) → (FQA, ◦). (2.10)

2.3.5. Symmetric monoidal category. A monoidal category C =(C, ⊗) equipped with a natural isomorphism σX,Y : X ⊗ Y −→∼ Y ⊗ X is called symmetric if for any X,Y,Z ∈ C we have σY,X σX,Y = idX⊗Y and the diagrams

σX,Y ⊗id X ⊗ Y ⊗ Z / Y ⊗ X ⊗ Z X ⊗ IC (2.11) ❚❚❚ ▲▲ σ ❚❚❚❚ ▲▲ X,IC ❚❚❚ id ⊗σX,Z ▲▲ σX,Y ⊗Z❚❚❚❚ ▲▲ ❚❚*  ▲▲& Y ⊗ Z ⊗ X X IC ⊗ X

commute. The natural transformation σ gives an additional structure on the monoidal category C.

10 All the monoidal categories considered in p. 2.3.2 are symmetric. For example, the symmetric structure of (Vect, ⊗) is given by the isomorphisms ∼ σV,W : V ⊗ W −→ W ⊗ V, σV,W (v ⊗ w)= w ⊗ v. (2.12)

One can check that it induces the symmetric structure of the monoidal categories (Alg, ⊗), (GrVect, ⊗), (GrAlg, ⊗), (GrVect, ◦), (GrAlg, ◦) and of their monoidal subcategories considered above. In fact, we already used the maps (2.12): the natural isomorphisms (2.9) ⊗l ⊗(2k−l−2) are compositions of some operators of the form id ⊗σV,W ⊗ id , l = 1,..., 2k − 3, k > 2. If a monoidal category C =(C, ⊗) is symmetric, then Cop =(Cop, ⊗) is also symmetric: the role of the structure isomorphisms V ⊗ W −→∼ W ⊗ V in Cop is played by the structure ∼ isomorphisms σW,V : W ⊗ V −→ V ⊗ W of the category C. 2.3.6. Symmetric monoidal functors. Let (C, ⊗) and (D, ⊙) be symmetric monoidal categories. A lax (or strong) monoidal functor F : (C, ⊗) → (D, ⊙) with the structure natural transformation φX,Y : FX ⊙ FY → F (X ⊗ Y ) is called symmetric if it preserves the symmetric structure in the sense that all the diagrams

σFX,FY FX ⊙ FY / FY ⊙ FX (2.13)

φX,Y φY,X

 F σX,Y  F (X ⊗ Y ) / F (Y ⊗ X)

commute. A colax monoidal functor F : (C, ⊗) → (D, ⊙) with the structure natural transfor- mation φX,Y : F (X ⊗ Y ) → FX ⊙ FY is called symmetric if the diagrams (2.13) with the re- versed vertical arrows commute (i.e. if the lax monoidal functor F op : (Cop, ⊗) → (Dop, ⊙) is symmetric). A composition of symmetric monoidal functors is symmetric. All the monoidal functors considered in p. 2.3.4 are symmetric.

2.4 Monoids and their actions We define a notion of monoid, bimonoid and Hopf monoid for a general (symmetric) monoidal category [MacLane], [Porst]. They generalise the notions of algebra, bialgebra and Hopf algebra (for this case see [Kass]). 2.4.1. Monoids and comonoids. A monoid in a monoidal category (C, ⊗) is a triple M = (X,µX , ηX ) of an object X ∈ C with structure morphisms µX : X ⊗ X → X and ηX : IC → X such that the diagrams

id ⊗µX ηX ⊗id id ⊗ηX X ⊗ X ⊗ X / X ⊗ X I ⊗ X / X ⊗ X o X ⊗ I (2.14) ▲ ▲▲▲ rrr ⊗ µ ▲▲▲ µX rrr µX id X ▲▲▲▲ rrrr ▲▲▲ rrrr  µX  ▲▲  rrr X ⊗ X / X X r

11 commute. Morphisms µX and ηX , which give a structure of monoid on X, are often called multiplication and unit. A comonoid in (C, ⊗) is a triple O = (X, ∆X ,εX) of an object X ∈ C and morphisms ∆X : X → X ⊗ X and εX : X → IC in C such that the reversed diagrams (2.14) commute, i.e. (idX ⊗∆X ) · ∆X = (∆X ⊗ idX ) · ∆X and (εX ⊗ idX ) · ∆X = idX = (idX ⊗εX ) · ∆X . For ∆X and εX one uses the terms ‘comultiplication’ and ‘counit’.

2.4.2. Morphisms of monoids and comonoids. A morphism of monoids (X,µX , ηX ) and (Y,µY , ηY ) or comonoids (X, ∆X ,εX ) and (Y, ∆Y ,εY ) is a morphism f : X → Y in C preserving the structure morphisms:

µX ηX ∆X εX X ⊗ X / X IC / X or X / X ⊗ X X / IC (2.15) ❆ > ❆❆ ⑥⑥ f⊗f f ❆ f f f⊗f f ⑥ η ❆❆ ⑥⑥ε Y ❆❆ ⑥⑥ Y  µY    ∆   ⑥ Y ⊗ Y / Y Y Y Y / Y ⊗ Y Y

Since the composition of two morphisms of (co)monoids is a morphisms of (co)monoids, the monoids and comonoids in a monoidal category (C, ⊗) form categories denoted by op Mon(C, ⊗) and Comon(C, ⊗) respectively. Note that Comon(C, ⊗)= Mon(Cop, ⊗) .  Proposition 2.1. Consider two monoids (X,µX , ηX ), (Y,µY , ηY ) ∈ Mon(C) or comonoids (X, ∆X ,εX), (Y, ∆Y ,εY ) ∈ Comon(C) in a monoidal category C =(C, ⊗). Let f : X −→∼ Y be an isomorphism in the category C. If f : X → Y preserves structure of these monoids or comonoids, then f −1 : Y → X also preserves structure of these monoids or comonoids. In particular, this implies an isomorphism of the corresponding monoids or comonoids in Mon(C) or Comon(C) respectively.

Proof. It is enough to consider the case of monoids. By using the commutativity of the left −1 −1 −1 −1 diagrams (2.15) one yields f · ηY = f · f · ηX = ηX and f · µY = f · µY · (f ⊗ f) · −1 −1 −1 −1 −1 −1 −1 (f ⊗ f )= f · f · µX · (f ⊗ f )= µX · (f ⊗ f ). 2.4.3. Monoidal products of monoids and comonoids. If C is symmetric then Mon(C) and Comon(C) are monoidal categories and they are also symmetric: the monoidal product of monoids (X,µX , ηX ) and (Y,µY , ηY ) or comonoids (X, ∆X ,εX ) and (Y, ∆Y ,εY ) is the object X ⊗ Y with the structure defined by the diagrams

X ⊗ Y ⊗ X ⊗ Y IC ❘ or ❯❯❯ ❘❘❘ ❯❯❯ µX⊗Y ❘❘❘ ηX⊗Y id ⊗σY,X ⊗id ❯❯❯ ❘❘ ❯❯❯ ❘❘❘ ❯❯❯❯ ❘❘  µX ⊗µY ❯* ηX ⊗ηY ❘( X ⊗ X ⊗ Y ⊗ Y / X ⊗ Y IC ⊗ IC / X ⊗ Y

X ⊗ Y ❱ X ⊗ Y ❱❱❱❱ PP ❱❱❱❱∆X⊗Y PPPεX⊗Y ∆X ⊗∆Y ❱❱❱ εX ⊗εY PP ❱❱❱❱ PPP ⊗ ⊗ ❱❱❱❱ PP  id σX,Y id +  PP( X ⊗ X ⊗ Y ⊗ Y / X ⊗ Y ⊗ X ⊗ Y IC ⊗ IC IC

12 A usual monoid is a monoid in the monoidal category Set = (Set, ×). A monoid in (Vect, ⊗) is an algebra: Mon(Vect, ⊗)=(Alg, ⊗). In its subcategory of finite-dimensional vector spaces we have Mon(FVect, ⊗)=(FAlg, ⊗). Comonoids in (Vect, ⊗) are called . Analogously, we have Mon(GrVect, ⊗)=(GrAlg, ⊗).

2.4.4. Eckmann–Hilton Principle. A monoid (X,µX , ηX ) in a symmetric monoidal cat- egory (C, ⊗) is called commutative if µX · σX,X = µX . Analogously, a comonoid (X, ∆X ,εX ) is called cocommutative if σX,X · ∆X = ∆X . Thus we obtain full monoidal subcategories cMon(C, ⊗) ⊂ Mon(C, ⊗) and cocComon(C, ⊗) ⊂ Mon(C, ⊗). For example, for the category C = Vect with the tensor product ⊗ we have cMon(Vect, ⊗)=(CommAlg, ⊗). The Eckmann–Hilton Principle was formulated in [EH, Th 1.12] for the groups: if a group homomorphism µ: G × G → G has a neutral element as a binary operation, then the group G is abelian and µ coincides with the group multiplication. This fact is generalised for any symmetric monoidal category (C, ⊗) (see e.g. [Porst, § 4]).

Proposition 2.2. The category Mon Mon(C, ⊗) is equivalent to cMon(C, ⊗). Namely, the monoid structure (µM, ηM) on an object M = (X,µX , ηX ) ∈ Mon(C, ⊗) exists iff the monoid M is commutative. This structure is unique: µM = µX , ηM = ηX . The category Comon Comon(C, ⊗) is equivalent to cocComon(C, ⊗) in the same way.  2.4.5. The functors Mon(F ) and Comon(F ). Any lax monoidal functor F : C → D between monoidal categories C =(C, ⊗) and D =(D, ⊙) induces the functor

Mon(F ): Mon(C) → Mon(D). (2.16)

Namely, if M = (X,µX , ηX ) is a monoid in C, then Mon(F )M = (FX,µFX, ηFX ) is a monoid in D, where µFX : FX ⊙ FX → FX and ηFX : ID → FX are the compositions

φX,X F µX ϕ F ηX FX ⊙ FX −−−→ F (X ⊗ X) −−→ FX and ID −→ FIC −−→ FX (2.17)

respectively. Dually, a colax monoidal functor F : C → D induces the functor

Comon(F ): Comon(C) → Comon(D). (2.18)

Note that the functors (2.16) and (2.18) depend on the monoidal structure on F , that is Mon(F )= Mon(F,ϕ,φ) and Comon(F )= Comon(F,ϕ,φ). A faithful lax or colax monoidal functor F induces a faithful functor (2.16) or (2.18) respectively. A strong monoidal functor F gives strong monoidal functors (2.16) and (2.18). If F is strong monoidal and fully faithful, then (2.16) and (2.18) are also fully faithful (the condition to be strong monoidal can be weakened: it is enough to require ϕ and all φX,X to be epimorphisms/monomorphisms for lax/colax case). Suppose the categories C and D are symmetric. If F is a symmetric lax/strong monoidal functor, then Mon(F ) is a symmetric lax/strong monoidal functor as well. If F is a sym- metric colax/strong monoidal functor, then Comon(F ) is also a symmetric colax/strong monoidal functor.

13 2.4.6. Bimonoids. Let C =(C, ⊗) be a symmetric monoidal category. A bimonoid in C is B = (X,µX , ηX , ∆X ,εX ) such that M = (X,µX , ηX ) and O = (X, ∆X ,εX ) are monoid and comonoid in C with compatible structures:

µX ∆ ηX X ⊗ X / X X / X ⊗ X I / X O C ❆ ❆❆❆❆ ⊗ ⊗ ❆❆❆ ∆X ∆X µX µX ❆❆❆ εX ❆❆❆❆  id ⊗σX,X ⊗id ❆  X ⊗ X ⊗ X ⊗ X / X ⊗ X ⊗ X ⊗ X IC (2.19) ηX µX IC / X X ⊗ X / X

∆X εX ⊗εX εX

ηX ⊗ηX    IC ⊗ IC / X ⊗ X IC ⊗ IC IC

These diagrams mean exactly that ∆X : X → X⊗X and εX : X → IC are monoid morphisms (X ⊗ X,µX⊗X , ηX⊗X ) → (X,µX , ηX ) and (X,µX , ηX ) → (IC, idC, idC) or, equivalently, that µX : X ⊗ X → X and εX : IC → X are comonoid morphisms (X ⊗ X, ∆X⊗X ,εX⊗X) → (X, ∆X ,εX) and (IC, idC, idC) → (X, ∆X ,εX). Bimonoids in C = (C, ⊗) are objects of the symmetric monoidal category Bimon(C, ⊗) = Bimon(C) = Comon Mon(C) = Mon Comon(C) .  Any symmetric strong monoidal functor F = (F,ϕ,φ): C → D between symmetric monoidal categories C and D induces the symmetric strong monoidal functor

Bimon(F ): Bimon(C) → Bimon(D), (2.20) where Bimon(F )= Mon Comon(F ) = Comon Mon(F ) .   2.4.7. Hopf monoids. For arbitrary bimonoid (X,µX , ηX , ∆X ,εX ) define a convolution on the set EndC(X). The convolution of two morphisms α: X → X and β : X → X in C is the morphism α ⋆ β := µX · (α ⊗ β) · ∆X : X → X. This is a multiplication on the set EndC(X) with the neutral element ηX · εX . A Hopf monoid in C is a bimonoid (X,µX , ηX , ∆X ,εX ) in C such that the morphism idX has an inverse ζX : X → X with respect to the convolution, that is ζX ⋆ idX = idX ⋆ζX = ηX · εX . The morphism ζX : X → X is called antipode. It is unique since the convolution is associative. The functor (2.20) maps a Hopf monoid in C to a Hopf monoid in D, the antipode ζX : X → X is mapped to the antipode F (ζX): FX → FX. Bimonoids and Hopf monoids in (Vect, ⊗) are bialgebras and Hopf algebras respectively.

2.4.8. Opposite monoid and coopposite comonoid. Let M =(X,µX , ηX ) be a monoid op in a symmetric monoidal category C =(C, ⊗). Let µX := µX ·σX,X : X ⊗X → X be opposite Mop op multiplication. Then := (X,µX , ηX ) is also monoid in C, it is called monoid opposite to M. Note that M is commutative in the sense of p. 2.4.4 iff Mop = M, but in general these are different monoids. If f : X → X is a morphism of the monoids M = (X,µX , ηX ) and op op M =(X,µ e , η e ), then it is a morphism M → M as well. X X e e e e 14 Dually, let O = (X, ∆X ,εX ) be a comonoid in a monoidal category C = (C, ⊗). The Ocop cop comonoid := (X, ∆X ,εX) ∈ Comon(C) with the so-called coopposite comultiplication cop ∆ = σX,X · ∆ and the same counit εX is called comonoid coopposite to O. For a bimonoid B =(X,µX , ηX , ∆X ,εX ) we can define the following tree bimonoids: oppo- Bop op Bcop cop site bimonoid = (X,µX , ηX , ∆X ,εX ), coopposite bimonoid = (X,µX , ηX , ∆X ,εX ) Bop,cop op cop and opposite coopposite bimonoid =(X,µX , ηX , ∆X ,εX ) (the commutativity of the corresponding diagrams (2.19) is check straightforwardly). Moreover, if the bimonoid B is a Hopf monoid, then Bop,cop is also a Hopf monoid with the same antipode.

2.4.9. Actions and coactions. A (left) action of a monoid (X,µX , ηX ) on an object V ∈ C is a morphism a: X ⊗ V → V making the diagrams

µX ⊗idV ηX ⊗idV X ⊗ X ⊗ V / X ⊗ V IC ⊗ V / X ⊗ V (2.21) ❘❘❘ ❘❘❘❘ idX ⊗a a ❘❘❘❘ a ❘❘❘❘  a  ❘❘❘❘  X ⊗ V / V ❘ V

commutative. A (left) coaction of a comonoid (X, ∆X ,εX ) on an object V ∈ C is a morphism δ : V → X ⊗ V making the reversed diagrams commutative:

δ δ V / X ⊗ V V / X ⊗ V (2.22) ●●● ●●●● ⊗ ●●● ⊗ δ ∆X idV ●●●● εX idV ●●●●  idX ⊗δ   X ⊗ V / X ⊗ X ⊗ V IC ⊗ V

The actions of a monoid M = (X,µX , ηX ) ∈ Mon(C) or a comonoid O = (X, ∆X ,εX) ∈ Comon(C) form a category Lact(M) or Lcoact(O) respectively. Their objects are pairs (V, a)and(V, δ). A morphism (V, a) → (V ′, a′) in Lact(M)or(V, δ) → (V ′, δ′) in Lcoact(M) is a morphism f : V → V ′ in C such that the diagram

a δ X ⊗ V / V or V / X ⊗ V (2.23)

idX ⊗f f f idX ⊗f  a′   δ′  X ⊗ V ′ / V ′ V ′ / X ⊗ V ′ commutes. For the case C = (Vect, ⊗) the monoid M is an algebra R ∈ Alg and the objects of Lact(R) are exactly R-modules, while Lcoact(C) is the category of comodules of a C ∈ Comon(Vect, ⊗). 2.4.10. Actions and coactions of a bimonoid. A bimonoid in a symmetric monoidal category C = (C, ⊗) is B = (X,µX , ηX , ∆X ,εX ) ∈ Bimon(C). By considering it as the monoid M = (X,µX , ηX ) we obtain the category Lact(B) := Lact(M). The comonoid structure (∆X ,εX) turns Lact(B) into a monoidal category in the following way [Par]. Let (V, a) and (W, b) be two objects of Lact(B). Their monoidal product (V, a) ⊗ (W, b) is the object V ⊗ W with the action

∆ ⊗id id ⊗σX,V ⊗id a⊗b X ⊗ (V ⊗ W ) −−−−→X X ⊗ X ⊗ V ⊗ W −−−−−−−→ X ⊗ V ⊗ X ⊗ W −−→ V ⊗ W. (2.24)

15 If f : V → V ′ and g : W → W ′ are morphisms in Lact(B) of the forms (V, a) → (V ′, a′) and (W, b) → (W ′, b′), then their monoidal product f ⊗ g : V ⊗ W → V ′ ⊗ W ′ is a morphism ′ ′ ′ ′ (V, a) ⊗ (W, b) → (V , a ) ⊗ (W , b ). The unit object of Lact(B) is IC with the action εX cop X ⊗ IC = X −→ IC. If B is cocommutative (B = B), then the symmetric structure on C gives a symmetric structure on the monoidal category Lact(B). Dually, the structure (µX , ηX ) turns the category Lcoact(B) := Lcoact(X, ∆X ,εX) into monoidal category. If δ and γ are coactions of the comonoid (X, ∆X ,εX ) on V and W respectively, then its coaction on V ⊗ W is the composition

δ⊗γ id ⊗σV,X ⊗id µ ⊗id V ⊗ W −−→ X ⊗ V ⊗ X ⊗ W −−−−−−−→ X ⊗ X ⊗ V ⊗ W −−−−→X X ⊗ (V ⊗ W ). (2.25)

ηX op The coaction on IC is the morphism IC −→ X = X ⊗ IC. If B is commutative (B = B), then Lcoact(B) is a symmetric monoidal category. 2.4.11. Translation of actions and coactions. Let F : (C, ⊗) → (D, ⊙) be a lax monoidal functor with structure morphisms φX,Y : FX ⊙ FY → F (X ⊗ Y ), ϕ: ID → FIC. The functor Mon(F ): Mon(C) → Mon(D) maps a monoid M =(X,µX , ηX ) to the monoid M = Mon(F )(M)=(FX,µFX, ηFX ), where µFX : FX ⊙ FX → FX, ηFX : ID → FX are (2.17). Let a: X ⊗ V → V be an action of the monoid M on V and consider the e morphism a: FX ⊙ FV → FV defined as a composition

φX,V F (a) e FX ⊙ FV −−−→ F (X ⊗ V ) −−→ FV. (2.26)

It is straightforward to check that a is an action of M on FV . If f : V → V ′ is a mor- phism (V, a) → (V ′, a′) in Lact(M) then F (f): FV → FV ′ is a morphism between the e corresponding actions in Lact(M). Thuse we obtain a functor

e Lact(M) → Lact(M) (2.27) induced by the lax monoidal functor F =(F,ϕ,φ). e Dually, a colax monoidal functor F =(F,ϕ,φ): C → D induces the functor

Lcoact(O) → Lcoact(O) (2.28) e for comonoids O ∈ Comon(C) and O = Comon(F )O ∈ Comon(D). If the monoidal categories C = (C, ⊗) and D = (D, ⊙) are symmetric, then any sym- e metric strong monoidal functor F =(F,ϕ,φ): C → D induces the strong monoidal functors

Lact(B) → Lact(B), Lcoact(B) → Lcoact(B) (2.29) e e for bimonoids B ∈ Bimon(C) and B = Bimon(F )B ∈ Bimon(D). e

16 2.5 Monoids and groups in a category with finite products A category with finite products is an important case of monoidal category. Here we consider this case in details (see [MacLane, § 3.5, 3.6]). 2.5.1. Category with finite products. Let C be a category. Suppose that there exists a terminal object E and a categorical product X ×Y of any two objects X,Y ∈ C. Then there n exist all the finite products Xi, where Xi ∈ C, n ∈ N0 (for n = 0 this product is equal i=1 to E). In this case (C, ×) is aQ symmetric monoidal category [MacLane, § 7.7]. For morphisms f : Y → X1 and g : Y → X2 we denote by (f,g) the unique morphism Y → X1 × X2 making the diagram

Y ❏ (2.30) tt ❏❏ f tt ❏❏ g tt (f,g) ❏❏ tt ❏❏ ztt  ❏$ X o X × X / X 1 p1 1 2 p2 2 commutative, where p1 : X1 × X2 → X1 and p2 : X1 × X2 → X2 are canonical projections of the product X1 × X2. More generally, for fi : Y → Xi there is a unique morphism n n f = (f1,...,fn): Y → Xi such that pi · f = fi, where pi : Xi → Xi are canonical i=1 i=1 projections. Q Q 2.5.2. Diagonal morphism. Any object X of a category with finite products C is equipped with a structure of comonoid in a unique way. Indeed, since the object E is terminal, there idX ×εX is a unique morphism εX : X → E. The morphisms X × X −−−−−→ X × E ∼= X and εX ×idX ∼ X × X −−−−→ E × X = X coincide with the canonical projections p1 and p2 respectively, so the morphism ∆X : X → X × X can be uniquely found from the commutative diagram

X ▲ (2.31) rrr ✤ ▲▲▲▲▲ rrr ✤ ▲▲▲▲ rrrr ∆X ▲▲▲ rrr ✤ ▲▲▲▲ rrr  ▲▲ X × Eo X × X / E × X idX ×εX εX ×idX

This is ∆X = (idX , idX ): X → X × X, it is called diagonal morphism for the object X. One can check that the composition of both morphisms (idX ×∆X ) · ∆X : X → X × X × X and (∆X × idX ) · ∆X : X → X × X × X with all three canonical projections X × X × X → X are equal to idX . Thus they coincide with the morphism (idX , idX , idX ) and hence equal to op each other. Note that the diagonal morphism is cocommutative: ∆X = ∆X . In this way we obtain a unique comonoid (X, ∆X ,εX ) ∈ Comon(C, ×) for any ob- ject X ∈ C. Moreover, any morphism f : X → Y in C is a morphism of comonoids (X, ∆X ,εX) → (Y, ∆Y ,εY ). This means that Comon(C, ×) coincide with C as a cate- gory. Since the monoidal product in Comon(C, ×) coincide with the direct product × in C, we have the strict monoidal equivalence

Comon(C, ×)= cocComon(C, ×)=(C, ×). (2.32)

17 2.5.3. Monoids and groups in the category C. A monoid M =(X,µX , ηX ) in (C, ×) is called monoid in the category C. Due to the equivalence (2.32) any structure of monoid on X ∈ C is compatible with the structure (∆X ,εX ). Hence the monoid M has a unique struc- ture of bimonoid B =(X,µX , ηX , ∆X ,εX ) ∈ Bimon(C, ×) given by the diagonal morphism ∆X : X → X × X and the unique morphism εX : X → E, so we obtain the strict monoidal equivalence Bimon(C, ×)= Mon(C, ×). (2.33) Note also that any bimonoid B ∈ Bimon(C, ×) is cocommutative. The category Set is a category with finite products. The terminal object in Set is the one-point set E = {∗}. For a set X ∈ Set the diagonal morphism ∆X : X → X × X has the form ∆X (x)=(x, x) and the morphism εX : X → E maps any element x ∈ X to the unique element ∗ of the set E. A monoid M = (X,µX , ηX ) in (Set, ×) is a monoid in the usual sense: µX (x, y) = xy is a multiplication map and ηX (∗) = e gives the neutral element e ∈ X. It becomes a bimonoid in a unique way: B = (X,µX , ηX , ∆X ,εX ). The convolution of two maps α, β : X → X is the point-wise product: (α ⋆ β)(x) = α(x)β(x); the neutral element for this convolution is the map ηX · εX : x 7→ e. Hence the monoid M = (X,µX , ηX ) ∈ Mon(Set, ×) is a group iff the bimonoid B = (X,µX , ηX , ∆X ,εX ) is a −1 Hopf monoid in (Set, ×). The role of antipode is played by the map ζX (x)= x . For general category with finite products C the monoid M = (X,µX , ηX ) in C is called group in the category C if the corresponding bimonoid B = (X,µX , ηX , ∆X ,εX ) is a Hopf monoid in (C, ×). It is enough to require one of the conditions µX · (ζX × idX ) · ∆X = ηX · εX or µX · (idX ×ζX ) · ∆X = ηX · εX , since they are equivalent in this case. 2.5.4. Algebraic monoids and algebraic groups. The category AlgSet also has all finite products. Indeed, the terminal object is A0, the product of two algebraic sets X ⊂ An and Y ⊂ Am is the set-theoretic product X ×Y canonically embedded into An ×Am = An+m. As we already mentioned in p. 2.3.2 the category CommAlg has all finite coproducts and hence AffSch = CommAlgop has all finite products. Since the categorical product of the affine schemes Spec R and Spec S is Spec(R ⊗ S), we have A(X × Y )= A(X) ⊗ A(Y ) for any X,Y ∈ AlgSet. An (affine) algebraic monoid is a monoid in AlgSet. Explicitly, this is a usual monoid that has a structure of algebraic set X such that the multiplication map X × X → X, (x, y) 7→ xy, is a morphism in AlgSet. An (affine) algebraic group is a group in AlgSet. This is a group with a structure of algebraic set X such that the maps X × X → X, (x, y) 7→ xy, and X → X, x 7→ x−1, are morphisms in AlgSet. More generally, one can define affine monoid/group scheme as a monoid/group in AffSch. Structure of monoid on an affine scheme X = Spec R is given by the homomorphisms ∆R : R → R ⊗ R and εR : R → K making R ∈ CommAlg into a commutative bialgebra. This is a structure of affine group scheme iff this bialgebra (R, ∆R,εR) is a Hopf algebra. Thus, a commutative bialgebra/Hopf algebra is the same as an affine monoid/group scheme (modulo direction of their morphisms).

18 3 Internal hom and representations

A linear representation of an algebra A ∈ Mon(Vect, ⊗) on a vector space V ∈ Vect is an algebra homomorphism ρ: A → End(V ), but from the categorical point of view the object End(V ) ∈ Mon(Set, ×) is a monoid without a structure of vector space. To require the linearity of ρ we need to equip End(V ) with a linear structure. This can be done by using the internal hom-functor. This functor and its generalised version (adjunction with a parameter) allow us to consider representations in more general monoidal categories.

