Hochschild cohomology for functors on linear symmetric monoidal categories
Abstract
Let be a commutative ring with unit. We develop a Hochschild cohomology theory in the category of linear functors defined from an essentially small symmetric monoidal category enriched in -Mod, to -Mod. The category is known to be symmetric monoidal too, so one can consider monoids in and modules over these monoids, which allows for the possibility of a Hochschild cohomology theory. The emphasis of the article is in considering natural hom constructions appearing in this context. These homs, together with the abelian structure of lead to nice definitions and provide effective tools to prove the main properties and results of the classical Hochschild cohomology theory.
Keywords: Hochschild cohomology, enriched monoidal categories, linear functors.
AMS MSC (2020): 18M05, 18D20, 18G90.
1 Introduction
We consider an essentially small symmetric monoidal category, , enriched in -Mod with a commutative ring with unit, and then -linear functors from to -Mod. This category of functors, denoted by , is an abelian, symmetric monoidal, closed category, via Day’s convolution. Given a monoid in , we have then a category a -modules, -Mod. We refer the reader to [9] or [7] for the generalities on monoids and modules over them. The main example we have in mind when considering these hypothesis is as the biset category and as the category of biset functors (see, for example, [3]). Monoids in this case are called Green biset functors and have been extensively studied in the last years, in particular, in [4] the commutant and the center of a Green biset functor are studied. As we will see, the commutant of a Green biset functor is the Hochschild cohomology functor of degree zero. So, it is a natural question to ask for a Hochschild cohomology theory of Green biset functors. The purpose of the paper is to develop the main results and properties of this theory for monoids in , in order to apply them to biset functors in a forthcoming paper in collaboration with Serge Bouc.
Even though there are recent articles about Hochschild cohomology in monoidal categories (see for example [1] and [6]), none of them is suited for a direct application to our case. So, we take a different and new approach by making use of the internal hom functor in , defined in Section 3. The advantage of working with functors from to -Mod is that we can give an explicit construction of the internal hom functor, via the Yoneda-Dress construction (see chapters 2 and 3 of [7] for the abstract definition). With this, the definition of the Hochschild cochain complex of appears in a natural way in the category . In particular, we can give an explicit description of its arrows, in terms of these homs, which allows for a better understanding of the Hochschild cohomology functors, denoted by , for an -bimodule and a natural number.
The article is inspired in Loday’s presentation of the classical Hochschild cohomology theory (see [8]). So, in Section 4 we introduce the bar resolution of a monoid in and we define the Hochschild cochain complex of . In Section 5 we deal with the important example of a separable monoid. The main result is Theorem 5.6, in which we show that if is a separable monoid in , then, for any -bimodule , the functors are zero for very and, conversely, if is a monoid in such that the first Hochschild cohomology functor is zero for any -bimodule , then is separable. As a corollary, we obtain that the Hochschild cohomology functors, for , of the Burnside biset functor, , as well as those of the biset functor of rational representations, , over a filed of characteristic , are all zero.
Finally, in Section 6, we describe the Hochschild cohomology functors of degrees 0, 1 and 2. As in the classical case, the degree 1 can be described in terms of derivations and the degree 2 is described in terms of extensions. Also, following [5], for a monoid in , we endow the coproduct with a structure of a graded monoid and call it the Hochschild cohomology monoid. We try too keep the proofs of the statements in this last section as brief as possible, since there are many straightforward calculations that work as in the classical case.
2 Preliminaries
For an object in a category , we denote the identity morphism of simply as .
In what follows is a commutative ring with identity and is an essentially small symmetric monoidal category enriched in -Mod, in the sense of Definition 3.1.51 of [7]. In particular the functor is -bilinear. The category of -linear functors from to -Mod is denoted by . Recall that is an abelian, symmetric monoidal, closed category with identity given by (see for example Section 3.3 in [7]). We denote the tensor product of two objects and in as . With this, the complete notation for is . As it is customary, we will avoid, whenever is possible, the parenthesis of association in tensor product of several objects in .
Given a monoid in , we denote its product by , in case of confusion we may write instead of . We recall that we have a morphism , or , and the following commuting diagrams,
in .
In what follows and are monoids in . A morphism of monoids from to is an arrow in such that the following diagram commutes
and .
The tensor product , of and , is a monoid with product given by
where is the symmetry . When needed, as it was in this case, we will denote the symmetry with subindex.