3.1 Internal (co)hom-functor and its generalisation 3.1.1. Closed categories. A symmetric monoidal category C = (C, ⊗) is called closed if for any Y ∈ C the functor − ⊗ Y : C → C has a right adjoint hom(Y, −): C → C. In this op case there exists a unique bifunctor hom = homC : C ⊗ C → C, which is called (internal) hom-functor, and an isomorphism ∼ θ = θX,Y,Z : Hom X, hom(Y,Z) −→ Hom X ⊗ Y,Z (3.1)   natural in X,Y,Z ∈ C (see details in [Bor2, § 6.1] or [MacLane, § 4.7]). The natural isomorphism (3.1) is also unique.2 The object hom(X,Y ) ∈ C is called (internal) hom- object. Due to symmetricity of C the functors − ⊗ Y and Y ⊗ − are isomorphic, so their right adjoints are the same (up to an isomorphism of functors). 3.1.2. Adjunction with a parameter. The classical example of a generalisation of the notion of internal hom appears in the theory of bimodules. Let A, B, C ∈ Alg. Let M and N be (A, B)- and (B, C)-bimodules respectively. Their tensor product is the (A, C)-module M ⊗B N. We have a natural isomorphism ∼ Hom(A,B)(M, homC(N,K)) = Hom(A,C)(M ⊗B N,K), (3.2) where K is a (A, C)-bimodule, Hom(A,B)(−, −) is the usual ‘external’ Hom-functor in the cat- egory of the (A, B)-bimodules and homC(N,K) is the set of right C-module homomorphisms N → K equipped with the structure of (A, B)-module in a natural way. In general, let F : C × P → C′ be a bifunctor such that F (−,Y ): C → C′ has a right ′ op ′ adjoint GY : C → C for any Y ∈ P. Then there exists a unique bifunctor G: P × C → C such that GY = G(Y, −) and the adjunction isomorphism

HomC′ F (X,Y ),Z ∼= HomC X,G(Y,Z) (3.3)   is natural in X ∈ C, Y ∈ P, Z ∈ C′ (see [MacLane, § 4.7, Th. 3]). This is so-called adjunction with a parameter with the parameter category P.

2If F : C → C′ and F ′ : C → C′ are two left adjoints of a functor G: C′ → C with adjunctions ∼ ′ ′ ∼ ϕX,Y : Hom(FX,Y ) −→ Hom(X, GY ) and ϕX,Y : Hom(F X, Y ) −→ Hom(X, GY ), then there exists a ′ ′ ∗ unique natural isomorphism ϑX : FX → F X such that ϕX,Y = ϕX,Y · ϑX (see the proof of [MacLane, § 4.1, Cor. 1]). In particular, by considering the case F = F ′ we deduce that the adjunction ϕ is unique.

19 3.1.3. Generalised internal hom. If one uses the notation F (X,Y ) = X ⊗ Y for the bifunctor F in (3.3), then we can denote G(Y,Z) by hom(Y,Z) in a generalised sense. The adjunction with a parameter has the form (3.1) where X ∈ C, Y ∈ P, Z ∈ C′. In some important cases the categories C, P and C′ are full subcategories of a symmetric monoidal category such that F (X,Y )= X⊗Y ∈ C′ for any X ∈ C and Y ∈ P. In particular, such situation takes place if P ⊂ C = C′ for a symmetric monoidal category C. 3.1.4. Evaluation. Let Y ∈ P and Z ∈ C′. Substitute X = hom(Y,Z) ∈ C to (3.1) and take idX in the left hand side, then we obtain so-called evaluation morphism

evY,Z : hom(Y,Z) ⊗ Y → Z (3.4)

in the right hand side, that is evY,Z = θ(idhom(Y,Z)). This morphism is the counit of the adjunction − ⊗Y, hom(Y, −), θ−1 and, in particular, it is natural in Z ∈ C′; the isomor- phism (3.1) can be expressed through the evaluation as

θX,Y,Z(f) = evY,Z ·(f ⊗ idY ), f ∈ Hom X, hom(Y,Z) , (3.5)  (see [MacLane, § 4.1]).

Proposition 3.1. Let X, X ∈ C, Y ∈ P, Z ∈ C′ and f : X → hom(Y,Z), g : X → X be morphisms in C. Then e e

θX,Y,Ze (f · g)= θX,Y,Z(f) · (g ⊗ idY ). (3.6) Proof. By virtue of (3.5) we derive

θX,Y,Ze (f · g) = evY,Z · (f · g) ⊗ idY = evY,Z ·(f ⊗ idY ) · (g ⊗ idY )= θX,Y,Z (f) · (g ⊗ idY ),  where we used the functoriality of −⊗−: C×P → C′ with respect to the first argument. 3.1.5. Internal composition. Remind that the usual ‘external’ composition is the map Hom(Y,Z) × Hom(X,Y ) → Hom(X,Z) (morphism in Set). One can define the analogous morphism for the internal hom objects as a morphism in C, where the role of × is played by ⊗ (see [Bor2, sec. 6.1]). Let us consider the case of full subcategory P ⊂ C = C′. For arbitrary objects X,Y ∈ P and Z ∈ C consider the morphism

id ⊗ ev ev hom(Y,Z) ⊗ hom(X,Y ) ⊗ X −−−−−−→X,Y hom(Y,Z) ⊗ Y −−−→Y,Z Z. (3.7)

By applying the map

θ−1 Hom hom(Y,Z) ⊗ hom(X,Y ) ⊗ X,Z −−→ Hom hom(Y,Z) ⊗ hom(X,Y ), hom(X,Z)   to the morphism (3.7) we obtain a morphism

cX,Y,Z : hom(Y,Z) ⊗ hom(X,Y ) → hom(X,Z),X,Y ∈ P,Z ∈ C.

20 It is called (internal) composition morphism. This composition is associative in the sense that cA,C,D · (id ⊗cA,B,C )= cA,B,D · (cB,C,D ⊗ id) for any A,B,C ∈ P, D ∈ C. 3.1.6. Internal end. Suppose again that P is a full subcategory of a symmetric monoidal category C = (C, ⊗,I) = C′. By substituting X = I and Z = Y ∈ P to (3.1) and taking the identification isomorphism I ⊗ Y = Y in the right hand side we obtain a morphism −1 uY := θ (idY ): I → hom(Y,Y ). The diagram

I ⊗ hom(X,Y ) hom(X,Y ) hom(X,Y ) ⊗ I (3.8) ❥❥4 j❚❚❚ ❥❥❥❥ ❚❚❚❚ uY ⊗id ❥❥❥ ❚❚❚ id ⊗uX ❥❥❥cX,Y,Y cX,X,Y ❚❚❚  ❥❥❥❥ ❚❚❚  hom(Y,Y ) ⊗ hom(X,Y ) hom(X,Y ) ⊗ hom(X,X) commutes for any X,Y ∈ P. In particular, for each X ∈ P the morphisms cX := cX,X,X and uX give a structure of monoid on the object hom(X,X). We denote this monoid by end(X). 3.1.7. Examples of closed categories.

• In (Set, ×) the internal hom-object hom(X,Y ) coincides with the ‘external’ Hom- object Hom(X,Y ). The evaluation is the map evX,Y : Hom(X,Y ) × X → Y that evaluates f ∈ Hom(X,Y ) on an element of X. Internal composition coincides with the usual composition and end(X) = End(X) is the usual monoid of the maps X → X.

• For the objects V and W of the monoidal category (Vect, ⊗) the object hom(V, W ) is the vector space of linear maps V → W . In this case it coincides with Hom(V, W )asa set. The evaluation morphism evV,W : hom(V, W ) ⊗ V → W acts as ev(f ⊗ v)= f(v), f : V → W , v ∈ V . The composition hom(W, Z) ⊗ hom(V, W ) → hom(V,Z) is given by the usual composition in the sense that it maps f ⊗ g to f · g. The monoid end(V ) ∈ Mon(Vect, ⊗) is the algebra of linear operators on V .

• The monoidal category (GrVect, ⊗) is also closed. As a vector space the internal hom in this category coincides with the internal hom in (Vect, ⊗), but it has additionally the structure of grading: the k-th component of the object hom(V, W ) consists of linear maps f : V → W such that f(Vl) ⊂ Wk+l. By considering the zero component of the graded space hom(V, W ) as a set we can identify this component with Hom(V, W ) with the zero component of hom(V, W ) (for this case the internal hom does not coincide with the external one as a set). The evaluation and composition look the same as for (Vect, ⊗). The monoid end(V ) is the graded algebra of linear operators on V . This example is directly generalized to the case of the category of A-graded vector spaces for any abelian monoid A ∈ cMon(Set, ×).

3.1.8. Coclosed categories. We also need dual notions to the notions of internal hom and end. Let C = (C, ⊗) be a symmetric monoidal category such that for any object Y ∈ C

21 the functor − ⊗ Y : C → C has a left adjoint cohom(Y, −): C → C, then there is a unique op bifunctor cohom = cohomC : C ⊗ C → C with a unique natural transformation ∼ ϑ = ϑX,Y,Z : Hom cohom(Y,X),Z −→ Hom X,Z ⊗ Y ) . (3.9)   A symmetric monoidal category C =(C, ⊗) satisfying this condition is called coclosed. Let us call the object cohom(Y,X) (internal) cohom-object. It coincides with the internal hom- op op object hom(Y,X) in the opposite category C . The bifunctor cohomC : C ⊗ C → C op coincides with the bifunctor (homCop ) , it is called (internal) cohom-functor. 3.1.9. Generalised cohom. Let us consider the dual version of the adjunction with a parameter from p. 3.1.2. Denote the bifunctor F : C × P → C′ as F (X,Y ) = X ⊗ Y . If − ⊗ Y : C → C′ has a left adjoint for any fixed Y ∈ P, then we obtain a bifunctor cohom: Pop × C′ → C with an isomorphism ∼ ϑ = ϑX,Y,Z : HomC′ cohom(Y,X),Z −→ HomC X,Z ⊗ Y (3.10)   natural in X ∈ C, Y ∈ P, Z ∈ C′. The coevaluation is defined as the morphism

coevY,X = ϑ(idhom(Y,X)): X → cohom(Y,X) ⊗ Y, (3.11) which is natural in X ∈ C. Then the isomorphism (3.10) can be expressed as

ϑ(f)=(f ⊗ idZ ) · coevY,X, f : cohom(Y,X) → Z. (3.12)

3.1.10. Internal coend. Suppose that P is a full subcategory of a symmetric monoidal category C = C′ =(C, ⊗). By using categorical duality to the morphisms defined in p. 3.1.5 and 3.1.6 we obtain the cocomposition

−1 dX,Y,Z = ϑ (id ⊗ coevZ,Y ) · coevY,X : cohom(Z,X) → cohom(Y,X) ⊗ cohom(Z,Y )  (3.13) and morphisms vY : cohom(Y,Y ) → I, where X ∈ C, Y,Z ∈ P. The pair (dY := dY,Y,Y , vY ) turns cohom(Y,Y ) into a comonoid in C. Denote it by coend(Y ) (not to confuse with ‘Coend of a functor’). 3.1.11. Internal cohom for vector spaces. Since the strong monoidal functor (−)∗ is an equivalence between FVect and FVectop, the monoidal category (FVect, ⊗) is coclosed. The internal cohom-functor can be extended for the case P = FVect, C = C′ = Vect.

Proposition 3.2. For any W ∈ FVect the functor −⊗W : Vect → Vect has a left adjoint cohom(W, −): Vect → Vect, which coincides with the internal hom-functor:

cohom(W, V )= hom(W, V ), V ∈ Vect. (3.14)

22 m i m ∗ i i Proof. Let (wi)i=1, (w )i=1 be dual bases of W and W , that is w (wj) = δj. Note that for any v ∈ V and ξ ∈ W ∗ the element ξ ⊗ v ∈ W ∗ ⊗ V can be considered as the linear i i operator W → V that maps w ∈ W to ξ(w)v ∈ V (in particular, (w ⊗ v)(wj) = δjv). This gives a natural isomorphism hom(W, V ) ∼= W ∗ ⊗ V . Let us define the linear operator ∗ m i ηV : V → hom(W, V ) ⊗ W = (W ⊗ V ) ⊗ W by the formula ηV (v) = i=1(w ⊗ v) ⊗ wi. It can be also defined in the form ηV (v) = (idW ⊗σW,V )(1end(W ) ⊗ v),P where 1end(W ) = ∗ uW (1) ∈ end(W, W )= W ⊗ W is the unity of the algebra end(W ), so it does not depend on the choice of the basis. Let us show that for any V,Z ∈ Vect and a linear operator f : V → Z ⊗ W there is a unique h: hom(W, V ) → Z such that the diagram

ηV V ❙ / hom(W, V ) ⊗ W (3.15) ❙❙❙❙ ❙❙❙❙ ❙❙❙ h⊗idW f ❙❙❙❙ ❙❙)  Z ⊗ W

m i commute. By decomposing the image f on v ∈ V we obtain f(v)= i=1 f (v) ⊗ wi, where ∗ ∗ f i(v) is the value of f(v) ∈ Z ⊗ W = hom(W ,Z) on wi ∈ W .P The commutativity of the diagram implies that h(wi ⊗ v)= f i(v)= f(v)(wi), so h is unique. It exists due to the ∗ linearity of f : V → Z ⊗ W and f(v): W → Z. We obtain the universal arrow ηV from V to the functor − ⊗ W and it is straightforward to check that ηV is natural in V . Hence the functor hom(W, −) is a left adjoint of − ⊗ W (see [MacLane, § 4.1, Th. 2 (i)]).

3.2 (Co)representations of (co)monoids 3.2.1. Representation of a usual monoid or a group. Remind that for any object V of a category C the set End(V ) = Hom(V,V ) is equipped with a structure of usual monoid (monoid in Set). In a wide generality we can say that a representation is a monoid homo- morphism ρ: M → End(V ). In particular, if M is a group, then ρ(m) is an automorphism of V for any m ∈ M, so we obtain a representation of a group M by automorphisms of the object V ∈ C. If the monoids M and End(V ) are equipped with some additional structure of the same type, one usually requires ρ to preserve this structure. A morphism from a representation ρ: M → End(V ) to a representation ρ′ : M → End(V ′) is a morphism f : V → V ′ in C such that f ·ρ(m)= ρ′(m)·f for any m ∈ M. In the pointless ∗ ′ form this condition reads f∗ · ρ = f · ρ . 3.2.2. Representation of an algebra. For example, let V ∈ Vect; if the monoid M has additionally a structure of vector space compatible with the monoid structure, then both M and End(V ) are algebras and we require ρ: M → End(V ) to be an algebra homomorphism, so we obtain the notion of ‘representation of an algebra’. More precisely, one should write there end(V ) instead of End(V ). As we will see below this case can be generalised to an arbitrary closed symmetric monoidal category (see p. 3.2.4). Remind that the representations of a group G (a group in Set) or of a g can be considered as particular cases of representations of algebras, these are the representations

23 of the corresponding group algebra K[G] or of the universal enveloping algebra U(g) respec- tively. In the same way representations of a monoid M ∈ Mon(Set, ×) can be identified with representations of the algebra K[M] consisting of the formal sums αmm, αm ∈ K. M mP∈ 3.2.3. Representation of an algebraic monoid/group. Another classical example is a structure of algebraic set on a monoid/group M. Suppose that this structure is compat- ible with the structure of monoid/group in the sense that M is an algebraic monoid/group (see p. 2.5.4). For V ∈ FVect the monoid End(V ) has a structure of algebraic monoid. In this way we obtain the representations of an algebraic monoid/group M on a vector space V . In particular, the representation ρ: M → End(Kn) gives an action a: M × An → An of a monoid/group M on the object An ∈ AlgSet. In other words, ‘linear’ representations of a monoid/group M ∈ Mon(AlgSet, ×) is a particular case of an action of M. 3.2.4. The case of a . Let C =(C, ⊗) be a closed symmetric monoidal category and M = (X,µX , ηX ) ∈ Mon(C) be a monoid in C. Define representation of M on an object V ∈ C as a morphism ρ: M → end(V ) in the category Mon(C). This is a morphism ρ: X → hom(V,V ) in C such that the diagrams

µX ηX X ⊗ X / X I / X (3.16) ❏❏ ❏❏ ❏❏ ρ⊗ρ ρ ❏❏ ρ uV ❏❏  c  %  hom(V,V ) ⊗ hom(V,V ) V / hom(V,V ) hom(V,V )

are commutative. 3.2.5. The case of generalised internal hom. Now consider more general case. Let C = (C, ⊗) be a symmetric monoidal category and let P be its full subcategory such that the functor − ⊗ Y : C → C has a right adjoint for any Y ∈ P, so we have the generalised hom: Pop × C → C in the sense of p. 3.1.3 (the case P ⊂ C′ = C). We get a monoid end(V ) ∈ Mon(C) for any V ∈ P. In this case we define representation of a monoid M = (X,µX , ηX ) ∈ Mon(C) on an object V ∈ P as a morphism ρ: M → end(V ) in Mon(C). Again, this is a morphism ρ: X → hom(V,V ) making the diagrams (3.16) commutative. 3.2.6. Representation as an action. By substituting Y = Z = V ∈ P to (3.1) we obtain the bijection

θ : Hom X, hom(V,V ) −→∼ Hom X ⊗ V,V . (3.17)   Let us prove that the representations and actions on V are in one-to-one correspondence via this bijection.

Lemma 3.3. A morphism ρ: X → hom(V,V ) is a representation of the monoid M iff a = θ(ρ): X ⊗ V → V is an action of M.

24 Proof. We check that the commutativity of (3.16) is equivalent to the commutativity of (2.21). Prop. 3.1 implies θ(ρ · µX ) = θ(ρ) · (µX ⊗ idV ) = a · (µX ⊗ idV ): X ⊗ X ⊗ V → V . By virtue of the same proposition we have the equalities θ cV · (ρ ⊗ ρ) = evV,V ·(idhom(V,V ) ⊗ evV,V ) · (ρ ⊗ ρ ⊗ idV ) = evV,V ·(ρ ⊗ idV ) · (idX ⊗ evV,V )· (idX ⊗ρ ⊗ id V ) of morphisms X ⊗ X ⊗ V → V . Due to (3.5) one can rewrite the latter one in the form θ cV · (ρ ⊗ ρ) = a · (idX ⊗a). Thus ρ · µX = cV · (ρ ⊗ ρ) iff a · (µX ⊗ idV ) = a · (idX ⊗a). Analogously,  we derive θ(ρ · ηX ) = a · (ηX ⊗ idV ). By taking into account θ(uV ) = idV we deduce the equivalence of the condition ρ · ηX = uV to the condition a · (ηX ⊗ idV ) = idV . M 3.2.7. Morphisms of representations. Denote by RepP( ) the category of the pairs (V, ρ), where V ∈ P and ρ is a representation of M on V ; a morphism (V, ρ) → (V ′, ρ′) ′ in this category is defined as a morphism f : V → V in P satisfying hom(idV , f) · ρ = ′ hom(f, idV ′ ) · ρ . The last equation can be written in the diagram form as

ρ X / hom(V,V ) (3.18)

′ ρ hom(idV ,f)

 hom(f,id ′ )  hom(V ′,V ′) V / hom(V,V ′)

To calculate hom(idV , f) and hom(f, idV ′ ) we will use the following general formula.

Proposition 3.4. Let f : V ′ → V and g : Z → Z′ be morphisms in P and C respectively. Then

θ hom(f,g) = g · evV,Z ·(idhom(V,Z) ⊗f). (3.19)  Proof. From the naturality of θX,V,Z in V and Z we obtain

θ Hom X, hom(V,Z) / Hom X ⊗ V,Z (3.20)

 ∗ hom(f,g)∗ (idX ⊗f) ·g∗

 θ  Hom X, hom(V ′,Z′) / Hom X ⊗ V ′,Z′   Let X = hom(V,Z). By taking idX in the left upper corner of the diagram (3.20) we obtain the formula (3.19). M By virtue of Lemma 3.3 we can regard the objects of the category RepP( ) as objects of Lact(M). The definitions of morphisms in these categories coincide due to the following statement.

Lemma 3.5. Let a = θ(ρ): X ⊗ V → V and a′ = θ(ρ): X ⊗ V ′ → V ′ be actions of M corresponding to the representations ρ: M → end(V ) and ρ′ : M → end(V ′), where V,V ′ ∈ P. Let f : V → V ′ be a morphism in P. Then the commutativity of (3.18) is equivalent to the commutativity of the left diagram (2.23).

25 Proof. Application of the formulae (3.6) and (3.19) gives

′ ′ θ hom(f, idV ′ ) · ρ = evV ′,V ′ ·(idhom(V ′,V ′) ⊗f) · (ρ ⊗ idV )= ′ ′ ′ evV ′,V ′ ·(ρ ⊗ idV ′ ) · (idX ⊗f)= θ(ρ ) · (idX ⊗f)=a · (idX ⊗f),

θ hom(idV , f) · ρ = f · evV,V ·(ρ ⊗ idV )= f · θ(ρ)=f · a,  where we also used (3.5). The Lemmas 3.3 and 3.5 have the following corollary.

Theorem 3.6. Let P be a full subcategory of a symmetric monoidal category C = (C, ⊗). Suppose that the functor − ⊗ V : C → C has a right adjoint for any V ∈ P. Let M = (X,µX , ηX ) ∈ Mon(C). Then the morphisms (3.17) gives a fully faithful functor embedding M M the category RepP( ) into Lact( ). In particular, if the symmetric monoidal category M C = (C, ⊗) is closed then we obtain an equivalence between the categories RepC( ) and Lact(M).

3.2.8. Corepresentations of comonoids. Let us dualise the notions and results of p. 3.2.4–3.2.7. Let P be a full subcategory of a symmetric monoidal category C = (C, ⊗) such that the functor −⊗V has left adjoint for any V ∈ P, so we have the generalised cohom- functor cohom: Pop × C → C with the adjunction (3.10). We call by corepresentation of a comonoid O = (X, ∆X ,εX) ∈ Comon(C) on V ∈ P a morphism ω : coend(V ) → O in Comon(C). This is a morphism ω : cohom(V,V ) → X in C such that the diagrams

ω ω cohom(V,V ) / X cohom(V,V ) / X ▼▼▼ ▼▼ ε dV ∆X v ▼▼ X V ▼▼▼  ω⊗ω  ▼▼&  cohom(V,V ) ⊗ cohom(V,V ) / X ⊗ X I

O are commutative. Denote by CorepP( ) the category whose objects are the pairs (V,ω), where V ∈ P and ω is a corepresentation of O on V , and morphisms (V,ω) → (V ′,ω′) are morphisms f : V → V ′ in P such that

cohom(f,idV ) cohom(V ′,V ) / cohom(V,V ) (3.21)

cohom(idV ′ ,f) ω  ω′  cohom(V ′,V ′) / X

By dualising Prop. 3.4 and Theorem 3.6 we obtain the following statements.

Proposition 3.7. Let f : V ′ → V and g : Z → Z′ be morphisms in P and C respectively. Then

ϑ cohom(f,g) = (idcohom(V ′,Z′) ⊗f) · coevV ′,Z′ ·g. (3.22)  26 Theorem 3.8. Let O =(X, ∆X ,εX) ∈ Comon(C). The isomorphisms

ϑV,V,X : Hom cohom(V,V ),X ∼= Hom V,X ⊗ V , V ∈ P, (3.23) O O  give a fully faithful functor CorepP( ) ֒→ Lcoact( ). In other words, the morphism ω : cohom(V,V ) → X is a corepresentation of O iff δ = ϑ(ω) is a coaction of O on V ∈ P, and a morphism f : V → V ′ in P makes the diagram (3.21) to be commutative for some corepresentations ω,ω′ iff it makes the right diagram (2.23) commutative with the coactions δ = ϑ(ω), δ′ = ϑ(ω′). In particular, if the symmetric monoidal category C =(C, ⊗) is coclosed then we obtain O O an equivalence between the categories CorepC( ) and Lcoact( ).

3.3 Translation of (co)representations under monoidal functors As we saw in p. 2.4.11 a (co)lax functor translates (co)actions to (co)actions. Theorems 3.6 and 3.8 imply that such functor translates the corresponding (co)representations to each other. Here we describe the translation of (co)representations explicitly. 3.3.1. Translation of internal hom. Let C = (C, ⊗) and D = (D, ⊙) be symmetric monoidal categories. Let P ⊂ C and Q ⊂ D be their full subcategories such that the functors −⊗V : C → C and −⊙W : D → D have right adjoints for each V ∈ P and W ∈ Q, so we have generalised internal hom-functors hom: Pop × C → C and hom: Qop × D → D. Let F : (C, ⊗) → (D, ⊗) be a lax monoidal functor with the monoidal structure mor- phisms φX,Y : FX ⊙ FY → F (X ⊗ Y ), ϕ: ID → FIC. Suppose F (P) ⊂ Q. For arbitrary V ∈ P and Z ∈ C consider the composition

φ F (evV,Z ) F hom(V,Z) ⊙ FV −→ F hom(V,Z) ⊗ V −−−−−→ FZ. (3.24)   By applying the adjunction θ−1 : Hom F hom(V,Z) ⊙ FV,FZ −→∼ Hom F hom(V,Z) , hom(FV,FZ) (3.25)     to (3.24) we obtain the following morphism in D:

−1 ΦV,Z = θ F (evV,Z) · φhom(V,Z),V : F hom(V,Z) → hom(FV,FZ). (3.26)   Thus we get a collection of morphisms ΦV,Z, V ∈ P, Z ∈ C, making the diagram

ΦV,Z ⊙idF V F hom(V,Z) ⊙ FV / hom(FV,FZ) ⊙ FV (3.27)  φ evFV,FZ

 F (evVZ )  F hom(V,Z) ⊗ FV / FZ  commute (the equivalence of the commutativity of (3.27) and the definition (3.26) follows from the formula (3.5)). The naturality of φX,Y in X ∈ C allows to prove the following properties of ΦV,Z.

27 Proposition 3.9. For any objects X,Z ∈ C, V ∈ P and a morphism f : X → hom(V,Z) in C we have the formula

θ ΦV,Z · F (f) = F θ(f) · φX,V . (3.28)   Proof. Note that (3.28) is equivalent to the commutativity of the diagram

F (f)⊙idF V ΦV,Z ⊙idF V FX ⊙ FV / F hom(V,Z) ⊙ FV / hom(FV,FZ) ⊙ FV (3.29)  φ evFV,FZ

 F (f⊗idV ) F (evVZ )  F (X ⊗ V ) / F hom(V,Z) ⊗ FV / FZ  By adding the vertical arrow φ to the centre we obtain two diagrams. Commutativity of the left one follows from the naturality of φ, while the right one is exactly the diagram (3.27). Warning. Even when F is strong monoidal, i.e. all φX,Y and φ are isomorphisms, we can not guarantee that ΦV,Z are also isomorphisms. 3.3.2. Translation of the internal end. Remind that the lax monoidal functor F : C → D induces the functor Mon(F ): Mon(C) → Mon(D) which translates each monoid M = (X,µX , ηX ) ∈ Mon(C) to the monoid Mon(F )M = (FX,µFX, ηFX ) ∈ Mon(D), where µFX : FX ⊙ FX → FX and ηFX : ID → FX are the compositions (2.17). In particular, if V ∈ P the monoid end(V ) is mapped to the monoid Mon(F ) end(V ) . The latter is the object F hom(V,V ) with the structure morphisms F (cV )·φhom (V,V ),hom(V,V ) and F (uV )·ϕ. On the other hand, since FV ∈ Q we have the monoid

end(FV )= hom(FV,FV ),cF V ,uF V ∈ Mon(D). (3.30)  We get a relationship between these two monoids by means of the morphisms (3.26).

Proposition 3.10. For any V ∈ P the morphism ΦV,V : F hom(V,V ) → hom(FV,FV ) in D gives a morphism Mon(F ) end(V ) → end(FV ) in Mon(D). In particular, if ΦV,V is an isomorphism in D then the monoids Mon(F ) end(V ) and end(FV ) are isomorphic to each other (as objects of Mon(D)).  Proof. The first sentence of the proposition will be proved in p. 3.3.3 in more general settings (see Remark 3.12). The second sentence follows from the first one and Prop. 2.1. 3.3.3. Translation of representations. Since F (P) ⊂ Q, the restriction of the func- tor (2.27) gives the functor M M RepP( ) → RepQ( ), (3.31) where M = Mon(F )M. Let us describe it explicitly. e M Propositione 3.11. The functor (3.31) maps an object (V, ρ) ∈ RepP( ) to (FV, ρ), where ρ: M → end(FV ) is the composition e F ρ ΦV,V e FX −→ F hom(V,V ) −−−→ hom(FV,FV ). (3.32)  28 Proof. Let a = θ(ρ). The functor (2.27) maps (V, a) to (FV, a), where a = F (a) · φX,V . By applying Prop. 3.9 to f = ρ we derive θ(ΦV,V · F ρ) = F θ(ρ) · φ = F (a) · φ = a, so the action a corresponds to the representation (3.32). e e e Remarke 3.12. The fact that (3.32) is a morphism of monoids M → end(FV ) follows im- mediately from Prop. 3.10 and functoriality of Mon(F ). Conversely, by applying Prop. 3.11 e to the case X = hom(V,V ), ρ = idX and by taking into account the statements written in p. 2.4.11 we see that ΦV,V is a representation. This implies the first sentence of Prop. 3.10.

3.3.4. Translation of internal cohom, coend and of corepresentations. In the same setting suppose that the functors − ⊗ V : C → C and − ⊙ W : D → D have left adjoints for each V ∈ P and W ∈ Q instead of right adjoints, so we have generalised internal cohom-functors cohom: Pop × C → C and cohom: Qop × D → D. Let F = (F,ϕ,φ): (C, ⊗) → (D, ⊙) be a colax monoidal functor such that F (P) ⊂ Q. For arbitrary V ∈ P and Z ∈ C consider the composition

F (coevV,Z ) φ FZ −−−−−−→ F cohom(V,Z) ⊗ V −→ F cohom(V,Z) ⊙ FV. (3.33)   By applying the adjunction we obtain

−1 ΦV,Z = ϑ φcohom(V,Z),V · F (coevV,Z) : cohom(FV,FZ) → F cohom(V,Z) , (3.34)   where V ∈ P, Z ∈ C. These morphisms satisfies

ϑ F (f) · ΦV,Z = φX,V · F ϑ(f) (3.35)   for any X,Z ∈ C, V ∈ P and f ∈ HomC cohom(V,Z),X .  Proposition 3.13. For any V ∈ P the morphism ΦV,V gives a morphism of comonoids

ΦV,V : coend(FV ) → Comon(F ) coend(V ) . (3.36)  In particular, if ΦV,V is an isomorphism in D, then these monoids are isomorphic.

Let O = (X, ∆X ,εX ) ∈ Mon(C) and O = Comon(F )O. Then the restriction of the functor (2.28) is the functor e O O CorepP( ) → CorepQ( ). (3.37) e It maps the object (V,ω) to (FV, ω), where ω : coend(FV ) → O is a corepresentation of O on FV given by the composition e e e e ΦV,V Fω coend(FV ) −−−→ F coend(V ) −−→ FX. (3.38) 

29 3.3.5. Translation of monoidal product of (co)representations. Let F : C → D be a symmetric strong monoidal functor such that F (P) ⊂ Q. Let B = (X,µX , ηX , ∆X ,εX ) be a bimonoid in C and B = Bimon(F )B be the corresponding bimonoid in D. Then the restriction of the functors (2.29) are the functors (3.31) and (3.37), where we have M = e (X,µX , ηX ) ∈ Mon(C), M = Mon(F )M ∈ Mon(D) and O = (X, ∆X ,εX) ∈ Comon(C), O = Comon(F )O ∈ Comon(D). Since the functors (2.29) are symmetric strong monoidal, e the restriction gives the symmetric strong monoidal functors e B B B B RepP( ) → RepQ( ), CorepP( ) → CorepQ( ). (3.39) For example, let ρ and π be representationse of the monoid M on W ∈eP and Z ∈ P. B Their monoidal product in RepP( ) is a representation τ on W ⊗Z. The functor (3.31) maps them to representations ρ, π and τ on F W , FZ and F (W ⊗ Z) respectively. The monoidal B product of ρ and π in RepQ( ) is a representation λ on F W ⊙ FZ. The isomorphism e e e φW,Z : F W ⊙ FZ −→∼ F (W ⊗ Z) gives the isomorphism F W ⊙ FZ,λ −→∼ F (W ⊗ Z), τ e e e in Rep (B). In the same way one can describe the isomorphism of the corresponding Q e corepresentations. e Remark 3.14. The strong monoidality of the functors (3.39) follows from the strong monoidal- ity of the functors (2.29). Alternatively, one can prove this in a direct way by using the formula (3.19) and Prop. 2.1.

4 Quantum linear spaces

Idea. Consider the sets Kn and Km as vector (linear) spaces, that is as objects of Vect or FVect. The morphisms between these vector spaces are linear maps Kn → Km (ma- trices m × n over K). If we (partially) forget linear structure of these spaces we ob- tain the affine spaces An and Am, which are the objects of AlgSet ⊂ AffSch. Since AffSch = CommAlgop and A(An) = K[x1,...,xn], the morphisms An → Am is in one- to-one correspondence with the algebra homomorphisms f : K[y1,...,ym] → K[x1,...,xn]. i 1 n 1 n Namely, if a homomorphism f is given by the images f(y )= Pi(x ,...,x ) ∈ K[x ,...,x ] n m for some polynomials Pi, then the corresponding morphism Φ: A → A is the map 1 n 1 n 1 n Φ(x ,...,x ) = P1(x ,...,x ),...,Pm(x ,...,x ) . Now let us recall back the linear structure of Kn = An and Km = Am. The map Φ preserves the linear structure iff all 1 n n j Pi are homogeneous polynomials of order 1, that is Pi(x ,...,x ) = j=1 aijx for some aij ∈ K. This means exactly that the homomorphism f preserves the gradingP of the algebras K[y1,...,ym] and K[x1,...,xn] (the k-th component of K[x1,...,xn] is the space of all the homogeneous polynomials of order k). Any linear map Kn → Km is arisen in this way: it is given by the m × n matrix (aij). Hence a natural candidates to a quantum analogue of a vector spaces are graded algebras or at least some of them, considered as objects of (a subcategory of) the category GrAlgop. In [Man87, Man88] Yuri Manin proposed a subcategory FQAop to generalise the finite- dimensional vector spaces. He introduced the notion ‘quantum linear space’ for this case.

30 4.1 Operations with quadratic algebras 4.1.1. Manin’s binary operations. Let A, B ∈ QA be arbitrary quadratic algebras (here we do not suppose that the quadratic algebras are finitely generated). They can be presented in the form A = TV/(R), B = TW/(S), (4.1) where V, W ∈ Vect and (R) ⊂ TV , (S) ⊂ T W are ideals generated by subspaces R ⊂ V ⊗2, S ⊂ W ⊗2 respectively. Define the following operations [Man87, Man88].

• Manin white product: A ◦ B = T (V ⊗ W )/(Rw), where

Rw = (idV ⊗σV,W ⊗ idW )(R ⊗ W ⊗ W + V ⊗ V ⊗ S) ⊂ V ⊗ W ⊗ V ⊗ W.

• Manin black product: A • B = T (V ⊗ W )/(Rb), where

Rb = (idV ⊗σV,W ⊗ idW )(R ⊗ S) ⊂ V ⊗ W ⊗ V ⊗ W.

• (Even) tensor product: A ⊗ B = T (V ⊕ W )/ R ⊕ [V, W ] ⊕ S , where  [V, W ] = {v ⊗ w − w ⊗ v | v ∈ V,w ∈ W } ⊂ (V ⊗ W ) ⊕ (W ⊗ V ) ⊂ (V ⊕ W )⊗2.

• Odd tensor product: A ⊗ B = T (V ⊕ W )/ R ⊕ [V, W ]+ ⊕ S , where  ⊗2 [V, W ]+ = {v ⊗ w + w ⊗ v | v ∈ V,w ∈ W } ⊂ (V ⊗ W ) ⊕ (W ⊗ V ) ⊂ (V ⊕ W ) .

4.1.2. Properties of Manin’s binary operations [Man88]. The Manin white product and the even tensor product coincide with the Manin product A ◦ B and the usual tensor product A ⊗ B defined in p. 2.3.2 for general graded algebras. All four operations are bifunctors QA × QA → QA equipping QA with monoidal structures. The unit objects of (QA, ◦) and (QA, •) are K[u] and K[ǫ]/(ǫ2) respectively. The unit object of (QA, ⊗) and (QA, ⊗) is K. The subcategory FQA ⊂ QA inherits all four monoidal structures. The inclusion Rb ⊂ Rw defines an epimorphism A•B→A◦B.

4.1.3. The functor (−)1. If we have a graded algebra A = A1 we can take its N kL∈ 0 first graded component A1. In this way we obtain the functor (−)1 : GrAlg → Vect, which maps a graded homomorphism f ∈ HomGrAlg(A, B) to its first order component f1 : A1 → B1. Restriction gives functors (−)1 : QA → Vect and (−)1 : FQA → FVect, which translate a quadratic algebra A = TV/(R) to the vector space V . If A ∈ QA, then a graded homomorphism f : A → B can be uniquely restored by the component f1. Hence the functors (−)1 : QA → Vect and (−)1 : FQA → FVect are full. By applying these functors to the result of the Manin’s binary operations we obtain the following natural isomorphisms:

(A ◦ B)1 =(A • B)1 = A1 ⊗ B1, (A ⊗ B)1 =(A ⊗ B)1 = A1 ⊕ B1. (4.2)

31 This means that the functors (−)1 : QA → Vect and (−)1 : FQA → FVect are equipped with the following strong monoidal structures

(QA, ◦) → (Vect, ⊗), (FQA, ◦) → (FVect, ⊗), (4.3) (QA, •) → (Vect, ⊗), (FQA, •) → (FVect, ⊗), (4.4) (QA, ⊗) → (Vect, ⊕), (FQA, ⊗) → (FVect, ⊕), (4.5) (QA, ⊗) → (Vect, ⊕), (FQA, ⊗) → (FVect, ⊕). (4.6)

The fullness of the functor (−)1 : QA → Vect and the left isomorphism (4.2) imply that the epimorphism A•B→A◦B is natural in A, B ∈ QA. Note also that the functors T : Vect → QA and T : FVect → FQA considered in p. 2.3.4 are left adjoints of (−)1 : QA → Vect and (−)1 : FQA → FVect respectively. We see that the binary operations ◦ and • are related with the tensor product of vector spaces, while ⊗ and ⊗ are related with the direct sum ⊕. Below we introduce two embedding of FVect into FQAop: ‘even’ and ‘odd’. The former one will relate the monoidal products ⊗ and ⊕ on FVect to ◦ and ⊗ on FQA, while the odd embedding will relate ⊗ and ⊕ to • and ⊗ respectively.

4.1.4. Purely even/odd quadratic super-algebras. Let Z2 = Z/2Z = {0¯, 1¯}. A quadratic super-algebra over K is a quadratic algebra A = TV/(R) with the structure of Z2-grading compatible with multiplication and with N0-grading in the sense that each com- ponent Ak is a direct sum of its subspaces of even and odd elements: A = Ak = N kL∈ 0 (Ak)0¯ ⊕(Ak)1¯. Such algebra is generated by (basis) elements of the subspaces V0¯ =(A1)0¯ N kL∈ 0 and V1¯ = (A1)1¯. We call it commutative if ab = −ba for a, b ∈ V1¯ and ab = ba for a ∈ V0¯, b ∈ V (this is exactly the commutativity defined in p. 2.4.4 for the monoidal category of super-vector spaces [Del, § 1.1, 1.2]). We call a quadratic super-algebra A = TV/(R) purely even or purely odd if V = V0¯ or V = V1¯ respectively. The commutative purely even and K odd quadratic super-algebras over form categories, which we denote by CommQSAeven and CommQSAodd. They can be identified with subcategories of QA consisting of com- mutative A ∈ QA (i.e. ab = ba ∀ a, b ∈ A) and of A ∈ QA satisfying a1b1 = −b1a1 ∀ a1, b1 ∈ A1 respectively. The additional condition that they are finitely generated gives the subcategories

FCommQSAeven = CommQSAeven ∩ FQA,

FCommQSAodd = CommQSAodd ∩ FQA.

4.1.5. The functors S and Λ. For a vector space V ∈ Vect and are the quadratic algebras

SV = TV/ {v1 ⊗ v2 − v2 ⊗ v1 | v1, v2 ∈ V } , (4.7)

ΛV = TV/{v1 ⊗ v2 + v2 ⊗ v1 | v1, v2 ∈ V }. (4.8)  32 We obtain the functors S : Vect → QA and Λ: Vect → QA and their ‘finite-dimensional’ versions S : FVect → FQA and Λ: FVect → FQA. The functors S and Λ are left ad- joint functors to the restrictions of the functor (−)1 : QA → Vect to the subcategories CommQSAeven and CommQSAodd respectively. Analogously, the finite-dimensional ver- sions are left adjoints to the restrictions of (−)1 : FQA → FVect to the subcategories FCommQSAeven and FCommQSAodd. Denote by SkV and ΛkV the k-th graded components of SV and ΛV . In these notations we can write the following natural isomorphisms

k k Sk(V ⊕ W ) ∼= (SlV ) ⊗ (Sk−lW ), Λk(V ⊕ W ) ∼= (ΛlV ) ⊗ (Λk−lW ) (4.9) Ml=0 Ml=0 in Vect. Summation over k ∈ N0 gives the natural isomorphisms

S(V ⊕ W ) ∼= (SV ) ⊗ (SW ), Λ(V ⊕ W ) ∼= (ΛV ) ⊗ (ΛW ) (4.10) in QA. They are the structure morphisms of the strong monoidal functors

S : (Vect, ⊕) → (QA, ⊗), S : (FVect, ⊕) → (FQA, ⊗), (4.11) Λ: (Vect, ⊕) → (QA, ⊗), Λ: (FVect, ⊕) → (FQA, ⊗). (4.12)

On the other hand the natural transformations

S(V ⊗ W ) → (SV ) ◦ (SW ), (ΛV ) • (ΛW ) → Λ(V ⊗ W ) (4.13) are not isomorphisms even for finite-dimensional V and W . This does not give strong monoidal functors. Nevertheless, we obtain colax monoidal functors

S : (Vect, ⊗) → (QA, ◦), S : (FVect, ⊗) → (FQA, ◦) (4.14) and lax monoidal functors

Λ: (Vect, ⊗) → (QA, •), Λ: (FVect, ⊗) → (FQA, •). (4.15)

The functors S and Λ are faithful since their compositions with the functor (−)1 coincide with the identical functor: (SV )1 = V , (ΛV )1 = V . Moreover, any morphism SV → SW or ΛV → ΛW is uniquely given by its first component and hence have the form Sf or, respectively, Λf for some linear map f : V → W , so that the functors S and Λ are fully faithful. 4.1.6. Koszul duality for quadratic algebras [Man87, Man88]. Consider a finitely generated quadratic algebra A = TV/(R), where V ∈ FVect and R ⊂ V ⊗2. Define its Koszul dual quadratic algebra as

A! = T (V ∗)/(R⊥), R⊥ = {ξ ∈ V ∗ ⊗ V ∗ | ξ(r)=0 ∀ r ∈ R}. (4.16)

33 The operation (4.16) is a contravariant functor (−)! : FQA → FQA. We have the following natural isomorphisms

(A!)! =∼ A, (A ◦ B)! =∼ A! • B!, (A • B)! ∼= A! ◦ B!, (4.17) ! ∗ ! ∼ ! ! ! ∼ ! ! (A )1 =(A1) , (A ⊗ B) = A ⊗ B , (A ⊗ B) = A ⊗ B . (4.18)

We see that this functor is an involutive anti-autoequivalence, which interchanges the Manin’s operations, and that the diagrams

(−)! (−)! FQA / FQAop FQAop / FQA (4.19)

(−)1 (−)1 (−)1 (−)1  (−)∗   (−)∗  FVect / FVectop FVectop / FVect are commutative. 4.1.7. The functors S∗ and Λ∗. Define contravariant functors S∗ : FVect → FQA and Λ∗ : FVect → FQA by the formulae S∗(V ) = S(V ∗), Λ∗(V ) = Λ(V ∗), V ∈ FVect. As functors FVect → FQAop they are the compositions of (−)∗ : FVect → FVectop with the functors Sop : FVectop → FQAop and Λop : FVectop → FQAop respectively, which are the opposite functors to ones defined in p. 4.1.5. The algebra S∗(V ) is the algebra of polynomial functions on V ∈ FVect. By choosing n ∼ Kn ∗ ∼ ∗ Kn a basis (xi)i=1 in V we obtain an isomorphism V = and hence S (V ) = S ( ) = K 1 n i n ∗ ∼ Kn ∗ [x ,...,x ], where (x )i=1 is the dual basis of V = ( ) . As a fully faithful functor the functor S∗ : FVect → FQAop embeds the finite-dimensional vector spaces into the category FQAop. This justifies the term ‘quantum linear spaces’ for the objects of FQAop, but this is only finite-dimensional quantum analogue. The functor Λ∗ also embeds FVect into FQAop, but in a different way. For V = Kn the algebra Λ∗(Kn) is isomorphic to the Grassmann algebra with generators ψ1,...,ψn and relations ψiψj + ψjψi = 0. Thus, the functors S∗ and Λ∗ are two different embeddings of the category FVect into FQAop. They identify the category FVect with some full subcategories of the categories op op op op FCommQSAeven ⊂ FQA and FCommQSAodd ⊂ FQA respectively. Hence the em- beddings S∗ and Λ∗ can be interpreted as follows. Any vector space V can be considered as a purely even or as a purely odd super-vector space. In the former case we apply S∗ to consider the vector space as quantum linear vector space S∗V = S(V ∗), in the latter case we apply Λ∗. Note that S∗(V ) = S(V ∗) = (ΛV )! and Λ∗(V ) = Λ(V ∗)=(SV )!. This means that the functor (−)! maps a vector space to its dual with change of its parity. The operations A ◦ B and A ⊗ B are analogues of tensor product and direct sum of vector spaces for the case of the quantum linear spaces A, B ∈ FQAop considered as purely even. The operations A • B and A ⊗ B are also analogues of the tensor product and direct sum, but in this case we need to consider A and B as purely odd.

34 4.1.8. Monoidal structures of S∗. Let us chose the ‘even’ interpretation: consider FVect op op ∗ as the subcategory of CommQSAeven ⊂ FQA by means of S . Since the monoidal func- tor (4.11) can be regarded as strict monoidal, the operation ⊕ in FVect coincides with the operation ⊗ in FQA (see the formula (4.10)). On the other hand, the functor (4.14) is colax monoidal, so its opposite Sop : (FVectop, ⊗) → (FQAop, ◦) is lax monoidal. By composing the latter one with the strong monoidal functor (−)∗ : (FVect, ⊗) → (FVectop, ⊗) in a proper order we obtain a lax monoidal functor. Thus we have strong monoidal functor S∗ : (FVect, ⊕) → (FQAop, ⊗), (4.20) lax monoidal functor S∗ : (FVect, ⊗) → (FQAop, ◦). (4.21) In particular, the tensor product of vector spaces does not coincide with the result of the Manin white product after the embedding S∗, because the monoidal functor (4.21) is not strong monoidal. Remark 4.1. One can define a functor S∗ : Vect → QAop as the composition of the functors (−)∗ : Vect → Vectop and Sop : Vectop → QAop. It has a structure of strong monoidal functor (Vect, ⊕) → (QAop, ⊗), but it does not get any monoidal structure (Vect, ⊗) → (QAop, ◦), since (−)∗ : (Vect, ⊗) → (Vectop, ⊗) is colax monoidal, while Sop : (Vectop, ⊗) → (QAop, ◦) is lax monoidal.

4.1.9. Coproduct of quadratic algebras. The product of quantum linear spaces is the coproduct in the category FQA or, more generally, in QA. For any two quadratic algebras A = TV/(R) ∈ QA and A′ = TV ′/(R′) ∈ QA their coproduct exists and has the form A∐A′ := T (V ⊕ V ′)/(R ⊕ R′). (4.22) Indeed, let B = TW/(S) ∈ QA and f : A → B, f ′ : A′ → B be two morphisms in QA. ′ ′ ′ Their first order components satisfy (f1 ⊗ f1)R ⊂ S, (f1 ⊗ f1)R ⊂ S. There exists a unique ′ linear map h1 : V ⊕ V → W such that the diagram V / V ⊕ V ′ o V ′ (4.23) ●● ✈ ● ✈✈ ●● h ✈ ● 1 ✈✈′ f1 ●● ✈✈ f ●#  z✈✈ 1 W ′ ′ ′ ′ commutes. Since (h1 ⊗h1)(R⊕R )=(f1 ⊗f1)(R)+(f1⊗f1)(R ) ⊂ S, there exists a morphism ′ h: A∐A → B with the first order component h1. Hence there is a unique morphism h making the diagram A / A∐A′ o A′ (4.24) ❍❍ ✈ ❍❍ ✈✈ ❍❍ h ✈✈ f ❍❍ ✈✈f ′ ❍❍  ✈✈ # B ✈z commutative. Note that the inclusion R ⊕ R′ ⊂ R ⊕ [V,V ′] ⊕ R′ implies the natural epimorphism A∐A′ ։ A⊗A′. (4.25)

35 Remark 4.2. The coproduct in the category the connected affinely generated graded alge- ′ ′ bras is different. Namely, the coproduct of A = Ak and A = Ak in this category is N N kL∈ 0 kL∈ 0 the graded algebra A A′ with the components c.a.g.` k+1 k k ′ ′ A A = (Al1 ⊗Al ⊗Al3 ⊗··· ). (4.26) k 2  c.a.g.a  Mp=1 Ml1=0 l2,...,lMp=1 l1+l2+...+lp=k 4.1.10. Functor T ∗. To extend the tensor product of vector spaces in an exact way one can consider another embedding such as the fully faithful functor T ∗ : FVect → FQAop defined as the composition of (−)∗ : FVect → FVectop and T op : FVectop → FQAop, that is T ∗(V ) = T (V ∗). It is a strong monoidal functor T ∗ : (FVect, ⊗) → (FQAop, ◦), but in this case the direct sum ⊕ is not strong-monoidally related with the Manin’s operation ⊗ with quadratic algebras. Indeed, T (V ) ⊗ T (W ) is a quotient algebra of T (V ⊕ W ) over the ideal generated by the subset {v⊗w−w⊗v | v ∈ V, w ∈ W }, so T : (FVect, ⊕) → (FQA, ⊗) is a colax (not strong) monoidal functor. The functor T preserves the finite coproducts, hence it has a structure of strong monoidal functor (Vect, ⊕) → (QA, ∐). If A, A′ ∈ FQA then A∐A′ ∈ FQA, so we also obtain the strong monoidal functor T : (FVect, ⊕) → (FQA, ∐). Thus we have:

lax monoidal functor T ∗ : (FVect, ⊕) → (FQAop, ⊗), (4.27) strong monoidal functor T ∗ : (FVect, ⊕) → (FQAop, ∐), (4.28) strong monoidal functor T ∗ : (FVect, ⊗) → (FQAop, ◦), (4.29) strong monoidal functor T ∗ : (Vect, ⊕) → (QAop, ∐), (4.30) colax monoidal functor T ∗ : (Vect, ⊗) → (QAop, ◦), (4.31)

where T ∗ : Vect → QAop is the composition of the functors (−)∗ : Vect → Vectop and T op : Vectop → QAop. ∗ 4.1.11. Dequantisation functor. Composition of the functors (−) and (−)1 gives the ∗ op ∗ ∗ functor (−)1 : FQA → FVect, A 7→ (A)1 = (A1) . By composing it with any of the functors S∗,Λ∗ and T ∗, which embed FVect into FQAop, we obtain the identical functor:

∗ ∗ ∗ ∗ ∗ ∗ S (V ) 1 = V, Λ (V ) 1 = V, T (V ) 1 = V. (4.32) ∗  op   The functor (−)1 : FQA → FVect is fully faithful. It has the following structures of strong monoidal functor:

(FQAop, ⊗) → (FVect, ⊕), (FQAop, ◦) → (FVect, ⊗), (4.33) (FQAop, ⊗) → (FVect, ⊕), (FQAop, •) → (FVect, ⊗), (4.34) (FQAop, ∐) → (FVect, ⊕). (4.35)

This functor extracts the classical part from a quantum linear space.

36 4.2 Manin matrices To define Manin matrices in terms of matrix idempotents we first describe the relationship between quadratic algebras and general idempotent operators. 4.2.1. Quadratic algebras for idempotents. Remind that an idempotent is an element A of a ring R satisfying A2 = A. Let W be a vector space and R, R′ ⊂ W be its subspaces such that W = R ⊕ R′. The composition W ։ W/R′ ∼= R ֒→ W is an idempotent in the algebra end(W ). Moreover, the idempotents A ∈ end(W ) are in one-to-one correspondence with the decompositions W = R ⊕ R′ into subspaces R, R′ ⊂ W such that R = Im A and ′ R = Im(1 − A), where 1 = idV ⊗V . The operator 1 − A ∈ end(W ) is the dual idempotent, it corresponds to the decomposition W = R′ ⊕ R. Let A ∈ QA have the form A = TV/(R) for a vector space V ∈ Vect and a subspace R ⊂ V ⊗2. As any subspace the subspace R can be given as an image of an idempotent A ∈ end(V ⊗2), since R always has a complement R′ ⊂ V ⊗2. Thus any connected quadratic algebra is isomorphic to TV/(Im A) for some idempotent A ∈ end(V ⊗2). ′ ′ Denote ΞA(K)= TV/(R ) where R = Im(1 − A). Any A ∈ QA is isomorphic to ΞA(K) for some idempotent operator A. More generally, any semi-connected quadratic algebra is isomorphic to ΞA(R) := R ⊗ ΞA(K) for some A and R ∈ Alg. If V ∈ FVect then (V ⊗2)∗ = (V ∗)⊗2, so any (idempotent) operator A: V ⊗2 → V ⊗2 gives the (idempotent) transpose operator A∗ : (V ∗)⊗2 → (V ∗)⊗2. Any quadratic algebra K ∗ ∗ K A ∈ FQA or A ∈ FQAsc isomorphic to the algebra XA( ) := TV /(Im A )=Ξ1−A∗ ( ) or ⊗2 XA(R) := R ⊗ XA(K)=Ξ1−A∗ (R) respectively for some idempotent A ∈ end(V ). The Koszul duality in these notations has the form:

! ! XA(K) =ΞA(K), ΞA(K) = XA(K). (4.36)

The semi-connected quadratic algebras XA(R) and ΞA(R) can be considered as values of ! K functors XA, ΞA : Alg → FQAsc coinciding with −⊗A and −⊗A , where A = XA( ). 1 − σ For the idempotent A = V,V (anti-symmetrizer of V ⊗ V ) we obtain V 2 K ∗ K XAV ( )= S (V ), ΞAV ( )=ΛV, (4.37)

The functors T ∗ and T give the quadratic algebras for the trivial idempotents 0, 1 ∈ end(V ⊗2):

∗ X0(K)= T (V ), Ξ1(K)= TV. (4.38)

n 4.2.2. The case of matrix idempotents. A basis (vi)i=1 of a vector space V ∈ FVect gives the isomorphism V =∼ Kn, v 7→ (x1,...,xn), where xi are components of v ∈ V defined n i ∼ Kn by the formula v = i=1 x vi. It induces the algebra isomorphism end(V ) = end( ), which maps operatorP on V to its n × n matrix in the basis (vi). The space V ⊗ V has the n 2 2 basis (vi ⊗ vj)i,j=1, so the operators on V ⊗ V correspond to the n × n matrices, whose entries have two pairs of indices. Namely, the operator A ∈ end(V ⊗2) corresponds to the ij n ij matrix A = (Akl) defined by the formula A(vk ⊗ vl) = i,j=1 Aklvi ⊗ vj. The transpose P 37 ∗ ∗ ∗ ∗ i j n ij k l i n operator A ∈ end(V ⊗ V ) has the form A (v ⊗ v )= k,l=1 Aklv ⊗ v , where (v )i=1 is the dual basis. P n Since any V ∈ FVect has a finite basis (vi)i=1, any algebra A = TV/(R) ∈ FQA is n n isomorphic to ΞA(K) for some matrix idempotent A ∈ end(K ⊗ K ). The latter is a matrix ij n ij kl ij K A =(Akl) such that k,l=1 AklAab = Aab. The algebra A =ΞA( ) is generated by v1,...,vn n ij with the quadratic commutationP relations i,j=1 Aklvjvj = vkvl, where we identified the n basis (vi) with the standard basis of K . P The same algebra is isomorphic to XA(K) = TW/(R) for another matrix idempotent A such that W = (Kn)∗ and R = Im A∗. Let (vi) be the basis of (Kn)∗ dual to the standard Kn K 1 n n ij k l basis of , then XA( ) is generated by v ,...,v with the relations k,l=1 Aklv v = 0. n n n n Consider, for instance, the anti-symmetrizer An = AKn : K ⊗ K →PK ⊗ K . It has the ij 1 i j i j K K entries (An)kl = 2 (δkδl − δl δk). The corresponding quadratic algebras XA( ) and ΞA( ) are i i j j i generated by v and vi with the relations v v = v v and vivj = −vjvi. The former one is the polynomial algebra K[x1,...,xn] with commutative variables xi = vi, while the latter one is the Grassmann algebra with the anti-commutative variables vi. 4.2.3. Manin matrix for a pair of idempotents. Any n × m matrix M with entries i Km Kn Mij = Mj ∈ R can be considered as an operator → R ⊗ . In terms of standard bases n m Kn Km n i (vi)i=1 and (wj)j=1 of the spaces and it acts as Mwj = i=1 Mj ⊗ vi. Denote by M (a) the operator M acting on the a-th matrix tensor factor Km Pof R ⊗ Km ⊗···⊗ Km. In the case of two such factors we have n (1) i M (r ⊗ wj ⊗ wl)= Mj r ⊗ vi ⊗ wl, r ∈ R, (4.39) Xi=1 n (2) k M (r ⊗ wj ⊗ wl)= Ml r ⊗ wj ⊗ vk, r ∈ R. (4.40) Xk=1 Definition 4.3. [S] Let A ∈ end(Kn ⊗Kn) and B ∈ end(Km ⊗Km) be matrix idempotents. The matrix M satisfying

AM (1)M (2)(1 − B) = 0 (4.41)

is called Manin matrix for the pair of idempotents (A, B) or simply (A, B)-Manin matrix. In the case A = B we will say B-Manin matrix instead of (B, B)-Manin matrix. Kn ∗ Km ∗ i m i j The matrix M gives the operator ( ) → R ⊗ ( ) acting as v 7→ j=1 Mj ⊗ w , i j where (v ) and (w ) are bases dual to (vi) and (wj). Any linear map XA(K)1 P→ XB(R)1 has such form for some matrix M. In fact, this map gives a homomorphism of graded algebras XA(K) → XB(R) iff M is an (A, B)-Manin matrix (see [S]), we denote this homomorphism of graded algebras by fM : XA(K) → XB(R). m n In the same way any linear map ΞB(K)1 → ΞA(R)1 has the form K → R ⊗ K , n i K i wj 7→ i=1 Mj vi. It gives a graded homomorphism ΞB( ) → ΞA(R) iff M = (Mj ) is an M (A, B)-ManinP matrix, we denote this homomorphism by f : ΞB(K) → ΞA(R).