A left -module is an object in together with an arrow such that the following diagrams commute
in . Right -modules are defined in an analogous way. Morphisms of -modules are defined in the obvious way and the subcategory of of left -modules is denoted by -Mod.
We denote the opposite of in as . That is, denotes the object with opposite monoidal structure, . The monoid is called commutative if it is equal to its opposite, i.e. .
Recall that -modules identify with right -modules and thus that -bimodules identify with -modules in the following way.
Remark 2.1.
If is an -bimodule, with compatible actions and , then the action of on is given by
The symmetry is actually , but in we consider with the opposite product, as above. On the other direction, if is an -module we obtain the actions of and by composing with and .
We denote the monoid by . We may indistinctly write -module, -bimodule or even -bimodule.
Example 2.2.
Let be the biset category for the class of finite groups over the ring , as defined in [3], and let be the category of biset functors, that is, -linear functors from to -Mod. Then , with the direct product of groups and the trivial group, satisfies the hypothesis of and so those of . In this case, the functor is the Burnside functor .
3 Hom functors
We begin by recalling the construction of the internal hom object in . The general definition of this object can be found in Section 3 of [7] but, since we are working with functors on -Mod, we follow the lines of [3], to give an explicit construction.
For each object , we have an -linear functor , given by the monoidal structure of , that is, it sends an object to and an arrow to . This in turn allows us to define the endofunctor called the Yoneda-Dress construction. In an object , it is defined as and in an arrow as , for an object of .
The following lemma is a generalization of Lemma 8.2.4. Its proof is straightforward.
Lemma 3.1.
Let , and be objects in .
-
1.
The functors and are naturally isomorphic.
-
2.
Given an arrow in , it induces a natural transformation . In a functor , the arrow , denoted by , is defined at an object in as .
-
3.
If and , then .
-
4.
The correspondence sending an object of to and an arrow in to is an -linear functor from to the category of -linear endofunctors of .
By point 4 of the previous lemma, for every object in , we have an -linear functor .
Remark 3.2.
For the biset category, the Yoneda-Dress construction at a group , satisfies also that is a self-adjoint functor. We will not need this extra condition in what follows.
Definition 3.3.
Let and be objects in , we denote by the object in defined by the following composition of functors
The correspondence is a bilinear functor, called the internal hom functor.
Lemma 3.4.
We have a Yoneda Lemma for the internal hom functor. If , then
Proof.
The known isomorphism of -modules, defines a natural equivalence between the functor and the evaluation functor, , going from to -Mod. By precomposition with the functor we obtain an equivalence between and in . Finally, by composing with the symmetry of , we obtain an equivalence between and . ∎
The rest of the section is devoted to define a hom functor for modules over monoids in . We begin by stating the universal property of tensor products. It is easy to see that the proof of Proposition 8.4.2 in [3] holds in our case, so Remark 8.4.3 in [3] can also be generalized.
Remark 3.5.
Let , and be objects in . Consider the functors and from to -Mod given by sending to and respectively, and in an obvious way in arrows. Then, there is a one-to-one correspondence between Hom and which is natural in every variable.
In other words, a natural transformation from to is given by a collection of -bilinear maps
satisfying obvious functoriality conditions.
Notice that this remark can be generalized to an arrow from a finite product to .
Let be a monoid in . The Yoneda Lemma,
together with the previous remark allows us to see the monoid as in Definition 8.5.1 of [3].
Let , and be objects in , consider , and . Then, is a monoid in if for every pair of objects in , there exist bilinear maps
denoted simply by , and an element satisfying the following conditions.
-
•
Associativity. For , and , objects in , and for any , and ,
-
•
Identity element. For an object in and any ,
-
•
Functoriality. Let and be arrows in . Then for any and ,
Also, a module over can be seen as an in Definition 8.5.5 of [3]. That is, a left -module is an object in such that for every pair of objects in , there exist bilinear maps
which we continue to denote by , satisfying corresponding conditions of associativity, identity element (on the left only) and functoriality. In a similar fashion we can define right modules and bimodules.
Morphisms of -modules can also be rewritten in these terms. Let and be -modules, a morphism of -modules from to is an arrow in such that for any objects and in , we have for all and .
The proof of the following lemma is straightforward.
Lemma 3.6.
Let be a monoid in and be an -module, then we have an isomorphism of -modules,
given by sending a morphism of -modules, , to .
Lemma 3.7.
Let and be objects in and be an arrow. Consider and as defined at the beginning of the section. Then defines a functor from to the category of -linear endofunctors of -.