38 Thus we have bijections between the sets Hom XA(K), XB(R) , Hom ΞB(K), ΞA(R) and the set of (A, B)-Manin matrices.   4.2.4. Usual Manin matrices. The Manin matrices introduced in the papers [CF, CFR] are n × m Manin matrices for the pair of idempotents (An, Am). They correspond to the graded homomorphisms of polynomial algebras K[x1,...,xn] → R[y1,...,ym] or, equiva- lently, to the graded homomorphisms S(V ∗) → R ⊗ S(W ∗), where V and W are n- and m-dimensional vector spaces.

4.3 Semi-linear spaces and their quantum analogue To interpret representations of algebraic monoids and groups on (finite-dimensional) linear spaces in terms of Subsection 3.2 one needs to include both algebraic sets and linear spaces to a bigger monoidal category. More generally, one needs to do the same for the affine schemes. In this way we will get categories of spaces which are partially linear in some sense. We call them semi-linear spaces. 4.3.1. Semi-linear algebraic sets. Describe the first case of semi-linear spaces. For an algebraic set X ∈ AlgSet and a linear space V ∈ FVect consider their set-theoretic product X × V as an algebraic set. Define morphisms between two such products X × V and Y × W as morphisms F : X × V → Y × W in AlgSet that have the form F (x, v)=(ϕ(x), f(x, v)), x ∈ X, v ∈ V, ϕ: X → Y, f : X × V → W, f(x, αv + v′)= αf(x, v)+ f(x, v′), x ∈ X, v,v′ ∈ V. (4.42)

In other words, the morphism F is given by arbitrary morphisms ϕ ∈ HomAlgSet(X,Y ) and f ∈ HomAlgSet(X ×V, W ) such that f(x, −): V → W is linear for any x ∈ X. In this way we obtain a category SLAlgSet with morphisms that partially satisfy the linearity condition, we call them semi-linear maps. We call the objects X ×V ∈ SLAlgSet semi-linear algebraic sets. 4.3.2. Monoidal product of semi-linear algebraic sets. Let X,X′ ∈ AlgSet and V,V ′ ∈ FVect. Define monoidal product of X × V,X′ × V ′ ∈ SLAlgSet as the object (X × X′) × (V ⊗ V ′). For morphisms F : X × V → Y × W and F ′ : X′ × V ′ → Y ′ × W ′ in SLAlgSet, given by maps ϕ: X → Y , f : X × V → W and ϕ′ : X′ → Y ′, f ′ : X′ × V ′ → W ′, their monoidal product is F ′′ = F ⊗ F ′ : (X × X′) × (V ⊗ V ′) → (Y × Y ′) × (W ⊗ W ′) given by ϕ′′ = ϕ × ϕ′ : X × X → Y × Y and f ′′ : (X × X′) × (V ⊗ V ′) → W ⊗ W ′, f ′′(x, x′, v ⊗ v′)= f(x, v) ⊗ f ′(x′, v′). In this way we obtain a symmetric monoidal category (SLAlgSet, ⊗) with the unit object 0 × K. The categories (AlgSet, ×) and (FVect, ⊗) can be regarded as monoidal subcategories of (SLAlgSet, ⊗) embedded via the faithful symmetric strict monoidal functors X 7→ X ×K and V 7→ {0}×V . The latter functor is full, the former is not, i.e. FVect is a full subcategory of SLAlgSet, while the subcategory AlgSet ⊂ SLAlgSet is not full. 4.3.3. Trivial vector bundles. The category SLAlgSet is equivalent to the category of trivial vector bundles over algebraic sets. The semi-linear algebraic set X × V corresponds

39 to the trivial bundle over X with the fibre V . A semi-linear map F : X × V → Y × W given by the formula (4.42) is the vector bundle morphism (ϕ, F ) in the sense that the diagram

F X × V / Y × W (4.43)

 ϕ  X / Y

is commutative (the vertical arrows mean the projections) and the morphism F induces a linear map on each fibre (see [Lang, III, § 1]). Category embeddings AlgSet ֒→ SLAlgSet and FVect ֒→ SLAlgSet identify an algebraic set X with the 1-dimensional trivial bundle over it, and a vector space V – with the trivial bundle over {0} = Spec K with the fibre V . The main difference between the notions of semi-linear algebraic set and trivial vector bundle is that we consider the former ones not over the structure of algebraic set, but together with this structure. In particular, the monoidal product of two semi-linear algebraic sets X × V and X × W is not the same as the tensor product of the corresponding bundles over X. One can consider more general vector bundles in the same manner, but the category SLAffSch is enough for out purposes, this is a minimal monoidal category that includes (AlgSet, ×) and (Vect, ⊗). 4.3.4. Algebras of functions on semi-linear algebraic sets. Let X ∈ AlgSet and V ∈ FVect. The algebra of regular functions on X × V as on algebraic set is A(X) ⊗ SV ∗. It inherits the grading from SV ∗ as a tensor product of an algebra with a graded algebra (see p. 2.1.5). Let us denote S∗(X × V ) = A(X) ⊗ SV ∗ ∈ GrAlg. For a morphism of the form (4.42) we obtain the morphism of graded algebras with the reverse direction: S∗(F ): A(Y ) ⊗ SW ∗ → A(X) ⊗ SV ∗, it is defined as

S∗(F ): s 7→ ϕ∗(s)= s · ϕ ∈ A(X), ϕ∗(s)(x)= s ϕ(x) , (4.44) S∗(F ): ξ 7→ f ∗(ξ)= ξ · f ∈ A(X) ⊗ V ∗, f ∗(ξ)(x, v)=ξ f(x, v) , (4.45)  where s ∈ A(Y ), ξ ∈ W ∗, x ∈ X, v ∈ V . The formula (4.44) gives the usual embedding op j i ∗ ∗ AlgSet ֒→ CommAlg . Let (vj), (v ) and (wi), (w ) be dual bases of V , V and W , W i i respectively. Then f(x, vj) = i fj (x)wi for some fj ∈ A(X), so the map (4.45) has the ∗ i i j ∗ form f (w )= j fj ⊗ v ∈ A(PX) ⊗ V . ∗ op We obtain aP lax monoidal functor S : (SLAlgSet, ⊗) → (FQAsc , ◦). In particular, any algebraic set X as the object X ×K corresponds to the graded algebra S∗(X)= A(X)⊗K[u] and we have the natural isomorphism S∗(X × V ) =∼ S∗(X) ◦ S∗(V ). 4.3.5. Semi-linear affine schemes. Let us generalise the semi-linear spaces to the case of affine schemes. Among the commutative semi-connected quadratic algebras consider the algebras of the form R ⊗ SV ∗ where R ∈ CommAlg, V ∈ FVect. Such algebra can be interpreted as the ‘function algebra’ on the affine scheme X × V , where X = Spec R. This is an affine scheme which corresponds to the algebra R ⊗ SV ∗ and has an additional structure given by the grading of this algebra. We call such schemes semi-linear affine schemes. By

40 op definition they are objects of the category FQAsc corresponding to the graded algebras ∗ op R ⊗ SV , where R ∈ CommAlg, V ∈ FVect. The embedding SLAffSch ֒→ FQAsc ∗ op ∗ ∗ extends the functor S : SLAlgSet → FQAsc , S (X × V )= R ⊗ SV , since SLAlgSet is a subcategory of SLAffSch, so we also denote it by S∗. Let S ∈ CommAlg, Y = Spec S and W ∈ FVect. A morphism X × V → Y × W in SLAffSch is given by a graded homomorphism S ⊗ SW ∗ → R ⊗ SV ∗ generated by an algebra homomorphism α: S → R and a linear map t: W ∗ → R ⊗ V ∗. In the order 1 it has the form

(α, t): S ⊗ W ∗ → R ⊗ V ∗, (α, t)(s ⊗ ξ)= α(s)t(ξ). (4.46)

For a morphism in SLAlgSet of the form (4.42) we have α = ϕ∗ and t = f ∗, (α, t)= F ∗. Remark 4.4. The semi-linear affine scheme X × V can be identified with a quasi-coherent sheaf corresponding to the free R-module R⊗V (or rather to its dual R⊗V ∗). The category SLAffSch is equivalent to the category of such free sheaves (R-modules) with appropriate morphisms. This generalises the equivalence described in p. 4.3.3. Let R, R′ ∈ CommAlg, X = Spec R, X′ = Spec R′, V,V ′ ∈ FVect. Define the monoidal product of the semi-linear affine schemes X × V and X′ × V ′ as the semi-linear affine scheme (X × X′) × (V ⊗ V ′) with the function algebra R ⊗ R′ ⊗ S∗(V ⊗ V ′). Note op that the embedding SLAffSch ֒→ FQAsc is a fully faithful symmetric lax monoidal functor ∗ op S : (SLAffSch, ⊗) → (FQAsc , ◦). 4.3.6. Semi-linear algebraic monoids and semi-linear affine monoid schemes. A monoid in (SLAlgSet, ⊗) is a semi-linear algebraic set X × V ∈ SLAlgSet with the semi- linear maps µX×V : (X × X) × (V ⊗ V ) → X × V , ηX×V : K → X × V which have the form

µX×V (x,y,u ⊗ v)= xy, f(x,y,u ⊗ v) , ηX×V (1) = e, 1V , (4.47)   where xy = µX (x, y), e = ηX (0) are defined by a monoid (X,µX , ηX ) ∈ Mon(AlgSet, ×) and f : X ×X ×(V ⊗V ) → V is a morphism in AlgSet linear in the last argument such that f(e, x, 1V ⊗v)= v = f(x, e, v ⊗1V ), f x,yz,u⊗f(y,z,v ⊗w) = f xy, z, f(x,y,u⊗v)⊗w , x, y, z ∈ X, u,v,w ∈ V . Denote by Vxthe copy of V considered as a fibre over x ∈ X, then for any x, y ∈ X we obtain the linear map fx,y : Vx ⊗ Vy → Vxy, fx,y(u ⊗ v)= f(x,y,u ⊗ v), and the last condition is equivalent to the commutativity of the diagram

id ⊗fy,z Vx ⊗ Vy ⊗ Vz / Vx ⊗ Vyz (4.48)

fx,y⊗id fx,yz

 fxy,z  Vxy ⊗ Vz / Vxyz while 1V belongs to Ve. If X = {e} is a one-point set, then to equip V = X × V with a monoid structure is the same as to choose the algebra (V, fe,e, 1) ∈ Mon(Vect, ⊗). If

41 V = K, then a monoid structure on X × V = X × K is given by an algebraic monoid (X,µX , e) ∈ Mon(AlgSet, ×) with collection of coefficients px,y ∈ K, while q ∈ K\{0} −1 satisfying px,yzpy,z = pxy,zpx,y and pe,x = px,e = q (the linear map fx,y : K ⊗ K → K is a multiplication by the coefficient px,y, while 1V = q). The case of a usual algebraic monoid corresponds to the values px,y = q = 1. In this way we see that the categorical -embeddings FVect ֒→ SLAlgSet and AlgSet ֒→ SLAlgSet induce the fully faithful func tor Mon(FVect, ⊗) → Mon(SLAlgSet, ⊗) and the faithful functor Mon(AlgSet, ×) → Mon(SLAlgSet, ⊗), which are symmetric strong monoidal due to p. 2.4.5. ∗ ∗ K Let R = A(X). The homomorphisms α = µX : R → R ⊗ R and β = ηX : R → gives the structure of coalgebra on R. The linear maps t = f ∗ : V ∗ → R ⊗ R ⊗ V ∗ ⊗ V ∗ and ∗ εV ∗ =1V : V → K make the diagrams

(α,t) R ⊗ V ∗ / R ⊗ R ⊗ V ∗ ⊗ V ∗ (4.49)

(α,t) σ(23)   R ⊗ R ⊗ V ∗ ⊗ V ∗ R ⊗ V ∗ ⊗ R ⊗ V ∗

σ(23) id ⊗ id ⊗(α,t)   R ⊗ V ∗ ⊗ R ⊗ V ∗ R ⊗ V ∗ ⊗ R ⊗ R ⊗ V ∗ ⊗ V ∗

(α,t)⊗id ⊗ id σ(34)σ(23)  σ(34)σ(45)  R ⊗ R ⊗ V ∗ ⊗ V ∗ ⊗ R ⊗ V ∗ / R ⊗ R ⊗ R ⊗ V ∗ ⊗ V ∗ ⊗ V ∗ R ⊗ V ∗ ❢❢❢ ❳❳❳❳ ❢❢❢❢❢❢❢ ❳❳❳❳❳❳❳ ❢❢❢❢❢❢❢ ❳❳❳❳❳❳❳ ❢❢❢❢❢❢❢ (α,t) ❳❳❳❳❳❳❳ ❢❢❢❢❢❢❢ ❳❳❳❳❳❳❳ ❢❢❢❢ β⊗id ⊗ε ∗ ⊗id  id ⊗β⊗id ⊗ε ∗ ❳❳❳ K ⊗ R ⊗ K ⊗ V ∗ o V R ⊗ R ⊗ V ∗ ⊗ V ∗ V / R ⊗ K ⊗ V ∗ ⊗ K

to be commutative, where σ(i,i+1) = id⊗(i−1) ⊗σ ⊗ id ⊗···⊗ id. Since (R,α,β) is a comonoid in (CommAlg, ⊗), it is enough to check the commutativity of the diagrams (4.49) by substituting the elements of the form 1 ⊗ ξ ∈ R ⊗ V ∗. Now consider monoids in (SLAffSch, ⊗), which we call semi-linear affine monoid schemes. Let R ∈ CommAlg, V ∈ FVect and X = Spec R. A structure of monoid on X × V is defined by the morphisms µX×V : (X × V ) ⊗ (X × V ) → X × V and ηX×V : K → X × V , which are given by graded homomorphisms R ⊗ S(V ∗) → R ⊗ R ⊗ S(V ∗) ⊗ S(V ∗) and R ⊗ S(V ∗) → K[u]. The zero and first order components of these homomorphisms are ∗ ∗ ∗ ∗ α: R → R ⊗ R, (α, t): R ⊗ V → R ⊗ R ⊗ V ⊗ V and β : R → K, (β,εV ∗ ): R ⊗ V → K, ∗ ∗ ∗ ∗ where t: V → R ⊗ R ⊗ V ⊗ V and εV ∗ : V → K are linear maps. They give a structure of a monoid on X × V iff (R,α,β) ∈ Comon(CommAlg, ⊗) and the diagrams (4.49) com- mute. The embedding AffSch ֒→ SLAffSch induces the faithful symmetric strong monoidal functor Mon(AffSch, ×) → Mon(SLAffSch, ⊗). 4.3.7. Quantum semi-linear spaces. The ‘algebra of functions’ on a semi-linear space X × V is the tensor product of a quadratic algebra S∗V with a commutative algebra that

42 is the ‘algebra of functions’ on X. By following the Manin’s idea described in the begin- ning of Section 3 we call quantum semi-linear space any semi-connected quadratic algebra op R ⊗A (where R ∈ Alg, A ∈ QA) considered as an object of the dual category QAsc . The op finite-dimensional case corresponds to the objects of FQAsc . The quantum linear spaces correspond to the connected case R = K. In Section 5 we will construct Quantum Represen- op tation Theory as the theory of representations of monoids in C =(QAsc , ◦) on the objects of a monoidal subcategory P = FQAop. op The composition of the embeddings AffSch ֒→ SLAffSch ֒→ FQAsc is a faithful op symmetric strong monoidal functor (AffSch, ×) → (FQAsc , ◦), its dual is the functor K (CommAlg, ⊗) → (FQAsc, ◦), which maps R to R ⊗ [u]. Thus, we consider an affine non-commutative space R ∈ Algop as the quantum semi-linear space R ⊗ K[u] ∈ FQAop.

Any morphism f : B→A in QAsc between arbitrary semi-connected quadratic algebras B = S ⊗ TW/(S) and A = R ⊗ TV/(R) is uniquely defined by its zero and first order components α = f0: S → R and f1 : S ⊗ W → R ⊗ V . We have f1 = (α, t) in the sense that f1(s ⊗ w) = α(s)t(w) = t(w)α(s), where t: W → R ⊗ V is defined by the formula t(w)= f1(1 ⊗ w). Since A2 is the quotient of (R ⊗ V ) ⊗R (R ⊗ V )= R ⊗ V ⊗ V over R ⊗ R, the linear map f2 : B2 →A2 is induced by

f1⊗Sf1 S ⊗ W ⊗ W =(S ⊗ W ) ⊗S (S ⊗ W ) −−−−→ (R ⊗ V ) ⊗R (R ⊗ V )= R ⊗ V ⊗ V, ′ ′ ′ ′ s ⊗ w ⊗ w =(s ⊗ w) ⊗S (1 ⊗ w ) 7→ α(s)t(w) ⊗R t(w )= α(s) t(w) ⊗R t(w ) , ′ ′ ′ 1 ⊗ w ⊗ w 7→ t(w) ⊗R t(w ) =(t ⊗˙ t)(w ⊗ w ),   where t ⊗˙ t: W ⊗ W → R ⊗ V ⊗ V is the composition

t⊗t idR ⊗σV,R⊗idV µR⊗id ⊗ id W ⊗ W −−→ R ⊗ V ⊗ R ⊗ V −−−−−−−−−→ R ⊗ R ⊗ V ⊗ V −−−−−−−−→V V R ⊗ V ⊗ V. An algebra homomorphism α: S → R and a linear map t: W → R ⊗ V define a homomor- phism f : B→A iff α(s)t(w)= t(w)α(s) ∀ s ∈ S, w ∈ W (4.50) and (t ⊗˙ t)S ⊂ R ⊗ R. In particular, (B, ∆,ε) is a comonoid in (QAsc, ◦) iff the zero and first order components of the graded homomorphisms ∆: B →B◦B and ε: B → K[u] satisfy the following conditions: α = ∆0 : S → S ⊗ S and β = ε0 : S → K are algebra homomorphisms, ∆1(s ⊗ w) = α(s)t(w) and ε1(s ⊗ w) = ε0(s)εW (w) for some linear maps t: W → S ⊗ S ⊗ W ⊗ W , εW : W → K subjected to the commutativity condition (4.50), (23) ⊗2 ⊗2 the inclusion conditions (t ⊗˙ t)S ⊂ S ⊗ S ⊗ σ (S ⊗ W + W ⊗ S), (εW ⊗ εW )S = 0 and making the diagrams (4.49) to be commutative (where V ∗ = W and R = S). O A coaction of the comonoid = (A, ∆,ε) ∈ Comon(QAsc, ◦) on an object B ∈ QAsc is a graded homomorphism δ : B→A◦B. Its zero component δ0 : B0 →A0 ⊗ B0 should be coaction of (A0, ∆0,ε0) ∈ Comon(Alg, ⊗) on B0 ∈ Alg and the first order component is δ1 =(δ0,p) for some linear map p: W →A0 ⊗B0 ⊗V ⊗W . To obtain the condition for p one needs to take first order components of the morphisms in the commutative diagrams (2.22), this leads to diagrams similar to (4.49).

43 4.4 Internal hom for quantum linear spaces The next ingredient, which we need to construct quantum representations, is a quantum analogue of the internal hom of vector spaces. This is an internal hom in (FQAop, ◦), i.e. an internal cohom in (FQA, ◦). We need also its generalised version in the sense of p. 3.1.3. 4.4.1. Internal cohom and universal Manin matrix. It was shown in [Man87, Man88] that the category FQA is coclosed with respect to the Manin product, namely, proved the existence of a natural isomorphism Hom(A • B!, C) ∼= Hom(A, B ◦ C). Hence the internal cohom for A, B ∈ FQA can be defined as3 cohom(B, A)= B! •A. (4.51)

⊗2 ⊗2 Let us calculate it for A = XA(K) and B = XB(K), where A ∈ end(V ) and B ∈ end(W ) n m ! are idempotents and V = K , W = K . Since B = ΞB(K) = TW/ Im(1 − B) and A = TV ∗/(Im A∗) we have  B! •A = T (W ⊗ V ∗)/ σ(23)(Im(1 − B) ⊗ Im A∗) , (4.52)  where σ(23) = id ⊗σ ⊗ id (the indices 2, 3 mean that σ acts on the second and third tensor i ∗ factors). Let (vi) and (wj) be bases of V and W respectively. Let (v ) be basis of V dual ∗ i i to (vi). Then the basis of W ⊗ V = Hom(V, W ) consists of Mj = wj ⊗ v . The subspace σ(23)(Im(1 − B) ⊗ Im A∗) ⊂ W ⊗ V ∗ ⊗ W ⊗ V ∗ is spanned by the vectors

(23) kl pq s t pq s t kl σ (1 − B)ij Ast wk ⊗ wl ⊗ v ⊗ v = Ast (Mk ⊗Ml)(1 − B)ij , k,l,s,tX k,l,s,tX

kl k l kl i where (1 − B)ij = δi δj − Bij . Therefore, the algebra cohom(B, A) is generated by Mj with the quadratic relation AM(1)M(2)(1 − B)=0, (4.53)

i where M is the matrix with entries Mj. The relation (4.53) means exactly that M is an (A, B)-Manin matrix over R = cohom XB(K), XA(K) . We call it universal (A, B)-Manin matrix. Note also that 

! ! cohom XB(K), XA(K) = XB(K) • XA(K)=ΞA(K) • ΞB(K)=

 = cohom ΞA(K), ΞB(K) . (4.54)  4.4.2. Connection with the Manin matrices. In [S] we denote by UA,B the algebra (4.54) considered as an object of Alg. This algebra was interpreted there via left adjoints to the functors XB : Alg → FQAsc and ΞA : Alg → FQAsc. In particular, we derived the natural bijections

HomGrAlg XA(K), XB(R) ∼= HomAlg(UA,B, R) ∼= HomGrAlg ΞB(K), ΞA(R) , (4.55)   3Manin denoted the cohom-functor cohom: FQAop × FQA → FQA as hom.

44 which can be described as follows. The left and right sets consists of the graded homomor- M i i phisms fM and f for (A, B)-Manin matrices M over R (see p. 4.2.3). A formula Mj 7→ Mj i gives a homomorphism UA,B → R (in Alg) iff M =(Mj ) is an (A, B)-Manin matrix. Consider the case when R is also a quadratic algebra: R = XC(K) for some idempotent K K K K C. Note that XB XC ( ) k = XC( )⊗XB ( )k, so the algebra XB XC ( ) does not coincide with XC (K)⊗XB(K) as graded algebra in the sense of the Manin’s operation ⊗ (see p. 4.1.1). Consider the subset of the set (4.55) corresponding to the (A, B)-Manin matrices M with i K K K entries of the first order: Mj ∈ XC ( )1. As a subset of HomGrAlg XA( ), XB XC ( ) it   consists of the graded homomorphisms XA(K) → XB XC (K) which factors through the morphism XC (K) ◦ XB(K) → XB XC(K) . This subset can be identified with the set  HomGrAlg XA(K), XC (K) ◦ XB(K) . As a subset of HomAlg UA,B, XB XC(K) it consists   of homomorphisms UA,B → XC(K) which preserve the grading, where the grading of the algebra UA,B arises, if one considers it as the cohom-object cohom XB(K), XA(K) . This leads to the adjunction from the beginning of p. 4.4.1, in the notation s of this paragraph it takes the form

Hom XA(K), XC (K) ◦ XB(K) ∼= Hom cohom XB(K), XA(K) , XC (K) , (4.56)     this isomorphism is induced by the isomorphism (4.55). 4.4.3. Internal hom for finite-dimensional vector spaces. Let us regard how the internal hom of quantum linear spaces agrees with the internal hom of vector spaces. Re- mind that the internal hom-object hom(W, V ) in Vect is the set Hom(W, V ) equipped with the usual structure of vector space (see p. 3.1.7). In the case W ∈ FVect we have natural linear isomorphism hom(W, V ) = W ∗ ⊗ V . By applying the strong monoidal ∗ op functor (−)1 : (FQA , ◦) → (FVect, ⊗) to the internal hom (4.51) we obtain exactly ∗ ∗ ∗ ∗ cohom(B, A) 1 = W ⊗ V = hom(W, V ), where W = B1, V = A1. In the nota- tions of Subsection 3.3 it means that the morphisms ΦB,A for the strong monoidal functor ∗ op (−)1 : (FQA , ◦) → (FVect, ⊗) are isomorphisms. On the other hand the quadratic algebra S∗ hom(W, V ) does not coincide with the al- ∗ ∗ ∗ ! ∗ ∗ gebra cohom(S W, S V )=(S W ) • S V = Λ(W ) • S (V ), which means that ΦW,V for the lax monoidal functor S∗ : (FVect, ⊗) → (FQAop, ◦) are not isomorphisms.

4.4.4. Coevaluation and cocomposition morphisms. Let (vi), (wj) be the bases of V, W ∈ FVect and let (vi), (wj) be the dual bases of V ∗, W ∗. The elements vi, wj and i i ∗ ∗ Mj = wj ⊗v are generators of the algebras A = TV /(R), B = T W /(S) and cohom(B, A). In these terms the coevaluation (3.11) (for X = A, Y = B) has the form (see [Man88])

i i j coev: A→ cohom B, A ◦ B, v 7→ Mj ⊗ w . (4.57)  Xj

∗ j j i i Let C ∈ FQA and (zl) be a basis of Z = C1 . Then Nl = zl ⊗w and Kl = zl ⊗v are the gen- erators of the algebras cohom(C, B) and cohom(C, A). The cocomposition morphism (3.13)

45 (for X = A, Y = B, Z = C) reads (see [S])

m i i j d = dA,B,C : cohom(C, A) → cohom(B, A) ◦ cohom(C, B), Kl 7→ Mj ⊗Nl . (4.58) Xj=1

m ij m ij Let B =(Bkl)i,j,k,l=1 be a matrix idempotent such that the elements Bklwi ⊗ wj span i,jP=1 the subspace S ⊂ W ⊗W , then B = XB(K) and cohom(B, B) is the quadratic algebra gener- i j (1) (2) i m ated by Mj = wj ⊗w with the relations BM M (1−B) = 0, where M =(Mj)i,j=1 is the universal B-Manin matrix. The comonoid coend(B) ∈ Comon(FQA, ◦) is the quadratic algebra cohom(B, B) with the structure morphisms dB : coend(B) → coend(B)◦coend(B) and vB : coend(B) → K[u] defined on generators as

m i i k i i dB(Mj)= Mk ⊗Mj , vB(Mj)= δju. (4.59) Xk=1 ! K K Since B =ΞB( )= TW/ Im(1−B) = X1−B⊤ ( ) is generated by w1,...,wm ∈ W , the ! j ∗ algebra coend(B ) is generated by the elements Mij = w ⊗ wi ∈ W ⊗ W . The latter ones ⊤ are entries of the universal (1 − B )-Manin matrixf M =(Mij) transposed to the universal B-Manin matrix M. Due to the formulae (4.59) this implies f f cop dB! = dB , vB! = vB. (4.60)

4.4.5. Infinite-dimensional case. Let A, A ∈ QA and B ∈ FQA. Then A = TV/(R), A = T V/(R) and B = TW/(S) for some V, V ∈ Vect, W ∈ FVect, R ⊂ V ⊗2, R ⊂ V ⊗2, e S ⊂ W ⊗2. Let us prove that there exists cohom(B, A) in the sense of p. 3.1.3, where the e e e e e e role of parameter is played by B ∈ FQA. Proposition 4.5. For any B ∈ FQA the functor − ◦ B : QA → QA has a left adjoint cohom(B, −): QA → QA. We have cohom(B, A) = TV ′/(R′) where V ′ = hom(W, V ) and

R′ = {U ∈ hom(W, V )⊗2 | U(W ⊗2) ⊂ R, U(S)=0}. (4.61)

m i m ∗ ∗ ∗ Proof. Let (wi)i=1 and (w )i=1 be dual bases of W and (W ) = W respectively. These bases give identifications W =(Km)∗ and W ∗ = Km. Let B ∈ end(Km ⊗ Km) be a matrix ∗ ∗ ∗ idempotent such that S = Im B , then B = XB(K)= TW/(Im B ). By identifying W ⊗ W with end(W ) we obtain

i j kl kl i j wi ⊗ w ⊗ wj ⊗ w = Bij + (1 − B)ij wk ⊗ w ⊗ wl ⊗ w , (4.62) Xi,j i,j,k,lX  ij K ij i j ij where the coefficients Bkl ∈ and (1 − B)kl = δkδl − Bkl are entries of the operators B and 1−B respectively. Consider the linear map ηV : V → hom(W, V )⊗W defined in the proof of

46 i Prop. 3.2, that is ηV (v)= i(wi ⊗v)⊗w . It generates a morphism η : A→ cohom(B, A)◦B with the first order componentP η1 = ηV . Indeed, for any element r = s xs ⊗ ys ∈ R we have P (23) i j σ (ηV ⊗ ηV )r = (wi ⊗ xs) ⊗ (wj ⊗ ys) ⊗ w ⊗ w = Xs,i,j kl kl i j = Bij + (1 − B)ij wk ⊗ xs ⊗ wl ⊗ ys ⊗ w ⊗ w . (4.63) s,i,j,k,lX  kl i j kl ′ Since Bij w ⊗ w ∈ S and (1 − B)ij wk ⊗ xs ⊗ wl ⊗ ys ∈ R , the element (4.63) belongs to hom(W, V )⊗2 ⊗ S + R′ ⊗ W ⊗2. Now we need to prove that for any A ∈ QA and f ∈ Hom(A, A◦B) there exists a unique h ∈ Hom cohom(B, A), A making the diagram e e  e η A ❙❙ / cohom(B, A) ◦ B (4.64) ❙❙❙ ❙❙❙ ❙❙❙ h◦id ❙❙❙ B f ❙❙❙  ❙❙) A ◦ B commutative. For purposes of the next paragraphe we do it for a slightly more general case A ∈ QAsc. Let R = A0, then A = R ⊗ T V/(R) for some vector space V ∈ Vect and subspace R ⊂ V ⊗ V . We have A = R ⊗ V and A = Te A /(R ⊗ R). Note that e e e e 1 e e e R 1 e any graded homomorphisme e e g : A → A ise uniquelye e determinede by its firste ordere component g : V → R ⊗ V and, conversely, this component gives the whole graded homomorphism 1 e g : A→ A if it is linear and satisfies (g1 ⊗˙ g1)R ⊂ R ⊗ R, where g1 ⊗˙ g1 : V ⊗ V → R ⊗ V ⊗ V e e (23) is defined as in p. 4.3.7, that is g ⊗˙ g = (µe ⊗ id e ⊗ id e ) · σ · (g ⊗ g ). By using the e 1 1 R e V e V 1 1 e e natural isomorphism (R⊗A)◦B = R⊗(A◦B) we obtain A◦B = R⊗T (V ⊗W )/(Rw), where R = σ(23)(V ⊗2 ⊗ S + R ⊗ W ⊗2). In particular, (A ◦ B) = R ⊗ V ⊗ W = hom(W ∗, R ⊗ V ). w 1 e e e Due to Prop. 3.2 there exists a unique linear map h : hom(W, V ) → R ⊗ V such that e e e 1 e e e e the diagram e e η1 V ❙❙ / hom(W, V ) ⊗ W (4.65) ❙❙❙ ❙❙❙ ❙❙❙ h ⊗id ❙❙❙ 1 W f1 ❙❙)  R ⊗ V ⊗ W commute, it has the form h1(wi ⊗ v) = f1(v)(ewi).e This implies the uniqueness of h. Let ′ ij us prove its existence. Let U ∈ R , it has the form U = Us wi ⊗ xs ⊗ wj ⊗ ys for some s,i,j ij K P ij kl Us ∈ and xs,ys ∈ V such that xs ⊗ys ∈ R. Since U(S) = 0 we have Us Bij xs ⊗ys = 0, s,i,j ij kl P which implies U = Us (1 − B)ij wk ⊗ xs ⊗ wl ⊗ ys. Hence s,i,j,k,lP ij kl (h1 ⊗˙ h1)U = Us (1 − B)ij (f1 ⊗˙ f1)(xs ⊗ ys)(wk ⊗ wl), (4.66) s,i,j,k,lX

47 ∗ ∗ where (f1 ⊗˙ f1)(xs ⊗ ys) belongs to R ⊗ V ⊗ W ⊗ V ⊗ W = hom(W ⊗ W , R ⊗ V ⊗ V ). Since (f ⊗˙ f )R ⊂ R , we obtain 1 1 w e e e e e e

(f1 ⊗ f1)(xs ⊗ ys) ∈ (f1 ⊗ f1)R ⊂ R ⊗ (idV ⊗σV,W ⊗ idW )(V ⊗ V ⊗ S + R ⊗ W ⊗ W ). ∗ ∗ e kl e The formula (1 − B )(Im B ) = 0 implies k,l(1 − B)ij (wk ⊗ wl)(S) = 0, so we derive

(h1 ⊗ h1)U ∈ R ⊗ R. P We proved that η is a universal arrow from B to the functor − ◦ B : QA → QA, hence e e this functor has a left adjoint cohom(B, −), which acts on an object A as defined in the proposition (see [MacLane, § 4.1, Th. 2 (ii)]). Remark 4.6. The proof of Prop. 4.5 can be directly generalised to the case of arbi- trary connected affinely generated graded algebras A, A, B such that B is finely generated: dim B < ∞. 1 e 4.4.6. Semi-connected case. The generalised internal cohom found in the previous para-

graph can be extended to the case C =(QAsc, ◦), P = FQA. Remind that any D ∈ QAsc has the form D = R ⊗A for R = D0 ∈ Alg and A ∈ QA.

Proposition 4.7. For any B ∈ FQA the functor − ◦ B : QAsc → QAsc has a left adjoint cohom(B, −): QAsc → QAsc. Let D = R ⊗A for some R ∈ Alg and A ∈ QA. Then cohom(B, D)= R ⊗ cohom(B, A), where cohom(B, A) ∈ QA is defined as in Prop. 4.5.

Proof. Let A ∈ QAsc. For an arbitrary graded homomorphism f : R ⊗A→ A consider its zero component α = f ∈ Hom (R, A ) and its restriction f ∈ Hom (A, A) to the e 0 Alg 0 b GrAlg e subalgebra A ⊂ R ⊗A. They satisfy the commutativity condition α(r)f(a ) = f(a )α(r) b e 1 e 1 ∀ r ∈ R, a1 ∈ A1. Any pair of such morphisms subjected to the commutativity condition uniquely defines a morphism f =(α, f) ∈ HomGrAlg(R ⊗A, A) (see [S, § 2.5], cf. p. 4.3.7). Let f = (α, f): D → A be a graded homomorphism to an algebra A ∈ QA , where b e sc α: R → A and f : A→ A are morphisms subjected to the commutativity condition. As it b 0 e e was proved in p. 4.4.5 there exists a unique graded homomorphism h: cohom(B, A) → A e e making the digram (4.64) commutative. By substituting f(v)= h (w ⊗ v) ⊗ wi, v ∈ V , i 1 i e to the commutativity condition α(r)f(v)= f(v)α(r) we obtain theP commutativity condition for the pair (α, h), so it gives a graded homomorphism h =(α, h): R ⊗ cohom(B, A) → A and we obtain the commutative diagram b e idR ⊗η R ⊗A / R ⊗ cohom(B, A) ◦ B (4.67) ❯❯❯❯ ❯❯❯❯ ❯❯❯ bh◦ α,h◦ b ❯❯❯ idB=( idB) f=(α,f) ❯❯❯ ❯❯❯❯  * A ◦ B

The uniqueness of h implies the uniqueness of h. e 4.4.7. Internal cohom on the morphisms.b Prop. 4.7 implies the existence of the op generalised internal cohom-functor cohom: FQA × QAsc → QAsc and gives its values on

48 objects. Let us calculate this functor on morphisms. Let B, B′ ∈ FQA, R, R′ ∈ Alg and ′ ′ ′ ′ ′ A, A ∈ QA. Set W = B1, W = B1, V = A1, V = A1. Consider arbitrary morphisms f : B′ → B and g : R ⊗A → R′ ⊗A′. They are uniquely determined by the components ′ ′ ′ ′ ′ ′ f1 : W → W , α = g0 : R → R and g1 = (α, t): R ⊗ V → R ⊗ V , where t: V → R ⊗ V . Then the morphism cohom(f,g): R ⊗ cohom(B, A) → R′ ⊗ cohom(B′, A′) is uniquely determined by the components

′ cohom(f,g)0 : R → R , (4.68) ′ ′ ′ cohom(f,g)1 : R ⊗ hom(W, V ) → R ⊗ hom(W ,V ). (4.69)

Proposition 4.8. The components (4.68), (4.69) equal

∗ cohom(f,g)0 = α, cohom(f,g)1 =(α, t∗f1 ), (4.70)

∗ ′ ′ ′ where g0 = α, g1 = (α, t) and the operator t∗f1 : hom(W, V ) → R ⊗ hom(W ,V ) is the composition

∗ f t hom(W, V ) −→1 hom(W ′,V ) −→∗ hom(W ′, R′ ⊗ V ′)= R′ ⊗ hom(W ′,V ′). (4.71)

Proof. Due to the formulae (3.12) and (3.22) we obtain

cohom(f,g) ◦ id · coev = ϑ cohom(f,g) = id ◦f · coev ·g, (4.72)    where coev coincides with the horizontal arrow in (4.67). This gives us the commuting diagram

g id ⊗η R ⊗A / R′ ⊗A′ / R′ ⊗ cohom(B′, A′) ◦ B′ (4.73)

id ⊗η id ⊗ id ◦f  cohom(f,g)◦id  R ⊗ cohom(B, A) ◦ B / R′ ⊗ cohom(B′, A′) ◦ B

By taking zero and first order components in this diagram we obtain (4.70). ′ Consider the case A, A ∈ FQA. Up to isomorphisms we have A = XA(K), A = XA′ (K), ′ ′ B = XB(K), B = XB′ (K) for some matrix idempotents A, A ,B,B . Then the morphisms f : B′ → B and g : R ⊗A→ R′ ⊗A′ have the form

′ f = fK : XB′ (K) → XB(K) and g =(α, fM ): XA(R) → XA′ (R ) (4.74)

respectively for a (B′, B)-Manin matrix K over K, an(A, A′)-Manin matrix M over R′ and ′ i i a homomorphism α: R → R such that α(r)Mj = Mj α(r) ∀ r ∈ R. Let M and N be the universal (A, B)- and (A′, B′)-Manin matrices respectively. Their entries are generators of cohom XB(K), XA(K) and cohom XB′ (K), XA′ (K) . The graded homomorphism

  ′ cohom(f,g): R ⊗ cohom XB(K), XA(K) → R ⊗ cohom XB′ (K), XA′ (K)   49 acts on the generators as

cohom(f,g): r ⊗ 1 7→ α(r) ⊗ 1, r ∈ R, (4.75) i i a b cohom(f,g): 1 ⊗Mj 7→ Ma ⊗Nb Kj . (4.76) Xa,b

Note that the matrix MN K appeared in (4.76) is a (A, B)-Manin matrix, which follows directly from [S, Prop. 2.26].

5 Quantum Representation Theory

op By taking the category C = QAsc with the Manin product ‘◦’ and parameter category P = FQAop we obtain a class of representations on quantum linear spaces. Quantum Representation Theory presented here investigates representations for this case and, more generally, for C = (GrAlgop, ◦). It can be regarded as a generalisation of Representation Theory on the usual vector spaces (classical representation theory). The latter one can be embedded into Quantum Representation Theory in two ways: by the functor S∗ or T ∗. The binary and duality operations with representations can be generalised to the quantum case.

5.1 Quantum representations First we describe objects which we are going to ‘represent’. In the classical representation theory these were algebraic monoids and algebras. In the quantum case we need to consider op op monoids in (QAsc , ◦) or, more generally, in (GrAlg , ◦). We call these monoids quantum algebras. 5.1.1. Quantum monoids. A quantum analogue of algebraic monoids is a comonoid in the monoidal category (Alg, ⊗). This is an algebra R ∈ Alg with comultiplication ∆R : R → R ⊗ R and counit εR : R → K. These are algebra homomorphisms satisfying the equations (εR ⊗ idR) · ∆R = idR = (idR ⊗εR) · ∆R and the coassociativity condition (∆R ⊗ idR) · ∆R = (idR ⊗∆R) · ∆R. Such comonoids (R, ∆R,εR) are exactly the bialgebras: Comon(Alg, ⊗)= Bimon(Vect, ⊗). To define representations of a quantum monoid we need to consider it as a quantum

,algebra, so we should translate it by means of the (not full) embedding Alg ֒→ QAsc K R 7→ R ⊗ [u]. This is a strong monoidal functor (Alg, ⊗) ֒→ (QAsc, ◦), it translates the bialgebra (R, ∆R,εR) to the comonoid O K R =(R ⊗ [u], ∆,ε) ∈ Comon(QAsc, ◦), ∆ = ∆R ⊗ idK[u], ε = εR ⊗ idK[u] . (5.1)

Component-wise we have (R ⊗ K[u])k = R, ∆k = ∆R and εk = εR. The comonoid OR is a quantum monoid considered as a comonoid in (QAsc, ◦). Note, however, that not all the comonoid structure on the semi-connected graded algebra R ⊗ K[u] has this form (because the embedding functor is not full).

50 5.1.2. Comonoids in (GrAlg, ◦). A linear version of an algebraic monoid is an algebra. Its quantum analogue is a comonoid in (QA, ◦). A quantum semi-linear monoid is a comonoid in

(QAsc, ◦). For wider generality we describe comonoids in (GrAlg, ◦), which we call quantum algebras. Such comonoid is a graded algebra A = Ak with graded homomorphisms N kL∈ 0 ∆: A→A◦A and ε: A → K[u], their components are linear maps ∆k : Ak → Ak ⊗Ak and εk : Ak → K satisfying the conditions (Ak, ∆k,εk) ∈ Comon(Vect, ⊗) ∀ k ∈ N0 and ∆k+l(albl) = ∆k(ak)∆l(bl) ∀ k,l ∈ N0, ak ∈Ak, bl ∈Al. By composing ∆ and ε with the embedding A◦A ֒→A⊗A and evaluation K[u] → K at u = 1 we get a comultiplication ∆A : A→A⊗A and a counit εA : A → K respectively (they are not graded). The conditions on the graded homomorphisms ∆ and ε imply that (A, ∆A,εA) is a bialgebra. Conversely, let a graded algebra A ∈ GrAlg has a bialgebra structure: (A, ∆A,εA) ∈ Comon(Alg, ⊗), then it gives a comonoid (A, ∆,ε) in (GrAlg, ◦) iff

∆A(Ak) ⊂Ak ⊗Ak k ∈ N0. (5.2)

5.1.3. Comonoids in (QAsc, ◦) and (QA, ◦). If A ∈ QAsc, then it is enough to re- quire (5.2) for k = 0 and k = 1:

∆A(A0) ⊂A0 ⊗A0, ∆A(A1) ⊂A1 ⊗A1. (5.3)

In this case ∆ and ε are completely determined by their zero and first order components ∆0 = α, ∆1 =(α, t), ε0 = β, ε1 = β ⊗ εV (see p. 4.3.7 for details). In the connected case A ∈ QA we have A0 = K, so it is enough to impose the con- dition (5.2) for k = 1, the graded homomorphisms ∆ and ε are uniquely determined by their first order components ∆1 = ∆V : V → V ⊗ V and ε1 = εV : V → K, where V = A1 ∈ Vect. Then the commutativity of the diagrams (4.49) reduces to the requirement that these components gives a structure of coalgebra (V, ∆V ,εV ) ∈ Comon(Vect, ⊗). The comonoid structure on a connected quadratic algebra A = TV/(R) is given by a coalgebra (23) ⊗2 ⊗2 (V, ∆V ,εV ) ∈ Comon(Vect, ⊗) such that (∆V ⊗ ∆V )R ⊂ σ (R ⊗ V + V ⊗ R) and (εV ⊗ εV )R = 0. O 5.1.4. Quantum representations. Let =(A, ∆,ε) be a comonoid in (QAsc, ◦), that is a semi-connected quadratic algebra A with a bialgebra structure (∆A,εA) satisfying (5.3). O Due to the propositions 4.5, 4.7 we have the category CorepFQA( ), which consists of corepresentations of O on finitely generated quadratic algebras B ∈ FQA. These corepre- op sentations can be regarded as representations of the corresponding monoid in (QAsc , ◦) on finite-dimensional quantum linear spaces. Let us define a slightly more general notion.

Definition 5.1. A quantum representation is a corepresentation ω : coend(B) → O of a comonoid O = (A, ∆,ε) ∈ Comon(GrAlg, ◦) on a quadratic algebra B ∈ FQA. A morphism between corepresentations ω : coend(B) → O and ω′ : coend(B′) → O is a

51 morphism f : B → B′ in FQA such that the diagram

cohom(f,idB) cohom(B′, B) / cohom(B, B) (5.4)

cohom(idB′ ,f) ω  ω′  cohom(B′, B′) / A

commute. The objects (B,ω), where B ∈ FQA and ω : coend(B) → O is a corepresentation, O form a category with such defined morphisms. We denote it by CorepFQA( ). O O Note that in the case ∈ Comon(QAsc, ◦) the category CorepFQA( ) coincides with one defined in p. 3.2.8, so in this case Theorem 3.8 implies that it is a full subcategory of Lcoact(O). More correctly, we should say that the category of quantum representations is rather M M op RepFQAop ( ), where ∈ Mon(GrAlg , ◦) is the monoid corresponding to the comonoid O O . This is the category opposite to CorepFQA( ). The morphisms of quantum representa- tions are reversed morphisms of corepresentations. However, it is more convenient for us to O work in terms of category CorepFQA( ). i 5.1.5. Multiplicative Manin matrices. An m × m matrix M = (Mj ) with entries in a bialgebra R =(R, ∆R,εR) (or, more generally in a coalgebra) is called multiplicative if

m i i k i i ∆R(Mj )= Mk ⊗ Mj , εR(Mj )= δj, i, j =1,...,m (5.5) Xk=1 (see [Man88, § 2.6], [Man91, § 4.1.1]). For example, the formulae (4.59) means exactly that i the universal B-Manin matrix M =(Mj) is a multiplicative m × m matrix over the algebra K i coend XB( ) . Recall that Mj are generators of this algebra. Let us establish a relationship between the notions of multiplicative Manin matrix and of quantum representation. Consider a comonoid O =(A, ∆,ε) in (GrAlg, ◦).

Theorem 5.2. Let B ∈ End(Km ⊗ Km) be a matrix idempotent. Any corepresentation ω : coend(B) → O of the comonoid O on the quadratic algebra B = XB(K) acts on the generators by the formula

i i ω(Mj)= Mj , (5.6)

i where M = (Mj ) is a multiplicative first order B-Manin matrix over A, that is an m × m i matrix with entries Mj ∈A1 satisfying (5.5) and

BM (1)M (2)(1 − B)=0. (5.7)

Conversely, any such matrix M defines a corepresentation of the comonoid O on B by the formula (5.6).

52 Proof. Any graded homomorphism ω : coend(B) →A is uniquely determined by its values i i Mj ∈ A1 on the generators Mj. The values (5.6) give a graded homomorphism ω iff they i belong to the component A1 and satisfy the same commutation relations that Mj do, i.e.the relations (5.7). The commutativity of the corresponding diagrams (2.15) is equivalent to the condition (5.5). Warning. The same multiplicative first order matrix M defines different corepresentations O K (different objects of the category CorepFQA( )) on different quadratic algebras B = XB( ) ′ K ′ ′ and B = XB′ ( ) if B1 = B1 and M is simultaneously B- and B -Manin matrix. In other words, a matrix M is not enough to fix a corepresentation, one also needs to know the quadratic algebra B (and also the comonoid O, of cause). By a quantum representation of a quantum monoid we understand a corepresentation of the comonoid, obtained by embedding of the bialgebra (R, ∆R,εR) via Alg ֒→ QAsc. This is a corepresentation of the comonoid OR =(R⊗K[u], ∆R ⊗idK[u],εR ⊗idK[u]) defined by (5.1). Theorem 5.2 implies in this case a one-to-one correspondence between corepresentations of OR on B = XB(K) and multiplicative B-Manin matrices over R. A trivial but important example of quantum representation is the identical morphism coend(B) −→id coend(B). It is a corepresentation of the comonoid O = coend(B) on the object B ∈ FQA. It is given by the universal B-Manin matrix M. The corresponding coaction coincides with the coevaluation (4.57) for A = B. 5.1.6. Morphisms of quantum representations. Consider two quantum representations ω : coend(B) → O and ω′ : coend(B′) → O of the same O = (A, ∆,ε) ∈ (GrAlg, ◦) on quadratic algebras B, B′ ∈ FQA. Suppose that these algebras are defined by matrix m m ′ m′ m′ ′ idempotents B ∈ End(K ⊗ K ), B ∈ End(K ⊗ K ), i.e. B = XB(K), B = XB′ (K). Due to Theorem 5.2 these corepresentations are given by multiplicative first order B- and B′-Manin matrices M and M ′ respectively: ω(M) = M, ω(M′) = M ′, where M and M′ are the universal B- and B′-Manin matrices. ′ ′ O Proposition 5.3. Any morphism f : (B,ω) → (B ,ω ) in the category CorepFQA( ) has ′ m′ m the form f = fK : XB(K) → XB′ (K) for an m × m matrix K ∈ Hom(K , K ) satisfying BK(1)K(2)(1 − B′)=0 and

KM ′ = MK. (5.8)

This gives a bijection between the morphisms f : (B,ω) → (B′,ω′) and (B, B′)-Manin matri- ces K over K satisfying (5.8).

Proof. The graded homomorphisms f : XB(K) → XB′ (K) is in one-to-one correspondence with the (B, B′)-Manin matrices over K (see p. 4.2.3). By taking the entries of the universal (B, B′)-Manin matrix in the left upper corner of the diagram (5.4) we obtain exactly the condition (5.8). Since these entries generate the algebra cohom(B′, B) the commutativity of the diagram (5.4) is equivalent to (5.8).

53 ′ ′ Corollary 5.4. Let B1 = B1. Suppose that this identification induces a morphism f : B → B ′ ′ in FQA, that is f1 = idB1 (this is not always true). Then f is a morphism (B,ω) → (B ,ω ) O ′ ′ in CorepFQA( ) iff the corepresentations ω and ω are given by the same matrix M = M .

Proof. Since f1 = idB1 we have f = fK for the identity matrix K = 1, so the condition (5.8) reduces to M = M ′ (the equation B(1 − B′) = 0 is equivalent to the requirement that the matrix K = 1 defines a morphism fK : XB(K) → XB′ (K) in the category FQA). 5.1.7. Coactions in (GrAlg, ◦). Let O be a comonoid in (GrAlg, ◦) corresponding to a bialgebra (A, ∆A,εA) in the sense of p. 5.1.2. Consider a coaction δ : B→A◦B of O on B ∈ GrAlg. By composing it with the embedding A ◦ B ֒→A⊗B we obtain a coaction δA : B→A⊗B of the bialgebra (A, ∆A,εA) in the monoidal category (Alg, ⊗). Conversely, a coaction δA : B→A⊗B of the bialgebra (A, ∆A,εA) gives a coaction δ : B→A◦B of O in (GrAlg, ◦) iff

δA(Bk) ⊂Ak ⊗ Bk, k ∈ N0. (5.9)

If B ∈ QA or B ∈ QAsc then it is enough to require this condition for k = 1 only or for k =0, 1 only respectively.

Proposition 5.5. Each coaction δ : XB(K) →A◦ XB(K) has the form

i i j δ(w )= Mj ⊗ w , (5.10) Xj

i j where M =(Mj ) is a multiplicative first order B-Manin matrix and (w ) is the standard basis m ∗ of (K ) . Any such matrix M define a coaction of O on XB(K). If ω is a corepresentation of O on XB(K) given by the matrix M, then the diagram

η X (K) / coend X (K) ◦ X (K) (5.11) B ❯ B B ❯❯❯❯ ❯❯❯❯  ❯❯❯❯ ω◦id δ ❯❯❯❯ ❯❯❯*  A◦ XB(K) commute. In particular, if A ∈ QAsc, then ϑ(ω)= δ.

Proof. Any graded homomorphism δ : XB(K) →A◦ XB(K) has the form (5.10) for a B- Manin matrix M (see p. 4.2.3). By taking wi in the left upper corners of the diagrams (2.22) we obtain the conditions (5.5). The commutativity of the diagram (5.11) is checked on the generators wi in a straightforward way. We established the bijection between corepresentations ω : cohom(B) → O and coac- tions δ : B→A◦B for an arbitrary comonoid O = (A, ∆,ε) ∈ Comon(GrAlg, ◦). This generalises the case C = QAsc of the bijection (3.23), so we denote it by the same letter: ϑ: ω ↔ δ.

54 ′ ′ Proposition 5.6. Let δ and δ be coactions of O on XB(K) and XB′ (K). Let M and M be the corresponding multiplicative first order B- and B′-Manin matrices. Then morphisms ′ ′ f : XB(K), δ → XB′ (K), δ in Lcoact(O) are homomorphisms f = fK for (B, B )-Manin K O O matrices K over satisfying (5.8). Hence CorepFQA( ) is a subcategory of Lcoact( ). Proof. By substituting wi to the left upper corner of the left diagram (2.23) we obtain the condition (5.8).

5.2 Operations with quantum representations We have tree operations with vector spaces: direct sum, tensor product and duality. They induce operations with representations (in the case of tensor product and duality we need an additional structure on the represented object). Similarly, the operations with quadratic algebras give operations with quantum representations. We construct these operations in terms of coactions and Manin matrices. 5.2.1. Direct sum of Manin matrices. According to the Manins’ ideas [Man88] the quantum analogue of the direct sum of vector spaces (in the ‘even’ interpretation) is the (even) tensor product of quadratic algebras (see p. 4.1.7 and p. 4.1.8). Therefore, we describe first the tensor product B ⊗ C for quadratic algebras B, C ∈ FQA. For arbitrary matrix idempotents B ∈ End(Km ⊗ Km) and C ∈ End(Kn ⊗ Kn) define the idempotent D = DiS(B,C) ∈ End(Km+n ⊗ Km+n) with the only non-zero entries

¯ 1 1 Dij = Bij,Da,¯ b = Cab Di,a¯ = Da,i¯ = ,Di,a¯ = Di,a¯ = − , (5.12) kl kl c,¯ d¯ cd i,a¯ a,i¯ 2 a,i¯ a,i¯ 2

where i,j,k,l =1,...,m, a,b,c,d =1,...,n,a ¯ = m+a. Then XB(K)⊗XC (K)= XD(K). ′ ′ ′ ′ Let B′ ∈ End(Km ⊗ Km ) and C′ ∈ End(Kn ⊗ Kn ) be idempotents as well. Denote ′ ′ ′ m′+n′ m′+n′ D = DiS(B ,C ) ∈ End(K ⊗ K ). Graded operators f : XB(K) → XB′ (R) and g : XC(K) → XC′ (R) give one more graded operator by the composition

f⊗g σ(23) XD(K)= XB(K) ⊗ XC(K) −−→ R ⊗ XB′ (K) ⊗ R ⊗ XC′ (K) −−−→

µR⊗id ⊗ id R ⊗ R ⊗ XB′ (K) ⊗ XC′ (K) −−−−−−→ R ⊗ XB′ (K) ⊗ XC′ (K)= XD′ (R). (5.13)

The condition that this operator is an algebra homomorphism can be written in terms of Manin matrices.