Proof.
Let and be an object and an arrow in . By Lemma 3.1, we only need to verify that restricts from -Mod to -Mod and that the components of the transformation , as defined before, are morphisms of -modules.
Let be an -module, then
given by the original action of on , endows with a structure of -module. Also, if is a morphism of -modules, then it is immediate to see that , is a morphism of -modules. Hence we can restrict --Mod. Now let be an -module, the functoriality of the action of on shows that is a morphism of -modules. ∎
Remark 3.8.
As said in Section 2, is a right -module if and only if it is an -module. The two structures are related by the following equation
for , and where is the symmetry in . Now, if we begin with a right -module, , then it is an -module and then is an -module, as in the previous lemma. After some straightforward computations we see that its structure of right -module is given by
A word of caution is needed, since, in general we will write for the action of in any right -module regardless of whether it is shifted or not. However, if is explicitly a shifted module , we will write the corresponding symmetry.
Given two -modules and , we have that is an -submodule of . Moreover, for an -module , we have a functor --.
Definition 3.9.
Let and be objects in -Mod, we denote by the object in defined by
Lemma 3.10.
The correspondence -- is a bilinear functor.
Proof.
This follows from Lemma 3.7. ∎
Corollary 3.11.
For and as before, we have an isomorphism in ,
Proof.
The bijection of Lemma 3.6 defines a natural equivalence between the functor and the evaluation functor, , going from -Mod to -Mod. By precomposing with we obtain an equivalence between and , but clearly the last one is equivalent to .
∎
4 The Hochschild cochain complex
4.1 Bar resolution of a monoid
Let be a monoid in . We omit the morphism of diagram 1 and write for the tensor product of with itself times.
Consider the following sequence in ,
where is in degree zero and , for , is defined in the following way. Given , define , for , as
where appears in the position of the tensor product. That is, , if and . With this, . When considering the augmented complex, having in degree -1, with morphism and 0 in all degrees , we will write .
Lemma 4.1.
The sequence is a presimplicial object in , that is, the arrows defined above satisfy
Moreover, is a complex in .
Proof.
Notice that if , then we clearly have . On the other hand, the commutativity of diagram 1 implies the previous equality if . It also implies that . With this, the proof of is standard (see for example Lemma 1.0.7 of [8]). ∎
The structure of -bimodule of is given by diagram 1. Now consider with . The structure of -bimodule of is given by , which is equal to .
Proposition 4.2.
Let be a monoid in , the complex is a resolution of the -module . It is called the bar resolution of .
Proof.
We will first show that the complex is exact and then we will verify that it is a complex of -modules.
Using diagram 1, one can show that is an epimorphism in but we prefer to define the operators of extra degeneracy and show that the complex is acyclic. Consider first as the morphism
from diagram 2. Next, for , define
It is easy to see that if , then . On the other hand, diagram 2 shows that for any we have . This implies for and . Hence is a contracting homotopy and is acyclic.
Now we see that is a complex of -modules. The arrow is a morphism of -modules thanks to the associativity of the product, i.e. diagram 1. Next, let and with . To see that the action of commutes with , we verify it commutes with the arrows , that is, the commutativity of the following diagram
where the horizontal arrows correspond to the structures of -bimodules of and . Notice that . By Lemma 4.1 we have
Hence, is a complex of -modules.
∎
4.2 Hochschild cochain complex
Definition 4.3.
Let be a monoid in and be an -bimodule. The Hochschild cochain complex is the complex in given by . Since is an abelian category, we define the Hochschild cohomology of with coefficients in as the object in given by
It is easy to see that defines a functor from -Mod to .
The next definition generalizes Definition 18 in [4].
Definition 4.4.
Let be a monoid in and be an -bimodule. The commutant of at is
where is the symmetry in . If the -bimodule is itself, we keep the notation of [4].
Since is a natural isomorphism, it is easy to see that is a subfunctor of . Also, we will see in the last section that the functors are all -modules.
To describe the Hochschild cochain complex we need the following lemma.
Lemma 4.5.
Let be a monoid in , be an -bimodule and be an integer. We have the following isomorphism in ,
where, if , then denotes . Also, in .
Proof.
For the first statement, consider . Using Remark 2.1, it is easy to see that and are isomorphic as -modules. Then, by lemmas 3.4 and 3.11, we have the following isomorphisms in ,
Hence .