Proposition 5.7. Let M and N be m × m′ and n × n′ matrices over R. Consider their direct sum M 0 L = M ⊕ N = . (5.14)  0 N

55 ′ ′ i i This is an (m + n) × (m + n ) matrix over R with the only non-zero entries Lj = Mj , a¯ a ′ L¯b = Nb , where i, j =1,...,m, a, b =1,...,n. The matrix L is a (D,D )-Manin matrix iff M is (B, B′)-Manin matrix, N is (C,C′)-Manin matrix and

i a a i Mj Nb = Nb Mj , ∀ i, j =1,...,m, a,b =1, . . . , n. (5.15)

In other words, the graded operator (5.13) is a graded homomorphisms iff f, g are algebra homomorphisms and the corresponding Manin matrices entry-wise commute. Proof. By writing the condition DL(1)L(2)(1−D′) = 0 in entries one sees that it is equivalent to the conditions BM (1)M (2)(1 − B′) = 0, CN (1)N (2)(1 − C′) = 0 and (5.15). 5.2.2. Direct sum of quantum representations. Consider coactions δ : B →A◦B and γ : C →A◦C of the comonoid O = (A, ∆,ε) ∈ Comon(GrAlg, ◦) on the objects B, C ∈ GrAlg. Let (A, ∆A,εA) ∈ Comon(Alg, ⊗) be the bialgebra corresponding to the comonoid O and δA : B→A⊗B, γA : C →A⊗C be the corresponding coactions of the bialgebra (A, ∆A,εA). The letter one is the bimonoid (A,µA,εA, ∆A,εA) in (Vect, ⊗), where µA : A⊗A→A and ηA : K → A are the multiplication and unity map in the algebra A, i.e. µA(a ⊗ b)= ab, ηA(1) = 1A ∈A0. We define an analogue of direct sum for the coactions δ and γ as a coaction of O on B ⊗ C by using the monoidal category (Alg, ⊗). According to p. 2.4.10 we need to introduce the structure of a bimonoid on the comonoid (A, ∆A,εA) ∈ Comon(Alg, ⊗), this is a pair ′ ′ ′ ′ (µA, ηA) such that (A,µA, ηA, ∆A,εA) is a bimonoid in (Alg, ⊗), where A is understood as the algebra (A,µA, ηA) ∈ Alg. But (Alg, ⊗) = Mon(Vect, ⊗), so due to Prop. 2.2 we obtain

Bimon(Alg, ⊗)= Comon Mon(Alg, ⊗) = Comon Mon Mon(Vect, ⊗) =    Comon cMon(Vect, ⊗) = Comon(CommAlg, ⊗).

′ ′  This implies (A,µA, ηA)=(A,µA, ηA) ∈ CommAlg, so we can define the direct sum of coactions by the formula (2.25) iff the algebra A is commutative. We need to use in this case the multiplication µA in (2.25). In the non-commutative case we define the direct sum in an immediate way, but not for every pair (δ, γ). Consider the linear map

δA⊗γA σ(23) µA⊗id ⊗ id B ⊗ C −−−−→A ⊗ B ⊗ A ⊗ C −−−→A ⊗ A ⊗ B ⊗ C −−−−−−→A ⊗ B ⊗ C. (5.16)

The conditions (5.9) and µA(Ak ⊗Al) ⊂ Ak+l imply that the image of Bk ⊗ Cl under the map (5.16) lies in Ak+l ⊗ Bk ⊗ Cl, so this map also satisfies (5.9). Hence it gives the graded map B⊗C→A◦ (B ⊗ C), which we denote by δ ∔ γ. Definition 5.8. We say that the direct sum of the coactions δ, γ exists if δ ∔ γ is an algebra homomorphism. In this case δ ∔ γ is a coaction of O on B ⊗ C called direct sum of the coactions δ and γ. Suppose B, C ∈ FQA, so we can consider the corresponding

56 corepresentations ω = ϑ−1(δ): coend(B) → O and ν = ϑ−1(γ): coend(C) → O. If the direct sum of δ, γ exists, then

ω ∔ ν := ϑ−1(δ ∔ γ): coend(B ⊗ C) → O

is a corepresentation of the comonoid O on the object B ⊗ C called direct product of corep- resentations ω and ν. In this case we say that the direct sum of corepresentations ω, ν exists.

Proposition 5.9. Let B = XB(K) and C = XC (K), D = DiS(B,C). Let δ and γ be coactions of O given by multiplicative first order B- and C-Manin matrices M and N respectively. The direct sum of the coactions δ, γ exists iff M and N entry-wise commute in the sense of (5.15). Thus the direct sum of the corepresentations ω, ν exists iff the corresponding Manin matrices M and N satisfy this condition. In this case the coaction δ ∔ γ : XD(K) → A◦ XD(K) and corepresentation ω ∔ ν : coend XD(K) → O are given by the multiplicative first order D-Manin matrix (5.14). 

Proof. Since B0 = C0 = K we have (B ⊗ C)1 = B1 ⊕ C1, so the first order component of δ ∔ γ coincides with δ1 ⊕ γ1. Hence the graded map δ ∔ γ : B⊗C→A◦ (B ⊗ C) is given by the matrix L = M ⊕ N. Since M and N are multiplicative first order matrices, so is L. The rest follows from Prop. 5.5 and 5.7. In general, the operation of direct sum does not give a monoidal structure on Lcoact(O) O ′ ′ O ′ ′ nor on CorepFQA( ). However if δ, γ, δ ,γ are coactions of on B, C, B , C such that the direct sums of the coactions δ, γ and of δ′,γ′ exist, then due to Prop. 5.9 and 5.6 any morphisms f : (B, δ) → (B′, δ′) and g : (C,γ) → (C′,γ′) in Lcoact(O) give a morphism

f ∔ g : (B ⊗ C, δ ∔ γ) → (B′ ⊗ C′, δ′ ∔ γ′).

Thus we have the functor from a full subcategory of Lcoact(O)×Lcoact(O) to Lcoact(O). The objects of this subcategory are the pairs of coactions for which the direct sum exists. Note that the direct sum of the coactions δ, γ exists iff it exists for γ, δ; in this case the isomorphism σB,C : B ⊗ C ∼= C ⊗ B gives the isomorphism (B ⊗ C, δ ∔ γ) ∼= (C ⊗ B,γ ∔ δ) in Lcoact(O). In particular, if the algebra A is commutative, then the direct sum functor (B, δ) ∔ (C,γ)=(B ⊗ C, δ ∔ γ) defines a symmetric monoidal structure on the category O O Lcoact( ) and on its subcategory RepFQA( ). 5.2.3. Categorical product of quantum representations. In p. 4.1.9 we constructed the coproduct of quadratic algebras, which corresponds to the product of quantum linear spaces in the category FQAop. Let us modify the definition of direct sum of corepresentations for the case of coproduct B ∐ C. m m Let XB(K) and XC (K) be quadratic algebras for idempotents B ∈ End(K ⊗ K ) and C ∈ End(Kn ⊗ Kn). Define the idempotent E = CoP(B,C) ∈ End(Km+n ⊗ Km+n) with the only non-zero entries

ij ij a,¯ ¯b ab Ekl = Bkl, Ec,¯ d¯ = Ccd , (5.17)

57 i,a¯ a,i¯ i,a¯ i,a¯ where i,j,k,l =1,...,m, a,b,c,d =1,...,n,a ¯ = m + a (the entries Ei,a¯ , Ea,i¯ , Ea,i¯ , Ea,i¯ are also zero in this case). Then XB(K) ∐ XC (K)= XE(K).

′ ′ ′ ′ Proposition 5.10. Consider idempotents B′ ∈ End(Km ⊗ Km ), C′ ∈ End(Kn ⊗ Kn ) and ′ ′ ′ ′ E′ = CoP(B′,C′) ∈ End(Km +n ⊗ Km +n ). Let M and N be m × m′ and n × n′ matrices over R. Let L = M ⊕ N be defined as in Prop. 5.7. Then L is a (E, E′)-Manin matrix iff M is (B, B′)-Manin matrix and N is (C,C′)-Manin matrix. Proof. The same as for Prop. 5.7. Note that in this case the commutativity condition (5.15) is not needed. Therefore we can define corepresentation of O on B ∐ C for any corepresentations ω, ν on B, C ∈ FQA.

Definition 5.11. Let B = XB(K) and C = XC (K). Then we have B ∐ C = XE(K) for E = CoP(B,C). Consider corepresentations ω : coend(B) → O and ν : coend(C) → O given by multiplicative first order B- and C-Manin matrices M and N over A. By taking into account Prop. 5.10 we see that L = M ⊕ N is a multiplicative first order E-Manin matrix over A. We call the corepresentation ω ⊔˙ ν : coend(B ∐ C) → O given by the matrix L coproduct of corepresentations ω and ν. Let δ = ϑ(ω), γ = ϑ(ν) be coactions of O on B and C given by the matrices M and N. By coproduct of coactions δ and γ we mean the coaction δ ⊔˙ γ = ϑ(ω ⊔˙ ν) on B ∐ C given by the matrix L. Note that the coproduct of corepresentations (coactions) is given by the same matrix L as the their direct sum, but this is a corepresentation (coaction) on another quadratic algebra (see the warning in p. 5.1.5). By using the results of p. 4.1.9 and Prop. 5.6 one can check that (B ∐ C,ω ⊔˙ ν) is a O coproduct of the objects (B,ω) and (C, ν) in the category CorepFQA( ). Hence the latter is a category with finite coproducts (its initial object is the algebra K with the trivial corepresentation). If we denote this coproduct by (B,ω) ⊔˙ (C, ν) := (B ∐ C,ω ⊔˙ ν), then we O O O obtain a bifunctor −⊔−˙ : CorepFQA( )×CorepFQA( ) → CorepFQA( ). The coproduct of corepresentations corresponds to the categorical product of quantum representations in O op CorepFQA( ) . 5.2.4. Tensor product of Manin matrices. The quantum analogue of the tensor product is the Manin product (see p. 4.1.7 and p. 4.1.8). Let us describe the Manin product of the connected finitely generated quadratic algebras B = XB(K) and C = XC (K), where B ∈ End(Km ⊗ Km) and C ∈ End(Kn ⊗ Kn) are matrix idempotents. Proposition 5.12. The operator F = TeP(B,C) ∈ End(Km ⊗ Kn ⊗ Km ⊗ Kn) defined as

F = σ(23)(B ⊗ 1+1 ⊗ C − B ⊗ C)σ(23), (5.18) is an idempotent. We have XB(K) ◦ XC (K)= XF (K). Proof. The idempotentness of (5.18) is equivalent to the idempotentness of the operator

F = σ(23)F σ(23) = B ⊗ 1+1 ⊗ C − B ⊗ C ∈ End(Km ⊗ Km ⊗ Kn ⊗ Kn), (5.19) e 58 which, in turn, directly follows from B2 = B and C2 = C. Let W =(Km)∗ and V =(Kn)∗, ∗ ∗ ∗ ⊗2 then we have XB(K)= TW/(Im B ) and XC(K)= TV/(Im C ), where B ∈ end(W ) and ∗ ⊗2 C ∈ end(V ) are transposed idempotents (see p. 4.2.1). The algebra XB(K)◦XC(K) is the (23) ∗ ⊗2 ⊗2 ∗ quotient T (W ⊗V /(Rw), where Rw = σ (Im B ⊗V +W ⊗Im C ). Hence it is enough to check that the subspace Im B∗⊗V ⊗2+W ⊗2⊗Im C∗ = Im(B∗⊗1)+Im(1⊗C∗) ⊂ W ⊗2⊗V ⊗2 coincides with σ(23) Im F ∗ = Im F ∗. We have Im F ∗ = Im(B∗ ⊗ 1+1 ⊗ C∗ − B∗ ⊗ C∗) ⊂ Im(B∗ ⊗ 1) + Im(1 ⊗ C∗). On the other hand, the formulae B∗ ⊗ 1 = F (B∗ ⊗ 1) and e e 1 ⊗ C∗ = F (1 ⊗ C∗) imply Im(B∗ ⊗ 1) ⊂ Im(F ∗) and Im(1 ⊗ C∗) ⊂ Im(F ∗). ′ ′ ′ e ′ Consider two more idempotents B′ ∈ End(Km ⊗ Km ), C′ ∈ End(Kn ⊗ Kn ). Define e ′ ′ ′ ′ e e F ′ = DiS(B′,C′) ∈ End(Km +n ⊗ Km +n ). We have F ′ = σ(23)F ′σ(23) with

F ′ = B′ ⊗ 1+1 ⊗ C′ − B′ ⊗ C′. e (5.20)

For graded homomorphisms f :eXB(K) → XB′ (R) and g : XC (K) → XC′ (R) consider a graded operator

f◦g (µR) XF (K)= XB(K) ◦ XC(K) −−→ XB′ (R) ◦ XC′ (R) −−→ XF ′ (R), (5.21)

(µR) where XB′ (R) ◦ XC′ (R) −−→ XD′ (R) is the multiplication in the tensor factors R, that is a graded operator with the components

σ(23) XB′ (R)k ⊗ XC′ (R)k = R ⊗ XB′ (K)k ⊗ R ⊗ XC′ (K)k −−−→

µR⊗id ⊗ id R ⊗ R ⊗ XB′ (K)k ⊗ XC′ (K)k −−−−−−→ R ⊗ XB′ (K)k ⊗ XC′ (K)k = XF ′ (R)k. The homomorphisms f and g are given by some (B, B′)- and (C,C′)-Manin matrices M and N. These are m × m′ and n × n′ matrices over R. The graded operator (5.21) is given by the matrix P = M (1)N (2), this is an (mn) × (m′n′) matrix over R with the entries

ia i a Pjb = Mj Nb . (5.22)

′ ′ In the notations of p. 4.3.7 this is the operator P = M ⊗˙ N : Km ⊗ Kn → R ⊗ Km ⊗ Kn. Proposition 5.13. Suppose M and N are (B, B′)- and (C,C′)-Manin matrices. The matrix P = M (1)N (2) = M ⊗˙ N is an (F, F ′)-Manin matrix iff

F M (1)(M (2)N (3) − N (3)M (2))N (4)(1 − F ′)=0, (5.23)

′ e e where F and F are given by the formulae (5.19), (5.20). Let f = fM and g = fN , then (5.23) is the condition for the graded operator (5.21) to be an algebra homomorphism. e e Proof. The matrix P is an (F, F ′)-Manin matrix iff F M (1)N (2)M (3)N (4)(1 − F ′) = 0. By means of the conjugation by σ(23) we obtain the equivalent condition

FM (1)N (3)M (2)N (4)(1 − F ′)=0. (5.24) e e 59 Due to BM (1)M (2) = BM (1)M (2)B′ and CN (1)N (2) = CN (1)N (2)C′ we derive

(B ⊗ 1)M (1)M (2)N (3)N (4) =(B ⊗ 1)M (1)M (2)N (3)N (4)(B′ ⊗ 1), (1 − B) ⊗ C M (1)M (2)N (3)N (4) = (1 − B) ⊗ C M (1)M (2)N (3)N (4)(1 ⊗ C′).   Addition and multiplication by 1 − F ′ = (1 − B′) ⊗ (1 − C′) on the right gives

FM (1)Me (2)N (3)N (4)(1 − F ′)=0. (5.25)

By subtracting (5.24) from (5.25)e we get (5.23). e Note that the condition (5.23) is valid for entry-wise commuting matrices M and N in the sense of Prop. 5.7. However, the condition (5.15) is not necessary. For example, if B, B′,C,C′ are 0, then any matrices M and N satisfy (5.23). 5.2.5. Tensor product of quantum representations. Let O =(A, ∆,ε) be a comonoid in (GrAlg, ◦), where A = (A,µA, ηA) ∈ GrAlg. By tensor product of its coactions or corepresentations on B and B′ we mean the corresponding coaction or, respectively, corepre- sentation on B ◦ B′. One can define it as a monoidal product in the category Lcoact(O) in- troduced in p. 2.4.10 if O has additionally a structure of bimonoid B =(A, µ, η, ∆,ε), where µ: A◦A→A and η : K[u] → A are morphisms of graded algebras compatible with the comonoid structure of O in the sense of p. 2.4.6 (the homomorphisms µ, η are not to be con- fused with the maps µA : A⊗A→A and ηA : K →A). This means the following conditions on the components µk : Ak ⊗Ak →Ak and ηk : K →Ak. First, they must give an associative k multiplication on each component Ak with the unities 1k = ηk(1) = η(u ) ∈ Ak such that (Ak,µk, ηk, ∆k,εk) ∈ Bimon(Vect, ⊗), where ∆k : Ak →Ak ⊗Ak and εk : Ak → K are the graded components of ∆ and ε. Second, they must be compatible with the multiplication ′ ′ ′ ′ µA(a⊗b)= ab and with the unity 1A = ηA(1), that is µk+l(akbl ⊗akbl)= µk(ak ⊗ak)µl(bl ⊗bl) ′ ′ for any ak, ak ∈ Ak, bl, bl ∈ Al and 1k+l = 1k1l, 10 = 1A. In particular, the vector space A0 has two multiplications (µA)0 and µ0 with the same unity. By Eckmann–Hilton Prin- ciple (Prop. 2.2) the compatibility condition µ0(ab ⊗ cd) = µ0(a ⊗ c)µ0(d ⊗ d) implies that µ0 =(µA)0 and this is a commutative multiplication on A0. Moreover, one can show that a monoid structure (µ, η) on a graded algebra R ⊗ S(V ) exists only if dim V 6 1.

Consider a bialgebra (R, ∆R,εR) ∈ Comon(Alg, ⊗). The embedding Alg ֒→ QAsc gives the comonoid OR =(R ⊗ K[u], ∆,ε) defined by the formula (5.1). If R ∈ CommAlg, then the comonoid OR has a unique bimonoid structure:

BR =(R ⊗ K[u], µ, η, ∆,ε), µ = µR ⊗ idK[u], η = ηR ⊗ idK[u] . (5.26)

In this case we get a tensor product of coactions as the monoidal product in Lcoact(BR). Here we have a situation similar to one from p. 5.2.2: we can define the tensor product without commutativity condition on µR, but this product not always exists. Namely, consider an arbitrary R ∈ Alg. Note that (R ⊗ K[u]) ◦ XB(K) = XB(R). Let δ : XB(K) → XB(R), γ : XC (K) → XC (R) be coactions of OR on B = XB(K) and C = XC (K) given by multi- plicative B- and C-Manin matrices M and N over R. By substituting the homomorphisms

60 f = fM = δ and g = fN = γ to (5.21) we obtain the graded operator

δ◦γ (µR) XF (K)= XB(K) ◦ XC (K) −−→ XB(R) ◦ XC (R) −−→ XF (R), (5.27) where F = TeP(B,C)= σ(23)(B ⊗ 1+1 ⊗ C − B ⊗ C)σ(23). Denote the composition (5.27) by δ ◦˙ γ. Due to Prop. 5.13 this operator is a coaction of OR on B ◦ C iff the matrices M and N satisfy (5.23). Definition 5.14. We say that the tensor product of the coactions δ, γ exists if δ ◦˙ γ is an algebra homomorphism, i.e. if

δ ◦˙ γ : B◦C→A◦B◦C = R ⊗B◦C is a coaction of OR on B ◦ C, we call it tensor product of the coactions δ and γ. Let −1 −1 ω = ϑ (δ): coend(B) → OR, ν = ϑ (γ): coend(C) → OR be the corresponding corepre- sentations of OR. If δ ◦˙ γ is a coaction, then

−1 ω ◦˙ ν := ϑ (δ ◦˙ γ): coend(B ◦ C) → OR is a corepresentation of OR on B ◦ C called tensor product of the corepresentations ω and ν. In this case we say that the tensor product of corepresentations ω, ν exists. It follows from Prop. 5.3 again, that the tensor product of corepresentations is a functor O O O − ◦˙ − from a full subcategory of CorepFQA( R) × CorepFQA( R) to CorepFQA( R). It has the form (B,ω) ◦˙ (C, ν)=(B ◦ C,ω ◦˙ ν). If R ∈ CommAlg, this functor is defined on the O O whole CorepFQA( R) × CorepFQA( R) and, consequently, gives a structure of monoidal O category CorepFQA( R), ◦˙ . It is symmetric due to commutativity of µR. Similarly we can define a tensor product of the quantum representations ω and ν on the (23) (23) black Manin product B • C = XG(K), where G = σ (B ⊗ C)σ is the corresponding idempotent. Since (B • C)1 = (B ◦ C)1, a corepresentation on the black Manin product can be defined by the same matrix P = M⊗˙ N = M (1)N (2), which is multiplicative with respect to ∆R. It is a G-Manin matrix iff

(B ⊗ C)M (1)(M (2)N (3) − N (3)M (2))N (4)(1 − B ⊗ C)=0. (5.28)

Under this condition we obtain black tensor product

ω •˙ ν : coend(B • C) → OR (5.29) defined by the matrix M⊗˙ N. This is a corepresentation of OR on the quadratic algebra B • C. We get one more partially defined binary operation with quantum representations. O O In the case R ∈ CommAlg, it is a bifunctor − •˙ −: CorepFQA( R) × CorepFQA( R) → O CorepFQA( R). 5.2.6. Opposite and coopposite quantum representations. For an arbitrary comonoid O = (A, ∆,ε) ∈ Comon(GrAlg, ◦) denote Oop = (Aop, ∆,ε), where Aop is the opposite

61 graded algebra (the opposite monoid in (GrVect, ⊗) in the sense of p. 2.4.8). The cor- op responding bialgebra (A , ∆A,εA) is the bimonoid in (Vect, ⊗) opposite to the bimonoid op op (A, ∆A,εA). This bialgebra satisfies (5.2), so the triple O =(A , ∆,ε) is also a comonoid in (GrAlg, ◦). The opposite graded algebra to a quadratic algebra A = TV/(R) ∈ QA has the form op m m A = TV/(σV,V R) (see [Man88]). Let B = XB(K) ∈ FQA for some B ∈ end(K ⊗K ) and (21) m m 2 2 denote B = σ ·B ·σ ∈ end(K ⊗K ), where σ = σKm,Km . This is an m ×m matrix over K (21) ij ji op K with entries (B )kl = Blk. The corresponding quadratic algebra is B = XB(21) ( ). Let i op op M = (Mj) be the universal B-Manin matrix. The algebra cohom(B , B ) is generated i (21) (1) (2) (21) by the same Mj, but with the relations B M M (1 − B ) = 0. In other words, the matrix M considered as a matrix over cohom(Bop, Bop) is the universal B(21)-Manin op i matrix. On the other hand the algebra hom(B, B) is generated by Mj with the relations BM(2)M(1)(1 − B) = 0, which is equivalent to the previous relations. Thus we obtain coend(Bop)= coend(B)op for any B ∈ FQA. Let ω : coend(B) → O be a corepresentation. The same linear map is a comonoid mor- phism coend(Bop)= coend(B)op → Oop, so this is a corepresentation of Oop on Bop. Let us denote it by ωop and call the corepresentation opposite to ω. Note that the corepresentations ω and ωop is given by the same multiplicative B-Manin matrix M, which is considered as B- or as B(21)-Manin matrix over A or over Aop respectively. We have the algebra isomorphism cohom(B, B) = B! • B = (B!)! • B! = cohom(B!, B!), so the formulae (4.60) implies the isomorphism coend(B!)= coend(B)cop. Any corepresen- tation ω : coend(B) → O can be regarded as a corepresentation ωcop : coend(B!) → Ocop of cop ! the coopposite comonoid O on the Koszul dual quadratic algebra B =ΞB(K), we call it coopposite corepresentation to the corepresentation ω. It corresponds to the (1 − B⊤)-Manin matrix M ⊤ over A, which is multiplicative with respect to ∆cop. By applying these two operations (in any order) to the corepresentation ω we obtain opposite coopposite corepresentation ωop,cop : coend (B!)op → Oop,cop. This is a corep- resentation of the comonoid Oop,cop = (Aop, ∆cop,ε), which corresponds to the opposite op cop ! op op ! K coopposite bialgebra (A , ∆A ,εA), on the quadratic algebra (B ) =(B ) =ΞB(21) ( ). 5.2.7. Dual and Koszul dual quantum representations. Assume that the bialgebra R = (R,µR, ηR, ∆R,εR) is a Hopf algebra and let ζR : R → R be its antipode. This is an algebra anti-homomorphism as well as a coalgebra anti-homomorphism in the sense

op ζR · µR = µR · (ζR ⊗ ζR), ζR · ηR = ηR, (5.30) cop (ζR ⊗ ζR) · ∆R = ∆R · ζR, εR · ζR = εR (5.31)

op,cop (see [Kass, Th. III.3.4]). This means that ζR is a morphism R → R in Bimon(Vect, ⊗), op,cop op cop K where R = (R,µR , ηR, ∆R ,εR). By tensoring with [u] we obtain the morphism Oop,cop O O K ζ = ζR ⊗ idK[u] : R → R, where R = (R ⊗ [u], ∆ = ∆R ⊗ idK[u],ε = εR ⊗ idK[u]) is Oop,cop op K cop the bialgebra R embedded to Comon(GrAlg, ◦) and R =(R ⊗ [u], ∆ ,ε). Let ω : coend(B) → OR be a corepresentation of the comonoid OR on a quadratic K i algebra B = XB( ) ∈ FQA given by a multiplicative B-Manin matrix M = (Mj ) over R.

62 Then the composition

op,cop D ! op ω Oop,cop ζ O ω : coend (B ) −−−−→ R −→ R (5.32)  is a corepresentation of the same comonoid OR on the opposite Koszul dual quadratic algebra ! op K (B ) =ΞB(21) ( ). Let us call it dual corepresentation to the corepresentation ω. This is a quantum analogue of dual (contragredient) representation on a dual vector space. Since M is a multiplicative matrix over a Hopf algebra, it is invertible. The entries of the −1 −1 i i inverse matrix M are (M )j = ζR(Mj ). Due to the fact that ζR is an anti-endomorphism of R the matrix M −1 is a B(21)-Manin matrix. The dual corepresentation (5.32) is given by the (1 − B(21))⊤-Manin matrix (M ⊤)−1 = (M −1)⊤ over R, which is multiplicative with respect to the comultiplication ∆R. Consider the corepresentation opposite to (5.32). It has the form

cop KD ! ω Ocop ζ Oop ω : coend(B ) −−→ R −→ R . (5.33) Oop ! This is a corepresentation of R on the Koszul dual quadratic algebra B . Let us call it Koszul dual corepresentation to the corepresentation ω : coend(B) → OR. It corresponds to the (1 − B⊤)-Manin matrix (M ⊤)−1 =(M −1)⊤ over Rop. Let ν : coend(C) → OR be a corepresentation of OR on C ∈ FQA and N be the corresponding multiplicative matrix. By using (4.17) we get B • C =(B! ◦ C!)!. One can show that the tensor product of ωKD, νKD exists iff the black tensor product of ω, ν exists (the condition (5.28) fulfils). In this case the Koszul dual to ωKD ◦˙ νKD is the corepresentation KD KD KD ω ◦˙ ν of OR on the black Manin product B • C, given by the multiplicative matrix M (1)N (2). It coincides with the corepresentation (5.29). Suppose the Hopf algebra R is commutative as an algebra. Then the Koszul dual to KD a corepresentation ω : coend(B) → OR is a corepresentation ω of the same comonoid Oop O ! O R = R on the object B . The comonoid R can be considered as a commutative bimonoid BR =(R⊗K[u],µ = µR⊗idK[u], η = ηR⊗idK[u], ∆,ε), it is a Hopf monoid with the antipode ζ. By virtue of the formulae (4.17), (4.51) we derive cohom(B, C) = B! • C =(B ◦ C!)!, so the homomorphism

KD KD KD hω, νi := ω •˙ ν = ω ◦˙ ν : coend cohom(B, C) → OR (5.34)   is a corepresentation of OR on the internal cohom of the corresponding quadratic algebras. It is given by the matrix (M −1)⊤ (1)N (2). Proposition 5.15. If the Hopf algebra R is commutative, then the symmetric monoidal B category CorepFQA( R) is coclosed and cohom (B,ω), (C, ν) is the pair consisting of the object cohom(B, C) and the corepresentation (5.34) . 

Proof. Let ω : end(B) → OR be a corepresentation of the comonoid OR on a quadratic algebra B ∈ FQA given by a multiplicative matrix M. Due to the definition of internal e cohom we havee the adjunction e f Hom(cohom(B, C), B) =∼ Hom(C, B ◦ B) (5.35) e e 63 i ∗ in the category FQA. Let (wi) and (w ) be dual bases of the vector spaces W =(B1) and ∗ l a ∗ ∗ W = B1. Let(z ) and (w ) be bases of Z = C1 and W = B1. The graded homomorphisms h: cohom(B, C) → B and f : C → B ◦ B related toeach other by the bijection (5.35) have f e the form e e e k k a k k a i h: wi ⊗ z 7→ Kiaw , f : z 7→ Kiaw ⊗ w (5.36) Xa Xi,a e e k K for the same coefficients Kia ∈ . The corepresentation hω, νi is given by the matrix (M −1)⊤ (1)N (2), hence h is a morphism from corepresentation hω, νi to the corepresenta- k a −1 j k l tion ω iff a KiaMb = j,l(M )i Nl Kjb. The condition that the homomorphism f is a morphismP from ν to ω ◦ ωPhas the form Kk M aM i = N kKl . Due to the equivalence e f i,a ia b j l l jb of these conditions the adjunction (5.35)P gives a natural bijectionP between the morphisms of f the corresponding corepresentations.e

5.3 S∗-embedding of classical representations Now we ‘embed’ the classical Representation Theory to the Quantum Representation Theory. Let us begin with representations of finite-dimensional algebra on finite-dimensional vector spaces, i.e. with representations in the closed category (FVect, ⊗). We embed them to the category of representations in (FQAop, ◦) via the functor S∗ : FVect ֒→ FQAop. Then we generalise this embedding for the monoidal category (SLAffSch, ⊗), which contains (AffSch, ×) and (AlgSet, ×) as (not full) monoidal subcategories. 5.3.1. Representations of a finite-dimensional algebra as quantum representa- tions. According to p. 2.4.5 the lax monoidal functor S∗ : (FVect, ⊗) → (FQAop, ◦) induces the contravariant functor Mon(S∗): FAlg → Comon(FQA, ◦), where we used Mon(FVect, ⊗) = FAlg, Mon(FQAop, ◦) = Comon(FQA, ◦)op (see p. 2.4.3, p. 2.4.2). Let A = (V,µV , ηV ) ∈ FAlg. As it is described in p. 2.4.11 and p. 3.3.3 the actions and representations of the algebra A (objects of Lact(A) = RepFVect(A)) are embedded con- O O O ∗ travariantly to Lcoact( ) = CorepFQA( ), where = Mon(S )A ∈ Comon(FQA, ◦). We call it S∗-embedding of finite-dimensional representations of A into the category of quan- tum representations. Let us describe this embedding in details. First we need to examine how the algebra A is translated to the quantum level. Re- call that S∗ is a composition of the contravariant and covariant functors (−)∗ and S. Since the duality functor (−)∗ : (FVect, ⊗) → (FVectop, ⊗) is a strong monoidal, the dual vector space V ∗ ∈ FVect is equipped with a structure of coalgebra by the maps µ∗ ∗ V ∗ ∼ ∗ ∗ ∗ ∗ K ∆V ∗ : V −→ (V ⊗ V ) = V ⊗ V and εV ∗ = ηV : V → . Further, the colax monoidal ∗ ∗ functor S : (FVect, ⊗) → (FQA, ◦) translates the coalgebra A = (V , ∆V ∗ ,εV ∗ ) to the Sµ∗ comonoid O = Comon(S)A∗ =(SV ∗, ∆,ε) with the morphisms ∆: SV ∗ −−→V S∗(V ⊗ V ) ∼= ∗ ∗ φV ∗,V ∗ ∗ ∗ ∗ ∗ K K S(V ⊗ V ) −−−−→ SV ◦ SV and ε = SηV : SV → S = [u], where φ is the colax monoidal structure of S.

64 ′ ′ Let vv = µV (v ⊗ v ) be the multiplication in the algebra A and 1A = ηV (1) be its unity. The quadratic algebras SV ∗ and SV ∗ ⊗ SV ∗ ∼= S∗(V ⊕ V ) are identified with the algebras of the polynomials on the spaces V and V × V = V ⊕ V considered as affine spaces (their ∗ linear structure gives the grading). Then the structure of the bialgebra (SV , ∆SV ∗ ,εSV ∗ ) corresponding to the comonoid O has the form ∗ ∗ ∗ ′ ′ ∆SV ∗ : SV → SV ⊗ SV , (∆SV ∗ p)(v, v )= p(vv ), (5.37) ∗ εSV ∗ : SV → K, εSV ∗ (p)= p(1A), (5.38) ∗ ′ n where p ∈ SV is a polynomial function on V and v, v ∈ V . In terms of dual bases (vi)i=1 i n ∗ k n k i j k k and (v )i=1 of V and V we have ∆(v ) = i,j=1 cijv ⊗ v , ε(v ) = d for coefficients k k K n k n k cij,d ∈ , determined from the formulae vivj P= k=1 cijvk, 1A = k=1 d vk. In particular, this implies the condition (5.3). P P Consider a representation ρ: A → end(W ) of the algebra A on a finite-dimensional vector space W ∈ FVect. It corresponds to an action a: V ⊗ W → W . Due to the natural isomorphism (V ⊗ W )∗ ∼= V ∗ ⊗ W ∗ the functor (−)∗ : FVect → FVectop gives the coaction a∗ : W ∗ → V ∗⊗W ∗ of the coalgebra A∗ ∈ Comon(FVect) on the vector space W ∗ ∈ FVect. ∗ Sa φV ∗,W ∗ Then, by applying S we get the coaction δ : SW ∗ −−→ S(V ∗ ⊗ W ∗) −−−−→ SV ∗ ◦ SW ∗ of the comonoid O on the quadratic algebra SW ∗ ∈ FQA. ∗ O Thus, the lax monoidal functor S induces the functor RepFVect(A) → CorepFQA( ), which maps (W, ρ) to (SW ∗,ω), where ω : coend(SW ∗) → O is a corepresentation corre- sponding to the coaction δ. This corepresentation has the form

∗ ΦW,W ∗ Sρ ω : coend(SW ∗) −−−→ S end(W ) −−→ O, (5.39)

∗ ∗ ∗  where ΦW,Z : cohom(SW ,SZ ) → S hom(W, Z) is the graded homomorphism whose first order component is the isomorphism hom(W∗,Z∗) ∼= W ⊗ Z∗ =∼ (W ∗ ⊗ Z)∗ ∼= hom(W, Z)∗ natural in W ∈ FVect, Z ∈ Vect. Thus, the ‘S∗-embedding’ of represen- tations is the functor O op ∗ RepFVect(A) → CorepFQA( ) , (W, ρ) 7→ (T W ,ω). (5.40) m i m ∗ Let (wi)i=1 and (w )i=1 be dual bases of the vector spaces W and W respectively. The ∗ K latter basis gives the isomorphism SW = XAm ( ). Any representation ρ: A → end(W ) i i K has the form ρ(v)wj = ρj(v)wi where ρj : V → are linear functions such that m i ′ i k ′ i i ρj(vv )= ρk(v)ρj (v ), ρj(1A)= δj. (5.41) Xk=1 ∗ i i ∗ ∗ These functions form an m×m matrix M over SV with the entries Mj = ρj ∈ V =(SV )1. Due to the formulae (5.37), (5.38) the conditions (5.41) means exactly that M is a mul- tiplicative matrix. Since SV ∗ is a commutative algebra this is a usual Manin matrix. i Thus M = (ρj) is the multiplicative first order Am-Manin matrix defining the corepre- sentation (5.39). Then, by using Prop. 5.3 one can show that the functor (5.40) is fully faithful.

65 Remark 5.16. The actions a: A⊗W → W of an infinite-dimensional algebra A =(V,µV , ηV ) on a vector space W ∈ FVect can be also translated to the graded homomorphisms δ : SW ∗ → S(V ∗ ⊗ W ∗) → SV ∗ ◦ SW ∗ due to the isomorphism (V ⊗ W )∗ ∼= V ∗ ⊗ W ∗ natural in V ∈ Vect and W ∈ FVect. However, the multiplication µV : V ⊗ V → V does not give a comultiplication V ∗ → V ∗ ⊗ V ∗ in general, since the natural embedding V ∗ ⊗ W ∗ ֒→ (V ⊗ W )∗ is not bijective for some infinite-dimensional vector spaces, even if V = W . To consider representations of this algebra A as quantum representations we need to introduce a topology on A and to replace Vect with an appropriate category of topological K-modules W . This could extend Quantum Representation Theory for infinitely gener- ated quadratic algebras and allow to consider the classical (including infinite-dimensional) representations of an arbitrary algebra at the quantum level.

5.3.2. Representations of a semi-linear affine monoid scheme as quantum repre- sentations. In p. 4.3.5 we extended the functor S∗ to all the semi-linear affine schemes. It ∗ op has a structure of a lax monoidal functor S : (SLAffSch, ⊗) → (FQAsc , ◦), so it translates actions and representations of a monoid S ∈ Mon(SLAffSch, ⊗) to actions and represen- O ∗ S tations of the corresponding comonoid = Mon(S ) ∈ Comon(QAsc, ◦). Recall that any semi-linear affine monoid scheme S ∈ Mon(SLAffSch, ⊗) is given by a bialgebra (R,α,β) ∈ Comon(CommAlg, ⊗) and a vector space V ∈ FVect with linear ∗ ∗ ∗ ∗ maps t: V → R ⊗ R ⊗ V ⊗ V , εV ∗ : V → K such that the diagrams (4.49) commute ∗ (see p. 4.3.6). Hence the monoid S can be regarded as the coalgebra C = (R ⊗ V , ∆C,εC) (23) ∗ ∗ ∗ with a comultiplication ∆C = σ · (α, t): R ⊗ V → R ⊗ V ⊗ R ⊗ V and a counit ∗ ∗ εC = β ⊗ εV ∗ : R ⊗ V → K. The comonoid O = Mon(S )S is the semi-connected quadratic algebra A = R ⊗ SV ∗ with morphisms ∆: A→A◦A and ε: A → K[u] whose first order components are ∆C and εC. Let us write more detailed formulae for the case S ∈ Mon(SLAlgSet, ⊗), i.e. for a semi-linear algebraic monoid. The role of the affine scheme Spec R is played by an algebraic set X with the algebra of functions A(X)= R. It is equipped with a structure of algebraic ∗ ∗ monoid (X,µX , ηX ) such that α = µX and β = ηX ; the maps t and εV ∗ are given by i f : X × X × (V ⊗ V ) → V and 1V ∈ V (see p. 4.3.6). Let (wi) and (w ) be dual bases of vector spaces W and W ∗. A representation ρ: X × V → end(W ) and the corresponding m i action a: X ×(V ⊗W ) → W have the form ρ(x, v)wj = a(x, v⊗wj)= i=1 ρj(x, v)wi, where i ∗ ρj ∈ A(X)⊗V are semi-linear functions on X×V . The commutativityP of the diagrams (2.21) i ′ m i k ′ i i is equivalent to ρj xy, fx,y(v ⊗ v ) = k=1 ρk(x, v)ρj (y, v ) and ρj(e, 1V )= δj. This means, i i m i k i i in turn, that the matrix M = (ρj) isP multiplicative: ∆(ρj) = k=1 ρk ⊗ ρj , ε(ρj) = δj. In K i i particular, for X = {0} = Spec we have the case of p. 5.3.1 withP ρj(v)= ρj(0, v). For general R ∈ CommAlg denote X = Spec R. An action a: X×(V ⊗W ) → W is given by a linear map a∗ : W ∗ → R ⊗ V ∗ ⊗ W ∗ that should be a coaction of the coalgebra C on the ∗ ∗ i m i j i ∗ i space W . It has the form a (w )= j=1 ρj ⊗w for some ρj ∈ R⊗V such that M =(ρj) is a multiplicative matrix. The correspondingP representation ρ = θ−1(a): X × V → end(W ) is ∗ ∗ i i given by the coalgebra morphism ρ : end(W ) → C, w ⊗wj 7→ ρj. The lax monoidal functor ∗ op S : (SLAffSch, ⊗) → (QAsc , ◦) translates the representation ρ to the corepresentation

66 ω : cohom(SW ∗) → O. It also has the form (5.39). In terms of bases it is given by the i ∗ Am-Manin matrix M = (ρj). We obtain an extension of the S -embedding (5.40) for the representations of S. This is the fully faithful functor

S O op ∗ (RepFVect( ) ֒→ CorepFQA( ) , (W, ρ) 7→ (SW ,ω), (5.42

where O = Mon(S∗)(S). 5.3.3. Representations of an algebraic monoid or group as quantum representa- tions. Consider an algebraic monoid M =(X,µX , ηX ), this is an algebraic set X equipped with the multiplication µX : X×X → X and the unity ηX : {0}→ X which give a structure of monoid in AlgSet. We use the notations xy = µX (x, y), e = ηX (0). The (not full) embedding ,(AlgSet ֒→ SLAlgSet gives a semi-linear algebraic monoid SM := (X×K,µX ×idK, ηX ×idK so we can apply the results of p. 5.3.2. The function algebra R = A(X) is a commu- tative bialgebra with the comultiplication (∆Rf)(x, y) = f(xy) and counit εR(f) = f(e). Let ρ: M → end(W ) be a representation of the algebraic monoid M on W ∈ FVect. i i In the basis (wi) it has the form ρ(x)wj = i ρj(x)wi, where ρj ∈ A(X). The corre- sponding action (x, w) 7→ x.w = ρ(x)w satisfiesP (xy).w = x.(y.w) and e.w = w, which i is equivalent to the condition of multiplicativity of the matrix M = (ρj) is multiplica- ∗ tive. This matrix gives the corresponding corepresentation ω : coend(SW ) → OR, where ∗ OR = Mon(S )SM =(R ⊗ K[u], ∆,ε) is defined by (5.1). If M is an algebraic group (a group in AlgSet), then R = A(X) is a commutative Hopf −1 i algebra with the antipode ζR : R → R, (ζRf)(x) = f(x ), hence the matrix M = (ρj) −1 i is invertible. The inverse matrix M has entries ζR(ρj). In this case we have the dual D ∗ D −1 representation ρ : M → end(W ). It is defined by the formula ρ (x)ξ = ξ∗ ρ(x ) . In the i D i i −1 j −1 ⊤ dual basis (w ) it reads ρ (x)w = j ρj(x )w , so it is given by the matrix (M ) . At the KD quantum level this matrix gives theP Koszul dual corepresentation ω : coend(ΛW ) → OR, which we defined in p. 5.2.7, since (SW ∗)! = ΛW . Note that it does not coincide with the quantum representation corresponding to the classical representation ρD, because this quantum representation is a corepresentation on SW , not on ΛW , despite they are defined by the same matrix (M −1)⊤ (see the warning in p. 5.1.5). Finite monoids and groups are particular cases of algebraic monoids and groups respec- tively. If the field K is infinite then we can embed a finite monoid X ∈ Mon(Set, ×) into A1 N as a set of isolated points given by the equation i=1(x−ai) = 0, where N is the number K K of elements in X and a1,...,aN are arbitrary pair-wiseQ different elements of . If is finite then one can embed a finite monoid into Kn for a big enough n. In any case we obtain A(X)= KN . Thus, the description given for general algebraic monoids and groups is valid for the finite case. 5.3.4. Translation of binary operations with representations by the S∗-embedding. In the classical representation theory there are two important binary operations with repre- sentations: direct sum and tensor product. Consider the question: how the S∗-embedding translates them to the quantum level?

67 Since the functor S∗ has a strong monoidal structure (4.20), it translates direct sum of classical representations on the vector spaces W, Z ∈ FVect to a corepresentation on the quadratic algebra S∗(W ⊕ Z) = S(W ∗ ⊕ Z∗) = SW ∗ ⊗ SZ∗. The corresponding quantum representations are corepresentations on B = SW ∗ and C = SZ∗, their direct sum defined in p. 5.2.2 is a corepresentation on the same quadratic algebra B ⊗ C = SW ∗ ⊗ SZ∗. Let us describe them in details. Consider a semi-linear monoid S as in p. 5.3.2. Let ρ: S → end(W ) i ∗ and π : S → end(Z) be its representations. In terms of dual bases (wi), (w ) of W , W and k ∗ i ∗ i ∗ (zk), (z ) of Z, Z these representations are given by the elements ρj = ρ (w ⊗wj) ∈ R⊗V k ∗ k ∗ ∗ O and πl = π (z ⊗ zl) ∈ R ⊗ V . The corresponding corepresentations ω : coend(SW ) → and ν : coend(SZ∗) → O of the comonoid O = Mon(S∗)S are defined by the multiplicative i k S matrices M =(ρj) and N =(πl ). The direct sum ρ ⊕ π : → end(W ⊕ Z) is described via the multiplicative matrix L = M ⊕ N as

∗ i i i k k k (ρ ⊕ π) : w ⊗ wj 7→ ρj, w ⊗ zl 7→ 0, z ⊗ wj 7→ 0, z ⊗ zl 7→ πl

(cf. the definition of L = M ⊕ N in Prop. 5.7). Since the entries of M and N belongs to the commutative algebra A = R ⊗ SV ∗, the direct sum of ω, ν exists. This is a corepresentation ω ∔ ν : coend(SW ∗ ⊗ SZ∗) → O given by the multiplicative matrix L = M ⊕ N. On the other hand, the S∗-embedding of the direct sum ρ⊗π : S → end(W ⊕Z) is a corepresentation coend(SW ∗ ⊗ SZ∗) → O defined by the same matrix L, and hence it coincides with the direct sum ω∔ν. Thus, the S∗-embedding translates a direct sum of classical representations to the direct sum of the corresponding quantum representations. Since the algebra A is O commutative, we have the symmetric monoidal category (CorepFQA( ), ∔) and we obtain the following statement.

Proposition 5.17. The S∗-embedding (5.42) has a structure of symmetric strict monoidal S O op functor RepFVect( ), ⊕ → CorepFQA( ) , ∔ .   The tensor product of the representations ρ and π is defined if S =(X ×V,µX×V , ηX×V ) is equipped with a structure of a bimonoid (X×V,µX×V , ηX×V , ∆X×V ,εX×V ) in (SLAffSch, ⊗). The comultiplication ∆X×V and the counit εX×V give the monoidal product and unit object in Lact(S) as it is described in p. 2.4.10. In order to translate the bimonoid structure to the quantum level we need a strong monoidal embedding functor, but unfortunately the monoidal structure (4.21) of the functor S∗ is only lax monoidal, so we can not translate the comultiplication ∆X×V to the quantum level in general. Suppose that the monoid S ∈ Mon(SLAffSch, ⊗) is obtained by the category embedding ,(× ,AffSch ֒→ SLAffSch from an affine monoid scheme M =(X,µX , ηX ) ∈ Mon(AffSch where the morphisms µX and ηX are given by algebra homomorphisms α: R → R ⊗ R and β : R → K, so that S = SM = (X × K,µX × idK, ηX × idK). Note that an action X×(K⊗W ) → W of SM on W is the same as an action X×W → W of M on W , hence we can identify the representations of M on W with the representations of SM of W . In particular, representations ρ: M → end(W ) and π : M → end(Z) are considered as representations of S = SM. According to p. 2.5.3 the monoid M = (X,µX , ηX ) has the unique structure

68 of bimonoid (X,µX , ηX , ∆X ,εX ), where the diagonal morphism ∆X : X → X × X and the unique morphism εX : X →{0} correspond to the multiplication µR : R ⊗ R → R and the unity ηR : K → R. In particular, ∆X defines the tensor product ρ ⊗ π : M → end(W ⊗ Z). In the case X ∈ AlgSet the diagonal morphism has the form ∆X (x)=(x, x), so we obtain (ρ ⊗ π)(x)= ρ(x) ⊗ π(x). For general X ∈ AffSch the representation ρ ⊗ π is a morphism corresponding to the homomorphism (ρ⊗π)∗ : S∗ coend(W ⊗Z) → R, w⊗z 7→ ρ∗(w)π∗(z), where w ∈ end(W )∗, z ∈ end(Z)∗, w ⊗ z ∈end(W )∗ ⊗ end (Z)∗ = end(W ⊗ Z)∗ ⊂ ∗ ∗ i kb b i bk b S coend(W ⊗ Z) . In terms of bases one yields (ρ ⊗ π) (w ⊗ wj ⊗ z ⊗ zl)= ρjπl . b b b b K The Bimon-functor of the embedding Alg ֒→ QAsc, R 7→ R ⊗ [u], translates the bimonoid (R,µR, ηR, ∆R = α,εR = β) to the bimonoid BR = (R ◦ K[u], µ, η, ∆,ε) defined by the formulae (5.1), (5.26). The commutative multiplication µ = µR ⊗ idK[u] gives the tensor product of corepresentations, which always exists in this case. Thereby, we get the ∗ ∗ tensor product ω ◦˙ ν : coend(SW ◦ SZ ) → BR. It corresponds to the multiplicative usual (1) (2) ik i k Manin matrix P = M ⊗˙ N = M N with the entries Pjl = ρjπl . On the other hand, ∗ ∗ ρ ⊗ π is translated to the corepresentation λ: coend S(W ⊗ Z ) → BR given by the same matrix P = M ⊗˙ N. Due to Corollary 5.4 this means that the graded homomorphism ∗ ∗ ∗ ∗ φW,Z : S(W ⊗ Z ) → SW ◦ SZ is a morphism between the corepresentations λ and ω ◦˙ ν. We can formulate it in the following form.

∗ Proposition 5.18. In the case S = SM the S -embedding (5.42) has a structure of symmetric S B op lax monoidal functor RepFVect( M), ⊗ → CorepFQA( R) , ◦˙ .   Remark 5.19. The S∗-embedding of ρ ⊕ π differs from the coproduct of ω and ν in general, since S∗(W ⊕ Z) =6 S∗(W ) ∐ S∗(Z). 5.3.5. Dualisation of a finite-dimensional bialgebra. For representations of a finite- dimensional algebra equipped with a bialgebra structure there is another way to translate them to the quantum level such that the representation spaces are translated via the same functor S∗. Let B = (V,µV , ηV , ∆V ,εV ) ∈ Bimon(FVect, ⊗) be a finite-dimensional bialgebra. This is an algebra A = (V,µV , ηV ) ∈ FAlg with a comultiplication ∆A = ∆V : V → V ⊗ V and a counit εA = εV : V → K. Application of the contravariant strong monoidal functor (−)∗ : (FVect, ⊗) → (FVect, ⊗) gives the dual bialgebra

∗ ∗ ∗ ∗ ∗ ∗ ∗ B = Bimon (−) B =(V ,µV ∗ = ∆V , ηV ∗ = εV , ∆V ∗ = µV ,εV ∗ = ηV ) (5.43)  (see e.g. [Kass, § 3.2, Ex. 1]). Consider the latter one as a comonoid in (Alg, ⊗). The O ∗ K (◦ ,embedding Alg ֒→ QAsc gives the comonoid B∗ =(B ⊗ [u], ∆,ε) ∈ Comon(FQAsc ∗ ∗ defined by (5.1). It differs from the comonoid O = Mon(S )A =(SV , ∆SV ∗ ,εSV ∗ ) used in p. 5.3.1, but sometimes we can translate a representation of A to a corepresentations of OB∗ . Suppose first that the bialgebra B is cocommutative. Then the dual bialgebra (5.43) is commutative, so by virtue of Prop. 2.2 the comonoid OB∗ has the structure of commutative B ∗ K bimonoid B∗ = (B ⊗ [u], µ, η, ∆,ε) ∈ Bimon(FQAsc, ◦) defined by (5.26). In this case any representation ρ: A → end(W ) on a vector space W ∈ FVect is translated to a

69 ∗ corepresentations of BB∗ on SW ∈ FQA. Indeed, due to the commutativity of µV ∗ the ∗ ∗ ∼ ∗ ∗ linear map (SV )1 = V −→ (BB∗ )1 induces the graded homomorphism SV → B ⊗ K[u]. In this way we obtain a comonoid morphism O → BB∗ . By composing it with (5.39) we get ′ ∗ the corepresentation ω of the bimonoid BB∗ on the quadratic algebra SW :

∗ ′ ∗ ΦW,W ∗ Sρ ω : coend(SW ) −−−→ S end(W ) −−→ O → BB∗ . (5.44)

m i ∗  i Let (wi)i=1 be a basis of W and ρj ∈ V be such that ρ(v)wj = i ρj(v)wi, then ρ is ′ ∗ translated to the corepresentation ω : coend(SW ) → BB∗ given by theP usual Manin matrix i M =(ρj). In this way we obtain the fully faithful functor B op ∗ ′ (RepFVect(A) ֒→ CorepFQA( B∗ ) , (W, ρ) 7→ (SW ,ω ). (5.45 Let π : A → end(Z) be a representation of A on a vector space Z ∈ FVect. The ′ ∗ functor (5.45) translates π to a corepresentation ν : coend(SZ ) → BB∗ . The direct sum ′ ′ ∗ ∗ ρ ⊕ π : A → end(W ⊕ Z) is translated to ω ∔ ν : coend(SW ⊗ SZ ) → BB∗ , so we can B op regard (5.45) as a strict monoidal functor RepFVect(A), ⊕ → CorepFVect( B∗ ) , ∔ . The translation of the tensor product by the functor (5.45) gives a lax monoidal functor B op RepFVect(A), ⊗ → CorepFVect( B∗ ) , ◦˙ . If B is a Hopf algebra with an antipode ∗ ∗ ∗ ζV : V → V , then (5.43) is also a Hopf algebra, its antipode is ζV : V → V (see [Kass, § 3.3, Prop. 3.3.3]), so BB∗ is a Hopf monoid and, hence, one can define the Koszul dual ′ KD corepresentation (ω ) : coend(ΛW ) → BB∗ . Let M =(X,µX , ηX ) ∈ Mon(Set, ×) be a finite monoid. The algebra B = K[M] defined in p. 3.2.2 is a cocommutative bialgebra with the coalgebra structure

∆V αxx = αx(x ⊗ x), εV αxx = αx. (5.46)  xX∈X  xX∈X  xX∈X  xX∈X It is a Hopf algebra iff M is a group. The dual bialgebra B∗ coincides with the bialgebra R = A(X) constructed in p. 5.3.3. A representation M → end(W ) is uniquely extended to a representation K[M] → end(W ). If we translate these representations to the corep- ∗ ′ ∗ resentations ω : cohom(SW ) → OR and ω : cohom(SW ) → OB∗ respectively, then the results coincide: ω = ω′, so in this case the functor (5.45) is just another description of the representation S∗-embedding obtained in p. 5.3.3. For a non-cocommutative bialgebra B ∈ Bimon(FVect, ⊗) we can not translate an arbi- trary representation ρ: A → end(W ) to a corepresentation of OB∗ , but if the corresponding i ∗ matrix M =(ρj) is an Am-Manin matrix as a matrix over B , then we have a corepresenta- ′ ∗ tion ω : coend(SW ) → OB∗ given by this matrix, since this matrix is multiplicative with respect to ∆V ∗ and εV ∗ anyway.

5.4 T ∗-embedding of classical representations An alternative way to translate a classical representation to the quantum level is to use the functor T ∗ : FVect → FQAop defined in p. 4.1.10 instead of S∗.