Now suppose and let be an -bimodule. Then, by Remark 3.5, an arrow , of -modules, is given by a corresponding collection of maps
linear and functorial in each entry. These maps also satisfy the following identity
for all objects , , , and and elements , , , and in the corresponding evaluations. In particular
and is functorial in and linear in the variable . Hence, if we consider the functors and from -Mod to -Mod, we obtain a natural isomorphism between them. Indeed, by sending to we define an isomorphism of -modules from to , which is clearly natural in . By precomposition with the functor -Mod, we obtain the result.
The second statement follows as in the classical case. That is, since is an epimorphism, we have a monomorphism in ,
defined in by sending to . Given that is a morphism of monoids and that , we have , hence . But, since is a morphism of -bimodules, it is easy to see that must be in . ∎
With this, the complex now looks like
with in degree zero. The second arrow from the left is just the inclusion of in and the arrows are described as follows.
Let be an object in . In the complex , an arrow in , with , is sent to in . We will see what this means in terms of the corresponding linear maps of Remark 3.5.
Suppose first . Denote by the -linear map that corresponds to by Remark 3.5. As in the previous lemma, the corresponding -linear map
is given by . In the variables we add the superscript to denote the position of in . If , with , is one of the morphisms defined in Section 4.1, then it is not hard to see that
sends an element to
Also,
and
Thus, if we let to be the arrow corresponding to and be the arrow in corresponding to , then
sends to
where the last symmetry comes from Remark 3.8.
For the case , we see immediately that is given by
5 Relative projectivity and separable monoids
Given a monoid in , we consider the functors
the first one is just the forgetful functor and the second one sends a functor to and an arrow to . It is easy to see that is a left adjoint of .
Following Definition 4.1 in [2] we say that an -module is projective with respect to the restriction if for any diagram
in -Mod such that is a split epimorphism, there exists a morphism in -Mod such that .
Now consider and . Lemma 4.6 in [2] translates in the following way.
Lemma 5.1.
For every object in -Mod there exists a resolution
where the are projective with respect to the restriction and such that
is an exact split complex. Moreover, such a resolution for is unique up to homotopy.
Remark 5.2.
This remark will also help us to prove the following result.
Lemma 5.3.
Let be a short exact sequence in -Mod,
such that is a split epimorphism. Then we have a short exact sequence of complexes in ,
Proof.
Observe first that, for an object in , the shifted sequence ,
is also a short exact sequence in -Mod. Also and for any -bimodule and any morphism of -bimodules . Finally, if is a split epimorphism, then is also a split epimorphism.
It is easy to see that we have a sequence of complexes in ,
Now, Let . The hom functor is clearly left exact, so to show that the sequence
(we abbreviate as ) is exact, it suffices to show that the morphism from to is surjective. Let . As in the previous remark, every with is projective with respect to , so, by definition, there exists in -Mod such that . ∎
We have immediately the following corollary.
Corollary 5.4.
If is a short exact sequence in -Mod as in the previous lemma, then we have a long exact sequence
in .
Since for any -bimodule , the functor is additive, we can calculate the Hochschild cohomology with coefficients in from any resolution that satisfies the conditions of Lemma 5.1, as it is the case in the following example.
Definition 5.5.
Let be a monoid in . We say that is separable if is a split epimorphism in -Mod.
Theorem 5.6.
If is a separable monoid in , then for any -bimodule we have for every . Conversely, if is a monoid in such that for any -bimodule , then is separable.
Proof.
If is separable, we have a split short exact sequence in -Mod
As in Remark 5.2, is projective with respect to the restriction and by Lemma 4.2 in [2], we have that is also projective with respect to . Hence, this short exact sequence gives a resolution for as in Lemma 5.1. Since this resolution is contractible, we have that the bar resolution is contractible and thus, the complex is contractible.
For the other direction, consider the short exact sequence in -Mod,
and let . By the first lines of the proof of Proposition 4.2, we have that is a split epimorphism. Then, by Corollary 5.4, we have the following exact sequence in
where the last zero corresponds to . But, by the last lines of the previous section, we have that for any -bimodule , hence, by Lemma 4.5, we have the following exact sequence in -Mod,
So, for the identity morphism, in , there exists a morphism such that . ∎
We will finish the section with an example but, first, we need the following lemmas. The first one generalizes what happens in the classical case.
Lemma 5.7.
Let be a monoid in . Then is separable if and only if there exists an element such that . Moreover, determines the morphism of separability .
Proof.
Suppose first that is separable. Then we have in -Mod that satisfies . Let . For an object in and , the action on the left gives
and on the right
but . This shows that and that it determines the morphism .
Now suppose we have as in the hypothesis of the lemma and define in an object as
where is the left action of on . Since this is equal to . This clearly defines a natural transformation from to and the functoriality of the bi-action of on shows that is a morphism of bimodules. Finally, we clearly have and thus . So, for any object in and any , since is a morphism of bimodules, we have
the last equality comes from the naturality of . Thus . ∎
Lemma 5.8.
Let and be monoids in and suppose is a morphism of monoids that is an epimorphism in . If is separable, then is separable.
Proof.
We notice first that (and we do not need it to be an epimorphism for this) induces a structure of -bimodule in , given by
Since is a morphism of monoids, it is easy to see it is also a morphism of -bimodules. Then clearly, is a -bimodule and is a morphism of bimodules.
Let be an element in as in the previous lemma and consider . We have
since is a morphism of monoids.
Now let be an object of and . Then there exists such that and the left action of is equal to . But this is precisely the left action of on and, since is a morphism of bimodules, we have that this is equal to
which, by functoriality, is equal to
which is equal to . Hence . ∎
Corollary 5.9.
For very object in and every , we have . Moreover if is a monoid such that there exists a morphism of monoids which is an epimorphism in , then for every -bimodule and every we have .
Example 5.10.
The previous corollary implies that for every biset functor and every , we have . Also, if we take as a field of characteristic and as the biset functor of rational representations, we know that is a commutative monoid and that the linearization morphism provides an arrow in as in the previous corollary. Hence, for every -module and every , we have .
6 Hochschild cohomology functors
Through this section is a monoid in , and are -bimodules and , , and are objects of .
Let , and be objects in . To simplify the notation, in what follows we will work indistinctly with and the corresponding set of bilinear maps given by Remark 3.5. In particular, given an arrow and in we will avoid the notation , used in Section 4.2.
6.1
As said before, in this case we clearly have
6.2
Definition 6.1.
An arrow is called a derivation from to if for any and , we have
The arrow is called an inner derivation if there exist such that the morphisms are given by sending to .
By the last paragraphs of Section 4.2, we have that the kernel of
corresponds to the derivations from to and that the image of corresponds to the inner derivations from to . In particular, we can define two subfunctors of in . First, that of derivations , defined in as the derivations from to , and then that of inner derivations , defined accordingly. With this
As in the classical case, we can endow with a Lie-type structure in the following way. If and , consider the arrow , given as
Defining as is well defined and the bracket satisfies the corresponding Lie-type axioms.
Remark 6.2.
Serge Bouc has suggested that if is a Green biset functor, then we can call a Lie biset functor.
6.3
Definition 6.3.
A square-zero extension of consists of a monoid in and a monoid homomorphism which is an epimorphism in and such that is an ideal of of square zero, that is, the bilinear maps
are zero for all and .
In this case, it is easy to see that becomes an -bimodule. If is isomorphic to , as -bimodules, we call this extension a square-zero extension of by .
Definition 6.4.
A square-zero extension of is called a Hochschild extension if it is a split extension in , i.e., there exists an arrow in such that .
Two Hochschild extensions, and , of by are said to be equivalent if there exists a monoid morphism such that the squares in the following diagram
commute. This defines an equivalence relation in the collection of all the Hochschild extensions of by , we denote the collection of equivalence classes by .
Notice that is not empty, since we have the class of the semidirect product . As an object in , it is given by the coproduct of with , that is . As a monoid, the product is given by
for and .
Theorem 6.5.
There exists a bijection
Proof.
By the last paragraphs of Section 4.2, we have that the kernel of
consists of arrows that satisfy that for any in , the expression
is equal to zero.
For such an , we consider in with the following product
for and . The equality to zero of the expression above allows us to show that this product is associative. Also, it shows that if we take , then is the identity element for the product. Finally, it is easy to verify that this product is functorial. With this, it is immediate to see that the extension
is a Hochschild extension. Now consider , where and
Then, the extensions given by and are equivalent. Indeed, the isomorphism of monoids is given by sending to .