70 5.4.1. T ∗-embedding of a finite-dimensional algebra and its representations. Let ∗ op A = (V,µV , ηV ) ∈ FAlg. The strong monoidal functor T : (FVect, ⊗) → (FQA , ◦) induces the functor Mon(T ∗), which maps the algebra A ∈ Mon(FVect, ⊗) to the comonoid O ∗ op ∗ ∗ ∗ ∗ = Mon(T )A = Comon(T )(V ,µV , ηV )=(TV , ∆,ε) in the category (FQA, ◦), where T µ∗ ∆: TV ∗ −−→V T ∗(V ⊗ V ) ∼= T (V ∗ ⊗ V ∗) ∼= TV ∗ ◦ TV ∗, ε = T η∗ : TV ∗ → T K = K[u]. V ∗ Due to Prop. 4.5 we obtain cohom(T W ∗,TZ∗)= T hom(W ∗,Z∗) ∼= T hom(W, Z) , so the morphisms (3.26) for the monoidal functor F = T ∗ : (FVect, ⊗)→ (FQA op, ◦) is the ∗ ∗ ∗ isomorphisms ΦW,Z : cohom(T W ,TZ ) −→∼ T hom(W, Z) given by the same identifica- tion hom(W ∗,Z∗) ∼= hom(W, Z)∗ as for S∗.  Let ρ: A → end(W ) be a representation of the algebra A on W ∈ FVect. In a basis m i ∗ (wi)i=1 of W it is given by the linear functions ρj ∈ V satisfying (5.41). By using the isomorphism ΦW,W we obtain the corresponding quantum representation

∗ T ρ∗ ω : coend(T W ∗) ∼= T end(W ) −−→ O. (5.47)  O ∗ ∗ K This is a corepresentation of = (TV , ∆,ε) on the quadratic algebra T W = X0m ( ), m m m where 0m ∈ end(K ⊗ K ) is the zero idempotent and the identification W = K is fixed by the basis (wi). Recall that any m × m matrix is a 0m-Manin matrix. The corepre- i sentation (5.47) is defined by the multiplicative first order matrix M = (ρj). Thus, the T ∗-embedding of the representations is the fully faithful functor

O op ∗ (RepFVect(A) ֒→ CorepFQA( ) , (W, ρ) 7→ (T W ,ω). (5.48

5.4.2. T ∗-embedding of a semi-linear affine monoid scheme and its representa- ∗ ∗ op tions. The functor T can be extended to the functor T : SLAffSch → QAsc by the formula T ∗ : (Spec R) × V 7→ R ⊗ TV ∗ with the obvious mapping of the morphisms. This is op a strong monoidal functor (SLAffSch, ⊗) → (QAsc , ◦). Let S ∈ Mon(SLAffSch, ⊗) be a semi-linear affine monoid scheme as in p. 5.3.2. Then the comonoid O = Mon(T ∗)S is the semi-connected quadratic algebra A = R ⊗ T ∗V with the graded homomorphisms ∆: A→A◦A and ε: A → K[u] such that their first order ∗ ∗ ∗ ∗ components ∆1 and ε1 coincide with ∆C : R ⊗ V → R ⊗ V ⊗ R ⊗ V and εC : R ⊗ V → K, which were described in p. 5.3.2. Let ρ: S → end(W ) be a representation of S on a space W ∈ FVect with a basis m ∗ ∗ i i (wi)i=1. It is given by the coalgebra morphism ρ : end(W ) → C, w ⊗ wj → ρj, where ∗ i C = (R ⊗ V , ∆C,εC) and M = (ρj) is a multiplicative first order matrix over the algebra ∗ ∗ op R ⊗ TV with respect to ∆ and ε. Since the functor T : (SLAffSch, ⊗) → (QAsc , ◦) is strong monoidal, it translates ρ to the corepresentation ω : coend(T W ∗) → O corresponding i ∗ to the 0m-Manin matrix M = (ρj). This extends T -embedding functor (5.48) to the fully faithful functor

S O op ∗ (RepFVect( ) ֒→ CorepFQA( ) , (W, ρ) 7→ (T W ,ω), (5.49

where O = Mon(T ∗)S.

71 5.4.3. Translation of binary operations under the T ∗-embedding. Let represen- S S i k ∗ tations ρ: → end(W ), π : → end(Z) be given by the elements ρj, πl ∈ R ⊗ V in ∗ bases (wi) and (zk). Denote the corresponding corepresentations coend(T W ) → O and ∗ O i k coend(TZ ) → given by the matrices M = (ρj) and N = (πl ) by ω and ν. The func- tor T ∗ : FVect → FQAop has a structure of strong monoidal functor (4.28). By using Corollary 5.4 again we see that the T ∗-embedding translates the direct sum representation ρ ⊕ π : S → end(W ⊕ Z) to the coproduct ω ⊔˙ ν : coend(T W ∗ ∐ TZ∗) → O given by the matrix L = M ⊕ N. This means that T ∗-embedding (5.49) preserves the finite products. In other words, the functor (5.49) has a structure of symmetric strong monoidal functor S O op RepFVect( ), ⊕ → CorepFQA( ) , ⊔˙ . By taking into account the lax monoidal structure (4.27) of the category embedding T ∗ : FVect ֒→ FQAop we see that the T ∗-embedding gives a symmetric lax monoidal functor S O op RepFVect( ), ⊕ → CorepFQA( ) , ∔ . ,(Since T ∗ : FVect ֒→ FQAop has a structure of symmetric strong monoidal functor (4.29 we can apply the results of p. 3.3.5. In this way we obtain a strong monoidal functor S O op ∗ RepFVect( ), ⊗ → CorepFQA( ) , ◦˙ given by the T -embedding (5.49).   5.4.4. Representations of a finite-dimensional bialgebra. Let B =(V,µV , ηV , ∆V ,εV ) be a bialgebra as in p. 5.3.5. Then we can translate any representation of the algebra ∗ A =(V,µV , ηV ) on W ∈ FVect to a quantum representation of the comonoid OB∗ on T W . We obtain the fully faithful functor

B op ∗ ′ (RepFVect(A) ֒→ CorepFQA( B∗ ) , (W, ρ) 7→ (T W ,ω ), (5.50

i i ∗ ′ ∗ O where ρ: A → end(W ), ρ(v)wj = i ρj(v)wi, ρj ∈ V and ω : coend(T W ) → B∗ is the i corepresentation given by the multiplicativeP matrix M =(ρj). Note that in this case we do not need to require the condition of cocommutativity of B nor any Manin condition on M. ′ ∗ Let π : A → end(Z) be representation and ν : coend(TZ ) → OB∗ be the corresponding corepresentation. The embedding (5.50) translates direct sum ρ ⊕ π : A → end(W ⊕ Z) to ′ ′ ∗ ∗ the coproduct ω ⊔˙ ν : coend(T W ∐ TZ ) → OB∗ . It translates the tensor product of representations ρ ⊗ π : A → end(W ⊗ Z) defined by means of ∆V to the tensor product ′ ′ ∗ ∗ ω ◦˙ ν : coend(T W ◦ TZ ) → OB∗ .

6 Examples

6.1 Corepresentations of Mq(m)

The relationship of some Manin matrices with the Lax operators of Uq(gln) type was de- scribed in details in [S]. It can be interpreted in terms of Quantum Representation Theory.

6.1.1. Matrix algebra. Consider the algebra A = Matm(K). It consists of m × m matrices over K. The multiplication in A is the usual matrix multiplication and the unity is the identity matrix. The quadratic algebra M(m) = M(m, K) = SA∗ is the algebra of regular

72 K i i functions on Matm( ) (as on an algebraic set). It is generated by the functions aj : M 7→ Mj , K 2 i ∗ M ∈ Matm( ). Thus M(m) is the graded algebra generated by m order 1 elements aj ∈ A . The multiplication and unity of Matm(K) gives the structure of comonoid on M(m). We obtain the comonoid O = M(m)=(M(m), ∆,ε) ∈ Comon(FQA, ◦), where the graded homomorphisms ∆: M(m) → M(m) ◦ M(m), ε: M(m) → K have the form

m i i k i i ∆(aj)= ak ⊗ aj , ε(aj)= δj. (6.1) Xk=1 The algebra M(m) with this structure can be considered as a bialgebra, but this is not a Hopf algebra. m The matrices from Matm(K) act on elements of W = K as on column vectors. The m corresponding representation ρ: Matm(K) → end(K ) is the isomorphism identifying op- erators on Km with the matrices. By the S∗-embedding described in p. 5.3.1 we get the K i corepresentation ω : coend XAm ( ) → M(m) given by the matrix M =(aj). m The right action of Matm(K) on the space K considered as the space of row-vectors ⊤ op m gives the representation ρ : Matm(K) → end(K ). The corresponding corepresentation ⊤ K cop ⊤ ω : coend XAm ( ) → M(m) is defined by the transposed matrix M . Note that the i k k i commutation relations ajal = al aj, which define the quadratic algebra M(m), are equivalent ⊤ to the requirement that both M and M are Am-Manin matrices (see e.g. [S, Prop. 3.1]). 6.1.2. A q-deformation of Km. Consider the K-plane K2. At the quantum level it is K K K described by the quadratic algebra XA2 ( ) = [x, y]. Let q ∈ \{0}. Replacement of the commutation relation yx = xy by yx = qxy gives a q-deformation of K[x, y] called quantum plane. It can be interpreted as a q-deformation of the vector space K2 in the category of quantum linear spaces (where the vector spaces considered as quadratic algebras via the .(embedding S∗ : FVect ֒→ FQAop For a general m a quantum q-deformation of Km is constructed as the quadratic algebra generated by x1,...,xm with the commutation relations xjxi = qxixj, i < j. This q K q Km Km 4 is the algebra XAm ( ) for the idempotent Am ∈ end( ⊗ ) defined by the formulae 1 Aq = (1 − P q ), (P q )ij = q δiδj , (6.2) m 2 m m kl ji l k

−1 qij = qji = q for i < j, qii = 1.

6.1.3. A q-deformation of the matrix algebra. Let Mq(m)= Mq(m, K) be the bialgebra i generated by aj, i, j =1,...,m, with the commutation relations

aj ai = qai aj , aiaj = aj ai, k k k k l k k l (6.3) i i i i i j j i −1 j i alak = qakal, akal − al ak =(q − q)akal,

4More generally, one can construct a multi-parametric deformation [Man89] by imposing the conditions −1 qij = qji , qii = 1 only (see [S, § 3.3]).

73 where i < j, k

∗ f ⊤ cop K q K T Matm( ) / coend XAm ( ) (6.5)

 ⊤  f ωq

 ωq  q K coend XAm ( ) / Mq(m)  in the category GrAlg and, consequently, in FQA (see [MacLane, § 3.3] for a definition ⊤ of pushout). Moreover, the comonoid Mq(m) is a pushout of f and f in the category Comon(FQA, ◦).

6.2 An extension of the Yangian Y (glm) and its corepresentations In [CF] the authors presented an example of a usual Manin matrix defined as a product of the Lax operator for the glm Yangian with a shift operator. We interpret this extension as K a corepresentation of some extension of this Yangian on the quadratic algebra XAm ( ).

74 6.2.1. Manin matrix from the Y (glm) Lax operator. Denote by Pm the operator acting m m m m by permutation of the tensor factors in K ⊗ K , that is Pm = σKm,Km ∈ end(K ⊗ K ). 1 − P It has entries (P )ij = δiδj and is related with the anti-symmetrizer as A = m . m kl l k m 2 Consider the rational R-matrix R(z) = z − Pm. The Yangian Y (glm) is the algebra K r Z generated (over ) by tij, i, j =1,...,m, r ∈ >1, with the relations

R(z − v)T (1)(z)T (2)(w)= T (2)(w)T (1)(z)R(z − w), (6.6)

−1 i i r −r where T (z) is the m×m matrix over Y (glm)[[z ]] with the entries T (z)j = δj + r>1 tijz . The Yangian is equipped with a structure of Hopf algebra: P m i i k i i −1 ∆ T (z)j = T (z)k ⊗ T (z)j , ε T (z)j = δj, ζ T (z) = T (z) . (6.7)  Xk=1  

− ∂ − ∂ Consider the shift operator e ∂z . The relation (6.6) implies that the matrix T (z)e ∂z is −1 − ∂ an Am-Manin matrix over the algebra Y (glm)[[z ]][e ∂z ] (see details in [CF], [S]). This fact is equivalent to the relation

AmT (z)T (z − 1)(1 − Am)=0. (6.8)

6.2.2. Extended Yangian and its corepresentation. Let us add one more generator τ to the Yangian and postulate the commutation relation

τT (z)= T (z − 1)τ (6.9)

r r r−1 k (via generators it can be written in the form τtij = k=1 k−1 tijτ). We obtain a bialgebra, where the comultiplication and counit of the newP generator  are defined by the formulae ∆(τ) = τ ⊗ τ, ε(τ) = 1. This is an extension of Y (glm) as a bialgebra, denote it by K Y (glm)[τ]. It has the form of the tensor product Y (glm) ⊗ [τ] as a vector space, but not r as an algebra, since τ does not commute with the generators tij. The relations (6.8), (6.9) imply that the matrix M = T (z)τ is an Am-Manin matrix over −1 the algebra R = Y (glm)[τ]((z )), where we extend the basic field to the field of formal Laurent series

N −1 k K((z )) = αkz | N ∈ Z,αk ∈ K . (6.10) n kX=−∞ o

The Am-Manin matrix M = T (z)τ is multiplicative with respect to the comultiplication −1 ∆R : R → R ⊗K((z−1)) R and counit εR : R → K((z )) obtained by the field extension, so it gives a corepresentation

K −1 O ω : end XAm ((z )) → R (6.11)   75 −1 of the comonoid OR = R◦K((z−1))K((z ))[u], ∆R⊗K((z−1))idK((z−1))[u],εR⊗K((z−1))idK((z−1))[u] , where the operation ◦K((z−1)) is defined over the extended field (6.10), we have the bialgebra −1 isomorphism R◦K((z−1)) K((z )))[u]= R◦K K[u] over K. The quantum representation ω has the classical representation space Km (with respect to the S∗-embedding), but the quantum algebra we represent is not classical. −1 In this construction we consider the elements of R = Y (glm)[τ]((z )) as elements of order 0, the graded algebra is R[u]. Alternatively one can set deg τ = 1. Then the bialgebra −1 Y (glm)[τ]((z )) itself becomes a comonoid in (GrAlg, ◦), and entries of M = T (z)τ become −1 elements of order 1. This is an affinely generated quadratic algebra over Y (glm)((z )), but it is not semi-connected.

−1 ±1 Remark 6.2. The localisation of Y (glm)[τ] by τ gives us a Hopf algebra Y (glm)[τ ] with the antipode extended as ζ(τ)= τ −1. However, the formula deg τ = 1 and its consequence −1 Z N ±1 deg τ = −1 define a -grading rather than an 0-grading on the algebra Y (glm)[τ ].

6.3 Corepresentations of an extended Y (som)

As we see in [S, § 7.1], the Lax operator of the Y (som) and Y (spm) also give Manin matrices for some idempotents, so we can construct quantum representations for them by analogy with Subsection 6.2. For simplicity we consider the orthogonal case only. The symplectic version is completely analogues. Here we suppose that char K = 0.

6.3.1. The algebras X(som) and Y (som). Consider again the collection of generators r Z tij, where i, j = 1,...,m, r ∈ >1. Let T (u) be the m × m matrix defined by the same formula as in p. 6.2.1. Denote by X(som) the algebra with these generators and commutation P Q relations (6.6) but for another R-matrix R(z)= R (z)=1− m + m , where Q som z z − m/2+1 m Km Km ij i+j m+1 is the operator on ⊗ with the entries (Qm)kl = δm+1δk+l . The Yangian Y (som) can m i m j i+j be defined as a quotient of X(som) by the relations k=1 T (z)kT (z + 2 − 1)m+1−k = δm+1. Both X(som) and Y (som) are Hopf algebras with theP comultiplication, counit and antipode of the same form (6.7) (see details in [AMR, § 2, § 3, Cor. 3.2], [Molev, § 11.1] and [S, § 7.1]).5 K 6.3.2. An algebra XBm ( ) and a Bm-Manin matrix. Consider the idempotent 1 − P Q Q B = m + m = A + m ∈ end(Km ⊗ Km) (6.12) m 2 m m m K (introduced in [S, § 7.1]). It gives the commutative quadratic algebra XBm ( ) isomorphic K K 1 m to the quotient of XAm ( )= [x ,...,x ] by the quadratic relations

m xixm+1−i =0. (6.13) Xi=1

5 The algebra X(som) is usually called ‘extended Yangian’, but we use this term for another algebra.

76 − ∂ It is proved in [S, § 7.1] that the matrix T (z)e ∂z is a Bm-Manin matrix over the algebra −1 − ∂ Y (som)[[z ]][e ∂z ]. This is equivalent to the relation

BmT (z)T (z − 1)(1 − Bm)=0. (6.14) K 6.3.3. A corepresentation on XBm ( ). Let us extend the Yangian Y (som) by a new generator τ and impose the relations (6.9). This is a bialgebra over K (the comultiplication K and counit has the same form as for the glm case). It is isomorphic to Y (som) ⊗ [τ] as a vector space, so we analogously denote it by Y (som)[τ]. The field extension gives the −1 bialgebra R = Y (som)[τ]((z )) over (6.10). Due to (6.14) the matrix M = T (z)τ is a Bm-Manin matrix over R. It is multiplicative with respect to the K((z−1))-linear comultiplication and counit of the bialgebra R, so it defines a corepresentation

K −1 O ω : end XBm ((z )) → R (6.15)   −1 of the comonoid OR =(R ◦ K((z )))[u], ∆R ⊗K((z−1)) idK((z−1))[u],εR ⊗K((z−1)) idK((z−1))[u]). Remark 6.3. Since the relation (6.14) follows from (6.6) only the matrix T (z)τ can be −1 considered as a Bm-Manin matrix over the bialgebra X(som)[τ]((z )). By using the locali- ±1 −1 sation we can regard it as a Bm-Manin matrix over the Hopf algebra Y (som)[τ ]((z )) or ±1 −1 X(som)[τ ]((z )).

Conclusion and further directions

Quantum linear spaces and quantum representations. The Manin product ‘◦’ allows us to construct Quantum Representation Theory on quantum linear spaces A ∈ FQAop. Namely, we applied the general approach described in Section 3 to the symmetric monoidal op op category C =(QAsc , ◦) and its subcategory P = FQA . By using the category embedding S∗ : FVect ֒→ FQAop we can consider the finite-dimensional vector spaces and as quantum ∗ op op -linear spaces. Its extension S : SLAffSch ֒→ QAsc ֒→ GrAlg helps us to interpret rep resentations of finite-dimensional algebras and algebraic groups as quantum representations. The binary operations with vector spaces and representations agree with their quantum versions (modulo some natural transformations φ). Semi-linear spaces. Introduction of semi-linear spaces gives the unified theory of finite- dimensional representations of finite-dimensional algebras and algebraic groups: one needs to apply the approach of Section 3 to C = (SLAffSch, ⊗) and P = FVect. We have the same situation at the quantum level, where the role of semi-linear spaces is played by op op semi-connected quadratic algebras: C =(QAsc , ◦), P = FQA . Multiplicative Manin matrices. It is shown that the quantum representations of a quantum monoid on XB(K) is in one-to-one correspondence with the multiplicative B-Manin matrices over the corresponding bialgebra. In the case of more general quantum algebra an

77 additional condition on the Manin matrix is imposed: its entries should belong to the first order graded component. (A, B)-Manin matrices. The multiplicative B-Manin matrices correspond to the comonoid morphisms from the graded algebra coend XB(K) equipped with the comultiplication and counit (4.59). A natural question arises: how to generalise the theory described here in order to include more general (A, B)-Manin matrices and cocomposition (4.58)? In this case we probably need to consider a family of Manin matrices compatible with the cocompositions and counits in some way. We plan to construct such a generalisation in future works. Another subcategories of graded algebras. The quantum linear spaces A ∈ FQAop can be generalised by considering more wide subcategory of GrAlg. The main condition we should keep is the connectivity: A0 = K. For example, one can take all the connected affinely generated graded algebras – this will allow to connect the (generalised) quantum linear spaces with a quantum analogue of the projective spaces. We hope to investigate this case in the future. Super-case. If one considers a vector space V ∈ FVect as a purely odd super-vector space, the embedding Λ∗ : FVect → FQAop should be applied (see p. 4.1.7). By combining the even and odd embeddings one can embed an arbitrary finite-dimensional super-vector ∗ ∗ space V = V0¯ ⊕ V1¯ into the category of quadratic super-algebras as S (V0¯) ⊗ Λ (V1¯). The general approach of Section 3 can be applied to the super-case as well. The super-version of Quantum Representation Theory is also a subject for future publications. S∗-embedding vs T ∗-embedding. Both functors S∗ and T ∗ lift the vector spaces and representations to the quantum level. These two ways to embed the vector spaces have advantages and disadvantages with respect to each other. The S∗-embedding has more intuitive interpretation and leads to the usual Manin matrices, which have a lot of ap- plications. It has a strong monoidal structure (FVect, ⊕) → (FQAop, ⊗), but it is not strong monoidal as a monoidal functor (FVect, ⊗) → (FQAop, ◦). The T ∗-embedding has a structure of strong monoidal functor (FVect, ⊗) → (FQAop, ◦). As a monoidal functor (FVect, ⊕) → (FQAop, ⊗) it is not strong monoidal, however it has a strong monoidal structure (FVect, ⊕) → (FQAop, ∐), so it is more natural to consider coproduct ∐ as an analogue of direct sum of vector spaces for the quantum case, when we use the T ∗-embedding (note that ∐ is exactly the categorical product in FQAop). Tensor product. If we translate a finite-dimensional algebra or an algebraic monoid/group to the quantum level by either S∗- or T ∗-embedding, then the tensor product of quantum representations of the obtained quantum algebra always exists due to the ‘commutativity’ of the classical algebra. For a more general comonoid O ∈ Comon(GrAlg, ◦) this existence is not guaranteed even if O ∈ Comon(FQA, ◦) or O = OR for a quantum monoid/group R = (R, ∆R,εR) ∈ Mon(Alg, ⊗). It means that the category of quantum representations loses the tensor product under a quantum deformation. We conjecture that the tensor product can be restored in some important cases by means of deformation of the symmetric

78 structure. Infinite-dimensional case. By considering the purely algebraic theory of quantum rep- resentations we are restricted by the finite-dimensional case by three reasons. First, the functor (−)∗ : Vect → Vectop is not full (it is only faithful). Second, it is not lax monoidal with respect to ⊗ (only colax monoidal). Finally, the monoidal category (QA, ◦) is not coclosed (the proof of Prop. 4.5 essentially uses the finiteness of the basis (wi)). It seems that a generalisation of the results given here to the infinite-dimensional case is possible, if we introduce some kind of topology on vector spaces and algebras.

References

[AMR] Arnaudon, D.; Molev, A.; Ragoucy, E.: On the R-matrix realization of Yan- gians and their representations, Annales Henri Poincar´e 7 (2006), 1269–1325; arXiv:math/0511481 [math.QA].

[Bor2] Borceux, F.: Handbook of Categorical Algebra: Volume 2, Categories and Structures, Encyclopedia of Mathematics and its Applications, Cambridge University Press, (1995).

[CF] Chervov, A.; Falqui, G.: Manin matrices and Talalaev’s formula, J. Phys. A: Math. Theor. 41 (2008), No. 19, 194006 (28pp.); arXiv:0711.2236 [math.QA]

[CM] Chervov, A.; Molev, A.: On higher order Sugawara operators, IMRN 2009, no.9, (2009), 1612–1635; arXiv:0808.1947 [math.RT].

[CFR] Chervov, A.; Falqui, G.; Rubtsov, V.: Algebraic properties of Manin matrices I, Adv. in Appl. Math. 43 (2009), no. 3, 239–315; arXiv:0901.0235 [math.QA].

[CFRS] Chervov, A.; Falqui, G.; Rubtsov, V.; Silantyev, A.: Algebraic properties of Manin matrices II: q-analogues and integrable systems, Adv. in Appl. Math. 60 (2014), 25–89; arXiv:1210.3529 [math.QA].

[Del] Deligne at al: Quantum fields and strings: A course for mathematicians, Vol. 1: Notes on (by Deligne P. and Morgan J. W.), American Mathematical Society, Providence, (1999).

[Dr] Drinfel’d, V. G.: Quantum groups, Zap. Nauchn. Sem. LOMI 155, ”Nauka”, Leningrad. Otdel., Leningrad, (1986), 18–49 (in Russian); translation in J. Soviet Math. 41:2 (1988), 898–915.

[EH] Eckmann, B.; Hilton, P.: Structure maps in group theory, Fund. Math. 50 (1961), 207–221.

[Ful] Fulton, W.: Algebraic Curves: An Introduction to Algebraic Geometry, Addison-Wesley Publishing Company, (1989); first published in 1969; revised in 2008, pdf.

79 [GLZ] Garoufalidis S.; Le, Thang TQ.; Zeilberger D.: The quantum MacMahon Mas- ter Theorem, Proc. Natl. Acad. of Sci. (PNAS) 103 (2006), No. 38, 13928-13931; arXiv:math.QA/0303319.

[Har] Hartshorne, R.: Algebraic Geometry, Graduate Texts in Mathematics 52, Springer Verlag, (1977).

[IO] Isaev, A. P.; Ogievetsky, O. V.: Half-quantum , Nankai Series in Pure, Applied Mathematics and Theoretical Physics Symmetries and Groups in Contemporary Physics (2013), 479–486; arXiv:1303.3991 [math.QA].

[Kass] Kassel, C.: Quantum groups, Graduate Texts in Mathematics, 155, Springer-Verlag, New York (1995).

[Lang] Lang, S.: Differential and Riemannian manifolds, Berlin, New York: Springer-Verlag, (1995).

[MacLane] MacLane, S.: Categories for the Working Mathematician (second edition), Grad- uate Texts in Mathematics 5, Springer, (1998).

[Man70] Manin, Yu. I.: Lekcii po algebraicheskoj geometrii 1966-1968 (in Russian), MSU, (1970); translation: Introduction to the Theory of Schemes, Springer International Pub- lishing, (2018).

[Man87] Manin, Yu. I.: Some remarks on Koszul algebras and quantum groups, Ann. de l’Inst. Fourier 37, no. 4 (1987), pp. 191–205.

[Man88] Manin, Yu. I.: Quantum Groups and Non-Commutative Geometry, University of Montreal, Centre de Recherches Math´ematiques, Montreal, QC, (1988), 91 pp.

[Man89] Manin, Yu. I.: Multiparametric quantum deformation of the general linear super- group, Comm. Math. Phys. 123 (1989), 163–175.

[Man91] Manin, Yu. I.: Topics in Non Commutative Geometry, M. B. Porter Lectures, Princeton University Press, (1991), 164 pp.

[Man92] Manin, Yu. I.: Notes on Quantum Groups and Quantum de Rham Complexes, The- oret. and Math. Phys., 92, no. 3 (1992), 997–1019.

[MR] Molev, A.; Ragoucy, E.: The MacMahon Master Theorem for right quantum superal- gebras and higher Sugawara operators for gl(m|n), Moscow Mathematical Journal 14 (2014), 83–119; arXiv:0911.3447 [math.RT].

[Molev] Molev, A.: Sugawara operators for classical Lie algebras, Providence, RI: American Mathematical Society, (2018), 321 pp.

80 [Par] Pareigis, B.: A non-commutative non-cocommutative Hopf algebra in “nature”, J. Al- gebra 70 (1981), 356–374.

[Porst] Porst, H.–E.: On Categories of Monoids, Comonoids, and Bimonoids, Quaest. Math. 31 (2008), 127–139.

[RTF] Reshetikhin, N. Y.; Takhtadzhyan, L. A.; Faddeev, L. D.: Quantization of Lie groups and Lie algebras, Algebra i Analiz 1 (1989), No.1, 178–206 (in Russian); translation in: Leningr. Math. J. 1 (1990), No.1, 193–225.

[RST] Rubtsov, V.; Silantyev, A.; Talalaev, D.: Manin Matrices, Quantum Elliptic Com- mutative Families and Characteristic Polynomial of Elliptic Gaudin model, SIGMA 5 (2009), 110, 22 pages; arXiv:0908.4064 [math-ph].

[S] Silantyev, A.: Manin matrices for quadratic algebras, SIGMA 17, (2021), 066, 81 pages; arXiv:2009.05993 [math.QA].

81