On the other direction, given a Hochschild extension
since there exists in such that , then in . Hence, we can obtain the monoid structure of from this direct sum,
for and . To simplify the notation, let , , and . Then, since the product is bilinear and is of square zero, it is easy to see that the product above is equal to
where is the unique element in that corresponds to the product . This allows us to define a map sending to . In turn, this defines an arrow and the associativity of the product of implies that is in the kernel of . Finally, suppose we have an equivalent extension
with splitting morphism . Let be the equivalence morphism. Then, as before, for there exist a unique that corresponds to . Hence we can define by sending to . This defines an arrow in . Also, using the fact that is a morphism of monoids, if we take the morphism defined before corresponding to , then we obtain . ∎
Corollary 6.6.
is a functor in .
Proof.
Indeed, since is an -module, the previous bijection endows with a structure of -module. The functorial structure is also given by the bijection. That is, given an arrow in , then the map
sends (the class of) the extension , corresponding to the cocycle , to . ∎
Remark 6.7.
It is not hard to see that the sum we obtain in , by the previous corollary, corresponds to the Baer sum of extensions. Given two Hochschild extensions, and , of by , their Baer sum is defined in by taking the pullback of and , that is
and then making the quotient over , if and are the corresponding morphisms. It is defined in an obvious way in arrows. This construction yields a Hochschild extension of by .
6.4 The Hochschild cohomology monoid
Definition 6.8.
An -graded monoid in is a lax monoidal functor , where is seen as a discrete monoidal category.
That is, an -graded monoid consists of the following:
-
•
For each , a functor in .
-
•
For every and every and objects in , a bilinear map
that is associative in an obvious way and functorial in and (analogous conditions to those appearing in Section 3 for a monoid).
-
•
An element such that for every and ,
In this case, is a monoid in and every is an -bimodule.
Since is an abelian category, we can consider the coproduct
in . We will endow with a structure of graded monoid. We begin by defining, for , a cup product,
sending , with and , to
which sends an -tuple to
here is the symmetry in sending to .
This product makes an -graded monoid in , with identity . Next, consider the complex with morphisms as in Section 4.2. A series of straightforward computations (in particular one must pay attention to the symmetries that appear in the expressions) show that for any and and as before,
This formula allows us to extend the cup product to , just as in the classical case. Also, it is easy to see that the product satisfies the conditions for a graded monoid, with .
Remark 6.9.
If is a Green biset functor, then we can call a graded Green biset functor.
Remark 6.10.
The cup product just described works exactly the same if we consider first
via the isomorphism and then
This shows that each is a -module.
6.5 Further results
The author believes that the following results, not considered in this paper, can also be extended to our context.
-
i)
The description of in terms of crossed bimodules (see E.1.5.1 in [8]).
-
ii)
The definition of a bracket giving a structure of graded Lie monoid, with the grading shifted by -1, as in the classical case.
Acknowledgments
The contents of this article were developed during a sabbatical year the author did at the laboratory LAMFA of the Université de Picardie, in Amiens, France, from August 2022 to July 2023. The author thanks the staff and colleagues at LAMFA for all the support she received during her stay, which led to the successful development of the project. Special thanks to Serge Bouc, host during the sabbatical, for all the ideas, suggestions and stimulating conversations.
References
- [1] A. Ardizzoni, C. Menini, and D. Ştefan. Hochschild cohomology and “smoothnes” in monoidal categories. Journal of pure and applied algebra, 208:297–330, 2007.
- [2] S. Bouc. Résolutions de foncteurs de Mackey. In Group representations: cohomology, group actions and topology (Seattle, WA, 1996), volume 63 of Proc. Sympos. Pure Math., pages 31–83. Amer. Math. Soc., Providence, RI, 1998.
- [3] Serge Bouc. Biset functors for finite groups. Springer, Berlin, 2010.
- [4] Serge Bouc and Nadia Romero. The center of a Green biset functor. Pacific J. Math., 303(2):459–490, 2019.
- [5] Murray Gerstenhaber. The cohomology structure of an associative ring. Annals of mathematics, 78:267–288, 1963.
- [6] Magnus Hellstrøm-Finnsen. Hochschild cohomology of ring objets in monoidal categories. Communications in algebra, 46:5202–5233, 2018.
- [7] Michael A. Hill, Michael J. Hopkins, and Douglas C. Ravenel. Equivariant stable homotopy theory and the Kervaire invariant problem, volume 40 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2021.
- [8] Jean-Louis Loday. Cyclic homology. Springer-Verlag, Berlin, 2nd edition, 1998.
- [9] Saunders Mac Lane. Categories for the working mathematician. Springer, Berlin, 1971.